Polypyrrole-Modified Prussian Blue Cathode Material for Potassium

May 31, 2019 - Potassium-ion batteries (PIBs) have received significant attention because of .... oxidative polymerization to obtain a potassium-rich ...
0 downloads 0 Views 4MB Size
Research Article www.acsami.org

Cite This: ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Polypyrrole-Modified Prussian Blue Cathode Material for Potassium Ion Batteries via In Situ Polymerization Coating Qing Xue,† Li Li,†,‡ Yongxin Huang,† Ruling Huang,† Feng Wu,†,‡ and Renjie Chen*,†,‡ †

Beijing Key Laboratory of Environmental Science and Engineering, Beijing Institute of Technology, Beijing 100081, China Collaborative Innovation Center of Electric Vehicles in Beijing, Beijing 100081, China



Downloaded via UNIV OF SOUTHERN INDIANA on July 25, 2019 at 13:24:04 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Potassium-ion batteries (PIBs) have received significant attention because of the abundant potassium reserves and similar electrochemistry of potassium to that of lithium. Because of the open framework and structural controllability, Prussian blue and its analogues (PB) are considered to be competitive cathodes of PIBs. However, the intrinsic lattice defects and poor electronic conductivity of PBs induce poor cycling performance and rate capability. Herein, we propose a polypyrrole-modified Prussian blue material (KHCF@PPy) via an in situ polymerization coating method for the first time. KHCF@PPy possesses a low defect concentration and improved electronic conductivity, and the electrode was found to exhibit 88.9 mA h g−1 discharge capacity at 50 mA g−1, with 86.8% capacity retention after 500 cycles. At a higher current density of 1000 mA g−1, the initial discharge capacity was 72.1 mA h g−1, which dropped slightly to 61.8 mA h g−1 after 500 cycles. The capacity decay rate was 0.03% per cycle. Detailed characterization showed a lack of phase transition during the charge and discharge processes and determined that K ions were not completely extracted from the monoclinic structure, possibly contributing to the excellent cycling stability. This simple surface modification method represents a promising means of mitigating issues currently associated with PB-based cathodes for PIBs. KEYWORDS: potassium ion batteries, cathode, Prussian blue, polypyrrole-modified, in situ polymerization coating

1. INTRODUCTION Lithium-ion batteries (LIBs) have been extensively used in consumer electronics and are attracting increasing attention in the field of electric vehicles. However, the high cost of lithium and its relative scarcity limit its applications in large-scale energy storage systems (EESs).1−6 For this reason, sodium-ion batteries (SIBs) and potassium-ion batteries (PIBs) have become promising candidates for EESs because of the rich abundance of Na and K and their similar electrochemical properties compared with lithium.7−12 Recently, PIBs have become the subject of much research because the K+/K redox pair has a more negative redox potential than the Na+/Na pair.13,14 Graphitic anodes have also been successfully employed with potassium, thus further promoting the study of PIBs.15,16 It has been reported that PIBs using graphite negative electrodes can obtain reversible capacities over 200 mA h g−1, which is comparable with that obtained from LIBs.16,17 However, compared with the progress in the anode design, appropriate PIB cathodes have rarely been reported, possibly because of the larger ionic radius of K ions, which possibly leads to structural collapse or distortion.18,19 Consequently, the development of PIBs is still in the infant stage. Prussian blue and its analogues (PBs) have been extensively studied as components of SIBs over the past few years.20,21 © 2019 American Chemical Society

This research interest stems from the structural stability, rich redox-active sites, and cost advantage of these materials, and has resulted in remarkable achievements.6,22 Even so, the electrochemical performance of PBs remains unsatisfactory because of the vacancies in the open frameworks of these materials accompanying the introduction of coordinated water molecules.23 To eliminate this effect, researchers have used many approaches to reduce the lattice defects and remove interstitial water in PB-based cathodes, including morphological tailoring, surface modification, nanostructure design, and composition optimization.24−28 The large interstices and ion-transport channels in PB lattices facilitate the accommodation of large alkali metal ions, suggesting that these materials could be used as cathodes for PIBs.29−32 In addition, PBs preferentially store K+ over Na+ ions because of the smaller Stokes radius of K+ in electrolyte solutions and the lower Gibbs free energy associated with the combination of K+ with the cavities in PBs.13,33 Moreover, the water content of Kbased PBs is very low as a result of the weak hydration ability of K + during the hydrothermal preparation of these materials.34−36 Generally speaking, it should be possible to Received: March 14, 2019 Accepted: May 31, 2019 Published: May 31, 2019 22339

