Realizing High Voltage Lithium Cobalt Oxide in Lithium-Ion Batteries

May 27, 2019 - Since the high potential or capacity LCO batteries have a lot of benefits, many ... Nontransition ion dopants used in LCO include Mg, A...
1 downloads 0 Views 3MB Size
Review Cite This: Ind. Eng. Chem. Res. 2019, 58, 10119−10139

pubs.acs.org/IECR

Realizing High Voltage Lithium Cobalt Oxide in Lithium-Ion Batteries Xiao Wang,† Xinyang Wang,† and Yingying Lu*

Downloaded via NOTTINGHAM TRENT UNIV on July 17, 2019 at 13:58:32 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

State Key Laboratory of Chemical Engineering, Institute of Pharmaceutical Engineering, College of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310027, China ABSTRACT: The combination of high voltage cathode and metal or graphite anodes provides a feasible way for future high-energy batteries. Among various battery cathodes, lithium cobalt oxide is outstanding for its excellent cycling performance, high specific capacity, and high working voltage and has achieved great success in the field of consumer electronics in the past decades. Recently, demands for smarter, lighter, and longer standby-time electronic devices have pushed lithium cobalt oxide-based batteries to their limits. To obtain high voltage batteries, various methods have been adopted to lift the cutoff voltage of the batteries above 4.45 V (vs Li/Li+). This review summarizes the mechanism of capacity decay of lithium cobalt oxide during cycling. Various modifications to achieve high voltage lithium cobalt oxide, including coating and doping, are also presented. We also extend the discussion of popular modification methods for electrolytes including electrolyte additives, quasi-solid electrolytes, and electrode/ electrolyte interfaces.



INTRODUCTION Lithium-ion batteries have attracted wide attention since the 1970s. In the research field of lithium-ion batteries, the cathode is a key component since its properties can largely influence the energy density and lifetime of the battery. To obtain high energy density batteries, various cathode materials have been studied, such as lithium cobalt oxide (LCO), lithium iron phosphate, manganese-based cathode, and so on.1−10 Figure 1 shows that different cathode materials have different volumetric energy densities with different anodes, among which LCO shows higher a volumetric energy density after the Li-rich one. In the field of consumer electronics products, because of the pursuit of lighter and thinner products, there is an increasing demand for higher volume energy density and the working voltage of batteries. To meet these requirements, it is necessary to make a proper selection for cathode materials with higher compaction density and a satisfactory working voltage range. Among several cathode materials, LCO has good compaction density (4.1−4.3 g cm−3) and a high working voltage range (3.7−3.9 V). To make use of the capacity of LCO, researchers have studied it for decades, and the cutting voltage has been lifted from 4.2 to 4.45 V, with the volumetric energy density of LCO-based batteries improved from 200 to 700 Wh/L. Although lithium-rich materials have higher capacity and voltage, they have serious oxygen evolution and irreversible voltage and capacity attenuation after the first cycle, making it difficult to be widely applied in the market. Therefore, in the foreseeable future, LCO cathodes are still the mainstream of cathode materials for Li-ion batteries for consumer electronic products. © 2019 American Chemical Society

A nickel−cobalt−manganese ternary material (NCM), more likely to be used in electrical vehicles because of its cost effectiveness, is a competing cathode material compared with LCO. It has different ratios, such as NCM111, NCM523, and NCM811. Although it has a good capacity density (3.6−3.8 g cm−3), the ternary material is not comparable to LCO in compaction density. This means that ternary materials have unsatisfactory volume energy density. In addition, the performance of ternary materials on the working voltage is not as high as that of LCO. Although ternary materials with high nickel ratios deliver high capacity density, the cycling performance of ternary materials is relatively poor due to the cationic rearrangement; thus, they are not suitable for the 3C market (computers, communications, and consumer electronics) with high requirements for battery life.11−13 Therefore, in general, LCO as a high voltage cathode material still has its advantages. The cutoff voltage of the current LCO electrode in practical use is only 4.2 V; that is, 50% of lithium ions are extracted. This is because when lithium ions in LCO are deintercalated more than 50%, the structure of LCO will be unstable, and capacity fading occurs. At the same time, at high voltage, a series of side reactions will occur at the interfaces between LCO particles and the electrolyte, resulting in irreversible element dissolution. LCO can be stabilized by modifications of the structure of LCO particles, such as surface modification or element doping. Received: Revised: Accepted: Published: 10119

March 5, 2019 May 23, 2019 May 27, 2019 May 27, 2019 DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 1. Calculated volumetric energy densities of lithium batteries.10 Reproduced from ref 10 with permission from the Royal Society of Chemistry. Copyright 2018.

Figure 2. Variation of lattice c (a) and a (b) during electrochemical lithium extraction of LiCoO2.19 Reproduced from ref 19 with permission from the The Electrochemical Society, copyright 1996.

lization of LCO in the space group R3̅m with cell constants a = 2.816 Å and c = 14.08 Å. With the deintercalation of Li+ ions from bulk LCO, the phase transition occurs from H1 to H2, M1, H3, M2, and O1 phases with different amounts of extraction of Li+ ions shown in Figure 2a and b. In terms of Figure 2, it is obvious that there exists a reduction of the c lattice around 50% of the Li+ extraction, which happens when overcharging above 4.2 V. Most groups2,7,11,16−38 believe that the decreased c lattice leads to capacity fading because of the changes in the c lattice, but almost no changes of the a lattice will induce nonuniform strain in particles and cause particle fractures. Some researchers thought that the ultrahigh voltage cannot be used with cycling cutoff voltage higher than 4.6 V. The transition from H3 to H1-3 near 4.55 V can greatly change the oxygen layer arrangement. Other than the structure changes during phase transition, there also exists Co dissolution which greatly lead to capacity fading. Amatucci et al.20 tested the Co dissolution, and they found that the Co dissolution was detected above 4.2 V. Chebiam et al.25 attributed it to the loss of oxygen from the lattice at deep lithium extraction. After studying the LCO structure at high voltage operation, researchers believe that the unstable structure and the losses of active elements will lead to the

Electrolytes are another bottleneck for practical application of high voltage LCO (HVLCO). Conventional nonaqueous electrolytes, which mainly refer to the carbonate solvent-based electrolytes, exhibit inferior anodic stability for high-voltage operation above 4.3 V vs Li/Li+. Developing solid electrolytes toward an all-solid-state battery is considered as the ultimate solution, whereas they still suffer from challenges such as low ionic conductivity, poor interfacial contact between electrolyte and electrode, and oversensitivity, which indicate that there is still a long way to go. Here, we highlight the recent progress in the use of surface modification, element doping, and electrolyte additives for HVLCO and quasi-solid electrolytes as appropriate alternatives at this time.



CHALLENGES FOR LCO LCO was first discovered as a viable layered rock-salt cathode material by Goodenough in 1980.14,15 After that, Sony Corporation applied it into the portable device in 1991. The structure of LCO is isostructural with α-NaFeO2. In the structure, Li+ and Co3+ ions are ordered on alternating (111) planes. This (111) ordering introduces a distortion of the lattice to hexagonal symmetry, which leads to the crystal10120

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 3. (a), (b) Structure and electrical properties of LiCo1−xMgxO2. (c), (d) Calculated CoO2 slab and interslab distances from Rietveld refinement of LiCo1−xMgxO2.44 (e), (f) Initial charge capacity and cycling stability of LiCo1−xMgxO2.45 Reproduced from refs 44 and 45 with permission from Elsevier and Elsevier, copyrights 2006 and 2014, respectively.

and use additives to inhibit the side reactions in LCO batteries, especially at high voltages, is worth studying. Researchers have found a lot of approaches to solve the problems when LCO is cycled above 4.2 V such as doping and coating. Doping can effectively change the structure of LCO and stabilize it. Coating can stabilize the electrolyte/electrode interface and prevent the side reactions between LCO and other battery components. All the modification methods are extensively discussed below.

incompatibility between host LCO and other battery component (such as electrolyte, binders, conductive additives, etc.), which deteriorates the electrochemical performance and causes safety issues at high voltages. Besides the intrinsic instability of LCO at high voltage, the interface stability between LCO and other battery component is another challenge. It is believed32 that the μc of the cathode should be higher than the highest occupied molecular orbital of the electrolyte unless the electrolyte will be oxidized. Thus, it is important to design a proper electrolyte to adjust to the high voltage. LiPF6, which is widely used as an electrolyte salt, has low thermal stability and is sensitive to hydrolysis. It will decompose and cause thermal runaway from side reactions. Cycling on high voltage will exacerbate the side reactions between electrolyte and cathode. So, how to design electrolytes



DOPING Doping is an effective way to improve the capacity and electrochemical stability window of LCO. Different element dopings have different effects. For example, transition metal ions can increase the capacity, while nontransition metal ions can increase the delithiation potential.2 Element doping can 10121

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 4. (a), (b) Initial capacity and cycling performance of LiCuxCo1−xO2.60 (c), (d) Zn-doped LCO (mol %).61 (e), (f) Ni-doped LCO.64 Reproduced from refs 60, 61, and 64 with permission from Elsevier, Springer, and Elsevier, copyrights 2011, 2012, and 2014, respectively.

is regarded as a promising substituent for Co in LCO. Tukamoto et al.39 found that Mg-doped LCO has a better electronic conductivity without an impurity phase. Though the capacity of Mg-doped LCO is slightly lower, the stability of cycling at 4.3 V can be achieved. Zou et al.40 synthesized LiMg0.05Co0.95O2 which had a charge/discharge capacity of around 160 mAh g−1 at a 0.2 C rate over 50 cycles. This is the first time that the stable capacity was obtained from doped LCO at a high voltage (4.5 V). Sathiyamoorthi et al.41studied the different amounts of Mg in LiCo1−xMgxO2 and found that LiMg0.2Co0.8O2 has the highest discharge capacity of 191.3

change the physical properties of the material’s crystal lattice. Since the high potential or capacity LCO batteries have a lot of benefits, many research groups have investigated doping effects through different nontransition ions, such as Mg, Ca, Al, and B, and transition metal ions, such as Ni, Zn, Cu, and Rh. Some others developed mixed element doping, like Ni + Mn, Li + Mn, Mn + Mg, Mg/Al/Zr + F, and La + Al. Nontransition Elements. Nontransition ion dopants used in LCO include Mg, Al, Ca, and B. The ionic radii of the ionic elements are 0.72 Å (rMg2+), 1.00 Å (rCa2+), 0.535 Å (rAl3+), and 0.27 Å(rB3+). Mg is a nontoxic, light, and cheap element, which 10122

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research mAh g−1 at a voltage of 4.5 V. Mladenov et al.42 found that the replacement of Co by Mg contributes to the enhancement of capacity retention, while the Mg dopants in the LiO2 layers has no effect. Since Mg-doped LCO exhibits better electrochemical performance than bare LCO, many groups are trying to find how Mg2+ leads to enhanced electrical performance of LCO. Tukamoto et al.39 thought this might be because of the presence of a small amount of Co4+. LCO exhibits p-type semiconductivity. They believed that the undoped LCO is Li deficient with a small concentration of Li vacancies and equal number of Co4+. The low concentration of Mg will lead to occupation of Li vacancies by Mg, which causes a reduced concentration of Co4+ and a decrease in conductivity. The high concentration of Mg will substitute the Co sites, which can contribute to the creation of Co4+ for charge balance. Levasseur et al.43 agreed with this, and suggested that intermediate spin Co3+(IS) ions are trapped in a square-based pyramidal environment due to the presence of oxygen vacancies. Kim et al.44 analyzed the crystal structure of LiCo1−xMgxO2 with X-ray diffraction and Rietveld refinement, shown in Figure 3. They found that the material displays a single phase when x ≤ 0.11, and second phase of MgO will occur when x ≥ 0.13. The higher the Mg concentration is, the larger the distance between CoO2 slabs is. Besides, they believed that the increased Mg content will increase the carrier concentration, while the carrier mobility is nearly unchanged, which leads to enhancement of electrical conductivity (Figure 3 a−d). The electrochemical performance of Mg-doped LCO is shown in Figure 3e and f.45 Ca is another popular element used as a dopant for LCO. Sathiyamoorthi et al.46,47 synthesized Ca-doped LCO materials by a low temperature solid state method. The lattice parameters a and c increase as the Ca content increases, which results from the bigger radius of Ca. The capacity of the doped LCO reached 178.5 mAh g−1 after 25 cycles at 4.5 V. Other than the divalent nontransition elements mentioned above, the trivalent nontransition elements were used as dopants as well, such as Al and B. Ceder et al.2 suggested that the Al substitution can increase the intercalation potential but decrease the capacity. This was later proved by several other groups.48 Huang et al.48 synthesized the solid solution LiAlxCo1−xO2 and found that when x = 0.15, the initial capacity can reach 160 mAh g−1, and no major changes could be observed through XRD. Jang et al.49 synthesized LiAl0.25Co0.75O2 and showed that the doped LCO had a better performance when cycled at 55 °C. They also found that the c/a ratio increases with the Al content, but the volume of the cell remains unchanged. Yoon et al.50 concluded that the Al substitution for Co leads to larger local structure distortion, and the capacity fading is associated with the existing Co4+ and the structure distortion. Also, they found that the temperature of LiAl0.25Co0.75O2 synthesis has an influence on the local structure distortion during cycling.51 Myung et al.52 studied relationships among the Al content, Co dissolution, capacity, and variation in c-axis and concluded that the capacity will decrease when Al replaces Co, but the cycling stability can be effectively enhanced. Yoon et al.53 manufactured Al-doped LCO by the sol−gel process, and the DSC result showed that the Al dopants can enhance the thermal stability of bulk LCO and reduce the heat generation when reacting with electrolytes, which was later proved by Zhou et al.54 as well. Yoon et al.55 concluded that

