Co-doped NiO nanoflake arrays toward superior anode materials for lithium ion batteries
Y.J. Mai,J.P. Tu,X.H. Xia,C.D. Gu,xiao ping wang
DOI: https://doi.org/10.1016/j.jpowsour.2011.03.089
IF: 9.2
2011-01-01
Journal of Power Sources
Abstract:Co-doped NiO nanoflake arrays with a cellular-like morphology are fabricated by low temperature chemical bath deposition. As anode material for lithium ion batteries (LIBs), the array film shows a capacity of 600 mAh g −1 after 50 discharge/charge cycles at low current density of 100 mA g −1 , and it retains 471 mAh g −1 when the current density is increased to 2 A g −1 . Appropriate electrode configuration possesses some unique features, including high electrode–electrolyte contact area, direct contact between each naonflake and current collector, fast Li + diffusion. The Co 2+ partially substitutes Ni 3+ , resulting in an increase of holes concentration, and therefore improved p-type conductivity, which is useful to reduce charge transfer resistance during the charge/discharge process. The synergetic effect of these two parts can account for the improved electrochemical performance. Keywords Nickel oxide Array film Metal-doping Electronic conductivity Lithium ion battery 1 Introduction Lithium ion batteries (LIBs) now dominate the portable electronic market, but they fall short of satisfying needs for high power and/or capacities for applications such as power tools, electronic vehicles or efficient use of renewable energies [1] . Conversion reactions of interstitial-free 3d metal oxides have been shown to be highly reversible, providing outstanding capacities to store lithium [2–12] . However, a major drawback of the conversion reactions is their poor kinetics related either to intrinsic poor electronic/ionic conductivity of the electrode materials or to inadequate means of configuring the electrodes. Recently, self-supported nano-material arrays directly growing on current collector represent the feasibility of improving the kinetics. These electrodes show several unique advantages, such as high accessible surface area, easy accessibility, good cohesion to current collector, and features on the nanometer scale. Li et al. demonstrated that the mesoporous Co 3 O 4 nanowire arrays on Ti foil as anodes for LIBs maintained a capacity of 700 mAh g −1 after 20 cycles at a current of 1 C, and retained 50% of the capacity when the current was increased to 50 C [13] . The self-supported Ni 3 S 2 /Ni nanoarchitectured electrode exhibited a sustained reversible capacity at 2 C with a loss of only 18% after 20 cycles [14] . Kim and co-workers reported that the self-supported SnS nanosheets possessed good cycling performance and superior rate capability of 380 mAh g −1 at 20 C [15] . Chemical bath deposition (CBD) is a prevalent low-temperature aqueous technique for directly depositing large-area thin film with self-supporting architecture of semiconductors [16–18] . In addition, it is theorized that if cation-deficient solid solutions could be retained, good p-type conductivity should result, which is well demonstrated by transition metal oxides [19–22] . In the special case of NiO, Co-doped NiO hexagonal nanoplatelets in the form of powder materials with improved electrochemical performance in the application of supercapacitor have been reported [23] . Moreover, the direct growth of self-support active materials on current collector is an important issue for their potential application in the next-generation LIBs. Here, we report a one-step method for producing Co-doped NiO thin film by CBD with an appropriate adding amount of CoSO 4 . The resulting electrode not only shows self-support nanoflake arrays but also possesses better electronic/ionic conductivities due to the doping of cobalt and nanoflake arrays. Using such electrodes, they show high capacity, good cycling performance and high rate capability. To the best of our knowledge, it is the first time to