In situ synthesis of amorphous titanium dioxide supported RuO2 as a carbon-free cathode for non-aqueous Li–O2 batteries

Jisu Kimab, Yeon Kimac, Mihye Wu*a, Dae-Ho Yoonb, Yongku Kangad and Ha-Kyun Jung*ad
aAdvanced Materials Division, Korea Research Institute of Chemical Technology, 141 Gajeongro, Yuseong, Daejeon 34114, Korea. E-mail: wumihye@krict.re.kr; hakyun@krict.re.kr
bSchool of Advanced Materials Science and Engineering, Sungkyunkwan University, Suwon 16419, Korea
cDepartment of Material Science and Engineering, Hanyang University, Seoul 04763, Korea
dDepartment of Chemical Convergence Materials, Korea University of Science and Technology (UST), 217 Gajeongro, Yuseong, Daejeon 34113, Korea

Received 28th June 2016 , Accepted 20th September 2016

First published on 21st September 2016


Abstract

Amorphous TiO2 supported crystalline RuO2 (a-TiO2/c-RuO2 hybrid) was prepared as a carbon-free cathode by in situ sol–gel synthesis, and its electrochemical performance in non-aqueous Li–O2 batteries was investigated for the first time. The a-TiO2/c-RuO2 hybrid enhanced battery performance, and this enhancement was attributed to the crystallinity of the TiO2. It was found that amorphous TiO2 is more electrochemically active toward the discharge/charge processes than crystalline TiO2. The a-TiO2/c-RuO2 hybrid exhibited good reversibility as well as cyclic stability, making it a promising carbon-free cathode material for non-aqueous Li–O2 batteries.


To meet the ever-growing demand for high energy density rechargeable batteries in various applications, Li–O2 batteries have attracted much attention and have been intensively studied as an effective energy storage system, because they possess an extremely high energy density.1 However, in practical use Li–O2 batteries also have limitations, such as poor cycle life, low round-trip efficiency, electrolyte degradation and decomposition of the carbon cathode.2 Among these shortcomings, the decomposition of the carbon cathode is responsible for poor battery performance because it results in the production of lithium carbonates (Li2CO3). Since the fundamental energy storage mechanisms in Li–O2 batteries are the formation and decomposition of lithium peroxide (Li2O2)3 upon discharge and charging, respectively, the formation of Li2CO3 is undesirable.4 Moreover, deposited Li2CO3 covers and blocks the available cathode surface and does not decompose, which further deteriorates battery performance, including the development of large overpotential.5 As a result, it has become necessary to develop alternative carbon-free cathode materials for Li–O2 batteries.

Titanium dioxide (TiO2) nanoparticles have been intensively investigated in many industrial and scientific areas due to their high chemical and physical stability, low-cost, non-toxicity and a straightforward synthesis route.6 In particular, TiO2 nanoparticles with diverse polymorphs, e.g., amorphous, anatase, brookite and rutile forms, have emerged in applications for energy storage devices, exhibiting excellent electrochemical activities.7

In this study, amorphous TiO2 was chosen as a catalyst support and was combined with ruthenium oxide (RuO2) as a catalyst in recognition of its great catalytic activity toward both discharge and charge.8 Although TiO2 has been widely applied in many research fields including energy storage systems, employing TiO2 as a catalyst support in Li–O2 batteries has been rarely reported since the first report from Zhao et al. that Pt modified anatase TiO2 nanotube arrays.9 Only few reports have been published on TiO2 as a carbon-free cathode including a TiO2(B) nanofiber@porous RuO2 composite from Guo et al.10 Different from previous reports on TiO2, which are crystalline in phase, the prepared TiO2 is amorphous in phase. To the best of our knowledge, amorphous TiO2 has not yet been reported as a catalyst support for Li–O2 batteries.

Herein, we report a simple in situ sol–gel synthesis of amorphous TiO2 supported crystalline RuO2 as a carbon-free cathode material, and its electrochemical performance in Li–O2 batteries is evaluated for the first time. The in situ synthesis is a convenient method for preparing TiO2/RuO2 hybrids owing to the simple and facile synthesis route. Moreover, the sol–gel process, which is one of the most commonly used methods for the preparation of metal oxide nanoparticles, provides chemical homogeneity at relatively low temperature.11 The experimental details are shown in the ESI.

