Robust Pitaya-Structured Pyrite as High Energy Density Cathode for

Aug 16, 2017 - Cycling in the voltage region of 1.0–3.0 V at 0.3 A g–1, the current conversion-based FeS2@C cathode displays a high and stable ene...
11 downloads 18 Views 4MB Size
Robust Pitaya-Structured Pyrite as High Energy Density Cathode for High-Rate Lithium Batteries Xijun Xu,†,‡ Jun Liu,*,†,‡ Zhengbo Liu,†,‡ Jiadong Shen,†,‡ Renzong Hu,†,‡ Jiangwen Liu,†,‡ Liuzhang Ouyang,†,‡ Lei Zhang,§ and Min Zhu*,†,‡ †

Guangdong Provincial Key Laboratory of Advanced Energy Storage Materials, School of Materials Science and Engineering, South China University of Technology, Guangzhou, 510641, China ‡ China-Australia Joint Laboratory for Energy & Environmental Materials, South China University of Technology, Guangzhou, 510641, China § School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou, 510640, China S Supporting Information *

ABSTRACT: To solve the serious problems (the agglomeration of nano-Fe0, dissolution of polysulfide, and low electronic conductivity of Li2S) of earth-abundant pyrite (FeS2) cathodes for lithium batteries, a simple in situ encapsulation and transformation route has been successfully developed to synthesis pitaya-structured porous carbon embedded with FeS2 nanoparticles. Due to such a hierarchical architecture design, this cathode of pitaya-structured FeS2@C can effectively avoid the serious agglomeration and coarsening of small Fe nanoparticles, reduce the dissolution of polysulfide, and provide superior kinetics toward lithium storage, resulting in enhanced reversibility and rate capability. Cycling in the voltage region of 1.0−3.0 V at 0.3 A g−1, the current conversionbased FeS2@C cathode displays a high and stable energy density (about 1100 Wh kg−1), ultrahigh rate capability (a reversible capability of 660, 609, 554, 499, 449, and 400 mA h g−1 at 0.2, 0.5, 1.0, 2.0, 5.0, and 10 A g−1, respectively), and stable cycling performance. KEYWORDS: FeS2 cathode, biomimetic structure, Li-ion batteries, rate capability, conversion reaction

U

intercalation LiNixCoyMnzO2 cathodes (less than 300 mA h g−1).14 During the discharging process, highly reactive nano-Fe0 and highly resistive Li2S will be formed through the conversion reaction of FeS2. As a consequence of the insulating feature of Li2S, the serious agglomeration of nano-Fe0, and the dissolution of polysulfide converted from the in situ generated sulfur, a rapid capacity loss is normally observed.21 Modifications (e.g., microstructure engineering, carbon/conductive polymer coating) of electroactive FeS2 are necessary to improve the reversibility of this kind of cathode.22−31 Especially, fully embedding electrode particles with uniform small size in a porous conductive matrix not only promotes the facile access of Li+/e− but also supplies sufficient space to accommodate the accumulated strain resulting from Li+ insertion/extraction.22−24 In addition, such hybrid structures can effectively avoid the serious agglomeration and coarsening of small metal nanoparticles, resulting in enhanced reversibility and rate capability.23

ntil now, the energy and power densities of commercial Li-ion batteries (LIBs) with conventional graphite anodes and LiTMO2 (TM = transition metal) cathodes have approached practical upper limits.1−3 The past two decades have witnessed much progress in searching for various kinds of anode materials for LIBs,4−13 such as alloying-type Si, Ge, Sb, and Sn and conversion-type metal oxides.9−13 However, for the cathode candidates, much fewer achievements have been obtained, which may be the nonflexible structure control of such complex LiMO2 oxides.14 Generally speaking, suitable cathode materials for LIBs should combine the merits of high energy density, long cycling life, and low cost. Over the past decade, some kinds of low-cost Fe-based conversion cathode materials (e.g., the fluoride FeF3 and sulfide FeS2) have attracted researchers’ intensive attention, and such kinds of simple compounds are possible candidates to replace the typical LiMO2 intercalation cathodes because they possess high theoretical energy densities and low-cost feature.15−20 Especially, the earth-abundant pyrite FeS2 is approved with a high theoretical capacity of 894 mA h g−1, corresponding to the generation of metallic Fe and insulating Li2S (FeS2 + 4Li+ + 4e− → Fe + 2Li2S),17−20 which is much higher than that of very best © 2017 American Chemical Society

