Uniform FexNiy Nanospheres: Cost-Effective ... - ACS Publications

Aug 7, 2017 - The energy density is 5−10 times higher than that of the intercalation cathode in the conventional lithium-ion batteries, and it is al...
0 downloads 10 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Uniform FexNiy Nanospheres: Cost-Effective Electrocatalysts for Nonaqueous Rechargeable Li−O2 Batteries Mengwei Yuan,† Caiyun Nan,† Yan Yang,† Genban Sun,*,†,‡ Huifeng Li,*,† and Shulan Ma*,† †

Beijing Key Laboratory of Energy Conversion and Storage Materials and College of Chemistry, Beijing Normal University, Beijing 100875, China ‡ Department of Materials Physics and Chemistry, University of Science and Technology Beijing, Beijing 100083, China S Supporting Information *

ABSTRACT: Uniform FexNiy nanospheres were synthesized via a simple solvothermal method and used as electrocatalysts for Li−O2 batteries. Fe7Ni3 nanospheres exhibited relatively high catalytic activities in the electrochemical tests. They delivered a reversible capacity of more than 7000 mAh/gKB and gave a discharge−charge voltage gap reduction of 250 mV compared with Ketjen Black.

1. INTRODUCTION To meet the requirements of energy storage and conversion, research efforts have been directed toward rechargeable nonaqueous Li−O2 batteries because of their extremely high theoretical specific energy of 11 400 Wh kg−1. The energy density is 5−10 times higher than that of the intercalation cathode in the conventional lithium-ion batteries, and it is also competitive to that of gasoline or diesel oil.1,2 However, lots of challenges still need to be overcome before Li−O2 batteries can find practical applications in electric vehicles, hybrid vehicles, and stationary energy storage. Among various issues, the sluggish kinetics of the oxygen reaction is an essential problem resulting in high overpotentials and affecting the performance of Li−O2 batteries, especially for the oxygen evolution reaction (OER).3,4 The high overpotential leads to poor round-trip efficiency and cyclability. One promising way of reducing the overpotentials is to add catalysts for the oxygen reduction reaction (ORR) and OER.5−7 Attention has been paid to explore effective catalysts to improve the energy efficiency of Li−O2 batteries. Cathode materials have been explored as effective catalysts in Li−O2 batteries, including porous carbon materials,8−10 noble metals,11−16 noble metal oxides,17,18 transitional metal oxides,19−24 and perovskite-based25,26 and pyrochlore-based27 compounds. Among those mentioned above, carbon materials, such as graphene,28,29 carbon nanotubes,30,31 and mesoporous carbon,9 show both high specific capacity and high ORR activity due to their large specific area and superior electronic conductivity. However, the OER activity is insufficient. The noble metal-based materials have been attracting much attention in this field because of their high activity for both ORR and OER, but they are too expensive for general use. Thus, the transitional metal oxides have been developed into promising electrocatalysts for ORR and OER process because of their low cost and high activity. However, there is still a huge © 2017 American Chemical Society

gap in comparison to noble metals, so a lower overpotential needs to be achieved by further study. Therefore, exploring new catalysts to reduce the overpotential and increase the round-trip efficiency is still an important challenge for Li−O2 batteries. The transitional metals and their alloys in nanoscale have good electronic conductivity, which is beneficial for the catalyst synergistic effect to possess high electrocatalytic activity, and various active sites. Previous studies have shown that Co32 and Ni33 nanoparticles have a higher electrocatalytic activity for OER or ORR. It has been proved that the transition metals could accelerate the reaction process. The current research also demonstrates that FeCo/CNT displays excellent electrocatalytic activity during the OER and ORR processes, which is more superior than that of the CNTs.34 Similarly, in comparison with the single metal, FeNi3@C presents superior electrocatalytic activity in the OER process with water system due to the optimized surface electronic structure as a result of the introduction of Fe, which also contributed to the synergistic effect of the alloy.35 All of these above reports suggest that the transition metal alloys in nanoscale are promising electrocatalysts in the Li−O2 batteries, as well as other electrocatalytic field. Moreover, the earth-enriched element of Fe and Ni were studied as transition metal alloys to prepare cost-effective electrocatalysts. Therefore, we have selected two stable structures Fe7Ni3 and FeNi3, which have a higher content of Fe atoms and synergistic effect. In this work, uniform FexNiy nanospheres, including the body-centered cubic Fe7Ni3 and the face-centered cubic FeNi3, were synthesized via a simple solvothermal method and employed in the Li−O2 battery for Received: April 22, 2017 Accepted: July 21, 2017 Published: August 7, 2017 4269

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

Figure 1. Schematic illustration of the formation of FexNiy nanospheres and their application in rechargeable Li−O2 batteries.

Figure 2. XRD patterns of (a) Fe7Ni3 nanospheres and (b) FeNi3 nanospheres.

the first time. The FexNiy nanospheres enhanced the performance of the Li−O2 batteries efficiently. In particular, the Fe7Ni3 nanosphere presents a superior performance with a higher initial discharge capacity of 7266 mAh/gKB and a lower overpotential during the charge−discharge process, wherein the voltage gap is reduced by 250 mV compared with that of the Ketjen Blank (KB) cathode.

with the body-centered cubic Fe7Ni3 phase (JCPDS 65-7251) and the face-centered cubic FeNi3 phase (JCPDS 38-0419). Through inductively coupled plasma-atomic emission spectrum (ICP-AES) measurement, the atomic ratios of the two samples were determined to be 7:3 and 1:3 (Table 1), corresponding to the compositions of Fe7Ni3 and FeNi3, consistent with the XRD results.

