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 13 毫秒
1.
The synthesis of substituted poly(p-phenylenenvinylene) (PPV) was carried out via metathesis polycondensation of 2,5-diheptyl-1,4-divinylbenzene (DHepDVB). A stable molybdenumcarbene complex served as catalyst. The preparation of the educt employed (DHepDPV, 3a ) is described. The obtained poly(2,5-diheptyl-1,4-phenylenevinylene) (DHepPPV, 5 ), an intensely yellow product, has an all-trans-configuration and, with a degree of polymerization of ≈ 10 is soluble in conventional organic solvents.  相似文献   

2.
A cylcopalladated complex, [Pd(Cl)(k2N,C-CH2C6H2(Me)2CHNC6H3(Pri)2]2 was found to be an excellent catalyst for Suzuki-Miyaura coupling of aryl halides with arylboronic acid in aqueous medium under aerobic conditions.  相似文献   

3.
4.
[reaction: see text] Glyoxal bis(N-methyl-N-phenylhydrazone) (1) and its related compounds such as 2-pyridinecarboxaldehyde N-methyl-N-phenylhydrazone (3) were prepared and examined as ligands for the Suzuki-Miyaura cross-coupling reaction of aryl halides and arylboronic acids. We found phosphine-free catalysts, such as Pd(OAc)(2)/hydrazone ligand 1 or 3, to be efficient catalysts for a variety of substrates to produce the coupling products in good yields.  相似文献   

5.
A series of diacetylene monomers with benzoyl, 4-hexylbenzoyl, 4-dodecylbenzoyl, and perfluorobenzoyl substituents were synthesized and investigated with respect to their crystal structures and polymerizability. In the absence of perfluorophenyl-phenyl interactions, the crystal structures of related alkylated and nonalkylated derivatives were substantially different and dominated by the phase segregation between the alkylated side chains and the diaryl-substituted diacetylene cores. By contrast, the perfluorophenyl-phenyl interactions served as a reliable supramolecular synthon in that they persisted in the crystal structures of different alkylated and nonalkylated derivatives. The packing of the diacetylene functions was appropriate for a topochemical polymerization in these cases, and the perfluorophenyl-phenyl interaction determined the polymerization direction. As a result, soluble alternating diacetylene copolymers were obtained which were further characterized with solution phase methods.  相似文献   

6.
Boc-protected (piperazin-1-ylmethyl)biaryls have been synthesised from (Boc-piperazin-1-ylmethyl)phenylboronic acid pinacol esters via a microwave-mediated Suzuki-Miyaura coupling with aryl bromides viz. 1-bromo-, 2-, 3- or 4-nitrobenzene or 2-bromo-5-nitropyridine. Judicial removal of the protecting group on the piperazine, or facile reduction of the nitro group on the biaryl system enabled the manipulation of two points of functionality in order to diversify the scope of the resulting biaryl library.  相似文献   

7.
Vanessa Cerezo 《Tetrahedron》2007,63(42):10445-10453
Microwave irradiation efficiently promoted the Suzuki-Miyaura reaction of a 5-bromohistidine with various arylboronic acids in the presence of a palladium catalyst. This methodology allowed the synthesis of histidines substituted at position 5 of the imidazole ring with a phenyl, a substituted phenyl, a pyridyl or a thienyl ring. The corresponding 5-arylhistidines were obtained in moderate to good yields.  相似文献   

8.
Stereoblock poly(lactic acid) consisting of D- and L-lactate stereosequences can be successfully synthesized by solid-state polycondensation of a 1:1 mixture of poly(L-lactic acid) and poly(D-lactic acid). In the first step, melt-polycondensation of L- and D-lactic acids is conducted to synthesize poly(L-lactic acid) and poly(D-lactic acid) with a medium-molecular-weight, respectively. In the next step, these poly(L-lactic acid) and poly(D-lactic acid) are melt-blended in 1:1 ratio to allow formation of their stereocomplex. In the last step, this melt-blend is subjected to solid-state polycondensation at temperature where the dehydrative condensation is allowed to promote chain extension in the amorphous phase with the stereocomplex crystals preserved. Finally, stereoblock poly(lactic acid) having high-molecular-weight is obtained. The stereoblock poly(lactic acid) synthesized by this way shows a higher melting temperature in consequence of the controlled block lengths and the resulting higher-molecular-weight. The product characterization as well as the optimization of the polymerization conditions is described. Changes in M(w) of stereoblock poly(lactic acid) (sb-PLA) as a function of the reaction time.  相似文献   

