ArticlePDF Available

N-Doping and Defective Nanographitic Domain Coupled Hard Carbon Nanoshells for High Performance Lithium/Sodium Storage

Wiley
Advanced Functional Materials
Authors:

Abstract and Figures

Hard carbons (HCs) possess high lithium/sodium storage capacities, which however suffer from low electric conductivity and poor ion diffusion kinetics. An efficient structure design with appropriate heteroatoms doping and optimized graphitic/defective degree is highly desired to tackle these problems. This work reports a new design of N-doped HC nanoshells (N-GCNs) with homogeneous defective nanographite domains, fabricated through the prechelation between Ni2+ and chitosan and subsequent catalyst confined graphitization. The as-prepared N-GCNs deliver a high reversible lithium storage capacity of 1253 mA h g−1, with outstanding rate performance (175 mA h g−1 at a high rate of 20 A g−1) and good cycling stability, which outperforms most state-of-the-art HCs. Meanwhile, a high reversible sodium storage capacity of 325 mA h g−1 is also obtained, which stabilizes at 174 mA h g−1 after 200 cycles. Density functional theory calculations are performed to uncover the coupling effect between heteroatom-doping and the defective nanographitic domains down to the atomic scale. The in situ Raman analysis reveals the “adsorption mechanism” for sodium storage and the “adsorption–intercalation mechanism” for lithium storage of N-GCNs.
This content is subject to copyright. Terms and conditions apply.
FULL PAPER
www.afm-journal.de
© 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1706294 (1 of 10)
N-Doping and Defective Nanographitic Domain Coupled
Hard Carbon Nanoshells for High Performance Lithium/
Sodium Storage
Shifei Huang, Zhiping Li, Bo Wang, Jiujun Zhang, Zhangquan Peng, Ruijuan Qi,
Jing Wang, and Yufeng Zhao*
Hard carbons (HCs) possess high lithium/sodium storage capacities, which
however suffer from low electric conductivity and poor ion diffusion kinetics.
An efficient structure design with appropriate heteroatoms doping and opti-
mized graphitic/defective degree is highly desired to tackle these problems.
This work reports a new design of N-doped HC nanoshells (N-GCNs) with
homogeneous defective nanographite domains, fabricated through the pre-
chelation between Ni2+ and chitosan and subsequent catalyst confined gra-
phitization. The as-prepared N-GCNs deliver a high reversible lithium storage
capacity of 1253 mA h g1, with outstanding rate performance (175 mA h g1
at a high rate of 20 A g1) and good cycling stability, which outperforms most
state-of-the-art HCs. Meanwhile, a high reversible sodium storage capacity
of 325 mA h g1 is also obtained, which stabilizes at 174 mA h g1 after
200 cycles. Density functional theory calculations are performed to uncover
the coupling effect between heteroatom-doping and the defective nanogra-
phitic domains down to the atomic scale. The in situ Raman analysis reveals
the “adsorption mechanism” for sodium storage and the “adsorption–
intercalation mechanism” for lithium storage of N-GCNs.
DOI: 10.1002/adfm.201706294
Dr. S. Huang, Prof. Z. Li, B. Wang, Prof. J. Wang, Prof. Y. Zhao
Key Laboratory of Applied Chemistry
Yanshan University
Qinhuangdao 066004, China
E-mail: yufengzhao@ysu.edu.cn
Prof. J. Zhang, Prof. Y. Zhao
College of Sciences
Shanghai University
Shanghai 200444, China
Prof. Z. Peng
State Key Laboratory of Electroanalytical Chemistry
Changchun Institute of Applied Chemistry Chinese Academy of Science
Changchun, Jilin 130022, China
Prof. R. Qi
Key Laboratory of Polar Materials and Devices
Ministry of Education
East China Normal University
Shanghai 200241, China
demanded for potential applications
in portable electronics, electrical vehi-
cles, and smart grids, etc.[1,2] Currently,
graphite is widely applied as the anode
material for commercial lithium-ion bat-
teries (LIBs), which however suffers
from limited energy and power perfor-
mance due to its low theoretical capacity
(372 mA h g1) and sluggish lithium dif-
fusion kinetics.[3–5] On the other hand,
graphite is inapplicable to sodium-ion bat-
teries (SIBs) due to the large ionic radius
of Na+ ions (1.02 vs 0.76 Å),[6,7] and the
sodium–graphite system is thermodynam-
ically instable.[6] Therefore, exploring new
anode materials with favorable voltage,
low cost, desired capacity, and high rate
of charge for LIBs/SIBs is of great impor-
tance for practical applications.