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Research Article

ACS Applied Materials & Interfaces

Figure 1. (a) Illustration scheme of the synthesis process; SEM of pure KHCF (b) and KHCF@PPy (c); TEM of KHCF@PPy (d) and KHCF (insert of d).

develop K+ insertion cathodes having good structural integrity based on PBs. Lei and co-workers29 investigated K0.22Fe[Fe(CN)6]0.805·□0.195·4.01H2O as a potential cathode material for nonaqueous PIBs. This material showed a reversible capacity of ∼73 mA h g−1 at 50 mA g−1 with a high discharge voltage between 3.1 and 3.4 V. However, the high potential polarization resulted from the low electronic conductivity of such cathodes, and the poor cycling performance caused by side reactions with the electrolyte should be ameliorated to allow the further development of PBs in PIBs.37−40 On the basis of the research regarding the use of PBs in SIBs, approaches, such as structural optimization and surface modification, could be examined to enhance the cycling performance and rate capability of PBs in PIBs. However, as far as we know, there are no reports regarding the surface modification of PBs for use in PIBs. Herein, we synthesized a polypyrrole (PPy)-coated K-rich iron hexacyanoferrate composite (KHCF@PPy) with a long cycle life and excellent rate capacity, for use as a PIB cathode material. In this synthesis, pyrrole was polymerized in situ on the surfaces of KHCF particles utilizing the intrinsic oxidizing ability of KHCF and with no additional oxidant. The introduction of this conductive polymer enhanced the electronic conductivity of KHCF, thus improving the rate capability. Simultaneously, KHCF itself was reduced during the oxidative polymerization to obtain a potassium-rich iron hexacyanoferrate. Moreover, excess [Fe(CN)6]4− ions in the solution were doped into the PPy chains during the polymerization process, resulting in improved electronic conductivity as well as increased capacity. It has been reported that the redox potential of [Fe(CN)6]4− ions is close to the Fermi level of PPy,41 such that these ions can not only act as dopants to enhance the electronic conductivity of PPy but also work as redox-active centers to

increase the capacity of the doped PPy. These effects mitigate the decrease in specific capacity caused by applying a conductive coating to KHCF. In summary, a modified PB was synthesized as a good candidate for the PIB cathode material because of its excellent cycling stability and rate capability. The one-pot coating method demonstrated herein is both simple and energy-effective, which is applicable to other electrode material modifications.

2. EXPERIMENTAL SECTION 2.1. Material Preparation. All the chemical reagents employed in this study were directly used without any purification. In the synthetic process, 3.315 g K4Fe(CN)6·H2O was dissolved in 135 mLof deionized water (DI) in a 250 mL three-necked flask to obtain solution A, whereas 1.48 g hydrochloric acid (37%) was added to 15 mL of DI to form solution B. Then, solution B was slowly added to solution A drop by drop with continuous stirring. The resulting light yellow mixture was kept at 60 °C with vigorous stirring, during which it gradually took on a blue coloration. After 4 h, 75 μL of pyrrole was added to the KHCF suspension with magnetic stirring, while bubbling nitrogen gas through the solution. Polymerization was carried out at 0−5 °C for 3 h, after which the PPy-coated KHCF precipitates were separated by centrifugation. Finally, KHCF@PPy was gained after drying in an oven at 80 °C overnight. As a contrast, pure KHCF material was prepared by a similar method but with no addition of pyrrole. 2.2. Structural Characterization. A Rigaku Ultima IV-185 type X-ray diffraction (XRD) instrument was used to characterize the crystal structures of the as-prepared powders. The radiation source is Cu Kα. Raman spectra at the range of 200−3200 cm−1 was obtained on a Renishaw inVia spectrometer with a laser wavelength of 633 nm. The chemical compositions of the synthesized powders were determined by assessing the mass ratios of K, Fe, C, N, and H using inductively coupled plasma (ICP) combined with elemental analysis. Additionally, thermogravimetric analysis (TGA) was conducted on a Netzsch STA 449F3 analyzer to determine the 22340

