Porous, Hyper-cross-linked, Three-Dimensional Polymer as Stable

Nov 22, 2016 - Considerable intake of Li+ ions giving rise to very high specific capacity of 1100 mA h g–1 at a discharge current of 50 mA g–1 and...
1 downloads 4 Views 4MB Size
Forum Article www.acsami.org

Porous, Hyper-cross-linked, Three-Dimensional Polymer as Stable, High Rate Capability Electrode for Lithium-Ion Battery Debdyuti Mukherjee,†,§ Guruprasada Gowda Y. K,†,§ Harish Makri Nimbegondi Kotresh,*,‡ and S. Sampath*,† †

Department of Inorganic and Physical Chemistry, Indian Institute of Science, Bangalore 560012, India Department of Chemistry, Acharya Institute of Technology, Soldevanahalli, Bangalore 560107, India



S Supporting Information *

ABSTRACT: Organic materials containing active carbonyl groups have attracted considerable attention as electrodes in Li-ion batteries due to their reversible redox activity, ability to retain capacity, and, in addition, their ecofriendly nature. Introduction of porosity will help accommodate as well as store small ions and molecules reversibly. In the present work, we introduce a mesoporous triptycene-related, rigid network polymer with high specific surface area as an electrode material for rechargeable Liion battery. The designed polymer with a three-dimensional (3D), rigid porous network allows free movement of ions/ electrolyte as well as helps in interacting with the active anhydride moieties (containing two carbonyl groups). Considerable intake of Li+ ions giving rise to very high specific capacity of 1100 mA h g−1 at a discharge current of 50 mA g−1 and ∼120 mA h g−1 at a high discharge current of 3 A g−1 are observed with excellent cyclability up to 1000 cycles. This remarkable rate capability, which is one of the highest among the reported organic porous polymers to date, makes the triptycene-related rigid 3D network a very good choice for Li-ion batteries and opens up a new method to design polymer-based electrode materials for metal-ion battery technology. KEYWORDS: dihydroanthracene succinic anhydride (DASA), hyper-cross-linking, porous 3D rigid polymer, Li-ion battery, organic electrode material



INTRODUCTION The current energy demand is mostly met by the limited, nonrenewable energy resources based on fossil fuels and nuclear energy thus causing various environmental concerns. Storage and utilization of clean and sustainable energy via ecofriendly routes becomes highly important in the present day context.1−3 In this direction, rechargeable Li-ion battery is undoubtedly the state-of-the-art technology in terms of both scientific and commercial importance for applications in portable electronics and possibly mobile vehicular applications.4−6 The achievable capacities of inorganic active materials are of the order of ∼170 mA h g−1 for lithium transition metal oxides and ∼370 mA h g−1 for graphite anode.7−10 On the other hand, organic electrode materials can address concerns related to safety, environmental friendliness, and possibly high specific capacity combined with rate capability. Organic carbonyl compounds have recently attracted attention as © 2016 American Chemical Society

electrode materials owing to their comparatively high theoretical gravimetric capacities, ease of synthesis under ambient conditions, reversible activity of carbonyl groups toward Li ions, and resource renewability. 1,11−13 The mechanism of lithiation/delithiation in organic compounds involves reversible redox reaction of respective organic functional groups such as keto, carboxylate, and anhydride.14−16 However, organic materials having low molecular weights usually exhibit relatively poor battery performance in terms of cyclability, and solubility in battery solvents is another concern.1,3 Various strategies involving anchoring new funcSpecial Issue: Focus on India Received: August 1, 2016 Accepted: November 10, 2016 Published: November 22, 2016 19446

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces tional groups, salt formation, polymerization, and use of solid electrolyte, etc., have been proposed to overcome some of the issues mentioned above.17−23 Use of polymers is a very good solution to address the solubility issue thereby improving the performance.1 Highly porous polymeric materials are found to be efficient owing to easy and faster movement of electrolyte species as well as insertion−deinsertion of ions (Li+) through the pores.24,25 Various porous carbon materials have been reported as electrodes for Li batteries.26,27 Metal−organic framework (MOF)-based electrode materials possess required porosity but they suffer from poor stability and processability. Recently, there have been reports based on porous organic polymers such as hexaazatrinaphthalene CMP (HATN-CMP), polydicarbazyl-benzothiadiazole (PDCzBT), polytriphenylamine (YPTPA), polybenzimidazole (TpDAB), and polyimides (PIs), etc., for storage applications.28−34 Herein, we introduce an anhydride-based, porous, rigid polymer as an efficient electrode material with excellent rate capability and stability. The triptycene-type anhydride monomer, dihydroanthracene succinic anhydride (DASA), is polymerized using an external cross-linker by knitting polymerization method.35 Gouda et al.36 have reviewed the chemistry of dibenzobarallene (aka DASA)-based compounds, particularly their synthetic routes. Triptycene-type molecules possess a 3D rigid framework and are known for their associated internal free volume. Network-based, 3D polymeric materials37 for applications in the area of membranes and separation of species38−46 have been reported. Considering the ability of fast transport and accommodation of various guest ions and molecules (for example, in adsorptive desulfurization),41,45 polymer based on DASA may be expected to exhibit excellent Li-ion storage behavior. Herein, for the first time, we use rigid polymers based on DASA and show that the polymer indeed is an excellent material for rechargeable lithium batteries.



