Ultrahigh-Capacity Organic Anode with High-Rate ... - ACS Publications

Aug 22, 2017 - Ultrahigh-Capacity Organic Anode with High-. Rate Capability and Long Cycle Life for. Lithium-Ion Batteries. Yan Wang,. †. Yonghong D...
0 downloads 0 Views 5MB Size
Ultrahigh-Capacity Organic Anode with HighRate Capability and Long Cycle Life for Lithium-Ion Batteries Yan Wang,† Yonghong Deng,§ Qunting Qu,† Xueying Zheng,† Jingyu Zhang,† Gao Liu,*,‡ Vincent S. Battaglia,‡ and Honghe Zheng*,† †

College of Physics, Optoelectronics and Energy & Collaborative Innovation Center of Suzhou Nano Science and Technology, Soochow University, Suzhou, Jiangsu 215006, PR China ‡ Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley, California 94720, United States § Department of Materials Sciences, South University of Science and Technologies of China, Shenzhen, 518000, China S Supporting Information *

ABSTRACT: Organic rechargeable batteries have attracted extensive attention as a potential alternative for the current lithium-ion batteries. However, most of the reports are limited to organic macromolecules or modified small organic molecules which exhibit low reversible capacity, poor rate capability, and very limited cycle life. Herein, a small organic compound, maleic acid, is adopted as the anode for lithium ion batteries without any modification. It exhibits an ultrahigh reversible capacity of ca. 1500 mAh g−1 at 46.2 mA g−1 current density. Even at a high current density of 46.2 A g−1, the electrode still delivers a capacity of 570.8 mAh g−1. When cycled at 2.31 A g−1, a capacity retention of 98.1% is obtained after 500 cycles. The excellent performance of the maleic acid organic anode is ascribed to its small volume effect and unique lithium-ion storage mechanisms. This new type of organic anode material may have a great opportunity for large-scale energy-storage systems with high-power properties. friendly, low cost, flexible, and structural diverse.10−15 However, most reports relate to organic cathode rather than anode because their redox potentials are usually between 2.0 and 4.0 V versus Li/Li+.16−18 Besides, all small-molecule organic compounds generally show very poor cycling performance. Pyrene4,5,9,10-tetraone is able to deliver 275 mAh g−1 in the first cycle, but it retains only ca. 120 mAh g−1 after 20 cycles at the rate of 0.2 C.19 2-Vinyl-4,8-dihydrobenzo[1,2-b:4,5-b′]-dithiophene-4,8-dione exhibits a reversible capacity of 225 mAh g−1 in the first cycle, and the capacity drops to 50 mAh g−1 after 25 cycles.20 Naphthoquinone has a reversible capacity of 190 mAh g−1 and a capacity retention of about 45% after 100 cycles at 0.2 C.21 Anthraquinone has a reversible capacity of about 250 mAh g−1 in the first cycle, and the capacity drops to 30 mAh g−1 after 100 cycles at 0.2 C.17 The origin of poor cycling performance of small-molecule organic compounds is generally believed to be due to the high solubility of the small organic molecules in aprotic electrolytes commonly used in LIBs.17,22,23 To address

W

ith increasing demand for electric vehicle (EV) and energy-storage system (ESS) applications, largescale lithium-ion batteries (LIBs) have emerged as the most promising choice for chemical power sources.1−3 However, conventional LIBs are not immediately suitable for large-scale applications because the energy density and power density for the commercially available electrodes are very limited. Graphite anode, for example, has a theoretical specific capacity of only 372 mAh g−1, and its rate performance is even worse than that of most of the typical commercial cathode materials such as LiMn2O4, LiCoO2, LiFePO4, etc. These drawbacks definitely hinder its application in EV and ESS applications.4−6 To overcome the limitations arising from the graphite anode, one of the major challenges is finding new electrodes with high energy density and power density. Although many kinds of high-capacity anode materials (e.g. Si, Sn, and transition-metal oxides) have been widely investigated in the latest two decades, the progress seems to be very sluggish, and these materials still exhibit poor rate capability and cycling behavior.7−9 Organic compounds offer new possibilities for developing large-scale LIBs, because they are renewable, environmentally © 2017 American Chemical Society

Received: July 16, 2017 Accepted: August 22, 2017 Published: August 22, 2017 2140

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

http://pubs.acs.org/journal/aelccp

Letter

ACS Energy Letters

Figure 1. Electrochemical performances of the maleic acid anode in the voltage range of 0.01−3.0 V. (a) Cyclic voltammograms; (b) charge and discharge profiles at 46.2 mA g−1; (c) rate capability at room temperature after 20 cycles at 46.2 mA g−1 (red and blue symbols indicate charge and discharge rate capability, respectively); and (d) cycling and Coulombic efficiency at current density of 2.31 A g−1 after rate test (the inset is the cycling and Coulombic efficiency at current density of 462 mA g−1 from initial cycles).

