Next Article in Journal
Polyaspartate-Ionene/Na+-Montmorillonite Nanocomposites as Novel Adsorbent for Anionic Dye; Effect of Ionene Structure
Next Article in Special Issue
Enhanced Photovoltaic Properties of Perovskite Solar Cells by Employing Bathocuproine/Hydrophobic Polymer Films as Hole-Blocking/Electron-Transporting Interfacial Layers
Previous Article in Journal
Mechanical Properties of Natural-Fiber-Reinforced Biobased Epoxy Resins Manufactured by Resin Infusion Process
Previous Article in Special Issue
Electrosynthesis and Electrochromism of a New Crosslinked Polydithienylpyrrole with Diphenylpyrenylamine Subunits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Terpyridine End-Modified Polystyrenes through ATRP for Facile Construction of Metallo-Supramolecular P3HT-b-PS Diblock Copolymers

Department of Chemistry, National Taiwan University, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
Polymers 2020, 12(12), 2842; https://doi.org/10.3390/polym12122842
Submission received: 9 November 2020 / Revised: 25 November 2020 / Accepted: 26 November 2020 / Published: 29 November 2020

Abstract

:
Complementary complexation between 2,2′:6′,2″-terpyridine (tpy) and 6,6″-dianthracenyl-substituted tpy in the presence of Zn(II) ions provided an efficient strategy for construction of metallo-supramolecular diblock copolymers. To synthesize well-defined tpy-modified polystyrenes (PSs), an Fe(II) bis(tpy) complex bearing α-bromoester as a metallo-initiator was applied to atom transfer radical polymerization (ATRP) to avoid poisoning the Cu(I) catalyst. Subsequently, a series of tpy-functionalized PSs was obtained after the decomplexation of <tpy-Fe(II)-tpy> junction by tetrakis(triethylammonium) ethylenediaminetetraacetate (TEA-EDTA) under mild conditions. The metallo-supramolecular poly(3-hexylthiophene) (P3HT)-block-PS diblock copolymers were prepared by simply mixing the corresponding terminally tpy-modified homopolymers with Zn(II) ions, and further characterized by 1H NMR and diffusion ordered spectroscopy (DOSY) experiments. The approach using metallo-initiators for ATRP offers an opportunity to construct tpy-functionalized polymers with controllable molecular weights and low polydispersities. Through the spontaneous heteroleptic complexation, a variety of metallo-supramolecular diblock copolymers with tunable block ratios can be easily constructed.

1. Introduction

Metallo-supramolecular block copolymers are generated through metal–ligand coordination of ligand end-modified homopolymers with proper metal ions [1]. Various ligands have been utilized in construction of diverse copolymers, such as 2,2′-bipyridine (bpy) [2,3,4], 2,2′:6′,2″-terpyridine (tpy) [5,6,7,8,9,10,11,12,13], and pincer-type ligands [14,15,16,17,18]. Among them, bis(tpy) complexes are commonly used for preparation of diblock copolymers in a stepwise manner [6,7,8,9,10,11,12]. However, the challenge remains to prevent unwanted homoleptic complexation and tedious purification processes. Recently, we demonstrated that the spontaneous heteroleptic complexation between unsubstituted tpy and 6,6″-dianthracenyl-substituted tpy could be achieved upon the addition of Zn(II) ions under ambient conditions, and was further applied to the construction of poly(3-hexylthiophene)-block-poly(ethylene oxide) (P3HT-b-PEO) copolymers [19]. The complementary ligand pairing indeed provides an efficient approach for synthesis of metallo-supramolecular diblock copolymers.
In order to incorporate tpy motifs into a wide variety of well-defined polymers, the tpy-modified chain-transfer agents or initiators have been feasibly utilized in controlled polymerizations, such as reversible addition-fragmentation chain transfer (RAFT) [20,21,22], nitroxide-mediated radical polymerization (NMP) [23,24,25,26], and ring-opening polymerization (ROP) [27,28,29]. However, since the atom transfer radical polymerization (ATRP) is often mediated by copper catalysts, the presence of uncomplexed tpy ligands could adversely affect the copper catalyst during the ATRP reaction [30,31]. To attenuate the ligand interference, bpy ligands have been successfully introduced into a polymeric structure via the ATRP initiated by bpy-based metallo-initiators [32,33,34,35,36,37,38,39,40]. For example, the Ru(II), Fe(II), and Zn(II) tris(bpy) complexes bearing chloromethylpyridine and α-bromoester have been employed in the ATRP of styrene and methyl methacrylate [41,42], where the use of metallo-initiators with coordinated bpy ligands efficiently prevented poisoning of copper catalysts. Therefore, inspired by the aforementioned researches, herein the use of an Fe(II) bis(tpy) complex as metallo-initiators in ATRP is realized for the synthesis of tpy end-modified polystyrenes (PSs), which is evident from the kinetic study as well as 1H NMR and gel permeation chromatography (GPC) analyses. Subsequently, the well-defined tpy-functionalized PSs obtained after decomplexation are successfully applied to the construction of metallo-supramolecular diblock copolymers (P3HT-Zn-PS), where the complexation and self-assembly characteristics are investigated by diffusion ordered spectroscopy (DOSY) NMR experiments.