Scheme 1. Synthesis of Hyper-cross-linked Poly(dihydroanthracene succinic anhydride) (PDASA) from the Corresponding Monomer, DASA

collector. The mass of the electrode material used was ∼2−3 mg, and the fabrication of cells was carried out in a glovebox under argon atmosphere, with less than 1 ppm O2 and H2O. Characterization. Both the monomer and polymer were analyzed by various techniques. X-ray diffraction (XRD) patterns were taken using Philips (PAN analytical) instrument with Cu Kα source. UV−vis spectra were recorded in diffused reflectance mode (using integrating sphere) using PerkinElmer (Lambda 750) double beam UV−vis− near-IR spectrophotometer. Fourier transform infrared (FT-IR) spectroscopy was carried out in transmittance mode using a PerkinElmer spectrum 1 spectrophotometer. 13C NMR spectra were recorded in the solid state mode using a CP/MAS instrument (JEOL) with a 4 mm probe and 10 kHz spinnng. Scanning electron microscopy (SEM) was carried out on Carl Zeiss ultra 55 instrument using an In-lens detector along with EDS. Surface area was measured using the Brunauer−Emmett−Teller (BET) method, recorded at 77 K using a surface area analyzer (ASAP 2020, Micromeritics), and the pore size distribution was calculated by the Barrett−Joyner−Halenda (BJH) method from the desorption branches. X-ray photoelectron spectroscopic analysis was performed on a Kratos Axis Ultra DLD spectrometer with monochromatic Al Kα (1486.708 eV) radiation. Cyclic voltammograms were carried out using an electrochemical workstation (CH 660A, CH Instruments, Austin, TX, USA). Galvanostatic charge−discharge measurements were conducted using an electrochemical workstation (Arbin) at 25 °C.



EXPERIMENTAL SECTION

RESULTS AND DISCUSSION The as-synthesized polymer, PDASA, is in powder form and is brown in color. The particle size varies in the range 20−30 nm. It is found to be insoluble in EC and DMC (solvents used for battery electrolyte) and other common solvents. The FT-IR spectra (Figure 1A) of the monomer give a doublet at ∼1860 and ∼1840 cm−1 due to coupling vibration of the two CO groups in the anhydride moiety and the bands at ∼1292 and ∼1227 cm−1 are due to C−O stretching.36 Aromatic stretching vibrations are observed at low frequencies (below 1000 cm−1). In the IR spectrum of the polymer, the band observed at ∼1712 cm−1 is due to the quinonoid (CCH) region47 which is absent in the monomer. The low frequency modes appear with very small intensity. The solid state UV−visible spectrum (Figure 1B) for PDASA shows a broadened peak as compared to that of DASA, and the position also shifts to higher wavelength region due to extended quinonoid conjugation after polymerization.36,47 The polymer is further characterized by solid state 13C NMR spectroscopy (Figure 1C). The 13C NMR spectrum of the monomer exhibits a peak at δ = 170 ppm (due to carbonyl C of the anhydride moiety), and two less intense peaks in the region, 40−50 ppm, which are due to aliphatic carbons. In the case of the polymer, all the signals are observed to be broadened and, in addition, extra peaks observed, one in the aromatic region ∼ 130 ppm (shown in the Figure 1C by red dot), which is due to the quinonoid ring carbon,48 and another at ∼108 ppm due to the carbon from free CH2 linker in the polymer (where another DASA is not linked). The broadening

Materials. Anthracene, maleic anhydride, and 1,3,5-trioxane were obtained from SD Fine (Mumbai, India) and Alfa Aesar (Heysham, U.K.), respectively. Anhydrous ferric chloride and 1,2-dichloroethane were of analytical grade. Battery fabrication was carried out using the following materials: carboxymethyl cellulose (CMC), acetylene black (AB), N-methyl-2-pyrrolidone (NMP), copper (Cu)/lithium foils (Li, 99.999%), commercial LP30 electrolyte [1 M LiPF6 in ethylene carbonate (EC) and dimethylcarbonate (DMC)], tetrabutylammonium perchlorate (TBAP), and Whatman GF/D borosilicate glass fiber sheet were purchased from Aldrich. Polymer Synthesis. Dihydroanthracene succinic anhydride was synthesized from anthracene and maleic anhydride by Diels−Alder reaction (Supporting Information Scheme S1). Polymerization of DASA monomer was carried out using a procedure similar to that reported for anthracene-based polymer35,47 with a slight modification. Briefly, 1 equiv of DASA was mixed with 6 equiv of 1,3,5-trioxane and anhydrous ferric chloride in 1,2-dichloroethane solvent and heated at 80 °C for 20 h. with constant stirring (Scheme 1). A brown colored solid compound (poly(dihydroanthracene succinic anhydride), designated from hereafter as PDASA) was formed, which was separated from the unreacted product by washing thoroughly with methanol and drying under vacuum. The as-synthesized polymer was characterized by various spectroscopic techniques. Electrode Preparation. Swagelok-type cells were used to assemble batteries to follow the electrochemical performance. The configuration, Li metal/separator (whatman GF/D borosilicate glass fiber) soaked with electrolyte (1 M LiPF6 in ethylene carbonate and dimethyl carbonate in 1:1 by volume)/working electrode was used. The working electrode consisted of PDASA (40 wt %) with 40 wt % acetylene black and 20 wt % binder, CMC, coated on a Cu foil current 19447