the substitution of Al for Co will increase the O2 participation in the charge compensation process during charging. Besides Al, B was used as an effective trivalent dopant as well. Alcantara et al.56 suggested that boron dopants can improve the reversibility of the material and contribute to the adaption of the lattice to lithium order−disorder in the depleted LiO2 layers, preventing the onset of the structural first-order transition related to the Verwey transition in Li0.5CoO2.57 Julien et al.58 realized that boron-doped LCO provides low polarization, but the capacity is not satisfying (only 130 mAh g−1 at 4.3 V). Transition Metal Elements. Transition metal elements used as LCO dopants include Ni, Zn, Cu, and Rh. The radii of the ionic elements are 0.74 Å (rZn2+), 0.73 Å (rCu2+), 0.56 Å (rNi3+), and 0.665 Å(rRh3+). Zou et al.59 synthesized Cu-, Zn-, and Mn-doped LCO and showed that the Cu- and Zn-doped LCO provided great initial capacity of around 165−170 mAh g−1 in the range of 3.5−4.5 V but faded slowly and retained ∼130 and ∼110 mAh g−1, respectively. Mn-doped LCO showed better capacity stability, which delivered ∼158 mAh g−1 after 50 cycles at 4.5 V. Nithya et al.60 used the microwave method to synthesize Cu-doped LCO and found that LiCu0.2Co0.8O1.9 delivered a capacity of around 150 mAh g−1 in the range of 2.7−4.6 V (Figure 4a, b). Valanarasu et al.61 found that the 3 mol % Zn-doped LCO delivered about 178 mAh g−1 after 30 cycles between 3.0−4.5 V (Figure 4c, d). The trivalent ions used as dopants can be Ni3+, Rh3+, and Fe3+. Madhavi et al.62 found that Rh can suppress the phase transition when 50% Li deintercalates and the voltage of the deintercalation decreases. Reddy et al.63 studied two kinds of LCO substituted by Ni, LiNi0.3Co0.7O2 and LiNi0.7Co0.3O2, and showed that the latter delivered a better capacity of 165 mAh g−1 after 60 cycles at 4.3 V. Liang et al.64 found that Ni located in Li sites can support the layered structure, which allows LCO to intercalate/deintercalate more Li ions (Figure 4e, f). Mixed Elements. Although single element doping can improve the electrochemical performance of LCO to some extent, its effectiveness is not comprehensive and is only beneficial for one or two aspects, such as the following: Mg mainly improves the conductivity of lithium cobalt oxide. Al can play a pillar effect to stabilize the material structure. Ni can make the material release more lithium ions at a specific voltage. To improve the electrochemical properties of materials more comprehensively, researchers tried to dope lithium cobalt oxide with two different elements. Compared with single element doping, binary doping can improve more properties of the materials, which is a good reference for future research on high voltage LCO. Ni + Mn. Researchers tried to take Mn and Ni as codopants because LiNiO2 and LiMnO2 are isostructural with LCO. The redox of Ni2+/Ni4+ is lower than that of Co3+/Co4+, which can improve the capacity of LCO, and Mn4+ can show pillar effects to stabilize the structure. Lu et al.65 reported the electrochemical performance and thermal stability of LiNixCo1−2xMnxO2 for the first time. When x = 0.25 or 0.375, the materials delivered 160 mAh g−1 at 4.4 V and showed better thermal stability. Ohzuku et al.66 synthesized LiNi1/3Co1/3Mn1/3O2 (NMC111), which delivered 150 mAh g−1 at 3.5−4.2 V or 200 mAh g−1 at 3.5−5.0 V. Shaju67 and Hwang68 studied the structure change of NMC111 and found that the redox processes around 3.8 and 4.6 V were ascribed to Ni2+/4+ and Co3+/4+, respectively. Kim et al.69 found that the great cycling stability of NMC111 can be attributed to no 10123

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 5. (a), (b) Structure model and cycling performance of LiNi1/3Mn1/3Co1/3O2 codoped LCO73,87 and (c), (d) xLi2MnO3·(1−x)LiCoO2.88,89 (e), (f) Electrochemical properties of La + Al codoped LCO. .86 Reproduced from refs 73, 87, 88, 89, and 86 with permission from Elsevier, Elsevier, Elsevier, The Royal Society of Chemistry, and Nature, copyrights 2004, 2003, 2015, 2012, and 2018, respectively.

rate charge and discharge, which results from the severe structural inhomogeneity and instability of NMC811. They discovered that the structural degradation is due to the reduction of the Ni ions and the consequent decrease in Li ions conductivity. These problems lead to capacity fading as well. Kim et al.78 designed and synthesized a stable Ni-rich cathode by a polymeric bead-assisted coprecipitation method. The prepared material showed a transition metal gradient with reduced nickel ions and internal pores in the secondary particles, which stabilizes the structure and accommodates the strain generated during cycling. Li + Mn. Other than Ni and Mn, researchers also studied Li and Mn as codopants. Because Li2MnO3 is isostructural with LCO, codoping is easy with the two elements.79 Numata et al.80,81 synthesized LiCo1−xLix/3Mn2x/3O2 and showed that the Co3+ ions located near Mn4+ require a higher oxidation voltage, which decreases the capacity. Thackeray et al.79 discovered that the xLiMnO3·(1−x)LiCoO2 can be activated by Li2O and Li from the structure and delivered a capacity of more than 250 mAh g−1 at high voltage. Rana et al.82 studied 0.5Li2MnO3· 0.5LiCoO2 and found that the Li2MnO3 and LiCoO2 exit as separate domains. The randomly distributed Li2MnO3 acts as an excess lithium source and leads to higher capacity and an upper operating voltage. Sun et al.83 discovered that the discharge capacity decreases with an increasing Li2MnO3 content, and the layered structure can remain unchanged during the charge process, which enhances cycling performance. After studying the electrochemical performance of the Li + Mn-codoped LCO, further studies regarding the structure

structural variation and volume change during cycling, and the cation mixing effect was greatly undermined low. Li et al.70 tested two different methods to synthesize NMC111 and found that the sample prepared by the spray dry method exhibited better performance. Other methods to synthesize NMC111 include the ultrasonic spray pyrolysis method,71 solid-state reaction of metal oxide and lithium hydroxide,72 and coprecipitation.73 The electrochemical performance and structure are shown in Figure 5a and b. Besides NMC111, researchers studied different proportions of Ni, Mn, and Co as well. Lee et al.74 synthesized material with a concentration gradient of metal elements inside particles. The electrochemical performance demonstrated continuous cycling for over 2500 cycles with only 16.7% capacity fading, but the capacity was not high, which suggests that the improved capacity can be reached through optimizing the concentration gradient to reduce the lattice mismatch. Lin et al.75 prepared LiNi0.4Mn0.4Co0.2O2 by the spray pyrolysis method. The prepared material exhibited elements segregationNi-poor and Mn-rich surfaces show great resistance against surface reconstructionwhich contributes to the superior electrochemical performance. Shen et al.76 investigated LiNi0.5Co0.2Mn0.3O2 and detected the formation of NiMn2O4 and LiOH after being charged at 4.5 V. The lithium ion mobility is limited, and the vulnerable surface film leads to the cracks on the surface. Also, stable NiMn2O4 generates after cycling, causing structural damage and decreasing the electrochemical performance. Hwang et al.77 found that the discharge capacities of LiNi0.8Co0.1Mn0.1O2 (NMC811) are low at high 10124

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 6. (a) Cycling performance of MgO-coated LCO and Al2O3-coated LCO.99 (b) Scheme of Al2O3 coating layer evolution on LCO under different annealing temperatures.98 (c) Near-0-ppm region of Al MAS NMR results of as-coated and annealed LCO samples compared to LiAlCoO2 solid solution.98 (d) Schematic view of the electronic structure at the MgO-LCO/electrolyte interface upon electrolyte immersion.101 (e) Schematic view of the surface of MgO-LCO during lithium ion extraction.101 Reproduced from refs99, 98, and101 with permission from The Electrochemical Society, American Chemical Society, and Wiley-VCH, copyrights 2002, 2017, and 2014, respectively.

changes were conducted. Xiang et al.84 found initial capacity changes with different x values in LiMnO3·(1−x)LiCoO2. The results showed that when x = 0.4/0.5, the initial capacity reaches the highest. The larger the x, the higher the oxygen loss. Though modified LCO can deliver high capacity and work at a high cutoff voltage, the irreversible capacity loss, which is caused by the loss of the oxygen and lithium, is still a problem for applying this modification method. Furthermore, the elements loss can result in safety issues which should be taken into consideration. The electrochemical performance and structure are shown in Figure 5c and d. Other Mixed Dopants. Jung et al.85 used metal and F as codopants. They found that Mg and F cosubstituting for Co had a better cycling stability, and the phase transition was not observed through differential capacity vs voltage profiles. The initial capacity of LiMg0.05Co0.95O1.95F0.05 (LMgCOF) was almost the same as LCO. They believed that metal ions could promote Li+ diffusion, and F could modify the surface of the material and prevent the HF attack. The effect of metal ions doped in LCO has been studied thoroughly, but the specific effect of the F ion substituted for O sites still needs further study. Liu et al.86 used La and Al as codopants to improve LCO and showed that the phase transition between 4.1−4.2 V was inhibited (Figure 5e and f), and the material could deliver 190 mAh g−1 at 4.5 V with 96% capacity retention after 50 cycles. They thought that La works as a pillar and Al as a positively charged center, which stabilize the structure, enhance Li+ diffusion, and suppresse phase transition. As discussed above, element-doped LCO performed better than the pristine material, and the effects of different elements have been studied for a long time, such as Mg, Al, Ni, and Mn. They can enhance the electronic conductivity, structure

stability, and structure reversibility. Researchers also tried to combine two elements to dope LCO, and the results are satisfying. But there are still some problems to be studied, such as the actual effect of F substitution for O.



COATING Doping is a great approach to improve the structure stability, electrochemical performance, and thermal stability. But it is hard to suppress the reactions between the electrolyte and cathode surface. The surface reaction is a main cause of the capacity decay. For example, the reaction of HF with hydroxylated LCO results in the formation of H2O and LiF. LiF, which is insoluble in the electrolyte, will inhibit the migration of Li+. The resultant H2O can react with PF5 in the electrolyte, which produces extra HF.90 Considering the shortcomings of the doping methods, coating as an effective way to protect the surface of the LCO particles. The coating materials include oxides, fluorides, phosphates, and others. Oxides. Due to great thermal stability and electrochemical stability, oxides were first used as materials for coatings to modify the surface of LCO. Researchers believed that oxides could act as a protective layer to prevent HF from reacting with the bulk cathode. Cho et al.23,91 synthesized the cathode with SnO2 and Al2O3 coatings. The former delivered a capacity of 166 mAh g−1 with 86% capacity retention after 47 cycles between 2.75−4.4 V at a 0.5 C rate, and the latter delivered capacity of 174 mAh g−1 with 97% after 50 cycles between 2.75−4.4 V. It is believed that the high concentration of Sn or Al on the surface can prevent the phase transition from the monoclinic phase, which results in better electrochemical performance. Scott et al.92 coated LCO with Al2O3 by atomic 10125

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 7. (a), (b) Electrochemical performance of AlF3-coated LCO at different voltages112 (c) Formation and synthetic process for constructing LAF-LCO. (d) Electrochemical performance of LAF-LCO. (e), (f) Electrochemical analyses of 2% LAF-LCO.117 Reproduced from refs 112 and117 with permission from The Electrochemical Society and Nature, copyrights 2009 and 2018, respectively.

4.3 V. Wang et al.99 showed that the MgO-coated LCO showed good electrochemical stability up to 4.7 V, which may result from the diffusion of metal ions from the coating layer and occupied Li vacancies preventing the vacancy ordering (Figure 6a). Shim et al.100 studied the effect of different annealing temperature on MgO-LCO and showed that the sample annealed at 810 °C led to a high doping level of Mg2+ in Li sites, which improved cycling stability through the pillar effect, and it was later proved by Orikasa101 as well (Figure 6d, e). ZrO2,102−104 ZnO,105−108 CuO,109 and Fe2O3110 were also used as coating materials. Though most oxide coatings can enhance cycling stability, the mechanisms for enhancement are controversial. Some researchers believed that the coating can stabilize the structure of LCO. They thought the coating can suppress the phase transition from H2 to M1, which is the main cause of structural degradation. However, other researchers thought that coating materials cannot suppress the phase transition but can make LCO more reversible. In conclusion, the mechanisms of effectiveness of coatings need to be studied further as well as the effect of the solid solution formed between coating materials and bulk LCO. Fluorides. Apart from oxides, fluorides are also effective coating materials. In studies of the oxide coating, researcher found that the surface composition of modified LCO contains F− ions, which aroused the interest for studying fluoride coating materials. Since Al2O3 performed well, AlF3 was widely studied by researchers. Sun et al.111,112 coated LCO with AlF3, and 0.5 mol % AlF3-coated LCO showed better electrochemical performance with capacity retention of 97.7% at 4.5 V and 94.3% at 4.54 V after 50 cycles (Figure 7a and b). They found that the coating layers can stabilize the interface between the cathode and electrolyte, preventing the attack from HF and Co dissolution, which delays the phase transition to the spinel phase during cycling. They also found that using more AlF3 will lead to a thicker inhomogeneous layer, which is not good for electrochemical performance. Lee et al.113 blended AlF3coated LCO and AlF3-coated LiNi1/3Co1/3Mn1/3O2, which

layer deposition and showed that the coated cathode delivered 133 mAh g−1 at 7.8 C, which was a great improvement compared with pristine LCO. This ultrathin film can protect the LCO particles and enable them to be cycled at high rates. Liu et al.93 studied Al2O3-coated LCO by the coprecipitation method and found that amorphous Al2O3 can prevent Co dissolution from LCO, which decreases capacity loss. Oh et al.94 thought that there exists a thin layer of Li-Al-Co-O, which protects the structure from corrosion of the electrolyte. The results suggested that the dissolution of Co decreases, and thermal stability is enhanced. After that, the same group95 observed that AlF3·3H2O was formed from the reaction of the Al2O3 coating with HF and H2O. They believed that the coating can scavenge the H2O impurities, which prevents corrosion and decreases Co dissolution. As a consequence, the electrochemical performance was promoted. Daheron et al.96 studied the LiCo1−xAlxO2 solid solution formed between bulk LCO and the Al2O3 coating via XPS. They found that the compound can make LCO less sensitive to acidic attack and prevents Co dissolution, which enhances cycling stability. Jung et al.97 found that Li3.4Al2O3 is the energetically most favorable composition of the lithiated oxide alumina. They found that Al in the Al2O3 coating layer can accept electrons from the Li atoms resulting in formation of different Al structures, and Li can diffuse faster than Al and O atoms when the concentration is optimized, leading to deeper insight of the Li diffusion in the coating layer. Han et al.98 compared the annealing temperature when coating LCO with Al2O3. They found that Al can diffuse into the LCO lattice, and the modified material showed better initial capacity but worse cyclability at high temperature. They concluded that the worse cyclability resulted from the intercalation of surface alumina into bulk LCO (Figure 6b, c). The coating layer of Al2O3 on LCO is generally believed to be effective to inhibit Co dissolution, as well as preserve the variation range of the lattice parameters during cycling, which greatly improved the capacity retention of the cathode. Other than Al2O3, there are other oxides used as coating materials. Mladenov et al.42 reported MgO-coated LCO, which showed good cyclability with 135 mAh g−1 after 30 cycles at 10126

