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 593 毫秒
1.
The ratio of the intensities of excimer and monomer emission, ID/IM, has been measured for Py? COO? (CH2)m? OOC? Py (m = 2–6, Py = 1-pyrenyl) in solvents of different viscosity, η. These compounds are dimeric models for the corresponding polymers, selected to eliminate the influence of energy migration on the population of an excimer. The solvents are H(CH2)nOH, n = 1–7, and ethylene glycol. The values of ID/IM decrease as η increases, and the rate of this decrease is a function of m. Therefore the qualitative dependence of ID/IM on m is a function of η. In the limit as η → ∞, ID/IM is an odd–even function of m for m = 2–5, and the values at m = 5 and 6 are nearly identical. A rotational-isomeric state analysis of the conformations of the diesters can rationalize the appearance of the odd–even effect. This result is insensitive to assumptions about the extent of overlap of the two pyrene ring systems that is required for the production of emission in the excimer band. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
Steady‐state fluorescence was used to measure the ratio of emission intensities, denoted ID/IM, for excited state complexes and excited monomers of five trichromophoric compounds, 2‐naphthyl‐COO‐(CH2)m‐OOC‐2,6‐dinaphthyl‐COO‐(CH2)m‐OOC‐2‐naphthyl, m = 2–6. The linear aliphatic alcohols H(CH2)nOH, n = 1–7, as well as mixtures of ethylene glycol and methanol, were used to change the viscosity of the medium, η. The values of ID/IM depend on η and m. A Rotational Isomeric State model and Molecular Dynamics simulations were used for interpretation of the experimental results. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 253–266, 1999  相似文献   

3.
Steady-state fluorescence measurements and molecular dynamics simulations have been used to study the intramolecular formation of excimers in five model compounds for polyesters containing naphthalene groups separated by flexible spacers. The model compounds are derived from 2-hydroxynaphthalene and HOOC (CH2)n COOH, n = 2–6. The ratio of the intensity of excimer and monomer emissions, ID/IM, is nearly independent of the viscosity of the medium, η, over the range covered in dilute solution. Although ID/IM is always very small, it shows an odd–even effect for the first four members of the series, with maxima when n is odd. Molecular dynamics simulations provide an explanation for the small values of ID/IM, their weak dependence on η, and the trend of ID/IM with n. The results for the present series of model compounds are compared with previous work, which reported larger values of ID/IM, and a stronger dependence of ID/IM on η, for bichromophoric compounds derived from 2-naphthoic acid and aliphatic glycols, where the direction of the ester groups is reversed. The origin of the difference in the behavior of ID/IM in the two series is identified. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35: 1127–1133, 1997  相似文献   

4.
Data on the viscosity η of moderately concentrated solutions of polystyrene are reported. Several solvents were investigated, including cyclopentane solutions over a temperature span between θU = 19.5°C and θL = 154.5°C. The data were analyzed in terms of a relation giving η as a function of αφM, where αφ is the expansion factor for the chain dimensions in a solution with volume fraction φ of polymer with molecular weight M. It is shown that values of αφ so determined decrease as ? lnαφ/? lnφ = (1 ? 2μ)/6μ for φ greater than φ* = 0.2M/s3 for moderately concentrated solutions, where s is the root-mean-square radius of gyration and μ = ? ln[η]/? lnM with [η] the intrinsic viscosity.  相似文献   

