Open Access Article
Mohamed Gamal
Mohamed
*ab,
Ahmed. F. M.
EL-Mahdy
a,
Mohammed G.
Kotp
a and
Shiao-Wei
Kuo
*ac
aDepartment of Materials and Optoelectronic Science, Center for Functional Polymers and Supramoleuclar Materials and Center of Crystal Research, National Sun Yat-Sen University, Kaohsiung 804, Taiwan. E-mail: mgamal.eldin34@gmail.com; kuosw@faculty.nsysu.edu.tw
bChemistry Department, Faculty of Science, Assiut University, Assiut, 71516, Egypt
cDepartment of Medicinal and Applied Chemistry, Kaohsiung Medical University, Kaohsiung 807, Taiwan
First published on 23rd November 2021
Porous organic polymers (POPs) are organic macromolecules that are considered emerging materials because of their high specific surface areas, tunable porosities, low densities, high chemical and thermal stabilities, variable compositions, convenient post-functionalization, extended π-conjugations, and their high contents of carbon, nitrogen, oxygen, and other non-metallic atoms. POPs have been classified into four types: covalent triazine frameworks (CTFs), hypercrosslinked polymers (HCPs), covalent organic frameworks (COFs), and conjugated microporous polymers (CMPs). These materials have potential applications in, for example, gas capture/separation, energy storage, H2 production from water, photocatalysis, chemical sensing, perovskite solar cells, water treatment, optical devices, and biomedicine. In this review, we provide an overview of recent reports describing the preparation and various applications of POPs.
![]() | ||
| Fig. 2 Schematic cartoon for incorporating metal species into HCP frameworks via two different knitting models. Reproduced from ref. 65 with permission from Wiley-VCH. | ||
![]() | ||
| Fig. 3 Formation of CTFs via direct or indirect approaches. Reproduced from ref. 66 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 4 (a) and (b) Synthesis of 2D-CTFs through polymerization at CH2Cl2/CF3SO3H interfaces, (b) and (d) represent the atomic structure and DFT band structure of the 2D-CTF. Reproduced from ref. 78 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 5 Topology diagrams representing a general basis for COF design and the construction of (A) 2D COFs and (B) 3D COFs.93 Reproduced from ref. 93 with permission from Elsevier. | ||
![]() | ||
| Fig. 6 Various linkages for COF formation. Reproduced from ref. 93 with permission from Elsevier. | ||
![]() | ||
| Fig. 7 Various topologies in 2D COFs. Reproduced from ref. 94 with permission from American Chemical Society. | ||
C and NH groups through a Scholl reaction of 3,5-diphenyl-1H-pyrazole in the presence of AlCl3 in CHCl3 as the solvent. They prepared the hybrid material MOP-PZ-Ag by embedding silver nanoparticles (NPs) into MOP-PZ, and found that MOP-PZ-Ag functioned as a heterogeneous catalyst for the production of propiolic acid through the carboxylation of some terminal alkynes; this stable hybrid material also displayed an excellent degree of CO2 adsorption (183.7 mg g−1).98 Wang and co-workers prepared nitrogen-enriched microporous polymers containing various contents of amino groups through condensation reactions of melamine with formohydrazide, formamide, N,N-dimethylformamide (DMF), and N-methylformamide. These four polymers possessed pore diameters of 0.52–1.10 nm and specific surface areas of 612–748 m2 g−1, with the BET method, and SEM, and TEM analysis revealing that they were agglomerates of tiny particles with an irregular shape and porous structures; the four polymers provided CO2 uptakes of up to 103 mg g−1 at 273 K.99 Liang et al. prepared four kinds of CMPs and used a post-knitting approach to improve their porosities, affording eight CMP-based HCPs (KCMPs) (Fig. 9).100 These KCMPs had high specific surface areas (up to 2267 m2 g−1), high pore volumes (up to 3.27 cm3 g−1), and excellent degrees of CO2 uptake (up to 175.12 mg g−1 at 273 K). Das et al. prepared