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 140 毫秒
1.
Copolymers of p-vinylphenol were prepared in bulk with heptafluorobutyl and pentadecafluorooctyl acrylates and trifluoroethyl, hexafluoroisopropyl, heptafluorobutyl, octafluoropentyl and pentadecafluorooctyl methacrylates using azobisisobutyronitrile as the initiator in sealed tubes. Intrinsic viscosities of the copolymers ranged from 0.44 to 1.85. Monomer reactivity ratios for copolymers of trifluoroethyl methacrylate (M1) were: with hydroxyethyl methacrylate (M2), r1 = 0.47, r2 = 1.0; with methyl methacrylate (M2), r1 = 0.82, r2 = 0.50; with styrene (M2), r1 = 0.29, r2, = 0.20; and with p-vinylphenol (M2), r1 = 0.096, r2 = 1.5. Q and e values of trifluoroethyl methacrylate were 1.30 and 0.92, respectively. Monomer reactivity ratios of octafluoropentyl methacrylate (M1) were: with styrene (M2), r1 = 0.26, r2 = 0.20; and with p-vinylphenol, r1 = 0.21, r2 = 1.5. Q and e values for octafluoropentyl methacrylate were 1.27 and 0.92, respectively. Critical surface tensions of the homopolymers ranged from 17.9 to 14.8 dyn/cm. A copolymer of hexafluoro-i-propyl methacrylate and p-vinylphenol exhibited a critical surface tension of 16.5 dyn/cm.  相似文献   

2.
To test the effect of NH−C=S groups (Scheme 1) on the stability of β-peptide secondary structures, we have synthesized three β-thiohexapeptide analogues of H-(β-HVal-β-HAla-β-HLeu)2-OH ( 1 ) with one, two, and three C=S groups in the N-terminal positions (cf. 2 – 4 and model in Fig. 1). The first C=S group was introduced selectively by treatment with Lawesson reagent of Boc-β-dipeptide esters ( 6 and 8 ). A series of fragment-coupling steps (with reagents as for the corresponding sulfur-free building blocks) and another thionation reaction led to the title compounds with a C=S group in residues 1, 1, and 3, as well as 1, 2, and 3 of the β-hexapeptide (Schemes 2 and 3). The sulfur derivatives, especially those with three C=S groups, were much more soluble in organic media than the sulfur-free analogues (>1000-fold in CHCl3; Table 1). The UV and CD spectra (in CHCl3, MeOH, and H2O) of the new compounds were recorded and compared with those of the parent β-hexapeptide 1 (Figs. 2 – 4); they indicate the presence of more than one secondary structure under the various conditions. Most striking is a pronounced exciton splitting (Δλ ca. 20 nm, amplitude up to +121000) of the ππ*C=S band near 270 nm with the β-trithiohexapeptide (with and without terminal protecting groups), and strong, so-called `primary solvent effects', in the CD spectra. The CD spectrum of the β-dithiohexapeptide 3 undergoes drastic changes upon irradiation with 266-nm laser light of a MeOH solution (Fig. 5). The NMR structure in CD3OH of the unprotected β-trithiohexapeptide 4 was determined to be an (M)-314-helix (Fig. 7), very similar to that of the non-thionated analogue (cf. 1 ). NMR and mass spectra of the β-hexapeptides with C=S and with C=O groups are compared (Figs. 6 and 8).  相似文献   