The TiO2/RuO2 hybrids were synthesized by an in situ sol–gel method followed by annealing at relatively low temperature. Although it is not easy to demonstrate the exact formation mechanisms of TiO2/RuO2 hybrids, hydrolysis and polycondensation are assumed to be the main steps of the reactions.12 Based on the hydrolysis of titanium oxysulfate and ruthenium chloride in aqueous solution, the resultant precursors are regarded as a composite of titanium hydroxide and ruthenium hydroxide. This in situ preparation may create close interfacial contact between the titanium and ruthenium, which facilitates the uniform distribution of the TiO2 and RuO2.13 The TiO2/RuO2 precursors were annealed at 300 and 600 °C for crystallization and phase transformation from hydroxide to oxide, and the resulting crystal structures of the TiO2/RuO2 hybrids are shown in Fig. 1a. When annealed at a given temperature both samples exhibited X-ray diffraction (XRD) patterns corresponding to the RuO2 (JCPDS#01-071-4825). On the other hand, there was a difference in the diffraction patterns associated with TiO2: the sample annealed at 600 °C exhibited a mixture of anatase (JCPDS#01-075-2546) and rutile (JCPDS#99-000-3236) phases of TiO2, whereas the sample annealed at 300 °C exhibited the amorphous phase. Since XRD is not able to reveal the presence of amorphous material, it is essential to confirm the presence of TiO2. In this regard, Raman spectroscopy was conducted on the a-TiO2/c-RuO2 hybrid, and the resultant spectra in Fig. 1b demonstrated the formation of TiO2.14 Therefore, it can be concluded that TiO2/RuO2 hybrids with different crystalline structure were achieved by changing the annealing temperature. A composite of amorphous TiO2 with crystalline RuO2 (a-TiO2/c-RuO2) was obtained with the 300 °C-annealing, while crystalline TiO2 with crystalline RuO2 (c-TiO2/c-RuO2) was obtained with the 600 °C-annealing.


image file: c6ra16645h-f1.tif
Fig. 1 (a) XRD patterns of TiO2/RuO2 hybrids based on the annealing temperature (b) Raman spectroscopy of a-TiO2/c-RuO2 hybrid (T: TiO2, R: RuO2).

The microstructural characteristics of a-TiO2/c-RuO2 and c-TiO2/c-RuO2 hybrids were studied using transmission electron microscopy (TEM). In the sol–gel approach, the relative reaction rates of hydrolysis and polycondensation greatly affect the size and morphology of TiO2/RuO2.15 As shown in Fig. 2, both a-TiO2/c-RuO2 and c-TiO2/c-RuO2 hybrids exhibit semi-spherical particles with diameters under 15 nm, indicating the formation of nanostructured materials, and this signifies that there is no significant morphological deviation or particle growth between the a-TiO2/c-RuO2 and c-TiO2/c-RuO2 hybrids. Therefore, the effect of annealing temperature on the particle size and morphology are negligible in the TiO2/RuO2 hybrid. For further investigation, TEM images were analyzed using energy dispersive X-ray spectroscopy (EDS) mapping, shown in Fig. 2c, to examine the elemental distribution of particles. From the results, it was observed that the elements including Ti, Ru and O, which originated from the TiO2 and RuO2, were uniformly distributed over the microstructure.


image file: c6ra16645h-f2.tif
Fig. 2 TEM images of (a) a-TiO2/c-RuO2 (b) c-TiO2/c-RuO2 hybrids (c) TEM-EDS mapping images of a-TiO2/c-RuO2 hybrid.

The electrochemical performances of a-TiO2/c-RuO2 and c-TiO2/c-RuO2 hybrids as air electrodes in Li–O2 batteries were investigated, and the results are shown in Fig. 3. The profiles of the galvanostatic discharge/charge were obtained based on the mass of TiO2 and were compared at a current density of 100 mA g−1. The a-TiO2/c-RuO2 hybrid showed high electrochemical activity toward ORR/OER in the given condition, whereas that of c-TiO2/c-RuO2 hybrid showed poor electrochemical activity. The capacity of the a-TiO2/c-RuO2 hybrid was 1200 mA h g−1, while the c-TiO2/c-RuO2 hybrid was not even effectively discharged in the given range, showing a capacity of only 71 mA h g−1. Since the important difference between a-TiO2/c-RuO2 and c-TiO2/c-RuO2 hybrids was their respective crystallinity, it is appropriate to assume that crystallinity is the key factor affecting their electrochemical properties in Li–O2 batteries.


image file: c6ra16645h-f3.tif
Fig. 3 Voltages profiles of TiO2/RuO2 hybrids.