5.
N–Isopropylacrylamide (NIPAM) was polymerized using 1‐pyrenyl 2‐chloropropionate (PyCP) as the initiator and CuCl/tris[2‐(dimethylamino)ethyl]amine (Me6TREN) as the catalyst system. The polymerizations were performed using the feed ratio of [NIPAM]0/[PyCP]0/[CuCl]0/[Me6TREN]0 = 50/1/1/1 in DMF/water of 13/2 at 20 °C to afford an end‐functionalized poly(N‐isopropylacrylamide) with the pyrenyl group (Py–PNIPAM). The characterization of the Py–PNIPAM using matrix‐assisted laser desorption ionization time‐of‐flight mass spectrometry provided the number–average molecular weight (Mn,MS). The lower critical solution temperature (LCST) for the liquid–solid phase transition was 21.7, 24.8, 26.5, and 29.3 °C for the Py–PNIPAMs with the Mn,MS's of 3000, 3400, 4200, and 5000, respectively; hence, the LCST was dramatically lowered with the decreasing Mn,MS. The aqueous Py–PNIPAM solution below the LCST was characterized using a static laser light scattering (SLS) measurement to determine its molar mass, Mw,SLS. The aqueous solutions of the Py–PNIPAMs with the Mn,MS's of 3000, 3400, 4200, and 5000 showed the Mw,SLS of 586,000, 386,000, 223,000, and 170,000, respectively. Thus, lowering the LCST for Py–PNIPAM should be attributable to the formation of the PNIPAM aggregates. The LCST of 21.7 °C for Py–PNIPAM with the Mn,MS of 3000 was effectively raised by adding β‐cyclodextrin (β‐CD) and reached the constant value of ~26 °C above the molar ratio of [β‐CD]/[Py–PNIPAM] = 2/1, suggesting that β‐CD formed an inclusion complex with pyrene in the chain‐end to disturb the formation of PNIPAM aggregates, thus raising the LCST. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 1117–1124, 2006  相似文献   

6.
Four new 1D spin‐Peierls‐type compounds, [D5]1‐(4′‐R‐benzyl)pyridinium bis(maleonitriledithiolato)nickelate ([D5]R‐Py; R=F, I, CH3, and NO2), were synthesized and characterized structurally and magnetically. These 1D compounds are isostructural with the corresponding non‐deuterated compounds, 1‐(4′‐R‐benzyl)pyridinium bis(maleonitriledithiolato)nickelate (R‐Py; R=F, I, CH3, and NO2). Compounds [D5]R‐Py and R‐Py (R=F, I, CH3, and NO2) crystallize in the monoclinic space group P21/c with uniform stacks of anions and cations in the high‐temperature phase and triclinic space group P$\bar 1$ with dimerized stacks of anions and cations in the low‐temperature phase. Similar to the non‐deuterated R‐Py compounds, a spin‐Peierls‐type transition occurs at a critical temperature for each [D5]R‐Py compound; the magnetic character of the 1D S=1/2 ferromagnetic chain for [D5]F‐Py and the 1D S=1/2 Heisenberg antiferromagnetic chain for others appear above the transition temperature. Spin‐gap magnetic behavior was observed for all of these compounds below the transition temperature. In comparison to the corresponding R‐Py compound, the cell volume is almost unchanged for [D5]F‐Py and shows slight expansion for [D5]R‐Py (R=I, CH3, and NO2) as well as an increase in the spin‐Peierls‐type transition temperature for all of these 1D compounds in the order of F>I≈CH3≈NO2. The large isotopic effect of nonmagnetic countercations on the spin‐Peierls‐type transition critical temperature, TC, can be attributed to the change in ω0 with isotope substitution.  相似文献   

7.
The equilibrium constant for the reaction CH2(COOH)2 + I3? ? CHI(COOH)2 + 2I? + H+, measured spectrophotometrically at 25°C and ionic strength 1.00M (NaClO4), is (2.79 ± 0.48) × 10?4M2. Stopped-flow kinetic measurements at 25°C and ionic strength 1.00M with [H+] = (2.09-95.0) × 10?3M and [I?] = (1.23-26.1) × 10?3M indicate that the rate of the forward reaction is given by (k1[I2] + k3[I3?]) [HOOCCH2COO?] + (k2[I2] + k4[I3?]) [CH(COOH)2] + k5[H+] [I3?] [CH2(COOH)2]. The values of the rate constants k1-k5 are (1.21 ± 0.31) × 102, (2.41 ± 0.15) × 101, (1.16 ± 0.33) × 101, (8.7 ± 4.5) × 10?1M?1·sec?1, and (3.20 ± 0.56) × 101M?2·sec?1, respectively. The rate of enolization of malonic acid, measured by the bromine scavenging technique, is given by ken[CH2(COOH)2], with ken = 2.0 × 10?3 + 1.0 × 10?2 [CH2(COOH)2]. An intramolecular mechanism, featuring a six-member cyclic transition state, is postulated to account for the results on the enolization of malonic acid. The reactions of the enol, enolate ion, and protonated enol with iodine and/or triodide ion are proposed to account for the various rate terms.  相似文献   

