Design Strategies for Promising Organic Positive Electrodes in Lithium

Sep 29, 2017 - Organic materials have attracted considerable attention as potential positive electrodes in lithium-ion batteries owing to their high d...
0 downloads 3 Views 9MB Size
Review Cite This: Ind. Eng. Chem. Res. 2017, 56, 12009-12023

pubs.acs.org/IECR

Design Strategies for Promising Organic Positive Electrodes in Lithium-Ion Batteries: Quinones and Carbon Materials Ki Chul Kim* Department of Chemical Engineering, Konkuk University, Seoul 05029, Republic of Korea ABSTRACT: Organic materials have attracted considerable attention as potential positive electrodes in lithium-ion batteries owing to their high densities of active surface sites, which can promote fast redox reactions. Rational design strategies for developing redox-active organic materials, however, have not been established systematically. In this work, recent approaches to rational development of organic positive electrodes are comprehensively reviewed to establish design strategies for organic positive electrodes. Three primary conclusions are highlighted: (i) The ability of pristine organic materials to exhibit excellent cell performance is limited. (ii) Incorporation of appropriate functionalities into organic materials can significantly improve their redox activities. (iii) Uniform dispersion of redox-active inorganic materials in a matrix of conductive carbon materials provides nanostructured organic−inorganic composites with optimized cell performance. The strategies outlined in this review will play a critical role in the design of promising organic positive electrodes with appropriate redox activities for lithiumion batteries.

1. INTRODUCTION The demand for reduced production of environmentally harmful gases such as carbon monoxide and NOx, generated by internal combustion of fossil fuels in vehicles, has resulted in increased interest in hybrid−electric and all-electric vehicles.1−3 The development of these alternative vehicles represents a paradigm shift in electricity productionaway from burning fuels and instead toward sustainable energy resources. In recent years, there have been significant efforts to develop technologies for stable harvesting of sustainable energy resources, in particular, solar,4−7 wind,8−10 wave,11,12 and geothermal energies,13,14 owing to the importance of realizing continuous and stable supplies for automobile applications. However, the existing energy-harvesting technologies are currently too immature for effective utilization in practical systems. Thus, suitable technologies for efficient storage of sustainable energy are required to overcome the major challenge of energy supply fluctuations. In this respect, lithium-ion batteries are attractive candidates for electrochemical energy storage processes. The potential of lithium-ion batteries to store energy derived from environmentally benign energy resources in a stable manner is determined by several key parameters, namely, charge density, energy density, power density, cell voltage, cyclic stability, safety, and cost effectiveness.15−18 Despite their superior performance (i.e., high energy and charge densities) over other battery technologies, lithium-ion batteries are associated with several limitations that need to be addressed.15,16 In particular, the low power densities of conventional transition metal oxides in positive electrodes, arising from the slow diffusion of lithium ions in the bulk phase, are a major challenge that must be resolved before they can be © 2017 American Chemical Society

employed in practical applications. Thus, many studies have concentrated on various strategies (e.g., doping, introducing additives, incorporating nanostructures, and enhancing porosity) to improve the slow diffusion of lithium ions in inorganic positive electrodes.19−27 Wu and co-workers, for example, reported that doping LiFePO4 with zinc enlarged the lattice volume without destroying the lattice structure, thus providing increased space for the movement of lithium ions.19 Goodenough and co-workers developed a novel solvothermal method and combined it with high-temperature calcination to generate LiFePO4 microspheres with three-dimensional porous microstructures, which displayed excellent rate capability and cyclic stability.20 Lou and co-workers employed a facile impregnation method, followed by a simple solid-state reaction to synthesize uniform LiNi0.5Mn1.5O4 hollow microspheres and microcubes with nanosized subunits. This method provided short Li diffusion distances and effective strain relaxation through slippage of the nanosized subunits, thus leading to excellent rate capabilities and cyclic stabilities.21 Recently, efforts to identify promising high-performance positive electrode materials have been extended to metal-free organic materials with redox-active sites on their surfaces, such as conducting polymers and quinone molecules.28−42 In addition to their potential for kinetic performance, organic materials are attractive because they can reduce the costs of electrochemical energy storage by replacing conventional transition metal oxides with abundant carbon materials. Received: Revised: Accepted: Published: 12009

July 27, 2017 September 26, 2017 September 29, 2017 September 29, 2017 DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

Table 1. Performance Parameters of Quinone- and Dopamine-Derived Materials, Carbon Materials, and Nanostructured Composites of Organic Materials with Other Components for Positive Electrodes in Lithium-Ion Batteriesa

12010

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research Table 1. continued

If no value was reported for the energy density of a molecule, the density was calculated by “energy density = theoretical charge capacity × redox potential” and is reported in red.

a

Specifically, the potential of conducting polymers28−30 and redox-active quinone derivatives31−36,38−40 as organic positive electrodes in lithium-ion batteries and catholytes in redox-flow batteries has been investigated intensively. For instance, composites of conducting polymers (e.g., polypyrrole and polyaniline) coated on materials such as LiFePO4 and carbon nanotubes have been suggested as promising positive electrodes with significantly enhanced electronic conductivities and electrochemical activities.28,29 Abruña and co-workers suggested that yolk−shell structures, consisting of sulfur coated with polyaniline and incorporating an internal void space within the shell, could accommodate the volume expansion of sulfur occurring during repeated lithiation/delithiation processes in lithium−sulfur batteries.30 Yokoji et al. exploited the structural diversity afforded by introducing electron-withdrawing perfluoroalkyl groups into benzoquinone to successfully increase the voltage of benzoquinone derivatives.32 Aspuru-Guzik and co-workers employed a high-throughput computational screening approach to investigate the redox potentials of 1710 quinone derivatives with the aim of identifying promising candidates for both negative and positive sides of redox flow batteries.35 This work highlighted that functionalization at

locations close to the ketone moiety impacted the reduction potential considerably, while functionalization at positions far from the ketone moiety improved solubility. Dunn and coworkers investigated the cycling performance and redox properties of a family of naphthalene diimide derivatives with tailored functionalities as positive electrode materials for lithium-ion batteries,37 demonstrating that the redox potentials varied from 2.3 to 2.9 V vs Li/Li+ depending on the functional groups. Huan et al. employed density functional theory (DFT) calculations to study the electrochemical properties of a novel pillar[5]arene derivative, pillar[5]quinone (P5Q), belonging to a family of multicarbonyl macrocyclic compounds.38 They reported that the macrocyclic skeletons in P5QLin structures were predicted to be distorted to different extents as a result of the interactions between Li atoms and P5Q. Genorio studied the redox properties of diquinone and dihydroquinone derivatives of calix[4]arene for positive electrodes in lithiumion batteries.39 Tuntulani and co-workers also studied the electrochemical properties of calix[4]quinones, derived from double calix[4]arenes, not only for lithium-ion batteries but also for sodium- and potassium-ion batteries.40 Yao and coworkers studied two cross-conjugated quinone oligomers to 12011

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

transfer reactions, thus achieving high specific charge capacities and even exceeding the specific charge capacities of inorganic materials. The design of redox-active small organic molecules is based on various chemistries, including petrochemicals, industrial organics, and biologically active organic compounds. For instance, benzoquinones are derived from benzene, which is one of the key petrochemicals, whereas the anthraquinone framework is based on anthracene, which is commonly used in the production of dye molecules (e.g., the red dye alizarin). Naphthalene is a commonly used hydrocarbon compound that is utilized in the design of another important quinone derivative, naphthoquinone. Dopamine, which is a biological organic molecule that is synthesized in a restricted set of cell types, mainly neurons and cells in the medulla of the adrenal glands, has likewise been employed in the design of redoxactive materials. For example, Lee and co-workers have recently found that dopamine molecules can be self-polymerized after a series of cyclization and oxidation reactions.53 The dopamine polymers equipped with redox-active carbonyl groups were used to coat the surfaces of CNTs to give free-standing positive electrodes for lithium-ion batteries. In contrast, redox-active carbon materials are designed primarily from graphenes and CNTs by incorporating oxygen-containing functional groups such as carbonyls, epoxides, carboxylates, and hydroxyl groups, either on top of the planes or at the edge of the materials. Small organic molecules such as quinones and dopamine, which incorporate an electron-rich carbonyl moiety in their framework, have a reductive ability and enable reversible electron-transfer reactions as described in Scheme 1.54 In

assess their potential as positive electrodes in lithium-ion batteries.43 Despite the rapidly growing importance of organic materials for positive electrodes in secondary batteries, most studies reported to date have focused only on the performance of individual materials, and this fact has hindered the identification of well-defined strategies for the rational design of promising organic materials for secondary battery applications. Hence, it is necessary to review comprehensively the vast number of organic materials that have been examined for such applications in the literature thus far. In this work, studies focusing on a wide range of organic materials, primarily quinones and carbon materials (e.g., graphenes, carbon nanotubes (CNTs), carbon nanoparticles, mesoporous carbons) that have been employed as positive electrodes in lithium-ion batteries are comprehensively reviewed, with the aim of delineating the strategies for the design of organic materials with optimal performance in such applications. Specifically, the primary focus of this review is on evaluating various organic materials using a selected set of performance factors to identify common rules that can guide the design and fabrication of suitable positive electrodes. The Introduction is followed by a critical review of organic materials that have been developed through unique design strategies to date.