2. Materials and Methods

2.1. Materials

Styrene (Showa, 99%, Tokyo, Japan) and acetonitrile (Fischer, HPLC grade, Loughborough, UK) were distilled over CaH2 under reduced pressure. N,N,N′,N″,N″-Pentamethyldiethylenetriamine (PMDETA) (Acros, 98%, Morris Plains, NJ, USA) was distilled over KOH under reduced pressure, and then degassed by three freeze-pump-thaw cycles in a Schlenk tube before use. CuBr (Alfa Aesar, 99%, Ward Hill, MA, USA) was purified by stirring overnight in AcOH, filtered, washed with absolute EtOH and diethyl ether, and then dried under vacuum. Unless otherwise noted, reagents and solvents were used as received from Fisher Scientific (Loughborough, UK) and Sigma-Aldrich (St. Louis, MO, USA) without further purification. 4′-(4-Hydroxyphenyl)-2,2′:6′,2″-terpyridine (1) [43], 9-anthraceneboronic acid [44], tetrakis(triethylammonium) ethylenediaminetetraacetate (TEA-EDTA) [45], 2,5-dibromo-3-hexylthiophene [46], mono-brominated P3HT (Br–P3HT) [19], 6,6″-di(anthracen-9-yl)-4′-(4-methoxyphenyl)-2,2′:6′,2″-terpyridine (L2) [19], 4′-(4-methoxyphenyl)-2,2′:6′,2″-terpyridine (L3) [47], and (L2–Zn–L3) complex [48] were prepared according to the reported procedures.

2.2. Methods

1H, 13C, and DOSY NMR experiments were performed at 25 °C on a Varian Mercury NMR 400 spectrometer, where chemical shifts (δ in ppm) were determined with respect to the nondeuterated solvents as a reference. Gel permeation chromatography (GPC) was conducted on the instrument equipped with two columns (Shodex KF-803 and KF-804, Tokyo, Japan), a Waters 515 HPLC pump, and a differential refractive index detector (LabAlliance RI2000, New York, USA). Tetrahydrofuran (THF) mixed with tetrabutylammonium bromide (TBAB) (1 wt%) was utilized as an eluent at a flow rate of 1 mL min−1 at 40 °C [49]. The calibration curve was established by linear polystyrene standards. MALDI-TOF-MS measurements were conducted on a Bruker Autoflex Speed MALDI-TOF-MS (Bruker Daltonics, Billerica, MA, USA). The polystyrene samples for MALDI-TOF-MS measurements were prepared by mixing the polymer (5.0 mg mL−1), dithranol (DIT) (10 mg mL−1), and trifluoroacetic acid (TFA) (1%) in THF.

2.3. General Procedure for Complexation Reactions

To a CHCl3 solution (5 mL) of PSn (n = 19, 33, 85, 106, 161, and 235) and P3HT54 in an equimolar ratio calculated from the corresponding number average molecular weights (Mn,NMR), 1 Eq of Zn(OTf)2 in MeOH (5 mL) was added. After the reaction mixture was stirred at room temperature for 5 min, the solvent was evaporated under reduced pressure to give the corresponding diblock copolymers.

