Na3V2O2(PO4)2F

sustainable and cost-effective grid-scale energy storage systems. Li-ion batteries (LIBs) have ... (the capacity ratio of negative/positive (N/P) elec...
2 downloads 4 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

Energy, Environmental, and Catalysis Applications

Improving the Performance of Hard Carbon//Na3V2O2(PO4)2F Sodiumion Full-cells by Utilizing the Adsorption Process of Hard Carbon Bolei Shen, Ya You, Yubin Niu, Yi Li, Chunlong Dai, Linyu Hu, Bingshu Guo, Jian Jiang, Shu-Juan Bao, and Maowen Xu ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b03986 • Publication Date (Web): 27 Apr 2018 Downloaded from http://pubs.acs.org on April 27, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Improving

the

Performance

of

Hard

Carbon//Na3V2O2(PO4)2F Sodium-ion Full-cells by Utilizing the Adsorption Process of Hard Carbon Bolei Shen,† Ya You *,‡ Yubin Niu, † Yi Li, † Chunlong Dai, †,‡ Linyu Hu, †,‡ Bingshu Guo,† Jian Jiang, † Shujuan Bao, † Maowen Xu *,† † Institute for Clean Energy & Advanced Materials, Southwest University, Chongqing 400715, P. R. China ‡ Materials Science and Engineering Program and Texas Materials Institute, The University of Texas at Austin, Austin, Texas 78712, United States * E-mail: [email protected] and [email protected]. ABSTRACT: Hard carbon has been regarded as promising anode materials for Na-ion batteries. Here we show, for the first time, the effects of two Na+ uptake/release routes, i.e., adsorption and intercalation process, on the electrochemical performance of half and full sodium batteries. Various Na+ storage processes are isolated in full-cells by controlling the capacity ratio of anode/cathode and the sodiation state of hard carbon anode. Full-cells utilizing adsorption region of hard carbon anode show better cycling stability and high-rate capability compare with those utilizing intercalation region of hard carbon anode. On the other hand, the intercalation region promises a high working voltage full cell because of the low Na+ intercalation potential. We believe this work is enlightening for the further practical application of hard carbon anode.

1 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

KEYWORDS: sodium-ion battery; sodium-ion full battery; hard carbon; adsorption mechanism; insertion mechanism

1. INTRODUCTION The worldwide demand for renewable and green energy sources requires the development of sustainable and cost-effective grid-scale energy storage systems. Li-ion batteries (LIBs) have been widely used in portable electronic devices and are expanding to the field of electronic vehicles.1 However, the resource availability and cost of lithium are becoming the major concern on its large-scale utilizations.2 Sodium-ion batteries (SIBs) show great promise as sustainable alternatives for LIBs due to its cost and resource advantage, which can meet the price and sustainability requirements of large-scale electrical energy storage. For the past several years, great efforts have been made in the search for suitable electrode materials.3-16 Although cathode materials have been extensively investigated in recent years, anode material is still an obstacle to the development of SIBs, because the commonly used anode of LIBs (such as graphite) is not suitable for SIBs on account of the larger ionic radius and orbital size of Na+. Among the various candidates reported,17-25 hard carbon20-25 (HC) has been regarded as one of the most promising anode materials for SIBs due to its high capacity, low operating potential, low price, rich resources, and simple synthesis. Nevertheless, some intrinsic disadvantages of HC anodes impede the practical use. On one hand, HC shows a low initial Coulombic efficiency, which consumes Na+ ions from the cathode to form solid electrolyte interface (SEI) layer and results in a severe capacity loss in full-batteries. On the other hand, the Na+ storage process of HC is more complicated compared with Li+ storage in graphite. The charge/discharge curves of HC anodes can be divided into two regions, a sloping voltage region