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces

Figure 1. (A) FT-IR spectra, (B) solid state UV−vis spectra taken in the diffused reflectance mode, (C) 13C NMR spectra of DASA (blue) and PDASA (red), and (D) SEM picture of the polymer.

of NMR signal around 130 ppm in the case of the polymer could be due to the presence of a quinonoid carbon along with aromatic carbon.48 Thus, both 13C NMR and FT-IR indicate that a hyper-crosslinked polymer has been formed via a CH- linker. Based on the observations, the expected structure of the polymer is given in Scheme 2. The microporous nature of the polymer is revealed in the SEM picture given in Figure 1D. The X-ray diffraction patterns reveal that the crystallinity of the monomer is not retained after polymerization (Figure S1), which could be due to the lack of distinct planes of the aromatic rings after forming the polymer.49 The surface area and pore structure of the polymer are obtained using nitrogen adsorption−desorption isotherm, measured at 77 K. The curve depicted in Figure 2A displays an adsorption isotherm that resembles type IV behavior, indicating the mesoporous nature of the polymer50 where, at relatively high pressure of the adsorbate, capillary condensation takes place together with multilayer adsorption at relatively low pressures. However, the observed adsorption isotherm being a combination of types I and IV, where the presence of micropores (size less than 2 nm51) along with mesopores cannot be excluded. The BET specific surface area and the total pore volume (Vtol) of PDASA are determined to be 511 m2 g−1 and 0.61 cm3 g−1, respectively. The micropore volume, Vmicro, and mesopore volume, Vmeso, are 0.07 cm3 g−1 and 0.54 cm3 g−1, respectively (Vmicro/Vtol = 0.12 and Vmeso/Vtol = 0.88), indicaing that both micro- and

Scheme 2. Schematic Representation (2D) of Hyper-crosslinked, Porous Structure of PDASA, Where Each Monomer Is Linked Through a CH- Linker

mesopores are present in PDASA, while the mesopores are dominant. The pore size distribution curve is given in Figure 2B and reveals the average pore size to be between 3 and 5 nm (according to IUPAC convention, pores with sizes below 2 nm 19448

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces

Figure 2. (A) N2 adsorption−desorption isotherm at 77 K and (B) pore size distribution of PDASA. High resolution X-ray photoelectron spectra for the C 1s (C and D) and O 1s regions (E and F) for DASA (C and E) and PDASA (D and F).

are referred to as micropores51), indicating uniform porosity in the polymer. The high surface area and inherent pores with large pore volume provide a sufficient number of active sites and are expected to be responsible for easy transport and storage of Li ions. The highly porous nature of PDASA is also observed from the SEM picture given in Figure 1D and Figure S2. Figure 2 shows the X-ray photoelectron spectra for C 1s and O 1s regions of the monomer and the polymer. DASA shows peaks corresponding to the C 1s level and is assigned to being due to sp2 (from the aromatic ring, 284 eV) and sp3 and O− CO carbons (286.4 and 288.5 eV), respectively. The peak observed at relatively high binding energy value (291−293 eV) may be attributed to the satellite C 1s peak (π−π* transition) due to delocalization of π-electrons. In the case of PDASA, an additional peak is observed at low binding energy (285.2 eV) and is assigned to the sp2 carbon of quinonoid-type linkage present in the polymer. As for the O 1s region, CO is

observed at high binding energy value of 533−533.4 eV and C−O is observed at a binding energy value of 531.5−532.5 eV in the case of monomer as well as the polymer. Further, the thermal stabilities of both DASA and PDASA indicate that the monomer degrades at 300 °C (Figure S3) while the polymer degradation is rather slow. The weight loss for the polymer being small probably indicates higher stability. The lithium storage performance of PDASA is followed using Swagelok cells as given in the Experimental Section. Based on the difference in the energy level of Li/Li+ and the HOMO of the polymer (determined using cyclic voltammetry), the thermodynamic open circuit potential is found to be 4.8 V. Experimentally, we observe an open circuit voltage of 3.1−3.2 V. The cyclic voltammogram depicted in Figure 3A shows a peak between 2.5 and 2 V along with a peak around 1.5−1.0 V vs Li/Li+ in the first cathodic scan. The second and subsequent cycles show peak currents that show reduced currents around 2.0, 1.30, and 0.8 V with corresponding reverse peaks in the 19449