3.
Using a Monte Carlo simulation in three dimensions, we studied the variation of the root-meansquare (rms) displacement (Rrms) of polymer chains with time and the rates of their mass transfer (j) as a function of biased field (B), polymer concentration (p), chain length (Lc), porosity (ps), and temperature (T). In homogeneous/annealed system, the rms displacement of the chains shows a drift-like behavior, Rrmst, in the asymptotic time regime preceded by a subdiffusive power-law (Rrmstk, with k < 1/2) at high p. The subdiffusive regime expands on increasing Lc and p but reduces on increasing T or B. In quenched porous media, the drift-like behavior of Rrms persists at low barrier concentration (pb) and high T. However, at high pb and/or low T, chains relax into a subdrift and/or subdiffusive behavior especially with high p or long Lc. Flow of chains is measured via an effective permeability (σ) using a linear response assumption. In annealed system, σ increases monotonically with B at high T and low p but varies nonmonotonically at low T, high p and high Lc. We find that σ decays with Lc, σ ∼ L, where α depends on B, p and T with a typical value a α ∼ 0.43−0.64 for p = 0.1-0.3 at B = 0.5. Further, σ decays with p, σ ∼ − Cp with a decay rate C sensitive to T and B. In quenched porous media, even at low pb and high T, σ varies nonmonotonically with bias, i.e., the increase of σ is followed by decay on increasing the bias beyond a characteristic value (Bc). This characteristic bias seems to decrease logarithmically with barrier concentration, Bc ∼ −klnpb. The prefactor k depends on the chain length, k ≈ 0.35 for shorter chains (Lc = 20, 40) and ≈ 0.15 for longer chains (Lc = 60). Scaling dependence of σ on Lc similar to that in annealed system is also observed in porous media with different values of exponent α. The current density shows a nonlinear power-law response, jBσ, with a nonuniversal exponent δ ≈ 1.10−1.39 at high temperatures and low barrier concentrations.  相似文献   

4.
Glass transition temperatures have been determined for polystyrenes crosslinked with 1–10% divinylbenzene and swollen with toluene, chloroform, N,N-dimethylformamide, and tetrahydrofuran to as high as 0.7 weight fraction solvent. The Tg′s depend approximately on the weight fractions and the Tg′s of the components according to the empirical equation 1nTg = m1 1nTg1 + m2 1nTg2 of Pochan. The Tg′s of the networks swollen with toluene also fit approximately a quasithermodynamic equation of Karasz based on the Tg′s and the ΔCp′s at Tg of the components.  相似文献   

5.

Mono(thio)substituted dienes 1 gave 3a–g , 5 , and 7 with piperazine derivatives in dichloromethane. Hexachlorobutadiene 14 in a water-ethanol mixture in the presence of sodium hydroxide reacted with thiol 15 to give the mono(thio)substituted thioether 16 and di(thio)substituted thioether 17 . 18 was obtained from the reaction of 16 with m-CPBA in chloroform. 9 was obtained from the reaction of l,2,3,4,4-pentachloro-(1-2-hydroxyethylthio)-1,3- butadiene 8 with 47% HI, and 11 was synthesized from the reaction of 8 with concentrated H2SO4 and KBr. Compounds 9 and 11 gave in the reaction with m-CPBA in chloroform 10 , 12 , and 13 , respectively.  相似文献   

6.
Hydrostannylation reactions of the phosphaalkenes 9,11, and 21 with the triorganotin hydrides 1 proceed by different routes. Whereas the trior-ganotin hydrides 1a,b undergo regioselective 1,2-addition to the P/C double bond of the P-aminophosphaalkene 9 to furnish the 2-stannylphosphanes 17a,b, the 1,2-addition products to the P-halophosphaalkenes 11 and 21 can only be postulated as the reactive intermediates 20 and 23, respectively. The reactions of 11 with 1a,b proceed with cleavage of the triorganotin halide via the diphosphene 15 to furnish the cyclophosphanes 18 and 19. On the other hand, the hydrostannylation reactions of the phosphaalkene 21 are not selective, and the 1,3-diphosphetane 22 is isolated as one of the reaction products. © 1998 John Wiley & Sons, Inc. Heteroatom Chem 9:453–460, 1998  相似文献   