2. ORGANIC MATERIALS FOR POSITIVE ELECTRODES: BACKGROUND AND ELECTROCHEMICAL REDOX MECHANISMS The electrochemical performance of the current lithium-ion batteries based on the intercalation concept of inorganic materials is not mature enough to support large-scale systems.44−52 In contrast, organic materials are regarded as important positive electrode materials for lithium-ion batteries owing to their potential to enhance electrochemical performance. As depicted in Tables 1 and 2, these materials have exceptionally high power, charge (100−1500 Ah/kg), and energy (300−1700 Wh/kg) densities as well as redox potentials (1.0−4.5 V Li/Li+) comparable to those of inorganic positive electrodes. The organic materials can be designed to contain a high number of redox-active sites using reversible multielectron

Scheme 1. Schematic Illustration of the Reversible Multielectron Transfer Redox Reaction Mechanism Occurring in Benzoquinone via Two Steps

Table 2. Performance Parameters of Conventional Inorganic Positive Electrodes in Lithium-Ion Batteriesa molecule LiCoO2 (layered) LiNiO2 (layered) LiMnO2 (layered) LiMn2O4 (spinel) LiCo2O4 (spinel) LiFePO4 (olivine) LiMnPO4 (olivine) LiFeSO4F (tavorite)

theoretical charge capacity (Ah/kg)

energy density (Wh/kg)

power density (kW/kg)

redox potential (V vs Li/ Li+)44

27445

∼1041

3.8

27546

∼1045

3.8

47

∼941

3.3

14848

∼607

4.1

14249

∼568

4.0

17050

∼578

3.4

51

∼650

3.8

15152

∼559

3.7

285

171

addition, organic molecules with multiple carbonyls lead to multielectron transfer redox reactions, thereby resulting in increased theoretical charge capacity. In Scheme 1, an example quinone, namely benzoquinone, undergoes a two-step electrontransfer redox reaction during the discharge process. In the first step, one of the carbonyls accepts one electron and forms a free-radical anion, which binds chemically with a Li cation. Next, the other carbonyl accepts one electron, forming a molecule with divalent anions, which binds chemically another Li cation. In contrast, carbon materials such as graphenes and CNTs, which contain oxygen-containing functional groups, can reversibly transfer electrons through a combined mechanism of redox reaction (via oxygen-containing functional groups) and faradaic charge transfer (on top of honeycomb planes) during

The energy densities were calculated by “energy density = theoretical charge capacity × redox potential.”

a

12012

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

and Table 1).54 Despite their relatively low theoretical charge capacities, carbon materials, such as graphenes and CNTs with oxygen-containing functional groups, are also highly attractive as free-standing electrode materials, without the need for any additives or binding materials, owing to their superior redox potentials and high electronic conductivities. Small organic molecules such as quinones and dopamine contain high densities of redox-active Li-binding sites that enable these molecules to exhibit excellent theoretical charge capacities and reasonable discharge voltages for lithium-ion batteries. However, two main shortcomings are associated with the construction of positive electrodes from small organic molecules, namely (1) the serious cycling fading behaviors of small molecules arising as a result of the dissolution of the reduced products and (2) their insulating properties, which typically require the use of two primary additive components (conductive additives and binders). In terms of the preparation of well-designed positive electrodes without any capacity loss arising due to the presence of additives and binders, carbon materials such as grapheme oxides and CNTs are more suitable than small organic materials. However, despite their excellent potential, the relatively low theoretical charge capacities of carbon materials in comparison to those of small organic molecules represent a primary disadvantage that needs to be addressed. Thus, appropriate design strategies need to be identified and adopted to fabricate additive-free, free-standing, redox-active positive electrodes by employing conductive carbon materials or small organic molecules. In contrast to their structural diversity, the classification of design strategies for organic materials is relatively simple. In the following sections, two main strategies employed in the design of quinone- and carbon-based organic positive electrodes, namely (1) incorporation of functional groups and (2) formation of nanostructured composites of organic materials with other components, will be reviewed and examples of organic materials fabricated using these strategies will be provided. 3.1. Strategy 1: Functionalization of Organic Materials. Recently, the functionalization strategy based on the incorporation of target functional groups into organic materials has been extensively studied for its potential to deliver

charge and discharge processes. The redox reaction mechanism for the carbon materials is similar to that of the small organic molecules; however, the faradaic charge transfer mechanism involves a charge transfer process without the creation of chemical bonds.

3. STRATEGIES FOR THE DESIGN OF PROMISING ORGANIC POSITIVE ELECTRODES Organic materials are structurally versatile, ranging from small organic molecules such as quinones and dopamine, to large organic frameworks such as graphene oxides and CNTs, and their properties depend on their size and chemical structure. Among the various organic materials, small organic molecules bearing a carbonyl moiety, e.g., benzoquinone, anthraquinone, and dopamine, are regarded as some of the most important redox-active electrode materials for lithium-ion batteries owing to their relatively high redox potentials and theoretical charge capacities when compared to other organic materials such as imides and conjugated carboxylate-based electrodes (Figure 1

Figure 1. Comparison of redox potentials and theoretical charge capacities for various organic materials.

Figure 2. Chemical structures of seven quinone derivatives: 1,4-benzoquinone, 1,4-naphthoquinone, 9,10-anthraquinone, 2-aminoanthraquinone, 2,6-diaminoanthraquinone, anthraquinone-2-carboxylic acid, and anthraquinone-2,6-dicarboxylic acid. The atoms in gray, white, red, and blue depict carbon, hydrogen, oxygen, and nitrogen, respectively.31 Reprinted with permission from ref 31. Copyright 2016, American Chemical Society. 12013

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research rationally designed organic materials for utilization as positive electrodes in lithium-ion batteries.31,34,35,43,55−61 Redox-active quinones with carbonyl moieties have been suggested in a variety of studies, not only for lithium-ion batteries but also for other electrochemical energy storage systems.31,34,35,43,55−61 In fact, these studies have examined a broad range of quinone derivatives dissolved in various electrolytes as potential redoxactive components in redox flow batteries.33−36 Aspuru-Guzik and co-workers employed a high-throughput computational screening approach to study the redox potentials of a selected set of quinones and hydroquinones, including the promising classes of 9,10-anthraquinones, 1,2-benzoquinones, 2,3-naphthoquinones, and 2,3-anthraquinones, for both anolytes and catholytes in redox flow batteries.35 Assary and co-workers studied the redox properties of 50 anthraquinone derivatives to evaluate their potential as anolytes in redox flow batteries.36 They incorporated a series of electron-donating methyl and electron-withdrawing chloro groups into the anthraquinone to understand the effect of functional groups on the redox window of anthraquinone, which could be used to predict new and improved redox-active molecules. Inspired by the studies examining the applicability of quinones to redox flow batteries, computational studies were carried out by Jang and co-workers to evaluate the potential of quinones as positive electrodes in lithium-ion batteries.31 They employed a DFT modeling approach to investigate the redox properties of seven quinone derivatives, namely 1,4-benzoquinone, 1,4-naphthoquinone, 9,10-anthraquinone, 2-aminoanthraquinone, 2,6-diaminoanthraquinone, anthraquinone-2-carboxylic acid, and anthraquinone-2,6-dicarboxylic acid, which were systematically designed as positive electrode materials for lithium-ion batteries (Figure 2).31 This study highlighted the following three conclusions about the redox properties of quinone derivatives (Figure 3). First, the redox potential of a quinone derivative could be tuned by modifying its structure either geometrically or chemically. For example, decreasing the backbone length of the quinone derivative or incorporating electron-withdrawing functional groups (e.g., −COOH) improved its redox potential, whereas incorporating electron-donating amine groups reduced its redox potential. Second, the redox potential generally decreased during a discharge process because the electron affinity of a quinone derivative decreased after binding Li. Finally, the two quinone derivatives with carboxylic functional groups maintained their cathodic activities with positive redox potentials, even after binding two Li atoms, whereas the remaining five quinone derivatives lacking the −COOH functionality exhibited only positive (cathodic) redox potentials up to the point of binding one Li atom. These results indicate that these materials can store Li atoms, until all the carbonyl moieties present within them are occupied by Li. As a consequence, the carboxylic-acid-containing anthraquinones exhibited charge capacities of at least three Li atoms, and those lacking −COOH displayed charge capacities of two Li atoms. Thus, this study revealed that modification of quinone derivatives with electron-withdrawing carboxylic functional groups would be a promising approach for improving both redox properties and charge capacities. Jang and co-workers extended this theoretical work on quinones to other redox-active carbonyl compounds such as ketone derivatives of phenalenyl and anthracene as well as dopamine.53,62 Two main findings were highlighted from their DFT analysis of the ketone derivatives of phenalenyl and anthracene:62 (1) the thermodynamic stability depended

