Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Energy Storage Materials 48 (2022) 325–334

Contents lists available at ScienceDirect

Energy Storage Materials


journal homepage: www.elsevier.com/locate/ensm

Nitrogen and sulfur co-doped graphene nanoribbons with well-ordered


stepped edges for high-performance potassium-ion battery anodes
Juhyung Choi a,‡, Aihua Jin b,‡, Hyun Dong Jung c,‡, Dongjin Ko a, Ji Hyun Um b,
Yoon Jeong Choi b, So Hee Kim d, Seoin Back c,∗, Seung-Ho Yu b,∗, Yuanzhe Piao a,e,∗
a
Department of Transdisciplinary Studies, Graduate School of Convergence Science and Technology, Seoul National University, 145 Gwanggyo-ro, Yeongtong-gu,
Suwonsi, Gyeonggi-do 16229, Republic of Korea
b
Department of Chemical and Biological Engineering, Korea University, 145 Anam-ro, Seongbuk-gu, Seoul 02841, Republic of Korea
c
Department of Chemical and Biomolecular Engineering, Institute of Emergent Materials, Sogang University, Seoul 04107, Republic of Korea
d
Advanced Analysis Center, Korea Institute of Science and Technology, Seoul 02792, Republic of Korea
e
Advanced Institutes of Convergence Technology, 145 Gwanggyo-ro, Yeongtong-gu, Suwon-si, Gyeonggi-do 16229, Republic of Korea

a b s t r a c t

Graphitic carbon materials, particularly few-layered graphene, exhibit great potentials as potassium-ion battery (PIBs) anodes. However, bulk graphene-based ma-
terials have the disordered structure owing to randomly stacked graphene layers, which causes the high migration barrier during K+ intercalation/deintercalation
reactions and thus the surface-dominated capacitive response. Here, we present a novel nanoarchitecture of nitrogen and sulfur co-doped graphene nanoribbons with
well-ordered stepped edges (NS–sGNR) via the electrochemical unzipping of multiwalled carbon nanotubes (MWCNTs) and the subsequent N/S co-doping process
for high-performance PIB anodes. As an anode material for PIBs, the prepared sample exhibits high initial capacity (329.1 mAh g−1 at 50 mA g−1 ), superior rate
capability (211.7 mAh g−1 at high current density, 2000 mA g−1 ), outstanding reversibility of K-staging, and stable long-term cyclability. Theoretical calculations
were conducted to demonstrate that sGNRs with NS co-doping (NS–sGNR) exhibit much improved K+ intercalation properties, such as the K+ adsorption energy,
charge transfer, and migration barriers, compared with the parallel-edged GNRs. Particularly, the migration barrier (the rate-determining step) can be substantially
reduced at the stepped edges during K+ intercalation.

1. Introduction driven capacitive behaviors reduced the energy density when the full-
battery device was assembled [24,25]. Furthermore, the short-range or-
Owing to their high energy density and cycle stability, as well as der of the carbon structure negatively influences the electrical conduc-
low self-discharge rate, lithium-ion batteries (LIBs) are the most uti- tivity in the electrochemical process [26]. Thus, graphite is ideal because
lized electrical energy storage (EES) devices [1–4]. However, LIBs are it exhibits a distinct plateau, which is associated with the staging process
still susceptible to issues relating to the lithium source, and this could that produces K-graphite intercalation compounds (K-GICs), at a low po-
limit the development of the EES field [5]. Therefore, the development tential [12]. In addition to graphite, few-layered graphene can increase
of attractive battery systems that are based on earth-abundant materials the storage capacity and rate capability of PIBs via the efficient forma-
is attracting enormous attention [6–11]. Potassium-ion batteries (PIBs) tion of GICs [27]. Further, it has been demonstrated that the doping of
are among the most promising alternatives to LIBs owing to the high graphitic networks with heteroatoms (e.g., N, S, O, and B) could suffi-
abundance, inexpensiveness, and low redox potential (−2.93 V vs. stan- ciently mitigate the K+ adsorption energy and enhance the electrochem-
dard hydrogen electrode) of potassium [12–20]. ical properties of PIBs [27,28]. However, the number of electroactive
Carbon-based materials are considered attractive anode materials sites could be substantially reduced by graphene stacking owing to the
for PIBs because of their inexpensiveness and high physical/chemical 𝜋–𝜋 interaction between the interlayers [29,30]. Additionally, the disor-
stabilities. Among these materials, ordered mesoporous carbon and N- dered graphene structure suffers from poor rate capability, which is due
doped porous carbon nanosheets with enlarged interlayer spacing exhib- to the high migration barrier that is the rate-determining step, during
ited increased specific capacity and enhanced reaction kinetics [21–23]. the intercalation/deintercalation process [31], and this could limit the
These materials improved the electrochemical activities of PIBs. How- K-staging capacity and induce surface-dominated capacitive charge stor-
ever, non-graphitic carbon anodes still display surface-driven capaci- age. Therefore, a rational design is required for graphene-based materi-
tive behaviors; they exhibit sloping curves rather than flat plateaus be- als to facilitate the reaction kinetics and consolidate the superior struc-
low 0.5 V in their voltage profiles. Notably, these predominant surface- tural advantages of high-performance anode materials. Furthermore, in-


Corresponding authors.
E-mail addresses: sback@sogang.ac.kr (S. Back), seunghoyu@korea.ac.kr (S.-H. Yu), parkat9@snu.ac.kr (Y. Piao).

These authors equally contributed to this work.

https://doi.org/10.1016/j.ensm.2022.03.041
Received 28 December 2021; Received in revised form 10 March 2022; Accepted 24 March 2022
Available online 26 March 2022
2405-8297/© 2022 Published by Elsevier B.V.
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