8.
The crystal structures of the M2NaIO6 series (M = Ca, Sr, Ba), prepared at 650 °C by ceramic methods, were determined from conventional laboratory X‐ray powder diffraction data. Synthesis and crystal growth were made by oxidizing I with O2(air) to I7+ followed by crystal growth in the presence of NaF as mineralizator, or by the reaction of the alkali‐metal periodate with the alkaline‐earth metal hydroxide. All three compounds are insoluble and stable in water. The barium compound crystallizes in the cubic space group Fm3m (no. 225) with lattice parameters of a = 8.3384(1) Å, whereas the strontium and calcium compounds crystallize in the monoclinic space group P21/c (no. 14) with a = 5.7600(1) Å, b = 5.7759(1) Å, c = 9.9742(1) Å, β = 125.362(1)° and a = 5.5376(1) Å, b = 5.7911(1) Å, c = 9.6055(1) Å, β = 124.300(1)°, respectively. The crystal structure consists of either symmetric (for Ba) or distorted (for Sr and Ca) perovskite superstructures. Ba2NaIO6 contains the first perfectly octahedral [IO6]5– unit reported. The compounds of the ortho‐periodates are stable up to 800 °C. Spectroscopic measurements as well as DFT calculations show a reasonable agreement between calculated and observed IR‐ and Raman‐active vibrations.  相似文献   

9.
[Mo3,OS3(dtp)4(H2O)] reacts with NaOAc·3H2O in Py to give the title compound. The crystal data are as follows: [Mo2OS3)(OAc)2(dtp)2·Py]?0.5H,O(dtp = [S3P(OC2H5)2]?, Py = C5H5N); M = 976.64; triclinic; space group P1 ; a=11.704(5), b=14.169(7), c= 11.688 (5) Å α=109.94(4) β = 91.53(4), γ = 91.93(4)°; V= 1819(1) Å2; Z=2; Dc = 1.78 g·cm?3 λ(Mo Kα) = 0.71069 Å μ=15.15 cm?1; F(000) = 970 T=296 K; final R=0.071 for 1652 reflections with I>3σ(I). In the molecule, the [Mo3OS3] core is surrounded by two bridging OAc groups and two terminal chelate dtp groups attached to the {Mo3} triangle in a symmetric style, and the Py ligand is coordinated to the Mo atom at the apex of {Mo3} triangle with the nitrogen. This novel configuration is obtained for the first time with Mo—N bond length being 2.27 (2) Å and three Mo—Mo bond lengths 2.584 (4), 2.587 (4) and 2.657(4) Å, respectively. As a whole, the molecule has a virtual C2 symmetry.  相似文献   

10.
The kinetics of the photoinitiated reductions of methyl iodide and carbon tetrachloride by tri-n-butylgermanium hydride in cyclohexane at 25°C have been studied and absolute rate constants have been measured. Rate constants for the combination of CH3? and CCl3? radicals are equal within experimental error and are also equal to the values found for the self-reactions of most non-polymeric radicals in low viscosity solvents, i.e. ~1–3 × 109 M?1 sec?1. Rate constants for hydrogen atom abstraction by CH3? and CCl3? radicals are both ~1?2 × 105 M?1 sec?1. Tri-n-butyltin hydride is about 10–20 times as good a hydrogen donor to alkyl radicals as is tri-n-butylgermanium hydride. The strength of the germanium–hydrogen bond, D(n-Bu3Ge–H) is estimated to be approximately 84 kcal/mole.  相似文献   