Figure 3. (a) Correlation between redox potential (as determined using the PBE0 functional) and backbone aromaticity. The number of aromatic rings in 1,4-benzoquinone, 1,4-naphthoquinone, and 9,10anthraquinone are zero, one, and two, respectively. (b) Changes in redox potentials (as determined using the PBE0 functional) as a function of the number of electron-donating (i.e., −NH2) or electronwithdrawing (i.e., −COOH)) groups incorporated in 9,10-anthraquinone. On the x axis, the number of electron-donating or electronwithdrawing groups is denoted by a negative or positive integer, respectively.31 Reprinted with permission from ref 31. Copyright 2016, American Chemical Society.

strongly on the geometric distribution of carbonyl groups, and (2) the redox potential increased with the number of incorporated carbonyl groups. In collaboration with Liu and Lee, they also developed free-standing flexible hybrid films of CNTs, spontaneously coated with polymerized dopamine chains, as positive electrodes in lithium-ion batteries (Figure 4).53 Their computational investigation revealed exceptionally high performance for the polydopamine-coated CNTs (∼133 mAh/g) when compared to that of pristine CNTs (∼40 mAh/ g) owing to the presence of highly redox-active carbonyl moieties in the polydopamine chains. As the polydopamine chains could contain either carbonyl or hydroxyl moieties (Figure 4), the analysis revealed that suppressing the hydroxyl moieties in polydopamine chains in favor of carbonyl groups within hybrid electrodes would be a suitable strategy for designing promising positive electrodes. Several studies have examined the redox properties of redoxactive carbonyl compounds for positive electrodes in lithiumion batteries via computational approaches.32,38,63−67 As a case study of short-chain carbonyl compounds, Zhao and co-workers employed DFT calculations to investigate the thermodynamic (i.e., redox potential and charge capacity) and dynamic (i.e., charge transfer rate) properties of a series of heteroatomsubstituted anthraquinone derivatives.63 They reported that two 12014

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

Figure 4. (a) Oxidative self-polymerization of dopamine in weak aqueous solution and continuous coating of polydopamine on the surface of fewwalled carbon nanotubes (FWNTs). (b) Digital image of a flexible hybrid film consisting of polydopamine-coated FWNTs.53 Reprinted with permission from ref 53. Copyright 2017, Royal Society of Chemistry.

of the anthraquinone derivatives, namely BDOZD and BDIOZD, with simultaneously high theoretical capacities (279 Ah/kg) and high working potentials (2.8 V vs Li/Li+) exhibited the largest energy densities of about 780 Wh/kg, which were 41% larger than that of anthraquinone itself. These results suggested that the functionality incorporated into the organic molecules via heteroatom substitution could improve the performance of anthraquinone derivatives as materials for positive electrodes significantly. Likewise, Yokoji et al. exploited a combination of experimental and computational techniques to examine the effect of structural diversity afforded by introduction of electron-withdrawing perfluoroalkyl groups into benzoquinone.32 This study revealed that benzoquinones modified with perfluoroalkyl groups exhibited high average discharge voltages (3.1 V vs Li/Li+). Sjödin and co-workers studied the redox properties of 16 isoindole-4,7-diones using DFT calculations64,65 and identified a set of promising organic molecule candidates in terms of the possibility of polymerization as well as redox potential and charge capacity. In addition to the reports examining short-chain carbonyl compounds, studies focusing on large-scale multicarbonyl compound families such as pillar-arene-based macrocyclic quinone derivatives and quinone-based organic crystalline materials have also been reported recently.38,66 Huan et al. employed DFT modeling to study the electrochemical properties of P5Q, belonging to a family of multicarbonyl macrocyclic compounds.38 They reported that the macrocyclic skeletons in P5QLin structures were predicted to undergo different levels of distortion as a result of the interactions between Li atoms and P5Q. The P5Q molecule was predicted to be structurally stable while storing up to 10 Li atoms. Sun et al. employed dispersion-corrected DFT calculations to investigate the electrochemical performance (redox properties and charge capacities) for 40 quinone-based organic crystalline materials with the aim of identifying appropriate organic materials for positive electrodes in lithium-ion batteries.66

Through this work, they were able to identify a set of organic materials that could exhibit high redox potentials and charge capacities. In addition to the various carbonyl-containing organic molecules, redox-active carbon materials bearing carbonyl moieties have also been suggested for lithium-ion batteries.68−70 Jang and co-workers reported that the incorporation of carbonyl groups, rather than hydroxyl groups, at the edge of graphene would be a promising strategy for designing thermodynamically stable positive electrode materials with high redox potentials.68 The scope of their study was extended further in collaboration with Liu and Lee, who studied experimentally the potential of reduced graphene oxides as positive electrodes in lithium-ion batteries.69 The effect of redox-active oxygen-containing functional groups in hydrothermally reduced graphene oxides on the redox properties was examined. They suggested that carbonyl and epoxide functional groups contributed the most to enhancing the redox potential of graphene, whereas hydroxyl groups had no impact on the redox potential. Lin and Kuo employed first-principles calculations to examine the lithiation mechanisms of functionalized graphene nanoribbons (GNRs) and the effect of various functional groups on the electrochemical performance of the graphene-based nanomaterials.70 This work revealed that only functionalization with ketone groups and their derivatives, such as pyrone and quinone, at the edge could effectively enhance Li adsorption on GNRs. They also highlighted that full lithiation of GNRs terminated with a ketone and ketone−ether pair would lead to a Li/O atomic ratio of 1.0 and 0.5, respectively, indicating that these edge-oxidized groups could effectively enhance the Li adsorption capacity of GNRs when compared to that of pristine graphite. As described above, computational studies have focused primarily on identifying promising approaches for improving the redox properties of organic materials through incorporation of carbonyl moieties. In parallel, significant experimental efforts 12015

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

Figure 5. Schematic illustration of layer-by-layer (LBL) assembly of functionalized multiwalled carbon nanotubes (MWNTs).80 Reprinted with permission from ref 80. Copyright 2011, Royal Society of Chemistry.

Figure 6. Electrochemical characteristics of LBL-MWNT electrodes in two-electrode lithium cells with 1 M LiPF6 in a mixture of ethylene carbonate and dimethyl carbonate (3:7, v/v). (a) Cyclic voltammogram data for a 0.3 mm LBL-MWNT electrode over a range of scan rates. The current at ∼3 V versus scan rate is shown in the inset. (b) Charge and discharge profiles of an electrode of 0.3 mm obtained over a wide range of gravimetric current densities between 1.5 and 4.5 V versus Li. Before each charge and discharge measurement shown in b, cells were held at 1.5 and 4.5 V for 30 min, respectively.79 Reprinted with permission from ref 79. Copyright 2010, Nature Publishing Group.

have been made toward designing surface-functionalized organic nanostructures through spray methods,71 chemical vapor deposition,72−75 contact printing,76 electrophoretic deposition,77 and layer-by-layer (LBL) assembly78,79 to improve cell performance parameters such as power density without compensating for their high energy and charge capacities. In particular, the LBL assembly technique has been extensively employed to prepare carbon-based nanostructured electrodes78−80 with the aim of improving cell performance by enhancing power, energy, and charge capacities. For example, Shao-Horn and co-workers studied the design of nanostructured electrode materials in which surface charge was introduced onto nanocarbon materials (e.g., multiwalled carbon nanotubes (MWNTs)-COO− and MWNTs-NH3+) via surface functionalization. Subsequently, these functionalized MWNTs were incorporated into assembled structures through the LBL electrostatic assembly technique (Figure 5).78−80 The resulting LBL-MWNT material, which was developed by sequential adsorption of MWNT-COO− and MWNT-NH3+ dispersions, exhibited a high gravimetric energy density (200 Wh/kg, Figure 6) with an exceptionally high power density (100 kW/kg) as a result of the faradaic reactions occurring at the surface functional groups in LBL-MWNTs. The LBL-MWNTs were extended into LBL-MWNT-RuO2 nanostructures by nanoscale conformal coating of RuO2 on the surface of LBL-MWNT.80