depth theoretical studies are required to avail valuable insights into the ers (≤ 6 layers) (Fig. 2b). The distance between the edges of each step
intrinsic energy kinetics of K+ intercalation in the desired structure. increased as the step approached the outer regions (ca. 1–5 nm). No-
Herein, we report N/S co-doped graphene nanoribbons with numer- tably, this morphological characteristic is unique, particularly when it
ous stepped edges (NS–sGNR), which were produced via the precise is compared with those of the carbon layers in pristine MWCNTs (Fig.
electrochemical unzipping of MWCNTs, followed by an in situ N/S co- S2a). Furthermore, the HR-TEM image also revealed that the interlayer
doping process. The systemic investigation of NS–sGNR indicated that it spacing of NS–sGNR (0.346 nm) was larger than that of pristine MWCNT
comprised few-layered GNR sheets (≤ 6 layers) containing highly meso- (0.334 nm) (refer to Fig. S2b for the TEM image of pristine MWCNT).
porous and long-range ordered stepped edges. The open MWCNT stems Before the co-doping process, sGNR exhibited much larger interlayer
are expected to accelerate ion and electron transport, and the unique spacing (0.361 nm) owing to the presence of the oxygen functional
structure of NS–sGNR could enhance the reaction kinetics. NS–sGNRs groups, which exhibited significant amorphous characters (Figs. S2c and
maintained a high reversible capacity (> 280 mAh g−1 over 150 cycles d). This result indicated that the interlayer spacing was reduced by re-
at a current density of 50 mA g−1 ) when they were applied as anodes moving the oxygen functional groups, thus rendering the carbon layers
in PIBs. Moreover, they exhibited remarkable rate performance (211.7 more ordered during NS co-doping. The energy-dispersive X-ray spec-
mAh g−1 at a high current density of 2000 mA g−1 ). Operando X-ray troscopy (EDX) mapping of NS–sGNR confirmed the uniform dispersions
diffraction (XRD) and ex situ transmission electron microscopy (TEM) of N and S in the carbon-based material (Fig. 2c). The N and S contents
were performed to demonstrate the highly reversible staging behaviors of the material, 0.80 and 1.15 wt%, respectively, were determined via
of GICs in NS–sGNR during potassiation and depotassiation. These ex- elemental analysis (EA) (Table S1). X-ray photoelectron spectroscopy
cellent properties might correlate with the enhanced kinetics due to the (XPS) was performed to elucidate the chemical and oxidation states,
presence of the stepped edges and additional N/S co-doping. Detailedly, as well as the chemical compositions of the as-prepared samples at the
density functional theory (DFT) calculations confirmed the significantly surface. Fig. 2d shows the high-resolution C 1 s spectrum of NS–sGNR,
decreased migration barrier of K+ intercalation at the stepped edges which revealed the existences of C–C/C = C (284.6 eV), C–O/C–S/C–N
compared with that at the parallel ones. (285.6 eV), C = O (287.4 eV), O–C = O (289.1 eV), and 𝜋–𝜋 ∗ (290.5 eV)
[35]. The high-resolution S 2p spectrum indicated that the fitted S 2p3/2
2. Results and discussion and S 2p1/2 peaks were located at 164.0 and 165.2 eV, respectively.
Another peak that was observed at 168.2 eV was ascribed to the SOx
2.1. Structural characterization peak (Fig. 2e) [35,36]. Contrarily, the S 2p spectrum of sGNR only re-
vealed the SOx peak (Fig. S3b), which originated from the intercalated
Fig. 1a shows the schematic of the experimental procedure for pro- sulfuric acid. Furthermore, the high-resolution N 1 s spectrum of NS–
ducing the quasi-one-dimensional NS–sGNR from commercial MWCNTs. sGNR could be deconvoluted into four peaks, namely the pyridinic N
Briefly, the sGNRs were synthesized through the electrochemical unzip- (398.4 eV), pyrrolic N (399.8 eV), graphitic N (401.6 eV), and oxidized
ping of the MWCNTs that were recently developed by our group [32]. N (403.8 eV) (Fig. 2f) [35,37]. Dissimilar to the graphitic N, which was
When a galvanostatic current of 0.2 A g−1 was applied to the electro- surrounded by carbon atoms, the pyridinic and pyrrolic N species were
chemical cell for 2.5 h, the bisulfate anions (HSO4 − ) were intercalated located at the edges of the carbon frameworks; thus, they would accept
into the MWCNT, thereby initiating the lengthwise unzipping of its con- more charges from the adjunct K atom to improve the reversible capac-
centric walls [33]. The constant anodic charging current further induced ities of PIBs [37,38]. Furthermore, the atomic fractions of the four N
the oxidation of the unzipped MWCNTs into exfoliated sGNRs. Finally, species were 37.3% (pyridinic N), 28.4% (pyrrolic N), 26.2% (graphitic
NS–sGNRs were subsequently produced via the co-pyrolysis of thiourea N), and 8.1% (oxidized N), and the ratio of the edge at N-6 and N-5
and the sGNR mixture in an Ar atmosphere at 800 °C. Following its ther- to those at N-G ((N-6 + N-5)/N-G) was 2.3, indicating that the nitro-
mal decomposition at a high temperature, thiourea can afford highly gen moiety dominated the exposed edge of the GNRs. Additionally, the
reactive N and S species, such as NH3 , H2 S, and CS2 , to execute the N/S calculated XPS survey spectra revealed that the oxygen/carbon (O/C)
co-doping process [34]. ratios of MWCNT, sGNR, and NS–sGNR were 0.04, 0.18, and 0.06, re-
Scanning electron microscopy (SEM) and transmission electron mi- spectively (Fig. S3a), indicating that the thermal annealing process of
croscopy (TEM) analyses were performed to characterize the morphol- co-doping N/S effectively decreases the number of oxygen species; this
ogy of the as-obtained products. It was observed that the MWCNTs were result correlates well with the TEM results and was further confirmed
directly unzipped into the quasi-one-dimensional sGNRs without appar- by the thermogravimetric analysis (TGA) results (Fig. S4).
ently shortening of the pristine tubes (Figs. 1b and c). Moreover, the The structural evolution of the as-prepared products was further in-
TEM image of sGNR revealed the lengthwise view of the GNRs com- vestigated via Raman spectroscopy, X-ray diffraction (XRD), and the
pared with those of the MWCNTs (Figs. 1f and e). This result confirms Brunauer–Emmett–Teller (BET) analyses. A sharp (002) peak was ob-
that the electrochemically unzipped MWCNTs could produce the edge- served in the XRD patterns of pristine MWCNTs, indicating their highly
created and opened GNR structures, which were attached to the MWCNT graphitic layered structure (Fig. 2g). Conversely, the XRD patterns of
core, and there was no significant morphological change after the an- sGNR revealed a broad (002) peak, which corresponds to the disor-
nealing process (Figs. 1d and g). Additionally, the histograms of the dered characteristics of carbon. Moreover, a (001) peak, which origi-
widths (Figs. 1h–j) revealed that the widths of the graphene layers in nated from the evolution of the C–O bonds at the basal planes, was ob-
sGNR and NS–sGNR are largely identical. Their widths were measured served [32]. Dissimilar to the oxidized sGNR, NS–sGNR exhibited a rela-
employing the TEM images (Fig. S1 in Supporting Information), and the tively strong (002) peak and restored the (004) peak, indicating that the
average widths of MWCNT, sGNR, and NS–sGNR were 149.5, 216.2, long-range ordered (≥ 4 layers) graphitic structure was recovered after
and 226.7 nm, respectively, indicating that lengthwise nanoribbon lay- the annealing process [39]. The Raman spectra of MWCNT, sGNR, and
ers were produced in sGNR and NS–sGNR. NS–sGNR revealed that the intensity ratios of their D (ca. 1338 cm−1 ) to
A closer observation of the surface of NS–sGNR (Fig. 2a) revealed G (ca. 1567 cm−1 ) bands (ID /IG ) were 0.05, 0.48, and 0.32, respectively
the opened nanostructure through which the outer GNR layers were (Fig. 2h). Notably, the higher ID /IG values of sGNR and NS–sGNR com-
seamlessly connected to the MWCNT stems. The detailed morphology pared with that of pristine MWCNT correspond to the new edges that
of NS–sGNR (the inset of Fig. 2a) indicated that the “stepped” edges were formed from the unzipped MWCNT [33]. However, the dominant
comprised thin nanoribbon sheets. It appeared that the stepped GNR D‘ (ca. 1609 cm−1 ) band of sGNR corresponds to the graphene oxide
sheets originated from different radial diameters of the rolled-up car- (GO) phase, and this agrees well with the XRD results. All the samples
bon sheets. The high-resolution TEM (HR-TEM) image of NS–sGNR re- exhibited intense 2D (ca. 2690 cm−1 ) bands, which were closely related
vealed the edge-rich step layer that exhibited the few-layered GNR lay- to the non-oxidized graphitic phase in the MWCNT core [33]. Concur-