two kinds of CTFs based on 1,2,3-triazolo units; their Tz-CTF polymeric frameworks exhibited excellent CO2 uptake capacities, with a high surface area, chemical and thermal stabilities, agglomerated tiny particles, and regular porous structures, based on SEM and TEM imaging.101 Our group also prepared two bicarbazole-based COFs (Cz-BD, and Cz-DHBD) through Schiff-base condensations of Cz-4CHO (as the C2-symmetric knot) with BD and DHBD.102 These Cz-BD and Cz-DHBD COFs possessed high surface areas (up to 2111 and 992.2 m2 g−1, respectively) as well as thermal stability up to 507 °C. The Cz-DHBD COF featured a kagome structure, due to intramolecular OH⋯N hydrogen bonding, that increased the steric bulk hinderance and decreased the nucleophilicity of the imino nitrogen atoms, thereby decreasing its ability to capture CO2 (110.59 mg g−1) relative to that of the Cz-BD COF (125.95 mg g−1). In addition, we have prepared hollow microspherical and microtubular carbazole-based COFs through condensations of Car-3NH2 and the triformyl linkers TPA-3CHO, TPP-3CHO, and TPT-3CHO with various degrees of planarity, obtaining high surface areas of 1334, 743, and 721 m2 g−1, respectively. Due to their high surface areas, these Car-TPA, Car-TPT, and Car-TPP COFs provided high degrees of CO2 capture, with recorded uptake rates of up to 61, 42, and 34 mg g−1, respectively, at 298 K. Although Car-TPT and Car-TPP had similar surface areas, the former provided a higher CO2 capture rate because of the higher nitrogen content of the triazine units in the Car-TPT COF.103 The effect of the nitrogen content on the CO2 capture ability of COFs was also observed in six 2D COFs displaying various planarities, symmetries, and nitrogen contents, synthesized through the condensation reaction of TPA-3NH2 and TPT-3NH2 (as triarylamine monomers) with the monomers TPA-3CHO, TPP-3CHO, and 2TPT-3CHO (with various degrees of planarity) (Fig. 10).104 Here, TPT-3NH2 is more planar than TPA-3NH2, and TPT-3CHO is more planar than TPP-3CHO and TPA-3CHO. The COFs with lower planarity possessed lower surface areas; those with higher nitrogen contents and higher nucleophilicities displayed higher degrees of CO2 capture. As expected, the higher nitrogen contents in the TPT-based COFs provided higher rates of CO2 capture (up to 92.38 mg g−1 at 273 K for TPT-COF-6), with the greater planarity of their building blocks also enhancing the morphology to favor CO2 capture.104 We have also prepared β-ketoenamine-linked COFs (TFP-TPA, TFP-Car, and TFP-TPP) through solvothermal Schiff-base [3+3] polycondensations of TFP-3OHCHO with three tris(aminophenyl)-presenting derivatives (possessing amino, carbazole, and pyridine units, respectively).105 The TFP-TPA, TFP-Car, and TFP-TPP COFs possessed different degrees of planarity and surface areas of 457, 362, and 686 m2 g−1, respectively. The higher degrees of planarity of the TFP-Car and TFP-TPP COFs resulted in higher thermal stabilities and stronger interactions with CO2 molecules at 273 K (up to 190 and 200 mg g−1, respectively) relative to those exhibited by the lower-planarity TPA-Car COF (183 mg g−1).105 Nagai et al. reported four types of PI-COF prepared through solvothermal reactions of TAPA and TAPB with PMDA and NTCDA (Fig. 11).106 These PI-COF materials had specific surface areas exceeding 500 m2 g−1 and were applied for CO2 uptake. We have used Schiff base formation, reduction, and Mannich and Sonogashira–Hagihara couplings to prepare two new CMPs (TPE-TPE-BZ, and Py-TPE-BZ) featuring benzoxazine-linked tetraphenylethylene and pyrene units (Fig. 12)107 Upon thermal treatment, the benzoxazine units in the backbones of the TPE-TPE-BZ and Py-TPE-BZ CMPs underwent ring opening polymerizations to form new materials displaying high performance for CO2 uptake, due to the presence of phenolic OH and Mannich bridges capable of hydrogen bonding with CO2 molecules.
