Metal Oxides and Oxysalts as Anode Materials ... - ACS Publications

Peng MeiJaewoo LeeMalay PramanikAbdulmohsen AlshehriJeonghun KimJoel ...... Mixed Oxides, (Ni1–xZnx)Fe2O4 (x = 0, 0.25, 0.5, 0.75, 1): Molten Salt S...
2 downloads 0 Views 66MB Size
Review pubs.acs.org/CR

Metal Oxides and Oxysalts as Anode Materials for Li Ion Batteries M. V. Reddy, G. V. Subba Rao, and B. V. R. Chowdari* Department of Physics, Solid State Ionics & Advanced Batteries Lab, National University of Singapore, Singapore- 117 542 3.4. Other Group IV Oxides 3.4.1. SiO and SiO2 3.4.2. GeO2 and Germanates 3.5. Group II and III Oxides, ZnO, CdO, and In2O3 4. Anodes Based on Conversion (Redox) Reaction 4.1. Binary Oxides 4.1.1. Oxides with Rock Salt Structure (MO; M = Mn, Fe, Co, Ni, or Cu) 4.1.2. Oxides with Spinel Structure (M3O4, M = Co, Fe, or Mn) 4.1.3. Oxides, M2O3, with Corundum Structure (M=Fe, Cr or Mn) 4.1.4. Metal Dioxides, MO2 (M=Mn, Mo or Ru) 4.2. Ternary Oxides 4.2.1. Oxides with Spinel Structure (AB2O4, A = Mn, Fe, Co, Ni, or Cu; B  Mn, Fe, Co, Ni, or Cu; A ≠ B) 4.3. Complex Oxides 4.3.1. Oxides of Vanadium Possessing Spinel and Other Structures 4.3.2. Oxides of Molybdenum with Scheelite and Other Structures 4.3.3. Oxides with CaFe2O4, Brownmillerite and Perovskite Structure 4.3.4. Metal Borates: MBO3 (M = Fe or Cr), Fe3BO6, and M3B2O6 (M = Co, Ni, or Cu) 4.3.5. Metal Oxysalts: Carbonates, Oxalates, Oxyhydroxides, and Oxyfluorides 5. Anodes based on Both Alloying−Dealloying and Conversion Reaction 5.1. Oxides with Spinel Structure: ZnM2O4 (M = Co, Fe) and CdFe2O4 5.2. Oxides of Tin Adopting the Spinel Structure 5.3. Metal Oxysalts: Carbonates and Oxalates 6. Li ion Batteries (Cells) with Oxide Anodes 6.1. LIBs with Li4Ti5O12 (LTO) as Anode 6.1.1. LIBs with 4 V cathodes 6.1.2. LIBs with 3.5 V cathode, LiFePO4 6.1.3. LIBs with 4 V Cathodes, LiVPO4F and Na3V2P2O8F3 6.1.4. LIBs with 5.0 V Cathode, Li(Ni0.5Mn1.5)O4 6.2. LIBs with TiO2-B as Anode with Cathodes, LiFePO4 and Li(Ni0.5Mn1.5)O4 6.3. LIBs with Amorphous Tin Composite Oxide, SnO2 and GeO x, with Cathodes LiCoO2, Li(Ni0.8Co0.2)O2, and Li(Ni1/3Co1/3Mn1/3)O2 6.4. LIBs with CoO and Co3O4 as Anodes with Cathodes, LiCoO2 and LiMn2O4

CONTENTS 1. Introduction 1.1. Lithium Ion Batteries 1.2. High-Energy and High-Power Applications of LIBs 1.3. Future-Generation Cathodes 1.4. Ideal Anode for LIBs 1.5. Future-Generation Anodes 1.6. Scope of the Review and Nomenclature 2. Anodes Based on Li Intercalation−Deintercalation Reaction 2.1. Binary Oxides 2.1.1. TiO2 2.1.2. Vanadium (V) and Molybdenum (Mo) Oxides 2.1.3. Nb2O5 2.2. Ternary Oxides of Ti and Nb 2.2.1. Li4Ti5O12 (LTO) 2.2.2. MgTi2O5, LiTiNbO5, TiNb2O7, and Other Oxides 3. Anodes Based on Alloying−Dealloying Reaction 3.1. Binary Tin Oxides 3.1.1. SnO 3.1.2. SnO2 3.2. Ternary Tin Oxides 3.2.1. M2SnO4 (M = Metal) 3.2.2. ASnO3 (A = Ca, Sr, Ba, Co, and Mg) 3.2.3. Li2SnO3 3.2.4. A2Sn2O7 (A = Y or Nd) 3.2.5. K2(M,Sn)8O16 (M = Li, Mg, Fe, Mn, Co, or In) 3.2.6. SnP 2 O 7 , LiSn 2 P 3 O 12 , Sn 2 P 2 O 7 , and Sn3P2O8 3.2.7. Amorphous Tin Composite Oxides (ATCOs) 3.3. Antimony Oxides and Mixed Oxides 3.3.1. Sb2O3 3.3.2. MSb2O6, M = Co, Ni, and Cu 3.3.3. VSbO4, (M1/2Sb1/2Sn)O4 (M = V, Fe, or In), BiSbO4, SbPO4, and MSb2O4 (M = Ni or Co) © 2013 American Chemical Society

5365 5365 5365 5366 5366 5367 5368 5368 5368 5368 5377 5377 5377 5378 5380 5382 5384 5384 5385 5390 5390 5390 5392 5392 5392 5393 5393 5393 5393 5393

5396 5396 5396 5397 5398 5398 5398 5404 5407 5412 5415

5415 5417 5417 5418 5423 5423 5424 5424 5424 5427 5428 5429 5429 5430 5431 5432 5432 5433

5435 5435

Received: March 29, 2011 Published: April 2, 2013

5394 5364

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews 6.5. LIB with Fe2O3 as the Anode and LiFePO4 as the Cathode 6.6. Mitigating the ICL Associated with the Oxide Anodes 7. Conclusions Author Information Corresponding Author Notes Biographies Acknowledgments References Note Added after ASAP Publication

Review

books1−5 have appeared periodically. As can be expected, a large number of review articles have also been published. The 2010 annual worldwide production of LIBs was worth about US $11.8 billion.9 Of these, about 60% were for use in mobile phones, and the rest were for the notebook computers, power tools, medical, and other uses. With the expected use of large-scale LIBs for vehicles and stationary off-peak energy storage systems and other applications, the industry-projected demand and growth is expected to dramatically increase to a market value of US $31.4 billion in 2015 and up to US $53.7 billion in 2020.9

5436 5436 5437 5437 5437 5437 5438 5438 5439 5457

1.2. High-Energy and High-Power Applications of LIBs

LIB packs, in principle, can satisfy the requirements of both high-power and high-energy-density applications. For the past 10 years, they have been explored and implemented for vehicular transport. The latter are broadly classified into pure electric (EV), plug-in hybrid (PHEV), and hybrid (HEV) vehicles and are expected to replace or complement the internal combustion (IC) engine, which uses fossil fuels, gasoline or diesel. Both EV and PHEV use electric motors for vehicular propulsion. The PHEV has an electric motor and IC engine combination, in the so-called serial configuration, and the IC engine is activated only after the battery power is used up, to run a generator and produce electricity for running the electric motor. The HEV also has an electric motor and IC engine combination, but in the so-called parallel configuration. While the electric motor is used for starting the vehicle and low-speed propulsion, the IC engine takes over for higher speed and long distance propulsion. Thus, the HEV needs a high-power LIB pack, whereas the EV and PHEV require a high-energy-density LIB pack. Generally, to match the IC engine-powered vehicle, the EV needs a battery pack that can deliver 25−45 kW·h. This is 2 times the battery size of a PHEV and 10 times that of the HEV. The commercial HEVs presently use Ni-MH battery packs due to safety and other considerations. Vigorous efforts are underway, at present, to replace them with LIB packs. In this regard, low-energy cathodes like LiFePO4 (3.5 V) and lowenergy anodes like Li4Ti5O12 (1.5 V) are the preferred choice for the LIB packs for HEVs.10 For high-power and high-energy-density applications, in addition to vehicular transport as described above, LIBs are also actively being considered for power tools, back-up power supply units, and off-peak energy storage (load leveling) from the electric grid. For all the above uses, the LIBs need to satisfy four important criteria: (1) Cost reduction. The LIBs are approximately three times more expensive than the nearest competitor, namely, Ni-metal hydride batteries. This is because of the expensive (and toxic) cobalt present in the cathode, the manufacturing process of the anode (microcrystalline purified graphite or specialty graphite, like mesocarbon microbeads (MCMB)) and the expensive Li salt, LiPF6. Thus, there is a need to find cheaper electrode and electrolyte materials, because these constitute 30−40% of the total cost of the final product. (2) Improvement in the energy density from ∼120 to ∼250 W·h kg−1. The energy density of the LIBs depends on the capacities and operating potentials of the respective electrode materials, and at present, only 50% of the theoretical capacity of the cathode, LiCoO2 is utilized in practice (∼140 mA·h g−1 vs theoretical, 274 mA·h g−1). While the theoretical capacity of the anode, graphite, is 372 mA·h g−1 (for LiC6), the practical capacity is ∼300−320 mA·h g−1. Thus, there is a need to find electrode materials that can yield higher capacities and,

1. INTRODUCTION 1.1. Lithium Ion Batteries

Commercial lithium ion batteries (LIBs), originally developed by Sony Co., Japan, more than 20 years ago, use layer-type compounds, lithium cobalt oxide (LiCoO2) as the cathode (positive electrode) and graphite (C) as the anode (negative electrode) material, and a nonaqueous Li ion conducting medium as the electrolyte. The electrolyte can be in the form of a solution of a Li salt, specifically, LiPF6 dissolved in a mixture of ethylene carbonate (EC) and diethyl carbonate (DEC), dimethyl carbonate (DMC), or both. It can also be in the form an immobilized gel polymer, which contains a mixture of the copolymer, poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP). The principle of operation involves intercalation−deintercalation of Li ions to and from the electrodes, to store and deliver dc electrical energy during the charge and discharge process, respectively. The reactions are LiCoO2 + 6C ↔ Li1−xCoO2 + LixC6 (0 < x ≤ 1). Charge neutrality for x ≠ 0 is maintained by the oxidation of Co3+ to Co4+ ion in the cathode. Similarly, the graphite lattice will be in reduced valency state. The LIBs are assembled in the discharged state. During charging by an external dc source, the electrical energy is converted to chemical energy in the form of charged products, Li1−xCoO2 and LixC6, whereas during discharge, under a load, the reverse reaction occurs. Studies have shown that for 0 < x ≤ 0.5 in Li1−xCoO2 and for 0 < x ≤ 1.0 in LixC6, the system is completely reversible for a large number of charge−discharge cycles. It is immediately clear that a LIB cathode must contain not only Li but also a transition metal in the form of a compound or composite. LIBs are lightweight and, with an operating voltage of ∼3.6 V and a deliverable capacity ranging from 700 to 2400 mA·h for a single cell (battery), are extensively used in the present-day portable electronic gadgets like cell phones, notebook computers, iPads, and video camcorders as has been documented in several books1−5 and general reviews.6−8 Other uses are in the medical industry in defibrillators, pace makers, etc., and in the aerospace industry for storing photovoltaic-generated dc electricity. LIBs are usually charged and discharged at 0.2−1C current rate, meaning that full capacity of the cell is stored or utilized in 5−1 h, respectively. The usual operating temperature (T) range of the LIBs is 15− 60 °C. At T < 15 °C, the capacity becomes low, whereas at T > 60 °C slow degradation of the electrode/electrolyte materials sets in over a period of time. Due to the importance of the topic dealing with the aspects of energy storage in the overall picture of the energy landscape and due to the large number of research groups worldwide working in the area of LIBs, five 5365

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

electrolyte. As a result, there are six prospective futuregeneration 4 V cathodes, namely, the Li-containing mixed oxides, Li(Ni0.8Co0.15Al0.05)O211−18 and Li(Ni1/3Co1/3Mn1/3)O2,19−24 which have a lower Co-content than LiCoO2,25−32 modified LiMn2O4,27,33−55 which is cheap and environmentally friendly, LiFePO4,56−81 which is very cheap but acts only as a 3.5 V cathode, and the vanadium-containing phosphates, LiVPO4F (4.2 V vs Li)82−86 and Li3V2P3O12,87−90 which are also relatively cheap and environmentally acceptable. The latter three compounds also show good T stability in the charged state. The power packs using LiMn2O4 and LiFePO410,91 as cathodes are already available for use in portable power tools, and demonstration of HEVs/PHEVs incorporating these have been field-tested during the past 5−7 years. It is certain that in the coming years, at least three of these cathodes will be employed in LIB power packs for HEVs/PHEVs and possibly for EVs. Mention must be made of the development of the 5 V cathodes, namely, Li[Ni0.5Mn1.5]O4,92−103 Li[Co,Mn]O4104 and Li[V,Ni]O4 (spinel structure),105−107 and LiCoPO479,108−114(isostructural to the orthorhombic LiFePO4), of which the first compound has been extensively studied and is a prospective candidate for future-generation LIBs.

preferably, operate at high voltages (≥4.0 V vs Li) for cathodes and low voltages (≤0.5 V vs Li) for anodes to obtain high energy density. (3) Safety-in-operation. This is a crucial requirement for high-power operation of LIBs at high current charge/discharge rates, such as, 2−10C. Occasional reports appear in the press about the LIBs used in cell phones or notebook computers catching fire or evolving smoke during charging or discharging operation. This is usually attributed to the absence or malfunctioning of the safety electronic circuit in the gadget, which restricts the charging voltage to 4.2 V and charging current to ≤1C. The charged cathode and anode are thermally unstable and act as powerful oxidizing and reducing agents, respectively, and decompose the solvents present in the electrolyte evolving gases and heat. For example, when LiCoO2 is charged to >4.3 V vs Li, the resulting composition, LixCoO2 (x < 0.5) contains large amounts of Co4+ ion in it, which is a powerful oxidizer. Further, the above composition is unstable and evolves oxygen gas when heated, either accidentally or intentionally, to temperature T ≥ 240 °C. Similarly, at high current rates of charge, there exists a distinct possibility of Li metal depositing on the graphite anode, many times as dendrites, which can penetrate the microporous separator and can give rise to short-circuiting of the LIB. When the latter happens, all the chemical energy stored in the LIB is converted to heat resulting in smoke evolution, fire, or explosion. This is one of the reasons the LIBs are not sold in the market as individual cells, such as, D, AA and AAA size, and are always supplied as hermetically sealed units as part of the electronic gadgets. The most important safety aspect in the usage of LIBs for the HEVs/PHEVs is “heat management”. Like all other batteries, LIBs are also prone to the 10−15% loss of electrical energy as heat, the so-called I2R-loss, where I is the current and R is the overall resistance or impedance of the LIBs. This is very familiar to all users of notebook computers, who always feel the warming of the LIB pack during discharge operation. In the large power packs needed for HEVs/PHEVs, heat management during discharge of the LIBs is a nontrivial problem, and unless extreme precautions are taken, explosion can result. The car manufacturers, no doubt, are aware of this, and hence, LIBs with LiCoO2 as the cathode for the HEV/PHEV power pack are completely ruled out, both due to safety and cost considerations. Thus, there is a need to come up with cheaper and more stable cathode and anode materials. (4) Improvements in the low- and high-temperature operation. The present-day LIBs show ∼50% and ∼75% capacity degradation when operated at T = 10 and 0 °C, respectively. This is due to the slower reaction kinetics of the Li ions and also the decrease in the ionic conductivity of the electrolyte. Similarly, when LIBs are operated at T > 60 °C, the opposite effects set-in giving a better performance, but at the same time, electrolyte degradation (drying-up) and cathode decomposition are increased, which are highly undesirable. Thus, for use in HEVs/PHEVs, there is a strong urge to extend the T-limits of operation of the LIBs, especially at T = 0−10 °C and at T = 60−80 °C.

1.4. Ideal Anode for LIBs

It will be of interest to enumerate the properties of an ideal anode material for LIBs and to compare them with those exhibited by the graphite. They are as follows: (1) It must contain elements or compounds with low atomic or formula weights, be low density, accommodate fairly large amounts of Li per formula unit, and be cyclable, to yield large, stable, and reversible gravimetric (mA·h g−1) and volumetric (mA·h cm−3) capacities. Graphite can intercalate/deintercalate Li up to the composition LiC6, at which a stage I intercalation product is realized, meaning that a Li atom (Li ion + electron) is present between every layer of the host graphite lattice.115−117 This gives, as mentioned earlier, a theoretical capacity of 372 mA h g−13. Thus, even though carbon, in the form of graphite, has a low atomic weight and low density, its reversible capacity is limited. Efforts to increase the Li content in LiC6 and reversibly cycle it have not been successful. (2) An ideal anode material must show a potential as close to that of Li metal as possible and must not show large variations in the potential with changes in the Li content. This is because, when combined with a 4 V cathode, the overall working voltage of the LIB will not be much lower than 4 V. Graphite satisfies this requirement in that it shows a potential of 0.15−0.25 V vs Li metal. (3) It must not be soluble in the solvents of the electrolyte and must not chemically react with the salt or solvents of the electrolyte. Studies have shown that in EC-based solvents, a protective film, the so-called solid electrolyte interphase (SEI) forms on the external surfaces of the graphite particles during the first few cycles of discharge−charge reactions. This is a result of chemical reaction of Li with the EC. The SEI prevents excessive solvent cointercalation and also acts as a good Li ion conductor and enables facile Li cycling.2,3,5,7,118 Further, the SEI protects the LiC6 (charged graphite), which is a strong reducing agent, from coming into direct contact with the solvents of the electrolyte and thereby suppresses the unwanted side reactions. (4) An ideal anode must possess good electronic and Li ionic conductivity (“mixed conduction”) so that the electrode will have small impedance for current pick-up and for the motion of the Li ions with in the active material. While graphite is a

1.3. Future-Generation Cathodes

To satisfy at least some of the above criteria, research has been and is being carried out worldwide to find alternative electrode materials based on cheap and environmentally friendly compounds, which may work on principles different from the Li intercalation−deintercalation reactions. Similarly, efforts are being made to replace the toxic and costly salt LiPF6 as the 5366

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 1. Classification of oxide anode materials based on the reversible Li insertion and extraction process: Intercalation−deintercalation, alloying− dealloying, and conversion (redox) reaction. In favorable cases, the latter two processes can act synergistically to give rise to large and stable capacities. Selected examples are given. Schematic diagram of the process is shown. The voltages shown are vs Li metal.

semiconductor (conductivity (σ300 K) ≈ 10−2 to 10−3 S cm−1), LiC6 is an excellent mixed conductor and, as a matter of fact, shows metallic-type electronic conductivity (σ300 K ≈ 102 to 103 S cm−1) and a high Li ion mobility (DLi+ at 300 K ≈ 10−8 to 10−10 cm2 s−1). (5) It must be cheap and environmentally friendly. Graphite meets this requirement, even though the specialty graphites, like MCMB or microcrystalline graphite are expensive due to the manufacturing process. Thus, it is clear that graphite satisfies most of the requirements of an ideal anode and, hence, has been the preferred choice for the commercial LIBs.

metal oxides and other compounds with a two-dimensional (2 D) layer structure or 3D network structure that can reversibly intercalate Li into the lattice, similar to graphite, without destroying the crystal structure have been explored. Oxides of titanium (Ti) have received the most attention. (2) Elements (A) or metals (M) that can form alloys with Li metal (strictly, intermetallic compounds, since LixA or LixM adopt distinct crystal structures and exhibit physical properties different from those of Li, A, and M). Elements like Si120−140 and Sb130,136,141−144 and metals like Sn,145,146 Zn,130 In,130 Bi,130,136 and Cd130 form alloys with Li metal and have been studied extensively as prospective anodes. The Li alloying− dealloying reactions, which usually occur at low potentials (≤1.0 V vs Li) contribute to the reversible capacity during Li cycling (e.g., Sn (or Si) + 4.4Li ↔ Li4.4Sn (or Si)). (3) The third mechanism by which a material can act as anode is the socalled redox or “conversion” reaction with Li. Normally, the stable lithium oxide, Li2O, is electrochemically inactive and cannot be decomposed to the metal and oxygen. However, in the presence of nanosize transition metal particles, which may or may not be electrochemically generated, it can be decomposed, as exemplified by the now well-known reaction: nano-CoO + 2Li ↔ nano-Co + Li2O. Thus, at suitable potentials, depending on the nature of the metal, Li cycling can occur, giving rise to large and reversible capacities, stable over a large number of discharge−charge cycles. This conversion reaction is a general phenomenon and is applicable to oxides, fluorides, oxyfluorides, sulfides, nitrides, phosphides, etc. Thus, a large number and wide variety of metal-containing compounds, including binary, ternary, and complex oxides and oxysalts have been explored for their Li

1.5. Future-Generation Anodes

Unfortunately, the picture of the prospective anodes to replace the graphite in LIBs is not as rosy as that of the prospective future-generation cathodes described earlier. Despite extensive research over the past 15 years, only one alternative anode has been successfully introduced to the market, namely, an amorphous/nanocrystalline composite of Sn/Co/C developed by Sony Co., Japan, in 2005.119 In this material, Sn is the electroactive metal, whereas cobalt and carbon act as the electroinactive “matrix” elements and help in better Li cycling in the voltage range of operation. A reversible and stable capacity of ∼350−450 mA·h g−1 has been achieved, and the major advantage is claimed to be safety-in-operation and relatively low cost in comparison to the specialty graphite anode. Nevertheless, there are very many materials that have been explored, especially the metal oxides, and the future prospects are bright. Research on anodes for LIBs is directed toward materials that can perform via three different mechanisms as shown in Figure 1. (1) Intercalation−deintercalation mechanism. Transition 5367

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

present review, in the half-cell configuration, Li insertion into the oxide is termed as the discharge reaction, whereas the Li extraction from the oxide or composite is termed as the charge reaction. However, in the Li ion cell (full cell) configuration, where a Li containing oxide (like, LiCoO2) forms the cathode and an oxide material forms the anode, the processes are named in the reverse fashion. That is, in this case, “discharge reaction” involves the Li extraction from the oxide anode, and the corresponding Li insertion into the cathode (like, Li1−xCoO2). Similarly, the “charge reaction” involves Li extraction from the oxide cathode (to form, say, Li1−xCoO2) and Li insertion into the oxide anode. Due to the vast amount of literature, it has not been possible to mention all the available references. However, it is hoped that important and pertinent papers have not been omitted and the interested reader may consult the references cited in these papers.

cyclability. In addition, as can be expected, the techniques and methodology of nanotechnology have been employed to prepare many simple and complex compounds, including oxides, and have been studied for Li cycling via various mechanisms. As a result of the research on alternative anodes for LIBs, the materials chemistry and electrochemistry of these materials have been enhanced and enriched very significantly during the past decade. 1.6. Scope of the Review and Nomenclature

Metal-containing compounds in the form of oxides and oxysalts, such as, oxyfluorides, oxyhydroxide, phosphates, carbonates, and oxalates, are discussed in this review. These include bulk (micrometer-size) particles, nanosize particles, or agglomerates with various morphologies, thin films, and carbon, CNT, or graphene/metal oxide composites. The latter may contain electrochemically -active or −inactive additives. Metalcontaining compounds in the form of fluorides,147,148 sulfides,148 selenides,136 nitrides,148−151 phosphides136,148 and antimonides136 are omitted, even though good amount of work is available in the literature. As mentioned earlier, several review articles have appeared over the years, summarizing the situation. While there are only nine reviews6,7,115,152−157 prior to 2006, there are more than 40 articles in the last 5 years that directly or indirectly described and discussed Li storage and cycling of oxide and oxide-related materials, in the form of Feature articles, Accounts, Perspectives, and Mini- and regular reviews. Many of these are listed in the references.8,73,102,136,146,148,158−199 In brief, discussions were made on layered vanadium oxides by Cavana et al.,165 molybdenum oxides by Cavana et al.165 and Ellefson et al.,200 TiO2-based nanostructures and their composite oxides by Arico et al.,7 Bruce et al.,162 Deng et al.,166 Yang et al.,169 Chen et al.,185 Djenizian et al,201 Zhu et al.,202 Su et al.,203 Berger et al.,204 and Froschl et al.205 Very recently, Froschl et al.205 reviewed the formation of crystalline nanoscale TiO2 anatase, rutile, and brookite particles via solution-based approaches, crystal structure, morphology, surface area, and particle dimensions and energy storage properties. Yang et al.183 and Yi et al.172 reported studies on Li4Ti5O12. Knoops et al.206 and Meng et al.207 reported nanostructured Li ion batteries fabricated using atomic layer deposition (ALD) technology. Cavaliere et al.182 nicely reviewed electrospun one-dimensional mesostructured organic, inorganic, and hybrid nanomaterials and their energy storage applications. Reviews on metal oxide with carbon, carbon nanotube (CNT), and graphene composites for Li ion batteries have also been published.189,194,208−214 Cabana et al.148 nicely reviewed different conversion reaction based anodes based on metal oxides, sulfides, nitrides, phosphides, and fluorides. In an excellent critical review by Baddour-Hadjean and Pereira-Ramos173 on Raman spectroscopy studies, they discussed Raman studies on carbon-based materials, layered LiCoxNiyO2 oxides, Mn-based oxides (MnO2, bare and doped LiMn2O4), LixV2O5 oxides, TiO2, and Li4Ti5O12 electrode materials. This amply attests to the importance of the LIBs as a very active research area of topical interest. A word about the nomenclature is in order: Evaluation of the electrochemical performance of oxide materials is usually carried out in the so-called “half-cell” configuration, in which the oxide acts as a “cathode” (positive electrode) with Li metal as the counter negative electrode (anode). Hence, in the

2. ANODES BASED ON LI INTERCALATION−DEINTERCALATION REACTION By definition, intercalation of an ion, atom, or molecule into a crystal lattice of the host compound is incorporation of the guest species without destroying the crystal structure. However, minor modifications of the crystal structure do occur. The process is also called “topotactic reaction”. Three important conditions need to be satisfied for the above reaction involving Li. First, the compound must be crystalline and there must be empty sites in the host crystal lattice, in the form of isolated vacancies or as one-dimensional (1D) channels, 2D layers (van der Waals gap), or channels in the 3D network. The compounds with 2D and 3D structures are most conducive for Li intercalation−deintercalation. On the other hand, compounds with isolated vacant sites or 1D channels may easily intercalate but not easily deintercalate Li ions and hence are not of much use as anodes for LIBs. Second, the host compound must contain a transition metal or rare earth metal, like Ce, Pr, Eu, or Tb, which can exhibit one or more stable valency states. This is because intercalation of Li ion (along with the electron) will reduce the valency state of the host metal ion by one unit, for example, Ti4+ to Ti3+ in Ti oxides. For example, α-Fe2O3 and MgFe2O4 can be intercalated with Li to form LixFe2O3 and LixMgFe2O4, x ≤ 0.5, whereas αAl2O3 and MgAl2O4 will not undergo Li intercalation. The third condition, which may not be very relevant in the present context, is that if the host compound has an unfavorable crystal structure and transition metals in low valency state, like TiO, VO, and NbO with the cubic rock salt structure, Li intercalation may not occur. Hence, metal oxides exhibiting semiconducting or insulating behavior, containing transition metal ions, and having 2D or 3D structures are the most suitable for Li cycling (Figure 1). 2.1. Binary Oxides

2.1.1. TiO2. Titanium dioxide, TiO2, is an attractive candidate for use as an anode for LIBs due to its low cost, ready availability, and ecofriendliness. TiO2 exists in several polymorphic modifications: anatase, rutile, brookite, TiO2-B (bronze), TiO2-R (ramsdellite), TiO2-H (hollandite), TiO2-II (columbite), and TiO2-III (baddeyite). All of them contain TiO6 octahedra (Figure 2). Yang et al.169 and Froschl et al.205 described and discussed, in detail, the Li cycling properties of different titanium oxides in their recent review. Raman studies on Li-intercalated TiO2 are 5368

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

intercalation−deintercalation.169 The crystal structure is shown in Figure 2b. It adopts a body-centered tetragonal crystal structure with the lattice parameters a = 3.784 Å and c = 9.546 Å (space group, I41/amd) and contains edge-shared TiO6 octahedra. The interstitial sites (octahedral holes) in this framework structure can intercalate/deintercalate Li ions via eq 1, and charge-neutrality is maintained by the reduction of the valency of the Ti ion. The theoretical capacity is 335 mA·h g−1 (for 1 mol of Li per mole of TiO2) when cycled to the lower cutoff voltage of 1.0 V vs Li. Studies have shown that during Li insertion, the bulk (microcrystalline) anatase TiO2 spontaneously phase separates into a Li-rich phase (Li∼0.6(TiO2)) and Li-poor phase (Li∼0.01(TiO2)) on a scale of several tens of nanometers in the material.169,232−234 This two-phase equilibrium is maintained until all of the Li-poor phase is converted into the Li-rich phase up to the completion of Li insertion. Similar behavior is seen upon Li extraction, and ∼0.5 mol of Li is found to be cyclable without significant capacity fading. The two-phase reaction is reflected as a voltage plateau in the galvanostatic discharge profile (voltage vs capacity curve) and as a relatively sharp peak in the cyclic voltammogram (current vs voltage profile). These studies showed that Li insertion occurs at ∼1.72−1.75 V and Li extraction occurs at ∼1.8−1.9 V vs Li. The voltage hysteresis is ∼0.1−0.2 V. Detailed studies including recent in situ X-ray absorption spectroscopy235 have shown that the Li-rich phase, namely, Li0.55(TiO2), adopts an orthorhombic structure236 with lattice parameters a = 3.814 Å, b = 4.084 Å, and c = 9.066 Å (space group, Imma). The change in the crystal symmetry is accompanied by a net increase in the unit cell volume by ∼4%, in comparison to the pristine anatase TiO2. The existence of a tetragonal to orthorhombic phase transition during Li insertion was clearly demonstrated by the Raman and XRD measurements by Baddour-Hadjean et al.237−239 and Hardwick et al.240 and analyzed in an excellent review by Baddour-Hadjean and Pereira-Ramos.173 The group of Wagemaker and Mulder233,234,241−252 carefully examined the effect of particle size, lithium diffusion, and structure on Li-poor and Lirich LixTiO2 phases by solid-state nuclear magnetic resonance, neutron diffraction, X-ray absorption spectroscopy, and molecular dynamics simulations techniques. By reduction of the particle size, for example, to nanosize, it is possible to insert more Li electrochemically,234 to obtain the composition Li(Ti3+)O2 (x = 1) via two-phase reaction. This phase adopts the anatase structure with a = 4.043 Å and c = 8.628 Å (space group, I41/amd), and there is no significant change in the unit cell volume in going from Li0.55TiO2 to LiTiO2, except for a change in the crystal symmetry. However, the long-term Li cyclability is found to be poor, and severe capacity fading occurs upon cycling in the entire range of x, from 0 to 1. This is due to changes in the crystal symmetry and unit cell volumes and poor Li ion kinetics of the phase with x = 1, since all the available octahedral interstices are filled by Li ions. As can be expected, decreasing the particle size to the nanometer regime alters the reactivity toward Li and the extent of Li cycling. Thus, nano-anatase TiO2 along with unique morphology, for example, rod-like, tube-like, fiber-like, reticular, mesoporous, or nanoporous, can cycle more than 0.5 mol of Li. Some of the recent data on the Li cycling properties of nanoanatase TiO2 are given in Table 1. High reversible capacities, close to ∼0.7 mol of cyclable Li per mole of TiO2, and good current (C) rate capabilities have been achieved.

Figure 2. Crystallographic representation of (a) rutile, (b) anatase, (c) brookite, and (d) bronze (B) TiO2. Blue and red spheres are Ti and O atoms respectively. The arrangement of TiO6 octahedra and outline of the unit cell are shown for each polymorph. Reproduced with permission from ref 169. Copyright 2009 Elsevier.

nicely reviewed by Baddour-Hadjean and Pereira-Ramos.173 Presently, a few more recent reports will be discussed. Generally, the Li storage and cycling performance of TiO2 polymorphs depends on the method of preparation, particle size, and shape and morphology. In general, smaller particle size (≤200 nm) with high surface area and porous morphology with interconnected particles can deliver stable and near theoretical capacity, according to eq 1. Ti4 +O2 + Li+ + e− ↔ Lix(Ti 3 + xTi4 +1 − x)O2 (e− = electron)

(x ≤ 1) (1)

TiO2 in the form of nanoparticles, nanotubes, nanowires, nanorods, and nanofibers and conductive carbon or metal oxide nanophase coated systems and composites have been studied for Li cyclability. The following are some of the techniques adopted to prepare nanophase TiO2: hydrothermal method,215,216 sol−gel method,217−221 soft-template method,222 precipitation or solid state method followed by ion exchange, 223,224 urea-mediated hydrolysis/precipitation route,225 anodization,226,227 molten salt method,228 synthesis by Ionic liquids,229 and electrospinning technique.228,230 Synthesis of Ti oxides nanostructures are discussed in the recent reviews by Zhou et al.231 and Froschl et al.205 2.1.1.1. Anatase TiO2. Among the various modifications, the anatase TiO2 is known to be the most electroactive host for Li 5369

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

5370

17 (XRD) 29 29 submicrometer size, 150 nm

21 18 100−200

rice grain/4 wt % CNT composite nanofibers hollow peanuts hollow ellipsoid hollow capsule hollow pseudocube solid spheres TiO2 nanoparticles TiO2−δ-H2-1h TiO2−δ-H2-7h N,F-co-doped TiO2

rice grain shape

6 9 10 20 (XRD)

15−20 2−3 9−11 100−300, sheet thickness ∼5 nm

34.2 37.2 9.5 8.3 6.9 6.2 4.5 174 153.7 150.8

123 110 150 34.1

135 598 80 170

189 50−400

127 297 202

18.8

5 (100 °C), 8 (300 °C), 10−25 (400−600 °C), 50 (500 °C) 28 10−12

524.5 221.9

5.1 (XRD) 9.5

92

223 10 116.49

surface area, m2 g−1

224

particle/crystallite size (nm)

30 diam, 5 thickness (TEM) 10

6 89 14 (XRD) ∼5−10 9 (XRD)

mesoprous hollow spheres nanoparticles

hollow nanospeheres nanorods and particles (bare and carboncoated) mesoporous (C16-TiO2) amorphous TiO2 spherical nanosheets

TiO2 nanosheets TiO2/graphene nanosheets mesoporous thin films

TiO2−Cu TiO2−Sn hollow microspheres with nanotubes mesoporous polyaniline/TiO2 nanocomposite nanoporpous as prepared nanoporous calcined (at 400 °C) commercial

mesoporous TiO2

TiO2 nanoparticles commercial sample mesoporous TiO2 sub-microspheres

morphology

Table 1. Physical Properties and Electrochemical Li Cycling Data of Anatase TiO2

0.09

1C

150 mA g−1 (0.45C) 0.45C 0.45C 0.1C

0.5C 250 mA g−1 (0.75C) 66 mA g−1 (0.2C) 33.5 66 mA g−1 (0.2C) 1C (1C = 170 mA g−1) 0.25C 1C

0.2C 0.2C 0.1 mA cm−2

0.2C

1C 2000 mA g−1

0.05C 0.05C 0.1C 0.5C 8C (1C = 168 mA g−1)

current ratea

139 144 156 (2nd discharge) 157 136 131 123 138 220 140 90

230 160 148 162

275 335 270 190

198 120−275

118 202 205 100−250

94 229

191 196 220 185

165 100 180 162 157

reversible capacity (mA·h g−1)

(n (n (n (n

= = = =

1−50) 1−50) 2−500) 1−70)

93% 92% 77% 77% 88% 89% 85% 30% 60% 63% 78%

80% 83% 78% 81%

93% 60% 91% 92%

(n (n (n (n (n (n (n (n (n (n (n

(n (n (n (n

(n (n (n (n

= = = = = = = = = = =

= = = =

= = = =

10−800) 10−800) 2−30) 2−30) 2−30) 2−30) 2−30) 1−20) 1−20) 1−20) 2−60)

40) 50) 50) 10−800)

1−50) 2−50) 2−30) 2−100)

94% (n = 50) 84−95% (n = 50)

90% (n = 100) 63% (n = 1−30) 89% (n = 1−30) 85−90% (n = 2−20)

95% (n = 100) 95% (n = 100)

82% 80% 68% 76%

84% (n = 1−50)

91% (n = 1−50; 1.5−3.0 V) 80% (n = 1−20; 1.5−3.0 V) 93% (n = 5−80; 1.5−3.0 V)

voltage, range = 1−3 V vs Li (cycling range, capacity retention after n cycles)

ref

268

267

266

265

263 264

222 260 261 262

258 259

257

256

255

215 216

221

254

253

Chemical Reviews Review

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

271

Mesoporous anatase TiO2 with an ordered 3D pore structure was synthesized by the group of Bruce272 using the silica KIT-6 as a hard template. The ordered pore structure consisted of 11 and 50 nm pores with walls of thickness 6.5 nm composed of anatase crystallites with a BET surface area of 205 m2 g−1. The galvanostatic discharge−charge cycling in the range 1−3 V vs Li at a current of 30 mA g−1 (0.09C) showed a first-discharge capacity of 322 mA·h g−1 (composition, Li0.96TiO2), excellent cycling performance (up to 1000 cycles), and very good C-rate capability. For example, at 35.7C, a reversible capacity of 125 mA·h g−1 was observed. The authors also reported and discussed the ex-situ XRD and Raman spectra of the phases with various x in LixTiO2. The group of Maier255 recently reported the Li cycling properties of hierarchical nanoporous anatase TiO2 (np-TiO2) and found that as much as 64% of the total reversible capacity is contributed by pseudocapacitive interfacial storage, with 36% coming from the bulk Li insertion/extraction. The interfacial storage is aided by high surface area and high porosity of the nanostructured morphology where additional Li (and electrons) can be accommodated by a synergistic storage mode of a Li ion accepting phase and an electron-accepting phase at the solid−solid and solid−liquid interfaces of the nanometer-sized composite electrode system. The np-TiO2 was prepared by a modified in situ hydrolysis route employing Ti-glycolate precursor, and the 400 °C heattreated (calcined) np-TiO2 showed a high surface area of 221.9 m2 g−1, crystallite size of 9.5 nm, average pore diameter of 6 nm, and pore volume of 0.298 cm3 g−1 and has 8 mol % residual H2O. The galvanostatic discharge−charge profiles of calcined np-TiO2 in the voltage range 1−3 V vs Li at current rate of 0.2C (1 Li per formula unit in 5 h), along with commercial TiO2 of two different particle sizes, are shown in Figure 3. The profiles of commercial anatase TiO2 (sizes 200

Unless otherwise stated, the current rate 1C = 335 mA g−1 and corresponds to insertion/extraction of 1 mol of Li per mol of TiO2.

89% (n = 2−150)

Figure 3. Galvanostatic curves for the two types of commercial TiO2 particles with different grain sizes (∼200 and ∼5−10 nm) and for the calcined np-TiO2 discharged at 0.2C. Reproduced with permission from ref 255. Copyright 2011 Wiley-VCH.

nm and ∼5−10 nm) are also shown. As can be seen, the firstdischarge curve can be divided into three different voltage regions. In region A, a monotonic decrease in the voltage occurs up to ∼1.75 V and is attributed to a homogeneous Li intercalation into the bulk up to a solid−solution limit of ∼Li0.15TiO2. Region B corresponds to the two-phase voltage plateau where the Li-poor and Li-rich (Li∼0.5TiO2) regions coexist. In region C, there is a small pseudoplateau at ∼1.5 V indicating Li intercalation into bulk with x > 0.5, but the continued linear decrease of the voltage profile up to 1 V cutoff

a

180 20

0.1C 0.091C

0.5C (1C = 160 mA g−1) 100

94% (n = 5−60) 250 99 nanosized particles rod-like morphology

34% (n = 2−27) 92% (n = 5−60) 93% (n = 5−60)

270 22

13 (XRD)

TiO2 prepared by molten salt method (MSM) at 280 °C MSM TiO2 nanoparticles reheated to 500 °C MSM−TiO2 nanoparticles reheated to 750 °C TiO2 commercial TiO2 (prepared by MSM) (1) 180 °C (0.62M LiNO3:0.38M LiOH) (2) urea treated-MSM-180 (3) MSM sample 1 reheated at 300 °C TiO2 nanoparticles

morphology

Table 1. continued

164 175 200 10 86.45 106 130 avg crystallite size,5−10 nm (XRD) rod-like morphology

81% (n = 2−58) 181 0.1C 12.5 91

72% (n = 2−60) 221 0.1C 80

227 0.1C 200

76% (n = 2−60)

269

Review

particle/crystallite size (nm)

surface area, m2 g−1

current ratea

reversible capacity (mA·h g−1)

voltage, range = 1−3 V vs Li (cycling range, capacity retention after n cycles)

ref

Chemical Reviews

5371

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

clearly indicates interfacial storage of Li. The first-discharge capacity was 388 mA·h g−1 (composition, Li1.16TiO2) whereas the first-charge capacity was 271 mA·h g−1. The 30% capacity loss was attributed to the consumption of Li by the residual H2O in the compound. After a few cycles, the capacity stabilized at 242 mA·h g−1 (Li0.72TiO2) with only a small loss up to 20 cycles and with ∼100% Coulombic efficiency. Rate capability, up to 60C, was tested and found to be very good. For example, at 30C, a capacity of 77 mA·h g−1 was stable at least up to 10 cycles. Further, capacities of 302 and 200 mA·h g−1 at 1C and 5C rates were realized after 100 cycles, with only a small capacity fading between 2 and 100 cycles as shown in Figure 4.

storage mechanism by Lafont et al.235 in their study on nanoTiO2 prepared by a modified sol−gel method using a surfactant. Also, as pointed out by the group of Maier255 (Table 1), the interfacial Li storage mechanism is a nonequilibrium phenomenon, in comparison to the thermodynamically stable bulk Li intercalation/deintercalation, and after an extended rest period, bulk degradation can occur leading to significant capacity loss. The authors255 state that this can be overcome by cycling the material, preferably at high C-rates, without further long rest periods, thereby suppressing bulk degradation and achieving high and sustainable reversible capacities via bulk and interfacial Li storage, as shown in their study (Figure 5). Additional recent references on the Li cycling of anatase TiO2 are by Serventi et al.273 and Xu et al.,274 porous and dense TiO2 nanospheres by Wang et al.,275 mesoporous rods by Jiang,276 TiO2 nanotubes by Ryu et al.,277 TiO2 nanotubes by Panda et al.,278 Ortiz et al.,279 Gonzalez et al.,280 Han et al.,281 and Li et al.,282 TiO2@carbon nanofibers by Yang et al.,283 mesoporous carbon TiO2 composites by Chang et al.,284 TiO2 graphene composites by Qiu et al.,285 Cao et al.,286 Tao et al.,287 and Shah et al.,288 porous single crystal TiO2 by Zhang et al.,289 Sn-doped TiO2 nanotubes by Kyeremateng et al.,290 amorphous TiO2 nanotubes by Xiong et al.,291 Ni/TiO2 nanowires by Wang et al.,292 oxygen-deficient TiO2−δ nanoparticles via hydrogen reduction by Shin et al.,267 nanostructured TiO2 by Xiao et al.293 and Hong et al.,294 (N,F)-codoped TiO2 by Cherian et al.,268 Cu-doped TiO2 by Barreca et al.,295 10 mol % M-doped TiO2 (M = Mn, Sn, Zr, V, Fe, and (Ni, Nb)) by molten salt at 410 °C for 2 h in air reported by Reddy et al.,270 kinetics of lithated TiO2 by Belak et al.,296 and defect chemistry and surface structures using atomistic simulation technique by the group of Islam.297 Sandwich-like graphene-based mesoporous anatase titania, G-TiO2 has been prepared by Yang et al.256 using G-SiO2 nanosheets as the template and (NH4)2TiF6 as the precursor using a sol−gel process, and they compared its Li cycling performance with that of anatase TiO2 nanosheets prepared without graphene. The former compound contained 9 wt % graphene, which is supposed to act as a mini-current collector. When cycled in the range 1−3 V vs Li at a current rate of 0.2C, the first-discharge capacity was 269 mA·h g−1 (composition Li0.8TiO2) whereas the first-charge capacity was 202 mA·h g−1.

Figure 4. Charge/discharge capacities versus number of cycles for the calcined np-TiO2 (dis)charged directly at rates of 1C (charge/ discharge to theoretical capacity within 1 h) (a) and 5C (charge/ discharge to theoretical capacity within 12 min) (b). The data labeled as (c) represent the charge/discharge capacities when the material was (dis)charged with rates ascending stepwise from 0.2C to 60C. Reproduced with permission from ref 255. Copyright 2011 WileyVCH.

The superior performance of calcined np-TiO2 was demonstrated by a comparison of the data on as-prepared np-TiO2 and commercial nano-TiO2 (size 5−10 nm) (Table 1). Based on the observed data, Shin et al.255 proposed a Li insertion mechanism for the nanoporous, high-surface area np-TiO2 emphasizing the regions A, B, and C (Figure 3), in which at high C-rates, the interfacial Li storage mechanism is dominant. This is shown in Figure 5. It must be mentioned that the solid−solution region A (Figures 3 and 5) is also partly attributed to the interfacial Li

Figure 5. A proposed overall Li insertion mechanism for the nanoporous TiO2 (anatase) structure with a high surface area. Reproduced with permission from ref 255. Copyright 2011 Wiley-VCH. 5372

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Table 2. Physical Properties and Electrochemical Li Cycling Data of Rutile TiO2 morphology needle-shaped morphology commercial nanoparticles rod-shaped particles commercial nanoparticles bulk nanoparticles nanoparticles TiO2/C nanofibers spherical particles nanoneedles MSM−TiO2 nanoparticles reheated to 850 °C ball-milled (N,F) codoped TiO2 ball-milled TiO2

surface area, m2 g−1

particle/crystallite size (nm) 10 nm diam, 30−40 nm length

current rate, mA g−1

reversible capacity, mA·h g−1

voltage, range 1−3 V vs Li (cycling range, capacity retention after n cycles)

0.05C

ref

205

80% (n = 2−50)

306

15 nm

50

264

90% (n = 2−20)

307

200 nm long, 10 nm diam

30

160

94% (n = 1−50)

308

50 nm

30

37

94% (n = 1−15)

181

30 33

10 225

98% (n = 1−15) 89% (n = 2−30)

220

180 267

33 1C

380 120

82% (n = 1−5; 0.1−3.0 V) 98% (n = 2−100)

303 309

200

90% (n = 2−35)

310

0.1C 33

177 22

77% (n = 2−15) 80% (n = 2−50)

311 269

30

210

80% (n = 2−60)

268

115

96% (n = 5−60)

1 μm whiskers 4−5 nm length 40−50 nm avg crystallite size, 6 nm (XRD); whiskers 4−5 nm, length 40−50 nm 10−40 nm size 1.5−3 μm, 20 nm nanorods

0.033C

length >100 nm, width 20−25 nm 200 nm (XRD)

135 0.74

nanosized particles, 10−20 nm

13 nm

which does not allow facile Li ion transport along the crystallographic ab-plane in comparison to that along the caxis. However, nanophase rutile TiO2 particles with various morphologies have demonstrated that ∼0.5 mol of Li per TiO2 can be cycled in the voltage range 1−3 V vs Li, as has been discussed in the review by Yang et al.169 It is now known that Li intercalation into nanorutile TiO2 proceeds through two-phase domains, and after x reaches unity in LixTiO2, the structure transforms to a disordered cubic rock salt-type LiTiO2 (a = 4.140 Å, V = 70.96 Å3). Other studies proposed that LiTiO2 adopts a hexagonal structure. Subsequent Li deintercalation (charging) and successive Li cycling occurs in this newly formed phase in a solid−solution domain, and a reversible capacity corresponding to x ≈ 0.5 (∼170 mA·h g−1) is normally observed. Some recent reports on the Li cycling of nanorutile TiO2 are described here, and capacity values are summarized in Table 2. Neutron diffraction and X-ray absorption spectroscopy studies of micro- and nanosized bare and Li intercalated rutile TiO2 are discussed by the group of Wagemaker and Mulder.299,300 The group of Wohlfahrt-Mehrens301−304 prepared nanorutile TiO2 by sol gel method using glycerol-modified Ti-precursor in the presence of an anionic surfactant and heat treatment at 400 °C in air. The rutile whiskers (needles) thus obtained are agglomerated to form cauliflower-like aggregates of several micrometer size. The whiskers had a diameter of 4−6 nm in the ab-plane and a length of ∼50 nm in the c-direction. The BET surface area was 181 m2 g−1. The Li cyclability was studied in the voltage ranges 1−3 V and 0.1−3 V vs Li at various current rates, 0.05C to 30C (1C = 335 mA g−1), and at various temperatures from 20 °C to −40 °C. Complementary cyclic voltammetry (CV), impedance spectra, TGA/DSC of the discharged and charged electrodes are presented. The authors also fabricated and tested the Li ion cells (pouch-type) with LiFePO4 as cathode.

The interfacial storage was also considered as contributing to the total capacity. However, the reversible capacity degraded slowly up to 15 cycles and stabilized at ∼190 mA·h g−1 in the range 17−30 cycles. The performance of anatase nanosheet of TiO2 without graphene was found to be inferior; even though the initial reversible capacity is 202 mA·h g−1, only 63% of it was retained after 30 cycles (Table 1). The C-rate capability was tested on both phases, up to 50C, and the G-TiO2 was found to perform better. Myung et al.298 prepared nanorods of anatase TiO2 by hydrothermal reaction of amorphous TiO2 in NaOH solution to form nanorod Na2Ti3O7, followed by ion exchange in HCl solution, to give H2Ti3O7 and dehydration at 400 °C. The nanorod morphology was retained through ion exchange and dehydration treatment. The nanorod anatase TiO2 particles are 40−50 nm in length and 10 nm in diameter and had a BET surface area of 185.5 m2 g−1. When cycled between 1 and 3 V vs Li at a current of 50 mA g−1, the first-discharge and -charge capacities were 320 (composition Li0.95TiO2) and 265 mA·h g−1, respectively, and the voltage-capacity profiles are typical of the two-phase reaction with voltage plateaus at ∼1.75 and ∼1.9 V. The cycling stability was good, even though there was a slight decrease in the extent of the plateau voltages, and a reversible capacity of 225 mA·h g−1 was observed after 50 cycles, corresponding to a capacity fade of 15%. 2.1.1.2. Rutile TiO2. The rutile polymorph is the stable form of TiO2. The anatase form transforms to rutile form by heating to temperatures T > 700 °C and also by high-energy ball milling. The tetragonal structure (a = b = 4.593 Å, c = 2.96 Å, cell volume V = 62.8 Å3) is made up of TiO6 octahedra that share edges to form single chains extending along the c-axis giving a (1 × 1) tunnel structure (Figure 2). The Li ions can occupy these tunnels during intercalation. Normally, microcrystalline rutile TiO2 does not show any significant amount of Li storage and cycling at ambient temperature. This is due to the peculiar tetragonal structure, 5373

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Main results are as follows: (1) CV curves showed two highintensity cathodic peaks at ∼1.4 and ∼1.1 V, which are ascribed to the irreversible phase transition from TiO2 to LiTiO2 (Figure 6). In addition, under deep discharge to 0.1 V, two low-

Figure 7. Specific capacities obtained at different temperatures (at C/5 rate) in two (0.1−3 V and 1−3 V) investigated potential windows. Reproduced with permission from ref 303. Copyright 2011 Elsevier.

first 200 cycles, from 150 to ∼120 mA·h g−1, a slow degradation up to 750 cycles and finally stabilized at 105 mA·h g−1 in the range 750−1000 cycles. This corresponds to a capacity retention of 70%. (5) TGA/DSC of the ex-situ electrodes discharged to 0.1 V vs Li showed two exothermic peaks at ∼130 and at ∼190 °C with low ΔH values of 18.3 and 3.0 J g−1, respectively, and are ascribed to the decomposition of the surface film formed on the electrode. These values are very small in comparison to the ΔH of 202.5 J g−1 shown by the well-known anode material, LixC (graphite) in the temperature range 220−290 °C. It must be mentioned that the mechanism of Li cycling of nanorutile TiO2 in the extended voltage range 0.1−3 V reported in the above studies has not been discussed. Normally, amorphization and crystal structure destruction of TiO2 is expected under deep discharge conditions to 0.1 V vs Li. The reasons for the first-cycle cathodic CV peaks below 1 V and their disappearance during subsequent cycling (Figure 6) and the unusually large galvanostatic first-discharge capacity (∼2 mol of Li at 0.2C and at 20 °C) mentioned above (Figure 7) are difficult to explain by the Li intercalation mechanism. It is possible that amorphization of nanorutile TiO2 occurs during the first discharge at least in the few surface layers, resulting in the formation of a composite, Li2O·TiOy (y ≤ 1.5), and Li cycling occurs through a “conversion reaction”. Detailed structural studies are needed to clarify the mechanism. Kinetics of Li intercalation into nanorutile TiO2 has been studied recently by Bach et al.305 by impedance spectroscopy. The compound was prepared by hydrolysis of TiCl4 and heating in air at 100 °C and consisted of aggregates of spherical particles with a mean diameter of 1.5−3 μm. Each particle was made up of a collection of thin nanorods of thickness 20 nm. The residual water content in the compound was not estimated. Li cycling in the range 1−3 V vs Li at current rates from 0.03C to 1C has been studied up to 40 cycles. Reversible and stable capacities were noted. For example, at 1C, a capacity of 80 mA·h g−1 was stable between 2 and 40 cycles. The Li ion diffusion coefficients (DLi+) were extracted from an analysis of the impedance spectra measured for various x in LixTiO2. The DLi+ at ambient temperature for x = 0.1 was found to be fairly high, 7 × 10−8 cm2 s−1, and the value decreased to 1 × 10−9 cm2 s−1 for x = 0.4. It linearly decreased with an increase in x and

Figure 6. Cyclic voltammograms of the TiO2 rutile between 1 and 3 V (a) and 0.1 and 3 V vs Li (b) collected at room temperature with a scan rate of 0.05 mV s−1. Reproduced with permission from ref 303. Copyright 2011 Elsevier.

intensity peaks were noted at ∼0.75 and at ∼0.4 V. The firstcycle oxidation and subsequent cycling showed broad reduction and oxidation peaks at ∼1.8 V, consistent with the Li deintercalation and intercalation in LiTiO2. (2) Galvanostatic cycling at 0.2C at 20 °C in both the voltage ranges reflected the behavior as shown in CV and gave large first-discharge capacities of 380 and 660 mA·h g−1 for the lower cutoff voltages of 1 V and 0.1 V, respectively. These values correspond to ∼1.1 and ∼2 mol of Li per mol of TiO2, respectively. Large irreversible capacity loss was also noted in the first discharge− charge cycle in the above voltage ranges. After five cycles, the Coulombic efficiency improved to ∼95−97% and reversible capacities of 183 mA·h g−1 (0.55 mol of Li) and 324 mA·h g−1 (0.97 mol of Li) were observed in the voltage ranges 1−3 V and 0.1−3 V vs Li, respectively. As can be seen from Figure 7, the cycling performance at temperatures from the ambient to −40 °C is very good. For example, at −20 °C, a capacity of 80 and 140 mA·h g−1 was noted in the above voltage ranges, corresponding to ∼40% capacity retention. (3) The C-rate capability was also found to be excellent, with ∼50% and ∼24% capacity retention at 20C rate, in comparison to the values at 0.05C rate. (4) Long-term cycling studies, up to 1000 cycles at 5C rate and at 20 °C, showed a fading of capacity during the 5374

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

5375

335 325

333 334

2−200) 2−40) 2−50) 2−50) 2−50) 2−80) = = = = = = (n (n (n (n (n (n 94% 76% 40% 78% 95% 93% particles nanotubes

7 216

10C 35 35 35 0.16C 20

bulk nanowires 93% TiO2-B, 7% TiO2 anatase and 5−10 nm sized gold particles nanosheets nanoparticles

11

151

5−10 nm thickness 10 nm (XRD) 16 nm 33 nm needle-like structure, 2 μm length and 250 nm diam diam 10 nm, thickness 2.5 nm

216 210 148 45 315 237

330 97% 84% (n = 2−100) 160 232 30 30−80 nm diam

2−100) 2−10) 2−10) 2−10) 2−10; 1.2−3.6 V) = = = = = (n (n (n (n (n 90% 82% 72% 86% 98% 98% 205 190 256 200 253 180 50 111 (0.33C) 0.1C 0.2C 100 140 126 200 66 77 20−40 nm 30 nm diam 1−2 μm length 1 μm spheres 5−10 nm nano grains 20−25 nm size nanosheets 1−10 Plate-like morphology, ∼1 μm in length, ∼200 nm width, ∼200−300 nm in thickness nanowires nanoribbons (95% TiO2-B + 5% anatase TiO2) mesoporous microspheres (87%TiO2-B + 13% anatase) nanoparticles KNS (potassium exploliated nanosheets) NS

particle/crystallite size (nm) morphology

Table 3. Physical Properties and Electrochemical Li Cycling Data of TiO2-B

surface area, m2 g−1

current rate, mAg−1 (1C = 335 mA g−1)

reversible capacity (mA·h g−1)

voltage, range 1−3 V vs Li (cycling range, capacity retention after n cycles)

ref

finally at x = 0.8, attained a value of ∼5 × 10−10 cm2 s−1. Thus, the DLi+ decreased by 2 orders of magnitude over the composition range x = 0.1−0.8. It must be mentioned that the above DLi+ are more than 2 orders of magnitude larger than the values reported in the literature, ∼1 × 10−13 cm2 s−1, for the micrometer-size rutile TiO2, thereby explaining the observed facile Li cycling in the case of nanorutile TiO2. For comparison, the DLi+ for the micrometer-size anatase TiO2 are found to be in the range 5 × 10−12 to 1 × 10−11 cm2 s−1. From the impedance spectra measured on Li0.3TiO2 in the T range 10−50 °C, the extracted DLi+ were found to obey the Arrhenius equation with an energy of activation, Ea = 0.35 eV, which is much less than the Ea = 0.85 eV reported for the micrometer-size rutile TiO2. Bach et al.305 also studied the impedance spectra as a function of cycle number. The extracted DLi+ in the first cycle was 4 × 10−9 cm2 s−1, and it increased to ∼2 × 10−8 cm2 s−1 at the second cycle and remained unchanged at the end of 10 and 20 cycles. This 5fold increase in DLi+ confirms the fact that after the initial cycle, Li cycling proceeds in the newly formed “LiTiO2” thereby giving stable and high reversible capacities. 2.1.1.3. TiO2-B. The polymorph, TiO2-B (bronze) is composed of corrugated sheets of edge- and corner-shared TiO6 octahedra, which form an open framework structure consisting of perovskite-like (ReO3-type) pathways along which Li can be reversibly intercalated/deintercalated (Figure 2). It adopts a monoclinic structure with the space group, C2/m and has a lower density (3.7 g cm−3) in comparison with the anatase (3.89 g cm−3) and rutile (4.25 g cm−3) polymorphs, and thus, the open framework structure enables easy Li ion transport within it. TiO2-B can only be synthesized by the acid (proton) exchange of sodium titanate, Na2Ti3O7, followed by heat treatment at temperatures ranging from 400 to 500 °C in air/ argon. Unfortunately, this leads to residual adsorbed/absorbed H2O in the TiO2-B and formation of non-negligible amounts of the anatase TiO2. Nevertheless, TiO2-B is an attractive anode material due to its lower intercalation/deintercalation voltage and realization of larger reversible capacity.169 Electrochemical Li cycling data of TiO2-B compounds are summarized in Table 3. Bruce and co-workers312 studied the Li cyclability of TiO2-B in the form of bulk (microcrystalline) and nanowires (20−40 nm diameter and several micrometers in length) and also fabricated and tested Li ion cells with TiO2-B as anode and LiFePO4 and Li[Ni0.5Mn1.5]O4 as the cathodes. The group of Gratzel313 also studied Li cyclability of TiO2-B and stressed the importance of pseudocapacitive Li storage in it. The capacity vs voltage profiles of TiO2-B at a current of 10 mA g−1 between 1 and 3 V vs Li from the work of Armstrong et al.312 are shown in Figure 8. It is clear from Figure 8a that the nanowire TiO2-B performs better than the bulk form, with first-discharge and -charge capacities of 305 (Li0.91TiO2) and 280 mA·h g−1 (Li0.84TiO2), respectively, with a loss of only 0.07 mol of Li in this cycle. The smoothly varying discharge and -charge profiles indicate a quasi-single-phase reaction with pseudocapacitive behavior. From the differential capacity (incremental capacity) vs voltage plot of Figure 8b, it can be seen that the main intercalation/deintercalation voltages are between 1.5 and 1.6 V. The additional oxidation/reduction peaks at ∼1.9 and ∼1.8 V could be due to minor structural readjustment in the lattice or the presence of anatase TiO2 impurity in the compound.312 A recent EXAFS study on the system Lix−TiO2B (micrometer-size) by Okumura et al.314,315 has shown that