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Research Article

ACS Applied Materials & Interfaces

Figure 2. (a) XRD of KHCF@PPy; (b) Raman spectra of KHCF and KHCF@PPy; (c) Fe 2p XPS spectra of KHCF@PPy; (d) iron-57 Mössbauer spectra of KHCF@PPy. water content in each sample. X-ray photoelectron spectroscopy (XPS) was performed on Uivac-PHI with Al Kα as the radiation source to assess the element valence values. Iron-57 Mössbauer spectra were determined by a Wissel accelerated drive Mössbauer spectrometer with 57Co/Pd as the radiation source at room temperature. The spectra were fitted and resolved using the leastsquares method. Morphology characterization was performed using a Hitachi S-4800 scanning electron microscope combined with a JEM2100F transmission electron microscope. 2.3. Computational Method. Density functional theory (DFT) calculations were performed employing the CASTEP software package (Materials Studio 2018). The exchange−correlation function was calculated according to the generalized gradient approximation with the parameter setting of Perdew−Burke−Ernzerhof. Hubbard U term values of 7.0 and 3.0 eV were applied for the high-spin and lowspin Fe 3d orbitals, and a cutoff value of 450 eV for kinetic energy of plane wave expansion was adopted to fit the valence electrons. Ultrasoft pseudopotentials were used to describe the core-valence interactions, and the Brillouin region was integrated with the k-point mesh density of 4 × 2 × 2. Additionally, the maximum stress was 0.02 GPa and the maximum force was 0.01 eV Å−1. 2.4. Electrochemical Characterization. The KHCF@PPy and KHCF electrodes were prepared by casting a slurry of active powder, conductive Super P, and polyvinylidene fluoride (with a weight ratio of 8:1:1) onto Al foil. The thickness of the resulting layer was controlled at 150 μm. After drying at 80 °C overnight, the as-prepared electrode piece was cut into small disks, with a material loading on each disk of approximately 2.1 mg cm−2. The electrochemical performances of the as-prepared electrodes were tested in CR2032type coin cells which were assembled in a glovebox filled with argon. In each cell, the K metal anode and the as-prepared cathode were separated by a Whatman glass fiber separator. The electrolyte used was 0.8 M KPF6 dissolved in ethylene carbonate and diethyl carbonate with a volume ratio of 1:1. The cells were cycled in the range of 2−4.2 V at different current densities using a Land Test System. Cyclic voltammetry (CV) was conducted with a 0.1 mV s−1 scan rate on a SP-150-type Biologic workstation. The specific capacity was calculated based on the combined KHCF and PPy mass. In the preparation for ex situ XRD analyses, the cathodes were separated from the disassembled cells of different charge/discharge states, washed with 1,3-dioxolane, and dried sufficiently under vacuum.

3. RESULTS AND DISCUSSION The reaction mechanism associated with the KHCF@PPy synthesis is schematically illustrated in Figure 1a. In an acidic environment, Fe2+ ions slowly dissociated from [Fe(CN)6]4− and were oxidized to Fe3+ by reaction with ambient air. In addition, the undecomposed [Fe(CN)6]4− reacted with Fe2+/ Fe3+ to form cubic KHCF nuclei that gradually grew as the reaction proceeded. After 4 h, pyrrole was added to the suspension containing Fe3+ ions, which oxidized pyrrole to form the polymer. Simultaneously, the residual [Fe(CN)6]4− in the solution was doped into PPy. In this manner, high-quality KHCF coated with [Fe(CN)6]4−-doped PPy was obtained. The morphological features were characterized by scanning electron microscopy (SEM) and transmission electron microscopy (TEM) and the resulting images are shown in Figure 1b−d. Numerous standard nanocubic structures with smooth surfaces can be observed in the pure KHCF, the edge lengths of which vary between 500 and 700 nm (Figures 1b and S1a). In comparison, the coated KHCF exhibits coarse surfaces with some small aggregations on the surfaces, and these features are related to the PPy coating (Figures 1c and S1b). The high-resolution TEM images further indicate the existence of a PPy coating layer on the KHCF surface. Compared with the pure crystalline KHCF (insert to Figure 1d), Figure 1d demonstrates that the edges of KHCF@PPy have a uniform, thin amorphous layer, indicative of a homogeneous PPy coating. The data used for the structural analysis of KHCF@PPy powders are presented in Figure 2. The crystal properties of KHCF@PPy were assessed by XRD, and the associated Rietveld refinement patterns are shown in Figure 2a. According to the Rietveld refinement (Rwp = 10.88%, Rp = 7.51%) based on the Reflex procedure, the lattice parameters were determined to be a = 10.091 Å, b = 7.341 Å, c = 7.048 Å, α = 90.00°, β = 89.58°, and γ = 90°, which can be readily indexed to a monoclinic phase with the space group P21/n. This is a typical highly stable, K-rich configuration, in which FeN6 and FeC6 octahedra are linked by (CN)− ligands. The rigid open 22341

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Research Article

ACS Applied Materials & Interfaces

Figure 3. CV curves of KHCF@PPy (a) and KHCF (b) at a scan rate of 0.1 mV s−1; charge/discharge profiles of KHCF@PPy (c) and KHCF (d) at different cycles under 50 mA g−1.