The resulting LBL-MWNT-RuO2 electrodes showed an enhanced rate capability that was attributable to full utilization of the pseudocapacitance of nanoscale RuO2 in addition to the double-layer capacitance of LBL-MWNTs, and faradaic reactions taking place between the surface functional groups of the LBL-MWNTs and lithium cations. Lee and co-workers studied the potential of various carbon materials incorporating oxygen-containing functional groups as positive electrodes in lithium-ion batteries.81−83 Specifically, they employed hydrothermal oxidation to prepare single-walled carbon nanotubes (SWNTs) with redox-active oxygen-containing functional groups, primarily carbonyls, epoxides, and carboxylates.81 The functionalized SWNT electrodes were observed to deliver high volumetric and gravimetric capacities of 154 Ah/L and 152 Ah/ kg, respectively (Figure 7).81 In another study, they assembled hierarchical networks of electrodes including reduced crumpled graphene oxides and functionalized few-walled carbon nanotubes (FWNTs) via a vacuum-filtration process.82 The prepared three-dimensional porous structures were observed to fully utilize the redox-active oxygen-containing functional groups, primarily carbonyls, epoxides, and carboxylates, and thus deliver high gravimetric charge capacities of up to 170 Ah/ kg with significantly enhanced rate capabilities when compared to those of conventional two-dimensional reduced graphene oxides. Further, using a hydrothermal process, they synthesized 12016

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

Figure 7. (a) Galvanostatic charge and discharge curves for a functionalized SWNT film at a concentration of 0.3 M HNO3 over 2 h (SWNT-0.3M2h). (b) Comparison of gravimetric and volumetric capacities of pristine SWNT film and SWNT-0.3M-2h at a current density of 0.05 A/g. Ratedependent (c) gravimetric and (d) volumetric capacities of the electrodes. The thicknesses of the pristine SWNT film, functionalized SWNT film at a concentration of 0.15 M HNO3 for 1 h (SWNT-0.15M-1h), and SWNT-0.3 M-2h film were 16, 10, and 15 μm, respectively.81 Reprinted with permission from ref 81. Copyright 2016, John Wiley & Sons Inc.

Figure 8. (a) Digital images of an aqueous glucose solution (3 mg/mL, left) and the colloidal dispersion of synthesized carbon spheres52 after the hydrothermal carbonization (HTC) process (right). SEM images of (b) a pristine few-walled carbon nanotube (FWNT) film and composite films consisting of (c) 40 wt % of CSs (CS-0.4) and (d) 68 wt % of CSs (CS-0.68). The digital image of the composite film CS-0.4 fabricated by vacuumfiltration is shown in the inset of c. (e) Low- and (f) high-magnification SEM images of composite film CS-0.68 after the microwave process at 1250 W for 30 s in Ar.83 Reprinted with permission from ref 83. Copyright 2016, Royal Society of Chemistry.

capacities of up to ∼155 Ah/kg, with a broad voltage range of 2.2−3.7 V vs Li/Li+, and maintained their high capacities for up to 10000 cycles. Complementing these studies, several reports have described the synthesis and characterization of multiscale multicarbonyl materials such as pillared-arene-based macrocyclic quinone

biomass-derived carbon spheres consisting of a dehydrated hydrophobic core and an oxygen-containing hydrophilic shell (Figure 8).83 Free-standing composite electrodes were fabricated by mixing the biomass-derived carbon spheres with submillimeter-long FWNTs and their redox properties were examined. The composite electrodes delivered high specific 12017

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

performance by enhancing the redox potentials and charge capacities of organic materials. (3) The presence of redoxinactive hydroxyl moieties in both small organic molecules and carbon materials bearing oxygen-containing functional groups is detrimental, and these moieties need to be removed in order to achieve promising performance as organic positive electrodes. (4) In general, the presence of other oxygen-containing functional groups such as epoxides and carboxylates in the carbon materials contributes positively to improving their redox activity. (5) The effect of nitrogen-containing functionalities in both small organic molecules and carbon materials on their cell performance in organic positive electrodes is somewhat controversial. For example, while the incorporation of amine groups into organic materials affects their redox potentials negatively, it improves moderately their charge capacities. 3.2. Strategy 2: Formation of Nanostructured Composites of Organic Materials with Other Components. The second main strategy for the development of welldesigned organic materials with optimum performance is based on combining organic materials with other components, primarily inorganic compounds, to realize nanostructured composites.97−99 One of the first reports exploiting an approach to synthesis of composite materials involved the incorporation of redox-active metal oxides such as SnO2, Fe3O4, and RuO2 into carbon matrices as a means of significantly improving redox activity through pseudocapacitance or electrochemical reactions.80,100−112 Sun and co-workers used SnO2 nanoparticle seeds on graphene layers to initiate seed-assisted SnO2 nanorod growth on graphene and subsequent carbon layer coating using the LBL assembly technique, with the aim of developing sandwiched graphene/SnO2 nanorod/carbon nanocomposites.100 The resultant nanocomposites, which consisted of active materials (i.e., metal oxides) homogeneously anchored in graphene, exhibited stably an ultrahigh reversible specific capacity of 1419 Ah/kg at the 150th cycle, with a high rate capability owing to the synergistic effect of SnO2 (pseudocapacitance) and graphene (double-layer capacitance). They suggested that the outermost carbon layer could inhibit direct contact between SnO2 and the electrolyte, thus reducing the decomposition of SnO2 nanorods. Hu et al.113 and Dong et al.114 studied the electrochemical performance of the RuO2/ activated carbon composite and MnO2/template mesoporous carbon composite electrodes, respectively, and highlighted two important conditions for optimizing the performance of the composite electrodes. Namely, (1) nanosized redox-active pseudocapacitive metal oxide particles should be used to sustain high capacitance values because pseudocapacitive reactions are confined to particle surfaces, and (2) a uniform distribution of metal oxides within the carbon matrix is required for full utilization of the redox activity of metal oxides. They also emphasized that the RuO2/carbon matrix composite showed better capacitance (>850 F/gRuO2) than the MnO2/ carbon matrix composite (∼600 F/gMnO2) owing to the relatively low electrical conductivity of MnO2.113,114 Manthiram and co-workers employed a facile microwave-based solvothermal method to synthesize nanostructured composites of Li2FeSiO4 decorated with conductive carbon materials.101 The structure of the obtained Li2FeSiO4/C nanoparticles, with an average particle size of around 20 nm, was verified to be the main cause of the enhanced ionic diffusivity and electronic conductivity as compared with those of crystalline Li2FeSiO4. Fan and co-workers fabricated binder-free positive electrodes consisting of superaligned carbon nanotubes (SACNTs) in