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 8. (a) Electrochemical performance of AlPO4-coated LCO.127 (b) Cross-sectional TEM images of AlPO4-coated LCO.129 (c) Proposed working mechanism of AlPO4-coated LCO.130 (d) Cycling performance of LiMgPO4-coated LCO.133 (e) Cycling performance of LiAlPO3.93F1.07coated LCO.138 Reproduced from refs 127, 129, 130, 133, and 138 with permission from The Electrochemical Society, American Society Chemistry, American Society Chemistry, Elsevier, and Elsevier, copyrights 2004, 2007, 2009, 2012, and 2017, respectively.

delivered a capacity of ∼185 mAh g−1 at 4.5 V and showed 95% capacity retention after 50 cycles. Zhou et al.114 coated LCO with AlF3 by atomic layer deposition. The fabricated electrode had a high active material loading of 20 mg cm−2 and delivered a capacity of 216 mAh g−1 at 4.7 V and preserved 75.7% of its initial capacity after 100 cycles. Yang et al.115 synthesized LaF3-coated LCO, which had a discharge capacity of 177.4 mAh g−1 at 4.5 V and showed 90.9% of initial capacity retention after 50 cycles. The obtained material also exhibited a good overcharge potential of 4.3, 4.5, and 4.7 V. Abaolaich et al.116 coated LCO with CeF3, which showed a reversibility capacity of 160 mAh g−1. Qian et al.117 synthesized Li-Al-Fcoated LCO, which delivered an initial capacity of 216.2 mAh g−1 and showed a capacity retention of 81.8% at 4.6 V (Figure 7c−f). They created a stable conductive solid solution layer via hydrothermal-assisted surface treatment. Through density function theory calculation, they predicted the most possible structure of the solid solution layer to be Li1/3Al1/3Co2/3O4/3F2/3. The layer retarded the reaction occurring at high voltage and alleviated the phase transition from O3 to H1-3. Even though the coatings with fluorides show better electrochemical performance, the Li diffusivity is likely to be low which may limit the rate capability of the modified cathode. Hao et al.118 studied the Li transport in LCO and found that the diffusivity of Li is low for α-AlF3 (α-Al2O3) due

to unfavorable Li binding sites and relatively high diffusion barriers. This diffusivity is lower than those in the benchmark materials, Li-β-alumina and LiFePO4. Therefore, it is also a challenge for researchers to discover how to enhance the Li diffusivity with fluoride coatings. Phosphates. Phosphates are used as coating material for better thermal stability and can be applied in harsh conditions because of the strong PO bonds and strong covalent bonds formed between polyanions and metal ions.37 Cho et al.119−126 first showed that coating with AlPO4 enhanced the thermal stability and cycling stability of bulk LCO. The modified LCO retained 135 mAh g−1 after 200 cycles at 60 °C.119 After that, they studied120 the correlation between the thickness of the coating layer and thermal stability. The results showed that the thicker the coating is, the better the thermal stability is. However, Li diffusivity decreased with the thicker layer. They concluded that the nanoscale coating layer (∼20 nm) was believed to be the optimized condition for modified LCO. The tests of overcharge behavior were also conducted and showed that the coated LCO did not exhibit thermal runaway even at a 3C rate, which enhanced the safety of Li-ion cells.123,124They122 compared the P2O5-coated LCO and AlPO4-coated LCO and showed that thermal stability was enhanced by the coating, while the AlPO4-coated LCO delivered more capacity during cycling. Compared with Al2O3-coated LCO, AlPO4 could protect LCO even at 4.8 V. 10127

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 9. (a) Scheme of formation solid superacids on LiCoO2 in LiPF6-based electrolyte.150 (b) Discharge capacity retention Li/LiCoO2 cell cycled between 3.3 and 4.5 V at 0.15 mA cm−2 using different electrolytes.150 (c) Schematic of SiO2-IL-TFSI/PC as additives in 1 M LiTFSI hybrid electrolyte system.151 Reproduced from refs 150 and151 with permission from The Electrochemical Society and Wiley-VCH, copyrights 2007 and 2014, respectively.

The irreversible capacity of Al2O3-coated LCO (∼34 mAh g−1) was larger than the AlPO4-coated LCO (∼24 mAh g−1).125 Lee et al.127 studied the concentration of coated AlPO4, and 1 wt % AlPO4 showed better electrochemical performance at 4.8 V (Figure 8a). Kim et al.128 found that the AlPO4 nanoparticles could effectively inhibit the nonuniform distribution of local strain, which greatly enhanced the cycling performance at an ultrahigh voltage of 4.6 V. They studied the size of particles and found that the particles smaller than 20 nm resulted in complete encapsulation, while those with sizes bigger than 150 nm formed partial encapsulation. The electrochemical performance was better with complete encapsulation. They concluded that the AlPO4 coating was better with a particle size smaller than 3 nm and thickness smaller than 10−15 nm. Appapillai et al.129 studied the microstructure change occurring at the surface of AlPO4-coated LCO. They found AlPO4 absent and formation of LiAlyCo1−yO2 and Li3PO4. The former was formed by the migration of Al3+ ions to the lattice of LCO, and the latter was formed through the reaction between LiCO3 and AlPO4. LiAlyCo1−yO2 was believed to stabilize the structure of LCO, and Li3PO4 was believed to act as a Li diffusion channel which enhanced Li ion conductivity. Also, Li3PO4 reduced the potential experienced by LixCoO2 and inhibited the decomposition of the electrolyte (Figure 8b). Lu et al.130 found a thin layer formed on the surface of pristine LCO during cycling containing LiF and LixPFyOz, while the layer formed on the surface of AlPO4-coated LCO included Co- and Al-containing fluorides, oxyfluorides, and PFx(OH)y-like species. They speculated that the coated LCO particles promoted the formation of Co-Al-O-F species, which prevented the side reaction with the electrolyte and reduced impedance growth when cycling at high voltage (Figure 8c).

Lee et al.131 synthesized LiCoPO4-coated LCO through the reaction between Co3(PO4)2 and the surface of LCO. The modified material showed better thermal stability than AlPO4coated LCO. Li et al.132 used FePO4 as a coating material, and the 3 wt % FePO4-coated LCO showed good electrochemical performance with an initial capacity of 146 mAh g−1 at 4.3 V and 88.7% capacity retention after 400 cycles. Morimoto et al.133 studied LiMgPO4-coated LCO which showed an initial capacity of ∼200 mAh g−1 and capacity of ∼175 mAh g−1 after 50 cycles at 4.5 V (Figure 8d). After that, they synthesized Li1+xAlxTi2−x(PO4)3 (LATP) through the Li2O−Al2O3− TiO2−P2O5 system. LATP-coated LCO exhibited a high capacity of 180 mAh g−1 at 4.5 V. The high Li diffusivity of LATP inhibited the increase in charge transfer resistance and decreased the activation energy.134 Since most coating materials have poor ionic mobility, Zhou et al.135 used Li3PO4 (LPO)-coated LCO to improve the performance. LPO acted as an efficient channel for Li+, which enhanced the rate performance of LCO. Shen et al. synthesized LiAlPO3.93F1.07 (LAPF)-coated LCO.138 It delivered a capacity of 206 mAh g−1 up to 0.5 C and showed great capacity retention of 91.7% after 50 cycles (Figure 8e). From the coating with phosphate, researchers found an interesting phenomenon. Bai et al.136 used YPO4 as a coating layer and found that YPO4 increased the acidity of the electrolyte by converting Lewis acid, which enhanced the conductivity of the SEI layers. They believed the effect could be realized through adding YPO4 or Al2O3 into the electrolyte and concluded that the Lewis acids purified the surface of LCO, increased the conductivity of formed SEI, and promoted the formation of a more stable surface solid solution with the substitution for Co and Li ions with metal ions from these additives.137 10128

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

as an electrolyte additive. We151,152 developed an organic− inorganic hybrid electrolyte by grafting SiO2 nanoparticles with ionic liquid 1-methy-3-propylimidazolium bis(trifluoromethanesulfone)imide in a conventional PC/LiTFSI liquid electrolyte (SiO2-IL-TFSI) (Figure 9c). The nanoparticles can serve as reservoirs and immobilize the anions, thus significantly retarding Li dendrite growth and delaying a cell short circuit by 1 order of magnitude in comparison to that using a pure PC electrolyte according to the space charge model. Also, the room-temperature electrochemical stability window of the silica nanoparticle-modified electrolyte was proved stable in the voltage range of 0−7 V. Inspired by the solid superacid model, the SiO2-combined nanoparticle additive may also act as a precursor of a superacid and be an appropriate HVLCO alternate, meanwhile avoiding the dendrites generated from the lithium metal anode. Benzene Derivatives and Aromatic Heterocyclic Additives. Biphenyl (BP) was first chosen to be an overcharge indicator, as H2 evolution would generate at higher voltage and stop the battery operation by opening the safety cap at a certain pressure. But the thick, non-Li+-conductive decomposition generated from too much BP will cover the cathode surface and make it inert. Replaced by forming a very thin Li+conductive polymer film on the surface of LCO using just a tiny amount (e.g.,0.1−0.2 wt %), Abe et al.153,154 first presented the effects of a series of benzene derivatives (BP and o-terphenyl) and heterocyclic compound (furan, thiophene, N-methylpyrrole, and 3,4-ethylenedioxyiophene) additives on improving the cycle performance of HVLCO above 4.2 V. Because these additives process higher HOMO energy levels than those of carbonate solvents, they can be oxidized ahead of the electrolytes. Also, because of the conjugated structure of the polymer species, these films are electroconductive and named as electro-conducting membrane (ECM). After this initial study, Takeuchi et al.155,156 examined OT, N,N-dimethyl-aniline, anisole, thioanisole, o/m/p-(1Adamantyl) toluene, furan, 2-methyl furan, 2,5-dimethyl furan, thiophene, 2-methyl thiophene, and 2,5-dimethyl thiophene for potential additives. They obtained the exact oxidation potentials (Eox) by linear scanning of the voltage method with Li/Pt cells at a scan rate of 5 mV/s and proposed the relatively lower oxidation potentials, making these additives oxidize quicker than the base electrolyte solution, thereby covering the catalytic and oxidative sites of LCO. The stabilized cathode/electrolyte interface allows excellent cycling performance up to 4.4 V. Xia et al.157 further found that 2,2′bithiophene (2TH) and 2,2′:5′,2″-terthiophene (3TH) can be electrochemically polymerized prior to the base electrolyte and even the monomolecular thiophene on the cathode surface, which blocks off severe electrolyte decomposition at high voltages. The discharge capacity retention is 50% in the base electrolyte, while the electrolyte containing 0.1 wt % 3TH displays a high retention of 84.8% in the voltage range of 3.0− 4.4 V. Phosphorus-Based Additives. It is generally recognized that the highly combustible carbonate-based organic electrolyte used in the current lithium-based battery technology has a potential hazards of firing and explosion under hostile conditions such as short circuit, high thermal impact, and especially electrical overcharging. A variety of phosphides have been confirmed useful as flame-retardant solvents or additives to improve the safety of batteries. Also, some of them such as (phenoxy)pentafluorocyclotriphosphazene N 3 P 3 (OPh)F 5

Coating materials have been studied for a long time, and many groups found various chemicals for coating. The solid solution between the coating layer and bulk LCO is considered as the protective layer for inhibition of side reactions between LCO and the electrolyte. It is common that solid solutions form during the annealing at high temperature. However, the exact properties and composition are still not clear, meaning that more in depth understanding is needed. Additionally, the different coating materials showed different improvement of LCO. For example, fluorides can enhance stability, and phosphates can tolerate harsh conditions. Therefore, selecting suitable coating materials in different conditions needs to be considered. Also, the binary coatings such as LiMn2O4,139,140 MgAl2O4,141 and LiMgPO4133 showed better electrochemical performance than the unitary coatings, but there are few studies about it. Furthermore, polymers have also been used as coating materials, but the studies and their working mechanism are rare.



ELECTROLYTE ADDITIVES As discussed above, many doping and coating methods have been used to improve the structural stability and the interfacial compatibility of LCO at high voltages. However, conventional carbonate solvent-based electrolytes such as ethylene carbonate exhibit inferior anodic stability lower than 4.3 V vs Li/Li+, which becomes another obstacle to hinder HVLCO for further commercialization. Moreover, the high flammability of nonaqueous electrolyte solvents is a big potential safety hazard. To overcome these annoying issues, many efforts have been focused on developing advanced electrolytes with unique characteristics such as fast ion transportation, less side reactions, higher operation voltage ranges, and working temperature limits. In terms of solvents, sulfone-based solvents,142−144 room temperature ionic liquids,145−147 dinitrile solvents,148 and plenty of functional group-modified carbonates146,149 with high anodic stabilities have been investigated. Unfortunately, these solvents still suffer from some issues including large viscosity, high price, low cathodic stability combining carbonaceous cathodes, and high toxicity. On the other hand, adding just a small amount (usually less than 5%) of functional additives, which can form a stable artificial solid electrolyte interphase layer on the cathode (CEI) and improve the anodic stability of the electrolyte, have been recently recognized as a promising means for HVLCO due to the lower cost and high effectivity. Nanocompound Additives. Different from a surface coating/modification strategy using metal oxides, Liu et al.150 proposed a novel solid superacid model (Figure 9a) by soaking nano-Al2O3 in a common LiPF6/EC/DMC electrolyte and then removing nano-Al2O3. They found that the added nanoAl2O3 scavenges HF, which is recognized to lead to capacity leakage and converts to solid superacids AlF3/Al2O3 and LiAlF6/AlF3 in the presence of a water residue. Furthermore, these solid superacids can corrode the insulating alkaline species such as Li2CO3 and LiOH on LiCoO2 and enhance the ionic conduction of CEI; thus, 85% of the initial discharge capacity (192 mAh g−1) remained after 100 cycles between 3.3 and 4.5 V at 0.15 mA cm−2 using the nano-Al2O3 soaked electrolyte, as shown in Figure 9b. Considering various superacids can be formed by a similar mechanism, the group speculated that other metal oxides such as MgO, ZnO, ZrO2, TiO2, and SiO2 can also work by forming solid superacids. Recently, our group also have studied some issues of SiO2 used 10129

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 10. (a) Illustration of the flammability testing of the electrolytes.159 (b) Cycle performance of LiCoO2/Li half cells in electrolytes without and with 5 wt % PFPN at current density of 40 mA g−1 and cutoff voltage of 4.5−3.0 V.159 (c) C 1s, O 1s, P 2p, and N 1s XPS spectra of LCO using the electrolyte without and with the EEEP additive after 50 cycles.162 Reproduced from refs 159 and 162 with permission from Elsevier and Elsevier, copyrights 2015 and 2017.

