Highly Reversible Mg Insertion in Nanostructured Bi for Mg Ion

Nov 26, 2013 - (1-5) This technology uses earth-abundant Mg element with a bivalent charge carrier Mg2+ which could potentially lead to low cost and h...
1 downloads 15 Views 4MB Size
Letter pubs.acs.org/NanoLett

Highly Reversible Mg Insertion in Nanostructured Bi for Mg Ion Batteries Yuyan Shao,* Meng Gu, Xiaolin Li, Zimin Nie, Pengjian Zuo, Guosheng Li, Tianbiao Liu, Jie Xiao, Yingwen Cheng, Chongmin Wang, Ji-Guang Zhang, and Jun Liu* Pacific Northwest National Laboratory, Richland, Washington 99352, United States S Supporting Information *

ABSTRACT: Rechargeable magnesium batteries have attracted wide attention for energy storage. Currently, most studies focus on Mg metal as the anode, but this approach is still limited by the properties of the electrolyte and poor control of the Mg plating/stripping processes. This paper reports the synthesis and application of Bi nanotubes as a highperformance anode material for rechargeable Mg ion batteries. The nanostructured Bi anode delivers a high reversible specific capacity (350 mAh/gBi or 3430 mAh/cm3Bi), excellent stability, and high Coulombic efficiency (95% initial and very close to 100% afterward). The good performance is attributed to the unique properties of in situ formed, interconnected nanoporous bismuth. Such nanostructures can effectively accommodate the large volume change without losing electric contact and significantly reduce diffusion length for Mg2+. Significantly, the nanostructured Bi anode can be used with conventional electrolytes which will open new opportunities to study Mg ion battery chemistry and further improve its properties. KEYWORDS: Energy storage, magnesium battery, anode, bismuth nanotube, insertion

M

The revolution and large scale commercialization of Li-ion batteries were facilitated by the successful development of the intercalation and insertion materials for both cathode and anode.25−31 Significant progresses have been made in the synthesis, understanding and utilization of such intercalation and insertion materials.27,32−35 Extensive knowledge has been gained on the chemistry, properties and optimization of electrolytes for intercalation and insertion electrode materials.15,36 If such mechanisms can be successfully used in Mg chemistry, the battery may not be limited by the Mg plating/ stripping reaction and the special electrolytes associated with such reactions. Instead, conventional electrolytes commonly used in Li-ion batteries could be used. The new approach will significantly widen the choices of the electrode and electrolyte materials, provide great opportunity to gain better understanding of the reaction mechanisms involved Mg chemistry, and may lead to new cost-effective materials and technologies. Recently, a few groups studied bismuth and tin as the anode materials for rechargeable Mg-ion batteries.37−39 These materials still have poor rate performance and fast capacity fading upon cycling. Taking tin as an example, only ∼20% of its theoretical capacity was obtained at the rate C/20 (discharge in 20 h), and the capacity fading is fast.38 The Coulombic

agnesium-based batteries have attracted increasing interest in the past few years.1−5 This technology uses earth-abundant Mg element with a bivalent charge carrier Mg2+ which could potentially lead to low cost and high energy density. To date, most efforts in this field have been focused on Mg metal anode based rechargeable Mg battery1,4 by developing high-performance electrolytes5−9 and cathodes.2,10−14 The use of Mg metal is still limited by the properties of the electrolyte and the poor control of the electrochemical reaction at the metal−electrolyte interfaces.1,4 Conventional electrolytes made by mixing simple Mg salts (e.g., Mg(ClO4)2) and nonaqueous solvents (e.g., propylene carbonate), as widely used in lithium batteries,15 do not produce reversible plating/ stripping of Mg.16,17 This is likely due to a nonconductive layer formed on Mg surface in these conventional electrolytes.18 Specially designed and synthesized electrolytes are needed for Mg metal anode.4,19 So far, there are only a limited number of electrolytes that show reversible Mg plating/stripping.5−7,20 The synthesis of these electrolytes usually requires strictly controlled chemical processes.5,6,9 Low electrochemical stable window,1,4,21 the corrosive nature of the electrolytes21−24 and high volatility, usually involving solvents such as tetrahydrofuran (THF),4,5,7,9 are also significant limiting factors. A new approach that does not rely on the special electrolytes associated with Mg plating/stripping could lead to significant breakthroughs in this field. © 2013 American Chemical Society

