亲爱的研友该休息了!由于当前在线用户较少,发布求助请尽量完整的填写文献信息,科研通机器人24小时在线,伴您度过漫漫科研夜!身体可是革命的本钱,早点休息,好梦!

Imine-Rich Poly( o -phenylenediamine) as High-Capacity Trifunctional Organic Electrode for Alkali-Ion Batteries

中国科学院 中国 北京 资源(消歧) 图书馆学 化学 计算机科学 历史 考古 计算机网络
作者
Tieheng Sun,Zongjun Li,Xu Yang,Sai Wang,Yingmei Zhu,Xinbo Zhang
出处
期刊:CCS Chemistry [Chinese Chemical Society]
卷期号:1 (4): 365-372 被引量:38
标识
DOI:10.31635/ccschem.019.20190003
摘要

Open AccessCCS ChemistryRESEARCH ARTICLE1 Oct 2019Imine-Rich Poly(o-phenylenediamine) as High-Capacity Trifunctional Organic Electrode for Alkali-Ion Batteries Tao Sun, Zong-Jun Li, Xu Yang, Sai Wang, Yun-Hai Zhu and Xin-Bo Zhang Tao Sun State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022 (China) University of Chinese Academy of Sciences, Beijing 100049 (China) Google Scholar More articles by this author , Zong-Jun Li State Key Laboratory of Electroanalytical Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022 (China) Google Scholar More articles by this author , Xu Yang State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022 (China) Google Scholar More articles by this author , Sai Wang State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022 (China) Google Scholar More articles by this author , Yun-Hai Zhu State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022 (China) Google Scholar More articles by this author and Xin-Bo Zhang *Corresponding author: E-mail Address: [email protected] State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022 (China) Google Scholar More articles by this author https://doi.org/10.31635/ccschem.019.20190003 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesTrack Citations ShareFacebookTwitterLinked InEmail Alkali-ion batteries, including potassium-ion batteries, lithium-ion batteries, and sodium-ion batteries are important energy storage devices; however, with the cation size increased, there exists a great challenge for an inorganic electrode material to accommodate the different properties of the alkali-ion. Herein, as a proof-of-concept experiment, an imine-rich poly(o-phenylenediamine) (PoPD) is synthesized through a rational controllable oxidization. Due to the abundance of active sites and ladder-conjugated structure, PoPD in the optimized oxidation state endows alkali-ion batteries with a stable cyclability at high capacity. The highly reversible redox performance of PoPD with alkali-ions is verified by theoretical calculations and demonstrated as a trifunctional electrode material (537 and 307 mAh·g−1 for Li and Na storage capacity after 300 cycles, respectively), especially for successful application in potassium storage (450 mAh·g−1 after 205 cycles), and provides compelling evidence for the wide application of organic electrode materials. Download figure Download PowerPoint Introduction In the past decades, batteries based on the reversible storage of alkali-ions have achieved great success. Among them, lithium-ion batteries (LIBs) have brought great convenience to our daily lives; however, their development still lags behind the requirements for sustainable energy storage devices.1,2 To solve the scarcity and uneven distribution of Li, the nonlithium battery systems based on alternative alkali metals ions, such as sodium-ion batteries (SIBs) and potassium-ion batteries (PIBs), have attracted great attention due to their Earth-abundant mineral resource requirements and similar operation principle as LIBs.3–5 Theoretically, the reduction potential of K+/K (−2.93 V) is close to Li+/Li (−3.04 V), which is even lower than its Na counterpart (−2.71 V), and yields PIBs with high voltage and energy.6–8 However, the large-scale applications of PIBs are still facing many challenges, including relative low capacity and bad cycling stability.9–12 Due to the large size of K ions, an inorganic electrode based on intercalation or an alloying reaction is often associated with drastic structural changes during the electrochemical cycling, which will deteriorate its performances.13–17 As a matter of fact, this phenomenon is not only limited to PIBs. As is well known, even though PIBs share a similar energy storage mechanism to that of their Li