Figure 11. (a) Linear sweep voltammograms of LiCoO2/Li.163 (b) Cycle performance of LiCoO2/graphite cells with different amounts of DPDS in voltage ranges of 3.0−4.4 V.163 (c) Discharge capacity of cells as a function of cycle number obtained at different cutoff voltages: 3.0−4.2 V (up), 3.0−4.5 V (down).164 TEM images of LiCoO2 electrode: (d) fresh, (e) without MMDS after 150 cycles, and (f) with MMDS after 150 cycles.164 Reproduced from refs 163 and 164 with permission from Elsevier and Elsevier, copyrights 2016 and 2012, resepctively.

(PFPN), 158,159 N-(triphenylphosphoranylidene) aniline (TPPA),160 tri(β-chloromethyl) phosphate (TCEP),161 and poly[bis(ethoxyethoxyethoxy)phosphazene] (EEEP)162 have been reported as electrolyte additives to improve the

performances of HVLCO. Xia et.al159 found that an addition of only 5% PFPN can make the electrolyte thoroughly nonflammable, as shown in Figure 10a. The discharge capacity of the cell with the PFPN additive in the electrolyte only 10130

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research showed a 10% capacity fade from 188.1 to 169 mAh g−1 for 60 cycles in contrast to a 34.6% fade using a blank electrolyte at a high cutoff voltage of 4.5 V (Figure 10b). However, since PFPN is difficult to commercialize for its low volatilization temperature, very recently the group162 synthesized and explored the EEEP additive. They found that a 5% EEEPcontaining electrolyte made the capacity retention of the LCO cathode elevate to 89.9% from 51.2% after 100 cycles at a high cutoff voltage of 4.4 V, meanwhile obviously reducing the flammability of the electrolyte. They ascribed the enhanced performance to a stable solid interface forming on the cathode which is related to the P−O bond of molecules based on extended analyses of SEM and XPS (Figure 10c). Sulfur-Based Additives and Other Additives. Sulfurbased additives, such as diphenyl disulfide (DPDS),163 methylene methanedisulfonate (MMDS),164 1,3,2-dioxathiolane-2,2-dioxide (DTD), 1,3-propanediol cyclic sulfate (TMS), and propylene sulfate (PLS)165 have also been used in the HVLCO cathode not only because they have a similar effect to the additives but also because they benefit the graphite or lithium metal anode owing to the sulfur-based functional groups. With the guidelines that SO2, vinylene trithiocarbonate, or butylene sulfite could be reduced prior to the carbonate solvent and form a positive SEI film on the graphite anode, Zhao et al.163 evaluated DPDS which consists of both a conjugated benzene component and sulfur atoms as a new bifunctional electrolyte additive. Linear sweep voltammograms proved the oxidation of DPDS (Figure 11a), while XRD, FTIR, and SEM analyses demonstrated that the SEI film is produced primarily on the graphite anode via the decomposition of DPDS at normal voltage, and the CEI can be formed simultaneously on LCO at higher potentials. The cell with 1.0 wt % DPDS exhibited the best capacity retention with a value of 83% of the initial discharge capacity and still retained 130 mAh g−1 after 200 cycles in the range of 3.0−4.4 V (Figure 11b). Zuo et al.164 found that the capacity retention of LCO cells in the range of 3.0−4.5 V is significantly increased from 32.0% to 69.6% after 150 cycles with the addition of 0.5 wt % MMDS into the base electrolyte, as shown in Figure 11c. They ascribed this improvement to the modification of components of the cathode surface layer in the presence of MMDS. TEM images clearly demonstrated that a thin surface layer which has a thickness of approximately 3−5 nm (Figure 11d−f) would effectively decrease the interfacial resistance thus improving the cycling performance. Boron-containing additives including lithium bis(oxalate)borate,166 tris (2H-hexafluoroisopropyl) borate,167 and trimethylboroxine168 have been found to be effective electrolyte additives for the formation of protective films on the HVLCO. Some nitrile-based electrolytes, such as fumaronitrile,169 heptyl cyanide,170 suberonitrile, succinonitrile (SN),171 and 4-(trifluoromethyl) benzonitrile172 have also been used as HVLCO additives because the electrochemically and thermally stable −CN group of the decomposition has a lone pair of two electrons leading to a strong bond with Co ions on the LCO surface. Mixed Additives. It is noted that single electrolyte additives, which are the focus of many academic studies, are unlikely to ever be competitive with multiple additives chosen to act synergistically. Wang et al.173 studied over 55 combinations of additives in which vinylene carbonate (VC), fluoroethylene (FEC), vinyl ethylene carbonate (VEC), LiBOB, and LiTFSI were highlighted. The group verified VC as a promising additive and set the performance of cells with 2

wt % VC as a baseline to evaluate the effectiveness of additive blends. The combination of VC, VEC, and FEC, 2 wt % for each component, was shown to have the highest Coulombic efficiency and the second lowest charge slippage but with the highest impedance and recognized suitable for cells destined only for low rate or long lifetime applications. Kim et al.171 found that the addition of both 2 wt % VC and 2 wt % SN leads to intermediate behavior, while single 2 wt % VC/SN, respectively, suppressed/increased the impedance growth of the cathode during cycling. Ji et al.174 revealed that LiBOB can be oxidized prior to the decomposition of base electrolytes but does not work well with Co-based LCO, while the use of SUN tends to form too thick CEI containing a large amount of LiF. However, the interface impedance was greatly suppressed thus enhancing the cycling performance of LCO when using a combination of additives containing both. Recently, Cui’s group167 developed a novel tree-component functional additive using tris(2H-hexafluoroisopropyl) borate (THFPB), adiponitrile (AND), and cyclohexylbenzene (CHB). By systematic analysis, it is demonstrated to be a synergistic effect, that THFPB can suppress the appearance of needle-shaped crystals by the addition of the dinitrile of AND and reduce the polarization and interfacial resistance by partially dissolving resistant components such as LiF. Furthermore, the addition of CHB, which consisted of benzene groups, participates in CEI modification at the high voltage by polymerization as mentioned above. The cells containing this three-component additive showed a higher capacity retention of 87.6%, 79.6%, and 60.0% than the base electrolytes at cutoff voltages of 4.4, 4.45, and 4.5 V, respectively. Quasi-Solid Electrolytes. Adding electrolyte additives is a facile and effective way to enhance the performance of LCO at a high cutoff voltage, but liquid electrolytes essentially still have potential safety concerns of electrolyte leakage, depletion, limited working temperature, and negligible mechanical strength, as well as the potential internal short circuit caused by dendrites combining with the lithium metal anode. The current solid electrolytes, which are recognized as an ultimate alternative toward an all-solid state battery, can be divided into two categories: polymer and inorganic ceramic. Usually, polymer electrolytes exhibit better flexibility, with relatively low ionic conductivity and a narrow voltage window, while the latter possesses a desirable high mechanism strength and a higher ionic conductivity but is very fragile. One strategy is to make a polymer/ceramic composite solid electrolyte, but unfortunately, things are still far from satisfactory because of some annoying issues, especially sluggish/unstable interfaces and low bulk ionic conductivity problems. Quasi-solid electrolytes are some normally solid composite systems with liquid electrolytes. Combining the advantages of both liquid and solid electrolytes, the quasi-solid electrolytes may become an appropriate choice for transitional period to all-solid-state high-voltage LCO batteries. Gel Polymer Electrolytes. Gel polymer electrolytes (GPEs) are formed by adding plasticizers or solvents into solid polymer electrolytes, which have a special phase between solid and liquid. Ether solvents, liquid oligomers, and ionic liquids are commonly used as plasticizers. The transport of Li+ is mainly supported by the solvation of plasticizers, while polymers play a holder role. Plenty of works have demonstrated that GPEs have higher acceptable ionic conductivity than solid electrolytes at ambient temperature and higher thermal and mechanical stability than liquid 10131

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 12. (a) Schematic representation of the PMM-CPE fabrication process. (b) Schematic illustration of flexible LiCoO2/Li metal cell in a bent state. (c) Cycle performance (1 C) of the 4.45 V LiCoO2/Li cell with LiODFB/PC electrolyte and PMM-GPE at 60 °C.183 Reproduced from ref 183 with permission from Wiley-VCH, copyright 2018.

(Figure 12c). By contrast, the basic battery based on the LiPF6/EC-DMC electrolyte only keeps 1% of the initial capacity at same test conditions. Rao et al.184 developed a GPE using an electro-spun polymer blending membrane of PAN and PMMA incorporated with N-methy-N-butylpyrrolidinium bis(trifluoromethanesulfonyl)imide ionic liquids. The electrospun membrane has 3D interconnected fibers with a porosity of 86% and a reversible cathodic stability up to 5 V. Inspired by additives used in liquid electrolytes, Choi’s group185 first studied the influence of electrolyte additives including EDOT, TH, and BP on the cycling performance of LCO based on GPE. PVdF-HFP, dibutyl phthalate, and acetone were mixed, ball milled, and then cast to a thin film of 500 μm. After being immersed in methanol to remove the plasticizer, homogeneous pores can be realized to contain a certain amount of liquid electrolyte without leakage. Their research demonstrated that the thiophene additive led to the best performance with a capacity retention and a high initial discharge capacity of 136 mAh g−1 between 2.8 and 4.3 V. Recently, bilayer electrolytes186 in single battery strategy were also suggested in which the high-voltage electrolyte near the cathode can protect the low-voltage electrolytes near the anode from oxidation. In summary, some efforts have made to improve the performance of high voltage, but the GPEs are still limited. It is suggested that novel high-performance GPEs and more effective bulk or interfacial modification methods should be developed. Electrolytes Combining Ceramic Nanoparticles. To achieve a quasi-solid state, ceramic nanoparticles are conveniently adapted to the electrolytes with an amount ≥50%. These nanoparticle fillers can improve the bulk conductivity while retaining the mechanism strength and flexibility. Furthermore, liquid electrolytes with ceramic fillers display wider electrochemical windows which offer the possibility of applying HVCLO and enabling good long-term cycling performance. Choi et al.187 developed a novel hybrid quasi-solid electrolyte composed of an ionic liquid (LITFSI/ Pyr14TFSI) and BaTiO3 ceramic nanoparticles without a polymer. Dispersing an amount of BaTiO3 ceramic powder in the liquid ionic electrolyte system can develop space charge layers on the boundaries of ceramic particles which tend to

electrolytes. A series of polymer hosts, such as poly(ethylene oxide) (PEO),175 poly(propylene oxide) (PPO),176 poly(acrylonitrile) (PAN), 177 poly(methyl methacrylate) (PMMA),178 poly(vinylidene fluoride) (PVdF),179 poly(vinylidene fluoride-co-hexafluoro propylene)(PVdF-coHFP),180 and poly(vinylpyrrolidone) (PVP)181 have been developed for many years. Though considered to have a relatively better surface contact due to the flexibility of the polymer, few of them are compatible with high-voltage cathodes caused by a limited voltage window. For example, the widely studied PEO holder will be slowly oxidized at a voltage over 3.9 V, which has restricted their use. Normally, blending, copolymerization, cross-linking, and surface modification are the main strategies to improve high voltage stability, ionic conductivity, and mechanical stability. Kim et al.182 developed a homogeneous gel polymer electrolyte consisting of 25 wt % PVdf-co-HFP, 65 wt % ECPC, and 10 wt % LiTFSI using a solvent-casting technique, and its ionic conductivity can reach 1.2 × 10−4 S cm−1. While combining the LCO cathode and Li metal anode, the cell delivers a specific discharge capacity of 147.6 mAh g−1 and a Coulombic efficiency of 99.8% at the initial cycle at 4.2 V. Recently, Cui’s group183 demonstrated for the first time a bacterial cellulose-supported copolymer poly(methy vinyl ether-alt-maleic anhydride) (P(MVE-MA) combining PC/ LiODFB liquid electrolytes as a gel polymer electrolyte (PMMGPE) for a 4.45 V LCO lithium metal battery (the highest voltage application in LCO for GPE that we know) The fabrication process is shown in Figure 12a. The produced P(MVE-MA) membrane consists of a sandwich cross-linking architecture of which the bacterial cellulose can work as a stress contributor, while P(MVE-MA) consisting of an ultrahigh Young’s modulus of 6.9 GPa works as a strain contributor so that a flexible electronic structure can be achieved (Figure 12b). Electrochemical measurements further demonstrated that PMM-GPE has a wide voltage window over 5.2 V and a high Li+ transference number of 0.54. Based on the superiority of all components, the LCO/Li cell with PMM-GPE exhibits an ultrahigh discharge capacity retention of 85% of the initial capacity (174.4 mAh g−1) after 700 cycles even at 60 °C 10132

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

Figure 13. (a), (b) Model of the lithium-ion conducting paths arising from the formation of the charge space. (c) Flammability testing of hybrid solid electrolyte. (d) Cycling performance for Li/LiCoO2 cells with liquid electrolyte and hybrid solid electrolyte at 80 °C.187 Reproduced from ref 187 with permission from Springer, copyright 2017.

floating of an interface during charging and discharging could make voids at the interface intensify the bad contact. Adding modified electrolytes is recognized as an effective method to improve the interfacial performance. Xu et al.191 used n-BuLi to construct a stable interface between the Li metal anode and the Li7La3Zr1.5Ta0.5O 12 ceramic solid electrolytes. The modified cells had been cycled for 400 cycles at 100 and 200 μA cm−2 with a constant resistance, while the interface resistance of the control group increased from 1056 to 2419 Ω cm2 after only 10 cycles. Wang et al.192 revealed that a small amount of liquid electrolyte (2 μL LiPF6 in EC/DMC/DEC) at the interface can eliminate the large interfacial resistance while preventing the reduction of Li1.4Al0.4Ti1.6(PO4)3 by a lithium metal. Kitaura et al.193 fabricated LCO interfaces by using the supercooled liquid state of glassy electrolytes. A method of LiNbO3 coating was used to further suppress the side reaction of LCO and electrolytes. Li4Ti5O12/80Li2S· 20P2S5 glass/LiNbO3-coated LiCoO2 prepared showed a reversible capacity of 120 mAh g−1 at 0.064 mA cm−2.