3. Results and Discussion

Inspired by the pioneering research of metallo-initiators for ATRP, we designed and synthesized an Fe(II) bis(tpy) complex bearing α-bromoester as a bifunctional metallo-initiator (4) (Scheme 1). The α-bromoester modified tpy-based initiator (3) was synthesized from the precursor 4’-(4-hydroxyphenyl)-tpy (1) in moderate yield. First, 1 was alkylated by 2-chloroethanol to give 4′-(4-(2-hydroxyethoxy)phenyl)-tpy (2). The following esterification was achieved by reaction of 2 with α-bromoisobutyryl bromide in the presence of triethylamine at room temperature to afford compound 3. Subsequently, the complexation was conducted by adding a MeOH solution of FeCl2 (0.5 Eq) into a CHCl3 solution of ligand 3 (1 Eq) at room temperature, followed by counter-anion exchange with NH4PF6 (10 Eq), to yield 4 as a dark purple powder in quantitative yield. The suitable crystals for single-crystal X-ray crystallography were obtained by slow diffusion of diethyl ether into an MeCN solution of 4. The structure of complex 4 was unequivocally established by its crystal structure (Figure S8), NMR spectroscopy (Figures S5 and S6) and ESI-MS (Figure S7).
Our attempts to conduct Cu(I)-mediated ATRP of styrene using the initiator 3 were unsuccessful possibly due to the strong chelating ability of tpy ligands acting as a catalyst poison. To investigate the controllability of ATRP of styrene initiated by the bifunctional metallo-initiator 4, a kinetic study was performed via a typical ATRP protocol as follows. Initiator 4 (45.2 mg, 34.9 μmol) and CuBr (15.0 mg, 104.8 μmol) were added to a degassed Schlenk flask equipped with a stir bar. Subsequently, MeCN (1.3 mL) and styrene (3.6 mL, 31.4 mmol) were added into the flask, which was degassed by three freeze-pump-thaw cycles, followed by the addition of PMDETA (21.9 μL, 104.8 μmol), and then stirred at 110 °C. The polymerization solution was periodically sampled via a pre-degassed syringe to monitor the conversion of monomer by 1H NMR and calculate the theoretical molecular weight (Mn,theo). The sampled solution was further treated with TEA-EDTA in dimethylformamide (DMF) for 1 day at room temperature to decomplex the <tpy–Fe(II)–tpy> junction to afford the tpy end-modified PSs. The corresponding Mn,GPC and PDI (Mw/Mn) were determined by GPC, and Mn,NMR was calculated from the 1H NMR peak integral ratios of polymerized styrene and terminal tpy.
The kinetic study on ATRP of styrene initiated by the metallo-initiator 4 was summarized in Figure 1. The semilogarithmic kinetic plot of ln([M]0/[M]) versus reaction time indicated the first-order radical polymerization process and the radical concentration was kept constant during the polymerization (Figure 1a). In addition, the experimental molecular weights (Mn,NMR) were in good agreement with the theoretical ones (Mn,theo) and linearly increased with respect to the monomer conversion (Figure 1b), implying the absence of significant chain transfer reactions. Moreover, the polydispersities were decreased with increasing conversion (Figure 1b), and a clear shift to higher molecular weights with a mono-distribution was evidenced by GPC traces (Figure 1c). These observations suggested that the well-controlled ATRP of styrene could be initiated by 4.
Based on the kinetic result of ATRP of styrene, a series of tpy-functionalized PSn (n = 19, 33, 85, 106, 161, and 235) with varying chain lengths was prepared via ATRP under optimized conditions (Table 1). The formation of well-defined PSn could be evident from the narrow molecular weight distributions and the consistency between Mn,NMR and Mn,GPC. It is noteworthy that the terminal bromide at PS chain-ends susceptible to elimination easily led to formation of a double bond during MALDI-TOF-MS measurements [50]. Nevertheless, the high fidelity in the tpy chain-end functionality was verified by the corresponding MALDI-TOF-MS peaks (Figures S10, S12, S14, S16, and S18). On the other hand, the well-defined P3HT54 (DP = 54, Mn,GPC = 8800 Da, Mw/Mn = 1.23) end-functionalized with a 4-(4′-(6,6″-dianthracenylterpyridyl))phenyl group was obtained through the Suzuki–Miyaura coupling reaction of mono-brominated P3HT (Br–P3HT) with 6,6″-dianthracenyl-4′-(4-boronophenyl)tpy (Figure 2a) [19]. Notably, Br–P3HT prepared by Grignard metathesis (GRIM) polymerization method possessed two isomeric chain-end structures, i.e., head-to-head and head-to-tail orientations, which could not be differentiated by the 1H NMR spectrum of Br–P3HT but clearly seen in that of P3HT54 (Figure 2b) [51,52]. Therefore, the 3-hexylthiophene coupled with two 6,6″-dianthracenyl-substituted tpys (L1) was synthesized as a model compound to ensure the proper 1H NMR assignments. The two sets of tpy signals of L1 corresponded to two types of chain-end connections, and the chain-end head-to-tail content of P3HT54 was estimated to be 22% (Figure S26). The single molecular weight distribution in the MALDI-TOF-MS spectra (Figure S27) strongly supported the high chain-end functionality for P3HT54.
We have demonstrated that the complementary complexation between 6,6″-substituted and unsubstituted tpy ligands with Zn(II) under ambient conditions could be applied to construction of the metallo-supramolecular diblock copolymers from two distinct tpy-modified homopolymers [19,48]. In the ligand design, the bulky 9-anthracenyl substituents effectively decelerated the formation rate of homoleptic complexes. Moreover, the X-ray single-crystal structure of [L2–Zn–L3] (Figure 3a) exhibited the π–π interactions between unsubstituted L3 and two anthracenyl substituents to facilitate the formation of heteroleptic complexes. Consequently, a series of metallo-supramolecular diblock copolymers of [P3HT54–Zn–PSn] (n = 19, 33, 85, 106, 161, and 235) could be readily constructed from homopolymers PSn and P3HT54 in the presence of Zn(II) ions (Scheme 2). Due to the labile coordination bonds, the intact copolymers could not be detected by MALDI-TOF-MS [53]. Hence, the resultant diblock copolymers were characterized by 1H NMR experiments (Figure 3a). The 1H NMR spectra of [P3HT54–Zn–PSn] strongly supported the formation of the desired heteroleptic junctions between P3HT54 and PSn as compared with that of the model complex [L2–Zn–L3]. In addition, the diffusion ordered spectroscopy (DOSY) of [P3HT54–Zn–PSn] (5 mg mL−1 in CDCl3) revealed all the relevant 1H NMR resonances have the identical diffusion coefficients (D) for each copolymer (Figure 3b), implying the single distribution of hydrodynamic radii in solution [54]. Accordingly, as the chain length of the PS segment was decreased, the diffusion coefficient was increased for [P3HT54–Zn–PS235] (D = 1.05 ×10−10 m2 s−1), [P3HT54–Zn–PS161] (D = 2.16 × 10−10 m2 s−1), [P3HT54–Zn–PS106] (D = 3.08 × 10−10 m2 s−1), and [P3HT54–Zn–PS85] (D = 7.99 × 10−10 m2 s−1), due to the shrinkage in molecular size. However, [P3HT54–Zn–PS19] (D = 1.80 × 10−10 m2 s−1) and [P3HT54–Zn–PS33] (D = 2.61 × 10−10 m2 s−1) showed much smaller diffusion coefficients than expected, presumably because of the severe intermolecular aggregation, where the shorter PS length attenuated the interference in the assembly of P3HT segments [55].

