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 562 毫秒
1.
Three aluminum complexes supported by salen ligands derived from cis-1,2-cyclohexanediamine and salicylaldehyde derivatives were synthesized. They were characterized by 1H, 13C NMR spectra, and elemental analysis. X-ray diffraction analysis revealed that aluminum was in distorted square pyramidal geometry in 2. These complexes were employed as catalysts for the ring-opening polymerization (ROP) of L-lactide and rac-lactide. Complex 2 showed the highest activity among these complexes with isopropanol for the ROP of L-lactide and 3 showed the highest stereoselectivity for the ROP of rac-lactide attaining partially isotactic polylactide with a Pm of 0.75. The kinetic data of the polymerization utilizing 3 as catalyst showed that the polymeric rate was first order to the monomer and catalyst.  相似文献   

2.
New polynuclear zinc complexes containing tridentate Schiff base ligands were successfully synthesized and fully characterized. The solid‐state structure of the complexes was determined using single crystal X‐ray diffraction. The complexes display a tetranuclear cubane‐like core structure [Zn4O4] and sowed good catalytic activity towards the ring‐opening polymerization (ROP ) of rac‐lactide (rac‐LA ) and ε‐caprolactone (ε‐CL ) under solvent‐free conditions. The polylactic acid (PLA ) obtained from rac‐LA showed isotactic enrichment, as proved by homonuclear decoupled 1H‐NMR analysis. These complexes also showed good activity and superior control towards the ROP of rac‐LA and ε‐CL in the presence of benzyl alcohol as a co‐initiator. Furthermore, kinetic studies demonstrated that the ROP of rac‐LA and ε‐CL has a first order dependence on both monomer (rac‐LA and ε‐CL ) and catalyst concentration.  相似文献   

3.
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   

4.
Aluminum‐based salen and salan complexes mediate the ring‐opening polymerization (ROP) of rac‐β‐butyrolactone (β‐BL), rac‐lactide, and ε‐caprolactone. Al‐salen and Al‐salan complexes exhibit excellent control over the ROP of rac‐β‐butyrolactone, yielding atactic poly(3‐hydroxybutyrate) (PHB) with narrow PDIs of <1.15 for Al‐salen and <1.05 for Al‐salan. Kinetic studies reveal pseudo‐first‐order polymerization kinetics and a linear relationship between molecular weight and percent conversion. These complexes also mediate the immortal ROP of rac‐β‐BL and rac‐lactide, through the addition of excess benzyl alcohol of up to 50 mol eq., with excellent control observed. A novel methyl/adamantyl‐substituted Al‐salen system further improves control over the ROP of rac‐lactide and rac‐β‐BL, yielding atactic PHB and highly isotactic poly(lactic acid) (Pm = 0.88). Control over the copolymerization of rac‐lactide and rac‐β‐BL was also achieved, yielding poly(lactic acid)‐co‐poly(3‐hydroxybutyrate) with narrow PDIs of <1.10. 1H NMR spectra of the copolymers indicate a strong bias for the insertion of rac‐lactide over rac‐β‐BL. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013  相似文献   

5.
A novel hydroxy‐, methoxy‐, and phenoxy‐bridge “Mitsubishi emblem” tetranuclear aluminum complex ( 1 ) is synthesized from an unsymmetric amine‐pyridine‐bis(phenol) N2O2‐ligand (H2L1) and a symmetric amine‐tris(phenol) NO3‐ligand (H2L2). Two same configuration chiral nitrogen atoms are formed in the tetranuclear Al complex upon coordination of the unsymmetric tertiary amine ligand to central Al. Complex 1 initiates controlled ring‐opening polymerization (ROP) of rac‐lactide and afford polylactide (PLA) with narrow molecular weight distributions (Mw/Mn = 1.05–1.19). The analysis of 1H NMR spectra of the oligomer indicates that the methoxy group is the initiating group and the ring‐opening polymerization of lactide follows a coordination‐insertion mechanism. The Homonuclear decoupled 1H NMR spectroscopy suggests the isotactic‐rich chains is preferentially formed in PLA. The study on kinetics of the ROP of lactide reveals the homopropagation rate is higher than the cross‐propagation rate, which is in agreement with the observed isotactic selectivity in the ROP of rac‐lactide. The stereochemistry of the polymerization was also supported by activation parameters. The introduction of unsymmetric ligand H2L1 has an effect on stereoslectivity of polymerization. This result may be of interest for the design of multinuclear metal complex catalysts containing functionalized ligands. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55, 2084–2091  相似文献   