![]() | ||
| Fig. 8 Synthesis of (a) PN-Nap-2 and HCP-PN-2, (b) and (c) CO2 adsorption of PN-Nap-2 and HCP-PN-2 at 0 and 25 °C. Reproduced from ref. 97 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 9 Schematic cartoon for the synthesis of CMPs via the post-knitting method. (b) The cross-linker structure. (c) The synthetic routes for the KCMP via palladium-catalyzed Suzuki coupling and a Lewis acid catalyzed Friedel–Crafts reaction. Reproduced from ref. 100 with permission from the Royal Society of Chemistry. | ||
![]() | ||
| Fig. 10 (a and b) Synthesis of TPA-COF-1, TPA-COF-2, TPA-COF-3, TPT-COF-1, TPT-COF-2, and TPT-COF-3 through Schiff base reactions. (c and d) Their color photos. Reproduced from ref. 104 with permission from the Royal Society of Chemistry. | ||
![]() | ||
| Fig. 11 Synthesis of the four polyimide COFs. Reproduced from ref. 106 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 12 Synthesis of the TPE-TPE-BZ CMP, and TPE-TPE-BZ CMPs through a multistep reaction. Reproduced from ref. 107 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 13 Synthesis routes for BCB-CMP, Py-CMP, CCBCB-CMP and CCPy-CMP. Reproduced from ref. 108 with permission from Elsevier. | ||
C units do not display any reactivity in acidic media, and increasing the number of triphenylamine groups enhanced the specific capacitance; as a result, the capacitance of TPA-TPA-COF-1, which featured six triphenylamine groups, was higher than those of the other COFs 2–4. Morphologies and chemical structures can also play a role in affecting the specific capacitances of supercapacitors. We confirmed these phenomena in a recent study of the previously mentioned Car-TPA, Car-TPP, and Car-TPT COFs, which possessed carbazole, pyridine, and triphenylamine units, respectively, making them ideal candidates for storing energy; at 0.2 A g−1, they displayed specific capacitances of 13.6, 14.5, and 17.4 F g−1, respectively.103 The Car-TPA and Car-TPP COFs had high surface areas and two redox groups, but provided lower capacitances; in contrast, the Car-TPT COF had a low surface area and only one redox group, but its microtubular structure and molecular design led to enhanced capacitance and retention during the charging/discharging process. Recently, Maha et al. used Sonogashira–Hagihara reactions to prepare a series of CMPs based on tetrabenzonaphthalene (TBN) moieties (TBN-TPE-CMP, TBN-Py-CMP, and TBN-Car-CMP) [Fig. 14(a)].121 Subsequently, we improved the conductivity of these materials by mixing them with single-walled carbon nanotubes (SWCNTs) [Fig. 14(b)]. Interestingly, the capacitance of the TBN-Py-CMP/SWCNT nanocomposite (430 F g−1 at 0.5 A g−1) was higher than those of the other samples, presumably because of strong π–π interactions between the pyrene units in the TBN-Py-CMP framework and the SWCNTs.122–125 Eddaoudi et al. prepared Hex-Aza-COF-2 and Hex-Aza-COF-3 through solvothermal condensations of benzoquinone with redox-functionalized aromatic tetramines and phenazine, respectively [Fig. 15(a)].126 They used powder X-ray diffraction (PXRD), solid state NMR spectroscopy, BET analysis, SEM, and TEM imaging to investigate the chemical structures, crystallinities, porosities, and morphologies of these materials [Fig. 15(b)–(i)]. The PXRD data and SEM images revealed that both materials had moderate crystallinities and aggregated spherical morphologies. Furthermore, based on electrochemical measurements, the specific capacitances of Hex-Aza-COF-2 (585 F g−1) and Hex-Aza-COF-3 (663 F g−1) were higher than those reported previously for porous polymeric and COF materials. Yang et al. prepared a new porous organic HCP through the reaction of pyrene with CHCl3 as the crosslinker and AlCl3 as the catalyst, with the subsequent hydrolysis affording HcPPy that possessed a high surface area (723 m2 g−1) and micropores (1 nm) and mesopores (2–3 and 3.5–4.5 nm).127 This material functioned as an anode in Li batteries and provided excellent current densities (up to 683 MA h g−1 at 1000 mA g−1). Wong et al. prepared materials named SP and HP through Sonogashira couplings of tri(4-ethynylphenylamine) with 9-ferrocenylidene-2,7-diiodo-9H-fluorene in the presence and absence of ZIF-67, respectively.128 They then prepared Fe-embedded magnetic carbon materials (SP and HP-MCMs) through the carbonization of SP and HP at 500 °C for 1 h. The obtained spherical SP-MCP had a microporous architecture, good redox performance, and good cycling stability when applied in lithium-ion batteries (LIBs) as an anode; its performance was superior to that of the HP-MCP. Zhao et al. prepared the 2D-COF-ETTA-ETTCA and COF-ETTA-ETTCA-S materials by loading 88.4 wt% of sulfur and applied these materials in lithium–sulfur batteries.129 The devices displayed high capacity (up to 1617 mA h g−1 at 0.1C), low capacitance decay after 528 cycles, and high coulombic performance (ca. 98%).