9.
Research on Chemical Intermediates - An expedient synthesis of spiro-fused heterocycles from multi-component reaction of urea, aryl aldehydes and Meldrum’s or barbituric acid by using polymer...  相似文献   

10.
The interfacial polycondensation method has been used for the preparation of brominated poly(arylcarboxylate)s. Brominated poly(arylcarboxylate)s can be prepared easily by mixing a solution of diacid chloride in a water-immiscible organic solvent with an aqueous alkaline solution of bisphenol in the presence of catalyst such as quaternary ammonium salts. First, in a dichloromethane-water system using triethylbenzylammonium chloride (TEBAC) as the catalyst a series of 2,2-bis(4-hydroxy-3,5-dibromophenyl)propane (3,3′,5,5′-tetrabromobisphenol A, TBBPA) with isophthaloyl (I) and terephthaloyl chlorides (T) has been prepared and some properties as inherent viscosity, solubility, crystallinity, and flammability have been measured. Copolymers prepared from TBBPA and mixed T/I with 33/67–67/33 molar ratios show good solubility and amorphous nature, and can be cast into transparent and tough films with limiting oxygen index of 58–59 (ANSI/ASTM D2683-77). Second, the effects of some variables as the nature of organic phase and catalysts, concentration of reactants, and basicity of aqueous phase on the interfacial polycondensation of TBBPA with equal parts of T and I [T/I (50/50)] was investigated in some detail. Among the solvents tested dichloromethane was found to be the best solvent and quaternary ammonium salts such as TEBAC and tetra-n-butylammonium bromide (TBAB) were highly efficient catalysts. Poly(arylcarboxylate)s with the highest molecular weights were obtained at an acid chloride concentration of 0.2 mol/L in dichloromethane and a concentration of TBBPA of 0.1 mol/L in alkali when TEBAC was used as catalyst. A maximum of inherent viscosity was obtained at two equivalent amounts of alkali corresponding to bisphenol. Polycondensation of several combinations of T/I (50/50) with some other tetrabromobisphenols, such as 3,3′,5,5′-tetrabromo-4,4′-biphenol, 3,3′,5,5′-tetrabromobisphenol S, 3,3′,5,5′-tetrabromo-4,4′-thiodiphenol, and 3,3′,5,5′-tetrabromophenolphthalein, were carried out with limited success. Whereas, a more favorable result could be obtained by the mixed copolycondensation of these tetrabromobisphenols and bisphenol A (BPA) with T/I (50/50). Finally, the copoly(arylcarboxylate)s from TBBPA, BPA, T, and I were prepared and characterized. The incorporation of bromine on the polymer backbone caused a decrease of inherent viscosity, glass transition temperature, crystallinity, and thermal stability of copolyarylates, whereas it caused a great enhancement of flame retardancy.  相似文献   

11.
We report that static quenching of a mannosylated conjugated polymer (sugar-PPE) by Concanavalin A is positively dependent upon sugar-PPE concentration, that is, the recorded Stern-Volmer constants increase with increasing sugar-PPE concentration. Comparison with data obtained from isothermal titration calorimetry (ITC) display the increased sensitivity of the quenching method when compared to ITC. The proposed mechanism suggests the interaction of two or more chains of PPE with one Con A molecule leading to a quenched sugar-PPE-Con A construct.  相似文献   

12.
A chiral harvesting transmission mechanism is described in poly(acetylene)s bearing oligo(p-phenyleneethynylene)s (OPEs) used as rigid achiral spacers and derivatized with chiral pendant groups. The chiral moieties induce a positive or negative tilting degree in the stacking of OPE units along the polymer structure, which is further harvested by the polyene backbone adopting either a P or M helix.

A chiral harvesting transmission mechanism is described in poly(acetylene)s bearing oligo(p-phenyleneethynylene)s (OPEs) used as rigid achiral spacers and derivatized with chiral pendant groups.