counterparts, similar inorganic materials that are widely used in LIBs cannot be successfully used in SIBs or in PIBs. Each type of alkali-ion requires a unique crystal structure and insertable sites. Restricted by the rigid structure, it is an almost impossible task to find a suitable inorganic electrode material for the reversible storage and release of alkali-ions with different properties. Increasing the size of the alkali-ion presents a great challenge for traditional inorganic electrode materials. Interestingly, an organic compound is a promising candidate for the intercalation of large metal ions.18–21 Assembled by the van der Waals force, an organic compound featuring a large spacing and flexible structure provides an elastic buffer matrix to accommodate the volume expansion/contraction during the K ion insertion/extraction process. However, the dissolution of small molecular material in the electrolyte is a serious problem, which severely disrupts its cycling stability. To overcome this issue, an efficient strategy is to construct a high-performance polymer electrode material with a stable skeleton and abundant electroactive groups.22 Among various redox-active polymers, poly(o-phenylenediamine) (PoPD), characterized by a highly reversible redox chemistry and unique doping mechanism, has attracted a great deal of attention in the electrochemistry field.23–25 In fact, the abundant nitrogen-containing functional groups and ladder-conjugated structures would endow PoPD with a high capacity and stable reversibility, and it would thus potentially form a good candidate electrode material. Importantly, the flexible host polymer matrix allows for frequent intercalation and deintercalation of K ions, which avoids the worry of volume change. The working principle of the organic material is based on the redox reaction of the functional group accompanied by the incorporation/release of alkali metal ions or electrolyte anions.1 To a certain extent, organic materials are not typically restricted by the choice of counter-ion, which indicates that the same organic material is practicable for different energy storage devices.26–28 Therefore, it could be predicted that PoPD is a versatile electrode, wherein K+, Na+, and Li+ could offer reversible storage and release. Doping state is a crucial factor determining the properties of an aromatic amine polymer.24,26 Therefore, to obtain satisfactory electrochemical performance, the oxidation state of PoPD should be rationally designed. Quinoid imine (–N=) unit is a characteristic of the oxidation state of PoPD, which is also the major active site to bond with alkali metal ions. A greater quinoid imine content in PoPD means more active sites and better electrochemical performance; however, excessive oxidation will deteriorate the electronic conductivity and stability.26 Accordingly, oxidation state optimization is crucial for PoPD to exhibit the optimized electrochemical properties, but to achieve this remains challenging. Herein, PoPD with a variable oxidation state is synthesized as a trifunctional electrode for alkali-ion batteries. Taking the potassium storage performance as an example, through a rational manipulation of oxidant in the synthesis procedure, the relationship between oxidation state and electrochemical behavior is investigated. Due to the abundance of active sites, PoPD in the optimized oxidation state exhibits excellent PIB performance, including a high specific capacity and cycling stability (450 mAh·g−1 after 205 cycles). As a versatile electrode, PoPD also exhibits a reversible capacity of 537 and 307 mAh·g−1 for LIBs and SIBs, respectively, which offer convincing evidence for the great potential application of organic electrode materials. Results and Discussion Poly(o-phenylenediamine was synthesized through chemical oxidative polymerization using ammonium persulfate (APS) [(NH4)2S2O8] as oxidant (Figure 1a).25 To tune the degree of oxidation and electrochemical performance, a series of samples with APS/oPD ratios varying from 0.5∶1 to 3∶1 have been fabricated, which were labeled PoPD-0.5 to PoPD-3, respectively (). Fourier-transform infrared spectroscopy (FTIR) clearly