Received: October 16, 2013 Revised: November 20, 2013 Published: November 26, 2013 255

dx.doi.org/10.1021/nl403874y | Nano Lett. 2014, 14, 255−260

Nano Letters

Letter

efficiency for Mg insertion/deinsertion is also very low.38 Such problems are largely associated with the slow solid-state diffusion of Mg ions in the host2 and large volume expansion/shrinkage during Mg insertion/deinsertion processes. Substantial improvements and structural optimization are still required. In the last few decades, nanostructured materials have shown great promise in improving the ion diffusion and structural stability of electrode materials in Li-ion batteries.40−51 In this paper we demonstrate that nanostructured material can significantly improve the properties of anode for Mg-ion batteries. Such anode is able to deliver 91% of the theoretical capacity of bismuth, up to 350 mAh/gBi or 3430 mAh/cm3Bi (the specific capacities are based Bi if not specifically stated; the density of Bi is 9.8 g/cm3). The high volumetric capacity is particularly attractive. The material also exhibits superior rate capabilities, maintaining a capacity as high as 216 mAh/g as the rate is increased to 5C (discharge/charge in 12 min), with excellent cycling stability and close to 100% Coulombic efficiency. More significantly, the Mg-ion chemistry is shown to be compatible with conventional electrolytes. Bi-NTs were synthesized by a hydrothermal reaction method to obtain materials with desirable properties (see the Supporting Information for synthesis details).52,53 The X-ray diffraction (XRD) pattern of the as-synthesized Bi-NT powder (Figure S1) reveals that the material is well-crystallized, and the diffraction patterns correspond well with the rhombohedral phase of bismuth metal (a = 4.535 Ǻ , b = 4.535 Ǻ , c = 5.000 Ǻ , α = 90.0, β = 90.0, γ = 120.0).52 Figure 1a shows a typical TEM

bismuth nanoparticles (Bi-NP, 30−50 nm) were also assembled and tested using identical protocols. For the half cell testing, we used an Mg electrolyte with highly efficient Mg plating/ stripping (with Mg(BH4)2, LiBH4, and diglyme) (the detailed study of this electrolyte has been reported in our previous paper,55 and Figure S2 shows 100% Coulombic efficiency of this electrolyte for Mg plating/stripping). Figure 2a compares cyclic voltammograms (CVs) of Mg insertion (0−0.25 V) and deinsertion (0.3−0.5 V) in Bi-NT and Bi-Micro. The Bi-NT electrode exhibited highly symmetric cathodic/anodic sharp peaks with narrower peak separation than Bi-Micro (0.264 V vs 0.313 V), indicating tthat he insertion/deinsertion of Mg2+ in Bi-NT is highly reversible and fast. Figure 2b shows the galvanostatic discharge/charge profile of an Mg−Bi half cell. The plateaus at ∼0.2 V are the Mg insertion process in Bi, corresponding to the 0−0.25 V peaks in CVs, and the plateaus at ∼0.35 V are the Mg deinsertion process in Bi, corresponding to the 0.3−0.5 V peaks in CVs. At the same discharge/charge rate (2C, discharge/charge in 30 min), the overvoltage is lower, and the specific capacity is higher for Bi-NT. The rate performance of electrodes made with Bi-NTs was further studied and compared with Bi-Micro. Their capacities at varied rates from 0.02C to 5C are summarized in Figure 2c. We found that the high rate performance of Bi-NTs was substantially better than Bi-Micro. The capacities of Bi-NTs and Bi-Micro were comparable at low rates (for example, both are ∼350 mAh/gBi at 0.05C, corresponding to 3430 mAh/ cm3Bi). As the charging/discharging rates were increased, the capacity of Bi-Micro anode decreased dramatically, retaining only 51 mAh/g at 5C (60% retention). Such observations clearly demonstrate the superior rate performance of Bi-NTs for Mg insertion/deinsertion. Cycling stability is another important factor for battery electrode materials. In this work the cycling stability of Bi-NTs was examined and compared with that of Bi-Micro in Figure 2d. The electrode with Bi-NTs exhibited excellent cycling stability, with only 7.7% of capacity fading after 200 cycles (maintaining 303 mAh/g at the 200th cycle). In contrast, the Bi-Micro electrodes experienced substantial capacity decay of 27.5% under the same testing conditions (only 188 mAh/g at the 200th cycle). The Coulombic efficiency of Bi-NTs is stable and close to 100% during the cycling; however, it is lower and keeps decreasing during cycling for Bi-Micro. Interestingly, we found that most of the capacity fading of both electrodes happened during the initial 20 cycles. Considering the fact that the insertion of Mg ions into Bi will lead to 100% volume expansion (Figure S5) and introduce substantial stress to the crystals, our results likely indicate that the Bi-Micros were unable to withstand the volume change, eventually leading to material pulverization and capacity fading. This agrees well with typical insertion type electrode materials with large volume expansion, such as silicon for lithium ion batteries.40 Compared with Bi-Micros, the substantially higher cycling stability and higher/stable Coulombic efficiency of Bi-NT indicates that the tubular structure is very effective in withstanding the volume expansion/shrinkage of Bi when hosting Mg2+.