higher than those obtained from many of the reported organic materials in the literature. At 2.31 A g−1 current density, a capacity retention of 98.1% is obtained after 500 electrochemical cycles, demonstrating superior cycling stability of the electrode. At an extreme high charge rate of 46.2 A g−1, corresponding to 36 s for a full charge, the electrode retained 570.8 mAh g−1 capacity. The superior electrochemical performances of the maleic acid organic anode are ascribed to its small volume effect and unique Li ion storage mechanisms. The promising results may open up a new approach for seeking small molecule organic electrodes for future LIBs of high energy, high power, and long life span. The electrochemical properties of the maleic acid anode were evaluated. Figure 1a shows the cyclic voltammogram (CV) curves within a voltage window of 0.01−3 V at a scan rate of 0.1 mV s−1. A broad and strong peak in the first cathodic process is

this issue, one feasible strategy is to reduce the solubility of the organic compounds through polymerization,24−27 salt formation,28−31 grafting,19,32−35 etc.20,29,36,37 Most of these approaches are, however, only partially successful as the capacity somewhat decreases because of the increase in the molecular weight. In addition, most of the experiments were conducted at very low charge−discharge current.24,26,33,35,36 High power performance of these materials has never been demonstrated to the best of our knowledge. Therefore, finding new organic compounds that allow for fast charge and discharge with superior cyclability is very essential for developing organic LIBs for EV and ESS purposes. In the present work, a small organic molecule, maleic acid, was adopted as the anode of LIBs without any modification. This new type of organic material exhibits a reversible capacity of ca. 1500 mAh g−1 at 46.2 mA g−1, which is almost five times 2141

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters

Figure 2. SEM images of the maleic acid electrode: (a, b) pristine electrode, (c, d) cross-section of pristine electrode, and (e, f) after 500 electrochemical cycles.

capacity of ca. 700 mAh g−1, the low Coulombic efficiency of maleic acid electrode can be explained. In the second cycle, the Coulombic efficiency attained 86.0%. In the following cycles, the reversible capacity of maleic acid electrode undergoes an increase with cycle number and finally stabilizes at ca. 1500 mAh g−1 after 20 electrochemical cycles. The reversible capacity of 1500 mAh g−1 obtained with the organic compound is almost five times higher than those with other different organic materials in the literature.10,12,19,30,39,42 Compared to the 372 mAh g−1 of the traditional graphite anode, the reversible capacity is also dramatically improved. Rate capability is one of the important properties for EV and ESS applications. Very different from the reported organic electrodes, maleic acid anode exhibits outstanding charge and discharge rate capability. As shown in Figure 1c, at a charge rate of 4.62 A g−1, it delivers a reversible capacity of 1025.2 mAh g−1. At 23.1 and 46.2 A g−1, it still delivers 736.7 and 570.8 mAh g−1, respectively. Meanwhile, at 4.62, 23.1, and 46.2 A g−1 of the discharge rate, the electrode is still able to deliver 919.6, 607.2, and 457.2 mA g−1, respectively. When the charge and discharge rate returns to 46.2 mA g−1, a total recovery of the initial capacity is obtained. It is worth noting that maleic acid also exhibits good high (55 °C) and low (−10 °C) temperature rate performance (Figure S2). The rate performance of the maleic acid anode is remarkably higher than that of most of the reported organic electrodes. Nokami and co-workers reported an aromatic carbonyl compound, which delivers a reversible capacity of only 225.5 mAh g−1 at 2.62 A g−1.19 Li4C8O6 nanosheets show a reversible capacity of 175 mAh g−1 at a current density of 1205 mA g−1.30 Wang reported an extended π-conjugated substance, which delivers a capacity of 72 mAh