4. Conclusions

A series of tpy end-modified polystyrenes with controllable molecular weights and narrow polydispersities was successfully prepared using a bifunctional tpy-based metallo-initiator via ATRP and subsequent decomplexation. Based on the pre-designed complementary ligand pairing, metallo-supramolecular diblock copolymers (P3HT-b-PS) were readily constructed from the well-defined P3HT and PS homopolymers end-functionalized with 6,6″-dianthracenyl-substituted tpy and unsubstituted tpy, respectively, upon the addition of Zn(II) ions. The DOSY NMR analysis not only supported the formation of the expected copolymers [P3HT–Zn–PS], but also revealed that the PS chain length would influence the assembly of P3HT segments in solution. We anticipate that the ATRP protocol using tpy-based metallo-initiators along with the complementary ligand pair will provide facile access to construction of various copolymers with enhanced topological diversity and complexity.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/12/12/2842/s1. Scheme S1. Synthesis of metallo-initiator 4; Figures S1–S6. NMR spectra of compounds 2-4; Figure S7. ESI-MS spectrum of 4; Figure S8. X-ray crystal structure of 4; Scheme S2. Synthesis of tpy-functionalized polystyrene PSn (n = 19, 33, 85, 106, 161, and 235) from the metallo-initiator 4; Figures S9–S20. NMR spectra and MALDI-TOF-MS spectra of PSn; Figure S21. GPC traces of PSn; Scheme S3. Synthesis of P3HT54; Figures S22–S25. NMR spectra of compounds 6 and 7; Figures S26–28. NMR, MALDI-TOF-MS, and GPC trace of P3HT54; Scheme S4. Synthesis of L1; Figures S29–S31. NMR spectra of L1; Table S1 and S2. Crystal data and experimental details for 4 and [L2–Zn–L3].

Author Contributions

Conceptualization, investigation, funding acquisition, writing—review and editing, and supervision, Y.-T.C.; methodology, data curation, and writing—original draft, T.-H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Technology of Taiwan (MOST 109-2113-M-002-019-MY3).

