Catalytic Oxidation of Thiols within Cavities of Phthalocyanine Network

Oct 5, 2017 - •S Supporting Information. ABSTRACT: Two three-dimensional (3D) network polymers (1 and 2), in which zinc(II) or cobalt(II) phthalocya...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/Macromolecules

Catalytic Oxidation of Thiols within Cavities of Phthalocyanine Network Polymers Rei Tamura, Takahiro Kawata, Yoshiyuki Hattori, Nagao Kobayashi, and Mutsumi Kimura* Department of Chemistry and Materials, Faculty of Textile Science and Technology, Shinshu University, Ueda 386-8567, Japan S Supporting Information *

ABSTRACT: Two three-dimensional (3D) network polymers (1 and 2), in which zinc(II) or cobalt(II) phthalocyanines were interconnected with twisted 9,9′-spirobifluorene linkers, were synthesized in order to investigate their performance as heterogeneous catalysts for thiol oxidations. From the spectroscopic analyses of two dimers (3 and 4) as component units of the network polymers, 3 connected with a short linker revealed electronic interaction between the two phthalocyanine units. Micrometer-sized polymer particles were formed due to the condensation of the twisted 9,9′-spirobifluorene linkers in the presence of zinc or cobalt ions. The dispersed solutions of 1 and 2 had sharp Q-bands, indicating the prevention of stacking among phthalocyanine moieties within the polymers. Powdered X-ray diffraction pattern and N2 adsorption− desorption analyses suggested that 1 created small and rigid cavities as compared with 2 through the regular spatially arrangement of the phthalocyanine moieties in the 3D networks. The photocatalytic and catalytic activities of 1 and 2 for thiol oxidations using molecular oxygen were examined. We found that the catalytic activity of 1 was higher than that of 2 having larger cavities.



INTRODUCTION Phthalocyanines (Pcs) and their metal complexes (MPcs) have received special attention as molecular components for constructing functional architectures.1−3 The functionalities of MPcs strongly depend on their intermolecular arrangement of MPcs. Planar-shaped MPcs can form a one-dimensional columnar stack through their strong intermolecular π−π interaction, and the stacking of MPcs enables efficient electron or energy transport along with the columnar axis.4,5 On the other hand, the stacking of MPcs diminishes catalytic activity because of the spatial blocking of the substrate approach to the central metal of the MPcs as an active site for catalytic reactions.6 Three-dimensional (3D) porous materials have recently been made by connecting MPcs with rigid linkers.7−12 The connection of MPcs with rigid linear linkers forms two-dimensional (2D) porous sheets comprised of square phthalocyanine networks, and the 2D sheets are crystallized into layered structures through the formation of noncovalent stacks of MPcs units.7−10 In contrast, the connection of MPcs with rigid contorted linkers leads to the creation of small cavities through the prevention of MPcs stacking within polymers.11,12 The prevention of MPcs stacking may enable exposure to the central metal sites of MPcs within solids, and the creation of cavities around MPcs can also provide favorable penetration pathways of substrates from outside of solids. Microporous polymers containing MPcs have great potential for the use as heterogeneous catalysts. McKeown et al. reported the enhancement of catalytic activity of MPcs within the microporous MPc network polymers made by the interconnection of MPc units through spirocyclic linkers.12 However, there are no other reports on the catalytic activity of MPcs in the MPc © XXXX American Chemical Society

network polymers. In this study, we synthesized two twisted 9,9′spirobifluorene linkers of different linker lengths and investigated the influence of cavity size within two MPc network polymers 1 and 2 (Figure 1) on catalytic activity for thiol oxidation reactions.

Figure 1. Chemical structures of MPc network polymers 1 and 2.