326
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

Fig. 1. Synthesis of NS–sGNR. (a) Schematic of the preparation of NS–sGNR. SEM and TEM images of (b, e) MWCNT, (c, f) sGNR, and (d, g) NS–sGNR. The histograms
of the width of the as-prepared samples of (h) MWCNT, (i) sGNR, and (j) NS–sGNR, as determined from the TEM images.

rently, the longitudinal opening in MWCNTs can induce additional pores respectively (the SSA and total pore volumes of all the samples are pre-
in its structure [32]. The nitrogen adsorption–desorption isotherms were sented in Table S2). The pore size distributions, which were determined
measured to determine the surface area and pore size distribution of from the isotherms, indicated that sGNR and NS–sGNR exhibited meso-
all the samples (Fig. 2i). Further, the specific surface areas (SSAs) of pores and macropores that mainly originated from the opened cap, as
MWCNT, sGNR, and NS–sGNR were 15.65, 29.79, and 33.18 m2 g−1 , well as the formation of edges, in their structures (Fig. S5). Moreover,

327
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

Fig. 2. Characterization of NS–sGNR. (a) TEM and (b) corresponding HR-TEM images of the selected area (brown box). (c) EDX mappings of the C, S, and N elements
in NS–sGNR (inset of Fig. 2a is the corresponding high-magnification TEM image). (d–f) High-resolution C 1 s, S 2p, and N 1 s XPS spectra of NS–sGNR. (g) XRD
patterns, (h) Raman spectra, and (i) N2 adsorption/desorption isotherms of MWCNT, sGNR, and NS-sGNR (inset of Fig. 2i is the pore size distribution below 6 nm,
as derived via the Barrett–Joyner–Halenda (BJH) method.

the increasing value of the pore volumes of NS–sGNR, which was cen- To further elucidate the mechanisms of the potassiation and depotas-
tered at 2.6 nm, was partly due to the long-range ordered GNR layers siation processes in NS–sGNR, operando XRD was performed at a cur-
(inset of Fig. 2i) [30]. rent density of 200 mA g−1 as shown in Fig. 3c (see Fig. S6 for en-
larged XRD patterns at the selected regions). In the potassiation region
(from the initial state to 0.32 V), the (002) peak of NS–sGNR at 26.2°
2.2. Electrochemical performance
gradually vanished mainly because of the initiation of K+ intercalation
into carbon. In Stage 3, a weak peak, which was attributed to KC36 ,
The electrochemical properties of NS–sGNR and sGNR were evalu-
appeared at 29.4°, and its strength increased until the end of Stage 3.
ated for PIB anodes in the voltage range of 0.01–3.0 V (vs. K+ /K). Cyclic
Following further potassiation (in Stage 2), the KC36 peak shifted con-
voltammetry (CV) was performed at a scan rate of 0.1 mV s−1 . Fig. 3a
tinuously toward 30.6° to form KC24 . The diffraction peaks of KC8 at
shows that three well-defined peaks were observed around 0.65, 0.32,
16.5° and 33.5° evolved during the initial parts of Stage 1, and their
and 0.10 V in the first cathodic process. The peak at 0.65 V was at-
dominance increased until the end of the potassiation. A precise re-
tributed to the decomposition of electrolyte and formation of the SEI
versible reaction proceeded during the depotassiation process, and it
layer [40]. This peak disappears in the subsequent cycles, showing that
confirmed that the (002) peak of NS–sGNR was recovered after the first
stable SEI layer is formed during the initial cycle. sGNR (Fig. 3b) also
cycle [12,38,41,42]. This high reversibility of producing K-GICs during
has similar irreversible peak at 0.78 V during the initial cathodic pro-
cycling is very essential for achieving stable and high-energy-density
cess. The redox couple of 1.42/1.91 V could only be observed in sGNR,
PIBs.
which is probably attributed to the oxygen-related functional groups.
The K+ intercalation/deintercalation properties were evaluated via
This peak gradually disappears during cycling, suggesting that the re-
the galvanostatic charging/discharging of the cells at a constant current
action occurred at this potential is also irreversible. The sharp peak at
(50 mA g−1 ) (Fig. 3d; Fig. S7 shows the corresponding voltage profiles).
about 0.1 V is found in NS–sGNR during the initial cathodic process,
NS–sGNR delivered an initial depotassiation capacity of 329.1 mAh g−1 ,
which is attributed to intercalation of K ions into NS–sGNR to form K–
whereas sGNR delivered 274.5 mAh g−1 . Additionally, NS–sGNR main-
GIC. This peak is maintained during the subsequent cycles, and they are
tained a high capacity (> 287.0 mAh g−1 ) for 150 cycles, whereas sGNR
well overlapped, showing high reversibility of NS–sGNR.

328
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

Fig. 3. Electrochemical performance of NS–sGNR. Cyclic voltammograms of (a) NS–sGNR and (b) sGNR at a scan rate of 0.1 mV s−1 . (c) Operando XRD patterns of
NS–sGNR at a current density of 200 mA g−1 . (d) Cycle performance of sGNR and NS–sGNR at a current density of 50 mA g−1 .

reached 183.5 mAh g−1 . The corresponding cycle retentions of NS–sGNR long cycle performance at a high current density of 500 mA g−1 is shown
and sGNR during 150 cycles were 87.0% and 66.8%, respectively. The in Fig. 4d. NS–sGNR and sGNR became stable after several initial cy-
voltage profiles (Fig. S7) indicated that the slope region to 0.50 V during cles (Coulombic efficiency ≥ 95%). sGNR exhibited initial Coulombic
the initial discharge process originated from the irreversible decompo- of 38.96%, but NS-sGNR exhibited much improved initial Coulombic
sition and formation of the electrolyte and the SEI film, respectively, as efficiency (55.06%). During the initial several cycles, Coulombic effi-
mentioned in CV curves, and a similar slope was also observed in the ciencies were stabilized in both electrodes and reached above 95%. At
voltage profiles of sGNR. NS–sGNR exhibited a lower voltage plateau 250th cycle, Coulombic efficiencies were 99.87 and 99.73% in NS-sGNR
with much smaller voltage hysteresis compared with sGNR, particu- and sGNR, respectively (see Figs. S8 and S9). The reversible capacity
larly in the K-GIC formation-related regions [43,44]. The widened low- of NS–sGNR was 224.0 mAh g−1 after 500 cycles, exhibiting cycle re-
voltage plateau was mainly attributed to the long-range ordered carbon tention of 84.8%. Even at 500th cycle, NS–sGNR exhibited very high
structure and doping effects. Coulombic efficiency (100%). These electrochemical properties of NS-
The rate capabilities of NS–sGNR and sGNR were also tested by in- sGNR electrode are highly competitive with those of recently reported
ducing a stepwise change in the current density stepwise from 50 to carbonaceous PIB anodes (Table S3).
2000 mA g−1 . Fig. 4a shows that the reversible charge capacities of NS– In order to clarify the effect of doping on the electrochemical prop-
sGNR were 274.6, 264.2, 257.5, 248.3, 237.4, and 211.7 mAh g−1 at erties, a sample was prepared by annealing sGNR in an Ar atmosphere
50, 100, 200, 500, 1000, and 2000 mA g−1 , respectively, indicating the without thiourea (this sample was hereafter referred to as sGNR@Ar)
excellent rate capabilities of NS–sGNR even at very high current den- for comparison. The TEM image of sGNR@Ar products clearly shows
sities (1000 and 2000 mA g−1 ). However, the reversible charge capac- the one-dimensional GNR structures with no structural deformation. In
ities of sGNR dropped rapidly to 51.8 and 26.0 mAh g−1 at those cur- addition, the corresponding HR-TEM image confirms that the lattice
rent densities. Moreover, both samples recovered their initial capacities spacing is 3.38 Å, supporting the removal of oxygen groups (Figs. S10a
when the current density returned to 50 mA g−1 . To further highlight and b). Fig. S10c presents the XRD patterns of sGNR and sGNR@Ar.
the excellent rate capability of NS–sGNR, the second voltage profiles of Compared with sGNR, sGNR@Ar has a relatively strong (002) peak,
sGNR and NS–sGNR at each current density were plotted in Fig. 4b and but no (001) peak, which comes from GO phase, indicating the restora-
Fig. 4c, respectively. The corresponding voltage profiles at various cur- tion of the graphitic structure in sGNR@Ar. Moreover, the surface area
rent densities clearly show that the shapes of the profiles of NS–sGNR and pore size distribution of sGNR@Ar obtained from the BET and BJH
were generally maintained even at high current densities, wheareas the analysis are presented in Fig. S10d and Table S2. sGNR@Ar exhibits
plateau in the K-GIC-related region was almost invisible in sGNR. The the largest total pore volume (0.2573 cm3 g−1 ) and specific surface area