However, it was not clear whether the amorphous TiO2 provides a higher electrochemical activity than that of crystalline TiO2, because it was reported in our previous study that RuO2 crystallinity can also influence battery performance, based on its different catalytic activity.16 To determine the effect of TiO2 crystallinity on battery performance, it is necessary to minimize the effect of the crystallinity of the RuO2 catalyst in the TiO2/RuO2 hybrid. For that purpose, we employed the ex situ method, which allowed the amorphous TiO2 and crystalline RuO2 to be synthesized separately, and then mixed together in an agate mortar. The a-TiO2/c-RuO2 hybrid compound prepared by the ex situ method was fabricated into cathode material for Li–O2 batteries, and the results were compared. As can be seen, the a-TiO2/c-RuO2 hybrid prepared by the ex situ method showed a capacity of 470 mA h g−1 with an overpotential of 1.4 V, which are much better values than those of the c-TiO2/c-RuO2 hybrid. Therefore, it was demonstrated that amorphous TiO2 is more electrochemically active as a catalyst support than crystalline TiO2.

In order to provide insights into the reversibility and stability of the a-TiO2/c-RuO2 hybrid for Li–O2 battery cathodes, the surface morphology at discharge/charge states and cycle life were examined. The FE-SEM image of a-TiO2/c-RuO2 hybrid taken after discharge demonstrates that the discharge products are deposited on the surface of the cathode, which clogs the oxygen penetration through the cathode (Fig. 4b). This blocked surface is recovered after subsequent charging, as presented in Fig. 4c, which shows that the deposited discharge products have completely disappeared, and the morphology of the surface is identical to pristine a-TiO2/c-RuO2 hybrid. To capture the details of discharge products, high-resolution X-ray diffraction (HR-XRD) and X-ray photoelectron spectroscopy (XPS) were applied and the results are shown in Fig. 4d and S1. The HR-XRD result exhibited the peak corresponding to Li2O2 appeared after the first discharge. In Fig. S1, the Li 1s spectrum obtained after discharge exhibited the Li2O2 relevant binding energy peak, which utterly disappeared after recharge.17 This strongly supports the result from FE-SEM that Li2O2 was produced as a main discharge product and was decomposed after recharge. Therefore, the reversibility of charge/discharge reaction was verified.


image file: c6ra16645h-f4.tif
Fig. 4 FE-SEM images of a-TiO2/c-RuO2 hybrid: (a) pristine (b) after first discharge (c) after recharge, (d) HR-XRD pattern after first discharge (e) terminal voltages and energy efficiency as a function of cycle numbers.

For further investigation of reversibility and stability, the cycle life of a-TiO2/c-RuO2 hybrid at limited capacity was measured, and the results are provided in Fig. 4e. The cell was cycled at a DOD of 300 mA h g−1 at a current density of 100 mA g−1. The cell was observed to maintain capacity without fading up to 132 cycles with no significant decrease in discharge terminal voltage, or increase in charge terminal voltage. The energy efficiency of the a-TiO2/c-RuO2 hybrid in accordance with cycle life indicated that energy efficiency was more than 82% after 132 cycles. Based on this account, it is clear that as a cathode material the a-TiO2/c-RuO2 hybrid facilitates reversible and rechargeable Li–O2 battery performance.

In summary, a a-TiO2/c-RuO2 hybrid was prepared by the in situ sol–gel route and its electrochemical performance as a new carbon-free cathode material for non-aqueous Li–O2 batteries was evaluated for the first time. The a-TiO2/c-RuO2 hybrid achieved an overpotential of 1.0 V at a specific capacity of 1200 mA h g−1, and these values were much better performances than those of the c-TiO2/c-RuO2 hybrid. The results strongly suggest that the electrochemical activity of TiO2 as a catalyst support differs based on whether it is crystalline, and amorphous TiO2 is more electrochemically active toward discharge/charge processes than crystalline TiO2. The reversibility and stability of the a-TiO2/c-RuO2 hybrid was demonstrated, confirming its applicability as an alternative to carbon material for non-aqueous Li–O2 batteries.

Acknowledgements

This research was supported by the Government-Funded General Research & Development Program by the Ministry of Trade, Industry and Energy, Republic of Korea.