demonstrates the formation of PoPD (Figure 1b). The stretching vibration of the quinoid ring is found at 1618 cm−1. Two signals at 1350 and 1221 cm−1 correspond to the C–N stretch mode in the quinoid and benzenoid rings, respectively. The band at 850 cm−1 can be assigned to the C–H out-of-plane bending vibration of the 1,2,4,5-tetrasubstituted benzenoid, indicating the existence of a rigid phenazine skeleton in PoPD. However, the peak located at 1134 cm−1 corresponds to the in-plane bending vibrations of the 1,2,4-trisubstituted benzenoid, suggesting the PoPD does not have a perfect ladder structure. Therefore, it can be concluded that the as-prepared PoPD has a rigid skeleton with some open rings.25 With increasing oxidant/monomer ratio, the stretching of the quinoid ring at 1618 cm−1 is increased, but the signal at 1517 cm−1 (benzenoid ring) becomes weaker and even disappears when the ratio reaches 3∶1, which indicates an decreased benzenoid content but increased quinoid content. Accordingly, the oxidation state of PoPD improves as the amount of oxidant increased. Figure 1 | Preparation and characterization of PoPD. (a) Synthesis of PoPD. (b) FTIR of PoPD samples. (c) XPS spectrum of PoPD. Download figure Download PowerPoint This conclusion can also be demonstrated by the elucidation of the N1s spectra of PoPD. According to the X-ray photoelectron spectroscopy (XPS) (Figure 1c and ), the peak of the N1s spectra can be fitted with three components: –N=, –NH–, and –NH2, which are centered at 398.1, 399.2, and 400.2 eV, respectively. For the PoPD-0.5, the content of quinoid imine is 34.8%, whereas for PoPD-3, this value increased to 58% (). The reduced benzenoid amine content means the benzenoid amine has been oxidized to quinoid imine, thus yielding a PoPD sample in high oxidation state. Generally, the area ratio of imine to amine peaks (–N=/–NH–) can be used to determine the oxidation state of aromatic amine/diamine polymers. As the XPS analysis results shown in Figure 1c, it is obvious that the ratio of imine to amine is improved as the oxidant/monomer ratio is increased, which substantiates the assumption that the proportion of imine increases as the oxidation state increases. However, it does not imply that a higher oxidation state presents a better electrochemical performance. Through the comparison of the cycling performance of different PoPD samples, the effect of oxidation state on electrochemical performance can be clearly demonstrated (Figure 2a and ). For the samples in a low oxidation state, PoPD-0.5 nearly cannot deliver any capacity. Unexpectedly, PoPD exhibited excellent potassium storage performance when the molar ratio increased to 1∶1. The capacity of PoPD-1 achieved 147 mAh·g−1 in the first cycle and increased gradually in the following 50 cycles. During the next 70 cycles, the reversible capacity of PoPD-1 became stable and reached 240 mAh·g−1. Even the discharge capacity of PoPD-2 reached 253 mAh·g−1 in the first cycle; however, this remarkable value decreased upon cycling and was maintained at 150 mAh·g−1 after 120 cycles. On increasing the molar ratio to 3∶1, there was a decrease in the capacity. A reasonable oxidation state is critical for the aromatic amine/diamine polymers to exhibit the optimized electrochemical activity. Because of the repulsion of charge between structural units, the excessive oxidation state is unstable and usually involves side reactions with the electrolyte. Therefore, 1∶1 is the optimized molar ratio, wherein PoPD electrodes exhibit the optimized electrochemical performance. Figure 2 | Electrochemical performance of PoPD as PIBs anode. (a) Cycling performance comparison of PoPD samples (charge capacity profiles vs cycle number under a current of 200 mA·g−1). (b) Cycling performance of PoPD-1. (c) Discharge/charge profiles of PoPD-1. (d) Rate performance of PoPD-1. Download figure Download PowerPoint Figure 2c shows the discharge/charge curves of the 1st, 2nd, 10th, and 100th cycle for the PoPD-1 electrode at 50 mA·g−1. The large capacity gap between the initial discharge and the next charge process corresponds to the formation of a solid electrolyte interphase