Received: May 20, 2017 Accepted: August 16, 2017 Published: August 16, 2017 9033

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

www.acsnano.org

Article

ACS Nano

into inorganic carbon via a simple annealing treatment, resulting in a uniform pitaya-structured Fe3O4@C precursor. Scanning electron microscopy (SEM) images of these Fe3O4@C particles (Figure 1b,c) clearly display that they have a uniform spherical shape and size (about 200 nm in average diameter). The structural feature of pitaya-shaped Fe3O4@C, i.e., tiny Fe3O4 nanoparticles fully embedded in a carbon sphere matrix, can be directly revealed by transmission electron microscopy (TEM) (Figure 1e). The pure cubic magnetite phase (JCPDS file no. 190629) of the inner capsulated oxide nanoparticles was also directly confirmed by X-ray diffractometry (XRD) analysis (Figure 1d). Hereafter, these uniform oxide-based nanospheres were used as a suitable template to form pitaya-structured particles by hydrothermal sulfuration. X-ray photoelectron spectroscopy (XPS) measurements were also performed to reveal the major component of the Fe3O4@C precursor (Figure S1, in the Supporting Information). As displayed in Figure S1a, the characteristic peaks of binding energies located at 284.7, 531.1, and 710.5 eV are associated with C 1s, O 1s, and Fe 2p, respectively. In particular, the peaks located at 710.5 and 724.5 eV shown in Figure S1b are attributed to the Fe 2p3/2 and Fe 2p1/2, respectively,25,26 and the ones at 529.8, 531.1, 531.8, and 532.9 eV shown in Figure S1c are related to Fe−O−Fe, Fe−O, Fe−O−C, and C−O, respectively.27,28 Figure S1d shows the C 1s XPS spectra of Fe3O4@C, in which the peak at 284.7 eV is related to C−C and CC and the ones at 286.0 and 288.9 eV are associated with C−O and O−CO, respectively.25,26 As clearly displayed in Figure 2a, the XRD pattern of the final FeS2@C spheres can be assigned to pyrite FeS2 (JCPDS no. 421340), without any peaks from marcasite, greigite FeS2, sulfur, or other impurities. The broadening intensity of the diffraction peaks in Figure 2a indicates the small crystallite size of these embedded FeS2 nanoparticles, which normally benefits fast Li+ diffusion, local strain release, and high utilization of electrode materials. Figure S2b (in the Supporting Information) shows the feature peak of 707.2 eV, associated with the FeS2. The peak of 710.5 eV is attributed to Fe 2p3/2, and the peaks at 719.9 and 724.5 eV are associated with Fe 2p1/2.25,29 In Figure S2c (in the Supporting Information), the S 2p peak at 162.6 eV is related to FeS2,30 the peaks at 164 and 165.1 eV are attributed to Sx2−,31,32 while the one located at 169.2 eV is related to S in the porous vesica-like carbon frameworks.31 The XPS results support that Fe3+ was reduced to Fe2+ and the formation of FeS2 in the hydrothermal process. The mass content of electroactive FeS2 in pitaya-structured FeS2@C composites was measured with thermogravimetric analysis (TGA) in an O2 atmosphere, which indicates a high content of about 86.1 wt % (Figure 2b), assuming the final product was Fe2O3. The microstructure and crystalline features of these FeS2@C particles were investigated by SEM and TEM measurements, as displayed in Figure 2c−f. The FeS2@C sample obtained via solution sulfuration appears to still be individual nanospheres similar as those of the Fe3O4@C precursor (Figure 2c). A high-magnification image (Figure 2d) displays that these spherical particles seem smooth and encapsulated a great quantity of granules. The corresponding typical TEM images of these nanocomposites are included in Figure 2e. It could be seen that a clear pitaya-like structure with dense small seeds (FeS2 particles with a size in the range of 10− 15 nm) fully embedded in porous vesica-like carbon frameworks (Figure S3, in the Supporting Information) was actually formed (inset shows the photographic cross-section of pitaya). This kind of pitaya-like mesoporous structure (Figure S4, in the Supporting Information) possesses a dispersion of tiny FeS2 particles in a

Herein, we demonstrate that this kind of pitaya-structured nanospheres (FeS2@C) can well improve the electrochemical performance of a conversion-type FeS2 cathode. A simple organometallic chemical compound, ferrocene, was chosen as the sole reactant (both as iron and carbon source), which can form homogeneously distributed Fe3O4 nanoparticles in the spherical matrix of carbon. After a convenient sulfuration treatment, these uniform pitaya-structured Fe3O4@C nanocomposites can be completely transformed into FeS2@C, with the pitaya-structured framework well retained. Such a 3D porous vesica-like carbon matrix can alleviate the large volume change, inhibit the coarsening of Fe nanoparticles, and reduce the dissolution of polysulfide. As a consequence, these pitayastructured FeS2@C nanospheres exhibit high energy density, long cycling life, and high rate capability for lithium batteries.

RESULTS AND DISCUSSION Figure 1a exhibits the detailed synthesis process of the pitayastructured FeS2@C uniform nanospheres. First, the precursor of

Figure 1. (a) Schematic illustration of the formation process of pitaya-structured FeS2 @C nanospheres via a facile in situ encapsulation and sulfuration route; (b, c) low- and highmagnification images of a uniform pitaya-structured Fe3O4@C precursor with an average diameter of 200 nm; (d) XRD pattern of a pitaya-structured precursor showing the pure phase of wellcrystallized Fe3O4; (e) TEM image of a typical Fe3O4@C nanosphere clearly displaying small Fe3O4 nanparticles fully embedded in a carbon framework.

uniform Fe3O4@C nanospheres was facilely obtained via a onepot solvothermal decomposition route with the simple organometallic chemical compound ferrocene acting as both iron and carbon source. During this ferrocene decomposition process, iron atoms were oxidized into magnetite and grew as nanoparticles. Accompanying the crystallization of Fe3O4, a carbon-rich macromolecular organic matrix derived from cyclopentadienyl rings in ferrocene was full coated on these nanoparticles via in situ pyrolysis of cyclopentadienyl rings. Subsequently, the carbon-rich organic matrix was transformed 9034

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

ACS Nano

According to previous reports,36,39 theoretically, two cathodic peaks should emerge during the first discharge process, belonging to lithiation (eq 2) and conversion (eq 3) reactions. (2)

Li 2FeS2 + 2Li+ + 2e− → Fe + 2Li 2S

(3)

Actually, these two reactions (lithiation and conversion) are almost simultaneously happening, as their potentials are very close and normally overlap with each other, thus forming only one big cathodic peak. The ex situ XRD patterns of the FeS2@C electrode during discharge/charge processes are illustrated in Figure S5 (in the Supporting Information). It can be seen that when discharged to 1.5 V, FeS2 was transformed into Li2FeS2. When further discharged to 1.0 V, only one fresh peak of Li2S emerged, supporting that the Li2FeS2 was fully converted into Li2S. In the subsequent charge to 3.0 V, the peaks of FeS2 and Fe7S8 can be observed, proving that the electrode during the charge process turns back to FeS2 with a part oxidized to Fe7S8 and S. The corresponding ex situ Fe 2p and S 2p XPS spectra are also provided in Figures S6 and S7 (in the Supporting Information), and the detailed variations of S 2p XPS spectra also support the current proposed lithiation/delithiation mechanism for FeS2. In the first cycle, the oxidation peaks are located at 1.82, 2.35, and 2.50 V, corresponding to the following reactions:36,39