RESULTS AND DISCUSSION First, we synthesized two phthalocyanine dimers (3 and 4) as component units of MPc network polymers 1 and 2 by mixing the condensation of 4-tert-butylphthalonitrile with two linker units (5 or 6) in the presence of metal salts (Zn(AcO)2 or CoCl2) Received: August 8, 2017 Revised: September 26, 2017

A

DOI: 10.1021/acs.macromol.7b01713 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

Figure 2. Synthetic procedure of phthalocyanine dimers 3 and 4 connected with twisted linkers. Reagents and conditions: (i) 4-tert-butylphthalonitrile, Zn(AcO)2 or CoCl2, N,N-dimethylaminoethanol (DMAE) and o-dichlorobenzene (o-DCB), 140 °C, 16 h.

(Figure 2). Two phthalonitriles in 5 and 6 were connected with different 9,9′-spirobifluorene linkers, and the twisting of fluorene rings prevented the intramolecular formation of a Pc ring. Figure 3 shows the ultraviolet−visible (UV−vis) absorption, fluorescence, and magnetic circular dichroism (MCD) spectra of zinc complexes Zn-3 and Zn-4 in THF; Table 1 summarizes the optical and electrochemical data of the zinc and cobalt complexes. While Zn-4 displayed a sharp Q-band at 675 nm, Zn-3 exhibited a split Q-band at 695 and 675 nm.13 The Q-bands for Zn-3 and Zn-4 followed the Lambert−Beer law, indicating that the aggregation behavior should be negligible under the experimental conditions. Furthermore, the emission maximum of Zn-3 was red-shifted by 10 nm as compared with that of Zn-4, indicating that Zn-3 had a narrower band gap than Zn-4. The MCD spectra of Zn-4 showed negative and positive signs at the Q-band, indicating that the chromophore symmetry of Zn-4 is close to D4h.14 In contrast, Zn-3 exhibited the overlapping of two Faraday B-terms at the split Q-band. This suggested splitting of the lowest unoccupied molecular orbitals’ (LUMO and

Figure 3. Absorption (a, c) and MCD (b, d) spectra of Zn-3 (a, b) and Zn-4 (c, d) in THF.

LUMO+1) energy levels in Zn-3. The splitting of the Q-band and the red-shifting of the emission maximum of Zn-3 were probably attributable to the through-space electronic interaction between two zinc phthalocyanine (ZnPc) moieties connected with the twisted spirobifluorene linker.15 The oxidation potentials of four dimers were determined by differential pulse B

DOI: 10.1021/acs.macromol.7b01713 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

LUMO and LUMO+1 energy levels of unsubstituted ZnPc were degenerate, the degeneracy of the LUMO and LUMO+2 levels for Zn-3 and Zn-4 was broken due to their asymmetrical structures. Moreover, the calculated highest occupied molecular orbital (HOMO) energy level of Zn-3 was destabilized as compared with that of ZnPc and Zn-4. Time-dependent (TD) DFT calculations reproduced the spectral differences between Zn-3 and Zn-4 very well. The Q-bands can be assigned to the HOMO → LUMO and HOMO → LUMO+1 transitions of the Pc ligand (Table 2). The lowest-lying electronic transition of

Table 1. Optical and Electrochemical Properties of Zn-3, Zn-4, Co-3, and Co-4 dimer

absorptiona/nm (log ε)

Zn-3 695 (5.19), 675 (5.25), 352 (4.96) Zn-4 675 (5.32), 348 (5.00) Co-3 678 (5.10), 666 (5.13), 335 (5.04) Co-4 666 (5.24), 336 (5.13)

fluorescencea/nm E1/2b/V vs Fc+/Fc 699

+0.30

689

+0.23 +0.22, +0.52 +0.15, +0.45

Absorption and fluorescence spectra were measured in THF solution. b DPVs were measured in THF with 0.1 M TBAP with a scan rate of 50 mV s−1. a