layer. When tested at 50 mA·g−1, the reversible capacity of PoPD-1 reached 450 mAh·g−1 after which it became stable from the 100th cycle and continued to sustain this high capacity for the next 105 cycles (Figure 2b). Such a high capacity and cycling stability are rarely seen in PIBs or in organic electrode materials even without using any special carbon-based composite techniques (). For PoPD-1, it should be noted that there exists an activation process during the cycling process (Figure 2a,b). On the one hand, due to the insufficient contact between the polymer and the electrolyte, the capacity increasing stage can be identified as an “activation process.”29 On the other hand, this phenomenon might probably be attributed to the oxidation of benzenoid amine to quinoid imine during charge/discharge process. In other words, the cycling process will improve the efficiency of redox reactions and expose more active sites.26 Excepting the high capacity, the rate capability of PoPD-1 was further investigated under various test conditions. As displayed in Figure 2d, PoPD-1 delivers a stable capacity of 313, 227, 149, 104, and 73 mAh·g−1 at different current rates of 100, 200, 500, 1000, and 2000 mA·g−1, respectively. When the current rate returned to 50 mA·g−1, the PoPD-1 electrode recovered a reversible capacity of 300 mAh·g−1. To gain insight into the energy storage mechanism of PoPD, density functional theory (DFT) calculations were applied to discuss the reaction process from the aspects of reaction kinetics and thermodynamics (Figure 3). To simplify the calculation, the project was initiated by designing a core structure of PoPD, which only contains benzenoid amine and quinoid imine. First, natural bond orbital (NBO) charge distributions are performed to analyze the process of potassium insertion (). When the PoPD is reduced to monoanion PoPD·–, the negative charge on quinoid imine nitrogen (N18 and N19) increased greatly (from −0.484 to −0.607) (). When the first potassium was inserted, imine nitrogen (N18) and potassium possess the most negative (−0.730) and positive charges (0.937), which indicates the formation of a chemical bond. On receipt of the second electron, the N19 atom showed the greatest increase of negative charge in monoanion PoPD-1K– (). Therefore, an N19 atom would be the most active site to bind with the second potassium. Based on the above analysis, it can be deduced that imine nitrogen is the most active site to bind with K, and this reflects the importance of imine (C=N bond) in PoPD. This conclusion is supported by the thermodynamics calculations, wherein the negative stabilization energy value indicates a favorable binding of the potassium ions (Figure 3c and ). It is worth noting that the binding energy of the second potassium (−0.75 eV) is stronger than the first one (−0.50 eV), which indicates that the insertion of the first potassium facilitates the second step. Figure 3 | DFT calculations and redox mechanism. (a) Electrochemical redox process of PoPD. (b) Major bond length change of PoPD during the reduction process. (c) Stabilization energies at various reduction stages of PoPD. (d–f) Three configurations of PoPD-K. Download figure Download PowerPoint During the stepwise redox process of PoPD, the evolution of the conjugated structure is clearly manifested by the changes in the major bond lengths (Figure 3b and ). With the aid of a ladder-conjugated system, the homogenization of bond length occurs, wherein the distinct length gap between double bond and single bond gradually decreased during the process of potassium insertion.30,31 In the process of reduction, the π-conjugated skeleton could effectively delocalize the negative charge, and thus achieve the stabilization of the reduced PoPD. The length of the C=N double bond is increased as the reduction process; in the fully reduced state, the length gap between the C=N (C7–N19 and C2–N18) and C–N (C1–N18 and C10–N19) bonds reaches a minimum, which implies the consumption of the C=N bond and is consistent with the above NBO charge analysis. The ladder-conjugated skeleton plays a significant role in determining the electrochemical performance of