In recent years, hard carbons (HCs)
have been emerging as one of the most
promising anodes for LIBs/SIBs,[8,9] which
can be divided into polymer derived HCs
and biomass derived HCs according to different types of the
carbon precursors.[6,7,10–14] These HCs are composed of plen-
tiful randomly packed single graphene layers as described by
the “house of cards” model,[15] which can provide numerous
nanopores and large surface area, thus offering significantly
enhanced capacity by facilitating both intercalation and adsorp-
tion of Li+ or Na+ ions.[16] Nevertheless, the low graphitization
degree of HCs normally cause low electric conductivity and
electrochemical instability.[1,17,18] Attempts have been made by
compositing HCs with graphitic carbon, which however cannot
overcome the disadvantage of the intrinsically disordered micro-
architecture.[18–21] Creating nanographitic domains in HCs
through catalyst assisted synthesis, demonstrates an efficient
way to improve the electric conductivity.[1,22–24] Such unique
structure possesses plenty of nanopore defects in the graphitic
domains, thereby enable stable charge storage through both
interlayer intercalation (nanographite layers) and interfacial
adsorption (nanopore defects). On the other hand, doping het-
eroatoms such as B, N, S, P, etc. to the carbon architecture, can
introduce more active sites and facilitate the charge transfer,
thus significantly enhance the Li/Na storage capability.[25–31]
Therefore, the synergistic effect between heteroatom-doping
and nanopore defects in the nanographitic domains are obliged
Lithium/Sodium Storage
1. Introduction
Electrochemical energy storage technologies based on recharge-
able batteries (e.g., lithium or sodium-ion batteries) with high
storage capacity and excellent rate performance are critically
Adv. Funct. Mater. 2018, 1706294
www.afm-journal.dewww.advancedsciencenews.com
1706294 (2 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
to be considered in further research efforts. An effective design
of hard carbons with optimized nanoarchitecture and corre-
sponding theoretical study are critically demanded for high per-
formance lithium/sodium storage.
In this work, we report a new design of N-doped hard carbon
nanoshells with tailored graphitic/defective degree (N-GCNs),
whereby Ni2+ ions are prechelated with chitosan carbon
source to enable a catalyst confined graphitization process and
then assure the homogeneous distribution of nanographitic
domains. As expected, the as-prepared N-GCNs perform out-
standing charge storage capacities and rate performances for
both LIBs and SIBs. The density functional theory (DFT) cal-
culations evidence the synergistic effect between N-doping and
nanopore defects in the nanographitic domains for Li storage,
which demonstrates that proper amount of defects and het-
eroatom doping will improve the Li binding ability, charge
transfer ability, and increase the density of states (DOS) around
Fermi level. In situ Raman analysis unravels the different
charge storage mechanism of N-GCNs for LIBs (adsorption–
intercalation mechanism) and SIBs (adsorption mechanism).
2. Results and Discussion
2.1. Characterization
The N-GCNs were fabricated through a one-pot catalyst con-
fined graphitization process as illustrated in Figure 1a. Typically,
Artemia cyst shell was used as carbon precursor and soaked
with aqueous solution containing Ni2+ ions and KOH, whereby
the Ni2+ is chelated with chitosan to make sure the uniform
distribution of Ni2+ ions in the carbon precursor.[32,33] The
KOH reacts immediately with Ni2+, forming homogeneously
dispersed Ni(OH)2 nanoparticles (NPs) (decomposition yields
NiO at 230 °C). The carbon precursors then go through a
pyrolysis-carbonization process at the elevated temperature,
where the overamount KOH functions as the activating agent
by reacting with carbon forming noncatalytic potassium com-
pounds (K2O, K2CO3, etc., Figure S1a, Supporting Informa-
tion)[32–34] and the NiO are reduced to metallic Ni NPs and
serve as the catalyst for the subsequent graphitization.[35–37]
Simultaneously, the carbon around Ni is catalyzed at 850 °C to
form a graphitic nanoshell (Figure S1b, Supporting Informa-
tion),[35] while the carbon away from Ni remains amorphous,
demonstrating a catalyst confined graphitization process.