Figure 1. (a) TEM image of Bi nanotubes showing a bundle of nanotubes; (b) high-resolution TEM image of Bi nanotubes showing the lattice fringes.

image of pristine Bi nanotubes with uniform diameters of ∼8 nm (wall thickness of about 2 nm). These nanotubes form bundle structure because of the van der Waals interaction between small tubes (like in the case of carbon nanotubes54). A high-resolution TEM image provided in Figure 1b shows the detailed structure of the nanotubes. The measured lattice spacing perpendicular to the tube direction is 3.2 Ǻ , corresponding to the (012) planes, consistent with the report from literature.52 On the basis of the Bi−Mg alloying reaction (2Bi + 3Mg2+ + 6e− → Mg3Bi2), bismuth has a theoretical capacity of 385 mAh/gBi. (For comparison, graphite, which enabled the practical application of and is still widely used in lithium ion battery, has a theoretical capacity of 372 mAh/g for lithium storage.) As we have discussed, due to the bivalent nature of magnesium, the solid-state diffusion of Mg is very slow.2 We believe that the nanostructured Bi could be beneficial for Mg diffusion because the diffusion length within the nanocrystal is significantly shorter than in bulk materials. We assembled the testing cells using Mg metal as the anode and a Bi-NT electrode as the cathode. For comparison, control electrodes using Bi microparticles (Bi-Micro, ∼100 μm) or 256

dx.doi.org/10.1021/nl403874y | Nano Lett. 2014, 14, 255−260

Nano Letters

Letter

Figure 2. (a) Cyclic voltammograms of Mg insertion/deinsertion in bismuth; (b) discharge/charge profile of an Mg−Bi cell; (c) rate performance of an Mg−Bi cell; (d) cycling stability and Coulombic efficiency (CE) of bismuth electrode for reversible Mg insertion/deinsertion. Cell configuration: Mg/0.1 M Mg(BH4)2−1.5 M LiBH4−diglyme/Bi.