7.
The C3‐symmetric propeller‐chiral compounds (P,P,P)‐ 1 and (M,M,M)‐ 1 with planar π‐cores perpendicular to the C3‐axis were synthesized in optically pure states. (P,P,P)‐ 1 possesses two distinguishable propeller‐chiral π‐faces with rims of different heights named the (P/L)‐face and (P/H)‐face. Each face is configurationally stable because of the rigid structure of the helicenes contained in the π‐core. (P,P,P)‐ 1 formed dimeric aggregates in organic solutions as indicated by the results of 1H NMR, CD, and UV/Vis spectroscopy and vapor pressure osmometry analyses. The (P/L)/(P/L) interactions were observed in the solid state by single‐crystal X‐ray analysis, and they were also predominant over the (P/H)/(P/H) and (P/L)/(P/H) interactions in solution, as indicated by the results of 1H and 2D NMR spectroscopy analyses. The dimerization constant was obtained for a racemic mixture, which showed that the heterochiral (P,P,P)‐ 1 /(M,M,M)‐ 1 interactions were much weaker than the homochiral (P,P,P)‐ 1 /(P,P,P)‐ 1 interactions. The results indicated that the propeller‐chiral (P/L)‐face interacts with the (P/L)‐face more strongly than with the (P/H)‐face, (M/L)‐face, and (M/H)‐face. The study showed the π‐face‐selective aggregation and π‐face chiral recognition of the configurationally stable propeller‐chiral molecules.  相似文献   

8.
Proton spin-spin relaxation times have been measured as a function of temperature for ultradrawn polypropylene with draw ratios λ up to 24. The three relaxation times T2a (the longest), T2i (intermediate), and T2c (the shortest), observed for all the samples, have been ascribed to the relaxations of the amorphous, constrained amorphous, and crystalline components, respectively. T2i and T2a, which reflect the changes in structure and mobility in the noncrystalline regions, decrease with increasing λ; T2i becomes saturated at λ > 9, whereas T2a shows a substantial decrease up to λ = 24. The continued decrease in T2a indicates that the constraint on the amorphous segments keeps increasing up to the highest λ. The associated mass fractions Fa, Fi, and Fc also change with λ. At λ < 9, the increasc in Fi with increasing λ is accompanied by a decrease in Fa, with Fc remaining unchanged. At higher λ, however, Fa is almost constant, and stepwise rises in Fc at about λ = 12 and 24 are accompanied by corresponding drops in Fi. It seems that, in this high draw ratio range, some of the taut molecules are fully extended and are in sufficiently good lateral register to transform into crystalline bridges. This conjecture is supported by the similarity in the λ dependence of Fc and the mass-fraction crystallinity obtained from the heat of fusion.  相似文献   

9.
The reaction of thiocarbohydrazide with carboxylic acids at the melting temperature allows an improved preparation of 5-substituted-4-amino-3-mercapto 1,2,4-triazoles 1 a ? g . Compound 1 a reacted with 2-bromopropionic acid to give acid derivative 2 . The latter was reacted with a mixture of acetic anhydride and triethylamine to afford the mesoionic compound 3 . Heating of compound 3 in ethanol gave the ester derivative 4 , which on alkaline hydrolysis in methanol gave ketone derivative 5 . Substituted 1,2,4-triazolo [3,4-b]-6H-1,3,4-thiadiazine 6 h,i and 7 were synthesized by reaction of 1 a with acetylacetone, ethyl acetoacetate and chloroacetamide. Heterocyclic systems 8 and 9 were prepared through the reaction of 1 a with 2,3-dichloro-1,4-naphthoquinone and 2,3-dichloroquinoxaline. In addition, thenoyl isothiocyanate, thenoyl chloride, 2-thiophenecarbaldehyde, and p-chlorophenyl isocyanate reacted with compound 1 a to afford 1,2,4-triazolo[3,4-b]-1,3,4-thiadiazole ring system 10 , 11 , and urea derivative 12 . 1,2,4-Triazolo[3,4-b]-5H-pyrazole derivatives 14 j,k were prepared through the reaction of compound 1 a with 3-chloro-2,4-pentandione and ethyl-2-chloroacetoacetate. Compound 14 j was treated with hydrazine to afford products 15 , 16 , and 17 depending on the type of hydrazine derivative and reaction conditions. Compound 19 was synthesized by refluxing of compound 14 j with hydroxylamine hydrochloride to afford the corresponding oxime derivative 18 followed by treatment with thenoyl chloride.  相似文献   

10.
The spread s(G) of a graph G is defined as s(G) = max i,j i − λ j |, where the maximum is taken over all pairs of eigenvalues of G. Let U(n,k) denote the set of all unicyclic graphs on n vertices with a maximum matching of cardinality k, and U *(n,k) the set of triangle-free graphs in U(n,k). In this paper, we determine the graphs with the largest and second largest spectral radius in U *(n,k), and the graph with the largest spread in U(n,k).   相似文献   