PoPD, in which the electron transfer not only will be easier, but also presents a stable backbone that can tolerate frequent redox in the discharge/charge process. More importantly, the rigid ladder-conjugated structures can also provide the potential binding active sites to storage K ions, except the C=N bond. In fact, structures of benzene and pyrazine are also strong binding sites for the electron-deficient K cations, which can be attributed to the interaction of potassium with the π electrons.30,32 presents the NBO charge distributions in the complexes of PoPD and potassium, wherein the potassium atom above the plane of the benzene ring and phenazine ring bearing the largest positive charge densities (0.957 and 0.943) are shown, indicating the strong interaction between potassium and six-membered rings. The distance of the centroid of the benzene ring and potassium is around at 2.672 Å, which is similar to the length of the K–N bond (Figure 3d,f). However, this distance is shortened to 2.585 Å when potassium is placed above the plane of the phenazine ring (Figure 3e), in which the most negative charge in the N19 and N20 atoms (−0.711 and −0.724) would exhibit a strong interaction with the potassium (). Based on above discussion, it can be concluded that PoPD is a promising electrode material for KIBs due to its diversified energy storage mechanism and abundant active sites. To investigate the structural change of PoPD at different stages of electrochemical reactions and to verify the theoretical studies, ex situ XPS analysis was performed (Figure 4a). For the fresh electrode, the characteristic peaks are in good accordance with the spectra of pristine PoPD. After being discharged to 0.01 V, the peak located at 398.6 eV (–N=) weakens, indicating the consumption of imine. Meanwhile, a new peak centered at 286.6 eV emerges, which implies the formation of a new bond between potassium and nitrogen (K–N). Recharged to 3.0 V, the K–N bond disappeared; all the peaks recovered to the pristine state. The reversible change of the imine vibration mode strongly supports the participation of imine in the reversible reaction with potassium. This conclusion gains support from the ex situ FTIR spectroscopy (Figure 4b), in which the stretching vibration of the quinoid ring located at 1665 cm−1 gradually weakens during the reduction process. The stretching vibrations of the C=C bond at 1505 cm−1 nearly disappeared in the discharged state, suggesting the reaction of the benzene ring and potassium. In the charge process, these peaks appear reversibly. In addition, the repeatability of the FTIR spectral signals in other regions also provides compelling evidence for the participation of the entire PoPD during the discharge/charge process. Figure 4 | Ex situ analysis of PoPD-1 electrode at different states. (a) Ex situ XPS local scan spectra of N1s regions. (b) Ex-situ FTIR characterizations. Download figure Download PowerPoint As mentioned above, the working principle of PoPD is based on the redox reaction of the functional group and is not typically restricted by the choice of counter-ion, which implies that PoPD can also be used as an electrode material for SIBs and LIBs. Fortunately, this assumption is demonstrated by the theoretical analysis, in which the stabilization energy profiles and bond length change explicitly demonstrate that the PoPD is also suitable to bind with Li and Na (). To confirm the calculation prediction, the electrochemical behavior of PoPD for LIBs and SIBs was examined (Figure 5 and ). After the initial activation process, PoPD delivered a reversible lithium storage capacity of 537 mAh·g−1 and sodium storage capacity of 307 mAh·g−1 even after the 300th cycle (Figure 5b, ()), which provides convincing evidence for the broad practicability of this organic electrode material. Figure 5 | Electrochemical performance of PoPD-1 as LIBs and SIBs anode at a current density of 500 mA·g−1. (a) Discharge/charge profiles. (b) Cycling performance. Download figure Download PowerPoint Conclusion In summary, in response to the challenges of using an inorganic electrode for the reversible accommodation of different size metal ions, an