Figure 2. (a) XRD pattern of the pure phase of pitaya-structured FeS2@C products; (b) TGA curve of pitaya-structured FeS2@C nanospheres under an O2 atmosphere from room temperature to 700 °C, indicating the carbon content is about 13.9%; (c, d) low- and high-magnification SEM images of FeS2@C products showing that they well retain the same spherical morphology of the Fe3O4@C precursor; (e, f) TEM and HRTEM images of pitaya-structured FeS2@C nanospheres showing that FeS2 nanoparticles are fully embedded in the amorphous carbon framework. The inset of Figure 2e shows a photo image of the cross-section of a pitaya.

Fe + 2Li 2S → Li 2FeS2 + 2Li+ + 2e−

(4)

Li 2FeS2 → Li 2 − xFeS2 + x Li+ + x e− (0.5 < x < 0.8) (5)

Li 2 − xFeS2 → 0.8ortho‐FeS2 + 0.2FeS8/7 + 0.175S + (2 − x)Li+ + (2 − x)e−

(6)

As the ortho-FeS2 and S are charged products in the FeS2 electrode from the second cycle onward, the discharge processes could be ascribed to a slightly different mechanism with a little different discharge profile (Figure 3b). The outstanding lithium storage properties of pitaya-structured FeS2@C nanospheres are certified by their superior cycling behavior, as depicted in Figure 3c−f. The FeS2@C cathode displays a large first discharge/ charge capacity of about 800/720 mA h g−1 at 0.1 A g−1, with an initial efficiency of approximate 90%. These robust pitayastructured FeS2@C particles are demonstrated with outstanding high-rate capacities ranging from 720 mA h g−1 at 0.1 A g−1 to 400 mA h g−1 at 10.0 A g−1 (Figure 3c,d). When the rate changes back to 0.1 A g−1, the specific capacity can be recuperated to 690 mA h g−1. Besides, good cycling stability is also obtained with this pitaya-structured FeS2@C electrode. After 100 cycles, a capacity of 614 mA h g−1 at 0.3 A g−1 (Figure 3e) and 455 mA h g−1 (Figure 3f) at 1.5 A g−1 is achieved, respectively. Such high rate capability of pitaya-structured FeS2@C is superior to recently reported research on a pyrite-based electrode for LIBs (Table S1, in the Supporting Information). The discharge energy density of the FeS2@C electrode initially reaches 1280 Wh kg−1 (calculated with the mass of FeS2) with a retention of 1074 Wh kg−1 after 100 cycles (Figure 4a), much higher than the LiCoO2 cathode (about 550 Wh kg−1).21 Figure 4b shows the Ragone plots for the current pitaya-structured FeS2@C and some typical cathodes. The as-obtained FeS2@C has an ultrahigh energy density of about 1200 W h kg−1 (the y-axis, the calculation method is shown in the Supporting Information), which is much higher than that

carbon buffer framework, which can alleviate the big volume variation of FeS2 during cycling. In addition, the spherical shape could endow the FeS2@C electrode with a relatively high volumetric density. Actually, the tapped density of these obtained pitaya-structured FeS2@C particles is about 1.3 g cm−3, comparable to a commercial LiFePO4 cathode. The highly crystallized nature of these embedded nanoparticles is also supported by high-resolution TEM (HRTEM) measurement. As displayed in Figure 2f, the clear space of the lattice fringes of 0.270 and 0.242 nm can be assigned to the (200) and (210) planes of the cubic FeS2, respectively. The detailed Li-ion storage performance of these pitayastructured FeS2@C spheres was characterized as the cathode of lithium batteries. It is should be noted that the component of carbon in FeS2@C will not provide capacity in the work voltage region of 1.0−3.0 V. Figure 3a illustrates the initial cyclic voltammetry (CV) profiles of the pitaya-structured FeS2@C composites at a 0.1 mV s−1 scan rate. As shown in this figure, the initial CV curve displays a much different shape as compared with that of the following three cycles. In detail, the first Li+ insertion CV curve shows only one strong cathodic peak located at about 1.24 V. This cathodic peak is associated with the conversion of FeS2 into Li2S and Fe, corresponding to the following reaction:21,33−36,38,39 FeS2 + 4Li+ + 4e− → Fe + 2Li 2S

FeS2 + 2Li+ + 2e− → Li 2FeS2

(1) 9035

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

ACS Nano

Figure 3. Superior Li-ion storage performance of pitaya-structured FeS2@C nanospheres: (a) CV curves at a scanning rate of 0.1 mV s−1 in the voltage range of 1.0−3.0 V; (b) the first three voltage−capacity curves at 0.1 A g−1; (c) voltage−capacity curves of an FeS2@C electrode at different rates (from 0.1 to 10 A g−1); (d) rate performance at different rates (from 0.1 to 10 A g−1); (e, f) long cycling performance of an FeS2@C cathode at 0.3 A g−1 (e) and 1.5 A g−1 (f).

of the modified LiNi 0.5 Mn 1.5 O 4 , LiMn 2 O 4 , and LiFePO 4 cathodes.40−43 Referring to the power density, FeS2@C still displays a comparable specific power density of nearly 10 000 W kg−1 (the x-axis), which endows this FeS2-based cathode with the possibility of fast charging and discharging for practical applications. To further fully investigate the capacitive behavior of the pitaya-structured FeS2@C composites, their kinetics were also analyzed with CV measurements. As clearly shown in Figure 5a, the CV profiles at different sweep rates increasing from 0.2 to 2.0 mV s−1 display a similar shape. Normally, the scan rate (v) and the measured current (i) obey the following relationship:44−46 i = avb