solid line in this figure represent the experimental data and fitted results, respectively, and the details of the fitted results are listed in Table S1. The isomer shift values of 0.07 and 0.95 mm s−1 are ascribed to Fe3+ and Fe2+, with the corresponding weight percentages of 0.44 and 0.56, respectively. These results further demonstrate that the K content in KHCF was high relative to previously reported values.42−44 We attempted to use elemental analysis and TGA results to determine the PPy content in the composite but, unfortunately, the exact value was difficult to obtain because C and N were present in both PPy and KHCF, and the temperature ranges over which pyrrole decomposed and coordinated water evaporated were overlapped. Therefore, a pyrrole content of about 5 wt % was roughly estimated based on the mass difference between KHCF and KHCF@PPy. The electrochemical behaviors of KHCF and KHCF@PPy were initially measured by CV in K-half cells and the results are displayed in Figure 3. The CV curves obtained from KHCF@ PPy (Figure 3a) show a 3.68/3.3 V couple peak during the first scan, ascribed to the redox potential of low-spin FeIII/FeII connected to C atoms. The following two CV profiles exhibit similar two peaks, implying identical redox mechanism and the high reversibility of depotassiation/potassiation. It is noteworthy that the sharp peak at approximately 4.15 V that appeared during the first cycle is attributed to the solid electrolyte interphase (SEI) formation, based on prior reports. Subsequently, the 4.15 V peak weakened and the polarization voltage gradually decreased to a steady value of 350 mV, suggesting a complete and stable SEI film formation. The CV behavior of KHCF as depicted in Figure 3b was basically the same as that of KHCF@PPy, except that its polarization voltage was slightly larger (approximately 500 mV).The charge/discharge profiles of KHCF@PPy at 50 mA g−1 are provided in Figure 3c. Two obvious plateaus can be observed at approximately 3.6 and 3.3 V, which is consistent with the CV measurements. The first charge and discharge capacities of KHCF@PPy were 122.2 and 88.8 mA h g−1, respectively, with a Coulombic efficiency (CE) of 72.67%. This value was higher than that obtained from KHCF (64%) in Figure 3d. This

framework of this structure, which contains large interstitial sites, can accommodate the relatively large K+ ions, while simultaneously providing suitable channels for the diffusion of these ions. In contrast, the pure KHCF exhibits a typical facecentered cubic structure (Figure S2), which is consistent with that of Fe-HCF synthesized via a traditional single iron source co-precipitation method.42 The KHCF@PPy Raman spectrum in Figure 2b contains four characteristic PPy peaks at 925.2, 1050.3, 1328.8, and 1601.9 cm−1, confirming the introduction of PPy onto the KHCF surface through the in situ polymerization coating method. Peaks related to the ν (CN−) band appear at 2115 and 2079 cm−1, suggesting that Fe in this compound was primarily bivalent. This was confirmed by the XPS data in Figure 2c. In the Fe 2p spectrum obtained from KHCF@PPy, the primary peaks centered at 708.63 and 721.43 eV are assigned to FeII 2p3/2 and FeII 2p1/2, respectively, indicating that Fe2+ was the primary ion. Considering the high concentration of divalent iron in the compound, it is likely that PB had a high potassium content, which is also demonstrated by the chemical composition analysis. On the basis of the ICP and elemental analysis results, the chemical composition of KHCF@PPy was characterized as K1.87Fe[Fe(CN)6]0.97·□0.03 (where □ represents a [Fe(CN)6]4− vacancy), which obviously corresponds to a K-rich PB compound. The TGA curves obtained from this material (Figure S3) show a mass loss of 4.2% below 250 °C, indicating 0.84 water molecules per molecule. Consequently, the chemical formula of KHCF@PPy is K 1 . 8 7 Fe[Fe(CN)6]0.97·□0.03·0.84H2O. Using the same method, the formula of KHCF was found to be K 0 . 6 8 Fe[Fe(CN)6]0.86·□0.14·1.68H2O, suggesting that there was a lower K content and a greater concentration of defects in the pure KHCF. It is assumed that the higher K concentration in KHCF@PPy can be attributed to the reduced number of vacancies in the framework as well as the increased level of Fe2+ resulting from the oxidation effect during the pyrrole polymerization. Iron-57 Mössbauer spectra were acquired at room temperature (Figure 2d) to identify the Fe valence as well as the K content in KHCF@PPy. The black stars and red 22342

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Research Article

ACS Applied Materials & Interfaces

Figure 4. Electrochemical performances of KHCF@PPy: (a) cycling performance and CE at 50 mA g−1; (b) rate capability and CE at various current densities; (c) long cycling performance at 1000 mA g−1; (d) electrochemical impedance spectra measured from 10 mHz to 100 kHz.