Acknowledgments

We gratefully thank S.-L. Huang at MOST (National Taiwan University) for the assistance in NMR experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schubert, U.S.; Eschbaumer, C. Macromolecules Containing Bipyridine and Terpyridine Metal Complexes: Towards Metallosupramolecular Polymers. Angew. Chem. Int. Ed. 2002, 41, 2892–2926. [Google Scholar] [CrossRef]
  2. Fraser, C.L.; Smith, A.P.; Wu, X. Metal Template-Assisted Block Copolymer Synthesis:  Use of Solvent Polarity to Control Chain Conformation and Reactivity at the Metal Core. J. Am. Chem. Soc. 2000, 122, 9026–9027. [Google Scholar] [CrossRef]
  3. Corbin, P.S.; Webb, M.P.; McAlvin, J.E.; Fraser, C.L. Biocompatible Polyester Macroligands:  New Subunits for the Assembly of Star-Shaped Polymers with Luminescent and Cleavable Metal Cores. Biomacromolecules 2001, 2, 223–232. [Google Scholar] [CrossRef] [PubMed]
  4. Bender, J.L.; Corbin, P.S.; Fraser, C.L.; Metcalf, D.H.; Richardson, F.S.; Thomas, E.L.; Urbas, A.M. Site-Isolated Luminescent Europium Complexes with Polyester Macroligands:  Metal-Centered Heteroarm Stars and Nanoscale Assemblies with Labile Block Junctions. J. Am. Chem. Soc. 2002, 124, 8526–8527. [Google Scholar] [CrossRef] [PubMed]
  5. Schubert, U.S.; Eschbaumer, C. Functionalized Oligomers and Copolymers with Metal Complexing Segments: A Simple and High Yield Entry Towards 2,2′:6′,2″-Terpyridine Monofunctionalized Telechelics. Macromol. Symp. 2001, 163, 177–188. [Google Scholar] [CrossRef]
  6. Gohy, J.-F.; Lohmeijer, B.G.G.; Schubert, U.S. Metallo-Supramolecular Block Copolymer Micelles. Macromolecules 2002, 35, 4560–4563. [Google Scholar] [CrossRef]
  7. Chiper, M.; Meier, M.A.R.; Kranenburg, J.M.; Schubert, U.S. New Insights into Nickel(II), Iron(II), and Cobalt(II) Bis-Complex-Based Metallo-Supramolecular Polymers. Macromol. Chem. Phys. 2007, 208, 679–689. [Google Scholar] [CrossRef]
  8. Ott, C.; Wouters, D.; Thijs, H.M.L.; Schubert, U.S. New Preparation and Purification Methods for Metallo-Supramolecular Block Copolymers. J. Inorg. Organomet. Polym. Mater. 2007, 17, 241–249. [Google Scholar] [CrossRef]
  9. Landsmann, S.; Winter, A.; Chiper, M.; Fustin, C.-A.; Hoeppener, S.; Wouters, D.; Gohy, J.-F.; Schubert, U.S. Poly(dimethylsiloxane)-Substituted 2,2′:6,2″-Terpyridines: Synthesis and Characterization of New Amphiphilic Supramolecular Diblock Copolymers. Macromol. Chem. Phys. 2008, 209, 1666–1672. [Google Scholar] [CrossRef]
  10. Gohy, J.-F.; Chiper, M.; Guillet, P.; Fustin, C.-A.; Hoeppener, S.; Winter, A.; Hoogenboom, R.; Schubert, U.S. Self-Organization of Rod–Coil Tri- and Tetra-Arm Star Metallo-Supramolecular Block Copolymers in Selective Solvents. Soft Matter 2009, 5, 2954–2961. [Google Scholar] [CrossRef]
  11. Gohy, J.-F.; Ott, C.; Hoeppener, S.; Schubert, U.S. Multicompartment Micelles From a Metallo-Supramolecular Tetrablock Quatercopolymer. Chem. Commun. 2009, 6038–6040. [Google Scholar] [CrossRef] [PubMed]
  12. Mugemana, C.; Guillet, P.; Hoeppener, S.; Schubert, U.S.; Fustin, C.-A.; Gohy, J.-F. Metallo-Supramolecular Diblock Copolymers Based on Heteroleptic Cobalt(III) and Nickel(II) Bis-terpyridine Complexes. Chem. Commun. 2010, 46, 1296–1298. [Google Scholar] [CrossRef] [PubMed]
  13. Brzeziński, M.; Kacprzak, A.; Calderón, M.; Seiffert, S. Metallo-Polymer Chain Extension Controls the Morphology and Release Kinetics of Microparticles Composed of Terpyridine-Capped Polylactides and Their Stereocomplexes. Macromol. Rapid Commun. 2017, 38, 1600790. [Google Scholar] [CrossRef] [PubMed]
  14. Ambade, A.V.; Yang, S.K.; Weck, M. Supramolecular ABC Triblock Copolymers. Angew. Chem. 2009, 121, 2938–2942. [Google Scholar] [CrossRef]
  15. Moughton, A.O.; Stubenrauch, K.; O’Reilly, R.K. Hollow Nanostructures from Self-Assembled Supramolecular Metallo-Triblock Copolymers. Soft Matter 2009, 5, 2361–2370. [Google Scholar] [CrossRef]
  16. Elacqua, E.; Manning, K.B.; Lye, D.S.; Pomarico, S.K.; Morgia, F.; Weck, M. Supramolecular Multiblock Copolymers Featuring Complex Secondary Structures. J. Am. Chem. Soc. 2017, 139, 12240–12250. [Google Scholar] [CrossRef]
  17. Lye, D.S.; Xia, Y.; Wong, M.Z.; Wang, Y.; Nieh, M.-P.; Weck, M. ABC Supramolecular Triblock Copolymer by ROMP and ATRP. Macromolecules 2017, 50, 4244–4255. [Google Scholar] [CrossRef]
  18. Deng, R.; Milton, M.; Pomarico, S.K.; Weck, M. Synthesis of a Heterotelechelic Helical Poly(methacrylamide) and Its Incorporation Into a Supramolecular Triblock Copolymer. Polym. Chem. 2019, 10, 5087–5093. [Google Scholar] [CrossRef]