329
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

Fig. 4. Rate properties of NS-sGNR. (a) Rate properties of sGNR and NS–sGNR at various current densities. Voltage profiles of (b) sGNR and (c) NS–sGNR at various
current densities. (d) Cycle performance of NS–sGNR and sGNR at a current density of 500 mA g−1 .

(99.63 m2 g−1 ), followed by those of MWCNT, sGNR, and NS–sGNR. This a current density of 15 and 500 mA g–1 and rate properties of graphite,
result suggests the removal of oxygen functional groups and highly in- respectively. In all these conditions, the electrochemical properties of
compact stacking of GNR layers in sGNR@Ar. The capacity of sGNR@Ar graphite are lower compared to those of NS–sGNR, which supports the
was much lower compared with that of NS–sGNR (Fig. S11), indicating superior properties of stepped edges beyond the parallel edges.
the contribution of NS doping to the electrochemical properties. The The excellent electrochemical performance of NS–sGNR was proba-
TEM images of NS–sGNR after 50 cycles (Fig. S12a) show that its mor- bly due to the enlarged interlayer distances, the “stepped” edges for ad-
phology was largely maintained after the cycling test. Electrochemical ditional active sites, and the excellent structural stability for long-term
impedance spectroscopy (EIS) was conducted before the cycling test cycling. Moreover, nitrogen and sulfur co-doping might have improved
and after the 1st cycle (Fig. S13). The semicircles of NS–sGNR before the electrical conductivity and availed additional active sites for K+ in-
and after the 1st cycle were much smaller than those of sGNR, demon- tercalation. The origins of the improved electrochemical performance
strating that the charge transfer in NS–sGNR had increased significantly of NS–sGNRs were further studied via DFT calculations in the following
[22,45,46] (see Fig. S14 for GITT curves of NS–sGNR and sGNR). In section.
both electrodes, the impedance increases after the 1st cycle compared
to that before cycle, which might be related to the formation of pas- 2.3. DFT calculations
sivation layer [47,48]. Furthermore, we evaluate the electrochemical
properties of commercial graphite and compare with those of NS–GNR DFT calculations were conducted to compare the thermodynamics
(See Fig. S15a for SEM image of graphite and Fig. S15b. for XRD pat- and kinetics of GNRs exhibiting parallel edges (pGNR), sGNR, and NS–
tern of graphite). Fig. S16a-c present cycle performance of graphite at sGNR and to elucidate the effects of the stepped edges and NS co-doping

330
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

Fig. 5. Theoretical calculations of NS–sGNR. (a) Schematic illustration of the intercalation of K+ . The migration pathways are shown in the inset Figs. Therein, the
transparent and solid green atoms correspond to the intermediate and final positions of K+ in the barrier calculations, respectively. (b) The largest migration barriers
among the three processes. (c) Adsorption energies of K+ at the IS and TS of the rate-limiting step, the internal migration, for the undoped parallel and stepped edges.
(d) Average local interlayer distances (dint ) between the carbon atoms in the top and the bottom layers that are adjacent to the migrating K+ . (e) Adsorption energies
of an additional K+ at the parallel and stepped edge sites with and without NS co-doping. (f) Charge density difference (Δ𝜌) of the additional adsorption of K+ at
the (left) sGNR and (right) NS-sGNR (N-doped site), where the yellow and blue areas represent the electron accumulation and depletion with an isosurface level of
0.005 e/Å3 , respectively. Δ𝜌 was calculated as 𝜌𝐾∗ − 𝜌∗ − 𝜌𝐾 , where 𝜌𝐾∗ , 𝜌∗ and 𝜌𝐾 represent the charge densities of the system with and without additional K and
isolated K, respectively.

on the battery performance. The adsorption energy of K+ was calculated, or (2) corresponded to the rate-limiting steps in all the cases, indicating
as follows: that the intercalation kinetics near the edges must be facile to maximize
( ) ( ) the kinetic performance of the PIB. The calculated MAX (ΔEa1 , ΔEa2 ,
ΔEK ∗ = E K ∗ − E ∗ − E(K ) (1) and ΔEa3 ) of the stepped edges were generally lower compared with
those of the parallel edges, indicating improved kinetics for the stepped
where E(K∗ ), E(∗ ), and E(K) are the electronic energies of the K+ -
edges (Fig. S19a). Furthermore, the maximum barriers of the N/S co-
adsorbed systems, K+ -removed systems, and body-centered cubic (BCC)
doped stepped edges were 0.51 eV (the N-doped site) and 0.42 eV (the
bulk K per atom, respectively (negative binding energies indicated en-
S-doped site), which were much lower than that of the undoped coun-
hanced adsorption; the Supplementary Methods section contains addi-
terpart (0.99 eV) (Fig. S19b). This could be correlated with the experi-
tional information on the computational modeling of each process).
mentally observed higher kinetics of NS–sGNR compared with those of
It could be assumed that K+ intercalation comprises three successive
sGNR (Fig. 4a). Moreover, the maximum barriers of the N/S co-doped
migration processes: (1) the approach of K+ to the exposed edge sites,
stepped edges were lower than those of their singly N-doped (0.95 eV)
(2) the internal migration from the edge sites to the bulk regions, and
or S-doped (0.80 eV) counterparts, indicating the synergistic effect of
(3) the migration in the bulk regions (Fig. 5a). Among these three, the
NS co-doping. These results confirmed that the generation of abundant
largest migration barriers of the parallel and stepped edges with and
stepped edge sites and N/S co-doping are effective strategies for enhanc-
without NS co-doping were plotted and compared since the migration
ing the kinetic properties of PIBs. The additional DFT calculations of the
step with the largest barrier represented the rate-limiting step of the en-
undoped and NS−doped models were performed to provide an insight
tire K+ intercalation process (Fig. 5b). It was observed that Process (1)