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces

Figure 3. (A) Cyclic voltammograms of PDASA (0.5 mV s−1) in the range of 3 to 0.2 V vs Li/Li+. The voltammograms are carried out from the open circuit voltage (∼3.1 V) to 0.2 V and back. The geometric area of the electrode (0.636 cm2) is used to calculate the current density given on the y-axis. (B) FT-IR spectra of PDASA electrode after 10 cycles: (red) discharged state and (blue) charged state. Spectrum of the as-prepared electrode is also shown.

Scheme 3. Possible Insertion Scheme of Lithium Ions into (A) DASA and (B) PDASA Polymer

However, the carbonyl stretch (around 1760 cm−1) totally disappears after discharge and reappears after charge giving credence to the enolization part of the mechanism. The ex situ IR spectra after different cycles were also recorded (Figure S4) It has been reported that organic materials based on 1,4,5,8naphthalenetetracarboxylic dianhydride (NTCDA)52 and graphene wrapped 3,4,9,10-perylenetetracarboxylic dianhydride (PTCDA),53 show initial Li insertion above 1.1 V attributed to (lithium) enolization reaction at the carbonyl oxygens. In addition, it is known that when the cells are (deep) discharged up to 0.5 V, electrochemical lithiation would take place at the unsaturated carbons as well.52,53 This is referred to as “superlithiation” since the capacity far exceeds the value based on the monomer. Further, the deep discharge is reported to result in irreversibility of the insertion−deinsertion process.53 Third,32 it is speculated that Li insertion and deinsertion may also take place inside the meso-/micropores of the porous polymer when it is deep discharged. It is very likely that multiplicity of sites with different activation energies are available for Li insertion−deinsertion reaction in the polymer matrix. The lithiation into pores and unsaturated carbons may occur at potentials close to and less than 1.2 V vs Li/Li+. Control experiments carried out in the absence of any Li salt in the electrolyte reveals the intrinsic redox behavior of the polymer toward lithiation and delithiation (Figure S5). Based

anodic scan. The observations are similar to the dianhydridebased electrode materials52,53 wherein about five peaks are observed during the first discharge. The large peak observed around 2.5−2.0 V in the first discharge is likely to be due to the interaction of Li+ with the carbonyl groups of anhydride moiety in the polymer.52 The second large peak that appears around 1.0 to 0.5 V vs Li/Li+ in the first discharge is possibly due to the formation of solid−electrolyte interface (SEI) on the electrode surface,32 along with the insertion of Li+ ions into unsaturated carbon subunits as given in Scheme 3. The anodic peaks in the range 1.5−3.0 V are due to the deinsertion of Li+ ions from the anhydride moieties. The low intense peaks observed below 1 V reflect doping−dedoping of Li+ ions inside the pores of the polymer network as has been reported for PDCzBT.32 Ex situ IR measurements have been carried out to understand the mechanism further. Figure 3B provides the IR spectra of the as-prepared material along with the electrode after first discharge (red) and charge (discharge followed by charge, blue). The peak below 1400 cm−1 for the as-prepared polymer may be attributed to C−H bending. The broad peak observed around 1420 cm−1 after discharge is likely to be due to O−C−O···Li bond.52 The broadness of the region possibly masks the C−H bend in the IR spectrum taken after discharge. After charging, the intensity of the 1420 cm−1 peak is reduced and is due to the removal of Li from part of the active sites. 19450

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces

Figure 4. (A) Galvanostatic charge−discharge curves of PDASA at applied currents of (A) 50 and (C) 3 A g−1. Capacity as a function of cycle number at (B) 50 and (D) 3 A g−1. Inset shows the corresponding Coulombic efficiency plots.

the charge and discharge capacity values show a small difference. This may be attributed to the so-called “shuttle effect” seen in Li−S systems and also in Li−organic batteries (Supporting Information).18 The performance in terms of high specific capacity and efficient cyclability of the polymer even at very high currents may be attributed to its high specific surface area and highly porous 3D network structure. This allows easy and fast diffusion of the electrolytes as well as Li+ ions inside the polymer chains throughout the pores thus providing a large number of active sites for Li storage. Along with the pores, the presence of anhydride moieties helps in reversible uptake of Li+ resulting in efficient battery performance. Table S2 gives a comparison of the cell performance in terms of specific capacity and cyclability with other reported porous polymer electrodes. The theoretical capacity has been calculated based on the molecular mass and number of active sites for Li insertion of the monomer unit. If we consider only the anhydride moiety as the active sites, then the number of active sites per monomer unit will be 2 and the theoretical capacity is 194.5 mA h g−1. When the cells are discharged up to 0.5 V, electrochemical lithiation takes place at the unsaturated carbon subunits, and if we consider this possibility, the number of active sites per monomer will be (12 + 2) and hence the theoretical capacity is ∼1170 mA h g−1. In addition to the above, the quinonoid cross-linker (from the trioxane unit during the synthesis) of PDASA may also

on the above discussion, the following general eq 1 may be proposed for lithium insertion and desinsertion as shown in Scheme 3. PDASA + x Li+ + x e− ↔ Lix +PDASA−