observed at around 0.98 V (from 0.7 to 1.3 V), and this process disappears in the following cycles. This strong irreversible reduction can be ascribed to two reasons: One is the irreversible reaction of maleic acid with solvated lithium ions,38−40 corresponding to the transformation from maleic acid to lithium maleate. The other one is the formation of solid electrolyte interface (SEI), which is a common character for organic electrodes.13,30,39−41 Both categories of reactions are irreversible, and this explains the disappearance of the reduction in the second cathodic sweep. In the first anodic scan, one pronounced anodic peak at around 1.58 V is observed, and it disappears in the following cycles. As the cycle continue, three typical cathodic peaks at around 2.04, 0.56, and 0 V and one typical anodic peak at around 2.64 V gradually formed, associated with the lithiation and delithiation process of maleic acid, respectively. From the 20th cycles, the redox peaks retain stable, and CV curves in the subsequent cycles almost overlap, indicating good reversibility and the stability of the maleic acid electrode. The initial 20 discharge−charge profiles of the maleic acid anode at a current density of 46.2 mA g−1 from 0.01 to 3 V are presented in Figure 1b. In the first discharge, onset of the high potential plateau at around 1.6 V is ascribed to the SEI formation and the irreversible reaction of maleic acid with the solvated lithium ions. This is identical with the cathodic process obtained from CV curves. After this reduction process, the electrode is lithiated with decrease of the electrode potential. The first discharge and charge capacities are obtained to be 2688.9 mAh g−1 and 1233.5 mAh g−1, respectively, corresponding to the first Coulombic efficiency of 45.9%. Considering that the plateau between 1.6 and 1.3 V contributes an irreversible 2142

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters

Figure 3. (a) EIS measurements of the maleic acid electrode conducted at different electrochemical states, (b) Nyquist plots at different DODs and SOCs in the first cycle, and (c) impedance change with cycle number after the electrodes are charged to 3 V.

g−1 at 10 A g−1.38 Rate capability of the maleic acid anode obtained in this study is also much higher than those of the traditional graphite anodes and the state-of-the-art silicon anodes.4,34,43−45 Figure 1d shows the long-term cycling behavior of the maleic acid anode at different current densities. When cycled at a current density of 462 mA g−1, the initial six Coulombic efficiencies are 41.2, 78.6, 87.7, 93.0, 95.7, and 97.7, respectively. Thereafter, the Coulombic efficiency of the electrode attains ca. 99.5% for each cycle. Meanwhile, the capacity shows a continuous increase and remains stable after 20 cycles. The capacity increase during the initial 20 cycles is attributed to the electrochemical activation of maleic acid electrode. After 100 cycles, a reversible capacity of 1027.7 mAh g−1 is obtained. When cycled at 2.31 A g−1 current density after the rate test, the electrode shows a reversible capacity of 851.7 mAh g−1 after 500 cycles with a capacity retention of 98.1%, revealing superior cycling stability of the maleic acid electrode. Figure 2 shows the SEM images of pristine maleic acid electrode and the electrode after long-term cycling test. For the pristine electrode (Figure 2a,b), the maleic acid particles of 600−800 nm diameter can be clearly observed, and the conductive carbon additive is homogeneously distributed in the electrode. Figure 2c,d shows the cross-sectional images of the pristine maleic acid electrode. It is seen that the conductive carbon is distributed around the maleic acid balls. The electrode thickness is about 10 um on the copper foil. After 500 electrochemical cycles (Figure 2e,f), the maleic acid particles were covered with a thick film of insoluble materials. This is believed to be attributed to the development of SEI film during the long-term cycling. The abundant functional carbonyl groups in the maleic acid are known to be very helpful for SEI formation during cell formation. Hence, it may play an important role in the SEI formation and thus prevent the dissolution of the active material in the organic electrolyte.

To get more details about the electrochemical performance of the maleic acid anode, the influences of the voltage range, AB content, and counter electrode on its electrochemical performances were studied. Figure S3 shows the electrochemical performance of maleic acid in the voltage range of 0.01−2.0 V. Compared to the electrode in the 0.01−3.0 V range, the reversible capacity in the 0.01−2.0 V range is significantly decreased. This is because the charging plateau at around 2.5 V (Figure 1b) contributes a lot to its reversible capacity. Nevertheless, the reversible capacity of the maleic acid anode in 0.01−2.0 V range is still significantly higher than that of most of the reported organic electrodes in the literature.4,34,43−45 Figure S4 presents electrochemical performance of the maleic acid electrodes with different AB contents (10−40 wt %) in the 0.01−3.0 V voltage range. Obviously, lack of AB leads to severe deterioration of rate capability, indicating electronic conductivity plays a crucial role controlling the reaction rate. SEM images of pristine maleic acid electrodes with different AB contents are shown in Figure S5. When AB content is 40 wt %, maleic acid balls are effectively surrounded by the conductive carbon. Good connectivity between AB particles and the maleic acid balls contributes to high electric conductivity of the electrode laminate and thus promotes the electrochemical reaction kinetics. Decreasing AB content results in fewer AB particles around the maleic acid balls, thereby affecting the rate performance of the electrode. In addition, the capacity of maleic acid electrode is greatly affected by the counter electrode. With LiFePO4 as the counter electrode in the 0.01−3.0 V range (Figure S6), reversible capacity of the maleic acid electrode is significantly reduced, indicating some subtle reactions may exist in the full cell. Nevertheless, good cyclability of the cell is obtained. To further verify the reversible Li storage capability of the maleic acid anode, a spin-coated maleic acid anode without addition of conductive carbon and polymeric binder were 2143

