13
Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buer Layers Rishabh Jain, ,Prateek Hundekar, ,Tao Deng, Xiulin Fan, Yashpal Singh, § Anthony Yoshimura, Varun Sarbada, Tushar Gupta, Aniruddha S. Lakhnot, Sang Ouk Kim, # Chunsheng Wang, and Nikhil Koratkar* ,,Department of Mechanical, Aerospace and Nuclear Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, United States Department of Chemical and Biomolecular Engineering, University of Maryland, College Park, Maryland 20742, United States § Material Science Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States Department of Physics, Applied Physics and Astronomy, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, United States Department of Materials Science and Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, United States # National Creative Research Initiative (CRI) Center for Multi-dimensional Directed Nanoscale Assembly, Department of Materials Science and Engineering, KAIST, Daejeon 34141, Republic of Korea * S Supporting Information ABSTRACT: High specic capacity materials that can store potassium (K) are essential for next-generation K-ion batteries. One such candidate material is phosphorene (the 2D allotrope of phosphorus (P)), but the potassiation capability of phosphorene has not yet been established. Here we systemati- cally investigate the alloying of few-layer phosphorene (FLP) with K. Unlike lithium (Li) and sodium (Na), which form Li 3 P and Na 3 P, FLP alloys with K to form K 4 P 3 , which was conrmed by ex situ X-ray characterization as well as density functional theory calculations. The formation of K 4 P 3 results in high specic capacity (1200 mAh g 1 ) but poor cyclic stability (only 9% capacity retention in subsequent cycles). We show that this capacity fade can be successfully mitigated by the use of reduced graphene oxide (rGO) as buer layers to suppress the pulverization of FLP. We studied the performance of rGO and single-walled carbon nanotubes (sCNTs) as buer materials and found that rGO being a 2D material can better encapsulate and protect FLP relative to 1D sCNTs. The half- cell performance of FLP/rGO could also be successfully reproduced in a full-cell conguration, indicating the possibility of high-performance K-ion batteries that could oer a sustainable and low-cost alternative to Li-ion technology. KEYWORDS: phosphorene anode, potassium-ion batteries, K 4 P 3 alloy, energy density, cycle stability R apid exploitation and depletion of nonrenewable energy resources and intermittent power generation from renewable resources makes ecient energy storage essential to our progress and prosperity. 1,2 Industry has to date primarily relied on lithium (Li)-ion batteries (LIBs) to meet the rapidly escalating global energy storage demand because of their higher energy density and longer cycle life relative to other competing technologies. 36 However, the battery industry is facing a bottleneck in widening the application of LIBs, which arises due to uneven distribution and scarcity of Li (0.0017 wt %) in the Earths crust. 7,8 This not only increases the production cost of LIBs but also raises a question on their ability to meet the burgeoning demand for energy storage. 9 This has stimulated interest in potassium (K)-ion batteries (KIBs) as a possible replacement to LIBs due to the following reasons: (i) low cost Received: August 22, 2019 Accepted: November 14, 2019 Published: November 14, 2019 Article www.acsnano.org Cite This: ACS Nano 2019, 13, 14094-14106 © 2019 American Chemical Society 14094 DOI: 10.1021/acsnano.9b06680 ACS Nano 2019, 13, 1409414106 Downloaded via UNIV OF MARYLAND COLG PARK on March 8, 2020 at 19:39:50 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

Reversible Alloying of Phosphorene withPotassium and Its Stabilization Using ReducedGraphene Oxide Buffer LayersRishabh Jain,†,¶ Prateek Hundekar,†,¶ Tao Deng,‡ Xiulin Fan,‡ Yashpal Singh,§ Anthony Yoshimura,∥

Varun Sarbada,⊥ Tushar Gupta,† Aniruddha S. Lakhnot,† Sang Ouk Kim,# Chunsheng Wang,‡

and Nikhil Koratkar*,†,⊥

†Department of Mechanical, Aerospace and Nuclear Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York12180, United States‡Department of Chemical and Biomolecular Engineering, University of Maryland, College Park, Maryland 20742, United States§Material Science Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States∥Department of Physics, Applied Physics and Astronomy, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180,United States⊥Department of Materials Science and Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, New York 12180, UnitedStates#National Creative Research Initiative (CRI) Center for Multi-dimensional Directed Nanoscale Assembly, Department of MaterialsScience and Engineering, KAIST, Daejeon 34141, Republic of Korea

*S Supporting Information

ABSTRACT: High specific capacity materials that can storepotassium (K) are essential for next-generation K-ion batteries.One such candidate material is phosphorene (the 2D allotropeof phosphorus (P)), but the potassiation capability ofphosphorene has not yet been established. Here we systemati-cally investigate the alloying of few-layer phosphorene (FLP)with K. Unlike lithium (Li) and sodium (Na), which form Li3Pand Na3P, FLP alloys with K to form K4P3, which was confirmedby ex situ X-ray characterization as well as density functionaltheory calculations. The formation of K4P3 results in highspecific capacity (∼1200 mAh g−1) but poor cyclic stability(only ∼9% capacity retention in subsequent cycles). We showthat this capacity fade can be successfully mitigated by the useof reduced graphene oxide (rGO) as buffer layers to suppressthe pulverization of FLP. We studied the performance of rGO and single-walled carbon nanotubes (sCNTs) as buffermaterials and found that rGO being a 2D material can better encapsulate and protect FLP relative to 1D sCNTs. The half-cell performance of FLP/rGO could also be successfully reproduced in a full-cell configuration, indicating the possibilityof high-performance K-ion batteries that could offer a sustainable and low-cost alternative to Li-ion technology.KEYWORDS: phosphorene anode, potassium-ion batteries, K4P3 alloy, energy density, cycle stability

Rapid exploitation and depletion of nonrenewable energyresources and intermittent power generation fromrenewable resources makes efficient energy storage

essential to our progress and prosperity.1,2 Industry has to dateprimarily relied on lithium (Li)-ion batteries (LIBs) to meet therapidly escalating global energy storage demand because of theirhigher energy density and longer cycle life relative to othercompeting technologies.3−6 However, the battery industry isfacing a bottleneck in widening the application of LIBs, whicharises due to uneven distribution and scarcity of Li (0.0017 wt

%) in the Earth’s crust.7,8 This not only increases the productioncost of LIBs but also raises a question on their ability to meet theburgeoning demand for energy storage.9 This has stimulatedinterest in potassium (K)-ion batteries (KIBs) as a possiblereplacement to LIBs due to the following reasons: (i) low cost

Received: August 22, 2019Accepted: November 14, 2019Published: November 14, 2019

Artic

lewww.acsnano.orgCite This: ACS Nano 2019, 13, 14094−14106

© 2019 American Chemical Society 14094 DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

Dow

nloa

ded

via

UN

IV O

F M

AR

YL

AN

D C

OL

G P

AR

K o

n M

arch

8, 2

020

at 1

9:39

:50

(UT

C).

See

http

s://p

ubs.

acs.

org/

shar

ingg

uide

lines

for

opt

ions

on

how

to le

gitim

atel

y sh

are

publ

ishe

d ar

ticle

s.

Page 2: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

and higher abundancy of potassium in the Earth’s crust (2.09 wt%), (ii) similar working mechanism (rocking chair concept) toLIBs, and (iii) almost similar reduction potential (−2.97 V vsE°) in comparison with Li (−3.04 V vs E°).10−13 However, thespecific capacity and energy density of KIBs are still lackingwhen compared to LIBs, which warrants the development ofimproved anode and cathode materials for KIB applica-tions.14−16

Researchers have to date explored several cathode materialsfor KIBs among which Prussian blue and its analogues,17−19

layered oxides based on manganese,20 vanadium,21 and cobaltanions,22 polyanionic compounds,23 and organic compounds24

show high specific capacity and long cycle life. In the case ofanode materials, various carbonaceous materials (e.g., graphite,hard carbons, soft carbons, hard−soft carbons, doped graphene)have been reported as active materials for intercalation/deintercalation of K+ ions and show reasonable specificcapacities of 200−300 mAh g−1.25−32 However, in order tomatch LIB performance, further improvements in anodeperformance are necessary. One way to achieve this is to explorealloys (e.g., tin (Sn), antimony (Sb), bismuth (Bi), andphosphorus (P)), since alloy-based chemistries offer superiorspecific capacity when compared to the traditional K-ionintercalation chemistry.33−36 Among alloy-based materials, Pshows the highest theoretical capacity of∼2500 mAh g−1, whichis attributed to K3P formation.37 This led to extensive researchover red phosphorus (RP) as anode materials for KIBs.38−40

However, to date there is large scatter in the reported data. Infact, depending on the conditions used for RP synthesis and

processing, a variety of alloys including KP,38,39 K2P3,39 K3P,

37

and K4P340 have been reported in the literature. The ambiguity

in electrochemical data of red P can be due to a variety of factorssuch as its amorphous structure, processing conditions, and thelevel of purity of the material (red P generally contains traces ofwhite phosphorus). It is difficult to pinpoint the relativeimportance of these factors, but in our view the amorphousnature of red P plays a significant role. In this work, as analternative to RP, we chose to focus on the potassiation anddepotassiation behavior of “phosphorene”, whose structure ishighly crystalline and well-defined.41 In fact, phosphorene hasproved to be a superior anode material and had shown higherspecific capacity and rate capability than RP in Li and Na ionbatteries, which is attributed to its higher specific surface area,higher electrical conductivity, and nanolayer (two-dimensional,2D) structure.42−45 However, to the best of our knowledge,phosphorene-based materials have not yet been explored inKIBs.Herein, we report the fabrication and electrochemical

characterization of KIBs based on few-layer phosphorene(FLP) anodes. FLP was synthesized in our work by liquidphase exfoliation of bulk black phosphorus (BP). Reducedgraphene oxide (rGO) and single-walled carbon nanotubes(sCNTs) were used as a buffer layer to accommodate thevolume expansion caused by alloy formation. Most interestingly,ex situ X-ray photoelectron spectroscopy (XPS), X-raydiffraction (XRD), and density functional theory (DFT)calculations all indicate that FLP reacts with P to form theK4P3 alloy phase, which is very distinct from what has been