6.
The ring-opening polymerization (ROP) of cyclic esters/carbonates is a crucial approach for the synthesis of biocompatible and biodegradable polyesters. Even though numerous efficient ROP catalysts have been well established, their toxicity heavily limits the biomedical applications of polyester products. To solve the toxicity issues relating to ROP catalysts, we report herein a biocompatible coordination network, CZU-1, consisting of Zn44-O)(COO)6 secondary building units (SBUs), biomedicine-relevant organic linkers and guest water, which demonstrates high potential for use in the catalytic ROP synthesis of biomedicine-applicable polyesters. Both experimental and computational results reveal that the guest water in CZU-1 plays crucial roles in the activation of the Zn44-O)(COO)6 SBUs by generating μ4-OH Brønsted acid centers and Zn–OH Lewis acid centers, having a synergistic effect on the catalytic ROP of cyclic esters. Different to the mechanism reported in the literature, we propose a new reaction pathway for the catalytic ROP reaction, which has been confirmed using density functional theory (DFT) calculations, in situ diffuse reflectance IR Fourier transform spectroscopy (DRIFTS), and matrix-assisted laser desorption/ionization time-of-flight mass spectroscopy (MALDI-TOF MS). Additionally, the hydroxyl end groups allow the polyester products to be easily post-modified with different functional moieties to tune their properties for practical applications. We particularly expect that the proposed catalytic ROP mechanism and the developed catalyst design principle will be generally applicable for the controlled synthesis of biomedicine-applicable polymeric materials.

A new ring-opening polymerization mechanism is unveiled based on synergistic catalysis involving Brønsted and Lewis acid centers in a coordination framework.  相似文献   

7.
Bivalent organolanthanides with unbridged substituted indenyl or fluorenyl ligands (1-SiMe3Ind)2YTHF ( I ), (9-SiMe3Flu)2YTHF ( II ), were found to efficiently catalyze the stereoregular polymerization of methyl methacrylate. The microstructure of resultant polymers was shown to be dependent of a conformation that the ligands adopt at the polymerization temperature. The formation of isotactic rich PMMAs from complex II was proposed to be associated with the fluctuation of the 9-trimethylsilylfluorenyls around a C2 symmetric twisted-conformation. The formation of the multi(syndioPMMA-block-iso-PMMA) polymers from the mixture of rac- and meso-isomers of I was rationalized on the basis of competing conjugate addition and inversion of the metallocene conformation. Surprisingly, both rac- and meso-isomers of I were found to operate with similar activity and stereospecifity, although the stereospecific operation of the meso-form was not completely understandable. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 1599–1606, 1998  相似文献   

8.
The dimethylamine in the adducts [(HNMe2)B(azolyl)3] (azolyl=methimazolyl, pyrazolyl), obtained by reaction of the azole with B(NMe2)3, can readily be substituted with a range of nitrogen donors to provide new charge‐neutral, tripodal ligands in high yield. This observation has led to a revision of an earlier interpretation of the mechanism of the formation of these species. The donor properties of the ligands [(nmi)B(azolyl)3] (nmi=N‐methylimidazole) have been compared with their anionic analogues [HB(azolyl)3]? by synthesis of their manganese(I)–tricarbonyl complexes and comparison of their infrared νCO energies. This comparison indicates that the new neutral ligands are only marginally weaker donors than the corresponding anionic hydrotris(azolyl)borate ligands. This may be explained by the ability of the attached nmi ring to stabilize a positive charge remotely from the coordinated metal, which may also account for the fact that the [(nmi)B(pyrazolyl)3] ligand is a substantially stronger donor than the similarly neutral tris(pyrazolyl)methane ligand.  相似文献   