312 223 316 326 328

Review

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Recently, Liu et al.316 reported the Li cycling properties of mesoporous microspheres of TiO2-B (Table 3). The phase was prepared by a five-step template-assisted ultrasonic spray pyrolysis method followed by refluxing, ion-exchange, and heat treatment at 500 °C for 1 h in flowing Ar-gas atmosphere. The product thus obtained was composed of spherical particles ∼1 μm diameter, uniform mesopores of size 12 nm, and crystal grains of ∼6 nm, showed a BET surface area of 126 m2 g−1, and contained ∼13 wt % anatase TiO2 as a second phase. When cycled between 1 and 2.5 V vs Li at 0.1C rate (1C = 335 mA g−1), the first-discharge and -charge capacities were 311 (93% theoretical) and 256 mA·h g−1, respectively. Even though the reversible capacity degraded somewhat at this current rate, the cycling stability improved remarkably for higher C-rates, from 0.5C to 60C. For example, at 10C, the mesoporous-microsphere TiO2-B showed a sixth cycle reversible capacity of 166 mA·h g−1 and retained a value of 149 mA·h g−1 at the end of 5000 cycles, showing only ∼10% capacity fading. The Coulombic efficiency approached ∼100%. Similarly, at 30C and 60C rates, capacities of 130 and 115 mA·h g−1, respectively, that were stable up to at least 10 cycles were noted. The authors stressed that the Li storage and cycling in TiO2-B occurs mainly via pseudocapacitive Faradaic reaction. Akimoto et al.317 prepared micrometer-sized TiO2-B by heating H2Ti3O7 at 350−400 °C in air. The latter compound was obtained by the proton exchange of Na2Ti3O7. In addition, they also prepared crystalline H2Ti6O13 and H2Ti12O25 by heat treatment of H2Ti3O7 at 140−200 °C and at 200−260 °C, respectively, in air. TGA, MAS NMR, and FT-IR spectra showed that these hydrogen titanates did not contain H2O or H3O+ species, and the local structure of H-atoms in them differ from each other. Li cycling has been investigated in the voltage range 1−3 V vs Li at 10 mA g−1, which showed that TiO2-B transforms to anatase TiO2 after a few cycles, and a reversible capacity of ∼150 mA·h g−1 was observed up to 10 cycles. Tests on the H2Ti3O7 showed that after three cycles, H+ to Li+ exchange occurs resulting in the formation of Li2Ti3O7. The reversible capacity slowly degraded from ∼120 to ∼100 mA·h g−1 during 2−10 cycles. A flat voltage plateau was observed at ∼1.55 V with a firstdischarge capacity of 283 mA·h g−1 (5.25 mol of Li per formula unit) in the case of H2Ti6O13.317 However, the plateau disappeared during the following Li extraction and the voltage-capacity profiles adopted a single sloping curve in the voltage range 1.4−2.1 V. The reversible capacity stabilized at ∼190 mA·h g−1 during 2−10 cycles, and the particles retained 92% of it after 50 cycles. The best Li cycling behavior was shown by H2Ti12O25 with a flat voltage plateau at ∼1.55 V with a first-discharge capacity of 236 mA·h g−1 (8.6 mol of Li per formula unit).317 With only a small capacity loss in the first cycle, the second cycle reversible capacity of ∼205 mA·h g−1 (0.62 mol of Li per Ti atom) remained stable during 2−10 cycles, and 96% of the capacity was retained after 50 cycles. Exsitu XRD of the cycled electrodes showed that the framework structure of the H2Ti6O13 and H2Ti12O25 was retained. Synthesis, structural, and electrochemical studies of the ionexchanged Li2Ti6O13 titanate have been reported by PerezFlores et al.318,319 Myung et al.298 prepared nanowires of TiO2-B by heating nanowires of H2Ti3O7 at 400 °C, which were obtained by proton exchange of hydrothermally synthesized nanowire Na2Ti3O7. The nanowire morphology was retained through all the operations. The TiO2-B nanowires are several micro-

Figure 8. (a) Variation of potential vs Li/Li+ (1 M) electrode with Li content (charge passed) for TiO2-B nanowires and normal TiO2-B cycled under identical conditions. Rate (current) 10 mA g−1 and voltage limits of 1 and 3 V. Inset shows an equivalent discharge curve for TiO2-B nanowires illustrating variation in irreversible capacity. (b) Incremental capacity plots for TiO2-B nanowires and normal TiO2-B cycled at 10 mA g−1 between voltage limits of 1 and 3 V. Reproduced with permission from ref 312. Copyright 2005 Elsevier.

the Li ions are inserted into the sites with 5-fold or distorted octahedral O-coordination distributed at the vicinity of O layers parallel to the ab plane for x ≤ 0.5, whereas the Li ions are accommodated into the 5-fold coordinated sites distributed at the vicinity of TiO2 layers parallel to the ab plane for x > 0.5. The interaction between the inserted Li ion and TiO6 octahedra affects the lattice distortion for x > 0.5 and could be the reason for the split peak at ∼1.5 and ∼1.6 V during discharge and charge, respectively in the incremental capacity vs voltage plot (Figure 8b). Beauvier et al.223 prepared nanoribbons of TiO2-B by the solution-refluxing method (120 °C for 7 days) followed by heat treatment at 400 and 500 °C to yield 96% nanoribbons, 4% nanotubes, and 1% nanospheres in the 500 °C heated product (Table 3). The nonporous nanoribbons were ∼30 nm width, ∼6 nm thick, and 1−2 μm in length, showed a BET surface area of 115 m2 g−1, and had some adsorbed H2O and CO2 and contained ∼5% anatase TiO2. The nanoribbon TiO2-B electrodes with 50 wt % carbon black loading showed the best Li cycling performance when cycled between 1.25 and 2.5 V vs Li: reversible capacities of 200, 150, and 100 mA·h g−1 were obtained at current rates of 0.33C, 3C, and 15C, respectively (1C = 330 mA g−1). Further, a capacity loss of only 5% was observed up to 500 cycles at the 3C rate. 5376

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 9. Crystal structure of Li4Ti5O12 (LTO). Pink spheres, Li; gray spheres, Ti; red spheres, O. Reproduced with permission from ref 172. Copyright 2010 Elsevier.

galvanostatic studies on the M-Nb2O5 showed main redox peak at ∼1.65/1.70 V vs Li and a flat two-phase region. The Li cycling behavior of micrometer size Nb2O5 phases were reported by the group of Kumagai346−348 and of Nb2O5 in the form of nanobelts349 and nanofibers/nuggets350,351 and as thin films.352−354 The group of Kumagai348 studied the Li cycling of O-, T-, and M-Nb2O5 in the voltage range 1.2−3.0 V vs Li at current rate of 0.2 mA cm−2 and found that O-Nb2O5 showed a stable capacity of 155 mA·h g−1 up to 40 cycles and T-Nb2O5 showed stable reversible capacity of 190 mA·h g−1 up to 30 cycles. The M-Nb2O5 showed an initial discharge capacity of 190 mA·h g−1, and it retained a capacity of 170 mA·h g−1 at the end of 20 cycles. The authors discussed the Li cycling mechanism using in situ X-ray diffraction and X-ray absorption spectroscopy techniques. Wei et al.349 reported the Li cycling of Nb2O5 nanobelts, which showed an initial discharge (intercalation) capacity of 250 mA·h g−1 in the voltage range of 1.2−3.0 V, at 100 mA g−1, which corresponds to x = 2.5 in LixNb2O5. A stable capacity of 180 mA·h g−1 was obtained at the end of 50 cycles. The cyclic voltammetry of Nb2O5 at 0.1 mV s−1 in the potential range 1.0−3.0 V vs Li showed the reduction of Nb5+ ↔ Nb4+ and Nb4+ ↔ Nb3+. The potentials corresponding to these reduction reactions are 1.62 and 1.33 V, respectively. During oxidation, peaks at 1.73 and 1.42 V were noticed, which correspond to the reverse reactions, respectively. Recently, the group of Chowdari350 prepared Nb 2O 5 nanofibers/nuggets by electrospinning followed by heating at temperatures in the range 500−1000 °C for 1 h in air. They reported that the M-Nb2O5 showed a discharge capacity of 242 and 218 (±3) mA·h g−1 at the end of second and 25th cycle, respectively, when cycled at a current of 50 mA g−1, in the voltage range 1.0−2.6 V vs Li. They also observed improved cycling stability with the heat-treated electrode (at 220 °C, in Ar for 6 h) and reported a reversible capacity of 228 mA·h g−1 (2nd cycle) with only 4% capacity fading between 10 and 100 cycles at the current rate of 400 mA g−1. The authors also discussed variation of peak potentials of the H-, O-, and MNb2O5 phases using combined cyclic voltammetry and differential capacity vs voltage plots studies.

meters in length and tens to hundreds of nanometers in thickness and had a BET surface area of 40.3 m2 g−1. When cycled between 1 and 3 V vs Li at a current of 50 mA g−1, the first-discharge and -charge capacities were 270 and 225 mA·h g−1, respectively. The cycling stability was found to be poor in that the reversible capacity degraded drastically to ∼100 mA·h g−1 in 20 cycles and finally to ∼75 mA·h g−1 after 50 cycles. The Li cycling behavior and changes in the Li ion−vacancy arrangement in the ramsdellite-type Li2+xTi3O7 were investigated by galvanostatic cycling, impedance spectroscopy, entropy measurement of the reaction, and Monte Carlo simulation by Cho et al.320,321 Computational studies based on density functional theory were reported by the group of Islam.322,323 Additional recent studies on the Li cycling of TiO2-B are by Saito et al.,324 Brutti et al.,325 Wessel et al.,326 Ren et al.,327 and Jang et al.,328 on TiO2-B carbon (37%) composites by Yang et al.,329 on TiO2-B@anatase hybrid nanowires by Yang et al.,330 on nanowires by Li et al.,331 on nano- and meso-TiO2-B by Dylla et al.,332 on porous nanosheets Liu et al.,333 on nanofiber bundles by Guo et al.,336 and the electrochemical cycling data are given in Table 3. 2.1.2. Vanadium (V) and Molybdenum (Mo) Oxides. There are many vanadium oxides, VO, V2O3, VO2, V2O5, and VnO2n−1 etc. Some of them, for example,VO2(B) (isostructural to TiO2(B)) and V2O5, have been investigated for their Li cyclability, via intercalation−deintercalation reactions. These have been recently reviewed by the group of Whittingham,159,165 and NMR studies on V2O5 are reviewed by the group of Grey.337 The average voltage at which Li insertion− extraction occurs varies from ∼2.2 to ∼2.6 V vs Li in these vanadium oxides; this voltage, unfortunately, is too low for use as a cathode and too high for use as an anode in the LIBs. Additional studies on V2O5 are reported in refs 338−343. Similar is the case with the layered-MoO3, with remarkably identical discharge−charge voltages.159,165,344,345 Thus, even though reversible capacities ranging from 150 to 250 mA·h g−1 have been observed, the vanadium and molybdenum oxides are unlikely to find use as anodes for LIBs by Li-intercalationdeintercalation mechanism, and hence they will not be discussed further here. 2.1.3. Nb2O5. Niobium pentoxide, Nb2O5, exists in different crystal structures, namely, pseudohexagonal (H), orthorhombic (O), tetragonal (T), and monoclinic (M) phases. Among these, the T and M phases have been investigated for Li cycling and can deliver high capacity. The cyclic voltammetry and

2.2. Ternary Oxides of Ti and Nb

Several ternary titanium and niobium oxides possessing 2Dlayer or 3D-network structure have been examined in the literature for their Li cyclability via intercalation−deintercalation reactions, without affecting the host structure (topotactic reaction).355 The potentials at which Li cycling occurs are in 5377

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

better accessibility of the electrolyte to the nanoparticles, in comparison to the micrometer-size particles. The present trend is for the synthesis of micrometer-size spherical particles of LTO with a porous structure, which are composed of nanosize particles. This will enable better packing of the LTO particles in the composite electrode and, hence, improve the volumetric energy density (mA h cm−3) in addition to the gravimetric energy density (mA·h g−1). The LIBs fabricated and tested with LTO as anode will be discussed in section 6.1. Some recent studies on LTO will be discussed here. Zhu et al.357 prepared composites of 1 wt % graphene and nano-LTO, which had a surface area of 170 m2 g−1, by electrospinning and heat treatment, and TEM showed nanosized crystalline clusters that were coated by a uniform graphene sheet adjacent to their crystal phase. The authors reported high C-rate capability and good cycling stability when the composites were cycled between 1 and 3 V vs Li in electrolyte containing LiClO4. For example, at 0.2C and at 8C (1C is not defined, but presumably = 170 mA g−1), specific capacities of 164 and 137 mA·h g−1, respectively, were noted. At 22C, the LTO−graphene composite gave a specific capacity of 110 mA·h g−1, which is more than twice the capacity of 47 mA·h g−1 measured on the elctrospun LTO fibers and three times the capacity of 34 mA·h g−1 obtained on the sol−gel derived micrometer-size LTO powder. After 1300 cycles, the composite retained 91% of the initial capacity. It is a bit surprising that 1 wt % of graphene vastly improved the cycling performance of LTO. Kim et al.358 prepared LTO−RGO (reduced graphite oxide) nanohybrid composite (72:38 wt %) by microwave-assisted solvothermal reaction followed by heat treatment at 700 °C in H2/Ar atmosphere. TEM showed uniform dispersion of LTO nanoplatelets, sizes of 10−20 nm, on the RGO nanosheets with a 3D-nanoporous structure. Li cycling of the electrodes fabricated without the addition of any conducting carbon on Ti-substrate was carried out in the range 1−2.5 V vs Li in the electrolyte LiClO4 dissolved in propylene carbonate at various C-rates. For example, reversible capacities of 154, 142, 128, and 101 mA·h g−1 were measured at current rates 1C, 10C, 50C, and 100C, respectively (1C is not defined, but presumably = 170 mA g−1). The capacity retention at 1C and 10C rates are 95% and 96%, respectively, after 100 cycles, indicating very good stability. The authors also found that the hysteresis between the discharge and charge voltage plateaus at 1C rate is small, ΔV = 0.028 V with a discharge voltage of 1.543 V. Jung et al.359,360 prepared bare and C-coated LTO by a twostep process involving solid state reaction and pitch carbon. The 5 μm sized spherical particles are composed of 100 nm sized primary particles, and the nominal 5 wt % C-coated LTO had a uniform 3 nm thick carbon coating (Figure 10). While the BET surface areas did not vary much before and after Ccoating (∼12 m2 g−1), the electronic conductivity (σelec) increased by 6 orders of magnitude in the C-coated LTO, from 1.58 × 10−9 to 5.24 × 10−3 S cm−1 at 300 K. Due to the spherical morphology and good conductivity, the 5.2 wt % Ccoated LTO showed excellent cycling stability up to 100 cycles when cycled at 1C, 5C, and 10C rates (1C = 170 mA g−1) in the voltage range 1−3 V vs Li (Figure 11). For example, an initial capacity of 165 mA·h g−1 at 1C rate remained constant up to 100 cycles, whereas the bare LTO showed much smaller capacity, 55 mA·h g−1, which also remained constant till 100 cycles. It is also clear from Figure 11 that the 5.2 wt % coated

the range 1.3−1.8 V vs Li, and hence they will be of interest as anodes for LIBs. Of course, one disadvantage with these oxides is that they are electronic insulators with very low electronic conductivity (at 300 K, σelec ≈ 1 × 10−9 to 10−13 S cm−1). 2.2.1. Li4Ti5O12 (LTO). Lithium titanium oxide with the cubic spinel structure, Li4Ti5O12 (LTO), has been extensively investigated for Li storage and cyclability by the groups of Dahn, Thackeray, Ohzuku, and others for the last 15 years in the form of micro- and nanocrystalline particles and with various morphologies, as composites and as those coated with conducting additives like carbon. Hundreds of papers have appeared on the cycling behavior of LTO. Yang et al.169 and Yi et al.172 have recently reviewed its outstanding Li cycling properties, which show that pure and modified LTO are viable anodes for LIBs. The main findings are summarized here: (1) The formula can be written as Li[Li1/3Ti5/3]O4 indicating that Li ions adopt both tetrahedral (occupy the 8a sites) and octahedral (occupy the 16d sites) oxygen coordination (Figure 9). It can accommodate 3 mol of Li per formula unit of LTO via intercalation giving a theoretical capacity of 175 mA·h g−1. Near-theoretical reversible capacities have been realized by many groups during initial discharge−charge cycles. (2) Analogous to the behavior of anatase TiO2, during Li insertion (discharge process), the virgin oxide phase separates into a Lirich phase (Li ∼3 (Li 4 Ti 5 O 12 )) and a Li-poor phase (Li∼0.02(Li4Ti5O12)), and this equilibrium persists until the completion of the reaction to yield pure Li-rich phase. Similarly, Li extraction (charge process) takes place through the above two-phase reaction. Accordingly, Li cycling occurs at ∼1.5 V vs Li, which is characteristic of LTO, with only a small hysteresis between the discharge and charge process.169,356 (3) Interestingly, Li cycling involves very little change in the cubic lattice parameter, namely, a = 8.3595 Å for Li4Ti5O12 and a = 8.3538 Å for Li7Ti5O12. Thus, it is a “zero-strain” material and ideally suited as an anode for LIBs. (4) The Li ion diffusion coefficient (DLi+) is fairly high (∼1 × 10−8 cm2 s−1 at 300 K) and thus can sustain high current (C) rates of discharge and charge. Stable and reversible capacities as high as 100 mA h g−1 have been achieved at 15−30C rate when cycled between 1 and 3 V vs Li. (5) Amorphization (crystal structure destruction) does not occur when cycled at high C-rates, as happens in many other titanium oxides. (6) The discharged compound shows good thermal stability at high temperatures. (7) The Li ion conductivity is small, σdc ≈ 3 × 10−10 S cm−1 at 300 K, and the electronic conductivity is still smaller by 2 orders of magnitude, σelec ≈1 × 10−12 to 10−13 S cm−1 at 300 K. This is in tune with the electronic band structure of LTO characterized by a filled O 2p band and an empty Ti 3d conduction band (t2g band or localized levels) with a band gap of 2−3 eV. Li intercalation will result in an increase in σelec by 5−6 orders of magnitude, due to the creation of Ti3+ ions (d1 system) and thereby enabling the hopping of electronic charge carriers between the Ti3+ and Ti4+ ions in the lattice. In order to improve the σelec of virgin LTO and thereby increase its C-rate capability, conductive coatings, such as carbon, Ag, TiN, or polyaniline, have been successfully applied. In addition, doping of LTO by aliovalent ions at the Ti site has also been tried to improve the electronic conductivity. (8) Using nanotechnology, nano-LTO of various morphologies has been prepared by a variety of methods, and these showed vastly improved reversible capacities and C-rate capability. This is due to the reduction in the Li ion diffusion pathway in the particles and 5378

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

N-doped carbon were obtained. The authors also prepared porous bare LTO, as well as sugar-derived C-coated LTO, which does not contain nitrogen. The surface areas of bare and N-doped C-coated LTO were 43.4 and 51.3 m2 g−1, respectively. The XPS analysis of the latter compound showed the presence, on the surface, of N-containing carbon species, C−N, CN, and N−O bonds, and a Ti−N−C-like compound. However, the amount of N-doping in the carbon is not known. The Li cycling properties of bare, sugar-C-coated, and N-doped C-coated LTO were carried out in the voltage range 1−2.2 V vs Li at various current (C) rates, 0.5C, 1C, 5C, and 10C (1C = 170 mA g−1). Results showed that the 7 wt % N-doped C-coated LTO performed excellently well in comparison to the 3.6 and 9.6 wt % doped compositions, as well as those of the bare and sugar-C-coated LTO. For example, at 5C and 10C rates, capacities of 145 and 129 mA·h g−1, respectively, which were stable for at least 20 cycles, were measured. Under similar conditions, bare LTO showed only 60 and 15 mA·h g−1 and the 7 wt % sugar-C-coated LTO showed 71 and 18 mA·h g−1, respectively. The long-term cycling performance, up to 2200 cycles, at 2C rate was found to be very good with an initial capacity of 150 mA·h g−1 that gradually decreased to 124 mA·h g−1, showing a capacity retention of 83% (Figure 12). On the other hand, bare LTO yielded a

Figure 10. SEM images of carbon-free (a) and 5 wt % pitch-coated Li4Ti5O12 powders (b). High-resolution TEM images of the 5 wt % pitch-coated Li4Ti5O12 after calcination at 750 ◦C (c). Reproduced with permission from ref 359. Copyright 2011 Elsevier.

Figure 12. Cycling performance of coated Li4Ti5O12 (sample 2). Inset, the cycling performance of pristine Li4Ti5O12. Reproduced with permission from ref 361. Copyright 2011 Wiley-VCH.

capacity of 135 mA·h g−1 at 2C, which decreased to 98 mA·h g−1 after 100 cycles, with a moderate capacity retention of 73% (Figure 12, inset). The authors claimed that the N-doped Ccoating yields a 3D mixed conducting network in the porous LTO and is responsible for the vastly superior Li cycling performance as an anode. Template-free solvothermal synthesis of mesoporous carboncoated LTO has been reported by Shen et al.362 using tetrabutyl titanate, lithium acetate, furfural, and ethanol in an autoclave at 180 °C and subsequent heat treatment at 550 °C in nitrogen atmosphere. The 5 wt % C-coated LTO was composed of 0.5−1 μm sized porous microspheres with a pore diameter ∼4.3 nm, and each sphere was composed of small crystallites (size ∼11 nm) uniformly coated with ∼2 nm thick carbon. The surface areas of C-coated LTO and bare LTO (prepared without the addition of furfural) were 159.4 and 61.7 m2 g−1, respectively. When cycled in the voltage range 1−2.5 V vs Li at various C-rates, 0.2−50C (1C = 170 mA g−1), the Ccoated LTO performed much better than bare LTO. For example, the former compound gave capacities of ∼158 and ∼100 mA·h g−1 at 1C and at 50C rates, respectively, without

Figure 11. (a) Rate capability of 5.2 wt % C-coated Li4Ti5O12 and (b) cyclability of 5.2 wt % C-coated Li4Ti5O12 from 1C rate (0.17 A g−1) to 10C rate (1.7 A g−1) in comparison with C-free Li4Ti5O12 cycled at 1C rate (0.15 A g−1). Charged at 1C rate. Reproduced with permission from ref 360. Copyright 2011 The Royal Society of Chemistry.

LTO gives high capacities at C-rates as high as 160C and very good cyclability at 5C and 10C rates up to 100 cycles. The group of Chen361 prepared porous LTO by spray drying method and coated it with nitrogen-doped carbon using an ionic liquid, 1-ethyl-3-methylimidazolium dicyanamide, and subsequent heat treatment at 600 °C in Ar atmosphere. Spherical particles, size 3−5 μm, composed of nanosize aggregates uniformly coated with a thin amorphous layer of 5379

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

by Prakash et al.;367 Nb-doped LTO by Tian et al.368 and Yi et al.;369 Zr-doped LTO by Gu et al.;370 V-doped LTO by Yu et al.;371 Carbon-coated Sn-doped LTO by Zhang et al.;372 sawtooth-like LTO nanosheets by Chen et al.;373 porous LTO by Lin and Duh;374 LTO-nanostructures by Yu et al.,375 Feckl et al.,376 and Lu et al.;377 nanoflower-like LTO by Lin et al.;378 dual-phase LTO-TiO2 by Li et al.;379LTO−TiO2−C by Rehman et al.;380 LTO microspheres by Chou et al.381 and Zhang et al.;382 hierarchical mesoporous nest-like LTO by Chen et al.;383 mesoporous LTO by Lin et al.;384 film shaped LTO by Mosa et al.;385 effect of current collectors on power performance by Wu et al.;386 electrospun LTO by Guo et al.387 and Wu et al.;388 supercritical synthesis of LTO by Laumann et al.;389 microwave-assisted zinc nanocoating by Hsieh et al.;390 spray dried LTO by He et al.;391 LTO−C composites by Cheng et al.,392 Wang et al.,393 Zhu et al.,394 Hu et al.,395 Li et al.,396 Kang et al.,397 He et al.,398 Ogihara et al.,399 Xie et al.,400 Li et al.,401 and Zhang et al.402 LTO/multiwalled carbon nanotube composite by Dominko et al.,403 Jhan and Duh,404 Shen et al.,405 and Ni et al.;406 LTO/graphene composites by Shi et al.,407 Xiang et al.,408 Han et al.,409 Jian et al.,410 and Tang et al.;411 local structure, Li dynamics, and diffusion coeffecients by NMR spectroscopy studies and electroanalytical techniques of LTO by Wilkening et al.,412 Vijayakumar et al.,413 Hain et al.,414 Rho et al.,415 and Wunde et al.416 The electrochemical performances of selected compounds are summarized in Table 4. From the above discussion, it is clear that studies on LTO will be continued in future to optimize its Li cycling properties and put it to use as anodes in commercial LIBs, especially for HEVs. 2.2.2. MgTi2O5, LiTiNbO5, TiNb2O7, and Other Oxides. Magnesium titanate, MgTi2O5, crystallizes in the pseudobrookite structure containing one-dimensional channels. Anji Reddy et al.432 studied the Li cycling properties of micrometer-size and nano-MgTi2O5. While micrometer-size particles did not show any reversible Li cycling, nano-MgTi2O5 showed a reversible capacity of 130 mA·h g −1 (1 mol of Li) stable up to 50 cycles when cycled in the voltage range 1.0−3.0 V at 0.1C rate. The average discharge−charge voltage is ∼1.7 V. The oxide LiTiNbO5 with a layered structure has been prepared by ion-exchange method and its Li cyclability was studied by Colin et al.433 They found that 0.8 mol of Li can be cycled at 1.67 V vs Li, leading to a capacity of ∼115 mA·h g−1 during the initial cycles, which slowly degraded to ∼90 mA·h g−1 after 40 cycles when cycled in the voltage range 1.0−3.0 V at 0.1C rate. Titanium phosphate, TiP2O7, adopts a network structure and its Li cyclability has been studied.434−437 Results showed that ∼0.8 mol of Li can be cycled at 0.1C between 1.5 and 3.5 V. However, the cycling occurs at ∼2.6 V vs Li, and because of this fairly large voltage, it is unlikely to find use as an anode for LIBs. Lithium titanium phosphate, LiTi2P3O12, adopts NASICONtype (sodium superionic conductor) framework structure, and its electrochemical properties were evaluated.434,438−442 Li cycling studies showed main reduction and oxidation peaks at ∼2.38 and ∼2.6 V involving a two-phase reaction mechanism smilar to TiP2O7. A discharge capacity of ∼120 mA·h g−1 was observed at a current rate of 0.1C (1C = 138 mA g−1) when cycled in the range 2.0−3.4 V. The mixed oxide AlNbO4 is isostructural to VO2 (B) and is thus amenable to Li cycling. Though earlier studies on the

any noticeable fading over 100 cycles. Bare LTO showed smaller but stable capacities of ∼150 mA·h g−1 at 1C and ∼95 mA·h g−1 at 20C. The impedance spectral data showed that the C-coated LTO had a smaller charge-transfer resistance, ∼50 Ω, in comparison to that shown by bare LTO, namely, ∼140 Ω. Raman microspectrometry and colorimetry have been used by Schneider et al.363 to characterize the Li insertion/extraction to/from LTO by in situ study of specially fabricated electrochemical cells. LTO exhibits five Raman-allowed phonon modes in the range 200−700 cm−1. Three of these bands are of high intensity. Studies showed that a pronounced fading of the intensities of all the bands occurs even at 20% state of charge (SOC), namely, x ≈ 0.5 in Li4+xTi5O12 for the samples studied by the authors. Only the bands due to the electrolyte remained in the spectrum (Figure 13). This continues till SOC reaches

Figure 13. In situ Raman spectroscopy of lithium titanate (LTO) electrodes, showing the evolution of Raman signals during charging (lithiation) from 2.2 V vs Li/Li+ (SOC = 0%, top curve) to 1.3 V vs Li/Li+ (SOC = 100%, middle) and back to 2.2 V vs Li/Li+ (SOC = 0%, bottom). Electrolyte signals are marked with an asterisk. Reproduced with permission from ref 363. Copyright 2011 Elsevier.

100% (x ≈ 2.6). Upon complete Li extraction, the Raman bands regained their original intensity, confirming good reversibility of Li cycling in the LTO. The rapid decrease in the intensities of the bands upon Li insertion is attributed to the drastic increase in the σelec by several orders of magnitude, which in turn reduces the optical skin depth of the laser used for recording the spectra and also due to the change in the color of the sample. Pristine LTO is white, and the composite electrode with ∼7 wt % graphite, used as a conducting additive, is light-gray in color. Upon Li insertion, the color changes to light-to-dark blue for SOC of ∼20−40% and finally to nearly black for SOC of 100%. Based on these observations, Schneider et al.363 developed a colorimetric method to determine the state of SOC in LTO. Most accurate estimates were obtained for SOC between 0% and ∼50%, the region in which the calibration curves showed the steepest slopes. Other reports on Li intercalated LTO by Raman spectroscopy are nicely summarized in an excellent review by Baddour-Hadjean and Pereira-Ramos.173 Kinetics of two phase reaction, size effects, and Li ion mobility studies of LTO using neutron diffraction and NMR techniques reported were by the group of Wagemaker and Mulder.364−366 Additional recent works on the Li cycling of LTO can be found in the following: solution-combustion synthesis of LTO 5380

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Table 4. Physical Properties and Summary of Electrochemical Cycling Data of Li4Ti5O12 morphology submicrosized LTO (C) LTO

Mg-substituted LTO spherical LTO 1D nanostructure LTO

particle/crystallite size (nm)

submicrometer 0.7 μm 8.5 μm nanowires, 10 nm diam, several micrometers length

spherical LTO/C composites

avg particle size: 5.5−14.1 μm

LTO nanowire LTO nanoflowers

150 nm diam 300−500 nm diam, sheets of several nanometers thick

mesoporous LTO microspheres

spheres, 300 nm, primary particles, 20 nm

porous 3D network LTO

crystallite size 20−50 nm

4 μm spheres, 33 nm paticles fibers 1 μm diam 1−2 μm spherical particles; 10 nm crystallites ∼2 μm (consists of 5−10 nm) 5 μm-size spherical particles, 100 nm-sized primary particles 200 nm sized particles 70 μm thick electrode ∼50 nm size 10 nm thick 100−300 nm size

reversible capacity (mA·h g−1)

cycling range, 1−3 V vs Li (capacity retention after n cycles)

ref

140

98% (n = 35−100; V = 1.5−2.5 V)

417

1 A·h at 1500 cyc.

418

0.25 mAcm−2 (C/2) 1C

87% 99% 94% 86%

2−8) 2−100) 2−100) 1−15)

419 420

0.3C

155 1.75 A·h at 1st cyc. 169 165 ∼150 142

421

1C 0.056C 0.5C

115 150 140

96% (n = 16−45) 86% (n = 1−70) 73% (n = 1−30)

422 423

1.5C 0.08 mAcm−2

87 153.1

51% (n = 1−30) 97% (n = 1−20)

424

0.32 1.6 3.2 0.1C 0.2C

136.9 128.8 114.8 165 167

91% 89% 87% 99% 77%

1−50) 1−50) 1−50) 1−30) 1−100)

425 426

166

8C 4C

152 137

92% (n = 1−100) 90% (n = 2−200)

427

12

20C 1C

125 158

100% (n = 2−200) 98% (n = 2−100)

367

57.5

5C 10C 2C

147 140 147.4

96% (n = 2−100) 95% (n = 2−100) 95% (n = 2−200)

428

22C

101

90% (n = 2−1300)

429

8

1C

155

95% (n = 2−200)

375

43.4

2C

150

82% (n = 12−2200)

361

12

1C

165

99% (n = 2−100)

360

5C 1C 1C 1C

160 148 155 170

99% 97% 80% 99%

388 430 406

1C 1C

144 150

94% (n = 2−300) 80% (n = 2−400)

600 nm (submicrometer size)

submicrometer size 150 nm diam, avg size 2 V) of SnO2 has been shown unequivocally, by careful in situ and ex situ studies during Li cycling (eq 5). The latter process will invariably give rise to capacity fading on long-term cycling, because of the very large

presence of electrochemically formed amorphous or nano-Li2O. The Li2O behaves as buffering domain during Li−Sn alloy formation and decomposition, thereby reducing the volume variation and the capacity fading by maintaining the electrode integrity. It also provides ionically conducting medium for the Li ion migration and helps to keep the electrochemically formed nano-Sn metal particles apart and prevents their agglomeration. Detailed studies have shown that the Li alloying−dealloying reactions of Sn, Si, Sb, etc. involve large changes in the unit cell volume, as high as 300% in some cases, and this is detrimental to the long-term Li cyclability since it will give rise to “electrochemical pulverization” of the active material of the electrode resulting in loss of electrical contact between the particles and the current collector. This eventually leads to the electrode disintegration and capacity fading upon long-term cycling. In order to mitigate or reduce the above electrode disintegration, the following four approaches have been adapted: First is the use of nanosize particles of the metals or oxides or other compounds, which will enable them to absorb some of the volume changes because of the smaller number of atoms present in the nanograins and the inherent large surface area of the nanosize particles.3,5,7,16,23,33,34 The latter will enable an increased access of Li ions and their short diffusion path lengths to the alloy-forming metal particles, thereby yielding better electrode kinetics, more stable SEI formation, and an enhanced current-rate capability. The second approach is the incorporation of one or more matrix elements, which are electrochemically active or inactive toward Li, such as Ca, Co, Al, Ti, and Ni.2−5,37,145,450,451 These can help in absorbing some of the volume changes of the main alloy-forming metal, help in improving the electronic conductivity of the composite, and also act as catalysts for better Li cycling. Of course, the presence of the electrochemically inactive matrix elements will reduce the obtainable 5383

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

was the case with ball-milled SnO. The cycling was carried out at 0.2 mA cm−2 in the range 0−2.0 V. Wang et al.465 studied Li cycling of SnO and Sn·Li2O composite prepared by ball milling of SnO and Li metal. They noticed less ICL in the case of composite and better capacity retention when cycled in the voltage range 0.02−1.5 V at 0.1 mA cm−2. The composite electrode showed 170 mA·h g−1 at the 100th cycle, whereas ball-milled SnO showed only 20 mA·h g−1. Ortiz et al.466 studied nanoarchitectured TiO2/SnO composite as anode in the voltage range 0.01−1.2 V at 50 and 100 μA cm−2. The composite of Sn/SnO2 of 2 μm thickness showed a reversible capacity of 140 μA·h cm−2 and retained ∼85% of the capacity after 50 cycles. Recently, Chowdari’s group467 studied the nanocomposites SnO(V2O3)x (x = 0.25 and 0.5) and SnO(VO)0.5 prepared from SnO and V2O3/VO by high-energy ball milling (HEB) and examined the Li cycling properties by galvanostatic cycling and cyclic voltammetry. Interestingly, SnO and SnO(VO)0.5 are not stable to HEB and undergo self-oxidation−reduction to give Sn and SnO2. Ball milling of SnO(V2O3)x gives rise to a nanocomposite of SnO2 and VOx. Because SnO is a good reducing agent, it forms SnO2 by reducing V2O3 to its lower valency state (VOx, x ≈ 1.0). No lines due to Sn metal are seen (Figure 15a). Thus, the composite consists of SnO2·VOx.

unit cell volume changes accompanying not only alloying− dealloying reactions but also the conversion reactions with Sn + Li2O. Accordingly, the nature and any amount of “matrix” element and any type of morphology of the host material will not suffice to completely suppress the capacity fading. 3.1. Binary Tin Oxides

3.1.1. SnO. Tin monoxide, SnO, adopts a tetragonal structure and is made up of SnO6 octahedra. It is commercially available in microcrystalline form but can also be prepared in micro- and nanoform by various chemical methods, and capacity values are summarized in Table 5. The group of Dahn452−454 has studied the Li cyclability of micrometer-sized SnO. Courtney and Dahn453,454 found that bulk SnO can give an initial reversible capacity of 825 mA·h g−1, but the capacity fades drastically on long-term cycling in the voltage range 0.0− 1.3 V vs Li. But, when the voltage range was restricted to 0.4− 1.3 V, somewhat stable capacity was obtained up to 10 cycles. Sakamoto et al.455 studied preliminary Li cying studies of SnO prepared by precipitation method (surface area 87 m2 g−1) and a commercial sample (surface area 28 m2 g−1); they showed a reversible capacity of 657 and 590 mA·h g−1 in the voltage range 0−1.2 V at current of 0.02 mA cm−2 . Branci et al.456 reported SnO thin films for microbattery applications. Chouvin et al.457,458 studied the Li−SnO system using X-ray powder diffraction (XRD), Sn Mossbauer spectroscopy, and X-ray absorption studies. They showed that during the first discharge, reduction of SnO occurs to form β-Sn and an intermetallic tin in strong interaction with the structural SnO. During the charging process up to 3.0 V, the mechanism is in part reversible with the re-formation of the SnO, but there is also the formation of Sn(IV) as SnO2. Aurbach et al.459 studied the Li cycling of sonochemically prepared nano-SnO along with complementary impedance and infrared spectral studies. They obtained near-theoretical reversible capacity in the first cycle, which slowly degraded upon cycling to 30 cycles in the voltage range 0.01−1.6 V at 0.1C rate. Uchiyama et al.460 reported a comparative Li cycling study of SnO of different morphologies with that of commercial SnO. They found that both meshed and flat-plate SnO showed initial capacities of 760 and 880 mA·h g−1, respectively, compared with 525 mA·h g−1 for commercial SnO, when cycled in the voltage range 0.1−1.0 V at 100 mA g−1. However, drastic capacity fading was noticed in all cases after 20 cycles. Yang et al.461 studied the Li cycling of SnO particles of different sizes in the voltage window 0.1−1.3 V; they found an initial capacity of 530 mA·h g−1 at 0.4 mA cm−2. Fairly high ICL was noted for fine-particle SnO but a comparatively better capacity retention till 50 cycles (∼350 mA·h g−1). Ning et al.462 studied the Li cycling of nanoflower SnO in the voltage range 0.01−2.0 V at 0.1C rate and showed that a fourth cycle charge capacity of ∼750 mA·h g−1 slowly decreased to 450 mA·hg−1 within 20 cycles (Table 5). Ning et al.463 prepared layered plate-like, nestlike, and stepwise-bipyramid-like SnO nanocrystals by the decomposition of tin oxide hydroxide (Sn6O4(OH)4) and examined the Li cycling properties. Results showed that irrespective of morphology, in all the cases, capacity fading was noted within 10 cycles. The group of Chen464 reported that the Li cycling properties of SnO were influenced by the particle size, as was clear from studies on ball-milled SnO. Unmilled SnO gave a first-charge capacity as high as 760 mA·h g−1, but capacity fading occurred giving only 420 mA·h g−1 at the end of the 11th cycle. Similar

Figure 15. Powder X-ray diffraction patterns: (a) nano-SnO(V2O3)0.25 and (b) nano-SnO(VO) 0.5 . Miller indices (hkl) are shown. Reproduced with permission from ref 467. Copyright 2011 Springer.

Figure 15b shows the XRD pattern of nano-SnO(VO)0.5, and as can be seen, lines due to both SnO2 and Sn metal are present, as can be expected because VO cannot be reduced to V metal by SnO. Thus, the composite consists of SnO2·Sn·VOx (x ≈ 1). In the XRD patterns of Figure 15, lines due to VOx (x ≤ 1) are not seen, possibly due to their amorphous nature as a result of HEB and also due to the low atomic scattering factor of V in comparison to Sn. The average crystallite sizes of the above composites are 12 ± 3 nm as estimated by Scherrer’s equation of the XRD data and HR-TEM studies. When cycled in the voltage range 0.005−0.8 V vs Li at a current of 60 mA g−1 (0.12C) the nano-SnO(V2O3)0.5 showed a first-charge capacity of 435 ± 5 mA·h g−1, which stabilized to 380 ± 5 mA·h g−1 with no noticeable fading in the range of 10−60 cycles (Figure 16c). Under similar cycling conditions, 5384

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 16. Capacity vs cycle number plots: (a) nano-SnO, (b) nano-SnO(V2O3)0.25 at 0.12C and 0.5C (1C = 500 mA g−1), (c) nano-SnO(V2O3)0.5, and (d) nano-SnO(VO)0.5. Voltage range 0.005−0.8 V vs Li at current density of 60 mA g−1 (0.12C). Filled symbols indicate discharge capacity, open symbols charge capacity. Reproduced with permission from ref 467. Copyright 2011 Springer.

composite and observed a first charge capacity of ∼780 mA·h g−1 in the voltage range 0−2.0 V at 40 mA g−1. However, capacity fading was noticed within 10 cycles, giving ∼450 mA·h g−1. Sakaushi et al.470 reported that nanosheet SnO/CNT composite delivers a reversible capacity of ∼250 mA·h g−1 when cycled at a current rate of 200 mA g−1 in the voltage range 0.1−1.0 V, but the value decreases to ∼170 mA·h g−1 after 20 cycles (Table 5). Recent additional Li-cycling studies on SnO are given refs 471−474 (Table 5). 3.1.2. SnO2. Tin dioxide, SnO2, adopts a tetragonal rutile structure (Figure 2a). Li cycling studies and the key factors controlling the reversibility of SnO2 have been studied by many groups. The group of Dahn452 reported a stable capacity of ∼300 mA·h g−1 up to 25 cycles in the cycling range 0.3−1.0 V at current of 37.2 mA g−1. However, with an upper cutoff voltage of 1.3 V, severe capacity fading was noted. Retoux et al.475 studied the Li cycling of SnO2 films up to 500 cycles along with complementary HR-TEM investigations. They found that SnO2 crystallites decompose to 10−50 nm size Sn grains during the first cycle, and these are surrounded by an amorphous 5−10 nm wide ring made of Sn and O. They reported that the size of the Sn crystallites increases from 40 to ∼110 nm from the first cycle to the 500th cycle. In addition, the structure of the amorphous compound made up of Sn and O surrounding the Sn particles changes after 500 cycles, suggesting that a beginning of crystallization has occurred. The authors noted that the possible reason for capacity fading during long-term cycling is due to expansion or the formation of this semicrystalline layer. Recent papers476,477 and a review478 discuss in more detail the in situ TEM studies of SnO2 during Li cycling. The Li cycling studies on micrometer-size SnO2 by Belliard et al.,479 bare and In-doped nanocrystalline SnO2 by Subramanian et al.,480−482 nanowires and nanotubes of SnO2 by Park et al.,483,484 and nanorods of SnO2 by Wang and Lee,485 Liu et al.,486 and Guo et al.487 showed that a reversible and stable capacity of ∼700 mA·h g−1 can be obtained when the system is cycled in the range 0.005−1.0 V up to 50−60 cycles. Larger capacities (∼1000 mA·h g−1) are obtained when the upper cutoff voltage is raised to 1.5 V (or 2.0 or 3.0 V), but this always gives rise to capacity fading on cycling beyond 10 cycles.

nano-SnO, nano-SnO(V2O3)0.25, and nano-SnO(VO)0.5 showed initial reversible capacities between 630 and 390 ± 5 mA·h g−1. Between 10 and 50 cycles, nano-SnO showed a capacity fade as high as 59%, in good agreement with literature reports, whereas the above two VOx-containing composites showed capacity fade ranging from 10% to 28% (Figure 16, Table 5). The Coulombic efficiency increased in all the nanocomposites to 96−98% after 10 cycles. Under the above conditions, both V2O3 and VO were found to be electrochemically inactive indicating that they do not contribute to the reversible capacities. The observed galvanostatic cycling, CV, and ex situ XRD data have been interpreted in terms of the alloying− dealloying reaction of Sn in the nanocomposites (eq 4), and the presence of matrix (VOx) enables better Li cycling behavior by buffering the unit cell volume variations and providing an electronically conducting network for Li diffusion in the nanocomposites. In all the nanocomposites, the average discharge (alloying) potential is 0.2−0.3 V and average charge (dealloying) potential is 0.5−0.6 V vs Li, in good agreement with literature reports. Xie et al.468 studied thin films of SnO prepared by rfmagnetron sputtering on Cu substrates and studied the electrochemical properties by galvanostatic cycling and cyclic voltammetry, up to 3 cycles. The films were 1.7 μm thick and are poorly crystallized as determined by XRD and Raman spectra. The first discharge capacity and charge−discharge efficiency are 1868 mA·h g−1 and 44% for SnO. After the first discharge, the capacities of the films are all close to their theoretical values where a Li4.4Sn composition is expected. It should be noted that during the first Li insertion process, SnO is irreversibly decomposed into Sn and Li2O followed by the formation of LiδSn (0 ≤ δ ≤ 4.4) in the Li2O matrix, as per eq 2. Slow capacity fading was noted upon cycling between 0.01 and 1.5 V vs Li at a current of 15 μA cm−2. The apparent Li ion chemical diffusion coefficient (DLi) was evaluated by galavanostatic intermittent titration technique (GITT) in the potential range 0.15−1.1 V and the values ranged from 1 × 10−15 to 5 × 10−14 cm2 s−1, showing a slightly increasing trend with an increase in the applied potential. There are a few reports on the Li cyclability of composites of SnO with carbon nanotubes (CNT) or with other oxides: Chen et al.469 reported the cycling behavior of SnO−CNT nano5385

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Zhu et al.488 studied the Li cycling of SnO2 nanoparticles with size ranging from 3 to 62 nm and obtained reversible capacities of 300−800 mA·h g−1, in the cycling range 0−1.0 V at current rate of 0.2 mA cm−2. Kim et al.489 reported that the SnO2 of particle size of ∼3 nm prepared by a hydrothermal method showed a reversible capacity of ∼740 mA·h g−1 stable up to 60 cycles when it was cycled at 60 mA g−1 in the voltage range 0−1.2 V. However, under similar cycling conditions, particles with sizes ∼4 and ∼8 nm showed capacity fading. Kwon et al.490 studied the anodic properties of F-doped SnO2 and reported a reversible capacity of 600 mA·h g−1. Studies on the bare and Al-doped SnO2 were reported by Tirado’s group491 and complemented by Al MAS NMR and Sn Mossbauer data. Severe capacity fading was observed when the systems was cycled to an upper cutoff voltage of 2.5 V. A good number of studies on the Li cycling performance of nano-SnO2 in the form of composites are reported: SnO2− CNT (carbon nanotubes) composites492−497 exhibit a reversible capacity ranging from 380 to 720 mA·h g−1, when cycled in the cycling range of 0.005−3.0 V. For example, Wang et al.497 reported a reversible capacity of 540 mA·h g−1, almost stable up to 200 cycles for the core−shell nanotubes of SnO2− CNT when cycled in the range 0.005−3.0 V at ∼0.5C rate. Li cycling studies on CuO nanotube−SnO2 and Cu/SnO 2 nanocomposites showed reversible capacities ranging from 380 to 900 mA·h g−1498,499 and Li2O−CuO−SnO2 composites delivered a high capacity of 950 mA·h g−1.500 Composites of SnO2 with carbon and carbon nanofibers have been prepared by several groups501−511 and studied as anodes. Reversible capacities of 250−630 mA·h g−1 were reported after 50 cycles, when the system was cycled in range 0−3.0 V. Composites of SnO2 with graphene have been studied recently, and these materials showed reversible capacities of 550−600 mA·h g−1 after 50 cycles upon cycling in the range 0.02−3.0 V at various current densities.512−515 SnO2 in the form of thin films on suitable substrates prepared by various methods have been studied for Li cyclability. Reversible capacities of 480−600 mA·h g−1 stable up to 100 cycles were reported in the cycling voltage range 0−1.0 V.516,517 Li et al.518,519 prepared nanostructured SnO2 films by the sol− gel method and showed that they possess high rate capability with a capacity of 460 mA·h g−1 at 60C rate. Nanocrystalline SnO2 thin films, prepared by electrodeposition technique, showed a reversible capacity of 400−780 mA·h g−1, stable up to 15 cycles.520,521 Nam et al.522 found that SnO2 films prepared using electron beam evaporation showed stable capacity of 380 mA·h g−1. Zhu et al.523,524 examined the Li cyclability of Sn− Co−O and Sn−Mn−O thin film composites with reticular structure prepared by electrospray deposition method. When cycled in the range 0.01−3.0 V at a current rate of 0.5C, the Sn−Co−O films showed a reversible capacity of ∼734 mA·h g−1, which gradually increased to 845 mA·h g−1 at the end of 50 cycles. Under similar cycling conditions, the Sn−Mn−O films showed a reversible and stable capacity of 656 mA·h g−1 up to 30 cycles524 (Figure 17). On the other hand, pure SnO2 films showed continuous capacity fading524 (Figure 17). There are a few reports on the anodic behavior of coated SnO2. Kim et al.525 prepared AlPO4-coated SnO2 and found a reversible capacity of 781 mA·h g−1 (near theoretical value) when it was cycled at 105 mA g−1 in the range 0−2.5 V. It showed 44% capacity retention compared with the bare SnO2 (8%) after 15 cycles. Recently, Lee et al.526 studied nano-Sicoated nanotube SnO2 with the composition 40:60 wt % in the

Figure 17. (a) Cycling performance of the SnO2−MnO film between 0.01 and 3.0 V at 0.5C and (b) Coulombic efficiency of the tin− manganese oxide film. Reproduced with permission from ref 524. Copyright 2010 Elsevier.

voltage range 0−1.2 V vs Li up to 90 cycles at various C-rates, 0.2−2C (1C = 2 A g−1). Here, both Sn and Si participate in the Li cycling, as shown by the cyclic voltammetry. At 0.5C rate, the above composite showed an initial reversible capacity of ∼1800 mA·h g−1, which slowly degraded to ∼1600 mA·h g−1 at the end of 90 cycles (∼11% fading). Under similar cycling conditions, bare nanotube SnO2 (that is, without the Si-coating on the inside nanotube walls) exhibited an initial reversible capacity of ∼800 mA·h g−1, which slowly degraded to ∼400 mA·h g−1 at the end of 90 cycles (∼50% fading). Also, the above composite showed good C-rate capability: 93% of the capacity of 1838 mA·h g−1 at 0.2C was retained at 2C. TEM studies of the cycled electrodes, after 90 cycles, showed substantial fragmentation of the bare nanotube SnO2, whereas the Si-coated nanotube SnO2 maintained the overall tubular morphology, with some deterioration of the tube wall structure along the length of the tube. Tin(II) oxyhydroxide, Sn3O2(OH)2, in nano/amorphous form was prepared and studied by Kim et al.527 When it was cycled in the range 0−2.0 V at 100 mA g−1, reversible capacity as high as 1030 mA·h g−1 was observed, which degraded to ∼650 mA·h g−1 after 20 cycles. When the upper cutoff voltage was reduced to 0.8 V, the performance improved. At 50 mA g−1, a first-cycle reversible capacity of ∼600 mA·h g−1 slowly decreased to ∼500 mA·h g−1 after 100 cycles. Additional recent works on the Li cycling of SnO2 and its composites can be found in the following: Hierarchical SnO2 nanostructures via self-assembly by Yin et al.;528 Sn/SnO2 composite oxides by Tirado group;529 SnO2 −NiO−C nanocomposites by Hassan et al.;530 polycrystalline electrospun SnO2 nanofibers by Yang et al.;531 electrospun carbon−SnO2 composite nanofibers by Bonino et al.,532 Kong et al.,533 and Mu et al.;534 carbon fiber supported SnO2 by Mi et al.;535 5386

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Table 6. Physical Properties and Electrochemical Li Cycling Data of SnO2 sample and morphology SnO2 commercial SnO2 prepared using tin acetate at 420 °C in air SnO2 prepared using tin acetate at 390 °C, in argon

SnO2 thin films

commercial sample nanoparticles In-doped SnO2 (Sn0.9In0.1)O2 nanoparticles nanotubes nanowires nanoparticles nanowires 1D nanowires coated with CNT nanoparticles nanosheets hollow spheres nanoparticles nanoparticles (prepared by molten salt method (MSM), 280 °C) MSM nanoparticles reheated to 500 °C MSM nanoparticles reheated to 700 °C nanoparticles 67 wt % SnO2 nanoparticles/33 wt % carbon SnO2 nanoparticles SnO2/graphene nanosheets SnO2 nanorods 44 wt % SnO2 nanorods/56 wt % graphene

sandwiched SnO2−graphene nanosheets (GNS)−SnO2 SnO2−PPy composites

particle/crystallite size

surface area, m2 g−1

current rate, mA g−1 37 mA g−1

>40 nm 5.8 nm

reversible capacity (mA·h g−1); voltage range

capacity retention after n cycles (cycling range)

510; (0.2−1.3 V) 520;(0.2−1.3 V)

20% (n = 2−50) 51% (n = 2−50)

4.5 nm

310; (0.3−1.0 V)

98% (n = 2−38)

crystallite size, 30−50 nm

525; (0.2−1.3 V) 300; (0.3−1.0 V) 460; (0−0.8 V) 500

60% 98% 98% 98%

400; (0.05−1.15 V) 660 665; (0.005−2.0 V) 603 740; (0−1.2 V) 636 1384; (0.05−1.5 V) 1134 1277 750; (0.1−3.0 V) 418; (0.01−1.5 V) 511 (bare) 887; (0.1−2.0 V) 936; (0.1−2.0 V) 930 762 632; (0.005−1.0 V)

60% (n = 1−8) 52% (n = 2−10) 95% (n = 2−10) 99% (n = 2−60) 73% (n = 2−60) 22% (n = 1−50) 26% (n = 1−50) 8% (n = 1−50) 68% (n = 2−10) 78% (n = 2−30) 15% (n = 2−20) 57% (n = 1−50) 57% (n = 1−50) 38% (n = 1−50) 23% (n = 1−50) 49% (n = 2−50)

0.1 mA cm−2 0.3 mA cm−2 0.2 mA cm−2 0.5 mA cm−2

micrometer size 50 nm 5 nm 3 nm 4 nm

30 nm 50 nm nanowires, 60 nm diam 17 nm 100 nm diam, 5 nm thick

60 231 146 97

100

0.2C 0.2C

= = = =

2−50) 2−25) 2−20) 5−100)

516, 604 464 481 489 484

605 606

7 nm (XRD)

55.5 180 101.8 199.7 129

12 nm (XRD)

39

628

77% (n = 2−50)

20 nm (XRD), ∼33 nm (TEM) 14 nm (XRD) 8.4 nm (XRD)

15

602

47% (n = 2−30)

500 500

450; (0.005−1.2 V) 400; (0.005−1.2 V)

33% (n = 2−50) 97% (n = 2−50)

594 594

100 100 100 100

1150; (0.005−3.0 V) 1000 1250; (0.005−3.0 V) 907

12% 75% 19% 78%

545

10−20 nm diam 100−200 nm length 30−40 nm

400

968; (0.01−3.0 V)

52% (n = 2−150)

571

nanowires, 70−120 nm diam

235

690; (0.005−3.0 V)

90% (n = 2−80)

609

4−6 nm 3−5 nm 10−20 nm diam 100−200 nm length

124 283

78 156

(n (n (n (n

ref 452

100 (0.12C; 1C = 780 mA g−1)

review of SnO2 hollow structures by Chen et al.;185 macroporous SnO2−carbon composites by Yim et al.;536 SnO2−carbon composites by Wang et al.537 and Li et al.;538 porous microspheres of SnO2 nanoparticles by Wang et al.;539 SnO2@carbon core−shell nanochains by Yu et al.;540 SnO2− Fe2O3 composites by Zhou et al.;541 SnO2−Fe2O3/rGO (reduced graphene oxide) composites by Zhu et al.;542 SnO2/ rGO composites by Zhu et al.;543 SnO2/graphene composites by Li et al.,544 Zhao et al.,545 Zhong et al.,546 Lian et al.,547 Huang et al.,548 Xie et al.,549 Baek et al.,550 Wang et al.,551 and Xu et al.;552 macroporous SnO2 with and without carbon by Wang et al.;553 SnO2/mesoporous carbon spheres composites by Di Lupo et al.;554 SnO2 deposited on MWCNT (multiwalled carbon nanotubes) by Ren et al.,555 Shao et al.,556 and Jin et al.;557 monodispersed SnO2@carbon composite hollow spheres

(n (n (n (n

= = = =

2−50) 2−50) 2−50) 2−50)

607 608

567

582

by Chen et al.;558 nanocarbon−SnO2 composite by Zhang et al.,559 SnO2 nanotubes by Wang et al.;560 SnO2−CMK-3 nanocomposites by Qiao et al.;561 antimony-doped SnO2 by Wang et al.,562 Wu et al.,563 and Volosin et al.;564 Ti-doped SnO2 by Issac et al.;565 porous SnO2/layered titanate nanohybrids by Kang et al.566 Very recently, Reddy et al.567 prepared SnO2 nanoparticles by a molten salt method using 0.88LiNO3:0.12LiCl eutectic salt at 280 °C and studied the effect of heat treatement on particle size and its electrochemical properties. They reported reversible capacities in the range 602−632 mA·h g−1, cycled in the voltage range 0.005−1.0 V and at a current rate of 100 mA g−1 (Table 6). Heat-treated sample at 500 °C retained a capacity of 480 mA·h g−1 at the end of the 50th cycle, and they proposed an electrochemical alloying/dealloying reaction mechanism (voltage range 0.005− 5387

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 18. SEM images for (a) GNS, (b) GNS/SnO2 composite (low magnification), and (c) GNS/SnO2 (high magnification). TEM images for (d) SnO2 nanoparticles, (e) GNS/SnO2(1.5), and (f) GNS/SnO2(0.34). HRTEM images for GNS/SnO2(0.34) (g) low magnification and (h) high magnification. Reproduced with permission from ref 545. Copyright 2011 Elsevier.

capacity−voltage profiles, up to 5 cycles, of GNS, SnO2, and GNS/SnO2(0.34), respectively. The capacity−cycle number plots, up to 50 cycles, are shown Figure 19d. It is clear that in all cases, there is a large irreversible capacity loss (ICL) between the first-discharge and first-charge capacity. For SnO2 and the composites, GNS/SnO2, the ICL can be understood as due to the reduction of SnO2 to Sn (eq 3) and also due to the formation of the SEI. In the case of GNS, SEI formation, and possibly the reduction of oxygen species on the surfaces of GNS might be contributing to the ICL. These aspects were not discussed by the authors. From Figure 19a, we note that there is no specific Li intercalation−deintercalation signatures into GNS at fixed potentials, in contrast to that shown by graphite at ∼0.15 to ∼0.25 V vs Li as reported in the literature.3 This shows that mostly interfacial Li storage occurs in GNS in the form of separated Li ions and electronic charges. Indeed, the authors state that bilayer adsorption theory predicts a theoretical capacity of 744 mA h g−1 for GNS. From Figure 19a,d, it can be seen that GNS shows a second cycle reversible capacity of ∼1000 mA·h g−1, which decreases to ∼650 mA·h g−1 after 15 cycles and stabilizes at that value up to 35 cycles, followed by a small degradation up to 50 cycles. Figure 19b

1.0 V) that was similar to previous original work by the group of Dahn452 and Retoux et al.475 Results from the recent studies on the Li cycling of nanostructured SnO2, carbon, CNT, and graphene composites568−609 and capacity values of a few slected references are summarized in Table 6. The work of Zhao et al.545 will be described in detail since it illustrates some interesting features. The authors prepared graphene nanosheet (GNS)/SnO2 composites with two different mass ratios of 1.5/1 and 0.34/1. These are called GNS/SnO2(1.5) and GNS/SnO2(0.34), respectively. For comparison, pure nano-SnO2 and GNS were also prepared for the study. Figure 18 shows the SEM and HR-TEM pictures of the phases. Nanocrystalline particles of SnO2 of grain size 4− 6 nm were found homogeneously distributed on the matrix of GNS (Figure 18f,g,h). Nitrogen adsorption tests showed an “ink-bottle-like” pore structure of GNS/SnO2 and the SnO2 nanoparticles were confined in the interlayer of GNS without agglomeration. BET surface areas of the GNS, pure SnO2, and GNS/SnO2(0.34) were found to be 430, 124, and 283 m2 g−1, respectively. Galvanostatic cycling tests were carried out at 100 mA g−1 in the voltage range 0.005−3 V vs Li. Figure 19a−c shows the 5388

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 19. Lithium insertion/extraction curves of (a) GNS, (b) SnO2, and (c) GNS/SnO2(0.34) and (d) cycle performances of GNS, SnO2, GNS/ SnO2(1.5), and GNS/SnO2(0.34). Reproduced with permission from ref 545. Copyright 2011 Elsevier.