During the last years, dynamic helical polymers have attracted the attention of the scientific community due to the possibility of tuning the helical sense and/or the elongation of the helical structure by using external stimuli.1–14In the case of a chiral dynamic helical polymer, modifications in its structure—helical sense enhancement or helix inversion—arise from conformational changes induced at its chiral pendants—usually, with just one stereocenter—, by stimuli such as variations in solvent polarity or temperature, the addition of certain ions, and so on (Fig. 1a).15 On the other hand, if a helical polymer is achiral (i.e., bearing achiral pendants), the chiral amplification phenomena can emerge from interactions between the polymer and external chiral molecules.16 In both the above cases, the changes produced in the helical structures are related to the spatial dispositions adopted by the substituents or associated species at the pendant groups.17–19Open in a separate windowFig. 1Several scenarios depicting conceptual representations of the transmission of chiral information. (a) Helical switch via chiral tele-induction. (b) Effect of distance on chiral tele-induction from multichiral pendants. (c) Helicity controlled by the conformational composition of achiral spacers.A step forward in the helical sense control of poly(phenylacetylene)s (PPA)s is to study different mechanisms of transmission of chiral information from the pendant to the polyene backbone by introducing achiral spacers. The goal is to demonstrate how far it is possible to place the chiral center and still have an effective chiral induction on the polyene backbone. Therefore, transmission of the chiral information from a remote position can occur through space, thus overpassing the distance generated by the spacer—tele-induction—(Fig. 1b),20–28 or through the achiral spacer itself, producing in it a preferred structure, such as a helical structure and where the orientation of the achiral helix is further transmitted to the polyene backbone—conformational switch—(Fig. 1c).29–31For the first mechanism—chiral tele-induction—, both flexible and rigid spacers have been designed.20–28 In all cases, supramolecular interactions, such as H bonding or π–π stacking, generate organized structures. As a result, the chiral center is located into a specific orientation, producing an effective helical induction. Additionally, those studies allow evaluating how distances and sizes have an effect on this phenomenon.In the second strategy, the helix induction is transmitted through conformational changes along an achiral spacer which is harvested by the polyene. For instance, an achiral peptide or an achiral polymeric helix derivatized at one end with a chiral residue and linked to the polymer main chain at the other end. In such cases, changes in the absolute configuration or even just a conformational change at the chiral center can induce an opposite helical structure into the achiral spacer, which in turn will be harvested by the polymer main chain (Fig. 1c).29–31Herein we will demonstrate another remote chiral induction mechanism based on a different chiral harvesting process. In this case, the chiral center does not produce a conformational change at the achiral spacer, but affects its array within the helical scaffold. Thus, to perform these studies we decided to introduce the use of oligo(p-phenyleneethynylene)s (m = 1, 2, 3) (OPEs) as rigid spacers to separate the distant chiral center from the polyene backbone. These OPE units have been used in the formation of benzene-1,3,5-tricarboxamide (BTA) based supramolecular helical polymers, demonstrating their ability to stack with a certain tilting degree commanded by the chiral center.32–34Hence, in our design, the chiral moiety will determine the supramolecular chiral orientation of the OPE groups used as spacers, which is further harvested by the polyene backbone. The overall process yields a helix with a preferred screw sense (Fig. 2).Open in a separate windowFig. 2Conceptual side view and top view of the chiral information transmission mechanism from stereocenters at the far end of oligo(p-phenyleneethynylene) spacers to the polyene backbone via chiral harvesting.To perform these studies, we used as model compounds two PPAs—poly-(R)-1 and poly-(S)-1—derived from the 4-ethynylanilide of (S)- and (R)-α-methoxy-α-phenylacetic acid (MPA, m-(S/R)-1), whose helical structures and dynamic behaviors have been deeply studied by our group—poly-(R)-1 and poly-(S)-1—(Fig. 3).35–46 By using these polymers as reference materials, four novel PPAs were designed introducing two OPE spacers—4-[(p-phenyleneethynylene)n]ethynylanilide (n = 1, 2)—between the phenyl acetylene group and the (S)- or (R)-α-methoxy-α-phenylacetic acid (MPA) chiral group. Thus, monomers m-(S)- and m-(R)-2 and m-(S)- and m-(R)-3 (Fig. 3a) were prepared and submitted to polymerization by using a Rh(i) catalyst poly-(S)- and poly-(R)-2 and poly-(S)- and poly-(R)-3 (Fig. 3b) were obtained in high yield and showed Raman spectra characteristic of cis polyene backbones (see Fig. S11 and S12).Open in a separate windowFig. 