Table 2. Calculated Transition Energies, Oscillator Strength f, and Configurations of Zn-3 and Zn-4

voltammetry (DPV) measurements in dry CH2Cl2 containing 0.1 M n-Bu4NPF6 as a supporting electrolyte (Table 1). The dimers Zn-3 and Zn-4 exhibited one oxidation peak at +0.30 and +0.23 V vs ferrocene/ferrocenium couple (Fc/Fc+), which was due to the Pc ring-based oxidation process of ZnPc(−1)/ ZnPc(−2).16 The oxidation potential of Zn-4 was slightly negatively shifted as compared with that of Zn-3 because of the effect of electron-donating ether units. The spectroscopic and electrochemical properties of Co complexes (Co-3 and Co-4) were almost identical to those of the corresponding Zn complexes (Table 1). Density functional theory (DFT) calculations at the CAMB3LYP/6-31G* level (Gaussian 09) were carried out for Zn-3 and Zn-4 to enhance our understanding of the spectroscopic and electrochemical data.17 In accordance with the optimized geometries of Zn-3 and Zn-4, twisted angles between two ZnPc units were found to be 91°, and the distance between two central Zn atoms in Zn-4 was greater than that in Zn-3 (Figure 4a). Figure 4b shows energy-level diagrams and calculated electronic absorption spectra of Zn-3 and Zn-4. Whereas the estimated

compounds

λmax/nm

f

configurations

Zn-3

635 606 618 606

0.894 0.794 0.801 0.845

H→L (50%), H-1→L+1 (41%) H→L+2 (46%), H-1→L+3 (37%) H→L+1 (48%), H-1→L (48%) H-1→L+2 (49%), H→L+3 (48%)

Zn-4

Zn-3 was red-shifted by 17 nm compared with that of Zn-4 due to the narrowing of the HOMO−LUMO band gap through the destabilization of the HOMO level and the stabilization of the LUMO level. Whereas Zn-4 exhibited the localized molecular orbitals (MOs) in each ZnPc moiety, as shown in Figure 4b, the MOs of Zn-3 were populated in the whole molecules, suggesting the electronic interaction between the two ZnPc moieties in the twisted dimer. The two orthogonally oriented ZnPc moieties in Zn-3 are electronically communicated through the 9,9′-spirobifluorene linker. In contrast, the linkage of ether bonds between ZnPcs and 9,9′-spirobifluorene linker in Zn-4 may induce the localization of MOs within individual ZnPcs. The Q-band of ZnPc consists of two transitions (Qx and Qy) that are polarized perpendicularly to each other. The split width of Qx and Qy was wider because of the interaction between two covalently linked ZnPc moieties in Zn-3 as shown in Figure 4b. The structural and electronic differences in 3 and 4 may have influenced the catalytic activity of microporous MPc network polymers. Network polymers 1 and 2 were prepared from 5 or 6 in the presence of metal ions.11 Dark green precipitates were collected and washed with methanol and THF to remove metal ions and low-molecular-weight compounds. Both polymers formed micrometer-sized particles as observed by scanning electron microscopy (SEM) images (Figure 5a,b). Transmission electron microscopy (TEM) images of Zn-1 and Zn-2 exhibited a fine contrast difference, suggesting the presence of small space within the particles. The absorption spectra of Zn-1 and Zn-2 dispersed into N,N-dimethylformamide (DMF) by applying ultrasonic treatment displayed a sharp Q-band at 697 and 683 nm, respectively (Figure 5c and Table 3). The peak positions of Zn-1 and Zn-2 were in fair agreement with the values of the dimers Zn-3 and Zn-4 in solution. Considering from the absorption spectral shapes and the peak positions of Q-bands, ZnPc moieties within network polymers do not aggregate due to the connection with twisted linkages. The internal structures of Zn-1 and Zn-2 were investigated by the powdered X-ray diffraction (XRD) pattern and N2 adsorption/desorption isotherms. The XRD pattern of Zn-1 showed sharp reflection peaks at d = 8.0, 6.9, 4.7, 4.0, and 3.1 Å. That of the other Zn-2 exhibited broad and weak reflections in the measured 2θ range (Figure 5d). Moreover, the dimers Zn-3 and Zn-4 showed broad XRD reflections as well as broadening of

Figure 4. (a) Optimized structures and (b) calculated energy diagrams and selected molecular orbitals of Zn-3 and Zn-4 derived from DFT calculation at CAM-B3LYP/6-31G* level (peripheral tert-butyl units were omitted for calculation). C