report Co-doped NiO self-support arrays fabricated by CBD toward electrode materials for LIBs. 2 Experimental Co-doped NiO nanoflake arrays were synthesized by chemical bath deposition method modified from our previous reports [24,25] . Specifically, 38.8 mmol NiSO 4 , 1.2 mmol CoSO 4 and 7.4 mmol K 2 S 2 O 8 were dissolved in 10 ml of concentrated ammonia and 90 ml of deionized water. For comparison, pure NiO thin film was also prepared with the same process without addition of CoSO 4 . After deposition on the pre-treated commercial Ti foil, the films were calcined in a tube furnace at 350 °C for 2 h in flowing argon. If the deposition time is 2 h, the thickness is about 980 nm for pure NiO film and about 360 nm for Co-doped NiO film, respectively. For comparison on their electrochemical performance, the deposition time is cut to 20 min for pure NiO film, resulting in nearly the same thickness of both films. Some samples were also deposited on silicon wafers and slides to facilitate characterization of XRD and measurement of Hall effect. CR-2025-type coin cells were assembled in a glove box under argon atmosphere, where the as-prepared NiO electrodes were used as working electrodes and lithium foil as counter electrode and reference electrode. The electrolyte was 1 M LiPF 6 in ethylene carbonate (EC)–dimethyl carbonate (DME) (1:1 in volume). The galvanostatic charge–discharge tests were conducted on LAND battery program-control test system. Cyclic voltammetric (CV) measurements of the electrodes were performed on a CHI660C electrochemical workstation with a scan rate of 0.1 mV s −1 between 0 and 3 V (vs. Li/Li + ). The electrochemical impedance spectroscopy (EIS) of the electrodes was also carried out on CHI660C with a frequency range of 0.01 Hz to 100 kHz. Before the measurement, the electrodes were cycled for eight cycles, then discharged to 2.7 V and kept until the open-circuit voltage stabilized. The microstructure, surface morphology and composition of the as-prepared NiO films were characterized by using X-ray diffraction (D/max 2550-PC), in which a glancing-angle mode with the incident beam at an angle of 4° was employed, X-ray photoelectron spectroscopy (XPS, AXIS UTLTRADLD equipped with a dual Mg Kα–Al Kα anode for photoexcitation) and field emission scanning electron microscope (FESEM, S-4800 coupled with EDX). Hall effect measurements were carried out on HL5500 Hall system. 3 Results and discussion Fig. 1 a shows the XRD patterns of NiO and Co-doped NiO films. The diffraction peaks of both films can be well assigned to face-centered-cubic (FCC) NiO (JCPDS 78-0643), and no obvious peaks for cobalt oxides or other impurities are detected, implying that the cobalt doping does not change the original NiO structure. Broader peaks in XRD pattern make it difficult to observe the obvious shift after the doping Co into NiO thin film. The corresponding XPS provides further structural information for the obtained films. Fig. 1 b shows the XPS wide spectrum of Co-doped NiO film. Characteristic peaks in the inset of Fig. 1 b represent Co 2p of CoO [26] . Ni 2p 3/2 peak includes two components, one at 854.4 eV due to Ni–O bonds and the other one at 855.6 eV due to Ni–OH bands, respectively. The Ni–OH bands mainly come from nickel hydroxide α-Ni(OH) 2 , β-Ni(OH) 2 and higher valence nickel oxides such as Ni 2 O 3 ·H 2 O, β-NiO(OH), or 4Ni(OH) 2 ·NiOOH· x H 2 O [16] . In O 1s spectrums, the peak at 531.3 eV for pure NiO as well as Co-doped NiO indicates the presence of Ni-OH bands, which are consistent with Ni 2p 3/2 spectrums. It is likely that the as-prepared films contain higher valence oxides because of the presence of persulfate. Comparing O 1s spectrum of pure NiO and Co-doped NiO film shown in Fig. 1 c and 1d, respectively, we found, first, that the Co-doped NiO gives rise to one more O 1s peak at 529.6 eV, corresponding to CoO. Second, the intensity ratio of Ni–OH bands/NiO (NiO and CoO) bands