Figure 3. Morphology and structure evolution of Bi-NTs during the discharge (Mg insertion) and charge (Mg deinsertion). (a) X-ray diffraction patterns of Bi-NTs electrode before and after discharge/charge; (b, c) TEM image (b = bright field, c = dark field) of Bi-NTs after discharge showing uniform nanoparticles; (d) TEM image (dark field) of Bi-NTs after discharge showing overall tube morphology; (e) HR-TEM image of Bi-NTs after discharge showing interconnected nanoparticles; (f) TEM image showing the region where Mg mapping (g) is acquired; (h) Mg K edge and Bi M edge as observed on Bi after Mg insertion; (i) TEM image showing the morphology of Bi after Mg deinsertion; (j) EELS spectra showing the Mg is removed from the sample after Mg deinsertion.

257

dx.doi.org/10.1021/nl403874y | Nano Lett. 2014, 14, 255−260

Nano Letters

Letter

Figure 4. Structural transformation of bismuth during the discharge/charge process. Bi-NTs may be fully magnesiated, and the structure evolves into porous interconnected nanoparticles which keeps electric contact; however, Bi large particles can only be partially magnesiated, and the structure pulverizes, losing electric contact. This leads to high performance of Bi-NTs and low performance of Bi-Micro in terms of rate capability and cycling stability.

To further demonstrate the unique advantages of Bi-NTs for hosting Mg2+, we also studied the behavior of Bi-NP for Mg insertion/deinsertion. The Bi-NP electrodes showed better performance than Bi-Micros in that their capacity fades 16.2% after 200 cycles. However, they are still inferior to Bi-NTs both in rate capability and cycling stability (Figure S6). These results indicate that the morphology and size of active Bi have a significant influence on its Mg2+ insertion/deinsertion behavior, and tubular Bi with a few nanometers performs best in our study. The Mg2+ insertion/deinsertion behavior in Bi was also studied with ex situ XRD and TEM. Figure 3a shows a set of XRD patterns of Bi-NT obtained before and after discharge/ charge. After discharge (Mg insertion), peaks corresponding to Mg3Bi2 were clearly observed, indicating formation of a Bi alloy with Mg. There are still small peaks of Bi metal, consistent with the discharge capacity which is 91.2% of the theoretical values at low rates. After charging (Mg deinsertion), Mg3Bi2 alloy peaks disappear, and Bi metal peaks corresponding to the original material reappear. This indicates high reversibility of Mg insertion/deinsertion in Bi-NT electrodes. To further understand the unique advantage of Bi-NTs, we used TEM to track its morphology and structure evolution upon the Mg insertion/deinsertion process. Figure 3b−j shows TEM images and electron energy-loss spectroscopy (EELS) spectra of Bi after Mg2+ insertion and deinsertion. After Mg insertion, the nanotubes were transformed into interconnected nanoparticles as shown in the bright and dark field images in Figure 3b and c. However, a careful examination suggested that some aggregated nanoparticles partially retained the general nanotube morphology (Figure 3d, dark field image). The highresolution TEM image in Figure 3f further reveals the interconnected nanoparticles. The EELS spectra and energy filtered image confirm the coexistence of Mg and Bi in the discharged sample (Figure 3f,g,h). After charging, the nanoparticle morphologies were retained (Figure 3i), but the Mg is almost completely removed (EELs spectrum in Figure 3j). Combining the XRD patterns, CV results, and the cycling data as discussed above, it is concluded that Bi-NTs synthesized in this work are able to store/release Mg ions reversibly without a noticeable nonreversible process. Similar phenomena have been observed in Si anode materials for Li-ion batteries. The use of nanotubes or porous Si significantly reduced the capacity fading