Figure 1. Characterization of as-produced few-layer phosphorene (FLP) sheets. (a) Schematic illustration of exfoliation of FLP in NMPsolution. (b) SEM image of FLP deposited on a Si/SiO2 wafer (inset shows AFM image; scale bar = 500 nm). (c) Typical TEM image of FLP. (d)Representative HRTEM image of FLP (inset shows SAED pattern). (e) Raman spectroscopy and (f) high-resolution P 2p XPS spectra of FLP.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14095

Page 3: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

reported for Li-ion (Li3P) and Na-ion (Na3P) batteries.42,43 In

the absence of a mechanically robust carbon buffer layer, thespecific capacity of the FLP anode drops precipitously after a fewcycles due to volume expansion driven pulverization of theelectrode material. By proper selection of the geometry of thecarbon buffer material (e.g., 2D rGO vs 1D sCNTs) and itsloading content, the FLPmaterial showed drastic improvementsin its cycle stability. Furthermore, we demonstrated the viabilityof FLP by employing the FLP/rGO anode in a working full-cellwherein spherical potassium cobalt oxide (s-KCO)46 was usedas a cathode. FLP/rGO anode in FLP/rGO||s-KCO full-cells

delivered high specific capacity (∼420 mAh g−1 at 0.5 C),reasonably good rate capability (∼180 mAh g−1 at 2.5C and∼150 mAh g−1 at 4C), and stable cycle life (∼230 mAh g−1 at0.5C for 100 cycles). Such a working full-cell demonstration of aKIB with an FLP anode indicates the promise of this material forpractical applications.

RESULTS AND DISCUSSIONSynthesis of Few-Layer Phosphorene. In order to

synthesize FLP, bulk black phosphorus was exfoliated inanhydrous N-methyl-2-pyrrolidone (NMP) as depicted in

Figure 2. Characterization of FLP/rGO and FLP/sCNTs composites. (a) Schematic illustration of synthesis of FLP/rGO and FLP/sCNTs.Representative TEM image of (b) FLP/rGO and (c) FLP/sCNTs (inset shows SAED pattern). High-resolution (d, e) P 2p andC 1s XPS spectraof FLP/rGO and (f, g) P 2p and C 1s XPS spectra of FLP/sCNTs. (h) Raman spectra of FLP/rGO and FLP/sCNTs.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14096

Page 4: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

Figure 1a. As water plays an important role in oxidation ofphosphorus, anhydrous organic solvent was used for exfoliation

in order to prevent degradation of the as-synthesizedphosphorene sheets.44 Among various organic solvents, NMP

Figure 3. Electrochemical performance of FLP/rGO (1:3) cycled in a half-cell configuration against K metal. (a) Schematic illustration of thehalf-cell configuration. (b) CV of the FLP/rGO (1:3) electrode at a scanning rate of ∼0.2 mV s−1. Broad peaks in the CV are indicated by theyellow bands. (c) Electrochemical charge/discharge voltage profile of FLP/rGO (1:3) at a current density of ∼0.1C. (d) Cyclic performanceand Coulombic efficiency of FLP/rGO (1:3), FLP/sCNTs (1:3), FLP/rGO (1:1), only rGO, and only FLP at a current density of ∼0.1C. (e)Rate capability of FLP/rGO (1:3) at different current densities, increasing from∼0.1C to∼1.2C. (f) Electrochemical charge/discharge voltageprofile of FLP/rGO (1:3) at different current densities, increasing from ∼0.1C to ∼1.2C. (g) Cycling performance of FLP/rGO (1:3) at acurrent density of ∼0.5 C. It should be noted that for all tests a charge−discharge rate of 1C corresponds to a current density of ∼1 A g−1.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14097

Page 5: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

was used for the exfoliation of BP due to its higher exfoliationefficacy and its ability to prevent rapid degradation ofphosphorene in ambient condition through a solvation shell.47

Scanning electron microscopy (SEM) and transmission electronmicroscopy (TEM) images (Figure 1b,c) show synthesis of few-layer phosphorene sheets with sharp edges, with an average sizeof 500 nm to 2 μm, and no visible signs of ambient degradation.An atomic force microscopy (AFM) image (inset of Figure 1b)further confirms production of FLP with an average thickness of3−8 nm, corresponding to ∼6−10 layers of phosphorene(assuming that the interlayer spacing in BP is ∼0.53 nm). Well-aligned lattice fringes observed through high-resolution trans-mission electron microscopy (HRTEM) and confirmation oforthorhombic crystal structure through selected area electrondiffraction pattern (SAED) proved that the liquid phaseexfoliation method did not alter the structure of phosphorene(Figure 1d). The lattice constants (3.26 and 4.4 Å) measuredthrough HRTEM are also consistent with previously reportedvalues.48 Scanning transmission electron microscopy (STEM)established that the two-dimensional sheets are composed ofphosphorus atoms (Figure S1). Raman spectroscopy (Figure1e) was further employed to confirm the successful exfoliation ofblack phosphorus. Bulk black phosphorus generally shows threemajor unique phonon modes at ∼359 cm−1 (Ag

1), ∼434 cm−1

(B2g), and ∼461 cm−1 (Ag2), which exhibit a slight right shift

after exfoliation (among which the Ag2 shift (∼3.0 cm−1) is

major), confirming the decrease in the thickness of BP.49 Thehigh ratio of Ag

1/Ag2 (>0.5) further confirms minimal oxidation

of the FLP sheets.49 Lastly, XPS was used to investigate thechemical composition of the as-synthesized phosphorene sheets.High-resolution P2p XPS analysis shows a major peak with ashoulder between 129 and 131 eV, which is attributed to the P−P bond, whereas a minor peak between 132 and 134 eV isattributed to the P−O bond (Figure 1f). This slight oxidation ofphosphorene is to be expected since the exfoliation andcentrifugation processes were carried out under ambientconditions.Synthesis of Phosphorene−Carbon Composites. Two

categories of graphitic materials with similar density (∼1.7−1.9g cm−3), reduced graphene oxide (2D) and single-walled carbonnanotubes (1D), were mixed with FLP. This was done for threereasons. One reason is to mechanically buttress the FLP sheetsand protect them from pulverization during alloying/dealloyingwith K. This is critically important because large volume changes(theoretical volume expansion of K4P3 is ∼411.78%) cangenerate significant stresses in the material, making it necessaryto protect the material. The other reasons are to increase theelectrical conductivity of the composite electrode and to providea stable solid electrolyte interface (SEI). Phosphorene−reducedgraphene oxide (FLP/rGO) and phosphorene−single-walledcarbon nanotubes (FLP/sCNTs) composites were synthesizedby mixing the graphitic materials and phosphorene in an NMPdispersion (Figure 2a), followed by vacuum filtration (seeMethods section). This results in a sandwich structure whereinphosphorene sheets are encapsulated with graphitic materials(Figure 2b,c). TEM images of FLP/rGO and FLP/sCNTsclearly show the presence of FLP sheets underneath wider rGOsheets (Figure 2b) and a network of sCNTs (Figure 2c). TheSAED pattern further verifies the presence of both phosphoreneand rGO/sCNTs in the composite structure (insets of Figure 2band c). The high degree of transparency of the FLP observed inTEM images indicates that self-assembly and restacking of theFLP was hindered by the rGO/sCNTs. High-resolution P 2p

and C 1s XPS analysis of FLP/rGO and FLP/sCNTs shows theabsence of any additional peaks, which confirms only secondary(i.e., van derWaals) interactions between FLP and rGO/sCNTsand eliminates the possibility of any chemical reaction betweenthem (Figure 2d−g). While the functional groups present inrGO and sCNT are almost identical, the relative intensity of thefunctional groups with respect to the C−C bond intensity isdifferent in rGO and sCNT. While this could affect thenonbonding (i.e., van der Waals) interaction between thematerials, XPS analysis clearly shows that there is no directchemical bonding between the FLP sheets and the nanocarbonmaterials in our electrode. The increase in intensity of the P−Osignature in FLP/rGO and FLP/sCNTs composites incomparison with as-synthesized phosphorene can be due tothe oxidation of phosphorene sheets during vacuum filtration,slurry coating, and electrode preparation (all processes wereconducted in ambient conditions). Raman spectroscopy ofFLP/rGO and FLP/sCNTs shows the presence of distinct D, G,2D, and G+Dmodes of rGO and sCNTs, as well as Ag

1, B2g, andAg

2 modes of phosphorene, which also confirms successfulsynthesis of the FLP/carbon composite structure (Figure 2h).