(7)

where a and b stand for empirical parameters. According to previous research,36,37 the b-value of 0.5 represents a diffusioncontrolled behavior, whereas the value of 1 indicates an ideal capacitive one.44 In the current work, the b-values of the pitayastructured FeS2@C for two cathodic peaks and one anodic peak can be determined to be 0.88, 0.69, and 0.81 (by plotting log i vs log v as displayed in Figure 5b), respectively. It can be seen that all these values of b indicate fast kinetics resulting from the pseudocapacitive effect. Furthermore, the total capacitive contribution at a given scan rate could be distinguished into two parts (the diffusion-controlled fraction k1v1/2 and capacitorlike one k2v) at a fixed potential (V) according to the following equation:47,48

Figure 4. (a) Discharge energy density vs cycling number at the material level (∼Wh/kg-FeS2); (b) Ragone plots of typical cathode materials in LIBs and the FeS2@C cathode in this work.

9036

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

ACS Nano

Figure 5. (a−d) Kinetic analysis of the anode of pitaya-structured FeS2@C nanospheres: (a) CV curves with scan rates from 0.2 to 5 mV s−1; (b) CV peak current (Ip) logarithmically potted as a function of the sweep rates (v) to give the slops (b); (c) separation of the capacitive and diffusion currents at a scan rate of 0.6 mV s−1; the capacitive contribution to the total current is shown by the shaded region; (d) contribution ratio of the capacitive and diffusion-controlled charge vs scan rate; (e) schematic illustration of the pitaya-structured nanospheres during lithiation/ delithiation processes.

Figure 6. Na-ion storage performance of pitaya-structured FeS2@C nanospheres: (a) the first three voltage−capacity curves at 0.1 A g−1; (b) long cycling performance of pitaya-structured FeS2@C at 0.6 A g−1 after activation with low current densities of 0.1 and 0.3 A g−1; (c) voltage−capacity curves of pitaya-structured FeS2@C nanospheres at different rates (from 0.1 to 3.2 A g−1); (d) rate performance at different rates (from 0.1 to 3.2 A g−1).

i(V ) = k1v1/2 + k 2v

i(V )/v1/2 = k1 + k 2v1/2

(8)

(9)

By simply transforming eq 8 to 9037

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

ACS Nano k1 and k2 can be facilely achieved by plotting i(V)/v1/2 vs v1/2, and thus the capacitive current ic(V) = k2v could be extracted from the total one with the value of k2.48 For example, at a given 0.6 mV s−1 sweep rate, the CV profile for the capacitive current compared with that of the total measured current is shown in Figure 5c, in which 80.5% is quantified as capacitive. On the basis of the same method, all contribution ratios of the capacitive capacity at different sweep rates (increasing from 0.2 to 2.0 mV s−1) were also determined. Figure 5d summarizes contributions of the pseudocapacitive behaviors at various scan rates. The contributions are 77.9%, 79.2%, 80.5%, 83.4%, 84.1%, 87.0%, and 90.5% at scan rates of 0.2, 0.4, 0.6, 0.8, 1.0, 1.5, and 2.0 mV s−1, respectively. These results clearly show that the pseudocapacitive charge-storage amount does occupy a high portion of the whole capacity. On the basis of the aforementioned microstructure characterization and the above analysis, the superior Li+ storage performance of the current FeS2@C could be closely related to the well-defined 3D architecture, as illustrated in Figure 5e. The 3D porous vesica-like carbon framework can effectively alleviate the large volume change, inhibit the coarsening of Fe nanoparticles, and reduce the dissolution of polysulfide. To confirm this, further SEM measurement after 100 cycles at 0.3 A g−1 was carried out. As exhibited in Figure S8 (Supporting Information), the 3D fully encapsulated spherical framework can also be basically maintained, indicating the good stability of the electrode structure. To strongly support the claim of enhancing electrochemical performance through this kind of structure engineering, the Naion storage performance of the current FeS2@C electrode was also investigated. Both CV (Figure S10a, Supporting Information) and galvanostatic charge/discharge (Figure 6a) measurements show a little larger irreversible specific capacity in the first cycle (122 vs 80 mA h g−1). As clearly exhibited in Figure 6a, the pitaya-structured FeS2@C displays an initial discharge/charge capacity of about 701/579 mA h g−1 at 0.1 A g−1. After the first cycle, both the CV and voltage−capacity profiles nearly overlap, indicating a good cycling performance. Figure 6b displays a long cycling performance of pitaya-structured FeS2@C at 0.6 A g−1 after activation with low rates of 0.1 and 0.3 A g−1. It is clear that a relatively high specific capacity of around 415 mA h g−1 can be well maintained during the whole 100 cycles, while the rate capability testing also shows that this structure-stable electrode has a good performance. As clearly shown in Figure 6c and d, a high and stable reversible capability of 580, 502, 463, 422, 374, and 307 mA h g−1 at 0.1, 0.2, 0.4, 0.8, 1.6, and 3.2 A g−1, respectively, was achieved for Na-ion storage. The Na-storage performance of this pitaya-structured FeS2@C is also superior to other FeS2-based electrodes.49−51 The electrochemical impedance spectroscopy (EIS) spectra of pitaya-structured FeS2@C for Li-ion and Na-ion half-cells are provided in Figure S11 (in the Supporting Information), directly revealing a higher diffusion resistance for the Na-ion half-cell. As compared with that of Liion storage, the electrochemical performance of these pitayastructured FeS2@C for Na-ion storage is a little inferior, which may be related to the larger radius of Na+ (0.9 vs 0.65 Å) influencing the diffusion rate and the higher diffusion resistance.52−54

treatment and a subsequent sulfuration process. The resultant uniform pitaya-structured nanohybrid cathode achieves a stable cycling stability and high rate capability for lithium batteries due to its bionic structure and desirable composition. Remarkably, they demonstrate an ultrahigh and stable specific energy density of about 1100 W h kg−1 at 300 mA g−1, which is much larger than that of the typical LiNi0.5Mn1.5O4, LiMn2O4, and LiFePO4 cathodes. The current simple synthesis approach opens up a way to synthesize other sulfide-based composites for energy storage systems.