DOI: 10.1021/acs.macromol.7b01713 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

that the ZnPc moieties in Zn-1 were arranged spatially with regular distances in the 3D networks. Longer bridges in Zn-2 caused conformational fluctuations among the ZnPc moieties within the polymer. The Brunauer−Emmett−Teller (BET) surface-area analysis of N2 adsorption−desorption isotherms at 77 K for Zn-2 gave a BET surface area of 340 m2 g−1, and Zn-2 exhibited high uptake at low relative pressure, which is typical of microporous materials (Figure 5e and Table 3). Whereas the BET surface area of Zn-2 was lower than the reported value of phthalocyanine network polymer CoPc20 linked by indan5,5′,6,6′-tetraol,11,12 the hysteresis between the N2 adsorption and desorption curves was observed at the same relative pressure range for CoPc20. This type of hysteresis is a characteristic of microporous polymer materials, in which covalent linkages of molecular units create a network of micropores with large cavities.11,12 Although Zn-1 was made from the twisted monomer, the N2 adsorption−desorption isotherm for Zn-1 showed a low amount of adsorption with BET surface area lower than 5 m2 g−1 (Figure 4e). Low N2 uptake has been previously reported for other microporous polymers.18,19 This implies a restricted adsorption of N2 molecules onto narrow micropores in the rigid polymer networks. The structural difference of the twisted linkers between Zn-1 and Zn-2 affected the pore structures within the 3D network polymers. Zinc and cobalt phthalocyanine complexes have been studied as highly efficient photocatalysts or catalysts in various oxidation reactions that use molecular oxygen.13 When ZnPcs absorbed visible light, energy-rich oxygen (singlet oxygen) can generate from the energy transfer from excited ZnPc to oxygen, and the generated singlet oxygen can attack substrates. In the case of cobalt phthalocyanines (CoPcs), oxygen and substrate can coordinate with the central cobalt ion in CoPcs to form sixcoordinated species, and the oxidized substrates produced by the intramolecular electron transfer from oxygen to the substrate.14 Insoluble heterogeneous catalysts have been developed from the hybridization of MPcs with organic and inorganic supports.13 Heterogeneous catalysts containing MPcs exhibited higher catalytic activity than low-molecular-weight MPcs and showed a high stability for reuse. Water-soluble ZnPcs show an efficient photocatalytic activity for the oxidation of 2-mercaptoethanol (RSH, R = HOCH2CH2−), and thiols can be converted to the corresponding sulfonic acids.20 We examined the photocatalytic activity of microporous Zn-1 and Zn-2 for the RSH oxidation in water by monitoring oxygen consumption at 25 °C under light irradiation (Figure 6a). Oxygen consumption was scarcely observed for the RSH aqueous solution in the presence of dimers Zn-3 and Zn-4 under light irradiation. This was due to the depression of substrate access to the catalytic centers in the aggregated ZnPcs. Moreover, oxygen concentration of the RSH solution with Zn-1 and Zn-2 did not change in the dark. Polymers Zn-1 and Zn-2 exhibited photocatalytic activity due to the RSH oxidation under light irradiation, and the initial reaction rate V0 of Zn-1 was higher than that of Zn-2 (V0: Zn-1: 8.0 × 10−5 M min−1; Zn-2: 4.0 × 10−5 M min−1).

Figure 5. (a) FE-SEM (left) and TEM (right) images of Zn-1. (b) FE-SEM (left) and TEM (right) images of Zn-2. (c) Absorption spectra of Zn-1 (solid line) and Zn-2 (dotted line) dispersed into DMF. (d) XRD patterns of Zn-1 (solid line) and Zn-2 (dotted line). (e) N2 adsorption (●, ▲) and desorption (○, △) isotherms at 77 K for Zn-1 (▲, △) and Zn-2 (●, ○).