2. RESULTS AND DISCUSSION The FexNiy nanospheres were prepared via a facile solvothermal method. During the slow processes of nucleation and growth, some critical factors needed to be controlled. First, before the reaction, the mixture must be continuously stirred to be homogeneous. Simultaneously, the oxygen of the system was removed by a swift Ar flow for about 5 min. In addition, the octadecylamine served as a surfactant and was significant for the reaction. It tended to form a polymer−metal complex prescursor in the presence of the long-chain organic compounds and restricted the nucleation and the growth of the crystal nucleus.36−38 The hydrazine hydrate acted as a reductant to reduce Fe3+ and Ni2+ and then formed an alloy in the alkaline system. The FexNiy nanospheres served as electrocatalysts for Li−O2 batteries, and the Li−O2 batteries performance were investigated, as shown in Figure 1. The Li− O2 batteries were assembled with a Li foil serving as an anode and FexNiy nanospheres mixed with KB was employed as the air cathode; the two ends were separated by a glass fiber filter separator. The X-ray diffraction (XRD) patterns of the asprepared samples are presented in Figure 2. As shown in Figure 2, the XRD results show the diffraction peaks that matched well

Table 1. Fe/Ni Atom Rate Investigated by ICP-AES sample

Fe (atom %)

Ni (atom %)

Fe7Ni3 FeNi3

0.706 0.252

0.294 0.748

Figure 3 presents the morphology and structure of Fe7Ni3 and FeNi3 nanospheres analyzed by scanning electron microscopy (SEM) and transmission electron microscopy (TEM) measurements. A large number of Fe7Ni3 spherelike secondary particles with a size of ca. 150 nm are observed in Figure 3a. The TEM and HRTEM images of Fe 7 Ni 3 nanospheres shown in Figures 3b and S1a give further evidence that the as-prepared secondary particles are composed of Fe7Ni3 primary nanoparticles with the size of less than 10 nm. The composition of FeNi3 was similar to that of Fe7Ni3, but the range of particle size was different. The HRTEM image for Fe7Ni3 nanospheres shown in Figure 3c confirms the unidirectional fringe pattern and the measured interlayer spacing of 0.203 nm, corresponding to the (110) plane. The corresponding selected area electron diffraction (SAED) pattern in this region is shown in the inset of Figure 3c, which shows that the clear diffraction ring can be indexed to the 4270

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

Figure 3. a) SEM image, (b) TEM image, (c) HRTEM image (the inset is the corresponding SAED pattern), and (d) elemental distribution mappings for Fe and Ni of Fe7Ni3 nanospheres. (e) SEM image, (f) TEM image, (g) HRTEM image (the inset is the corresponding SAED pattern), and (h) elemental distribution mappings for Fe and Ni of FeNi3 nanospheres.

Figure 4. XPS spectrum of the as-prepared FexNiy nanospheres: (a, d) survey spectrum of Fe7Ni3 and FeNi3; (b, e) Fe (2p) binding energy spectrum of Fe7Ni3 and FeNi3; (c, f) Ni (2p) binding energy spectrum of Fe7Ni3 and FeNi3.

cubic Fe7Ni3(110) plane. Moreover, element mapping (Figure 3d) showed that Fe and Ni elements are distributed uniformly, which indicated the formation of an alloy.13,39 Similarly, FeNi3 spherelike secondary particles with a size of ca. 200 nm consisted of primary particles (Figure 3e,f) with the size of less than 10 nm (Figure S1b). The interlayer spacings were 0.208 and 0.180 nm, corresponding to (111) and (200) planes, respectively, corresponding to the two clear diffraction rings shown in the inset of Figure 3g. The elemental mapping images shown in Figure 3h offer the evidence for the existence of FeNi3 alloy.39

X-ray photoelectron spectroscopy (XPS) was employed to further distinguish the formation of Fe7Ni3 and FeNi3 alloys, whose spectra are illustrated in Figure 4. The peaks imply the presence of iron (Fe 2p), nickel (Ni 2p), and oxygen (O 1s) coming from Fe7Ni3, FeNi3, and some surface oxide of the alloys. The peaks shown in the spectra of Fe at 706.7 and 706.5 eV are attributed to the Fe 2p of metallic Fe, which indicates the presence of a zero-valent Fe.39−41 Peaks at 710.2 and 710.7 eV are attributed to Fe3+.39 Peaks at 852.0 and 852.6 eV for Fe7Ni3 and FeNi3 indicate the existence of a zero-valent Ni.42,43 Peaks at 855.1 and 855.4 in Figure 4c,f, respectively, are assigned to Ni2+.39 These results suggest a large amount of Fe 4271