2 Environment ACS Paragon Plus

Page 2 of 20

Page 3 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

( > 0.2 V) with hysteresis between charge/discharge profiles and a low-voltage plateau ( < 0.2 V) region. Cao’s group demonstrates an “adsorption–insertion” mechanism for Na+ ions storage in HC materials. The sloping potential region is assigned to Na+ ions adsorption on surface active sites and low potential plateau is attributed to the insertion of Na+ ions into graphite-like layers to form NaCx compounds.26 However, so far there have been few studies into the effect of each Na storage mechanism on the thermodynamic stability of the HC anode vs. Na and further, the practical electrochemical performance of a rechargeable Na full cell. In this work, for the first time, we reveal the influence of the above two Na+ uptake/release routes on the electrochemical properties of the HC anode both in sodium-ion half and full cells. To isolate the two Na+ storage process from each other, an excess amount of HC was employed (the capacity ratio of negative/positive (N/P) electrodes is ~3) in the full-cells and the sodiation state of HC anode was adjusted. It is found that Na+ adsorption process ( > 0.2 V vs. Na+/Na) in HC is thermodynamic stable, which ensures a good cycling stability and rate performance; while the intercalation process promises a high working voltage in full-cells because of the low Na+ intercalation potential. This work demonstrates the feasibility of selecting suitable HC anodes for high performance practical SIBs.

2. EXPERIMENTAL SECTION Materials preparation All chemical reagents were analytical grade and used without further purification. HC anode was obtained from Kureha company and used as received. The Na3V2O2(PO4)2F/graphene cathode was prepared via a hydrothermal method as described in our previously reported work.27 Graphene oxide (GO) was prepared by a modified Hummer’s method. 40 mg GO was dispersed in 50 mL N,N-dimethylformamide (DMF) by sonication for 15 min. 1 mmol NH4VO3, 1.5 mmol NaF and 1 mmol (NH4)2HPO4 was dissolved into 20 mL 3 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

deionized water and the obtained solution was slowly added into the above-mentioned GO suspension. The mixture was stirred at 80 oC for 4 hours to ensure GO is homogeneously dispersed. Afterwards, the suspension was transferred into a 100 mL Teflon autoclaves and heated at 180 oC for 12 hours. The obtained product was washed by deionized water and dehydrated ethanol and dried at 120 oC. Materials characterization Powder X-ray diffraction (XRD) data were collected on a MAXima-X XRD-7000 diffractometer equipped with filtered Cu Kα radiation (λ = 1.5416 Å) to identify the crystalline phase of the material. The morphologies of the samples were observed with a field-emission scanning electron microscopy (SEM; JEOL6300F). Electrochemical measurements The electrochemical performance of cathode and anode is tested in half cells. For half cells, the cathode electrode was prepared by mixing 80 wt% Na3V2O2(PO4)2F/rGO active materials, 10 wt% carbon black, and 10 wt% polyvinylidenefluoride (PVDF) suspended in N-methyl-2-pyrrolidone (NMP) solvent. The obtained slurry was casted onto aluminum foil and dried in a vacuum oven overnight at 120 °C. The resulting electrode has an average mass loading of 2 mg cm-2. The anode electrode was prepared in the same way as the cathode electrode with commercial hard carbon as active materials. The average mass loading of anode electrode is 1.5 mg cm-2. The mass loading of the cathode electrode in full-cells is the same with that in half cells. The capacity ratio of negative/positive (N/P) electrodes is set to be ~3 for adsorption and intercalation full cells. The average mass loading of anode electrode is 2.3 mg cm-2.

4 Environment ACS Paragon Plus

Page 4 of 20

Page 5 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The as-prepared anode and cathode electrodes were assembled into CR2032-type coin cells in a glove box under high-purity argon atmosphere (H2O and O2 < 1 ppm). For all half-cells, a sodium foil was used as the counter electrode and Celgard 2400 was used as separators. The electrolyte was 1 M NaClO4 in a mixture of ethylene carbonate (EC) and diethyl carbonate (DEC) (1 : 1 in volume) with 5 wt% fluoroethylene carbonate (FEC) as an additive. The Na//Na3V2O2(PO4)2F/rGO half-cells were tested within 2.0 − 4.2 V. Na//Hard carbon half-cells were first cycled between 0.001 and 3 V for 5 cycles, then cycled between 0.2−3 V and 0.2−0.001 V for different Na+ storage process. For full-cells, pre-sodiated hard carbon was used as the anode and Na3V2O2(PO4)2F/rGO was used as the cathode. The HC anodes were first electrochemically passivated in half-cells and then discharged to a desired voltage (~1.2 V for adsorption full cells and ~0.2 V for insertion full cells). Then these batteries were disassembled in the glove box to obtain anode electrode at different sodiated status. The as-prepared full-cells were tested within the potential range of 2.0−4.2 V at rates of 0.5C, 1C, 2C and 3C (1C = 120 mA g-1). The specific capacity of full cells is calculated based on the mass of the positive electrode material. The galvanostatic charge and discharge measurements were carried out on a LAND battery test system (Wuhan Land Electronics Co., Ltd. China). The cyclic voltammetry (CV) measurements were carried out on an Arbin instrument battery testing system at a scan rate of 0.1 mV s-1.