From the above discussion, it is clear that SnO2 has attracted much more attention as an anode in comparison to SnO, because it is the stable tin oxide. Bulk as well as nano-SnO2 in the form of thin films, nanowires, nanorods, nanospheres, and other nanostructures and its composites with carbon (or graphite), CNT, graphene, etc. can yield near-theoretical reversible capacities (782 mA·h g−1) for Li alloying−dealloying reaction of eq 4 only when it is cycled in the range 0.005 (or 0.1) to below 1 V vs Li. Cycling to higher voltages, for example, to 1.5, 2, or 3 V always leads to the formation and decomposition of SnOx (x ≤ 2) (eq 5), and the changes in the unit cell volumes associated with eqs 4 and 5 are too large for the nanostructured SnO2 and its composite to accommodate and always lead to capacity fading. Thus, reversible capacities greater than 800 mA·h g−1 are not sustained under these conditions upon long-term cycling (50−100 cycles) at all current rates, 0.1−5C with upper cutoff voltages ≥1.5 V (Table 6). These aspects are not considered by many authors in the recent literature, who simply claim large reversible capacities in the initial few cycles, stabilizing at capacity values ≤800 mA·h g−1 during long-term cycling with upper cutoff voltages ≥1.5 V and at C rates below 1C. Many claims of larger and stable capacities in the nanostructured SnO2 have not been replicated. It must also be mentioned, however, that even though these nanostructured SnO2 composites are prospective anodes for LIBs, the challenges of scaling-up the preparation of the above materials to multigram quantities with reproducible Li cycling properties are nontrivial tasks.

shows the typical discharge−charge profiles of pure SnO2 involving Li alloying−dealloying reactions of Sn occurring below ∼0.7 V (eq 4). In addition, the profiles also show small voltage plateaus upon charging and discharging at ∼1.0−1.2 V, indicating partial oxidation of Sn to SnOx (eq 5). These characteristics are also seen in the discharge−charge profiles of the composite GNS/SnO2(0.34) shown in Figure 19c. From Figure 19b,d, it is clear that pure SnO2 shows a drastic and continuous capacity fading, from ∼1180 mA·h g−1 at the second cycle to 136 mA·h g−1 after 50 cycles (88% capacity fading). The composite with higher graphene content, namely, GNS/SnO2(1.5) shows a second cycle capacity of ∼720 mA·h g−1, which stabilizes at ∼650 mA·h g−1 in the range 15−50 cycles (Figure 19d). Interestingly, the capacity vs cycle number plots overlap very well for both GNS and GNS/SnO2(1.5) in the range 15−50 cycles (Figure 19d). On the other hand, the composite with higher tin oxide content, namely, GNS/SnO2(0.34) shows a second cycle discharge capacity of ∼1030 mA·h g−1, which degrades to ∼800 mA·h g−1 at the end of 35 cycles (22% capacity fading) and stabilizes at ∼775 mA·h g−1 in the range 40−50 cycles (Figure 19c,d). Even though the authors, Zhao et al.,545 claim that the latter value is comparable to the calculated capacity value as a sum of the contributions from the Li cycling of GNS and SnO2, the fact remains that the observed high capacity values in the range 2−35 cycles for this composite, as well as those shown by GNS/SnO2(1.5) and pure SnO2, can only be explained as due to the additional capacity contribution from the conversion reactions of Sn with Li2O (eq 5). 5389

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

600 °C for 6 and 24 h, respectively (Figure 20). The hydroxide precursor was prepared using coprecipitation from salt

3.2. Ternary Tin Oxides

As mentioned in section 3, incorporation of one or more electrochemically inactive “matrix” elements or oxides will enable better Li cyclability of alloy-forming metal oxides. These matrices can buffer to some extent the unit cell volume changes that occur during alloying−dealloying reactions thereby minimizing the capacity fading. In favorable cases, these matrices do improve the electronic conductivity and also act as catalysts for better Li cycling. The ideal way to incorporate these matrices is to start with ternary or complex oxides, rather than mechanical mixing of inactive and active oxides. With this in view, a large number of tin oxides of different crystal structures have been studied over the years. By nanotechnology methods, nanosize mixed oxides also have been studied. One obvious disadvantage of the mixed tin oxides is that the theoretical reversible capacity will be less than that of the binary tin oxides. The Li cycling mechanism for ternary/quaternary oxides, MxSnOy, during the first discharge reaction with Li metal always involves the destruction of the crystal structure or amorphization of the lattice of the virgin mixed oxide, similar to the binary tin oxides. This results in the formation of metal (M) or metal oxides (MOy) and formation of nano-Sn-metal dispersed in amorphous Li2O. This is followed by alloy formation, Li4.4Sn. Reduction of the counter metal ions to metals depends upon the magnitude of free energy of formation of respective oxide in comparison to Li2O. In cases where M is alkali, alkaline earth, or rare earth metal, studies have shown that only metal oxides (MO or MOy) are formed during the first discharge. If M = Fe, Co, Ni, Cu, Zn, or Cd, the respective nanoparticles of M are formed along with Li2O. 3.2.1. M2SnO4 (M = Metal). The compounds M2SnO4 (M = Mg, Mn, Co, or Zn) adopt a spinel structure with Sn in tetrahedral O-coordination. The group of Irvine479,610 reported preliminary Li cycling behavior of ternary tin oxides, M2SnO4 (M = Mg, Mn, Co, Zn), in the range 0.02−1.5 V vs Li at low current rates. Best reversibility is noted for M = Mn and Zn and the worst when M = Mg. They found a correlation between the initial reduction potential of the spinel oxides and enthalpy of formation of the metal oxide, MO. A few studies of the above compounds when cycled in the range 0−3.0 V are discussed in section 4.1. Chowdari’s group611 studied the Li cycling of Ca2SnO4, which adopts the Sr2PbO4 structure. The orthorhombic unit cell is comprised of infinite chains of edge-shared SnO6 octahedra that are connected by a perpendicular network of CaO7 monocapped trigonal prisms. A reversible capacity of ∼220 mA·h g−1 stable up to 40 cycles in cycling range 0.005− 1.0 V at 60 mA g−1 current rate was observed. This is 50% smaller than the theoretical value. 3.2.2. ASnO3 (A = Ca, Sr, Ba, Co, and Mg). Chowdari’s group611,612 reported the Li cycling studies of ASnO3 (A = Ca, Sr, or Ba), which contains SnO6 octahedra and adopts a perovskite structure. They have shown excellent cycling performance of CaSnO3 with a reversible capacity of 380 mA·h g−1 stable up to 100 cycles in the voltage range 0.005−1.0 V at 60 mA g−1. This capacity value corresponds to 3.0 mol of Li per mole of CaSnO3 compared with the theoretically obtainable value of 4.4 mol of Li via alloying−dealloying of Sn. Under similar cycling conditions, SrSnO3 and BaSnO3 gave smaller but stable capacities, ranging from 150 to 220 mA·h g−1. Recently, Chowdari’s group613 prepared X-ray amorphous nanocomposite CaO−SnO2 and X-ray crystalline nanophase CaSnO3 by the thermal decomposition of CaSn(OH)6 in air at

Figure 20. X-ray diffraction patterns of as-prepared cubic-CaSn(OH)6 and of the products calcined in air at 400, 500, and 600 °C for 6 h at each temperature and at 600 °C for 24 h. Formation of X-ray crystalline CaSnO3 is seen in the latter case. Miller indices (hkl) are shown. Reproduced with permission from ref 613. Copyright 2008 American Chemical Society.

solutions at ambient temperature. X-ray diffraction, HR-TEM, and SAED analysis revealed that the X-ray amorphous nanoCaO−SnO2 consists of nanosize (3−6 nm) grains of CaO and SnO2 whereas nano-CaSnO3 consists of ∼60 nm size crystallites. Galvanostatic cycling studies at room temperature at the current 60 mA g−1 (0.12C) were performed on both nanoCaO−SnO2 and nano-CaSnO3 in two voltage windows, 0.005− 1.0 V and 0.005−1.3 V vs Li. The cycling performance of nanoCaO−SnO2 is better than that of the nano-CaSnO3 and showed almost theoretical capacity, 550 ± 5 mA h g−1 (4.2 mol of cyclable Li vs the theoretical 4.4 mol of Li) stable up to 50 cycles when cycled between 0.005 and 1.3 V. In the range of 0.005−1.0 V, nano-CaO−SnO2 and nano-CaSnO3 showed reversible capacities of 490 ± 5 mA h g−1 (∼3.8 mol of Li) and 445 ± 5 mA h g−1 (∼3.4 mol of Li), respectively, stable up to 50 cycles with a Coulombic efficiency of 96−98% (Figure 21). At 0.4C, nano-CaO−SnO2 showed a capacity of 420 ± 5 mA h g−1 (∼3.2 mol of Li). Nano-CaSnO3 gave an initial reversible capacity of 570 ± 5 mA h g−1 (4.4 mol of Li, theoretical value) when cycled in the voltage window 0.005−1.3 V at 60 mA g−1. However, the capacity degraded on cycling, relatively slowly up to 40 cycles and fairly rapidly in the range 40−70 cycles, and retained only 380 ± 5 mA h g−1 at the end of the 70th cycle (Figure 21). The better cycling performance of nano-CaO−SnO2 has been attributed to the presence of smaller nanosize regions (3− 6 nm) of CaO and SnO2 in it and thus a better buffering effect of CaO, as an electrochemically inactive matrix, along with Li2O. The observed capacity fading beyond 50 cycles shown by nano-CaO−SnO2 with an upper cutoff voltage of 1.3 V may be attributed to the SnOx (x ≤ 1) formation (eq 5). A similar explanation holds for nano-CaSnO3 when cycled with the upper cutoff voltage 1.3 V. Average discharge and charge potentials for both nano-CaO−SnO2 and nano-CaSnO3, observed by the cyclic voltammetry (CV) in the voltage 5390

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

First-discharge, Composite

Figure 21. Capacity vs cycle number plots at the current of 60 mA g−1 (0.12C) (a) for nano-CaO−SnO2 and (b) for nano-CaSnO3. Voltage windows are indicated. (c) The capacity vs cycle number plot at various current rates of nano-CaO−SnO2 in the voltage window 0.005−1.0 V. The filled and open symbols represent discharge and charge capacities, respectively. The C values corresponding to different current rates are shown, assuming 1C = 500 mA g−1. Reproduced with permission from ref 613. Copyright 2008 American Chemical Society.

window 0.005−1.0 V (Figure 22) and complemented by galvanostatic cycling, are ∼0.2 and ∼0.5 V, respectively.

Figure 23. Ex situ TEM of nano-CaO−SnO2 charged to 1.0 V after 30 cycles. (a) Lattice image. The nanocrystalline regions marked 1 and 2 correspond to metallic tin. The other regions (circled) show the overlapping of various planes of crystalline metallic tin. (b) The SAED pattern. The Miller indices corresponding to the diffuse spots are assigned to metallic tin. Scale bar is 5 nm. Reproduced with permission from ref 613. Copyright 2008 American Chemical Society.

CaOSnO2 + 4Li+ + 4e− → CaO + 2Li 2O + Sn

(6)

First-discharge, Compound CaSnO3 + 4Li+ + 4e− → CaO + 2Li 2O + Sn

(7)

Impedance spectra were measured on nano-CaO−SnO2 vs Li, during the first discharge−charge cycle and also during the 11th discharge cycle, and the relevant impedance parameters along with the “apparent” chemical diffusion coefficients (DLi+) have been evaluated and interpreted to complement the CV and galvanostatic results. Bulk impedance of the electrode (Rb) dominates at low voltages (0.5 V) during Li cycling of nanoCaO−SnO2. The observed value of DLi+ was ∼1.0 × 10−14 cm2 s−1 at voltages ≤1 V vs Li during the first cycle and the 11th discharge cycle. Zhao et al.614 prepared nanoflower CaSnO3 by a hydrothermal method, and they showed that the reversible capacities range from 250 to 545 mA·h g−1 at a current rate of 60 mA g−1, when the system is cycled in the voltage range 0.005−1.0 V. Recently, Mouyane et al.615 prepared CaSnO3 and CaSnSiO5 by solid state reaction and studied the Li cycling mechanism by galavanostatic tests, XRD, and Mossbauer spectra. They

Figure 22. Cyclic voltammograms of nano-CaO−SnO2 in the voltage window (a) 0.005−1.0 V and (b) 0.005−1.3 V vs Li at the slow scan rate of 58 μV s−1. Numbers indicate the cycle number. Reproduced with permission from ref 613. Copyright 2008 American Chemical Society.

Ex situ XRD on nano-CaSnO3 during the first cycle and ex situ TEM and SAED on nano-CaO−SnO2 after 30 cycles were carried out, and the results were interpreted to support the proposed reaction mechanism involving Li alloying−dealloying reaction of Sn (eq 4) after the initial formation of Sn metal and Li2O as per eqs 6 and 7 (Figure 23). 5391

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

showed that during the first discharge, progressive transformation of the pristine materials leads to nanocomposites formed by particles of Li2O and Sn(0) species (and SiO2 in the case of CaSnSiO5) as per eq 7. Mossbauer spectral data in the discharged-state suggested the existence of Ca−Sn bonds, in addition to the usual Li−Sn and Sn−Sn interactions found in LixSn (x ≤ 4.4) alloys, thereby showing the importance of Ca as the matrix element. Very recently, Hu et al.616 reported Li cycling studies on core−shell carbon-coated CaSnO3 nanotubes. Cadmium stannate, CdSnO3, adopts a rhombohedralhexagonal structure, which is related to the perovskite structure. The group of Chowdari617 prepared and studied CdSnO3 (30 nm) and reported a reversible and stable capacity of 475 ± 5 mA·h g−1 (2−40 cycles) in the voltage range 0.005−1.0 V at 0.13C rate. This corresponds to ∼5 mol of Li per mole, since Cd metal also can alloy with Li in addition to Sn. But, at a higher cutoff voltage (1.3 V), the capacity degraded from 580 (1 cycle) to 245 ± 5 mA·h g−1 (25 cycles). There are reports on the Li cycling of ASnO3, A = Co,618 Mn,619 and Mg.620 However, no such well-defined compounds exist, and they may be composed of composites, AO·SnO2 or SnO2 + A2SnO4. High reversible capacities (800−1200 mA·h g−1) were obtained when these systems were cycled in the range 0−3.0 V for A = Co618 and in the range 0−2.0 V for A = Mn619 and Mg.620 However, capacity fading was noted in all cases. 3.2.3. Li2SnO3. Lithium stannate, Li2SnO3, contains SnO6 octahedra and adopts a monoclinically distorted layer structure, similar to that of Li2MnO3. Its formula can be written as Li(Li1/3Sn2/3)O2, and it contains LiO6 octahedra in both the Li layers and the (Li−Sn) layers. Preliminary Li cycling studies were reported by Courtney and Dahn453 and Belliard and Irvine.621 The latter authors prepared both low-temperature and high-temperature forms of Li2SnO3 by adjusting the synthesis temperature. They observed a first cycle reversible capacity of 685 and 596 mA·h g−1 for the Li2SnO3 prepared at 650 and 1000 °C, respectively, when cycled in the range 0−2.0 V at a current of 0.05 mA. However, severe capacity fading was noted after 20 cycles. Zhang et al.622 prepared 200−300 nm sized particles of Li2SiO3 using a sol−gel method and reported a first-cycle reversible capacity of ∼410 mA·h g−1, which gradually decreased to ∼380 mA·h g−1 at the end of the 50th cycle, when it was cycled in the range 0−1.0 V at 60 mA g−1. On the other hand, the performance of Li2SnO3 prepared by the solid-state method (micrometer-size) was slightly inferior. 3.2.4. A2Sn2O7 (A = Y or Nd). Li cyclability of tin oxides with the pyrochlore structure, A2Sn2O7 (A = Y or Nd) composed of SnO6 octahedra were carried out by Chowdari’s group.623 They reported first-discharge capacities of 913 and 722 mA·h g−1 and first-charge capacities of 350 and 265 mA·h g−1 for A = Y and Nd, respectively. Crystal structure destruction occurs during the first discharge leading to the formation of nanoparticles of Sn metal and finally Li4.4Sn alloy in a matrix of Li2O + A2O3. After 50 cycles, both compounds showed a capacity retention of 89% of the 10th cycle reversible capacity. The average charge and discharge voltages in both compounds are 0.5 and 0.25 V, respectively. 3.2.5. K2(M,Sn)8O16 (M = Li, Mg, Fe, Mn, Co, or In). The Sn-based hollandite-type compounds containing SnO6 octahedra with chemical formula A2(M,Sn)8O16, where A = large size alkali or alkaline earth cation like K or Ba and M = bivalent or trivalent metal ion, are of interest to study as anode materials. Chowdari’s group.624 studied the Li cyclability of K2(M,Sn)8O16

(M = Li, Mg, Fe, and Mn) and observed first-cycle charge capacities of 602, 505, 481, and 418 ± 3 mA·h g−1 for M = Li, Mg, Fe, and Mn, respectively, when the system was cycled in the voltage range of 0.005−1.0 V vs Li at the current of 60 mA g−1. The phases with M = Li showed 73% and M = Fe 83% capacity retention after 50 cycles. They mentioned that the M ion plays a role on the Li cyclability with M = Li and Fe acting as “good” matrices, whereas M = Mg and Mn are “bad” matrices and yielded drastic capacity fading. The group of Chowdari625,626 extended the studies on Sn hollandites with M = Co and In in the form of micrometer-size and nanosize particles. The cycling studies of the nanophase (≤10 nm) carried out in the voltage range 0.005−0.8 V showed a stable and reversible capacities of 500 ± 5 mA·h g−1 up to 60 cycles for M = Co. However, capacity fading of 4% was noticed in the range 60−80 cycles (Figure 24a). The phase also

Figure 24. Capacity vs cycle number plots: (a, b) nano-(K−Co) in different voltage ranges; (c) at various current (C) rates of nano-(K− Co); (d) nano-(K−In). Voltage range 0.005−0.8 V vs Li at current of 60 mA g−1 (0.12C) assuming 1C = 500 mA g−1. Filled symbols, discharge capacity; open symbols, charge capacity. Reproduced with permission from ref 626. Copyright 2011 The Royal Society of Chemistry.

exhibited good C-rate capability as is clear from Figure 24c. At 1C rate, the nano-(K,Co) showed a capacity of 410 ± 5 mA·h g−1 stable up to at least 100 cycles. When cycled in the range 0.005−1.0 V, a stable capacity of 525 ± 5 mA·h g−1 was noted up to 40 cycles followed by small capacity fading. The phase with M = In (K,In) showed an initial capacity of 570 mA·h g−1, which degraded to 485 mA·h g−1 at the end of 60 cycles (15% loss), when it was cycled in the range 0.005−0.8 V at 0.12C (Figure 24d). However, the heat-treated electrode (300 °C, 12 h, Ar) showed a significant improvement and gave a stable capacity of 570 ± 5 mA·h g−1 in the range of 5−50 cycles. This improvement is ascribed to a better binder distribution, caused by melting and spreading of the binder, followed by adhesion to the active material particles and to the current collector of the electrode. The micrometer-sized particles of the above two compounds showed small but continuous capacity fading when cycled in both voltage ranges. Thus, the results showed that M = Co and In are also good matrices for Li cycling via alloying− dealloying of Sn, with In also contributing to the Li cycling via alloying−dealloying (3Li + In ↔Li3In). The “apparent” chemical diffusion coefficients (DLi+) estimated from the 5392

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

cycling is carried out in the range 0.005−0.8 V or 1.0 V vs Li. These are stable for a fairly good number of cycles, at least 25 cycles, in most of the ATCO compositions. However, when the upper cutoff voltage is ≥1.2 V, slow capacity fading occurs due to tin nanoparticle aggregation to form large grains during longterm cycling; with a 2 or 3 V cutoff voltage, drastic capacity fading will occur within 10−20 cycles at 0.2−0.5C rates. This is related to the formation and decomposition of SnOx during cycling. (4) It is worth continuing studies on ATCOs preferably incorporating additional “good” matrix elements, like Co and Ca. Also, it is worthwhile to extend studies to other Li alloy forming elements, especially those containing silicon or germanium.

impedance spectra of nano-(K,Co) was found to be (2.0−2.6) × 10−14 cm2 s−1 between 0.25 and 0.45 V vs Li during the first cycle. 3.2.6. SnP2O7, LiSn2P3O12, Sn2P2O7, and Sn3P2O8. The group of Irvine627−629 reported Li cycling behavior of tincontaining phosphates possessing different crystal structures, namely, SnP2O7, LiSn2P3O12, Sn2P2O7, and Sn3P2O8 in the range 0.02−1.2 V at 0.2 mA cm−2. The cubic SnP2O7 phase containing isolated SnO6 octahedra (and corner-linked PO4 tetrahedra) showed higher reversible capacity and 96% capacity retention at the end of 50 cycles, compared with the SnP2O7 with the layered structure, which showed only 53% capacity retention at the end of 50 cycles. This shows the importance of the starting crystal structure of the compound on the Li cyclability, even though amorphization and crystal structure destruction occurs during the initial discharge cycle. All of the other Sn−P−O compounds showed more severe capacity fading. Xiao et al.630 studied the Li cycling performance of crystalline and amorphous Sn2P2O7 and obtained a reversible capacity of 461 and 487 mA·h g−1 at the end of the fifth cycle, when it was cycled in the range 0−1.4 V at a current rate of 20 mA g−1. Li and Li631 prepared and studied carbon-coated Sn2P2O7 and reported a reversible capacity of 600 mA·h g−1, which slowly degraded to 392 mA·h g−1 after 20 cycles, when it was cycled in the voltage range 0−3.0 V at a current density of 100 mA g−1. Ren et al.632 reported Li cycling studies on SnP2O7 thin films. 3.2.7. Amorphous Tin Composite Oxides (ATCOs). In 1997, Idota et al.633 reported that amorphous tin-based composite oxides (ATCO) can be viable anodes for LIBs. These contain SnO as active material and glass-forming oxides of boron and phosphorus, which form a network. The ATCO of the composition Sn1.0B0.56P0.40Al0.42O3.6 showed a reversible capacity of 650 mA·h g−1 in the cycling range 0−1.2 V at ∼50 mA g−1. The capacity was found to be stable over many cycles. They also fabricated and tested LIBs with LiCoO2 cathode and ATCO as anode. The above report led to detailed investigations of the Sn−O glasses of a variety of compositions containing B, P, Si, Ge, etc. with Sn in 2+ and 4+ oxidation states by the research groups of Dahn,454,634 Nazar,635 Minami,636 Yamaki,471 Kumta,637 Lee,630,638 Biensan,639,640 Mansour et al.,641 Gejke et al.,642,643 Conte et al.,644,645 and Mouyane et al.646 These ATCOs were prepared by various methods, characterized, and studied for their Li cyclability. Care must be taken to ensure that Sn2+ ion is stabilized in the ATCO. In addition, complementary microscopic techniques were employed, such as, Li NMR, Sn Mossbauer, and X-ray absorption spectra and neutron diffraction studies to obtain information regarding the nature of the reactions occurring and the role played by the Li2O, B, P, and other ions. The following conclusions can be drawn on the Li cyclability of ATCO: (1) The Li insertion during the first-discharge reaction modifies the Sn environment of the ATCO, the Li acting as glassy-network modifier by inducing the formation of nonbridging oxygen atoms. (2) The Li cycling mechanism involves alloying−dealloying reaction of Sn, but the structural nature of the dispersed alloy formed after the first discharge and that of the Sn metal formed after the first charge (in the glassy matrix) differ from those of the corresponding crystalline phases observed when pure SnO or SnO2 or Sn are used as the active materials. Thus, the vitreous (glassy) nature of the matrix helps in dispersing the Sn and alloy particles. (3) Reversible capacities ranging from 400 to 550 mA·h g−1 are obtained when

3.3. Antimony Oxides and Mixed Oxides

Antimony has received attention, similar to Sn, due to its high specific capacity via alloying−dealloying reaction (Li3Sb; theoretical capacity is 660 mA·h g−1). However, the operating voltage is ∼1.0 V vs Li, almost twice that of Sn (∼0.5 V), and there is the inherent problem of large volume variation (e.g., 137% per Sb atom), which are the disadvantages. Nevertheless, binary and a few ternary antimony oxides have been studied for their Li cyclability. Binary oxides are Sb2O3, Sb2O4 (mixed valent oxide), and Sb2O5 of which Sb2O3 is most stable. During the first discharge cycle, Sb2O3 undergoes structure destruction and formation of the elemental Sb and Li2O. This is followed by alloy formation (eqs 8 and 9), and the reversible capacity comes from eq 9. Sb2 O3 + 6Li+ + 6e− → 3Li 2O + 2Sb

(8)

Sb + 3Li+ + 3e− ↔ Li3Sb

(9) 464

3.3.1. Sb2O3. The group of Chen studied the Li cycling of Sb2O3 in the range 0−2.0 V at a current of 0.2 mA cm−2, and they found a first charge capacity of 500 mA·h g−1, which slowly degraded to ∼300 mA·h g−1 after 10 cycles. Tarascon’s group647 reported preliminary Li cycling properties of Sb2O3, Sb2O4, and Sb2O5 and found that the latter compound is electrochemically inactive. A few studies on thin films and composites of Sb2O3 have been reported. Xue and Fu648 prepared thin films of Sb2O3 by pulsed laser deposition and observed a first-cycle reversible capacity of ∼790 mA·h g−1, which slowly decreased and stabilized after 15 cycles to give capacity of ∼690 mA·h g−1 at the end of 70 cycles in the cycling range 0−3.0 V at a current of 273 mA g−1. The alloying and dealloying voltages of Sb were found to be 0.8 and 1.05 V, respectively. Complementary XRD, XPS, and HR-TEM were carried out to support the reaction mechanism (eqs 8 and 9). Bryngelsson et al.142,649 prepared Sb/Sb2O3 thin films on Nisubstrates by electrodeposition technique, and these exhibited a capacity of 660 mA·h g−1 stable for more than 50 cycles when they were cycled at a current rate of 0.1C in the range 0−1.5 V. They observed that thin films of Sb2O3-coating on the nanoparticle Sb in the composite helps in the cycling stability. Simonin et al. 650 prepared 10−20 nm sized Sb/Sb2 O 3 nanocomposites by spark discharge generation method, reported preliminary Li cycling behavior in the range 0.5−1.4 V, and observed a reversible capacity of 530 mA·h g−1. Li cycling studies on the compound LiSbO3 with the orthorhombic structure showed a reversible capacity of 580 mA·h g−1 when it was cycled in the voltage range 0.2−1.2 V.651 3.3.2. MSb2O6, M = Co, Ni, and Cu. The group of Tarascon647 studied the Li cycling of the compounds MSb2O6, 5393

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

voltammetry (CV). N-(V1/2Sb1/2Sn)O4 showed stable cycling performance at 24 °C (ambient temperature) with capacities 450, 435, and 360 ± 5 mA·h g−1 at 0.43C, 1.0C, and 3.5C rates, respectively, in the range 0.005−1.0 V vs Li (1C = 580 mA g−1) (Figures 25 and 26).

M = Co, Ni, and Cu, which possess the trirutile structure. They showed that for M = Co and Ni, structure destruction occurs during the first discharge with the formation of Co or Ni metal and Sb particles at ∼1.5 V vs Li (eq 10). (Co/Ni)Sb5 +2 O6 + 12Li → Co/Ni(0) + 2Sb(0) + 6Li 2O (10)

This is followed by the alloying reaction involving Sb, which occurs at potentials of ∼1.0−1.2 V (eq 9). The compounds with M = Co and Ni showed first-cycle reversible capacities of 400−450 mA·h g−1 (6−7 mol of Li per mole of MSb2O6) when cycled in the range 0−3.0 V at 0.05C. However, the cyclability was poor, and CoSb2O6 showed only ∼200 mA·h g−1 after six cycles (0−2.0 V; 0.05C). Interestingly, under similar cycling conditions, CuSb2O6 did not show any measurable first-charge capacity. Reasons for this surprising result were discussed based on the in situ XRD, TEM, and Mossbauer spectral studies, showing that during the first discharge, Cu-metal cluster particles are formed by the reduction of Cu ion and incorporation of Li into the lattice to form Li2Sb2O6 as per eq 11 CuSb5 +2 O6 + 2Li → Cu(0) + Li 2Sb5 +2 O6

(11)

Due to the highly insulating (electronically) nature of Li2Sb2O6, only a small fraction of it is reduced to Sb(0) upon further discharge resulting in a negligibly small amount of Li3Sb alloy formation. The authors647 confirmed the above reaction mechanism by demonstrating that Li/Sb2O5 cells also do not show any electrochemical reversibility. On the other hand, the good reversibility of Li/Sb2O3 cells proves that in the case of (Co/Ni)Sb2O6, the discharge reaction of eq 10 operates by the reduction of Sb5+ ion to Sb3+ ion and finally to Sb(0). It may be possible to improve the cycling stability of the Li/(Co/ Ni)Sb2O6 system by restricting the voltage range from 0.005− 1.2 V ensuring only eq 9 contributes to the reversible capacity, with Co/Ni nanoparticles acting as matrix metals. Also, studies employing nanosize (Co/Ni)Sb2O6 of various morphologies will help in the improvement of Li cycling performance. 3.3.3. VSbO4, (M1/2Sb1/2Sn)O4 (M = V, Fe, or In), BiSbO4, SbPO4, and MSb2O4 (M = Ni or Co). VSbO4 is a cation-deficient mixed-valent oxide and has the formula [(V3+0.28,V4+0.64)(Sb5+0.92,□0.16)O4], where □ = vacancy. It has a tetragonal crystal structure. Morales et al.652 studied the Li cycling of VSbO4 prepared by mechanical milling, and they reported a reversible capacity of 650 and 450 mA·h g−1 in the cycling ranges of 0.5−2.5 and 0.5−1.5 V, respectively, at 0.08C rate. However, both the capacities degraded on cycling to ∼250 mA·h g−1 after 30 cycles. The group of Chowdari653 studied Li cycling of VSbO4 prepared by urea combustion and hightemperature methods in the range 0.005−3.0 V at a current of 60 mA g−1 (∼0.1C). They reported a stable capacity of 510 ± 10 mA·h g−1 for the VSbO4 prepared by the first method in the range 2−30 cycles, with a Coulombic efficiency ∼98%. The VSbO4 showed average discharge and charge potentials at ∼0.7 and ∼1.3 V, respectively, as per the reaction mechanism (eq 9). Recently, Chowdari’s group654 prepared the mixed vanadium antimony tin oxide, M (micrometer size)-(V1/2Sb1/2Sn)O4 with the tetragonal rutile structure by the high-temperature solidstate reaction. The N (nanosize, 10−20 nm) analogue was obtained by high-energy ball milling. Li cycling properties were studied at 24 and at 54 °C in cells with Li metal as the counter electrode by galvanostatic charge−discharge cycling and cyclic

Figure 25. Voltage vs capacity profiles of (V1/2Sb1/2Sn)O4 vs Li at ambient temperature (24 °C): (a) M-(V1/2Sb1/2Sn)O4 at 60 mA g−1 (0.1C rate), (b) M-(V1/2Sb1/2Sn)O4 at 250 mA g−1 (0.43C rate), and (c) N-(V1/2Sb1/2Sn)O4 at 250 mA g−1 (0.43C rate). (d) N(V1/2Sb1/2Sn)O4 at 0.43C rate cycled at 54 °C. The numbers refer to the cycle number. 1C = 580 mA g−1. M refers to micrometer size particles, and N refers to nanosize particles. Reproduced with permission from ref 654. Copyright 2011 The Royal Society of Chemistry.

These values are stable up to 100 cycles at 0.43C and up to at least 50 cycles at 1C and 3.5C. The capacity of 450 ± 5 mA·h g−1 corresponds to 4.5 mol of cyclable Li per mole of the compound. At 54 °C and at 0.43C and 1C, capacities of 440 ± 5 mA·h g−1 (stable up to 100 cycles) and 440 ± 5 mA·h g−1 (stable up to 70 cycles), respectively, were observed (Figure 26). The Coulombic efficiency, given by the ratio of discharge to the charge capacity at a given cycle number, stabilized and improved to almost 96−98% after five cycles. M-(V1/2Sb1/2Sn)O4 showed large reversible capacities at 0.1C with the upper cutoff voltages of 1.0 and 1.3 V, but slow capacity fading was noted when it was cycled both at 24 and at 54 °C and also at current rates of 0.43C and 0.72C (Figures 25 and 26). The M- and N-(V1/2Sb1/2Sn)O4 showed main cathodic and anodic peaks at ∼0.2 and ∼0.5 V, respectively, by CV. To understand electrode kinetics, the Chowdari group also carried out impedance studies of N-(V1/2 Sb1/2 Sn)O4 with Li metal as the counter electrode at various voltages and cycle number in the range 0.005−1.0 V at 0.43C rate. The Nyquist plots (Z′ vs −Z″) are shown in Figures 27 and 28. The equivalent electrical circuit to fit the impedance data is shown in Figure 28c. This consists of resistances (electrolyte, surface film (sf), charge-transfer (ct), and bulk resistance (b)), constant phase element (CPEi; sf, double layer (dl), and b) used instead of pure capacitor due to the composite nature of electrode, Warburg impedance (Ws), and intercalation capacitance (Ci). Results showed R(sf+ct) values during the first discharge and first 5394

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

systematically with an increase in the cycle number from 57 ± 3 Ω at the fifth cycle to 10 ± 3 Ω after 80 cycles. This is due to a systematic decrease in the contribution from the surface-film resistance to the total R(sf+ct) and possibly becomes negligibly small after 60 cycles. On the other hand, the R(sf+ct) values in the fully charged state (at 1.0 V) vary from 8 to 11 ± 3 Ω in the range 5−80 cycles, indicating only contribution from the charge-transfer resistance. Further, the bulk resistance (Rb) is almost constant, 18 ± 1 Ω in the discharged state, whereas no Rb contribution is seen in the charged state during 5−80 cycles Complementary ex-situ XRD data were reported to support the reaction mechanism involving Li alloying−dealloying reactions of Sn and Sb in the nanocomposite formed after the first-discharge reaction, with electronically conducting VO (vanadium oxide) acting as a beneficial matrix. To understand the effect of matrix element, very recently the group of Chowdari655 extended Li cycling studies on bare and ball-milled (M1/2Sb1/2Sn)O4 (M = In or Fe), and they also studied the effect of electrode heat treatment and carbon nanotube (CNT). In all cases, initial reversible capacities in the range 425−550 mA g−1 and observed capacity values are slightly lower compared with M = V.654 Li cycling studies on the layered oxide, BiSbO4 in the voltage range 1.2−3.0 V showed a first-charge capacity of ∼150 mA·h g−1, which degraded to ∼60 mA·h g−1 at the end of 14 cycles.656 Pena et al.657 reported the Li cycling of SbPO4 with the layered structure and found a reversible capacity of ∼220 mA·h g−1, which degraded to ∼165 mA·h g−1 after 20 cycles in the voltage range 0.25−1.25 V at a current rate of 0.05C. Under similar cycling conditions but with upper cutoff voltage of 2.0 V, higher reversible capacity was noted (∼330 mA·h g−1), but

Figure 26. Capacity vs cycle number plots: (a) M-(V1/2Sb1/2Sn)O4 at 24 °C and 60 mA g−1 (0.1C) or 250 mA g−1 (0.43C rate; 1C = 580 mA g−1), voltage range = 0.005−1.0 or 0.005−1.3 V; (b) N-(V1/2Sb1/2Sn)O4 at 24 °C and 0.43C, 1C, and 3.5C rates; (c) M- and N(V1/2Sb1/2Sn)O4 at 54 °C and 0.43C. (d) M- and N-(V1/2Sb1/2Sn)O4 at 54 °C at the current rates of 0.72C and 1C, respectively. Open symbols, charge capacity; closed symbols, discharge capacity. M refers to micrometer size particles, and N refers to nanosize particles. Reproduced with permission from ref 654. Copyright 2011 The Royal Society of Chemistry.

charge and at various cycle numbers from 93 to 11 ± 3 Ω.654 Interestingly, the R(sf+ct) in the discharged state decreases

Figure 27. Nyquist plots (Z′ vs −Z″) of N-(V1/2Sb1/2Sn)O4 at 0.43C and at various voltages: (a) first discharge cycle, (b) first charge cycle, (c) 100th discharge cycle, and (d) 100th charge cycle. Li metal was the counter electrode. Data were collected after stabilizing at each voltage for 2 h. Geometrical area of the electrode is 2.0 cm2. Selected frequencies are indicated. Reproduced with permission from ref 654. Copyright 2011 The Royal Society of Chemistry. 5395

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 28. Nyquist plots (Z′ vs −Z″) of N-(V1/2Sb1/2Sn)O4 in the (a) discharged state (0.005 V) and (b) charged state (1.0 V) at 0.43C and at various cycle numbers. Li metal was the counter electrode. Data were collected after stabilizing at each voltage for 2 h. Geometrical area of the electrode is 2.0 cm2. Selected frequencies are shown. (c) Equivalent electrical circuit used to fit the data of Figures 27 and 28. Re, electrolyte resistance; R(sf+ct), surface film + charge transfer resistance; CPE(sf+dl), constant phase element due to surface film + double layer capacitance; Rb, bulk resistance; CPEb, bulk capacitance; Ws, Warburg resistance; Ci, intercalation capacitance. Reproduced with permission from ref 654. Copyright 2011 The Royal Society of Chemistry.

the capacity value dropped to ∼165 mA·h g−1 after 20 cycles. The proposed reaction mechanism involves eq 9, and the measured potentials of alloying and dealloying process are at 0.7 and 1.3 V, respectively. Very recently, Jibin et al.658 reported Li cycling studies on tetragonal Pb3O4-type compound MSb2O4 (M = Co or Ni). The cyclic voltammetry studies showed a main cathodic and anodic peak at ∼0.57 and ∼1.06 V, respectively, for both compounds. CoSb2O4 and NiSb2O4 exhibit a reversible capacity (first cycle) of 490 and 412 mA·h g−1, respectively, in the voltage range 0.005−1.3 V at a current of 30 mA g−1 when cycled up to 25 or 35 cycles. The capacity slowly degraded to 348 mA·h g−1 at the end of 35th cycle for CoSb2O4 and to 290 mA·h g−1 at the end of 25th cycle for NiSb2O4.

Recent interesting studies by Komaba et al.664 reported Li cycling studies of SiO using poly(acrylic acid) (PAA) based binder, which showed a high reversible capacity of 826 mA·hg−1 (Figure 29) and retained a capacity of 720 mA·h g−1 at the end of the 50th cycle in the voltage range 0.00−2.0 V at a current rate of 50 mA g−1.

3.4. Other Group IV Oxides

3.4.1. SiO and SiO2. The so-called silicon monoxide, SiO, is a mixture of amorphous silicon (Si) and amorphous silica (SiO2). There are a good number of reports on the Li cyclability of “SiO” and its composites, where the electrochemically active species is only the Si, giving reversible capacity via alloying−dealloying reaction with Li in a matrix of lithium silicates formed during the first discharge by the reaction of SiO2 and Li. SiO and its composites are nicely reviewed by Park et al.136 and other recent additional studies are by Park et al.,659 Yamada et al.,660 Lu et al.,661 Jeong et al.662 Yamamura et al.,663 Komaba et al.,664 Si et al.,665 Yoo et al.,666 Yao et al.,667 Kim et al.,668 Miyuki et al,669 Ren et al.,670 Ban et al.,671 GuO et al.,672 Jeong et al.,673 and Song et al.674 The group of Chen675 reported electrochemical reduction of nano-SiO2 in hard carbon, prepared by the hydrothermal method, and showed evidence, by way of HR-TEM, Si NMR, and XPS spectra, that SiO2 is reduced to Si along with the formation of Li2O, Li4SiO4, or both. The alloying−dealloying of Si (Si + 4.4Li+ + 4.4 e− ↔ Li4.4Si) contributed to the observed reversible capacity of 630 mA·h g−1, which was found to be stable up to 14 cycles when cycled at 0.1 mA cm−2 in the range 0−3.0 V.

Figure 29. Relation between reversible capacity and cycle number for the SiO composite electrodes prepared with (a) PVdF, (b) PVA, (c) CMCNa, and (d) PAA as a binder. Reproduced with permission from ref 664. Copyright 2011 American Chemical Society.

3.4.2. GeO2 and Germanates. Pena et al.676 studied the electrochemical properties of the Li−GeO2 system in the voltage range 0.05−2.0 V at 0.05C or 0.005C rate. During the first discharge reaction, they showed that an amorphous lithium germanate, Li2O·GeO2, phase forms at 0.65 V, and in the range 0.55−0.35 V, crystal structure destruction occurs to form the nano omposite Li2O·Ge and (LiGe). This is followed by the formation of the alloy, Li4.2Ge in the voltage range 0.35−0.05 V. Cycling at 0.05C showed a drastic capacity fading, from ∼740 mA·h g−1 at the first charge to ∼220 mA·h g−1 at the 10th cycle. The reasons for capacity fading were discussed in terms of the large unit cell volume changes as a result of alloying− dealloying reactions, Ge + 4.2Li+ + 4.2e− ↔ Li4.2Ge. 5396

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Kim et al.677 studied Li cycling of GeO2 glasses prepared by melt quenching and obtained a reversible capacity of 550 mA·h g−1 when they were cycled in the voltage range 0−1.5 V at 0.1C rate. The capacity degraded slowly to 330 mA·h g−1 at the end of the 25th cycle. Preliminary Li cycling studies of GeO2, MGeO3 (M = Cu, Fe, and Co), and CuO·GeO2, composites were reported by Kim et al.678 A first cycle reversible capacity of 908, 949, 1007, 1038, and 932 mA·h g−1, respectively, was obtained for the above compounds at a current of 100 mA g−1 in the voltage range 0−3.0 V. During the first discharge, structure destruction followed by the formation of Ge and M metal nanoparticles and the alloy formation, Li4.2Ge, was established by X-ray absorption spectroscopy and X-ray diffraction. The CuGeO3 showed lower electrode polarization in comparison to Cu/GeO2 or GeO2. The group of Lu679,680 examined the Li cycling of LiGe2P3O12 (LGP) and Li1.5(Al0.5Ge1.5)P3O12 (LAGP), which possess the NASICON-type 3D structure containing GeO6 octahedra, PO4 tetrahedra, and Li ions occupyng the 3D channels. A reversible capacity of 460 mA·h g−1 was obtained when LGP was cycled in the voltage range 0.001−1.5 V at a current rate of 150 mA g−1, and 92% of the capacity was retained after 25 cycles. At a high current of 1500 mA g−1, 77% capacity retention was noted after 1000 cycles.679 When LAGP was cycled in the range 0−2.5 V at a current density of 0.1 mA cm−2, a first-charge capacity of ∼340 mA·h g−1 was obtained, which slowly degraded to ∼220 mA·h g−1 after 100 cycles. The current rate capability was also examined, and ∼60% of the capacity was obtained for a 10-fold increase of the current rate during initial cycles.680 Additional recent Li cycling studies on GeO2 thin films and CuGeO3 by the group Lu Li are reported in refs 681 and 682.

Thereafter, slow capacity fading was observed, leading to a capacity of ∼310 mA·h g−1 after 40 cycles. Under similar cycling conditions, bulk powder ZnO electrode showed an initial reversible capacity of ∼250 mA·h g−1, which degraded to ∼90 mA·h g−1 after 20 cycles. The Li cycling studies of bare and Ni-coated ZnO prepared by electroless plating showed initial reversible capacities of ∼630 and ∼890 mA·h g−1, respectively, when cycled at current of 80 mA g−1 in the range 0.02−3.0 V.686 Severe capacity degradation was noted up to ∼10 cycles and a slow degradation up to 30 cycles. However, the performance of Ni-coated ZnO was better in comparison to uncoated ZnO, since the reversible capacities at the end of 30 cycles were ∼490 and ∼100 mA·h g−1, respectively. Pan et al.687 reported the Li cycling properties of flower-like thin films of ZnO−NiO−C, which were heat-treated at 300−500 °C. They observed first-cycle reversible capacities of 488, 805, and 846 mA·h g−1 for the films heat-treated at 300, 400, and 500 °C, respectively, when they were cycled at 0.5C rate in the range 0− 3.0 V. Good cyclability was observed in all cases, and a capacity of 485 mA·h g−1 was retained after 50 cycles for the 500 °Ctreated composition. Good C-rate capability was also observed when the films were tested up to 4C and up to 170 cycles (400 °C-treated composition). Liu et al.688 studied carbon-coated ZnO nanorods on Ni-substrates and reported initial reversible capacities of 550, 471, and 385 mA·h g−1 at 0.35C, 0.75C, and 2C rates, respectively, when cycled between 0 and 2.5 V (1C = 987.8 mA g−1). At 0.75C, the discharge and charge capacities were, 360 and 324 mA·h g−1, respectively, after 25 cycles. Recently, Huang et al.689 reported the Li cycling of porous ZnO nanosheets grown directly on copper substrates by the chemical bath deposition technique followed by a heat treatment. They showed initial discharge and charge capacities of 1120 and 750 mA·h g−1 and retained a stable capacity of ∼400 mA·h g−1 at the end of 100 cycles, when cycled at current of 0.5 A g−1 in the voltage range 0.005−3.0 V. They also found good C-rate capabilities up to 2 A g−1. Wu et al.690 reported flower-like ZnO−CoO−C nanowall arrays prepared through solution-immersion steps followed by heat treatment 400 °C. When cycled in the voltage range 0.02−2.5 V, they showed a reversible capacity of 438 mA h g−1 at the end of the 50th cycle at a current rate of 0.5C and 224 mA h g−1 at 2.0C rate. The probable reason for the improvement in cycling performance in comparison to bare ZnO is ascribed to the catalytic effect of Co and conductive carbon layer coating. Ahmad et al.691 prepared hierarchical flower-like ZnO nanostructures and functionalized by Au nanoparticles using a combination of solution and electrodeposition techniques. They showed an initial discharge capacity of 1280 mA h g−1 and a reversible capacity of 392 mA h g−1 at the end of 50 cycles when the composites were cycled in the range 0.02−3.0 V vs Li at current rate 120 A g−1. Recent additional studies on ZnO composites are given in refs 692−696. The theoretical reversible capacity of ZnO via the alloying− dealloying reaction is 375 mA·h g−1, and it occurs at potentials 0 to ∼0.6 V vs Li.464,479,683 It is clear from the above discussion that when ZnO is cycled in the voltage range 0−3 V, in addition to the above alloying−dealloying reaction, oxidation of Zn, formed after the first-discharge reaction, will also occur due to the redox (conversion) reaction with the electrochemically formed Li2O. This is the reason for the observed high capacities. Since both of the above processes involve large volume changes, capacity fading will inevitably take place, as is observed. Use of matrix elements (Ni, carbon) and particle size

3.5. Group II and III Oxides, ZnO, CdO, and In2O3

Binary oxides ZnO, CdO, and In2O3 have been studied for their Li cycling behavior via the alloying−dealloying reaction since the respective metals can form alloys with Li, such as, LiZn, Li3Cd, and Li3In at potentials V ≤ 0.7 V, and thus large reversible capacities are expected. The group of Chen464 reported preliminary Li-cycling studies of ZnO, which showed a first-discharge and -charge capacity of 640 and 390 mA·h g−1, respectively, in the voltage range 0−2.0 V at a current of 0.2 mA cm−2. However, the reversible capacity degraded to ∼250 mA·h g−1 after 13 cycles. The group of Irvine479,683 studied the Li cycling of bare and ball-milled ZnO and ZnO−SnO2 composites. A reversible capacity of 359 and 412 mA·h g−1 for the bare and ball-milled ZnO, respectively, was observed in the cycling range 0.02−1.0 V at a current of 0.11 mA cm−2. Compositions of bare and ball-milled ZnO− SnO2 (1:1 and 1:2) showed reversible capacities ranging from 428 to 613 mA·h g−1. However, capacity fading was observed in all cases when they were tested up to 20 cycles. Fu et al.684 prepared ZnO thin films by pulsed laser deposition technique on stainless steel substrates and studied their Li cyclability in the range 0.1−3.0 V at a current density of 20 μA cm−2. Initial reversible capacities in the range 380−450 mA·h g−1 were observed depending on the deposition conditions of the films. These values slowly degraded to 200−250 mA·h g−1 after 40 cycles. Wang et al.685 have grown ZnO nanorods on Cu substrates and examined their Li cyclability in the voltage range 0−3.0 V at 0.1 mA cm−2. Initial reversible capacity as high as ∼1000 mA·h g−1 was observed, which rapidly decayed to 420 mA·h g−1 after 5 cycles. 5397

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

matrix. During charge, the re-formation of MO is seen as a consequence of decomposition of Li2O. Binary oxides, MO, with M = Mn, Co, and Ni adopt the cubic rock salt structure, whereas FeO adopts a cubic defect structure. CuO adopts a monoclinically distorted rock salt structure due to the presence of Jahn−Teller ion. 4.1.1.1. CoO. The group of Tarascon702−705 first reported systematic studies on MO with M = Fe, Co, Ni, and Cu as anode materials via the “conversion/displacement reaction” (eq 12). The voltage vs composition (x in LixMO) profiles of various MO (M = Fe, Co, or Ni) are shown in Figure 30a. As

reduction can suppress the capacity fading to some extent, as has been noted in the cases of ZnO nanorods, films with or without Ni, composites, etc. Careful and detailed studies are needed to improve the performance of ZnO as a viable anode material. Preliminary Li cycling of CdO showed a reversible capacity of ∼350 mA·h g−1 when cycled in the voltage range 0−3.0 V at current of 110 mA g−1.697 Since cadmium is highly toxic there will not be much interest in its Li cyclability and use as an anode in LIBs. Li cycling studies of In2O3 were reported by the group of Chen.464 Initial reversible capacity of ∼360 mA·h g−1 slowly degraded to ∼200 mA·h g−1 after 11 cycles when it was cycled in the voltage range 0−2.0 V at current rate of 0.2 mA cm−2. Studies of In2O3 in the form of thin films have been reported. Nanostructured thin films (180 nm thick) deposited onto stainless steel substrates were examined for their Li cyclability by Zhou et al.698 When cycled at 10 μA cm−2 in the range 0.01−3.5 V, a second discharge capacity as high as 883 mA·h g−1 (8.9 mol of Li per mole of In2O3) was obtained, which gradually decayed to ∼520 mA·h g−1 at the end of 30 cycles. Since the theoretical reversible capacity of In2O3 via alloying− dealloying of In (3Li + In ↔ Li3In) is only 578 mA·h g−1, the authors suggested that “conversion reaction” of In with Li2O also occurs during cycling. Ho et al.699 studied nano-In2O3 films (40 nm), grown on Pt substrates by the cathode electrodeposition method, in the voltage range 0.2−1.2 V at a current of 50 μA cm−2. The first-discharge capacity was as high as 1400 mA·h g−1, which degraded to ∼200 mA·h g−1 and stabilized in the range 2−10 cycles. Recently, Yang et al.700 reported the Li cyclability of In2O3 nanobelts of thickness 20 nm obtained by plasma-enhanced chemical vapor deposition. When they were cycled in the range 0.005−2.0 V at a current density of 250 mA g−1, the first-charge capacity of ∼1700 mA·h g−1 rapidly degraded to 680 mA·h g−1 at the end of 30 cycles. However, in the range 30−100 cycles, the reversible capacity decreased slowly, resulting in a value of 580 mA·h g−1 after 100 cycles. The Coulombic efficiency ranged from 95% to 98% during 30− 100 cycles. Recently Liu et al.701 reported a discharge capacity of 569 mA·h g−1 at the end of 50 cycles at a current rate of 0.1C for In2O3/carbon core−shell nanospheres.

Figure 30. Properties of MO/Li cells (MO = transition-metal oxides): (a) the voltage−composition profile for various MO/Li cells cycled between 0.01 and 3 V at a rate of C/5 (one lithium in 5 h). The capacity fading for the same cells under similar conditions is shown in panel b, which also shows data for a Co3O4/Li cell in order to show that the reported behavior is not specific to divalent oxides. Inset, the rate capability of a CoO electrode. Reproduced with permission from ref 702. Copyright 2000 Nature Publishing Group.