Electrochemical Testing in Half-Cell Configuration.Electrodes were prepared by slurry coating a mixture ofcomposite powder, super-P carbon black, and polyvinylidenefluoride (PVDF) on copper foil (see Methods section). Withpotassium hexafluorophosphate (KPF6) in ethylene carbonate/diethyl carbonate (EC/DEC) as electrolyte and potassium foilas reference/counter electrode, half-cells were configured asshown in Figure 3a. In addition to pure FLP, FLP/rGO andFLP/sCNTs electrodes in the same weight proportion (1:3)were prepared. Electrodes with 1:1 weight proportion of FLPand rGO were also synthesized. The K-ion storage mechanismof the FLP/rGO (1:3) electrode was investigated through cyclicvoltammetry (CV) in the voltage range of 0.03−2.3 V at 0.2 mVs−1 scan rate (Figure 3b). Broad peaks located at∼0.4 and∼0.65V appeared in the first cycle during the discharge process butdisappeared in subsequent cycles. This is attributed todecomposition of the electrolyte, which leads to the formationof the SEI layer over the FLP/rGO electrode. Formation of theSEI layer is further verified by the voltage profile of the dischargecurve in the first (formation) cycle wherein similar plateaus andslope shapes appeared as in the CV scan (Figure S2a). Insubsequent cycles, CV peaks located at ∼0.1, ∼0.2, and ∼0.7 V(Figure S2b) are observed during the discharge process and areattributed to the stepwise potassiation process of FLP and theintercalation of K in rGO (to form KC8). Similarly, in the chargeprocess, the appearance of three broad peaks at ∼0.15, ∼0.45,and ∼0.75 V (Figure S2b) are due to deintercalation ofpotassium from rGO and stepwise depotassiation of FLP. Theoverlapping of CV profiles in Figure 3b (after the third cycle iscompleted) indicates good electrochemical stability of the FLP/rGO electrode. Stable and reversible potassiation/depot-assiation in the FLP/rGO electrode is further verified byoverlapped charge/discharge curves in subsequent cycles afterthe initial capacity loss (see Figure 3c). The mechanisms fordepotassiation/potassiation of phosphorene and intercalation/deintercalation of rGO are further verified by the presence ofsimilar plateaus and slope shape in the voltage profiles of thecharge/discharge curves in subsequent cycles, as indicated inFigure 3c. It should be noted that the charge/discharge voltageprofile of a pure rGO electrode (without FLP) (Figure S3),fabricated using the same procedure as FLP/rGO, showedintercalation and deintercalation at ∼0.2 and ∼0.45 V,

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14098

Page 6: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

respectively. This allows us to differentiate between theelectrochemical signatures related to K-alloy formation andthose caused by K-intercalation in rGO. Similarly CV data takenon pure rGO (without FLP) and pure FLP (without rGO)electrodes (Figure S4) allow us to distinguish the electro-chemical signatures associated with K-alloying and intercalationwith FLP vs K-intercalation into rGO.It should be noted that due to the bulkiness and slow kinetics

of the larger K ion, broad peaks are observed in the CV plots.This is in contrast to Li, for which much sharper peaks have beenreported.43 The broad peaks that we report in our work are alsoconsistent with what has been reported in the literature forpotassium.44,45 The broadness of the response makes itchallenging to precisely identify the peaks for our K-ion batterysystem. Therefore, for better interpretation of the results wehave computed the dQ/dV plots for the FLP/rGO (1:3) andpure rGO electrodes. The dQ/dV plots for FLP/rGO (1:3) at0.1C for discharge and charge steps are shown in Figure S5a,b.During the discharge step, a broad prominent peak is observedbetween 0.1 and 0.25 V, and a small peak is observed at ∼0.7 V.Similarly, during the charge step a broad prominent peak isobserved between 0.15 and 0.5 V and a small peak is observed at∼0.75 V. The corresponding dQ/dV plots for pure rGO (at0.1C) for discharge and charge steps are shown in Figure S5c,d.We can observe peaks at ∼0.2 V during discharge and ∼0.45 Vduring charge. By comparing Figure S5a,b and Figure S5c,d, weconclude that potassium intercalation and deintercalation intoFLP occurs at ∼0.7 and ∼0.75 V respectively, while the alloyingand dealloying reactions of FLP with potassium take place at∼0.1 and ∼0.15 V, respectively. Intercalation and deintercala-tion of potassium into rGO appears to take place at ∼0.2 and∼0.45 V, respectively. We emphasize that given the broadness ofthe peaks these are ballpark estimates rather than precise values.In addition to CV and dQ/dV analysis, we also used the

galvanostatic intermittent titration technique (GITT) to furtherstudy the reaction chemistry. More specifically, we haveperformed GITT measurement of an FLP/rGO (1:3) half-cellat a C rate of 0.1C after four cycles of activation (Figure S6a).During GITT measurement, pulse current was provided for 1 hfollowed by a considerably long relaxation time of 5 h forobtaining the quasi-open-circuit potential. The reactionresistance was evaluated from GITT by measuring theoverpotential (i.e., potential drop between termination ofrelaxation step and initiation of current step) at each currentpulse or step and dividing it with the pulse current. Figure S6bindicates that the reaction resistance of the FLP/rGO electrodecontinually decreases during the discharge process. This is dueto the formation of the KxP alloy, which has better conductivitythen FLP. As expected, the reaction resistance in the chargeprocess increases due to reversibility of alloy formation (i.e.,conversion of KxP alloy back to FLP). We approximated the K+

chemical diffusion coefficient for potassiation and depot-assiation of FLP/rGO (1:3) based on a simplified version ofFick’s second law of diffusion (which assumes that the cellvoltage varies linearly with τ1/2; see inset of Figure S6c):50

τ=πτ

ΔΔ

≪τ

Dm VM S

EE

LD

4 B m

B

2s

2 2ikjjjjj

y{zzzzz

ikjjjjj

y{zzzzz

ikjjjj

y{zzzz

where D is the diffusion coefficient, mB is the active mass of theelectrode, Vm is the molar volume of the electrode, MB is themolar mass of the electrode, S is the electrode area,ΔEs andΔEτ

are obtained from GITT measurements (inset of Figure S6a),

and τ is the duration of the current pulse. Due to the bigger sizeof the K ion, the DK‑ion value, calculated based on the geometricarea of the electrode, for both discharge and charge processes isquite low and lies in the range of 10−11−10−13 cm2s−1 (FigureS6c,d). This DK‑ion value is 2−3 orders of magnitude lower thanthe average diffusion coefficient for Li- and Na-ion diffusion intophosphorus and rGO-based materials (10−9−10−11

cm2s−1),51−56 which confirms the sluggish K-ion diffusionkinetics in the FLP/rGO electrode material.During discharge, the diffusion coefficient shows three

minima at ∼0.1, 0.2, and 0.7 V, which is consistent with ourCV (Figure S4) and dQ/dV (Figure S5) results. The diffusionslows down at these voltages due to alloying of FLP with K (at∼0.1 V), intercalation of K into FLP (at ∼0.7 V), andintercalation of K into rGO (at ∼0.2 V). Similarly, in the chargeprocess, the diffusion coefficient is low at ∼0.2, ∼0.4, and ∼0.8V, which is also consistent with the aforementioned CV and dQ/dV results. The drop in diffusion rate at these voltages suggestsdealloying of K from FLP (at∼0.2 V), deintercalation of K fromFLP (at ∼0.8 V), and deintercalation of K from rGO (at ∼0.4V).To confirm the GITT prediction of sluggish diffusion kinetics

of K, we also carried out electrochemical impedance spectros-copy (EIS) on an FLP/rGO (1:3) half-cell after five cycles ofactivation. Figure S7 shows the Nyquist plot of the FLP/rGO(1:3) cell within a frequency range of 0.1 Hz to 100 kHz (insetshows the equivalent circuit model used to fit the data). Thedepressed semicircle in the Nyquist plot is due to impedancefrom the solid electrolyte interface (resistance R2 andcapacitance C2) at high frequency and charge transfer(resistance R3 and capacitance C3) at middle frequency. Theintercept of the Nyquist plot at the x axis in the high-frequencyrange determines the electrolyte resistance (R1). The linearslope in the Nyquist plot in the low-frequency range is attributedto Warburg impedance (W1). Using the equivalent circuitdiagram, electrolyte resistance (∼20 Ω), solid electrolyteresistance (∼465 Ω), and charge transfer resistance (∼705 Ω)were obtained by fitting the model to the test data. Both solidelectrolyte resistance and charge transfer resistance are muchhigher in our battery when compared to phosphorus- andcarbon-based Li-ion and Na-ion batteries (∼50−150 Ω).42,57,58This is due to the relatively low ionic conductivity of K+ ions inthe SEI layer and relatively slower diffusion of K+ in the FLP/rGO electrode, which can be attributed to the larger size of Kcompared to Li and Na ions.59 Note that such high resistanceswere also observed in other K-ion batteries based on carbon andother alloy materials (e.g., red P, Bi, Sn, Sn4P3, SnSb, andFeP).43,59−62