3. RESULTS AND DISCUSSION A high-voltage Na3V2O2(PO4)2F/rGO cathode material was synthesized by a hydrothermal method as reported in our previous work.27 The crystal structure of the assynthesized material was determined by XRD and all the Bragg diffraction peaks can be 5 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

well indexed to Na3V2O2(PO4)2F (PDF#97-041-1950; space group: I4/mmm (139)) with no additional peaks from impurities were detected (Figure S1a). SEM images show that monodispersed Na3V2O2(PO4)2F cubes with average sizes of ~ 400 nm were uniformly distributed among rGO substrate (Figure S1b). XRD patterns of HC (Figure 1a) presented two broad diffraction peaks at 23.42o and 44o, which corresponds to (002) and (100) planes in a disordered carbon structure.28 The interlayer spacing of (002) plane was calculated to be ~0.39 nm. Such a large interlayer spacing is beneficial for Na+ insertion/extraction due to its large ionic size. HC appears to have a particle size of 1−5 µm with irregular shape (Figure 1b).

Figure 1. Morphological and structural characterization of the hard carbon anode. (a) XRD patterns and (b) SEM images of the commercial hard carbon.

Figure S2a shows the galvanostatic charge-discharge (GCD) curves of the Na3V2O2(PO4)2F/rGO cathode cycled at 0.5C in a voltage range of 2−4.2 V. The battery exhibits two charge/discharge voltage plateaus with a reversible capacity of 120 mA h g-1. CV curves (inset of Figure S2a) show two pairs redox peaks, which are consistent with the GCD profiles. Na3V2O2(PO4)2F/rGO cathode also exhibits impressive rate capabilities. The capacity values are 121, 118, 114, 106 and 96 mA h g-1, respectively, at constant

6 Environment ACS Paragon Plus

Page 6 of 20

Page 7 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

rates of 0.5, 1, 2, 5 and 10 C (Figure S2b), indicating a good reversibility of the Na3V2O2(PO4)2F/rGO within a wide current range. At 0.5C the discharge capacity is 111 mA h g-1 after 50 cycles, demonstrating a good cycling stability with 95% capacity retention (Figure S2c). Figure 2a shows the GCD potential profiles of the HC anode at a current density of 50 mA g-1 between 0.001−3 V. HC anode exhibits an initial discharge and charge capacity of 438 mA h g-1 and 332 mA h g-1, respectively, corresponding to a Coulombic efficiency of 75%. The irreversible specific capacity is attributed to the decomposition of electrolyte and SEI formation in the first discharge process. Once a stable SEI is formed, the HC anode exhibits a reversible capacity of ~320 mA h g-1 in the following cycles. The rate performance of the hard carbon was displayed in Figure 2b. The reversible capacities are 325, 210, 134, 111, 97, 94 mA h g-1 at current densities of 50, 100, 200, 300, 400, 500 mA g-1, respectively. At a current density of 50 mA g-1, the HC anode exhibits 80% capacity retention over 170 cycles (Figure 2c). The SEM images and XRD patterns were collected on the cycled HC electrodes and the results are included in Figure S3. It is observed from Figure S3a that the cycled HC anode still preserves the pristine morphology without apparent particle cracking and pulverization. According to the XRD results (Figure S3b), the (002) peak shifts to a lower angle after 150 cycles, indicating that the interlayer spacing of (002) plane expands during the substantial Na+ insertion/extraction.

7 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Electrochemical performance of the HC anode. (a) GCD profiles of the HC anode at a current density of 50 mA g-1. The inset shows the corresponding CV curves at a scan rate of 0.1 mV s-1. (b) Rate capabilities and (c) cycling performance of the HC anode.