9.
Crystallization behaviors and spherulitic morphology of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) [P(3HB-co-4HB)] with different 4-hydroxybutyrate (4HB) molar fraction were investigated by differential scanning calorimetry and polarized optical microscopy. Crystallization behaviors of P(3HB-co-4HB) are significantly affected by 4HB molar fraction. The melting temperature (T m), glass transition temperature (T g), and crystallinity (X c) decrease with the increase of 4HB molar fraction. Banded spherulites are observed in poly (3-hydroxybutyrate) (PHB) and P(3HB-co-4HB) copolymers. The band spacing decreases with the increase of 4HB molar fraction. The morphology and growth rate of the spherulites strongly depend on 4HB molar fraction and the crystallization temperatures. The introduction of 4HB unit can inhibit the emergence of cracks in PHB spherulites.  相似文献   

10.
Heterometallic cooperativity is an emerging strategy to elevate polymerisation catalyst performance. Here, we report the first heterotrimetallic Na/Zn2 and K/Zn2 complexes supported by a ProPhenol ligand, which deliver “best of both” in cyclic ester ring-opening polymerisation, combining the outstanding activity (Na/K) and good control (Zn2) of homometallic analogues. Detailed NMR studies and density-functional theory calculations suggest that the Na/Zn2 and K/Zn2 complexes retain their heterometallic structures in the solution-state. To the best of our knowledge, the K/Zn2 analogue is the most active heterometallic catalyst reported for rac-lactide polymerisation (kobs = 1.7 × 10−2 s−1), giving activities five times faster than the Na/Zn2 complex. These versatile catalysts also display outstanding performance in ε-caprolatone and δ-valerolactone ring-opening polymerisation. These studies provide underpinning methodologies for future heterometallic polymerisation catalyst design, both in cyclic ester polymerisation and other ring-opening (co)polymerisation reactions.

Cooperative heterotrimetallic Na/Zn2 and K/Zn2 complexes combine the excellent activities and control of the homometallic analogues, giving “best of both” in cyclic ester ring-opening polymerisation.  相似文献   

11.
Ring opening polymerization (ROP) of (R,S)-β-butyrolactone (BL) using dibutylmagnesium (Bu2Mg) as initiator was investigated both in bulk and in solution. The synthetic poly-3-hydroxybutyrates (P3HB) were characterized by 1H NMR, 13C NMR, FT-IR and GPC. Effects of molar ratio of initiator to monomer, reaction temperature and time on the monomer conversion and the polymer molecular weight and its distribution were discussed. The kinetics of the solution polymerization of BL was examined and showed a first order both in monomer concentration and initiator concentration. The end groups analysis suggested that the monomer inserted into the growing chain proceeding through the coordination-insertion mechanism based on the acyl-oxygen bond scission rather than the alkyl-oxygen bond cleavage of the BL ring. Furthermore, a possible mechanism for the initiation and propagation procedures of P3HB synthesized from BL with Bu2Mg was proposed.  相似文献   

12.
Colin Baillie 《Tetrahedron》2004,60(19):4159-4168
Monodentate, biphenyl-type phosphines have emerged as a powerful class of ligands in homogeneous catalysis. Synthetic methods for these ligands are limited, however. We report that the palladium-catalysed Suzuki coupling of OPR2(o-C6H4X) (R=Ph, t-Bu; X=Br, I) with arylboronic acids affords a variety of biaryl phosphine oxides including those that contain heterocycles. The corresponding phosphines are readily obtained by treatment with HSiCl3. The methodology provides an easy entry to monodentate biaryl and heterobiaryl P?X (X=N, O, S) phosphines with diverse steric and electronic properties.  相似文献   

13.
Isotactic polypropylene block copolymers, isotactic-polypropylene-block-poly (methyl methacrylate) (i-PP-b-PMMA) and isotactic-polypropylene-block-polystyrene (i-PP-b-PS), were prepared by atom transfer radical polymerization (ATRP) using a brominated styrene-terminated isotactic polypropylene macroinitiator synthesized from bromination of styrene-terminated isotactic polypropylene. The styrene-terminated isotactic polypropylene can be obtained by polymerization of propylene in the presence of styrene and hydrogen chain transfer agents using a rac-Me2Si[2-methyl-4-(1-naphyl)Ind]2ZrCl2 as catalyst. The molecular weights of isotactic polypropylene block copolymers were controlled by altering the amount of hydrogen used in the polymerization of propylene and the amount of monomer used in the blocking reaction. The effect of i-PP-b-PS block copolymer on PP-PS blends and that of i-PP-b-PMMA block copolymer on PP-PMMA blends were studied by scanning electron microscopy.  相似文献   