References

  1. K. R. Yoon, G. Y. Lee, J.-W. Jung, N.-H. Kim, S. O. Kim and I.-D. Kim, Nano Lett., 2016, 16, 2076 CrossRef CAS PubMed; F. Li, T. Zhang and H. Zhou, Energy Environ. Sci., 2013, 6, 1125 Search PubMed; T. Ogasawara, A. Débart, M. Holzapfel, P. Novák and P. G. Bruce, J. Am. Chem. Soc., 2006, 128, 1390 CrossRef PubMed; P. G. Bruce, S. A. Freunberger, L. J. Hardwick and J.-M. Tarascon, Nat. Mater., 2012, 11, 19 CrossRef PubMed.
  2. G. Girishkumar, B. McCloskey, A. C. Luntz, S. Swanson and W. Wilcke, J. Phys. Chem. Lett., 2010, 1, 2193 CrossRef CAS; R. R. Mitchell, B. M. Gallant, C. V. Thompson and Y. Shao-Horn, Energy Environ. Sci., 2011, 4, 2952 Search PubMed; M. Piana, J. Wandt, S. Meini, I. Buchberger, N. Tsiouvaras and H. A. Gasteiger, J. Electrochem. Soc., 2014, 161, A1992 CrossRef.
  3. A. Débart, J. Bao, G. Armstrong and P. G. Bruce, J. Power Sources, 2007, 174, 1177 CrossRef; A. Débart, A. J. Paterson, J. Bao and P. G. Bruce, Angew. Chem., 2008, 120, 4597 CrossRef.
  4. B. D. McCloskey, A. Speidel, R. Scheffler, D. C. Miller, V. Viswanathan, J. S. Hummelshøj, J. K. Nørskov and A. C. Luntz, J. Phys. Chem. Lett., 2012, 3, 997 CrossRef CAS PubMed.
  5. F. Li, H. Kitaura and H. Zhou, Energy Environ. Sci., 2013, 6, 2302 Search PubMed; M. M. Ottakam Thotiyl, S. A. Freunberger, Z. Peng and P. G. Bruce, J. Am. Chem. Soc., 2013, 135, 494 CrossRef CAS PubMed.
  6. J.-B. Park, I. Belharouak, Y. J. Lee and Y.-K. Sun, J. Power Sources, 2015, 295, 299 CrossRef CAS; D. Chu, Y. Hou, J. He, M. Xu, Y. Wang, S. Wang, J. Wang and L. Zha, J. Nanopart. Res., 2009, 11, 1805 CrossRef.
  7. A. G. Dylla, G. Henkelman and K. J. Stevenson, Acc. Chem. Res., 2013, 46, 1104 CrossRef CAS PubMed; S. Yoon, E. S. Lee and A. Manthiram, Inorg. Chem., 2012, 51, 3505 CrossRef PubMed.
  8. L. Y. Chen, X. W. Guo, J. H. Han, P. Liu, X. D. Xu, A. Hirata and M. W. Chen, J. Mater. Chem. A, 2015, 3, 3620 Search PubMed; H.-G. Jung, Y. Jeong, J.-B. Park, Y.-K. Sun, B. Scrosati and Y. J. Lee, ACS Nano, 2013, 7, 3532 CrossRef CAS PubMed; G. Zhao, Y. Niu, L. Zhang and K. Sun, J. Power Sources, 2014, 270, 368 Search PubMed; C. S. Park, J. H. Kim and Y. J. Park, J. Electroceram., 2013, 31, 224 CrossRef.
  9. G. Zhao, R. Mo, B. Wang, L. Zhang and K. Sun, Chem. Mater., 2014, 26, 2551 CrossRef CAS.
  10. Z. Guo, C. Li, J. Liu, X. Su, Y. Wang and Y. Xia, J. Mater. Chem. A, 2015, 3, 21123 CAS.
  11. D. P. Macwan and P. N. Dave, J. Mater. Sci., 2011, 46, 3669 CrossRef CAS.
  12. V. Houšková, V. Štengl, S. Bakardjieva, N. Murafa and V. Tyrpekl, Appl. Catal., B, 2009, 89, 613 CrossRef.
  13. S. Liu, H. Sun, S. Liu and S. Wang, Chem. Eng. J., 2013, 214, 298 CrossRef CAS.
  14. W. Zhu, Y. Xu, H. Li, B. Dai, H. Xu, C. Wang, Y. Chao and H. Liu, Korean J. Chem. Eng., 2014, 31, 211 CrossRef CAS; H. Shin, H. S. Jung, K. S. Hong and J.-K. Lee, J. Solid State Chem., 2005, 178, 15 CrossRef; X. Cheng, L. Zhang, L. Chen, Y. Zhang and Q. Fan, ECS Trans., 2009, 19, 1 Search PubMed.
  15. B. E. Yoldas, J. Mater. Sci., 1986, 21, 1087 CrossRef CAS.
  16. M. Wu, J. Y. Jo, S. Choi, Y. Kang and H.-K. Jung, RSC Adv., 2015, 5, 24175 RSC.
  17. Z. Zhang, J. Lu, R. S. Assary, P. Du, H.-H. Wang, Y.-K. Sun, Y. Qin, K. C. Lau, J. Greeley, P. C. Redfern, H. Iddir, L. A. Curtiss and K. Amine, J. Phys. Chem. C, 2011, 115, 25535 Search PubMed; Z. H. Cui, W. G. Fan and X. X. Guo, J. Power Sources, 2013, 235, 251 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Details of the synthesis. See DOI: 10.1039/c6ra16645h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.