331
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

into the effect of C − O bond on the K+ migration at the basal plains. the stepped vs. parallel edges and (2) the NS co-doped stepped edges
Fig. S20a highlights the position of adsorbed O on the top layer (dashed vs. the N or S singly doped stepped edges. Compared with the parallel
red circle) based on the bulk region system (i. e. Process 3), which is edges, the migration barrier of K+ intercalation (the rate-determining
located near the migration pathway of K+ ion. As shown in Fig. S20b, step) could be lowered at the stepped edges, which could also play a
the calculated migration energies with a C − O bond indicate 0.524 and crucial role in providing the additional K+ adsorption site. The study
0.344 eV of undoped and NS−doped, respectively, while those without a also demonstrated the extensive decrease in the migration barrier of the
C − O bond reveal the lower migration energies of 0.479 and 0.292 eV. N/S co-doped stepped edges compared with that of the N or S singly
Accordingly, the effect of the C − O bond on the migration barrier is doped sGNRs, and the enhanced interlayer spacing between the bilayer
negligible, as manifested by the low energy differences (about 0.05 eV) NS–sGNR enhanced the intercalation dynamics. Our findings avail new
between with C − O bond and without C − O bond of undoped (0.045 eV) insights into the rational design of carbon-based electrodes for high-
and NS-doped (0.052 eV), respectively. performance PIBs, which can be extended to other energy storage sys-
To elucidate the source of such improvements, the atomic structures tems.
and ∆EK∗ were analyzed at the initial states (IS) and the transition states
(TS) of the rate-limiting step, where the energy difference determines 4. Experimental section
the migration barrier. It was observed that the ∆EK∗ of IS in the un-
doped parallel edge (−1.39 eV) was much stronger than that of the un- 4.1. Chemicals and materials
doped stepped edge (−0.79 eV), while the differences in the ∆EK∗ of
TSs in the parallel and stepped edges were much smaller (−0.13 and MWCNTs with an outer diameter and length of 110–170 nm and 5–
0.20 eV, respectively), thus increasing the migration barrier of the par- 9 𝜇m, respectively (Sigma-Aldrich,); thiourea (CH4 N2 S, Sigma-Aldrich,
allel edge (Fig. 5c). This can be attributed to the significantly different >99%); and concentrated sulfuric acid (H2 SO4 , Samchun Chemicals,
local atomic structures of the IS of the parallel and stepped edges, while >96%) were employed without, as received.
their TS atomic structures were nearly equivalent. To further investigate
the effect of N/S co-doping on the migration barriers, the average local 4.2. Preparation of sGNR
interlayer distances (dint ) between the carbon atoms in the top and bot-
tom layers adjacent to the migrating K+ were compared (Fig. 5d). It was MWCNTs were electrochemically unzipped performed employing
observed that ∆dint (dint,TS − dint,IS ) increased very evidently with the co- the two-electrode configuration in 18 M H2 SO4 [32]. First, to prepare
doping of NS at the stepped edge, indicating that a very flexible stepped the working electrode, 50 mg of MWCNTs was poured into a glass fiber
edge would reduce the migration barriers near the edge sites (Fig. 5d membrane via vacuum-assisted filtration. Thereafter, the MWCNT mem-
and Table S4). Notably, regarding the layered carbon materials, such as brane was placed in a homemade jig and connected to a current collector
pGNR/sGNR/NS–sGNR in this work, the increased interlayer distances (a platinum wire that was attached to a platinum disk). The prepared
generally reduced the migration barriers [22]. DFT calculations of paral- MWCNT film and a platinum ring were employed as the working and
lel and stepped edge, including pGNR, NS−pGNR, sGNR, and NS−sGNR, counter electrodes, respectively. A galvanostatic charging process was
in the bi-layer graphene system models were conducted to confirm that performed on a ZIVE Lab electrochemical workstation. A constant cur-
the interlayer distance increases according to NS co-doping in Fig. S21. rent of 10 mA (0.2 A g–1 ) was applied to the assembled cell for 2.5 h.
As summarized in Table S5, for the N/S co-doped systems, the interlayer Afterward, the collected product was washed several times with deion-
distances (Å) of pGNR and sGNR is 3.396 and 3.40 Å, respectively, when ized water and dried overnight at 60 °C.
they were doped with N and S, suggesting the reduced migration barri-
ers during K+ intercalation reaction. 4.3. Preparation of NS–sGNR
After the full intercalation of K+ between the pGNR/sGNR/NS–sGNR
bilayers, additional K+ could be adsorbed at the edge sites to increase the The as-obtained sGNR (0.1 g) was mixed well with 1 g of thiourea
total capacity (Fig. S22) [49]. The stepped edges generally bound the ad- and vacuum dried for 3 h at 60 °C. Next, the solid mixture was an-
ditional K+ more strongly than the parallel edges, and the effect of N/S nealed for 2 h at 800 °C and a heating rate of 5 °C min−1 in an Ar
co-doping on the adsorption strength of the additional K+ was signifi- atmosphere. For comparison, sGNR was also annealed for 2 h at 600 °C
cant (Fig. 5e) probably owing to the charge transfer from the adsorbed without thiourea, and the obtained sample was denoted as sGNR@Ar.
K+ to nearby adsorption sites. The calculated Bader charges [50] indi-
cated that the adsorbed K+ at the parallel and stepped edges lost 0.81– 4.4. Material characterization
0.83 e− and 0.86–0.89 e− , respectively, thus ensuring enhanced binding
of the additional K+ at the latter (Fig. S23), as visualized in the charge The morphologies and structures of the materials were observed via
density difference plot (Fig. 5f and Fig. S24). Consequently, DFT calcu- SEM (Hitachi Regulus 8230 and Hitachi S-4800) and TEM (JEOL JEM-
lations confirmed that the stepped edges and N/S co-doping effectively ARM200F), as well as EDX. The XRD measurements were conducted
enhanced the kinetics and total capacity of PIBs. Particularly, the atomic on a Bruker D8 Advance diffractometer exhibiting Cu K𝛼 radiation
configurations at the beginning of K+ intercalation were crucial in de- (𝜆 = 1.5406 Å). The Raman spectra were obtained at an excitation wave-
termining the aforementioned properties. length of 532 nm (LabRAM HV Evolution). XPS analysis was performed
via AXIS-His-spectroscopy employing Al K𝛼 as the X-ray source. The ni-
3. Conclusions trogen adsorption/desorption isotherms were obtained with a Belsorp-
mini II instrument. EA was performed on a Thermo Fisher Scientific
NS–sGNR was fabricated via the electrochemical unzipping of MWC- Flash 2000. TGA was performed on a thermal analyzer (TGA/DSC1) in
NTs, followed by an in situ co-doping process. The prepared NS–sGNR a nitrogen atmosphere at a heating rate of 5 °C min−1 .
exhibited the following excellent electrochemical properties: (1) a high
capacity of 329.1 mAh g−1 at a current density of 50 mA g−1 , (2) supe- 4.5. Electrochemical characterization
rior cycle retention of 87.0% during 150 charging/discharging cycles,
(3) remarkable rate performance (211.7 mAh g−1 ) at a very high cur- All the electrode slurries were prepared by adding the active ma-
rent density of 2000 mA g−1 , and (4) outstanding reversibility of the terials (70 wt%), super P (15 wt%), and poly(acrylic acid) (15 wt%)
K-GICs formation process during the cycling, as confirmed via operando to N-methyl-2-pyrrolidinone. Thereafter, the slurries were coated on
XRD. The in-depth theoretical simulations adequately demonstrate the Cu foil and vacuum dried overnight. KPF6 (0.8 M) in ethylene carbon-
fast intercalation kinetics of the two major counterparts, namely (1) ate/diethyl carbonate (volume ratio = 1:1) and a Whatman GF/C glass