The GCD potential profiles of HC anode can be divided into two distinct regions; a slope region within 0.2−1.2 V and a voltage plateau region below 0.2 V (Figure 3a). The sloping region between 0.2 and 1.2 V was caused by Na+ ions adsorption at surface or active cites due to a wide distribution of adsorption energies and the plateau region ( < 0.2 V) is dominated by Na+ insertion between parallel or nearly parallel graphene layers.26 CV curves of HC half cells (inset of Figure 2a) show a pair of sharp redox peaks around 0 −0.2

V vs. Na+/Na and broad sodiation/desodiation peaks around 0.2−1.2 V, which are

associated with insertion and adsorption process, respectively, well consistent with the GCD profiles. To illustrate the effects of each Na+ uptake/release routes on electrochemical performance, HC anode was galvanostatically cycled within various voltage domains. The HC//Na half-cells were first cycled between 0.001 and 3 V for 5 cycles at 50 mA g-1 to fully activate HC and form a stable SEI, then cycled between 0.2− 3.0 V (adsorption region) and 0.001−0.2 V (intercalation region), respectively (Figure S4). HC shows a capacity of 104 mA h g-1 and 211 mA h g-1 in adsorption and intercalation region, respectively. As seen from Figure 3b, HC anode cycled within the adsorption voltage region demonstrates good cycling stability with a capacity retention 90% after 180 cycles, while a 67% capacity retention over 70 cycles is observed if Na+ storage is achieved through a insertion/extraction process. Figure 3c shows the rate capability of HC anodes cycled within various voltage regions. For adsorption process, the capacity retentions are 97%, 82%, 57% and 46% at current densities of 50, 100, 300, 500 mA g-1,

8 Environment ACS Paragon Plus

Page 8 of 20

Page 9 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

respectively (The specific capacity vs. cycle number chart is displayed in Figure S5). The capacity of HC declines to zero at 300 mA g-1 when it is cycled within the intercalation voltage region. Based on the results above, intercalation region promises high capacity and low working potential, while the adsorption region provides high cycling stability and high-rate capability.

Figure 3. The schematic illustration and electrochemical performance of the HC anode within various voltage regions in HC//Na half cells. (a) Schematic illustration of the adsorptionintercalation mechanisms for Na+ storage in HC. (b) Cycling performance and (c) rate capability of HC cycled within adsorption, intercalation and whole voltage regions. For capacity retention at various rates, the initial capacity at 50 mA g-1 is set to be 100%.

Na-ion full cells were assembled using the above mentioned Na3V2O2(PO4)2F/rGO positive electrodes and HC negative electrodes to further reveal the influence of two Na+ storage routes in HC on full-cells. All the HC anodes were cycled in half-cells first to passivate the surface and form a stable SEI. Without the passivation process, the assembled full-cell shows a low initial Coulombic efficiency (46%) and a low reversible

9 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

capacity (60 mA h g-1) resulting from the formation of SEI in the HC anode in the initial charge process (Figure S6). However, the electrochemical performance is notably improved if the HC anode is passivated first in half-cells (Figure S7). The schematic illustration of the pre-sodiated method is displayed in Figure 4a and 4b. To isolate two different Na+ storage processes, an excess amount of HC anode is employed with the capacity ratio of negative/positive (N/P) to be ~3. The soidation state of HC was adjusted as well. HC//Na half-cells were first cycled between 0.001 and 3 V for 5 cycles, then discharged to a desired voltage (1.2 V for adsorption process and 0.2 V for intercalation process). The schematic illustrations of the first charge process for adsorption and intercalation full battery are shown in Figure 4c and 4d, respectively. The discharge schematic illustrations are displayed in Figure S8. When HC is discharged to 1.2 V, both the adsorption and intercalation region are available to accommodate Na+. During the first charge process of full-cell, Na+ ions are extracted from cathode and inserted into the HC anode. Because of the excess anode (N/P is 3), Na+ only fill the adsorption region (Figure 4c). On the contrary, Na+ will occupy the adsorption region if the HC anode is discharged to 0.2 V. As a result, Na+ ions can only be inserted into the intercalation region of HC due to the filled adsorption region (Figure 4d).

10 Environment ACS Paragon Plus

Page 10 of 20

Page 11 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 4. Schematic illustration of the working principle of the adsorption and intercalation full cells. (a) The schematic illustration of the passivation technology. (b) The internal state of the pre-sodiated HC. (c) The schematic illustration of the first charging process of the adsorption full cell. (d) The schematic illustration of the first charging process of intercalation full cell.