3(a) Monomers and (b) polymers synthetized in this study.X-ray structures of the monomers show a preferred antiperiplanar (ap) orientation between the carbonyl and methoxy groups (O Created by potrace 1.16, written by Peter Selinger 2001-2019 C–C–OMe) for m-(R)-2 and m-(S)-3, whereas in the case of m-(S)-1 a synperiplanar (sp) geometry is favoured (Fig. 4a).35 In complementary studies, CD spectra of monomers m-(S)-[1–3] in CHCl3 show negative Cotton effects, indicative of major ap conformations in solution (Fig. 4b),35 further corroborated by theoretical calculations (see Fig. S10). Interestingly, the maximums of the Cotton effects in CD undergo a bathochromic shift—from 266 nm in m-1 to 327 nm in m-3—due to a larger conjugation of the π electrons (from the anilide to the alkyne group) when the length of the spacer increases (Fig. 4b).Open in a separate windowFig. 4(a) X-ray structures of m-(S)-1, m-(R)-2 and m-(S)-3. (b) CD traces of m-(S)- and m-(R)-1; m-(S)- and m-(R)-2; m-(S)- and m-(R)-3 in CHCl3 (0.1 mg mL−1). (c) CD spectra for poly-(S)- and poly-(R)-1 in CHCl3 (0.1 mg mL−1); poly-(S)- and poly-(R)-2 in DMSO (0.1 mg mL−1); poly-(S)- and poly-(R)-3 in DMSO (0.1 mg mL−1).CD studies of the polymer series bearing OPE spacers—poly-(R)- and poly-(S)-[2–3]—in different solvents show the formation of a PPA helical structure with a preferred helical sense, while the parent polymer, poly-1, devoid of the OPE unit, has a poor CD. This is a very interesting phenomena that indicates that the OPE spacers work as transmitters of the chiral information from remote chiral centers to the polyene backbone—placed at 1.7 nm for poly-2 and at 2.4 nm for poly-3—(Fig. 4a). These large distances between the chiral center and the polymer main chain mean that other mechanisms of chiral induction, such as chiral tele-induction effect, should be almost null in these cases.In these two polymers (poly-2 and poly-3), the chiral information transmission mechanism must occur in different sequential steps. First, the chiral centers possessing a major (ap) conformation induce a certain tilting degree (θ) in the achiral spacer array. This step resembles the helical induction mechanism found in supramolecular helical polymers bearing OPE units.32–34 Next, the chiral array induced in the OPE units is harvested by the polyene backbone, resulting in an effective P or M helix induction (Fig. 2).34,47Additional structural studies were carried out in poly-(S)-2 and poly-(S)-3 to obtain an approximated secondary structure of these polymers and determine their dynamic behaviour.From literature it is known that the conformational equilibrium of poly-1 can be altered in solution by the presence of metal ions. The addition of monovalent ions (e.g., Li+) stabilizes the ap conformer at the pendant group by cation–π interactions, while divalent ions (e.g., Ca2+) stabilize the sp conformations by chelation with the methoxy and carbonyl groups.36,38,39,43 As a result, both the P or M helical senses can be selectively induced in poly-1 by the action of metal ions.Therefore, we decided to add different perchlorates of monovalent and divalent metal ions to solutions of poly-(S)-2 and poly-(S)-3 with the aim of determining the conformational composition at the pendant groups. Thus, when monovalent metal ions (Li+, Ag+ and Na+) are added to a chloroform solution of poly-(S)-2, a chiral enhancement is observed (Fig. 5d for Li+ and Fig. S16 for Na+ and Ag+). IR and 7Li-NMR studies show that those ions stabilize the ap conformer at the pendant group in a similar fashion to poly-1, this is by coordination to the carbonyl group of the MPA (Fig. 5g) and the presence of a cation–π interaction with the aryl ring of the chiral (|Δδ| 7Li ca., 3.75 ppm) (Fig. 5f and ESI). Therefore, addition of Li+ produces a larger number of pendant groups with ap conformation among poly-2, which triggers a chiral enhancement effect through a cooperative process.Open in a separate windowFig. 5(a) Conceptual representation of the chiral information harvesting and top view of the 3D model for poly-(S)-2. (b) CD spectra of poly-(S)-2 (0.2 mg mL−1) in DMSO vs. calculated ECD spectra. Full width at half-maximum (FWHM) equals 20 nm. (c) Low-resolution AFM image from a poly-(S)-2 monolayer and profile depicting the chain separation of the yellow highlighted area in the AFM image. (d) CD spectra showing the chiral enhancement after the addition of Li+ (50 mg mL−1, THF) to a poly-(S)-2 solution (0.1 mg mL−1, THF). (e) CD trace of poly-(S)-2 before and after the addition of a Ca2+ solution (50 mg mL−1, THF). (f) 7Li-NMR spectra substantiating the cation–π interaction. (g) IR shifts observed for carbonyl and methoxy groups after the addition of LiClO4 and Ca(ClO4)2 (50 mg mL−1, THF) to a poly-(S)-2 solution (3 mg mL−1, CHCl3). The coordination modes of the MPA moiety with Li+ and Ca2+ are shown vertically in the middle of the figure.On the contrary, the addition of perchlorates of divalent metal ions, such as Ca2+and Zn2+, produced an inversion of the third Cotton band—310 nm—associated to the MPA moiety and the disappearance of both first and second Cotton effects (Fig. 5e for Ca2+ and Fig. S17 for Zn2+). This is a very interesting outcome because, although the conformational equilibrium at the MPA group changes from ap to sp after the addition of Ca2+, the number of pendant groups with sp conformation do not reach the number needed to trigger the helix inversion process and in fact, a mixture of P and M helices at the polyene backbone is obtained.The helical structures adopted by both polymer systems, PPAs (poly-1) and poly[oligo(p-phenyleneethynylene)phenylacetylene]s (POPEPAs) (poly-2 and poly-3), are defined by two coaxial helices, one formed by the polyene backbone (internal helix, CD active) and the other constituted by the pendants (external helix, observed by AFM).These two helices can rotate in