Figure 5. (a) First galvanostatic charge/discharge curves of KHCF@PPy electrodes; (b) ex situ XRD patterns and the enlarged patterns (right column) at different charge and discharge states as marked in (a); (c) crystal structural evolution of KHCF@PPy based on the calculation results and the illustration of the calculation of energy barriers (table).

subsequent process, the decreased amount of residual water would give an enhanced CE value. The rate performance of KHCF@PPy at different current densities ranged from 50 to 1000 mA g−1 are displayed in Figure 4b. The reversible capacity at 100 mA g−1 was 80 mA h g−1, equal to approximately 95% of its capacity at a current density of 50 mA g−1. For larger current densities, such as 200, 500, and 1000 mA g−1, the specimen was still able to discharge 77, 69, and 60 mA h g−1, respectively, all of which are higher than the values obtained from KHCF (Figure S5). After cycling at high rates, the capacity of KHCF@PPy returned to 83 mA h g−1 when the current density was lowered back to 50 mA g−1, demonstrating minimal damage to the structure and the framework interface. The initial low CE gradually increased and stabilized at 98% along with an increase in the current density. Prolonged cycling at a high rate of 1000 mA g−1 was further investigated, and the results are presented in Figures 4c and S6. The KHCF@PPy material owned an initial discharge

enhanced CE may have originated from the lower coordinated water content and reduced defect concentration in KHCF@ PPy, as noted in the structural characterization discussion. During the subsequent long cycles, there was no significant voltage decay or polarization voltage increase in the case of KHCF@PPy, whereas KHCF exhibited poor kinetics by comparison (Figure 3d). The detailed electrochemical performance of KHCF@PPy is shown in Figure 4. After 500 cycles at a current density of 50 mA g−1, the KHCF@PPy electrode maintained 86.8% of its initial discharge capacity, with a value of 77.1 mA h g−1 (Figure 4a), whereas KHCF only delivered a capacity of 54.5 mA h g−1, equal to 60% of its initial value (Figure S4). The CE gradually increased to a stable level of above 98% after several cycles, which may have been related to the presence of interstitial water in the material.45 During the first several cycles, the decomposition of interstitial water during the charge process would be expected to lead to a low CE. Throughout the 22343

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Research Article

ACS Applied Materials & Interfaces capacity of 72.1 mA h g−1 and retained 61.8 mA h g−1 after 500 cycles with a capacity retention of 85% and a capacity decay rate of 0.03% per cycle. The CE was greater than 98% after several cycles, indicating a remarkable reversibility of K+ extraction and insertion. The enhanced rate capability can possibly be ascribed to the improved electronic conductivity conferred by the PPy coating, as demonstrated by the electrochemical impedance spectra in Figure 4d. In these data, KHCF@PPy produced a smaller semicircle than KHCF, indicating an enhanced electronic response and a decreased charge-transfer resistance. Ex situ XRD analyses were conducted to assess the phase evolution of KHCF@PPy during the charge/discharge process, and Figure 5b shows the detailed XRD profiles acquired at different charge/discharge states (as marked in Figure 5a). It can be seen that there are no obvious changes in the diffraction peaks during the entire charge/discharge process, suggesting no phase transition during the electrochemical redox process. In the enlarged patterns between the 2θ values of 34° and 37°, the (400) peaks shift to larger angles, suggesting that the lattice parameter was reduced during the charge process (A−C) because of K+ extraction. Conversely, when K+ ions were inserted into the crystal (discharge process, D,E), the lattice parameter increased, as reflected in the shift of the (400) peaks to smaller angles. Therefore, K+ storage in KHCF@PPy proceeded via a reversible interaction mechanism without twophase reactions. This was confirmed by theoretical calculations. On the basis of the DFT calculation results, if K+ ions are completely extracted from the monoclinic structure, it will transition to a cubic phase (Figure 5c), which is in conflict with the ex situ XRD results. Moreover, the energy barrier for a single K+ insertion into a K-free unit cell is calculated as 6.99 eV, which is higher than that for a K-rich structure (6.5 eV), as shown in the table in Figure 5. Consequently, it can be inferred that K+ ions are instead partially extracted from/inserted into the monoclinic crystal during the charge/discharge process. The residual K+ in the crystal structure can preserve the structural stability of the PB framework, which may be beneficial for the excellent cycling stability of KHCF@PPy.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Renjie Chen: 0000-0002-7001-2926 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 was supported by the National Key Research and Development Program of China “New Energy Project for Electric Vehicle” (2016YFB0100204), the National Natural Science Foundation of China (51772030), the Joint Funds of the National Natural Science Foundation of China (U1564206), Major achievements Transformation Project for Central University in Beijing, and Beijing Key Research and Development Plan (Z181100004518001).