The possible reaction equations are listed in the Supporting
Information.[32–37] The graphitic degree of the HCs is tailored
through adjusting the calcination temperature to 750 and
950 °C, respectively. The as-formed noncatalytic potassium
compounds together with the inorganic impurities from carbon
precursor also affect the catalysis reaction, leading to the dis-
continuous feature of the graphitic stripes. These impurities
are removed with acid washing, and form nanopores between
the microcrystal stripes in the graphitic domains. The scanning
electron microscopic (SEM) images and the corresponding
elemental mapping of the N-GCNs before acid washing
(Figure 1b–e; Figure S1c, Supporting Information) illustrate the
interconnected graphitic carbon nanoshells and the imbedded
Ni NPs with noncatalytic potassium compound impurities,
which agree well with the above explanations.
Raman spectra demonstrate (Figure 2a) three characteristic
peaks including D-band (1345 cm1), G-band (1570 cm1),
and 2D-band (2700 cm1). The D-band reflects the disorder
degree of the crystal structure due to the breathing motion of
sp2 atoms in the amorphous carbon or at edge planes/defects
of the graphitic domains, while G-band represents a scattering
E2g order vibration mode of atoms in rings and chains.[38,39]
2D-band is a two-phonon resonance Raman spectra peak,
reflecting the extent of the stacked graphene layers.[40] The
ID/IG ratio of N-GCNs is calculated as 0.85, indicating the
Adv. Funct. Mater. 2018, 1706294
Figure 1. The synthetic mechanism inference of N-GCNs. a) Schematic illustration of the synthesis process of N-GCNs. b) SEM images of the
unwashed sample and c–e) the corresponding element mapping images.
www.afm-journal.dewww.advancedsciencenews.com
1706294 (3 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
highly ordered structure of the as prepared HC. The well-
crystalline structure of N-GCNs is confirmed by the X-ray dif-
fraction (XRD) patterns (Figure S2a, Supporting Information),
which exhibits a sharp peak at 2
θ
–26.4°, corresponding to the
(002) plane of the graphitic domains with expanded interlayer
distance of 3.373 Å,[41] while the broad hump is attributed to
the amorphous structure in the partially graphitized samples.
For comparison, the diffraction patterns (Figure S2a, Sup-
porting Information) and Raman spectra (Figure S2b, Sup-
porting Information) of the control sample prepared at 750 °C
(N-GCNs-750) and 950 °C (N-GCNs-950) were also obtained,
which suggest the enhanced crystallinity with the increasing
heat temperature. The surface chemical composition of the
N-GCNs was examined by X-ray photoelectron spectrum (XPS),
indicating the successful doping of N (1.3 at%) in the N-GCNs
(Figure S2c, Supporting Information). The high-resolution N1s
spectrum (Figure 2b) can be deconvoluted into three peaks
ascribed to the pyridinic N (398.3 eV), pyrrolic N (399.9 eV), and
graphitic N (401.2 eV), with a proportion of 24.5%, 57%, and
18.5%, respectively.[32,42] The doping of N to the carbon archi-
tecture can introduce more active sites and facilitate the charge
transfer, thus significantly enhance the Li/Na storage capa-
bility.[25,26] In addition, the very high content of pyridinic and
pyrollic N in the N-GCNs, are believed more beneficial to the
storage of Li/Na ions.[43] The Fourier-transform infrared (FTIR)
spectra (Figure S2f, Supporting Information) demonstrate
a strong characteristic peak at 3435 cm1, which is attributed
to the stretching vibration of the OH bond. Small peaks at
2920 and 2840 cm1 arise from the stretching vibration of the
CH (CH and HCO) bond. The peaks at 1700, 1620,
1380, 1210, and 1090 cm1, can be assigned to the stretching
vibration of carboxyl groups CO, CC, CO, COC, and
Adv. Funct. Mater. 2018, 1706294
Figure 2. a) Raman spectra, b) deconvoluted XPS spectra (N1s), c) porosity distribution curve by original density functional theory of N-GCNs, d) SEM
image, e,g) amplified SEM, f,h,i) TEM image, and j–l) HRTEM images of N-GCNs (inset: enlarged image).