METHODS Preparation of Pitaya-Structured FeS2@C. The pitaya-structured FeS2@C composite was prepared via a typical solvothermal method. First, 0.75 g of ferrocene was dissolved in the 75 mL of acetone with continuous stirring for 30 min. Then, the mixture was poured into a Teflon-lined autoclave and moved to an oven at 210 °C for 24 h. The obtained precipitate was centrifuged with ultrapure water and ethanol three times; the collected powder was dried at 80 °C overnight. To get the Fe3O4@C, 0.3 g of collected precursor powder was annealed at 500 °C for 4 h. To prepare the FeS2@C, 0.2 g of Fe3O4@C and 0.4 g of thioactamide (TAA) were added into a Teflon-lined autoclave and dissolved with 80 mL of ultrapure water, then maintained at 200 °C for 15 h.55−58 Finally, the obtained product was centrifuged with ultrapure water and ethanol three times and dried at 80 °C overnight under vacuum conditions to get the FeS2@C composite. Materials Characterization. The X-ray diffraction patterns of the obtained Fe3O4@C and FeS2@C were recorded by a Rigaku-DMax 2400 diffractometer equipped with graphite-monochromatized Cu Kα radiation flux at a scanning rate of 0.02 deg s−1. The morphology and microstructure information was characterized by a Carl Zeiss Supra 40 scanning electron microscope, while TEM and HRTEM were characterized with a JEM-2100 microscope. The content of FeS2 in the FeS2@C composites was determined by a thermogravimetric analysis in an air atmosphere from 40 to 700 °C at a rate of 10 °C min−1 by using a TGA system (TG 209 F3 Tarsus). The nitrogen adsorption/ desorption isotherms of pitaya-structured FeS2@C spheres were measured at 77 K with a Quadrachrome adsorption instrument. XPS analysis was recorded on an XPS/ESCA (electron spectroscopy for chemical analysis) instrument (Axis Ultra DLD from Kratos Inc.) using a monochromatized Al Kα X-ray source. Electrochemical Measurements. The homogeneous slurry was prepared by mixing FeS2@C, polyvinylidene fluoride (PVDF), and Super P with a weight ratio of 80:10:10 in a mortar with appropriate Nmethyl-2-pyrrolidone (NMP). Then the slurry was coated on Cu foil and transferred to a vacuum oven at 80 °C for 12 h. The Li//FeS2@C batteries were assembled in the glovebox (with the O2 and H2O < 0.1 ppm) with Celgard2400 as separator and 1 M lithium bistrifluoromethanesulfonylimide (LiTFSI) dissolved in 1,3-dioxolane (DOL)/diglyme (DME) (1:1 in volume) as electrolyte. For the Na// FeS2@C batteries, 1 M NaCF3SO3 in DME was adopted as electrolyte and Whatman GF/D was used as separator. The galvanostatic charge− discharge tests were performed on CT2001A, LAND test system, with cutoff potentials from 1.0 to 3.0 V for Li-ion batteries and 0.5 to 3.0 V for Na-ion batteries. Cyclic voltammetry profiles were carried out on a Gamry Interface 1000 workstation with a voltage region of 1.0−3.0 V and a scan rate from 0.1 to 2 mV s−1.

ASSOCIATED CONTENT S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsnano.7b03530. Supplementary TEM images of porous vesica-like carbon framework, BET data of uniform pitaya-structured FeS2@ C nanospheres, SEM images of pitaya-structured FeS2@C cathode after 100 cycles, ex situ XPS and XRD results of FeS2@C cathode during charge/discharge processes,

CONCLUSION In summary, we have fabricated pitaya-structured nanocomposites with conversion-type FeS2 nanoparticles embedded in porous vesica-like carbon frameworks via a facile in situ encapsulation route, which consists of a first solvothermal 9038