overlap, yielding favorable pathways for Li+ transport (Figure 13a, b). The hybrid quasi-solid electrolyte exhibits excellent flame retardancy (Figure 13c), an excellent ionic conductivity of 1.3 × 10−3 S cm−1 at 30 °C, and a high anodic stable voltage up to 6.0 V. Also, a stable discharge capacity of 128 mAh g−1 (0.1 C) at 80 °C (Figure 13d) was available due to the high thermal stability of this electrolyte. Appetecchi et al.188 reported that adding selected Al2O3 ceramic powders can also solve the liquid leakage problem. Li et al.189 used the insulting 2D graphene-analogues boron nitride (g-BN) nanosheets as an electrolyte host. The rich porosity enables g-BN to incorporate a large amount of ionic liquids, as much as 10 times its initial volume. The hybrid electrolytes can achieve a high ionic conductivity of 3.85 × 10−3 S cm−1 at 25 °C (even 2.32 × 10−4 S cm−1 at −20 °C) due to the smooth Li+ transport pathway of uniform and wide channels between the outside g-BN layers. Tsao et al.190 prepared a novel organic− inorganic hybrid gel polymer electrolyte with immobilized ionic liquid components by a combination of a sol−gel method and a chemical cross-linking reaction. This quasi-solid electrolyte exhibits a low glass transition temperature, high ionic conductivity of 5.9 mS cm−1 with the lithium ion transference number increasing from 0.28 to 0.57, and a high electrochemical window of 5.0 V. In general, ceramic nanoparticles combined with quasi-solid electrolytes are very promising candidates for LCO-based batteries with higher energy density, but the choices are still limited. Furthermore, the conductivity and oxidation stability should be further enhanced to match HVLCO. Electrolyte Wetting on the Interface. Ceramic electrolytes usually possess higher modulus and ionic conductivity. However, the intrinsic rigidity always makes an inferior interfacial contact known as a heterogeneous solid−solid contact, thus leading to sluggish ion transport. Some study also showed that the major voltage drop from cathode to anode is mainly derived from the huge interfacial impedance. Moreover,



CONCLUSION AND PERSPECTIVE To meet the demand of high energy density batteries in portable electronic devices, charging LCO to a higher voltage is a satisfying choice. However, due to the intrinsic properties of LCO and instability of the interfaces between LCO particles and other battery components, researchers have studied LCO and found a lot of strategies to solve the problem, including doping, coating, and modification of the electrolyte. Doping was the earliest proposed and effective way to stabilize the LCO particles. Element doping, with Al, Mg, and Ca, has been proved by many groups. The widely agreed mechanism of dopants is the pillaring effect, meaning that the elements have larger ionic radii than Co ions so that they can stabilize the layered structure during the process of lithium ion intercalation/deintercalation. Some elements like Mg were also believed to change the electronic conductivity by increasing 10133

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research the amount of Co4+. Though the doping method has been studied for a long time, there are still a lot of unknown areas remaining to be studied, such as the effects of F−, effects of multiple element doping, and mechanism of phase transition inhibition. So, for doping, more studies are needed to understand the effects of different elements, especially the mechanism of inhibition of phase transition and the synergetic effects of two or more elements, especially for combinations of anions and cations. Coating is another effective way to enhance the electrochemical performance of HVLCO. It can construct a protective layer to prevent corrosion from the electrolyte which can produce HF with a trace of water. The coating layer is believed to make the phase transition more reversible rather than inhibit the phase transition, which is different from doping. It is also widely agreed that there exists a solid solution between bulk LCO and the coating layer. The solid solution was believed to be synthesized through transition metals diffusing to the surface of LCO. However, there is no direct evidence to prove it or characterize it. Thus, finding a way to detect the properties of a solution layer is a focus of researchers. Since lithium ion mobility should be considered when designing coating materials, more studies should focus on lithium ion conductive materials instead of the stability of materials only. Therefore, a polycation substance with Li ions can be considered as a great candidate, and a polyanion may be effective as well. In terms of liquid electrolytes, adding electrolyte additives is generally more attractive for its tiny amount of use and great effectiveness of enhancing the high voltage performance of LCO through the modified cathode/electrolyte interface and the improved cathodic stability of the electrolyte. Additives with a high HOMO can be preferentially oxidized or electrochemically oxidized to form functional CEI films, which covers the active sites and prevents further decomposition of the electrolytes. Multifunctional phosphorus-based additives can largely improve the thermal stability of HVLCO and lead to flame-retardant electrolytes. Sulfide-based additives can simultaneously enhance the anodic stability of graphite or lithium metal anodes. Mixed additives, which are chosen to work synergistically and usually outperform a single additive, should be the focus of future studies. The quasi-solid electrolyte which is recognized as an appropriate alternative nowadays displays higher safety and mechanical stability than liquid electrolytes, but it still suffers from poor interface stability. So, strategies about stable and compatible interface constructions are still needed, and the loading of liquid electrolytes should be further reduced. In this review, most of the common modification strategies are summarized. Since many groups have studied LCO for decades, the LCO cathode has been widely used in lithium-ion batteries for portable electronic devices. HVLCO-based batteries have advantages in terms of capacity, voltage, and compaction density compared to other cathode materials.



Author Contributions †

Xiao Wang and Xinyang Wang contributed equally to this work. Xiao Wang wrote the Introduction and the following sections: Challenges for LCO; Doping; and Coating. Xinyang Wang wrote the Introduction and the following sections: Electrolyte Additives and Conclusion and Perspective. Notes

The authors declare no competing financial interest. Biographies

Xiao Wang received his B.S. degree in 2018 from the Zhejiang University. He is currently a M.S. student in the College of Chemical and Biological Engineering, Zhejiang University. His work focuses on high-voltage cathode materials in rechargeable lithium-ion batteries.

Xinyang Wang is currently an M.S. student in the College of Chemical and Biological Engineering, Zhejiang University. He received his B.S. degree in applied chemistry from the China University of Mining and Technology in 2018. His research is focused on electrochemical conversion and storage materials.

AUTHOR INFORMATION

Corresponding Author

*Tel.: +86-0571-87953906. E-mail: [email protected]. ORCID

Yingying Lu: 0000-0001-9713-8441 10134

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

(18) Ohzuku, T.; Ueda, A. Solid-state redox reactions of LiCoO2 (R(3)over-bar-m) for 4 V secondary lithium cells. J. Electrochem. Soc. 1994, 141, 2972. (19) Amatucci, G. G.; Tarascon, J. M.; Klein, L. C. CoO2, the end member of the LixCoO2 solid solution. J. Electrochem. Soc. 1996, 143, 1114. (20) Amatucci, G. G.; Tarascon, J. M.; Klein, L. C. Cobalt dissolution in LiCoO2-based non-aqueous rechargeable batteries. Solid State Ionics 1996, 83, 167. (21) Van der Ven, A.; Aydinol, M. K.; Ceder, G. First-principles evidence for stage ordering in Li(x)CoO(2). J. Electrochem. Soc. 1998, 145, 2149. (22) Van der Ven, A.; Aydinol, M. K.; Ceder, G.; Kresse, G.; Hafner, J. First-principles investigation of phase stability in LixCoO2. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58, 2975. (23) Cho, J.; Kim, Y. J.; Park, B. Novel LiCoO2 cathode material with Al2O3 coating for a Li ion cell. Chem. Mater. 2000, 12, 3788. (24) Yang, X. Q.; Sun, X.; McBreen, J. New phases and phase transitions observed in Li1‑xCoO2 during charge: In situ synchrotron X-ray diffraction studies. Electrochem. Commun. 2000, 2, 100. (25) Chebiam, R. V.; Kannan, A. M.; Prado, F.; Manthiram, A. Comparison of the chemical stability of the high energy density cathodes of lithium-ion batteries. Electrochem. Commun. 2001, 3, 624. (26) Cho, J.; Kim, Y. J.; Kim, T. J.; Park, B. Zero-strain intercalation cathode for rechargeable Li-ion cell. Angew. Chem., Int. Ed. 2001, 40, 3367. (27) Sun, X.; Yang, X. Q.; McBreen, J.; Gao, Y.; Yakovleva, M. V.; Xing, X. K.; Daroux, M. L. New phases and phase transitions observed in over-charged states of LiCoO2-based cathode materials. J. Power Sources 2001, 97, 274. (28) Chen, Z. H.; Lu, Z. H.; Dahn, J. R. Staging phase transitions in LixCoO2. J. Electrochem. Soc. 2002, 149, A1604. (29) Chen, Z. H.; Dahn, J. R. Improving the capacity retention of LiCoO2 cycled to 4.5 V by heat-treatment. Electrochem. Solid-State Lett. 2004, 7, A11. (30) Chen, Z. H.; Dahn, J. R. Methods to obtain excellent capacity retention in LiCoO2 cycled to 4.5 V. Electrochim. Acta 2004, 49, 1079. (31) Marianetti, C. A.; Kotliar, G.; Ceder, G. A first-order Mott transition in LixCoO2. Nat. Mater. 2004, 3, 627. (32) Goodenough, J. B.; Kim, Y. Challenges for rechargeable Li batteries. Chem. Mater. 2010, 22, 587. (33) Ensling, D.; Cherkashinin, G.; Schmid, S.; Bhuvaneswari, S.; Thissen, A.; Jaegermann, W. Nonrigid band behavior of the electronic structure of LiCoO2 thin film during electrochemical Li deintercalation. Chem. Mater. 2014, 26, 3948. (34) Hoang, K.; Johannes, M. D. Defect chemistry in layered transition-metal oxides from screened hybrid density functional calculations. J. Mater. Chem. A 2014, 2, 5224. (35) Kalluri, S.; Yoon, M.; Jo, M.; Park, S.; Myeong, S.; Kim, J.; Dou, S. X.; Guo, Z.; Cho, J. Surface engineering strategies of layered LiCoO2 cathode material to realize high-energy and high-voltage Liion cells. Adv. Energy Mater. 2017, 7, 1601507. (36) Min, S. H.; Jo, M. R.; Choi, S.-Y.; Kim, Y.-I.; Kang, Y.-M. A layer-structured electrode material reformed by a PO4-O-2 hybrid framework toward enhanced lithium storage and stability. Adv. Energy Mater. 2016, 6, 1501717. (37) Kalluri, S.; Yoon, M.; Jo, M.; Park, S.; Myeong, S.; Kim, J.; Dou, S. X.; Guo, Z.; Cho, J. Surface engineering strategies of layered LiCoO2 cathode material to realize high-energy and high-voltage Liion cells. Adv. Energy Mater. 2017, 7, 1601507. (38) Fan, L.; Wei, S.; Li, S.; Li, Q.; Lu, Y. Recent progress of the solid-state electrolytes for high-energy metal-based batteries. Adv. Energy Mater. 2018, 8, 1702657. (39) Tukamoto, H.; West, A. R. Electronic conductivity of LiCoO2 and its enhancement by magnesium doping. J. Electrochem. Soc. 1997, 144, 3164. (40) Zou, M. J.; Yoshio, M.; Gopukumar, S.; Yamaki, J. Synthesis and electrochemical performance of high voltage cycling (LiMCoO2)-

Yingying Lu is now a Principal Investigator in the College of Chemical and Biological Engineering, Zhejiang University. She received her B.S. degree in Chemical Engineering from Zhejiang University in 2010 and her Ph.D. from Cornell University in 2014. Afterward, she worked as a postdoctoral fellow in the Department of Materials Science and Engineering at Stanford University. Her research interests include nanomaterials, ionic liquids, and electrochemical energy storage and conversion.



ACKNOWLEDGMENTS



REFERENCES

We acknowledge financial supports from the National Key R&D Program of China (2018YFA0209600, 2016YFA0202900) and the Natural Science Foundation of China (NSFC, Grants 21878268 and 21676242).

(1) Armstrong, A. R.; Bruce, P. G. Synthesis of layered LiMnO2 as an electrode for rechargeable lithium batteries. Nature 1996, 381, 499. (2) Ceder, G.; Chiang, Y. M.; Sadoway, D. R.; Aydinol, M. K.; Jang, Y. I.; Huang, B. Identification of cathode materials for lithium batteries guided by first-principles calculations. Nature 1998, 392, 694. (3) Tarascon, J. M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359. (4) Armand, M.; Tarascon, J. M. Building better batteries. Nature 2008, 451, 652. (5) Aykol, M.; Kim, S.; Hegde, V. I.; Snydacker, D.; Lu, Z.; Hao, S.; Kirklin, S.; Morgan, D.; Wolverton, C. High-throughput computational design of cathode coatings for Li-ion batteries. Nat. Commun. 2016, 7, 13779. (6) Choi, J. W.; Aurbach, D. Promise and reality of post-lithium-ion batteries with high energy densities. Nat. Rev. Mater. 2016, 1, 16013. (7) Janek, J.; Zeier, W. G. A solid future for battery development. Nat. Energy 2016, 1, 16141. (8) Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical energy storage for the grid: A battery of choices. Science 2011, 334, 928. (9) Li, H.; Wang, Z.; Chen, L.; Huang, X. Research on advanced materials for Li-ion batteries. Adv. Mater. 2009, 21, 4593. (10) Wang, L.; Chen, B.; Ma, J.; Cui, G.; Chen, L. Reviving lithium cobalt oxide-based lithium secondary batteries-toward a higher energy density. Chem. Soc. Rev. 2018, 47, 6505. (11) Radin, M. D.; Hy, S.; Sina, M.; Fang, C.; Liu, H.; Vinckeviciute, J.; Zhang, M.; Whittingham, M. S.; Meng, Y. S.; Van der Ven, A. Narrowing the gap between theoretical and practical capacities in Liion layered oxide cathode materials. Adv. Energy Mater. 2017, 7, 1602888. (12) Jung, S.-K.; Gwon, H.; Hong, J.; Park, K.-Y.; Seo, D.-H.; Kim, H.; Hyun, J.; Yang, W.; Kang, K. Understanding the degradation mechanisms of LiNi0.5Co0.2Mn0.3O2 cathode material in lithium ion batteries. Adv. Energy Mater. 2014, 4, 1300787. (13) Yan, P.; Zheng, J.; Gu, M.; Xiao, J.; Zhang, J.-G.; Wang, C.-M. Intragranular cracking as a critical barrier for high-voltage usage of layer-structured cathode for lithium-ion batteries. Nat. Commun. 2017, 8, 14101. (14) Mizushima, K.; Jones, P. C.; Wiseman, P. J.; Goodenough, J. B. LiXCoO2 ″(oless-thanxless-than-or-equal-t01) - a new cathode material for batteries of high-energy density. Mater. Res. Bull. 1980, 15, 783. (15) Goodenough, J. B.; Mizushima, K.; Takeda, T. Solid-solution oxides for storage-battery electrodes. Jpn. J. Appl. Phys. 1980, 19, 305. (16) Reimers, J. N.; Dahn, J. R. Electrochemical and insitu x-raydiffraction studies of lithium intercalation in LiXCoO2. J. Electrochem. Soc. 1992, 139, 2091. (17) Reimers, J. N.; Dahn, J. R.; Vonsacken, U. Effects of impurities on the electrochemical properties of LiCoO2. J. Electrochem. Soc. 1993, 140, 2752. 10135