11.
《Chemphyschem》2002,3(12):1014-1018
In general, sensitization of lanthanide(III ) ions by organic sensitizers is regarded to take place via the triplet state of the sensitizers. Herein, we show that in dansyl‐ and lissamine‐functionalized Nd3+ complexes energy transfer occurs from the singlet state of the sensitizers to the Nd3+ center. No sensitized emission was observed in the corresponding complexes with Er3+, Yb3+, and Gd3+ ions. Furthermore, the fluorescence of the sensitizers was quenched only in the Nd3+ complex and not in the complexes with the other ions. Only Nd3+ centers can accept energy from the singlet state of the dyes, because the excited states of Nd3+ have a high spectral overlap with the fluorescence of the dansyl and lissamine sensitizers, and because the selection rules allow a fast energy transfer, which apparently is competitive with the fluorescence.  相似文献   

12.
20S-Protopanaxadiol (3β,12β,20S-trihydroxydammar-24-ene) 3-, 12-, and 20-O-β-D-galactopyranosides were synthesized for the first time. Condensation of 12β-acetoxy-3β,20S-dihydroxydammar-24-ene (1) and 2,3,4,6-tetra-O-acetyl-α-D-galactopyranosylbromide (α-acetobromogalactose) (2) under Koenigs–Knorr conditions with subsequent removal of the protecting groups resulted in regio- and stereoselective formation of 20S-protopanaxadiol 3-O-β-D-galactopyranoside, an analog of the natural ginsenoside Rh2. Glycosylation of 12β,20S-dihydroxydammar-24-en-3-one (5) by 2 with subsequent treatment of the reaction products with NaBH4 in isopropanol and deacetylation with NaOMe gave 20S-protopanaxadiol 12- and 20-O-β-Dgalactopyranosides.  相似文献   

13.
Crosslinks are introduced by γ irradiation into 1,2-polybutadiene while strained in uniaxial extension near Tg with stretch ratio λ0, thereby trapping a proportion of the entanglements originally present. The stress at any subsequent strain λ is accurately given by the sum σN + σx, where σN is the stress contributed by a trapped entanglement network with λ = 1 as reference and a Mooney–Rivlin stress-strain relation, and σx is that contributed by a crosslink network with λ = λ0 as reference and neo-Hookean stress-strain relation. The birefringence is accurately given as δn = ?NσN + ?xσx, where the ?'s are the respective stress-optical coefficients. From measurements at λ = λ0 where σx = 0, ?N can be determined separately. For polymer with 88% 1,2 microstructure, ?N and ?x are nearly equal and independent of irradiation dose, though strongly dependent on temperature. For polymer with (95–96)% 1,2, ?N and ?x are different (even opposite in sign) and dependent on dose. This behavior is associated with a side reaction of cyclization by the γ irradiation, which is inhibited by the 1,4 moiety in the polymer with lesser 1,2 content. It is responsible for residual birefringence in the state of ease (λ = λs) where σN = –σx and the stress is zero.  相似文献   

14.
A new dinuclear manganese complex, [Mn2L2(N3)2], was prepared by reaction of bis-Schiff base N,N'-bis(5-fluoro-2-hydroxybenzylidene)ethane-1,2-diamine (H2L) with manganese acetate and sodium azide in methanol. Both the Schiff base and the complex were characterized by physico-chemical methods and single crystal X-ray determination. The Schiff base crystallized in the orthorhombic space group Pbca with unit cell dimensions a?=?7.3191(7)?Å, b?=?6.0948(6)?Å, c?=?35.382(3)?Å, V?=?1578.3(3)?Å3, Z?=?4, R1?=?0.0481, wR2?=?0.1488. The manganese complex crystallized in the monoclinic space group P21/c with unit cell dimensions a?=?8.802(1)?Å, b?=?14.928(2)?Å, c?=?14.478(2)?Å, β?=?105.517(2)°, V?=?1833.0(4)?Å3, Z?=?2, R1?=?0.0768, wR2?=?0.1640. There are crystallographic inversion centers in both the ligand and the complex. The ligand coordinates to Mn through all the phenolate oxygens and imino nitrogens. The two Mn in the complex are bridged by two phenolate oxygens with separation of 3.549(1)?Å. Each Mn is octahedral with four donors of the Schiff base ligand defining the equatorial plane, and with one azido nitrogen and one phenolate oxygen of an adjacent Schiff base ligand occupying the axial positions. The Schiff base and the complex were tested in vitro for their antibacterial activities.  相似文献   