Table 3. Optical Properties and BET Surface Areas (SBET) of Zn-1, Zn-2, Co-1, and Co-2 dimer

absorptiona/nm (fluorescencea/nm)

SBET/m2 g−1

pore volb/cm3 g−1

Zn-1 Zn-2 Co-1 Co-2

697 (703) 683 (685) 680 (−) 670 (−)

4 340 5 350

0.01 0.24 0.01 0.25

4RSH + O2 → 2RSSR + 2H 2O

a Absorption and fluorescence spectra were measured in dispersed DMF solutions. bPore volume determined by P/P0 = 0.9 exclude external surface area effects.

(1)

Cobalt(II) phthalocyanines have been employed as effective catalysts for the RSH oxidation.20,21 This oxidation obeys eq 1, in which the consumption of four RSH is accompanied by that of one oxygen. Figure 6b shows the rate curves for oxygen consumption in the RSH oxidation catalyzed by Co-1 and Co-2. While both polymers catalyzed the RSH oxidation in the dark, the V0 of Co-2

Q-bands in spin-coated thin films, suggesting the formation of aggregates in the amorphous solids (Figure S28). This suggests D

DOI: 10.1021/acs.macromol.7b01713 Macromolecules XXXX, XXX, XXX−XXX

Macromolecules

Article



EXPERIMENTAL SECTION



ASSOCIATED CONTENT

General. NMR spectra were recorded on a Bruker AVANCE 400 FT NMR spectrometer at 399.65 MHz for 1H and 13C in CDCl3 solution. Chemical shifts are reported relative to internal TMS. UV−vis spectra and fluorescence spectra were measured on a JASCO V-650 and a JASCO FP-750. MALDI-TOF mass spectra were obtained on a Bruker autoflex spectrometer with dithranol as matrix. High-resolution mass spectra with electrospray ionization were obtained on a Bruker Daltonics micrOTOF II. DPV data were recorded with an ALS 720C potentiostat, and electrochemical experiments were performed under purified nitrogen gas. A reference electrode was Ag/AgCl, corrected for junction potentials by being referenced to the ferrocenium/ferrocene (Fc+/Fc) couple. DFT and TD-DFT calculations were performed using the Coulomb-attenuating B3LYP (CAM-B3LYP) and 6-31G* basis set as implemented in the Gaussian 09 software suit.11 The nitrogen adsorption and desorption isotherms at 77 K were measured using a Micromeritics TriStar 3000 system. All chemicals were purchased from commercial supplies used without further purification. Column chromatography was performed with silica gel or activated alumina (200 mesh). Recycling preparative gel permeation chromatography was carried out by a JAI recycling preparative HPLC using CHCl3 as an eluent. Analytical thin layer chromatography was performed with commercial Merck plates coated with aluminum oxide 60 F254. Catalytic Activities. The catalytic activities for the oxidation of 2-mercaptoethanol by catalysts were measured according to the following procedure: Catalyst was dispersed by stirring in 0.9 mL of water in a thermostated reaction vessel fitted with a dioxygen microelectrode at 25 °C. A solution of 2-mercaptoethanol (0.1 mL) was added into the vessel by means of a syringe. The consumption of dioxygen was monitored in the dark or under light irradiation (LED light, HAYASHI LA-HDF158AS (600 000 lx); distance between light and vessel: 5 cm). During the measurements at 25 °C, the reaction solution was stirred with a magnetic bar (500 rpm). The initial rate of dioxygen consumption was determined from the slope of the consumption curve vs time. The exact concentration of 2-mercaptoethanol in the stock solution was determined by the reaction with I2, followed by a titration of the excess I2 with a NaS2O3 solution.