(1)

Figure 4A displays typical galvanostatic charge−discharge behavior of PDASA, cycled in the potential region of 3.2 to 0.5 V vs Li/Li+. A large part of the specific capacity in the low potential region below 1 V is due to lithiation into unsaturated carbons (superlithiation) and possibly Li+ doping into polymer pores.28 The specific capacity in the high potential region (above 1 V) is due to Li+ insertion into the anhydride moiety as observed in cyclic voltammetry. Adsorption of Li ions on the surface of the electrode material may also contribute.54 It is observed that the polymer (PDASA) exhibits very high specific discharge capacity of ∼1100 mA h g−1 at an applied current of 50 mA g−1 (Figure 4A,B). This is the highest value observed at a discharge current of 50 mA g−1 to date, among the reported porous, organic polymer materials.29−34 One of the most challenging issues in using organic materials is their poor cyclability at high rate (or current) of discharge. At a high current of 3 A g−1, PDASA exhibits reasonably good and highly stable specific discharge capacity of 120 mA h g−1 for ∼1000 cycles with Coulombic efficiency of ∼85%. The Coulombic efficiencies at different current densities (Table S1) are in the range of 85−90% except for a few odd points where 80% is observed. We believe that the data points are within the statistical error limit. However, as observed from Figure 4B,D, 19451

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces

Figure 5. (A) Galvanostatic charge−discharge behavior and (B) cycling performance of PDASA at different discharge currents.



facilitate electron transfer by delocalization.29 The capacity will be accordingly high (Scheme 3). Further, the cell performance is studied at different applied currents and the stability of the material is examined by cycling the material at different currents. Panels A and B of Figure 5 indicate that, with increasing discharge currents, the specific capacity value decreases as expected, and by changing the discharge currents back and forth, the cell performance is almost retained, which is an indication of the excellent stability of the electrode material. One of the reasons attributed for poor cyclability of (organic) electrode materials in LiPF6-based electrolyte in carbonate solvent55 is the possibility of formation of LiF-type species and subsequent degradation. The degradation is not observed in the present case. The percentage of capacity retention for the present material with respect to both first and second cycles at 50 mA g−1 discharge current density are given in Table 1. It is clear that the capacity

CONCLUSIONS A highly porous, metal-free hyper-cross-linked rigid polymer (PDASA) based on triptycene backbone is synthesized from anthracene and maleic anhydride and has been shown to be an excellent electrode material for rechargeable lithium-ion batteries. This 3D rigid polymer (PDASA) opens up a new class of organic material that may be tuned for other battery systems as well.



The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.6b09575. Synthetic procedure of DASA, additional characterization of DASA and PDASA, and comparisons of performance of different porous electrode materials (PDF)



Table 1. Percentage of Capacity Retention with Respect to First and Second Cycles 100th cycle

500th cycle

1000th cycle

66 87

62 83

59 80

AUTHOR INFORMATION

Corresponding Authors

*(H.M.N.K.) E-mail: [email protected]. *(S.S.) E-mail: [email protected]. Web: http://ipc.iisc. ernet.in/~sampath/new.

percentage of capacity retention with respect to first cycle with respect to second cycle

ASSOCIATED CONTENT

S Supporting Information *

Author Contributions §

D.M. and G.G.Y.K. contributed equally to this work.

Notes

The authors declare no competing financial interest.

■ ■ ■

retention is rather good from the second cycle onward though the irreversible capacity with respect to the first cycle is rather large. This is similar to various electrode materials including inorganics and attributed to the formation of SEI as well as equilibration of the interface. The insoluble nature combined with the porosity and reversible Li+ insertion capacity due to anhydride moieties result in remarkable rate capability of the present polymer material. The SEM pictures of the electrode taken after first discharge, after first charge, and after 1000 cycles (Figure S7) show that the morphology is retained and a highly porous nature is observed. The low magnification micrograph shows no cracks on the electrode surface after 1000 cycles. Also, it is observed that there is no change in UV−vis spectra after cycling (Figure S8)

ACKNOWLEDGMENTS This work was supported by DST, New Delhi, India. D.M. thanks CSIR for a research fellowship. ABBREVIATIONS DASA, dihydroanthracene succinic anhydride PDASA, poly(dihydroanthracene succinic anhydride) REFERENCES