of Si anode due to volume expansion caused by Li insertion.57,58 The structural transformation and the charge/discharge process are schematically shown in Figure 4. The nanotube structures are first converted to nanoparticles during discharge. These nanoparticles are interconnected and roughly retain the overall morphologies of the nanotubes. During the subsequent charge and discharge processes, the Mg ions are inserted and deinserted without causing structural collapse. It is interesting to notice that the stability of aggregated nanoparticles directly prepared from similar solution synthesis is not as stable as the nanotube-based materials. The good cycling stability, high/ stable Coulombic efficiency, and high rate performance of the nanotube derived materials are likely due to the existence of the hollow space in the nanotubes which enabled the formation of highly interconnected Bi nanoparticles along the nanotubes; the conductivity between the particles is maintained through such connections. The in situ formed very small Bi nanoparticles (3−5 nm, Figure 3b and e) and the nanopores between them provide fast ion transport. However, for preformed Bi particles without hollow space inside, they just pulverized and lost the connection. Similar phenomena have been observed on Si for lithium insertion/deinsertion.40,57,59 It is important to show that the Bi anode material can be used with a conventional electrolyte. As discussed earlier, the formation of the nonconducting layer in conventional electrolytes that blocks Mg2+ is likely due to the reaction of freshly deposited Mg metal with the conventional electrolytes.16,60 As we can see from the CVs (Figure 2a), the Mg insertion/ deinsertion in Bi happens to be higher than 0 V (vs Mg0). This high potential suggests that Mg deposition on Bi in the form of Mg0 could be effectively avoided60 and the incompatibility issue with conventional electrolytes might not exist. To evaluate the performance of the proof-of-concept Mg ion battery with conventional electrolytes, an Mg intercalation material Mo6S8 was selected as the cathode,10 a premagnesiated Bi electrode was used as the anode, and a cathode-limited cell was assembled with a conventional electrolyte 0.4 M Mg(TFSI)2diglyme for test (Cell A). Mo6S8 is chosen as the cathode to demonstrate the compatibility of the anode material with the conventional electrolyte because it is currently most widely studied cathode material,10 even though the specific capacity is still quite limited now. For comparison, we also assemble the cell with the Mg electrolyte Mg(BH4)2−LiBH4−diglyme (Cell 258

dx.doi.org/10.1021/nl403874y | Nano Lett. 2014, 14, 255−260

Nano Letters

Letter

Figure 5. (a) Discharge/charge profile of an Mg3Bi2−Mo6S8 cell and (b) cycling stability of an Mg3Bi2−Mo6S8 cell. Cell configuration: (A) Mg3Bi2/ 0.4 M Mg(TFSI)2−diglyme/Mo6S8, (B) Mg3Bi2/0.1 M Mg(BH4)2−1.5 M LiBH4−diglyme/Mo6S8.

Sciences and Engineering, under Award KC020105FWP12152. G.L. and T.L. would like to acknowledge the support from Pacific Northwest National Laboratory (PNNL) Laboratory Directed Research and Development program for synthesizing the cathode and understanding the electrolyte. The TEM and XRD work were performed using EMSL, a national scientific user facility sponsored by the Department of Energy’s Office of Biological and Environmental Research and located at PNNL. PNNL is a multiprogram national laboratory operated for DOE by Battelle.

B). Figure 5 shows the discharge/charge profiles and cycling stability of these two cells (the capacity is based on the weight of cathode Mo6S8). The two cells show comparable specific capacity and good cycling stability. The discharge/charge plateaus are also similar, with the one with conventional electrolyte showing slightly large overvoltage. The XRD patterns of Bi electrode before and after discharge/charge also show the reversible Mg insertion/deinsertion (Figure S7). These indicate that the Bi anode is compatible with conventional electrolyte. We believe this opens up a new strategy for Mg battery development“Mg ion battery”. The properties of the full cell can be significantly improved if better cathode materials can be developed. In summary, we have developed a high-performance anode material based on Bi nanotubes for a rechargeable Mg battery. Bi nanotubes evolved into interconnected nanoporous Bi during Mg insertion/deinsertion. The interconnected nanoporous Bi anode delivers superior cycling stability and rate performance. XRD and TEM characterizations show the formation of Mg3Bi2 during discharge and reversible Mg insertion/deinsertion in Bi. The Mg cell by coupling premagnesiated a Bi (Mg3Bi2) anode with Mo6S8 cathode in a conventional electrolyte 0.4 M Mg(TFSI)2−diglyme shows the discharge/charge and stable cycling behavior similar to that in Mg electrolyte, indicating its compatibility with the conventional electrolyte. This work suggests that other high energy density alloy compounds may also be considered for Mg-ion chemistry, which, coupled with high-capacity highvoltage cathodes using conventional electrolytes, could lead to a high energy density Mg-ion battery. This opens up a new approach to develop Mg ion batteries.