4. ANODES BASED ON CONVERSION (REDOX) REACTION 4.1. Binary Oxides

4.1.1. Oxides with Rock Salt Structure (MO; M = Mn, Fe, Co, Ni, or Cu). The reaction mechanism involved during Li storage and cycling is different from that of intercalation/ deintercalation and alloying/dealloying and involves the formation and decomposition of Li2O, along with the reduction to and oxidation of metal nanoparticles, the so- called “conversion/displacement reaction” (eq 12).

can be seen, for all M, a flat voltage region during the first discharge at a voltage of 1.0−0.8 V is seen, typical of a twophase reaction, that is, coexistence of virgin CoO and the reaction products, Co and Li2O, up to the end of the voltage plateau region. This is followed by a sloping region up to the deep discharge limit of 0.01 V, indicating a single-phase Li insertion reaction (Figure 30a). CoO showed a reversible capacity of ∼700 mA·h g−1 (theoretical value, 715 mA·h g−1 as per eq 12) in the voltage range of 0.01−3.0 V when the system was cycled at 0.2C (1 mol of Li in 5 h), and the capacity remained stable for several discharge−charge cycles (Figure 30b).702−704 The voltage plateau noticed corresponds to the crystal structure destruction (amorphization) and formation of Co/Li2O, where the nanosize Co metal particles (∼4−5 nm)

MO + 2Li+ + 2e− ↔ M + Li 2O (M = Mn, Fe, Co, Ni, Cu)

(12)

Normally, Li2O is electrochemically inert, but it can participate during electrochemical cycling due to the “in situ” electrochemically formed nanoparticles, because the latter catalyze the reaction. First-discharge reaction with Li metal involves crystal structure destruction (amorphization of lattice) followed by the formation of nanoparticles of metal embedded into the Li2O 5398

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

are embedded in a matrix of X-ray amorphous Li2O. The extra capacity (consumption of more than 2.0 mol of Li per mole of CoO noticed during the first-discharge process) is ascribed to factors, such as (i) formation of the solid electrolyte interphase (SEI) and (ii) formation of a polymeric gel-type layer on the Co metal nanoparticles upon deep discharge (0.01 V). Both of the above are due to the catalytic decomposition of the solvents of the electrolyte.702,703 During the charging process, reformation of CoO is noticed, which is supported by in situ XRD,702−704,706 TEM,702−704,706 X-ray photoelectron spectroscopy (XPS),707 and X-ray absorption spectroscopy studies.708,709 Studies showed that charging to voltage >2.0 V vs Li is necessary in order to achieve a stable capacity because dissolution of the polymeric layer (formed during discharge) occurs above 2.0 V.710 A brief discussion of the SEI and the polymeric gel-type layer on the CoO electrode during Li cycling is in order. The SEI forms on the anode material particles, whether it is graphite, alloy-forming elements like Si and Sn, or metal oxides like CoO at voltages less than 1 V vs Li during the first-discharge process (Li insertion). This is due to the reduction of the solvents present in the electrolyte, namely, ethylene carbonate (EC)/ diethyl carbonate (DEC), aided by the presence of the salt LiPF6. Studies have shown184,186 that the SEI is ∼3−6 nm thick and consists of a heterogeneous mixture of Li2CO3, ROLi, ROCO2Li (R = alkyl group), and polycarbonates. It may also contain F-containing species (moieties with Li−C and C−F bonds.186 The formation of SEI leads to irreversible capacity loss (ICL) during the initial discharge−charge cycle due to the consumption of Li ions and electrons. But the SEI fully develops and becomes stable during subsequent cycles and is beneficial since it allows the unsolvated Li ions to permeate through it and enables good ion transfer, while it is electronically insulating.707 Detailed studies by the group of Tarascon employing TEM707,710 on CoO/Li cells in the discharged state (at 0.7 and 0.02 V) and in the charged state (at 1.8 and 3 V) at 0.2C rate have shown that, in addition to the SEI layer, a polymeric gel-type film of thickness ranging from 50 to 100 nm forms on the electrode in the fully discharged state and it either remains modified or disappears when charged to 1.8 V but completely disappears upon charging to 3 V. The C 1s XPS spectra of CoO discharged and charged to various voltages are shown in Figure 31. The starting material, CoO, showed a peak at 285 eV (binding energy) attributed to the contaminating hydrocarbon layers (adventitious carbon) (Figure 31a). In the spectrum of the CoO discharged to ∼0.7 V (consumption of 1 mol of Li per mole of CoO), a distinct peak at 289 eV is seen indicative of COO-like carbons (Figure 31b). After Ar ion etching, the sample showed the presence of Li2CO3 and alkyl carbonate species, clearly indicating that the SEI formed is composed of these heterogeneous species. After the system is discharged to 0.02 V, a high-intensity peak appears at 286.0 eV and corresponds to the presence of CO-like carbons (Figure 31c). In addition, a new peak appears at 290.0 eV with a small shoulder peak at 289.0 eV. The latter two peaks are ascribed to the SEI, as mentioned above. The 286.0 eV peak can be assigned to poly(ethylene oxide)-type polymer or oligomers, which are formed as the gel-like layer on the Co particles. This layer is in addition to the already-formed SEI layer. The spectra obtained after charging to 1.8 and 3 V clearly show the disappearance of the peak at 286.0 eV, while both the peaks at 289 and 290 eV remain unchanged except for slight broadening

Figure 31. C 1s XPS spectra of (a) starting CoO powder, (b) discharged at x = 1 Li (in LixCoO), (c) discharged at 0.02 V, (d) charged at 1.8 V, and (e) charged at 3.0 V. Reproduced with permission from ref 707. Copyright 2004 American Chemical Society.

and overlap (Figure 31d,e). These results show that the organic species corresponding to the peak at 286 eV disappear during charge, while the SEI layer remains unchanged. The TEM bright field images of the CoO electrode after first discharge to 0.02 V and at various voltages after cycling are shown in Figure 32. As can be seen, the image in the fully discharged state shows a veil-type polymeric film (thickness ∼100 nm) on the Co particles (Figure 32a). The image of the electrode discharged to 0.02 V after 10 cycles between 0.02 and 1.8 V still shows the polymeric film (∼50 nm) but with a different texture (Figure 32b). Similar is the case with the electrode that was charged to 1.8 V after 10 cycles between 0.02 and 1.8 V indicating that the ∼50 nm film may have been modified but is stable (Figure 32c). On the other hand, the electrode charged to 3 V after cycling shows the quasi-complete disappearance of the thick polymer gel-like coating and only the very thin SEI layer (∼3 nm) is seen surrounding the reoxidized (CoO) nanoparticles (Figure 32d). Laruelle et al.710 mention that the measured thicknesses of the polymeric gel-type layer is accurate to only ±20%, the reason being the sensitivity of the layer to long-term electron beam exposure. Based on the preliminary IR,710 mass spectrometry,704 and detailed XPS data,707 the Tarascon group proposed that the gel-type layer is formed as a result of the nanoparticle (Co) driven electrolyte (solvent) decomposition presumably leading to free radicals, part of which polymerize or oligomerize to elastic fragments on the electrode surface, and this is made up of similar functional groups as the SEI layer. Formation of a polymeric gel-type layer has also been observed on other anode electrodes, like, CoF2, CoS, and Cu3N.704 However, more careful and detailed studies are needed to understand the mechanism of the formation of the polymeric gel-type layer and its role in the reversible cycling of CoO and other oxide anodes. The current (C)-rate capability of CoO has also been tested from 0.2C to 2C up to 250 cycles.711 As expected, due to kinetic 5399

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 32. TEM bright field images realized (a) on a CoO-based electrode after its first discharge down to 0.02 V, (b) on an electrode that was at 0.02 V after 10 cycles between 0.02 and 1.8 V, (c) on a electrode that was at 1.8 V after 10 cycles between 0.02 and 1.8 V, and (d) on a electrode that was cycled and fully reoxidized up to 3 V. Note that only the polymeric layer disappears. The 30 Å SEI inorganic layer can be seen on the fully charged sample (d). Reproduced with permission from ref 710. Copyright 2002 The Electrochemical Society.

interface electrochemistry in conversion reactions, as applied to the CoO electrode, through first-principles DFT (density functional theory) or DFT + U calculations. This method describes the main interface effects by means of three interdependent descriptors, namely, the chemical (interface bonding), the mechanical (stress), and the electrical/electrochemical (electric bias) descriptors. On the basis of these results, the most probable interfaces occurring during the CoO conversion have been determined, and a probable electrode morphology has been proposed, with the sequence, Li2O/ CoO/Co0/CoO/Li2O (Figure 33). Furthermore, the addition

considerations, there was a 33% drop in the reversible capacity going from 0.5C to 2C, namely, from ∼600 to ∼400 mA·h g−1. However, the capacity values remained more or less stable at 0.2C up to 50 cycles and at 0.5C, 1C, and 2C up to 250 cycles.704 Reversible capacities in the range 750−1000 mA·h g−1, which are much larger than the theoretical values, observed by Grugeon et al.704 both at 55 and at 75 °C, have been ascribed to the decomposition of the polymeric layer that is formed during the discharge reaction. Since the original work of Poizot et al.,702 CoO has been studied as anode in the form of micrometer-size particles,712 nanosize particles,709 nanoplatelets,713 porous nanowires,714 and composites with carbon/graphene715−719 and found to perform well. Yu et al.720 prepared composite films of CoO− Li2O, both as dense films and as Ni-foam-supported reticular films, by electrostatic spray deposition and examined their Li cycling behavior. They found that the dense films showed capacities in the range 450−550 mA·h g−1 when cycled at 0.2C in the voltage range 0.01−3.0 V vs Li up to 100 cycles. Under similar cycling conditions, the reticular films showed capacities in the range 650−780 mA·h g−1, and in both cases, the capacities increased with an increase in the cycle number. The formation of higher oxides (Co3O4 and Co2O3) by the decomposition of Li2O in the composite was invoked to explain the observed data. Yao et al.713 prepared multilayered CoO platelets and nanoparticles, and they compared the Li cycling behavior of the oxides (range 0.01−3.0 V vs Li) with different surface areas ranging from 20 to 9 m2 g−1. The CoO platelets with surface area 10 m2 g−1 delivered a high and stable capacity of 800 mA·h g−1 up to 100 cycles at a current of 100 mA g−1. CoO−mesoporous carbon composites exhibited a reversible capacity in the range 500−700 mA·h g−1 after 20 cycles in the voltage range 0.01−3.0 V at a current of 100 mA g−1,715 whereas CoO/C hybrid microspheres delivered a capacity of 345 mA·h g−1.716 Recently, Dalverny et al.721 presented a new methodology based on a multi-interface superlattice approach to investigate

Figure 33. Schematic representation of the most probable electrode morphology as deduced from the computation of interface energies and energy stress. Reproduced with permission from ref 721. Copyright 2011 The Royal Society of Chemistry.

of an external redox potential to the multi-interface superlattice has revealed asymmetric electrochemical interface phenomena in charge and discharge, shedding some light into the voltage hysteresis experimentally observed for the CoO conversion/ reconversion. Recent additional studies on the Li cyclability CoO are core−shell Co@CoO nanocomposites by Zhang et al.,722 echinus-like nanostructures of mesoporous CoO nanorod@CNT by Wu et al.,723 Cu-doped h-CoO nanorods by 5400

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Nam et al.,724 CoO−PVDF nanocomposites by Valvo et al.,725 CoO−MnO nanostructures by Kukubu et al.,726 CoO/Cu nanostructures by Qi et al.,727 CoO nanodisks,728 octahedral nanocages Guan et al.,729 MgO/CoO and Al2O3/CoO composites by Yu et al.,730 plate-like CoO/carbon nanofiber composites by Yao et al.,731 and porous hierarchical-like CoO@ C microsheets by Liu et al.732 (Table 7). Very recently, Reddy et al.733 reported submicrometer-sized CoO made by a carbothermal reduction method. They showed a reversible capacity of 875 mA·hg−1 at the end of the 60th cycle in the voltage range 0.005−3.0 V at a current rate of 60 mA g−1 (Table 7). 4.1.1.2. NiO. Li cyclability of NiO through conversion reaction was first reported by Poizot et al.,702 who showed that an initial reversible capacity of ∼600 mA·h g−1 slowly degrades to ∼200 mA·h g−1 at the end of 50 cycles (Figure 30b). Since then several groups reported studies on nanoparticles and thin films of NiO, and composites with carbon in the voltage range 0.01−3.0 V. Huang et al.734 studied Li cycling of NiO−Ni composites and found reversible capacities of ∼500 mA·h g−1 for bare NiO and 720 mA·h g−1 for Ni−NiO composite at the end of the 30th cycle at a current rate of 100 mA g−1. Hosono et al.735 studied mesoporous Ni/NiO structures and observed a capacity of 695 mA·h g−1 at a high current density of 10 A g−1. The group of Chowdari736 reported the Li cyclability of NiO nanowalls prepared on Ni foil and found a stable and reversible capacity of 635 mA·h g−1 (at 1.25C rate) in the range 2−80 cycles. At ∼1.9C (1343 mA g−1), a stable capacity of ∼490 mA·h g−1 was observed in the range 2−50 cycles (Figure 34; Table 7). NiO prepared by spray pyrolysis followed by heating at different temperatures showed reversible capacities ranging from 390 to 720 mA·h g−1 at the 10th cycle at current rate of 1 mA cm−2.737 But, in all cases drastic capacity fading was noted after 20 cycles. Net-structured NiO and NiO−C (carbon) composites have been studied by Huang et al.738 in the voltage range 0.02−3.0 V vs Li at 0.1C (1C = 718 mA g−1) up to 40 cycles. An initial reversible capacity of ∼750 mA·h g−1 of NiO continuously degraded to 178 mA·h g−1 after 40 cycles. On the other hand, the NiO−C performed better, with an initial reversible capacity of ∼650 mA·h g−1 slowly degrading to 429 mA·h g−1 after 40 cycles. Ni/NiO core−shell particles,739 NiO hollow spheres,740 NiO microsperes,741 NiO nanotubes,742 NiO/carbon nanocomposites, 743−747 NiO porous thin films,748−750 and NiO/polyaniline, poly(3,4-ethylenedioxythiophene) (PEDOT), and Ag composites751−753 have been studied, and they showed reversible capacities ranging from 250 to 650 mA·h g−1 after 20−50 cycles at current rates of 0.1− 1C in voltage range 0.005−3.0 V. Other recent reports on the Li cycling of NiO are mesoporous NiO by Li et al.754 and Ma et al.,755 nanoporous NiO films by Chen et al.756,757 and Yao et al.,758 nanostructured NiO by Wang et al.,759 NiO nanocone by Wang et al.,760 NiO nanosheets grown on nanoflowers by Gao et al.,761 NiO/Co−P nanocomposites by Huang et al.,762 self-assembled sandwichlike NiO film by Zhong et al.,763 Co-doped NiO nanoflake arrays by Mai et al.,764 NiO nanoflakes by Ni et al.,765 NiO/Ni composites by Wen et al.,766 Li et al.,767 Mai et al.,768 and Wu et al.,769 NiO/RuO2 carbon nanofibers by Wu et al.,770 NiO/ multiwalled carbon nanotube composites by Xu et al.,771 NiO/ graphene composites by Kottegoda et al.,772 Zou and Wang,773 Huang et al.,774 Mai et al.,775,776 Qiu et al.,777 and Zhou et

al.,778 and NiO/reduced graphene oxide by Zhu et al.,779 and capacity values are summarized in Table 7. 4.1.1.3. CuO and Cu2O. Copper oxides, Cu2O and CuO, are attractive materials as anodes via conversion reactions, since they are cheap and environmentally acceptable with theoretical reversible capacities of 375 and 674 mA·h g−1, respectively. The group of Tarascon702,711,780 have studied these two oxides and found that the particle size plays a significant role in giving stable and reversible capacities, and they established the mechanism of Li cycling. Stable capacities of ∼400 mA·h g−1 up to 70 cycles at 0.2C rate were obtained when these were cycled in the range 0.02−3.0 V for both Cu2O and CuO with a particle size of 1 μm. On the other hand, continuous capacity fading was noted when the particle size was 0.15 μm. With the CuO−Li system, an intermediate phase (Cu2O) always forms during the charge reaction. Studies on nanorod Cu2O are reported by Lee et al.781 and on 1−2 μm size cube-shaped Cu2O by Fu et al.,782 who found near-theoretical and stable capacities. Studies on CuO in the form of thin films,783,784 sisal-like CuO/C films,785 polycrystalline powders,737 nanowires,786 nanoribbon arrays,787 leaf-like nanoplates788,789 and hierarchical nanostructures790 have been reported. In the latter two cases, continuous capacity fading was noted irrespective of the morphology when cycled up to 50 cycles. The nanowires of CuO gave an initial reversible capacity of 720 mA·h g−1, and a capacity of ∼650 mA·h g−1 was stable up to 100 cycles, when it was cycled in the voltage range 0.02−3.0 V at 0.5C rate.786 The nanoribbon arrays of CuO also perform well: an initial reversible capacity of ∼500 mA·h g−1 slowly increased to ∼610 mA·h g−1 upon cycling to 275 cycles, when the nanoribbons were cycled in the voltage range 0.02−3.0 V at a current rate of 175 mA g−1. Good C-rate capability, tested up to 800 mA g−1, was also noted.787 Recent additional references on the synthesis and electrochemical Li cycling of CuO are CuO hierarchical hollow microspheres by Guan et al.,791 CuO nanostructures by Zhang et al.,792 CuO hollow nanospheres by Kong et al.,793 porous CuO by Wan et al.794 and Chen et al.,795 hollow structures of CuO by Ju et al.,796 nanoplates of CuO by Xiang et al.,797 nanocrystal thin film of CuO by Feng et al.,798 CuO microscale cog-like superstructures by Zhang et al.,799 CuO nanocomposite films by Lamberti et al.,800 CuO/polypyrrole core−shell nanocomposites by Yin et al.,801 pine-needle-like array structures of CuO by Chen et al.,802 CuO leaf-like mesocrystals by Xu et al.,803 CuO sponge-like architecture by Choi et al.,804 CuO nanowalnuts and nanoribbons by Yu et al.,805 micro-CuO/C microspheres by Huang et al.,806 CuO/C core−shell nanowires by Liu et al.,807 carbon/CuO composites by Kim et al.808 and Zhong et al.,809 CuO fibers by Sahai et al.,810 CuO/TiO2 composite thin films by Barreca et al.,295 CuO/CNT composites by Ko et al.,811 and CuO/graphene composites by Mai et al.,812 Zhou et al.,813 Wang,814,815 Qi et al.,816 and Guo et al.817 (Table 7). Very recently, Reddy et al.270,818 reported Li cycling studies on CuO micrometer/ nanoflake walls prepared using molten salt synthesis in the temperature range 410−950 °C. The reversible capacity at the end of the 40th cycle is 236, 250, 620, 551, and 432 mA·h g−1, respectively, at a current rate of 60 mA g−1 in the voltage range 0.005−3.0 V for the samples prepared at 410, 510, 750, 850, and 950 °C (Table 7). 4.1.1.4. MnO. There are a few reports on the Li cyclability of the MnO−Li system. While the initial results by Poizot et al.819 5401

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

5402

CuO nanofibers CuO/CNT composites CuO (prepared by molten salt method) at 410 °C 510 °C (slow cooling) 510 °C (quenched) 750 °C 850 °C 950 °C carbon-coated MnO MnO/graphene nanosheets MnO/reduced graphene oxide sheet (MnO/RGOS)

NiO flakes NiO nanospheres ammonia-induced nanoporous NiO films NiO nanoplates/graphene NiO/graphene composites (37 wt %) NiO/RuO2 carbon nanofibers CuO particles (prepared by polyol method at 160 °C) CuO/graphene composites CuO CuO/graphene CuO/graphene composites CuO/graphene composites CuO CuO/G

avg crystallite size 19 nm (XRD); particle size of 1−2 μm nanoparticles with diam ranging from 20 to 250 nm nanoparticles

submicrometer sized particles and flakes

fiber-like morphology avg dimension, 110 nm long axis and 34 nm short axis micrometer sized particles

0.19 0.61 17.34 11 25.37

0.2

50 100 100

65 60 600 60 600 100 0.1C 60

urchin-like feature with an avg diam of about 1.5 μm porous sheet-like flakes

72 0.2C

100 300

67

65

16

30 nm particles on GO sheets

nanofibers with 80 nm diam 1 μm

100 nm length, 10 nm thickness 20−120 nm sized particles on graphene

20 nm flake thickness, pore size 20−200 nm 200−500 nm diam spheres, avg size 20 nm porous film thickness 400−500 nm, 30−60 nm particles

895 (1.25C) 100 100 143

60

submicrometer size

100 nm length, 40 nm thick

50 11.8

200 200

Cu. Complementary HR-TEM and B-NMR studies were also reported to confirm the conversion reactions in these borates. From the above discussion, it is clear that the borate ion is not a good “matrix” for the conversion reaction in terms of the capacity retention over a large number of cycles. 4.3.5. Metal Oxysalts: Carbonates, Oxalates, Oxyhydroxides, and Oxyfluorides. The group of Tirado820,1160,1161 was the first to explore the Li cyclability of metal carbonates and oxalates. Submicrometer size particles of MnCO3 showed a first-charge capacity of ∼600 mA·h g−1, when cycled between 0.0 and 3 V at 0.25C rate. This value is larger than the theoretical capacity (466 mA·h g−1) assuming the conversion reaction involving Mn0 + Li2CO3. However, the observed capacity slowly degraded to ∼460 mA·h g−1 after 25 cycles. The metal oxalates, Fe(II)C2O4 and Co(II)C2O4, were prepared in the form of mesoporous nanoribbons, and their Li cyclability was examined in both the hydrated and dehydrated form.1160,1161 Fe(II)C2O4 showed capacities ranging from 500 to 700 mA·h g−1 between 5 and 75 cycles when cycled at 2C rate (1C = 1 Li h−1 mol−1) in the voltage range 0−2 V vs Li. At 5C rate, the capacity values ranged from 400 to 500 mA·h g−1. Complementary Mossbauer and FT-IR spectral studies showed that the original compound does not re-form upon charging to 2 V after the first deep discharge to 0 V. Similarly, Co(II)C2O4 also showed reversible capacities ranging from 400 to 800 mA·h g−1 between 10 and 80 cycles, when cycled at 2C rate (1C = 1 Li h−1 mol−1) in the voltage range 0.01−3.0 V vs Li. Good Crate capability was also shown.658,659 Aragon et al.1162 recently reported the Li cycling of the oxalate solid solutions, (FexCo1−x)C2O4, x = 0.25, 0.5, and 0.75 in the form of 40 nm wide nanoribbons with porous structure and found a synergistic effect of Fe and Co for x = 0.75 in displaying a reversible capacity of ∼600 mA·h g−1 at 5C rate with very good capacity retention after 75 cycles. Results also showed that besides the usual Faradaic contribution coming from the conversion reaction, the total reversible capacity of the iron and cobalt oxalate electrodes contains a significant contribution of capacitive effects. Iron oxyhydroxide, FeOOH (mineral, goethite), in the form of nanorods (40 nm diameter and 100−200 nm in length) was prepared by Lou et al.1163 and studied for its Li cyclability by conversion reactions in the range 0.01−3.0 V at a current of 100 mA g−1. They reported a first-charge capacity as high as 1044 mA·h g−1, which quickly faded to 600 mA·h g−1 after 20 cycles and finally stabilized at ∼500 mA·h g−1. Titanium and niobium oxyfluorides, TiOF2 and NbO2F, are well-defined compounds with the cubic ReO3-type structure. In them, the O and F ions are randomly distributed in the MO6 octahedra, the latter being corner-linked to form a 3D-network. Chowdari’s group1164 studied the Li cyclability of the above

compounds. A stable capacity of 400 ± 5 mA·h g−1 (1.5 mol of cyclable Li per mole) was exhibited by TiOF2 when cycled between 5 and 100 cycles at a current of 65 mA g−1 in the voltage range 0.005−3.0 V vs Li (Figure 53). On the other

Figure 53. (a) Galvanostatic charge−discharge curves of TiOF2 in the voltage range 0.005−3.0 V vs Li at 30 mA g−1. The cycle number is indicated. (b) Capacity vs cycle number for TiOF2. Voltage ranges and specific current are indicated. Filled and open symbols represent discharge and charge capacity, respectively. Reproduced with permission from ref 1164. Copyright 2006 Elsevier.

hand, NbO2F gave a stable capacity of only 180 ± 5 mA·h g−1 (∼1 mol of cyclable Li per mole) between 10 and 40 cycles at a current of 30 mA g−1. Conversion reactions involving the nanocomposite “LixTiOy +LiF” were invoked to explain the observed excellent Li cyclability. Chen et al.1165 studied TiOF2 nanocubes prepared by a hydrothermal method, and they reported a discharge capacity of 510 mA·h g−1 after 40 cycles at a current of 30 mA g−1, when cycled in the voltage range 0.25− 3.0 V vs Li. Recently Zeng et al.1166 reported Li cycling studies of TiOF2 nanotubes prepared by a solvothermal method, and they showed reversible capacity of 580 mA·hg−1 at a current rate of 30 mA g−1 at the end of 60th cycle.

5. ANODES BASED ON BOTH ALLOYING−DEALLOYING AND CONVERSION REACTION Since binary iron and cobalt oxides show large and stable reversible capacities upon Li cycling via conversion reactions, it is an interesting idea to explore simpler compounds in which a Li alloying−dealloying element is also present along with a transition metal, and synergistic effect may enable Li cycling of both metals to yield larger and stable reversible capacities. Studies carried out in recent years, indeed, show that the concept works in practice as has been shown by employing mixed oxides ZnM2O4 (M = Fe, Co, and Mn), M−Sn oxides, and metal carbonates. 5.1. Oxides with Spinel Structure: ZnM2O4 (M = Co, Fe) and CdFe2O4

Zinc cobalt oxide, ZnCo2O4, is a “normal” spinel oxide with Zn ion occupying the tetrahedral sites and Co ions occupying the octahedral sites. The Zn metal can form an alloy with Li, and in 5424

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

electrode thereby corroborating the reaction mechanisms of eqs 16 and 17. The observed long-term cycling stability can be attributed to the ability of cobalt to act as a good “matrix” in supporting the conversion reaction of the nontransition metal (Zn) and, also, relatively small unit cell volume variation during the alloying−dealloying reaction with Zn metal. The average discharge and charge voltages were ∼1.2 and ∼1.9 V, respectively. Studies by Qiu et al.1168 on hexagonal nanodisks of ZnCo2O4 prepared by the hydrothermal method, followed by heating in air at 400 °C, showed a reversible capacity of 750 mA·h g−1 after 40 cycles at a current of 80 mA g−1 in the range 0.01−3.0 V. Porous ZnCo2O4 1D nanowires were synthesized by Du et al.1169 by the microemulsion method followed by annealing at 500 and 700 °C in air. They reported cycling up to 20 cycles and obtained a reversible capacity of 1197 and 957 mA·h g−1, respectively, for the 500 and 700 °C synthesized samples at current of 100 mA g−1. Recently, the group of Chowdari1170 prepared ZnCo2O4 by the molten salt method at 280 °C in air using various zinc and cobalt salts as starting materials and studied their electrochemical performance. The ZnCo2O4-(II) prepared using ZnSO4 4H2O and Co(OH)2 showed the best Li cycling performance, showing a capacity of 974 mA·h g−1 at the end of the 40th cycle, when cycled at current of 60 mA g−1 in the voltage range 0.005−3.0 V vs Li. Additional recent studies on ZnCo2O4 are hierarchical three-dimensional nanowire arrays by Liu et al.1171 and electrospun porous nanotubes by Luo et al.1172 Another normal spinel oxide, ZnFe2O4, is cost-competitive and environmentally friendly in comparison to ZnCo2O4 as the oxide anode. The group of Chowdari1173 prepared submicrometer size ZnFe2O4 by the urea combustion method, followed by heating in air at 900 °C, and reported a stable capacity of 615 ± 10 mA·h g−1 (5.5 mol of Li per mole of ZnFe2O4) between 15 and 60 cycles, when cycled in the range 0.005−3.0 V vs Li at a current of 60 mA g−1 (0.1C). NuLi et al.1174 studied the Li cycling of bare and “Ag-doped” ZnFe2O4 films (300 nm thick) and reported a reversible capacity 556 and 700 mA·h g−1, respectively, at the current rate of 10 μA cm−2. Recently, GuO et al.1175 prepared hollow-spherical ZnFe2O4 by hydrothermal method followed by heating at 600 °C . The submicrometer size hollow spheres had a wall thickness of 100 nm and are composed of 10−20 nm size primary phase-pure ZnFe2O4 particles (Figure 55). The XRD pattern was indexed as the spinel structure. The Li cycling properties were studied in the voltage range 0.005−3.0 V vs Li at 65 mA g−1 (0.07C) up to 50 cycles. The discharge−charge profiles are shown in Figure 56. The first-discharge (Li insertion) profile is interpreted as due to the formation of Lix−ZnFe2O4 (x = 0.5 at ∼1.4 V and 2.0 at ∼0.9 V; Li intercalation), followed by crystal structure destruction (amorphization of the lattice), formation of SEI, and formation of Zn and Fe metal nanoparticles and, subsequently, the formation of the alloy (Li−Zn) in a matrix of Li2O. The HRTEM and SAED pattern data, shown in Figure 57a confirm the discharge reaction mechanism. The first-charge (Li extraction) profile is interpreted as due to the dealloying reaction of (Li−Zn) to give Zn metal and conversion reactions of Zn and Fe with Li2O eventually yielding ZnO and Fe2O3 after charging to 3.0 V. Again, the HRTEM and SAED pattern data of the charged electrode, shown in Figure 57b, prove the existence of the ZnO and Fe2O3, thereby confirming the charge reaction mechanism.

addition, it can participate in the conversion reaction along with cobalt as per the eqs 13−17. ZnCo2O4 + x Li+ + x e− → Lix(ZnCo2O4 )

(13)

Lix(ZnCo2O4 ) + (8 − x)Li+ + (8 − x)e− → Zn + 2Co + 4Li 2O

(14)

Zn + Li+ + e− ↔ ZnLi

(15)

Zn + 2Co + 3Li 2O ↔ ZnO + 2CoO + 6Li+ + 6e− (16) +

2CoO + 2/3Li 2O ↔ 2/3Co3O4 + 4/3Li + 4/3e



(17)

A reversible capacity corresponding to 8.3 mol of Li per mole of ZnCo2O4 is expected, assuming x ≤ 0.5 for Li intercalation into the spinel lattice (eq13), crystal structure destruction, followed by metal particle formation, and alloy formation with Zn (eqs 14 and 15). The conversion reactions occur during charging (Li extraction, eq 16). Part of the CoO can also form Co3O4 during the charging process (eq 17). The group of Chowdari1167 has realized near theoretical capacity (∼900 mA·h g−1) in nanosized ZnCo2O4 (20 nm) prepared by the urea combustion technique, which was stable up to at least 60 cycles at a current of 60 mA g−1 (0.07C) when cycled in the range 0.005−3.0 V (Figure 54). Good C-rate capability was also established, showing a stable capacity of ∼400 mA·h g−1 at 0.7C. Ex-situ HR-TEM and SAED (selected area electron diffraction) studies showed the formation of ZnO and Co3O4 in the charged

Figure 54. (a) Voltage versus capacity profiles of ZnCo2O4, for the 1st, 2nd, and 15th discharge and charge cycle in the voltage window 0.005−3.0 V versus Li at 60 mA g−1. Inset shows the Y versus voltage plots, where Y = differential capacity, dq/dV (×103), mA·h g−1 V−1. The numbers indicate cycle number. (b) Capacity versus cycle number plot for ZnCo2O4 in the voltage range 0.005−3.0 V at current rate, 60 mA g−1. Filled and open symbols correspond to discharge and charge capacities, respectively. Reproduced with permission from ref 1167. Copyright 2007 Wiley-VCH. 5425

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 55. The XRD pattern (a) and (b) morphology of the asprepared ZnFe2O4 hollow spheres. Reproduced with permission from ref 1175. Copyright 2010 Elsevier.

Figure 57. HRTEM and the SAED patterns of the electrode: (a) initially discharged to 0.005 V (selected XRD patterns of Li−Zn, Fe, and Zn in the JCPDS cards are shown for comparison) and (b) recharged to 3.0 V after several complete discharge−charge cycles. Reproduced with permission from ref 1175. Copyright 2010 Elsevier. Figure 56. The capacity−voltage profiles of ZnFe2O4 in the voltage range 0.005−3.0 V (the vertical axis in both insets is for the specific capacity, mA·h g −1. The numbers in inset 2 are for the various C values; 0.07C = 65 mA g −1 for the cell in inset 1. Reproduced with permission from ref 1175. Copyright 2010 Elsevier.

polymer pyrolysis method, namely, metal−polyacrylate decomposition at 600 °C in air. When the particles were cycled between 0.01 and 3 V vs Li at a current of 116 mA g−1, firstcycle discharge and charge capacities of 1419.6 and 957.7 mA·h g−1, respectively, were measured. After 50 cycles, the reversible capacity dropped to ∼800 mA·h g−1 (∼80% retention). The compound also showed good C-rate capability, ranging from 0.2C to 4C (1C = 928 mA g−1). For example, at 4C, the nanoZnFe2O4 showed a reversible capacity over 400 mA·h g−1. These data clearly show the advantages of realizing the proper favorable morphology in giving rise to high and stable capacities in the spinel oxides. Recently, the group of Chowdari1095 reported the Li cycling of solid solutions of nanocrystalline spinel phases, (Ni1−xZnx)Fe2O4 (0 ≤ x ≤ 1), prepared by the

From Figure 56, it is clear that a reversible capacity of over 900 mA·h g−1 is obtained between 5 and 50 cycles at 0.07C when cycled in the range 0.01−3.0 V vs Li. This value compares well with the theoretical value of 1006 mA·h g−1 (9 mol of Li per formula unit, assuming formation of ZnO and Fe2O3 during charge reaction by conversion reaction). At 0.7C, a reversible capacity of ∼500 mA·h g−1 is seen (Figure 56, inset). Ding et al.1176 recently reported the Li cycling behavior of nanostructured ZnFe2O4 (30−70 nm particles) prepared by the 5426

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

al.1185 on the synthesis of flake-shaped nano-ZnMn2O4 by the calcination of the single-source precursor, Zn−Mn citrate complex at 700 °C, and its anodic performance in terms of reversible capacity, C-rate, and cycling up to 200 cycles. Very recent additional studies on ZnMn 2 O 4 with different morphologies are nanoplates by Zhao et al.,1186 hollow microspheres by Zhou et al.1187 and Zhang et al.,1188 and nanoparticles by Courtel et al.1189 Clew-like hollow spheres of the normal spinel ZnV2O4 with vanadium in 3+ oxidation state were synthesized by the solvothermal method, and their Li cyclability was studied by Xiao et al.1190 When the particles were cycled at 50 mA g−1 in the voltage range 0.01−3.0 V, initial reversible capacity of 548 mA·h g−1 was observed. It exhibited almost a stable capacity of 524 mA·h g−1 over 50 cycles. Assuming a redox reaction of V2+ ↔ V3+, along with those of Zn, involving alloying−dealloying and conversion reaction, the theoretical capacity is 579 mA·h g−1, and the observed capacity is in fairly good agreement. Recently Zheng et al.1191 reported Li cycling studies on ZnV2O4/mesoporous carbon by a carbothermal reduction route, and they showed improved reversible capacity of 575 mA·h g−1 at the end of 200 cycles at a current density of 100 mA g−1.

sol−gel autocombustion method. The compounds showed a second discharge capacity above 1000 mA·h g−1 for all x when cycled at current of 50 mA g−1 in the range 0.005−3.0 V vs Li. The capacity fading between 10 and 50 cycles was found to be greater than 52% for x ≤ 0.4 and for x = 0.8. For x = 0.6 and x = 1, the respective values were 40% and 18%, and a capacity of 570 and 835 mA·h g−1 is retained after 50 cycles. Recent additional studies have been reported on MWCNT−ZnFe2O4 nanocomposites by Sui et al.,1177 porous ZnFe2O4/graphene nanosheets by Chen et al.,1178 ZnFe2O4 nano-octahedrons by Xing et al.,1179 and nanoparticles and nanospheres by Xu et al.1180 Cadmium can also form an alloy with Li similar to Zn, and under optimal conditions, Li3Cd is formed during Li cycling. Thus, CdFe2O4 with the normal spinel structure is worthy of investigation, even though Cd is toxic. Chowdari’s group1181 prepared and examined the Li cycling of submicrometer sized CdFe2O4. It showed a first-cycle reversible capacity of 870 mA·h g−1 at 0.07C rate when cycled between 0.005 and 3.0 V, but the capacity degraded slowly, and it retained a capacity of 680 mA·h g−1 after 50 cycles. Significantly, heat-treated electrode of CdFe2O4 (300 °C, 12 h, Ar) showed a significantly improved cycling performance under the above cycling conditions and a stable capacity of 810 ± 10 mA·h g−1 corresponding to 8.7 mol of Li per mole of CdFe2O4 (vs theoretical, 11.0 mol of Li) was maintained up to 60 cycles, with a Coulombic efficiency of 96−98%. Rate capability of heattreated CdFe2O4 was also good: reversible capacities of 650 and 450 mA·h g−1 at 0.5 and 1.4C (1C = 840 mA g−1), respectively, were observed. The reasons for the improved cycling performance upon heat treatment of the electrode are similar to those proposed for the Fe2O3 electrode and are discussed in section 4.1.3.1. CdFe2O4 showed an average discharge potential of 0.9 V, whereas the charge potential is 2.1 V, similar to that shown by ZnFe2O4 and ZnCo2O4. The underlying reaction mechanism, based on ex-situ HR-TEM and SAED data, is the combination of dealloying−alloying and conversion reactions of “Li−Cd−Fe−Li2O” composite. Another normal spinel oxide is ZnMn2O4, which possesses a tetragonal structure due to the presence of Jahn−Teller ion, Mn3+ in the lattice. Yang et al.1182 and Xiao et al.1183 recently prepared and studied the Li cycling properties of this cheap and nontoxic ZnMn2O4, both as nanoparticles (size 30−60 nm) and as flower-like superstructures. Galvanostatic cycling of the nano-ZnMn2O4 at 100 mA g−1 in the range 0.01−3.0 V showed a first-charge capacity of 766 mA·h g−1, which matches well with the theoretical value of 784 mA·h g−1 (7 mol of cyclable Li, assuming that Mn0 is oxidized only to Mn2+ ions during the conversion reaction). Almost stable cycling performance was observed between 10 and 50 cycles, with an average capacity fading of 0.2% per cycle, and a capacity of 569 mA·h g−1 was retained after 50 cycles. Under similar cycling conditions, flower-like ZnMn2O4 performed better, retaining a capacity of 626 mA·h g−1 after 50 cycles. Significantly, the average discharge and charge voltages were ∼0.5 and ∼1.2 V, respectively. These are much less than those encountered in ZnCo2O4. Additional recent references are Courtel et al.1184 on the synthesis of ZnMn2O4 by coprecipitation method and its Li cycling performance as a function of synthesis temperature, effect of binder, performance at 24 °C and at 60 °C, and fabrication and testing of the full Li ion cell with ZnMn2O4 as anode and Li[Ni1/2Mn3/2]O4 as the 5 V cathode and Deng et

5.2. Oxides of Tin Adopting the Spinel Structure

A good number of tin oxides with the inverse spinel structure, M2SnO4, M = Co, Mn, Mg, and Zn, are known. Here, the tetravalent Sn ions occupy the tetrahedral sites and the bivalent M ions occupy the octahedral sites. Since Sn can reversibly cycle 4.4 mol of Li via alloying−dealloying process, it will be of interest to examine whether additional capacity can be realized by the participation of Sn (and Li2O) by conversion reaction to form SnO and SnO2. Toward this end, preliminary studies were conducted by Belliard et al.479 on Zn2SnO4, and by Connor and Irvine1192 and the group of Tirado1193 on Co2SnO4. Large capacities, over and above those expected from the cycling of the alloy Li4.4Sn, were encountered upon charging to 1.5 V for Zn2SnO4 and to 3.0 or 4.5 V for Co2SnO4, after the first discharge. The complementary EXAFS and Sn Mossbauer data on Co2SnO4 indeed confirm that both alloying−dealloying process and conversion reactions of Sn and Co with Li2O are involved in the Li cycling.1192,1193 But, long-term cycling was not carried out. Wang et al.1194 reported results on the hydrothermally synthesized nano-HT-Co2SnO4 (size, 80−120 nm) and micrometer-sized particles (SS) prepared by high-temperature solid-state reaction. Upon cycling at 30 mA g−1 between 0.01 and 3.0 V vs Li, initial charge capacity of 1089 mA·h g−1 (12.2 mol of Li per mole) was noted for the HT-Co2SnO4, which matches well with the theoretical value, 12.4 mol of Li. However, continuous capacity fading occurred, and at the end of 50 cycles, the capacity value was 556 mA·h g −1 , corresponding to a drop of 49.7%. The performance of SSCo2SnO4 was found to be inferior to that of HT-Co2SnO4: an initial charge capacity of 900 mA·h g−1 dropped to 113 mA·h g−1 at the end of 50 cycles (a fading of 87.8%). Recently, Qi et al.1195 studied the Li cycling of 10.5 wt % (coating thickness 2−3 nm) and 25.2 wt % (5−10 nm thick) core−shell carbon−Co2SnO4 nanostructures prepared through a glucose hydrothermal method and subsequent carbonization at 500 °C. The phases showed reversible capacities of 474, 173, and 56 mA·h g−1 for thin, thick carbon-coated, and bare Co2SnO4, respectively, after 75 cycles, when cycled in the range 5427

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

of 0.01−2.5 V and at a current of 100 mA g−1. Improved capacity retention, about 60.4%, was noted for the 25.2 wt % carbon-coated sample in comparison to others. It is not clear whether the authors have taken the contribution of carbon in the capacity calculations. The above studies show that irrespective of carbon coating or otherwise, capacity-fading in Co2SnO4 is not significantly suppressed. In recent years, studies on the Li cycling of Zn2SnO4 have been reported on the phases prepared by solid state reaction as well as by hydrothermal method, either in pure form or as composites with carbon.1196−1206 It has a large theoretical reversible capacity of 14.4 mol of Li per mole of Zn2SnO4 (1231 mA·h g−1) as a result of contributions from the alloying− dealloying of Sn and Zn and the conversion reactions involving Sn and Zn and Li2O. When cycled at 60 or 100 mA g−1 in the voltage range 0.01−3.0 V, the reported initial reversible capacities ranged from 990 to 1045 mA·h g−1 depending on the particle size, morphology, with or without carbon etc. However, continuous capacity degradation was found in all cases upon cycling to 40 or 50 cycles, at the end of which the observed reversible capacities ranged from 400 to 660 mA·h g−1. Very recently, Cherian et al.1206 studied Zn2SnO4 nanowires (NW) on stainless steel substrates by chemical vapor deposition technique and nanoplates by hydrothermal method. They studied Li cycling properties in the voltage range 0.005−3 V and 0.005−1.5 V. For the first few cycles, Zn2SnO4 NW electrodes showed 100% capacity retention whereas Zn2SnO4 nanoplate composite electrodes showed drastic capacity fading. In the voltage range 0.005−3 V, the nanowire electrode retains a capacity of 660 mA·h g−1 after 50 cycles, and in the lower cutoff voltage of 1.5 V, a capacity of 390 mA·h g−1 is delivered after 50 cycles. The capacity fading behavior observed in Zn2SnO4 NW electrodes, after 10 cycles, can be attributed to the complete destruction of nanowires into nanoparticles and subsequent disconnection from the stainless steel substrate during further cycling. Very recently, Becker1200 and Sepelak et al.1205 reported studies on Zn2SnO4 nanoparticles by Sn-119 MAS NMR, Raman spectroscopy, Sn-119 Mossbauer spectroscopy, XPS, and TEM studies. Lei et al.1207 studied the Li cycling properties of nanoMn2SnO4 prepared by the thermal decomposition of MnSn(OH)6 and found a reversible capacity of ∼750 mA·h g−1 when cycled between 0.01 and 3.0 V, which degraded and finally stabilized to ∼150 mA·h g−1 after 80 cycles. Satya Kishore et al.1208 studied the Li cycling behavior of ramsdellite-type orthorhombic phases, LiMSnO4, M = Fe and In, at 0.2C rate in the voltage range 0.05−2.0 V. For M = Fe, an initial reversible capacity of 655 mA·h g−1 decreased to ∼100 mA·h g−1 at the end of 25 cycles. The performance of the compound with M = In was very poor under similar cycling conditions: The firstcharge capacity of 375 mA·h g−1 degraded to zero after five cycles. However, a stable capacity of ∼90 mA·h g−1 up to 25 cycles was observed when the lower cutoff voltage was increased to 0.75 V. It is clear from the above discussion that even though it is attractive to employ tin oxides to obtain very large reversible capacities by exploiting both alloying−dealloying and conversion reactions, by cycling in the range 0.005−3.0 V vs Li, due to the very large unit cell volume changes of the nanocomposites involved in the processes, drastic capacity fading occurs upon long-term cycling, despite the presence of good “matrix” ions like, Co.

5.3. Metal Oxysalts: Carbonates and Oxalates

The Li storage and cycling behavior of the mixed-metal carbonate nano-(Cd1/3Co1/3Zn1/3)CO3 (CCZC) prepared under ambient conditions by the coprecipitation method have been reported by the group of Chowdari.1209 Flower-like agglomerates of nanosize flakes or needles of nano-CCZC with length 100−200 nm and thickness 5−10 nm are obtained (Figure 58). The compound adopts the rhombohedral-

Figure 58. (a) SEM photograph of nano-(Cd1/3Co1/3Zn1/3)CO3 (CCZC). Flower-like agglomerates of nanosize flakes and needles are seen. Scale bar is 100 nm. (b) TEM image showing the needleshape morphology. Scale bar is 100 nm. Reproduced with permission from ref 1209. Copyright 2009 Royal Society of Chemistry.

hexagonal structure, similar to CdCO3. A reversible capacity of 680 ± 10 mA·h g−1, corresponding to 3.5 mol of cyclable Li per mole of the CCZC (theoretical 3.33 mol of Li), stable in the range 8−60 cycles, was observed when cycled at 0.09C (1C = 680 mA g−1) in the range 0.005− 3.0 V vs Li (Figure 59). The nano-CCZC also showed stable and reversible capacities at various C-rates up to 170 cycles (Figure 59d). On the basis of galvanostatic cycling, cyclic voltammetry, and ex-situ XRD, TEM, and SAED studies (Figure 60), the proposed reaction mechanism involved the reduction of the CCZC by Li to 5428

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 59. The voltage vs capacity profiles of nano-(Cd1/3Co1/3Zn1/3)CO3 (CCZC) in the voltage window 0.005−3.0 V vs Li at the current of 60 mA g−1 (0.09C) at room temperature. (a) The first cycle. (b) The profiles during the 2−25 cycles. Only selected cycles are shown for clarity. The numbers refer to the cycle number. (c) The capacity vs cycle number plots in the voltage range 0.005−3.0 V and 0.005−1.5 V at the current of 60 mA g−1 (0.09C) (1C = 680 mA g−1). The filled and open symbols represent discharge and charge capacities, respectively. (d) Capacity vs cycle number plot in the voltage range 0.005−3.0 V at various current (C) rates of CCZC. For the sake of clarity, data for selected C-rates are shown (1C = 680 mA g−1). Filled and open symbols correspond to discharge and charge capacities, respectively. Reproduced with permission from ref 1209. Copyright 2009 Royal Society of Chemistry.

nanometal (M = Cd, Co, and Zn) particles embedded in amorphous Li2CO3 during the first discharge, and this is followed by the formation of alloys (Li−Zn and Li−Cd). Upon charging of the electrode, the dealloying reaction and metal carbonate (MCO3) formation occurs, thereby contributing to the reversible capacity. This shows that the carbonate is as good as the oxide, fluoride, or oxyfluoride ion in enabling the reversible “conversion” and alloying−dealloying reactions involving both transition and nontransition metal ions. It is worth investigating other pure and mixed carbonates, prepared by precipitation as well as by hydrothermal methods, for their Li cyclability. Recently group of Tirado1210,1211 reported Li cycling studies on bare and Co-substituted MnCO3 prepared by a reverse micelle method. The group of Tirado1212 examined the Li cycling behavior of tin oxalate, Sn(II)C2O4, in the voltage range 0.1−1.5 V vs Li at a current rate of 0.5C. The initial reversible capacity of ∼650 mA·h g−1 continuously degraded to ∼150 mA·h g−1 at the end of 50 cycles. The authors proposed the formation of Sn metal and also the reduction of Li2C2O4 to Li2O (and formation of CO) during the first-discharge reaction. The charging reaction will involve dealloying of Li4.4Sn and, possibly, SnO formation by the conversion reaction. But, this needs to be confirmed by complementary studies as well as studies on other metal oxalates.

Figure 60. An ex-situ TEM study of nano-(Cd1/3Co1/3Zn1/3)CO3 (CCZC) charged to 3.0 V after 30 cycles. (a) A high-resolution lattice image. The nanocrystalline regions dispersed in amorphous regions are clearly seen. The derived d-values (interplanar spacings) corresponding to the crystalline regions are shown. Scale bar is 5 nm. (b) The corresponding SAED pattern. The Miller indices corresponding to the diffuse spots are assigned to the metal carbonates (M′CO3, M′ = Cd, Co, or Zn). Scale bar is 5 nm. Reproduced with permission from ref 1209. Copyright 2009 Royal Society of Chemistry.

6. LI ION BATTERIES (CELLS) WITH OXIDE ANODES 6.1. LIBs with Li4Ti5O12 (LTO) as Anode

As mentioned in the Introduction, many prospective futuregeneration cathodes have been developed as an alternative to the 4 V cathode, LiCoO2, for use in the LIBs. Specifically, the 3.5 and 4 V cathodes are Li(Ni1/3Co1/3Mn1/3)O2, modified LiMn2O4, LiFePO4, LiVPO4F, and Li3V2P3O12, and the 5 V cathode is Li[Ni0.5Mn1.5]O4. As mentioned in section 2.2.1, 5429

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

6.1.1. LIBs with 4 V cathodes. 6.1.1.1. LiMn2O4. LiMn2O4 adopts a cubic spinel structure, and it is a 4.0 V cathode. It is cheap and environmentally acceptable, and modified compounds (e.g., doping with Li and Al at the Mn site) have been employed as cathodes in LIBs with Li4Ti5O12 as the anode. The group of Ozhuku1216,1217 studied the cell LTO/(Li,Al)-doped LiMn2O4 (LAMO), which has a cell voltage of 2.5 V and shows a capacity of 67 mA·h g−1 (260 mA·h cm−3) and energy density of 165 W·h kg−1 or 645 W·h dm −3. The group of Amine1218,1219 studied the cycling, up to 160 cycles, of the cell LTO/LiMn2O4 and reported a reversible capacity 90−100 mA·h g−1 when cycled at 55 °C at 1C rate (Figure 62). They

Li4Ti5O12 (LTO) is a viable 1.5 V anode material, due to several advantages over the graphite anode presently used in LIBs. Hence, it is natural to expect that LIBs with the above combination of cathode and anodes will be investigated. Indeed, as early as 1994, Thackeray’s group1213 fabricated and tested the Li ion cell (battery) with the cathode LiMn2O4 and the anode Li4Ti5O12. This was followed by Peramunage and Abraham417 in 1998, on the same system, who employed LiPF6 dissolved in polyacrylonitrile (PAN)/EC/propylene carbonate (PC) as the solid polymer electrolyte. Later, in 1999, Thackeray’s group1214 also tested the cells LTO/Li(Ni0.8Co0.2)O2 and LTO/LiCoO2 and showed their viability for high-power applications. But, these were not pursued seriously due to the lower operating voltage. However, in recent years, interest in LIBs with Li4Ti5O12 as anode has been growing due to the safety and cost considerations, especially for use in HEV and in load-shedding. Stux and Swider-Lyons1215 reported the performance of the cells LTO/LiCoO 2 and LTO/LiCoO2+Li2RuO3. The cathode-limited cells showed greater discharge capacities when Li2RuO3-containing cathodes were cycled up to 2C rates up to 50 cycles. Optimal LIB design requires that the maximum deliverable capacity of the positive electrode (cathode) of a given mass (QPE) must be matched with that of the maximum deliverable capacity of the negative electrode (anode) of a given mass (QNE) to obtain the maximum deliverable specific capacity of the cell (Qcell), and the following relationship holds good: 1/ Qcell = 1/QPE + 1/QNE. Energy density of the cell (Wcell) can be calculated from the values of Qcell and the average cell voltage, Ecell. In conventional LIBs with graphite anode, the optimum cell design is positive-limited, that is, QPE < QNE. This is to avoid deposition of Li metal during high current rates of charge. The group of Ohzuku1216,1217 proposed that conceptually, the 12 V lead−acid batteries for automobile and stationary applications can be replaced by LIBs with Li4Ti5O12 as anode and a suitable Li mixed oxide as cathode, and gravimetric energy densities as high as 250 W·h kg−1 (and volumetric energy densities of 970 W·h dm−3) can be realized (Figure 61).

Figure 62. Cycling performance of the Li4Ti5O12/LiMn2O4 coin cell cycled under 1C rate at 50 °C in the presence of 1.2 M LiPF6 in an EC/EMC 3:7 wt % electrolyte. The inset illustrates the discharge profile curves for selected cycles. Reproduced with permission from ref 1219. Copyright 2007 Electrochemical Society.

also reported thermal degradation studies of the charged LTO and LiMn2O4 and observed less heat generation compared with graphite anode. Du Pasquier et al.1220 studied the high-power capabilities of nano-LTO/LiMn2O4 batteries, using specially prepared 20 nm LTO. Cycling data up to 1000 cycles at 25 and at 55 °C, effect of electrode-capacity matching, current rates from 1C to 80C, and fast-charge capability were reported. Ionica-Bousquet et al.1221 studied cycling of LTO (cast)/ LiMn2O4 (powder) using polyfluorinated boron cluster based salts as the electrolytes and compared their performance with the cells containing the standard salt (LiPF6) electrolyte. They reported a reversible capacity of 85−110 mA·h g−1 at 0.05C rate using Li2B12F8H4, Li2B12F12, and LiPF6 electrolytes, and at the end of 15 cycles, all the cells with different electrolytes showed a capacity of 85 mA·h g−1. Recently, Amine’s group1222 reported the performance characteristics of nanostructured LTO/LiMn2O4 cells. The micrometer size (∼0.5−2 μm) secondary particles composed of nanosize (≤10 nm) primary particles (MSNP) LTO were combined with spherical particles of LiMn2O4 (LMO) as the anode and cathode, respectively. Near 100% capacity retention was noticed up to 1000 cycles when cycled at 5C rate at 55 °C. The cells also showed unprecedented power capability, low (∼10 Ω cm2) area specific impedance, good calendar life, low heat generation during high-rate operation, excellent lowtemperature (−30 °C) performance, and unmatched abuse tolerance. The latter constitute the over charge, overheating, cold-cranking and nail-penetration tests. The authors con-

Figure 61. Gravimetric and volumetric energy densities in W·h kg−1 and W·h dm−3, respectively, calculated for possible lead-free accumulators based on Li4Ti5O12 (LTO) negative electrode. In calculating the values, weight and volume calculated from structural data together with observed reversible voltage are used. The values for lead−acid batteries are also shown for comparison. For lead−acid batteries, Pb, PbO2, and 5 M H2SO4 are considered in calculating energy density. Reproduced with permission from ref 1217. Copyright 2007 Elsevier. 5430

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

cluded that the above LIB system is suitable for HEV and energy-storage alternative for electric grids. Additional recent work on the LIB (full cell) with LTO as anode and Li[Li0.1Al0.1Mn1.8]O4 as cathode is by the group of Ohzuku1223 on the performance of the 2.5 V cells at −10, 25, and 50 °C. 6.1.1.2. Layered Oxides, LiCoO2, Li(Ni0.8Co0.2)O2, Li(Co1/2Ni1/2)O2, and Li(Co1/3Ni1/3Mn1/3)O2. As mentioned earlier, the group of Thackeray1214 reported long-term cycling studies of the LIBs with LiCoO2 and Li(Ni0.8Co0.2)O2 as cathodes and LTO as anode and discussed pulse power capability and specific impedance data. LIBs with the 4 V cathodes, Li(Co1/2Ni1/2)O21224 and Li(Co1/3Ni1/3Mn1/3)O2,1217 are also reported by the group of Ohzuku. Data on the cell LTO/Li(Co1/2Ni1/2)O2 cycled in the voltage range 1.0−2.7 V at a current rate of 0.17 mA cm−2 showed good performance. Studies of impedance spectra as a function of cycle number and accelerated cycling tests up to 50 000 cycles were also carried out.1224 The cell, LTO/Li(Ni1/3Co1/3Mn1/3)O2 showed a capacity ∼85 mA·h g−1 (340 mA·h cm−3) and an energy density of 215 W·h kg−1 or 970 W·h dm−3 with an average voltage value of 2.5 V.1217 Cycling studies on the cell LTO/Li1+x(Ni1/3Co1/3Mn1/3)1−xO2 are also reported by the group of Amine,1225 who found an excellent cycling performance at high C-rates and at both room and elevated temperature (55 °C) and reported the impedance behavior. 6.1.2. LIBs with 3.5 V cathode, LiFePO4. As mentioned earlier, LiFePO4 is a 3.5 V cathode and adopts an orthorhombic olivine (mineral with the chemical formula (Mg,Fe)2SiO4) structure. The group of Scrosati1226 studied the cell LTO/ LiFePO4 using gel-polymer electrolyte and obtained a capacity of 120 mA·h g−1 at the end of 80 cycles at 0.2C when it was cycled in the range 1−2.5 V. The average voltage of the cell is ∼1.75 V. The same group1227 also suggested that the cell LTO/ LiFePO4 employing a Li salt in an ionic liquid, namely, lithium bis(trifluoromethane) sulfonamide, LiTFSI in Py24TFSI, is also viable. Preliminary studies on LTO/LiFePO4 showed a cell capacity of 82 mA·h g−1 (290 mA·h cm−3) and energy density of 165 W·h kg−1 or 585 W·h dm−3 by the group of Ohzuku.1217 Sun et al.1228 reported the cycling of the cell LTO/LiFePO4 (nanocrystalline polyacene (PAS)-coated), up to 1000 cycles. The group of Doeff1229 reported on nanoscale LTO/ LiFePO4 system and obtained capacities of 155 and 122 mA·h g−1 at 0.1C and 5C rates. When cycled between 0.5 and 3.0 V, the average cell voltage is ∼1.8 V. The size of the LTO particles ranged from 50 to 200 nm whereas the size of the LiFePO4 particles ranged from 50 to 100 nm. The latter were coated uniformly with a 2−4 nm thick carbon. The performance was found to be stable over 200 cycles even at 5C and 10C rates (Figure 63). The group of Dahn1230 studied the effect of the redox additives, 2,5-di-tert-butyl-1,4-dimethoxybenzene (DDB) and 4-tert-butyl-1,2-dimethoxybenzene (TDB) for overcharge protection of the cell, LTO/LiFePO4. They found that DDB can support over 200 shuttle-protected overcharge cycles of 100% cell capacity, whereas TDB can only support between 3 and 15 overcharge cycles. A safe and fast-charging LIB with long shelf life for power applications has been described recently by Zaghib et al.,10 who used LTO (particle size, 150 nm) as anode and 2 wt % carboncoated nano-LiFePO4 (particle size, 25 nm) as cathode. Both coin-type cells and those of size 18650 (18 mm diameter and 65 mm in length) have been fabricated and tested up to 30 000 charge−discharge cycles at high C-rates, namely, charging at

Figure 63. Discharge/charge curves for a Li4Ti5O12/LiFePO4 full cell at several different current rates. For 5C and 10C rates, the voltage limits are 0.3 and 3.2 V. Reproduced with permission from ref 1229. Copyright 2009 Electrochemical Society.