obviously decreases after doping of cobalt. The Hall effect measurements suggested the obtained NiO film without cobalt doping was n-type which could be mainly attributed to the present of Ni 3+ [21] , although NiO was not n-dopable because of the spontaneous formation of electron killer [27] . The carrier concentration of pure NiO film was 1.93 × 10 12 cm −3 with a mobility of 40.5 cm 2 V −1 s −1 as well as a resistivity of 8 × 10 4 Ω cm. After doping with Co, the as-prepared film transferred to p-typed due to Co 2+ partially substitute Ni 3+ . The concentration of majority carrier was 6.95 × 10 15 cm −3 , coupled with a mobility of 17.2 cm 2 V −1 s −1 as well as a resistivity of 52.28 Ω cm. Based on the above results, it is reasonable to deduce that the as-prepared NiO films contain higher valence oxides consisting of Ni 2 O 3 ·H 2 O, β-NiO(OH), or 4Ni(OH) 2 ·NiOOH· x H 2 O, and their amount is so small that beyond the detection limitation of XRD but XPS. After doping with Co, the Co 2+ partially substituted Ni 3+ , resulting in an increase of hole concentration, and therefore an enhanced p-type conductivity, which is useful to reduce the charge transfer resistance during the charge/discharge process. It is expected that the electrode materials with an enhanced intrinsic electronic conductivity should result in good electrochemical performance when they are applied to lithium ion batteries. Fig. 2 shows the SEM image of Co-doped NiO nanoflake arrays film. The NiO film is uniform in appearance, exhibiting nanoflake arrays and cellular like morphology, which is the result of a competition between heterogeneous film growth on the substrate and homogeneous growth and aggregation leading to particle formation [16] . However, it is observed that no film appears under the same preparation condition of pure NiO thin film if NiSO 4 is replaced by CoSO 4 . Neither raising temperature nor increasing the amount of oxidizer (K 2 S 2 O 4 ) can obtain film. Unless CoSO 4 decreases to a certain amount, the Co-doped NiO thin film will not be obtained, but particles. Second, the adding of CoSO 4 significantly affects the growth of NiO film, which can be confirmed by the film thickness obtained from the SEM images of cross section of both samples. This can be attributed that cobalt (II) hydroxide additive inhibits the transition from β-Ni(OH) 2 to γ-NiOOH [28] , while ions with higher valence states are more prone to undergo hydrolysis and precipitate as hydroxides or oxides than those with lower valence states [29] . In addition, EDX analysis indicated the atomic ratio of Ni and Co is 9:1. It is noted that cobalt doping and post heat treatment do not change the morphology of the prepared films. For comparison on the electrochemical performance, especially rate capability, we deliberately decreased the deposition time of pure NiO to 20 min to obtain an approximate thickness of both films (about 360 nm). The reaction mechanism of NiO by Li was well demonstrated in previous reports [2,30–33] . It first involved the formation of highly reactive nickel nanograins, which was the key to enable the formation-decomposition of Li 2 O upon subsequent cycles, dispersed into Li 2 O matrix, followed by the growth of an organic coating that partially dissolved upon the subsequent charge while Ni converted back to NiO nanograins. It can be described by the electrochemical conversion reactions [2] : NiO + 2 L i + + 2 e ⇄ charge discharge L i 2 O + Ni . The CV curves of pure NiO and Co-doped NiO electrodes, as shown in Fig. 3 , reflect the above reaction mechanism. For the Co-doped NiO electrode in the first cycle ( Fig. 3 a), a main cathodic peak is observed at about 0.45 V and a low-intensity peak near 1.3 V correspond the reduction of NiO to metallic Ni nanoparticles and the formation of a partially reversible SEI layer whose composition includes Li 2 CO 3 , ethylene-oxide-based oligomers, LiF, and lithium alkyl