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters

Figure 4. SEM images of the maleic acid electrode at different electrochemical states.

prepared. Electrochemical properties of the electrode are shown in Figure S7. Under this condition, the maleic acid electrode still has more than 170 mAh g−1 reversible capacity at 46.2 mA g−1, indicating that pure maleic acid is capable of storing lithium ions reversibly. Electrochemical performance of the spin-coated maleic acid electrode without AB and PVDF is dramatically decreased. Like most other organic electrode materials, the poor electronic connectivity is a main cause hindering the reaction kinetics. Theoretical capacity of maleic acid anode is dependent on the transferred electron number of the compound.17 As the acetylene black can also reversibly store lithium in the 0.01−3 V range (real capacity of 270 mAh g−1 as seen in Figures 1c and S8), the real capacity of the electrode should be the sum capacity of the maleic acid and the acetylene black, i.e. Ct (maleic acid) × 50 wt % + 270 mAh g−1 × 40 wt %. For the real capacity of the maleic acid electrode (maleic acid: AB: PVDF = 5:4:1) obtained to be 1459.7 mAh g−1 at 231 mA g−1, about 12 electrons are involved in the electrochemical reactions (see Table S1). Electrochemical impedance spectroscopy (EIS) studies were performed at different electrochemical states, and the results are depicted in Figure 3. Figure 3b shows the Nyquist plots of the maleic acid electrode at different depth of discharge (DOD) and state of charge (SOC) in the first cycle. For the asassembled maleic acid electrode (point A), there is only one semicircle in the high-to-medium frequency region, reflecting the interfacial impedance between the electrode and the electrolyte.13,36 When discharged to 0.8 V (point B), two depressed semicircles are observed in high and intermediate frequency regions, showing the solid electrolyte interface resistance (RSEI) and the charge-transfer (Rct) process, respectively. When discharged to 0.01 V potential (point C), an increase of the electrode impedance is probably due to the slow kinetic process at the discharge limit of the electrode. Afterward, when the electrode is charged to 1, 2, and 3 V (points D, E, and F), respectively, the impedance undergoes a decrease at first and a negligible change later, implying that the maleic acid electrode is quite stable in the lithiation− delithiation process. Figure 3c shows the Nyquist plots of maleic acid electrode after different electrochemical cycles. By comparison, the overall impedance of the maleic acid electrode

is significantly lower than those of many other reported organic electrodes in the literature.13,41,46 This explains the superior high rate capability. Also, it should be noted that the diameter of the semicircle shows only a slight increase after 1700 charge−discharge cycles (for RSEI and Rct values, see Table S2). This implies that the particle surface of the maleic acid anode is stable during the long-term cycles. A slight increase of chargetransfer resistance is believed to be associated with the evolution of the electrode−electrolyte interface. The morphologies of maleic acid electrode at different electrochemical states (the points in Figure 3a) are compared. As shown in Figure 4, when discharged to 0.8 V (B), an insoluble coating is clearly observed on the maleic acid surface. This is mainly attributed to the irreversible reactions of maleic acid with solvated lithium ions. At 0.01 V potential (C), an increase of the maleic acid ball is related to the volume expansion of the active material as a result of full lithiation and SEI formation on the maleic acid surface. When the electrode is charged to 1, 2, and 3 V (points D, E, and F) respectively, no significant morphological and volume change is observed, revealing a stable morphology of the maleic acid electrode in the electrochemical cycles. SEM images of the maleic acid anodes at different DODs and SOCs (Figure 4) show that there are no obvious morphological changes at different electrochemical stages, indicating a small volume change of the organic material during charge and discharge. After 500 charge−discharge cycles (Figure 2e, f), the maleic acid balls are still well-preserved. It can be concluded that the small volume effect and stable electrode−electrolyte interface of the maleic acid anode explains the excellent cycling performance of the electrode. To investigate the Li storage mechanisms of the maleic acid, ex-situ X-ray photoelectron spectroscopy (XPS) analyses were conducted. All the samples were retrieved from the dismantled cells at different DOD and SOC states (see Figure 5a). As seen in Figure 5b, there are significant differences in the collected C 1s and O 1s spectra between the pristine maleic acid electrode and the electrode at various electrochemical states, indicating structural changes of the compound during the electrochemical process. Figure 5c shows the collected Li 1s spectra of the maleic acid at different electrochemical states. For the asassembled electrode (A), no Li 1s signal is discerned. With 2144