10C (charging time 6 min) and at 15C (4 min) and discharge at 5C (12 min; 1C = 150 mA g−1). The cells were cathodecapacity-limited and yielded a maximum capacity of ∼150 mA·h g−1. The fabricated 18650 cell showed a capacity of 910 mA·h in the first few cycles. The voltage vs capacity profiles are shown in Figure 64. The cells with a voltage of 1.9 V were

Figure 64. Voltage−capacity cycles of the LiFePO4/EC-DEC-1 M LiPF6/Li4Ti5O12. Cycle charge rate is 15C (4 min) and the discharge rate is 5C (12 min). The spikes are due to computer buffer saturation and the creation of new files because of the great number of data. Reproduced with permission from ref 10. Copyright 2011 Elsevier.

cycled between 1.2 and 2.5 V at ambient temperature. The cells retained full capacity after 20 000 cycles performed at charge rate of 10C and discharge rate of 5C, and retained 95% of the capacity after 30 000 cycles at charge rate 15C and discharge rate 5C, both at 100% DOD (depth of discharge) and 100% SOC (state of charge) (Figure 64). The authors stated that this breakthrough performance of LIB (LTO−LiFePO4 combination) is attractive for use in EV/PHEV and energy storage for wind and solar energy. The group of Aurbach1231 fabricated and tested the LIB full cell with LTO as anode and LiMnPO4 as the cathode and reported that it has an operating voltage of 2.4 V and expected energy density greater than 100 W·h kg−1. They suggested that this cell can be a suitable battery system for load-leveling applications. LiMnPO4 is isostructural to LiFePO4 and performs as the 4.2 V cathode, in contrast to LiFePO4, which 5431

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 65. Charge−discharge profiles for (a) Li4Ti5O12, and (b) C−LiMnPO4 composite electrodes at potential range of 1.1−2.4 and 2.7−4.4 V (as indicated), respectively, in EC-DMC 1:2/LiPF6 1.5 M solution (coin-type cell, T = 30 °C) at C/20 rate. Cycling protocol was constant current− constant voltage for LiMnPO4 providing potentiostatic steps at 4.4 V during 5 h for C/20 rate and constant current for Li4Ti5O12 electrodes. (c) Charge−discharge profiles of Li4Ti5O12/LiMnPO4 coin cell at C/10 rate (T = 30 °C). (d) Voltage−capacity profiles of the Li4Ti5O12/LiMnPO4 coin cells measured galvanostatically at various discharge rates at 30 °C. Reproduced with permission from ref 1231. Copyright 2011 Electrochemical Society.

discharge voltage of 2.6 V and, at low C-rates (≤1C), performed well over 400 cycles with ≤8% capacity fade when cycled between 2.0 and 3 V. At a 10C discharge rate, the cell delivered 80% of the rated capacity (relative to 0.13C rate).1234 Studies on the cell LTO/Na3V2P2O8F3 employing the electrolytes of salts LiPF6 and mixed (NaPF6 + LiPF6) showed an average voltage of 2.4 V and encouraging cycling performance, up to 70 cycles with low capacity fading, when cycled between 1.5 and 3.2 V at 0.2C rate.1235 6.1.4. LIBs with 5.0 V Cathode, Li(Ni0.5Mn1.5)O4. As mentioned earlier, Li(Ni0.5Mn1.5)O4 (LNMO) with the spinel structure is extensively studied as a 4.7 V cathode vs Li metal. The group of Ozhuku1217,1236,1237 fabricated and studied the cell LTO/LNMO. They reported an average flat operating voltage of 3.2 V and measured a discharge capacity in the range 5.4−4.8 mA·h (for mass balanced cells) at current rates of 12.7−509 mA g−1 when the cell was cycled in the range 3.5−0 V. Accelerated cycle tests up to 1100 cycles showed 83% of the initial capacity was delivered. Xiang et al.1238,1239 studied the cycling performance, choice of electrolyte (LiPF6 and dimethyl methyl phosphonate), electrode-mass matching, and overcharge tolerances of the cell LTO/LNMO. The group of Amine1240 studied the effect of cell design and cathode and anode electrode mass-balance on long-term cycling performance of the LIB LTO/LNMO (Figure 68b). They showed 86% capacity retention, when the cell was cycled at C-rates from

is a 3.5 V cathode. Figure 65a,b shows the charge−discharge profiles of LTO and 18 wt % carbon-coated LiMnPO4 at 0.05C rate. Figure 65c shows the first-cycle charge−discharge profiles of the LIB (full cell) at 0.1C rate, and Figure 65d shows the discharge profiles at various C-rates. As can be seen, at 0.05C and 0.2C, capacities of 130 and 120 mA·h g−1, respectively, were observed. Studies showed that the capacity at 0.5C remained almost constant up to 300 cycles. The authors also found that no appreciable changes occur in the thermal stability of the electrodes in their pristine and in the charged state in contact with the electrolyte. Very recently, the group of Wakihara1232 studied full cell (laminate type) LTO/LiFePO4 using a solid polymer electrolyte (SPE). They showed a reversible capacity of ∼135 mA·h g−1 at the end of the 50th cycle when cycled at 0.1C rate at 50 °C with an discharge voltage of 1.83 V (Figure 66). SEM crosssection photograph (Figure 67) showed, after 50 cycles, stable electrolyte and electrode interface and no corrosion of electrolyte with current collectors and the epoxy resin. Recent additional studies on carbon-coated LTO/LiFePO4 were reported by Zaghib et al.1233 6.1.3. LIBs with 4 V Cathodes, LiVPO 4 F and Na3V2P2O8F3. The group of Barker1234,1235 developed the vanadium phosphate fluorides, LiVPO4F and Na3V2P2O8F3, as 4 V cathodes and examined the performance of the LIBs with LTO as the anode. The cell LTO/LiVPO4F gave an average 5432

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 68. Rate capabilities at room temperature of (a) LNMO 11.2 mg/Li and LTO 7.6 mg/Li half-cells and (b) LNMO 11.2 mg/LTO 7.6 mg full cells. Reproduced with permission from ref 1240. Copyright 2009 Electrochemical Society.

Figure 66. Charge/discharge curves for cycling at 0.1C (50 °C) (a) and cycle performance of the cell up to the 50th cycle at 50 °C (b). Reproduced with permission from ref 1232. Copyright 2012 Springer.

Figure 69. Capacity retention of LNMO/LTO cells cycled at room temperature. The inset is the absolute capacity cycling performance. (N/P ratio represents the mass of negative and positive electrode material.) Reproduced with permission from ref 1240. Copyright 2009 Electrochemical Society.

cycles at current rates of 0.5C and 1C (Figure 71). The weight ratio of anode to cathode was 1.36 in the laminated-type cell. Figure 67. Photograph of the cross-section of the present cell (of Figure 66) after 50th cycle. Reproduced with permission from ref 1232. Copyright 2012 Springer.

6.2. LIBs with TiO2-B as Anode with Cathodes, LiFePO4 and Li(Ni0.5Mn1.5)O4

The groups of Bruce and of Scrosati158,1242 studied the cycling performance of cells TiO2-B (nanowire) as anode and LiFePO4 and Li(Ni0.5Mn1.5)O4 as cathodes using a gel-polymer electrolyte (GPE), consisting of LiPF6 dissolved in ethylene carbonate (EC) and propylene carbonate (PC) suitably gelled in poly(vinylidene fluoride) (PVDF) matrix. As described in section 2.1.1.3, TiO2(B) is a 1.5 V anode material with a theoretical capacity of 330 mA·h g−1. The nanowires of TiO2(B) had a diameter in the range 20−40 nm and up to several micrometers in length. The cell TiO2(B)/LiFePO4 showed an average voltage of 2.0 V and capacity of 225

0.5C to 10C. Tests up to 1000 cycles, between 2.0 and 3.5 V at room temperature and at 55 °C, and area specific impedance showed very good performance (Figure 69). Recently, the group of Scrosati1241 reported the rate capability studies of the cells LTO/Li(Ni0.45Co0.1Mn1.45)O4 at current rates of 0.2C to 10C (Figure 70). They obtained a reversible capacity of 124 mA·h g−1 at a current rate of 1C when cycled up to 500 cycles and also reported the operating temperature range extending from −20 to 55 °C up to 100 5433

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Figure 70. Performance of the LTO/LNMO battery at various rates. (a) Shows the voltage profiles of the charge (lithium transfer from LNMO to LTO)−discharge (back transfer of lithium from LTO to LNMO) cycles of the battery run at various rates. As expected from their respective voltages, the combination of LTO with LNMO gives rise to a battery operating in the 3 V range. (b) Reports the capacity delivery at increasing rate regimes demonstrating that at rates as high as 10C still 83.5% of the initial capacity is provided by the battery. (c) Shows that the battery may sustain a life extending to 500 cycles with only 14.6% capacity loss. Room temperature. Reproduced with permission from ref 1241. Copyright 2011 Nature Publishing Group.

Figure 71. Performance of the LTO/LNMO battery at various temperatures: (a) voltage profiles of the charge−discharge cycles of the battery in a temperature range extending from −20 °C to +55 °C demonstrating excellent voltage and capacity retention; (b) capacity vs number of cycles, showing that even at these extreme temperature limits, the battery cycles with a very stable capacity delivery and at comparatively high rates. Reproduced with permission from ref 1241. Copyright 2011 Nature Publishing Group.

mA·h g−1 when cycled at 0.2C in the range 1−2.5 V. The value slowly degraded to ∼190 mA·h g−1 at the end of 100 cycles. Good C-rate capacity was observed, in comparison to the cell Li4Ti5O12/LiFePO4. For example, at 6.5C rate, 55% of the capacity at 0.2C is observed. The performance of the cell TiO2(B)/Li(Ni1/2Mn3/2)O4 (LNMO) with a gel polymer electrolyte (GPE) is shown in Figure 72.158,1242 The voltage vs capacity profiles showed an average voltage of 3.0 V when cycled between 2 and 3.5 V at 0.2C. The capacity vs cycle number plot (Figure 72b) shows a reversible capacity of ∼220 mA·h g−1, which slowly degrades to ∼200 mA·h g−1 after 100 cycles. As can be seen in Figure 72c, the cell shows very good C-rate capability. Recently, Cao et al.1243 reported the cycling studies on the cell with TiO2−C (carbon) composite in the form of mesoporous nanospheres as the anode and with LiFePO4− carbon as the cathode. They observed a reversible capacity of ∼105 mA·h g−1 at current rate of 1.5C when cycled in the range 1−3.0 V. The average cell voltage is ∼1.6 V and the cell exhibited good C-rate capability up to 7.5C. Ding et al.1244 reported LIBs with VO2 as the anode and cathodes like LiCoO2, LiFePO4, and Li(Ni0.5Mn1.5)O4. The average cell voltages were 1.5, 1.0, and 2.0 V, respectively. When cycled between 0 and 1.5 V for the cell with LiFePO4 cathode and between 0 and 2.5 V for the LiCoO2 cathode at

0.5C and 1C rates, good cycling stability was observed up to 50 cycles, with the observed capacities ranging from 125 to 140 mA·h g−1. Recently, Brutti et al.1245 reported a LIB (full cell) with anatase TiO2 as the anode and Li(Ni1/2Mn3/2)O4 as the cathode to give an operating voltage of ∼3 V. To mitigate the irreversible capacity loss and electrolyte decomposition, the authors coated the cathode with 1.5 wt % ZnO and the nanoTiO2 (crystallite size, 6 nm) was preloaded (mixed) with 2 wt % nano-Li metal particles. The latter are commercially available. Results showed a very good cycling performance. The mean charge and discharge voltages were 2.93 and 2.78 V, respectively. The energy density reached a stable value of ∼125 W·h kg−1 and the η (Coulombic efficiency) >90% after 30 cycles. On the other hand, the performance of LIB with bare cathode and bare anode was not satisfactory. Recent additional work on TiO2 nanotube anode with LiNi0.5Mn1.5O4 cathode was reported by Xiong et al.291 Han and Goodenough1246 described the fabrication and testing of a LIB with carbon-coated TiNb2O7 as the 1.5 V anode and Li(Ni1/2Mn3/2)O4 as the cathode. The cell showed an operating voltage of 3 V. Test results showed that the anodecapacity-limited cells perform better than the cathode-capacitylimited cells. For example, an anode-specific capacity of ∼200 5434

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

electrode materials. They also found that when SnO was used as the anode, the cell with the LiCoO2 cathode showed only ∼35% of the initial capacity after 100 cycles. The group of Scrosati1247 reported cycling studies of the cell, with SnO2 as the anode and the 4 V cathode Li(Ni0.8Co0.2)O2, using gel-type electrolyte, that is, LiClO4 dissolved in EC− DMC in a poly(acrylonitrile) polymer matrix. They showed a reversible capacity of 250 mA·h g−1 at a current rate of 0.25 mA cm−2 in the voltage range 3.7−1.5 V. It is worthwhile to investigate the LIBs with other tin oxides as anodes. Very recently, Wong et al.1248 studied the cycling performance of the LIB full cell, comprising amorphous hierarchical porous GeOx as anode and Li(Ni1/3Co1/3Mn1/3)O2 as the cathode. The GeOx (x ≈ 0.67), consisting of primary nanoparticles of size 3.5 ± 1.0 nm in the form of ∼50 nmsized aggregates, was prepared by the reduction of GeO2 in aqueous solution using NaBH4 as the reducing agent. The anode GeOx was prelithiated, meaning that it was coated with air-stable Li metal powder (commercially available) to compensate for the Li loss due to SEI formation etc., during the initial charge−discharge cycles. Results showed an average voltage of ∼3.4 V for the cell. The cathode-limited initial discharge capacity was 164 mA·h g−1 in the constant current/ constant voltage (CCCV) mode at a current of 80 mA g−1 (C/ 20 rate) between 2.5 and 4.2 V with an initial η of 85% (Figure 73a). Under similar cycling conditions and at 0.5C rate, a

Figure 73. Cell performance of the initial GeOx anode. (a) Initial profiles of the Li-compensated GeOx/NCM full cell in comparison with those of the Li metal/NCM half-cell. (b) Reversible battery discharge capacity of NCM in the full cell. Reproduced with permission from ref 1248. Copyright 2011 American Chemical Society. Figure 72. (a) Charge/discharge voltage profile for the TiO2(B)− GPE−Li[Ni0.5Mn1.5]O4 battery at room temperature and at a rate of C/5. (b) Variation of charge and discharge capacities versus cycle number for the TiO2(B)−GPE−LiNi0.5Mn1.5O4 battery cycled between 2 and 3.5 V at room temperature and at a rate of C/2. (c) Variation of discharge capacity as a function of rate, cycled between 2 and 3.5 V, expressed in terms of percentage of the maximum capacity obtained at low rate for the TiO2(B)−GPE−LiNi0.5Mn1.5O4 battery at room temperature. Reproduced with permission from ref 1242. Copyright 2006 Wiley-VCH.

capacity of 144 mA·h g−1 was measured, with an average loss of 0.028% per cycle over 200 cycles (Figure 73b). Recent additional studies on TiO2 nanoparticles with Li(Ni1/3 Co1/3 Mn1/3)O4 cathode were reported by Moretti et al.271 6.4. LIBs with CoO and Co3O4 as Anodes with Cathodes, LiCoO2 and LiMn2O4

The group of Tarascon702,703,705 reported more than 10 years ago the fabrication and testing of LIBs with CoO and Co3O4 as anodes and with LiMn2O4 and LiCoO2 as the cathode. They showed an average voltage of 2 V and a stable and reversible specific energy density of 120 W·h kg−1 during extended cycling at ambient and elevated temperatures. They also found a capacity of ∼100−120 mA·h when the cells were cycled at a rate of 0.1C or 0.2C in the voltage range 0.9−4.1 V at temperatures of 25 and 55 °C703,705 (Figure 74). Cycling studies of the cells CoO/LiMn2O4 and CoO/Li1.33Mn2O4 showed a reversible capacity of ∼80 and ∼110 mA·h g−1, respectively, at 0.2C rate and in the cycling range 1−4.0 V (Figure 75). Thus, it will be of interest to study LIBs with other metal oxide anodes, especially those containing cheaper and environmentally friendly transition metals.

mA·h g−1 at 0.1C rate was maintained up to 50 cycles with a η > 95%. 6.3. LIBs with Amorphous Tin Composite Oxide, SnO2 and GeOx, with Cathodes LiCoO2, Li(Ni0.8Co0.2)O2, and Li(Ni1/3Co1/3Mn1/3)O2

As early as 1997, Idota et al.633 reported cycling studies of the LIB with the amorphous tin composite oxide (ATCO; Sn1.0B0.56P0.40 Al0.42O3.6) anode and with LiCoO2 cathode. They showed that 90% of the initial reversible capacity was retained after 100 cycles, when the cell was cycled in the range 2.8−4.1 V using LiPF6 (EC/DEC) as the electrolyte, at a current rate of 0.5C, with an optimized mass ratio of the 5435

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

6.5. LIB with Fe2O3 as the Anode and LiFePO4 as the Cathode

Scrosati and co-workers1249 recently reported the LIB (full cell) with α-Fe2O3 as the anode and LiFePO4 as the cathode to give a working voltage of ∼2 V. To compensate for the irreversible capacity loss in the first-cycle, the authors prelithiated the anode (to yield Li2Fe2O3) and predelithiated (charged) the cathode (to yield □FePO4, □ = vacancy) by using an additional electrode (counter) consisting of Li metal. In addition, the maximum anode capacity limit was adjusted to be 300 mA·h g−1 (theoretical capacity is 1008 mA·h g−1). The resulting LIB, namely, Li2Fe2O3/□FePO4 was cycled between 1.5 and 3.3 V at various current rates, 0.1C to 1C (1C = 170 mA g−1). The measured cathode-limited capacities ranged from 120 to 150 mA·h g−1, depending on the C-rate, over 200 discharge− charge cycles. The practical energy density was estimated to be ∼100 W·h kg−1. 6.6. Mitigating the ICL Associated with the Oxide Anodes

As discussed in earlier sections, oxide anodes with a working voltage below 1 V vs Li suffer from the major disadvantage of large ICL (irreversible capacity loss) during the first or a few initial discharge−charge cycles. The main reason for this is wellunderstood, as due to the formation of SEI (and polymeric geltype layer, in addition to the SEI, in some cases) on the anode particles, and this consumes Li ions (and electrons). The latter has to come from the cathode material, despite the fact that the SEI is highly beneficial for the long-term cycling of the LIBs. Additional Li consumption occurs in the case of oxide anodes that function on the basis of Li alloying−dealloying reaction mechanism. This is due to the necessity of reducing the oxide to the metal nanoparticles before the alloy formation and depends on the oxidation state of the metal ion, like Sn2+, Sn4+, Zn2+, Sb3+ etc. Several approaches have been tried, over the years, to mitigate the ICL. With regard to cathode, use of excess cathode material or prelithiated cathode, like Li1.33Mn2O4, or use of Lirich cathodes like Li(LixNiyMnzCop)O2 or its composite with Li2MnO3/LiMn2O4 have been suggested or implemented. From Figure 75 (inset), it is clear that the use of Li1.33Mn2O4 as the Li reservoir to compensate the ICL during the first charge−discharge cycle of the full cell CoO/Li1.33Mn2O4 has a large beneficial effect on the reversible capacity and cyclability, in comparison to bare cathode, LiMn2O4 at least up to 14 cycles. With regard to anode, use of prelithiated anode, like, LiGeOx by Wang et al.1248 and Li2Fe2O3 by Scrosati and coworkers1249 has been successfully tried to mitigate the ICL. Other methods suggested/being tried are admixing the anode (e.g., SnO, Co 3 O 4 ) with Li-yielding compounds like (Li2.6Co0.4)N, which gives rise to prelithiated anode, formation of artificial SEI (using, for example, carboxymethyl celluslose), carbon-coating, and coating the oxide anodes with Li4Ti5O12 (1.5 V anode and zero-strain material).184 The group of Tarascon1250 recently examined “sacrificial salts”, the Li salts of azide, oxocarbons, dicarboxylates, and hydrazides, as a means to compensate the ICL. The anions of these salts lose electrons during the first charge and convert to N2, CO, or CO2 gases within an acceptable potential range of 3−4.5 V vs Li. These possess specific capacities ranging from 430 to 567 mA·h g−1. Specifically, Li3N, Li squirate (Li2C4O4), Li oxalate (Li2 C 2 O 4 ), Li ketomelonate (Li 2 C 3 O 5 ), Li diketosuccinate (Li2C4O6), and Li poly(oxalylhydrazide)

Figure 74. Capacity retention at room temperature for (a) a plastic CoO/LiCoO2 Li ion cells (Mr = 6.4) and (b) a plastic Co3O4/LiCoO2 Li ion cell assembled at a Mr value of 8.4. The 55 °C capacity retention for this Co3O4 /LiCoO2 ion cell is shown in panel c. The cells were cycled at a C/5 rate between 0.9 and 4.1 V. To stress the reproducibility of the experiments, the data for four and two different cells has been reported in panels a and b and c, respectively. Mr refers to optimum mass ratio of the electrode materials. The C/5 rate corresponds to 1 Li ion per formula unit in 5 h. Reproduced with permission from ref 705. Copyright 2002 Electrochemical Society.

Figure 75. Voltage vs composition profiles for a CoO/LiMn2O4 Li ion cell demonstrating the possible use of Li1.33Mn2O4 as a Li reservoir. The capacity at a 3 V plateau is used to compensate for the first cycle irreversibility observed with CoO. The inset shows the capacity retention for this cell and illustrates the beneficial effect of using the Li reservoir on cell capacity compared with CoO/LiMn2O4 Li ion cells. Reprined with permission from ref 705. Copyright 2002 Electrochemical Society.

5436

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

very good chance of getting accepted for EVs/PHEVs by car manufacturers in the next few years. Alloying−dealloying reaction involving Sn (at potentials ≤0.5 V vs Li) in the tin oxides can give large and stable capacities with good current (C)-rate capability, provided they are associated with good “matrix” elements and are operated in the restricted voltage range 0.005 to 0.8 or to 1.0 V. However, they suffer from large irreversible capacity loss (ICL) during the first cycle, due to the necessity of reducing the oxide to the metal, before alloy formation and SEI formation. Similarly, oxide anodes based on conversion reactions also suffer from large ICL and, in addition, exhibit large electrode polarization. That is, they show large voltage hysteresis between the discharge and charge reaction (∼1 V). However, these need not be treated as great disadvantages: For example, the ICL can be compensated, to some extent, by the use of “sacrificial” cathodes or Li reservoir compounds in the cathode mix, such as Li1.33Mn2O4 or LixMnO2, by the use of sacrificial salts, such as Li3N and Li salts of oxocarbons, by use of prelithiated anode with stabilized Li metal powder, and by suitable mass-balance of the anode and cathode. Efforts are being made to reduce the electrode polarization in the oxide anodes by using some catalysts, either by intimate mixing or by surface coating. The LIBs with the tin oxide anodes and with many of the cheap and environmentally friendly mixed oxide anodes, such as, iron, zinc, or manganese oxides are worth studying by suitably combining with 4 V and 5 V cathodes. The coming 5 years will see the development and adaption of at least three or more oxide anodes, in addition to LTO, in LIBs for use in highenergy-density and high-power applications, mainly due to the cost and safety-in-operation considerations in comparison to graphite anodes. The conventional LIBs with the graphite anodes will, however, continue to be used for low-power portable electronic appliances, due to their convenience and nearly perfected technology. It is concluded that the field of research of the oxide anodes for LIBs has a very bright future because of the existence of a wide and distinct variety of compounds with various morphologies that have been studied, as reported in this review, and will continue to be investigated, thanks to the extensive use of the methodology of nanotechnology and use of the above-mentioned three Li cycling mechanisms for discovering new and novel oxide anodes, as well as optimizing the performance of the already established ones. As a result, the materials chemistry and electrochemistry of these oxide materials will be further enhanced and enriched.

[COCON(Li)N(Li)]n were studied, of which Li3N and the two Li keto-acid salts showed very encouraging results. A recent report1251 on the availability of stabilized lithium metal powder (SLMP) commercially, to be used as a Li reservoir to compensate for the ICL, is worthy of consideration. As mentioned in sections 6.3 and 6.5, Wang et al.1248 and Hassoun et al. 1249 used SLMP as a Li reservoir for compensating the ICL of the anodes GeOx and Fe2O3, respectively. The SLMP particles are spherical with a diameter of 29 μm and consist of 97 at. % Li metal, coated with a protective layer of, for example, Li2CO3 (3%). The protective layer gives both air and solvent stability to SLMP. Studies showed that it has a reversible capacity of 3860 mA·h g−1, when cycled between 0 and 3 V vs Li at a current of 0.1 mA. The oxide anodes can thus be prelithiated with calculated amounts of SLMP, either by a surface application technique employing a slurry in xylene or by spray method onto an already fabricated composite electrode, followed by roller-pressing. The viability of this method of mitigating the ICL by admixing of SLMP with graphite and hard-carbon anodes, in combination with LiMn2O4 cathode has been demonstrated by Li and Fritch.1251

7. CONCLUSIONS The electrochemical Li storage and cycling properties of binary, ternary, and complex metal oxides and oxysalts are reviewed. The motivation for the studies of the oxides as negative electrode (anode) materials came from the urge to replace the graphite anode and increase the energy density, decrease the cost, and most importantly, ensure safety-in-operation of the existing rechargeable lithium ion batteries (LIBs), especially for use in hybrid electric/plug-in-hybrid (HEV/PHEV) vehicles and for off-peak energy storage. Efforts to develop alternative anode (negative electrode) materials to replace graphite for use in the LIBs for the past 15 years have resulted in the investigations of a large number and wide variety of metal oxides and oxysalts for their Li cyclability. There are three mechanisms by which Li can be stored and cycled electrochemically: the first one is the Li intercalation− deintercalation reaction, notably exhibited by the titanium oxides. This is similar to that shown by the 2D-layer-structured graphite. The second mechanism is the alloying−dealloying reactions of suitable metals, notably those containing Sn, and the third mechanism is the conversion (redox) reaction involving Li2O and a transition metal, notably Co and Fe. There are both advantages and disadvantages with each of these mechanisms when dealing with metal oxides. The titanium oxides suffer from limited capacity, much less than the theoretical reversible capacity of graphite, namely, ∼170−300 vs 372 mA·h g−1. Also, the voltage at which Li cycling occurs in the titanium oxides is fairly high, namely, 1.3−1.6 vs 0.1−0.25 V for graphite, which is a large disadvantage since LIBs with these as anodes and a suitable 4 V cathode will have lower operating voltage and lower energy density. Advantages are the obvious low cost, environmental compatibility, and very good stability in both the discharged state and charged state. In recent years, LIBs with Li4Ti5O12 (LTO) as the anode and with optimized 3.5−5 V cathodes, such as LiMn2O4, LiFePO4, and Li[Ni0.5Mn1.5]O4, have been fabricated and tested, especially employing nanometric LTO. These showed outstanding performance at ambient temperature and at 55 °C, with regard to the C-rate capability, cycling stability, calendar life, pulsepower characteristics, and abuse tolerance. Thus, they stand a

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Tel: (+65) 6516 2531. Fax: (+65) 6777 6126. Notes

The authors declare no competing financial interest. 5437

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

Biographies

Prof. G. V. Subba Rao has been working on Li ion battery materials for the past 13 years. Earlier, he worked in the areas of perovskite and pyrochlore oxides, intercalated layered chalcogenides, and low- and high-Tc superconductors, including copper oxide superconductors, photoelectrochemical cells for solar energy conversion, NASICONtype fast ion conductors, and oxide phosphors. He is a Fellow of the Indian Academy of Sciences and the Indian National Science Academy. He is a member of the editorial board of the Journal of Solid State Electrochemistry.

Dr. M. V. Reddy obtained his M.Sc. from Bangalore University, Bangalore, India, and Ph.D. in 2003 (mention with highest honors) from Institute of Condensed Matter Chemistry of Bordeaux (ICMCBCNRS)/National School of Chemistry and Physics of Bordeaux (ENSCPB), University of Bordeaux, France, and for the past 10 years, he has been working as a Research Fellow (2003−2012), and from Jan. 2013 working as Senior Research Fellow at the National University of Singapore (NUS), Singapore. He has been working on Li ion battery materials (cathodes, anodes, and solid electrolytes), including novel methods of synthesis, characterization, and evaluation of the electrochemical properties. He has published more than 85 papers in various International Journals and gave plenary, keynote and invited talks at various conferences. He is serving as editorial advisory board member for 20 open access journals, referee for various international journals and project proposals, Member in International Centre for Diffraction data (ICDD), USA, International Society of Electrochemistry (Switzerland), Electrochemical Society (ECS), USA,

Prof. B.V. R. Chowdari has been working on Li ion battery materials for the past 14 years. Earlier, he worked in the area of amorphous, crystalline, glassy, and polymer electrolytes exhibiting Li, Cu, and Ag ion conduction and EPR studies of optoelectronic solids. He is the President of the Materials Research Society of Singapore and of the Asian Society for Solid State Ionics and last year assumed charge as the President of the International Union of Materials Research Societies (IUMRS) for a two-year term. He is a member of the editorial boards of Solid State Ionics and the Journal of Solid State Electrochemistry.

International Conference of Young Researchers on Advanced Materials (ICYRAM-IUMRS) 2012, Materials Research Society Singapore (MRSS), Global Materials Network (GMN-USA), and Institute of Physics Singapore (IPSS). He is one of the organizer for inaugural ICYRAM-IUMRS 2012 conference, served as Theme chair in Energy and the Environment, and a session chair in Batteries & Super capacitors, Fuel cells and Materials for Environmental protection. (http://www.mrs.org.sg/icyram2012/). Having trained

ACKNOWLEDGMENTS Thanks are due to Mr. Christie T. Cherian, Mr. M. J. Silvister Raju, Mr. Shaul and Mr.Wu Yongzi for help in the preparation of the manuscript. M.V.R. thanks Assoc. Prof. S. Adams, Department of Materials Science & Engineering, NUS and Prof. K. P. Loh, Dept. of Chemistry, NUS, for their help and encouragement. Part of the work is supported by Defence Advanced Research Projects Agency (DARPA), USA (Grant No. R-144-000-226-597), Ministry of Education (MOE), Singapore (Grant No. WBS-R-284-000-076-112), and National Research Foundation (NRF) Singapore NRF-CRP grant (Grant Nos. R-144-000-295-281 and R-143-000-360-281). Thanks are due to the anonymous referees for helpful

many high school, college, polytechnic and international exchange students in the area lithium ion batteries research projects, he won Outstanding Science Mentorship Award (2010, 2011, and 2012) from Ministry of Education (MOE), Gifted Education Branch (GEB), Singapore, and Inspiring Research Mentor Award (2011, 2012 and 2013) from NUS High School, Singapore. For other details, see http://www.researcherid.com/rid/B-3524-2010; http://orcid.org/ 0000-0002-6979-5345; http://www.physics.nus.edu.sg/ solidstateionics/. Contact E-mail addresses: [email protected]; [email protected]; [email protected]. 5438

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(35) Gummow, R. J.; Dekock, A.; Thackeray, M. M. Solid State Ionics 1994, 69, 59. (36) Jang, D. H.; Shin, Y. J.; Oh, S. M. J. Electrochem. Soc. 1996, 143, 2204. (37) Amatucci, G. G.; Blyr, A.; Sigala, C.; Alfonse, P.; Tarascon, J. M. Solid State Ionics 1997, 104, 13. (38) Thackeray, M. M. Prog. Solid State Chem. 1997, 25, 1. (39) Blyr, A.; Sigala, C.; Amatucci, G.; Guyomard, D.; Chabre, Y.; Tarascon, J. M. J. Electrochem. Soc. 1998, 145, 194. (40) Song, D.; Ikuta, H.; Uchida, T.; Wakihara, M. Solid State Ionics 1999, 117, 151. (41) Pistoia, G.; Zane, D.; Zhang, Y. J. Electrochem. Soc. 1995, 142, 2551. (42) Jiang, Z. P.; Abraham, K. M. J. Electrochem. Soc. 1996, 143, 1591. (43) Myung, S. T.; Komaba, S.; Kumagai, N. J. Electrochem. Soc. 2001, 148, A482. (44) Aurbach, D.; Markovsky, B.; Salitra, G.; Markevich, E.; Talyossef, Y.; Koltypin, M.; Nazar, L.; Ellis, B.; Kovacheva, D. J. Power Sources 2007, 165, 491. (45) Xia, Y. Y.; Yoshio, M. J. Power Sources 1997, 66, 129. (46) Matsuo, Y.; Kostecki, R.; McLarnon, F. J. Electrochem. Soc. 2001, 148, A687. (47) Kannan, A. M.; Manthiram, A. Electrochem. Solid-State Lett. 2002, 5, A167. (48) Taniguchi, I.; Song, D.; Wakihara, M. J. Power Sources 2002, 109, 333. (49) Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. J. Mater. Chem. 2003, 13, 106. (50) Fergus, J. W. J. Power Sources 2010, 195, 939. (51) Sakunthala, A.; Reddy, M. V.; Selvasekarapandian, S.; Chowdari, B. V. R.; Selvin, P. C. Electrochim. Acta 2010, 55, 4441. (52) Reddy, M. V.; Raju, M. J. S.; Sharma, N.; Quan, P. Y.; Nowshad, S. H.; Emmanuel, H. E. C.; Peterson, V. K.; Chowdari, B. V. R. J. Electrochem. Soc. 2011, 158, A1231. (53) Sharma, N.; Reddy, M. V.; Du, G. D.; Adams, S.; Chowdari, B. V. R.; Guo, Z. P.; Peterson, V. K. J. Phys. Chem. C 2011, 115, 21473. (54) Zhao, X.; Reddy, M. V.; Liu, H.; Ramakrishna, S.; Subba Rao, G. V.; Chowdari, B. V. R. RSC Adv. 2012, 2, 7462. (55) Prabhu, M.; Reddy, M. V.; Selvasekarapandian, S.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochim. Acta 2013, 88, 745. (56) Padhi, A. K.; Nanjundaswamy, K. S.; Goodenough, J. B. J. Electrochem. Soc. 1997, 144, 1188. (57) Huang, H.; Yin, S. C.; Nazar, L. F. Electrochem. Solid-State Lett. 2001, 4, A170. (58) Ravet, N.; Chouinard, Y.; Magnan, J. F.; Besner, S.; Gauthier, M.; Armand, M. J. Power Sources 2001, 97−98, 503. (59) Yamada, A.; Chung, S. C.; Hinokuma, K. J. Electrochem. Soc. 2001, 148, A224. (60) Chen, Z. H.; Dahn, J. R. J. Electrochem. Soc. 2002, 149, A1184. (61) Doeff, M. M.; Hu, Y. Q.; McLarnon, F.; Kostecki, R. Electrochem. Solid-State Lett. 2003, 6, A207. (62) Herle, P. S.; Ellis, B.; Coombs, N.; Nazar, L. F. Nat. Mater. 2004, 3, 147. (63) Hsu, K. F.; Tsay, S. Y.; Hwang, B. J. J. Mater. Chem. 2004, 14, 2690. (64) Myung, S. T.; Komaba, S.; Hirosaki, N.; Yashiro, H.; Kumagai, N. Electrochim. Acta 2004, 49, 4213. (65) Islam, M. S.; Driscoll, D. J.; Fisher, C. A. J.; Slater, P. R. Chem. Mater. 2005, 17, 5085. (66) Belharouak, I.; Johnson, C.; Amine, K. Electrochem. Commun. 2005, 7, 983. (67) Chen, J. J.; Whittingham, M. S. Electrochem. Commun. 2006, 8, 855. (68) Delacourt, C.; Poizot, P.; Levasseur, S.; Masquelier, C. Electrochem. Solid-State Lett. 2006, 9, A352. (69) Choi, D.; Kumta, P. N. J. Power Sources 2007, 163, 1064. (70) Martin, J. F.; Yamada, A.; Kobayashi, G.; Nishimura, S. I.; Kanno, R.; Guyomard, D.; Dupre, N. Electrochem. Solid-State Lett. 2008, 11, A12.

suggestions and constructive comments to improve the manuscript.

REFERENCES (1) Wakihara, M., Yamamoto, O., Eds. Li-Ion Batteries; Wiley-VCH: New York, 1998. (2) Van Schalkwijk, W. A., Scrosati, B., Eds. Advances in Lithium-Ion Batteries; Kluwer Academic /Plenum: New York, 2002. (3) Nazri, G.-A., Pistoia, G., Eds. Lithium Batteries: Science and Technology; Kluwer Academic Publishers: New York, 2003. (4) Ozawa, K., Ed. Lithium Ion Rechargeable Batteries; Wiley-VCH: Weinheim, Germany, 2009. (5) Alifantis, K. E., Hackney, S. A., Vasant Kumar, R., Eds. High Energy Density Lithium Batteries: Materials, Engineering, Applications; Wiley VCH: Weinheim, Germany, 2010. (6) Tarascon, J. M.; Armand, M. Nature 2001, 414, 359. (7) Arico, A. S.; Bruce, P.; Scrosati, B.; Tarascon, J. M.; Van Schalkwijk, W. Nat. Mater. 2005, 4, 366. (8) Armand, M.; Tarascon, J. M. Nature 2008, 451, 652. (9) http://www.isuppli.com/semiconductor-value-chain/pages/ strong-growth-to-drive-lithium-ion-battery-market-to-61-billion-by2020.aspx (Accessed on Feb. 29, 2012). (10) Zaghib, K.; Dontigny, M.; Guerfi, A.; Charest, P.; Rodrigues, I.; Mauger, A.; Julien, C. M. J. Power Sources 2011, 196, 3949. (11) Delmas, C.; Menetrier, M.; Croguennec, L.; Saadoune, I.; Rougier, A.; Pouillerie, C.; Prado, G.; Grune, M.; Fournes, L. Electrochim. Acta 1999, 45, 243. (12) Yang, H.; Prakash, J. J. Electrochem. Soc. 2004, 151, A1222. (13) Amine, K.; Liu, J.; Belharouak, I.; Kang, S. H.; Bloom, I.; Vissers, D.; Henriksen, G. J. Power Sources 2005, 146, 111. (14) Abraham, D. P.; Reynolds, E. M.; Sammann, E.; Jansen, A. N.; Dees, D. W. Electrochim. Acta 2005, 51, 502. (15) Kerlau, M.; Reimer, J. A.; Cairns, E. J. J. Electrochem. Soc. 2005, 152, A1629. (16) Kondo, H.; Takeuchi, Y.; Sasaki, T.; Kawauchi, S.; Itou, Y.; Hiruta, O.; Okuda, C.; Yonemura, M.; Kamiyama, T.; Ukyo, Y. J. Power Sources 2007, 174, 1131. (17) Ramana, C. V.; Zaghib, K.; Julien, C. M. Appl. Phys. Lett. 2007, 90, No. 021916. (18) Zheng, S. J.; Huang, R.; Makimura, Y.; Ukyo, Y.; Fisher, C. A. J.; Hirayama, T.; Ikuhara, Y. J. Electrochem. Soc. 2011, 158, A357. (19) Ohzuku, T.; Makimura, Y. Chem. Lett. 2001, 642. (20) Lu, Z. H.; MacNeil, D. D.; Dahn, J. R. Electrochem. Solid-State Lett. 2001, 4, A200. (21) Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochim. Acta 2002, 48, 145. (22) Choi, J.; Manthiram, A. Electrochem. Solid-State Lett. 2005, 8, C102. (23) Myung, S. T.; Izumi, K.; Komaba, S.; Sun, Y. K.; Yashiro, H.; Kumagai, N. Chem. Mater. 2005, 17, 3695. (24) Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2006, 159, 263. (25) Ohzuku, T.; Ueda, A.; Nagayama, M.; Iwakoshi, Y.; Komori, H. Electrochim. Acta 1993, 38, 1159. (26) Caurant, D.; Baffier, N.; Garcia, B.; PereiraRamos, J. P. Solid State Ionics 1996, 91, 45. (27) Koksbang, R.; Barker, J.; Shi, H.; Saidi, M. Y. Solid State Ionics 1996, 84, 1. (28) Saadoune, I.; Delmas, C. J. Solid State Chem. 1998, 136, 8. (29) Levasseur, S.; Menetrier, M.; Suard, E.; Delmas, C. Solid State Ionics 2000, 128, 11. (30) MacNeil, D. D.; Lu, Z. H.; Chen, Z. H.; Dahn, J. R. J. Power Sources 2002, 108, 8. (31) Chen, Z. H.; Dahn, J. R. Electrochim. Acta 2004, 49, 1079. (32) Tan, K. S.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2005, 147, 241. (33) Abraham, K. M. Electrochim. Acta 1993, 38, 1233. (34) Tarascon, J. M.; Guyomard, D. Electrochim. Acta 1993, 38, 1221. 5439

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(71) Delmas, C.; Maccario, M.; Croguennec, L.; Le Cras, F.; Weill, F. Nat. Mater. 2008, 7, 665. (72) Saravanan, K.; Reddy, M. V.; Balaya, P.; Gong, H.; Chowdari, B. V. R.; Vittal, J. J. J. Mater. Chem. 2009, 19, 605. (73) Scrosati, B.; Garche, J. J. Power Sources 2010, 195, 2419. (74) Tarascon, J. M.; Recham, N.; Armand, M.; Chotard, J. N.; Barpanda, P.; Walker, W.; Dupont, L. Chem. Mater. 2010, 22, 724. (75) Fongy, C.; Gaillot, A. C.; Jouanneau, S.; Guyomard, D.; Lestriez, B. J. Electrochem. Soc. 2010, 157, A885. (76) Saravanan, K.; Balaya, P.; Reddy, M. V.; Chowdari, B. V. R.; Vittal, J. J. Energy Environ. Sci. 2010, 3, 457. (77) Gong, Z. L.; Yang, Y. Energy Environ. Sci. 2011, 4, 3223. (78) Martin, J. F.; Cuisinier, M.; Dupre, N.; Yamada, A.; Kanno, R.; Guyomard, D. J. Power Sources 2011, 196, 2155. (79) Dimesso, L.; Forster, C.; Jaegermann, W.; Khanderi, J. P.; Tempel, H.; Popp, A.; Engstler, J.; Schneider, J. J.; Sarapulova, A.; Mikhailova, D.; Schmitt, L. A.; Oswald, S.; Ehrenberg, H. Chem. Soc. Rev. 2012, 41, 5068. (80) Kandhasamy, S.; Nallathamby, K.; Minakshi, M. Prog. Solid State Chem. 2012, 40, 1. (81) Seid, K. A.; Badot, J. C.; Dubrunfaut, O.; Levasseur, S.; Guyomard, D.; Lestriez, B. J. Mater. Chem. 2012, 22, 2641. (82) Barker, J.; Saidi, M. Y.; Swoyer, J. L. J. Electrochem. Soc. 2003, 150, A1394. (83) Zhou, F.; Zhao, X. M.; Dahn, J. R. Electrochem. Commun. 2009, 11, 589. (84) Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2010, 195, 5768. (85) Mueller, T.; Hautier, G.; Jain, A.; Ceder, G. Chem. Mater. 2011, 23, 3854. (86) Mba, J. M. A.; Masquelier, C.; Suard, E.; Croguennec, L. Chem. Mater. 2012, 24, 1223. (87) Saidi, M. Y.; Barker, J.; Huang, H.; Swoyer, J. L.; Adamson, G. Electrochem. Solid-State Lett. 2002, 5, A149. (88) Huang, H.; Yin, S. C.; Kerr, T.; Taylor, N.; Nazar, L. F. Adv. Mater. 2002, 14, 1525. (89) Yin, S. C.; Grondey, H.; Strobel, P.; Anne, M.; Nazar, L. F. J. Am. Chem. Soc. 2003, 125, 10402. (90) Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. Future Generation Electrode Materials for Lithium Batteries Solid State Ionics: Fundamental Researches and Technological Applications. In proceedings of the 12th Asian Conf. on Solid State Ionics; Chowdari, B. V. R., et al., Eds.; Wuhan University of Technology Press: Wuhan, China, 2010; p 497. (91) Forman, J. C.; Moura, S. J.; Stein, J. L.; Fathy, H. K. J. Power Sources 2012, 210, 263. (92) Zhong, Q. M.; Bonakdarpour, A.; Zhang, M. J.; Gao, Y.; Dahn, J. R. J. Electrochem. Soc. 1997, 144, 205. (93) Myung, S. T.; Komaba, S.; Kumagai, N.; Yashiro, H.; Chung, H. T.; Cho, T. H. Electrochim. Acta 2002, 47, 2543. (94) Lee, Y. S.; Sun, Y. K.; Ota, S.; Miyashita, T.; Yoshi, M. Electrochem. Commun. 2002, 4, 989. (95) Alcantara, R.; Jaraba, M.; Lavela, P.; Tirado, J. L. Electrochim. Acta 2002, 47, 1829. (96) Markovsky, B.; Talyossef, Y.; Salitra, G.; Aurbach, D.; Kim, H. J.; Choi, S. Electrochem. Commun. 2004, 6, 821. (97) Kunduraci, M.; Amatucci, G. G. Electrochim. Acta 2008, 53, 4193. (98) Shaju, K. M.; Bruce, P. G. Dalton Trans. 2008, 5471. (99) Liu, J.; Manthiram, A. Chem. Mater. 2009, 21, 1695. (100) Liu, J.; Manthiram, A. J. Electrochem. Soc. 2009, 156, A66. (101) Santhanam, R.; Rambabu, B. J. Power Sources 2010, 195, 5442. (102) Kim, M. G.; Cho, J. Adv. Funct. Mater. 2009, 19, 1497. (103) Reddy, M. V.; Cheng, H. Y.; Tham, J. H.; Yuan, C. Y.; Goh, H. L.; Chowdari, B. V. R. Electrochim. Acta 2012, 62, 269. (104) Kawai, H.; Nagata, M.; Tukamoto, H.; West, A. R. J. Power Sources 1999, 81, 67. (105) Fey, G. T. K.; Li, W.; Dahn, J. R. J. Electrochem. Soc. 1994, 141, 2279.

(106) Prabaharan, S. R. S.; Michael, M. S.; Radhakrishna, S.; Julien, C. J. Mater. Chem. 1997, 7, 1791. (107) Fey, G. T. K.; Huang, D. L. Electrochim. Acta 1999, 45, 295. (108) Amine, K.; Yasuda, H.; Yamachi, M. Electrochem. Solid-State Lett. 2000, 3, 178. (109) Okada, S.; Sawa, S.; Egashira, M.; Yamaki, J.; Tabuchi, M.; Kageyama, H.; Konishi, T.; Yoshino, A. J. Power Sources 2001, 97−98, 430. (110) Lloris, J. M.; Vicente, C. P.; Tirado, J. L. Electrochem. Solid-State Lett. 2002, 5, A234. (111) Ramana, C. V.; Ait-Salah, A.; Utsunomiya, S.; Becker, U.; Mauger, A.; Gendron, F.; Julien, C. M. Chem. Mater. 2006, 18, 3788. (112) Bramnik, N. N.; Nikolowski, K.; Baehtz, C.; Bramnik, K. G.; Ehrenberg, H. Chem. Mater. 2007, 19, 908. (113) Wolfenstine, J.; Read, J.; Allen, J. L. J. Power Sources 2007, 163, 1070. (114) Prabu, M.; Selvasekarapandian, S.; Reddy, M. V.; Chowdari, B. V. R. J. Solid State Electrochem. 2012, 16, 1833. (115) Winter, M.; Besenhard, J. O.; Spahr, M. E.; Novak, P. Adv. Mater. 1998, 10, 725. (116) Aurbach, D.; Markovsky, B.; Weissman, I.; Levi, E.; Ein-Eli, Y. Electrochim. Acta 1999, 45, 67. (117) Flandrois, S.; Simon, B. Carbon 1999, 37, 165. (118) Verma, P.; Maire, P.; Novak, P. Electrochim. Acta 2010, 55, 6332. (119) Sony press news (2005) (www.Sony.net/SonyInfo/News/ Press/200502/05-006E/index.html) (120) Hatchard, T. D.; Dahn, J. R. J. Electrochem. Soc. 2004, 151, A838. (121) Kim, I. S.; Kumta, P. N. J. Power Sources 2004, 136, 145. (122) Liu, Y.; Matsumura, T.; Imanishi, N.; Hirano, A.; Ichikawa, T.; Takeda, Y. Electrochem. Solid-State Lett. 2005, 8, A599. (123) Takamura, T.; Uehara, M.; Suzuki, J.; Sekine, K.; Tamura, K. J. Power Sources 2006, 158, 1401. (124) Yin, J. T.; Wada, M.; Yamamoto, K.; Kitano, Y.; Tanase, S.; Sakai, T. J. Electrochem. Soc. 2006, 153, A472. (125) Li, J.; Lewis, R. B.; Dahn, J. R. Electrochem. Solid-State Lett. 2007, 10, A17. (126) Ma, H.; Cheng, F. Y.; Chen, J.; Zhao, J. Z.; Li, C. S.; Tao, Z. L.; Liang, J. Adv. Mater. 2007, 19, 4067. (127) Larcher, D.; Beattie, S.; Morcrette, M.; Edstroem, K.; Jumas, J. C.; Tarascon, J. M. J. Mater. Chem. 2007, 17, 3759. (128) Lestrie, B.; Bahri, S.; Sandu, I.; Roue, L.; Guyomard, D. Electrochem. Commun. 2007, 9, 2801. (129) Sandu, I.; Moreau, P.; Guyomard, D.; Brousse, T.; Roue, L. Solid State Ionics 2007, 178, 1297. (130) Obrovac, M. N.; Christensen, L.; Le, D. B.; Dahn, J. R. J. Electrochem. Soc. 2007, 154, A849. (131) Cui, L. F.; Ruffo, R.; Chan, C. K.; Peng, H. L.; Cui, Y. Nano Lett. 2009, 9, 491. (132) Mazouzi, D.; Lestriez, B.; Roue, L.; Guyomard, D. Electrochem. Solid-State Lett. 2009, 12, A215. (133) Chan, C. K.; Patel, R. N.; O’Connell, M. J.; Korgel, B. A.; Cui, Y. ACS Nano 2010, 4, 1443. (134) Hertzberg, B.; Alexeev, A.; Yushin, G. J. Am. Chem. Soc. 2010, 132, 8548. (135) Magasinski, A.; Dixon, P.; Hertzberg, B.; Kvit, A.; Ayala, J.; Yushin, G. Nat. Mater. 2010, 9, 353. (136) Park, C. M.; Kim, J. H.; Kim, H.; Sohn, H. J. Chem. Soc. Rev. 2010, 39, 3115. (137) Foll, H.; Hartz, H.; Ossei-Wusu, E.; Carstensen, J.; Riemenschneider, O. Phys. Status Solidi RRL 2010, 4, 4. (138) Choi, N. S.; Yao, Y.; Cui, Y.; Cho, J. J. Mater. Chem. 2011, 21, 9825. (139) Liu, N.; Wu, H.; McDowell, M. T.; Yao, Y.; Wang, C. M.; Cui, Y. Nano Lett. 2012, 12, 3315. (140) Phan, V. P.; Pecquenard, B.; Le Cras, F. Adv. Funct. Mater. 2012, 22, 2580. 5440

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(141) Ionica, C. M.; Lippens, P. E.; Fourcade, J. O.; Jumas, J. C. J. Power Sources 2005, 146, 478. (142) Bryngelsson, H.; Eskhult, J.; Edstrom, K.; Nyholm, L. Electrochim. Acta 2007, 53, 1062. (143) Sun, H.; He, X. M.; Ren, J. G.; Li, J. J.; Jiang, C. Y.; Wan, C. R. Rare Met. Mater. Eng. 2007, 36, 1313. (144) Roberts, M.; Johns, P.; Owen, J.; Brandell, D.; Edstrom, K.; El Enany, G.; Guery, C.; Golodnitsky, D.; Lacey, M.; Lecoeur, C.; Mazor, H.; Peled, E.; Perre, E.; Shaijumon, M. M.; Simon, P.; Taberna, P. L. J. Mater. Chem. 2011, 21, 9876. (145) Huggins, R. A. J. Power Sources 1999, 81, 13. (146) Todd, A. D. W.; Ferguson, P. P.; Fleischauer, M. D.; Dahn, J. R. Int. J. Energy Res. 2010, 34, 535. (147) Li, H.; Balaya, P.; Maier, J. J. Electrochem. Soc. 2004, 151, A1878. (148) Cabana, J.; Monconduit, L.; Larcher, D.; Palacin, M. R. Adv. Mater. 2010, 22, E170. (149) Nishijima, M.; Kagohashi, T.; Takeda, Y.; Imanishi, M.; Yamamoto, O. J. Power Sources 1997, 68, 510. (150) Das, B.; Reddy, M. V.; Malar, P.; Osipowicz, T.; Subba Rao, G. V.; Chowdari, B. V. R. Solid State Ionics 2009, 180, 1061. (151) Das, B.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Mater. Chem. 2012, 22, 17505. (152) Nazar, L. F.; Goward, G.; Leroux, F.; Duncan, M.; Huang, H.; Kerr, T.; Gaubicher, J. Int. J. Inorg. Mater. 2001, 3, 191. (153) Wakihara, M. Mater. Sci. Eng., R 2001, 33, 109. (154) Dell, R. M.; Rand, D. A. J. J. Power Sources 2001, 100, 2. (155) Tirado, J. L. Mater. Sci. Eng., R 2003, 40, 103. (156) Winter, M.; Brodd, R. J. Chem. Rev. 2004, 104, 4245. (157) Maier, J. Nat. Mater. 2005, 4, 805. (158) Bruce, P. G. Solid State Ionics 2008, 179, 752. (159) Whittingham, M. S. Dalton Trans. 2008, 5424. (160) Guo, Y. G.; Hu, J. S.; Wan, L. J. Adv. Mater. 2008, 20, 2878. (161) Balaya, P. Energy Environ. Sci. 2008, 1, 645. (162) Bruce, P. G.; Scrosati, B.; Tarascon, J. M. Angew. Chem., Int. Ed. 2008, 47, 2930. (163) Cheng, F.; Tao, Z.; Liang, J.; Chen, J. Chem. Mater. 2008, 20, 667. (164) Shukla, A. K.; Kumar, T. P. Curr. Sci. 2008, 94, 314. (165) Chernova, N. A.; Roppolo, M.; Dillon, A. C.; Whittingham, M. S. J. Mater. Chem. 2009, 19, 2526. (166) Deng, D.; Kim, M. G.; Lee, J. Y.; Cho, J. Energy Environ. Sci. 2009, 2, 818. (167) Centi, G.; Perathoner, S. Eur. J. Inorg. Chem. 2009, 3851. (168) Li, H.; Wang, Z. X.; Chen, L. Q.; Huang, X. J. Adv. Mater. 2009, 21, 4593. (169) Yang, Z. G.; Choi, D.; Kerisit, S.; Rosso, K. M.; Wang, D. H.; Zhang, J.; Graff, G.; Liu, J. J. Power Sources 2009, 192, 588. (170) Liu, C.; Li, F.; Ma, L. P.; Cheng, H. M. Adv. Mater. 2010, 22, E28. (171) Su, D. S.; Schlogl, R. ChemSusChem 2010, 3, 136. (172) Yi, T. F.; Jiang, L. J.; Shu, J.; Yue, C. B.; Zhu, R. S.; Qiao, H. B. J. Phys. Chem. Solids 2010, 71, 1236. (173) Baddour-Hadjean, R.; Pereira-Ramos, J. P. Chem. Rev. 2010, 110, 1278. (174) Jiang, J.; Li, Y.; Liu, J.; Huang, X. Nanoscale 2011, 3, 45. (175) Kamali, A. R.; Fray, D. J. Rev. Adv. Mater. Sci. 2011, 27, 14. (176) Lee, K. T.; Cho, J. Nano Today 2011, 6, 28. (177) Li, J.; Daniel, C.; Wood, D. J. Power Sources 2011, 196, 2452. (178) Song, M.-K.; Park, S.; Alamgir, F. M.; Cho, J.; Liu, M. Mater. Sci. Eng., R 2011, 72, 203. (179) Zhang, X.; Ji, L.; Toprakci, O.; Liang, Y.; Alcoutlabi, M. Polym. Rev. 2011, 51, 239. (180) Jeong, G.; Kim, Y. U.; Kim, H.; Kim, Y. J.; Sohn, H. J. Energy Environ. Sci. 2011, 4, 1986. (181) Ji, L.; Lin, Z.; Alcoutlabi, M.; Zhang, X. Energy Environ. Sci. 2011, 4, 2682. (182) Cavaliere, S.; Subianto, S.; Savych, I.; Jones, D. J.; Roziere, J. Energy Environ. Sci. 2011, 4, 4761.