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

ACS Nano

Route from Reed Plants to a Silicon Anode for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2015, 54, 9632−9636. (12) Liu, J.; Wen, Y.; van Aken, P. A.; Maier, J.; Yu, Y. Facile Synthesis of Highly Porous Ni−Sn Intermetallic Microcages with Excellent Electrochemical Performance for Lithium and Sodium Storage. Nano Lett. 2014, 14, 6387−392. (13) Li, X.; Dhanabalan, A.; Gu, L.; Wang, C. Three-Dimensional Porous Core-Shell Sn@Carbon Composite Anodes for High-Performance Lithium-Ion Battery Applications. Adv. Energy Mater. 2012, 2, 238−244. (14) Whittingham, M. S. Ultimate Limits to Intercalation Reactions for Lithium Batteries. Chem. Rev. 2014, 114, 11414−11443. (15) Li, C.; Gu, L.; Tong, J.; Maier, J. Carbon Nanotube Wiring of Electrodes for High-Rate Lithium Batteries Using an ImidazoliumBased Ionic Liquid Precursor as Dispersant and Binder: A Case Study on Iron Fluoride Nanoparticles. ACS Nano 2011, 5, 2930−2938. (16) Li, C.; Yin, C.; Gu, L.; Dinnebier, R. E.; Mu, X.; van Akan, P. A.; Maier, J. An FeF3·0.5H2O Polytype: A Microporous Framework Compound with Intersecting Tunnels for Li and Na Batteries. J. Am. Chem. Soc. 2013, 135, 11425−11428. (17) Yersak, T. A.; Macpherson, H. A.; Kim, S. C.; Le, V. D.; Kang, C. S.; Son, S. B.; Kim, Y. H.; Trevey, J. E.; Oh, K. H.; Stoldt, C.; Lee, S. H. Solid State Enabled Reversible Four Electron Storage. Adv. Energy Mater. 2013, 3, 120−127. (18) Evans, T.; Piper, D. M.; Kim, S. C.; Han, S. S.; Bhat, V.; Oh, K. H.; Lee, S. H. Ionic Liquid Enabled FeS2 for High-Energy-Density LithiumIon Batteries. Adv. Mater. 2014, 26, 7386−7392. (19) Son, S. B.; Yersak, T. A.; Piper, D. M.; Kim, S. C.; Kang, C. S.; Cho, J. S.; Suh, S. S.; Kim, Y. U.; Oh, K. H.; Lee, S. H. A Stabilized PAN-FeS2 Cathode with an EC/DEC Liquid Electrolyte. Adv. Energy Mater. 2014, 4, 1300961. (20) Whiteley, J. M.; Hafner, S.; Han, S. S.; Kim, S. C.; Oh, K. H.; Lee, S. H. FeS2-Imbedded Mixed Conducting Matrix as a Solid Battery Cathode. Adv. Energy Mater. 2016, 6, 1600495. (21) Liu, J.; Wen, Y.; Wang, Y.; van Aken, P. A.; Maier, J.; Yu, Y. Carbon-Encapsulated Pyrite as Stable and Earth-Abundant High Energy Cathode Material for Rechargeable Lithium Batteries. Adv. Mater. 2014, 26, 6025−6030. (22) Zhu, Y.; Fan, X.; Suo, L.; Luo, C.; Gao, T.; Wang, C. Electrospun FeS2@Carbon Fiber Electrode as a High Energy Density Cathode for Rechargeable Lithium Batteries. ACS Nano 2016, 10, 1529−1538. (23) Zhu, C.; Wen, Y.; van Aken, P. A.; Maier, J.; Yu, Y. High Lithium Storage Performance of FeS Nanodots in Porous Graphitic Carbon Nanowires. Adv. Funct. Mater. 2015, 25, 2335−2342. (24) Xu, L.; Hu, Y.; Zhang, H.; Jiang, H.; Li, C. Confined Synthesis of FeS2 Nanoparticles Encapsulated in Carbon Nanotube Hybrids for Ultrastable Lithium-Ion Batteries. ACS Sustainable Chem. Eng. 2016, 4, 4251−4255. (25) Wei, R.; Wang, J.; Wang, Z.; Tong, L.; Liu, X. Magnetite-Bridged Carbon Nanotubes/Graphene Sheets Three-Dimensional Network with Excellent Microwave Absorption. J. Electron. Mater. 2017, 46, 2097−2105. (26) Jiao, J.; Qiu, W.; Tang, J.; Chen, L.; Jing, L. Synthesis of WellDefined Fe3O4 Nanorods/N-doped Graphene for Lithium-Ion Batteries. Nano Res. 2016, 9, 1256−1266. (27) Kumar, R.; Singh, R. K.; Vaz, A. R.; Savu, R.; Moshkalev, S. A. SelfAssembled and One-Step Synthesis of Interconnected 3D Network of Fe3O4/Reduced Graphene Oxide Nanosheets Hybrid for HighPerformance Supercapacitor Electrode. ACS Appl. Mater. Interfaces 2017, 9, 8880−8890. (28) Divyapriya, G.; Nambi, I. M.; Senthilnathan, J. An Innate Quinone Functionalized Electrochemically Exfoliated Graphene/Fe3O4 Composite Electrode for the Continuous Generation of Reactive Oxygen Species. Chem. Eng. J. 2017, 316, 964−977. (29) Li, Y.; Van Santen, R.; Weber, T. High-Temperature FeS−FeS2 Solid-State Transitions: Reactions of Solid Mackinawite with Gaseous H2S. J. Solid State Chem. 2008, 181, 3151−3162. (30) Velasquez, P.; Leinen, D.; Pascual, J.; Ramos-Barrado, J. R.; Grez, P.; Gomez, H.; Schrebler, R.; Del Río, R.; Cordova, R. A Chemical,

electrochemical performance of the pitaya-structured FeS2@C electrode for SIBs (PDF)