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters

Figure 5. (a) Points selected for XPS measurements during the first charge and discharge; XPS spectra of (b) C 1s, O 1s, and (c) Li 1s of maleic acid electrodes at different electrochemical states; and (d) area ratios of Li 1s/C 1s peaks and Li 1s/O 1s peaks; O 1s XPS spectra with fitted components of (e) fresh maleic acid electrode and maleic acid electrode after being discharged to (f) 0.8 V, (g) 0.01 V, and charged to (h) 3 V.

lithiation of the electrode (B → C), intensity of the Li 1s peak increases. With the delithiation of the electrode (D → F), intensity of the Li 1s peak deceases. However, the Li 1s peak is

still clearly visible when the electrode is charged to 3 V (at fully delithiated state). This is because a large amount of lithium is not extracted from maleic acid electrode because of the strong 2145

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters Scheme 1. Structure and Proposed Lithium Storage Mechanisms of the Maleic Acid Electrode

irreversible reactions in the first discharge. Area ratios of both Li 1s/C 1s and Li 1s/O 1s for the maleic acid electrode at different electrochemical stages verify the reversible lithiation− delithiation processes, as shown in Figure 5d. Figure 5e−h presents the O 1s-fitted XPS spectra. For the pristine maleic acid electrode (Figure 5e), the O1s peaks with binding energy of 533.3 and 532.4 eV are attributed to O−H and CO of maleic acid, respectively (structure I).47 When discharged to 0.8 V (Figure 5f), the O−H peak disappeared while the CO peak is remarkably decreased. Meanwhile, a new peak centered at 531.4 emerges, corresponding to O−Li,48 indicating the transformation from maleic acid to lithium maleate. At 0.01 V potential (Figure 5g), the CO signal completely disappeared. When charged to 3 V (Figure 5h), CO peak appeared again, indicating the formation of lithium maleate, reflecting the C O bond is the main active site for reversibly storing lithium. Hence, the XPS analyses provide clear evidence of the lithium storage mechanism of the maleic acid electrode. The conjugated carbonyl groups act as the active sites for lithium ions. To clearly reveal the reduction process of the maleic acid, Scheme 1 is proposed, which is divided into five steps. In the first step, two carboxyl hydrogen atoms of maleic acid (structure I) are irreversibly replaced by two lithium ions, leading to the formation of structure II. In the second step, a two-electron reduction reaction happens on the two carbonyl groups, resulting in the formation of structure III. In the third step, another two-electron reduction happens, producing structure IV. Structure IV is not stable, OLi− may leave from the structure producing a more stable structure V by eliminating two irreversible lithium oxide (Li2O). In the fourth step, two-electron reduction continues on the two carbonyl groups, leading to the formation of structure VI. In the fifth step, another two-electron reduction results in the formation of structure VII. According to this proposed reduction process of

the maleic acid electrode, eight electrons are involved in the reversible electrochemical reactions. Considering the real capacity of the maleic acid electrode is obtained to be as high as 1500 mAh g−1 at 46.2 mA g−1, this means there are about 12 electrons involved in the reaction (see Table S1). We believe surface adsorption of lithium may be another important contribution to the reversible capacity of the maleic acid. Of course, this requires further confirmation in future work. To better understand the electrochemical redox mechanism of maleic acid, lithium storage properties of lithium/sodium maleate were also investigated to compare with that of the maleic acid. Details of the preparation of lithium/sodium maleate are described in the Supporting Information (sample synthesis). Briefly, maleic salts were synthesized through a simple acid−base reaction between maleic acid and lithium/ sodium hydroxide in a solution of methanol (Scheme S1). FTIR and Raman spectra (Figure S1) clearly illustrate the formation of lithium/sodium maleate. The charge−discharge profiles of the lithium/sodium maleate electrodes at 46.2 mA g−1 are presented in Figure S9a. The first discharge capacity of lithium maleate (705.9 mAh g−1) and sodium maleate (865.0 mAh g−1) are significantly lower than that of the maleic acid electrode (2688.9 mAh g−1). Figure S9b presents the differential capacity plots (dQ/dV versus cell potential) for the lithium/sodium maleate electrodes in the first two cycles. Compared to the maleic acid electrode (Figure 1a), there is no sharp anodic peak around 2.6 V. Rate capabilities for both lithium maleate and sodium maleate (Figure S9c) are significantly lower than that of the maleic acid electrode (Figure 1c). This implies that the H proton on the carbonyl functional group plays an important role in the high capacity of the maleic acid. Further studies are required to explain the detailed mechanisms of the lithiation and delithiation for the maleic acid electrode. 2146