either the same or the opposite sense, depending on the dihedral angle between conjugated double bonds. Thus, internal and external helices rotate in the same direction in cis-cisoidal polymers, while they rotate in opposite directions in cis-transoidal ones.14,42,48,49In order to find out an approximated helical structure for poly-(S)-2, DSC studies were performed. The thermogram shows a compressed cis-cisoidal polyene skeleton (see Fig. S13a), similar to the one obtained for poly-1.42 Moreover, AFM studies on a 2D crystal of poly-(S)-2 did not produce high-resolution AFM images, although some parameters such as helical pitch (c.a., 2.8 nm) and packing distance between helices of (c.a., 6 nm) could be extracted from the well-ordered monolayer analyzed (Fig. 5c).Previous structural studies in PPAs found that it is possible to correlate the internal helical sense with the Cotton band associated to the polyene backbone—CD (+), Pint; CD (−), Mint—.50,51 Herein, the positive Cotton effect observed for the polyene backbone [CD365 nm = (+)] in poly-(S)-2 is indicative of a P orientation of the internal helix, which correlates with a P orientation of the external helix in a cis-cisoidal polyene scaffold. To summarize, DSC, AFM and CD studies agree that poly-(S)-2 is made up of a cis-cisoidal framework with Pint and Pext helicities (Fig. 5a).Computational studies [TD-DFT(CAM-B3LYP)/3-21G] were carried out on a P helix of an n = 9 oligomer of poly-(S)-2, possessing a cis-cisoidal polyene skeleton (ω1 = +50°, ω3 = −40°) and an antiperiplanar orientation of the carbonyl and methoxy groups at the pendants. The theoretical ECD spectrum obtained from these studies (Fig. 5b and see ESI for additional information) is in good agreement with the experimental one, indicating that our model structure is a good approximation of the helical structure adopted by poly-(S)-2.Next, a similar set of DSC and AFM studies were carried out for poly-(S)-3, that bears an OPE spacer with n = 2. The data showed that this polymer presents a compressed cis-cisoidal polyene skeleton, similar to those obtained for poly-1 and poly-2 (see Fig. S13b), with a helical pitch of 3.8 nm and a Pext helical sense (Fig. 6a and c).Open in a separate windowFig. 6(a) Conceptual representation of the chiral information harvesting and top view of the 3D model for poly-(S)-3. (b) CD spectrum of poly-(S)-3 in THF (0.2 mg mL−1) and comparison to the calculated ECD spectra. Full width at half-maximum (FWHM) equals 20 nm. (c) AFM image obtained from a poly-(S)-3 monolayer. (d) CD traces for poly-(S)-3 in THF polymerized at different temperatures.UV studies indicate that, in poly-(S)-3, the polyene backbone absorbs at ca. 380 nm, coincident with the first Cotton effect, that is positive (see Fig. S15b). Therefore, it reveals that poly-(S)-3 adopts a Pint helicity (Fig. 6b). Thus, as expected for cis-cisoidal scaffolds, the orientations of the two coaxial helices are coincident.Computational studies [TD-DFT(CAM-B3LYP)/3-21G] were carried out on a P helix of an n = 9 oligomer of poly-(S)-3, possessing a cis-cisoidal polyene skeleton (ω1 = +63°, ω3 = −40°) and an antiperiplanar orientation of the carbonyl and methoxy groups at the pendants. The theoretical results (Fig. 6b and see ESI for additional information) match with the experimental data, indicating that our model structure is a good approximation to the helical structure adopted by poly-(S)-3.Finally, the stimuli response properties of poly-(S)-3 were explored by CD. These experiments revealed that the addition of monovalent or divalent metal ions to a chloroform solution of poly-(S)-3 does not produce any significant effect in the structural equilibrium of this polymer (see Fig. S18). This fact, in addition to the previous results obtained from the interaction of poly-(S)-2 with divalent metal ions, corroborates the decrease of the dynamic character of helical PPAs when large OPEs are used as spacers.The poor dynamic behaviour was further demonstrated by polymerizing m-(S)-3 at a lower temperature (0 °C) (Fig. 6d). In this case, the region around 240–350 nm remains unaffected, indicating that the pendant is ordered in a similar manner in both batches of polymers, regardless of the temperature at which they were synthesized (20 °C and 0 °C). Interestingly, the magnitude of the first Cotton band is duplicated when the polymer is obtained at low temperature due to a stronger helical sense induction at the polyene backbone. This result indicates that a preorganization process may occur during polymerization, affecting the screw sense excess of the PPA.In conclusion, a novel chiral harvesting transmission mechanism has been described in poly(acetylene)s bearing oligo(p-phenylenethynylene)s as rigid spacers that place the chiral pendant group away from the polyene backbone, at a distance around ca. 1.7 nm for poly-2, and 2.4 nm for poly-3. Hence, the disposition of the chiral moiety affects the stacking of the OPE units within the helical structure, inducing a specific positive or negative tilting degree, which is further harvested by the polyene backbone inducing either a P or M internal helix.We believe that these results open new horizons in the development of novel helical structures by combining information from the helical polymers and supramolecular helical polymers fields, which leads to the formation of novel materials with applications in important fields such as asymmetric synthesis, chiral recognition or chiral stationary phases among others.  相似文献   