imine-rich PoPD is obtained through a rational controllable oxidization, acting as a trifunctional electrode in alkali-ion batteries. Based on substantial characterization techniques and DFT calculations, PoPD with abundant active sites exhibits a stable cyclability at high capacity when acting as a PIB anode (450 mAh·g−1 after 205 cycles). At the same time, its successful application as a versatile electrode in lithium and sodium storage (537 and 307 mAh·g−1after 300 cycles for LIBs and SIBs, respectively) extends the application range of PoPD and may even be useful for multivalent- and dual-ion batteries in the future. We anticipate that these findings will enlighten the design and application of organic electrode materials for cheap, green, sustainable, and versatile energy storage devices. Supporting Information Supporting information is available. Conflicts of Interest The authors declare no competing financial interests. Acknowledgments This work was financially supported by the National Natural Science Foundation of China (21725103 and 51472232), JCKY2016130B010, Jilin Province Science and Technology Development Plan Funding Project (20180101203JC and 20160101289JC), and Changchun Science and Technology Development Plan Funding Project (18DY012). T.S. and Z.J.L. contributed equally to this work. References 1. Song Z.; Zhou H.Towards Sustainable and Versatile Energy Storage Devices: An Overview of Organic Electrode Materials.Energy Environ. Sci.2013, 6, 2280–2301. Google Scholar 2. Wang Y.; Chen R.; Chen T.; Lv H.; Zhu G.; Ma L.; Wang C.; Jin Z.; Liu J.Emerging Non-Lithium Ion Batteries.Energy Storage Mater.2016, 4, 103–129. Google Scholar 3. Zhao Q.; Lu Y.; Chen J.Advanced Organic Electrode Materials for Rechargeable Sodium-Ion Batteries.Adv. Energy Mater.2016, 7, 1601792. Google Scholar 4. Wang H.-G.; Yuan S.; Ma D.-L.; Huang X.-L.; Meng F.-L.; Zhang X.-B.Tailored Aromatic Carbonyl Derivative Polyimides for High-Power and Long-Cycle Sodium-Organic Batteries.Adv. Energy Mater.2014, 4, 1301651. Google Scholar 5. Lee M.; Hong J.; Kim H.; Lim H.-D.; Cho S. B.; Kang K.; Park C. B.Organic Nanohybrids for Fast and Sustainable Energy Storage.Adv. Mater.2014, 26, 2558–2565. Google Scholar 6. Chen Y.; Luo W.; Carter M.; Zhou L.; Dai J.; Fu K.; Lacey S.; Li T.; Wan J.; Han X.; Bao Y.; Hu L.Organic Electrode for Non-Aqueous Potassium-Ion Batteries.Nano Energy2015, 18, 205–211. Google Scholar 7. Eftekhari A.; Jian Z.; Ji X.Potassium Secondary Batteries.ACS Appl. Mater. Interfaces2017, 9, 4404–4419. Google Scholar 8. Kim H.; Kim J. C.; Bianchini M.; Seo D. H.; Rodriguez-Garcia J.; Ceder G.Recent Progress and Perspective in Electrode Materials for K-Ion Batteries.Adv. Energy Mater.2017, 8, 1702384. Google Scholar 9. Jian Z.; Luo W.; Ji X.Carbon Electrodes for K-Ion Batteries.J. Am. Chem. Soc.2015, 137, 11566–11569. Google Scholar 10. Jian Z.; Xing Z.; Bommier C.; Li Z.; Ji X.Hard Carbon Microspheres: Potassium-Ion Anode Versus Sodium-Ion Anode.Adv. Energy Mater.2016, 6, 1501874. Google Scholar 11. Luo W.; Wan J.; Ozdemir B.; Bao W.; Chen Y.; Dai J.; Lin H.; Xu Y.; Gu F.; Barone V.; Hu L.Potassium Ion Batteries With Graphitic Materials.Nano Lett.2015, 15, 7671–7677. Google Scholar 12. Zhao J.; Zou X.; Zhu Y.; Xu Y.; Wang C.Electrochemical Intercalation of Potassium Into Graphite.Adv. Funct. Mater.2016, 26, 8103–8110. Google Scholar 13. Zhang W.; Mao J.; Li S.; Chen Z.; Guo Z.Phosphorus-Based Alloy Materials for Advanced Potassium-Ion Battery Anode.J. Am. Chem. Soc.2017, 139, 3316–3319. Google Scholar 14. Sultana I.; Ramireddy T.; Rahman M. M.; Chen Y.; Glushenkov A. M.Tin-Based Composite Anodes for Potassium-Ion Batteries.Chem. Commun.2016, 52, 9279–9282. Google Scholar 15. Han J.; Niu Y.; Bao S.-J.; Yu Y.-N.; Lu S.