14.
Heteroligand europium cinnamates with nitrogen- and phosphorus-containing neutral ligands were studied by luminescent and X-ray electron spectroscopy. A tendency toward an increase in the electron density on Eu3+ was observed when neutral ligands were coordinated by donor nitrogen atoms. The Δν0, ΔF 1, and I(F 0)/I(F 2) values monotonically decreased as the donor properties of neutral ligands increased because of strengthening of the interaction of Eu3+ with the ligands.  相似文献   

15.
Bulk propylene polymerization in the presence of ansa-metallocenes with C 2 and C 1 symmetries has been studied. The catalytic activity, polymerization kinetics, and the molecular weight of polypropylene (PP) depend strongly on catalyst formation conditions. Mixtures of rac and meso isomers of metallocenes make it possible to rapidly produce a high-molecular-weight isotactic PP with high stereoregularity and mechanical characteristics and thus skip the stage of the isolation of pure rac isomer in the catalyst synthesis. The ability of triisobutylaluminum to serve as a cocatalyst is studied for ansa-metallocenes of C 1 symmetry. In this case, the molecular weight of PP is higher, indicating that organoaluminum compounds participate in chain termination reactions. An increase in the reaction temperature results in an increase in the stereoregularity and crystallinity of PP. Polypropylene synthesized using ansa-metallocenes of C 1 symmetry has good elastomeric properties.  相似文献   

16.
Cyclopolymerization of nonconjugated dienes produces poly(methylene‐1,3‐cycloalkanes) and provides a pathway to a number of stereochemically complex polymers. Activation of a diastereomeric mixture of a six‐membered metallacycle complex (rac‐ 1 ) in the presence of 1,5‐hexadiene produced poly(methylene‐1,3‐cyclopentane) (PMCP) with >98% cyclization of the diene monomer. The catalyst was found to cyclopolymerize 1,5‐hexadiene with relatively high activity. The microstructure of the PMCP furnished by rac‐ 1 was found to contain a high proportion of cis‐cyclopentane rings (σ = 0.70–0.74) and a relatively high isotactic content (α = 0.93–0.96). These are the first cis‐enriched isotactic cyclopolymers of 1,5‐hexadiene. Cyclopolymerization of 1,6‐heptadiene with rac‐ 1 /B(C6F5)3 produced poly(methylene‐1,3‐cyclohexane) containing 97% cis‐isotactic rings. This is the first report of this highly isotactic and diastereomerically‐pure microstructure.

  相似文献   


17.
The rhenium complex, [K(DME)(18-c-6)][ReH4(Bpin)(η2-HBpin)(κ2-H2Bpin)] 1, comprising hydride and boron ligands only, has been synthesized by exhaustive deoxygenation of the commercially available perrhenate anion (ReO4) with pinacol borane (HBpin). The structure of 1 was analysed by X-ray crystallography, NMR spectroscopy, and DFT calculations. While no hydrides were located in the X-ray crystal structure, it revealed a trigonal arrangement of pinacol boron ligands. Variable-temperature NMR spectroscopy supported the presence of seven hydride ligands but further insight was hindered by the fluxionality of both hydride and boron ligands at low temperature. Further evaluation of the structure by Ab Initio Random Structure Searching (AIRSS) identified the presence of hydride, boryl, σ-borane, and dihydroborate ligands. This complex, either isolated or prepared in situ, is a catalyst for the 1,4-hydroboration of N-heteroaromatic substrates under simple operating procedures. It also acts as a reagent for the stoichiometric C–H borylation of toluene, displaying high meta regioselectivity in the borylated products. Reaction of 1 with 9-BBN resulted in HBpin substitution to form the new anionic tetra(dihydroborate) complex [K(DME)(18-c-6)][Re(κ2-H-9-BBN)4] 4 for which the hydride positions were clearly identified by X-ray crystallography. The method used to generate these isolable yet reactive boron–hydride complexes is direct and straightforward and has potential utility for the exploitation of other metal oxo compounds in operationally simple catalytic reactions.