REFERENCES

(1) Armand, M.; Tarascon, J.-M. Building Better Batteries. Nature 2008, 451, 652−657. (2) Choi, J. W.; Aurbach, D. Promise and Reality of Post-LithiumIon Batteries with High Energy Densities. Nat. Rev. Mater. 2016, 1, 16013. (3) Chen, R.; Zhao, T.; Zhang, X.; Li, L.; Wu, F. Advanced Cathode Materials for Lithium-Ion Batteries Using Nanoarchitectonics. Nanoscale Horiz. 2016, 1, 423−444. (4) Winter, M.; Barnett, B.; Xu, K. Before Li Ion Batteries. Chem. Rev. 2018, 118, 11433−11456. (5) Wang, F.; Wu, X.; Li, C.; Zhu, Y.; Fu, L.; Wu, Y.; Liu, X. Nanostructured Positive Electrode Materials for Post-Lithium Ion Batteries. Energy Environ. Sci. 2016, 9, 3570−3611. (6) Hwang, J.-Y.; Myung, S.-T.; Sun, Y.-K. Sodium-Ion Batteries: Present and Future. Chem. Soc. Rev. 2017, 46, 3529−3614. (7) Wang, H.-G.; Yuan, S.; Ma, D.-L.; Zhang, X.-B.; Yan, J.-M. Electrospun Materials for Lithium and Sodium Rechargeable Batteries: From Structure Evolution to Electrochemical Performance. Energy Environ. Sci. 2015, 8, 1660−1681. (8) Yuan, S.; Liu, Y.-B.; Xu, D.; Ma, D.-L.; Wang, S.; Yang, X.-H.; Cao, Z.-Y.; Zhang, X.-B. Pure Single-Crystalline Na1.1V3O7.9 Nanobelts as Superior Cathode Materials for Rechargeable Sodium-Ion Batteries. Adv. Sci. 2015, 2, 1400018. (9) Li, T.; Zhang, Q. Advanced Metal Sulfide Anode for Potassium Ion Batteries. J. Energy Chem. 2018, 27, 373−374. (10) Yuan, S.; Zhu, Y.-H.; Li, W.; Wang, S.; Xu, D.; Li, L.; Zhang, Y.; Zhang, X.-B. Surfactant-Free Aqueous Synthesis of Pure SingleCrystalline Snse Nanosheet Clusters as Anode for High Energy- and Power-Density Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1602469. (11) Sun, T.; Li, Z.-j.; Wang, H.-g.; Bao, D.; Meng, F.-l.; Zhang, X.-b. A Biodegradable Polydopamine-Derived Electrode Material for HighCapacity and Long-Life Lithium-Ion and Sodium-Ion Batteries. Angew. Chem., Int. Ed. 2016, 128, 10820−10824. (12) Tian, B.; Tang, W.; Su, C.; Li, Y. Reticular V2O5·H2O Xerogel as Cathode for Rechargeable Potassium Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 642−650.

4. CONCLUSIONS In this work, a K-rich Prussian blue cathode material (KHCF@ PPy) was synthesized by an in situ polymerization coating method. As a result of the PPy coating, the composite had a low defect concentration with enhanced electronic conductivity and exhibited excellent cycling stability and rate capability. The as-prepared electrode made from this compound showed a high capacity retention of 86.8% after 500 cycles at 50 mA g−1 and delivered an initial discharge capacity of 72.1 mA h g−1 while retaining 61.8 mA h g−1 after 500 cycles at 1000 mA g−1, with a capacity decay rate of 0.03% per cycle. Ex situ XRD combined with DFT calculations revealed that K+ storage in KHCF@PPy is associated with a reversible interaction mechanism without a phase transition, which is beneficial with regard to structural stability. This novel modification method provides a promising route to the application of PBbased cathodes in PIBs.