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

(61) Valanarasu, S.; Chandramohan, R.; Thirumalai, J.; Vijayan, T. A. Structural and electrochemical investigation of Zn-doped LiCoO2 powders. Ionics 2012, 18, 39. (62) Madhavi, S.; Subba Rao, G. V.; Chowdari, B. V. R.; Li, S. F. Y. Synthesis and cathodic properties of LiCo1‑yRhyO2 (0 ≤ y ≤ 0.2.) and LiRhO2. J. Electrochem. Soc. 2001, 148, A1279. (63) Reddy, M.V.; Tung, B. D.; Yang, L.; Quang Minh, N. D.; Loh, K.P.; Chowdari, B.V.R. Molten salt method of preparation and cathodic studies on layered-cathode materials Li(Co0.7Ni0.3)O2 and Li(Ni0.7Co0.3)O2 for Li-ion batteries. J. Power Sources 2013, 225, 374. (64) Liang, J.; Wu, D.; Hu, M.; Tian, Y.; Wei, J.; Zhou, Z. Could Li/ Ni disorder be utilized positively? Combined experimental and computational investigation on pillar effect of Ni at Li sites on LiCoO2 at high voltages. Electrochim. Acta 2014, 146, 784. (65) Lu, Z. H.; MacNeil, D. D.; Dahn, J. R. Layered LiNixCo1−2xMnxO2 cathode materials for lithium-ion batteries. Electrochem. Solid-State Lett. 2001, 4, A200. (66) Ohzuku, T.; Makimura, Y. Layered lithium insertion material of LiCo1/3Ni1/3Mn1/3O2 for lithium-ion batteries. Chem. Lett. 2001, 30, 642. (67) Shaju, K.M; Subba Rao, G.V; Chowdari, B.V.R Performance of layered Li(Ni1/3Co1/3Mn1/3)O2 as cathode for Li-ion batteries. Electrochim. Acta 2002, 48, 145. (68) Hwang, B. J.; Tsai, Y. W.; Carlier, D.; Ceder, G. A combined computational/experimental study on LiNi1/3Co1/3Mn1/3O2. Chem. Mater. 2003, 15, 3676. (69) Kim, J. M.; Chung, H. T. The first cycle characteristics of LiNi1/3Co1/3Mn1/3O2 charged up to 4.7 V. Electrochim. Acta 2004, 49, 937. (70) Li, D. C.; Muta, T.; Zhang, L. Q.; Yoshio, M.; Noguchi, H. Effect of synthesis method on the electrochemical performance of LiNi1/3Mn1/3Co1/3O2. J. Power Sources 2004, 132, 150. (71) Park, S. H.; Yoon, C. S.; Kang, S. G.; Kim, H. S.; Moon, S. I.; Sun, Y. K. Synthesis and structural characterization of layered LiNi1/3Co1/3Mn1/3O2 cathode materials by ultrasonic spray pyrolysis method. Electrochim. Acta 2004, 49, 557. (72) Wang, Z. X.; Sun, Y. C.; Chen, L. Q.; Huang, X. J. Electrochemical characterization of positive electrode material LiNi1/3Co1/3Mn1/3O2 and compatibility with electrolyte for lithiumion batteries. J. Electrochem. Soc. 2004, 151, A914. (73) Lee, M. H.; Kang, Y.; Myung, S. T.; Sun, Y. K. Synthetic optimization of Li Ni1/3Co1/3Mn1/3O2 via co-precipitation. Electrochim. Acta 2004, 50, 939. (74) Lee, E.-J.; Chen, Z.; Noh, H.-J.; Nam, S. C.; Kang, S.; Kim, D. H.; Amine, K.; Sun, Y.-K. Development of microstrain in aged lithium transition metal oxides. Nano Lett. 2014, 14, 4873. (75) Lin, F.; Nordlund, D.; Li, Y.; Quan, M. K.; Cheng, L.; Weng, T.-C.; Liu, Y.; Xin, H. L.; Doeff, M. M. Metal segregation in hierarchically structured cathode materials for high-energy lithium batteries. Nat. Energy 2016, 1, 15004. (76) Shen, C.-H.; Wang, Q.; Chen, H.-J.; Shi, C.-G.; Zhang, H.-Y.; Huang, L.; Li, J.-T.; Sun, S.-G. In situ multitechnical investigation into capacity fading of high-voltage LiNi0.5Co0.2Mn0.3O2. ACS Appl. Mater. Interfaces 2016, 8, 35323. (77) Hwang, S.; Jo, E.; Chung, K. Y.; Hwang, K. S.; Kim, S. M.; Chang, W. Structural evolution of LixNiyMnzCo1‑y‑zO2 cathode materials during high-rate charge and discharge. J. Phys. Chem. Lett. 2017, 8, 5758. (78) Kim, J.; Cho, H.; Jeong, H. Y.; Ma, H.; Lee, J.; Hwang, J.; Park, M.; Cho, J. Self-induced concentration gradient in nickel-rich cathodes by sacrificial polymeric bead clusters for high-energy lithium-ion batteries. Adv. Energy Mater. 2017, 7, 1602559. (79) Thackeray, M. M.; Johnson, C. S.; Vaughey, J. T.; Li, N.; Hackney, S. A. Advances in manganese-oxide ’composite’ electrodes for lithium-ion batteries. J. J. Mater. Chem. 2005, 15, 2257. (80) Numata, K.; Sakaki, C.; Yamanaka, S. Synthesis and characterization of layer structured solid solutions in the system of LiCoO2-Li2MnO3. Solid State Ionics 1999, 117, 257.

Co-0.05-O-0.95 as cathode material for lithium rechargeable cells. Electrochem. Solid-State Lett. 2004, 7, A176. (41) Sathiyamoorthi, R.; Shakkthivel, P.; Gangadharan, R.; Vasudevan, T. Layered LiCo1‑xMgxO2 (x = 0.0, 0.1, 0.2, 0.3 and 0.5) cathode materials for lithium-ion rechargeable batteries. Ionics 2007, 13, 25. (42) Mladenov, M.; Stoyanova, R.; Zhecheva, E.; Vassilev, S. Effect of Mg doping and MgO-surface modification on the cycling stability of LiCoO2 electrodes. Electrochem. Commun. 2001, 3, 410. (43) Levasseur, S.; Menetrier, M.; Delmas, C. On the dual effect of Mg doping in LiCoO2 and Li1+deltaCoO2: Structural, electronic properties, and Li-7 MAS NMR studies. Chem. Mater. 2002, 14, 3584. (44) Kim, H.-J.; Jeong, Y. U.; Lee, J.-H.; Kim, J.-J. Crystal structures, electrical conductivities and electrochemical properties of LiCo1‑xMgxO2 (0 ≤ x ≤ 0.11). J. Power Sources 2006, 159, 233. (45) Zhu, X.; Shang, K.; Jiang, X.; Ai, X.; Yang, H.; Cao, Y. Enhanced electrochemical performance of Mg-doped LiCoO2 synthesized by a polymer-pyrolysis method. Ceram. Int. 2014, 40, 11245. (46) Sathiyamoorthi, R.; Shakkthivel, P.; Vasudevan, T. New solidstate synthesis routine and electrochemical properties of calcium based ceramic oxide battery materials for lithium battery applications. Mater. Lett. 2007, 61, 3746. (47) Sathiyamoorthi, R.; Chandrasekaran, R.; Gopalan, A.; Vasudevan, T. Synthesis and electrochemical performance of high voltage cycling LiCo0.8M0.2O2 (M = Mg, Ca, Ba) as cathode material. Mater. Res. Bull. 2008, 43, 1401. (48) Huang, H. T.; Rao, G. V. S.; Chowdari, B. V. R. LiAlxCo1‑xO2 as 4 V cathodes for lithium ion batteries. J. Power Sources 1999, 81, 690. (49) Jang, Y. I.; Huang, B. Y.; Wang, H. F.; Sadoway, D. R.; Ceder, G.; Chiang, Y. M.; Liu, H.; Tamura, H. LiAlyCo1‑yO2 (R(3)over-barm) intercalation cathode for rechargeable lithium batteries. J. Electrochem. Soc. 1999, 146, 862. (50) Yoon, W. S.; Lee, K. K.; Kim, K. B. Structural and electrochemical properties of LiAlyCo1‑yO2 cathode for Li rechargeable batteries. J. Electrochem. Soc. 2000, 147, 2023. (51) Yoon, W. S.; Lee, K. K.; Kim, K. B. X-ray absorption spectroscopic study of LiAlyCo1‑yO2 cathode for Li rechargeable batteries. J. Electrochem. Soc. 2002, 149, A146. (52) Myung, S. T.; Kumagai, N.; Komaba, S.; Chung, H. T. Effects of Al doping on the microstructure of LiCoO2 cathode materials. Solid State Ionics 2001, 139, 47. (53) Yoon, W. S.; Lee, K. K.; Kim, K. B. Synthesis of LiA1(y)Co(1‑y)O(2) using acrylic acid and its electrochemical properties for Li rechargeable batteries. J. Power Sources 2001, 97, 303. (54) Zhou, F.; Luo, W.; Zhao, X.; Dahn, J. R. Relative impact of Al or Mg substitution on the thermal stability of LiCo(1‑z)M(z)O(2) (M = Al or Mg) by accelerating rate calorimetry. J. Electrochem. Soc. 2009, 156, A917. (55) Yoon, W. S.; Kim, K. B.; Kim, M. G.; Lee, M. K.; Shin, H. J.; Lee, J. M. Oxygen contribution on Li-ion intercalation-deintercalation in LiAlyCo1‑yO2 investigated by O K-edge and Co L-edge X-ray absorption spectroscopy. J. Electrochem. Soc. 2002, 149, A1305. (56) Alcantara, R.; Lavela, P.; Tirado, J. L.; Stoyanova, R.; Zhecheva, E. Structure and electrochemical properties of boron-doped LiCoO2. J. Solid State Chem. 1997, 134, 265. (57) Julien, C. M.; Mauger, A.; Groult, H.; Zhang, X.; Gendron, F. LiCo1‑yByO2 as cathode materials for rechargeable lithium batteries. Chem. Mater. 2011, 23, 208. (58) Julien, C.; Nazri, G. A.; Rougier, A. Electrochemical performances of layered LiM1‑yMy ′ O-2 (M = Ni, Co; M ′ = Mg, Al, B) oxides in lithium batteries. Solid State Ionics 2000, 135, 121. (59) Zou, M. J.; Yoshio, M.; Gopukumar, S.; Yamaki, J. Synthesis of high-voltage (4.5 V) cycling doped LiCoO2 for use in lithium rechargeable cells. Chem. Mater. 2003, 15, 4699. (60) Nithya, C.; Thirunakaran, R.; Sivashanmugam, A.; Gopukumar, S. Microwave synthesis of novel high voltage (4.6V) high capacity LiCuxCo1‑xO2 /‑delta cathode material for lithium rechargeable cells. J. Power Sources 2011, 196, 6788. 10136