  19. He, Y.-J.; Tu, T.-H.; Su, M.-K.; Yang, C.-W.; Kong, K.V.; Chan, Y.-T. Facile Construction of Metallo-Supramolecular Poly(3-hexylthiophene)-block-Poly(ethylene oxide) Diblock Copolymers via Complementary Coordination and Their Self-Assembled Nanostructures. J. Am. Chem. Soc. 2017, 139, 4218–4224. [Google Scholar] [CrossRef]
  20. Zhou, G.; Harruna, I.I. Synthesis and Characterization of Bis(2,2′:6′,2″-terpyridine)ruthenium(II)-Connected Diblock Polymers via RAFT Polymerization. Macromolecules 2005, 38, 4114–4123. [Google Scholar] [CrossRef]
  21. Zhang, L.; Zhang, Y.; Chen, Y. Synthesis of Bis(2,2′:6′,2″-terpyridine)-Terminated Telechelic Polymers by RAFT Polymerization and Ruthenium–Polymer Complexation Thereof. Eur. Polym. J. 2006, 42, 2398–2406. [Google Scholar] [CrossRef]
  22. Elkin, T.; Copp, S.M.; Hamblin, R.L.; Martinez, J.S.; Montaño, G.A.; Rocha, R.C. Synthesis of Terpyridine-Terminated Amphiphilic Block Copolymers and Their Self-Assembly into Metallo-Polymer Nanovesicles. Materials 2019, 12, 601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Aamer, K.A.; Tew, G.N. Synthesis of Terpyridine-Containing Polymers with Blocky Architectures. Macromolecules 2004, 37, 1990–1993. [Google Scholar] [CrossRef]
  24. Lohmeijer, B.G.G.; Schubert, U.S. The LEGO Toolbox: Supramolecular Building Blocks by Nitroxide-Mediated Controlled Radical Polymerization. J. Polym. Sci. A Polym. Chem. 2005, 43, 6331–6344. [Google Scholar] [CrossRef]
  25. Tew, G.N.; Aamer, K.A.; Shunmugam, R. Incorporation of Terpyridine into the Side Chain of Copolymers to Create Multi-Functional Materials. Polymer 2005, 46, 8440–8447. [Google Scholar] [CrossRef]
  26. Ott, C.; Lohmeijer, B.G.G.; Wouters, D.; Schubert, U.S. Terpyridine-Terminated Homo and Diblock Copolymer LEGO Units by Nitroxide-Mediated Radical Polymerization. Macromol. Chem. Phys. 2006, 207, 1439–1449. [Google Scholar] [CrossRef]
  27. Heller, M.; Schubert, U.S. Optically Active Supramolecular Poly(L-lactide)s End-Capped with Terpyridine. Macromol. Rapid Commun. 2001, 22, 1358–1363. [Google Scholar] [CrossRef]
  28. Hofmeier, H.; Hoogenboom, R.; Wouters, M.E.L.; Schubert, U.S. High Molecular Weight Supramolecular Polymers Containing Both Terpyridine Metal Complexes and Ureidopyrimidinone Quadruple Hydrogen-Bonding Units in the Main Chain. J. Am. Chem. Soc. 2005, 127, 2913–2921. [Google Scholar] [CrossRef]
  29. Winter, A.; Schubert, U.S. New Polyester-Based Terpyridine Macroligands and Their Blue Iron(II) Complexes. Macromol. Chem. Phys. 2007, 208, 1956–1964. [Google Scholar] [CrossRef]
  30. Tzanetos, N.P.; Andreopoulou, A.K.; Kallitsis, J.K. Side-Chain Terpyridine Polymers through Atom Transfer Radical Polymerization and Their Ruthenium Complexes. J. Polym. Sci. A 2005, 43, 4838–4848. [Google Scholar] [CrossRef]
  31. Efkianakis, E.K.; Tzanetos, N.P.; Chochos, C.L.; Andreopoulou, A.K.; Kallitsis, J.K. End-Functionalization of Semiconducting Species with Dendronized Terpyridine–Ru(II)–Terpyridine Complexes. J. Polym. Sci. A Polym. Chem. 2009, 47, 1939–1952. [Google Scholar] [CrossRef]
  32. Lamba, J.J.S.; Fraser, C.L. Synthesis of Metal-Centered Star-Shaped Polyoxazolines Using Fe(II) and Ru(II) Tris-bipyridine Derivatives as Multifunctional Initiators. J. Am. Chem. Soc. 1997, 119, 1801–1802. [Google Scholar] [CrossRef]
  33. McAlvin, J.E.; Fraser, C.L. Polymerization of 2-Ethyl-2-oxazoline Using Di-, Tetra-, and Hexafunctional Ruthenium Tris(bipyridine) Metalloinitiators. Macromolecules 1999, 32, 6925–6932. [Google Scholar] [CrossRef]
  34. Fraser, C.L.; Smith, A.P. Metal Complexes with Polymeric Ligands: Chelation and Metalloinitiation Approaches to Metal Tris(bipyridine)-Containing Materials. J. Polym. Sci. A 2000, 38, 4704–4716. [Google Scholar] [CrossRef]
  35. Johnson, R.M.; Corbin, P.S.; Ng, C.; Fraser, C.L. Poly(methyl methacrylates) with Ruthenium Tris(bipyridine) Cores via NiBr2(PR3)2-Catalyzed Atom Transfer Radical Polymerization (ATRP). Macromolecules 2000, 33, 7404–7412. [Google Scholar] [CrossRef]
  36. Wu, X.; Fraser, C.L. Architectural Diversity via Metal Template-Assisted Polymer Synthesis:  A Macroligand Chelation Approach to Linear and Star-Shaped Polymeric Ruthenium Tris(bipyridine) Complexes. Macromolecules 2000, 33, 4053–4060. [Google Scholar] [CrossRef]
  37. Wu, X.; Collins, J.E.; McAlvin, J.E.; Cutts, R.W.; Fraser, C.L. Ruthenium Tris(bipyridine)-Centered Linear and Star-Shaped Polystyrenes:  Making Atom Transfer Radical Polymerization and Metal Complex Initiators Compatible. Macromolecules 2001, 34, 2812–2821. [Google Scholar] [CrossRef]