332
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

fiber were employed as the electrolyte and separator, respectively. All Supplementary materials
the 2032-coin cells were assembled in an Ar-filled glove box and cycled
by a WBCS3000 cycler (WanATech, Korea) in the range of 0.01–3.0 V Supplementary material associated with this article can be found, in
(vs. K+ /K) at 25 °C. Further, the operando XRD data were obtained from the online version, at doi:10.1016/j.ensm.2022.03.041.
the beamline 3D at Pohang Light Source II (PLS II), Korea. EIS was per-
formed with a ZIVE MP1 (WonATech, Korea) in the frequency range of
References
100 kHz–10 mHz.
[1] B. Dunn, H. Kamath, J.-.M. Tarascon, Electrical energy storage for the grid: a battery
4.6. Theoretical calculation of choices, Science 334 (2011) 928–935, doi:10.1126/science.1212741.
[2] J. Lu, Z. Chen, Z. Ma, F. Pan, L.A. Curtiss, K. Amine, The role of nanotechnology
in the development of battery materials for electric vehicles, Nat. Nanotechnol. 11
The DFT calculations were conducted employing the Vienna ab ini-
(2016) 1031–1038, doi:10.1038/nnano.2016.207.
tio simulation package (VASP version 5.4.4) [51,52]. The generalized [3] Y. Sun, N. Liu, Y. Cui, Promises and challenges of nanomaterials for lithium-based
gradient approximation employing the Perdew–Burke–Ernzerhof func- rechargeable batteries, Nat. Energy 1 (2016) 16071, doi:10.1038/nenergy.2016.71.
tional (GGA-PBE) [53] was employed to describe the exchange correla- [4] X.-.Q. Zhang, X. Chen, X.-.B. Cheng, B.-.Q. Li, X. Shen, C. Yan, J.-.Q. Huang,
Q. Zhang, Highly stable lithium metal batteries enabled by regulating the solva-
tions of the electron interactions. The D2 correction by Grimme [54] was tion of lithium ions in nonaqueous electrolytes, Angew. Chem. Int. Ed. 57 (2018)
applied to include the van der Waals interactions to the systems com- 5301–5305, doi:10.1002/anie.201801513.
prising GNR and the K atoms. The ionic cores were described by the [5] J.-.M. Tarascon, Is lithium the new gold? Nat. Chem. 2 (2010) 510-510,
doi:10.1038/nchem.680.
projector augmented-wave (PAW) pseudopotentials [55] with an energy [6] Z. Jian, Z. Xing, C. Bommier, Z. Li, X. Ji, Hard carbon microspheres: potassium-
cutoff of 400 eV. The energy and force convergence criteria were set to ion anode versus sodium-ion anode, Adv. Energy Mater. 6 (2016) 1501874,
10−4 eV and 0.05 eV/Å, respectively. (6 × 6 × 6) and (12 × 12 × 1) doi:10.1002/aenm.201501874.
[7] S. Komaba, T. Hasegawa, M. Dahbi, K. Kubota, Potassium intercalation
Monkhorst-Pack k-meshes [56] were sampled for the BCC bulk K (MPID: into graphite to realize high-voltage/high-power potassium-ion batteries
mp-58) and bulk graphene, respectively, while (1 × 1 × 1) was em- and potassium-ion capacitors, Electrochem. Comm. 60 (2015) 172–175,
ployed for all the other calculations because of the large number of doi:10.1016/j.elecom.2015.09.002.
[8] Z. Zhang, Y. Du, Q.-.C. Wang, J. Xu, Y.-.N. Zhou, J. Bao, J. Shen, X. Zhou, A yolk–
computational cells. The climbing image-nudged elastic band (CI-NEB)
shell-structured FePO4 cathode for high-rate and long-cycling sodium-ion batteries,
method [57] with 3–9 intermediate images was employed to calculate Angew. Chem. Int. Ed. 59 (2020) 17504–17510, doi:10.1002/anie.202008318.
the energy barriers of the migration processes of K+ . The computational [9] X. Xu, L. Si, X. Zhou, F. Tu, X. Zhu, J. Bao, Chemical bonding between antimony and
ionic liquid-derived nitrogen-doped carbon for sodium-ion battery anode, J. Power
details are available in the Supplementary Methods.
Sources 349 (2017) 37–44, doi:10.1016/j.jpowsour.2017.03.026.
[10] S. Huang, Y. Lv, W. Wen, T. Xue, P. Jia, J. Wang, J. Zhang, Y. Zhao, Three-
Declaration of Competing Interest dimensional hierarchical porous hard carbon for excellent sodium/potassium
storage and mechanism investigation, Mater. Today Energy 20 (2021) 100673,
The authors declare that they have no known competing financial doi:10.1016/j.mtener.2021.100673.
[11] H. Long, X. Yin, X. Wang, Y. Zhao, L. Yan, Bismuth nanorods confined in hollow car-
interests or personal relationships that could have appeared to influence bon structures for high performance sodium- and potassium-ion batteries, J. Energy
the work reported in this paper. Chem. 67 (2022) 787–796, doi:10.1016/j.jechem.2021.11.011.
[12] Z. Jian, W. Luo, X. Ji, Carbon electrodes for K‑ion batteries, J. Am. Chem. Soc. 137
CRediT authorship contribution statement (2015) 11566–11569, doi:10.1021/jacs.5b06809.
[13] W. Zhang, Y. Liu, Z. Guo, Approaching high-performance potassium-ion batter-
ies via advanced design strategies and engineering, Sci. Adv. 5 (2019) eaav7412,
Juhyung Choi: Conceptualization, Data curation, Formal analysis, doi:10.1126/sciadv.aav7412.