ASSOCIATED CONTENT

S Supporting Information *

Additional information about experimental details and additional figures. This material is available free of charge via the Internet at http://pubs.acs.org.



REFERENCES

(1) Yoo, H. D.; Shterenberg, I.; Gofer, Y.; Gershinsky, G.; Pour, N.; Aurbach, D. Energy Environ. Sci. 2013, 6, 2265−2279. (2) Levi, E.; Gofer, Y.; Aurbach, D. Chem. Mater. 2010, 22, 860−868. (3) Aurbach, D.; Lu, Z.; Schechter, A.; Gofer, Y.; Gizbar, H.; Turgeman, R.; Cohen, Y.; Moshkovich, M.; Levi, E. Nature 2000, 407, 724−727. (4) Muldoon, J.; Bucur, C. B.; Oliver, A. G.; Sugimoto, T.; Matsui, M.; Kim, H. S.; Allred, G. D.; Zajicek, J.; Kotani, Y. Energy Environ. Sci. 2012, 5, 5941−5950. (5) Guo, Y. S.; Zhang, F.; Yang, J.; Wang, F. F.; NuLi, Y. N.; Hirano, S. I. Energy Environ. Sci. 2012, 5, 9100−9106. (6) Mizrahi, O.; Amir, N.; Pollak, E.; Chusid, O.; Marks, V.; Gottlieb, H.; Larush, L.; Zinigrad, E.; Aurbach, D. J. Electrochem. Soc. 2008, 155, A103−A109. (7) Kim, H. S.; Arthur, T. S.; Allred, G. D.; Zajicek, J.; Newman, J. G.; Rodnyansky, A. E.; Oliver, A. G.; Boggess, W. C.; Muldoon, J. Nat. Commun. 2011, 2, 427. (8) Mohtadi, R.; Matsui, M.; Arthur, T. S.; Hwang, S. J. Angew. Chem., Int. Ed. 2012, 51, 9780−9783. (9) Pour, N.; Gofer, Y.; Major, D. T.; Aurbach, D. J. Am. Chem. Soc. 2011, 133, 6270−6278. (10) Levi, E.; Gershinsky, G.; Aurbach, D.; Isnard, O.; Ceder, G. Chem. Mater. 2009, 21, 1390−1399. (11) Rasul, S.; Suzuki, S.; Yamaguchi, S.; Miyayama, M. Electrochim. Acta 2012, 82, 243−249. (12) Gershinsky, G.; Yoo, H. D.; Gofer, Y.; Aurbach, D. Langmuir 2013, 29, 10964−10972. (13) Imamura, D.; Miyayama, M.; Hibino, M.; Kudo, T. J. Electrochem. Soc. 2003, 150, A753−A758. (14) Zhang, R. G.; Yu, X. Q.; Nam, K. W.; Ling, C.; Arthur, T. S.; Song, W.; Knapp, A. M.; Ehrlich, S. N.; Yang, X. Q.; Matsui, M. Electrochem. Commun. 2012, 23, 110−113. (15) Xu, K. Chem. Rev. 2004, 104 (10), 4303−4417. (16) Lu, Z.; Schechter, A.; Moshkovich, M.; Aurbach, D. J. Electroanal. Chem. 1999, 466, 203−217. (17) Gregory, T. D.; Hoffman, R. J.; Winterton, R. C. J. Electrochem. Soc. 1990, 137, 775−780. (18) Aurbach, D.; Gofer, Y.; Schechter, A.; Chusid, O.; Gizbar, H.; Cohen, Y.; Moshkovich, M.; Turgeman, R. J. Power Sources 2001, 97− 8, 269−273.