(183) Yang, Z.; Zhang, J.; Kintner-Meyer, M. C. W.; Lu, X.; Choi, D.; Lemmon, J. P.; Liu, J. Chem. Rev. 2011, 111, 3577. (184) Ji, G.; Ma, Y.; Lee, J. Y. J. Mater. Chem. 2011, 21, 9819. (185) Chen, J. S.; Archer, L. A.; Lou, X. W. J. Mater. Chem. 2011, 21, 9912. (186) Marom, R.; Amalraj, S. F.; Leifer, N.; Jacob, D.; Aurbach, D. J. Mater. Chem. 2011, 21, 9938. (187) Dunn, B.; Kamath, H.; Tarascon, J. M. Science 2011, 334, 928. (188) Liu, Y. Y.; Liu, D. W.; Zhang, Q. F.; Cao, G. Z. J. Mater. Chem. 2011, 21, 9969. (189) Singh, V.; Joung, D.; Zhai, L.; Das, S.; Khondaker, S. I.; Seal, S. Prog. Mater. Sci. 2011, 56, 1178. (190) Zhou, Z. Y.; Tian, N.; Li, J. T.; Broadwell, I.; Sun, S. G. Chem. Soc. Rev. 2011, 40, 4167. (191) Cheng, F. Y.; Chen, J. J. Mater. Chem. 2011, 21, 9841. (192) Tartaj, P.; Morales, M. P.; Gonzalez-Carreno, T.; VeintemillasVerdaguer, S.; Serna, C. J. Adv. Mater. 2011, 23, 5243. (193) Dillon, S. J.; Sun, K. Curr. Opin. Solid State Mater. Sci. 2012, 16, 153. (194) Li, Y.; Yang, X. Y.; Feng, Y.; Yuan, Z. Y.; Su, B. L. Crit. Rev. Solid State Mater. Sci. 2012, 37, 1. (195) Kim, T.-H.; Park, J.-S.; Chang, S. K.; Choi, S.; Ryu, J. H.; Song, H.-K. Adv. Energy Mater. 2012, 2, 860. (196) Devan, R. S.; Patil, R. A.; Lin, J. H.; Ma, Y. R. Adv. Funct. Mater. 2012, 22, 3326. (197) Liu, J. H.; Liu, X. W. Adv. Mater. 2012, 24, 4097. (198) Nishihara, H.; Kyotani, T. Adv. Mater. 2012, 24, 4473. (199) Tan, C. W.; Tan, K. H.; Ong, Y. T.; Mohamed, A. R.; Zein, S. H. S.; Tan, S. H. Environ. Chem. Lett. 2012, 10, 265. (200) Ellefson, C. A.; Marin-Flores, O.; Ha, S.; Norton, M. G. J. Mater. Sci. 2012, 47, 2057. (201) Djenizian, T.; Hanzu, I.; Knauth, P. J. Mater. Chem. 2011, 21, 9925. (202) Zhu, G. N.; Wang, Y. G.; Xia, Y. Y. Energy Environ. Sci. 2012, 5, 6652. (203) Su, X.; Wu, Q. L.; Zhan, X.; Wu, J.; Wei, S. Y.; Guo, Z. H. J. Mater. Sci. 2012, 47, 2519. (204) Berger, T.; Monllor-Satoca, D.; Jankulovska, M.; LanaVillarreal, T.; Gomez, R. ChemPhysChem 2012, 13, 2824. (205) Froschl, T.; Hormann, U.; Kubiak, P.; Kucerova, G.; Pfanzelt, M.; Weiss, C. K.; Behm, R. J.; Husing, N.; Kaiser, U.; Landfester, K.; Wohlfahrt-Mehrens, M. Chem. Soc. Rev. 2012, 41, 5313. (206) Knoops, H. C. M.; Donders, M. E.; van de Sanden, M. C. M.; Notten, P. H. L.; Kessels, W. M. M. J. Vac. Sci. Technol., A 2012, 30, No. 010801. (207) Meng, X. B.; Yang, X. Q.; Sun, X. L. Adv. Mater. 2012, 24, 3589. (208) Inagaki, M.; Yang, Y.; Kang, F. Y. Adv. Mater. 2012, 24, 2547. (209) Yan, L.; Zheng, Y. B.; Zhao, F.; Li, S. J.; Gao, X. F.; Xu, B. Q.; Weiss, P. S.; Zhao, Y. L. Chem. Soc. Rev. 2012, 41, 97. (210) Liu, X. M.; Huang, Z. D.; Oh, S. W.; Zhang, B.; Ma, P. C.; Yuen, M. M. F.; Kim, J. K. Compos. Sci. Technol. 2012, 72, 121. (211) Tiwari, J. N.; Tiwari, R. N.; Kim, K. S. Prog. Mater. Sci. 2012, 57, 724. (212) Dai, L. M.; Chang, D. W.; Baek, J. B.; Lu, W. Small 2012, 8, 1130. (213) de las Casas, C.; Li, W. Z. J. Power Sources 2012, 208, 74. (214) Luo, B.; Liu, S. M.; Zhi, L. J. Small 2012, 8, 630. (215) Chen, J. Z.; Yang, L.; Tang, Y. F. J. Power Sources 2010, 195, 6893. (216) Lai, C.; Li, G. R.; Dou, Y. Y.; Gao, X. P. Electrochim. Acta 2010, 55, 4567. (217) Gnanasekar, K. I.; Subramanian, V.; Robinson, J.; Jiang, J. C.; Posey, F. E.; Rambabu, B. J. Mater. Res. 2002, 17, 1507. (218) Subramanian, V.; Karki, A.; Gnanasekar, K. I.; Eddy, F. P.; Rambabu, B. J. Power Sources 2006, 159, 186. (219) Wang, J. P.; Bai, Y.; Wu, M. Y.; Yin, J.; Zhang, W. F. J. Power Sources 2009, 191, 614. 5441

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(220) Kubiak, P.; Pfanzelt, M.; Geserick, J.; Hormann, U.; Husing, N.; Kaiser, U.; Wohlfahrt-Mehrens, M. J. Power Sources 2009, 194, 1099. (221) Mancini, M.; Kubiak, P.; Geserick, J.; Marassi, R.; Husing, N.; Wohlfahrt-Mehrens, M. J. Power Sources 2009, 189, 585. (222) Saravanan, K.; Ananthanarayanan, K.; Balaya, P. Energy Environ. Sci. 2010, 3, 939. (223) Beuvier, T.; Richard-Plouet, M.; Mancini-Le Granvalet, M.; Brousse, T.; Crosnier, O.; Brohan, L. Inorg. Chem. 2010, 49, 8457. (224) Inaba, M.; Oba, Y.; Niina, F.; Murota, Y.; Ogino, Y.; Tasaka, A.; Hirota, K. J. Power Sources 2009, 189, 580. (225) Jin, Y. H.; Lee, S. H.; Shim, H. W.; Ko, K. H.; Kim, D. W. Electrochim. Acta 2010, 55, 7315. (226) Wei, Z.; Liu, Z.; Jiang, R. R.; Bian, C. Q.; Huang, T.; Yu, A. S. J. Solid State Electrochem. 2010, 14, 1045. (227) Kim, H. S.; Kang, S. H.; Chung, Y. H.; Sung, Y. E. Electrochem. Solid-State Lett. 2010, 13, A15. (228) Reddy, M. V.; Jose, R.; Teng, T. H.; Chowdari, B. V. R.; Ramakrishna, S. Electrochim. Acta 2010, 55, 3109. (229) Li, H. Q.; Martha, S. K.; Unocic, R. R.; Luo, H. M.; Dai, S.; Qu, J. J. Power Sources 2012, 218, 88. (230) Nam, S. H.; Shim, H. S.; Kim, Y. S.; Dar, M. A.; Kim, J. G.; Kim, W. B. ACS Appl. Mater. Interfaces 2010, 2, 2046. (231) Zhou, W. J.; Liu, H.; Boughton, R. I.; Du, G. J.; Lin, J. J.; Wang, J. Y.; Liu, D. J. Mater. Chem. 2010, 20, 5993. (232) Lunell, S.; Stashans, A.; Ojamae, L.; Lindstrom, H.; Hagfeldt, A. J. Am. Chem. Soc. 1997, 119, 7374. (233) Wagemaker, M.; Kentgens, A. P. M.; Mulder, F. M. Nature 2002, 418, 397. (234) Wagemaker, M.; Borghols, W. J. H.; Mulder, F. M. J. Am. Chem. Soc. 2007, 129, 4323. (235) Lafont, U.; Carta, D.; Mountjoy, G.; Chadwick, A. V.; Kelder, E. M. J. Phys. Chem. C 2010, 114, 1372. (236) Luca, V.; Hunter, B.; Moubaraki, B.; Murray, K. S. Chem. Mater. 2001, 13, 796. (237) Baddour-Hadjean, R.; Bach, S.; Smirnov, M.; Pereira-Ramos, J. P. J. Raman Spectrosc. 2004, 35, 577. (238) Baddour-Hadjean, R.; Pereira-Ramos, J. P. J. Power Sources 2007, 174, 1188. (239) Smirnov, M.; Baddour-Hadjean, R. J. Chem. Phys. 2004, 121, 2348. (240) Hardwick, L. J.; Holzapfel, M.; Novak, P.; Dupont, L.; Baudrin, E. Electrochim. Acta 2007, 52, 5357. (241) Wagemaker, M.; van de Krol, R.; Kentgens, A. P. M.; van Well, A. A.; Mulder, F. M. J. Am. Chem. Soc. 2001, 123, 11454. (242) Wagemaker, M.; Kearley, G. J.; van Well, A. A.; Mutka, H.; Mulder, F. M. J. Am. Chem. Soc. 2003, 125, 840. (243) Wagemaker, M.; Lutzenkirchen-Hecht, D.; Keil, P.; van Well, A. A.; Frahm, R. Phys. B (Amsterdam, Neth.) 2003, 336, 118. (244) Wagemaker, M.; van de Krol, R.; van Well, A. A. Phys. B (Amsterdam, Neth.) 2003, 336, 124. (245) Wagemaker, M.; Lutzenkirchen-Hecht, D.; van Well, A. A.; Frahm, R. J. Phys. Chem. B 2004, 108, 12456. (246) Wagemaker, M.; van Well, A. A.; Kearley, G. J.; Mulder, F. M. Solid State Ionics 2004, 175, 191. (247) Lutzenkirchen-Hecht, D.; Wagemaker, M.; van Well, A. A.; Frahm, R. Phys. Scr. 2005, T115, 390. (248) Wagemaker, M.; Van Der Ven, A.; Morgan, D.; Ceder, G.; Mulder, F. M.; Kearley, G. J. Chem. Phys. 2005, 317, 130. (249) Gubbens, P. C. M.; Wagemaker, M.; Sakarya, S.; Blaauw, M.; Yaouanc, A.; de Reotier, P. D.; Cottrell, S. P. Solid State Ionics 2006, 177, 145. (250) Wagemaker, M.; Borghols, W. J. H.; van Eck, E. R. H.; Kentgens, A. P. M.; Kearley, G. L.; Mulder, F. M. Chem.Eur. J. 2007, 13, 2023. (251) Borghols, W. J. H.; Lutzenkirchen-Hecht, D.; Haake, U.; van Eck, E. R. H.; Mulder, F. M.; Wagemaker, M. Phys. Chem. Chem. Phys. 2009, 11, 5742.

(252) Ganapathy, S.; van Eck, E. R. H.; Kentgens, A. P. M.; Mulder, F. M.; Wagemaker, M. Chem.Eur. J. 2011, 17, 14811. (253) Sudant, G.; Baudrin, E.; Larcher, D.; Tarascon, J. M. J. Mater. Chem. 2005, 15, 1263. (254) Jung, H. G.; Oh, S. W.; Ce, J.; Jayaprakash, N.; Sun, Y. K. Electrochem. Commun. 2009, 11, 756. (255) Shin, J. Y.; Samuelis, D.; Maier, J. Adv. Funct. Mater. 2011, 21, 3464. (256) Yang, S. B.; Feng, X. L.; Mullen, K. Adv. Mater. 2011, 23, 3575. (257) Kang, J. W.; Kim, D. H.; Mathew, V.; Lim, J. S.; Gim, J. H.; Kimz, J. J. Electrochem. Soc. 2011, 158, A59. (258) Sasidharan, M.; Nakashima, K.; Gunawardhana, N.; Yokoi, T.; Inoue, M.; Yusa, S.; Yoshio, M.; Tatsumi, T. Chem. Commun. 2011, 47, 6921. (259) Park, S. J.; Kim, H.; Kim, Y. J.; Lee, H. Electrochim. Acta 2011, 56, 5355. (260) Borghols, W. J. H.; Lutzenkirchen-Hecht, D.; Haake, U.; Chan, W.; Lafont, U.; Kelder, E. M.; van Eck, E. R. H.; Kentgens, A. P. M.; Mulder, F. M.; Wagemaker, M. J. Electrochem. Soc. 2010, 157, A582. (261) Das, S. K.; Bhattacharyya, A. J. J. Electrochem. Soc. 2011, 158, A705. (262) Chen, J. S.; Tan, Y. L.; Li, C. M.; Cheah, Y. L.; Luan, D.; Madhavi, S.; Boey, F. Y. C.; Archer, L. A.; Lou, X. W. J. Am. Chem. Soc. 2010, 132, 6124. (263) Zhang, F.; Zhang, Y.; Song, S. Y.; Zhang, H. J. J. Power Sources 2011, 196, 8618. (264) Kubiak, P.; Froschl, T.; Husing, N.; Hormann, U.; Kaiser, U.; Schiller, R.; Weiss, C. K.; Landfester, K.; Wohlfahrt-Mehrens, M. Small 2011, 7, 1690. (265) Zhu, P.; Wu, Y.; Reddy, M. V.; Nair, A. S.; Chowdari, B. V. R.; Ramakrishna, S. RSC Adv. 2012, 2, 531. (266) Wang, Y.; Su, X. W.; Lu, S. J. Mater. Chem. 2012, 22, 1969. (267) Shin, J. Y.; Joo, J. H.; Samuelis, D.; Maier, J. Chem. Mater. 2012, 24, 543. (268) Cherian, C. T.; Reddy, M. V.; Magdaleno, T.; Sow, C. H.; Ramanujachary, K. V.; Subba Rao, G. V.; Chowdari, B. V. R. CrystEngComm 2012, 14, 978. (269) Reddy, M. V.; Teoh, X. W. V.; Nguyen, T. B.; Lim, Y. Y. M.; Chowdari, B. V. R. J. Electrochem. Soc. 2012, 159, A762. (270) Reddy, M. V.; Pei Theng, L.; Soh., H.; Beichen, Z.; Jiahuan, F.; Yu, C.; Yen Ling, A.; Andreea, L. Y.; Justin Ng, C. H.; Galen Liang, T. J. L.; Ian, M. F.; An, H. V. T.; Ramanathan, K.; Kevin, C. W. J.; Daryl, T. Y. W.; Yi Hao, T.; Loh, K. P.; Chowdari, B. V. R. Solid State Ionics: Ionics for Sustainable World, Proceedings of the 13th Asian Conference; Chowdari, B. V. R., Kawamura, J., Mizusaki, J., Eds.; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2012; p 265. (271) Moretti, A.; Kim, G. T.; Bresser, D.; Renger, K.; Paillard, E.; Marassi, R.; Winter, M.; Passerini, S. J. Power Sources 2013, 221, 419. (272) Ren, Y.; Hardwick, L. J.; Bruce, P. G. Angew. Chem., Int. Ed. 2010, 49, 2570. (273) Serventi, A. M.; Rodrigues, I. R.; Trudeau, M. L.; Antonelli, D.; Zaghib, K. J. Power Sources 2012, 202, 357. (274) Xu, R.; Li, J. R.; Tang, Z. L.; Zhang, Z. T. Electrochim. Acta 2011, 56, 6330. (275) Wang, H. E.; Cheng, H.; Liu, C. P.; Chen, X.; Jiang, Q. L.; Lu, Z. G.; Li, Y. Y.; Chung, C. Y.; Zhang, W. Y.; Zapien, J. A.; Martinu, L.; Bello, I. J. Power Sources 2011, 196, 6394. (276) Jiang, Y. M.; Wang, K. X.; Guo, X. X.; Wei, X.; Wang, J. F.; Chen, J. S. J. Power Sources 2012, 214, 298. (277) Ryu, W. H.; Nam, D. H.; Ko, Y. S.; Kim, R. H.; Kwon, H. S. Electrochim. Acta 2012, 61, 19. (278) Panda, S. K.; Yoon, Y.; Jung, H. S.; Yoon, W. S.; Shin, H. J. Power Sources 2012, 204, 162. (279) Ortiz, G. F.; Hanzu, I.; Lavela, P.; Tirado, J. L.; Knauth, P.; Djenizian, T. Int. J. Nanotechnol. 2012, 9, 260. (280) Gonzalez, J. R.; Alcantara, R.; Nacimiento, F.; Ortiz, G. F.; Tirado, J. L.; Zhecheva, E.; Stoyanova, R. J. Phys. Chem. C 2012, 116, 20182. 5442

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(281) Han, H.; Song, T.; Lee, E. K.; Devadoss, A.; Jeon, Y.; Ha, J.; Chung, Y. C.; Choi, Y. M.; Jung, Y. G.; Paik, U. ACS Nano 2012, 6, 8308. (282) Li, H. Q.; Martha, S. K.; Unocic, R. R.; Luo, H. M.; Dai, S.; Qu, J. J. Power Sources 2012, 218, 88. (283) Yang, Z. X.; Du, G. D.; Meng, Q.; Guo, Z. P.; Yu, X. B.; Chen, Z. X.; Guo, T. L.; Zeng, R. J. Mater. Chem. 2012, 22, 5848. (284) Chang, P. Y.; Huang, C. H.; Doong, R. A. Carbon 2012, 50, 4259. (285) Qiu, J. X.; Zhang, P.; Ling, M.; Li, S.; Liu, P. R.; Zhao, H. J.; Zhang, S. Q. ACS Appl. Mater. Interfaces 2012, 4, 3636. (286) Cao, H. Q.; Li, B. J.; Zhang, J. X.; Lian, F.; Kong, X. H.; Qu, M. Z. J. Mater. Chem. 2012, 22, 9759. (287) Tao, H. C.; Fan, L. Z.; Yan, X. Q.; Qu, X. H. Electrochim. Acta 2012, 69, 328. (288) Shah, M.; Park, A. R.; Zhang, K.; Park, J. H.; Yoo, P. J. ACS Appl. Mater. Interfaces 2012, 4, 3893. (289) Zhang, D. Q.; Wen, M. C.; Zhang, P.; Zhu, J.; Li, G. S.; Li, H. X. Langmuir 2012, 28, 4543. (290) Kyeremateng, N. A.; Hornebecq, V.; Knauth, P.; Djenizian, T. Electrochim. Acta 2012, 62, 192. (291) Xiong, H.; Yildirim, H.; Shevchenko, E. V.; Prakapenka, V. B.; Koo, B.; Slater, M. D.; Balasubramanian, M.; Sankaranarayanan, S.; Greeley, J. P.; Tepavcevic, S.; Dimitrijevic, N. M.; Podsiadlo, P.; Johnson, C. S.; Rajh, T. J. Phys. Chem. C 2012, 116, 3181. (292) Wang, W.; Tian, M.; Abdulagatov, A.; George, S. M.; Lee, Y. C.; Yang, R. G. Nano Lett. 2012, 12, 655. (293) Xiao, P. F.; Lai, M. O.; Lu, L. Electrochim. Acta 2012, 76, 185. (294) Hong, Z. S.; Xu, Y. X.; Liu, Y. B.; Wei, M. D. Chem.Eur. J. 2012, 18, 10753. (295) Barreca, D.; Carraro, G.; Gasparotto, A.; Maccato, C.; CruzYusta, M.; Gomez-Camer, J. L.; Morales, J.; Sada, C.; Sanchez, L. ACS Appl. Mater. Interfaces 2012, 4, 3610. (296) Belak, A. A.; Wang, Y. Z.; Van der Ven, A. Chem. Mater. 2012, 24, 2894. (297) Olson, C. L.; Nelson, J.; Islam, M. S. J. Phys. Chem. B 2006, 110, 9995. (298) Myung, S. T.; Takahashi, N.; Komaba, S.; Yoon, C. S.; Sun, Y. K.; Amine, K.; Yashiro, H. Adv. Funct. Mater. 2011, 21, 3231. (299) Lutzenkirchen-Hecht, D.; Wagemaker, M.; Keil, P.; van Well, A. A.; Frahm, R. Surf. Sci. 2003, 538, 10. (300) Borghols, W. J. H.; Wagemaker, M.; Lafont, U.; Kelder, E. M.; Mulder, F. M. Chem. Mater. 2008, 20, 2949. (301) Kubiak, P.; Pfanzelt, M.; Geserick, J.; Hormann, U.; Husing, N.; Kaiser, U.; Wohlfahrt-Mehrens, M. J. Power Sources 2009, 194, 1099. (302) Pfanzelt, M.; Kubiak, P.; Wohlfahrt-Mehrens, M. Electrochem. Solid-State Lett. 2010, 13, A91. (303) Marinaro, M.; Pfanzelt, M.; Kubiak, P.; Marassi, R.; WohlfahrtMehrens, M. J. Power Sources 2011, 196, 9825. (304) Pfanzelt, M.; Kubiak, P.; Fleischhammer, M.; WohlfahrtMehrens, M. J. Power Sources 2011, 196, 6815. (305) Bach, S.; Pereira-Ramos, J. P.; Willman, P. Electrochim. Acta 2010, 55, 4952. (306) Hu, Y. S.; Kienle, L.; Guo, Y. G.; Maier, J. Adv. Mater. 2006, 18, 1421. (307) Jiang, C. H.; Honma, I.; Kudo, T.; Zhou, H. S. Electrochem. Solid-State Lett. 2007, 10, A127. (308) Baudrin, E.; Cassaignon, S.; Koesch, M.; Jolivet, J. P.; Dupont, L.; Tarascon, J. M. Electrochem. Commun. 2007, 9, 337. (309) Zhao, B.; Cai, R.; Jiang, S.; Sha, Y.; Shao, Z. Electrochim. Acta 2012, 85, 636. (310) Bach, S.; Pereira-Ramos, J. P.; Willman, P. Electrochim. Acta 2010, 55, 4952. (311) Khomane, R. B. J. Colloid Interface Sci. 2011, 356, 369. (312) Armstrong, A. R.; Armstrong, G.; Canales, J.; Bruce, P. G. J. Power Sources 2005, 146, 501. (313) Zukalova, M.; Kalbac, M.; Kavan, L.; Exnar, I.; Graetzel, M. Chem. Mater. 2005, 17, 1248.

(314) Okumura, T.; Fukutsuka, T.; Yanagihara, A.; Orikasa, Y.; Arai, H.; Ogumi, Z.; Uchimoto, Y. J. Mater. Chem. 2011, 21, 15369. (315) Okumura, T.; Fukutsuka, T.; Yanagihara, A.; Orikasa, Y.; Arai, H.; Ogumi, Z.; Uchimoto, Y. Chem. Mater. 2011, 23, 3636. (316) Liu, H. S.; Bi, Z. H.; Sun, X. G.; Unocic, R. R.; Paranthaman, M. P.; Dai, S.; Brown, G. M. Adv. Mater. 2011, 23, 3450. (317) Akimoto, J.; Chiba, K.; Kijima, N.; Hayakawa, H.; Hayashi, S.; Gotoh, Y.; Idemoto, Y. J. Electrochem. Soc. 2011, 158, A546. (318) Perez-Flores, J. C.; Kuhn, A.; Garcia-Alvarado, F. J. Power Sources 2011, 196, 1378. (319) Perez-Flores, J. C.; Baehtz, C.; Hoelzel, M.; Kuhn, A.; GarciaAlvarado, F. Phys. Chem. Chem. Phys. 2012, 14, 2892. (320) Cho, W.; Kashiwagi, T.; Ra, W.; Nakayama, M.; Wakihara, M.; Kobayashi, Y.; Miyashiro, H. Electrochim. Acta 2009, 54, 1842. (321) Cho, W.; Park, M. S.; Kim, J. H.; Kim, Y. J. Electrochim. Acta 2012, 63, 263. (322) Arrouvel, C.; Parker, S. C.; Islam, M. S. Chem. Mater. 2009, 21, 4778. (323) Armstrong, A. R.; Arrouvel, C.; Gentili, V.; Parker, S. C.; Islam, M. S.; Bruce, P. G. Chem. Mater. 2010, 22, 6426. (324) Saito, M.; Murota, Y.; Takagi, M.; Tajima, M.; Asao, T.; Inoue, H.; Tasaka, A.; Inaba, M. J. Electrochem. Soc. 2012, 159, A49. (325) Brutti, S.; Gentili, V.; Menard, H.; Scrosati, B.; Bruce, P. G. Adv. Energy Mater. 2012, 2, 322. (326) Wessel, C.; Zhao, L. A.; Urban, S.; Ostermann, R.; Djerdj, I.; Smarsly, B. M.; Chen, L. Q.; Hu, Y. S.; Sallard, S. Chem.Eur. J. 2011, 17, 775. (327) Ren, Y.; Liu, Z.; Pourpoint, F.; Armstrong, A. R.; Grey, C. P.; Bruce, P. G. Angew. Chem. 2012, 124, 2206. (328) Jang, H.; Suzuki, S.; Miyayama, M. J. Power Sources 2012, 203, 97. (329) Yang, Z. X.; Du, G. D.; Guo, Z. P.; Yu, X. B.; Chen, Z. X.; Guo, T. L.; Liu, H. K. J. Mater. Chem. 2011, 21, 8591. (330) Yang, Z. X.; Du, G. D.; Guo, Z. P.; Yu, X. B.; Chen, Z. X.; Guo, T. L.; Sharma, N.; Liu, H. K. Electrochem. Commun. 2011, 13, 46. (331) Li, J. M.; Wan, W.; Zhou, H. H.; Li, J. J.; Xu, D. S. Chem. Commun. 2011, 47, 3439. (332) Dylla, A. G.; Lee, J. A.; Stevenson, K. J. Langmuir 2012, 28, 2897. (333) Liu, S. H.; Jia, H. P.; Han, L.; Wang, J. L.; Gao, P. F.; Xu, D. D.; Yang, J.; Che, S. N. Adv. Mater. 2012, 24, 3201. (334) Reddy, M. A.; Pralong, V.; Varadaraju, U. V.; Raveau, B. Electrochem. Solid-State Lett. 2008, 11, A132. (335) Saito, M.; Murota, Y.; Takagi, M.; Tajima, M.; Asao, T.; Inoue, H.; Tasaka, A.; Inaba, M. J. Electrochem. Soc. 2012, 159, A49. (336) Guo, Z. Y.; Dong, X. L.; Zhou, D. D.; Du, Y. J.; Wang, Y. G.; Xia, Y. Y. RSC Adv. 2013, 3, 3352. (337) Grey, C. P.; Dupre, N. Chem. Rev. 2004, 104, 4493. (338) Baddour-Hadjean, R.; Navone, C.; Pereira-Ramos, J. P. Electrochim. Acta 2009, 54, 6674. (339) Sakunthala, A.; Reddy, M. V.; Selvasekarapandian, S.; Chowdari, B. V. R.; Selvin, P. C. Energy Environ. Sci. 2011, 4, 1712. (340) Sathiya, M.; Prakash, A. S.; Ramesha, K.; Tarascon, J. M.; Shukla, A. K. J. Am. Chem. Soc. 2011, 133, 16291. (341) Dewangan, K.; Sinha, N. N.; Chavan, P. G.; Sharma, P. K.; Pandey, A. C.; More, M. A.; Joag, D. S.; Munichandraiah, N.; Gajbhiye, N. S. Nanoscale 2012, 4, 645. (342) Semenenko, D. A.; Kozmenkova, A. Y.; Itkis, D. M.; Goodilin, E. A.; Kulova, T. L.; Skundin, A. M.; Tretyakov, Y. D. CrystEngComm 2012, 14, 1561. (343) Wu, Y.; Zhu, P.; Zhao, X.; Reddy, M. V.; Peng, S.; Chowdari, B. V. R.; Ramakrishna, S. J. Mater. Chem. A 2013, 1, 852. (344) Wei, X.; Jiao, L. F.; Sun, J. L.; Liu, S. C.; Yuan, H. T. J. Solid State Electrochem. 2010, 14, 197. (345) Chen, J. S.; Cheah, Y. L.; Madhavi, S.; Lou, X. W. J. Phys. Chem. C 2010, 114, 8675. (346) Kumagai, N.; Koishikawa, Y.; Komaba, S.; Koshiba, N. J. Electrochem. Soc. 1999, 146, 3203. 5443

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(347) Lantelme, F.; Groult, H.; Kumagai, N. Electrochim. Acta 2000, 45, 3171. (348) Kodama, R.; Terada, Y.; Nakai, I.; Komaba, S.; Kumagai, N. J. Electrochem. Soc. 2006, 153, A583. (349) Wei, M. D.; Wei, K.; Ichihara, M.; Zhou, H. Electrochem. Commun. 2008, 10, 980. (350) Le Viet, A.; Reddy, M. V.; Jose, R.; Chowdari, B. V. R.; Ramakrishna, S. J. Phys. Chem. C 2010, 114, 664. (351) Viet, A. L.; Reddy, M. V.; Jose, R.; Chowdari, B. V. R.; Ramakrishna, S. Electrochim. Acta 2011, 56, 1518. (352) Kumagai, N.; Tateshita, Y.; Takatsuka, Y.; Baba, M.; Ikeda, T.; Tanno, K. J. Power Sources 1995, 54, 175. (353) Nakazawa, H.; Sano, K.; Abe, T.; Baba, M.; Kumagai, N. J. Power Sources 2007, 174, 838. (354) Wen, H.; Liu, Z. F.; Wang, J.; Yang, Q. B.; Li, Y. X.; Yu, J. Appl. Surf. Sci. 2011, 257, 10084. (355) Pralong, V. Prog. Solid State Chem. 2009, 37, 262. (356) Ariyoshi, K.; Yamato, R.; Ohzuku, T. Electrochim. Acta 2005, 51, 1125. (357) Zhu, N.; Liu, W.; Xue, M. Q.; Xie, Z. A.; Zhao, D.; Zhang, M. N.; Chen, J. T.; Cao, T. B. Electrochim. Acta 2010, 55, 5813. (358) Kim, H.-K.; Bak, S.-M.; Kim, K.-B. Electrochem. Commun. 2010, 12, 1768. (359) Jung, H. G.; Kim, J.; Scrosati, B.; Sun, Y. K. J. Power Sources 2011, 196, 7763. (360) Jung, H. G.; Myung, S. T.; Yoon, C. S.; Son, S. B.; Oh, K. H.; Amine, K.; Scrosati, B.; Sun, Y. K. Energy Environ. Sci. 2011, 4, 1345. (361) Zhao, L.; Hu, Y. S.; Li, H.; Wang, Z. X.; Chen, L. Q. Adv. Mater. 2011, 23, 1385. (362) Shen, L.; Yuan, C.; Luo, H.; Zhang, X.; Chen, L.; Li, H. J. Mater. Chem. 2011, 21, 14414. (363) Schneider, H.; Maire, P.; Novák, P. Electrochim. Acta 2011, 56, 9324. (364) Wagemaker, M.; Simon, D. R.; Kelder, E. M.; Schoonman, J.; Ringpfeil, C.; Haake, U.; Lutzenkirchen-Hecht, D.; Frahm, R.; Mulder, F. M. Adv. Mater. 2006, 18, 3169. (365) Borghols, W. J. H.; Wagemaker, M.; Lafont, U.; Kelder, E. M.; Mulder, F. M. J. Am. Chem. Soc. 2009, 131, 17786. (366) Wagemaker, M.; van Eck, E. R. H.; Kentgens, A. P. M.; Mulder, F. M. J. Phys. Chem. B 2009, 113, 224. (367) Prakash, A. S.; Manikandan, P.; Ramesha, K.; Sathiya, M.; Tarascon, J. M.; Shukla, A. K. Chem. Mater. 2010, 22, 2857. (368) Tian, B. B.; Xiang, H. F.; Zhang, L.; Li, Z.; Wang, H. H. Electrochim. Acta 2010, 55, 5453. (369) Yi, T. F.; Xie, Y.; Shu, J.; Wang, Z. H.; Yue, C. B.; Zhu, R. S.; Qiao, H. B. J. Electrochem. Soc. 2011, 158, A266. (370) Gu, F.; Chen, G.; Wang, Z. H. J. Solid State Electrochem. 2012, 16, 375. (371) Yu, Z. J.; Zhang, X. F.; Yang, G. L.; Liu, J.; Wang, J. W.; Wang, R. S.; Zhang, J. P. Electrochim. Acta 2011, 56, 8611. (372) Zhang, B. A.; Huang, Z. D.; Oh, S. W.; Kim, J. K. J. Power Sources 2011, 196, 10692. (373) Chen, J. Z.; Yang, L.; Fang, S. H.; Tang, Y. F. Electrochim. Acta 2010, 55, 6596. (374) Lin, C.-Y.; Duh, J.-G. J. Alloys Compd. 2011, 509, 3682. (375) Yu, S. H.; Pucci, A.; Herntrich, T.; Willinger, M. G.; Baek, S. H.; Sung, Y. E.; Pinna, N. J. Mater. Chem. 2011, 21, 806. (376) Feckl, J. M.; Fominykh, K.; Doblinger, M.; FattakhovaRohlfing, D.; Bein, T. Angew. Chem., Int. Ed. 2012, 51, 7459. (377) Lu, J.; Nan, C. Y.; Peng, Q.; Li, Y. D. J. Power Sources 2012, 202, 246. (378) Lin, Y. S.; Tsai, M. C.; Duh, J. G. J. Power Sources 2012, 214, 314. (379) Li, X.; Lai, C.; Xiao, C. W.; Gao, X. P. Electrochim. Acta 2011, 56, 9152. (380) Rahman, M. M.; Wang, J. Z.; Hassan, M. F.; Wexler, D.; Liu, H. K. Adv. Energy Mater. 2011, 1, 212. (381) Chou, S. L.; Wang, J. Z.; Liu, H. K.; Dou, S. X. J. Phys. Chem. C 2011, 115, 16220.

(382) Zhang, A.; Zheng, Z. M.; Cheng, F. Y.; Tao, Z. L.; Chen, J. Sci. China: Chem. 2011, 54, 936. (383) Chen, J. Z.; Yang, L.; Fang, S. H.; Hirano, S.; Tachibana, K. J. Power Sources 2012, 200, 59. (384) Lin, Y. S.; Duh, J. G. J. Power Sources 2011, 196, 10698. (385) Mosa, J.; Velez, J. F.; Lorite, I.; Arconada, N.; Aparicio, M. J. Power Sources 2012, 205, 491. (386) Wu, H. C.; Lee, E.; Wu, N. L.; Jow, T. R. J. Power Sources 2012, 197, 301. (387) Guo, B. K.; Li, Y.; Yao, Y. F.; Lin, Z.; Ji, L. W.; Xu, G. J.; Liang, Y. Z.; Shi, Q.; Zhang, X. W. Solid State Ionics 2011, 204, 61. (388) Wu, Y. Z.; Reddy, M. V.; Chowdari, B. V. R.; Ramakrishna, S. Electrochim. Acta 2012, 67, 33. (389) Laumann, A.; Bremholm, M.; Hald, P.; Holzapfel, M.; Fehr, K. T.; Iversen, B. B. J. Electrochem. Soc. 2012, 159, A166. (390) Hsieh, C. T.; Chang, B. S.; Lin, J. Y.; Juang, R. S. J. Alloys Compd. 2012, 513, 393. (391) He, Z. J.; Wang, Z. X.; Wu, F. X.; Guo, H. J.; Li, X. H.; Xiong, X. H. J. Alloys Compd. 2012, 540, 39. (392) Cheng, L.; Yan, J.; Zhu, G.-N.; Luo, J.-Y.; Wang, C.-X.; Xia, Y.Y. J. Mater. Chem. 2010, 20, 595. (393) Wang, G. X.; Yan, K. P.; Yu, Z. L.; Qu, M. Z. J. Appl. Electrochem. 2010, 40, 821. (394) Zhu, G. N.; Liu, H. J.; Zhuang, J. H.; Wang, C. X.; Wang, Y. G.; Xia, Y. Y. Energy Environ. Sci. 2011, 4, 4016. (395) Hu, X. B.; Lin, Z. J.; Yang, K. R.; Huai, Y. J.; Deng, Z. H. Electrochim. Acta 2011, 56, 5046. (396) Li, X.; Qu, M. Z.; Huai, Y. J.; Yu, Z. L. Electrochim. Acta 2010, 55, 2978. (397) Kang, E.; Jung, Y. S.; Kim, G. H.; Chun, J.; Wiesner, U.; Dillon, A. C.; Kim, J. K.; Lee, J. Adv. Funct. Mater. 2011, 21, 4349. (398) He, Y. B.; Ning, F.; Li, B. H.; Song, Q. S.; Lv, W.; Du, H. D.; Zhai, D. Y.; Su, F. Y.; Yang, Q. H.; Kang, F. Y. J. Power Sources 2012, 202, 253. (399) Ogihara, T.; Yamada, M.; Fujita, A.; Akao, S.; Myoujin, K. Mater. Res. Bull. 2011, 46, 796. (400) Xie, G. W.; Ni, J. F.; Liao, X. F.; Gao, L. J. Mater. Lett. 2012, 78, 177. (401) Li, B. H.; Han, C. P.; He, Y. B.; Yang, C.; Du, H. D.; Yang, Q. H.; Kang, F. Y. Energy & Environ. Sci. 2012, 5, 9595. (402) Zhang, B. A.; Liu, Y. S.; Huang, Z. D.; Oh, S.; Yu, Y.; Mai, Y. W.; Kim, J. K. J. Mater. Chem. 2012, 22, 12133. (403) Dominko, R.; Gaberscek, M.; Bele, A.; Mihailovic, D.; Jamnik, J. J. Eur. Ceram. Soc. 2007, 27, 909. (404) Jhan, Y. R.; Duh, J. G. J. Power Sources 2012, 198, 294. (405) Shen, L. F.; Yuan, C. Z.; Luo, H. J.; Zhang, X. G.; Xu, K.; Zhang, F. J. Mater. Chem. 2011, 21, 761. (406) Ni, H. F.; Fan, L. Z. J. Power Sources 2012, 214, 195. (407) Shi, Y.; Wen, L.; Li, F.; Cheng, H. M. J. Power Sources 2011, 196, 8610. (408) Xiang, H. F.; Tian, B. B.; Lian, P. C.; Li, Z.; Wang, H. H. J. Alloys Compd. 2011, 509, 7205. (409) Han, S. Y.; Kim, I. Y.; Jo, K. Y.; Hwang, S. J. J. Phys. Chem. C 2012, 116, 7269. (410) Jian, Z. L.; Zhao, L.; Wang, R.; Hu, Y. S.; Li, H.; Chen, W.; Chen, L. Q. RSC Adv. 2012, 2, 1751. (411) Tang, Y. F.; Huang, F. Q.; Zhao, W.; Liu, Z. Q.; Wan, D. Y. J. Mater. Chem. 2012, 22, 11257. (412) Wilkening, M.; Amade, R.; Iwaniak, W.; Heitjans, P. Phys. Chem. Chem. Phys. 2007, 9, 1239. (413) Vijayakumar, M.; Kerisit, S.; Rosso, K. M.; Burton, S. D.; Sears, J. A.; Yang, Z. G.; Graff, G. L.; Liu, J.; Hu, J. Z. J. Power Sources 2011, 196, 2211. (414) Hain, H.; Scheuermann, M.; Heinzmann, R.; Wunsche, L.; Hahn, H.; Indris, S. Solid State Nucl. Magn. Reson. 2012, 42, 9. (415) Rho, Y. H.; Kanamura, K. J. Solid State Chem. 2004, 177, 2094. (416) Wunde, F.; Berkemeier, F.; Schmitz, G. J. Power Sources 2012, 215, 109. 5444

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(417) Peramunage, D.; Abraham, K. M. J. Electrochem. Soc. 1998, 145, 2609. (418) Zaghib, K.; Simoneau, M.; Armand, M.; Gauthier, M. J. Power Sources 1999, 81, 300. (419) Chen, C. H.; Vaughey, J. T.; Jansen, A. N.; Dees, D. W.; Kahaian, A. J.; Goacher, T.; Thackeray, M. M. J. Electrochem. Soc. 2001, 148, A102. (420) Nakahara, K.; Nakajima, R.; Matsushima, T.; Majima, H. J. Power Sources 2003, 117, 131. (421) Li, J. R.; Tang, Z. L.; Zhang, Z. T. Electrochem. Commun. 2005, 7, 894. (422) Yu, Y.; Shui, J. L.; Chen, C. H. Solid State Commun. 2005, 135, 485. (423) Lu, H. W.; Zeng, W.; Li, Y. S.; Fu, Z. W. J. Power Sources 2007, 164, 874. (424) Gao, J.; Ying, J. R.; Jiang, C. Y.; Wan, C. R. J. Power Sources 2007, 166, 255. (425) Kim, J.; Cho, J. Electrochem. Solid-State Lett. 2007, 10, A81. (426) Tang, Y. F.; Yang, L.; Qiu, Z.; Huang, J. S. Electrochem. Commun. 2008, 10, 1513. (427) Tang, Y. F.; Yang, L.; Qiu, Z.; Huang, J. S. J. Mater. Chem. 2009, 19, 5980. (428) Shen, L. F.; Yuan, C. Z.; Luo, H. J.; Zhang, X. G.; Xu, K.; Xia, Y. Y. J. Mater. Chem. 2010, 20, 6998. (429) Zhu, N.; Liu, W.; Xue, M. Q.; Xie, Z. A.; Zhao, D.; Zhang, M. N.; Chen, J. T.; Cao, T. B. Electrochim. Acta 2010, 55, 5813. (430) Izumi, A.; Sanada, M.; Furuichi, K.; Teraki, K.; Matsuda, T.; Hiramatsu, K.; Munakata, H.; Kanamura, K. Electrochim. Acta 2012, 79, 218. (431) Nugroho, A.; Kim, S. J.; Chung, K. Y.; Kim, J. Electrochim. Acta 2012, 78, 623. (432) Anji Reddy, M.; Kishore, M. S.; Pralong, V.; Caignaert, V.; Varadaraju, U. V.; Raveau, B. Chem. Mater. 2008, 20, 2192. (433) Colin, J. F.; Pralong, V.; Caignaert, V.; Hervieu, M.; Raveau, B. Inorg. Chem. 2006, 45, 7217. (434) Patoux, S.; Masquelier, C. Chem. Mater. 2002, 14, 5057. (435) Shi, Z. C.; Wang, Q.; Ye, W. L.; Li, Y. X.; Yang, Y. Microporous Mesoporous Mater. 2006, 88, 232. (436) Uebou, Y.; Okada, S.; Egashira, M.; Yamaki, J. I. Solid State Ionics 2002, 148, 323. (437) Satya Kishore, M.; Pralong, V.; Caignaert, V.; Varadaraju, U. V.; Raveau, B. J. Power Sources 2007, 169, 355. (438) Delmas, C.; Nadiri, A.; Soubeyroux, J. L. Solid State Ionics 1988, 28, 419. (439) Aatiq, A.; Menetrier, M.; Croguennec, L.; Suard, E.; Delmas, C. J. Mater. Chem. 2002, 12, 2971. (440) Liu, X. H.; Saito, T.; Doi, T.; Okada, S.; Yamaki, J. J. Power Sources 2009, 189, 706. (441) Wessells, C.; Huggins, R. A.; Cui, Y. J. Power Sources 2011, 196, 2884. (442) Aravindan, V.; Chuiling, W.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R.; Madhavi, S. Phys. Chem. Chem. Phys. 2012, 14, 5808. (443) Anji Reddy, M.; Varadaraju, U. V. Chem. Mater. 2008, 20, 4557. (444) Patoux, S.; Dolle, M.; Rousse, G.; Masquelier, C. J. Electrochem. Soc. 2002, 149, A391. (445) Saritha, D.; Pralong, V.; Varadaraju, U. V.; Raveau, B. J. Solid State Chem. 2010, 183, 988. (446) Colin, J. F.; Pralong, V.; Hervieu, M.; Caignaert, V.; Raveau, B. Chem. Mater. 2008, 20, 1534. (447) Pralong, V.; Reddy, M. A.; Caignaert, V.; Malo, S.; Lebedev, O. I.; Varadaraju, U. V.; Raveau, B. Chem. Mater. 2011, 23, 1915. (448) Reddy, M. A.; Varadaraju, U. V. J. Phys. Chem. C 2011, 115, 25121. (449) Han, J.-T.; Huang, Y.-H.; Goodenough, J. B. Chem. Mater. 2011, 23, 2027. (450) Besenhard, J. O.; Wachtler, M.; Winter, M.; Andreaus, R.; Rom, I.; Sitte, W. J. Power Sources 1999, 81, 268.

(451) Zhang, W. J. J. Power Sources 2011, 196, 877. (452) Courtney, I. A.; Dahn, J. R. J. Electrochem. Soc. 1997, 144, 2943. (453) Courtney, I. A.; Dahn, J. R. J. Electrochem. Soc. 1997, 144, 2045. (454) Courtney, I. A.; McKinnon, W. R.; Dahn, J. R. J. Electrochem. Soc. 1999, 146, 59. (455) Sakamoto, J. S.; Huang, C. K.; Surampudi, S.; Smart, M.; Wolfenstine, J. Mater. Lett. 1998, 33, 327. (456) Branci, C.; Benjelloun, N.; Sarradin, J.; Ribes, M. Solid State Ionics 2000, 135, 169. (457) Chouvin, J.; Branci, C.; Sarradin, J.; Olivier-Fourcade, J.; Jumas, J. C.; Simon, B.; Biensan, P. J. Power Sources 1999, 81, 277. (458) Chouvin, J.; Olivier-Fourcade, J.; Jumas, J. C.; Simon, B.; Biensan, P. H.; Madrigal, F. J. F.; Tirado, J. L.; Vicente, C. P. J. Electroanal. Chem. 2000, 494, 136. (459) Aurbach, D.; Nimberger, A.; Markovsky, B.; Levi, E.; Sominski, E.; Gedanken, A. Chem. Mater. 2002, 14, 4155. (460) Uchiyama, H.; Hosono, E.; Honma, I.; Zhou, H.; Imai, H. Electrochem. Commun. 2008, 10, 52. (461) Yang, J.; Takeda, Y.; Imanishi, N.; Xie, J. Y.; Yamamoto, O. J. Power Sources 2001, 97−98, 216. (462) Ning, J. J.; Dai, Q. Q.; Jiang, T.; Men, K. K.; Liu, D. H.; Xiao, N. R.; Li, C. Y.; Li, D. M.; Liu, B. B.; Zou, B.; Zou, G. T.; Yu, W. W. Langmuir 2009, 25, 1818. (463) Ning, J. J.; Jiang, T.; Men, K. K.; Dai, Q. Q.; Li, D. M.; Wei, Y. J.; Liu, B. B.; Chen, G.; Zou, B.; Zou, G. T. J. Phys. Chem. C 2009, 113, 14140. (464) Li, H.; Huang, X. J.; Chen, L. Q. Solid State Ionics 1999, 123, 189. (465) Wang, X. Y.; Wen, Z. Y.; Yang, X. L.; Lin, B. Solid State Ionics 2008, 179, 1238. (466) Ortiz, G. F.; Hanzu, I.; Lavela, P.; Knauth, P.; Tirado, J. L.; Djenizian, T. Chem. Mater. 2010, 22, 1926. (467) Das, B.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Solid State Electrochem. 2011, 15, 259. (468) Xie, J.; Imanishi, N.; Hirano, A.; Takeda, Y.; Yamamoto, O.; Zhao, X. B.; Cao, G. S. Solid State Ionics 2010, 181, 1611. (469) Chen, M. H.; Huang, Z. C.; Wu, G. T.; Zhu, G. M.; You, J. K.; Lin, Z. G. Mater. Res. Bull. 2003, 38, 831. (470) Sakaushi, K.; Oaki, Y.; Uchiyama, H.; Hosono, E.; Zhou, H. S.; Imai, H. Small 2010, 6, 776. (471) Machill, S.; Shodai, T.; Sakurai, Y.; Yamaki, J. J. Power Sources 1998, 73, 216. (472) Ortiz, G. F.; Hanzu, I.; Knauth, P.; Lavela, P.; Tirado, J. L.; Djenizian, T. Electrochem. Solid-State Lett. 2009, 12, A186. (473) Iqbal, M. Z.; Wang, F.; Zhao, H.; Rafique, M. Y.; Wang, J.; Li, Q. Scripta Mater. 2012, 67, 665. (474) Cherian, C. T.; Reddy, M. V.; Sow, C. H.; Chowdari, B. V. R. RSC Adv. 2013, 3, 3118. (475) Retoux, R.; Brousse, T.; Schleich, D. M. J. Electrochem. Soc. 1999, 146, 2472. (476) Huang, J. Y.; Zhong, L.; Wang, C. M.; Sullivan, J. P.; Xu, W.; Zhang, L. Q.; Mao, S. X.; Hudak, N. S.; Liu, X. H.; Subramanian, A.; Fan, H. Y.; Qi, L. A.; Kushima, A.; Li, J. Science 2010, 330, 1515. (477) Zhong, L.; Liu, X. H.; Wang, G. F.; Mao, S. X.; Huang, J. Y. Phys. Rev. Lett. 2011, 106, No. 248302. (478) Liu, X. H.; Huang, J. Y. Energy Environ. Sci. 2011, 4, 3844. (479) Belliard, F.; Connor, P. A.; Irvine, J. T. S. Solid State Ionics 2000, 135, 163. (480) Subramanian, V.; Jiang, J. C.; Smith, P. H.; Rambabu, B. J. Nanosci. Nanotechnol. 2004, 4, 125. (481) Subramanian, V.; Gnanasekar, K. I.; Rambabu, B. Solid State Ionics 2004, 175, 181. (482) Subramanian, V.; Burke, W. W.; Zhu, H. W.; Wei, B. Q. J. Phys. Chem. C 2008, 112, 4550. (483) Park, M. S.; Wang, G. X.; Kang, Y. M.; Wexler, D.; Dou, S. X.; Liu, H. K. Angew. Chem., Int. Ed. 2007, 46, 750. (484) Park, M. S.; Kang, Y. M.; Wang, G. X.; Dou, S. X.; Liu, H. K. Adv. Funct. Mater. 2008, 18, 455. (485) Wang, Y.; Lee, J. Y. J. Phys. Chem. B 2004, 108, 17832. 5445

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(521) Chou, S. L.; Wang, J. Z.; Liu, H. K.; Dou, S. X. Electrochem. Commun. 2009, 11, 242. (522) Nam, S. C.; Paik, C. H.; Cho, W. I.; Cho, B. W.; Chun, H. S.; Yun, K. S. J. Power Sources 1999, 84, 24. (523) Zhu, X. J.; Guo, Z. P.; Zhang, P.; Du, G. D.; Zeng, R.; Chen, Z. X.; Li, S.; Liu, H. K. J. Mater. Chem. 2009, 19, 8360. (524) Zhu, X. J.; Guo, Z. P.; Zhang, P.; Du, G. D.; Poh, C. K.; Chen, Z. X.; Li, S.; Liu, H. K. Electrochim. Acta 2010, 55, 4982. (525) Kim, T. J.; Son, D.; Cho, J.; Park, B.; Yang, H. Electrochim. Acta 2004, 49, 4405. (526) Lee, W. J.; Park, M. H.; Wang, Y.; Lee, J. Y.; Cho, J. Chem. Commun. 2010, 46, 622. (527) Kim, J. H.; Jeong, G. J.; Kim, Y. W.; Sohn, H. J.; Park, C. W.; Lee, C. K. J. Electrochem. Soc. 2003, 150, A1544. (528) Yin, X. M.; Chen, L. B.; Li, C. C.; Hao, Q. Y.; Liu, S. A.; Li, Q. H.; Zhang, E. D.; Wang, T. H. Electrochim. Acta 2011, 56, 2358. (529) Ortiz, G. F.; Lavela, P.; Knauth, P.; Djenizian, T.; Alcantara, R.; Tirado, J. L. J. Electrochem. Soc. 2011, 158, A1094. (530) Hassan, M. F.; Rahman, M. M.; Guo, Z. P.; Chen, Z. X.; Liu, H. K. J. Mater. Chem. 2010, 20, 9707. (531) Yang, Z. X.; Du, G. D.; Feng, C. Q.; Li, S. A.; Chen, Z. X.; Zhang, P.; Guo, Z. P.; Yu, X. B.; Chen, G. N.; Huang, S. Z.; Liu, H. K. Electrochim. Acta 2010, 55, 5485. (532) Bonino, C. A.; Ji, L. W.; Lin, Z.; Toprakci, O.; Zhang, X. W.; Khan, S. A. ACS Appl. Mater. Interfaces 2011, 3, 2534. (533) Kong, J. H.; Wong, S. Y.; Zhang, Y.; Tan, H. R.; Li, X.; Lu, X. H. J. Mater. Chem. 2011, 21, 15928. (534) Mu, J. B.; Chen, B.; Guo, Z. C.; Zhang, M. Y.; Zhang, Z. Y.; Shao, C. L.; Liu, Y. C. J. Colloid Interface Sci. 2011, 356, 706. (535) Mi, H. Y.; Xu, Y. L.; Shi, W.; Yoo, H. D.; Park, S. J.; Park, Y. W.; Oh, S. M. J. Mater. Chem. 2011, 21, 19302. (536) Yim, C. H.; Baranova, E. A.; Courtel, F. M.; Abu-Lebdeh, Y.; Davidson, I. J. J. Power Sources 2011, 196, 9731. (537) Wang, J.; Zhao, H. L.; Liu, X. T.; Wang, C. M. Electrochim. Acta 2011, 56, 6441. (538) Li, M. Y.; Liu, C. L.; Wang, Y.; Dong, W. S. J. Electrochem. Soc. 2011, 158, A296. (539) Wang, H.; Wu, Y. M.; Bai, Y. S.; Zhou, W.; An, Y. R.; Li, J. H.; Guo, L. J. Mater. Chem. 2011, 21, 10189. (540) Yu, X. Y.; Yang, S. Y.; Zhang, B. H.; Shao, D.; Dong, X. M.; Fang, Y. P.; Li, Z. S.; Wang, H. Q. J. Mater. Chem. 2011, 21, 12295. (541) Zhou, W.; Cheng, C.; Liu, J. P.; Tay, Y. Y.; Jiang, J.; Jia, X. T.; Zhang, J. X.; Gong, H.; Hng, H. H.; Yu, T.; Fan, H. J. Adv. Funct. Mater. 2011, 21, 2439. (542) Zhu, J.; Lu, Z.; Oo, M. O.; Hng, H. H.; Ma, J.; Zhang, H.; Yan, Q. J. Mater. Chem. 2011, 21, 12770. (543) Zhu, X. J.; Zhu, Y. W.; Murali, S.; Stoller, M. D.; Ruoff, R. S. J. Power Sources 2011, 196, 6473. (544) Li, Y. M.; Lv, X. J.; Lu, J.; Li, J. H. J. Phys. Chem. C 2010, 114, 21770. (545) Zhao, B.; Zhang, G. H.; Song, J. S.; Jiang, Y.; Zhuang, H.; Liu, P.; Fang, T. Electrochim. Acta 2011, 56, 7340. (546) Zhong, C.; Wang, J. Z.; Chen, Z. X.; Liu, H. K. J. Phys. Chem. C 2011, 115, 25115. (547) Lian, P. C.; Zhu, X. F.; Liang, S. Z.; Li, Z.; Yang, W. S.; Wang, H. H. Electrochim. Acta 2011, 56, 4532. (548) Huang, X. D.; Zhou, X. F.; Zhou, L. A.; Qian, K.; Wang, Y. H.; Liu, Z. P.; Yu, C. Z. ChemPhysChem 2011, 12, 278. (549) Xie, J.; Liu, S. Y.; Chen, X. F.; Zheng, Y. X.; Song, W. T.; Cao, G. S.; Zhu, T. J.; Zhao, X. B. Int. J. Electrochem. Sci. 2011, 6, 5539. (550) Baek, S.; Yu, S. H.; Park, S. K.; Pucci, A.; Marichy, C.; Lee, D. C.; Sung, Y. E.; Piao, Y.; Pinna, N. RSC Adv. 2011, 1, 1687. (551) Wang, X. Y.; Zhou, X. F.; Yao, K.; Zhang, J. G.; Liu, Z. P. Carbon 2011, 49, 133. (552) Xu, C. H.; Sun, J.; Gao, L. J. Mater. Chem. 2012, 22, 975. (553) Wang, F.; Yao, G.; Xu, M. W.; Zhao, M. S.; Sun, Z. B.; Song, X. P. J. Alloys Compd. 2011, 509, 5969. (554) Di Lupo, F.; Gerbaldi, C.; Meligrana, G.; Bodoardo, S.; Penazzi, N. Int. J. Electrochem. Sci. 2011, 6, 3580.