www.afm-journal.dewww.advancedsciencenews.com
1706294 (4 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
CC bond. The peak at 1310 cm1 is corresponding to the
stretching vibration of CN band.[32] It has been reported that,
proper amount of oxygen functional groups on the carbon
framework can react with lithium/sodium ions, resulting in
improved storage capacity.[44] For example, the CO groups can
be reduced by Li+/Na+ when the voltage versus Li/Na is higher
than 1.5 V (Equation (S7), Supporting Information), thus con-
tributing a certain degree of capacity.[45,46] The type I isotherm
plot from nitrogen adsorption–desorption test (Figure S2g,
Supporting Information) suggests the dominant amount of
micropores,[47] with a total Brunauer–Emmett–Teller (BET) spe-
cific surface area of 1490 m2 g1 and pore volume of 2.03 cm3
g1, and the porosity distribution by original density functional
theory (Figure 2c) reveals the hierarchical porous structure of
N-GCNs. The SEM image (Figure 2d) illustrates the unique
3D interconnected carbon nanoshell architecture of N-GCNs
with diameter sizes from 200 to 600 nm. The amplified SEM
images (Figure 2e,g) and the corresponding transmission
electron microscopic (TEM) images (Figure 2f,h) clearly illus-
trate the open hollow characteristic of the nanoshell, with the
wall thickness of 20–50 nm (Figure 2i). The high-resolution
TEM (HRTEM) images (Figure 2j–l) demonstrate the tur-
bostratic carbon structure of N-GCNs, with ordered nanogra-
phitic domains (the nanovoids surrounded by some parallel
carbon hexagonal layers). The graphite microcrystal stripes in
the ordered nanodomains are discontinuous, and some gra-
phitic layers are crisscrossed with each other in the ordered
nanodomains, giving rise to defects such as edges and pores,
which will provide sufficient active sites for Li+/Na+ storage.[14]
The HRTEM confirms the expanded interlayer distance of
3.47–4.0 Å, which would favor the intercalation of Li+/Na+ ions.
By contrast, the detailed characterization of N-GCNs-750 and
N-GCNs-950 is also provided (Figures S2c–f and S3, Tables S1
and S2, Supporting Information).
2.2. Electrochemical Performances
The Li+ storage behavior in N-GCNs was evaluated using cyclic
voltammetry (CV) and galvanostatic charge–discharge (GCD)
techniques (Figure 3; Figure S4, Supporting Information).
Figure S4a shows the CV curves for the first five cycles of the
electrode at a scanning rate of 0.1 mV s1 in the voltage range
of 0.01–3.0 V versus Li+/Li. In the first discharge cycle, the irre-
versible peaks at 0.75 V of LIBs should be attributed to the solid
electrolyte interphase (SEI) film formation and the decomposi-
tion of the electrolyte, and the peak at 1.7 V is caused by the
loss of some irreversible lithium storage sites during the initial
lithiation process.[48,49] While the sharp peak at 0.01 V can also
be imputed to the irreversible reaction of electrolyte with sur-
face functional groups and the formation of SEI film.[50] The
three pronounced peaks at 0.23 V in the corresponding charge
cycles (Figure S4a, Supporting Information) are attributed to
the delithiation of the graphitic carbon layers,[51] which become
more intense with the increase of cycle count attributed to the
electrochemical activation process.[52] The overlap of the CV
curves in the subsequent four cycles indicates the reversible
electrochemical reactions in the electrode. The N-GCNs display
a high discharge capacity of 1253 mA h g1 in the second cycle
at 0.1 A g1 for N-GCNs, which stabilizes at 1236 mA h g1 after
100 cycles (Figure 3a). The corresponding GCD voltage profiles
are illustrated in Figure 3b. The initial coulombic efficiency is
calculated as 53.4%, the capacity loss should be attributed to the
formation of an SEI layer and the decomposition of the elec-
trolyte on the surfaces of the HC nanoshells.[3] A plateau at