AUTHOR INFORMATION

Corresponding Authors

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

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work is supported by the U.S. Department of Energy (DOE), Office of Basic Energy Sciences, Division of Materials 259

dx.doi.org/10.1021/nl403874y | Nano Lett. 2014, 14, 255−260

Nano Letters

Letter

(19) Aurbach, D.; Schechter, A.; Moshkovich, M.; Cohen, Y. J. Electrochem. Soc. 2001, 148, A1004−A1014. (20) Aurbach, D.; Gizbar, H.; Schechter, A.; Chusid, O.; Gottlieb, H. E.; Gofer, Y.; Goldberg, I. J. Electrochem. Soc. 2002, 149, A115−A121. (21) Muldoon, J.; Bucur, C. B.; Oliver, A. G.; Zajicek, J.; Allred, G. D.; Boggess, W. C. Energy Environ. Sci. 2013, 6, 482−487. (22) Yagi, S.; Tanaka, A.; Ichikawa, Y.; Ichitsubo, T.; Matsubara, E. J. Electrochem. Soc. 2013, 160, C83−C88. (23) Lv, D. P.; Xu, T.; Saha, P.; Datta, M. K.; Gordin, M. L.; Manivannan, A.; Kumta, P. N.; Wang, D. H. J. Electrochem. Soc. 2013, 160, A351−A355. (24) Cheng, G.; Xu, Q.; Ding, F.; Sang, L.; Liu, X. J.; Cao, D. X. Electrochim. Acta 2013, 88, 790−797. (25) Ellis, B. L.; Lee, K. T.; Nazar, L. F. Chem. Mater. 2010, 22, 691− 714. (26) Winter, M.; Besenhard, J. O.; Spahr, M. E.; Novak, P. Adv. Mater. 1998, 10, 725−763. (27) Whittingham, M. S. Chem. Rev. 2004, 104, 4271−4301. (28) Li, H.; Huang, X. J.; Chen, L. Q.; Wu, Z. G.; Liang, Y. Electrochem. Solid-State Lett. 1999, 2, 547−549. (29) Winter, M.; Besenhard, J. O. Electrochim. Acta 1999, 45, 31−50. (30) Scrosati, B.; Garche, J. J. Power Sources 2010, 195, 2419−2430. (31) Dahn, J. R.; Zheng, T.; Liu, Y. H.; Xue, J. S. Science 1995, 270, 590−593. (32) Goodenough, J. B.; Kim, Y. Chem. Mater. 2010, 22, 587−603. (33) Wang, Y.; Cao, G. Z. Adv. Mater. 2008, 20, 2251−2269. (34) Hu, Y. S.; Demir-Cakan, R.; Titirici, M. M.; Muller, J. O.; Schlogl, R.; Antonietti, M.; Maier, J. Angew. Chem., Int. Ed. 2008, 47, 1645−1649. (35) Jiao, F.; Bao, J. L.; Hill, A. H.; Bruce, P. G. Angew. Chem., Int. Ed. 2008, 47, 9711−9716. (36) Dudley, J. T.; Wilkinson, D. P.; Thomas, G.; Levae, R.; Woo, S.; Blom, H.; Horvath, C.; Juzkow, M. W.; Denis, B.; Juric, P.; Aghakian, P.; Dahn, J. R. J. Power Sources 1991, 35, 59−82. (37) Arthur, T. S.; Singh, N.; Matsui, M. Electrochem. Commun. 2012, 16, 103−106. (38) Singh, N.; Arthur, T. S.; Ling, C.; Matsui, M.; Mizuno, F. Chem. Commun. 2013, 49, 149−151. (39) Malyi, O. I.; Tan, T. L.; Manzhos, S. J. Power Sources 2013, 233, 341−345. (40) Chan, C. K.; Peng, H. L.; Liu, G.; McIlwrath, K.; Zhang, X. F.; Huggins, R. A.; Cui, Y. Nat. Nanotechnol. 2008, 3, 31−35. (41) Bruce, P. G.; Scrosati, B.; Tarascon, J. M. Angew. Chem., Int. Ed. 2008, 47, 2930−2946. (42) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. M. Nature 2000, 407, 496−499. (43) Kim, M. G.; Cho, J. Adv. Funct. Mater. 2009, 19, 1497−1514. (44) Sun, Y. K.; Chen, Z. H.; Noh, H. J.; Lee, D. J.; Jung, H. G.; Ren, Y.; Wang, S.; Yoon, C. S.; Myung, S. T.; Amine, K. Nat. Mater. 2012, 11, 942−947. (45) Guo, Y. G.; Hu, Y. S.; Sigle, W.; Maier, J. Adv. Mater. 2007, 19, 2087−2091. (46) Malik, R.; Burch, D.; Bazant, M.; Ceder, G. Nano Lett. 2010, 10, 4123−4127. (47) Meethong, N.; Huang, H. Y. S.; Carter, W. C.; Chiang, Y. M. Electrochem. Solid-State Lett. 2007, 10, A134−A138. (48) Lou, X. W.; Wang, Y.; Yuan, C. L.; Lee, J. Y.; Archer, L. A. Adv. Mater. 2006, 18, 2325−2329. (49) Lee, S. W.; Yabuuchi, N.; Gallant, B. M.; Chen, S.; Kim, B. S.; Hammond, P. T.; Shao-Horn, Y. Nat. Nanotechnol. 2010, 5, 531−537. (50) Amine, K.; Belharouak, I.; Chen, Z. H.; Tran, T.; Yumoto, H.; Ota, N.; Myung, S. T.; Sun, Y. K. Adv. Mater. 2010, 22, 3052−3057. (51) Cao, Y. L.; Xiao, L. F.; Wang, W.; Choi, D. W.; Nie, Z. M.; Yu, J. G.; Saraf, L. V.; Yang, Z. G.; Liu, J. Adv. Mater. 2011, 23, 3155−3160. (52) Yang, B. J.; Li, C.; Hu, H. M.; Yang, X. G.; Li, Q. W.; Qian, Y. T. Eur. J. Inorg. Chem. 2003, 20, 3699−3702. (53) Li, Y. D.; Wang, J. W.; Deng, Z. X.; Wu, Y. Y.; Sun, X. M.; Yu, D. P.; Yang, P. D. J. Am. Chem. Soc. 2001, 123, 9904−9905.