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research (81) Numata, K.; Yamanaka, S. Preparation and electrochemical properties of layered lithium-cobalt-manganese oxides. Solid State Ionics 1999, 118, 117. (82) Rana, J.; Kloepsch, R.; Li, J.; Scherb, T.; Schumacher, G.; Winter, M.; Banhart, J. On the structural integrity and electrochemical activity of a 0.5Li(2)MnO(3)center dot 0.5LiCoO(2) cathode material for lithium-ion batteries. J. Mater. Chem. A 2014, 2, 9099. (83) Sun, Y.; Shiosaki, Y.; Xia, Y.; Noguchi, H. The preparation and electrochemical performance of solid solutions LiCoO2-Li2MnO3 as cathode materials for lithium ion batteries. J. Power Sources 2006, 159, 1353. (84) Xiang, X.; Knight, J. C.; Li, W.; Manthiram, A. Understanding the influence of composition and synthesis temperature on oxygen loss, reversible capacity, and electrochemical behavior of xLi(2)MnO(3)-(1-x)LiCoO2 Cathodes in the First Cycle. J. Phys. Chem. C 2014, 118, 23553. (85) Jung, H.-G.; Gopal, N. V.; Prakash, J.; Kim, D.-W.; Sun, Y.-K. Improved electrochemical performances of LiM0.05Co0.95O1.95F0.05 (M = Mg, Al, Zr) at high voltage. Electrochim. Acta 2012, 68, 153. (86) Liu, Q.; Su, X.; Lei, D.; Qin, Y.; Wen, J. G.; Guo, F. M.; Wu, Y. M. A.; Rong, Y. C.; Kou, R. H.; Xiao, X. H.; Aguesse, F.; Bareno, J.; Ren, Y.; Lu, W. Q.; Li, Y. X. Approaching the capacity limit of lithium cobalt oxide in lithium ion batteries via lanthanum and aluminium doping. Nat. Energy 2018, 3, 936. (87) Koyama, Y.; Tanaka, I.; Adachi, H.; Makimura, Y.; Ohzuku, T. Crystal and electronic structures of superstructural Li1‑xCo1/3Ni1/3Mn1/3O2 (0 ≤ x ≤ 1). J. Power Sources 2003, 119, 644. (88) Wang, Z.; Wang, Z.; Guo, H.; Peng, W.; Li, X. Synthesis of Li2MnO3-stabilized LiCoO2 cathode material by spray-drying method and its high-voltage performance. J. Alloys Compd. 2015, 626, 228. (89) Kim, S.; Kim, C.; Jhon, Y.-I.; Noh, J.-K.; Vemuri, S. H.; Smith, R.; Chung, K. Y.; Jhon, M. S.; Cho, B.-W. Synthesis of layered-layered 0.5Li(2)MnO(3)center dot 0.5LiCoO(2) nanocomposite electrode materials by the mechanochemical process and first principles study. J. Mater. Chem. 2012, 22, 25418. (90) Tebbe, J. L.; Holder, A. M.; Musgrave, C. B. Mechanisms of LiCoO2 cathode degradation by reaction with HF and protection by thin oxide coatings. ACS Appl. Mater. Interfaces 2015, 7, 24265. (91) Cho, J.; Kim, C. S.; Yoo, S. I. Improvement of structural stability of LiCoO2 cathode during electrochemical cycling by sol-gel coating of SnO2. Electrochem. Solid-State Lett. 1999, 3, 362. (92) Scott, I. D.; Jung, Y. S.; Cavanagh, A. S.; Yan, Y.; Dillon, A. C.; George, S. M.; Lee, S.-H. Ultrathin coatings on nano-LiCoO2 for Liion vehicular applications. Nano Lett. 2011, 11, 414. (93) Liu, L. J.; Wang, Z. X.; Li, H.; Chen, L. Q.; Huang, X. J. Al2O3coated LiCoO2 as cathode material for lithium ion batteries. Solid State Ionics 2002, 152-153, 341. (94) Oh, S.; Lee, J. K.; Byun, D.; Cho, W. I.; Cho, B. W. Effect of Al2O3 coating on electrochemical performance of LiCoO2 as cathode materials for secondary lithium batteries. J. Power Sources 2004, 132, 249. (95) Oh, Y.; Ahn, D.; Nam, S.; Park, B. The effect of Al2O3-coating coverage on the electrochemical properties in LiCoO2 thin films. J. Solid State Chem. 2010, 14, 1235. (96) Daheron, L.; Dedryvere, R.; Martinez, H.; Flahaut, D.; Menetrier, M.; Delmas, C.; Gonbeau, D. Possible explanation for the efficiency of Al-based coatings on LiCoO2: Surface properties of LiCo1‑xAlxO2 solid solution. Chem. Mater. 2009, 21, 5607. (97) Jung, S. C.; Han, Y.-K. How do Li atoms pass through the Al2O3 coating layer during lithiation in Li-ion batteries? J. Phys. Chem. Lett. 2013, 4, 2681. (98) Han, B.; Paulauskas, T.; Key, B.; Peebles, C.; Park, J. S.; Klie, R. F.; Vaughey, J. T.; Dogan, F. Understanding the role of temperature and cathode composition on interface and bulk: Optimizing aluminum oxide coatings for Li-ion cathodes. ACS Appl. Mater. Interfaces 2017, 9, 14769. (99) Wang, Z. X.; Wu, C. A.; Liu, L. J.; Wu, F.; Chen, L. Q.; Huang, X. J. Electrochemical evaluation and structural characterization of

commercial LiCoO2 surfaces modified with MgO for lithium-ion batteries. J. Electrochem. Soc. 2002, 149, A466. (100) Shim, J.-H.; Lee, S.; Park, S. S. Effects of MgO coating on the structural and electrochemical characteristics of LiCoO2 as cathode materials for lithium ion battery. Chem. Mater. 2014, 26, 2537. (101) Orikasa, Y.; Takamatsu, D.; Yamamoto, K.; Koyama, Y.; Mori, S.; Masese, T.; Mori, T.; Minato, T.; Tanida, H.; Uruga, T.; Ogumi, Z.; Uchimoto, Y. Origin of surface coating effect for MgO on LiCoO2 to improve the interfacial reaction between electrode and electrolyte. Adv. Mater. Interfaces 2014, 1, 1400195. (102) Chen, Z. H.; Dahn, J. R. Effect of a ZrO2 coating on the structure and electrochemistry of LixCoO2 when cycled to 4.5 V. Electrochem. Solid-State Lett. 2002, 5, A213. (103) Miyashiro, H.; Yamanaka, A.; Tabuchi, M.; Seki, S.; Nakayama, M.; Ohno, Y.; Kobayashi, Y.; Mita, Y.; Usami, A.; Wakihara, M. Improvement of degradation at elevated temperature and at high state-of-charge storage by ZrO2 coating on LiCoO2. J. Electrochem. Soc. 2006, 153, A348. (104) Takamatsu, D.; Mori, S.; Orikasa, Y.; Nakatsutsumi, T.; Koyama, Y.; Tanida, H.; Arai, H.; Uchimoto, Y.; Ogumi, Z. Effects of ZrO2 coating on LiCoO2 thin-film electrode studied by in situ x-ray absorption spectroscopy. J. Electrochem. Soc. 2013, 160, A3054. (105) Fang, T.; Duh, J. G.; Sheen, S. R. Improving the electrochemical performance of LiCoO2 cathode by nanocrystalline ZnO coating. J. Electrochem. Soc. 2005, 152, A1701. (106) Dai, X.; Wang, L.; Xu, J.; Wang, Y.; Zhou, A.; Li, J. Improved electrochemical performance of LiCoO2 electrodes with ZnO coating by radio frequency magnetron sputtering. ACS Appl. Mater. Interfaces 2014, 6, 15853. (107) Dai, X.; Zhou, A.; Xu, J.; Yang, B.; Wang, L.; Li, J. Superior electrochemical performance of LiCoO2 electrodes enabled by conductive Al2O3-doped ZnO coating via magnetron sputtering. J. Power Sources 2015, 298, 114. (108) Shen, B.; Zuo, P.; Li, Q.; He, X.; Yin, G.; Ma, Y.; Cheng, X.; Du, C.; Gao, Y. Lithium cobalt oxides functionalized by conductive Al-doped ZnO coating as cathode for high-performance lithium ion batteries. Electrochim. Acta 2017, 224, 96. (109) Hao, Q.; Ma, H.; Ju, Z.; Li, G.; Li, X.; Xu, L.; Qian, Y. NanoCuO coated LiCoO2: Synthesis, improved cycling stability and good performance at high rates. Electrochim. Acta 2011, 56, 9027. (110) Hao, Q.; Xu, C.; Jia, S.; Zhao, X. Improving the cycling stability of LiCoO2 at 4.5 V through surface modification by Fe2O3 coating. Electrochim. Acta 2013, 113, 439. (111) Sun, Y. K.; Cho, S. W.; Myung, S. T.; Amine, K.; Prakash, J. Effect of AlF(3) coating amount on high voltage cycling performance of LiCoO2. Electrochim. Acta 2007, 53, 1013. (112) Sun, Y.-K.; Yoon, C. S.; Myung, S.-T.; Belharouak, I.; Amine, K. Role of AlF3 coating on LiCoO2 particles during cycling to cut-off voltage above 4.5 V. J. Electrochem. Soc. 2009, 156, A1005. (113) Lee, K.-S.; Myung, S.-T.; Kim, D.-W.; Sun, Y.-K. AlF3-coated LiCoO2 and Li[Ni1/3Co1/3Mn1/3]O2 blend composite cathode for lithium ion batetries. J. Electrochem. Soc. 2011, 196, 6974. (114) Zhou, Y.; Lee, Y.; Sun, H.; Wallas, J. M.; George, S. M.; Xie, M. Coating solution for high-voltage cathode: AlF3 atomic layer deposition for freestanding LiCoO2 electrodes with high energy density and excellent flexibility. ACS Appl. Mater. Interfaces 2017, 9, 9614. (115) Yang, Z.; Qiao, Q.; Yang, W. Improvement of structural and electrochemical properties of commercial LiCoO2 by coating with LaF3. Electrochim. Acta 2011, 56, 4791. (116) Aboulaich, A.; Ouzaouit, K.; Faqir, H.; Kaddami, A.; Benzakour, I.; Akalay, I. Improving thermal and electrochemical performances of LiCoO2 cathode at high cut-off charge potentials by MF3 (M = Ce, Al) coating. Mater. Res. Bull. 2016, 73, 362. (117) Qian, J.; Liu, L.; Yang, J.; Li, S.; Wang, X.; Zhuang, H. L.; Lu, Y. Electrochemical surface passivation of LiCoO2 particles at ultrahigh voltage and its applications in lithium-based batteries. Nat. Commun. 2018, 9, 4918. 10137

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research

(138) Shen, B.; Liu, Q.; Wang, L.; Yin, G.; Zuo, P.; Ma, Y.; Cheng, X.; Du, C.; Gao, Y. Mixed lithium ion and electron conducting LiAlPO3.93F1.07-coated LiCoO2 cathode with improved electrochemical performance. Electrochem. Commun. 2017, 83, 106. (139) Cho, J. P. Direct micron-sized LiMn2O4 particle coating on LiCoO2 cathode material using surfactant. Solid State Ionics 2003, 160, 241. (140) Yoon, W.-S.; Chung, K. Y.; Nam, K.-W.; Kim, K.-B. Characterization of LiMn2O4-coated LiCoO2 film electrode prepared by electrostatic spray deposition. J. Power Sources 2006, 163, 207. (141) Fey, G. T. K.; Wang, Z. F.; Lu, C. Z.; Kumar, T. P. MgAl2O4 spinel-coated LiCoO2 as long-cycling cathode materials. J. Power Sources 2005, 146, 245. (142) Xu, K.; Angell, C. A. Sulfone-based electrolytes for lithium-ion batteries. J. Electrochem. Soc. 2002, 149, A920. (143) Watanabe, Y.; Kinoshita, S.-I.; Wada, S.; Hoshino, K.; Morimoto, H.; Tobishima, S.-I. Electrochemical properties and lithium ion solvation behavior of sulfone-ester mixed electrolytes for high-voltage rechargeable lithium cells. J. Power Sources 2008, 179, 770. (144) Abouimrane, A.; Belharouak, I.; Amine, K. Sulfone-based electrolytes for high-voltage Li-ion batteries. Electrochem. Commun. 2009, 11, 1073. (145) Zhou, Q.; Henderson, W. A.; Appetecchi, G. B.; Montanino, M.; Passerini, S. Physical and electrochemical properties of N-alkyl-Nmethylpyrrolidinium bis(fluorosulfonyl)imide ionic liquids: PY13FSI and PY14FSI. J. Phys. Chem. B 2008, 112, 13577. (146) Sato, T.; Maruo, T.; Marukane, S.; Takagi, K. Ionic liquids containing carbonate solvent as electrolytes for lithium ion cells. J. Power Sources 2004, 138, 253. (147) Lin, X.; Varela, J. C.; Grinstaff, M. W. Synthesis of ionic liquid based electrolytes, assembly of Li-ion batteries, and measurements of performance at high temperature. J. Visualized Exp. 2016, 118, e54864. (148) Xie, B.; Mai, Y.; Wang, J.; Luo, H.; Yan, X.; Zhang, L. Dinitrile compound containing ethylene oxide moiety with enhanced solubility of lithium salts as electrolyte solvent for high-voltage lithium-ion batteries. Ionics 2015, 21, 909. (149) Kim, K.; Park, I.; Ha, S.-Y.; Kim, Y.; Woo, M.-H.; Jeong, M.H.; Shin, W. C.; Ue, M.; Hong, S. Y.; Choi, N.-S. Understanding the thermal instability of fluoroethylene carbonate in LiPF6-based electrolytes for lithium ion batteries. Electrochim. Acta 2017, 225, 358. (150) Liu, J.; Liu, N.; Liu, D.; Bai, Y.; Shi, L.; Wang, Z.; Chen, L.; Hennige, V.; Schuch, A. Improving the performances of LiCoO2 cathode materials by soaking nano-alumina in commercial electrolyte. J. Electrochem. Soc. 2007, 154, A55. (151) Lu, Y.; Korf, K.; Kambe, Y.; Tu, Z.; Archer, L. A. Ionic-liquidnanoparticle hybrid electrolytes: Applications in lithium metal batteries. Angew. Chem., Int. Ed. 2014, 53, 488. (152) Lu, Y.; Das, S. K.; Moganty, S. S.; Archer, L. A. Ionic liquidnanoparticle hybrid electrolytes and their application in secondary lithium-metal batteries. Adv. Mater. 2012, 24, 4430. (153) Abe, K.; Ushigoe, Y.; Yoshitake, H.; Yoshio, M. Functional electrolytes: Novel type additives for cathode materials, providing high cycleability performance. J. Power Sources 2006, 153, 328. (154) Abe, K.; Takaya, T.; Yoshitake, H.; Ushigoe, Y.; Yoshio, M.; Wang, H. Functional electrolyte: Additives for improving the cyclability of cathode materials. Electrochem. Solid-State Lett. 2004, 7, A462. (155) Takeuchi, T.; Kyuna, T.; Morimoto, H.; Tobishima, S.-i. Influence of surface modification of LiCoO2 by organic compounds on electrochemical and thermal properties of Li/LiCoO2 rechargeable cells. J. Power Sources 2011, 196, 2790. (156) Lee, K.-S.; Sun, Y.-K.; Noh, J.; Song, K. S.; Kim, D.-W. Improvement of high voltage cycling performance and thermal stability of lithium−ion cells by use of a thiophene additive. Electrochem. Commun. 2009, 11, 1900.