11.
Diphenyl bismuth bromide (Ph2BiBr) allows for polymerizations of ε‐caprolactone in bulk at temperatures as low as 40 °C. Time conversion curves indicate a lower reactivity than tin(II) 2‐ethyl hexanoate (SnOct2) plus alcohol at 120 °C and also at 60 °C. Ph2BiBr also proved to be less reactive than Ph2BiOEt, but more reactive than BiBr3 and Bi(III)n‐hexanoate. Small amounts (≤1 wt %) of cyclic oligoester were detectable by MALDI‐TOF mass spectrometry even at a polymerization temperature of 40 °C. The molar masses depend on the monomer–initiator ratio (M/I) but not in a simple parallel manner. With M/I = 600/1 number average molecular weights (Mns, corrected values) around 500 kDa were obtained. Even at low M/Is high molar mass polylactones were found and CH2Br endgroups were not detectable. However, upon addition of tetra(ethylene glycol) the coinitiator was completely incorporated yielding telechelic polylactones and the molar mass increased with the monomer–coinitiator ratio. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 851–859, 2008  相似文献   

12.
The effect of pressure on the viscosity of dilute solutions of anionically polymerized polystyrene (M?w = 209,000; Mw/Mn = 1.12) in toluene has been studied at different temperatures and concentrations using a falling-body viscometer. Measurements were performed in the concentration range from 0.0025 to 0.02 g/mL and at temperatures from 25 to 45°C under pressure up to 1057 bars. The viscosity coefficient η increases exponentially with pressure at a given temperature and concentration, while the apparent volume of activation V? decreases with increasing temperature. The hypothesis that the pressure dependence of η is given by the pressure dependence of the activation energy holds true under the prevailing thermodynamic conditions. Log η increases linearly with increasing concentration at a given pressure. Intrinsic viscosity increases with increasing pressure, whereas the Huggins constant decreases.  相似文献   

13.
The title compound, also known as β‐erythroadenosine, C9H11N5O3, (I), a derivative of β‐adenosine, (II), that lacks the C5′ exocyclic hydroxymethyl (–CH2OH) substituent, crystallizes from hot ethanol with two independent molecules having different conformations, denoted (IA) and (IB). In (IA), the furanose conformation is OT1E1 (C1′‐exo, east), with pseudorotational parameters P and τm of 114.4 and 42°, respectively. In contrast, the P and τm values are 170.1 and 46°, respectively, in (IB), consistent with a 2E2T3 (C2′‐endo, south) conformation. The N‐glycoside conformation is syn (+sc) in (IA) and anti (−ac) in (IB). The crystal structure, determined to a resolution of 2.0 Å, of a cocrystal of (I) bound to the enzyme 5′‐fluorodeoxyadenosine synthase from Streptomyces cattleya shows the furanose ring in a near‐ideal OE (east) conformation (P = 90° and τm = 42°) and the base in an anti (−ac) conformation.  相似文献   

14.
The title compound {[Co2(NIT4Py)4(fum)2(H2O)4] · (H2O)2} n , where NIT4Py = 2-(4′-pyridinyl)-4,4,5,5-tetra-methylimidazoline-1-oxyl-3-oxide and fum = fumarate, was prepared and structurally characterized by X-ray diffraction. It crystallizes in triclinic, space group P 1 with a = 9.8853(9) Å, b = 13.2167(13) Å, c = 13.8481(12) Å, α = 68.5400(10)°, β = 73.467(2)°, γ = 74.596(2)°, V = 1587.6(3) Å3, Z = 1, and R 1 [I > 2σ(I)] = 0.0662. The complex consists of two kinds of neutral infinite 1-D chains, in which [Co(NIT4Py)2(H2O)2] moieties are connected by fumarate anions in a bis-monodentate mode. These 1-D chains are further connected by hydrogen bonds, generating a 2-D network. The variable temperature magnetic susceptibility measurement reveals that the magnetic couplings between Co(II) and NIT4Py are weak antiferromagnetic interactions.  相似文献   