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters

Materials for Lithium-Ion Batteries. Appl. Surf. Sci. 2016, 389, 240− 246. (6) Deng, T.; Zhou, X. Porous Graphite Prepared by Molybdenum Oxide Catalyzed Gasification as Anode Material for Lithium Ion Batteries. Mater. Lett. 2016, 176, 151−154. (7) Demirkan, M. T.; Trahey, L.; Karabacak, T. Cycling Performance of Density Modulated Multilayer Silicon Thin Film Anodes in Li-Ion Batteries. J. Power Sources 2015, 273, 52−61. (8) Dou, P.; Cao, Z.; Zheng, J.; Wang, C.; Xu, X. Solid Polymer Electrolyte Coating Three-Dimensional Sn/Ni Bimetallic Nanotube Arrays for High Performance Lithium-Ion Battery Anodes. J. Alloys Compd. 2016, 685, 690−698. (9) Ng, S. H.; Wang, J.; Wexler, D.; Konstantinov, K.; Guo, Z. P.; Liu, H. K. Highly Reversible Lithium Storage in Spheroidal Carbon-Coated Silicon Nanocomposites as Anodes for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2006, 45, 6896−6899. (10) Hanyu, Y.; Sugimoto, T.; Ganbe, Y.; Masuda, A.; Honma, I. Multielectron Redox Compounds for Organic Cathode Quasi-Solid State Lithium Battery. J. Electrochem. Soc. 2014, 161, A6−A9. (11) 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. (12) Liu, K.; Zheng, J.; Zhong, G.; Yang, Y. Poly(2,5-Dihydroxy-1,4Benzoquinonyl Sulfide) (PDBS) as a Cathode Material for Lithium Ion Batteries. J. Mater. Chem. 2011, 21, 4125−4131. (13) Wang, C.; Xu, Y.; Fang, Y.; Zhou, M.; Liang, L.; Singh, S.; Zhao, H.; Schober, A.; Lei, Y. Extended π-Conjugated System for FastCharge and -Discharge Sodium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 3124−3130. (14) Zhan, L. Z.; Song, Z. P.; Zhang, J. Y.; Tang, J.; Zhan, H.; Zhou, Y. H.; Zhan, C. M. Synthesis and Properties of Novel Organic Thiolane Polymer as Cathode Material for Rechargeable Lithium Batteries. J. Appl. Electrochem. 2008, 38, 1691−1694. (15) 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. (16) Gao, X. P.; Yang, H. X. Multi-Electron Reaction Materials for High Energy Density Batteries. Energy Environ. Sci. 2010, 3, 174−189. (17) Song, Z.; Zhou, H. Towards Sustainable and Versatile Energy Storage Devices: an Over View of Organic Electrode Materials. Energy Environ. Sci. 2013, 6, 2280−2301. (18) Wang, Y.; Zhang, L.; Zhang, L.; Zhang, F.; He, P.; Zheng, H.; Zhou, H. Reversible Lithium-Ion Uptake in Poly(methylmethacrylate) Thin-Film via Lithiation/De-lithiation at In Situ Formed Intramolecular Cyclopentanedione. Adv. Energy Mater. 2016, 6, 1601375. (19) 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. (20) Haupler, B.; Hagemann, T.; Friebe, C.; Wild, A.; Schubert, U. S. Dithiophenedione-Containing Polymers for Battery Application. ACS Appl. Mater. Interfaces 2015, 7, 3473−3479. (21) Lee, J.; Kim, H.; Park, M. J. Long-Life, High-Rate LithiumOrganic Batteries Based on Naphthoquinone Derivatives. Chem. Mater. 2016, 28, 2408−2416. (22) Xie, J.; Zhang, Q. Recent Progress in Rechargeable Lithium Batteries with Organic Materials as Promising Electrodes. J. Mater. Chem. A 2016, 4, 7091−7106. (23) Liang, Y.; Tao, Z.; Chen, J. Organic Electrode Materials for Rechargeable Lithium Batteries. Adv. Energy Mater. 2012, 2, 742−769. (24) 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. (25) Le Gall, T.; Reiman, K. H.; Grossel, M. C.; Owen, J. R. Poly(2,5Dihydroxy-1,4-Benzoquinone-3,6-Metyle-ne): a New Organic Polymer as Positive Electrode Material for Rechargeable Lithium Batteries. J. Power Sources 2003, 119−121, 316−320.