(486) Liu, J. P.; Li, Y. Y.; Huang, X. T.; Ding, R. M.; Hu, Y. Y.; Jiang, J.; Liao, L. J. Mater. Chem. 2009, 19, 1859. (487) Guo, Z. P.; Du, G. D.; Nuli, Y.; Hassan, M. F.; Liu, H. K. J. Mater. Chem. 2009, 19, 3253. (488) Zhu, J. J.; Lu, Z. H.; Aruna, S. T.; Aurbach, D.; Gedanken, A. Chem. Mater. 2000, 12, 2557. (489) Kim, C.; Noh, M.; Choi, M.; Cho, J.; Park, B. Chem. Mater. 2005, 17, 3297. (490) Kwon, C. W.; Campet, G.; Portier, J.; Poquet, A.; Fournes, L.; Labrugere, C.; Jousseaume, B.; Toupance, T.; Choy, J. H.; Subramanian, M. A. Int. J. Inorg. Mater. 2001, 3, 211. (491) Alcantara, R.; Fernandez-Madrigal, F. J.; Perez-Vicente, C.; Tirado, J. L.; Jumas, J. C.; Olivier-Fourcade, J. Chem. Mater. 2000, 12, 3044. (492) Du, G. D.; Zhong, C.; Zhang, P.; Guo, Z. P.; Chen, Z. X.; Liu, H. K. Electrochim. Acta 2010, 55, 2582. (493) Zhu, C. L.; Zhang, M. L.; Qiao, Y. J.; Gao, P.; Chen, Y. J. Mater. Res. Bull. 2010, 45, 437. (494) Yang, S. B.; Song, H. H.; Yi, H. X.; Liu, W. X.; Zhang, H. J.; Chen, X. H. Electrochim. Acta 2009, 55, 521. (495) Wen, Z. H.; Wang, Q.; Zhang, Q.; Li, J. H. Adv. Funct. Mater. 2007, 17, 2772. (496) Chen, G.; Wang, Z. Y.; Xia, D. G. Chem. Mater. 2008, 20, 6951. (497) Wang, Y.; Zeng, H. C.; Lee, J. Y. Adv. Mater. 2006, 18, 645. (498) Li, C.; Wei, W.; Fang, S. M.; Wang, H. X.; Zhang, Y.; Gui, Y. H.; Chen, R. F. J. Power Sources 2010, 195, 2939. (499) Xu, W.; Canfield, N. L.; Wang, D. Y.; Xiao, J.; Nie, Z. M.; Zhang, J. G. J. Power Sources 2010, 195, 7403. (500) Yu, Y.; Chen, C. H.; Shi, Y. Adv. Mater. 2007, 19, 993. (501) Tirado, J. L.; Santamaria, R.; Ortiz, G. F.; Menendez, R.; Lavela, P.; Jimenez-Mateos, J. M.; Garcia, F. J. G.; Concheso, A.; Alcantara, R. Carbon 2007, 45, 1396. (502) Park, M. S.; Kang, Y. M.; Kim, J. H.; Wang, G. X.; Dou, S. X.; Liu, H. K. Carbon 2008, 46, 35. (503) Yuan, L.; Konstantinov, K.; Wang, G. X.; Liu, H. K.; Dou, S. X. J. Power Sources 2005, 146, 180. (504) Qiao, H.; Zheng, Z.; Zhang, L. Z.; Xiao, L. F. J. Mater. Sci. 2008, 43, 2778. (505) Liu, J.; Li, W.; Manthiram, A. Chem. Commun. 2010, 46, 1437. (506) Lou, X. W.; Deng, D.; Lee, J. Y.; Archer, L. A. Chem. Mater. 2008, 20, 6562. (507) Chen, J. S.; Cheah, Y. L.; Chen, Y. T.; Jayaprakash, N.; Madhavi, S.; Yang, Y. H.; Lou, X. W. J. Phys. Chem. C 2009, 113, 20504. (508) Fan, J.; Wang, T.; Yu, C. Z.; Tu, B.; Jiang, Z. Y.; Zhao, D. Y. Adv. Mater. 2004, 16, 1432. (509) Yang, Z. X.; Du, G. D.; Guo, Z. P.; Yu, X. B.; Li, S. A.; Chen, Z. X.; Zhang, P.; Liu, H. K. Nanoscale 2010, 2, 1011. (510) Liu, B.; Guo, Z. P.; Du, G. D.; Nuli, Y.; Hassan, M. F.; Jia, D. Z. J. Power Sources 2010, 195, 5382. (511) Lou, X. W.; Li, C. M.; Archer, L. A. Adv. Mater. 2009, 21, 2536. (512) Paek, S. M.; Yoo, E.; Honma, I. Nano Lett. 2009, 9, 72. (513) Yao, J.; Shen, X. P.; Wang, B.; Liu, H. K.; Wang, G. X. Electrochem. Commun. 2009, 11, 1849. (514) Wang, D. H.; Kou, R.; Choi, D.; Yang, Z. G.; Nie, Z. M.; Li, J.; Saraf, L. V.; Hu, D. H.; Zhang, J. G.; Graff, G. L.; Liu, J.; Pope, M. A.; Aksay, I. A. ACS Nano 2010, 4, 1587. (515) Zhang, L. S.; Jiang, L. Y.; Yan, H. J.; Wang, W. D.; Wang, W.; Song, W. G.; Guo, Y. G.; Wan, L. J. J. Mater. Chem. 2010, 20, 5462. (516) Brousse, T.; Retoux, R.; Herterich, U.; Schleich, D. M. J. Electrochem. Soc. 1998, 145, 1. (517) Mohamedi, M.; Lee, S. J.; Takahashi, D.; Nishizawa, M.; Itoh, T.; Uchida, I. Electrochim. Acta 2001, 46, 1161. (518) Li, N. C.; Martin, C. R. J. Electrochem. Soc. 2001, 148, A164. (519) Li, N. C.; Martin, C. R.; Scrosati, B. Electrochem. Solid-State Lett. 2000, 3, 316. (520) Chang, S. T.; Leu, I. C.; Liao, C. L.; Yen, J. H.; Hon, M. H. J. Mater. Chem. 2004, 14, 1821. 5446

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(587) Li, G. D.; Xu, L. Q.; Hao, Q.; Wang, M.; Qian, Y. T. RSC Adv. 2012, 2, 284. (588) Zhang, H. K.; Song, H. H.; Chen, X. H.; Zhou, J. S.; Zhang, H. J. Electrochim. Acta 2012, 59, 160. (589) Chen, L. B.; Yin, X. M.; Mei, L.; Li, C. C.; Lei, D. N.; Zhang, M.; Li, Q. H.; Xu, Z.; Xu, C. M.; Wang, T. H. Nanotechnology 2012, 23, No. 035402. (590) Kim, J. G.; Lee, S. H.; Nam, S. H.; Choi, S. M.; Kim, W. B. RSC Adv. 2012, 2, 7829. (591) Noerochim, L.; Wang, J. Z.; Chou, S. L.; Wexler, D.; Liu, H. K. Carbon 2012, 50, 1289. (592) Wang, F.; Song, X. P.; Yao, G.; Zhao, M. S.; Liu, R.; Xu, M. W.; Sun, Z. B. Scr. Mater. 2012, 66, 562. (593) Yin, Y. X.; Xin, S.; Wan, L. J.; Li, C. J.; Guo, Y. G. J. Nanosci. Nanotechnol. 2012, 12, 2581. (594) Nam, S.; Kim, S.; Wi, S.; Choi, H.; Byun, S.; Choi, S. M.; Yoo, S. I.; Lee, K. T.; Park, B. J. Power Sources 2012, 211, 154. (595) Chen, J. S.; Ng, M. F.; Wu, H. B.; Zhang, L.; Lou, X. W. CrystEngComm 2012, 14, 5133. (596) Chen, S. Q.; Wang, Y.; A·hn, H.; Wang, G. X. J. Power Sources 2012, 216, 22. (597) Chen, X. C.; Kierzek, K.; Wilgosz, K.; Machnikowski, J.; Gong, J.; Feng, J. D.; Tang, T.; Kalenczuk, R. J.; Chen, H. M.; Chu, P. K.; Mijowska, E. J. Power Sources 2012, 216, 475. (598) Ding, X. L.; Sun, Q.; Lu, F.; Fu, Z. W. J. Power Sources 2012, 216, 117. (599) Kweon, S. C.; Oh, I. H.; Lee, J. K.; Khodin, A. A. Mater. Lett. 2012, 83, 88. (600) Liu, Z. Y.; Bai, H. W.; Sun, D. D. Chem.Asian J. 2012, 7, 2381. (601) Sun, W. J.; Luo, C. H.; Wang, G. X.; Yan, K. P. J. Alloys Compd. 2012, 535, 114. (602) Zhang, L. Q.; Liu, X. H.; Perng, Y. C.; Cho, J.; Chang, J. P.; Mao, S. X.; Ye, Z. Z.; Huang, J. Y. Micron 2012, 43, 1127. (603) Zhao, Y.; Li, J. X.; Wang, N.; Wu, C. X.; Dong, G. F.; Guan, L. H. J. Phys. Chem. C 2012, 116, 18612. (604) Brousse, T.; Retoux, R.; Herterich, U.; Schleich, D. M. SnO2 Thin Films As Possible Anodes for Li-Ion Batteries; Electrochemical Society Inc: Pennington, NJ, 1997, Vol 97, p 34. (605) Kim, D. W.; Hwang, I. S.; Kwon, S. J.; Kang, H. Y.; Park, K. S.; Choi, Y. J.; Choi, K. J.; Park, J. G. Nano Lett. 2007, 7, 3041. (606) Zhao, N. H.; Yang, L. C.; Zhang, P.; Wang, G. J.; Wang, B.; Yao, B. D.; Wu, Y. P. Mater. Lett. 2010, 64, 972. (607) Ui, K.; Kawamura, S.; Kumagai, N. Electrochim. Acta 2012, 76, 383. (608) Wang, C.; Du, G. H.; Stahl, K.; Huang, H. X.; Zhong, Y. J.; Jiang, J. Z. J. Phys. Chem. C 2012, 116, 4000. (609) Du, Z.; Zhang, S.; Jiang, T.; Wu, X.; Zhang, L.; Fang, H. J. Power Sources 2012, 219, 199. (610) Connor, P. A.; Irvine, J. T. S. J. Power Sources 2001, 97−98, 223. (611) Sharma, N.; Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2005, 139, 250. (612) Sharma, N.; Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochem. Commun. 2002, 4, 947. (613) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. Chem. Mater. 2008, 20, 6829. (614) Zhao, S.; Bai, Y.; Zhang, W. F. Electrochim. Acta 2010, 55, 3891. (615) Mouyane, M.; Womes, M.; Jumas, J. C.; Olivier-Fourcade, J.; Lippens, P. E. J. Solid State Chem. 2011, 184, 2877. (616) Hu, X. Y.; Xiao, T.; Huang, W.; Tao, W.; Heng, B. J.; Chen, X. Q.; Tang, Y. W. Appl. Surf. Sci. 2012, 258, 6177. (617) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2009, 192, 627. (618) Huang, F.; Yuan, Z. Y.; Zhan, H.; Zhou, Y. H.; Sun, J. T. Mater. Lett. 2003, 57, 3341. (619) Yuan, Z. Y.; Huang, F.; Sun, J. T.; Zhou, Y. H. J. Mater. Sci. Lett. 2003, 22, 143.

(555) Ren, J. G.; Yang, J. B.; Abouimrane, A.; Wang, D. P.; Amine, K. J. Power Sources 2011, 196, 8701. (556) Shao, Q. G.; Chen, W. M.; Wang, Z. H.; Qie, L.; Yuan, L. X.; Zhang, W. X.; Hu, X. L.; Huang, Y. H. Electrochem. Commun. 2011, 13, 1431. (557) Jin, Y. H.; Min, K. M.; Seo, S. D.; Shim, H. W.; Kim, D. W. J. Phys. Chem. C 2011, 115, 22062. (558) Chen, Y.; Huang, Q. Z.; Wang, J.; Wang, Q.; Xue, J. M. J. Mater. Chem. 2011, 21, 17448. (559) Zhang, C. F.; Quince, M.; Chen, Z. X.; Guo, Z. P.; Liu, H. K. J. Solid State Electrochem. 2011, 15, 2645. (560) Wang, J. Z.; Du, N.; Zhang, H.; Yu, J. X.; Yang, D. R. J. Phys. Chem. C 2011, 115, 11302. (561) Qiao, H.; Li, J.; Fu, J. P.; Kumar, D.; Wei, Q. F.; Cai, Y. B.; Huang, F. L. ACS Appl. Mater. Interfaces 2011, 3, 3704. (562) Wang, Y. D.; Djerdj, I.; Smarsly, B.; Antonietti, M. Chem. Mater. 2009, 21, 3202. (563) Wu, F. D.; Wu, M. H.; Wang, Y. Electrochem. Commun. 2011, 13, 433. (564) Volosin, A. M.; Sharma, S.; Traverse, C.; Newman, N.; Seo, D. K. J. Mater. Chem. 2011, 21, 13232. (565) Issac, I.; Scheuermann, M.; Becker, S. M.; Bardaji, E. G.; Adelhelm, C.; Wang, D.; Kubel, C.; Indris, S. J. Power Sources 2011, 196, 9689. (566) Kang, J. H.; Paek, S. M.; Choy, J. H. Chem. Commun. 2012, 48, 458. (567) Reddy, M. V.; Andreea, L. Y. T.; Ling, A. Y.; Hwee, J. N. C.; Lin, C. A.; Adams, S.; Loh, K. P.; Ozoemena, K. I.; Mathe, M.; Chowdari, B.V.R. Electrochim. Acta 2013, DOI:10.1016/j.electacta.2013.05.073. (568) Chen, C. M.; Zhang, Q.; Huang, J. Q.; Zhang, W.; Zhao, X. C.; Huang, C. H.; Wei, F.; Yang, Y. G.; Wang, M. Z.; Su, D. S. J. Mater. Chem. 2012, 22, 13947. (569) Chen, Z. X.; Zhou, M.; Cao, Y. L.; Ai, X. P.; Yang, H. X.; Liu, J. Adv. Energy Mater. 2012, 2, 95. (570) Hsieh, C. T.; Lin, J. S.; Chen, Y. F.; Teng, H. S. J. Phys. Chem. C 2012, 116, 15251. (571) Li, X. F.; Meng, X. B.; Liu, J.; Geng, D. S.; Zhang, Y.; Banis, M. N.; Li, Y. L.; Yang, J. L.; Li, R. Y.; Sun, X. L.; Cai, M.; Verbrugge, M. W. Adv. Funct. Mater. 2012, 22, 1647. (572) Li, Y.; Zhu, S. M.; Liu, Q. L.; Gu, J. J.; Guo, Z. P.; Chen, Z. X.; Feng, C. L.; Zhang, D.; Moon, W. J. J. Mater. Chem. 2012, 22, 2766. (573) Liang, J. F.; Wei, W.; Zhong, D.; Yang, Q. L.; Li, L. D.; Guo, L. ACS Appl. Mater. Interfaces 2012, 4, 454. (574) Liu, H. D.; Huang, J. M.; Li, X. L.; Liu, J.; Zhang, Y. X. Ceram. Int. 2012, 38, 5145. (575) Liu, H. D.; Huang, J. M.; Li, X. L.; Liu, J.; Zhang, Y. X.; Du, K. Appl. Surf. Sci. 2012, 258, 4917. (576) Liu, H. D.; Huang, J. M.; Li, X. L.; Liu, J.; Zhang, Y. X.; Du, K. Phys. E (Amsterdam, Neth.) 2012, 44, 1931. (577) Liu, X. H.; Liu, Y.; Kushima, A.; Zhang, S. L.; Zhu, T.; Li, J.; Huang, J. Y. Adv. Energy Mater. 2012, 2, 722. (578) Park, S. K.; Yu, S. H.; Pinna, N.; Woo, S.; Jang, B.; Chung, Y. H.; Cho, Y. H.; Sung, Y. E.; Piao, Y. J. Mater. Chem. 2012, 22, 2520. (579) Sathish, M.; Mitani, S.; Tomai, T.; Unemoto, A.; Honma, I. J. Solid State Electrochem. 2012, 16, 1767. (580) Wang, X.; Cao, X. Q.; Bourgeois, L.; Guan, H.; Chen, S. M.; Zhong, Y. T.; Tang, D. M.; Li, H. Q.; Zhai, T. Y.; Li, L.; Bando, Y.; Golberg, D. Adv. Funct. Mater. 2012, 22, 2682. (581) Xing, L. L.; Ma, C. H.; Cui, C. X.; Xue, X. Y. Solid State Sci. 2012, 14, 111. (582) Xu, C. H.; Sun, J.; Gao, L. J. Mater. Chem. 2012, 22, 975. (583) Xu, C. H.; Sun, J.; Gao, L. Nanoscale 2012, 4, 5425. (584) Zhang, C. F.; Peng, X.; Guo, Z. P.; Cai, C. B.; Chen, Z. X.; Wexler, D.; Li, S.; Liu, H. K. Carbon 2012, 50, 1897. (585) Zhou, X. S.; Yin, Y. X.; Wan, L. J.; Guo, Y. G. J. Mater. Chem. 2012, 22, 17456. (586) Kong, J. H.; Liu, Z. L.; Yang, Z. C.; Tan, H. R.; Xiong, S. X.; Wong, S. Y.; Li, X.; Lu, X. H. Nanoscale 2012, 4, 525. 5447

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(620) Huang, F.; Yuan, Z. Y.; Zhan, H.; Zhou, Y. H.; Sun, J. T. Mater. Chem. Phys. 2004, 83, 16. (621) Belliard, F.; Irvine, J. T. S. Ionics 2001, 7, 16. (622) Zhang, D. W.; Zhang, S. Q.; Jin, Y.; Yi, T. H.; Xie, S.; Chen, C. H. J. Alloys Compd. 2006, 415, 229. (623) Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2006, 159, 340. (624) Sharma, N.; Plevert, J.; Subba Rao, G. V.; Chowdari, B. V. R.; White, T. J. Chem. Mater. 2005, 17, 4700. (625) Das, B.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. Solid State Ionics: New Materials for Pollution free Energy Devices, Proceedings of the 11th Asian Conference on Solid State Ionics, Chowdari, B .V. R. et al., Eds.; MacMillan, India Ltd.: New Delhi, 2008, p 69. (626) Das, B.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Mater. Chem. 2011, 21, 1171. (627) Behm, M.; Irvine, J. T. S. Electrochim. Acta 2002, 47, 1727. (628) Attidekou, P. S.; Connor, P. A.; Wormald, P.; Tunstall, D. P.; Francis, S. M.; Irvine, J. T. S. Solid State Ionics 2004, 175, 185. (629) Attidekou, P. S.; Garcia-Alvarado, F.; Connor, P. A.; Irvine, J. T. S. J. Electrochem. Soc. 2007, 154, A217. (630) Xiao, Y. W.; Lee, J. Y.; Yu, A. S.; Liu, Z. L. J. Electrochem. Soc. 1999, 146, 3623. (631) Li, Y. M.; Li, J. H. J. Phys. Chem. C 2008, 112, 14216. (632) Ren, Q. Q.; Zhou, Y. N.; Sun, Q.; Fu, Z. W. J. Power Sources 2012, 199, 336. (633) Idota, Y.; Kubota, T.; Matsufuji, A.; Maekawa, Y.; Miyasaka, T. Science 1997, 276, 1395. (634) Courtney, I. A.; Dunlap, R. A.; Dahn, J. R. Electrochim. Acta 1999, 45, 51. (635) Goward, G. R.; Leroux, F.; Power, W. P.; Ouvrard, G.; Dmowski, W.; Egami, T.; Nazar, L. F. Electrochem. Solid-State Lett. 1999, 2, 367. (636) Morimoto, H.; Nakai, M.; Tatsumisago, M.; Minami, T. J. Electrochem. Soc. 1999, 146, 3970. (637) Kim, J. Y.; King, D. E.; Kumta, P. N.; Blomgren, G. E. J. Electrochem. Soc. 2000, 147, 4411. (638) Lee, J. Y.; Xiao, Y. W.; Liu, Z. L. Solid State Ionics 2000, 133, 25. (639) Robert, F.; Morato, F.; Chouvin, J.; Aldon, L.; Lippens, P. E.; Fourcade, J. O.; Jumas, J. C.; Simon, B.; Biensan, P. J. Power Sources 2003, 119, 581. (640) Chouvin, J.; Vicente, C. P.; Olivier-Fourcade, J.; Jumas, J. C.; Simon, B.; Biensan, P. Solid State Sci. 2004, 6, 39. (641) Mansour, A. N.; Mukerjee, S.; Yang, X. Q.; McBreen, J. J. Electrochem. Soc. 2000, 147, 869. (642) Gejke, C.; Nordstrom, E.; Fransson, L.; Edstrom, K.; Haggstrom, L.; Borjesson, L. J. Mater. Chem. 2002, 12, 2965. (643) Gejke, C.; Swenson, J.; Delaplane, R. G.; Borjesson, L. Phys. Rev. B 2002, 65, No. 212201. (644) Conte, D. E.; Aboulaich, A.; Robert, F.; Olivier-Fourcade, J.; Jumas, J. C.; Jordy, C.; Willmann, P. J. Solid State Chem. 2010, 183, 65. (645) Conte, D. E.; Stievano, L.; Olivier-Fourcade, J.; Jumas, J. C.; Willmann, P. J. Power Sources 2011, 196, 6644. (646) Mouyane, M.; Womes, M.; Jumas, J. C.; Olivier-Fourcade, J.; Lippens, P. E. J. Power Sources 2012, 204, 139. (647) Larcher, D.; Prakash, A. S.; Laffont, L.; Womes, M.; Jumas, J. C.; Olivier-Fourcade, J.; Hedge, M. S.; Tarascon, J. M. J. Electrochem. Soc. 2006, 153, A1778. (648) Xue, M. Z.; Fu, Z. W. Electrochem. Commun. 2006, 8, 1250. (649) Bryngelsson, H.; Eskhult, J.; Nyholm, L.; Herranen, M.; Alm, O.; Edstrom, K. Chem. Mater. 2007, 19, 1170. (650) Simonin, L.; Lafont, U.; Tabrizi, N.; Schmidt-Ott, A.; Kelder, E. M. J. Power Sources 2007, 174, 805. (651) Kundu, M.; Mahanty, S.; Basu, R. N. Electrochem. Commun. 2009, 11, 1389. (652) Morales, J.; Sanchez, L.; Martin, F.; Berry, F. J. Solid State Chem. 2006, 179, 2554. (653) Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. In Solid State Ionics: New Materials for Pollution Free Energy Devices, Proceedings

of the 11th Asian Conference on Solid State Ionics; Chowdari, B. V. R., et al., Eds.; MacMillan, India Ltd.: New Delhi. 2008, p 187. (654) Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Mater. Chem. 2011, 21, 10003. (655) Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Solid State Electrochem. 2012, DOI: 10.1007/s10008-012-1791-7. (656) Perez-Flores, J. C.; Kuhn, A.; Garcia-Alvarado, F. J. Power Sources 2008, 182, 365. (657) Pena, J. S.; Pascual, J. C.; Caballero, A.; Morales, J.; Sanchez, L. J. Solid State Chem. 2004, 177, 2920. (658) Jibin, A. K.; Reddy, M. V.; Subba Rao, G. V.; Varadaraju, U. V.; Chowdari, B. V. R. Electrochim. Acta 2012, 71, 227. (659) Park, C. M.; Choi, W.; Hwa, Y.; Kim, J. H.; Jeong, G.; Sohn, H. J. J. Mater. Chem. 2010, 20, 4854. (660) Yamada, M.; Ueda, A.; Matsumoto, K.; Ohzuku, T. J. Electrochem. Soc. 2011, 158, A417. (661) Lu, Z. W.; Zhang, L. Q.; Liu, X. J. J. Power Sources 2010, 195, 4304. (662) Jeong, G.; Kim, Y. U.; Krachkovskiy, S. A.; Lee, C. K. Chem. Mater. 2010, 22, 5570. (663) Yamamura, H.; Nobuhara, K.; Nakanishi, S.; Iba, H.; Okada, S. J. Ceram. Soc. Jpn. 2011, 119, 855. (664) Komaba, S.; Shimomura, K.; Yabuuchi, N.; Ozeki, T.; Yui, H.; Konno, K. J. Phys. Chem. C 2011, 115, 13487. (665) Si, Q.; Hanai, K.; Ichikawa, T.; Phillipps, M. B.; Hirano, A.; Imanishi, N.; Yamamoto, O.; Takeda, Y. J. Power Sources 2011, 196, 9774. (666) Yoo, H.; Lee, J. I.; Kim, H.; Lee, J. P.; Cho, J.; Park, S. Nano Lett. 2011, 11, 4324. (667) Yao, Y.; Zhang, J. J.; Xue, L. G.; Huang, T.; Yu, A. S. J. Power Sources 2011, 196, 10240. (668) Kim, J. H.; Park, C. M.; Kim, H.; Kim, Y. J.; Sohn, H. J. J. Electroanal. Chem. 2011, 661, 245. (669) Miyuki, T.; Okuyama, Y.; Sakamoto, T.; Eda, Y.; Kojima, T.; Sakai, T. Electrochemistry 2012, 80, 401. (670) Ren, Y. R.; Ding, J. N.; Yuan, N. Y.; Jia, S. Y.; Qu, M. Z.; Yu, Z. L. J. Solid State Electrochem. 2012, 16, 1453. (671) Ban, C. M.; Kappes, B. B.; Xu, Q.; Engtrakul, C.; Ciobanu, C. V.; Dillon, A. C.; Zhao, Y. F. Appl. Phys. Lett. 2012, 100, No. 243905. (672) Guo, C. F.; Wang, D. L.; Wang, Q. M.; Wang, B.; Liu, T. F. Int. J. Electrochem. Sci. 2012, 7, 8745. (673) Jeong, G.; Kim, J. H.; Kim, Y. U.; Kim, Y. J. J. Mater. Chem. 2012, 22, 7999. (674) Song, J. W.; Nguyen, C. C.; Song, S. W. RSC Adv. 2012, 2, 2003. (675) Guo, B. K.; Shu, J.; Wang, Z. X.; Yang, H.; Shi, L. H.; Liu, Y. N.; Chen, L. Q. Electrochem. Commun. 2008, 10, 1876. (676) Pena, J. S.; Sandu, I.; Joubert, O.; Pascual, F. S.; Arean, C. O.; Brousse, T. Electrochem. Solid-State Lett. 2004, 7, A278. (677) Kim, Y.; Hwang, H.; Lawler, K.; Martin, S. W.; Cho, J. Electrochim. Acta 2008, 53, 5058. (678) Kim, C. H.; Jung, Y. S.; Lee, K. T.; Ku, J. H.; Oh, S. M. Electrochim. Acta 2009, 54, 4371. (679) Feng, J. K.; Xia, H.; Lai, M. O.; Lu, L. J. Phys. Chem. C 2009, 113, 20514. (680) Feng, J. K.; Lu, L.; Lai, M. O. J. Alloys Compd. 2010, 501, 255. (681) Feng, J. K.; Lai, M. O.; Lu, L. Mater. Res. Bull. 2012, 47, 1693. (682) Feng, J. K.; Lai, M. O.; Lu, L. Electrochim. Acta 2012, 62, 103. (683) Belliard, F.; Irvine, J. T. S. J. Power Sources 2001, 97−98, 219. (684) Fu, Z. W.; Huang, F.; Zhang, Y.; Chu, Y.; Qin, Q. Z. J. Electrochem. Soc. 2003, 150, A714. (685) Wang, H. B.; Pan, Q. M.; Cheng, Y. X.; Zhao, J. W.; Yin, G. P. Electrochim. Acta 2009, 54, 2851. (686) Zhang, C. Q.; Tu, J. P.; Yuan, Y. F.; Huang, X. H.; Chen, X. T.; Mao, F. J. Electrochem. Soc. 2007, 154, A65. (687) Pan, Q. M.; Qin, L. M.; Liu, J.; Wang, H. B. Electrochim. Acta 2010, 55, 5780. (688) Liu, J. P.; Li, Y. Y.; Ding, R. M.; Jiang, J.; Hu, Y. Y.; Ji, X. X.; Chi, Q. B.; Zhu, Z. H.; Huang, X. T. J. Phys. Chem. C 2009, 113, 5336. 5448

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(689) Huang, X. H.; Xia, X. H.; Yuan, Y. F.; Zhou, F. Electrochim. Acta 2011, 56, 4960. (690) Wu, Z.; Qin, L. M.; Pan, Q. M. J. Alloys Compd. 2011, 509, 9207. (691) A·hmad, M.; Shi, Y. Y.; Nisar, A.; Sun, H. Y.; Shen, W. C.; Wei, M.; Zhu, J. J. Mater. Chem. 2011, 21, 7723. (692) Song, W. T.; Xie, J.; Liu, S. Y.; Zheng, Y. X.; Cao, G. S.; Zhu, T. J.; Zhao, X. B. Int. J. Electrochem. Sci. 2012, 7, 2164. (693) Zhou, Y. N.; Li, W. J.; Fu, Z. W. Electrochim. Acta 2012, 59, 435. (694) Huang, X. H.; Wu, J. B.; Lin, Y.; Guo, R. Q. Int. J. Electrochem. Sci. 2012, 7, 6611. (695) Usui, H.; Kono, T.; Sakaguchi, H. Int. J. Electrochem. Sci. 2012, 7, 4322. (696) Wang, X.; Liao, M. Y.; Zhong, Y. T.; Zheng, J. Y.; Tian, W.; Zhai, T. Y.; Zhi, C. Y.; Ma, Y.; Yao, J. N. A.; Bando, Y.; Golberg, D. Adv. Mater. 2012, 24, 3421. (697) Li, Y. G.; Tan, B.; Wu, Y. Y. Chem. Mater. 2008, 20, 567. (698) Zhou, Y. N.; Zhang, H.; Xue, M. Z.; Wu, C. L.; Wu, X. J.; Fu, Z. W. J. Power Sources 2006, 162, 1373. (699) Ho, W. H.; Li, C. F.; Liu, H. C.; Yen, S. K. J. Power Sources 2008, 175, 897. (700) Yang, R.; Zheng, J.; Huang, J.; Zhang, X. Z.; Qu, J. L.; Li, X. G. Electrochem. Commun. 2010, 12, 784. (701) Liu, J.; Zhu, Y. C.; Liang, J. W.; Qian, Y. T. Int. J. Electrochem. Sci. 2012, 7, 5574. (702) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. M. Nature 2000, 407, 496. (703) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. M. J. Power Sources 2001, 97−98, 235. (704) Grugeon, S.; Laruelle, S.; Dupont, L.; Tarascon, J. M. Solid State Sci. 2003, 5, 895. (705) Badway, F.; Plitz, I.; Grugeon, S.; Laruelle, S.; Dolle, M.; Gozdz, A. S.; Tarascon, J. M. Electrochem. Solid-State Lett. 2002, 5, A115. (706) Obrovac, M. N.; Dunlap, R. A.; Sanderson, R. J.; Dahn, J. R. J. Electrochem. Soc. 2001, 148, A576. (707) Dedryvere, R.; Laruelle, S.; Grugeon, S.; Poizot, P.; Gonbeau, D.; Tarascon, J. M. Chem. Mater. 2004, 16, 1056. (708) Choi, H. C.; Jung, Y. M.; Noda, I.; Kim, S. B. J. Phys. Chem. B 2003, 107, 5806. (709) Chen, C. H.; Hwang, B. J.; Do, J. S.; Weng, J. H.; Venkateswarlu, M.; Cheng, M. Y.; Santhanam, R.; Ragavendran, K.; Lee, J. F.; Chen, J. M.; Liu, D. G. Electrochem. Commun. 2010, 12, 496. (710) Laruelle, S.; Grugeon, S.; Poizot, P.; Dolle, M.; Dupont, L.; Tarascon, J. M. J. Electrochem. Soc. 2002, 149, A627. (711) Grugeon, S.; Laruelle, S.; Herrera-Urbina, R.; Dupont, L.; Poizot, P.; Tarascon, J. M. J. Electrochem. Soc. 2001, 148, A285. (712) Wang, G. X.; Chen, Y.; Konstantinov, K.; Lindsay, M.; Liu, H. K.; Dou, S. X. J. Power Sources 2002, 109, 142. (713) Yao, W. L.; Yang, J.; Wang, J. L.; Nuli, Y. J. Electrochem. Soc. 2008, 155, A903. (714) Jiang, J.; Liu, J. P.; Ding, R. M.; Ji, X. X.; Hu, Y. Y.; Li, X.; Hu, A. Z.; Wu, F.; Zhu, Z. H.; Huang, X. T. J. Phys. Chem. C 2010, 114, 929. (715) Zhang, H. J.; Tao, H. H.; Jiang, Y.; Jiao, Z.; Wu, M. H.; Zhao, B. J. Power Sources 2010, 195, 2950. (716) Qiao, H.; Xiao, L. F.; Zheng, Z.; Liu, H. W.; Jia, F. L.; Zhang, L. Z. J. Power Sources 2008, 185, 486. (717) Sun, B.; Liu, H.; Munroe, P.; A·hn, H.; Wang, G. X. Nano Res. 2012, 5, 460. (718) Xiong, S. L.; Chen, J. S.; Lou, X. W.; Zeng, H. C. Adv. Funct. Mater. 2012, 22, 861. (719) Peng, C. X.; Chen, B. D.; Qin, Y.; Yang, S. H.; Li, C. Z.; Zuo, Y. H.; Liu, S. Y.; Yang, J. H. ACS Nano 2012, 6, 1074. (720) Yu, Y.; Chen, C. H.; Shui, J. L.; Xie, S. Angew. Chem., Int. Ed. 2005, 44, 7085. (721) Dalverny, A. L.; Filhol, J. S.; Doublet, M. L. J. Mater. Chem. 2011, 21, 10134.

(722) Zhang, L. J.; Hu, P.; Zhao, X. Y.; Tian, R. L.; Zou, R. Q.; Xia, D. G. J. Mater. Chem. 2011, 21, 18279. (723) Wu, F. D.; Wang, Y. J. Mater. Chem. 2011, 21, 6636. (724) Nam, K. M.; Choi, Y. C.; Jung, S. C.; Kim, Y. I.; Jo, M. R.; Park, S. H.; Kang, Y. M.; Han, Y. K.; Park, J. T. Nanoscale 2012, 4, 473. (725) Valvo, M.; Garcia-Tamayo, E.; Lafont, U.; Kelder, E. M. J. Power Sources 2011, 196, 10191. (726) Kokubu, T.; Oaki, Y.; Hosono, E.; Zhou, H. S.; Imai, H. Adv. Funct. Mater. 2011, 21, 3673. (727) Qi, Y.; Du, N.; Zhang, H.; Wang, J. Z.; Yang, Y.; Yang, D. R. J. Alloys Compd. 2012, 521, 83. (728) Sun, Y. M.; Hu, X. L.; Luo, W.; Huang, Y. H. J. Mater. Chem. 2012, 22, 13826. (729) Guan, H.; Wang, X.; Li, H. Q.; Zhi, C. Y.; Zhai, T. Y.; Bando, Y.; Golberg, D. Chem. Commun. 2012, 48, 4878. (730) Yu, B. C.; Lee, J. O.; Song, J. H.; Park, C. M.; Lee, C. K.; Sohn, H. J. J. Solid State Electrochem. 2012, 16, 2631. (731) Yao, W. L.; Chen, J. Q.; Cheng, H. W. J. Solid State Electrochem. 2011, 15, 183. (732) Liu, J.; Zhou, Y. C.; Liu, C. P.; Wang, J. B.; Pan, Y.; Xue, D. F. CrystEngComm 2012, 14, 2669. (733) Reddy, M. V.; Prithvi, G.; Loh, K. P.; Chowdari, B. V. R. J. Phys. Chem. C 2013, submitted for publication. (734) Huang, X. H.; Tu, J. P.; Zhang, B.; Zhang, C. Q.; Li, Y.; Yuan, Y. F.; Wu, H. M. J. Power Sources 2006, 161, 541. (735) Hosono, E.; Fujihara, S.; Honma, I.; Zhou, H. S. Electrochem. Commun. 2006, 8, 284. (736) Varghese, B.; Reddy, M. V.; Yanwu, Z.; Lit, C. S.; Hoong, T. C.; Subba Rao, G. V.; Chowdari, B. V. R.; Wee, A. T. S.; Lim, C. T.; Sow, C. H. Chem. Mater. 2008, 20, 3360. (737) Oh, S. W.; Bang, H. J.; Bae, Y. C.; Sun, Y. K. J. Power Sources 2007, 173, 502. (738) Huang, X. H.; Tu, J. P.; Zhang, C. Q.; Xiang, J. Y. Electrochem. Commun. 2007, 9, 1180. (739) Li, X. F.; Dhanabalan, A.; Bechtold, K.; Wang, C. L. Electrochem. Commun. 2010, 12, 1222. (740) Zhong, C.; Wang, J. Z.; Chou, S. L.; Konstantinov, K.; Rahman, M.; Liu, H. K. J. Appl. Electrochem. 2010, 40, 1415. (741) Liu, L.; Li, Y.; Yuan, S. M.; Ge, M.; Ren, M. M.; Sun, C. S.; Zhou, Z. J. Phys. Chem. C 2010, 114, 251. (742) Needham, S. A.; Wang, G. X.; Liu, H. K. J. Power Sources 2006, 159, 254. (743) Cheng, M. Y.; Hwang, B. J. J. Power Sources 2010, 195, 4977. (744) Qiao, H.; Wu, N.; Huang, F. L.; Cai, Y. B.; Wei, Q. F. Mater. Lett. 2010, 64, 1022. (745) Rahman, M. M.; Chou, S. L.; Zhong, C.; Wang, J. Z.; Wexler, D.; Liu, H. K. Solid State Ionics 2010, 180, 1646. (746) Nuli, Y. N.; Zhang, P.; Guo, Z. P.; Liu, H. K.; Yang, J.; Wang, J. L. Mater. Res. Bull. 2009, 44, 140. (747) Lu, H. W.; Li, D.; Sun, K.; Li, Y. S.; Fu, Z. W. Solid State Sci. 2009, 11, 982. (748) Wang, C.; Wang, D. L.; Wang, Q. M.; Chen, H. J. J. Power Sources 2010, 195, 7432. (749) Yuan, Y. F.; Xia, X. H.; Wu, J. B.; Yang, J. L.; Chen, Y. B.; Guo, S. Y. Electrochem. Commun. 2010, 12, 890. (750) Huang, X. H.; Tu, J. P.; Xia, X. H.; Wang, X. L.; Xiang, J. Y.; Zhang, L.; Zhou, Y. J. Power Sources 2009, 188, 588. (751) Huang, X. H.; Tu, J. P.; Xia, X. H.; Wang, X. L.; Xiang, J. Y. Electrochem. Commun. 2008, 10, 1288. (752) Huang, X. H.; Tu, J. P.; Zeng, Z. Y.; Xiang, J. Y.; Zhao, X. B. J. Electrochem. Soc. 2008, 155, A438. (753) Huang, X. H.; Tu, J. P.; Xia, X. H.; Wang, X. L.; Xiang, J. Y.; Zhang, L. J. Power Sources 2010, 195, 1207. (754) Liu, H.; Wang, G. X.; Liu, J.; Qiao, S. Z.; A·hn, H. J. J. Mater. Chem. 2011, 21, 3046. (755) Ma, J. M.; Yang, J. Q.; Jiao, L. F.; Mao, Y. H.; Wang, T. H.; Duan, X. C.; Lian, J. B.; Zheng, W. J. CrystEngComm 2012, 14, 453. (756) Chen, X.; Zhang, N. Q.; Sun, K. N. Electrochem. Commun. 2012, 20, 137. 5449

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(757) Chen, Z.; Xiao, A.; Chen, Y.; Zuo, C.; Zhou, S.; Li, L. Mater. Res. Bull. 2012, 47, 1987. (758) Yao, Y.; Zhang, J. J.; Wei, Z.; Yu, A. S. Int. J. Electrochem. Sci. 2012, 7, 1433. (759) Wang, X. H.; Li, X. W.; Sun, X. L.; Li, F.; Liu, Q. M.; Wang, Q.; He, D. Y. J. Mater. Chem. 2011, 21, 3571. (760) Wang, X. H.; Yang, Z. B.; Sun, X. L.; Li, X. W.; Wang, D. S.; Wang, P.; He, D. Y. J. Mater. Chem. 2011, 21, 9988. (761) Cao, F.; Zhang, F.; Deng, R.; Hu, W.; Liu, D.; Song, S.; Zhang, H. CrystEngComm 2011, 13, 4903. (762) Huang, X. H.; Yuan, Y. F.; Wang, Z.; Zhang, S. Y.; Zhou, F. J. Alloys Compd. 2011, 509, 3425. (763) Zhong, J.; Wang, X. L.; Xia, X. H.; Gu, C. D.; Xiang, J. Y.; Zhang, J.; Tu, J. P. J. Alloys Compd. 2011, 509, 3889. (764) Mai, Y. J.; Tu, J. P.; Xia, X. H.; Gu, C. D.; Wang, X. L. J. Power Sources 2011, 196, 6388. (765) Ni, S. B.; Li, T.; Yang, X. L. Mater. Chem. Phys. 2012, 132, 1108. (766) Wen, W.; Wu, J. M. ACS Appl. Mater. Interfaces 2011, 3, 4112. (767) Li, X.; Dhanabalan, A.; Wang, C. J. Power Sources 2011, 196, 9625. (768) Mai, Y. J.; Xia, X. H.; Chen, R.; Gu, C. D.; Wang, X. L.; Tu, J. P. Electrochim. Acta 2012, 67, 73. (769) Wu, H.; Xu, M.; Wu, H. Y.; Xu, J. J.; Wang, Y. L.; Peng, Z.; Zheng, G. F. J. Mater. Chem. 2012, 22, 19821. (770) Wu, Y.; Balakrishna, R.; Reddy, M. V.; Nair, A. S.; Chowdari, B. V. R.; Ramakrishna, S. J. Alloys Compd. 2012, 517, 69. (771) Xu, C. H.; Sun, J.; Gao, L. A. J. Power Sources 2011, 196, 5138. (772) Kottegoda, I. R. M.; Idris, N. H.; Lu, L.; Wang, J. Z.; Liu, H. K. Electrochim. Acta 2011, 56, 5815. (773) Zou, Y. Q.; Wang, Y. Nanoscale 2011, 3, 2615. (774) Huang, Y.; Huang, X. L.; Lian, J. S.; Xu, D.; Wang, L. M.; Zhang, X. B. J. Mater. Chem. 2012, 22, 2844. (775) Mai, Y. J.; Shi, S. J.; Zhang, D.; Lu, Y.; Gu, C. D.; Tu, J. P. J. Power Sources 2012, 204, 155. (776) Mai, Y. J.; Tu, J. P.; Gu, C. D.; Wang, X. L. J. Power Sources 2012, 209, 1. (777) Qiu, D. F.; Xu, Z. J.; Zheng, M. B.; Zhao, B.; Pan, L. J.; Pu, L.; Shi, Y. J. Solid State Electrochem. 2012, 16, 1889. (778) Zhou, G. M.; Wang, D. W.; Yin, L. C.; Li, N.; Li, F.; Cheng, H. M. ACS Nano 2012, 6, 3214. (779) Zhu, X. J.; Hu, J.; Dai, H. L.; Ding, L.; Jiang, L. Electrochim. Acta 2012, 64, 23. (780) Debart, A.; Dupont, L.; Poizot, P.; Leriche, J. B.; Tarascon, J. M. J. Electrochem. Soc. 2001, 148, A1266. (781) Lee, Y. H.; Leu, I. C.; Liao, C. L.; Chang, S. T.; Wu, M. T.; Yen, J. H.; Fung, K. Z. Electrochem. Solid-State Lett. 2006, 9, A207. (782) Fu, L.; Gao, J.; Zhang, T.; Cao, Q.; Yang, L. C.; Wu, Y. P.; Holze, R.; Wu, H. Q. J. Power Sources 2007, 174, 1197. (783) Wang, H. B.; Pan, Q. M.; Zhao, H. W.; Yin, G. P.; Zuo, P. J. J. Power Sources 2007, 167, 206. (784) Pan, Q.; Jin, H.; Wang, H.; Yin, G. Electrochim. Acta 2007, 53, 951. (785) Wang, H. B.; Pan, Q. M.; Zhao, J. W.; Chen, W. T. J. Alloys Compd. 2009, 476, 408. (786) Chen, L. B.; Lu, N.; Xu, C. M.; Yu, H. C.; Wang, T. H. Electrochim. Acta 2009, 54, 4198. (787) Ke, F. S.; Huang, L.; Wei, G. Z.; Xue, L. J.; Li, J. T.; Zhang, B.; Chen, S. R.; Fan, X. Y.; Sun, S. G. Electrochim. Acta 2009, 54, 5825. (788) Xiang, J. Y.; Tu, J. P.; Zhang, J.; Zhong, J.; Zhang, D.; Cheng, J. P. Electrochem. Commun. 2010, 12, 1103. (789) Dar, M. A.; Nam, S. H.; Kim, Y. S.; Kim, W. B. J. Solid State Electrochem. 2010, 14, 1719. (790) Xiang, J. Y.; Tu, J. P.; Zhang, L.; Zhou, Y.; Wang, X. L.; Shi, S. J. J. Power Sources 2010, 195, 313. (791) Guan, X. F.; Li, L. P.; Li, G. S.; Fu, Z. W.; Zheng, J.; Yan, T. J. J. Alloys Compd. 2011, 509, 3367.

(792) Zhang, C. M.; Chen, J.; Zeng, Y.; Rui, X. H.; Zhu, J. X.; Zhang, W. Y.; Xu, C.; Lim, T. M.; Hng, H. H.; Yan, Q. Y. Nanoscale 2012, 4, 3718. (793) Kong, M.; Zhang, W. X.; Yang, Z. H.; Weng, S. Y.; Chen, Z. X. Appl. Surf. Sci. 2011, 258, 1317. (794) Wan, M.; Jin, D. L.; Feng, R.; Si, L. M.; Gao, M. X.; Yue, L. H. Inorg. Chem. Commun. 2011, 14, 38. (795) Chen, X.; Zhang, N. Q.; Sun, K. N. J. Mater. Chem. 2012, 22, 13637. (796) Ju, J. H.; Ryu, K. S. J. Electrochem. Soc. 2011, 158, A814. (797) Xiang, J. Y.; Tu, J. P.; Qiao, Y. Q.; Wang, X. L.; Zhong, J.; Zhang, D.; Gu, C. D. J. Phys. Chem. C 2011, 115, 2505. (798) Feng, J. K.; Xia, H.; Lai, M. O.; Lu, L. Mater. Res. Bull. 2011, 46, 424. (799) Zhang, W. X.; Li, M.; Wang, Q.; Chen, G. D.; Kong, M.; Yang, Z. H.; Mann, S. Adv. Funct. Mater. 2011, 21, 3516. (800) Lamberti, A.; Destro, M.; Bianco, S.; Quaglio, M.; Chiodoni, A.; Pirri, C. F.; Gerbaldi, C. Electrochim. Acta 2012, 70, 62. (801) Yin, Z. G.; Ding, Y. H.; Zheng, Q. D. H.; Guan, L. Electrochem. Commun. 2012, 20, 40. (802) Chen, X.; Zhang, N. Q.; Sun, K. N. J. Mater. Chem. 2012, 22, 15080. (803) Xu, M. W.; Wang, F.; Ding, B. J.; Song, X. P.; Fang, J. X. RSC Adv. 2012, 2, 2240. (804) Choi, C. S.; Park, Y. U.; Kim, H.; Kim, N. R.; Kang, K.; Lee, H. M. Electrochim. Acta 2012, 70, 98. (805) Yu, Q.; Huang, H. W.; Chen, R.; Wang, P.; Yang, H. S.; Gao, M. X.; Peng, X. S.; Ye, Z. Z. Nanoscale 2012, 4, 2613. (806) Huang, X. H.; Wang, C. B.; Zhang, S. Y.; Zhou, F. Electrochim. Acta 2011, 56, 6752. (807) Liu, X. Q.; Li, Z.; Zhang, Q.; Li, F.; Kong, T. Mater. Lett. 2012, 80, 37. (808) Kim, D. W.; Rhee, K. Y.; Park, S. J. J. Alloys Compd. 2012, 530, 6. (809) Zhong, C.; Wang, J. Z.; Gao, X. W.; Chou, S. L.; Konstantinov, K.; Liu, H. K. J. Nanosci. Nanotechnol. 2012, 12, 1314. (810) Sahay, R.; Kumar, P. S.; Aravindan, V.; Sundaramurthy, J.; Ling, W. C.; Mhaisalkar, S. G.; Ramakrishna, S.; Madhavi, S. J. Phys. Chem. C 2012, 116, 18087. (811) Ko, S.; Lee, J. I.; Yang, H. S.; Park, S.; Jeong, U. Adv. Mater. 2012, 24, 4451. (812) Mai, Y. J.; Wang, X. L.; Xiang, J. Y.; Qiao, Y. Q.; Zhang, D.; Gu, C. D.; Tu, J. P. Electrochim. Acta 2011, 56, 2306. (813) Zhou, J. S.; Ma, L. L.; Song, H. H.; Wu, B.; Chen, X. H. Electrochem. Commun. 2011, 13, 1357. (814) Lu, L. Q.; Wang, Y. Electrochem. Commun. 2012, 14, 82. (815) Lu, L. Q.; Wang, Y. J. Mater. Chem. 2011, 21, 17916. (816) Qi, Y.; Eskelsen, J. R.; Mazur, U.; Hipps, K. W. Langmuir 2012, 28, 3489. (817) Guo, Z.; Reddy, M. V.; Bee Min, G.; Angela Koh, P. S.; Bao, Q.; Zhang, H.; Loh, K. P. J. Mater. Chem. A 2013, submitted for publication. (818) Reddy, M. V.; Yu, C.; Jiahuan, F.; Loh, K. P.; Chowdari, B. V. R. ACS Appl. Mater. & Inter. 2013, 5, 4361. (819) Poizot, P.; Laruelle, S.; Grugeon, S.; Tarascon, J. M. J. Electrochem. Soc. 2002, 149, A1212. (820) Aragon, M. J.; Perez-Vicente, C.; Tirado, J. L. Electrochem. Commun. 2007, 9, 1744. (821) Ji, L. W.; Medford, A. J.; Zhang, X. W. J. Mater. Chem. 2009, 19, 5593. (822) Yu, X. Q.; He, Y.; Sun, J. P.; Tang, K.; Li, H.; Chen, L. Q.; Huang, X. J. Electrochem. Commun. 2009, 11, 791. (823) Liu, J.; Pan, Q. M. Electrochem. Solid-State Lett. 2010, 13, A139. (824) Zhong, K. F.; Xia, X.; Zhang, B.; Li, H.; Wang, Z. X.; Chen, L. Q. J. Power Sources 2010, 195, 3300. (825) Sun, B.; Chen, Z. X.; Kim, H. S.; A·hn, H.; Wang, G. X. J. Power Sources 2011, 196, 3346. (826) Zhong, K. F.; Zhang, B.; Luo, S. H.; Wen, W.; Li, H.; Huang, X. J.; Chen, L. Q. J. Power Sources 2011, 196, 6802. 5450

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(859) Ma, J. M.; Manthiram, A. RSC Adv. 2012, 2, 3187. (860) Lu, Y.; Wang, Y.; Zou, Y. Q.; Jiao, Z.; Zhao, B.; He, Y. Q.; Wu, M. H. Electrochem. Commun. 2010, 12, 101. (861) Liu, B.; Zhang, X. B.; Shioyama, H.; Mukai, T.; Sakai, T.; Xu, Q. J. Power Sources 2010, 195, 857. (862) Wang, X.; Wu, X. L.; Guo, Y. G.; Zhong, Y. T.; Cao, X. Q.; Ma, Y.; Yao, J. N. Adv. Funct. Mater. 2010, 20, 1680. (863) Wang, Y.; Zhang, H. J.; Lu, L.; Stubbs, L. P.; Wong, C. C.; Lin, J. Y. ACS Nano 2010, 4, 4753. (864) Barreca, D.; Cruz-Yusta, M.; Gasparotto, A.; Maccato, C.; Morales, J.; Pozza, A.; Sada, C.; Sanchez, L.; Tondello, E. J. Phys. Chem. C 2010, 114, 10054. (865) Wang, J. Q.; Du, G. D.; Zeng, R.; Niu, B.; Chen, Z. X.; Guo, Z. P.; Dou, S. X. Electrochim. Acta 2010, 55, 4805. (866) Liu, J.; Xia, H.; Lu, L.; Xue, D. F. J. Mater. Chem. 2010, 20, 1506. (867) Zheng, J.; Liu, J.; Lv, D. P.; Kuang, Q.; Jiang, Z. Y.; Xie, Z. X.; Huang, R. B.; Zheng, L. S. J. Solid State Chem. 2010, 183, 600. (868) Wang, C.; Wang, D. L.; Wang, Q. M.; Wang, L. Electrochim. Acta 2010, 55, 6420. (869) Yan, J.; Wei, T.; Qiao, W. M.; Shao, B.; Zhao, Q. K.; Zhang, L. J.; Fan, Z. J. Electrochim. Acta 2010, 55, 6973. (870) Guo, B. K.; Kong, Q. Y.; Zhu, Y.; Mao, Y.; Wang, Z. X.; Wan, M. X.; Chen, L. Q. Chem.Eur. J. 2011, 17, 14878. (871) Li, B. J.; Cao, H. Q.; Shao, J.; Li, G. Q.; Qu, M. Z.; Yin, G. Inorg. Chem. 2011, 50, 1628. (872) Li, C. C.; Li, Q. H.; Chen, L. B.; Wang, T. H. J. Mater. Chem. 2011, 21, 11867. (873) Li, C. C.; Yin, X. M.; Li, Q. H.; Chen, L. B.; Wang, T. H. Chem.Eur. J. 2011, 17, 1596. (874) Wang, B.; Wang, Y.; Park, J.; A·hn, H.; Wang, G. X. J. Alloys Compd. 2011, 509, 7778. (875) Wang, G. L.; Liu, J. C.; Tang, S.; Li, H. Y.; Cao, D. X. J. Solid State Electrochem. 2011, 15, 2587. (876) Xu, M. W.; Wang, F.; Zhao, M. S.; Yang, S.; Song, X. P. Electrochim. Acta 2011, 56, 4876. (877) Huang, H.; Zhu, W. J.; Tao, X. Y.; Xia, Y.; Yu, Z. Y.; Fang, J. W.; Gan, Y. P.; Zhang, W. K. ACS Appl. Mater. Interf. 2012, 4, 5974. (878) Zhou, W. W.; Liu, J. P.; Chen, T.; Tan, K. S.; Jia, X. T.; Luo, Z. Q.; Cong, C. X.; Yang, H. P.; Li, C. M.; Yu, T. Phys. Chem. Chem. Phys. 2011, 13, 14462. (879) Ponrouch, A.; Palacín, M. R. J. Power Sources 2011, 196, 9682. (880) Li, Y. L.; Zhao, J. Z.; Dan, Y. Y.; Ma, D. C.; Zhao, Y.; Hou, S. N.; Lin, H. B.; Wang, Z. C. Chem. Eng. J. 2011, 166, 428. (881) Lee, H. S.; Choi, J.; Jin, J.; Chun, J.; Lee, S. M.; Kim, H. J.; Son, S. U. Chem. Commun. 2012, 48, 94. (882) Donders, M. E.; Knoops, H. C. M.; Kessels, W. M. M.; Notten, P. H. L. J. Power Sources 2012, 203, 72. (883) Jayaprakash, N.; Jones, W. D.; Moganty, S. S.; Archer, L. A. J. Power Sources 2012, 200, 53. (884) Wang, L. M.; Liu, B.; Ran, S. H.; Huang, H. T.; Wang, X. F.; Liang, B.; Chen, D.; Shen, G. Z. J. Mater. Chem. 2012, 22, 23541. (885) Ponrouch, A.; Taberna, P. L.; Simon, P.; Palacin, M. R. Electrochim. Acta 2012, 61, 13. (886) Tao, L. Q.; Zai, J. T.; Wang, K. X.; Zhang, H. J.; Xu, M.; Shen, J.; Su, Y. Z.; Qian, X. F. J. Power Sources 2012, 202, 230. (887) Wen, W.; Wu, J. M.; Tu, J. P. J. Alloys Compd. 2012, 513, 592. (888) Yao, X. Y.; Xin, X.; Zhang, Y. M.; Wang, J.; Liu, Z. P.; Xu, X. X. J. Alloys Compd. 2012, 521, 95. (889) Zhang, F.; Yuan, C. Z.; Lu, X. J.; Zhang, L. J.; Che, Q.; Zhang, X. G. J. Power Sources 2012, 203, 250. (890) Chen, J.; Xia, X. H.; Tu, J. P.; Xiong, Q. Q.; Yu, Y. X.; Wang, X. L.; Gu, C. D. J. Mater. Chem. 2012, 22, 15056. (891) Fu, Y. J.; Li, X. W.; Sun, X. L.; Wang, X. H.; Liu, D. Q.; He, D. Y. J. Mater. Chem. 2012, 22, 17429. (892) Rosant, C.; Avalle, B.; Larcher, D.; Dupont, L.; Friboulet, A.; Tarascon, J. M. Energy & Environ. Sci. 2012, 5, 9936. (893) Lin, Z. Z.; Yue, W. B.; Huang, D. Z.; Hu, J. Y.; Zhang, X. Y.; Yuan, Z. Y.; Yang, X. J. RSC Adv. 2012, 2, 1794.

(827) Li, X. W.; Li, D.; Qiao, L.; Wang, X. H.; Sun, X. L.; Wang, P.; He, D. Y. J. Mater. Chem. 2012, 22, 9189. (828) Ding, Y. L.; Wu, C. Y.; Yu, H. M.; Xie, J.; Cao, G. S.; Zhu, T. J.; Zhao, X. B.; Zeng, Y. W. Electrochim. Acta 2011, 56, 5844. (829) Liu, Y. M.; Zhao, X. Y.; Li, F.; Xia, D. G. Electrochim. Acta 2011, 56, 6448. (830) Li, S. R.; Sun, Y.; Ge, S. Y.; Qiao, Y.; Chen, Y. M.; Liebervvirth, I.; Yu, Y.; Chen, C. H. Chem. Eng. J. 2012, 192, 226. (831) Xu, G. L.; Xu, Y. F.; Sun, H.; Fu, F.; Zheng, X. M.; Huang, L.; Li, J. T.; Yang, S. H.; Sun, S. G. Chem. Commun. 2012, 48, 8502. (832) Zhang, X.; Xing, Z.; Wang, L. L.; Zhu, Y. C.; Li, Q. W.; Liang, J. W.; Yu, Y.; Huang, T.; Tang, K. B.; Qian, Y. T.; Shen, X. Y. J. Mater. Chem. 2012, 22, 17864. (833) Yang, G.; Li, Y. H.; Ji, H. M.; Wang, H. Y.; Gao, P.; Wang, L.; Liu, H. D.; Pinto, J.; Jiang, X. F. J. Power Sources 2012, 216, 353. (834) Hsieh, C. T.; Lin, C. Y.; Lin, J. Y. Electrochim. Acta 2011, 56, 8861. (835) Zhang, K. J.; Han, P. X.; Gu, L.; Zhang, L. X.; Liu, Z. H.; Kong, Q. S.; Zhang, C. J.; Dong, S. M.; Zhang, Z. Y.; Yao, J. H.; Xu, H. X.; Cui, G. L.; Chen, L. Q. ACS Appl. Mater. Interfaces 2012, 4, 658. (836) Mai, Y. J.; Zhang, D.; Qiao, Y. Q.; Gu, C. D.; Wang, X. L.; Tu, J. P. J. Power Sources 2012, 216, 201. (837) Qiu, D. F.; Ma, L. Y.; Zheng, M. B.; Lin, Z. X.; Zhao, B.; Wen, Z.; Hu, Z. B.; Pu, L.; Shi, Y. Mater. Lett. 2012, 84, 9. (838) Tang, Q. W.; Shan, Z. Q.; Wang, L.; Qin, X. Electrochim. Acta 2012, 79, 148. (839) Kokubu, T.; Oaki, Y.; Hosono, E.; Zhou, H.; Imai, H. Adv. Funct. Mater. 2011, 21, 3673. (840) Reddy, M. V.; Xiaoning; Loh, K. P.; Chowdari, B. V. R. 2012, manuscript in preparation. (841) Sun, Y.; Hu, X.; Luo, W.; Huang, Y. J. Mater. Chem. 2012, 22, 13826. (842) Sun, Y. M.; Hu, X. L.; Luo, W.; Huang, Y. H. J. Phys. Chem. C 2012, 116, 20794. (843) Zhang, G. H.; Chen, Y. J.; Qu, B. H.; Hu, L. L.; Mei, L.; Lei, D. N.; Li, Q.; Chen, L. B.; Li, Q. H.; Wang, T. H. Electrochim. Acta 2012, 80, 140. (844) Hwang, S. G.; Kim, G. O.; Yun, S. R.; Ryu, K. S. Electrochim. Acta 2012, 78, 406. (845) Wang, B.; Wu, X. L.; Shu, C. Y.; Guo, Y. G.; Wang, C. R. J. Mater. Chem. 2010, 20, 10661. (846) Binotto, G.; Larcher, D.; Prakash, A. S.; Urbina, R. H.; Hegde, M. S.; Tarascon, J. M. Chem. Mater. 2007, 19, 3032. (847) Liu, Y.; Mi, C. H.; Su, L. H.; Zhang, X. G. Electrochim. Acta 2008, 53, 2507. (848) Reddy, M. V.; Beichen, Z.; Nicholette, L. J.; Kaimeng, Z.; Chowdari, B. V. R. Electrochem. Solid-State Lett. 2011, 14, A79. (849) Pralong, V.; Leriche, J. B.; Beaudoin, B.; Naudin, E.; Morcrette, M.; Tarascon, J. M. Solid State Ionics 2004, 166, 295. (850) Varghese, B.; Hoong, T. C.; Yanwu, Z.; Reddy, M. V.; Chowdari, B. V. R.; Wee, A. T. S.; Vincent, T. B. C.; Lim, C. T.; Sow, C. H. Adv. Funct. Mater. 2007, 17, 1932. (851) Wang, G. X.; Chen, Y.; Konstantinov, K.; Yao, J.; A·hn, J. H.; Liu, H. K.; Dou, S. X. J. Alloys Compd. 2002, 340, L5. (852) Wu, Z. S.; Ren, W. C.; Wen, L.; Gao, L. B.; Zhao, J. P.; Chen, Z. P.; Zhou, G. M.; Li, F.; Cheng, H. M. ACS Nano 2010, 4, 3187. (853) Du, N.; Zhang, H.; Chen, B.; Wu, J. B.; Ma, X. Y.; Liu, Z. H.; Zhang, Y. Q.; Yang, D.; Huang, X. H.; Tu, J. P. Adv. Mater. 2007, 19, 4505. (854) Shaju, K. M.; Jiao, F.; Debart, A.; Bruce, P. G. Phys. Chem. Chem. Phys. 2007, 9, 1837. (855) Nam, K. T.; Kim, D. W.; Yoo, P. J.; Chiang, C. Y.; Meethong, N.; Hammond, P. T.; Chiang, Y. M.; Belcher, A. M. Science 2006, 312, 885. (856) Li, Y. G.; Tan, B.; Wu, Y. Y. Nano Lett. 2008, 8, 265. (857) Wang, Y.; Xia, H.; Lu, L.; Lin, J. Y. ACS Nano 2010, 4, 1425. (858) Zhu, J.; Sharma, Y. K.; Zeng, Z.; Zhang, X.; Srinivasan, M.; Mhaisalkar, S.; Zhang, H.; Hng, H. H.; Yan, Q. J. Phys. Chem. C 2011, 115, 8400. 5451

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(894) Mei, W. M.; Huang, J.; Zhu, L. P.; Ye, Z. Z.; Mai, Y. J.; Tu, J. P. J. Mater. Chem. 2012, 22, 9315. (895) Tummala, R.; Guduru, R. K.; Mohanty, P. S. J. Power Sources 2012, 199, 270. (896) Yang, X. L.; Fan, K. C.; Zhu, Y. H.; Shen, J. H.; Jiang, X.; Zhao, P.; Li, C. Z. J. Mater. Chem. 2012, 22, 17278. (897) Zhang, W. Y.; Zeng, Y.; Xiao, N.; Hng, H. H.; Yan, Q. Y. J. Mater. Chem. 2012, 22, 8455. (898) Zhu, J. J.; Ouyang, X. Y.; Lee, M. Y.; Davis, R. C.; Scott, S. L.; Fischer, A.; Thomas, A. RSC Adv. 2012, 2, 121. (899) Kim, K. S.; Park, Y. J. Solid State Ionics 2012, 225, 513. (900) Yoon, T. H.; Park, Y. J. Solid State Ionics 2012, 225, 498. (901) Ponrouch, A.; Palacin, M. R. J. Power Sources 2012, 212, 233. (902) Wang, H. G.; Ma, D. L.; Huang, X. L.; Huang, Y.; Zhang, X. B. Sci. Rep. 2012, 2. (903) Zhan, L.; Wang, Y. L.; Qiao, W. M.; Ling, L. C.; Yang, S. B. Electrochim. Acta 2012, 78, 440. (904) Li, W. Y.; Xu, L. N.; Chen, J. Adv. Funct. Mater. 2005, 15, 851. (905) Lavela, P.; Ortiz, G. F.; Tirado, J. L.; Zhecheva, E.; Stoyanova, R.; Ivanova, S. J. Phys. Chem. C 2007, 111, 14238. (906) Yang, S.; Feng, X.; Ivanovici, S.; Müllen, K. Angew. Chem., Int. Ed. 2010, 49, 8408. (907) Reddy, M. V.; Beichen, Z.; Loh, K. P.; Chowdari, B. V. R. CrystEngComm 2013, 15, 3568. (908) Reddy, M. V.; Yu, C.; Jiahuan, F.; Loh, K. P.; Chowdari, B. V. R. RSC Adv. 2012, 2, 9619. (909) Taberna, L.; Mitra, S.; Poizot, P.; Simon, P.; Tarascon, J. M. Nat. Mater. 2006, 5, 567. (910) Liu, H.; Wang, G. X.; Wang, J. Z.; Wexler, D. Electrochem. Commun. 2008, 10, 1879. (911) Wang, L.; Yu, Y.; Chen, P. C.; Zhang, D. W.; Chen, C. H. J. Power Sources 2008, 183, 717. (912) Zhang, W. M.; Wu, X. L.; Hu, J. S.; Guo, Y. G.; Wan, L. J. Adv. Funct. Mater. 2008, 18, 3941. (913) Muraliganth, T.; Murugan, A. V.; Manthiram, A. Chem. Commun. 2009, 47, 7360. (914) Wang, S. Q.; Zhang, J. Y.; Chen, C. H. J. Power Sources 2010, 195, 5379. (915) Zhou, G. M.; Wang, D. W.; Li, F.; Zhang, L. L.; Li, N.; Wu, Z. S.; Wen, L.; Lu, G. Q.; Cheng, H. M. Chem. Mater. 2010, 22, 5306. (916) He, Y.; Huang, L.; Cai, J. S.; Zheng, X. M.; Sun, S. G. Electrochim. Acta 2010, 55, 1140. (917) Zhang, M.; Yin, X.; Du, Z.; Liu, S.; Chen, L.; Li, Q.; Jin, H.; Peng, K.; Wang, T. Solid State Sci. 2010, 12, 2024. (918) Deng, Y. F.; Zhang, Q. M.; Shi, Z. C.; Han, L. J.; Peng, F.; Chen, G. H. Electrochim. Acta 2012, 76, 495. (919) Zhang, W. D.; Wang, X. Y.; Zhou, H. H.; Chen, J. T.; Zhang, X. X. J. Alloys Compd. 2012, 521, 39. (920) Lim, H. S.; Jung, B. Y.; Sun, Y. K.; Suh, K. D. Electrochim. Acta 2012, 75, 123. (921) Duan, H.; Gnanaraj, J.; Liang, J. Y. J. Power Sources 2011, 196, 4779. (922) Kim, H.; Seo, D.-H.; Kim, H.; Park, I.; Hong, J.; Park, K.-Y.; Kang, K. Chem. Mater. 2012, 24, 720. (923) Su, D. W.; Horvat, J.; Munroe, P.; A·hn, H.; Ranjbartoreh, A. R.; Wang, G. X. Chem.Eur. J. 2012, 18, 488. (924) Kim, H. S.; Baek, S. H.; Jang, M. W.; Sun, Y. K.; Yoon, C. S. J. Electrochem. Soc. 2012, 159, A325. (925) Zhao, X. Y.; Xia, D. G.; Zheng, K. J. Alloys Compd. 2012, 513, 460. (926) Yang, Z.; Shen, J.; Archer, L. A. J. Mater. Chem. 2011, 21, 11092. (927) Zhang, Q. M.; Shi, Z. C.; Deng, Y. F.; Zheng, J.; Liu, G. C.; Chen, G. H. J. Power Sources 2012, 197, 305. (928) Huang, Y.; Dong, Z.; Jia, D.; Guo, Z.; Cho, W. I. Electrochim. Acta 2011, 56, 9233. (929) Das, B.; Reddy, M. V.; Chowdari, B. V. R. J. Alloys Compd. 2013, DOI: 10.1016/j.jallcom.2013.02.072.