carbonate [31,34] . Anodic peaks at around 0.5 V and 1.41 V correspond to the partial decomposition of the polymeric coating on the NiO surface, while the decomposition of Li 2 O and nickel converted back to NiO nanograins for another anodic peak at 2.32 V. These results are consistent with previous reports [25,35] . For the pure NiO electrode, the cathodic peak shifts to 0.39 V, and these anodic peaks shift to 1.56 and 2.33 V, respectively. It is found that in the first cycle, the intensity of cathodic peaks for Co-doped NiO electrodes at near 1.3 V is much stronger than that for pure NiO, which are mainly attributed to the improved electronic conductivity of Co-doped NiO electrode, facilitating the charge transfer needed for redox processes of solvent molecules in electrolyte relating to the formation of SEI layer. The integral area of anodic peak locating at near 1.41 V is larger than that for pure NiO electrode, indicating more dissolution of organic coating upon the subsequent charge for Co-doped NiO electrode. The separation between the reduction and oxidation peaks of the Co-doped NiO electrode decreases as compared to that of pure NiO electrode, indicating weaker polarization and better reversibility. This separation is also observed in the following cycles, as shown in Fig. 3 b. In addition, the peak intensity and integral areas of the Co-doped NiO electrode are larger than those of pure NiO, indicating that more reactions toward Li + occur in the Co-doped NiO electrode. The incorporation of cobalt affects both the ionic and the electronic conductivity of the active material, result in a better electrode utilization and improvement of reaction kinetics [36] , which is favorable in the application of supercapacitors [36,37] . The increased electronic conductivity after cobalt doping is further demonstrated by EIS. The Nyquist plots are shown in Fig. 4 , whose data are analyzed by fitting to an equivalent electrical circuit (inset of Fig. 4 ). R e , R sf and R ct represent the electrolyte resistance, surface film resistance and Li-ion charge transfer resistance at the interface between electrolyte and electrode, respectively. Given to the porosity of the measured materials, two constant phase element, CPE sf and CPE ct , were employed, instead of pure capacitances [38,39] . W s and C int correspond to diffusion components like Warburg impedance and the intercalation capacitance, respectively [40] . In Fig. 4 , the symbols are the experimental data whereas the continuous lines represent the fitted spectra. For the pure NiO electrode, two semicircles are observed, including the depressed one located at near 9670 Hz, corresponding mainly to surface film resistance, and the other at about 300 Hz, corresponding mainly to Li-ion charge transfer resistance. The impedance spectra for Co-doped NiO electrode also show two depressed semicircle. However, the diameter of the second semicircle, reflecting the resistant of R ct , reduces drastically, indicating that the latter possesses lower charge transfer resistance than pure NiO electrode. The curve fitting further confirms this assumption. It reveals the impedance of R ct is 28.5 ± 4.8 Ω for pure NiO electrode, whereas it reduces drastically to 19.6 ± 0.3 Ω for Co-doped NiO electrode. R sf is 38.3 ± 3.4 Ω for pure NiO and 32.9 ± 0.9 Ω for Co-doped NiO electrode. The discharge/charge curves of pure NiO and Co-doped NiO electrode at low current density of 100 mA g −1 are shown in Fig. 5 a and b , respectively. For the Co-doped NiO electrode, as shown in Fig. 5 b, two plateaus and a sloping potential range during the first discharge/charge are well resolved. A narrow plateau at 1.5–0.9 V and a sloping part at 0.6–0.01 V during discharge correspond to the plateau at 1.0–1.5 V and the sloping part at 0.01–1.0 V during charge, indicating the formation and partially dissolution of a SEI layer, respectively; while a long discharge plateau at 0.6 V and a