Figure 3d shows galvanostatic charge/discharge cyclingresponses of FLP/rGO (1:3), FLP/sCNTs (1:3), FLP/rGO(1:1), pure rGO, and pure FLP electrodes at a charge/dischargerate (C rate) of ∼0.1C (where, 1C = ∼1A g−1). After SEIformation, FLP/rGO (1:3) showed a high first dischargecapacity of ∼910 mAh g−1 (initial Coulombic efficiency (CE)∼81%), which is superior to FLP/sCNTs (1:3) (613 mAh g−1;initial CE ∼50%) and rGO (∼200 mAh g−1; initial CE ∼70%)but lower than FLP/rGO (1:1) and FLP, which achievedcapacities of∼950 mAh g−1 (initial CE∼40%) and ∼1200 mAhg−1 (initial CE ∼30%), respectively. In all our work, thegravimetric capacity of FLP/rGO and FLP/sCNTs has beencalculated by subtracting the contribution of rGO and sCNTs,respectively (Figures S8 and S9) and then normalizing by FLPmass. The FLP/rGO (1:3) electrode delivers a specific capacity

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14099

Page 7: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

of ∼650 mAh g−1 in the first 15 cycles and maintains a capacityof ∼550 mAh g−1 up to 50 cycles at a current density of 0.1C,showing ∼74% of capacity retention. The slight capacity fade inthe first few cycles for the FLP/rGO (1:3) electrode can be dueto irreversible processes (alloying/dealloying is not 100%reversible, especially in first few cycles). In contrast to FLP/rGO (1:3), FLP/sCNTs (1:3) shows rapid decay in capacity to∼200 mAh g−1 over the initial 15 cycles and then maintains acapacity of ∼160 mAh g−1 up to 50 cycles at the C rate of 0.1C,showing only 43% capacity retention. This large difference incapacity retention between the FLP/rGO (1:3) and FLP/sCNTs (1:3) electrode can be attributed to the difference insandwiching/packing of 2D−2D sheets and 1D−2D sheets. Inthe case of FLP/rGO, rGO tends to cover (Figure 2b) the entireFLP sheet (sizes of rGO and FLP sheets were ∼10 μm and 500nm to 2 μm, respectively) and hence is able to suppress thefracture/pulverization of the FLP by providing a stronger elasticbuffer layer. In the case of FLP/sCNTs, due to their 1Dstructure, sCNT networks that form over the FLP surface areunable to adequately protect the FLP sheets. The sCNTs form aporous structure (Figure 2c), which allows pulverized FLPparticles to escape from the electrode, resulting in a loss of activematerial.40 A similar effect was also observed when we reducedthe proportion of rGO in the FLP/rGO electrode. For instance,reducing the rGO:FLP weight ratio to 1:1 results in a rapidcapacity fade with only∼9% capacity retention after 50 cycles ata C rate of ∼0.1C. This can be attributed to unavailability ofsufficient buffer layer material to buttress the FLP sheets.Without the use of rGO, the capacity of the pure FLP electrodedecayed precipitously from 1200 mAh g−1 to below 50 mAh g−1

in less than 10 charge−discharge steps, indicating the pivotalrole played by the rGO buffer material in stabilizing theperformance of the FLP. In order to verify that rGO bufferingcan protect the FLP material, ex situ SEM imaging before andafter cycling for FLP/rGO (1:3) and FLP/sCNT (1:3) wascarried out. FLP/rGO (1:3) showed high structural stability, asthere was no indication of cracks or delamination of materialfrom the copper current collector (Figure S10). By contrast, theFLP/sCNTs (1:3) showed poor structural stability, as largecracks were observed and the electrode material delaminated,revealing the underlying copper current collector substrate(Figure S11). This confirmed that the 2D rGO buffer layer isable to suppress the pulverization and delamination of FLPmuch more effectively than 1D sCNTs.It should be noted that the CE of FLP/rGO (1:3) running at

0.1C is low during initial cycles, but quickly increase after 5cycles. This is indicative of SEI instability especially at lowervoltages (below 0.1 V) during initial cycles. Owing to the lowelectrochemical potential of K+/K, the solvent in the electrolytecan be readily reduced on the electrode surface.63 That beingsaid, the CE of FLP/rGO (1:3) did climb to ∼97% at the end of50 cycles of charge and discharge at 0.1C rate (Figure 3d). Infact, among all the electrodes that we tested, FLP/rGO (1:3)showed by far the highest CE. Further, the effect of unstable SEIduring the initial cycles can be mitigated at higher C rates asobserved in our work (CE of FLP/rGO reached∼99%within 25cycles at a faster rate of 0.5C (Figure 3g)). This is because athigher rates there is less time available for electrolyte breakdownand undesirable side reactions to take place. The CE duringinitial cycling could potentially be further improved byimprovements to the electrolyte, which should be pursued aspart of future work.

The rate capability of the FLP/rGO (1:3) anode wasinvestigated at different C rates as shown in Figure 3e, and itscorresponding charge/discharge voltage profiles are shown inFigure 3f. The FLP/rGO (1:3) anode delivered reversiblecapacities of ∼710, ∼600, and ∼550 mAh g−1 at rates of ∼0.1C,∼0.2C, and ∼0.4C, respectively. Even at higher rates of ∼0.6Cand ∼1.2C, FLP/rGO displayed a reasonable capacity of ∼400and ∼230 mAh g−1, respectively. When the C rate was revertedback to ∼0.1C, a charge capacity of ∼700 mAh g−1 was retainedafter 30 cycles, revealing reversible alloying/dealloying of theFLP and stable cycle life. On the other hand, the pure rGOelectrode showed much lower reversible capacities of ∼150,∼110, ∼80, ∼65, and ∼40 mAh g−1 at current rates of ∼0.1C,∼0.2C, ∼0.4C, ∼0.6C, and ∼1.2 C, respectively (Figure S8b).As mentioned previously, for all the specific capacitycalculations, the contribution of rGO was deducted in orderto establish the capacity delivered by the FLP. Similar shape andslope of charge/discharge voltage profiles, with reducedoverpotential, at different current rates further confirms thereversibility of the alloying/dealloying reaction. The FLP/rGO(1:3) electrode also delivers relatively stable performance(Figure 3g) at higher C rates. For example, at a current rate of∼0.5 C and after initial capacity losses, the FLP/rGO electrodedisplayed a high charge capacity of ∼450 mAh g−1 with >99%Coulombic efficiency. After cycling for 300 cycles, FLP/rGO(1:3) was still able to deliver a capacity of∼230 mAh g−1, whichindicates ∼0.16% capacity decay per cycle (Figure 3g). On thebasis of these results, we conclude that the rGO material servestwo important functions in our electrode: (1) rGO acts as anelastic buffer layer helping to accommodate the high volumeexpansion and prevent the pulverization/delamination of theFLP electrode and (2) rGO facilitates the transport of electrons(generated during the alloying reaction) to the copper currentcollector.One observation from the rate capability testing is that the

high rate performance of FLP/rGO (1:3) is not exceptional andneither is that of rGO alone. The main reason for this is thesluggish intercalation/deintercalation and alloying/dealloyingkinetics of K ions as compared to that of Li. The reason for thissluggishness arises from the larger size of the K ion relative tothat of Li. In fact, the rate capability performance of FLP/sCNTs(Figure S9) was similar to that of FLP/rGO. This confirms thatthe “intrinsic sluggishness” of alloying and intercalation of K isthe primary reason that the specific capacity drops significantlywith the C rate.

Alloying Mechanism in Phosphorene. Phosphorus hasthe ability to form various alloy states with potassium (e.g., KP,K2P3, K3P, and K4P3). Initial testing of RP in KIBs reportedformation of only the KP phase.38,39 However, Wu et al. recentlyreported the existence of the K4P3 phase while investigating RPencapsulated in dual-doped porous carbon nanofibers.40 Incontrast to this, other studies (e.g., Li et al.37) have reported theexistence of the K3P phase. A possible reason for the largediversity in alloying phases that have been reported to date couldbe the predominantly amorphous nature of RP, which lackslong-range order, and hence a variety of alloying states arepossible based on the experimental conditions, purity, initialstructure, and degree of crystallinity of the RP that was used.Indeed a mix of different alloying states is also possible in suchpredominantly amorphous systems. On the other hand, FLP,which is highly crystalline, and has a well-defined initialstructure, may offer a more deterministic alloying response.To study this, ex situ XPS and XRD measurements were carried

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14100

Page 8: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

out. Two types of samples, electrodes after discharge to∼0.03 Vand after charge to ∼2.3 V, were initially sputtered with Ar+ forthe removal of the SEI layer on the top of the electrode, so thatdetection of alloy formation beneath it could be possible.Figure 4a−f show high-resolution C 1s, K 2p, and P 2p XPS

analysis of discharged and charged FLP/rGO (1:3) electrodes.Two important observations can be made while comparing XPSdata of the discharged and charged electrodes: (i) shifting ofmajor peaks and (ii) the appearance of additional peaks. In C 1sXPS spectra (Figure 4a and d) of both discharged and charged

materials, C−C, CO/C−O, and C−F bonds appear at∼284.6, ∼286, and ∼288 eV, respectively which is mainlyattributed to rGO and PVDF binder. The small broad peakappearing between 282 and 284 eV is attributed to theintercalation of K ions in rGO (forming KC8). The relativeintensity of the KC8 peak decreases in the charged electrode,indicating deintercalation of K from rGO. However, thepresence of the KC8 peak in the charged electrode indicatesthat intercalation/deintercalation is not 100% reversible, andthis can contribute to capacity loss. In K 2p XPS spectra (Figure

Figure 4. Ex situ XPSmeasurement and DFT calculations. (a−f) High-resolution XPS analysis of (a and d) C 1s, (b and e) K 2p, and (c and f) P2p after (a−c) discharge (0.03 V) and (d−f) charge (2.3 V). (g) Formation energies of different alloys calculated by DFT.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14101