15.
Azo pigment yellow 14 (P.Y.14) was encapsulated into copolymer of styrene and maleic acid (PSMA) via phase separation technique followed by the preparation of composite dispersions. Herein, we mainly investigate its rheological properties. Our results showed that the apparent viscosity (n a ) of composite dispersion first decreased and then increased with an increase of molar content of maleic acid in PSMA (F M ), intrinsic viscosity of PSMA ([n]), and the weight ratio of PSMA to P.Y.14 (R C/P ), respectively. The composite dispersion with low n a was more close to Newtonian fluid when F M , [n] and R C/P were equal to 0.53, 79.65 mL/g, and 12%, respectively. n a of the composite would increase with increasing the pH value, and first decreased and then increased with a raising of the electrolyte and alcohol concentration, respectively, especially with AlCl3 and glycerol.  相似文献   

16.

The refluxing of 3-amino-6,8-dibromo-2-thioxo-2,3-dihydro-1H-quinazolin-4-one (5) with ethyl chloroformate and/or ethyl chloroacetate afforded compounds 6 and 7 . The reaction of 5 with ethyl bromobutyrate, chloroacetyl chloride, phenacyl chloride, and phenyl isocyanate yielded compounds 8 , 9 , 11 , and 12 . The coupling of 5 with (2,3,4,6-tetra-O-acetyl-α -D-gluopyranosyl)bromide( ABG ) in DMF at r.t. gave 3-amino-6,8-dibromo-2-(2′,3′,4′,6′-tetra-O-acetyl-β-D-glucopyranosyl)thioxo-2,3-dihydro-1H-quinazolin-4-one ( 14 ). The deblocking of 14 in sodium methoxide gave 5 . 3-Amino-6,8-dibromo-2-methylthio-3H-quinazolin-4-one ( 16 ) was prepared by stirring 5 with methyl iodide in methanol. The treatment of 16 with hydrazine hydrate afforded 4 . The condensation of 4 with aldehydes furnished 3,5-dibromo-2-arylaminobenzoic acid hydrazide ( 18a–c ). The refluxing of 18a with acetic anhydride gave 3-(benzylideneamino)-6,8-dibromo-2-methyl-3H-quinazolin-4-one ( 19 ). Hydrazones 20a–f were prepared by the condensation of 4 with pentoses and/or hexoses. The acetylation of ( 20a–f ) with acetic anhydride gave the acetyl derivatives 21a–f .  相似文献   

17.
The polymerization of N-vinylcarbazole (NVC) initiated by PhMgBr in benzene was studied at 32°C. Rp is second order with respect to PhMgBr concentration but increases with NVC concentration (up to 0.06 M) and falls thereafter. Rp and P n are depressed by the addition of thiophene and water. Modifiers such as benzaldehyde, butanone, and ethylene glycol practically inhibit the polymerization. Carbon tetrachloride and carbon dioxide, when passed through the NVC solution first, enhance the Rp and P n increases with increasing PhMgBr and NVC concentrations, respectively. Rp increases with temperature, but P n shows a maxium at a certain temperature. A cationic mechanism has been proposed where the polymerization is initiated by RMg+ cations produced from the ionization of PhMgBr by the Ashby and Smith mechanism.  相似文献   