derivatives, cross-conjugated quinone oligomers, and quinonebased polymers.39,40,43,84−96 For example, Genorio studied the redox properties of diquinone and dihydroquinone derivatives of calix[4]arene for positive electrodes in lithium-ion batteries,39 showing that the molecules were redox-active when adsorbed on the Au(111) surfaces. Tuntulani and coworkers also studied the electrochemical properties of calix[4]quinones derived from double calix[4]arenes not only for lithium-ion batteries but also for sodium- and potassium-ion batteries.40 Yao and co-workers designed two cross-conjugated quinone oligomers, namely, poly(benzo[1,2-b:4,5-b′]dithiophene-4,8-dione-2,6-diyl) (PBDTD) and poly(benzo[1,2-b:4,5-b′]dithiophene-4,8-dione-2,6-diyl sulfide) (PBDTDS), and their nanocomposites with carbon nanotubes, to evaluate their potential as low-cost organic electrode materials in lithium-ion batteries.43 They found that both quinone derivatives showed excellent specific charge capacities of over 200 mAh/g, while PBDTD displayed a higher rate performance than PBDTDS. The differences in their rate capabilities were explained by two primary causes. First, the electron transport in PBDTD was enhanced by conjugation generated during the reduction, whereas PBDTDS was always cross-conjugated. Second, the planar conformation of PBDTD was more conductive to electron-transfer than the helical conformation of PBDTDS. Pirnat et al. employed a combination of infrared spectroscopy and nuclear magnetic resonance spectroscopy to investigate the structural properties of cross-linked polymers assembled from hydrobenzoquinone− formaldehyde and oxidized benzoquinone−formaldehyde as positive electrodes in lithium-ion batteries.84 Song et al. synthesized and characterized poly(benzoquinonyl sulfide) (PBQS) for high-energy positive electrodes in lithium- and sodium-ion batteries.85 They reported that PBQS displayed remarkably high energy densities of 734 Wh/kg for lithium-ion batteries and 557 Wh/kg for sodium-ion batteries. Yoshida and co-workers prepared pyrene-4,5,9,10-tetraone, containing two redox-active six-membered-ring-1,2-diketone units bound to polymethacrylate, for fast-charge and fast discharge lithium-ion batteries.89 In summary, the improved redox properties of carbonylcontaining organic molecules, such as quinones and dopamine, and carbon materials with oxygen-containing functional groups, namely carbonyls, carboxylates, and epoxides, can be understood by the electron inductive effect of neighboring groups. Specifically, the electron-withdrawing carbonyl moieties and oxygen-containing functional groups generate electron-deficient spots on the backbones of carbonyl-containing organic molecules and main bodies of carbon materials, which facilitate the reduction of organic materials. Subsequently, Li cations bind the electron-rich carbonyls and oxygen-containing functional groups strongly and preferentially. Overall, the studies exploiting the first targeted functionalization strategy highlight the following primary rules for the design of promising organic structures with appropriate functional groups as positive electrodes in lithium-ion batteries. (1) Pristine carbon materials such as pure graphenes and CNTs rarely exhibit excellent cell performance as a result of their limited double-layer capacitance. The limited cell performance, however, can be improved by incorporation of redox-active functionalities into these materials. (2) Carbonyl moieties in both small organic molecules, such as quinones and dopamine, and carbon materials with oxygen-containing functional groups provide a redox-active environment that can significantly improve cell 12018

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

Figure 9. Schematic illustrations of the structures of (a) a binder-free LiCoO2−SACNT electrode and (b) a conventional LiCoO2−Super P-PTFE electrode. In the binder-free LiCoO2−SACNT electrode, LiCoO2 particles are uniformly distributed in a continuous superaligned carbon nanotube (SACNT) network. In contrast, in the LiCoO2−Super P-PTFE electrode, aggregates of Super P powder are separated from LiCoO2 particles by the binder (polytetrafluoroethylene, PTFE).102 Reprinted with permission from ref 102. Copyright 2012, John Wiley & Sons Inc.

Figure 10. Rate performance of (a) LiCoO2−Super P-PTFE composite electrodes and (b) binder-free LiCoO2−SACNT composite electrodes.102 Reprinted with permission from ref 102. Copyright 2012, John Wiley & Sons Inc.

which redox-active LiCoO2 nanoparticles were embedded to form a conductive and flexible three-dimensional network (Figure 9a).102 In their study, a composite of LiCoO2 mixed with a conductive additive (super P) and polytetrafluoroethylene (PTFE) as a binder was also prepared as a reference positive electrode (Figure 9b). LiCoO2 was uniformly distributed in the highly conductive SACNT network, whereas the Super P powders, which aggregated significantly, were separated from the LiCoO2 particles. The different structural arrangement in the binder-free LiCoO2−SACNT composite made it a promising material for positive electrodes. The rate performance of binder-free LiCoO2/SACNT electrodes (137.4 mAh/g at 2 C) was considerably better than that of LiCoO2/ Super P/PTFE electrodes (62.7 mAh/g at 2 C), as shown in Figure 10. Wei and co-workers employed a sol−gel route with subsequent hydrothermal treatment to prepare core−shell Li3V2(PO4)3@C nanocomposites as positive electrode materials.103 Compared to pure Li3V2(PO4)3 electrodes, the core− shell electrodes, which consisted of redox-active Li3V2(PO4)3 core particles encapsulated by an amorphous carbon shell, exhibited enhanced discharge capacities and cyclic stabilities (127.8 mAh/g in the first cycle and 125.9 mAh/g in the 50th cycle at a current density of 28 mA/g). As a second approach for designing organic−inorganic nanostructured complexes for positive electrodes in lithiumion batteries, redox-active organic materials have been incorporated into solid matrices. As shown in Figure 11, Gaberscek and co-workers grafted a redox-active quinone derivative of calix[4]arene onto the surfaces of nanosized silica particles to create a material for positive electrodes in lithiumion batteries.115 They claimed that immobilization of redoxactive organic molecules on an insoluble solid substrate

Figure 11. A quinone derivative of calix[4]arene grafted on a silica nanoparticle and the proposed redox reaction.115 Reprinted with permission from ref 115. Copyright 2010, John Wiley & Sons Inc.

prevented dissolution of the organic molecules in organic electrolytes. Further, they emphasized that their approach represented a significant step forward toward batteries with increased ordering of constituent phases at the molecular levelin this case, self-assembled monolayers of redox active molecules on a large surface area of solid-state substrates. In summary, the studies exploiting the second strategy emphasize a common critical requirement that must be satisfied to optimize the redox activity of nanostructured organic− inorganic composite electrodes. Namely, the redox-active component in a nanostructured organic−inorganic composite needs to be uniformly distributed in the matrix of the other component in order to fully exploit the functions of both components. For instance, redox-active metal oxide nanoparticles distributed uniformly in a conductive carbon matrix are expected to exhibit enhanced cell performance (i.e., improved electronic conductivity, rate capability, and cyclic stability) as compared with that of conventional metal oxides as a result of high electronic conductivity, which is mediated by 12019

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research

(2) Dijk, M.; Orsato, R. J.; Kemp, R. The emergence of an electric mobility trajectory. Energy Policy 2013, 52, 135−145. (3) Lave, L. B.; Hendrickson, C. T.; McMichael, F. C. Environmental implications of electric cars. Science 1995, 268, 993−995. (4) Mekhilef, S.; Saidur, R.; Safari, A. A review on solar energy use in industries. Renewable Sustainable Energy Rev. 2011, 15, 1777−1790. (5) Solangi, K. H.; Islam, M. R.; Saidur, R.; Rahim, N. A.; Fayaz, H. A review on global solar energy policy. Renewable Sustainable Energy Rev. 2011, 15, 2149−2163. (6) Grätzel, M. Solar energy conversion by dye-sensitized photovoltaic cells. Inorg. Chem. 2005, 44, 6841−6851. (7) Kamat, P. V. Meeting the clean energy demand: nanostructure architectures for solar energy conversion. J. Phys. Chem. C 2007, 111, 2834−2860. (8) Joselin-Herbert, G. M.; Iniyan, S.; Sreevalsan, E.; Rajapandian, S. A review of wind energy technologies. Renewable Sustainable Energy Rev. 2007, 11, 1117−1145. (9) Ackermann, T.; Söder, L. Wind energy technology and current status: a review. Renewable Sustainable Energy Rev. 2000, 4, 315−374. (10) Saidur, R.; Islam, M. R.; Rahim, N. A.; Solangi, K. H. A review on global wind energy policy. Renewable Sustainable Energy Rev. 2010, 14, 1744−1762. (11) Langhamer, O.; Haikonen, K.; Sundberg, J. Wave powersustainable energy or environmentally costly? a review with special emphasis on linear wave energy converters. Renewable Sustainable Energy Rev. 2010, 14, 1329−1335. (12) Westwood, A. Ocean power: wave and tidal energy review. Refocus 2004, 5, 50−55. (13) Lund, J. W.; Freeston, D. H.; Boyd, T. L. Direct application of geothermal energy: 2005 worldwide review. Geothermics 2005, 34, 691−727. (14) Lund, J. W.; Freeston, D. H.; Boyd, T. L. Direct utilization of geothermal energy 2010 worldwide review. Geothermics 2011, 40, 159−180. (15) Goodenough, J. B.; Park, K.-S. The Li-ion rechargeable battery: a perspective. J. Am. Chem. Soc. 2013, 135, 1167−1176. (16) 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. (17) Marom, R.; Amalraj, S. F.; Leifer, N.; Jacob, D.; Aurbach, D. A review of advanced and practical lithium battery materials. J. Mater. Chem. 2011, 21, 9938−9954. (18) Sun, Y.; Liu, N.; Cui, Y. Promises and challenges of nanomaterials for lithium-based rechargeable batteries. Nat. Energy 2016, 1, 16071. (19) Liu, H.; Cao, Q.; Fu, L. J.; Li, C.; Wu, Y. P.; Wu, H. Q. Doping effects of zinc on LiFePO4 cathode material for lithium ion batteries. Electrochem. Commun. 2006, 8, 1553−1557. (20) Sun, C.; Rajasekhara, S.; Goodenough, J. B.; Zhou, F. Monodisperse porous LiFePO4 microspheres for a high power Liion battery cathode. J. Am. Chem. Soc. 2011, 133, 2132−2135. (21) Zhou, L.; Zhao, D.; Lou, X. LiNi0.5Mn1.5O4 hollow structures as high-performance cathodes for lithium-ion batteries. Angew. Chem. 2012, 124, 243−245. (22) Zhang, X.; Cheng, F.; Yang, J.; Chen, J. LiNi0.5Mn1.5O4 porous nanorods as high-rate and long-life cathodes for Li-ion batteries. Nano Lett. 2013, 13, 2822−2825. (23) Taberna, P. L.; Mitra, S.; Poizot, P.; Simon, P.; Tarascon, J.-M. High rate capabilities Fe3O4-based Cu nano-architectured electrodes for lithium-ion battery applications. Nat. Mater. 2006, 5, 567−573. (24) Kunduraci, M.; Al-Sharab, J. F.; Amatucci, G. G. High-power nanostructured LiMn2-xNixO4 high-voltage lithium-ion battery electrode materials: electrochemical impact of electronic conductivity and morphology. Chem. Mater. 2006, 18, 3585−3592. (25) Croguennec, L.; Palacin, M. R. Recent achievements on inorganic electrode materials for lithium-ion batteries. J. Am. Chem. Soc. 2015, 137, 3140−3156. (26) Fergus, J. W. Recent developments in cathode materials for lithium ion batteries. J. Power Sources 2010, 195, 939−954.