(930) Kang, E.; Jung, Y. S.; Cavanagh, A. S.; Kim, G.-H.; George, S. M.; Dillon, A. C.; Kim, J. K.; Lee, J. Adv. Funct. Mater. 2011, 21, 2430. (931) Chen, Y. J.; Xiao, G.; Wang, T. S.; Ouyang, Q. Y.; Qi, L. H.; Ma, Y.; Gao, P.; Zhu, C. L.; Cao, M. S.; Jin, H. B. J. Phys. Chem. C 2011, 115, 13603. (932) Jin, S. L.; Deng, H. G.; Long, D. H.; Liu, X. J.; Zhan, L. A.; Liang, X. Y.; Qiao, W. M.; Ling, L. C. J. Power Sources 2011, 196, 3887. (933) Wu, P.; Du, N.; Zhang, H.; Yu, J. X.; Yang, D. R. J. Phys. Chem. C 2011, 115, 3612. (934) Latorre-Sanchez, M.; Primo, A.; Garcia, H. J. Mater. Chem. 2012, 22, 21373. (935) Xue, X. Y.; Ma, C. H.; Cui, C. X.; Xing, L. L. Solid State Sci. 2011, 13, 1526. (936) Li, B.; Cao, H.; Shao, J.; Qu, M.; Warner, J. H. J. Mater. Chem. 2011, 21, 5069. (937) Lian, P.; Zhu, X.; Xiang, H.; Li, Z.; Yang, W.; Wang, H. Electrochim. Acta 2010, 56, 834. (938) Chen, D. Y.; Ji, G.; Ma, Y.; Lee, J. Y.; Lu, J. M. ACS Appl. Mater. Interfaces 2011, 3, 3078. (939) Chen, Y.; Song, B. H.; Tang, X. S.; Lu, L.; Xue, J. M. J. Mater. Chem. 2012, 22, 17656. (940) Cheng, J. P.; Yu, J.; Shi, D.; Wang, D. S.; Liu, Y. F.; Liu, F.; Zhang, X. B.; Li, J. G. Appl. Phys. A: Mater. Sci. Process. 2012, 106, 837. (941) Li, L.; Wang, T. T.; Zhang, L. Y.; Su, Z. M.; Wang, C. G.; Wang, R. S. Chem.Eur. J. 2012, 18, 11417. (942) Lu, L.; Wang, J. Z.; Gao, X. W.; Zhu, X. B.; Liu, H. K. J. Nanosci. Nanotechnol. 2012, 12, 1246. (943) Ma, Y.; Zhang, C.; Ji, G.; Lee, J. Y. J. Mater. Chem. 2012, 22, 7845. (944) Si, Y.; Ren, T.; Ding, B.; Yu, J. Y.; Sun, G. J. Mater. Chem. 2012, 22, 4619. (945) Xie, K. Y.; Lu, Z. G.; Huang, H. T.; Lu, W.; Lai, Y. Q.; Li, J.; Zhou, L. M.; Liu, Y. X. J. Mater. Chem. 2012, 22, 5560. (946) Xiong, Q. Q.; Lu, Y.; Wang, X. L.; Gu, C. D.; Qiao, Y. Q.; Tu, J. P. J. Alloys Compd. 2012, 536, 219. (947) Zhao, X. Y.; Xia, D. G.; Zheng, K. ACS Appl. Mater. Interfaces 2012, 4, 1350. (948) Pasero, D.; Reeves, N.; West, A. R. J. Power Sources 2005, 141, 156. (949) Fan, Q.; Whittingham, M. S. Electrochem. Solid-State Lett. 2007, 10, A48. (950) Wang, H. L.; Cui, L. F.; Yang, Y. A.; Casalongue, H. S.; Robinson, J. T.; Liang, Y. Y.; Cui, Y.; Dai, H. J. J. Am. Chem. Soc. 2010, 132, 13978. (951) Gao, J.; Lowe, M. A.; Abruña, H. D. Chem. Mater. 2011, 23, 3223. (952) Li, L.; Guo, Z. P.; Du, A. J.; Liu, H. K. J. Mater. Chem. 2012, 22, 3600. (953) Wang, D. W.; Li, Y. Q.; Wang, Q. H.; Wang, T. M. Eur. J. Inorg. Chem. 2012, 628. (954) Lavoie, N.; Malenfant, P. R. L.; Courtel, F. M.; Abu-Lebdeh, Y.; Davidson, I. J. J. Power Sources 2012, 213, 249. (955) Wang, C. B.; Yin, L. W.; Xiang, D.; Qi, Y. X. ACS Appl. Mater. Interfaces 2012, 4, 1636. (956) Wang, Z. H.; Yuan, L. X.; Shao, Q. G.; Huang, F.; Huang, Y. H. Mater. Lett. 2012, 80, 110. (957) Lee, Y. F.; Chang, K. H.; Hu, C. C.; Chu, Y. H. J. Power Sources 2012, 206, 469. (958) Sathish, M.; Tomai, T.; Honma, I. J. Power Sources 2012, 217, 85. (959) Zhang, J. J.; Yao, Y.; Huang, T.; Yu, A. S. Electrochim. Acta 2012, 78, 502. (960) Xiong, Q. Q.; Tu, J. P.; Lu, Y.; Chen, J.; Yu, Y. X.; Qiao, Y. Q.; Wang, X. L.; Gu, C. D. J. Phys. Chem. C 2012, 116, 6495. (961) Christie, C. T.; Reddy, M. V.; Sow, C. H.; Chowdari, B. V. R. 2012, manuscript in preparation. (962) Ke, F. S.; Huang, L.; Zhang, B.; Wei, G. Z.; Xue, L. J.; Li, J. T.; Sun, S. G. Electrochim. Acta 2012, 78, 585. 5452

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(963) Reddy, M. V.; Christie, T. C.; Ramanathan, K.; Daryl, T. Y. W.; Jie, K. C. W.; Hao, T. Y.; Chowdari, B. V. R. RSC Adv. 2013, submitted for publication. (964) Larcher, D.; Masquelier, C.; Bonnin, D.; Chabre, Y.; Masson, V.; Leriche, J. B.; Tarascon, J. M. J. Electrochem. Soc. 2003, 150, A133. (965) Larcher, D.; Bonnin, D.; Cortes, R.; Rivals, I.; Personnaz, L.; Tarascon, J. M. J. Electrochem. Soc. 2003, 150, A1643. (966) Morales, J.; Sanchez, L.; Martin, F.; Berry, F.; Ren, X. L. J. Electrochem. Soc. 2005, 152, A1748. (967) Hang, B. T.; Doi, T.; Okada, S.; Yamaki, J. I. J. Power Sources 2007, 174, 493. (968) Hang, B. T.; Hayashi, H.; Yoon, S. H.; Okada, S.; Yamaki, J. J. Power Sources 2008, 178, 393. (969) Hang, B. T.; Okada, S.; Yamaki, J. J. Power Sources 2008, 178, 402. (970) Chen, J.; Xu, L. N.; Li, W. Y.; Gou, X. L. Adv. Mater. 2005, 17, 582. (971) Jiao, F.; Bao, J. L.; Bruce, P. G. Electrochem. Solid-State Lett. 2007, 10, A264. (972) Li, J.; Dahn, H. M.; Krause, L. J.; Le, D. B.; Dahn, J. R. J. Electrochem. Soc. 2008, 155, A812. (973) Wang, P. C.; Ding, H. P.; Bark, T.; Chen, C. H. Electrochim. Acta 2007, 52, 6650. (974) Yanna, N.; Zhang, P.; Guo, Z.; Munroe, P.; Liu, H. Electrochim. Acta 2008, 53, 4213. (975) Liu, H.; Wang, G. X.; Park, J.; Wang, J.; Zhang, C. Electrochim. Acta 2009, 54, 1733. (976) Chou, S. L.; Wang, J. Z.; Wexler, D.; Konstantinov, K.; Zhong, C.; Liu, H. K.; Dou, S. X. J. Mater. Chem. 2010, 20, 2092. (977) Hassan, M. F.; Rahman, M. M.; Guo, Z. P.; Chen, Z. X.; Liu, H. K. Electrochim. Acta 2010, 55, 5006. (978) Wu, M. S.; Ou, Y. H.; Lin, Y. P. Electrochim. Acta 2010, 55, 3240. (979) Kim, J.; Chung, M. K.; Ka, B. H.; Ku, J. H.; Park, S.; Ryu, J.; Oh, S. M. J. Electrochem. Soc. 2010, 157, A412. (980) Hariharan, S.; Saravanan, K.; Balaya, P. Electrochem. Solid-State Lett. 2010, 13, A132. (981) Sarradin, J.; Guessous, A.; Ribes, M. J. Power Sources 1996, 62, 149. (982) Hosono, E.; Fujihara, S.; Honma, I.; Ichihara, M.; Zhou, H. S. J. Electrochem. Soc. 2006, 153, A1273. (983) Reddy, M. V.; Yu, T.; Sow, C. H.; Shen, Z. X.; Lim, C. T.; Subba Rao, G. V.; Chowdari, B. V. R. Adv. Funct. Mater. 2007, 17, 2792. (984) Wang, L.; Xu, H. W.; Chen, P. C.; Zhang, D. W.; Ding, C. X.; Chen, C. H. J. Power Sources 2009, 193, 846. (985) Ortiz, G. F.; Hanzu, I.; Lavela, P.; Tirado, J. L.; Knauth, P.; Djenizian, T. J. Mater. Chem. 2010, 20, 4041. (986) Ma, Y.; Ji, G.; Lee, J. Y. J. Mater. Chem. 2011, 21, 13009. (987) Hao, Q. Y.; Liu, S. A.; Yin, X. M.; Du, Z. F.; Zhang, M.; Li, L. M.; Wang, Y. G.; Wang, T. H.; Li, Q. H. CrystEngComm 2011, 13, 806. (988) Luo, Y. S.; Luo, J. S.; Jiang, J.; Zhou, W. W.; Yang, H. P.; Qi, X. Y.; Zhang, H.; Fan, H. J.; Yu, D. Y. W.; Li, C. M.; Yu, T. Energy Environ. Sci. 2012, 5, 6559. (989) Shi, C.; Cao, T.; Cao, C. B.; Wu, D. J.; Li, Y. A.; Wang, M. N.; Mo, W. Mater. Lett. 2012, 83, 35. (990) Ding, Y. H.; Li, J. X.; Zhao, Y.; Guan, L. H. Mater. Lett. 2012, 81, 105. (991) Shao, J.; Zhang, J. X.; Jiang, J. J.; Zhou, G. M.; Qu, M. Z. Electrochim. Acta 2011, 56, 7005. (992) Mi, H. Y.; Xu, Y. L.; Shi, W.; Yoo, H. D.; Chae, O. B.; Oh, S. M. Mater. Res. Bull. 2012, 47, 152. (993) Cherian, C. T.; Sundaramurthy, J.; Kalaivani, M.; Ragupathy, P.; Kumar, P. S.; Thavasi, V.; Reddy, M. V.; Sow, C. H.; Mhaisalkar, S. G.; Ramakrishna, S.; Chowdari, B. V. R. J. Mater. Chem. 2012, 22, 12198. (994) Zhao, Y.; Li, J. X.; Ding, Y. H.; Guan, L. H. Chem. Commun. 2011, 47, 7416.

(995) Ji, L. W.; Toprakci, O.; Alcoutlabi, M.; Yao, Y. F.; Li, Y.; Zhang, S.; Guo, B. K.; Lin, Z.; Zhang, X. W. ACS Appl. Mater. Interfaces 2012, 4, 2672. (996) Liu, Z. L.; Tay, S. W. Mater. Lett. 2012, 72, 74. (997) Zhou, G. M.; Wang, D. W.; Hou, P. X.; Li, W. S.; Li, N.; Liu, C.; Li, F.; Cheng, H. M. J. Mater. Chem. 2012, 22, 17942. (998) Li, J. X.; Zhao, Y.; Ding, Y. H.; Guan, L. H. RSC Adv. 2012, 2, 4205. (999) Zou, Y. Q.; Kan, J.; Wang, Y. J. Phys. Chem. C 2011, 115, 20747. (1000) Zhang, M.; Qu, B. H.; Lei, D. N.; Chen, Y. J.; Yu, X. Z.; Chen, L. B.; Li, Q. H.; Wang, Y. G.; Wang, T. H. J. Mater. Chem. 2012, 22, 3868. (1001) Shou, Q. L.; Cheng, J. P.; Zhang, L.; Nelson, B. J.; Zhang, X. B. J. Solid State Chem. 2012, 185, 191. (1002) Chen, D. Z.; Wei, W.; Wang, R. N.; Zhu, J. C.; Guo, L. New J. Chem. 2012, 36, 1589. (1003) Tian, L. L.; Zhuang, Q. C.; Li, J.; Wu, C.; Shi, Y. L.; Sun, S. G. Electrochim. Acta 2012, 65, 153. (1004) Hu, J.; Li, H.; Huang, X. J. Electrochem. Solid-State Lett. 2005, 8, A66. (1005) Grugeon, S.; Laruelle, S.; Dupont, L.; Chevallier, F.; Taberna, P. L.; Simon, P.; Gireaud, L.; Lascaud, S.; Vidal, E.; Yrieix, B.; Tarascon, J. M. Chem. Mater. 2005, 17, 5041. (1006) Dupont, L.; Grugeon, S.; Laruelle, S.; Tarascon, J. M. J. Power Sources 2007, 164, 839. (1007) Hu, J.; Li, H.; Huang, X. J.; Chen, L. Q. Solid State Ionics 2006, 177, 2791. (1008) Sun, J. P.; Tang, K.; Yu, X. Q.; Hu, J.; Li, H.; Huang, X. J. Solid State Ionics 2008, 179, 2390. (1009) Li, J. T.; Maurice, V.; Swiatowska-Mrowiecka, J.; Seyeux, A.; Zanna, S.; Klein, L.; Sun, S. G.; Marcus, P. Electrochim. Acta 2009, 54, 3700. (1010) Jiang, L. Y.; Xin, S.; Wu, X. L.; Li, H.; Guo, Y. G.; Wan, L. J. J. Mater. Chem. 2010, 20, 7565. (1011) Li, W. J.; Zhou, Y. N.; Fu, Z. W. J. Electrochem. Soc. 2010, 157, A957. (1012) Liu, H.; Du, X. W.; Xing, X. R.; Wang, G. X.; Qiao, S. Z. Chem. Commun. 2012, 48, 865. (1013) Guo, B. K.; Chi, M. F.; Sun, X. G.; Dai, S. J. Power Sources 2012, 205, 495. (1014) Liu, X. J.; Yasuda, H.; Yamachi, M. J. Power Sources 2005, 146, 510. (1015) Lavela, P.; Tirado, J. L.; Vidal-Abarca, C. Electrochim. Acta 2007, 52, 7986. (1016) Qiu, Y. C.; Xu, G. L.; Yan, K. Y.; Sun, H.; Xiao, J. W.; Yang, S. H.; Sun, S. G.; Jin, L. M.; Deng, H. J. Mater. Chem. 2011, 21, 6346. (1017) Deng, Y. F.; Li, Z. N.; Shi, Z. C.; Xu, H.; Peng, F.; Chen, G. H. RSC Adv. 2012, 2, 4645. (1018) Zhang, P.; Guo, Z. P.; Liu, H. K. Electrochim. Acta 2010, 55, 8521. (1019) Shao, J.; Zhang, J. X.; Jiang, J. J.; Zhou, G. M.; Qu, M. Z. Electrochim. Acta 2011, 56, 7005. (1020) Lin, Y. M.; Abel, P. R.; Heller, A.; Mullins, C. B. J. Phys. Chem. Lett. 2011, 2, 2885. (1021) Zou, Y. Q.; Kan, J.; Wang, Y. J. Phys. Chem. C 2011, 115, 20747. (1022) Zhu, X. J.; Zhu, Y. W.; Murali, S.; Stollers, M. D.; Ruoff, R. S. ACS Nano 2011, 5, 3333. (1023) Du, M.; Xu, C. H.; Sun, J.; Gao, L. Electrochim. Acta 2012, 80, 302. (1024) Xu, X.; Cao, R.; Jeong, S.; Cho, J. Nano Lett. 2012, 12, 4988. (1025) Jia, X.; Chen, J. J.; Xu, J. H.; Shi, Y. N.; Fan, Y. Z.; Zheng, M. S.; Dong, Q. F. Chem. Commun. 2012, 48, 7410. (1026) Yue, W. B.; Lin, Z. Z.; Jiang, S. H.; Yang, X. J. J. Mater. Chem. 2012, 22, 16318. (1027) Qiu, Y. C.; Xu, G. L.; Yan, K. Y.; Sun, H.; Xiao, J. W.; Yang, S. H.; Sun, S. G.; Jin, L. M.; Deng, H. J. Mater. Chem. 2011, 21, 6346. 5453

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(1028) Deng, Y.; Li, Z.; Shi, Z.; Xu, H.; Peng, F.; Chen, G. RSC Adv. 2012, 2, 4645. (1029) Wu, M. S.; Chiang, P. C. J.; Lee, J. T.; Lin, J. C. J. Phys. Chem. B 2005, 109, 23279. (1030) Zhao, J. Z.; Tao, Z. L.; Liang, J.; Chen, J. Cryst. Growth Des. 2008, 8, 2799. (1031) Reddy, A. L. M.; Shaijumon, M. M.; Gowda, S. R.; Ajayan, P. M. Nano Lett. 2009, 9, 1002. (1032) Xia, H.; Lai, M. O.; Lu, L. J. Mater. Chem. 2010, 20, 6896. (1033) Li, J. F.; Xi, B. J.; Zhu, Y. C.; Li, Q. W.; Yan, Y.; Qian, Y. T. J. Alloys Compd. 2011, 509, 9542. (1034) Guo, C. X.; Wang, M.; Chen, T.; Lou, X. W.; Li, C. M. Adv. Energy Mater 2011, 1, 736. (1035) Liang, Y. G.; Yang, S. J.; Yi, Z. H.; Lei, X. F.; Sun, J. T.; Zhou, Y. H. Mater. Sci. Eng., B 2005, 121, 152. (1036) Liang, Y. G.; Yang, S. J.; Yi, Z. H.; Sun, J. T.; Zhou, Y. H. Mater. Chem. Phys. 2005, 93, 395. (1037) Yang, L. C.; Gao, Q. S.; Zhang, Y. H.; Tang, Y.; Wu, Y. P. Electrochem. Commun. 2008, 10, 118. (1038) Yang, L. C.; Gao, Q. S.; Tang, Y.; Wu, Y. P.; Holze, R. J. Power Sources 2008, 179, 357. (1039) Shi, Y. F.; Guo, B. K.; Corr, S. A.; Shi, Q. H.; Hu, Y. S.; Heier, K. R.; Chen, L. Q.; Seshadri, R.; Stucky, G. D. Nano Lett. 2009, 9, 4215. (1040) Ku, J. H.; Jung, Y. S.; Lee, K. T.; Kim, C. H.; Oh, S. M. J. Electrochem. Soc. 2009, 156, A688. (1041) Gao, Q. S.; Yang, L. C.; Lu, X. C.; Mao, J. J.; Zhang, Y. H.; Wu, Y. P.; Tang, Y. J. Mater. Chem. 2010, 20, 2807. (1042) Wang, Z. Y.; Chen, J. S.; Zhu, T.; Madhavi, S.; Lou, X. W. Chem. Commun. 2010, 46, 6906. (1043) Yoon, S.; Manthiram, A. J. Mater. Chem. 2011, 21, 4082. (1044) Bhaskar, A.; Deepa, M.; Rao, T. N.; Varadaraju, U. V. J. Power Sources 2012, 216, 169. (1045) Reddy, M. V.; Yiming, W.; Loh, K. P.; Chowdari, B. V. R. 2012, manuscript in preparation. (1046) Guo, B. K.; Fang, X. P.; Li, B.; Shi, Y. F.; Ouyang, C. Y.; Hu, Y. S.; Wang, Z. X.; Stucky, G. D.; Chen, L. Q. Chem. Mater. 2012, 24, 457. (1047) Lei, Y. Z.; Hu, J. C.; Liu, H. W.; Li, J. L. Mater. Lett. 2012, 68, 82. (1048) Sun, Y. M.; Hu, X. L.; Luo, W.; Huang, Y. H. J. Mater. Chem. 2012, 22, 425. (1049) Xia, F. F.; Hu, X. L.; Sun, Y. M.; Luo, W.; Huang, Y. H. Nanoscale 2012, 4, 4707. (1050) Xu, Y.; Yi, R.; Yuan, B.; Wu, X. F.; Dunwell, M.; Lin, Q. L.; Fei, L.; Deng, S. G.; Andersen, P.; Wang, D. H.; Luo, H. M. J. Phys. Chem. Lett. 2012, 3, 309. (1051) Yang, L. C.; Liu, L. L.; Zhu, Y. S.; Wang, X. J.; Wu, Y. P. J. Mater. Chem. 2012, 22, 13148. (1052) Yoon, S.; Jung, K. N.; Jin, C. S.; Shin, K. H. J. Alloys Compd. 2012, 536, 179. (1053) Zhang, H. X.; Li, Y. F.; Hong, Z. S.; Wei, M. D. Mater. Lett. 2012, 79, 148. (1054) Zhao, X. Y.; Cao, M. H.; Liu, B.; Tian, Y.; Hu, C. W. J. Mater. Chem. 2012, 22, 13334. (1055) Balaya, P.; Li, H.; Kienle, L.; Maier, J. Adv. Funct. Mater. 2003, 13, 621. (1056) Delmer, O.; Balaya, P.; Kienle, L.; Maier, J. Adv. Mater. 2008, 20, 501. (1057) Bekaert, E.; Balaya, P.; Murugavel, S.; Maier, J.; Menetrier, M. Chem. Mater. 2009, 21, 856. (1058) Ji, H.-X.; Wu, X.-L.; Fan, L.-Z.; Krien, C.; Fiering, I.; Guo, Y.G.; Mei, Y.; Schmidt, O. G. Adv. Mater. 2010, 22, 4591. (1059) Alcantara, R.; Jaraba, M.; Lavela, P.; Tirado, J. L. Chem. Mater. 2002, 14, 2847. (1060) Thissen, A.; Ensling, D.; Madrigal, F. J. F.; Jaegermann, W.; Alcantara, R.; Lavela, P.; Tirado, J. L. Chem. Mater. 2005, 17, 5202.

(1061) Chadwick, A. V.; Savin, S. L. P.; Fiddy, S.; Alcantara, R.; Lisbona, D. F.; Lavela, P.; Ortiz, G. F.; Tirado, J. L. J. Phys. Chem. C 2007, 111, 4636. (1062) NuLi, Y. N.; Zhang, P.; Guo, Z. P.; Liu, H. K.; Yang, J. Electrochem. Solid-State Lett. 2008, 11, A64. (1063) Sharma, Y.; Srinivasan, M. In Solid State Physics, Pts 1 and 2; Mittal, R., Chauhan, A. K., Mukhopadhyay, R., Eds.; American Institute of Physics: Melville, NY, 2012; Vol. 1447. (1064) Reddy, M. V.; Yiming., X.; Tianyu., O.; Rajarajan., V.; Loh, K. P.; Chowdari, B. V. R. 2012, manuscript in preparation. (1065) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. Solid State Ionics 2008, 179, 587. (1066) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2007, 173, 495. (1067) Lavela, P.; Tirado, J. L.; Womes, M.; Jumas, J. C. J. Electrochem. Soc. 2009, 156, A589. (1068) Lavela, P.; Tirado, J. L. J. Power Sources 2007, 172, 379. (1069) Chu, Y. Q.; Fu, Z. W.; Qin, Q. Z. Electrochim. Acta 2004, 49, 4915. (1070) Jin, Y. H.; Seo, S. D.; Shim, H. W.; Park, K. S.; Kim, D. W. Nanotechnology 2012, 23, No. 125402. (1071) Wang, Y.; Park, J.; Sun, B.; A·hn, H.; Wang, G. Chem.Asian J. 2012, 7, 1940. (1072) NuLi, Y. N.; Qin, Q. Z. J. Power Sources 2005, 142, 292. (1073) Lavela, P.; Tirado, J. L.; Womes, M.; Jumas, J. C. J. Phys. Chem. C 2009, 113, 20081. (1074) Li, Z. H.; Zhao, T. P.; Zhan, X. Y.; Gao, D. S.; Xiao, Q. Z.; Lei, G. T. Electrochim. Acta 2010, 55, 4594. (1075) Vidal-Abarca, C.; Lavela, P.; Tirado, J. L. Solid State Ionics 2010, 181, 616. (1076) Vidal-Abarca, C.; Lavela, P.; Tirado, J. L. J. Power Sources 2011, 196, 6978. (1077) Wang, Y.; Park, J.; Sun, B.; Agn, H.; Wang, G. Chem.-Asian J. 2012, 7, 1940. (1078) Li, M. X.; Yin, Y. X.; Li, C. J.; Zhang, F. Z.; Wan, L. J.; Xu, S. L.; Evans, D. G. Chem. Commun. 2012, 48, 410. (1079) Wang, Y.; Su, D. W.; Ung, A.; A·hn, J. H.; Wang, G. X. Nanotechnology 2012, 23, No. 055402. (1080) Wu, L. J.; Xiao, Q. Z.; Li, Z. H.; Lei, G. T.; Zhang, P.; Wang, L. Solid State Ionics 2012, 215, 24. (1081) Xia, H.; Zhu, D. D.; Fu, Y. S.; Wang, X. Electrochim. Acta 2012, 83, 166. (1082) Alcantara, R.; Jaraba, M.; Lavela, P.; Tirado, J. L.; Jumas, J. C.; Olivier-Fourcade, J. Electrochem. Commun. 2003, 5, 16. (1083) Vidal-Abarca, C.; Lavela, P.; Tirado, J. L. J. Phys. Chem. C 2010, 114, 12828. (1084) Zhao, H. X.; Zheng, Z.; Wong, K. W.; Wang, S. M.; Huang, B. J.; Li, D. P. Electrochem. Commun. 2007, 9, 2606. (1085) Bomio, M.; Lavela, P.; Tirado, J. L. J. Solid State Electrochem. 2008, 12, 729. (1086) Jin, L. M.; Qiu, Y. C.; Deng, H.; Li, W. S.; Li, H.; Yang, S. H. Electrochim. Acta 2011, 56, 9127. (1087) Bomio, M.; Lavela, P.; Tirado, J. L. ChemPhysChem 2007, 8, 1999. (1088) Jin, L.; Qiu, Y.; Deng, H.; Li, W.; Li, H.; Yang, S. Electrochim. Acta 2011, 56, 9127. (1089) Ding, Y.; Yang, Y.; Shao, H. Solid State Ionics 2012, 217, 27. (1090) Zhou, L.; Zhao, D.; Lou, X. W. Adv. Mater. 2012, 24, 745. (1091) Yang, T.; Xia, D. G.; Wang, Z. L.; Chen, Y. Mater. Lett. 2009, 63, 5. (1092) Reddy, M. V.; Nicholas, T. Z. Y.; Soh, H.; Chowdari, B. V. R. 2012, manuscript in preparation. (1093) Li, X. D.; Yang, W. S.; Li, F.; Evans, D. G.; Duan, X. J. Phys. Chem. Solids 2006, 67, 1286. (1094) Fu, Y. S.; Wan, Y. H.; Xia, H.; Wang, X. J. Power Sources 2012, 213, 338. (1095) Cherian, C. T.; Reddy, M. V.; Subba Rao, G. V.; Sow, C. H.; Chowdari, B. V. R. J. Solid State Electrochem. 2012, 16, 1823. 5454

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(1096) Cherian, C. T.; Sundaramurthy, J.; Reddy, M. V.; Kalaivani, M.; Pliszka, D.; Ramakrishna, S.; Chowdari, B. V. R. ACS Applied Mater. & Inter. 2013, submitted for publication. (1097) Guyomard, D.; Sigala, C.; La Salle, A. L. G.; Piffard, S. Y. J. Power Sources 1997, 68, 692. (1098) Denis, S.; Baudrin, E.; Orsini, F.; Ouvrard, G.; Touboul, M.; Tarascon, J. M. J. Power Sources 1999, 81, 79. (1099) Orsini, F.; Baudrin, E.; Denis, S.; Dupont, L.; Touboul, M.; Guyomard, D.; Piffard, Y.; Tarascon, J. M. Solid State Ionics 1998, 107, 123. (1100) Han, X. Y.; Tang, W. C.; Yi, Z. H.; Sun, J. T. J. Appl. Electrochem. 2008, 38, 1671. (1101) Shirakawa, J.; Nakayama, M.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. Electrochem. Solid-State Lett. 2004, 7, A27. (1102) Shirakawa, J.; Nakayama, M.; Uchimoto, Y.; Wakihara, M. Electrochem. Solid-State Lett. 2006, 9, A200. (1103) Rossignol, C.; Ouvrard, G.; Baudrin, E. J. Electrochem. Soc. 2001, 148, A869. (1104) Lee, S. J.; Lee, H. Y.; Ha, T. S.; Baik, H. K.; Lee, S. M. Electrochem. Solid-State Lett. 2002, 5, A138. (1105) Reddy, M. V.; Wannek, C.; Pecquenard, B.; Vinatier, P.; Levasseur, A. J. Power Sources 2003, 119, 101. (1106) Tang, S. B.; Xia, H.; Lai, M. O.; Lu, L. J. Electrochem. Soc. 2006, 153, A875. (1107) Reddy, M. V.; Pecquenard, B.; Vinatier, P.; Levasseur, A. J. Phys. Chem. B 2006, 110, 4301. (1108) Reddy, M. V.; Pecquenard, B.; Vinatier, P.; Levasseur, A. Electrochem. Commun. 2007, 9, 409. (1109) Reddy, M. V.; Levasseur, A. J. Electroanal. Chem. 2010, 639, 27. (1110) Reddy, M. V.; Xing, L. H.; Chowdari, B. V. R. 2012, manuscript in preparation. (1111) Denis, S.; Baudrin, E.; Touboul, M.; Tarascon, J. M. J. Electrochem. Soc. 1997, 144, 4099. (1112) Denis, S.; Dedryvere, R.; Baudrin, E.; Laruelle, S.; Touboul, M.; Olivier-Fourcade, J.; Jumas, J. C.; Tarascon, J. M. Chem. Mater. 2000, 12, 3733. (1113) Poizot, P.; Baudrin, E.; Laruelle, S.; Dupont, L.; Touboul, M.; Tarascon, J. M. Solid State Ionics 2000, 138, 31. (1114) Ali, E. B.; Bernede, J. C.; Guyomard, D. Thin Solid Films 2002, 402, 215. (1115) Reddy, M. V.; Lee, B.; Loh, K. P.; Chowdari, B. V. R. ACS Applied Mater. & Inter. 2013, submitted for publication. (1116) Piffard, Y.; Leroux, F.; Guyomard, D.; Mansot, J. L.; Tournoux, M. J. Power Sources 1997, 68, 698. (1117) Leroux, F.; Piffard, Y.; Ouvrard, G.; Mansot, J. L.; Guyomard, D. Chem. Mater. 1999, 11, 2948. (1118) Kim, S. S.; Ikuta, H.; Wakihara, M. Solid State Ionics 2001, 139, 57. (1119) Hara, D.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. J. Mater. Chem. 2002, 12, 2507. (1120) Hara, D.; Shirakawa, J.; Ikuta, H.; Uchimoto, Y.; Wakihara, M.; Miyanaga, T.; Watanabe, I. J. Mater. Chem. 2002, 12, 3717. (1121) Morishita, T.; Konno, H.; Izumi, Y.; Inagaki, M. Solid State Ionics 2006, 177, 1347. (1122) Lei, S. J.; Tang, K. B.; Jin, Y.; Chen, C. H. Nanotechnology 2007, 18, No. 175605. (1123) Liu, R. S.; Wang, C. Y.; Drozd, V. A.; Hu, S. F.; Sheu, H. S. Electrochem. Solid-State Lett. 2005, 8, A650. (1124) Kim, S. S.; Ogura, S.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. Chem. Lett. 2001, 760. (1125) Kim, S. S.; Ogura, S.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. Solid State Ionics 2002, 146, 249. (1126) Leyzerovich, N. N.; Bramnik, K. G.; Buhrmester, T.; Ehrenberg, H.; Fuess, H. J. Power Sources 2004, 127, 76. (1127) Sharma, N.; Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R.; Dong, Z. L.; White, T. J. Chem. Mater. 2004, 16, 504. (1128) Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochim. Acta 2005, 50, 5305.

(1129) Liang, Y. G.; Han, X. Y.; Yi, Z. H.; Tang, W. C.; Zhou, L. Q.; Sun, J. T.; Yang, S. J.; Zhou, Y. H. J. Solid State Electrochem. 2007, 11, 1127. (1130) Cherian, C. T.; Reddy, M. V.; Sow, C. H.; Chowdari, B. V. R. ACS Appl. Mater. & Inter. 2013, 5, 918. (1131) Leroux, F.; Goward, G. R.; Power, W. P.; Nazar, L. F. Electrochem. Solid-State Lett. 1998, 1, 255. (1132) Leroux, F.; Nazar, L. F. Solid State Ionics 2000, 133, 37. (1133) Lee, S. H.; Kim, Y. H.; Deshpande, R.; Parilla, P. A.; Whitney, E.; Gillaspie, D. T.; Jones, K. M.; Mahan, A. H.; Zhang, S. B.; Dillon, A. C. Adv. Mater. 2008, 20, 3627. (1134) Jung, Y. S.; Lee, S.; A·hn, D.; Dillon, A. C.; Lee, S. H. J. Power Sources 2009, 188, 286. (1135) Riley, L. A.; Lee, S. H.; Gedvilias, L.; Dillon, A. C. J. Power Sources 2010, 195, 588. (1136) Riley, L. A.; Cavanagh, A. S.; George, S. M.; Jung, Y. S.; Yan, Y. F.; Lee, S. H.; Dillon, A. C. ChemPhysChem 2010, 11, 2124. (1137) Hassan, M. F.; Guo, Z. P.; Chen, Z.; Liu, H. K. J. Power Sources 2010, 195, 2372. (1138) Meduri, P.; Clark, E.; Kim, J. H.; Dayalan, E.; Sumanasekera, G. U.; Sunkara, M. K. Nano Lett. 2012, 12, 1784. (1139) Sasidharan, M.; Gunawardhana, N.; Noma, H.; Yoshio, M.; Nakashima, K. Bull. Chem. Soc. Jpn. 2012, 85, 642. (1140) Wang, Z. Y.; Madhavi, S.; Lou, X. W. J. Phys. Chem. C 2012, 116, 12508. (1141) Das, B.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. J. Solid State Electrochem. 2008, 12, 953. (1142) Das, B.; Reddy, M. V.; Krishnamoorthi, C.; Tripathy, S.; Mahendiran, R.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochim. Acta 2009, 54, 3360. (1143) Das, B.; Reddy, M. V.; Subba Rao, G. V.; Chowdari, B. V. R. Solid State Ionics: Fundamental Researches and Technological Applications; Chowdari, B.V.R. et al., Eds.; Proceedings of the 12th Asian Conference on Solid State Ionics; Wuhan University of Technology Press: Wuhan, China, 2010; p 820. (1144) Sun, Y. M.; Hu, X. H.; Luo, W.; Huang, Y. H. J. Mater. Chem. 2011, 21, 17229. (1145) Duncan, M. J.; Nazar, L. F. Mater. Res. Soc. Symp. Proc. 1999, 548, 71. (1146) Sharma, N.; Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2003, 124, 204. (1147) Sharma, N.; Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochim. Acta 2004, 49, 1035. (1148) Kim, D. W.; Ko, Y. D.; Park, J. G.; Kim, B. K. Angew. Chem., Int. Ed. 2007, 46, 6654. (1149) Zhu, X. B.; Chou, S. L.; Wang, L.; Li, Q.; Shi, D. Q.; Wang, J. Z.; Chen, Z. X.; Sun, Y. P.; Liu, H. K.; Dou, S. X. Electrochem. SolidState Lett. 2009, 12, A176. (1150) Ko, Y. D.; Kang, J. G.; Choi, K. J.; Park, J. G.; A·hn, J. P.; Chung, K. Y.; Nam, K. W.; Yoon, W. S.; Kim, D. W. J. Mater. Chem. 2009, 19, 1829. (1151) Li, J.; Dahn, H. M.; Sanderson, R. J.; Todd, A. D. W.; Dahn, J. R. J. Electrochem. Soc. 2008, 155, A975. (1152) Rowsell, J. L. C.; Gaubicher, J.; Nazar, L. F. J. Power Sources 2001, 97−98, 254. (1153) Palos, A. I.; Morcrette, M.; Strobel, P. J. Solid State Electrochem. 2002, 6, 134. (1154) Ibarra-Palos, A.; Darie, C.; Proux, O.; Hazemann, J. L.; Aldon, L.; Jumas, J. C.; Morcrette, M.; Strobel, P. Chem. Mater. 2002, 14, 1166. (1155) Okada, S.; Tonuma, T.; Uebo, Y.; Yamaki, J. J. Power Sources 2003, 119, 621. (1156) Kim, D. Y.; Song, M. S.; Eom, J. Y.; Kwon, H. S. J. Alloys Compd. 2012, 542, 132. (1157) Rowsell, J. L. C.; Nazar, L. F. J. Mater. Chem. 2001, 11, 3228. (1158) Shi, X. X.; Chang, C. X.; Xiang, J. F.; Xiao, Y.; Yuan, L. J.; Sun, J. T. J. Solid State Chem. 2008, 181, 2231. (1159) Debart, A.; Revel, B.; Dupont, L.; Montagne, L.; Leriche, J. B.; Touboul, M.; Tarascon, J. M. Chem. Mater. 2003, 15, 3683. 5455

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(1160) Aragon, M. J.; Leon, B.; Vicente, C. P.; Tirado, J. L. Inorg. Chem. 2008, 47, 10366. (1161) Aragon, M. J.; Leon, B.; Vicente, C. P.; Tirado, J. L.; Chadwick, A. V.; Berko, A.; Beh, S. Y. Chem. Mater. 2009, 21, 1834. (1162) Aragon, M. J.; Leon, B.; Serrano, T.; Vicente, C. P.; Tirado, J. L. J. Mater. Chem. 2011, 21, 10102. (1163) Lou, X. M.; Wu, X. Z.; Zhang, Y. X. Electrochem. Commun. 2009, 11, 1696. (1164) Reddy, M. V.; Madhavi, S.; Subba Rao, G. V.; Chowdari, B. V. R. J. Power Sources 2006, 162, 1312. (1165) Chen, L.; Shen, L.; Nie, P.; Zhang, X.; Li, H. Electrochim. Acta 2012, 62, 408. (1166) Zeng, Y.; Zhang, W. Y.; Xu, C.; Xiao, N.; Huang, Y. Z.; Yu, D. Y. W.; Hng, H. H.; Yan, Q. Y. Chem.Eur. J. 2012, 18, 4026. (1167) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. Adv. Funct. Mater. 2007, 17, 2855. (1168) Qiu, Y. C.; Yang, S. H.; Deng, H.; Jin, L. M.; Li, W. S. J. Mater. Chem. 2010, 20, 4439. (1169) Du, N.; Xu, Y. F.; Zhang, H.; Yu, J. X.; Zhai, C. X.; Yang, D. R. Inorg. Chem. 2011, 50, 3320. (1170) Reddy, M. V.; Kenrick, K. Y. H.; Wei, T. Y.; Chong, G. Y.; Leong, G. H.; Chowdari, B. V. R. J. Electrochem. Soc. 2011, 158, A1423. (1171) Liu, B.; Zhang, J.; Wang, X. F.; Chen, G.; Chen, D.; Zhou, C. W.; Shen, G. Z. Nano Lett. 2012, 12, 3005. (1172) Luo, W.; Hu, X. L.; Sun, Y. M.; Huang, Y. H. J. Mater. Chem. 2012, 22, 8916. (1173) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. Electrochim. Acta 2008, 53, 2380. (1174) NuLi, Y. N.; Chu, Y. Q.; Qin, Q. Z. J. Electrochem. Soc. 2004, 151, A1077. (1175) Guo, X. W.; Lu, X.; Fang, X. P.; Mao, Y.; Wang, Z. X.; Chen, L. Q.; Xu, X. X.; Yang, H.; Liu, Y. N. Electrochem. Commun. 2010, 12, 847. (1176) Ding, Y.; Yang, Y. F.; Shao, H. X. Electrochim. Acta 2011, 56, 9433. (1177) Sui, J. H.; Zhang, C.; Hong, D.; Li, J.; Cheng, Q.; Li, Z. G.; Cai, W. J. Mater. Chem. 2012, 22, 13674. (1178) Chen, X. L.; Cheng, B.; Xu, H. Y.; Yang, J.; Qian, Y. T. Chem. Lett. 2012, 41, 639. (1179) Xing, Z.; Ju, Z. C.; Yang, J.; Xu, H. Y.; Qian, Y. T. Nano Res. 2012, 5, 477. (1180) Xu, H. Y.; Chen, X. L.; Chen, L.; Li, L. E.; Xu, L. Q.; Yang, J.; Qian, Y. T. Int. J. Electrochem. Sci. 2012, 7, 7976. (1181) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. Bull. Mater. Sci. 2009, 32, 295. (1182) Yang, Y. Y.; Zhao, Y. Q.; Xiao, L. F.; Zhang, L. Z. Electrochem. Commun. 2008, 10, 1117. (1183) Xiao, L. F.; Yang, Y. Y.; Yin, J.; Li, Q.; Zhang, L. Z. J. Power Sources 2009, 194, 1089. (1184) Courtel, F. M.; Duncan, H.; Abu-Lebdeh, Y.; Davidson, I. J. J. Mater. Chem. 2011, 21, 10206. (1185) Deng, Y.; Tang, S.; Zhang, Q.; Shi, Z.; Zhang, L.; Zhan, S.; Chen, G. J. Mater. Chem. 2011, 21, 11987. (1186) Zhao, J.; Wang, F. Q.; Su, P. P.; Li, M. R.; Chen, J.; Yang, Q. H.; Li, C. J. Mater. Chem. 2012, 22, 13328. (1187) Zhou, L.; Wu, H. B.; Zhu, T.; Lou, X. W. J. Mater. Chem. 2012, 22, 827. (1188) Zhang, G. Q.; Yu, L.; Bin Wu, H.; Hoster, H. E.; Lou, X. W. Adv. Mater. 2012, 24, 4609. (1189) Courtel, F. M.; Abu-Lebdeh, Y.; Davidson, I. J. Electrochim. Acta 2012, 71, 123. (1190) Xiao, L. F.; Zhao, Y. Q.; Yin, J.; Zhang, L. Z. Chem.Eur. J. 2009, 15, 9442. (1191) Zeng, L. X.; Xiao, F. Y.; Wang, J. C.; Gao, S. K.; Ding, X. K.; Wei, M. D. J. Mater. Chem. 2012, 22, 14284. (1192) Connor, P. A.; Irvine, J. T. S. Electrochim. Acta 2002, 47, 2885. (1193) Alcantara, R.; Ortiz, G. F.; Lavela, P.; Tirado, J. L. Electrochem. Commun. 2006, 8, 731.

(1194) Wang, G.; Gao, X. P.; Shen, P. W. J. Power Sources 2009, 192, 719. (1195) Qi, Y.; Du, N.; Zhang, H.; Wu, P.; Yang, D. R. J. Power Sources 2011, 196, 10234. (1196) Rong, A.; Gao, X. P.; Li, G. R.; Yan, T. Y.; Zhu, H. Y.; Qu, J. Q.; Song, D. Y. J. Phys. Chem. B 2006, 110, 14754. (1197) Zhu, X. J.; Geng, L. M.; Zhang, F. Q.; Liu, Y. X.; Cheng, L. B. J. Power Sources 2009, 189, 828. (1198) Hou, X. H.; Cheng, Q. A.; Bai, Y.; Zhang, W. F. Solid State Ionics 2010, 181, 631. (1199) Yuan, W. S.; Tian, Y. W.; Liu, G. Q. J. Alloys Compd. 2010, 506, 683. (1200) Becker, S. M.; Scheuermann, M.; Sepelak, V.; Eichhofer, A.; Chen, D.; Monig, R.; Ulrich, A. S.; Hahn, H.; Indris, S. Phys. Chem. Chem. Phys. 2011, 13, 19624. (1201) Yuan, W. S.; Tian, Y. W.; Liu, G. Q. Russ. J. Electrochem 2011, 47, 829. (1202) Feng, N.; Peng, S. L.; Sun, X. L.; Qiao, L.; Li, X. W.; Wang, P.; Hu, D. K.; He, D. Y. Mater. Lett. 2012, 76, 66. (1203) Kim, K.; Annamalai, A.; Park, S. H.; Kwon, T. H.; Pyeon, M. W.; Lee, M. J. Electrochim. Acta 2012, 76, 192. (1204) Zheng, X. Z.; Li, Y. F.; Xu, Y. X.; Hong, Z. S.; Wei, M. D. CrystEngComm 2012, 14, 2112. (1205) Sepelak, V.; Becker, S. M.; Bergmann, I.; Indris, S.; Scheuermann, M.; Feldhoff, A.; Kubel, C.; Bruns, M.; Sturzl, N.; Ulrich, A. S.; Ghafari, M.; Hahn, H.; Grey, C. P.; Becker, K. D.; Heitjans, P. J. Mater. Chem. 2012, 22, 3117. (1206) Christie, T. C.; Zheng, M.; Reddy, M. V.; Sow, C. H.; Chowdari, B. V. R. ACS Applied Materials & Interfaces 2013, DOI: 10.1021/am400802j. (1207) Lei, S. J.; Tang, K. B.; Chen, C. H.; Jin, Y.; Zhou, L. Mater. Res. Bull. 2009, 44, 393. (1208) Satya Kishore, M. V. V. M.; Varadaraju, U. V.; Raveau, B. J. Solid State Chem. 2004, 177, 3981. (1209) Sharma, Y.; Sharma, N.; Subba Rao, G. V.; Chowdari, B. V. R. J. Mater. Chem. 2009, 19, 5047. (1210) Aragon, M. J.; Leon, B.; Vicente, C. P.; Tirado, J. L. J. Power Sources 2011, 196, 2863. (1211) Mirhashemihaghighi, S.; Leon, B.; Vicente, C. P.; Tirado, J. L.; Stoyanova, R.; Yoncheva, M.; Zhecheva, E.; Puche, R. S.; Arroyo, E. M.; de Paz, J. R. Inorg. Chem. 2012, 51, 5554. (1212) Alcantara, R.; Madrigal, F. J. F.; Lavela, P.; Perez-Vicente, C.; Tirado, J. L. J. Solid State Electrochem. 2001, 6, 55. (1213) Ferg, E.; Gummow, R. J.; Dekock, A.; Thackeray, M. M. J. Electrochem. Soc. 1994, 141, L147. (1214) Jansen, A. N.; Kahaian, A. J.; Kepler, K. D.; Nelson, P. A.; Amine, K.; Dees, D. W.; Vissers, D. R.; Thackeray, M. M. J. Power Sources 1999, 81, 902. (1215) Stux, A. M.; Swider-Lyons, K. E. J. Power Sources 2007, 165, 635. (1216) Ohzuku, T.; Yamato, R.; Kawai, T.; Ariyoshi, K. J. Solid State Electrochem. 2008, 12, 979. (1217) Ariyoshi, K.; Ohzuku, T. J. Power Sources 2007, 174, 1258. (1218) Lu, W.; Belharouak, I.; Liu, J.; Amine, K. J. Power Sources 2007, 174, 673. (1219) Belharouak, I.; Sun, Y. K.; Lu, W.; Amine, K. J. Electrochem. Soc. 2007, 154, A1083. (1220) Du Pasquier, A.; Huang, C. C.; Spitler, T. J. Power Sources 2009, 186, 508. (1221) Ionica-Bousquet, C. M.; Munoz-Rojas, D.; Casteel, W. J.; Pearlstein, R. M.; GirishKumar, G.; Pez, G. P.; Palacin, M. R. J. Power Sources 2010, 195, 1479. (1222) Amine, K.; Belharouak, I.; Chen, Z. H.; Tran, T.; Yumoto, H.; Ota, N.; Myung, S. T.; Sun, Y. K. Adv. Mater. 2010, 22, 3052. (1223) Imazaki, M.; Wang, L. N.; Kawai, T.; Ariyoshi, K.; Ohzuku, T. Electrochim. Acta 2011, 56, 4576. (1224) Sawai, K.; Yamato, R.; Ohzuku, T. Electrochim. Acta 2006, 51, 1651. 5456

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457

Chemical Reviews

Review

(1225) Lu, W.; Liu, J.; Sun, Y. K.; Amine, K. J. Power Sources 2007, 167, 212. (1226) Reale, P.; Panero, S.; Scrosati, B.; Garche, J.; WohlfahrtMehrens, M.; Wachtler, M. J. Electrochem. Soc. 2004, 151, A2138. (1227) Reale, P.; Fernicola, A.; Scrosati, B. J. Power Sources 2009, 194, 182. (1228) Sun, L. Q.; Cui, R. H.; Jalbout, A. F.; Li, M. J.; Pan, X. M.; Wang, R. S.; Xie, H. M. J. Power Sources 2009, 189, 522. (1229) Jaiswal, A.; Horne, C. R.; Chang, O.; Zhang, W.; Kong, W.; Wang, E.; Chern, T.; Doeff, M. M. J. Electrochem. Soc. 2009, 156, A1041. (1230) Moshuchak, L. M.; Bulinski, M.; Lamanna, W. M.; Wang, R. L.; Dahn, J. R. Electrochem. Commun. 2007, 9, 1497. (1231) Martha, S. K.; Haik, O.; Borgel, V.; Zinigrad, E.; Exnar, I.; Drezen, T.; Miners, J. H.; Aurbach, D. J. Electrochem. Soc. 2011, 158, A790. (1232) Wakihara, M.; Kadoma, Y.; Kumagai, N.; Mita, H.; Araki, R.; Ozawa, K.; Ozawa, Y. J. Solid State Electrochem. 2012, 16, 847. (1233) Zaghib, K.; Dontigny, M.; Guerfi, A.; Trottier, J.; HamelPaquet, J.; Gariepy, V.; Galoutov, K.; Hovington, P.; Mauger, A.; Groult, H.; Julien, C. M. J. Power Sources 2012, 216, 192. (1234) Barker, J.; Gover, R. K. B.; Burns, P.; Bryan, A. J. Electrochem. Solid-State Lett. 2007, 10, A130. (1235) Barker, J.; Gover, R. K. B.; Burns, P.; Bryan, A. J. J. Electrochem. Soc. 2007, 154, A882. (1236) Ohzuku, T.; Ariyoshi, K.; Yamamoto, S.; Makimura, Y. Chem. Lett. 2001, 1270. (1237) Ariyoshi, K.; Yamamoto, S.; Ohzuku, T. J. Power Sources 2003, 119, 959. (1238) Xiang, H. F.; Jin, Q. Y.; Wang, R.; Chen, C. H.; Ge, X. W. J. Power Sources 2008, 179, 351. (1239) Xiang, H. F.; Zhang, X.; Jin, Q. Y.; Zhang, C. P.; Chen, C. H.; Ge, X. W. J. Power Sources 2008, 183, 355. (1240) Wu, H. M.; Belharouak, I.; Deng, H.; Abouimrane, A.; Sun, Y. K.; Amine, K. J. Electrochem. Soc. 2009, 156, A1047. (1241) Jung, H. G.; Jang, M. W.; Hassoun, J.; Sun, Y. K.; Scrosati, B. Nat. Commun 2011, 2, 516. (1242) Armstrong, G.; Armstrong, A. R.; Bruce, P. G.; Reale, P.; Scrosati, B. Adv. Mater. 2006, 18, 2597. (1243) Cao, F. F.; Wu, X. L.; Xin, S.; Guo, Y. G.; Wan, L. J. J. Phys. Chem. C 2010, 114, 10308. (1244) Ding, N.; Feng, X. Y.; Liu, S. H.; Xu, J.; Fang, X.; Lieberwirth, I.; Chen, C. H. Electrochem. Commun. 2009, 11, 538. (1245) Brutti, S.; Gentili, V.; Reale, P.; Carbone, L.; Panero, S. J. Power Sources 2011, 196, 9792. (1246) Han, J.-T.; Goodenough, J. B. Chem. Mater. 2011, 23, 3404. (1247) Appetecchi, G. B.; Croce, F.; Marassi, R.; Panero, S.; Ronci, F.; Savo, G.; Scrosati, B. Solid State Ionics 2001, 143, 73. (1248) Wang, X. L.; Han, W. Q.; Chen, H. Y.; Bai, J. M.; Tyson, T. A.; Yu, X. Q.; Wang, X. J.; Yang, X. Q. J. Am. Chem. Soc. 2011, 133, 20692. (1249) Hassoun, J.; Croce, F.; Hong, I.; Scrosati, B. Electrochem. Commun. 2011, 13, 228. (1250) Shanmukaraj, D.; Grugeon, S.; Laruelle, S.; Douglade, G.; Tarascon, J. M.; Armand, M. Electrochem. Commun. 2010, 12, 1344. (1251) Li, Y.; Fitch, B. Electrochem. Commun. 2011, 13, 664.

NOTE ADDED AFTER ASAP PUBLICATION This paper was published to the Web on April 2, 2013, with minor errors in Dr. Reddy’s biography and the reference section. The corrected version of this paper was published to the Web on June 14, 2013.

5457

dx.doi.org/10.1021/cr3001884 | Chem. Rev. 2013, 113, 5364−5457