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

the plateau of Fe7Ni3 electrode was about 200 mV lower than that of pure KB electrode. In addition, Fe7Ni3 electrode showed about 180 mV lower charge plateau than FeNi3 electrode. Fe7Ni3 electrode exhibited a high round-trip efficiency of 63.8%, which was higher than that of FeNi3 electrode (60.0%) and KB electrode (59.8%), indicating a higher energy density than what could be reached by Fe7Ni3 electrode.45 It suggested that Fe7Ni3 had more efficient active sites than FeNi3 for the OER process. As seen from Figure 5a, the initial specific discharge capacity of Fe7Ni3, FeNi3, and KB electrode is 7608, 5507, and 5487 mAh/gKB, respectively. The corresponding coulombic efficiency of Fe7Ni3, FeNi3, and KB electrodes was 95.5, 80.6, and 44.6%, which indicated a high reversibility of Fe7Ni3 electrode.38 All of these results demonstrated that Fe7Ni3 nanospheres can reduce the discharge−charge overpotential and increase the specific capacity of the Li−O2 batteries to a great extent. To avoid serious polarization of the electrode during the discharge process, the cycle stability of the FexNiy electrodes was tested under a capacity-limited strategy of 1000 mAh/gKB and a current density of 150 mA/ gKB. As shown in Figure 5b, Fe7Ni3 electrode can remain discharged with no decay after 15 cycles, whereas FeNi3 electrode only operated for 10 cycles during the capacitylimited cycle test. This result proved that Fe7Ni3 electrode exhibited better performance with more number of cycles. The details of the discharge−charge curves for the Fe x Ni y nanospheres are illustrated in Figure 6. The correlation between the discharge-medium voltage and the cycle numbers is presented in Figure 5b. The discharge-medium voltage was reduced with increasing cycle numbers, indicating the overpotential was generally enhanced, which could be caused by the increased impedance of the Li−O2 batteries after cycling Figure 7.45 In rechargeable Li−O2 batteries, the formation and decomposition of Li2O2 were the dominant reactions during discharge and charge processes. Simultaneously, some byproducts were formed due to the reaction of electrolyte or carbon materials. Undesirable byproducts, such as Li2CO3, cannot be decomposed completely during the charge process and were accumulated on the air electrode after cycling and finally resulted in an increase in the overpotential.45,46 In addition, the electrocatalytic activity of OER was the significant factor to affect the overpotential and the charge capacity. The linear sweep voltammetry (LSV) of OER and chronoamperometry profiles were measured after the discharge capacity depth of 1000 mAh g−1 (Figure 8) and in the Li−O2 battery with the Li2O2 cathode (Figure S2), respectively. As shown in Figure 8a, it is obvious that the current density of Fe7Ni3 is higher than that of FeNi3 electrode, which corresponded with Figure S2a (a lower potential of oxidation and a higher current density), indicating a higher electrocatalytic acitivty of Fe7Ni3 electrode.38 Figure 8b presents the chronoamperometry profiles of Fe7Ni3 and FeNi3, which was used to investigate the OER kinetics. The current density of FeNi3 electrode was lower than that of Fe7Ni3 electrode at the potential of 4.0 V, corresponding with the result of Figure S2b (a lower current density of 0.05 mA cm−2), implying a fast kinetics process for OER on Fe7Ni3 electrode.47 On the basis of these results, the possibility of the capacity decade was related with the OER activity, which corresponded with the discharge−charge profiles (Figure 5a). As we all know, the generally accepted mechanism of the Li− O2 battery during the discharge process is that O2 undergoes a one-electron reduction to form LiO2 as an intermediate, which is adsorbed on the electrocatalyst. Due to the different

and Ni on the surface of the as-prepared FexNiy nanospheres. The oxides were formed due to exposure of the active metal nanoparticles to air, which caused the surface oxidation. Combining the XPS and XRD results, we can prove that FexNiy alloys were formed rather than the individual metals of Fe and Ni.44 The FexNiy nanospoheres generally reveal a higher electrical conductivity than metal oxides, which could decrease the resistance and facilitate the transfer of electrons in the oxygen cathodes. Besides, the size of the as-prepared samples is tiny which is endowed with large surface area, resulting in abundant active sites in catalytic reactions. Therefore, these FexNiy nanospheres are promising electrocatalysts for Li−O2 batteries, and their performance is being further studied. To prove the electrocatalytic activity of the FexNiy nanospheres, they were assembled as the oxygen electrode in Li−O2 batteries. For a comparison, the pure Ketjen Blank (KB) electrode was used in the same conditions. Figure 5a displays the discharge−charge

Figure 5. a) Discharge and charge curves obtained with the electrodes of KB, FeNi3, and Fe7Ni3 at a current density of 150 mA/gKB, (b) discharge capacity and discharge-medium voltage of Fe7Ni3 and FeNi3 electrodes vs cycle numbers.

curves of the electrodes of KB, Fe7 Ni3 (KB + Fe7 Ni 3 nanospheres), and FeNi3 (KB + FeNi3 nanospheres) at the current density of 150 mA/gKB. Fe7Ni3 nanospheres showed the best electrochemical property. The discharge plateau of Fe7Ni3 electrode was 2.72 V, 50 mV than that of KB (2.67 V) and 50 mV more than that of FeNi3 (2.67 V) electrodes, which indicated that Fe7Ni3 nanospheres can efficiently facilitate the ORR activity in the Li−O2 batteries. As for the charge process, 4272

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

Figure 6. Discharge−charge curves of (a) Fe7Ni3 and (b) FeNi3 electrodes with a capacity limitation of 1000 mAh/gKB at the current density of 150 mA/gKB.

Figure 7. (a) LSV and (b) chronoamperometry profiles for Fe7Ni3 and FeNi3 electrodes. The LSV profiles were tested at a sweep rate of 2 mV s−1. The chronoamperometry profiles were tested at the potential of 4.0 V.

Figure 8. Raman spectra of (a−c) Fe7Ni3 and (d−f) FeNi3 cathode at different states.

this work, the battery operated through a different process in which LiO2 was the main product of the discharge. To investigate the mechanism of the reaction process, the Raman spectra were employed to detect the intermediate product. Figure 7 illustrates the Raman spectra of Fe7Ni3 and FeNi3

adsorption energy, the LiO2 displays evident selective adsorption on special crystal planes of electrocatalysts. LiO2 with an oxygen-rich surface is beneficial to the formation of Li2O2. Simultaneously, after continuously charging to 4.5 V, the discharge product is decomposed completely.48,49 However, in 4273

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

Figure 9. SEM images of the fresh (a), first discharged (b), and first recharged (c) Fe7Ni3 electrodes and the fresh (d), first discharged (e), and first recharged (f) FeNi3 electrodes. XRD patterns of Fe7Ni3 electrodes (g) and FeNi3 electrodes (h) at different states, and the inset shows the partial magnified patterns of the discharged electrodes.

decomposition of Li2O2 and LiO2. The similar results were observed for FeNi3 cathode in the discharge−charge process. Thus, the FexNiy nanospheres tended to catalyze the formation of LiO2, which could increase the battery life. The morphology of the FexNiy electrodes at different states was characterized by ex situ SEM. As shown in Figure 7b, the first discharged Fe7Ni3 electrode seemed like a ball-like composite with a diameter of approximately 1 μm packed on the electrode. The product disappeared and the electrode surface recovered the same as the fresh one (Figure 7a) after the battery was recharged (Figure 7c), indicating Fe7Ni3 electrode has good reversibility over cycling. The discharged product was further investigated by ex situ XRD. The weak diffraction peaks at 32.9, 34.9, and 58.7° (magnified in the inset in Figure 9g) corresponded to Li2O2 (JCPDS 74-0115) and little Li2CO3 that appeared after the first discharge process. In addition, no obvious diffraction peaks of LiO2 were detected by XRD measurement due to the amorphous nature of the discharge products.51 As for FeNi3 electrode, there was no

cathodes at different states, including discharge to 2.6 V, discharge to 2.0 V, and charge to 4.5 V. Figure 7a−e shows the Raman spectra for Fe7Ni3. Figure 7a,b demonstrates the presence of peaks at 1123 and 1505 cm−1, which is associated with the formation of LiO2. The peak at 1123 cm−1 was a signature of the LiO2-like component. The peak at 1505 cm−1 was attributed to the distortion of the graphite ring stretching model, which is a consequence of the strong interaction between LiO2 and graphite carbon surface.15,50−52 There was a weak peak at about 820 cm−1 in Figure 7b, which could be ascribed to Li2O2. The typical peak of the commercial Li2O2 was appropriately 790 cm−1, which was lower than that given in this work. We and other authors suspected that the electrochemically formed Li2O2 differed from the commercial Li2O2, resulting from the interactions with the FexNiy and graphite carbon surface.49,53,54 The relatively intensity of LiO2 was much higher than that of Li2O2, which indicated that LiO2 was the dominant product. After continuously charging at 4.5 V, the peaks of Li2O2 and LiO2 disappeared, indicating the 4274

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

spectrometer (Thermo Fisher) with Al Kα radiation as the Xray source for excitation. 4.3. Electrochemical Measurements. The air electrode was assembled in the composition of FexNiy nanospheres 45 wt %, Ketjen Black (KB) 45 wt %, and poly(vinylidene fluoride) (PVDF) binder 10 wt %. The pure KB electrode consisted of KB 90 wt % and PVDF 10 wt %. N-Methyl-pyrrolidone was added to the powder mixture to make a uniform slurry. Then, the slurry was pasted on the carbon papers and dried at 80 °C overnight in a vacuum oven to form air electrodes. Each of the composite electrode paste was ca. 0.8 mg. In comparison, the KB electrode was ca. 0.4 mg. Electrochemical measurements were carried out in CR2032 coin cells with holes to allow O2 diffusion. The coin cells consisted of a Li foil anode, a glass fiber filter separator, an electrolyte (1 M LiTFSI-TEGDME) and air electrode, were assembled in an Argon-filled glove box with water and oxygen levels less than 0.5 ppm. Galvanostatic discharge−charge measurements were carried out on a Neware battery test system, and the current density and specific capacity was calculated based on the amount of KB. The linear sweep voltammetry (LSV) and chronoamperometry were measured on an electrochemical workstation (Zennium IM6 station) after discharge at the depth of 1000 mAh g−1.

evident difference at different states in comparison to Fe7Ni3 electrode at different states. The first charged electrode of FeNi3 seemed blurry. It suggested that some product was decomposed incompletely or some byproducts cannot be decomposed on the surface of the electrode. This result was coincident with the different performance, including the reversibility and overpotentials, between Fe7Ni3 and FeNi3 electrodes in discharge−charge curves. In this work, Fe7Ni3 cathode presents a higher specific capacity and a lower overpotential, especially in the OER process. Previous researches reported that increasing the mass of Fe in alloy enhanced the OER process, which was attributed to the optimized electronic structure due to the introduction of Fe.35,42,55 It endowed the active sites with appropriate adsorption energy, which benefited LiO2 adsorption in the ORR process and improved LiO2 desorption in the OER process. Fe7Ni3 and FeNi3 electrodes exhibited big difference in the performances. Therefore, this result may be caused by the mass of Fe, which could optimize the electron structure of the material surface.35

3. CONCLUSIONS In summary, body-centered cubic Fe7Ni3 and face-centered cubic FeNi3 nanospheres were successfully synthesized by a simple solvothermal method. Fe7Ni3 nanospheres exhibited a much higher activity and durability than FeNi3 nanospheres and served as electrocatalysts for nonaqueous rechargeable Li−O2 batteries. Significantly, the FexNiy nanospheres employed as electrocatalysts tend to change the discharge products in nature, differing from other transition metal electrocatalysts. This work would provide a new perspective for transition metals electrocatalysts in Li−O2 batteries and broaden the scope for future research.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b00497. TEM images in high magnification; LSV and chronoamperometry profiles for the Li2O2-containing cells (PDF)



AUTHOR INFORMATION

Corresponding Authors

4. EXPERIMENTAL SECTION 4.1. Preparation of FexNiy Nanospheres. KOH (1 g) and octadecylamine (2 g) were added into 35 mL N-methylpyrrolidone, and then the certain amounts of iron(III) acetylacetonate and nickel(II) acetylacetonate were added successively with continuous stirring. When the solution became homogeneous, the mixture was transferred to a Teflon-lined autoclave followed by adding 5 mL hydrazine hydrate and removing the air by Ar flow for about 5 min. Then, the autoclave was treated at a temperature of 180 °C in an electric oven for 2 h to produce FexNiy nanospheres. Finally, the products were obtained by centrifuging at 12 000 rpm. The as-prepared samples were washed alternately several times with n-hexane, acetone, and dried at 40 °C under vacuum. 4.2. Characterization. The crystalline structures of the assynthesized samples were analyzed by X-ray powder diffraction (XRD) using a Phillips X’pert Pro MPD diffractometer (Cu Kα, λ = 1.54056 Å). The size distribution and morphology of the products were characterized by a field emission scanning electron microscopy (FESEM, acceleration voltage of 5 kV, S4800, Hitachi) and a high-resolution transmission electron microscopy (HRTEM, an acceleration voltage of 200 kV, JEM2010, JEOL). The compositions of the as-prepared samples were investigated by inductively coupled plasma-atomic emission spectrum (ICP-AES, Jarrel-ASH, ICAP-9000). The surface composition of the samples were characterized by X-ray photoelectron spectroscopy (XPS) using ESCALAB 250Xi

*E-mail: [email protected] (G.S.). *E-mail: [email protected] (H.L.). *E-mail: [email protected] (S.M.). ORCID

Genban Sun: 0000-0001-9005-8123 Shulan Ma: 0000-0002-8326-3134 Author Contributions

The manuscript was written through contributions of all of the authors. All of the authors approved the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Science Foundation of China (NSFC, Nos. 21271001, 21471020, 21271028, and 51272030).



REFERENCES

(1) Bruce, P. G.; Freunberger, S. A.; Hardwick, L. J.; Tarascon, J. M. Li-O2 and Li-S Batteries with High Energy Storage. Nat. Mater. 2012, 11, 19−29. (2) Park, M.; Sun, H.; Lee, H.; Lee, J.; Cho, J. Lithium-Air Batteries: Survey on the Current Status and Perspectives Towards Automotive Applications from a Battery Industry Standpoint. Adv. Energy Mater. 2012, 2, 780−800. 4275