18.
Several solid phases with the general formula xM[XHgSO3yHgX2·zMX·nH2O were obtained from aqueous solutions during phase formation studies in the systems M2SO3/HgX2 (M = NH4, K; X = Cl, Br). All phases were structurally characterized on the basis of single crystal X‐ray diffraction data and adopt new structure types. Compounds with x, y, z = 1 and n = 0 are isostructural (structure type I ) and crystallise with two formula units in space group P21/m and lattice parameters of a ≈ 9.7, b ≈ 6.2, c ≈ 10.4Å, β ≈ 111°. Compounds with x, y = 1 and z, n = 0 (structure type II ) crystallize in space group Cmc21 with four formula units and lattice parameters of a ≈ 5.9, b ≈ 22.0, c ≈ 6.9Å. The structures with x = 2, y, z = 1 and n = 0 are likewise isostructural (stucture type III ) and consist of four formula units in space group Pnma with lattice parameters of a ≈ 22.2, b ≈ 6.1, c ≈ 12.4Å. K[HgSO3Cl]·KCl·H2O is the only representative where x = 1, y = 0, z = 1 and n = 1 (structure type IV ). It is triclinic (space group ) with four formula units and lattice parameters of a = 6.1571(8), b = 7.1342(9), c = 10.6491(14) Å, α = 76.889(2), β = 88.364(2), γ = 69.758(2)°. Characteristic for all structures types is the segregation of the M+ cations and the anions and/or HgX2 molecules into layers. The [XHgSO3] anions are present in all structures and have m symmetry, except for K[HgSO3Cl]·KCl·H2O with 1 symmetry (but very close to m symmetry). The different [XHgSO3] units exhibit very similar Hg‐S distances (average 2.372Å) and are more or less bent with ∠(X‐Hg‐S) angles ranging from 159.7 to 173.7°. The molecular HgX2 entities present in structure types I ‐ III deviate only slightly from linearity with ∠(X‐Hg‐X) angles ranging from 174 to 179°. The structures are stabilised by interaction of the K+ or NH4+ cations that are located between the anionic layers or in the vacancies of the framework, by K‐O contacts or, in case of ammonium compounds, by medium to weak hydrogen bonding interactions of the type N‐H···O.  相似文献   

19.
The glass transition temperature Tg of nylon 6 decreases monotonically toward a finite value Tgl upon increase of the moisture content. The mechanism of this decrease entails the reversible replacement of intercaternary hydrogen bonds in the accessible regions of the polyamide. The limiting glass transition temperature Tgl is approached when the moisture content approaches Wl, which corresponds to the amount of water required for complete interaction with all accessible amide groups. Denoting with Tg0 the glass transition temperature of the dry polymer, the effect of water on Tg is represented by the equation, Tg = (ΔTg)0 exp{?[ln(ΔTg)0]W/τWl} + Tgl, where (ΔTg)0 = Tg0 ?Tgl, and τ = W(Tgl+1)/Wl. This equation appears to be generally applicable to hydrophilic polymers, since correspondingly calculated data are also in very good agreement with experimental data for polymers such as nylon 66, poly(vinyl alcohol), and polyN-vinylpyrrolidone. The effect of water of Young's modulus E of nylon 6 is represented by an analogous relationship, and the quantity In[(E?El)/(Tg?Tgl)] is a linear function of the moisture content.  相似文献   

20.
The octahedral title complex, [PtI4(C2H6S)2] or trans‐PtI4(dms)2 (dms is dimethyl sulfide), crystallizes in the monoclinic space group P21/n (Z = 2), with molecular symmetry Ci, which is the most frequently occurring point group for trans‐PtX4L2 {56%, 28 structures in the Cambridge Structural Database (CSD) [Allen (2002). Acta Cryst. B 58 , 380–388]}, followed by C1 (22%, 11 structures). The complexes form a puckered pseudo‐hexagonal layer in the (10) plane, and the layers are stacked with an interplanar distance of 7.10 Å. Density functional theory (DFT) calculations on an isolated complex with the observed parameters as a starting structure converged to C2h. Constraints to Ci on the observed geometry give 3–4 kJ mol−1 higher energy compared with C2h. DFT calculations on [PtCl4(PzH)2] (PzH is pyrazole), reported in the CSD in both the cis and trans forms, show an energy difference of 21 kJ mol−1 in favour of the trans complex. A CSD search for PtX4L2‐type complexes, where X is a halogen and L is a ligand with a donor atom from group 14, 15 or 16, indicated a preferred trans geometrical arrangement, with a total fraction of 68%. The dominating crystal packing operators for the trans complexes are an inversion centre combined with a screw axis/glide plane (48%), followed by an inversion centre alone (28%).  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