In conclusion, a novel organic material of high capacity, high power, and long cycle life is reported in this work. Maleic acid as a LIB anode exhibits a reversible capacity of ca. 1500 mAh g−1 at 46.2 mA g−1, which is five times greater than those of different organic compounds reported in the literature. This electrode displays high-power performance, which is able to deliver reversible capacities of 736.7 and 570.8 mAh g−1 at charge current densities of 23.1 and 46.2 A g−1, respectively. At 2.31 A g−1 charge and discharge rate, it retains 98.1% of its initial capacity after 500 electrochemical cycles. The organic material is not prone to dissolving in the organic electrolyte. Small volume effect and the unique lithium-ion storage mechanisms explain its superior cycling performance. All these results reveal that the small organic compound has a very promising future for next-generation LIBs of high energy and power density and long cycle life. This work may challenge the conventional concept that small organic compounds are not readily suitable for storing lithium ions.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsenergylett.7b00622. Experimental methods, electrochemical performance data, and additional figures and tables as described in the text (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Qunting Qu: 0000-0003-2590-2695 Gao Liu: 0000-0001-6886-0507 Honghe Zheng: 0000-0001-9115-0669 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are greatly indebted to 863 project of the Department of Science and Technologies, China (Contract No. 2015AA034601) and the funding of Natural Science Foundation of China (NSFC, Contract Nos. 21473120 and 21403148).



REFERENCES

(1) Goodenough, J. B.; Kim, Y. Challenges for Rechargeable Li Batteries†. Chem. Mater. 2010, 22, 587−603. (2) Suzuki, T.; Hayashiya, H.; Yamanoi, T.; Kawahara, K. Introduction and Practical Use of Energy Storage System with Lithium-ion Battery for DC Traction Power Supply System. IEEJ. Journal IA 2016, 5, 20−25. (3) Tran, T. H.; Harmand, S.; Desmet, B.; Filangi, S. Experimental Investigation on the Feasibility of Heat Pipe Cooling for HEV/EV Lithium-Ion Battery. Appl. Therm. Eng. 2014, 63, 551−558. (4) Wang, Y.; Zheng, H.; Qu, Q.; Zhang, L.; Battaglia, V. S.; Zheng, H. Enhancing Electrochemical Properties of Graphite Anode by Using Poly(Methylmethacrylate)−Poly(Vinylidene Fluoride) Composite Binder. Carbon 2015, 92, 318−326. (5) Wang, X.; Liu, Y.; Arandiyan, H.; Yang, H.; Bai, L.; Mujtaba, J.; Wang, Q.; Liu, S.; Sun, H. Uniform Fe3O4 Microflowers Hierarchical Structures Assembled with Porous Nanoplates as Superior Anode 2147