The GCD curves of various HC//Na3V2O2(PO4)2F/rGO full-cells, which utilize different voltage domains of HC anode (adsorption and intercalation regions), are shown in Figure 5a and 5b. All the full-cells were cycled at 0.5C in a voltage range of 2.0−4.2 V. A capacity close to 120 mA h g-1 was observed for all cells no matter which working region of HC is utilized. However, the working voltages of these full batteries are quite different. The full batteries based on Na+ intercalation mechanism in HC anode (intercalation full-cells) show a high average working voltage (3.1 V), which represents the highest value for reported full-cells (Figure S9)8,29-37. However, the average working voltage of full battery based on Na+ adsorption mechanism (adsorption full-cells) is much lower than that of the intercalation full-cells. Figure 5c displays the cycling performance of these full batteries. Upon cycling at 0.5C for 70 cycles, adsorption full-cell

11 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

demonstrate a capacity retention of 94%, while the capacity retention of intercalation fullcells is 84%. Due to the poor reversibility of intercalation region for HC, the accommodation of Na+ are expected to gradually change from intercalation to adsorption region with the cycling going on, which can be proved by the significantly decreased average output voltage profiles of intercalation full battery after 60 cycles (Figure S10). Figure 5d shows the rate capability of these full-cells under various current densities (0.5, 1, 2, and 3C). For adsorption full battery, the capacities are 121 mA h g-1 at 0.5C and 109 mA h g-1 at 2C. Even at a rate of 3C, the discharge capacity remains at 99 mA h g-1, approximately 81% of the reversible capacity at 0.5C, while intercalation full-cell only delivers a discharge capacity of 8.1 mA h g-1 at 3C. All the above results show the adsorption full-cells exhibit good cycling stability and rate capability while intercalation full-cells ensures high working voltage, which are consistent with Figure 3b and 3c.

12 Environment ACS Paragon Plus

Page 12 of 20

Page 13 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 5. Electrochemical performance of the adsorption and intercalation full cells. (a) Galvanostatic charge profiles and (b) galvanostatic discharge profiles at 0.5C between 2.0 and 4.2 V. (c) Cycling performance at 0.5C. (d) Rate capability.

4. CONCLUSIONS In summary, the effects of two different Na+ storage process of hard carbon, i.e., adsorption and intercalation routes, on electrochemical performance of Na-ion full-cells were investigated for the first time. The adsorption process (> 0.2 V vs. Na+/Na) in HC is found to be thermodynamic stable, which ensures a good cycling stability and rate performance; while the intercalation process ( < 0.2 V vs. Na+/Na) shows slow reaction kinetics, leading to an unsatisfactory cycling stability and high-rate capability. This result shows that excessive hard carbon may lead to a better cycling stability in the industrialization of SIFBs. However, because of the low Na+ intercalation potential and high capacity, the intercalation process promises a high working voltage full-cell. It is expected that this strategy can be applied to other cathode materials (e.g., NASICON-type cathodes and layered transition-metal oxides) and the finding in this work will provide new avenues for designing advanced high-energy and long-life room temperature SIBs.



ASSOCIATED CONTENT

Supporting Information

13 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 20

Morphological and structural characterization of the Na3V2O2(PO4)2F/rGO; Electrochemical properties of the Na3V2O2(PO4)2F/rGO; The morphological and structural characterization of the cycled hard carbon; The galvanostatic charge/discharge (GCD) profiles of different regions of hard carbon; The rate performance of adsorption, intercalation and whole region of hard carbon; Schematic illustration of the working principle of adsorption and intercalation full batteries; Comparison of the working potential and specific capacity of full-cells in this work and other reported works; The discharge voltage profiles of intercalation full battery at various cycles; The electrochemical properties of whole region full-cells; The electrochemical performance of nonpre-sodiated full batteries. 

AUTHOR INFORMATION

Corresponding Author * E-mail: [email protected] and [email protected]. Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes The authors declare no competing financial interest. 

ACKNOWLEDGMENTS

This work is financially supported by grants from the National Natural Science Foundation of China (No. 21773188), Basic and frontier research project of Chongqing (cstc2015jcyjA50031) and

Fundamental

Research

Funds

for

the

Central

Universities

(XDJK2017A002,

XDJK2017B048) and Program for Innovation Team Building at Institutions of Higher Education in Chongqing (CXTDX201601011).