(118) Hao, S.; Wolverton, C. Lithium transport in amorphous Al2O3 and AlF3 for discovery of battery coatings. J. Mater. Chem. C 2013, 117, 8009. (119) Cho, J. Improved thermal stability of LiCoO2 by nanoparticle AlPO4 coating with respect to spinel Li1.05Mn1.95O4. Electrochem. Commun. 2003, 5, 146. (120) Cho, J. Correlation between AlPO4 nanoparticle coating thickness on LiCoO2 cathode and thermal stability. Electrochim. Acta 2003, 48, 2807. (121) 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 AlPO(4) nanoparticles. Angew. Chem., Int. Ed. 2003, 42, 1618. (122) Cho, J.; Lee, J. G.; Kim, B.; Park, B. Effect of P2O5 and AlPO4 coating on LiCoO2 cathode material. Chem. Mater. 2003, 15, 3190. (123) Cho, J. Dependence of AlPO4 coating thickness on overcharge behaviour of LiCoO2 cathode material at 1 and 2 C rates. J. Power Sources 2004, 126, 186. (124) Cho, J.; Kim, H.; Park, B. Comparison of overcharge behavior of AlPO4-coated LiCoO2 and Li(Ni)(Co)(Mn)(O)(0.8)(0.1)(0.1)(2) cathode materials in Li-ion cells. J. Electrochem. Soc. 2004, 151, A1707. (125) Cho, J.; Kim, T. G.; Kim, C.; Lee, J. G.; Kim, Y. W.; Park, B. Comparison of Al2O3- and AlPO4-coated LiCoO2 cathode materials for a Li-ion cell. J. Power Sources 2005, 146, 58. (126) Cho, J.; Lee, J. G.; Kim, B.; Kim, T. G.; Kim, J.; Park, B. Control of AlPO4-nanoparticle coating on LiCoO2 by using water or ethanol. Electrochim. Acta 2005, 50, 4182. (127) Lee, J. G.; Kim, B.; Cho, J.; Kim, Y. W.; Park, B. Effect of AlPO4-nanoparticle coating concentration on high-cutoff-voltage electrochemical performances in LiCoO2. J. Electrochem. Soc. 2004, 151, A801. (128) Kim, J.; Noh, M.; Cho, J.; Kim, H.; Kim, K. B. Controlled nanoparticle metal phosphates (metal = Al, Fe, Ce, and Sr) coatings on LiCoO2 cathode materials. J. Electrochem. Soc. 2005, 152, A1142. (129) Appapillai, A. T.; Mansour, A. N.; Cho, J.; Shao-Horn, Y. Microstructure of LiCoO2 with and without ″AlPO(4)″ nanoparticle coating: Combined STEM and XPS studies. Chem. Mater. 2007, 19, 5748. (130) Lu, Y.-C.; Mansour, A. N.; Yabuuchi, N.; Shao-Horn, Y. Probing the origin of enhanced stability of ″AlPO(4)″ nanoparticle coated LiCoO2 during cycling to high voltages: Combined XRD and XPS studies. Chem. Mater. 2009, 21, 4408. (131) Lee, H.; Kim, M. G.; Cho, J. Olivine LiCoPO4 phase grown LiCoO2 cathode material for high density Li batteries. Electrochem. Commun. 2007, 9, 149. (132) Li, G.; Yang, Z.; Yang, W. Effect of FePO4 coating on electrochemical and safety performance of LiCoO2 as cathode material for Li-ion batteries. J. Power Sources 2008, 183, 741. (133) Morimoto, H.; Awano, H.; Terashima, J.; Nakanishi, S.; Hirama, Y.; Ishikawa, K.; Tobishima, S.-i. Charge-discharge properties of LiCoO2 electrodes modified by olivine-type compounds of LiMgPO4 for lithium secondary batteries. J. Power Sources 2012, 211, 66. (134) Morimoto, H.; Awano, H.; Terashima, J.; Shindo, Y.; Nakanishi, S.; Ito, N.; Ishikawa, K.; Tobishima, S.-i. Preparation of lithium ion conducting solid electrolyte of NASICON-type Li1+xAlxTi2‑x(PO4)(3) (x = 0.3) obtained by using the mechanochemical method and its application as surface modification materials of LiCoO2 cathode for lithium cell. J. Power Sources 2013, 240, 636. (135) Zhou, A.; Xu, J.; Dai, X.; Yang, B.; Lu, Y.; Wang, L.; Fan, C.; Li, J. Improved high-voltage and high temperature electrochemical performance of LiCoO2 cathode by electrode sputter-coating with Li3PO4. J. Power Sources 2016, 322, 10. (136) Bai, Y.; Liu, N.; Liu, J.; Wang, Z.; Chen, L. Coating materialinduced acidic electrolyte improves LiCoO2 performances. Electrochem. Solid-State Lett. 2006, 9, A552. (137) Bai, Y.; Yin, Y.; Liu, N.; Guo, B.; Shi, H.; Liu, J.; Wang, Z.; Chen, L. New concept of surface modification to LiCoO2. J. Power Sources 2007, 174, 328. 10138

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139

Review

Industrial & Engineering Chemistry Research (157) Xia, L.; Xia, Y.; Liu, Z. Thiophene derivatives as novel functional additives for high-voltage LiCoO2 operations in lithium ion batteries. Electrochim. Acta 2015, 151, 429. (158) Ji, Y.; Zhang, P.; Lin, M.; Zhao, W.; Zhang, Z.; Zhao, Y.; Yang, Y. Toward a stable electrochemical interphase with enhanced safety on high-voltage LiCoO2 cathode: A case of phosphazene additives. J. Power Sources 2017, 359, 391. (159) Xia, L.; Xia, Y.; Liu, Z. A novel fluorocyclophosphazene as bifunctional additive for safer lithium-ion batteries. J. Power Sources 2015, 278, 190. (160) Lee, J.-N.; Han, G.-B.; Ryou, M.-H.; Lee, D. J.; Song, J.; Choi, J. W.; Park, J.-K. N-(triphenylphosphoranylidene) aniline as a novel electrolyte additive for high voltage LiCoO2 operations in lithium ion batteries. Electrochim. Acta 2011, 56, 5195. (161) He, Y.-B.; Liu, Q.; Tang, Z.-Y.; Chen, Y.-H.; Song, Q.-S. The cooperative effect of tri(β-chloromethyl) phosphate and cyclohexyl benzene on lithium ion batteries. Electrochim. Acta 2007, 52, 3534. (162) Zhou, M.; Qin, C.; Liu, Z.; Feng, L.; Su, X.; Chen, Y.; Xia, L.; Xia, Y.; Liu, Z. Enhanced high voltage cyclability of LiCoO2 cathode by adopting poly bis-(ethoxyethoxyethoxy)phosphazene with flameretardant property as an electrolyte additive for lithium-ion batteries. Appl. Surf. Sci. 2017, 403, 260. (163) Zhao, M.; Zuo, X.; Ma, X.; Xiao, X.; Yu, L.; Nan, J. Diphenyl disulfide as a new bifunctional film-forming additive for high-voltage LiCoO2/graphite battery charged to 4.4 V. J. Power Sources 2016, 323, 29. (164) Zuo, X.; Fan, C.; Xiao, X.; Liu, J.; Nan, J. High-voltage performance of LiCoO2/graphite batteries with methylene methanedisulfonate as electrolyte additive. J. Power Sources 2012, 219, 94. (165) Xia, J.; Petibon, R.; Sinha, N. N.; Dahn, J. R. One sulfonate and three sulfate electrolyte additives studied in graphite/LiCoO2 pouch cells. J. Electrochem. Soc. 2015, 162, A2227. (166) Xu, K.; Lee, U.; Zhang, S. S.; Allen, J. L.; Jow, T. R. Graphite/ electrolyte interface formed in LiBOB-based electrolytes - I. Differentiating the roles of EC and LiBOB in SEI formation. Electrochem. Solid-State Lett. 2004, 7, A273. (167) Pang, C.; Xu, G.; An, W.; Ding, G.; Liu, X.; Chai, J.; Ma, J.; Liu, H.; Cui, G. Three-component functional additive in a LiPF6based carbonate electrolyte for a high-voltage LiCoO2/graphite battery system. Energy Technology 2017, 5, 1979. (168) Wang, X.; Xing, L.; Liao, X.; Chen, X.; Huang, W.; Yu, Q.; Xu, M.; Huang, Q.; Li, W. Improving cyclic stability of lithium cobalt oxide based lithium ion battery at high voltage by using trimethylboroxine as an electrolyte additive. Electrochim. Acta 2015, 173, 804. (169) Wang, X.; Zheng, X.; Liao, Y.; Huang, Q.; Xing, L.; Xu, M.; Li, W. Maintaining structural integrity of 4.5V lithium cobalt oxide cathode with fumaronitrile as a novel electrolyte additive. J. Power Sources 2017, 338, 108. (170) Xie, Y.; Gao, H.; Gim, J.; Ngo, A. T.; Ma, Z. F.; Chen, Z. Identifying Active Sites for Parasitic Reactions at the CathodeElectrolyte Interface. J. Phys. Chem. Lett. 2019, 10, 589. (171) Kim, G.-Y.; Petibon, R.; Dahn, J. R. Effects of succinonitrile (SN) as an electrolyte additive on the impedance of LiCoO2/graphite pouch cells during cycling. J. Electrochem. Soc. 2014, 161, A506. (172) Wang, X.; Liao, X.; Huang, W.; Xing, L.; Liao, Y.; Huang, Q.; Xu, M.; Li, W. Improved cyclic stability of layered lithium cobalt oxide at high potential via cathode electrolyte interphase formed by 4(trifluoromethyl) benzonitrile. Electrochim. Acta 2015, 184, 94. (173) Wang, D. Y.; Sinha, N. N.; Petibon, R.; Burns, J. C.; Dahn, J. R. A systematic study of well-known electrolyte additives in LiCoO2/ graphite pouch cells. J. Power Sources 2014, 251, 311. (174) Ji, Y.; Li, S.; Zhong, G.; Zhang, Z.; Li, Y.; McDonald, M. J.; Yang, Y. Synergistic effects of suberonitrile-LiBOB binary additives on the electrochemical performance of high-voltage LiCoO2 electrodes. J. Electrochem. Soc. 2015, 162, A7015. (175) Kadir, M. F. Z. A.; Teo, L. P.; Majid, S. R.; Arof, A. K. Conductivity studies on plasticised PEO/chitosan proton conducting polymer electrolyte. Mater. Res. Innovations 2009, 13, 259.

(176) Ran, J.; Wu, L.; Ru, Y.; Hu, M.; Din, L.; Xu, T. Anion exchange membranes (AEMs) based on poly(2,6-dimethyl-1,4phenylene oxide) (PPO) and its derivatives. Polym. Chem. 2015, 6, 5809. (177) Voigt, N.; van Wuellen, L. The mechanism of ionic transport in PAN-based solid polymer electrolytes. Solid State Ionics 2012, 208, 8. (178) Kufian, M. Z.; Aziz, M. F.; Shukur, M. F.; Rahim, A. S.; Ariffin, N. E.; Shuhaimi, N. E. A.; Majid, S. R.; Yahya, R.; Arof, A. K. PMMALiBOB gel electrolyte for application in lithium ion batteries. Solid State Ionics 2012, 208, 36. (179) Janakiraman, S.; Surendran, A.; Ghosh, S.; Anandhan, S.; Venimadhav, A. A new strategy of PVDF based Li-salt polymer electrolyte through electrospinning for lithium battery application. Mater. Res. Express 2019, 6, 035303. (180) Sundaram, N. T. K.; Musthafa, O. T. M.; Lokesh, K. S.; Subramania, A. Effect of porosity on PYdF-co-HFP-PMMA-based electrolyte. Mater. Chem. Phys. 2008, 110, 11. (181) Dyartanti, E. R.; Purwanto, A.; Widiasa, I. N.; Susanto, H. Ionic conductivity and cycling stability improvement of PVDF/nanoclay using PVP as polymer electrolyte membranes for LiFePO4 batteries. Membranes 2018, 8, 36. (182) Kim, H. S.; Periasamy, P.; Moon, S. I. Electrochemical properties of the Li-ion polymer batteries with P(VdF-co-HFP)-based gel polymer electrolyte. J. Power Sources 2005, 141, 293. (183) Dong, T.; Zhang, J.; Xu, G.; Chai, J.; Du, H.; Wang, L.; Wen, H.; Zang, X.; Du, A.; Jia, Q.; Zhou, X.; Cui, G. A multifunctional polymer electrolyte enables ultra-long cycle-life in a high-voltage lithium metal battery. Energy Environ. Sci. 2018, 11, 1197. (184) Rao, M. M.; Geng, X. Y.; Liao, Y. H.; Hu, S. J.; Li, W. S. Preparation and performance of gel polymer electrolyte based on electrospun polymer membrane and ionic liquid for lithium ion battery. J. Membr. Sci. 2012, 399, 37. (185) Choi, J. A.; Eo, S. M.; MacFarlane, D. R.; Forsyth, M.; Cha, E.; Kim, D. W. Effect of organic additives on the cycling performances of lithium metal polymer cells. J. Power Sources 2008, 178, 832. (186) Zhou, W.; Wang, Z.; Pu, Y.; Li, Y.; Xin, S.; Li, X.; Chen, J.; Goodenough, J. B. Double-layer polymer electrolyte for high-voltage all-solid-state reachargeble batteies. Adv. Energy Mater. 2019, 31, 1805574. (187) Choi, H.; Kim, H. W.; Ki, J.-K.; Lim, Y. J.; Kim, Y.; Ahn, J.-H. Nanocomposite quasi-solid-state electrolyte for high-safety lithium batteries. Nano Research 2017, 10, 3092. (188) Appetecchi, G. B.; Romagnoli, P.; Scrosati, B. Composite gel membranes: A new class of improved polymer electrolytes for lithium batteries. Electrochem. Commun. 2001, 3, 281. (189) Li, M.; Zhu, W.; Zhang, P.; Chao, Y.; He, Q.; Yang, B.; Li, H.; Borisevich, A.; Dai, S. Graphene-analogues boron nitride nanosheets confining ionic liquids: A high-performance quasi-liquid solid electrolyte. Small 2016, 12, 3535. (190) Tsao, C. H.; Su, H. M.; Huang, H. T.; Kuo, P. L.; Teng, H. Immobilized cation functional gel polymer electrolytes with high lithium transference number for lithium ion batteries. J. Membr. Sci. 2019, 572, 382. (191) Xu, B.; Duan, H.; Liu, H.; Wang, C. A.; Zhong, S. Stabilization of garnet/liquid electrolyte interface using superbase additives for hybrid Li batteries. ACS Appl. Mater. Interfaces 2017, 9, 21077. (192) Wang, C.; Sun, Q.; Liu, Y.; Zhao, Y.; Li, X.; Lin, X.; Banis, M. N.; Li, M.; Li, W.; Adair, K. R.; Wang, D.; Liang, J.; Li, R.; Zhang, L.; Yang, R.; Lu, S.; Sun, X. Boosting the performance of lithium batteries with solid-liquid hybrid electrolytes: Interfacial properties and effects of liquid electrolytes. Nano Energy 2018, 48, 35. (193) Kitaura, H.; Hayashi, A.; Ohtomo, T.; Hama, S.; Tatsumisago, M. Fabrication of electrode-electrolyte interfaces in all-solid-state rechargeable lithium batteries by using a supercooled liquid state of the glassy electrolytes. J. Mater. Chem. 2011, 21, 118.

10139

DOI: 10.1021/acs.iecr.9b01236 Ind. Eng. Chem. Res. 2019, 58, 10119−10139