13.
The synthesis of nanocomposites via emulsion polymerization was investigated using methyl methacrylate (MMA) monomer, 10 wt % montmorillonite (MMT) clay, and a zwitterionic surfactant octadecyl dimethyl betaine (C18DMB). The particle size of the diluted polymer emulsion was about 550 nm, as determined by light scattering, while the sample without clay had a diameter of about 350 nm. The increase in the droplet size suggests that clay was present in the emulsion droplets. X-ray diffraction indicated no peak in the nanocomposites. Transmission electron microscopy showed that emulsion polymerization of MMA in the presence of C18DMB and MMT formed partially exfoliated nanocomposites. Differential scanning calorimetry showed an increase of 18 degrees C in the glass transition temperature (Tg) of the nanocomposites. A dynamic mechanical thermal analyzer also verified a similar Tg increase, 16 degrees C, for the partially exfoliated nanocomposites over poly(methyl methacrylate) (PMMA). Thermogravimetric analysis indicated a 37 degrees C increase in the decomposition temperature for a 20 wt % loss. A PMMA nanocomposite with 10 wt % C18DMB-MMT was also synthesized via in situ polymerization. This nanocomposite was intercalated and had a Tg 10 degrees lower than the emulsion nanocomposite. The storage modulus of the partially exfoliated emulsion nanocomposite was superior to the intercalated structure at higher temperatures and to the pure polymer. The rubbery plateau modulus was over 30 times higher for the emulsion product versus pure PMMA. The emulsion technique produced nanocomposites of the highest molecular weight with a bimodal distribution. This reinstates that exfoliated structures have enhanced thermal and mechanical properties over intercalated hybrids.  相似文献   