SEM images of KHCF and KHCF@PPy with low magnifications, XRD pattern of KHCF, TGA curves of KHCF@PPy and KHCF, cycling performance of pure KHCF, rate performance of pure KHCF, and iron-57 Mössbauer spectra parameters of KHCF@PPy (PDF)

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.9b04579. 22344

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345

Research Article

ACS Applied Materials & Interfaces (13) Eftekhari, A.; Jian, Z.; Ji, X. Potassium Secondary Batteries. ACS Appl. Mater. Interfaces 2017, 9, 4404−4419. (14) Lu, Y.; Chen, J. Robust Self-Supported Anode by Integrating Sb2S3 Nanoparticles with S,N-Codoped Graphene to Enhance KStorage Performance. Sci. China Chem. 2017, 60, 1533−1539. (15) Jian, Z.; Luo, W.; Ji, X. Carbon Electrodes for K-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 11566−11569. (16) Zhao, J.; Zou, X.; Zhu, Y.; Xu, Y.; Wang, C. Electrochemical Intercalation of Potassium into Graphite. Adv. Funct. Mater. 2016, 26, 8103−8110. (17) Luo, W.; Wan, J.; Ozdemir, B.; Bao, W.; Chen, Y.; Dai, J.; Lin, H.; Xu, Y.; Gu, F.; Barone, V.; Hu, L. Potassium Ion Batteries with Graphitic Materials. Nano Lett. 2015, 15, 7671−7677. (18) Pramudita, J. C.; Sehrawat, D.; Goonetilleke, D.; Sharma, N. An Initial Review of the Status of Electrode Materials for Potassium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1602911. (19) Deng, T.; Fan, X.; Luo, C.; Chen, J.; Chen, L.; Hou, S.; Eidson, N.; Zhou, X.; Wang, C. Self-Templated Formation of P2-Type K0.6CoO2 Microspheres for High Reversible Potassium-Ion Batteries. Nano Lett. 2018, 18, 1522−1529. (20) Ma, D.-l.; Wang, H.-g.; Li, Y.; Xu, D.; Yuan, S.; Huang, X.-l.; Zhang, X.-b.; Zhang, Y. In Situ Generated FeF3 in Homogeneous Iron Matrix toward High-Performance Cathode Material for Sodium-Ion Batteries. Nano Energy 2014, 10, 295−304. (21) Yuan, S.; Huang, X.-l.; Ma, D.-l.; Wang, H.-g.; Meng, F.-z.; Zhang, X.-b. Engraving Copper Foil to Give Large-Scale Binder-Free Porous Cuo Arrays for a High-Performance Sodium-Ion Battery Anode. Adv. Mater. 2014, 26, 2273−2279. (22) Qian, J.; Wu, C.; Cao, Y.; Ma, Z.; Huang, Y.; Ai, X.; Yang, H. Prussian Blue Cathode Materials for Sodium-Ion Batteries and Other Ion Batteries. Adv. Energy Mater. 2018, 8, 1702619. (23) Wang, L.; Lu, Y.; Liu, J.; Xu, M.; Cheng, J.; Zhang, D.; Goodenough, J. B. A Superior Low-Cost Cathode for a Na-Ion Battery. Angew. Chem., Int. Ed. 2013, 52, 1964−1967. (24) Huang, Y.; Xie, M.; Zhang, J.; Wang, Z.; Jiang, Y.; Xiao, G.; Li, S.; Li, L.; Wu, F.; Chen, R. A Novel Border-Rich Prussian Blue Synthetized by Inhibitor Control as Cathode for Sodium Ion Batteries. Nano Energy 2017, 39, 273−283. (25) Liu, Y.; Qiao, Y.; Zhang, W.; Li, Z.; Ji, X.; Miao, L.; Yuan, L.; Hu, X.; Huang, Y. Sodium Storage in Na-Rich NaxFeFe(CN)6 Nanocubes. Nano Energy 2015, 12, 386−393. (26) Song, J.; Wang, L.; Lu, Y.; Liu, J.; Guo, B.; Xiao, P.; Lee, J.-J.; Yang, X.-Q.; Henkelman, G.; Goodenough, J. B. Removal of Interstitial H2O in Hexacyanometallates for a Superior Cathode of a Sodium-Ion Battery. J. Am. Chem. Soc. 2015, 137, 2658−2664. (27) Wang, L.; Song, J.; Qiao, R.; Wray, L. A.; Hossain, M. A.; Chuang, Y.-D.; Yang, W.; Lu, Y.; Evans, D.; Lee, J.-J.; Vail, S.; Zhao, X.; Nishijima, M.; Kakimoto, S.; Goodenough, J. B. Rhombohedral Prussian White as Cathode for Rechargeable Sodium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 2548−2554. (28) Jiang, Y.; Yu, S.; Wang, B.; Li, Y.; Sun, W.; Lu, Y.; Yan, M.; Song, B.; Dou, S. Prussian Blue@C Composite as an Ultrahigh-Rate and Long-Life Sodium-Ion Battery Cathode. Adv. Funct. Mater. 2016, 26, 5315−5321. (29) Zhang, C.; Xu, Y.; Zhou, M.; Liang, L.; Dong, H.; Wu, M.; Yang, Y.; Lei, Y. Potassium Prussian Blue Nanoparticles: A Low-Cost Cathode Material for Potassium-Ion Batteries. Adv. Funct. Mater. 2017, 27, 1604307. (30) Bie, X.; Kubota, K.; Hosaka, T.; Chihara, K.; Komaba, S. A Novel K-Ion Battery: Hexacyanoferrate(Li)/Graphite Cell. J. Mater. Chem. A 2017, 5, 4325−4330. (31) Eftekhari, A. Potassium Secondary Cell Based on Prussian Blue Cathode. J. Power Sources 2004, 126, 221−228. (32) Chong, S.; Chen, Y.; Zheng, Y.; Tan, Q.; Shu, C.; Liu, Y.; Guo, Z. Potassium Ferrous Ferricyanide Nanoparticles as a High Capacity and Ultralong Life Cathode Material for Nonaqueous Potassium-Ion Batteries. J. Mater. Chem. A 2017, 5, 22465−22471. (33) Chen, L.; Chen, J.; Mizuno, F. First-Principles Study of Alkali and Alkaline Earth Ion Intercalation in Iron Hexacyanoferrate: The