the conductive carbon matrix, and fast diffusion of lithium ions through surface redox reactions. In addition, the volumetric change of redox-active metal oxide nanoparticles during repeated charge−discharge processes can be buffered by the flexible carbon matrix.

4. CONCLUSIONS Highly flexible organic materials with high densities of redoxactive moieties can promote surface electrochemical reactions between redox-active moieties and lithium ions, and therefore, these materials have been regarded as promising positive electrode candidates for lithium-ion batteries. Active sites in a variety of organic materials have been extensively modified to tune their electronic structures with the aim of improving the performance of these materials as positive electrodes. However, despite these considerable efforts, common design strategies that could be applied to a broad range of organic materials and assist the development of promising organic positive electrodes have not been established, until now. In this work, studies on organic materials employed as positive electrodes in lithium-ion batteries were comprehensively reviewed to delineate optimal design strategies that could be applied to improve cell performance by enhancing redox activity, charge capacity, rate capability, and cyclic stability. A review of these studies highlights three key conclusions for the design of positive electrode materials in lithium-ion batteries. First, pristine carbon materials need to be systematically modified by introducing appropriate functionalities or nanostructured organic−inorganic composite networks in order to realize high redox activities. Second, although carbonyls (or epoxides, carboxylates, etc.) have been confirmed as promising redox-active moieties in organic materials, rigorous efforts to identify other appropriate functionalities are still needed. Third, redox-active components should be uniformly distributed in nanostructured organic−inorganic composite networks in order to exploit fully the functions of both the organic and inorganic components. Both redox-active metal oxides distributed uniformly in conductive carbon matrices and redox-active organic molecules immobilized uniformly on the surfaces of inorganic substrates represent noteworthy approaches for achieving high-performance materials. In conclusion, this review provides insight into the efforts to establish optimal design strategies for organic materials with high cell performance. A combination of the outlined strategies will likely be necessary to realize organic materials that provide a breakthrough in cell performance.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Ki Chul Kim: 0000-0002-9359-9811 Notes

The author declares no competing financial interest.

■ ■

ACKNOWLEDGMENTS This paper was supported by Konkuk University in 2017 (Grant No. 2017-A019-0122). REFERENCES

(1) Simonsson, D. Electrochemistry for a cleaner environment. Chem. Soc. Rev. 1997, 26, 181−189. 12020

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research (27) Barpanda, P.; Ati, M.; Melot, B. C.; Rousse, G.; Chotard, J.-N.; Doublet, M.-L.; Sougrati, M. T.; Corr, S. A.; Jumas, J.-C.; Tarascon, J.M. A 3.90 V iron-based fluorosulphate material for lithium-ion batteries crystallizing in the triplite structure. Nat. Mater. 2011, 10, 772−779. (28) Chen, J.; Liu, Y.; Minett, A. I.; Lynam, C.; Wang, J.; Wallace, G. G. Flexible, aligned carbon nanotube/conducting polymer electrodes for a lithium-ion battery. Chem. Mater. 2007, 19, 3595−3597. (29) Wang, G. X.; Yang, L.; Chen, Y.; Wang, J. Z.; Bewlay, S.; Liu, H. K. An investigation of polypyrrole-LiFePO4 composite cathode materials for lithium-ion batteries. Electrochim. Acta 2005, 50, 4649− 4654. (30) Zhou, W.; Yu, Y.; Chen, H.; DiSalvo, F. J.; Abruña, H. D. Yolkshell structure of polyaniline-coated sulfur for lithium-sulfur batteries. J. Am. Chem. Soc. 2013, 135, 16736−16743. (31) 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. (32) Yokoji, T.; Matsubara, H.; Satoh, M. Rechargeable organic lithium-ion batteries using electron-deficient benzoquinones as positive-electrode materials with high discharge voltages. J. Mater. Chem. A 2014, 2, 19347−19354. (33) Wang, W.; Xu, W.; Cosimbescu, L.; Choi, D.; Li, L.; Yang, Z. Anthraquinone with tailored structure for a nonaqueous metal-organic redox flow battery. Chem. Commun. 2012, 48, 6669−6671. (34) Lin, K.; Chen, Q.; Gerhardt, M. R.; Tong, L.; Kim, S. B.; Eisenach, L.; Valle, A. W.; Hardee, D.; Gordon, R. G.; Aziz, M. J.; Marshak, M. P. Alkaline quinone flow battery. Science 2015, 349, 1529−1532. (35) Er, S.; Suh, C.; Marshak, M. P.; Aspuru-Guzik, A. Computational design of molecules for an all-quinone redox flow battery. Chem. Sci. 2015, 6, 885−893. (36) Bachman, J. E.; Curtiss, L. A.; Assary, R. S. Investigation of the redox chemistry of anthraquinone derivatives using density functional theory. J. Phys. Chem. A 2014, 118, 8852−8860. (37) Vadehra, G. S.; Maloney, R. P.; Garcia-Garibay, M. A.; Dunn, B. Naphthalene diimide based materials with adjustable redox potentials: evaluation for organic lithium-ion batteries. Chem. Mater. 2014, 26, 7151−7157. (38) Huan, L.; Xie, J.; Huang, Z.; Chen, M.; Diao, G.; Zuo, T. Computational electrochemistry of pillar[5]quinone cathode material for lithium-ion batteries. Comput. Mater. Sci. 2017, 137, 233−242. (39) Genorio, B. The synthesis of diquinone and dihydroquinone derivatives of calix[4]arene and electrochemical characterization on Au(111) surface. Acta Chim. Slov. 2016, 63, 496−508. (40) Kerdpaiboon, N.; Tomapatanaget, B.; Chailapakul, O.; Tuntulani, T. Calix[4]quinones derived from double calix[4]arenes: synthesis, complexation, and electrochemical properties toward alkali metal ions. J. Org. Chem. 2005, 70, 4797−4804. (41) Xie, J.; Zhang, Q. Recent progress in rechargeable lithium batteries with organic materials as promising electrodes. J. Mater. Chem. A 2016, 4, 7091−7106. (42) Xie, J.; Wang, Z.; Gu, P.; Zhao, Y.; Xu, Z. J.; Zhang, Q. A novel quinone-based polymer electrode for high performance lithium-ion batteries. Sci. China Mater. 2016, 59, 6−11. (43) Jing, Y.; Liang, Y.; Gheytani, S.; Yao, Y. Cross-conjugated oligomeric quinones for high performance organic batteries. Nano Energy 2017, 37, 46−52. (44) Jain, A.; Ong, S. P.; Hautier, G.; Chen, W.; Richards, W. D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; Persson, K. A. Commentary: the materials project: a materials genome approach to accelerating materials innovation. APL Mater. 2013, 1, 011002. (45) Cho, J.; Kim, Y.-W.; Kim, B.; Lee, J.-G.; Park, B. A breakthrough in the safety of lithium secondary batteries by coating the cathode material with AlPO4 nanoparticles. Angew. Chem., Int. Ed. 2003, 42, 1618−1621.