Exhaustive deoxygenation of perrhenate by pinacol borane forms a new Re anion of boron and hydride ligands only that undergoes borane ligand exchange, stoichiometric C–H boration, and catalytic pyridine hydroboration.  相似文献   

18.
Yttrium [amino‐alkoxy‐bis(phenolate)]amido complexes have been used for the ring‐opening polymerization (ROP) of racemic alkyl β‐malolactonates (4‐alkoxycarbonyl‐2‐oxetanones, rac‐MLARs) bearing an allyl (All), benzyl (Bz) or methyl (Me) lateral ester function. The nature of the ortho‐substituent on the phenolate rings in the metal ancillary dictated the stereocontrol of the ROP, and consequently the syndiotactic enrichment of the resulting polyesters. ROP promoted by catalysts with halogen (Cl, Br)‐disubstituted ligands allowed the first reported synthesis of highly syndiotactic PMLARs (Pr ≥ 0.95); conversely, catalysts bearing bulky alkyl and aryl ortho‐substituted ligands proved largely ineffective. All polymers have been characterized by 1H and 13C{1H} NMR spectroscopy, MALDI‐ToF mass spectrometry and DSC analyses. Statistical and thermal analyses enabled the rationalization of the chain‐end control mechanism. Whereas the stereocontrol of the polymerization obeyed a Markov first‐order (Mk1) model for the ROP of rac‐MLABz and rac‐MLAAll, the ROP of rac‐MLAMe led to a chain end‐control of Markov second‐order type (Mk2). DFT computations suggest that the high stereocontrol ability featured by catalysts bearing Cl‐ and Br‐substituted ligands does not likely originate from halogen bonding between the halogen substituent and the growing polyester chain.  相似文献   

19.
The role of Lewis bases in MgCl2 supported catalysts for olefin polymerization is a subject of continuous interest and discussion in order to obtain more and more active and stereospecific catalysts and to explain their stereoregulating mechanism. Through molecular calculation and conformational analysis it was possible to identify chelating diethers that have the correct oxygen-oxygen distance necessary to tightly coordinate with the Mg ions of the support, even in the presence of other strong Lewis acids, and unable to give secondary reactions with TiCl4, AlR3, Ti-C and Ti-H bonds. The use of these donors has allowed the synthesis of catalytic systems that are both highly active and stereospecific even in the absence of external donors. Kinetic data of propylene polymerization with these catalyst systems are reported. The importance of the distance between the donor atoms in bifunctional Lewis bases has been proved also in the case of new classes of internal donors. Molecular modelling studies have enabled us to formulate models of active sites, located on some corners of MgCl2 crystallites, whose chirality is induced by the presence of a donor molecule in their environment. These models could explain, at least in part, the exceptional increase of isotactic polymer productivity observed for stereospecific catalyst systems, containing only the internal donor, with respect to catalysts lacking the Lewis base and could account for the influence of the donor on the molecular properties of the obtained polymers.  相似文献   

20.
Chiral nonbonding interaction with N-protected amino acid methyl esters used as chiral additives in achiral solvents allows dynamic induction of single-handed helical conformation in poly(quinoxaline-2,3-diyl)s (PQX) bearing only achiral substituents. Ac-l-Pro-OMe, for instance, allows induction of energy preference of 0.16 kJ mol−1 per monomer unit for the M-helical structure over the P-helix in t-butyl methyl ether (MTBE). With this new mode of screw-sense induction, homochiral screw-sense has been induced in virtually achiral poly(quinoxaline-2,3-diyl)s 1000-mer containing phosphine pendants (PQXphos). Use of PQXphos as a helically dynamic ligand along with Ac-Pro-OMe (l or d) as a chiral additive in MTBE allowed a highly enantioselective Suzuki–Miyaura coupling reaction with up to 95% enantiomeric excess.

Achiral poly(quinoxaline-2,3-diyl) containing Ar2P groups undergo dynamic induction of M-helical conformation through nonbonding interaction with protected AA such as Ac-l-Pro-OMe, serving as a chiral ligand in asymmetric cross-coupling with up to 95% ee.  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