(1) Liang, Y.; Tao, Z.; Chen, J. Organic Electrode Materials for Rechargeable Lithium Batteries. Adv. Energy Mater. 2012, 2, 742−769. (2) Tarascon, J.-M. Towards Sustainable and Renewable Systems for Electrochemical Energy Storage. ChemSusChem 2008, 1, 777−779. (3) Chen, H.; Armand, M.; Demailly, G.; Dolhem, F.; Poizot, P.; Tarascon, J.-M. From Biomass to a Renewable LixC6O6 Organic Electrode for Sustainable Li-Ion Batteries. ChemSusChem 2008, 1, 348−355. 19452

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces (4) Armand, M.; Tarascon, J.-M. Building Better Batteries. Nature 2008, 451, 652−657. (5) Tarascon, J.-M. Key Challenges in Future Li-Battery Research. Philos. Trans. R. Soc., A 2010, 368, 3227−3241. (6) Reddy, A. L. M.; Gowda, S. R.; Shaijumon, M. M.; Ajayan, P. M. Hybrid Nanostructures for Energy Storage Applications. Adv. Mater. 2012, 24, 5045−5064. (7) Scrosati, B.; Garche, J. Lithium Batteries: Status, Prospect and Future. J. Power Sources 2010, 195, 2419−2430. (8) Ellis, B. L.; Lee, K. T.; Nazar, L. F. Positive Electrode Materials for Li-ion and Li-batteries. Chem. Mater. 2010, 22, 691−714. (9) Hernandez-Burgos, K.; Rodriguez-Calero, G. G.; Zhou, W.; Burkhardt, S. E.; Abruna, H. D. Increasing the Gravimetric Energy Density of Organic Based Secondary Battery Cathodes using Small Radius Cations (Li+ and Mg2+). J. Am. Chem. Soc. 2013, 135, 14532− 14535. (10) Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the Development of Advanced Li-Ion Batteries: A Review. Energy Environ. Sci. 2011, 4, 3243−3262. (11) Liang, Y.; Zhang, P.; Yang, S.; Tao, Z.; Chen, J. Fused Heteroaromatic Organic Compounds for High-Power Electrodes of Rechargeable Lithium Batteries. Adv. Energy Mater. 2013, 3, 600−605. (12) Goriparti, S.; Harish, M. N. K.; Sampath, S. Ellagic Acid - A Novel Organic Electrode Material for High Capacity Lithium Ion Batteries. Chem. Commun. 2013, 49, 7234−7236. (13) Schon, T. B.; McAllister, B. T.; Li, P.-F.; Seferos, D. S. The Rise of Organic Electrode Materials for Energy Storage. Chem. Soc. Rev. 2016, 45, 6345. (14) Liang, Y.; Zhang, P.; Chen, J. Function-Oriented Design of Conjugated Carbonyl Compound Electrodes for High Energy Lithium Batteries. Chem. Sci. 2013, 4, 1330−1337. (15) Walker, W.; Grugeon, S.; Mentre, O.; Laruelle, S.; Tarascon, J.M.; Wudl, F. Ethoxycarbonyl-Based Organic Electrode for Li-Batteries. J. Am. Chem. Soc. 2010, 132, 6517−6523. (16) Walker, W.; Grugeon, S.; Vezin, H.; Laruelle, S.; Armand, M.; Wudl, F.; Tarascon, J.-M. Electrochemical Characterization of Lithium4,4′-Tolane-Dicarboxylate for Use As a Negative Electrode in Li-Ion Batteries. J. Mater. Chem. 2011, 21, 1615−1620. (17) Chen, H.; Armand, M.; Courty, M.; Jiang, M.; Grey, C. P.; Dolhem, F.; Tarascon, J.-M.; Poizot, P. Lithium Salt of Tetrahydroxybenzoquinone: Toward the Development of a Sustainable Li-Ion Battery. J. Am. Chem. Soc. 2009, 131, 8984−8988. (18) Song, Z.; Qian, Y.; Otani, M.; Zhou, H. Stable Li-Organic Batteries with Nafion-Based Sandwich-Type Separators. Adv. Energy Mater. 2016, 6, 1501780. (19) Han, X.; Chang, C.; Yuan, L.; Sun, T.; Sun, J. Aromatic Carbonyl Derivative Polymers as High-Performance Li-Ion Storage Materials. Adv. Mater. 2007, 19, 1616−1621. (20) Nokami, T.; Matsuo, T.; Inatomi, Y.; Hojo, N.; Tsukagoshi, T.; Yoshizawa, H.; Shimizu, A.; Kuramoto, H.; Komae, K.; Tsuyama, H.; Yoshida, J. Polymer-Bound Pyrene-4,5,9,10-Tetraone for Fast-Charge and Discharge Lithium-Ion Batteries with High Capacity. J. Am. Chem. Soc. 2012, 134, 19694−19700. (21) Song, Z.; Zhan, H.; Zhou, Y. Anthraquinone Based Polymer as High Performance Cathode Material for Rechargeable Batteries. Chem. Commun. 2009, 448−450. (22) Zhu, Z.; Hong, M.; Guo, D.; Shi, J.; Tao, Z.; Chen, J. All-solidState Lithium Organic Battery with Composite Polymer Electrolyte and Pillar[5]Quinone Cathode. J. Am. Chem. Soc. 2014, 136, 16461− 16464. (23) Sharma, P.; Damien, D.; Nagarajan, K.; Shaijumon, M. M.; Hariharan, M. Perylene-Polyimide-Based Organic Materials for Rechargeable Lithium Batteries. J. Phys. Chem. Lett. 2013, 4, 3192− 3197. (24) Vu, A.; Qian, Y.; Stein, A. Porous Electrode Materials for Lithium-Ion Batteries − How to Prepare Them and What Makes Them Special. Adv. Energy Mater. 2012, 2, 1056−1085. (25) Li, S.-L.; Xu, Q. Metal−Organic Frameworks as Platforms for Clean Energy. Energy Environ. Sci. 2013, 6, 1656−1683.