(46) Ohzuku, T.; Ueda, A.; Nagayama, M. Electrochemistry and structural chemistry of LiNiO2 (R3̅m) for 4 V secondary lithium cells. J. Electrochem. Soc. 1993, 140, 1862−1870. (47) Bruce, P. G.; Armstrong, A. R.; Gitzendanner, R. L. New intercalation compounds for lithium batteries: layered LiMnO2. J. Mater. Chem. 1999, 9, 193−198. (48) Zhang, Z.-J.; Chou, S.-L.; Gu, Q.-F.; Liu, H.-K.; Li, H.-J.; Ozawa, K.; Wang, J.-Z. Enhancing the high rate capability and cycling stability of LiMn2O4 by coating of solid-state electrolyte LiNbO3. ACS Appl. Mater. Interfaces 2014, 6, 22155−22165. (49) Choi, S.; Manthiram, A. Synthesis and electrochemical properties of LiCo2O4 spinel cathodes. J. Electrochem. Soc. 2002, 149, A162−A166. (50) Yamada, A.; Chung, S. C.; Hinokuma, K. Optimized LiFePO4 for lithium battery cathodes. J. Electrochem. Soc. 2001, 148, A224− A229. (51) Choi, D.; Wang, D.; Bae, I.-T.; Xiao, J.; Nie, Z.; Wang, W.; Viswanathan, V. V.; Lee, Y. J.; Zhang, J.-G.; Graff, G. L.; Yang, Z.; Liu, J. LiMnPO4 nanoplate grown via solid-state reaction in molten hydrocarbon for Li-ion battery cathode. Nano Lett. 2010, 10, 2799− 2805. (52) Sobkowiak, A.; Roberts, M. R.; Younesi, R.; Ericsson, T.; Häggström, L.; Tai, C.-W.; Andersson, A. M.; Edström, K.; Gustafsson, T.; Björefors, F. Understanding and controlling the surface chemistry of LiFeSO4F for an enhanced cathode functionality. Chem. Mater. 2013, 25, 3020−3029. (53) Liu, T.; Kim, K. C.; Lee, B.; Chen, Z.; Noda, S.; Jang, S. S.; Lee, S. W. Self-polymerized dopamine as an organic cathode for Li- and Na-ion batteries. Energy Environ. Sci. 2017, 10, 205−215. (54) Wu, Y.; Zeng, R.; Nan, J.; Shu, D.; Qiu, Y.; Chou, S.-L. Quinone electrode materials for rechargeable lithium/sodium ion batteries. Adv. Energy Mater. 2017, 1700278. (55) 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 (24), 8336−8340. (56) Schmidt, D.; Haeupler, B.; Hager, M. D.; Schubert, U. S. Poly(DCAQI): synthesis and characterization of a new redox-active polymer. J. Polym. Sci., Part A: Polym. Chem. 2016, 54, 1998−2003. (57) Yokoji, T.; Kameyama, Y.; Sakaida, S.; Maruyama, N.; Satoh, M.; Matsubara, H. Steric effects on the cyclability of benzoquinone-type organic cathode active materials for rechargeable batteries. Chem. Lett. 2015, 44, 1726−1728. (58) Zeng, R.; Xing, L.; Qiu, Y.; Wang, Y.; Huang, W.; Li, W.; Yang, S. Polycarbonyl(quinonyl) organic compounds as cathode materials for sustainable lithium ion batteries. Electrochim. Acta 2014, 146, 447− 454. (59) Shimizu, A.; Tsujii, Y.; Kuramoto, H.; Nokami, T.; Inatomi, Y.; Hojo, N.; Yoshida, J.-i. Nitrogen-containing polycyclic quinones as cathode materials for lithium-ion batteries with increased voltage. Energy Technol. 2014, 2, 155−158. (60) 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. (61) 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. (62) Park, J. H.; Liu, T.; Kim, K. C.; Lee, S. W.; Jang, S. S. Systematic molecular design of ketone derivatives of aromatic molecules for lithium-ion batteries: first-principles DFT modeling. ChemSusChem 2017, 10, 1584−1591. (63) Yang, S.-J.; Qin, X.-Y.; He, R.; Shen, W.; Li, M.; Zhao, L.-B. A density functional theory study on the thermodynamic and dynamic properties of anthraquinone analogue cathode materials for rechargeable lithium ion batteries. Phys. Chem. Chem. Phys. 2017, 19, 12480− 12489. (64) Karlsson, C.; Jamstorp, E.; Strømme, M.; Sjödin, M.; Dincă, M. Computational electrochemistry study of 16 isoindole-4,7-diones as 12021

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research candidates for organic cathode materials. J. Phys. Chem. C 2012, 116, 3793−3801. (65) Karlsson, C.; Gogoll, A.; Strømme, M.; Sjödin, M. Investigation of the redox chemistry of isoindole-4,7-diones. J. Phys. Chem. C 2013, 117, 894−901. (66) Sun, S.; Chen, Y.; Yu, J. High throughput screening of organic electrode materials for lithium battery by theoretical method. J. Phys. Chem. C 2015, 119, 25770−25777. (67) Shimizu, A.; Kuramoto, H.; Tsujii, Y.; Nokami, T.; Inatomi, Y.; Hojo, N.; Suzuki, H.; Yoshida, J.-i. Introduction of two lithiooxycarbonyl groups enhances cyclability of lithium batteries with organic cathode materials. J. Power Sources 2014, 260, 211−217. (68) Kim, S.; Kim, K. C.; Lee, S. W.; Jang, S. S. Thermodynamic and redox properties of graphene oxides for lithium-ion battery applications: a first principles density functional theory modeling approach. Phys. Chem. Chem. Phys. 2016, 18, 20600−20606. (69) Liu, T.; Kim, K. C.; Kavian, R.; Jang, S. S.; Lee, S. W. Highdensity lithium-ion energy storage utilizing the surface redox reactions in folded graphene films. Chem. Mater. 2015, 27, 3291−3298. (70) Lin, K.-H.; Kuo, C.-L. Lithiation mechanisms and lithium storage capacity of reduced graphene oxide nanoribbons: a firstprinciples study. J. Mater. Chem. A 2017, 5, 4912−4922. (71) Kaempgen, M.; Chan, C. K.; Ma, J.; Cui, Y.; Gruner, G. Printable thin film supercapacitors using single-walled carbon nanotubes. Nano Lett. 2009, 9, 1872−1876. (72) Futaba, D. N.; Hata, K.; Yamada, T.; Hiraoka, T.; Hayamizu, Y.; Kakudate, Y.; Tanaike, O.; Hatori, H.; Yumura, M.; Iijima, S. Shapeengineerable and highly densely packed single-walled carbon nanotubes and their application as super-capacitor electrodes. Nat. Mater. 2006, 5, 987−994. (73) Zhang, H.; Cao, G.; Yang, Y. Carbon nanotube arrays and their composites for electrochemical capacitors and lithium-ion batteries. Energy Environ. Sci. 2009, 2, 932−943. (74) Pushparaj, V. L.; Shaijumon, M. M.; Kumar, A.; Murugesan, S.; Ci, L.; Vajtai, R.; Linhardt, R. J.; Nalamasu, O.; Ajayan, P. M. Flexible energy storage devices based on nanocomposite paper. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 13574−13577. (75) Chmiola, J.; Largeot, C.; Taberna, P.-L.; Simon, P.; Gogotsi, Y. Monolithic carbide-derived carbon films for micro-supercapacitors. Science 2010, 328, 480−483. (76) Izadi-Najafabadi, A.; Yasuda, S.; Kobashi, K.; Yamada, T.; Futaba, D. N.; Hatori, H.; Yumura, M.; Iijima, S.; Hata, K. Extracting the full potential of single-walled carbon nanotubes as durable supercapacitor electrodes operable at 4 V with high power and energy density. Adv. Mater. 2010, 22, E235−E241. (77) Pech, D.; Brunet, M.; Durou, H.; Huang, P.; Mochalin, V.; Gogotsi, Y.; Taberna, P.-L.; Simon, P. Ultrahigh-power micrometresized supercapacitors based on onion-like carbon. Nat. Nanotechnol. 2010, 5, 651−654. (78) Lee, S. W.; Kim, B.-S.; Chen, S.; Shao-Horn, Y.; Hammond, P. T. Layer-by-layer assembly of all carbon nanotube ultrathin films for electrochemical applications. J. Am. Chem. Soc. 2009, 131, 671−679. (79) Lee, S. W.; Yabuuchi, N.; Gallant, B. M.; Chen, S.; Kim, B.-S.; Hammond, P. T.; Shao-Horn, Y. High-power lithium batteries from functionalized carbon-nanotube electrodes. Nat. Nanotechnol. 2010, 5, 531−537. (80) Lee, S. W.; Gallant, B. M.; Byon, H. R.; Hammond, P. T.; ShaoHorn, Y. Nanostructured carbon-based electrodes: bridging the gap between thin-film lithium-ion batteries and electrochemical capacitors. Energy Environ. Sci. 2011, 4, 1972−1985. (81) Liu, T.; Davijani, A. A. B.; Sun, J.; Chen, S.; Kumar, S.; Lee, S. W. Hydrothermally oxidized single-walled carbon nanotube networks for high volumetric electrochemical energy storage. Small 2016, 12, 3423−3431. (82) Lee, B.; Lee, C.; Liu, T.; Eom, K.; Chen, Z.; Noda, S.; Fuller, T. F.; Jang, H. D.; Lee, S. W. Hierarchical networks of redox-active reduced crumpled graphene oxide and functionalized few-walled carbon nanotubes for rapid electrochemical energy storage. Nanoscale 2016, 8, 12330−12338.