  38. Johnson, R.M.; Fraser, C.L. Iron Tris(bipyridine)-Centered Star Block Copolymers:  Chelation of Triblock Macroligands Generated by ROP and ATRP. Macromolecules 2004, 37, 2718–2727. [Google Scholar] [CrossRef]
  39. Johnson, R.M.; Fraser, C.L. Metalloinitiation Routes to Biocompatible Poly(lactic acid) and Poly(acrylic acid) Stars with Luminescent Ruthenium Tris(bipyridine) Cores. Biomacromolecules 2004, 5, 580–588. [Google Scholar] [CrossRef]
  40. Hoogenboom, R.; Schubert, U.S. The Use of (Metallo-)Supramolecular Initiators for Living/Controlled Polymerization Techniques. Chem. Soc. Rev. 2006, 35, 622–629. [Google Scholar] [CrossRef]
  41. Collins, J.E.; Fraser, C.L. Transition Metals as Templates for Multifunctional Initiators:  Bulk Atom Transfer Radical Polymerization of Styrene Using Di-, Tetra- and Hexafunctional Ruthenium Tris(bipyridine) Reagents. Macromolecules 1998, 31, 6715–6717. [Google Scholar] [CrossRef]
  42. Viau, L.; Even, M.; Maury, O.; Haddleton, D.M.; Le Bozec, H. New Star-Shaped Metallo-Polymeric Chromophores. Macromol. Rapid Commun. 2003, 24, 630–635. [Google Scholar] [CrossRef]
  43. Bahrami, K.; Khodaei, M.M.; Meibodi, F.S. Suzuki and Heck Cross-Coupling Reactions Using Ferromagnetic Nanoparticle-Supported Palladium Complex as an Efficient and Recyclable Heterogeneous Nanocatalyst in Sodium Dodecylsulfate Micelles. Appl. Organomet. Chem. 2017, 31, e3627. [Google Scholar] [CrossRef]
  44. Kong, L.; Han, X.; Jiao, P. Catalytic Asymmetric Diels–Alder Reactions Involving Aryl Vinyl Ketones. Chem. Commun. 2014, 50, 14113–14116. [Google Scholar] [CrossRef]
  45. Scott, J.E.; Kyffin, T.W. Demineralization in Organic Solvents by Alkylammonium Salts of Ethylenediaminetetra-Acetic Acid. Biochem. J. 1978, 169, 697–701. [Google Scholar] [CrossRef]
  46. Campo, B.J.; Bevk, D.; Kesters, J.; Gilot, J.; Bolink, H.J.; Zhao, J.; Bolsée, J.-C.; Oosterbaan, W.D.; Bertho, S.; D’Haen, J.; et al. Ester-Functionalized Poly(3-alkylthiophene) Copolymers: Synthesis, Physicochemical Characterization and Performance in Bulk Heterojunction Organic Solar Cells. Org. Electron. 2013, 14, 523–534. [Google Scholar] [CrossRef]
  47. Wang, S.-Y.; Fu, J.-H.; Liang, Y.-P.; He, Y.-J.; Chen, Y.-S.; Chan, Y.-T. Metallo-Supramolecular Self-Assembly of a Multicomponent Ditrigon Based on Complementary Terpyridine Ligand Pairing. J. Am. Chem. Soc. 2016, 138, 3651–3654. [Google Scholar] [CrossRef]
  48. Tu, T.-H.; Sakurai, T.; Seki, S.; Ishida, Y.; Chan, Y.-T. Towards Macroscopically Anisotropic Functionality: Oriented Metallo-supramolecular Polymeric Materials Induced by Magnetic Fields. Angew. Chem. Int. Ed. 2020. [Google Scholar] [CrossRef]
  49. Meier, M.A.R.; Lohmeijer, B.G.G.; Schubert, U.S. Characterization of Defined Metal-Containing Supramolecular Block Copolymers. Macromol. Rapid Commun. 2003, 24, 852–857. [Google Scholar] [CrossRef]
  50. Francis, R.; Lepoittevin, B.; Taton, D.; Gnanou, Y. Toward an Easy Access to Asymmetric Stars and Miktoarm Stars by Atom Transfer Radical Polymerization. Macromolecules 2002, 35, 9001–9008. [Google Scholar] [CrossRef]
  51. Laird, D.W.; Loewe, R.S.; Ewbank, P.C.; Liu, J.; Zhai, L.; McCullough, R. Mechanistic Aspects of Regioregularity in Head-to-Tail Coupled Poly(3-alkylthiophes). Polymer Prepr. 2001, 42, 556–557. [Google Scholar]
  52. Tkachov, R.; Senkovskyy, V.; Komber, H.; Kiriy, A. Influence of Alkyl Substitution Pattern on Reactivity of Thiophene-Based Monomers in Kumada Catalyst-Transfer Polycondensation. Macromolecules 2011, 44, 2006–2015. [Google Scholar] [CrossRef]
  53. Huang, K.-H.; Tu, T.-H.; Wang, S.-C.; Chan, Y.-T.; Hsu, C.-C. Micelles Protect Intact Metallo-Supramolecular Block Copolymer Complexes from Solution to Gas Phase during Electrospray Ionization. Anal. Chem. 2018, 90, 7691–7699. [Google Scholar] [CrossRef] [PubMed]
  54. Bakkour, Y.; Darcos, V.; Li, S.; Coudane, J. Diffusion Ordered Spectroscopy (DOSY) as a Powerful Tool for Amphiphilic Block Copolymer Characterization and for Critical Micelle Concentration (CMC) Determination. Polym. Chem. 2012, 3, 2006–2010. [Google Scholar] [CrossRef]
  55. Yu, X.; Xiao, K.; Chen, J.; Lavrik, N.V.; Hong, K.; Sumpter, B.G.; Geohegan, D.B. High-Performance Field-Effect Transistors Based on Polystyrene-b-Poly(3-hexylthiophene) Diblock Copolymers. ACS Nano 2011, 5, 3559–3567. [Google Scholar] [CrossRef]