Page 9: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

4b), the discharged electrode shows two peaks in doublets(2p3/2 and 2p1/2) between 292−296 eV and 294−297 eV,corresponding to KxO/KC8 and KxF, respectively which isattributed to trace amounts of SEI leftover on the FLP/rGOelectrode. This is because Ar+ sputtering is not 100% effective inremoving the SEI layer. Similar peaks in doublets also appear inthe charged electrode (Figure 4e) but are right shifted, whichindicates the loss of electrons by potassium while going from thedischarged to the charged state. A decrease in the relative

intensity of KxO/KC8 in the charged state can be attributed todeintercalation of K ions from rGO and is consistent with the C1s XPS spectra (Figure 4a and d).Most importantly, the discharged electrode exhibits a

prominent peak (also in a doublet) between 289−292 eV,which is attributed to “alloy formation (KxP)” during thedischarge process (Figure 4b). According to the NIST database,such a peak can correspond to either K4X or K3X, wherein X is acounterion. This indicates the possibility of only K4P3 or K3P

Figure 5. Demonstration of full-cell KIBs based on s-KCO||FLP/rGO (1:3). (a) Charge/discharge curves for the FLP/rGO (1:3)||K and s-KCO||K half-cell configurations. (b) Electrochemical charge/discharge voltage profile of the full-cell KIB at a current density of ∼0.5C. (c)Charge−discharge cycling performance of the full-cell KIB at a current density of ∼0.5C. (d) Rate capability of the full-cell KIB at differentcurrent densities, increasing from 0.5C up to 4C. (e) Electrochemical charge/discharge voltage profile of the full-cell KIB at different currentdensities, from 0.5C to 4C. It should be noted that for all tests a charge−discharge rate of 1C corresponds to a current density of ∼1 A g−1.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14102

Page 10: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

alloy formation. Such a peak is not prominently observed in thecharged electrode (Figure 4e), confirming the reversibility of thealloying/dealloying mechanism, which is consistent with theelectrochemical performance of FLP/rGO. Lastly, in P 2p XPSspectra (Figure 4c and f), three peaks, in a doublet (2p3/2 and2p1/2), appear between 128 and 130 eV, 131−133 eV, and 134−136 eV, which corresponds to alloy formation (K4P3 or K3P), aP−P bond, and a P−O bond, respectively. The presence of theP−P bond indicates that only the K4P3 alloy (Figure S12g) ispresent and rules out K3P as a possibility. This is evident, sinceP−P bond formation is not present in K3P (Figure S12f). Ourconclusion that K4P3 is being formed is also in accordance withthe specific capacity numbers obtained from electrochemicaltests (Figure 3c and d); K3P alloy provides a much higherspecific capacity of ∼2500 mAh g−1, which is 2- to 3-fold higherthan what we havemeasured. As a final proof, we also carried outex situ XRD characterization of the charged and dischargedelectrodes, which revealed the presence of peaks correspondingto the K4P3 alloy phase (Figure S13).It should also be noted that on alloying phosphorus loses

electrons, which leads to a right shift of the P−P bond peak incomparison with unalloyed FLP. In the charged state, two majorpeaks, in the doublet, are observed between 129−131 eV and132−134 eV, which corresponds to the P−P bond and the P−Obond, respectively. Such a left shift in the P 2p XPS spectrum isdue to a gain of electrons by P after dealloying. This shiftbetween the charged and discharged states confirms thereversibility of the K4P3 alloying/dealloying reaction. The F 1sand O 1s XPS spectra of the discharged and charged stateconfirm the existence of the SEI layer over the FLP/rGOelectrode, PVDF binder, and oxide groups present onphosphorene and graphene sheets (Figure S14).First-principles DFT calculations were performed to develop

a deeper understanding as to why K4P3 alloy formation ispreferred over other competing alloy states. For this, theformation energy per P atom of possible alloys (KP, K2P3, K3P,and K4P3) of phosphorene with potassium is computed (seeMethods), wherein the K3P alloy shows a formation energy of−0.98 eV, which is very similar to the K4P3 alloy (−0.97 eV) butlower than the formation energy of KP (−0.90 eV) and K2P3alloy (−0.79 eV) (Figure 4g and Table S1). This indicates thatK2P3 is stable at higher potentials and then converts into KPwithfurther potassiation at lower potentials. Upon furtherpotassiation, KP will change to K4P3 at a potential of ∼0.03 V.If the K4P3 phase is potassiated to 0.0 V or below 0.0 V, K4P3 canchange into K3P, but kinetics permitting will get converted backto K4P3 and K based on the K−P phase diagram (MaterialsProject, https://materialsproject.org/). The formation of K4P3at∼0.03 V is also consistent with our experimental observations,since (i) the pure FLP anode shows a specific capacity after theformation cycle of ∼1200 mAh g−1 (Figure 3d), which closelymatches the theoretical capacity of K4P3, and (ii) P 2p XPSanalysis shows the presence of the P−P bond, which rules outK3P since P−P bonds are absent in K3P (see Figure S12f).Full-Cell Potassium-Ion Battery Demonstration.Devel-

oping a working full-cell device is indispensable to demonstratethe proof-of-concept in a realistic operational environment.Here we report a full-cell KIB device with the FLP/rGO (1:3) asthe anode and s-KCO as the cathode. The s-KCOmicrosphereswere fabricated using a facile self-templating method, whereincobalt carbon trioxide (CoCO3) microspheres, synthesizedthrough a solvent-thermal method, were initially calcined in airto form porous cobalt oxide (Co3O4) microspheres followed by

further calcination in the presence of potassium hydroxide (seeMethods).46 XRD data reveal that the s-KCO has a hexagonalstructure with lattice parameters of a = b = 2.8 Å and c = 12.4 Å.Peaks between 10° and 30° are attributed to scattering from theKapton film, which was used to prepare samples for XRD(Figure S15). SEM and TEM imaging indicates the sphericalmorphology of the s-KCO with an average size of 8−10 μm(Figures S16 and S17). Well-aligned lattice fringes (d = 0.62nm) observed through HRTEM are in accordance with XRDdata and prove the highly crystalline structure of the s-KCO(Figure S17f). Energy dispersive X-ray spectroscopy alsoconfirms the uniform distribution of potassium, oxygen, andcobalt in the s-KCO (Figure S18).46

Full-cells were configured with FLP/rGO (1:3) as the anode,s-KCO as the cathode, and KPF6 in EC/DEC as the electrolyte(see Methods). Before configuring the full-cell, FLP/rGO (1:3)and s-KCO were initially cycled, in their respective voltagewindows (0.03−2.3 V for FLP/rGO and 1.7−3.7 V for s-KCO),in a half-cell configuration for the removal of irreversiblecapacity and activation of both the materials. Both FLP/rGO(1:3) and s-KCO showed highly reversible capacities of ∼610and ∼65 mAh g−1 at a C rate of ∼0.2 C, respectively (see Figure5a). These activated materials were then used to fabricate s-KCO||FLP/rGO (1:3) full-cell KIBs (at a cathode-to-anodecapacity ratio of 0.95:1) and were operated in the 0.03−3.3 Vvoltage window. Charge/discharge voltage profiles and cyclingperformance of FLP/rGO (1:3) show a discharge capacity of∼420 mAh g−1 after the second cycle and retained a dischargecapacity of ∼250 mAh g−1 after the 70th cycle when operated ata current rate of ∼0.5 C (see Figure 5b,c). The specific capacityof the full-cell is calculated based on the phosphorene masspresent in the anode (i.e., after removing the contribution ofrGO). The s-KCO||FLP/rGO (1:3) full-cell delivers an averageCoulombic efficiency of >99% at a current rate of ∼0.5 C(Figure 5c). The rate capability of FLP/rGO (1:3) in the full-celldevice at different current rates and its corresponding charge/discharge voltage profile are shown in Figure 5d,e. The FLP/rGO anode delivered high discharge capacities of ∼420, ∼310,and ∼190 mAh g−1 at current rates of ∼0.5C, ∼1C, and ∼2C,respectively. Even at a high current rate of ∼2.5C and ∼4C,FLP/rGO showed discharge capacities of ∼180 and ∼150 mAhg−1, respectively. A stepwise decrease in current rate from 4C to0.5C resulted in the FLP/rGO regaining a discharge capacity of∼410 mAh g−1 after 45 cycles, revealing highly reversiblepotassium storage in FLP/rGO with stable performance.Voltage profiles at different current rates (Figure 5e) areconsistent with the rate capability results, as they show almostsimilar shape and slope with slightly different overpotential. Theabove test results demonstrate that the FLP/rGO (1:3)electrode retains its performance attributes in a full-cell settingand therefore shows promise as a high specific capacity anodematerial in KIBs.One interesting observation in all of our testing is that the

capacity retention of the FLP/rGO electrode improves at high Crates. This is because at high C rates a relatively smaller fractionof phosphorene is taking part in the alloying/dealloying reactiondue to the slow kinetics of alloy formation. In fact, the larger sizeof the K ion as compared to Li makes the alloying−dealloyingreaction for K more sluggish compared to Li, and hence morereaction time is required to fully utilize the entire phosphorene.This leads to a decrease in the stress buildup, since the volumeexpansion/contraction of the electrode is mitigated due toincomplete participation of phosphorene in the alloying/

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14103

Page 11: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

dealloying process. Therefore, at higher C rates, the possibility ofpulverization and delamination of the electrode is lower incomparison with slower C rates.