Writing – original draft. Aihua Jin: Conceptualization, Data curation, [14] J.C. Pramudita, D. Sehrawat, D. Goonetilleke, N. Sharma, An initial review of the sta-
tus of electrode materials for potassium-ion batteries, Adv. Energy Mater. 7 (2017)
Formal analysis, Writing – original draft. Hyun Dong Jung: Conceptual- 1602911, doi:10.1002/aenm.201602911.
ization, Data curation, Formal analysis, Writing – original draft. Dongjin [15] T. Li, Q. Zhang, Advanced metal sulfide anode for potassium ion batteries, J. Energy
Ko: Data curation, Formal analysis. Ji Hyun Um: Data curation, For- Chem. 27 (2018) 373–374, doi:10.1016/j.jechem.2017.12.009.
[16] J. Liao, C. Chen, Q. Hu, Y. Du, Y. He, Y. Xu, Z. Zhang, X. Zhou, A low-strain phosphate
mal analysis. Yoon Jeong Choi: Data curation, Formal analysis. So Hee
cathode for high-rate and ultralong cycle-life potassium-ion batteries, Angew. Chem.
Kim: Data curation, Formal analysis. Seoin Back: Conceptualization, Int. Ed. 60 (2021) 25575–25582, doi:10.1002/anie.202112183.
Funding acquisition, Supervision, Writing – review & editing. Seung-Ho [17] J. Wang, Y. Bo, T. Gao, X. Wang, W. Li, X. Hong, Z. Wang, H. He, Reduced graphene
oxide modified few-layer exfoliated graphite to enhance the stability of the negative
Yu: Conceptualization, Funding acquisition, Supervision, Writing – re-
electrode of a graphite-based potassium ion battery, Acta Phys. -Chim. Sin. 38 (2022)
view & editing. Yuanzhe Piao: Conceptualization, Funding acquisition, 2012088, doi:10.3866/pku.Whxb202012088.
Supervision, Writing – review & editing. [18] Y. Du, Z. Yi, B. Chen, J. Xu, Z. Zhang, J. Bao, X. Zhou, Sn4 P3 nanoparticles confined in
multilayer graphene sheets as a high-performance anode material for potassium-ion
Acknowledgments batteries, J. Energy Chem. 66 (2022) 413–421, doi:10.1016/j.jechem.2021.08.043.
[19] Q. Yu, B. Jiang, J. Hu, C.-.Y. Lao, Y. Gao, P. Li, Z. Liu, G. Suo, D. He, W.(Alex). Wang,
G. Yin, Metallic octahedral CoSe2 threaded by N-doped carbon nanotubes: a flex-
J. Choi, A. Jin, and H. D. Jung contributed equally to this work. ible framework for high-performance potassium-ion batteries, Adv. Sci. 5 (2018)
Y. P. acknowledges the support of the Basic Science Research Program 1800782, doi:10.1002/advs.201800782.
[20] S.M. Ahmed, G. Suo, W.A. Wang, K. Xi, S.B. Iqbal, Improvement in potassium ion
through the National Research Foundation of Korea (NRF) funded by batteries electrodes: recent developments and efficient approaches, J. Energy Chem.
the Ministry of Education (NRF-2021R1A2C1008380) and the Nano Ma- 62 (2021) 307–337, doi:10.1016/j.jechem.2021.03.032.
terial Technology Development Program (NRF-2015M3A7B6027970) [21] J. Ding, H. Zhang, H. Zhou, J. Feng, X. Zheng, C. Zhong, E. Paek, W. Hu, D. Mitlin,
Sulfur-grafted hollow carbon spheres for potassium-ion battery anodes, Adv. Mater.
of MSIP/NRF. S.-H. Y. acknowledges the support of the National Re- 31 (2019) 1900429, doi:10.1002/adma.201900429.
search Foundation of Korea (NRF) through a grant funded by the Ko- [22] W. Wang, J. Zhou, Z. Wang, L. Zhao, P. Li, Y. Yang, C. Yang, H. Huang, S. Guo,
rean government (MSIT) (NRF-2020R1C1C1012308). S. B. acknowl- Short-range order in mesoporous carbon boosts potassium-ion battery performance,
Adv. Energy Mater. 8 (2018) 1701648, doi:10.1002/aenm.201701648.
edges the support from the National Research Foundation of Korea
[23] W. Zhang, Z. Cao, W. Wang, E. Alhajji, A.H. Emwas, P.M.F.J. Costa, L. Cav-
(NRF) through a grant that was funded by the Ministry of Science allo, H.N. Alshareef, A site-selective doping strategy of carbon anodes with re-
and ICT (2015M3D3A1A01064929) and the generous supercomputing markable K-ion storage capacity, Angew. Chem. Int. Ed. 59 (2020) 4448–4455,
doi:10.1002/anie.201913368.
time from KISTI (KSC-2021-CRE-0060). A. J. acknowledges the support
[24] A. Eftekhari, Low voltage anode materials for lithium-ion batteries, Energy Storage
of theNational Research Foundation of Korea (NRF) through a grant Mater. 7 (2017) 157–180, doi:10.1016/j.ensm.2017.01.009.
funded by the Korean government (MSIT) (NRF-2020R1A2C1012342). [25] Y. Liu, Y.X. Lu, Y.S. Xu, Q.S. Meng, J.C. Gao, Y.G. Sun, Y.S. Hu, B.B. Chang, C.T. Liu,
J. H. U. acknowledges the support provided by Basic Science Research A.M. Cao, Pitch-derived soft carbon as stable anode material for potassium ion bat-
teries, Adv. Mater. 32 (2020) 2000505, doi:10.1002/adma.202000505.
Program through the National Research Foundation of Korea (NRF) [26] Y. Wang, Z. Wang, Y. Chen, H. Zhang, M. Yousaf, H. Wu, M. Zou, A. Cao,
funded by the Ministry of Education (NRF-2021R1I1A1A01044891). R.P.S. Han, Hyperporous sponge interconnected by hierarchical carbon nanotubes as

333
J. Choi, A. Jin, H.D. Jung et al. Energy Storage Materials 48 (2022) 325–334

a high-performance potassium-ion battery anode, Adv. Mater. 30 (2018) 1802074, [41] Z. Ju, P. Li, G. Ma, Z. Xing, Q. Zhuang, Y. Qian, Few layer nitrogen-doped graphene
doi:10.1002/adma.201802074. with highly reversible potassium storage, Energy Storage Mater. 11 (2018) 38–46,
[27] K. Share, A.P. Cohn, R. Carter, B. Rogers, C.L. Pint, Role of nitrogen-doped graphene doi:10.1016/j.ensm.2017.09.009.
for improved high-capacity potassium ion battery anodes, ACS Nano 10 (2016) [42] J. Zhao, X. Zou, Y. Zhu, Y. Xu, C. Wang, Electrochemical intercalation
9738–9744, doi:10.1021/acsnano.6b05998. of potassium into graphite, Adv. Funct. Mater. 26 (2016) 8103–8110,
[28] Y. Gao, J. Zhang, N. Li, X. Han, X. Luo, K. Xie, B. Wei, Z. Xia, Design principles doi:10.1002/adfm.201602248.
of pseudocapacitive carbon anode materials for ultrafast sodium and potassium-ion [43] L. Fan, R. Ma, Q. Zhang, X. Jia, B. Lu, Graphite anode for a potassium-ion battery
batteries, J. Mater. Chem. A 8 (2020) 7756–7764, doi:10.1039/D0TA01821J. with unprecedented performance, Angew. Chem. Int. Ed. 58 (2019) 10500–10505,
[29] J.H. Lee, N. Park, B.G. Kim, D.S. Jung, K. Im, J. Hur, J.W. Choi, Restacking-inhibited doi:10.1002/anie.201904258.
3D reduced graphene oxide for high performance supercapacitor electrodes, ACS [44] Y. Xu, C. Zhang, M. Zhou, Q. Fu, C. Zhao, M. Wu, Y. Lei, Highly nitrogen doped
Nano 7 (2013) 9366–9374, doi:10.1021/nn4040734. carbon nanofibers with superior rate capability and cyclability for potassium ion
[30] H. Ma, H. Geng, B. Yao, M. Wu, C. Li, M. Zhang, F. Chi, L. Qu, Highly ordered batteries, Nat. Commun. 9 (2018) 1720-1711, doi:10.1038/s41467-018-04190-z.
graphene Solid: an efficient platform for capacitive sodium-ion storage with ultra- [45] X. Hu, Y. Liu, J. Chen, L. Yi, H. Zhan, Z. Wen, Fast redox kinetics in
high volumetric capacity and superior rate capability, ACS Nano 13 (2019) 9161– bi-heteroatom doped 3D porous carbon nanosheets for high-performance hy-
9170, doi:10.1021/acsnano.9b03492. brid potassium-ion battery capacitors, Adv. Energy Mater. 9 (2019) 1901533,
[31] F.J. Sonia, M.K. Jangid, B. Ananthoju, M. Aslam, P. Johari, A. Mukhopadhyay, Un- doi:10.1002/aenm.201901533.
derstanding the Li-storage in few layers graphene with respect to bulk graphite: ex- [46] Y. Li, R.A. Adams, A. Arora, V.G. Pol, A.M. Levine, R.J. Lee, K. Akato,
perimental, analytical and computational study, J. Mater. Chem. A 5 (2017) 8662– A.K. Naskar, M.P. Paranthaman, Sustainable potassium-ion battery anodes de-
8679, doi:10.1039/C7TA01978E. rived from waste-tire rubber, J. Electrochem. Soc 164 (2017) A1234–A1238,
[32] D. Ko, J. Choi, B. Yan, T. Hwang, X. Jin, J.M. Kim, Y. Piao, A facile and scal- doi:10.1149/2.1391706jes.
able approach to develop electrochemical unzipping of multi-walled carbon nan- [47] E. Zhang, X. Jia, B. Wang, J. Wang, X. Yu, B. Lu, Carbon dots@rGO paper as free-
otubes to graphene nanoribbons, J. Mater. Chem. A 8 (2020) 22045–22053, standing and flexible potassium-ion batteries anode, Adv. Sci. 7 (2020) 2000470,
doi:10.1039/D0TA03782F. doi:10.1002/advs.202000470.
[33] A.M. Dimiev, A. Khannanov, I. Vakhitov, A. Kiiamov, K. Shukhina, J.M. Tour, [48] B. Cao, Q. Zhang, H. Liu, B. Xu, S. Zhang, T. Zhou, J. Mao, W.K. Pang, Z. Guo,
Revisiting the mechanism of oxidative unzipping of multiwall carbon nanotubes A. Li, J. Zhou, X. Chen, H. Song, Graphitic carbon nanocage as a stable and high
to graphene nanoribbons, ACS Nano 12 (2018) 3985–3993, doi:10.1021/ac- power anode for potassium-ion batteries, Adv. Energy Mater 8 (2018) 1801149,
snano.8b01617. doi:10.1002/aenm.201801149.
[34] X. Wang, J. Wang, D. Wang, S. Dou, Z. Ma, J. Wu, L. Tao, A. Shen, C. Ouyang, Q. Liu, [49] F.J. Sonia, M.K. Jangid, M. Aslam, P. Johari, A. Mukhopadhyay, Enhanced and faster
S. Wang, One-pot synthesis of nitrogen and sulfur co-doped graphene as efficient potassium storage in graphene with respect to graphite: a comparative study with
metal-free electrocatalysts for the oxygen reduction reaction, Chem. Commun. 50 lithium storage, ACS Nano 13 (2019) 2190–2204, doi:10.1021/acsnano.8b08867.
(2014) 4839–4842, doi:10.1039/C4CC00440J. [50] W. Tang, E. Sanville, G. Henkelman, A grid-based dader analysis algo-
[35] G. Zhou, E. Paek, G.S. Hwang, A. Manthiram, Long-life Li/polysulphide bat- rithm without lattice bias, J. Phys. Condens. Matter. 21 (2009) 084204,
teries with high sulphur loading enabled by lightweight three-dimensional doi:10.1088/0953-8984/21/8/084204.
nitrogen/sulphur-codoped graphene sponge, Nat. Commun. 6 (2015) 7760, [51] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for metals
doi:10.1038/ncomms8760. and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6 (1996) 15–
[36] A.M. El-Sawy, I.M. Mosa, D. Su, C.J. Guild, S. Khalid, R. Joesten, J.F. Rusling, 50, doi:10.1016/0927-0256(96)00008-0.
S.L. Suib, Controlling the active sites of sulfur-doped carbon nanotube-graphene [52] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy cal-
nanolobes for highly efficient oxygen evolution and reduction catalysis, Adv. En- culations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169–11186,
ergy Mater. 6 (2016) 1501966, doi:10.1002/aenm.201501966. doi:10.1103/PhysRevB.54.11169.
[37] L. Tao, Y. Yang, H. Wang, Y. Zheng, H. Hao, W. Song, J. Shi, M. Huang, D. Mitlin, [53] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made sim-
Sulfur-nitrogen rich carbon as stable high capacity potassium ion battery anode: ple, Phys. Rev. Lett. 77 (1996) 3865–3868, doi:10.1103/PhysRevLett.77.3865.
performance and storage mechanisms, Energy Storage Mater. 27 (2020) 212–225, [54] S. Grimme, Semiempirical GGA-type density functional constructed with a
doi:10.1016/j.ensm.2020.02.004. long-range dispersion correction, J. Comput. Chem. 27 (2006) 1787–1799,
[38] W. Zhang, J. Ming, W. Zhao, X. Dong, M.N. Hedhili, P.M.F.J. Costa, H.N. Alshareef, doi:10.1002/jcc.20495.
Graphitic nanocarbon with engineered defects for high-performance potassiumi bat- [55] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (1994) 17953–
tery anodes, Adv. Funct. Mater. 29 (2019) 1903641, doi:10.1002/adfm.201903641. 17979, doi:10.1103/PhysRevB.50.17953.
[39] C.-.J. Shih, A. Vijayaraghavan, R. Krishnan, R. Sharma, J.-.H. Han, M.-.H. Ham, [56] H.J. Monkhorst, J.D. Pack, Special points for brillouin-zone integrations, Phys. Rev.
Z. Jin, S. Lin, G.L.C. Paulus, N.F. Reuel, Q.H. Wang, D. Blankschtein, M.S. Strano, B 13 (1976) 5188–5192, doi:10.1103/PhysRevB.13.5188.
Bi- and trilayer graphene solutions, Nat. Nanotechnol. 6 (2011) 439–445, [57] G. Henkelman, B.P. Uberuaga, H. Jónsson, Climbing image nudged elastic band
doi:10.1038/nnano.2011.94. method for finding saddle points and minimum energy paths, J. Chem. Phys. 113
[40] D. Li, Q. Sun, Y. Zhang, X. Dai, F. Ji, K. Li, Q. Yuan, X. Liu, L. Ci, Fast and stable K-ion (2000) 9901–9904, doi:10.1063/1.1329672.
storage enabled by synergistic interlayer and pore-structure engineering, Nano Res
14 (2021) 4502–4511, doi:10.1007/s12274-021-3324-0.

334

You might also like