  56. Kato, H.; Saito, T.; Nabeshima, M.; Shimada, K.; Kinugasa, S. Assessment of Diffusion Coefficients of General Solvents by PFG-NMR: Investigation of the Sources Error. J. Magn. Reson. 2006, 180, 266–273. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of metallo-initiator 4 and PSn. Reagents and conditions: (a) K2CO3, dimethylformamide (DMF), 2-chloroethanol, 50 °C; (b) NEt3, α-bromoisobutyryl bromide, dichloromethane (DCM), 0 °C; (c) (1) FeCl2, MeOH/CHCl3, r.t., (2) NH4PF6; (d) styrene, CuBr, pentamethyldiethylenetriamine (PMDETA), MeCN, 110 °C; (e) tetrakis(triethylammonium) ethylenediaminetetraacetate (TEA-EDTA), DMF, r.t.
Scheme 1. Synthesis of metallo-initiator 4 and PSn. Reagents and conditions: (a) K2CO3, dimethylformamide (DMF), 2-chloroethanol, 50 °C; (b) NEt3, α-bromoisobutyryl bromide, dichloromethane (DCM), 0 °C; (c) (1) FeCl2, MeOH/CHCl3, r.t., (2) NH4PF6; (d) styrene, CuBr, pentamethyldiethylenetriamine (PMDETA), MeCN, 110 °C; (e) tetrakis(triethylammonium) ethylenediaminetetraacetate (TEA-EDTA), DMF, r.t.
Polymers 12 02842 sch001
Figure 1. (a) Semilogarithmic kinetic plot, (b) Mn and Mw/Mn versus conversion, and (c) gel permeation chromatography (GPC) traces for the atom transfer radical polymerization (ATRP) of styrene ([4]:[styrene]:[CuBr]:[PMDETA] = 1:900:3:3, styrene/MeCN = 3/1 (v/v), 110 °C).
Figure 1. (a) Semilogarithmic kinetic plot, (b) Mn and Mw/Mn versus conversion, and (c) gel permeation chromatography (GPC) traces for the atom transfer radical polymerization (ATRP) of styrene ([4]:[styrene]:[CuBr]:[PMDETA] = 1:900:3:3, styrene/MeCN = 3/1 (v/v), 110 °C).
Polymers 12 02842 g001
Figure 2. (a) Synthesis of P3HT54. Reagents and conditions: (i) (1) t-BuMgCl, LiCl, tetrahydrofuran (THF), (2) Ni(dppp)Cl2, THF, 0 °C; (ii) 6,6″-dianthracenyl-4′-(4-boronophenyl)tpy, Pd(PPh3)4, NaOH, THF/H2O (3/1, v/v), reflux. (b) 1H NMR spectra of ligand L1 and P3HT54.
Figure 2. (a) Synthesis of P3HT54. Reagents and conditions: (i) (1) t-BuMgCl, LiCl, tetrahydrofuran (THF), (2) Ni(dppp)Cl2, THF, 0 °C; (ii) 6,6″-dianthracenyl-4′-(4-boronophenyl)tpy, Pd(PPh3)4, NaOH, THF/H2O (3/1, v/v), reflux. (b) 1H NMR spectra of ligand L1 and P3HT54.
Polymers 12 02842 g002
Scheme 2. Construction of metallo-supramolecular diblock copolymers [P3HT54–Zn–PSn].
Scheme 2. Construction of metallo-supramolecular diblock copolymers [P3HT54–Zn–PSn].
Polymers 12 02842 sch002
Figure 3. (a) 1H NMR spectra of [L2–Zn–L3] and [P3HT54–Zn–PSn] (n = 19, 33, 85, 106, 161, and 235). The upper left figure shows the X-ray crystal structure of [L2–Zn–L3] (gray, carbon; blue, nitrogen; red, oxygen; yellow, zinc). Hydrogen atoms and OTf ions are omitted for clarity. (b) Stacked diffusion ordered spectroscopy (DOSY) spectra of [P3HT54–Zn–PSn]. The inset table shows the corresponding diffusion coefficients determined with respect to the D of CHCl3 (2.33 × 10−9 m2 s−1) at 298 K [56].
Figure 3. (a) 1H NMR spectra of [L2–Zn–L3] and [P3HT54–Zn–PSn] (n = 19, 33, 85, 106, 161, and 235). The upper left figure shows the X-ray crystal structure of [L2–Zn–L3] (gray, carbon; blue, nitrogen; red, oxygen; yellow, zinc). Hydrogen atoms and OTf ions are omitted for clarity. (b) Stacked diffusion ordered spectroscopy (DOSY) spectra of [P3HT54–Zn–PSn]. The inset table shows the corresponding diffusion coefficients determined with respect to the D of CHCl3 (2.33 × 10−9 m2 s−1) at 298 K [56].
Polymers 12 02842 g003
Table 1. Results and optimized conditions for ATRP of styrene using 4.
Table 1. Results and optimized conditions for ATRP of styrene using 4.
Entry[4]:[styrene]:[CuBr]:[PMDETA] aConversion (%)Mn,NMRbDP bMn,GPCcMw/Mnc
PS191:100:3:33127001919001.14
PS331:200:3:34840003336001.13
PS851:200:3:35590008579001.21
PS1061:400:3:34611,10010611,9001.14
PS1611:800:3:33316,70016116,2001.17
PS2351:800:3:36424,50023522,0001.26
a Styrene/MeCN = 3/1 (v/v) for PS19, PS33, PS85, PS106, and PS161. PS235 was obtained from the bulk polymerization. b Mn,NMR and degree of polymerization (DP) were calculated by 1H NMR. c Mn,GPC and Mw/Mn were determined by GPC.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tu, T.-H.; Chan, Y.-T. Synthesis of Terpyridine End-Modified Polystyrenes through ATRP for Facile Construction of Metallo-Supramolecular P3HT-b-PS Diblock Copolymers. Polymers 2020, 12, 2842. https://doi.org/10.3390/polym12122842

AMA Style

Tu T-H, Chan Y-T. Synthesis of Terpyridine End-Modified Polystyrenes through ATRP for Facile Construction of Metallo-Supramolecular P3HT-b-PS Diblock Copolymers. Polymers. 2020; 12(12):2842. https://doi.org/10.3390/polym12122842

Chicago/Turabian Style

Tu, Tsung-Han, and Yi-Tsu Chan. 2020. "Synthesis of Terpyridine End-Modified Polystyrenes through ATRP for Facile Construction of Metallo-Supramolecular P3HT-b-PS Diblock Copolymers" Polymers 12, no. 12: 2842. https://doi.org/10.3390/polym12122842

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop