12
2004457 (1 of 12) © 2020 Wiley-VCH GmbH www.small-journal.com FULL PAPER 2D Sandwiched Nano Heterostructures Endow MoSe 2 / TiO 2x /Graphene with High Rate and Durability for Sodium Ion Capacitor and Its Solid Electrolyte Interphase Dependent Sodiation/Desodiation Mechanism Cai Liu, Miaoxin Zhang, Xin Zhang, Biao Wan, Xiaona Li, Huiyang Gou,* Yexin Wang, Fuxing Yin, and Gongkai Wang* C. Liu, M. Zhang, X. Zhang, F. Yin, G. Wang School of Material Science and Engineering Tianjin Key Laboratory of Materials Laminating Fabrication and Interface Control Technology Hebei University of Technology Tianjin 300130, P. R. China E-mail: [email protected] B. Wan Key Laboratory of Materials Physics of Ministry of Education School of Physics Zhengzhou University Zhengzhou 450001, P. R. China DOI: 10.1002/smll.202004457 1. Introduction Massive fossil resource consumption gives rise to a “butterfly effect,” such as global warming and environmental contamination. Such issues can be effectively circumvented by the develop- ment of intermittent renewable energy sources (wind, solar and so on), in which advanced energy storage technologies play an important part. Obviously, this irreplaceable role has already made the present day an era of lithium ion batteries (LIBs). [1] However, as another one of the energy storage stars, supercapacitors (SCs) appear to have been sidelined, due to the very limited energy density (5 Wh kg 1 ), even though SCs can achieve unparalleled power density and long lifespan. [2] Sodium ion capacitors (SICs) represent one of the promising asymmetric SCs, in which the charge storage occurs through faradaic reactions on the battery-anode side and electrical double-layers on the capacitor- cathode side. [3,4] Therefore, they could deliver a much improved energy density Nano heterostructures relying on their versatile construction and the breadth of combined functionality have shown great potential in energy storage fields. Herein, 2D sandwiched MoSe 2 /TiO 2x /graphene nano heterostructures are designed by integrating structural and functional effects of each component, aiming to address the rate capability and cyclic stability of MoSe 2 for sodium ion capacitors (SICs). These 2D nano heterostructures based on graphene platform can facilitate the interfacial electron transport, giving rise to fast reaction kinetics. Meanwhile, the 2D open structure induces a large extent of surface capacitive contribution, eventually leading to a high rate capability (415.2 mAh g 1 @ 5 A g 1 ). An ultrathin oxygen deficient TiO 2x layer sandwiched in these nano heterostructures provides a strong chemical-anchoring regarding the products generated during the sodiation/desodiation process, securing the entire cyclic stability. The associated sodiation/desodiation mechanism is revealed by operando and ex situ characterizations, which exhibits a strong solid electrolyte interphase (SEI) dependence. The simulations verify the dependent sodiation products and enhanced heterostructural chemical-anchoring. Assembled SICs based on these nano heterostructures anode exhibit high initial Coulombic efficiency, energy/power densities, and long cycle life, shedding new light on the design of nano heterostructure electrodes for high performance energy storage application. X. Li School of Chemical Engineering Hebei University of Technology Tianjin 300130, P. R. China H. Gou Center for High Pressure Science and Technology Advanced Research Beijing 100094, P. R. China E-mail: [email protected] Y. Wang Beijing National Laboratory of Molecular Science Beijing Key Laboratory of Magnetoelectric Materials and Devices College of Chemistry and Molecular Engineering Peking University Beijing 100871, P. R. China The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/smll.202004457. Small 2020, 2004457

2D Sandwiched Nano Heterostructures Endow MoSe2/TiO2 …sscms.hpstar.ac.cn/.../files/2020/12/59f2492db1ca061.pdf · 2020. 12. 14. · and ,

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • 2004457 (1 of 12) © 2020 Wiley-VCH GmbH

    www.small-journal.com

    Full PaPer

    2D Sandwiched Nano Heterostructures Endow MoSe2/TiO2−x/Graphene with High Rate and Durability for Sodium Ion Capacitor and Its Solid Electrolyte Interphase Dependent Sodiation/Desodiation Mechanism

    Cai Liu, Miaoxin Zhang, Xin Zhang, Biao Wan, Xiaona Li, Huiyang Gou,* Yexin Wang, Fuxing Yin, and Gongkai Wang*

    C. Liu, M. Zhang, X. Zhang, F. Yin, G. WangSchool of Material Science and EngineeringTianjin Key Laboratory of Materials Laminating Fabrication and Interface Control TechnologyHebei University of TechnologyTianjin 300130, P. R. ChinaE-mail: [email protected]. WanKey Laboratory of Materials Physics of Ministry of EducationSchool of PhysicsZhengzhou UniversityZhengzhou 450001, P. R. China

    DOI: 10.1002/smll.202004457

    1. Introduction

    Massive fossil resource consumption gives rise to a “butterfly effect,” such as global warming and environmental contamination. Such issues can be effectively circumvented by the develop-ment of intermittent renewable energy sources (wind, solar and so on), in which advanced energy storage technologies play an important part. Obviously, this irreplaceable role has already made the present day an era of lithium ion batteries (LIBs).[1] However, as another one of the energy storage stars, supercapacitors (SCs) appear to have been sidelined, due to the very limited energy density (≈5 Wh kg−1), even though SCs can achieve unparalleled power density and long lifespan.[2] Sodium ion capacitors (SICs) represent one of the promising asymmetric SCs, in which the charge storage occurs through faradaic reactions on the battery-anode side and electrical double-layers on the capacitor-cathode side.[3,4] Therefore, they could deliver a much improved energy density

    Nano heterostructures relying on their versatile construction and the breadth of combined functionality have shown great potential in energy storage fields. Herein, 2D sandwiched MoSe2/TiO2−x/graphene nano heterostructures are designed by integrating structural and functional effects of each component, aiming to address the rate capability and cyclic stability of MoSe2 for sodium ion capacitors (SICs). These 2D nano heterostructures based on graphene platform can facilitate the interfacial electron transport, giving rise to fast reaction kinetics. Meanwhile, the 2D open structure induces a large extent of surface capacitive contribution, eventually leading to a high rate capability (415.2 mAh g−1@ 5 A g−1). An ultrathin oxygen deficient TiO2−x layer sandwiched in these nano heterostructures provides a strong chemical-anchoring regarding the products generated during the sodiation/desodiation process, securing the entire cyclic stability. The associated sodiation/desodiation mechanism is revealed by operando and ex situ characterizations, which exhibits a strong solid electrolyte interphase (SEI) dependence. The simulations verify the dependent sodiation products and enhanced heterostructural chemical-anchoring. Assembled SICs based on these nano heterostructures anode exhibit high initial Coulombic efficiency, energy/power densities, and long cycle life, shedding new light on the design of nano heterostructure electrodes for high performance energy storage application.

    X. LiSchool of Chemical EngineeringHebei University of TechnologyTianjin 300130, P. R. ChinaH. GouCenter for High Pressure Science and Technology Advanced ResearchBeijing 100094, P. R. ChinaE-mail: [email protected]. WangBeijing National Laboratory of Molecular ScienceBeijing Key Laboratory of Magnetoelectric Materials and DevicesCollege of Chemistry and Molecular EngineeringPeking UniversityBeijing 100871, P. R. China

    The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/smll.202004457.

    Small 2020, 2004457

    http://crossmark.crossref.org/dialog/?doi=10.1002%2Fsmll.202004457&domain=pdf&date_stamp=2020-11-05

  • 2004457 (2 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    as compared with that of traditional symmetric SCs (electrical double-layer capacitors, EDLCs), while keeping the advantages of power density and lifespan.[5] However, the faradaic reactions are dominated by the semi-infinite diffusion throughout the bulk anode, which usually leads to a large kinetics gap between the anode and cathode. As a result, this bottleneck makes the development of advanced anode materials for SICs a popular imperative.[6–8]

    Recent exploration of 2D layered transition metal dichal-cogenides (TMDs) suggests a great potential in the field of energy storage.[9–12] As one typical member of TMDs, MoSe2 is obtaining considerable attention. Notably, MoSe2 with van der Waals dominating layered structure exhibits a larger interlayer spacing (6.4 Å) than these of famous graphene (3.5 Å) and ana-logue MoS2 (6.2 Å), as well as the relatively narrow and tun-able band gap (1.4 eV). These features endow MoSe2 with more attraction, particularly as anode for sodium ion storage.[13] Kang et al reported yolk–shell MoSe2 microspheres by a solid-state selenization as the anode for sodium ion storage, in which the maximum specific capacity can reach 527 mAh g−1 at the cur-rent density of 0.2 A g−1, however, the capacity decreases down to 433 mAh g−1 after 50 cycles. In order to improve the perfor-mance, MoSe2/carbon nanotube (CNT)/reduced graphene oxide (rGO) composites were reported by Kang subsequently,[14,15] much better rate performance (255  mAh  g−[email protected]  A  g−1, 173 mAh g−1@30 A g−1) and cyclic stability (capacity retention of 95%@1.0 A g−1 after 400 cycles) were obtained. Afterward, sev-eral literatures reported hierarchical MoSe2 with continuously improved performances, including the cycle life.[16–20] Even so, most of the reported cycle life is around hundreds to one thou-sand, which is way behind the expectation and requirement of anode, especially for the application of SICs.

    Meanwhile, the sodiation/desodiation mechanism of MoSe2 still remains controversial. For example, Jiang and co-workers Luo and co-workers proposed a reversible intercalation mecha-nism between MoSe2 and Mo+Na2Se, namely, when the Na ion is inserted, MoSe2 is converted to amorphous Mo and Na2Se, during the desodiation, the crystalline MoSe2 can be fully recov-ered.[19,21] However, Yang and co-workers reported a two-step reaction mechanism, during the first sodiation/desodiation process, amorphous Mo and Se is precipitated from MoSe2, which is an irreversible reaction. Afterward, beginning with the second sodiation/desodiation process, a reversible reaction occurs between Se and Na2Se.[22] In addition, it is worth noting that the intermediate product of Na2Se is identified in both pro-posed mechanisms. This kind of sodium selenides tends to dis-solve into the electrolyte, deteriorating the long-term stability severely.[23] Therefore, the underlying mechanisms involving in phase change process, reaction product, and the correlation to electrolyte should be further unveiled, which determine the performance of sodium ion storage eventually.

    Nano heterostructures have attracted great attention and shown potential in many fields, including electronics, and energy storage applications.[24–26] The reasons for this stem from their versatile construction of layered materials and the breadth of combined functionality. In this work, MoSe2/TiO2−x/graphene (MoSe2–TiO2−x–G) nano heterostructures are designed as the anode for SICs. For one thing, these 2D nano heterostructures based on graphene platform can facilitate

    the interfacial electron transport, giving rise to fast reaction kinetics during sodiation/desodiation. Meanwhile, this 2D open structure induces a strong surface capacitive contribution, eventually leading to a high rate capability, a specific capacity of 415.2 mAh g−1 at the current density of 5 A g−1 can be deliv-ered. For another, an ultrathin oxygen deficient TiO2−x layer is sandwiched in these nano heterostructures by an atomic layer deposited (ALD) technology, which provides a strong heter-ostructural-anchoring regarding to the intermediate product generated during the sodiation/desodiation, so as to secure the reversibility of entire sodiation/desodiation process to a large extent. A specific capacity of 382.2 mAh g−1 at the current den-sity of 2  A  g−1 still can be delivered stably after 2000 cycles, which surpasses most of the state-of-the-art MoSe2 reported in literatures. At the same time, the sodiation/desodiation mecha-nism correlated to different electrolytes is proposed based on operando characterizations of Raman and X-ray diffraction (XRD), and ex situ transmission electron microscopy (TEM), in which the induced solid electrolyte interphase (SEI) confirms a significantly influence on the reversibility of sodiation/deso-diation. Our atomics simulations also provide evidences about the heterostructural chemical-anchoring and sodiation/deso-diation mechanism. SICs based on the MoSe2–TiO2−x–G nano heterostructures anode exhibit high initial Coulombic efficiency (ICE), energy/power densities, and cycle life, shedding new light on the design of TMDs based materials for high perfor-mance sodium ion storage application.

    2. Results and Discussion

    2.1. Structure and Morphology of Nano Heterostructures

    The synthetic procedure is schematically shown in Figure  1. Anatase TiO2 nanoislands were firstly deposited on graphene nanosheets by the ALD technique.[27,28] Subsequently, the heat treatment with fast cooling is designed to prepare the oxygen deficient TiO2−x. On the one hand, the oxygen vacancy induced Ti3+ species would provide abundant unpaired electrons, which can dramatically increase the electrical conductivity of TiO2−x.[29] On the other hand, the chemical affinity of TiO2−x to the selenides can be enhanced,[30] making the positive influence on the electrochemical reactions correlated to MoSe2. Finally, MoSe2 was uniformly coated on the 2D TiO2−x/graphene (ALD TiO2−x–G), during the charge/discharge process, the electrons can be transported smoothly throughout the entire nano heter-ostructures, leading to a superior electrochemical performance. The phase structure of typical samples is identified by XRD (Figure 2a). All the peaks of bare MoSe2 have a good agreement with that of the hexagonal MoSe2 (JCPDS No. 29-0914), dem-onstrating the pure phase of MoSe2. No obvious difference is observed between MoSe2/graphene (MoSe2–G) and bare MoSe2, except a “hump” region belonging to graphene nanosheets. A small peak at 25.7° corresponds to the ALD anatase TiO2 of MoSe2/TiO2/graphene (MoSe2–TiO2–G), and this character-istic peak in MoSe2–TiO2−x–G becomes sharp after the heat treatment due to the enhanced crystallization, which is also confirmed in the ALD TiO2/graphene (ALD TiO2–G) sample (Figure  S1, Supporting Information). No apparent changes

    Small 2020, 2004457

  • 2004457 (3 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    Figure 1. Schematic illustration of synthesized MoSe2–TiO2−x–G nano heterostructures.

    Figure 2. Structure and morphology. a) XRD patterns of MoSe2–TiO2−x–G nano heterostructures and control samples. b–d) XPS spectra of O 1s, Ti 2p, and C 1s before and after oxygen deficiency treatment. e,f) TEM images of MoSe2–TiO2−x–G nano heterostructures. g) SEAD pattern. h) EDX mapping.

    Small 2020, 2004457

  • 2004457 (4 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    occur to the phase of MoSe2 after the heat treatment, demon-strating the pure MoSe2 coated on the ALD TiO2−x–G composite. Figure S2 in the Supporting Information and Figure 1b–d show the X-ray photoelectron spectroscopy (XPS) spectra of MoSe2–TiO2–G before and after the heat treatment. Compared to the pristine spectrum of O1s (Figure 1b), after the heat treatment, the spectrum of O1s has a significant left-shift of about 1.2 eV. At the same time, the similar left-shift occurs to the spectrum of Ti2p (Figure 1c), and an additional small peak is identified at about 456.5  eV, corresponding to the trivalent Ti (Ti3+). These results prove the formation of oxygen deficient TiO2. In addi-tion, the oxygen containing functional groups (bindings of CO and OCO) on graphene nanosheets are reduced after the heat treatment (Figure  1d), which is favorable to the elec-trochemical performance of nano heterostructures due to the improved electrical conductivity of graphene nanosheets. The heat treatment has no obvious influence on the surface compo-sition of MoSe2, in which the pair peaks of Mo 3d3/2(232.1 eV)/Mo 3d5/2(228.9  eV), and Se 3d3/2 (54.0  eV)/Se 3d5/2 (53.2  eV), correspond to the Mo4+ and, Se2−, respectively (Figure S2b,c, Supporting Information). The Raman spectra confirm not only the phases of TiO2 and MoSe2 (Figure S3, Supporting Information), but also the decreased D/G band ratio, which is consistent with the XPS results. Herein, the intensity of char-acteristic peaks of TiO2 in XPS and Raman spectra is relatively low due to the outer covered MoSe2.

    The morphology and microstructure of bare MoSe2 and MoSe2–TiO2−x–G nano heterostructures are characterized by scanning electron microscopy (SEM) and TEM. The lam-inar edges are observed in the agglomerated MoSe2 clusters (Figure  S4, Supporting Information), because the bare MoSe2 without any substrates tends to aggregate during the nuclea-tion and growth of the hydrothermal reaction. For the MoSe2–TiO2−x–G nano heterostructures (Figure  1e–h and Figures S5 and S6, Supporting Information), firstly, the large graphene plate with wrinkles is observed, which serves as the substrate for nucleation and growth of ALD TiO2 and subsequent MoSe2. Secondly, thin TiO2−x nanoislands with the radial size less than 20 nm cover the graphene substrate uniformly. Thirdly, the lay-ered MoSe2 is coated on the wrinkled graphene substrates and TiO2−x nanoislands, to form the MoSe2–TiO2−x–G nano hetero-structures. The high resolution TEM image reveals the inter-layer distances of 0.35 and 0.68 nm (Figure 1f), corresponding to the (101) and (002) planes of TiO2 and MoSe2, respectively. Meanwhile, the selected area electron diffraction (SAED) pat-tern clearly shows the diffraction rings of (002), (004), and (100) planes belonging to MoSe2, and diffraction spots of (101), (103), (200) belonging to TiO2, suggesting the polycrystallinity of MoSe2 and certain single crystal features of TiO2. In addition, the energy-dispersive X-ray spectrometry (EDX) mappings dem-onstrate the evenly distributed elements of C, O, Se, Mo, and Ti, without any other impurities (Figure 1h). In addition, high reso-lution TEM of ALD TiO2 and ALD TiO2−x was also performed as shown in Figure S7 in the Supporting Information, the dis-ordered crystalline structure and higher crystallization was obtained in the TiO2−x obviously, whereas the untreated TiO2 exhibited clearly resolved, well defined lattice fringes, and lower crystallization, which are in a good agreement with the XRD and XPS results. The specific surface areas of pristine graphene

    nanosheets, ALD TiO2@graphene, and MoSe2–TiO2−x–G were measured (Figure S8, Supporting Information), and the calcu-lated specific surface areas are 594.9, 159.4, and 130.8 m2 g−1, respectively.[31] The 2D open structure of the nano heterostruc-tures with high specific surface area could provide abundant of reaction sites and shorten the diffusion path, leading to a fast reaction kinetics. The electron paramagnetic resonance (EPR) spectra of ALD TiO2@graphene and ALD TiO2−x@graphene at the room temperature were obtained (Figure S9, Supporting Information). Based on the cw-EPR spectra, a signal with a g factor of 2.002 can be observed for the ALD TiO2@graphene. Generally, there is no EPR signal for pure anatase TiO2,[32,33] therefore, the obtained EPR signal is supposed to be from gra-phene nanosheets. The EPR signal of graphene is attributed to the widespread free carbon-centered radicals and oxygen radi-cals of residual oxygen-containing functional groups.[34] After the heat treatment, more apparent signal at g  = 2.002 occurs to the ALD TiO2−x@graphene, which was enhanced by the electrons trapped on the oxygen vacancy of TiO2−x. This result confirms the presence of oxygen vacancy of TiO2−x, consistent with the results of XPS. The identified weight fraction in the nano heterostructures by thermal gravimetric analysis (TGA) is 71.7%, 18.3%, and 10%, corresponding to MoSe2, TiO2−x, and graphene, respectively (Figure S10, Supporting Information). With respect to the 2D heterostructures of MoSe2–TiO2−x–G, on the one hand, the graphene nanosheets offer not only a substrate for the extensive nucleation and growth of TiO2 and MoSe2, eventually preventing the agglomeration of MoSe2, but also a highly conductive platform to facilitate the heterostruc-tural electron transport. On the other hand, the ultrathin ALD TiO2−x sandwiched in the nano heterostructures has a close interaction with both graphene substrates and outer MoSe2, which can avoid the negative influence of TiO2−x due to the rela-tively large band gap on the efficient electron transport between graphene and the outer MoSe2. Furthermore, this kind of TiO2−x with a planar geometry can improve the efficiency of hetero-structural chemical affinity to byproducts generated during sodiation/desodiation, without deteriorating the comprehensive electrochemical performance of nano heterostructures.

    2.2. Sodium Ion Storage Performance

    The sodium ion storage performances of MoSe2–TiO2−x–G nano heterostructures as well as other control samples are evaluated in sodium ion half-cells, and the effect of different electrolytes on performance and mechanism is also considered. Figure 3a exhibits the initial five cyclic voltammetry (CV) profiles of MoSe2–TiO2−x–G in the ether electrolyte (dimethyl ether, DME), in which a dominating pair of oxidation–reduction peaks around 1.4–1.7 V is associated with the sodiation/desodiation of MoSe2. Meanwhile, a weak pair of peaks around 0.1 V is related to the intercalation between sodium ions and graphene nanosheets. Additionally, the side reactions including the decomposition of electrolyte and formation of SEI layer below 0.5 V are actu-ally very limited. Generally speaking, the CV curves of our MoSe2–TiO2−x–G nano heterostructures in the DME electrolyte demonstrate a great electrochemical reversibility. By contrast, the dramatically different electrochemical behavior in the ester

    Small 2020, 2004457

  • 2004457 (5 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    electrolyte (ethylene carbonate, EC/diethyl carbonate, DEC) is exhibited, particularly the initial cycle (Figure S11, Supporting Information), in which a large amount of irreversible reac-tions occurs around 1.2 V and below 0.75 V. These reactions are mainly ascribed to the formation of SEI layer. The same features are obtained in the initial five galvanostatic charge/discharge curves of the MoSe2–TiO2−x–G nano heterostructures in these two electrolytes (Figure  3b; Figure S12, Supporting Informa-tion). This result may predict the different structures of SEI layer in these two electrolytes, which have a significant influence on the corresponding electrochemical performance. Figures S13 and S14 in the Supporting Information show the compar-ison of initial two galvanostatic charge/discharge curves of the MoSe2–TiO2−x–G and MoSe2–TiO2–G nano heterostructures, MoSe2–G and bare MoSe2 at the same current density in each electrolyte. Apparently, the specific capacity of nano heterostruc-tures is significantly improved as compared to the bare MoSe2, which is contributed from the efficient heterostructural electron transport originated from the highly conductive graphene plat-form and ultrathin TiO2−x layers with abundant unpaired elec-trons. The rate performance of these samples is compared in Figure 3c. The MoSe2–TiO2−x–G nano heterostructures in DME electrolyte deliver average specific capacities of 484.2, 461.3, and 413.9 mAh g−1, at the current densities of 0.1, 0.5, and 5 A g−1, respectively. The specific capacity can be fully recovered since the current density returns back to 0.1  A  g−1. This typical rate performance is much higher than these of control samples, for

    example, at the current density of 5 A g−1, it is about 24%, 79.3%, and 172.8% higher than that of MoSe2–TiO2–G, MoSe2–G, and bare MoSe2, respectively, demonstrating a superior rate capability. It is worth noting that the performance of the same MoSe2–TiO2−x–G nano heterostructures in EC/DEC electrolyte is not as good as that in the DME electrolyte, its specific capacity at 5  A  g−1 is only about 52% of that in the DME electrolyte, indicating the strong electrolyte dependent the electrochemical performance. Furthermore, the MoSe2–TiO2−x–G nano het-erostructures exhibit a remarkable capacity retention relative to current density, the specific capacity at 5 A g−1 still remains 85.7% of the initial one at 0.1  A  g−1, which outperforms those of many MoSe2 based anode in previous literatures (Figure 3d), such as MoSe2/C,[18,35,36] MoSe2@CoSe/N-C,[20] CNT/MoSe2,[17] MoSe2/dual C,[37] MoSe2/N-C,[38] MoSe2/G,[19,39] MoSe2/SnO2,[40] MoSe2/3D-C,[41] TiO2@MoSe2 P-C,[42] MoSe2.[43] Meanwhile, the MoSe2–TiO2−x–G nano heterostructures exhibit a long term cycle stability in the DME electrolyte (Figure  3e), the specific capacity is still 407.1 mAh g−1 with the retention of 92.5% after 2000 cycles at 2  A  g−1 (recorded after the rate measurement), which is much superior than most of the state-of-the-art MoSe2 anode reported in literatures (Table S1, Supporting Informa-tion). As expected, the performance degradation in the EC/DEC electrolyte is severe.

    The sodiation/desodiation kinetics and mechanism are investigated to give insights into the excellent rate capability and cyclic stability of MoSe2–TiO2−x–G nano heterostructures.

    Figure 3. Electrochemical performance. a,b) Initial five CV curves and galvanostatic charge/discharge profiles of MoSe2–TiO2−x–G nano heterostruc-tures in DME. c) Rate performances of samples in DME. d) Comparison of rate capability with literatures. e) Cyclic stability of samples in DME and EC/DEC. f) Contribution ratio of capacity of MoSe2–TiO2−x–G nano heterostructures in DME at various scan rates. g) Nyquist plots of samples in DME. h,i) Fitted results of RS, RSEI, and RCT of MoSe2 in DME and EC/DEC electrolytes, respectively.

    Small 2020, 2004457

  • 2004457 (6 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    The primary factor that controls the electrochemical reaction is evaluated according to the Power-law based on the corre-sponding CV profiles at series of scan rates (Figure S15a, Sup-porting Information).[44] The calculated b value of ≈0.89 sug-gests that the sodiation/desodiation process relative to MoSe2–TiO2−x–G nano heterostructures in DME electrolyte is strongly influenced by the capacitive confined process (Figure S15b, Supporting Information). The capacitive contribution ratio can be quantified based on the k factor method according to the equation

    i k v k vV 1 21/2( ) = + (1)

    where all the k factors can be obtained from CV profiles at a given potential under different scan rates (v).[45–48] Based on the calculated results, this capacitive confined process occurs throughout the entire voltage range (0.1–3.0  V vs Na+/Na) (Figure S15c, Supporting Information), and contrib-utes a majority to the total capacity (76%@0.1 mV s−1, and up to 93%@2 mV s−1) (Figure 3f), enabling a fast reaction kinetics and rate capability. The high capacitive contribution is ben-efited from the 2D open structure of the nano heterostructures, in which graphene acts as the conductive platform, the outer MoSe2 with layered structure delivers the intrinsic rate con-tribution, and the sandwiched TiO2−x may provide interfacial sites for capacitive adsorptions as well.[49–52] By contrast, it can be seen that the CV curves recorded in the EC/DEC electrolyte become distorted as the increase of scan rates, demonstrating the deteriorated electron and ion transport efficiency (Figure S16a, Supporting Information). Even though the similar b value and the capacitive confined process of full voltage range are obtained (Figure S16b,c, Supporting Information), the contribution ratio of this capacitive confined process is about 20% lower than that in the DME electrolyte (Figure  3f), leading to the dete-riorated performance in the EC/DEC electrolyte. The Nyquist plots of these samples in the DME electrolyte are shown in Figure  3g. The decreased semicircle represents the decreased charge transfer resistance (RCT), revealing the much faster charge transfer kinetics in the MoSe2–TiO2−x–G nano hetero-structures. In situ electrochemical impedance spectroscopy (EIS) technique is adopted to analyze the reaction kinetics of MoSe2–TiO2−x–G nano heterostructures at typical points during the charge/discharge process in both DME and EC/DEC elec-trolytes. The recorded Nyquist plots consist of semicircles and straight line parts, corresponding to impedance of SEI film (RSEI), charge transfer resistance (RCT), and Warburg diffusion impedance (ZW), respectively (Figure S17 and S18, Supporting Information).[53,54] The intersection at the real axis represents the equivalent series resistance (RS). The RS of electrode in these two electrolytes is much stable at ≈5 Ω (Figure 3h), how-ever, the RCT and RSEI exhibit dramatically different tendencies. The RCT and RSEI in DME remain relatively small and extremely stable, they are around ≈75 and ≈45 Ω during the entire ini-tial two discharge/charge processes, whereas the initial RCT in EC/DEC is about 105 Ω (Figure 3i), which is two and a half times larger than that in DME (42 Ω), furthermore, it increases steeply along with the sodiation (discharge) starting around 0.9 V, and reaches the maximum value of 300 Ω, then decreases gradually with the completion of sodiation. Subsequently, the

    RCT keeps stable, but it is still two times larger than that in DME. The similar tendency of RSEI is obtained along with RCT, which is overall larger than that in DME. This result not only confirms the occurrence of abundant side reactions relative to the formation of SEI film in EC/DEC, but also provides evi-dences to the higher efficient and faster kinetics in DME than that in EC/DEC. Therefore, apart from the sodiation/desodia-tion mechanism, the superior electrochemical performance of MoSe2–TiO2−x–G nano heterostructures in the DME electrolyte is ascribed to the following aspects: 1) the nano heterostruc-tures with the close and strong interactions among graphene platform, oxygen deficient TiO2−x, and MoSe2 facilitate the het-erostructural electron transport throughout the entire nano het-erostructures; 2) the outer MoSe2 exhibits extensive and open active sites for sodiation, enabling the fast reaction kinetics and rate capability; 3) the stable behavior in the DME rather than EC/DEC electrolyte secures the smooth and fast ion transport.

    2.3. SEI Dependent Sodiation/Desodiation Mechanism

    The sodiation/desodiation mechanisms of MoSe2 in these two electrolytes are also probed. Figure 4a,b exhibit the in situ Raman contour plots of MoSe2 obtained at several typical voltages during the initial discharge/charge process in DME electrolyte. The corresponding Raman patterns are shown in Figure S19a in the Supporting Information. In the pattern of open cir-cuit voltage (OCV), a main characteristic peak at 240 cm−1 is attributed to A1g vibration mode of MoSe2 phase. Along with the insertion of sodium ions, this main peak becomes weak, and eliminates at the full sodiation state (0.1  V). During the sub-sequent desodiation process, this characteristic peak recovers gradually after voltage plateau (1.5  V), and can be identified apparently at the full desodiation state (3  V). This operando Raman result can demonstrate a great reversibility of MoSe2 during sodiation/desodiation in DME. By contrast, at the full desodiation state of MoSe2 in EC/DEC, an apparently left-shifted peak at 235 cm−1 is confirmed, which can be indexed into the first order A1 symmetric bond-stretching modes belonging to amorphous Se,[55] rather than MoSe2 (Figure S19b, Supporting Information). This dramatically different result exhibits a dif-ferent sodiation/desodiation mechanism of MoSe2 in EC/DEC, namely, MoSe2 is not reformed, whereas amorphous Se is segregated after the initial desodiation. In order to explore further evidences about the two different mechanisms, ex situ TEM is carried out to analyze the possible phase transforma-tion process (Figure  4e–h). As compared with the high reso-lution TEM image of pristine MoSe2 that shows the clear and ordered layer structure (Figure S20, Supporting Information), after full sodiation (at 0.1 V), the short range ordered structure with a little larger interlayer spacing (0.72 nm) than that of pris-tine MoSe2 (0.68 nm) is observed in DME (Figure 4e), coupling with the uniform distribution of Na element in the EDX ele-ment mapping (Figure 4f), this phase is assigned to NaxMoSe2. In addition, a few lattice fringes with the interlayer spacing of 0.33  nm correspond to the (200) plane of Na2Se phase. The phases of NaxMoSe2 and Na2Se can also be confirmed in the corresponding SEAD pattern (Figure  4e inset). Some local areas exhibiting amorphous features may be the segregated Mo

    Small 2020, 2004457

  • 2004457 (7 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    Figure 4. SEI dependent sodiation/desodiation mechanism. a,b) In situ Raman contour plots of MoSe2 in DME. c,d) In situ Raman contour plots of MoSe2 in EC/DEC. e) Ex situ TEM image and corresponding SEAD pattern of MoSe2 at 0.1 V in DME. f) Corresponding EDX mapping at 0.1 V. g) Ex situ TEM image and corresponding SEAD pattern of MoSe2 at 3 V in DME. h) Corresponding EDX mapping at 3 V. i) Schematic illustration of sodiation/desodiation mechanism of MoSe2 in DME. j,k) SEM images of cycled MoSe2 electrode in DME and EC/DEC, respectively. l,m) Schematic illustration of SEI dependent sodiation/desodiation mechanism of MoSe2 in DME and EC/DEC, respectively.

    Small 2020, 2004457

  • 2004457 (8 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    phase derived from the decomposition of NaxMoSe2. Then, at the full desodiation state of 3 V (Figure 4g), lattice fringes with the interlayer spacing of 0.68 nm correspond to the (002) plane of MoSe2 phase, in which the recovered MoSe2 exhibits a short range ordered structure, demonstrating the apparent structure refinement and rearrangement. This structure refinement can be also confirmed in the in situ XRD profiles (Figure S21a–c, Supporting Information). The SEAD image indicates the (002) and (004) planes of MoSe2, as well as some amorphous fea-tures that may be Se and Mo (Figure 4g inset). Small amount of Na element is traced in the EDX mapping, suggesting the extraction of Na ion from active materials and the formation of thin SEI layer (Figure  4h). Herein, a schematic illustration is proposed to summarize the sodiation/desodiation mecha-nism of MoSe2 in DME electrolyte (Figure 4i), namely, during the initial sodiation, MoSe2 converts into NaxMoSe2 accompany with slight segregation of Mo and Na2Se phases. Then, during the initial desodiation, MoSe2 is recovered from the NaxMoSe2 phase to a large extent, accompanied with a deep structure refinement, at the same time, small amount of Se with amor-phous feature is formed extracted from Na2Se. Subsequently, MoSe2 and slight Se follow the individual electrochemical pro-cess reversibly.

    By contrast, the main phases are similar during the sodia-tion/desodiation process in EC/DEC electrolyte. However, much larger amount of Na2Se phases and amorphous regions (supposed to be the segregated Mo) are identified in the ini-tial sodiation state than that in DME (ex situ TEM results in Figure S22a–c, Supporting Information). Subsequently, large amount of amorphous phases (supposed to be the segregated Mo and Se) are observed at the full desodiation state (3  V), whereas the phase of MoSe2 is rarely identified (Figure S22d,e, Supporting Information), which is consistent with the operando Raman results in EC/DEC (Figure  4d; Figure S19b, Supporting Information). In addition, a heavy Na element distribution is traced in the EDX mapping at the complete desodiation state (Figure S22f, Supporting Information). These results indicate that the very thick SEI layer is formed in the EC/DEC electrolyte (Cycled MoSe2 electrodes in DME and EC/DEC elec-trolytes are compared shown in Figure 4j–k), more importantly, this thick SEI layer has a significant influence on the revers-ible reaction between MoSe2 and NaxMoSe2, and induces the segregation of Mo and Na2Se phases due to the interaction (such as physical and chemical anchoring) between NaxMoSe2 and the thick SEI layer. Therefore, in the EC/DEC electrolyte, the pristine MoSe2 tends to convert to Se to a large extent, then undergoes the reversible reaction between Na2Se and Se, which eventually affects the cyclic stability of MoSe2 owing to the severe dissolution of sodium selenides in the electrolyte. In addition, the composition of each SEI layers in DME and EC/DEC was probed by XPS (Figure S23, Supporting Information). In the spectra of C 1s, considerable parts of OCC, Na2CO3, OCO, CCO, and CC can be identified. In the spectra of O 1s, main parts of Na–KLL, COC, CONa, and Na2O can be probed. However, it is worth noting that the contents of these parts are dramatically different, indicating different compositions of SEI layers in DME and EC/DEC, consistent with the above results. The mechanism and electrolyte effect are illustrated in Figure 4l,m. Actually, the bare MoSe2 can just

    afford a hundred of cycles in the EC/DEC electrolyte (Figure S24, Supporting Information), which highlights the importance of our designed MoSe2–TiO2−x–G nano heterostructures, as well as the SEI dependent sodiation/desodiation mechanism.

    2.4. Mechanism Discussion by Density Functional Theory (DFT) Calculations

    To understand the formation of nonequilibrium phases during the sodiation process, the structure predictions are performed using the state-of-the-art structure searching algorithm. The potentially stable phases in Na and MoSe2 system can be obtained by the insertion of sodium into MoSe2 lattice without constrains of atomic sodium numbers. The thermos-dynamical stability of obtained phases is assessed by comparing the total energy of sodium and MoSe2. Figure 5a shows the formation energy of obtained structures with the low-energy structures staying on the convex hull. Four NaxMoSe2 (x  = 0.5, 2.5, 3, and 4) phases are found to be stable with the increasing sodium concentration, which are likely to present in the sodiation pro-cess of the anode during cycling. This result proves that the possible sodiation products could be the NaxMoSe2 (x = 0.5, 2.5, 3, and 4) phases, which is consistent with the result obtained in DME electrolyte. From the structural identification in Figure 5b, we found that Na prefers the interstitial site of layered MoSe2 in Na0.5MoSe2, as we expected. When more Na atoms insert into MoSe2 lattice, Na atoms show the strong chemical affinity to Se atoms, whereas Mo atoms tend to stay together in Na2.5MoSe2, Na3MoSe2 to Na4MoSe2, which may indicate the dissolving out of Mo from these compounds. Then, the structural stability of obtained phases is evaluated by the calculations of formation energy using different reaction routes. Interestingly, although these phases could be stable relative to Na and MoSe2, we found that these lowest energy structures are not favorable in comparison with different reaction routes, in particular with Na2Se+MoSe2+Mo (Figure 5c), as a result, it is possible that the sodiation product could decompose into the relatively stable constitute during the charge/discharge cycling. This decom-position may depend on the external condition, such as the intercalation with SEI layer, which is consistent with the result obtained in the EC/DEC electrolyte. Furthermore, the hetero-structural chemical affinity between the intermediate product of Na2Se and the nano heterostructures are further calculated to reveal the assistant influence on the cyclic performance. The adsorption models of Na2Se/TiO2 and Na2Se/defective TiO2−x on graphene substrate (referred to as Na2Se/TiO2/G and Na2Se/TiO2−x/G) are built for DFT calculations, as shown in Figure 5d, the calculated binding energy (Eb) of optimized Na2Se/TiO2/G (Eb  =  −3.06  eV) is much greater than that of Na2Se/TiO2−x/G (Eb = −4.21 eV), which indicates the enhanced chemical affinity between Na2Se/TiO2−x and graphene substrate. We also noted that Se atoms in Na2Se have the preference to stay in the posi-tion of oxygen in defective TiO2, and thus Se atoms can also obtain more electrons from neighboring Ti atoms. The calcu-lated density of states (Figure S25, Supporting Information) indicate that TiO2−x–G maintains the metallic feature with a slight increased stability having lower values density of states at Fermi level. The ultrathin ALD TiO2−x layer with unpaired

    Small 2020, 2004457

  • 2004457 (9 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    electrons contributes to this metallic feature, improving the efficient electron transport between graphene platform and the outer MoSe2, and leading to the fast reaction kinetics.

    2.5. SICs Performance

    The SICs are constructed using the MoSe2–TiO2−x–G nano het-erostructures as the anode, activated carbon (AC) as the cathode, and 1.0 m NaPF6 in DME as the electrolyte, respectively. The mass ratio of the cathode to anode and operation voltage range is optimized in advance.[56] Figure S26 in the Supporting Infor-mation exhibits the typical performance of AC in DME. The SICs based on a certain mass ratio (2:1) are measured in dif-ferent voltage ranges (0.5–3.5, 0.5–3.0, 0.5–3.8, and 0.2–3.5  V). An obvious decomposition occurs to the DME electrolyte when the voltage is up to 3.8 V (Figure S27a, Supporting Information), based on the calculation equation of energy/power densities, an appropriate voltage range of 0.5–3.5 V is selected in our experi-mental condition. Based on this voltage range, the SICs with various mass ratios are tested, apparently, the one with the mass ratio of 3:1 exhibits an optimized performance (Figure S27b, Supporting Information). The CV and galvanostatic charge/discharge profiles demonstrate the hybrid energy storage mech-anism in the SICs (Figure 6a,b; Figure S27c, Supporting Infor-mation). The ICE of SICs is a very important metric, which can

    be effectively improved by a presodiation treatment of anode. As a result, the ICE of SICs with presodiation under different mass ratios is about 89% (Figure 6c), which is two times larger than that without presodiation treatment (Figure S27d, Sup-porting Information). Therefore, the optimized SICs can deliver a maximum energy density of 109 Wh kg−1 at the power density of 100 W kg−1, and exhibit an excellent energy density retention, when the power density reaches 8  kW kg−1, the energy density remains 64  Wh kg−1, surpassing many results reported in lit-eratures (Figure  6d), such as VS2/C,[8] MoSe2/N, P@Carbon,[22] MoSe2/graphene,[39] MoSe2 and SnO2,[40] S/Ni@C,[57] MCNFs-V,[58] TiS2/MXene,[59] Na–Mn–Co–Mg–O,[60] MoS2/C,[61] metal oxide-based (Mn2+/Nb5+).[62] As expected, the SICs also show a great capacity retention of 88% at the current density of 2 A g−1 after 3000 cycles (Figure  6e), demonstrating the superior and stable performance of MoSe2–TiO2−x–G nano heterostructures anode. Meanwhile, the fully charged SICs in 29 s can power a LEDs pattern (Figure 6f), further proving the high energy/power density feature of SICs, as well as the promising perspective.

    3. Conclusions

    In summary, MoSe2–TiO2−x–G nano heterostructures have been demonstrated as the high performance anode for SICs. A maximum specific capacity of 415.2  mAh  g−1 at the current

    Figure 5. a) Obtained stoichoimetries in the Na and MoSe2 system through state-of-art structural searching, lowest energy structures stay on the convex hull. b) Crystal structures of stable stoichoimetries. c) Formation energy evaluations using different reaction routes. d) Adsorption structure models of Na2Se/TiO2/graphene (top) and Na2Se/TiO2−x/graphene (bottom).

    Small 2020, 2004457

  • 2004457 (10 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    density of 5  A  g−1 can be delivered, and a specific capacity of 382.2  mAh  g−1 at the current density of 2  A  g−1 still can be delivered stably after 2000 cycles, which surpasses most of the state-of-the-art MoSe2 reported in literatures. The remarkable sodium storage performance can be attributed to the combined functionalities of each component in these nano heterostruc-tures. For one thing, the 2D nano heterostructures based on graphene platform can facilitate the interfacial electron trans-port, giving rise to a fast reaction kinetics. Meanwhile, this 2D open structure induces a strong surface capacitive contribu-tion, eventually leading to a high rate capability. For another, an ultrathin oxygen deficient TiO2−x layer is sandwiched in these nano heterostructures by ALD technology, which provides a strong chemical-anchoring regarding to the intermediate product generated during the sodiation/desodiation, so as to secure the reversibility of entire sodiation/desodiation process to a large extent. More significantly, the operando and ex situ characterizations reveal that the sodiation/desodiation mecha-nism varies depending on various SEI structures. Our atomics simulations prove the possibility of sodiation products during

    charge/discharge cycling, which may be influenced by the external condition. Meanwhile, the enhancement of chemical affinity between the intermediate product and defective nano heterostructures facilitates the cyclic stability of nano hetero-structures. Assembled SICs based on these nano heterostruc-tures anode exhibit high ICE, energy/power densities, and cycle life, shedding new light on the design of nano heterostructure electrodes for high performance energy storage application.

    4. Experimental SectionMaterials: Graphene nanosheets were purchased from Shanghai

    SIMBATT Energy Technology Company, China. TiO2–G was obtained using an ALD rotary reactor (Zhiliande Tech Co. Ltd., Zhenjiang, China). Titanium tetrachloride (TiCl4, 99.8%) and high performance liquid chromatography (HPLC) grade H2O were the precursors supplied by Zhiliande Tech Co. Ltd., Zhenjiang, China. Se powder (96%), N2H4∙H2O (98%), and Na2MoO4∙2H2O (98%) were purchased from Aladdin Co. Ltd, China. The AC (YP-50F) was purchased from Kuraray Chemical Co., Ltd. Other solvents with analytical grade were used as received.

    Figure 6. SICs performance. a) Typical CV curves of anode, cathode, and full cell. b) Galvanostatic charge/discharge profiles of SICs at various current densities. c) Rate performance and ICE of SICs with pre-sodiation treatment under different mass ratios of cathode to anode. d) Comparison of Ragone plot with literatures. e) Cyclic stability of SICs. f) Digital image of SICs powering a LEDs pattern.

    Small 2020, 2004457

  • 2004457 (11 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    ALD TiO2−x/Graphene: First, the ALD TiO2–G was prepared according to the previous report.[27,28] Subsequently, a certain amount (≈15 mg) of ALD TiO2–G was heated in a tube furnace under H2 at 600 °C for 1 h, then quenched by a fast cooling rate to room temperature. Finally, the oxygen deficient TiO2−x/graphene was obtained.

    MoSe2–TiO2−x–G nano Heterostructures: MoSe2 was grown on TiO2−x/graphene via a hydrothermal process. First, 157.94  mg selenium was dissolved in 5  mL N2H4∙H2O (98%) under stirring at 90  °C for 24  h. Afterward, 247.95  mg Na2MoO4∙2H2O was added into 12.5  mL deionized (DI) water to obtain a transparent solution, then above as-obtained solutions and 77  mg TiO2−x/graphene were mixed and stirred for 3 h followed by transferring into a 200  mL Teflon-lined stainless steel autoclave. The autoclave was heated at 200 °C for 24 h, then cooled down to room temperature. The precipitate was washed by DI water several times followed by calcination in a mixed atmosphere of H2 (8%) and Ar at 400 °C for 2 h, finally, the MoSe2–TiO2−x–G nano heterostructures were obtained. By contrast, the bare MoSe2, MoSe2/graphene composite, and MoSe2–TiO2–G nano heterostructures were also prepared as control samples following the same hydrothermal procedures with different substrates.

    Material Characterizations: The general XRD was performed on an X-ray diffraction system (D8 Focus, Bruker, Germany). Raman spectra were recorded on a spectrometer (inVia Reflex, Renishaw, UK) with a laser excitation wavelength of 532 nm. The incident source wavelength is 1.542 Å under operation parameters of 40  kV and 40  mA at room temperature. The corresponding operando characterizations were performed using home-made sample holders. XPS was measured on a Thermo Scientific ESCALAB 250XI Probe system. The morphology and microstructure of materials were observed by a field emission scanning electron microscope (FESEM, JEOL JSM-7100F) performed with an EDX spectrometer and a TEM (Tecnai LaB6). TGA (SDTQ-600, TA Instruments-Waters LLC) was performed in air at a ramping rate of 10° min−1 from room temperature to 700  °C. Specific surface area was measured on V-Sorb 2800P Surface Area instrument by the adsorption/desorption isotherm of Brunauer–Emmett–Tellter (BET) analyses. The EPR spectra were obtained on a Bruker Elexsys E580 spectrometer with a superhigh sensitivity probe head ( f = 9.376 GHz).

    Calculation Method: The structural searching of Nax(MoSe2)1−x was performed using evolutionary structure prediction method (USPEX).[63,64] During the searching process, the composition of Mo and Se was fixed by 1:2 and the total atoms in the unit cell were set up to 48 atoms. 50 generations were searched and each generation contained 60 individual candidates. Afterward, structural searches with fixed compositions with low formation enthalpy were carried out by CALYPSO method.[65,66] The structural relaxations and electronic structure calculations were carried out using ab initio DFT as implemented in the Vienna ab initio simulation package (VASP).[67–69] The PAW potentials of C_sv: 2s2, 2p2; Ti_pv: 2p6, 3d2, 4s2; O_sv: 2s2, 2p6; Se: 4s2, 4p4, Na_pv:2p6, 3s1, and Mo: 4d5, 5s1were adopted.[70,71] The electron exchange–correlation interaction was treated by generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE).[72] In the calculation of Na–MoSe2, The energy cutoff was set to 600  eV and Monkhorst–Pack scheme was used to sample the Brillouin zone with a resolution of 2π  × 0.03 Å−1.[73] The energy and force convergence character were set as 10−6  eV and 1 meV Å−1, respectively. In the calculation of slab models, the energy cutoffs were set to 400 eV. The binding energies between Na2Se and substrate were calculated by the following equation[74]

    b sub/Na Se sub Na Se2 2E E E E= − − (2)

    where the / , andsub Na Se sub Na Se2 2E E E are the total energies of substrate bonding with Na2Se, substrate, and Na2Se, respectively. The graphene (G) substrate was constructed by 6 × 5 × 1 unit cell of graphene, which contained 50 atoms with lattice parameter of a  = 12.34 Å. The [1 0 1] crystal planes for TiO2 were used to model G–TiO2 substrate. Both the lattice of graphene and TiO2 in [1 0 1] slab were redefined to make the lattice vectors perpendicular to each other. Afterward, the G–TiO2

    slab was constructed using 5 × 6 × 1 unit cells for redefined graphene (120 atoms) and 2 × 4 × 1 unit cells for redefined TiO2 (288 atoms). The lattice mismatch between G and TiO2 was lower than 2.6%, and the lattice parameters in G–TiO2 slab model was a  = 20.83 Å and b  = 15.29  Å. The vacuum was set as 20 Å for all the slab models. Only Γ point in the first Brillouin zone was calculated.[73]

    Electrochemical Measurements: For the sodium ion half-cell measurement, the as-obtained samples were evaluated using the type CR2025 coin cells. Firstly, the working electrode was prepared by mixing 80 wt% active materials, 10 wt% carbon black (CB) and 10 wt% binder (poly-vinylidene difluoride, PVDF) dissolved in N-methylpyrrolidone (NMP). The resultant slurry was uniformly coated onto a copper current collector and vacuum dried at 80  °C for 12 h, then punched into 10  mm discs and pressed under a pressure of 24  MPa to obtain a densified electrode. The mass loading of electrode materials was between 1.3 and 2.0 mg cm−2. Sodium foil was the counter and reference electrode. Glass fiber paper was used as the separator. There were two types of electrolytes, they were 1.0 m NaClO4 in ethylene carbonate/diethyl carbonate (EC/DEC, 1:1 by volume) with 5 vol% fluoroethylene carbonates (FEC) and 1.0 m NaPF6 in dimethyl ether, respectively. For the sodium ion half-cell of AC, the weight ratio of AC, CB, and PVDF was also 8:1:1, the slurry of cathode materials was coated on aluminum foil, and then the same processes was followed. In the case of SICs, the MoSe2–TiO2−x–G nano heterostructure electrode and AC were used as the anode and cathode, respectively. The anode was preactivated in the half-cell under the current density of 100 mA g−1 for two cycles. All the cells were assembled and disassembled in the Ar filled glove box with O2 and H2O lower than 0.01 ppm. The anode half-cell was investigated in a voltage window of 0.1–3 V (vs Na+/Na) by using an electrochemical workstation (Ametek, Princeton Applied Research, Versa STAT 4) for the CV and EIS measurements. The cathode half-cell was investigated in a voltage window of 2.5–4 V (vs Na+/Na). All the rate performances and long-term cycling performances were recorded on a multichannel battery testing system (Neware BTS4000) at room temperature. The calculation of energy/power density can be found in the supporting information.

    Supporting InformationSupporting Information is available from the Wiley Online Library or from the author.

    AcknowledgementsThe authors acknowledge the financial support from the Natural Science Foundation of Tianjin City, China (No. 19JCYBJC17900), the Natural Science Foundation of Hebei Province, China (No. E2018202123), and the Jian-Hua Research Foundation of Hebei University of Technology (No. HB1921).

    Conflict of InterestThe authors declare no conflict of interest.

    Keywordselectrochemical performances, MoSe2, nano heterostructures, sodiation/desodiation mechanisms, sodium ion capacitors

    Received: July 23, 2020Revised: September 8, 2020

    Published online:

    Small 2020, 2004457

  • 2004457 (12 of 12)

    www.advancedsciencenews.com

    © 2020 Wiley-VCH GmbH

    www.small-journal.com

    [1] M. Li, J. Lu, Z. Chen, K. Amine, Adv. Mater. 2018, 30, 1800561.[2] C.  Choi, D. S.  Ashby, D. M.  Butts, R. H.  DeBlock, Q.  Wei, J.  Lau,

    B. Dunn, Nat. Rev. Mater. 2020, 5, 5.[3] D. Xu, D. Chao, H. Wang, Y. Gong, R. Wang, B. He, X. Hu, H. J. Fan,

    Adv. Energy Mater. 2018, 8, 1702769.[4] Y. Li, H. Wang, L. Wang, Z. Mao, R. Wang, B. He, Y. Gong, X. Hu,

    Small 2019, 15, 1804539.[5] K.  Zou, P.  Cai, Y.  Tian, J.  Li, C.  Liu, G.  Zou, H.  Hou, X.  Ji,

    Small Methods 2020, 4, 1900763.[6] Y. Jiang, J. Liu, Energy Environ. Mater. 2019, 2, 30.[7] Z. Xia, H. Sun, X. He, Z. Sun, C. Lu, J. Li, Y. Peng, S. Dou, J. Sun,

    Z. Liu, Nano Energy 2019, 60, 385.[8] D. Xu, H. Wang, R. Qiu, Q. Wang, Z. Mao, Y. Jiang, R. Wang, B. He,

    Y. Gong, D. Li, X. Hu, Energy Storage Mater. 2020, 28, 91.[9] W. Choi, N. Choudhary, G. H. Han, J. Park, D. Akinwande, Y. H. Lee,

    Mater. Today 2017, 20, 116.[10] Y. Fang, D. Luan, Y. Chen, S. Gao, X. W. Lou, Angew. Chem., Int. Ed.

    2020, 59, 7178.[11] S. H. Yang, S.-K. Park, J. K. Kim, Y. C. Kang, J. Mater. Chem. A 2019, 7, 13751.

    [12] Z. Li, K. Jiang, F. Khan, A. Goswami, J. Liu, A. Passian, T. Thundat, Sci. Adv. 2019, 5, eaav2820.

    [13] Y.  Yi, Z.  Sun, C.  Li, Z.  Tian, C.  Lu, Y.  Shao, J.  Li, J.  Sun, Z.  Liu, Adv. Funct. Mater. 2020, 30, 1903878.

    [14] S. H. Choi, Y. C. Kang, Nanoscale 2016, 8, 4209.[15] G. D. Park, J. H. Kim, S.-K. Park, Y. C. Kang, ACS Appl. Mater. Inter-

    faces 2017, 9, 10673.[16] E. Xu, Y. Zhang, H. Wang, Z. Zhu, J. Quan, Y. Chang, P. Li, D. Yu,

    Y. Jiang, Chem. Eng. J. 2020, 385, 123839.[17] M.  Yousaf, Y.  Wang, Y.  Chen, Z.  Wang, A.  Firdous, Z.  Ali,

    N. Mahmood, R. Zou, S. Guo, R. P. S. Han, Adv. Energy Mater. 2019, 9, 1900567.

    [18] P.  Ge, H.  Hou, C. E.  Banks, C. W.  Foster, S.  Li, Y.  Zhang, J.  He, C. Zhang, X. Ji, Energy Storage Mater. 2018, 12, 310.

    [19] H.-N. Fan, Q. Zhang, Q.-F. Gu, Y. Li, W.-B. Luo, H.-K. Liu, J. Mater. Chem. A 2019, 7, 13736.

    [20] J. Chen, A. Pan, Y. Wang, X. Cao, W. Zhang, X. Kong, Q. Su, J. Lin, G. Cao, S. Liang, Energy Storage Mater. 2019, 21, 97.

    [21] H. Wang, X. Lan, D. Jiang, Y. Zhang, H. Zhong, Z. Zhang, Y. Jiang, J. Power Sources 2015, 283, 187.

    [22] F.  Niu, J.  Yang, N.  Wang, D.  Zhang, W.  Fan, J.  Yang, Y.  Qian, Adv. Funct. Mater. 2017, 27, 1700522.

    [23] D.  Ma, Y.  Li, J.  Yang, H.  Mi, S.  Luo, L.  Deng, C.  Yan, P.  Zhang, Z. Lin, X. Ren, J. Li, H. Zhang, Nano Energy 2018, 43, 317.