Figure 6. (a) Photo-oxidation of 2-mercaptoethanol catalyzed by Zn-1 (○) and Zn-2 (●) in 1.0 mL water at 25 °C under light irradiation. (b) Oxidation of 2-mercaptoethanol catalyzed by Co-1 (○) and Co-2 (●) in 1.0 mL of water at 25 °C. The oxygen consumptions were also monitored in the absence of catalyst (▲) or RSH (△). [Zn-1 and Co-1] = 14 μg/mL, [Zn-2 and Co-2] = 20 μg/mL, [RSH] = 1.0 × 10−2 mol L−1, [O2] = 2.2 × 10−4 mol L−1.

was 1/3 of the value of Co-1 (V0: Co-1: 1.2 × 10−4 M min−1; Co-2: 4.0 × 10−5 M min−1). The turnover frequencies for Co-1 and Co-2 (Co-1: 88 turnovers per CoPc per minute; Co-2: 29 turnovers per CoPc per minute) were lower than reported values of water-soluble CoPcs.20,21 After recovering the polymers from the solution by filtration, the recovered polymers dispersed in new solutions exhibited similar catalytic activity (Figures S29 and S30). As seen from these results, oxygen and RSH penetrated into the porous network polymers 1 and 2, and nonaggregated CoPcs and ZnPcs in the polymers showed the photocatalytic and catalytic activity due to the RSH oxidations. Furthermore, 1 having a rigid framework and small cavities displayed higher catalytic activity than 2. In summary, we investigated the catalytic activity of two MPc network polymers 1 and 2, in which CoPc or ZnPc units were linked with the rigid twisted spirobifluorene linkers 5 or 6. From the spectroscopic analyses of dimers 3 and 4 as component units of MPc network polymers, 3 connected with the short linker demonstrated the electronic interaction between the two MPc units. The micrometer-sized polymer particles were prepared from 5 or 6 in the presence of Zn or Co ions. The sharp Q-bands for the dispersed solutions of 1 and 2 indicated the prevention of MPcs stacking within polymers. The XRD and N2 adsorption− desorption analyses suggested the creation of smaller cavities in 1 relative to 2 by the regular spatially arrangement of MPcs in the 3D networks. The photocatalytic and catalytic RSH oxidations were examined for network polymers 1 and 2 containing ZnPcs or CoPcs. Both polymers displayed catalytic activity for RSH oxidations, indicating that the networking of MPcs by the rigid twisted linkers could create cavities, enabling substrates to gain access and reactions to occur. The catalytic activity of 1, which had rigid and small cavities, was higher than that of 2 having larger cavities. The higher catalytic activity of 1 suggests that the substrate diffusion in 1 is not restrictive compared with 2. We are now continuing the construction of MPc network polymers with different rigid linkers to enhance the catalytic activity by using the cooperative effect among MPc units as well as the detail analyses of micropore structures of MPc network polymers.

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.macromol.7b01713. Synthetic details of 1−6 (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (M.K.). ORCID

Mutsumi Kimura: 0000-0003-3050-8254 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work has been partially supported by Grants-in-Aid for Scientific Research (A) (No. 15H02172) from the Japan Society for the Promotion of Science (JSPS) of Japan.



REFERENCES

(1) Phthalocyanines: Properties and Applications; Lezonoff, C. C., Lever, A. B. P., Eds.; VCH: Weinheim, Germany, 1989, 1993, 1996; Vols. 1−4. (2) McKeown, N. B. Phthalocyanine Materials: Synthesis, Structure and Function; Cambridge University Press: Cambridge, 1998. (3) de la Torre, G.; Claessens, C. G.; Torres, T. Phthalocyanines: old dyes, new materials. Putting color in nanotechnology. Chem. Commun. 2007, 2000−2015. E