0.23 V is detected in all charge voltage profiles, corresponding
to the delithiation of the graphitic carbon layers. Most promis-
ingly, the specific capacity of N-GCNs remains 175 mA h g1
even at a ultrahigh rate of 20 A g1 (Figure 3c,d), and when
the current density directly reduced from 20 to 0.1 A g1,
the capacity quickly recovered to 965 mA h g1 (increased to
1160 mA h g1 gradually), demonstrating excellent charge
transfer kinetics within the N-GCNs architecture. Promis-
ingly, the N-GCNs also exhibit impressive electrochemical per-
formance in SIBs (Figure 3e–h), with a reversible discharge
capacity of 325 mA h g1 at 100 mA g1, which stabilizes at
174 mA h g1 after 200 cycles. A capacity of 63 mA h g1 can still
be maintained at high rate of 5 A g1. Interestingly, there is no
pronounced peak within the potential range of 0.01–3.0 V (vs
Na+/Na) for SIBs (Figure S4b, Supporting Information), indi-
cating that Na+ cannot be intercalated to the graphitic carbon
layers because of its larger ionic radius. Note that, fluctuation
is observed from the cycling curves for LIBs (Figure 3a), which
does not happen for NIBs (Figure 3e). This maybe because of
that, during the charge–discharge process, the intercalation
of lithium ions would cause the internal structure changes of
carbon materials, and formation of new electrode/electrolyte
interfaces, providing more active sites for lithium storage.
Meanwhile, the uninterrupted reformation of SEI on the new
interface will consume the electrolyte and has a certain nega-
tive influence on the energy storage of the materials. While, the
sodium ions are mainly stored through absorption/desorption
processes, which have little influence on the structure of the
material, resulting in a stable cycling performance. This result
agrees well with the literatures.[1,22,53] For comparison, the
corresponding LIBs/SIBs test for the samples prepared at
750 and 950 °C with varied graphitic degree was also conducted
(Figures S4 and S5, Supporting Information). A comparison
between the electrochemical performance of N-GCNs and
carbon materials previously reported are provided for both LIBs
(Table S3, Supporting Information) and SIBs (Table S4, Sup-
porting Information), indicating the superior lithium/sodium
storage capability of N-GCNs. The outperforming performance
of N-GCNs in terms of reversible capacity, cycling stability, and
rate performance, etc., implies that the unique nanostructure
of the as-prepared HCs plays a critical role in electrochemical
applications, which inspires us to further investigate it with in
situ Raman spectroscopy and theoretical calculations.
2.3. In Situ Raman Spectroscopy Study
The in situ Raman spectroscopy study was conducted during
the charge/discharge process of both LIBs and SIBs, to
moni tor the ion interclation or adsorption in the carbon mate-
rials,[48,54,55] and unravel the critical potential for ion adsorp-
tion on the carbonaceous surface and intercalation into the
graphitic layers.[56,57] Figure 4a–d shows the in situ Raman
Adv. Funct. Mater. 2018, 1706294
www.afm-journal.dewww.advancedsciencenews.com
1706294 (5 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
Adv. Funct. Mater. 2018, 1706294
Figure 3. Electrochemical properties. Electrochemical performance of N-GCNs tested against Lithium and Sodium: cycling performance and GCD
profiles of N-GCNs at a current density of 0.1 A g1 in LIBs. a,b) Cycling performance and GCD profiles of N-GCNs at a current density of 0.1 A g1 in
LIBs. c,d) GCD profiles of N-GCNs at different rates of 0.1 to 20 A g1 and rate performance gradually increasing from 0.1 to 20 A g1 and then back
to 0.1 A g1 in LIBs. e,f) Cycling performance and GCD profiles of N-GCNs at a current density of 0.1 A g1 in SIBs. g,h) GCD profiles of N-GCNs at
different rates of 0.1 to 5 A g1 and rate performance gradually increasing from 0.1 to 5 A g1 and then back to 0.1 A g1 in SIBs.