Important Role of Ionic Radius. J. Phys. Chem. C 2013, 117, 21158− 21165. (34) He, G.; Nazar, L. F. Crystallite Size Control of Prussian White Analogues for Nonaqueous Potassium-Ion Batteries. ACS Energy Lett. 2017, 2, 1122−1127. (35) Xue, L.; Li, Y.; Gao, H.; Zhou, W.; Lü, X.; Kaveevivitchai, W.; Manthiram, A.; Goodenough, J. B. Low-Cost High-Energy Potassium Cathode. J. Am. Chem. Soc. 2017, 139, 2164−2167. (36) Su, D.; McDonagh, A.; Qiao, S.-Z.; Wang, G. High-Capacity Aqueous Potassium-Ion Batteries for Large-Scale Energy Storage. Adv. Mater. 2016, 29, 1604007. (37) Zhu, Y.-h. Transformation of Rusty Stainless-Steel Meshes into Stable, Low-Cost, and Binder-Free Cathodes for High Performance Potassium-Ion Batteries. Angew. Chem., Int. Ed. 2017, 56, 7881. (38) Wessells, C. D.; Peddada, S. V.; Huggins, R. A.; Cui, Y. Nickel Hexacyanoferrate Nanoparticle Electrodes for Aqueous Sodium and Potassium Ion Batteries. Nano Lett. 2011, 11, 5421−5425. (39) Deng, L.; Yang, Z.; Tan, L.; Zeng, L.; Zhu, Y.; Guo, L. Investigation of the Prussian Blue Analog Co3[Co(CN)6]2 as an Anode Material for Nonaqueous Potassium-Ion Batteries. Adv. Mater. 2018, 30, 1802510. (40) Padigi, P.; Thiebes, J.; Swan, M.; Goncher, G.; Evans, D.; Solanki, R. Prussian Green: A High Rate Capacity Cathode for Potassium Ion Batteries. Electrochim. Acta 2015, 166, 32−39. (41) Zhou, M.; Qian, J.; Ai, X.; Yang, H. Redox-Active Fe(CN)64‑Doped Conducting Polymers with Greatly Enhanced Capacity as Cathode Materials for Li-Ion Batteries. Adv. Mater. 2011, 23, 4913− 4917. (42) You, Y.; Yu, X.; Yin, Y.; Nam, K.-W.; Guo, Y.-G. Sodium Iron Hexacyanoferrate with High Na Content as a Na-Rich Cathode Material for Na-Ion Batteries. Nano Res. 2014, 8, 117−128. (43) Samain, L.; Gilbert, B.; Grandjean, F.; Long, G. J.; Strivay, D. Redox Reactions in Prussian Blue Containing Paint Layers as a Result of Light Exposure. J. Anal. At. Spectrom. 2013, 28, 524−535. (44) Shadike, Z.; Shi, D.-R.; Tian-Wang, T.-W.; Cao, M.-H.; Yang, S.-F.; Chen, J.; Fu, Z.-W. Long Life and High-Rate Berlin Green FeFe(CN)6 Cathode Material for a Non-Aqueous Potassium-Ion Battery. J. Mater. Chem. A 2017, 5, 6393−6398. (45) Lu, Y.; Wang, L.; Cheng, J.; Goodenough, J. B. Prussian Blue: A New Framework of Electrode Materials for Sodium Batteries. Chem. Commun. 2012, 48, 6544−6546.

22345

DOI: 10.1021/acsami.9b04579 ACS Appl. Mater. Interfaces 2019, 11, 22339−22345