![]() | ||
| Fig. 14 (a) Synthesis of TBN-Py-CMP, TBN-TPE-CMP, and TBN-Car-CMP and (b) formation of TBN-CMP/SWCNTs. Reproduced from ref. 121 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 15 (a) Synthesis of Hex-Aza-2 and Hex-Aza-COF-3, (b) and (c) PXRD pattern for Hex-Aza-COF-2 and Hex-Aza-COF-3. (d) Solid-state 13C NMR for Hex-Aza-COF-2 and Hex- Aza-COF-3. (e) BET isotherms for Hex-Aza-COF-2 and Hex-Aza-COF-3. (f and g) SEM images for Hex-Aza-COF-2 and Hex-Aza-COF-3. (h and i) TEM images for Hex-Aza-COF-2 and Hex-Aza-COF-3. Reproduced from ref. 126 with permission from Wiley-VCH. | ||
![]() | ||
| Fig. 16 (a) Schematic for the preparation of U@TDEn core–shell hetero frameworks. (b) SEM image of NH2-UiO-66; (c) SEM, (d) TEM image and (e) EDX mapping of U@TDE4. Reproduced from ref. 133 with permission from the Royal Society of Chemistry. | ||
![]() | ||
| Fig. 17 Synthesis of Tp(BTxTP1−x)-COFs. (b)–(d) PXRD, N2 adsorption–desorption and FT-IR profiles of Tp(BTxTP1−x)-COFs. (e) Solid-state 13C CP/MAS NMR spectrum of Tp(BT0.05TP0.95)-COF. Reproduced from ref. 134 with permission from the Royal Society of Chemistry. | ||
N units functioned as electron donating groups that facilitated ICT between the carbazole BCTB-4CH
N units and the acceptors; in contrast, the weakly electron donating BCTA-4N
CH groups resulted in weak fluorescence for the BCTA-TP COF. The BCTB-BCTA COF displayed high fluorescence sensitivity for the detection of HCl, due to the increasing planarity of its constituent units upon protonation of the imino nitrogen atoms in the BCTB-BCTA COF architecture, resulting in red-shifting of the signal. The fluorescence lifetime of the BCTB-BCTA COF increased from 4.13 to 4.89 ns upon exposure to 1 mmol L−1 HCl, in addition to displaying an LOD of 10 nmol L−1. Increasing the planarity of the units in the COFs and converting them into quinoid structures were exploited by our group for HCl sensing using 2D PyTA-BC and PyTA-BC-Ph COFs, prepared through Schiff base condensations of PyTA-4NH2 with BC-4CHO and BC-Ph-4CHO, respectively, and separately under solvothermal conditions.125 The PyTA-BC and PyTA-BC-Ph COFs had thermal stabilities of up to 403 and 421 °C, respectively, and high surface areas (520 and 1445 m2 g−1, respectively); we attribute the superior thermal stability and surface area of the PyTA-BC-Ph COFs to their longer building blocks. Both the PyTA-BC and PyTA-BC-Ph COFs exhibited solvatochromism phenomena and red-shifting of their signals occurred, due to strong hydrogen bonding between the amino groups on the COF surfaces and the C
O groups of the polar solvents, which facilitated ICT. Interactions of the PyTA-BC and PyTA-BC-Ph COFs with HCl also induced red-shifting from their original yellow color; the signals reverted back after exposure to NH3 vapor. This process, which could be repeated without considerable performance loss, resulted from the increase in planarity upon protonation and the formation of quinoid structures. The PyTA-BC and PyTA-BC-Ph COFs had very low LODs for HCl of 24 and 20 nmol L−1, respectively.132 Guo et al. obtained a new 1D-COF with a specific surface area of 426 m2 g−1 and thermal stability up to 360 °C through the reaction of TFPPy and DABP; they used this COF as a chemical H+ sensor in acidic solutions.139 Zhu et al. prepared a dual-luminescent COF (DL-COF) through Schiff base formation between ETTA and 9,10-anthracenedicarboxaldehyde and used it for the chemical sensing of explosive nitro compounds with high selectivity and sensitivity.140 Mohamed et al. used Heck reactions to prepare the four ultrastable microporous polymers An-HPP, TPT-HPP, Car-HPP, and TPE-HPP from octavinylsilsesquioxane (OVS) and brominated anthracene triphenyltriazine, bicarbazole, and tetraphenylethene, respectively; we revealed that these materials displayed good thermal stabilities because of the presence of the inorganic POSS units and the high crosslinking densities.141–144 Furthermore, these four fluorescent materials could be used for the detection of Fe3+ ions.