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

(22) Black, R.; Lee, J. H.; Adams, B.; Mims, C. A.; Nazar, L. F. The Role of Catalysts and Peroxide Oxidation in Lithium-Oxygen Batteries. Angew. Chem., Int. Ed. 2013, 52, 392−396. (23) Oh, D.; Qi, J.; Han, B.; Zhang, G.; Carney, T. J.; Ohmura, J.; Zhang, Y.; Shao-Horn, Y.; Belcher, A. M. M13 Virus-Directed Synthesis of Nanostructured Metal Oxides for Lithium-Oxygen Batteries. Nano Lett. 2014, 14, 4837−4845. (24) Hu, X.; Cheng, F.; Han, X.; Zhang, T.; Chen, J. Oxygen BubbleTemplated Hierarchical Porous ε-MnO2 as a Superior Catalyst for Rechargeable Li-O2 Batteries. Small 2015, 11, 809−813. (25) Lu, F.; Wang, Y.; Jin, C.; Li, F.; Yang, R.; Chen, F. Microporous La0.8Sr0.2MnO3 Perovskite Nanorods as Efficient Electrocatalysts for Lithium−Air Battery. J. Power Sources 2015, 293, 726−733. (26) Liu, G.; Chen, H.; Xia, L.; Wang, S.; Ding, L. X.; Li, D.; Xiao, K.; Dai, S.; Wang, H. Hierarchical Mesoporous/Macroporous Perovskite La0.5Sr0.5CoO3‑x Nanotubes: A Bifunctional Catalyst with Enhanced Activity and Cycle Stability for Rechargeable Lithium Oxygen Batteries. ACS Appl. Mater. Interfaces 2015, 7, 22478−22486. (27) Oh, S. H.; Nazar, L. F. Oxide Catalysts for Rechargeable HighCapacity Li-O2 Batteries. Adv. Energy Mater. 2012, 2, 903−910. (28) Xiao, J.; Mei, D.; Li, X.; Xu, W.; Wang, D.; Graff, G. L.; Bennett, W. D.; Nie, Z.; Saraf, L. V.; Aksay, I. A.; Liu, J.; Zhang, J. G. Hierarchically Porous Graphene as a Lithium-Air Battery Electrode. Nano Lett. 2011, 11, 5071−5078. (29) Zhou, W.; Zhang, H.; Nie, H.; Ma, Y.; Zhang, Y.; Zhang, H. Hierarchical Micron-Sized Mesoporous/Macroporous Graphene with Well-Tuned Surface Oxygen Chemistry for High Capacity and Cycling Stability Li-O2 Battery. ACS Appl. Mater. Interfaces 2015, 7, 3389− 3397. (30) Chen, Y.; Li, F.; Tang, D.; Jian, Z.; Liu, C.; Golberg, D.; Yamada, A.; Zhou, H. Multi-walled Carbon Nanotube Papers as Binder-Free Cathodes for Large Capacity and Reversible Non-Aqueous Li−O2 Batteries. J. Mater. Chem. A 2013, 1, 13076−13081. (31) Thomas, M. L.; Yamanaka, K.; Ohta, T.; Byon, H. R. A Perfluorinated Moiety-Grafted Carbon Nanotube Electrode for the Non-Aqueous Lithium-Oxygen Battery. Chem. Commun. 2015, 51, 3977−3980. (32) Zhang, Z.; Su, L.; Yang, M.; Hu, M.; Bao, J.; Wei, J.; Zhou, Z. A Composite of Co Nanoparticles Highly Dispersed on N-Rich Carbon Substrates: an Efficient Electrocatalyst for Li-O2 Battery Cathodes. Chem. Commun. 2014, 50, 776−778. (33) Huang, B. W.; Liao, X. Z.; Wang, H.; Wang, C. N.; He, Y. S.; Ma, Z. F. Nanofibrous MnNi/CNF Composite Catalyst for Rechargeable Li-O2 Cell. J. Electrochem. Soc. 2013, 160, A1112−A1117. (34) Kwak, W.; Kang, T.; Sun, Y.; Lee, Y. J. Iron−Cobalt Bimetal Decorated Carbon Nanotubes as Cost-Effective Cathode Catalysts for Li−O2 Batteries. J. Mater. Chem. A 2016, 4, 7020−7026. (35) Cui, X.; Ren, P.; Deng, D.; Deng, J.; Bao, X. Single Layer Graphene Encapsulating Non-Precious Metals as High-Performance Electrocatalysts for Water Oxidation. Energy Environ. Sci. 2016, 9, 123−129. (36) Chen, T.; Deng, F.; Zhu, J.; Chen, C.; Sun, G.; Ma, S.; Yang, X. Hexagonal and Cubic Ni Nanocrystals Grown on Graphene: PhaseControlled Synthesis, Characterization and Their Enhanced Microwave Absorption Properties. J. Mater. Chem. 2012, 22, 15190−15197. (37) Pan, G.; Zhu, J.; Ma, S.; Sun, G.; Yang, X. Enhancing the Electromagnetic Performance of Co Through the Phase-Controlled Synthesis of Hexagonal and Cubic Co Nanocrystals Grown on Graphene. ACS Appl. Mater. Interfaces 2013, 5, 12716−12724. (38) Yuan, M.; Lin, L.; Yang, Y.; Nan, C.; Ma, S.; Sun, G.; Li, H. InSitu Growth of Ultrathin Cobalt Monoxide Nanocrystals on Reduced Graphene Oxide Substrates: an Efficient Electrocatalyst for Aprotic LiO2 Batteries. Nanotechnology 2017, 28, No. 185401. (39) Ma, T.; Yuan, M.; Islam, S. M.; Li, H.; Ma, S.; Sun, G.; Yang, X. FeNi3 Alloy Nanocrystals Grown on Graphene: Controllable Synthesis, In-Depth Characterization and Enhanced Electromagnetic Performance. J. Alloys Compd. 2016, 678, 468−477.