AUTHOR INFORMATION Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Jun Liu: 0000-0002-7078-8046 Liuzhang Ouyang: 0000-0003-2392-2801 Min Zhu: 0000-0001-5018-2525 Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS This work was supported by the National Key Research and Development Program (no. 2016YFA0202603), the National Natural Science Foundation of China (no. 51771076), the Foundation for Innovative Research Groups of the National Natural Science Foundation of China (no. 51621001), the “1000 Plan” from the Chinese Government, and the Project of Public Interest Research and Capacity Building of Guangdong Province (no. 2017A010104004). REFERENCES (1) Larcher, D.; Tarascon, J. M. Towards Greener and More Sustainable Batteries for Electrical Energy Storage. Nat. Chem. 2015, 7, 19−29. (2) Kovalenko, I.; Zdyrko, B.; Magasinski, A.; Hertzberg, B.; Milicev, Z.; Burtovyy, R.; Luzinov, I.; Yushin, G. A Major Constituent of Brown Algae for Use in High-Capacity Li-Ion Batteries. Science 2009, 334, 75− 79. (3) Choi, N. S.; Chen, Z.; Freunberger, S. A.; Ji, X.; Sun, Y. K.; Amine, K.; Yushin, G.; Nazar, L. F.; Cho, J.; Bruce, P. G. Challenges Facing Lithium Batteries and Electrical Double-Layer Capacitors. Angew. Chem., Int. Ed. 2012, 51, 9994−10024. (4) Liu, J.; Song, K.; van Aken, P. A.; Maier, J.; Yu, Y. Self-Supported Li4Ti5O12−C Nanotube Arrays as High-Rate and Long-Life Anode Materials for Flexible Li-Ion Batteries. Nano Lett. 2014, 14, 2597−2603. (5) Liu, J.; Gu, M.; Ouyang, L. Z.; Wang, H.; Yang, L.; Zhu, M. Sandwich-like SnS/Polypyrrole Ultrathin Nanosheets as High-Performance Anode Materials for Li-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 8502−8510. (6) Zhao, Y.; Li, X. F.; Yan, B.; Xiong, D.; Li, D.; Lawes, S.; Sun, X. Recent Developments and Understanding of Novel Mixed TransitionMetal Oxides as Anodes in Lithium Ion Batteries. Adv. Energy Mater. 2016, 6, 1502175. (7) Liu, J.; Xu, X.; Hu, R.; Yang, L.; Zhu, M. Uniform Hierarchical Fe3O4@Polypyrrole Nanocages for Superior Lithium Ion Battery Anodes. Adv. Energy Mater. 2016, 6, 1600256. (8) Liu, J.; Yu, L.; Wu, C.; Wen, Y.; Yin, K.; Chiang, F.; Hu, R.; Liu, J.; Sun, L.; Gu, L.; Maier, J.; Yu, Y.; Zhu, M. New Nanoconfined Galvanic Replacement Synthesis of Hollow Sb@C Yolk−Shell Spheres Constituting a Stable Anode for High-Rate Li/Na-Ion Batteries. Nano Lett. 2017, 17, 2034−2042. (9) Yu, L.; Liu, J.; Xu, X.; Zhang, L.; Hu, R.; Liu, J.; Yang, L.; Zhu, M. Metal-Organic Framework-Derived NiSb Alloy Embedded in Carbon Hollow Spheres as Superior Lithium-Ion Battery Anodes. ACS Appl. Mater. Interfaces 2017, 9, 2516−2525. (10) Liu, J.; Song, K.; Zhu, C.; Chen, C.; van Aken, P. A.; Maier, J.; Yu, Y. Ge/C Nanowires as High-Capacity and Long-Life Anode Materials for Li-Ion Batteries. ACS Nano 2014, 8, 7051−7059. (11) Liu, J.; Kopold, P.; van Aken, P. A.; Maier, J.; Yu, Y. Energy Storage Materials from Nature through Nanotechnology: A Sustainable 9039

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040

Article

ACS Nano Morphological, and Electrochemical (XPS, SEM/EDX, CV, and EIS) Analysis of Electrochemically Modified Electrode Surfaces of Natural Chalcopyrite (CuFeS2) and Pyrite (FeS2) in Alkaline Solutions. J. Phys. Chem. B 2005, 109, 4977−4988. (31) Niu, S.; Lv, W.; Zhou, G.; Shi, H.; Qin, X.; Zheng, C.; Zhou, T.; Luo, C.; Deng, Y.; Li, B. Electrostatic-Spraying an Ultrathin, Multifunctional and Compact Coating onto a Cathode for a LongLife and High-Rate Lithium-Sulfur Battery. Nano Energy 2016, 30, 138− 145. (32) Liu, L.-J.; Chen, Y.; Zhang, Z.-F.; You, X.-L.; Walle, M. D.; Li, Y.J.; Liu, Y.-N. Electrochemical Reaction of Sulfur Cathodes with Ni Foam Current Collector in Li-S Batteries. J. Power Sources 2016, 325, 301− 305. (33) Fan, X.; Zhu, Y.; Luo, C.; Suo, L.; Lin, Y.; Gao, T.; Xu, K.; Wang, C. Pomegranate-Structured Conversion-Reaction Cathode with a Builtin Li Source for High-Energy Li-Ion Batteries. ACS Nano 2016, 10, 5567−5577. (34) Walter, M.; Zund, T.; Kovalenko, M. V. Pyrite (FeS2) Nanocrystals as Inexpensive High Performance Lithium-Ion Cathode and Sodium-Ion Anode Materials. Nanoscale 2015, 7, 9158−9163. (35) Tan, R.; Yang, J.; Hu, J.; Wang, K.; Zhao, Y.; Pan, F. Core−Shell Nano-FeS2@N-Doped Graphene as an Advanced Cathode Material for Rechargeable Li-Ion Batteries. Chem. Commun. 2016, 52, 986−989. (36) Pan, G. X.; Cao, F.; Xia, X. H.; Zhang, Y. J. Exploring Hierarchical FeS2/C Composite Nanotubes Arrays as Advanced Cathode for Lithium Ion Batteries. J. Power Sources 2016, 332, 383−388. (37) Douglas, A.; Carter, R.; Oakes, L.; Share, K.; Cohn, A. P.; Pint, C. L. Ultrafine Iron Pyrite (FeS2) Nanocrystals Improve Sodium-Sulfur and Lithium-Sulfur Conversion Reactions for Efficient Batteries. ACS Nano 2015, 9, 11156−11165. (38) Hu, Z.; Zhang, K.; Zhu, Z.; Tao, Z.; Chen, J. FeS2 Microspheres with an Ether-Based Electrolyte for High-Performance Rechargeable Lithium Batteries. J. Mater. Chem. A 2015, 3, 12898−12904. (39) Zhang, F.; Wang, C.; Huang, G.; Yin, D.; Wang, L. FeS2@C Nanowires Derived from Organic-Inorganic Hybrid Nanowires for High-Rate and Long-Life Lithium-Ion Batteries. J. Power Sources 2016, 328, 56−64. (40) Cheng, F.; Wang, H.; Zhu, Z.; Wang, Y.; Zhang, T.; Tao, Z.; Chen, J. Porous LiMn2O4 Nanorods with Durable High-Rate Capability for Rechargeable Li-Ion Batteries. Energy Environ. Sci. 2011, 4, 3668−3675. (41) Zhang, X.; Cheng, F.; Yang, J.; Chen, J. LiNi0.5Mn1.5O4 Porous Nanorods as High-Rate and Long-Life Cathodes for Li-Ion Batteries. Nano Lett. 2013, 13, 2822−2825. (42) Lee, Y.; Kim, M. G.; Cho, J. Layered Li0.88[Li0.18Co0.33Mn0.49]O2 Nanowires for Fast and High Capacity Li-Ion Storage Material. Nano Lett. 2008, 8, 957−961. (43) Kang, B.; Ceder, G. Battery Materials for Ultrafast Charging and Discharging. Nature 2009, 458, 190−193. (44) Yang, L.; Li, X.; He, S.; Du, G.; Yu, X.; Liu, J.; Gao, Q.; Hu, R.; Zhu, M. Mesoporous Mo2C/N-doped carbon heteronanowires as highrate and long-life anode materials for Li-ion batteries. J. Mater. Chem. A 2016, 4, 10842−10849. (45) Zhang, K.; Park, M.; Zhou, L.; Lee, G. H.; Shin, J.; Hu, Z.; Chou, S. L.; Chen, J.; Kang, Y. M. Cobalt-Doped FeS2 Nanospheres with Complete Solid Solubility as a High-Performance Anode Material for Sodium-Ion Batteries. Angew. Chem., Int. Ed. 2016, 55, 12822−12826. (46) Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive Oxide Materials for High-Rate Electrochemical Energy Storage. Energy Environ. Sci. 2014, 7, 1597−1614. (47) Augustyn, V.; Come, J.; Lowe, M. A.; Kim, J. W.; Taberna, P. L.; Tolbert, S. H.; Abruna, H. D.; Simon, P.; Dunn, B. High-Rate Electrochemical Energy Storage through Li+ Intercalation Pseudocapacitance. Nat. Mater. 2013, 12, 518−522. (48) Chao, D.; Zhu, C.; Yang, P.; Xia, X.; Liu, J.; Wang, J.; Fan, X.; Savilov, S. V.; Lin, J.; Fan, H.; Shen, Z. Array of Nanosheets Render Ultrafast and High-Capacity Na-Ion Storage by Tunable Pseudocapacitance. Nat. Commun. 2016, 7, 12122.