DOI: 10.1021/acs.macromol.7b01713 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules (4) Hanack, M.; Lang, M. Conducting stacked metallophthalocyanines and related compounds. Adv. Mater. 1994, 6, 819−833. (5) Elemans, J. A. A. W.; Rowan, A. E.; van Hameren, R.; Nolte, R. J. M. Molecular materials by self-assembly of porphyrins, phthalocyanines, and perylenes. Adv. Mater. 2006, 18, 1251−1266. (6) Sorokin, A. B. Phthalocyanine metal complexes in catalysis. Chem. Rev. 2013, 113, 8152−8191. (7) Spitler, E. L.; Dichtel, W. R. Lewis acid-catalyzed formation of twodimensional phthalocyanine covalent organic frameworks. Nat. Chem. 2010, 2, 672−677. (8) Ding, X.; Guo, J.; feng, X.; Honsho, Y.; Guo, J.; Seki, S.; Maitarad, P.; Saeki, A.; Nagase, S.; Jiang, D. Synthesis of metallophthalocyanine covalent organic frameworks that exhibit high carrier mobility and photoconductivity. Angew. Chem. 2011, 123, 1325. (9) Colson, J. W.; Woll, A. R.; Mukherjee, A.; Levendorf, M. P.; Spitler, E. L.; Shields, V. B.; Spencer, M. G.; Park, J.; Dichtel, W. R. Oriented 2D covalent organic framework thin films on single-layer graphene. Science 2011, 332, 228−231. (10) Spitler, E. L.; Colson, J. W.; Uribe-Romo, F. J.; Woll, A. R.; Giovino, M. R.; Saldivar, A.; Dichtel, W. R. Lattice expansion of highly oriented 2D phthalocyanine covalent organic framework films. Angew. Chem., Int. Ed. 2012, 51, 2623−2627. (11) Maffei, A. V.; Budd, P. M.; McKeown, N. B. Adsorption studies of a microporous phthalocyanine network polymer. Langmuir 2006, 22, 4225−4229. (12) Mackintosh, H. J.; Budd, P. M.; McKeown, N. B. Catalysis by microporous phthalocyanine and porphyrin network polymers. J. Mater. Chem. 2008, 18, 573−578. (13) Mack, J.; Kobayashi, N. Low symmetry phthalocyanines and their analogues. Chem. Rev. 2011, 111, 281−321. (14) Mack, J.; Stillman, M. J.; Kobayashi, N. Applications of MCD spectroscopy to porphrinoids. Coord. Chem. Rev. 2007, 251, 429−453. (15) Asano, Y.; Muranaka, A.; Fukasawa, A.; Hatano, T.; Uchiyama, M.; Kobayashi, N. anti-[2,2](1,4)Phthalocyaninophane: Spectroscopic evidence for transannular interaction in the excited states. J. Am. Chem. Soc. 2007, 129, 4516−4517. (16) Lever, A. B. P.; Milaeva, E. R.; Speier, G. Phthalocyanines: Properties and Applications; Lezonoff, C. C., Lever, A. B. P., Eds.; VCH: Weinheim, Germany, 1993; Vol. 3, pp 1−70. (17) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J.; J., A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pommelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, J. J. Gaussian 09, Revision A.1 ed.; Gaussian, Inc.: Wallingford, CT, 2009. (18) Weber, J.; Su, Q.; Antonietti, M.; Thomas, A. Exploring polymers of intrinsic microporosity − Microporous, soluble polyamide and polyimide. Macromol. Rapid Commun. 2007, 28, 1871−1876. (19) Jimenez-Solomon, M. F.; Song, Q.; Munoz-Ibanez, M.; Livingston, A. G. Polymer nanofilms with enhanced microporosity by interfacial polymerization. Nat. Mater. 2016, 15, 760−767. (20) Wöhrle, D.; Suvorova, O.; Gerdes, R.; Bartels, O.; Lapok, L.; Baziakina, N.; Makarov, S.; Slodek, A. Efficient oxidations and photooxidations with molecular oxygen using metal phthalocyanines as catalysts and photocatalysts. J. Porphyrins Phthalocyanines 2004, 8, 1020−1041. (21) Shirai, H.; Tsuiki, H.; Masuda, E.; Koyama, T.; Hanabusa, H.; Kobayashi, N. Functional Metallomacrocycles and Their Polymers. 25. Kinetic and Mechanism of the Biomimetic Oxidation of Thiol by

Oxygen Catalyzed by Homogeneous (Polycarboxyphthalocyaninato) metals. J. Phys. Chem. 1991, 95, 417−423.

F

DOI: 10.1021/acs.macromol.7b01713 Macromolecules XXXX, XXX, XXX−XXX