    [24] J. Wang, B. Wang, B. Lu, Adv. Energy Mater. 2020, 10, 2000884.[25] Z.  Huang, A.  Alharbi, W.  Mayer, E.  Cuniberto, T.  Taniguchi,

    K. Watanabe, J. Shabani, D. Shahrjerdi, Nat. Commun. 2020, 11, 3029.[26] Y. Fang, X.-Y. Yu, X. W. Lou, Matter 2019, 1, 90.[27] G. Wang, C. Lu, X. Zhang, B. Wan, H. Liu, M. Xia, H. Gou, G. Xin,

    J. Lian, Y. Zhang, Nano Energy 2017, 36, 46.[28] Z. Liu, X. Zhang, D. Huang, B. Gao, C. Ni, L. Wang, Y. Ren, J. Wang,

    H. Gou, G. Wang, Chem. Eng. J. 2020, 379, 122418.[29] S.-T. Myung, M. Kikuchi, C. S. Yoon, H. Yashiro, S.-J. Kim, Y.-K. Sun,

    B. Scrosati, Energy Environ. Sci. 2013, 6, 2609.[30] H.-E.  Wang, K.  Yin, N.  Qin, X.  Zhao, F.-J.  Xia, Z.-Y.  Hu, G.  Guo,

    G. Cao, W. Zhang, J. Mater. Chem. A 2019, 7, 10346.[31] K. Zou, P. Cai, B. Wang, C. Liu, J. Li, T. Qiu, G. Zou, H. Hou, X. Ji,

    Nano-Micro Lett. 2020, 12, 121.[32] F.  Zuo, L.  Wang, T.  Wu, Z.  Zhang, D.  Borchardt, P.  Feng,

    J. Am. Chem. Soc. 2010, 132, 11856.[33] M. Long, Y. Qin, C. Chen, X. Guo, B. Tan, W. Cai, J. Phys. Chem. C

    2013, 117, 16734.[34] O.  Marciano, S.  Gonen, N.  Levy, E.  Teblum, R.  Yemini,

    G. D. Nessim, S. Ruthstein, L. Elbaz, Langmuir 2016, 32, 11672.[35] S. Y. Jeong, S.-K. Park, Y. C. Kang, J. S. Cho, Chem. Eng. J. 2018, 351,

    559.

    [36] B.  Li, Y.  Liu, Y.  Li, S.  Jiao, S.  Zeng, L.  Shi, G.  Zhang, ACS Appl. Mater. Interfaces 2020, 12, 2390.

    [37] Q.  Su, X.  Cao, T.  Yu, X.  Kong, Y.  Wang, J.  Chen, J.  Lin, X.  Xie, S. Liang, A. Pan, J. Mater. Chem. A 2019, 7, 22871.

    [38] F.  Zheng, W.  Zhong, Q.  Deng, Q.  Pan, X.  Ou, Y.  Liu, X.  Xiong, C. Yang, Y. Chen, M. Liu, Chem. Eng. J. 2019, 357, 226.

    [39] X.  Zhao, W.  Cai, Y.  Yang, X.  Song, Z.  Neale, H.-E.  Wang, J.  Sui, G. Cao, Nano Energy 2018, 47, 224.

    [40] X. Zhao, Y. Zhao, Z. Liu, Y. Yang, J. Sui, H.-E. Wang, W. Cai, G. Cao, Chem. Eng. J. 2018, 354, 1164.

    [41] Y. Liu, N. Wang, X. Zhao, Z. Fang, X. Zhang, Y. Liu, Z. Bai, S. Dou, G. Yu, J. Mater. Chem. A 2020, 8, 2843.

    [42] Y. Wang, Y. Wang, W. Kang, D. Cao, C. Li, D. Cao, Z. Kang, D. Sun, R. Wang, Y. Cao, Adv. Sci. 2019, 6, 1801222.

    [43] X.  Zhao, Y.  Zhao, B.  Huang, W.  Cai, J.  Sui, Z.  Yang, H.-E.  Wang, Chem. Eng. J. 2020, 382, 123047.

    [44] V.  Augustyn, J.  Come, M. A.  Lowe, J. W.  Kim, P.-L.  Taberna, S. H. Tolbert, H. D. Abruña, P. Simon, B. Dunn, Nat. Mater. 2013, 12, 518.

    [45] S.  Guo, S.  Liang, B.  Zhang, G.  Fang, D.  Ma, J.  Zhou, ACS Nano 2019, 13, 13456.

    [46] W.  Zhang, S.  Liang, G.  Fang, Y.  Yang, J.  Zhou, Nano-Micro Lett. 2019, 11, 69.

    [47] Y.  Li, H.  Wang, L.  Wang, R.  Wang, B.  He, Y.  Gong, X.  Hu, Energy Storage Mater. 2019, 23, 95.

    [48] P. Cai, K. Zou, G. Zou, H. Hou, X. Ji, Nanoscale 2020, 12, 3677.[49] Y. Fang, X.-Y. Yu, X. W. Lou, Adv. Mater. 2018, 30, 1706668.[50] Y. Fang, B. Y. Guan, D. Luan, X. W. Lou, Angew. Chem., Int. Ed. 2019,

    58, 7739.[51] Y. Fang, X.-Y. Yu, X. W. Lou, Angew. Chem., Int. Ed. 2019, 58, 7744.[52] Y. Fang, D. Luan, Y. Chen, S. Gao, X. W. Lou, Angew. Chem., Int. Ed.

    2020, 59, 2644.[53] H. Hou, C. E. Banks, M. Jing, Y. Zhang, X. Ji, Adv. Mater. 2015, 27, 7861.[54] D.  Cao, W.  Kang, S.  Wang, Y.  Wang, K.  Sun, L.  Yang, X.  Zhou,

    D. Sun, Y. Cao, J. Mater. Chem. A 2019, 7, 8268.[55] G.  Lucovsky, A.  Mooradian, W.  Taylor, G. B.  Wright, R. C.  Keezer,

    Solid State Commun. 1967, 5, 113.[56] C. Li, X. Zhang, K. Wang, X. Sun, Y. Ma, J. Power Sources 2018, 400,

    468.[57] S.  Li, W.  He, B.  Liu, J.  Cui, X.  Wang, D.-L.  Peng, B.  Liu, B.  Qu,

    Energy Storage Mater. 2020, 25, 636.[58] L.  Wang, G.  Yang, S.  Peng, J.  Wang, W.  Yan, S.  Ramakrishna,

    Energy Storage Mater. 2020, 25, 443.[59] J. Tang, X. Huang, T. Lin, T. Qiu, H. Huang, X. Zhu, Q. Gu, B. Luo,

    L. Wang, Energy Storage Mater. 2020, 26, 550.[60] H.-J.  Kim, H. V.  Ramasamy, G.-H.  Jeong, V.  Aravindan, Y.-S.  Lee,

    ACS Appl. Mater. Interfaces 2020, 12, 10268.[61] J.  Gao, Y.  Li, Y.  Liu, S.  Jiao, J.  Li, G.  Wang, S.  Zeng, G.  Zhang,

    J. Mater. Chem. A 2019, 7, 18828.[62] J. Qin, H. M. K. Sari, X. Wang, H. Yang, J. Zhang, X. Li, Nano Energy

    2020, 71, 104594.[63] A. R. Oganov, C. W. Glass, J. Chem. Phys. 2006, 124, 244704.[64] C. W.  Glass, A. R.  Oganov, N.  Hansen, Comput. Phys. Commun.

    2006, 175, 713.[65] Y. Wang, J. Lv, L. Zhu, Y. Ma, Comput. Phys. Commun. 2012, 183, 2063.[66] Y. Wang, J. Lv, L. Zhu, Y. Ma, Phys. Rev. B 2010, 82, 094116.[67] G. Kresse, J. Furthmüller, Comput. Mater. Sci. 1996, 6, 15.[68] G. Kresse, J. Hafner, Phys. Rev. B 1994, 49, 14251.[69] G. Kresse, J. Furthmüller, Phys. Rev. B 1996, 54, 11169.[70] G. Kresse, D. Joubert, Phys. Rev. B 1999, 59, 1758.[71] P. E. Blöchl, Phys. Rev. B 1994, 50, 17953.[72] J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1996, 77, 3865.[73] H. J. Monkhorst, J. D. Pack, Phys. Rev. B 1976, 13, 5188.[74] X.  Zhao, L.  Yin, T.  Zhang, M.  Zhang, Z.  Fang, C.  Wang, Y.  Wei,

    G. Chen, D. Zhang, Z. Sun, F. Li, Nano Energy 2018, 49, 137.

    Small 2020, 2004457