CONCLUSION

In conclusion, we have demonstrated successful application ofFLP as anode materials for KIBs. Pure FLP showed a high initialcapacity of ∼1200 mAh g−1 with rapid capacity decay insubsequent cycles, which is attributed to pulverization of theelectrode due to large volume changes during the alloying/dealloying of phosphorene. While red phosphorus has beenshown to form a variety of alloys with potassium (KP, K2P3, K3P,and K4P3) at different potentials, FLPwas observed to formK4P3at ∼0.03 V. This was confirmed by ex situ XPS and XRD as wellas first-principles DFT calculations and galvanostatic charge−discharge cycling. In order to prevent pulverization and capacityfade, composites of rGO and sCNTs with FLP were synthesizedwherein rGO and sCNTs were used as buffer layers tomechanically buttress the FLP sheets. We found that rGOexhibited a highermechanical stability (∼74% capacity retentionafter 50 cycles) than sCNTs (∼43% capacity retention after 50cycles). This is due to the 2D structure of rGO, which allows it tobetter support the 2D FLP sheets. Owing to its superiormechanical integrity, FLP/rGO not only delivered a highspecific capacity (710 mAh g−1 at 0.1C) and a reasonable ratecapability (400 and∼200mAh g−1 at 0.6 and 1.2 C rate) but alsoshowed good cycle stability (∼230 mAh g−1 at 0.5 C for 300cycles). A very similar performance was observed whileinvestigating FLP/rGO in a working full-cell KIB, wherein s-KCO was used as the cathode. On the basis of our findings, weconclude that FLP materials show significant promise as high-performance anode materials for next-generation KIBs. SuchKIBs could play an important role in reducing our currentreliance on lithium-ion technology.

METHODSSynthesis of Few-Layer Phosphorene. Black phosphorus,

purchased from Smart Elements, was immersed in anhydrous NMPat a concentration of ∼0.5 mg/mL and ultrasonicated (tip sonication)for ∼5 h in an ice bath wherein the on/off ratio was 3 s/2 s. In order toseparate unexfoliated BP, centrifugation at 500−2000 rpm wasperformed for ∼30 min. This led to the appearance of a brownsolution consisting of FLP sheets. The unexfoliated BP was furtherexfoliated using the tip sonication method in order to increase theefficiency of exfoliation.Synthesis of FLP, FLP/rGO, and FLP/sCNTs Electrodes.

Reduced graphene oxide and single-walled carbon nanotubes werepurchased from ACSMaterials. rGOwas mixed in anhydrous NMP at aconcentration of ∼1 mg/mL and was bath sonicated for ∼30 min,which led to a black-colored solution. The as-synthesized FLP solutionwas then mixed with a rGO solution followed by bath sonication for∼90 min, which led to a gray-colored solution. Lastly, the FLP/rGOsolution in anhydrous NMP was vacuum filtered and dried at ∼70 °Covernight inside an argon-filled glovebox in order to remove theremaining NMP. After complete removal of NMP, FLP/rGO powderwas mixed with super-P carbon black and PVDF binder in 80:10:10%proportion by weight. After mixing, the FLP/rGO electrode slurry wasprepared by adding a small amount of NMP to the mixture, followed bystirring at∼2000 rpm for∼20 min. The slurry was then coated on a Cufoil, heated in ambient conditions for 2−3 h, and then transferred to avacuum chamber and further heated at∼45 °Covernight. Pure FLP andthe FLP/sCNTs electrodes were also synthesized using the aboveprocedure.Synthesis of s-KCO Cathode. Initially, cobalt carbon trioxide

(CoCO3) microspheres were synthesized using the solvothermal

method46 wherein a solution containing ∼0.95 g of cobalt chloride(CoCl2·6H2O), ∼2.5 g of urea, ∼55 mL of glycerol, and ∼20 mL ofH2O was stirred for ∼2 h at room temperature and transferred to aTeflon-lined autoclave, which was then heated at ∼180 °C for ∼12 h.As-synthesized CoCO3 was washed with ethanol and water multipletimes before calcination at∼500 °C for∼4 h, which led to formation ofcobalt oxide (Co3O4)microspheres. Co3O4 (∼3mmol) was thenmixedwith potassium hydroxide (KOH, ∼6 mmol) in water (∼1 mL) anddried at ∼80 °C in order to make a solid mixture, which was thenpreheated at ∼350 °C for ∼2 h. After pretreatment, the solid mixturewas calcined at ∼700 °C for ∼10 h in an oxygen (O2) environmentfollowed by natural cooling to ∼200 °C. This led to the formation46 ofs-KCO, which was then taken out in an argon-filled glovebox to avoidany contamination in ambient conditions. For the preparation of the s-KCO electrode, the slurry was initially prepared by properly mixing s-KCO powder, super P-carbon black, and PVDF binder (mass ratio of5:1:1) in small amount of NMP, which was then coated on aluminumfoils and dried at ∼100 °C for ∼12 h in vacuum.

K-Ion Half-Cell Measurement. For the half-cell configuration,2032-type coin cells were used wherein potassiummetal was used as thereference/counter electrode and Celgard 2340 polypropylene mem-brane as separator. The electrode coated on copper foil was punched ina circular shape with an area of ∼1.27 cm2. Mass loading of the wholematerial (electrode materials (FLP/rGO, FLP/sCNTs, rGO) + carbonblack + PVDF) ranged from 1 to 2 mg, depending on the coatingtechnique used. Potassium hexaflorophosphate (0.9 Μ, KPF6) in ECand DEC, mixed in 1:1 ratio by volume, was used as the electrolyte.Assembly of all the batteries was done inside an argon-filled glovebox(MBraun Labstar). Charge/discharge testing and GITT wereperformed within voltage window of 0.03−2.3 V using an ArbinBT200 battery instrument. Cyclic voltammogram testing and EIS wereconducted at room temperature using a Gamry Instrumentspotentiostat. Similarly, the half-cell for s-KCO with a mass loading of∼1 mg/cm2 was also configured. Charge/discharge testing for s-KCOwas performed within a voltage window of 1.7−3.7 V using the sameArbin BT200 battery instrument.

K-Ion Full-Cell Measurement. For the full-cell test, the half-cell ofFLP/rGO (1:3) was cycled for one or two cycles between 0.03 and 2.3V in order to remove irreversible capacity, whereas s-KCO was cycledbetween 1.7 and 2.3 V for activating the cathode. The full-cell was thenassembled by taking out the anode material from the FLP/rGO||K half-cell and cathode from the s-KCO||K half-cell inside an argon-filledglovebox. A fresh cell was then assembled with cycled FLP/rGO (1:3)as anode, activated s-KCO as cathode, and 0.9 M KPF6 in EC/DEC(1:1) as the electrolyte. The active material mass ratio of the cathodeand anode was around 4:1.

Materials Characterization. Phosphorene was characterized usingSEM (Hitachi S4800), XPS (K-alpha Thermos VG Scientific), TEM(Technai G2 F30, 300 kV), STEM (Talso F200X, 200 kV), Raman(Aramis, France), and AFM (Bruker, USA). FLP/rGO and FLP/sCNTs were characterized using TEM (JEOL JEM 2010, 200 kV),XRD (D-MAX/2400 with Cu Kα radiation), XPS (PHI 500Versaprobe with Al Kα radiation), and Raman (labRAM HR800,632.8 nm He−Ne laser).

DFT Calculations. The Vienna ab Initio Simulation Package(VASP) was used to carry out all the DFT computations with core−valence interaction using the projector-augmented-wave (PAW)method. For all the calculations, exchange−correlation interactionswere described by the generalized gradient approximation (GGA) withthe Perdew−Burke−Ernzerhof (PBE) functional. A cut-off energy of500 eV was used to generate the plane-wave basis set, and the firstBrillouin zone was sampled using a Monkhorst and Pack k-pointscheme. All the structures were relaxed until the self-consistent fieldenergies were converged to 10−5 eV with a force on each atom of lessthan 0.05 eV Å−1. The details of the atomic position and cell dimensions(POSCAR file) for all the considered systems are given in thesupplementary details. Formation energy of the alloy was calculatedaccording to the following equation:

= { − *\ _ + * }E E X Y E n Y( mu K / ) /f (KxPy) P

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14104

Page 12: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

wherein Ef refers to the formation energy; E(KxPy) and EP refer to DFTrelaxed energy of the KxPy alloy, and phosphorene, respectively; \mu_Krefers to the chemical potential of K with respect to metallic K; n refersto numbers of phosphorus atoms in phosphorene; and X and Y refer tothe number of potassium and phosphorus atoms, respectively, in thealloy.

ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acsnano.9b06680.

Additional information (PDF)

AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] Wang: 0000-0002-8626-6381Nikhil Koratkar: 0000-0002-4080-3786Author Contributions¶R. Jain and P. Hundekar contributed equally.NotesThe authors declare no competing financial interest.

ACKNOWLEDGMENTSThis work was supported by the USA National ScienceFoundation (award numbers 1608171 and 1922633). N.K.also acknowledges funding support from the John A. Clark andEdward T. Crossan endowed chair professorship at RensselaerPolytechnic Institute (RPI). S.O.K. was financially supported bythe National Creative Research Initiative (CRI) Center forMult i -Dimensional Directed Nanoscale Assembly(2015R1A3A2033061) through the National Research Founda-tion of Korea (NRF). A.Y. acknowledges the guidance ofProfessor Vincent Meunier at RPI for carrying out DFTcalculations.