charge plateau at 2.0–2.5 V correspond to the reduction of NiO to metallic nickel nanoparticles dispersed into Li 2 O matrix and its oxidation to NiO nanograins and decomposition of Li 2 O. These results are agreed with the above CV results. Wang and co-workers reported that SEI layer had already started above 2.0 V during the first cathodic scan [41] , and Tarascon and co-workers also observed the formation of SEI layer by ex situ TEM at the sloping potential range of 1 to 0.02 V [31] . What's more, it is well known that Ni and Co nanograins can act as the catalyst toward the formation/dissolution of the organic coating. So, it is reasonable to deduce that for the present electrode materials, the SEI layer starts at relative high potential of 1.3 V and further grows at lower potential (less than 0.5 V). And with the help of highly catalyst reactive of Co and Ni nanograins, the formed SEI layer partially dissolves upon the subsequent charge. The pure NiO electrode delivers 832.8 mAh g −1 , 637.5 mAh g −1 and 463.3 mAh g −1 for the 1st, 2nd and 50th cycle, respectively. The initial discharge specific capacity is 1201 mAh g −1 for the Co-doped NiO electrode. The discharge specific capacity decreases to 882.6 mAh g −1 in the 2nd cycle and remains at 589.5 mAh g −1 after 50 cycles. As shown in Fig. 6 , the Co-doped NiO electrode has high specific capacities in spite of at a low current density (100 mA g −1 ) or at a high current density (2 A g −1 ) compared to the pure NiO electrodes. What's more, the Co-doped NiO electrode possesses excellent rate capability and shows good capacity retention even at higher current rates. Reversible and steady capacity of 471 mAh g −1 was achieved at a high current density of 2 A g −1 , whereas 174 mAh g −1 for the pure NiO electrode. To improve electrochemical performance of NiO in the application of LIBs, many attempts were operated, including ZnO–NiO–C film (230 mAh g −1 ) [42] , three-dimensional porous NiO (450 mAh g −1 ) [41] , core-shell structured Ni/NiO (about 200 mAh g −1 ) [43] , hierarchically ordered porous NiO (518 mAh g −1 ) [44] . Compared with these previous attempts, the present method synchronously considers optimization design of electrode configuration and improvement of the lattice electronic conductivity of active materials and the Co-doped NiO electrodes show comparable capacities at low current density (100 mA g −1 ), but better capacities and cycling performance at high current density (2 A g −1 ). It is suggested that the high capacity and rate capability of the Co-doped NiO electrode come from the unique hierarchical architecture combined with improved electronic conductivity after doping of Co 2+ . The arrays configuration can facilitate the nanoflakes to participate in the electrochemical reaction because each nanoflake is directly in electric contact with the current collector. The open space between neighboring nanoflake allows for easy contact and diffusion of the electrolyte into the inner region of the electrode, and the thin thickness of flakes enables the high rate of lithium insertion/removal. In addition, the cobalt doping enhances p-type electronic conductivity of NiO, resulting in improved kinetics. 4 Conclusions Co-doped NiO nanoflake arrays with a cellular-like morphology grown on Ti foil have been synthesized by CBD method. The addition of CoSO 4 significantly affects the growth of NiO film and an advisable adding amount of CoSO 4 is important to obtain the Co-doped NiO film. The Co-doped NiO electrode showed excellent rate capability and good capacity retention even at a high current rate of 2 A g −1 , with about 1.7 times improvement in the reversible specific discharge capacities compared to pure NiO. These improved electrochemical performances can be attributed to the synergetic effect coming from nanoflake array configuration, the open space and enhanced electronic conductivity after doping of cobalt. It also highlight our method is simple to synchronously obtain adequate electrode configuration and enhancement in the lattice electronic conductivity of NiO. References [1] P.L. Taberna S. Mitra P. Poizot P. Simon J.M. Tarascon Nat. Mater. 