(26) Zhu, Y.; Xiang, X.; Liu, E.; Wu, Y.; Xie, H.; Wu, Z.; Tian, Y. An Activated Microporous Carbon Prepared from Phenol-MelamineFormaldehyde Resin for Lithium Ion Battery Anode. Mater. Res. Bull. 2012, 47, 2045−2050. (27) Yang, X.; Li, C.; Zhang, G.; Yang, C. Polystyrene-Derived Carbon with Hierarchical Macro-Meso-Microporous Structure for High-Rate Lithium-Ion Batteries Application. J. Mater. Sci. 2015, 50, 6649−6655. (28) Sakaushi, K.; Hosono, E.; Nickerl, G.; Gemming, T.; Zhou, H.; Kaskel, S.; Eckert, J. Aromatic Porous-Honeycomb Electrodes for a Sodium-Organic Energy Storage Device. Nat. Commun. 2013, 4, 1485. (29) Xu, F.; Chen, X.; Tang, Z.; Wu, D.; Fu, R.; Jiang, D. RedoxActive Conjugated Microporous Polymers: A New Organic Platform for Highly Efficient Energy Storage. Chem. Commun. 2014, 50, 4788− 4790. (30) Patra, B. C.; Khilari, S.; Satyanarayana, L.; Pradhan, D.; Bhaumik, A. A New Benzimidazole Based Covalent Organic Polymer Having High Energy Storage Capacity. Chem. Commun. 2016, 52, 7592−7595. (31) Zhang, C.; Yang, X.; Ren, W.; Wang, Y.; Su, F.; Jiang, J.-X. Microporous Organic Polymer-Based Lithium Ion Batteries with Improved Rate Performance and Energy Density. J. Power Sources 2016, 317, 49−56. (32) Zhang, S.; Huang, W.; Hu, P.; Huang, C.; Shang, C.; Zhang, C.; Yang, R.; Cui, G. Conjugated Microporous Polymers with Excellent Electrochemical Performance for Lithium and Sodium storage. J. Mater. Chem. A 2015, 3, 1896−1901. (33) Tian, D.; Zhang, H.-Z.; Zhang, D.-S.; Chang, Z.; Han, J.; Gao, X.-P.; Bu, X.-H. Li-Ion Storage and Gas Adsorption Properties of Porous Polyimides (PIs). RSC Adv. 2014, 4, 7506−7510. (34) Xu, F.; Jin, S.; Zhong, H.; Wu, D.; Yang, X.; Chen, X.; Wei, H.; Fu, R.; Jiang, D. Electrochemically Active, Crystalline, Mesoporous Covalent Organic Frameworks on Carbon Nanotubes for Synergistic Lithium-Ion Battery Energy Storage. Sci. Rep. 2015, 5, 8225. (35) Li, B.; Gong, R.; Wang, W.; Huang, X.; Zhang, W.; Li, H.; Hu, C.; Tan, B. A New Strategy to Microporous Polymers: Knitting Rigid Aromatic Building Blocks by External Cross-Linker. Macromolecules 2011, 44, 2410−2414. (36) Gouda, M. A.; Berghot, M. A.; Elattar, K. M.; Khalil, A. E. G. M. Chemistry of Dibenzobarallene. Turk J. Chem. 2011, 35, 663−697. (37) Swager, T. M. Iptycenes in the Design of High Performance. Acc. Chem. Res. 2008, 41, 1181−1189. (38) Carta, M.; Croad, M.; Malpass-Evans, R.; Jansen, J. C.; Bernardo, P.; Clarizia, G.; Friess, K.; Lanč, M.; McKeown, N. B. Triptycene Induced Enhancement of Membrane Gas Selectivity for Microporous Tröger’s Base Polymers. Adv. Mater. 2014, 26, 3526− 3531. (39) Ghanem, B. S.; Swaidan, R.; Litwiller, E.; Pinnau, I. UltraMicroporous Triptycene-based Polyimide Membranes for HighPerformance Gas Separation. Adv. Mater. 2014, 26, 3688−3692. (40) Sydlik, S. A.; Chen, Z.; Swager, T. M. Triptycene Polyimides: Soluble Polymers with High Thermal Stability and Low Refractive Indices. Macromolecules 2011, 44, 976−980. (41) Liang, Q.; Jiang, G.; Zhao, Z.; Li, Z.; MacLachlan, M. J. CdSDecorated Triptycene-Based Polymer: Durable Photocatalysts for Hydrogen Production under Visible-Light Irradiation. Catal. Sci. Technol. 2015, 5, 3368−3374. (42) He, Y.; Zhu, X.; Li, Y.; Peng, C.; Hu, J.; Liu, H. Efficient CO2 Capture by Triptycene-Based Microporous Organic Polymer with Functionalized Modification. Microporous Mesoporous Mater. 2015, 214, 181−187. (43) Rabbani, M. G.; Reich, T. E.; Kassab, R. M.; Jackson, K. T.; ElKaderi, H. M. High CO2 Uptake and Selectivity by Triptycene-Derived Benzimidazole-Linked Polymers. Chem. Commun. 2012, 48, 1141− 1143. (44) Wang, H.; Pan, L.; Deng, W.; Yang, G.; Liu, X. One-Pot Synthesis of Triptycene-Based Porous Organic Frameworks with Tailored Micropore Environments for Highly Efficient and Selective Amine Adsorption. Polym. J. 2016, 48, 787−792. 19453