14 Environment ACS Paragon Plus

Page 15 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

 REFERENCES (1) Wang, Y.; Feng, Z.; Yang, S.-Z.; Gagnon, C.; Gariépy, V.; Laul, D.; Zhu, W.; Veillette, R.; Trudeau, M.-L.; Guerfi, A.; Zaghib, K., Layered oxides-LiNi1/3Co1/3Mn1/3O2 as anode electrode for symmetric rechargeable lithium-ion batteries. J. Power Sources. 2018, 378, 516−521. (2) Komaba, S.; Murata, W.; Ishikawa, T.; Yabuuchi, N.; Ozeki, T.; Nakayama, T.; Ogata, A.; Gotoh, K.; Fujiwara, K., Electrochemical Na Insertion and Solid Electrolyte Interphase for Hard Carbon Electrodes and Application to Na-Ion Batteries. Adv. Funct. Mater. 2011, 21, 3859−3867. (3) Berthelot, R.; Carlier, D.; Delmas, C., Electrochemical investigation of the P2NaxCoO2 phase diagram. Nature Mater. 2011,10, 74−80. (4) Didier, C.; Guignard, M.; Denage, C.; Szajwaj, O.; Ito, S.; Saadoune, I.; Darriet, J.; Delmas, C., Electrochemical Na-Deintercalation from NaVO2. Electrochem. Solid ST. 2011, 14, A75−A78. (5) Wang, Y.; Liu, J.; Lee, B.; Qiao, R.; Yang, Z.; Xu, S.; Yu, X.; Gu, L.; Hu, Y.-S.; Yang, W.; Kang, K.; Li, H.; Yang, X.-Q.; Chen, L.; Huang. X., Ti-substituted tunnel-type Na0.44MnO2 oxide as a negative electrode for aqueous sodium-ion batteries. Nat. Commun. 2015, 6, 6401. (6) Zhu, C.; Wu, C.; Chen, C.-C.; Kopold, P.; Aken, P.-A.; Maier, J.; Yu Y., A High Power-High Energy Na3V2(PO4)2F3 Sodium Cathode: Investigation of Transport Parameters, Rational Design and Realization. Chem. Mater. 2017, 29, 5207−5215. (7) Guo, S.; Yu, H.; Liu, P.; Ren, Y.; Zhang, T.; Chen, M.; Ishidab, M.; Zhou, H., New Insight into Structural Evolution in Layered NaCrO2 during Electrochemical Sodium Extraction. Energy Environ. Sci. 2015, 8, 1237−1244.

15 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(8) Wang, H.; Xiao, Y.; Sun, C.; Lai, C.; Ai, X., A type of sodium-ion full-cell with a layered NaNi0.5Ti0.5O2 cathode and a pre-sodiated hard carbon anode. RSC Adv. 2015, 5, 106519−106522. (9) Heubner, C.; Heiden, S.; Schneider, M.; Michaelis, A., In-situ preparation and electrochemical characterization of submicron sized NaFePO4 cathode material for sodium-ion batteries. Electrochim. Acta. 2017, 233, 78−84. (10) Shen, B.; Xu, M.; Niu, Y.; Han, J.; Lu, S.; Jiang, J.; Li, Y.; Dai, C.; Hu, L.; Li, C.-M., Sodium-Rich Ferric Pyrophosphate Cathode for Stationary Room-Temperature SodiumIon Batteries. ACS Appl. Mater. Interfaces 2018, 10, 502−508. (11) Niu, Y.; Xu, M.; Shen, B.; Dai, C.; Li, C.-M., Exploration of Na7Fe4.5(P2O7)4 as a cathode material for sodium-ion batteries. J. Mater. Chem. A. 2016, 4, 16531−16535. (12) Barpanda, P.; Liu, G.; Avdeev, M.; Yamada, A., t-Na2(VO)P2O7: A 3.8 V Pyrophosphate Insertion Material for Sodium-Ion Batteries. ChemElectroChem 2016, 1, 1488−1491. (13) Kim, H.; Park, I.; Lee, S.; Kim, H.; Park, K.-Y.; Park, Y.-U.; Kim, H.; Kim, J.; Lim, H.-D.; Yoon, W.-S.; Kang, K., Understanding the electrochemical mechanism of the new iron-based mixed-phosphate Na4Fe3(PO4)2(P2O7) in a Na rechargeable battery. Chem. Mater. 2013, 25, 3614 −3622.