(3) Wang, J.; Li, Y.; Sun, X. Challenges and Opportunities of Nanostructured Materials for Aprotic Rechargeable Lithium−Air Batteries. Nano Energy 2013, 2, 443−467. (4) Shao, Y.; Park, S.; Xiao, J.; Zhang, J.; Wang, Y.; Liu, J. Electrocatalysts for Nonaqueous Lithium−Air Batteries: Status, Challenges, and Perspective. ACS Catal. 2012, 2, 844−857. (5) Ma, Z.; Yuan, X.; Li, L.; Ma, Z.; Wilkinson, D. P.; Zhang, L.; Zhang, J. A Review of Cathode Materials and Structures for Rechargeable Lithium−Air Batteries. Energy Environ. Sci. 2015, 8, 2144−2198. (6) Feng, N.; He, P.; Zhou, H. Critical Challenges in Rechargeable Aprotic Li-O2 Batteries. Adv. Energy Mater. 2016, 6, No. 1502303. (7) Cheng, F.; Chen, J. Metal-Air Batteries: from Oxygen Reduction Electrochemistry to Cathode Catalysts. Chem. Soc. Rev. 2012, 41, 2172−2192. (8) Guo, Z.; Zhou, D.; Dong, X.; Qiu, Z.; Wang, Y.; Xia, Y. Ordered Hierarchical Mesoporous/Macroporous Carbon: a High-Performance Catalyst for Rechargeable Li-O2 Batteries. Adv. Mater. 2013, 25, 5668− 5672. (9) Sun, B.; Chen, S.; Liu, H.; Wang, G. Mesoporous Carbon Nanocube Architecture for High-Performance Lithium-Oxygen Batteries. Adv. Funct. Mater. 2015, 25, 4436−4444. (10) Wang, Z.; Xu, D.; Xu, J.; Zhang, L.; Zhang, X. Graphene Oxide Gel-Derived, Free-Standing, Hierarchically Porous Carbon for HighCapacity and High-Rate Rechargeable Li-O2 Batteries. Adv. Funct. Mater. 2012, 22, 3699−3705. (11) Luo, W. B.; Gao, X. W.; Chou, S. L.; Wang, J. Z.; Liu, H. K. Porous AgPd-Pd Composite Nanotubes as Highly Efficient Electrocatalysts for Lithium-Oxygen Batteries. Adv. Mater. 2015, 27, 6862− 6869. (12) Peng, Z.; Freunberger, S. A.; Chen, Y.; Bruce, P. G. A Reversible and Higher-Rate Li-O2 Battery. Science 2012, 337, 563−566. (13) Su, D.; Kim, H.; Kim, W.; Wang, G. A Study of PtxCoy Alloy Nanoparticles as Cathode Catalysts for Lithium-Air Batteries with Improved Catalytic Activity. J. Power Sources 2013, 244, 488−493. (14) Cui, Z.; Li, L.; Manthiram, A.; Goodenough, J. B. Enhanced Cycling Stability of Hybrid Li-Air Batteries Enabled by Ordered Pd3Fe Intermetallic Electrocatalyst. J. Am. Chem. Soc. 2015, 137, 7278−7281. (15) Lu, J.; Lee, Y. J.; Luo, X.; Lau, K. C.; Asadi, M.; Wang, H. H.; Brombosz, S.; Wen, J.; Zhai, D.; Chen, Z.; Miller, D. J.; Jeong, Y. S.; Park, J. B.; Fang, Z. Z.; Kumar, B.; Salehi-Khojin, A.; Sun, Y. K.; Curtiss, L. A.; Amine, K. A Lithium-Oxygen Battery Based on Lithium Superoxide. Nature 2016, 529, 377−382. (16) Tong, S.; Zheng, M.; Lu, Y.; Lin, Z.; Zhang, X.; He, P.; Zhou, H. Binder-Free Carbonized Bacterial Cellulose-Supported Ruthenium Nanoparticles for Li-O2 Batteries. Chem. Commun. 2015, 51, 7302− 7304. (17) Guo, Z.; Zhou, D.; Liu, H.; Dong, X.; Yuan, S.; Yu, A.; Wang, Y.; Xia, Y. Synthesis of Ruthenium Oxide Coated Ordered Mesoporous Carbon Nanofiber Arrays as a Catalyst for Lithium Oxygen Battery. J. Power Sources 2015, 276, 181−188. (18) Jiang, J.; He, P.; Tong, S.; Zheng, M.; Lin, Z.; Zhang, X.; Shi, Y.; Zhou, H. Ruthenium Functionalized Graphene Aerogels with Hierarchical and Three-Dimensional Porosity as a Free-Standing Cathode For Rechargeable Lithium-Oxygen Batteries. NPG Asia Mater. 2016, 8, No. e239. (19) Mohamed, S. G.; Tsai, Y. Q.; Chen, C. J.; Tsai, Y. T.; Hung, T. F.; Chang, W. S.; Liu, R. S. Ternary Spinel MCo2O4 (M = Mn, Fe, Ni, and Zn) Porous Nanorods as Bifunctional Cathode Materials for Lithium-O2 Batteries. ACS Appl. Mater. Interfaces 2015, 7, 12038− 12046. (20) Wu, B.; Zhang, H.; Zhou, W.; Wang, M.; Li, X.; Zhang, H. Carbon-Free CoO Mesoporous Nanowire Array Cathode for HighPerformance Aprotic Li-O2 Batteries. ACS Appl. Mater. Interfaces 2015, 7, 23182−23189. (21) Hu, Y.; Zhang, T.; Cheng, F.; Zhao, Q.; Han, X.; Chen, J. Recycling Application of Li−MnO2 Batteries as Rechargeable Lithium−Air Batteries. Angew. Chem., Int. Ed. 2015, 54, 4338−4343. 4276