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454

Forum Article

ACS Applied Materials & Interfaces (45) Jin, T.; An, S.; Yang, X.; Hu, J.; Wang, H.; Liu, H.; Tian, Z.; Jiang, D.; Mehio, N.; Zhu, X. Efficient Adsorptive Desulfurization by Task-Specific Porous Organic Polymers. AIChE J. 2016, 62, 1740− 1746. (46) Zhang, C.; Zhu, P.-C.; Tan, L.; Liu, J.-M.; Tan, B.; Yang, X.-L.; Xu, H.-B. Triptycene-Based Hyper-Cross-Linked Polymer Sponge for Gas Storage and Water Treatment. Macromolecules 2015, 48, 8509− 8514. (47) Vinodh, R.; Hemalatha, P.; Ganesh, M.; Peng, M. M.; Abidov, A.; Palanichamy, M.; Cha, W. S.; Jang, H.-T. Novel Microporous Hypercross-Linked Conjugated Quinonoid Chromophores with Broad Light Absorption and CO2 Sorption Characteristics. RSC Adv. 2014, 4, 3668−3674. (48) Vinodh, R.; Ganesh, M.; Peng, M. M.; Abidov, A.; Palanichamy, M.; Cha, W. S.; Jang, H.-T. Microporous Hypercross-linked Conjugated Quinonoid Chromophores of Anthracene: Novel Polymers for CO2 Adsorption. Chin. J. Polym. Sci. 2015, 33, 224−235. (49) Zhang, C.; Liu, Y.; Li, B.; Tan, B.; Chen, C.-F.; Xu, H.-B.; Yang, X.-L. Triptycene-Based Microporous Polymer: Synthesis and Its Gas Storage Properties. ACS Macro Lett. 2012, 1, 190−193. (50) Schneider, P. Adsorption Isotherms of Microporous-Mesoporous Solid Revisited. Appl. Catal., A 1995, 129, 157−165. (51) Gregg, S. J.; Sing, K. S. W. Adsorption, Surface Area and Porosity, 2nd ed.; Academic Press: London, 1982. (52) Han, X.; Qing, G.; Sun, J.; Sun, T. How Many Lithium Ions can be Inserted onto Fused C6 Aromatic Ring Systems? Angew. Chem., Int. Ed. 2012, 51, 5147−5151. (53) Cui, D.; Tian, D.; Chen, S.; Yuan, L. Graphene Wrapped 3,4,9,10-Perylenetetracarboxylic Dianhydride as a High-Performance Organic Cathode for Lithium Ion Batteries. J. Mater. Chem. A 2016, 4, 9177−9183. (54) Wang, G.; Shen, X.; Yao, J.; Park, J. Graphene Nanosheets for Enhanced Lithium Storage in Lithium Ion Battery. Carbon 2009, 47, 2049−2053. (55) Lux, S. F.; Lucas, I. T.; Pollak, E.; Passerini, S.; Winter, M.; Kostecki, R. The Mechanism of HF Formation in LiPF6 Based Organic Carbonate Electrolytes. Electrochem. Commun. 2012, 14, 47−50.

19454

DOI: 10.1021/acsami.6b09575 ACS Appl. Mater. Interfaces 2017, 9, 19446−19454