(14) Li, S.; Guo, J.; Ye, Z.; Zhao, X.; Wu, S.; Mi, J. X.; Wang, C. Z.; Gong, Z.; McDonald, M.-J.; Zhu, Z.; Ho, K.-M.; Yang. Y., Zero-Strain Na2FeSiO4 as Novel Cathode Material for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 17233− 17238. (15) Deng, C.; Zhang, S., Nanostructured Na7V4(P2O7)4(PO4) as high-potential and superior-

16 Environment ACS Paragon Plus

Page 16 of 20

Page 17 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

performance cathode material for sodium-ion batteries. ACS Appl. Mater. Interfaces 2014, 6, 9111−9117. (16) Kim, H.; Yoon, G.; Park, I.; Park, K.-Y.; Lee, B.; Kim, J.; Park, Y.-U.; Jung, S.-K.; Lim, H.-D., Anomalous Jahn-Teller behavior in a manganese-based mixed-phosphate cathode for sodium ion batteries. Energy Environ. Sci. 2015, 8, 3325−3335. (17) Sun, Y.; Zhao, L.; Pan, H.; Lu, X.; Gu, L.; Hu, Y.-S.; Li, H.; Armand, M.; Ikuhara, Y.; Chen, L.; Huang, X., Direct atomic-scale confirmation of three-phase storage mechanism in Li4Ti5O12 anodes for room-temperature sodium-ion batteries. Nat. Commun. 2013, 4, 1870. (18) Niu, Y.; Xu, M.; Zhang, Y.; Han, J.; Wang, Y.; Li, C.-M., Detailed investigation of a NaTi2(PO4)3 anode prepared by pyro-synthesis for Na-ion batteries. RSC Adv. 2016, 6, 45605− 45611. (19) Yu, D.-Y.-W.; Prikhodchenko, P.-V.; Mason, C.-W.; Batabyal, S.-K.; Gun, J.; Sladkevich, S.; Medvedev, A.-G.; Lev, O., High-capacity antimony sulphide nanoparticle-decorated graphene composite as anode for sodium-ion batteries. Nat. Commun. 2013, 4, 2922. (20) Stevens, D.-A.; Dahn, J.-R., High capacity anode materials for rechargeable sodium-ion batteries. J. Electrochem. Soc. 2000, 147, 1271−1273. (21) Komaba, S.; Murata, W.; Ishikawa, T.; Yabuuchi, N.; Ozeki, T.; Nakayama, T.; Ogata, A.; Gotoh, K.; Fujiwara, K., Electrochemical Na insertion and solid electrolyte interphase for hard‐ carbon electrodes and application to Na-Ion batteries. Adv. Funct. Mater. 2011, 21, 3859− 3867. (22) Zhang, B.; Ghimbeu, C.-M.; Laberty, C.; Guterl C.-V.; Tarascon, J., Correlation Between

17 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Microstructure and Na Storage Behavior in Hard Carbon. Adv. Energy Mater. 2016, 6, 1501588. (23) Wang, Q.; Zhu, X.; Liu, Y.; Fang, Y.; Zhou, X.; Bao J., Rice husk-derived hard carbons as high-performance anode materials for sodium-ion batteries. Carbon. 2018, 127, 658−666. (24) Wang, P.; Zhu, X.; Wang, Q.; Xu, X.; Zhou, X.; Bao, J., Kelp-derived hard carbons as advanced anode materials for sodium-ion batteries. J. Mater. Chem. A. 2017, 5, 5761−5769. (25) Wang, P.; Qiao, B.; Du, Y.; Li, Y.; Zhou, X.; Dai, Z.; Bao, J., Fluorine-Doped Carbon Particles Derived from Lotus Petioles as High-Performance Anode Materials for Sodium-Ion Batteries. J. Phys. Chem. C. 2015, 119, 21336−21344. (26) Qiu, S.; Xiao, L.; Sushko, M.-L.; Han, K.-S.; Shao, Y.; Yan, M.; Liang, X.; Mai, L.; Feng, J.; Cao, Y.; Ai, X.; Yang, H.; Liu, J., Manipulating Adsorption-Insertion Mechanisms in Nanostructured Carbon Materials for High-Efficiency Sodium Ion Storage. Adv. Energy Mater. 2017, 1700403. (27) Xu, M.; Wang, L.; Zhao, X.; Song, J.; Xie, H.; Lu, Y.; Goodenough, J.-B., Na3V2O2(PO4)2F/graphene sandwich structure for high-performance cathode of a sodium-ion battery. Phys. Chem. Chem. Phys. 2013, 15, 13032−13037. (28) Bommier, C.; Luo, W.; Gao, W.-Y.; Greaney, A.; Ma, S.; Jia, X., Predicting capacity of hard carbon anodes in sodium-ion batteries using porosity measurements. Carbon. 2014, 76, 165 −174.