(83) Liu, T.; Kavian, R.; Chen, Z.; Cruz, S. S.; Noda, S.; Lee, S. W. Biomass-derived carbonaceous positive electrodes for sustainable lithium-ion storage. Nanoscale 2016, 8, 3671−3677. (84) Pirnat, K.; Mali, G.; Gaberscek, M.; Dominko, R. Quinoneformaldehyde polymer as an active material in Li-ion batteries. J. Power Sources 2016, 315, 169−178. (85) Song, Z.; Qian, Y.; Zhang, T.; Otani, M.; Zhou, H. Poly(benzoquinonyl sulfide) as a high-energy organic cathode for rechargeable Li and Na batteries. Adv. Sci. 2015, 2, 1500124. (86) Song, Z.; Zhan, H.; Zhou, Y. Anthraquinone based polymer as high performance cathode material for rechargeable lithium batteries. Chem. Commun. 2009, 448−450. (87) Kassam, A.; Burnell, D. J.; Dahn, J. R. Lithiated 1,4,5,8naphthalenetetraol formaldehyde polymer, an organic cathode material. Electrochem. Solid-State Lett. 2011, 14, A22−A23. (88) Zhao, L.; Wang, W.; Wang, A.; Yuan, K.; Chen, S.; Yang, Y. A novel polyquinone cathode material for rechargeable lithium batteries. J. Power Sources 2013, 233, 23−27. (89) Matsuo, T. N. T.; Inatomi, Y.; Hojo, N.; Tsukagoshi, T.; Yoshizawa, H.; Shimizu, A.; Kuramoto, H.; Komae, K.; Tsuyama, H.; Yoshida, J.-i.; Nokami, T. 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. (90) 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. (91) 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. (92) Xiang, J.; Chang, C.; Li, M.; Wu, S.; Yuan, L.; Sun, J. A novel coordination polymer as positive electrode material for lithium ion battery. Cryst. Growth Des. 2008, 8, 280−282. (93) Barrès, A.-L.; Geng, J.; Bonnard, G.; Renault, S.; Gottis, S.; Frayret, O. M. C.; Dolhem, F.; Poizot, P.; Mentre, O. High-potential reversible Li deintercalation in a substituted tetrahydroxy-p-benzoquinone dilithium salt: an experimental and theoretical study. Chem. - Eur. J. 2012, 18, 8800−8812. (94) 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. (95) Hanyu, Y.; Ganbe, Y.; Honma, I. Application of quinonic cathode compounds for quasi-solid lithium batteries. J. Power Sources 2013, 221, 186−190. (96) Song, Z.; Qian, Y.; Gordin, M. L.; Tang, D.; Xu, T.; Otani, M.; Zhan, H.; Zhou, H.; Wang, D. Polyanthraquinone as a reliable organic electrode for stable and fast lithium storage. Angew. Chem., Int. Ed. 2015, 54, 13947−13951. (97) 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. (98) Xie, J.; Zhao, C.-E.; Lin, Z.-Q.; Gu, P.-Y.; Zhang, Q. Nanostructured conjugated polymers for energy-related applications beyond solar cells. Chem. - Asian J. 2016, 11, 1489−1511. (99) Xie, J.; Gu, P.; Zhang, Q. Nanostructured conjugated polymers: toward high-performance organic electrodes for rechargeable batteries. ACS Energy Lett. 2017, 2, 1985−1996. (100) Wang, D.; Yang, J.; Li, X.; Geng, D.; Li, R.; Cai, M.; Sham, T.K.; Sun, X. Layer by layer assembly of sandwiched graphene/SnO2 nanorod/carbon nanostructures with ultrahigh lithium ion storage properties. Energy Environ. Sci. 2013, 6, 2900−2906. (101) Muraliganth, T.; Stroukoff, K. R.; Manthiram, A. Microwavesolvothermal synthesis of nanostructured Li2MSiO4/C (M = Mn and Fe) cathodes for lithium-ion batteries. Chem. Mater. 2010, 22, 5754− 5761. (102) Luo, S.; Wang, K.; Wang, J.; Jiang, K.; Li, Q.; Fan, S. Binderfree LiCoO2/carbon nanotube cathodes for high-performance lithium ion batteries. Adv. Mater. 2012, 24, 2294−2298. 12022

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023

Review

Industrial & Engineering Chemistry Research (103) Ren, M. M.; Zhou, Z.; Gao, X. P.; Peng, W. X.; Wei, J. P. Coreshell Li3V2(PO4)3@C composites as cathode materials for lithiumion batteries. J. Phys. Chem. C 2008, 112, 5689−5693. (104) Wang, G.; Liu, H.; Liu, J.; Qiao, S.; Lu, G. M.; Munroe, P.; Ahn, H. Mesoporous LiFePO4/C nanocomposite cathode materials for high power lithium ion batteries with superior performance. Adv. Mater. 2010, 22, 4944−4948. (105) Li, Y.-X.; Gong, Z.-L.; Yang, Y. Synthesis and characterization of Li2MnSiO4/C nanocomposite cathode material for lithium ion batteries. J. Power Sources 2007, 174, 528−532. (106) Zhang, X.-F.; Wang, K.-X.; Wei, X.; Chen, J.-S. Carbon-coated V2O5 nanocrystals as high performance cathode material for lithium ion batteries. Chem. Mater. 2011, 23, 5290−5292. (107) Wang, J.; Sun, X. Understanding and recent development of carbon coating on LiFePO4 cathode materials for lithium-ion batteries. Energy Environ. Sci. 2012, 5, 5163−5185. (108) Rao, C. V.; Reddy, A. L. M.; Ishikawa, Y.; Ajayan, P. M. LiNi1/ 3Co1/3Mn1/3O2−graphene composite as a promising cathode for lithium-ion batteries. ACS Appl. Mater. Interfaces 2011, 3, 2966−2972. (109) Doherty, C. M.; Caruso, R. A.; Smarsly, B. M.; Adelhelm, P.; Drummond, C. J. Hierarchically porous monolithic LiFePO4/carbon composite electrode materials for high power lithium ion batteries. Chem. Mater. 2009, 21, 5300−5306. (110) Vu, A.; Stein, A. Multiconstituent synthesis of LiFePO4/C composites with hierarchical porosity as cathode materials for lithium ion batteries. Chem. Mater. 2011, 23, 3237−3245. (111) Feng, C. Q.; Chew, S. Y.; Guo, Z. P.; Wang, J. Z.; Liu, H. K. An investigation of polypyrrole−LiV3O8 composite cathode materials for lithium-ion batteries. J. Power Sources 2007, 174, 1095−1099. (112) Chen, Q.; Qiao, X.; Wang, Y.; Zhang, T.; Peng, C.; Yin, W.; Liu, L. Electrochemical performance of Li3-xNaxV2(PO4)3/C composite cathode materials for lithium ion batteries. J. Power Sources 2012, 201, 267−273. (113) Hu, C.-C.; Chen, W.-C.; Chang, K.-H. How to achieve maximum utilization of hydrous ruthenium oxide for supercapacitors. J. Electrochem. Soc. 2004, 151, A281−A290. (114) Dong, X.; Shen, W.; Gu, J.; Xiong, L.; Zhu, Y.; Li, H.; Shi, J. MnO2-embedded-in-mesoporous-carbon-wall structure for use as electrochemical capacitors. J. Phys. Chem. B 2006, 110, 6015−6019. (115) 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.

12023

DOI: 10.1021/acs.iecr.7b03109 Ind. Eng. Chem. Res. 2017, 56, 12009−12023