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148

Letter

ACS Energy Letters (26) Song, Z.; Zhan, H.; Zhou, Y. Anthraquinone Based Polymer as High Performance Cathode Material for Rechargeable Lithium Batteries. Chem. Commun. 2009, 448−450. (27) Song, Z.; Zhan, H.; Zhou, Y. Polyimides: Promising EnergyStorage Materials. Angew. Chem., Int. Ed. 2010, 49, 8444−8448. (28) 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. (29) Renault, S.; Geng, J.; Dolhem, F.; Poizot, P. Evaluation of Polyketones with N-Cyclic Structure as Electrode Material for Electrochemical Energy Storage: Case of Pyromellitic Diimide Dilithium Salt. Chem. Commun. 2011, 47, 2414−2416. (30) Wang, S.; Wang, L.; Zhang, K.; Zhu, Z.; Tao, Z.; Chen, J. Organic Li4C8H2O6 Nanosheets for Lithium-Ion Batteries. Nano Lett. 2013, 13, 4404−4409. (31) Yao, M.; Araki, M.; Senoh, H.; Yamazaki, S. i.; Sakai, T.; Yasuda, K. Indigo Dye as a Positive-Electrode Material for Rechargeable Lithium Batteries. Chem. Lett. 2010, 39, 950−952. (32) Pirnat, K.; Dominko, R.; Cerc-Korosec, R.; Mali, G.; Genorio, B.; Gaberscek, M. Electrochemically Stabilised Quinone Based Electrode Composites for Li-Ion Batteries. J. Power Sources 2012, 199, 308−314. (33) Oyama, N.; Sarukawa, T.; Mochizuki, Y.; Shimomura, T.; Yamaguchi, S. Significant Effects of Poly(3,4-Ethylenedioxythiophene) Additive on Redox Responses of Poly(2,5-Dihydroxy-1,4-Benzoquinone-3,6-Methylene) Cathode for Rechargeable Li Batteries. J. Power Sources 2009, 189, 230−239. (34) Song, Z.; Xu, T.; Gordin, M. L.; Jiang, Y. B.; Bae, I. T.; Xiao, Q.; Zhan, H.; Liu, J.; Wang, D. Polymer-Graphene Nanocomposites as Ultrafast -Charge and -Discharge Cathodes for Rechargeable Lithium Batteries. Nano Lett. 2012, 12, 2205−2211. (35) Genorio, B.; Pirnat, K.; Cerc-Korosec, R.; Dominko, R.; Gaberscek, M. Electroactive Organic Molecules Immobilized onto Solid Nanoparticles as a Cathode Material for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2010, 49, 7222−7224. (36) Bhosale, M. E.; Krishnamoorthy, K. Chemically Reduced Organic Small-Molecule-Based Lithium Battery with Improved Efficiency. Chem. Mater. 2015, 27, 2121−2126. (37) Yao, M.; Senoh, H.; Yamazaki, S.-i.; Siroma, Z.; Sakai, T.; Yasuda, K. High-Capacity Organic Positive-Electrode Material Based on a Benzoquinone Derivative for Use in Rechargeable Lithium Batteries. J. Power Sources 2010, 195, 8336−8340. (38) Wang, C.; Xu, Y.; Fang, Y.; Zhou, M.; Liang, L.; Singh, S.; Zhao, H.; Schober, A.; Lei, Y. Extended π-Conjugated System for FastCharge and -Discharge Sodium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 3124−3130. (39) Wang, H. g.; Yuan, S.; Si, Z.; Zhang, X. b. Multi-Ring Aromatic Carbonyl Compounds Enabling High Capacity and Stable Performance of Sodium-Organic Batteries. Energy Environ. Sci. 2015, 8, 3160− 3165. (40) Zhao, R. R.; Cao, Y. L.; Ai, X. P.; Yang, H. X. Reversible Li and Na Storage Behaviors of Perylenetetracarboxylates as Organic Anodes for Li- and Na-Ion Batteries. J. Electroanal. Chem. 2013, 688, 93−97. (41) Li, Z.; Zhou, J.; Xu, R.; Liu, S.; Wang, Y.; Li, P.; Wu, W.; Wu, M. Synthesis of Three Dimensional Extended Conjugated Polyimide and Application as Sodium-Ion Battery Anode. Chem. Eng. J. 2016, 287, 516−522. (42) Hu, P.; Wang, H.; Yang, Y.; Yang, J.; Lin, J.; Guo, L. RenewableBiomolecule-Based Full Lithium-Ion Batteries. Adv. Mater. 2016, 28, 3486−3492. (43) Koo, B.; Kim, H.; Cho, Y.; Lee, K. T.; Choi, N. S.; Cho, J. A Highly Cross-Linked Polymeric Binder for High-Performance Silicon Negative Electrodes in Lithium Ion Batteries. Angew. Chem., Int. Ed. 2012, 51, 8762−8767. (44) Park, S. J.; Zhao, H.; Ai, G.; Wang, C.; Song, X.; Yuca, N.; Battaglia, V. S.; Yang, W.; Liu, G. Side-Chain Conducting and PhaseSeparated Polymeric Binders for High-Performance Silicon Anodes in Lithium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 2565−2571.

(45) Park, S. H.; Kim, H. J.; Lee, J.; Jeong, Y. K.; Choi, J. W.; Lee, H. Mussel-Inspired Polydopamine Coating for Enhanced Thermal Stability and Rate Performance of Graphite Anodes in Li-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 13973−13981. (46) Wang, Y.; Ding, Y.; Pan, L.; Shi, Y.; Yue, Z.; Shi, Y.; Yu, G. Understanding the Size-Dependent Sodium Storage Properties of Na2C6O6-Based Organic Electrodes for Sodium-Ion Batteries. Nano Lett. 2016, 16, 3329−3334. (47) Lu, W.; Xie, K.; Chen, Z.; Xiong, S.; Pan, Y.; Zheng, C. A New Co-Solvent for Wide Temperature Lithium Ion Battery Electrolytes: 2,2,2-Trifluoroethyl n-Caproate. J. Power Sources 2015, 274, 676−684. (48) Zhuang, G.; Chen, Y.; Ross, P. N., Jr The Reaction of Lithium with Carbon Dioxide Studied by Photoelectron Spectroscopy. Surf. Sci. 1998, 418, 139−149.

2148

DOI: 10.1021/acsenergylett.7b00622 ACS Energy Lett. 2017, 2, 2140−2148