-Y.; Xu M.Nanocubic KTi2(PO4)3 Electrodes for Potassium-Ion Batteries.Chem. Commun.2016, 52, 11661–11664. Google Scholar 16. Kishore B.; Venkatesh G.; Munichandraiah N.K2Ti4O9: A Promising Anode Material for Potassium Ion Batteries.J. Electrochem. Soc.2016, 163, A2551–A2554. Google Scholar 17. Ren X.; Zhao Q.; McCulloch W. D.; Wu Y.MoS2 as a Long-Life Host Material for Potassium Ion Intercalation.Nano Res.2017, 10, 1313–1321. Google Scholar 18. Jian Z.; Liang Y.; Rodríguez-Pérez I. A.; Yao Y.; Ji X.Poly(anthraquinonyl sulfide) Cathode for Potassium-Ion Batteries.Electrochem. Commun.2016, 71, 5–8. Google Scholar 19. Deng Q.; Pei J.; Fan C.; Ma J.; Cao B.; Li C.; Jin Y.; Wang L.; Li J.Potassium Salts of Para-Aromatic Dicarboxylates as the Highly Efficient Organic Anodes for Low-Cost K-Ion Batteries.Nano Energy2017, 33, 350–355. Google Scholar 20. Lei K.; Li F.; Mu C.; Wang J.; Zhao Q.; Chen C.; Chen J.High K-Storage Performance Based on the Synergy of Dipotassium Terephthalate and Ether-Based Electrolytes.Energy Environ. Sci.2017, 10, 552–557. Google Scholar 21. Xing Z.; Jian Z.; Luo W.; Qi Y.; Bommier C.; Chong E. S.; Li Z.; Hu L.; Ji X.A Perylene Anhydride Crystal as a Reversible Electrode for K-Ion Batteries.Energy Storage Mater.2016, 2, 63–68. Google Scholar 22. Muench S.; Wild A.; Friebe C.; Häupler B.; Janoschka T.; Schubert U. S.Polymer-Based Organic Batteries.Chem. Rev.2016, 116, 9438–9484. Google Scholar 23. Zhao R.; Zhu L.; Cao Y.; Ai X.; Yang H. X.An Aniline-Nitroaniline Copolymer as a High Capacity Cathode for Na-Ion Batteries.Electrochem. Commun.2012, 21, 36–38. Google Scholar 24. Li X.-G.; Huang M.-R.; Duan W.; Yang Y.-L.Novel Multifunctional Polymers From Aromatic Diamines by Oxidative Polymerizations.Chem. Rev.2002, 102, 2925–3030. Google Scholar 25. Li X.-G.; Ma X.-L.; Sun J.; Huang M.-R.Powerful Reactive Sorption of Silver(I) and Mercury(II) Onto Poly(o-phenylenediamine) Microparticles.Langmuir2009, 25, 1675–1684. Google Scholar 26. Sun T.; Li Z. J.; Wang H. G.; Bao D.; Meng F. L.; Zhang X. B.A Biodegradable Polydopamine-Derived Electrode Material for High-Capacity and Long-Life Lithium-Ion and Sodium-Ion Batteries.Angew. Chem. Int. Ed.2016, 55, 10662–10666. Google Scholar 27. Yang A.; Wang X.; Lu Y.; Miao L.; Xie W.; Chen J.Core-Shell Structured 1,[email protected]2 Cathode for Lithium Batteries.J. Energy Chem.2018, 27, 1644–1650. Google Scholar 28. Lu Y.; Zhang Q.; Li L.; Niu Z.; Chen J.Design Strategies Toward Enhancing the Performance of Organic Electrode Materials in Metal-Ion Batteries.Chem2018, 4, 2786–2813. Google Scholar 29. Wu J.; Rui X.; Long G.; Chen W.; Yan Q.; Zhang Q.Pushing Up Lithium Storage Through Nanostructured Polyazaacene Analogues as Anode.Angew. Chem. Int. Ed.2015, 54, 7354–7358. Google Scholar 30. Kim K. C.; Liu T.; Lee S. W.; Jang S. S.First-Principles Density Functional Theory Modeling of Li Binding: Thermodynamics and Redox Properties of Quinone Derivatives for Lithium-Ion Batteries.J. Am. Chem. Soc.2016, 138, 2374–2382. Google Scholar 31. Lee M.; Hong J.; Seo D. H.; Nam D. H.; Nam K. T.; Kang K.; Park C. B.Redox Cofactor From Biological Energy Transduction as Molecularly Tunable Energy-Storage Compound.Angew. Chem. Int. Ed.2013, 52, 8322–8328. Google Scholar 32. Lee H. H.; Park Y.; Shin K.-H.; Lee K. T.; Hong S. Y.Abnormal Excess Capacity of Conjugated Dicarboxylates in Lithium-Ion Batteries.ACS Appl. Mater. Interfaces2014, 6, 19118–19126. Google Scholar Previous article FiguresReferencesRelatedDetailsCited ByMa K, Liu Y, Jiang H, Hu Y, Si R, Liu H and Li C (2020) Multivalence-Ion Intercalation Enables Ultrahigh 1T Phase MoS2 Nanoflowers to Enhanced Sodium-Storage Performance, CCS Chemistry, 3:5, (1472-1482), Online publication date: 1-May-2021.Jiang C, Meng X, Zheng Y, Yan J, Zhou Z and Tang Y (2021) High-Performance Potassium-Ion-Based Full Battery Enabled by an Ionic-Drill Strategy, CCS Chemistry, 3:9, (85-94), Online publication date: 1-Sep-2021. Issue AssignmentVolume 1Issue 4Page: 365-372Supporting Information Copyright & Permissions© 2019 Chinese Chemical SocietyKeywordspoly(o-phenylenediamine)sodium-ion batteriesorganic electrodepotassium-ion batterieslithium-ion batteriesAcknowledgmentsThis work was financially supported by the National Natural Science Foundation of China (21725103 and 51472232), JCKY2016130B010, Jilin Province Science and Technology Development Plan Funding Project (20180101203JC and 20160101289JC), and Changchun Science and Technology Development Plan Funding Project (18DY012). T.S. and Z.J.L. contributed equally to this work. Downloaded 1,861 times Loading ...