REFERENCES(1) Grey, C. P.; Tarascon, J. M. Sustainability and In SituMonitoringin Battery Development. Nat. Mater. 2017, 16, 45−56.(2) Chu, S.; Cui, Y.; Liu, N. The Path Towards Sustainable Energy.Nat. Mater. 2017, 16, 16−22.(3) Suresh, S.; Wu, Z. P.; Bartolucci, S. F.; Basu, S.; Mukherjee, R.;Gupta, T.; Hundekar, P.; Shi, Y.; Lu, T. M.; Koratkar, N. ProtectingSilicon Film Anodes in Lithium-Ion Batteries Using an Atomically ThinGraphene Drape. ACS Nano 2017, 11, 5051−5061.(4) Hu, J. W.; Wu, Z. P.; Zhong, S. W.; Zhang, W. B.; Suresh, S.;Mehta, A.; Koratkar, N. Folding Insensitive, High Energy DensityLithium-Ion Battery Featuring Carbon Nanotube Current Collectors.Carbon 2015, 87, 292−298.(5) Li, L.; Wu, Z. P.; Sun, H.; Chen, D.; Gao, J.; Suresh, S.; Chow, P.;Singh, C. V.; Koratkar, N. A Foldable Lithium-Sulfur Battery. ACSNano 2015, 9, 11342−11350.(6) Lin, D.; Liu, Y.; Cui, Y. Reviving the Lithium Metal Anode forHigh-Energy Batteries. Nat. Nanotechnol. 2017, 12, 194−206.(7) Tarascon, J.-M. Is Lithium the New Gold? Nat. Chem. 2010, 2,510−510.(8) Hwang, J. Y.; Myung, S. T.; Sun, Y. K. Sodium-Ion Batteries:Present and Future. Chem. Soc. Rev. 2017, 46, 3529−3614.(9) Zhou, X.; Liu, Q.; Jiang, C.; Ji, B.; Ji, X.; Tang, Y.; Cheng, H.-M.Beyond Conventional Batteries: Strategies Towards Low-Cost Dual-Ion Batteries with High Performance. Angew. Chem., Int. Ed. 2019,DOI: 10.1002/anie.201814294.

(10) Pramudita, J. C.; Sehrawat, D.; Goonetilleke, D.; Sharma, N. AnInitial Review of the Status of Electrode Materials for Potassium-IonBatteries. Adv. Energy Mater. 2017, 7, 1602911.(11) Liu, Y.; Fan, F.; Wang, J.; Liu, Y.; Chen, H.; Jungjohann, K. L.;Xu, Y.; Zhu, Y.; Bigio, D.; Zhu, T.; Wang, C. In Situ TransmissionElectron Microscopy Study of Electrochemical Sodiation andPotassiation of Carbon Nanofibers. Nano Lett. 2014, 14, 3445−3452.(12) Eftekhari, A. Potassium Secondary Cell Based on Prussian BlueCathode. J. Power Sources 2004, 126, 221−228.(13) Jian, Z.; Luo, W.; Ji, X. Carbon Electrodes for K-Ion Batteries. J.Am. Chem. Soc. 2015, 137, 11566−11569.(14) Sonia, F. J.; Jangid, M. K.; Aslam, M.; Johari, P.; Mukhopadhyay,A. Enhanced and Faster Potassium Storage in Comparative Study withLithium Storage. ACS Nano 2019, 13, 2190−2204.(15) Ding, J.; Zhang, H.; Zhou, H.; Feng, J.; Zheng, X.; Zhong, C.;Paek, E.; Hu, W.; Mitlin, D. Sulfur-Grafted Hollow Carbon Spheres forPotassium-Ion Battery Anodes. Adv. Mater. 2019, 1900429.(16) Son, J.; Oh, S.; Bae, S.; Nam, S.; Oh, I. A Pair of NiCo 2O 4 and V2 O 5 Nanowires Directly Grown on Carbon Fabric for HighlyBendable Lithium-Ion Batteries. Adv. Energy Mater. 2019, 9, 1900477.(17) Eftekhari, A.; Jian, Z.; Ji, X. Potassium Secondary Batteries. ACSAppl. Mater. Interfaces 2017, 9, 4404−4419.(18) Wessells, C. D.; Peddada, S. V.; Huggins, R. A.; Cui, Y. NickelHexacyanoferrate Nanoparticle Electrodes for Aqueous Sodium andPotassium Ion Batteries. Nano Lett. 2011, 11, 5421−5425.(19) Zhang, C.; Xu, Y.; Zhou, M.; Liang, L.; Dong, H.; Wu, M.; Yang,Y.; Lei, Y. Potassium Prussian Blue Nanoparticles: A Low-CostCathode Material for Potassium-Ion Batteries. Adv. Funct. Mater. 2017,27, 1604307.(20) Vaalma, C.; Giffin, G. A.; Buchholz, D.; Passerini, S. Non-Aqueous K-Ion Battery Based on Layered K0.3MnO2 andHard Carbon/Carbon Black. J. Electrochem. Soc. 2016, 163, A1295−A1299.(21) Park, W. B.; Han, S. C.; Park, C.; Hong, S. U.; Han, U.; Singh, S.P.; Jung, Y. H.; Ahn, D.; Sohn, K. S.; Pyo, M. KVP2O7 as a Robust High-Energy Cathode for Potassium-Ion Batteries: Pinpointed by a FullScreening of the Inorganic Registry under Specific Search Conditions.Adv. Energy Mater. 2018, 8, 1703099.(22) Kim, H.; Kim, J. C.; Bo, S. H.; Shi, T.; Kwon, D. H.; Ceder, G. K-Ion Batteries Based on a P2-Type K0.6CoO2 Cathode. Adv. EnergyMater. 2017, 7, 1700098.(23) Xue, L.; Li, Y.; Gao, H.; Zhou, W.; Lu, X.; Kaveevivitchai, W.;Manthiram, A.; Goodenough, J. B. Low-Cost High-Energy PotassiumCathode. J. Am. Chem. Soc. 2017, 139, 2164−2167.(24) Zhao, Q.; Wang, J.; Lu, Y.; Li, Y.; Liang, G.; Chen, J. OxocarbonSalts for Fast Rechargeable Batteries. Angew. Chem., Int. Ed. 2016, 55,12528−12532.(25) Ju, Z.; Zhang, S.; Xing, Z.; Zhuang, Q.; Qiang, Y.; Qian, Y. DirectSynthesis of Few-Layer F-Doped Graphene Foam and Its Lithium/Potassium Storage Properties. ACS Appl. Mater. Interfaces 2016, 8,20682−20690.(26) Li, D.; Ren, X.; Ai, Q.; Sun, Q.; Zhu, L.; Liu, Y.; Liang, Z.; Peng,R.; Si, P.; Lou, J.; Feng, J.; Ci, L. Facile Fabrication of Nitrogen-DopedPorous Carbon as Superior Anode Material for Potassium-IonBatteries. Adv. Energy Mater. 2018, 8, 1802386.(27) Cao, B.; Zhang, Q.; Liu, H.; Xu, B.; Zhang, S.; Zhou, T.; Mao, J.;Pang, W. K.; Guo, Z.; Li, A.; Zhou, J.; Chen, X.; Song, H. GraphiticCarbon Nanocage as a Stable and High Power Anode for Potassium-Ion Batteries. Adv. Energy Mater. 2018, 8, 1801149.(28) Jian, Z.; Hwang, S.; Li, Z.; Hernandez, A. S.; Wang, X.; Xing, Z.;Su, D.; Ji, X. Hard−Soft Composite Carbon as a Long-Cycling andHigh-Rate Anode for Potassium-Ion Batteries. Adv. Funct. Mater. 2017,27, 1700324.(29) Ma, G.; Huang, K.; Ma, J. S.; Ju, Z.; Xing, Z.; Zhuang, Q. C.Phosphorus and Oxygen Dual-Doped Graphene as Superior AnodeMaterial for Room-Temperature Potassium-Ion Batteries. J. Mater.Chem. A 2017, 5, 7854−7861.(30) 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 withGraphitic Materials. Nano Lett. 2015, 15, 7671−7677.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14105

Page 13: Reversible Alloying of Phosphorene with Article Potassium ... · Reversible Alloying of Phosphorene with Potassium and Its Stabilization Using Reduced Graphene Oxide Buffer Layers