15.
A new access to cationic zirconium and hafnium compounds [L2MCH2PR2][MeB(C6F5)3] (L = Cp, Ind; R = iso‐Pr, tert‐Bu; M = Zr, Hf) exhibiting an intramolecular donor‐acceptor system was established by treating the precursors L2M(Me)CH2PR2 with B(C6F5)3 (BCF). Precursors 1 – 6 [L2M(Me)CH2PR2 with L = Cp, Ind; R = iso‐Pr, tert‐Bu; M = Zr, Hf] were fully characterized. The crystal structures of these compounds revealed large M–CH2–P bond angles with values of about 134° indicating the absence of interactions between the Lewis‐acid and Lewis‐base. The cationic compounds [L2MCH2PR2][MeB(C6F5)3] ( 7 – 12 ) were obtained by treatment of 1 – 6 with BCF. They were characterized by NMR spectroscopy, mass spectrometry, and elemental analyses; in H/D‐scrambling experiments with H2/D2 mixtures 7 – 12 disclosed their reactivity towards cleavage of hydrogen.  相似文献   

16.
The title compound, 4-n-nonylbenzoic acid, CH3(CH2)8C6H4COOH (NBA), has been characterized thus: triclinic, P1, a = 13.514(4)Å, b = 23.4672(5)Å, c = 7.658(3)Å, alpha = 90.914(3)°, beta = 100.403(3)°, gamma = 77.781(2)°, V = 2334(2)Å3, Z = 2, and F.W = 745.06, Dm = 1.161 g cm-3, Dc = 1.061 g cm-3, F000 = 816.00, lambda = 0.71069Å, mu = 0.6 cm-1, goodness-of-fit is 1.029, final R1 = 0.063 and wR2 = 0.16.  相似文献   

17.
The reaction of methylammonium halides and cobalt halides yielded the organic‐inorganic hybrid compounds of general formula (CH3NH3)2CoX4. By varying the different halides, we were able to synthesize the whole row from Cl to I as well as some mixed halides compounds and to determinate the crystal structures. (CH3NH3)2CoX4 (X = Cl, Br, Cl0.5Br0.5, Br0.5I0.5) crystallize isotypic to (CH3NH3)2HgCl4 in space group P21/c with Z = 4 [X = Cl: a = 7.6483(9), b = 12.6885(18), c = 10.8752(12) Å, β = 96.639(9)°; X = Cl0.5Br0.5: a = 7.8271(9), b = 12.9543(9), c = 11.1007(11) Å, β = 96.320(8)°; X = Br: a = 7.9782(2), b = 13.1673(2), c = 11.2602(2) Å, β = 96.3260(10)° and X = Br0.5I0.5: a = 8.2435(12), b = 13.645(2), c = 11.5856(18) Å, β = 95.54(2)°]. The mixed halides show a statistic distribution in both cases. In (CH3NH3)2CoCl2I2 an ordered variant is realized representing a new structure type [C2/m, Z = 4, a = 18.808(4), b = 7.3604(7), c = 10.4109(17) Å, β = 120.364(13)°]. (CH3NH3)2CoI4 crystallizes again isotypic to the respective mercury compound [(CH3NH3)2HgCl4] [Pbca, Z = 8, a = 10.9265(5), b = 12.1552(5), c = 20.9588(9) Å]. All structures are build up by inorganic tetrahedral [CoX4]2– anions and organic (CH3NH4)+ cations. Additionally the Raman spectra as well as the optical reflectance spectra are discussed.  相似文献   

18.
Perylene (Py)‐containing polyacetylenes with different skeleton structures ? [HC?C(C6H4)CO2? Py]n? (P 1 ), ? [HC?C(CH2)8CO2? Py]n? (P 2 ), and ? {[(C6H5) C?C(CH2)9NH2]? co? [(C6H5)C?C(CH2)9? Py]}n? (P 3 ) are synthesized in satisfactory yields by Rh‐catalyzed polymerization (for P 1 and P 2 ) and polymer reaction (for P 3 ). All the polymers are soluble and possess high molecular weights (Mw up to 2.8 × 105). Their structures and properties are characterized and evaluated by IR, NMR, UV, TGA, PL, and photovoltaic (PV) analyses. The polymers are thermally stable, losing little of their weights when heated to 330 °C. When their solutions are irradiated, their perylene pendants emit intense red fluorescence at 610 nm. PV cells with a configuration of ITO/PEDOT:PSS/polymer/LiF/Al are fabricated, which show maximum current density of 10.3 μA/cm2. The external quantum efficiency is sensitive to the polymer structure, with P 3 exhibiting the highest value of 0.23%. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2025–2037, 2008  相似文献   