(29) Li, H.; Peng, L.; Zhu, Y.; Chen, D.; Zhang, X.; Yu, G., An Advanced High-Energy Sodium Ion Full Battery Based on Nanostructured Na2Ti3O7/VOPO4 Layered Materials. Energy Environ. Sci., 2016, 9, 3399.

18 Environment ACS Paragon Plus

Page 18 of 20

Page 19 of 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(30) Wang, Y.; Yu, X.; Xu, S.; Bai, J.; Xiao, R.; Hu, Y.-S.; Li, H.; Yang, X.-Q.; Chen, L.; Huang, X., A zero-strain layered metal oxide as the negative electrode for long-life sodium-ion batteries. Nat. Commun. 2013, 4, 2365. (31) Ming, J.; Ming, H.; Yang, W.; Kwak, W.-J.; Park, J.-B.; Zheng, J.; Sun, Y.-K., A sustainable iron-based sodium ion battery of porous carbon-Fe3O4/Na2FeP2O7 with high performance. RSC Adv. 2015, 5, 8793−8800. (32) Lûpez, M.-C.; Aragûn, M.-J.; Ortiz, G.-F.; Lavela, P.; Alcantara, R.; Tirado, J.-L., High Performance Full Sodium-Ion Cell Based on a Nanostructured Transition Metal Oxide as Negative Electrode. Chem. Eur. J. 2015, 21, 14879−14885. (33) Xiong, H.; Slater, M.-D.; Balasubramanian, M.; Johnson, C.-S.; Rajh, T. Amorphous TiO2 nanotube anode for rechargeable sodium ion batteries. J. Phys. Chem. Lett. 2011, 2, 2560−2565. (34) Wan, F.; Guo, J.-Z.; Zhang, X.-H.; Zhang, J.-P.; Sun, H.-Z.; Yan, Q.; Han, D.-X.; Niu, L.; Wu, X.-L., In Situ Binding Sb Nanospheres on Graphene via Oxygen Bonds as Superior Anode for Ultrafast Sodium-Ion Batteries. ACS Appl. Mater. Interfaces. 2016, 8, 7790−7799. (35) Guo, J.-Z.; Yang, Y.; Liu, D.-S.; Wu, X.-L.; Hou, B.-H.; Pang, W.-L.; Huang, K.-C.; Zhang, J.-P.; Su, Z.-M., A Practicable Li/Na-Ion Hybrid Full Battery Assembled by a High-Voltage Cathode and Commercial Graphite Anode: Superior Energy Storage Performance and Working Mechanism. Adv. Energy Mater. 2018, 1702504. (36) Wang, Y.-Y.; Hou, B.-H.; Guo, J.-Z.; Ning, Q.-L.; Pang, W.-L.; Wang, J.; Lü, C.-L.; Wu, X.-L., An Ultralong Lifespan and Low-Temperature Workable Sodium-Ion Full

19 Environment ACS Paragon Plus

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Battery for Stationary Energy Storage. Adv. Energy Mater. 2018, 1703252. (37) Hou, B.-H.; Wang, Y.-Y.; Guo, J.-Z.; Zhang, Y.; Ning, Q.-L.; Yang, Y.; Li, W.-H.; Zhang, J.-P.; Wang, X.-L.; Wu, X.-L., A Scalable Strategy To Develop Advanced Anode for Sodium-Ion Batteries: Commercial Fe3O4-Derived Fe3O4@FeS with Superior FullCell Performance. ACS Appl. Mater. Interfaces. 2018, 10, 3581−3589.

Table of Contents

20 Environment ACS Paragon Plus

Page 20 of 20