(31) Share, K.; Cohn, A. P.; Carter, R.; Rogers, B.; Pint, C. L. Role ofNitrogen-DopedGraphene for ImprovedHigh-Capacity Potassium IonBattery Anodes. ACS Nano 2016, 10, 9738−9744.(32) Xu, Y.; Zhang, C.; Zhou, M.; Fu, Q.; Zhao, C.; Wu, M.; Lei, Y.Highly Nitrogen Doped Carbon Nanofibers with Superior RateCapability and Cyclability for Potassium Ion Batteries. Nat. Commun.2018, 9, 1720.(33) Zhang, Q.; Mao, J.; Pang, W. K.; Zheng, T.; Sencadas, V.; Chen,Y.; Liu, Y.; Guo, Z. Boosting the Potassium Storage Performance ofAlloy-Based Anode Materials Via Electrolyte Salt Chemistry. Adv.Energy Mater. 2018, 8, 1703288.(34) An, Y.; Tian, Y.; Ci, L.; Xiong, S.; Feng, J.; Qian, Y. Micron-SizedNanoporous Antimony with Tunable Porosity for High-PerformancePotassium-Ion Batteries. ACS Nano 2018, 12, 12932−12940.(35) Han, C.; Han, K.; Wang, X.; Wang, C.; Li, Q.; Meng, J.; Xu, X.;He, Q.; Luo, W.; Wu, L.; Mai, L. Three-Dimensional Carbon NetworkConfined Antimony Nanoparticle Anodes for High-Capacity K-IonBatteries. Nanoscale 2018, 10, 6820−6826.(36) Sultana, I.; Rahman, M. M.; Chen, Y.; Glushenkov, A. M.Potassium-Ion Battery Anode Materials Operating through theAlloying−Dealloying Reaction Mechanism. Adv. Funct. Mater. 2018,28, 1703857.(37) Li, S.; Chen, Z.; Guo, Z.; Mao, J.; Zhang, W. Phosphorus-BasedAlloy Materials for Advanced Potassium-Ion Battery Anode. J. Am.Chem. Soc. 2017, 139, 3316−3319.(38) Chang, W. C.; Wu, J. H.; Chen, K. T.; Tuan, H. Y. RedPhosphorus Potassium-Ion Battery Anodes. Adv. Sci. 2019, 6, 1801354.(39) Xiong, P.; Bai, P.; Tu, S.; Cheng, M.; Zhang, J.; Sun, J.; Xu, Y. RedPhosphorus Nanoparticle@3D Interconnected Carbon NanosheetFramework Composite for Potassium-Ion Battery Anodes. Small2018, 14, 1802140.(40) Wu, Y.; Hu, S.; Xu, R.; Wang, J.; Peng, Z.; Zhang, Q.; Yu, Y.Boosting Potassium-Ion Battery Performance by Encapsulating RedPhosphorus in Free-Standing Nitrogen-Doped Porous Hollow CarbonNanofibers. Nano Lett. 2019, 19, 1351−1358.(41) Kavalsky, L.; Mukherjee, S.; Singh, C. V. Phosphorene as aCatalyst for Highly Efficient Nonaqueous Li − Air Batteries. ACS Appl.Mater. Interfaces 2019, 11, 499−510.(42) Sun, J.; Lee, H.W.; Pasta, M.; Yuan, H.; Zheng, G.; Sun, Y.; Li, Y.;Cui, Y. A Phosphorene-Graphene Hybrid Material as a High-CapacityAnode for Sodium-Ion Batteries.Nat. Nanotechnol. 2015, 10, 980−985.(43) Zhang, Y.; Wang, H.; Luo, Z.; Tan, H. T.; Li, B.; Sun, S.; Li, Z.;Zong, Y.; Xu, Z. J.; Yang, Y.; Khor, K. A.; Yan, Q. An Air-Stable DenselyPacked Phosphorene−Graphene Composite Toward AdvancedLithium Storage Properties. Adv. Energy Mater. 2016, 6, 1600453.(44) Sultana, I.; Ramireddy, T.; Rahman, M.; Chen, Y.; Glushenkov,A.M. Tin-Based Composite Anodes for Potassium-Ion Batteries.Chem.Commun. 2016, 52, 9279−9282.(45) Zhao, J.; Zou, X.; Zhu, Y.; Xu, Y.; Wang, C. ElectrochemicalIntercalation of Potassium into Graphite. Adv. Funct. Mater. 2016, 26,8103−8110.(46) Deng, T.; Fan, X.; Luo, C.; Chen, J.; Chen, L.; Hou, S.; Eidson,N.; Zhou, X.; Wang, C. Self-Templated Formation of P2-TypeK0.6CoO2 Microspheres for High Reversible Potassium-Ion Batteries.Nano Lett. 2018, 18, 1522−1529.(47) Phosphorus, T. B.; Kang, J.; Wood, J. D.; Wells, S. A.; Lee, J.; Liu,X.; Chen, K.; Hersam, M. C. Solvent Exfoliation of Electronic-Grade,Two-Dimensional Black Phosphorous.ACSNano 2015, 9, 3596−3604.(48) Li, L.; Chen, L.; Mukherjee, S.; Gao, J.; Sun, H.; Liu, Z.; Ma, X.;Gupta, T.; Singh, C. V.; Ren, W.; Cheng, H. M.; Koratkar, N.Phosphorene as a Polysulfide Immobilizer and Catalyst in High-Performance Lithium − Sulfur Batteries. Adv. Mater. 2017, 29,1602734.(49) Favron, A.; Gaufres, E.; Fossard, F.; Heureux, A. P.; Tang, N. Y.;Levesque, P. L.; Loiseau, A.; Leonelli, R.; Francoeur, S.; Martel, R.Photooxidation and Quantum Confinement Effects in Exfoliated BlackPhosphorus. Nat. Mater. 2015, 14, 826−832.

(50) Jian, Z.; Xing, Z.; Bommier, C.; Li, Z.; Ji, X. Hard CarbonMicrospheres: Potassium-Ion Anode Versus Sodium-Ion Anode. Adv.Energy Mater. 2016, 6, 1501874.(51) Li, Y.; Wu, W.; Ma, F. Blue Phosphorene/GrapheneHeterostructure as a Promising Anode for Lithium-Ion Batteries: AFirst-Principles Study with Vibrational Analysis Techniques. J. Mater.Chem. A 2019, 7, 611−620.(52) Levi, M. D.; Aurbach, D. Diffusion Coefficients of Lithium Ionsduring Intercalation into Graphite Derived from the SimultaneousMeasurements and Modeling of Electrochemical Impedance andPotentiostatic Intermittent Titration Characteristics of Thin GraphiteElectrodes. J. Phys. Chem. B 1997, 101, 4641−4647.(53) Wang, K.; Jin, Y.; Sun, S.; Huang, Y.; Peng, J.; Luo, J.; Zhang, Q.;Qiu, Y.; Fang, C.; Han, J. Low-Cost and High-Performance HardCarbon Anode Materials for Sodium-Ion Batteries. ACS Omega 2017,2, 1687−1695.(54) Liu, H.; Zhang, S.; Zhu, Q.; Cao, B.; Zhang, P.; Sun, N.; Xu, B.;Wu, F.; Chen, R. Fluffy Carbon-Coated Red Phosphorus as a HighlyStable and High-Rate Anode for Lithium-Ion Batteries. J. Mater. Chem.A 2019, 7, 11205−11213.(55) Cai, X.; Lai, L.; Shen, Z.; Lin, J. Graphene and Graphene-BasedComposites as Li-Ion Battery Electrode Materials and TheirApplication in Full Cells. J. Mater. Chem. A 2017, 5, 15423−15446.(56) Xin, F. X.; Tian, H. J.; Wang, X. L.; Xu, W.; Zheng, W. G.; Han,W. Q. Enhanced Electrochemical Performance of Fe0.74Sn5@ReducedGraphene Oxide Nanocomposite Anodes for Both Li-Ion and Na-IonBatteries. ACS Appl. Mater. Interfaces 2015, 7, 7912−7919.(57) Zhou, J.; Jiang, Z.; Niu, S.; Zhu, S.; Zhou, J.; Zhu, Y.; Liang, J.;Han, D.; Xu, K.; Zhu, L.; Liu, X.; Wang, G.; Qian, Y. Self-StandingHierarchical P/CNTs@rGO with Unprecedented Capacity andStability for Lithium and Sodium Storage. Chem. 2018, 4, 372−385.(58) Prabakar, S. J. R.; Jeong, J.; Pyo, M. Nanoporous Hard CarbonAnodes for Improved Electrochemical Performance in Sodium IonBatteries. Electrochim. Acta 2015, 161, 23−31.(59) Adams, R. A.; Varma, A.; Pol, V. G. Temperature DependentElectrochemical Performance of Graphite Anodes for K-Ion and Li-IonBatteries. J. Power Sources 2019, 410−411, 124−131.(60)Wang, Z.; Dong, K.; Wang, D.; Luo, S.; Liu, Y.; Wang, Q.; Zhang,Y.; Hai, A.; Shi, C.; Zhao, N. A Nanosized SnSb Alloy Confined in N-Doped 3D Porous Carbon Coupled with Ether-Based Electrolytestoward High-Performance Potassium-Ion Batteries. J. Mater. Chem. A2019, 7, 14309−14318.(61) Li, W.; Yan, B.; Fan, H.; Zhang, C.; Xu, H.; Cheng, X.; Li, Z.; Jia,G.; An, S.; Qiu, X. FeP/C Composites as an Anode Material for K IonBatteries. ACS Appl. Mater. Interfaces 2019, 11, 22364−22370.(62) Prabakar, S. J. R.; Han, S. C.; park, C.; Bhairuba, I. A.; Reece, M.J.; Sohn, K. S.; Pyo, M. Spontaneous Formation of Interwoven PorousChannels in Hard-Wood-Based Hard-Carbon for High-PerformanceAnodes in Potassium-Ion Batteries. J. Electrochem. Soc. 2017, 164,A2012−A2016.(63) Zhang, W.; Liu, Y.; Guo, Z. Approaching High-PerformancePotassium-Ion Batteries Via Advanced Design Strategies and Engineer-ing. Sci. Adv. 2019, 5, 7412.

ACS Nano Article

DOI: 10.1021/acsnano.9b06680ACS Nano 2019, 13, 14094−14106

14106