14.
Poly(spiropyran)s were synthesized by polycondensation of a bis(indoline) monomer with bis(o-positioned formyl and hydroxy)-substituted aromatic monomers in alcoholic solvents. The structures of the products and their molecular weights were determined by the 1H NMR and GPC measurements, respectively. Furthermore, photoisomerization behaviors of the poly(spiropyran)s were analyzed by the UV–Vis and 1H NMR measurements of their solutions in dichloromethane upon the UV and visible light irradiations.  相似文献   

15.
熔融聚合法合成生物材料聚(乳酸-丙氨酸)   总被引:2,自引:0,他引:2  
直接以乳酸、丙氨酸为原料,通过熔融聚合法合成生物降解材料聚(乳酸-丙氨酸)[P(LA-co-Ala)],并用特性粘数、IR、1H-NMR等手段对目标产物进行了表征。研究催化剂的种类和用量、聚合温度、聚合时间等对P(LA-co-Ala)合成的影响,得到最佳实验条件为:分步脱水后,以SnO为催化剂(质量分数为0.7%),170℃、70Pa下反应8h,P(LA-co-Ala)的特性粘数为0.7888 dL/g。  相似文献   

16.
Two hydrophilic branched oligo(ethylene glycol)-substituted PPV derivatives, poly(2,5-bis(1,3-bis(triethoxymethoxy)propan-2-yloxy)-1,4-phenylene vinylene) (BTEMP-PPV) and poly(2-methoxy-5-(1,3-bis(triethoxymethoxy)propan-2-yloxy)-1,4-phenylene vinylene) (MTEMP-PPV), are presented. Polymerizations have been performed via the dithiocarbamate precursor route, using lithium hexamethyldisilazide (LHMDS) as a base, to obtain high molecular weight precursor polymers. After thermal conversion of the precursor polymers into the fully conjugated systems, the solubility of the polymers has been examined. The polar nonionic side chains of MTEMP-PPV and BTEMP-PPV render the PPV backbone soluble in a variety of solvents, including alcohols and even water, making these polymers suitable candidates to be used in optoelectronic devices that can be processed from environmentally friendly solvent systems.  相似文献   

17.
A copper-catalyzed aerobic oxidative CN bond cleavage reaction was developed for the synthesis of 4-substituted-NH-1,2,3-triazoles. Diverse β-ketotriazoles derivatives, which are the starting materials for the aerobic oxidative CN bond cleavage reaction, were prepared from nine aryl and seven alkyl alkynes and α-azidoacetophenone by a copper(I)-catalyzed [3 + 2]cycloaddition reaction. The aerobic oxidation of α-(1,2,3-triazol-1-yl)acetophenones using a catalytic amount of copper(II) acetate in the presence oxygen under neutral conditions gave the title compounds in high yield.  相似文献   

18.
1,4‐Dibromobenzenes carrying nonpolar hexoxy and polar oligo(ethylene glycol) side chains were subjected to Suzuki polycondensation with a benzene‐1,4‐bisboronic acid ester to produce high‐molar‐mass poly(para‐phenylene)s. The molar masses were determined with size exclusion chromatography with conventional polystyrene and universal calibration. These novel amphiphilically equipped rigid‐rod polymers have the potential to segregate lengthwise into polar and nonpolar domains, a property that has only rarely been described, and promise to exhibit novel interesting supramolecular properties. The oligo(ethylene gylcol) side chains terminate with a silyl‐protected alcohol group, and its deprotection on the polymer was proven to proceed quantitatively. This not only led to a further polarity increase but allows us to attach even more polar (e.g., charged) units in future projects. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 2879–2889, 2003  相似文献   

19.
High molecular weight aromatic poly(amide-ester)s were prepared by the direct polycondensation reactions between aromatic dicarboxylic acids and aminophenols under mild conditions in pyridine. The condensing agents examined in this study were diphenyl chlorophosphate (DPCP), DPCP/LiCl, and DPCP/DMF. Addition time of the aminophenols, depending on their nucleophilicities, affected the ηinh values and monomer sequence of the resulting polymer. Their thermal properties were studied in terms of the sequences in the polymer backbones.  相似文献   

20.
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