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277

ACS Omega

Article

(40) Liang, Y.; Liu, P.; Yang, G. Fabrication of One-Dimensional Chain of Iron-Based Bimetallic Alloying Nanoparticles with Unique Magnetizations. Cryst. Growth Des. 2014, 14, 5847−5855. (41) Abellán, G.; Prima-García, H.; Coronado, E. Graphene Enhances the Magnetoresistance of FeNi3 Nanoparticles in Hierarchical FeNi3−Graphene Nanocomposites. J. Mater. Chem. C 2016, 4, 2252−2258. (42) Wang, X.; Liu, X.; Tong, C.; Yuan, X.; Dong, W.; Lin, T.; Liu, L.; Huang, F. An Electron Injection Promoted Highly Efficient Electrocatalyst of FeNi3@GR@Fe-NiOOH for Oxygen Evolution and Rechargeable Metal−Air Batteries. J. Mater. Chem. A 2016, 4, 7762− 7771. (43) Kurlyandskaya, G. V.; Madinabeitia, I.; Murzakaev, A. M.; Sanchez-Ilarduya, M. B.; Beketov, V.; Medvedev, A. I.; Larranaga, A.; Safronov, A. P.; Schegoleva, N. N. Core-Shell Fine Structure of FeNi Magnetic Nanoparticles Produced by Electrical Explosion of Wire. IEEE Trans. Magn. 2014, 50, 1−4. (44) Wang, W.; Wang, Z.; Ding, Y.; Xi, J.; Lu, G. Partial Oxidation of Ethanol to Hydrogen over Ni-Fe Catalysts. Catal. Lett. 2002, 81, 63− 68. (45) Sun, B.; Huang, X.; Chen, S.; Munroe, P.; Wang, G. Porous Graphene Nanoarchitectures: an Efficient Catalyst for Low ChargeOverpotential, Long Life, and High Capacity Lithium-Oxygen Batteries. Nano Lett. 2014, 14, 3145−3152. (46) Qiu, F.; He, P.; Jiang, J.; Zhang, X.; Tong, S.; Zhou, H. Ordered Mesoporous TiC−C Composites as Cathode Materials for Li−O2 Batteries. Chem. Commun. 2016, 52, 2713−2716. (47) Jiang, Y.; Cheng, J.; Lu, Z.; Li, X.; Gong, Y.; Chi, B.; Pu, J.; Li, J. In-Situ Growth of CeO2 Nanoparticles on N-doped Reduced Graphene Oxide for Anchoring Li2O2 Formation in Lithium-Oxygen Batteries. Electrochim. Acta 2016, 210, 712−719. (48) Zhao, Q.; Wu, C.; Cong, L.; Zhang, Y.; Sun, G.; Xie, H.; Sun, L.; Liu, J. Yolk−Shell Co2CrO4 Nanospheres as Highly Active Catalysts for Li−O2 Batteries: Understanding the Electrocatalytic Mechanism. J. Mater. Chem. A 2017, 5, 544−553. (49) Yilmaz, E.; Yogi, C.; Yamanaka, K.; Ohta, T.; Byon, H. R. Promoting Formation of Noncrystalline Li2O2 in the Li-O2 Battery with RuO2 Nanoparticles. Nano Lett. 2013, 13, 4679−4684. (50) Zhai, D.; Wang, H. H.; Lau, K. C.; Gao, J.; Redfern, P. C.; Kang, F.; Li, B.; Indacochea, E.; Das, U.; Sun, H. H.; Sun, H. J.; Amine, K.; Curtiss, L. A. Raman Evidence for Late Stage Disproportionation in a Li-O2 Battery. J. Phys. Chem. Lett. 2014, 5, 2705−2710. (51) Jeong, Y. S.; Park, J. B.; Jung, H. G.; Kim, J.; Luo, X.; Lu, J.; Curtiss, L.; Amine, K.; Sun, Y. K.; Scrosati, B.; Lee, Y. J. Study on the Catalytic Activity of Noble Metal Nanoparticles on Reduced Graphene Oxide for Oxygen Evolution Reactions in Lithium-Air Batteries. Nano Lett. 2015, 15, 4261−4268. (52) Yang, J.; Zhai, D.; Wang, H. H.; Lau, K. C.; Schlueter, J. A.; Du, P.; Myers, D. J.; Sun, Y. K.; Curtiss, L. A.; Amine, K. Evidence for Lithium Superoxide-Like Species in the Discharge Product of a Li-O2 Battery. Phys. Chem. Chem. Phys. 2013, 15, 3764−3771. (53) Yao, X.; Cheng, Q.; Xie, J.; Dong, Q.; Wang, D. Functionalizing Titanium Disilicide Nanonets with Cobalt Oxide and Palladium for Stable Li Oxygen Battery Operations. ACS Appl. Mater. Interfaces 2015, 7, 21948−21955. (54) Eysel, H. H.; Thym, S. Raman Spectra of Peroxides. Z. Anorg. Allg. Chem. 1975, 411, 97−102. (55) Yang, Y.; Zhuang, L.; Lin, R.; Li, M.; Xu, X.; Rufford, T. E.; Zhu, Z. A Facile Method to Synthesize Boron-Doped Ni/Fe Alloy NanoChains as Electrocatalyst for Water Oxidation. J. Power Sources 2017, 349, 68−74.

4277

DOI: 10.1021/acsomega.7b00497 ACS Omega 2017, 2, 4269−4277