![]() | ||
| Fig. 18 Schematic representation of Cu@TFPB-DHTH COF as a chemical sensor for Cys and L-His. Reproduced from ref. 137 with permission from the Royal Society of Chemistry. | ||
![]() | ||
| Fig. 19 Preparation of TDFB-TEB and TCT-TEB. Reproduced from ref. 145 with permission the Royal Society of Chemistry. | ||
![]() | ||
| Fig. 20 Synthesis of perfluoroalkyl-functionalized CMPs. Reproduced from ref. 148 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 21 (a) Preparation of the pcCMP (b) PXRD for pcCMP-O; (c) N2 adsorption (77 K) for pcCMP-O; (d) CO2 adsorption for pcCMP-O and pcCMP-C; (e) FESEM and (f) TEM images for pcCMPO; (g and h) solid-state 13C NMR for pcCMP-O and pcCMP-C, respectively; and (i) UV-vis absorption spectra in the solid state for pcCMP-O and pcCMP-C. Reproduced from ref. 152 with permission from American Chemical Society. | ||
![]() | ||
| Fig. 22 (a) Synthesis of BT-COFs; (b) top and (c) side view of TPB-BT-COF; (d) top and (e) side view of TAPT-BT-COF; (f) photographs of TPB-BT-COF, TAPT-BT-COF and TPB-TP-COF. Reproduced from ref. 156 with permission from the Royal Society of Chemistry. | ||
![]() | ||
| Fig. 23 Synthesis of microporous TAPP-TFPP-COF. Reproduced from ref. 158 with permission from American Chemical Society. | ||
| An | Anthracene |
| BTD | Benzothiadiazole |
| Ben | Benzene |
| BD | Benzidine |
| BT | Benzothiadiazole |
| Cz-4CHO | Bi-carbazole-4CHO |
| Car-4CN | [9,9′-Bicarbazole]-3,3′,6,6′-tetracarbonitrile |
| An-4Ph | 9,10-Bis(diphenylmethylene)-9,10-dihydroanthracene |
| BFTB-4CHO | 4,4′,4″,4‴-([9,9′-Bifluorenylidene]-3,3′,6,6′-tetrayl) tetrabenzaldehyde |
| BFTB-4NH2 | 4,4′,4″,4‴-([9,9′-Bifluorenylidene]-3,3′,6,6′-tetrayl)tetraaniline |
| BCTA-4NH2 | 4,4′,4″,4‴-([9,9′-Bicarbazole]-3,3′,6,6′-tetrayl)tetraaniline |
| BCTB-4CHO | 4,4′,4″,4‴-([9,9′-Bicarbazole]-3,3′,6,6′-tetrayl)tetrabenzaldehyde |
| BC-Ph-4CHO | 4,4′,4″,4‴-([9,9′-Bicarbazole]-3,3″,6,6″-tetrayl)tetrabenzaldehyde |
| TCNPy | 1,3,6,8-Cyanopyrene |
| CO2 | Carbon dioxide |
| OVS | Cubic octavinylsilsesquioxane |
| CTFs | Covalent triazine frameworks |
| COFs | Covalent organic frameworks |
| CMPs | Conjugated microporous polymers |
| DHBD | 3,3′-Dihydroxybenzidine |
| DAHQ-2HCl | 2,5-Diaminohydroquinone dihydrochloride |
| DHTH | 2,5-Dihydroxyterephthalohydrazide |
| BMOB | Dimethoxybenzene |
| DABP | 4,4′-Diaminobenzophenone |
| γ-CD | γ-Cyclodextrin |
| ETTA | 4,4′,4″,4‴-(Ethane-1,1,2,2-tetrayl)tetranilino |
| H2 | Hydrogen |
| H2O2 | Hydrogen peroxide |