(49) Kitajou, A.; Yamaguchi, J.; Hara, S.; Okada, S. Discharge/Charge Reaction Mechanism of a Pyrite-Type FeS2 Cathode for Sodium Secondary Batteries. J. Power Sources 2014, 247, 391−395. (50) Chen, K.; Zhang, W.; Xue, L.; Chen, W.; Xiang, X.; Wan, M.; Huang, Y. Mechanism of Capacity Fade in Sodium Storage and the Strategies of Improvement for FeS2 Anode. ACS Appl. Mater. Interfaces 2017, 9, 1536−1541. (51) Chen, W.; Qi, S.; Guan, L.; Liu, C.; Cui, S.; Shen, C.; Mi, L. Pyrite FeS2 Microspheres Anchoring on Reduced Graphene Oxide Aerogel as an Enhanced Electrode Material for Sodium-Ion Batteries. J. Mater. Chem. A 2017, 5, 5332−5341. (52) Xu, X.; Liu, J.; Hu, R.; Liu, J.; Ouyang, L.; Zhu, M. Self-Supported CoP Nanorod Arrays Grafted on Stainless Steel as an Advanced Integrated Anode for Stable and Long-Life Lithium-Ion Batteries. Chem. - Eur. J. 2017, 23, 5198−5204. (53) Hu, Z.; Niu, Z.; Cheng, F.; Zhang, K.; Wang, J.; Chen, C.; Chen, J. Pyrite FeS2 for High-Rate and Long-Life Rechargeable Sodium Batteries. Energy Environ. Sci. 2015, 8, 1309−1316. (54) Yu, L.; Liu, J.; Xu, X.; Zhang, L.; Hu, R.; Liu, J.; Ouyang, L.; Yang, L.; Zhu, M. Ilmenite Nanotubes for High Stability and High Rate Sodium-Ion Battery Anodes. ACS Nano 2017, 11, 5120−5129. (55) Li, T.; Liu, H.; Wu, Z.; Liu, Y.; Guo, Z.; Zhang, H. Seeded Preparation of Ultrathin FeS2 Nanosheets from Fe3O4 Nanoparticles. Nanoscale 2016, 8, 11792−11796. (56) Li, T.; Li, H.; Wu, Z.; Hao, H.; Liu, J.; Huang, T.; Sun, H.; Zhang, J.; Zhang, H.; Guo, Z. Colloidal Synthesis of Greigite Nanoplates with Controlled Lateral Size for Electrochemical Applications. Nanoscale 2015, 7, 4171−4178. (57) Liang, C.; Dudney, N. J.; Howe, J. Y. Hierarchically Structured Sulfur/Carbon Nanocomposite Material for High-Energy Lithium Battery. Chem. Mater. 2009, 21, 4724−4730. (58) Ding, N.; Lum, Y.; Chen, S.; Chien, S. W.; Hor, T. A.; Liu, Z.; Zong, Y. Sulfur−Carbon Yolk−Shell Particle Based 3D Interconnected Nanostructures as Cathodes for Rechargeable Lithium−Sulfur Batteries. J. Mater. Chem. A 2015, 3, 1853−1857.

9040

DOI: 10.1021/acsnano.7b03530 ACS Nano 2017, 11, 9033−9040