(54) O’Connell, M. J.; Bachilo, S. M.; Huffman, C. B.; Moore, V. C.; Strano, M. S.; Haroz, E. H.; Rialon, K. L.; Boul, P. J.; Noon, W. H.; Kittrell, C.; Ma, J. P.; Hauge, R. H.; Weisman, R. B.; Smalley, R. E. Science 2002, 297, 593−596. (55) Shao, Y. Y.; Liu, T. B.; Li, G. S.; Gu, M.; Nie, Z. M.; Engelhard, M. H.; Xiao, J.; Lv, D. P.; Wang, C. M.; Zhang, J. G.; Liu, J. Sci. Rep. 2013, 3, 3130. (56) McDowell, M. T.; Ryu, I.; Lee, S. W.; Wang, C. M.; Nix, W. D.; Cui, Y. Adv. Mater. 2012, 24, 6034−6041. (57) Park, M. H.; Kim, M. G.; Joo, J.; Kim, K.; Kim, J.; Ahn, S.; Cui, Y.; Cho, J. Nano Lett. 2009, 9, 3844−3847. (58) Magasinski, A.; Dixon, P.; Hertzberg, B.; Kvit, A.; Ayala, J.; Yushin, G. Nat. Mater. 2010, 9, 353−358. (59) Kasavajjula, U.; Wang, C. S.; Appleby, A. J. J. Power Sources 2007, 163, 1003−1039. (60) Lossius, L. F.; Emmenegger, F. Electrochim. Acta 1996, 41, 445− 447.

260

dx.doi.org/10.1021/nl403874y | Nano Lett. 2014, 14, 255−260