| HCPs | Hypercrosslinked polymers |
| htb | Hexagonal tungsten bronze |
| hxl | Hexagonal layer |
| kgm | Kagome |
| IUPAC | International Union of Pure and Applied Chemistry |
| ICT | Intramolecular charge transfer |
| Li–S | Lithium–sulfur batteries |
| LiOH | Lithium hydroxide |
| MOFs | Metal–organic frameworks |
| NTCDA | 1,4,5,8-Naphthalenetetracarboxylic dianhydride |
| PS | Polystyrene |
| Py | Pyrene |
| POPs | Porous organic polymers |
| PAFs | Porous aromatic frameworks |
| PIMs | Polymers of intrinsic microporosity |
| PMDA | Pyromellitic dianhydride |
| PA | Phenylamine |
| PDA | Phenylenediamine |
| PD | p-Phenylenediamine |
| PyTA-4NH2 | 4,4′,4″,4‴-(Pyrene-1,3,6,8-tetrayl)tetraaniline |
| PyTA-4NH2 | 4,4′,4″,4‴-Pyrene-1,3,6,8-tetrayl)tetraaniline |
| SEM | Scanning electron microscope |
| TEM | Transmission electron microscope |
| TGA | Thermogravimetry analyses |
| TfOH | Trifluoromethanesulfonic acid |
| Car-3NH2 | Triamine 9-(4-aminophenyl)-carbazole-3,6-diamine |
| TPA-3CHO | Tris(4-formylphenyl)amine |
| TPP-3CHO | 2,4,6-Tris(4-formylphenyl)pyridine |
| TPT-3CHO | 2,4,6-Tris(4-formylphenyl)triazine |
| TPA-3NH2 | Tris(4-aminophenyl)amine |
| TPT-3NH2 | 2,4,6-Tris(4-aminophenyl)triazine |
| TFP-3OHCHO | 1,3,5-Triformylphloroglucinol |
| TAPA | Tris(4-aminophenyl)amine |
| TAPB | 1,3,5-Tris(4-aminophenyl)benzene |
| TPPDA(NH2)4 | Tetraphenyl-p-phenylenediamine |
| TPPyr(CHO)4 | 1,3,6,8-Tetrakis(4-formylphenyl)pyrene |
| TPTPE(CHO)4 | 1,1,2,2-Tetrakis[4-formyl-(1,1′-biphenyl)]ethane) |
| TBN | Tetrabenzonaphthalene |
| Tp | 1,3,5-Triformylphloroglucinol |
| TPE | Tetraphenylethene |
| Pyr-4Ph | Tetraphenylpyrazine |
| TNT | Trinitrotoluene |
| TFPB-3CHO | 1,3,5-Tris(4-formylphenyl)benzene |
| BC-4CHO | 3,3′,6,6″-Tetraformyl-9,9″-bicarbazole |
| TFPPy | 1,3,6,8-Tetrakis(p-formylphenyl)pyrene |
| TP | Terephthalaldehyde |
| TPT | Triphenyltriazine |
| Ben-T | 1,3,5-Tris(4-ethynylphenyl)benzene |
| TM | 2,4,6-Trimethyl-1,3,5-trizaine |
| Car-4CHO | 3,3′,6,6′-Tetraformyl-9,9′-bicarbazole |
| B(OMe)3 | Trimethyl borate |
| sql | Square lattice |
| SBUs | Secondary building units |
| SWCNTs | Single walled carbon nanotubes |
| TAPP | Zinc 5,10,15,20-tetra(4-aminophenyl)porphyrin |
| TFPP | Zinc 5,10,15,20-tetra(4-formylphenyl)porphyrin |
C and C–F Bonds: Achieving Remarkable Stability and Super Anhydrous Proton Conductivity, ACS Appl. Mater. Interfaces, 2021, 13, 15536–15541 CrossRef CAS PubMed.| This journal is © The Royal Society of Chemistry 2022 |