最长约 10秒,即可获得该文献文件

科研通智能强力驱动
Strongly Powered by AbleSci AI
更新
大幅提高文件上传限制,最高150M (2024-4-1)

科研通是完全免费的文献互助平台,具备全网最快的应助速度,最高的求助完成率。 对每一个文献求助,科研通都将尽心尽力,给求助人一个满意的交代。
实时播报
从容芮完成签到,获得积分0
2分钟前
CharlotteBlue应助迷糊的橙子采纳,获得30
2分钟前
5分钟前
侯小菊发布了新的文献求助10
6分钟前
hx完成签到,获得积分20
6分钟前
这不是我本名完成签到,获得积分0
7分钟前
xiewuhua完成签到,获得积分10
7分钟前
顾矜应助侯小菊采纳,获得10
8分钟前
黑蚊子多完成签到 ,获得积分10
8分钟前
lixuebin完成签到 ,获得积分10
8分钟前
w1x2123完成签到,获得积分10
11分钟前
YYYYYYYYY完成签到,获得积分10
12分钟前
12分钟前
zeer0707发布了新的文献求助10
12分钟前
12分钟前
风起云涌龙完成签到 ,获得积分0
13分钟前
14分钟前
四氧化三铁完成签到,获得积分10
14分钟前
yanhua发布了新的文献求助10
14分钟前
郑牛牛发布了新的文献求助10
14分钟前
香蕉觅云应助郑牛牛采纳,获得10
14分钟前
搜集达人应助海绵采纳,获得10
14分钟前
lzxbarry完成签到,获得积分0
14分钟前
HS完成签到,获得积分10
16分钟前
tt发布了新的文献求助10
17分钟前
qqJing完成签到,获得积分10
17分钟前
大个应助qqJing采纳,获得10
17分钟前
大个应助薄荷味的soda采纳,获得10
18分钟前
丘比特应助科研通管家采纳,获得10
19分钟前
LXM完成签到 ,获得积分10
19分钟前
柯语雪完成签到 ,获得积分10
20分钟前
CharlotteBlue应助Hayat采纳,获得30
21分钟前
等于几都行完成签到 ,获得积分10
22分钟前
23分钟前
23分钟前
卑微老大完成签到 ,获得积分10
24分钟前
称心嫣娆完成签到,获得积分10
24分钟前
小二郎应助称心嫣娆采纳,获得10
25分钟前
CharlotteBlue应助pugongying采纳,获得20
25分钟前
秋雪瑶应助qiuxuan100采纳,获得20
25分钟前
高分求助中
Teaching Social and Emotional Learning in Physical Education 900
Gymnastik für die Jugend 600
Chinese-English Translation Lexicon Version 3.0 500
Electronic Structure Calculations and Structure-Property Relationships on Aromatic Nitro Compounds 500
マンネンタケ科植物由来メロテルペノイド類の網羅的全合成/Collective Synthesis of Meroterpenoids Derived from Ganoderma Family 500
[Lambert-Eaton syndrome without calcium channel autoantibodies] 440
Plesiosaur extinction cycles; events that mark the beginning, middle and end of the Cretaceous 400
热门求助领域 (近24小时)
化学 材料科学 医学 生物 有机化学 工程类 生物化学 纳米技术 物理 内科学 计算机科学 化学工程 复合材料 遗传学 基因 物理化学 催化作用 电极 光电子学 量子力学
热门帖子
关注 科研通微信公众号,转发送积分 2384351
求助须知:如何正确求助?哪些是违规求助? 2091281
关于积分的说明 5257866
捐赠科研通 1818169
什么是DOI,文献DOI怎么找? 906953
版权声明 559082
科研通“疑难数据库(出版商)”最低求助积分说明 484248