19.
Time-resolved light scattering was employed to investigate kinetics of phase separation in mixtures of poly (ethylene glycol monomethylether) (PEGE)/poly (propylene glycol) (PPG) oligomers. Phase diagrams for PEGE/PPG of varying molecular weights were established by means of cold point measurements. The oligomer mixtures reveal an upper critical solution temperature (UCST). Several temperature quench experiments were carried out with a 60/40 PEGE/PPG blend by rapidly quenching from a single phase (69°C) to two-phase temperatures (66–61°C) at 1°C intervals. As is typical for oligomer mixtures, the early stage of spinodal decomposition (SD) was not detected. The kinetics of phase decomposition was found to be dominated by the late stage of SD. Time-evolution of scattering intensity was analyzed in accordance with nonlinear and dynamical scaling theories. The time dependence of the peak intensity Im and the corresponding peak wavenumber qm was found to follow the power-law {Im(t)? tα, qm(t)? t} with the values of α = 3 ± 0.3 and β = 1 ± 0.2, which are very close to the values predicted by Siggia. This process has been attributed to a coarsening mechanism driven by surface tension. In the temporal scaling analysis, the structure function reveals university with time, suggesting self-similarity. Phase separation dynamics in 60/40 PEGE/PPG resembles the behavior predicted for off-critical mixtures.  相似文献   

20.
Crystal Structures, Vibrational Spectra, and Normal Coordinate Analyses of the Chloro-Iodo-Rhenates(IV) (CH2Py2)[ReCl5I], cis -(CH2Py2)[ReCl4I2] · 2 DMSO, trans -(CH2Py2)[ReCl4I2] · 2 DMSO, and fac -(EtPh3P)2[ReCl3I3] [ReCl5I]2–, cis-[ReCl4I2]2–, trans-[ReCl4I2]2–, and fac-[ReCl3I3]2– have been synthesized by ligand exchange reactions of [ReI6]2– with HCl and are separated by ion exchange chromatography on diethylaminoethyl cellulose. X-ray structure determinations have been performed on single crystals of (CH2Py2)[ReCl5I] ( 1 ) (triclinic, space group P1 with a = 7.685(2), b = 9.253(2), c = 12.090(4) Å, α = 90.06(2), β = 101.11(2), γ = 95.07(2)°, Z = 2), cis-(CH2Py2)[ReCl4I2] · 2 DMSO ( 2 ) (triclinic, space group P1 with a = 8.662(2), b = 12.109(2), c = 12.9510(12) Å, a = 97.533(11), β = 96.82(2), γ = 89.90(2)°, Z = 2) , trans-(CH2Py2)[ReCl4I2] · 2 DMSO ( 3 ) (triclinic, space group P1 with a = 9.315(7), b = 9.663(3), c = 15.232(3) Å, α = 80.09(2), β = 81.79(4), γ = 83.99(5)°, Z = 2) and fac-(EtPh3P)2[ReCl3I3] ( 4 ) (monoclinic, space group P21/a with a = 17.453(2), b = 13.366(1), c = 19.420(1) Å, β = 112.132(8)°, Z = 4). The crystal structure of ( 1 ) reveals a positional disorder of the anion sublattice along the asymmetric axis. Due to the stronger trans influence of I compared with Cl on asymmetric axes Cl˙–Re–I′ is caused a mean lenghthening of the Re–Cl˙ distances of 0.020 Å (0.8%) and a shortening of the Re–I′ distances of 0.035 Å (1.3%) with regard to symmetrically coordinated axes Cl–Re–Cl and I–Re–I, respectively. Using the molecular parameters of the X-Ray determinations the low temperature (10 K) IR and Raman spectra of the (n-Bu4N) salts of all four chloro-iodo-rhenates(IV) are assigned by normal coordinate analyses. The weakening of the Re–Cl˙ bonds and the strengthening of the Re–I′ bonds is indicated by a decrease or increase of the valence force constants each by 9%.  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