5 2006 567 573 [2] P. Poizot S. Laruelle S. Grugeon L. Dupont J.M. Tarascon Nature 407 2000 496 499 [3] M.K. Gulbinska S.L. Suib J. Power Sources 196 2011 2149 2154 [4] P. Tartaj J.M. Amarilla J. Power Sources 196 2011 2164 2170 [5] Y. Sun X.-Y. Feng C.-H. Chen J. Power Sources 196 2011 784 787 [6] P. Zhang, Z.P. Guo, Y.D. Huang, D.Z. Jia, H.K. Liu, J. Power Sources, doi:10.1016/j.jpowsour.2010.10.090 . [7] K.F. Zhong, B. Zhang, S.H. Luo, W. Wen, H. Li, X.J. Huang, L.Q. Chen, J. Power Sources, doi:10.1016/j.jpowsour.2010.10.031 . [8] X.H. Huang J.P. Tu B. Zhang C.Q. Zhang Y. Li Y.F. Yuan H.M. Wu J. Power Sources 161 2006 541 544 [9] M.-Y. Cheng B.-J. Hwang J. Power Sources 195 2010 4977 4983 [10] J.Y. Xiang J.P. Tu L. Zhang Y. Zhou X.L. Wang S.J. Shi J. Power Sources 195 2010 313 319 [11] X.H. Huang J.P. Tu C.Q. Zhang X.T. Chen Y.F. Yuan H.M. Wu Electrochim. Acta 52 2007 4177 4181 [12] K.F. Zhong X. Xia B. Zhang H. Li Z.X. Wang L.Q. Chen J. Power Sources 195 2010 3300 3308 [13] Y.G. Li B. Tan Y.Y. Wu Nano Lett. 8 2008 265 270 [14] Q. Wang R. Gao J.H. Li Appl. Phys. Lett. 90 2007 143107 [15] J.-G. Kang J.-G. Park D.-W. Kim Electrochem. Commun. 12 2010 307 310 [16] S.Y. Han D.H. Lee Y.J. Chang S.O. Ryu T.J. Lee C.H. Chang J. Electrochem. Soc. 153 2006 C382 C386 [17] J.C. Park J. Kim H. Kwon H. Song Adv. Mater. 21 2009 803 807 [18] H.Y. Xu S.L. Xu H. Wang H. Yan J. Electrochem. Soc. 152 2005 C803 C807 [19] H.L. Pan B. Yao T. Yang Y. Xu B.Y. Zhang W.W. Liu D.Z. Shen Appl. Phys. Lett. 97 2010 142101 142103 [20] Y. Tsur I. Riess Ionics 1 1995 488 490 [21] T. Dutta P. Gupta A. Gupta J. Narayan J. Appl. Phys. 108 2010 [22] D. Pasero N. Reeves A.R. West J. Power Sources 141 2005 156 158 [23] Z. Zheng L. Huang Y. Zhou X.M. Hu X.M. Ni Solid State Sci. 11 2009 1439 1443 [24] X.H. Xia J.P. Tu J. Zhang X.L. Wang W.K. Zhang H. Huang Electrochim. Acta 53 2008 5721 5724 [25] X.H. Huang J.P. Tu X.H. Xia X.L. Wang J.Y. Xiang L. Zhang J. Power Sources 195 2010 1207 1210 [26] S.C. Petitto E.M. Marsh G.A. Carson M.A. Langell J. Mol. Catal. A-Chem. 281 2008 49 58 [27] S. Lany J. Osorio-Guillen A. Zunger Phys. Rev. B 75 2007 241203 [28] B.B. Ezhov O.G. Malandin J. Electrochem. Soc. 138 1991 885 889 [29] H. Unuma Y. Saito K. Watanabe M. Sugawara Thin Solid Films 468 2004 4 7 [30] L. Yuan Z.P. Guo K. Konstantinov P. Munroe H.K. Liu Electrochem. Solid State Lett. 9 2006 A524 A528 [31] S. Grugeon S. Laruelle R. Herrera-Urbina L. Dupont P. Poizot J.M. Tarascon J. Electrochem. Soc. 148 2001 A285 A292 [32] P. Poizot S. Laruelle S. Grugeon J.M. Tarascon J. Electrochem. Soc. 149 2002 A1212 A1217 [33] B. Varghese M.V. Reddy Z. Yanwu C.S. Lit T.C. Hoong G.V. Subba Rao B.V.R. Chowdari A.T.S. Wee C.T. Lim C.-H. Sow Chem. Mater. 20 2008 3360 3367 [34] G. Gachot S. Grugeon M. Armand S. Pilard P. Guenot J.-M. Tarascon S. Laruelle J. Power Sources 178 2008 409 421 [35] X.H. Huang J.P. Tu Z.Y. Zeng J.Y. Xiang X.B. Zhao J. Electrochem. Soc. 155 2008 A438 A441 [36] A.B. Yuan S.A. Cheng J.Q. Zhang C.N. Cao J. Power Sources 77 1999 178 182 [37] J.H. Park S. Kim O.O. Park J.M. Ko Appl. Phys. A-Mater. 82 2006 593 597 [38] M.V. Reddy S. Madhavi G.V. Subba Rao B.V.R. Chowdari J. Power Sources 162 2006 1312 1321 [39] M.D. Levi D. Aurbach J. Phys. Chem. B 108 2004 11693 11703 [40] M.V. Reddy T. Yu C.H. Sow Z.X. Shen C.T. Lim G.V. Subba Rao B.V.R. Chowdari Adv. Funct. Mater. 17 2007 2792 2799 [41] C. Wang D.L. Wang Q.M. Wang H.J. Chen J. Power Sources 195 2010 7432 7437 [42] Q.M. Pan L.M. Qin J. Liu H.B. Wang Electrochim. Acta 55 2010 5780 5785 [43] X. Li A. Dhanabalan K. Bechtold C. Wang Electrochem. Commun. 12 2010 1222 1225 [44] Y.F. Yuan X.H. Xia J.B. Wu J.L. Yang Y.B. Chen S.Y. Guo Electrochem. Commun. 12 2010 890 893