www.afm-journal.dewww.advancedsciencenews.com
1706294 (6 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
spectra and the corresponding discharge–charge profiles
at the first cycle of N-GCNs for LIBs (Figure 4a,b) and SIBs
(Figure 4c,d) with a current density of 0.25 A g1. The intercala-
tion of guest ions into carbonaceous host strongly affects the
position, shape, and intensity of the D and G-band (1330 and
1600 cm1 at the open-circuit voltage (OCV, 3.0 V vs Li/Li+
or Na/Na+)). The decreased intensity of D-band from OCV to
0.01 V can be attributed to the introduction of Li+/Na+, which
occupy the defective sites, i.e., the nanopores in the nanogra-
phitic domains and the amorphous structure of N-GCNs,
thus the breathing motion of sp2 atoms in the rings at edge
planes is limited by the absorbed Li+/Na+ ions. Meanwhile, the
G-band of N-GCNs shows unconspicuous red-shifts from OCV
to 0.2 V in LIBs (Figure 4a) and from OCV to 0.01 V in SIBs
(Figure 4c) during discharge process. This can be accounted
for the charge transfer effects from the adsorption of Li+/Na+
ions and the according formation of SEI layer (below 0.8 V
in LIBs (Figure 4b) and 0.6 V in SIBs (Figure 4d)).[49] These
Li+/Na+ ions randomly accommodate the nanoporous defective
sites and fully cover the N-GCNs surface at relatively higher
voltage, thus causing the early disappearance of G-band. The
intensity of G-band steadily decreases with potential and
finally fades into the signal noise from 0.2 to 0.01 V in LIBs
because of the weaken of resonance caused by Li+ intercalation,
which indicates the lithiation of the nanographitic domains
at 0.2 V.[49,58] This result is further confirmed with ex situ
XRD test (Figure S6, Supporting Information). Meanwhile,
a large amount of nanovoids filling is in accordance with the
plateau region close to 0.01 V in the corresponding discharge
curves (Figure 4b). However, when the N-GCNs are explored
as SIB electrode, the discharge curves exhibit smooth slope
from OCV to 0.6 V and 0.6 to 0.01 V, neither evident inten-
sity decrease from G-band, nor corresponding plateau region
in the discharge and charge curves are observed (Figure 4b).
This indicates no obvious ion intercalation happened in the
graphitic domains due to the larger ionic radius of Na+. Thus,
the Na+ can only be adsorbed on the disordered carbon archi-
tectures of N-GCNs, and the nanovoids filling region is also
close to 0.01 V.[50] So the “adsorption–intercalation mechanism
for Li+ storage and “adsorption mechanism” for Na+ storage
of N-GCNs are verified through the in situ Raman analysis.
The position, shape, and intensity of the D and G-band was
recovered progressively because of the extraction of the Li+/Na+
during the charge process which also indicates that the good
reversibility of the materials.[57]
2.4. The First-Principles Calculations
The first-principles calculations were performed to unveil the
origin of excellent electrochemical performance of N-GCNs,
based on DFT as implemented in the Vienna ab initio simula-
tion package. The defective nanographitic domain of N-GCNs
was modeled using a periodically repeated slab of N-doped
graphene with nanopore defects, denoted as NDGs. By con-
trast, control systems including NGs (N-doped nondefective
graphene sheet), DGs (doping free defective graphene sheets),
and PGs (pristine graphene sheets) were also constructed. The
most stable electronic structures for Li absorbed systems were
proposed and optimized (Figure 5a–d) and the corresponding
binding energy (EB) is calculated by Equation (1) (Table S5,
Supporting Information)
=−EEE
BPR
(1)
where EP is the total energy of the Li absorbed graphene and ER
is the total energy of an isolated Li atom and the substrate. The
results demonstrate that, the most stable site for Li adsorption
is at the center of a carbon hexagon (hollow site) for both PGs
and NGs, demonstrating that the Li atom prefers C rather than
N atoms in the graphene structures. The EB between absorbed
Li atom and substrate NGs/PGs is calculated as 0.49/0.95 eV
(Table S5, Supporting Information), indicating weaker interac-
tion between Li and NGs. Therefore, simple nitrogen doping
cannot effectively improve the Li atom adsorption capa-
bility of graphene. Interestingly, the most stable sites of DGs
(Figure 5c) and NDGs (Figure 5d) are at the center of nano-
pore defects. This implies that, the electron-accepting tendency
of the defective nanoarchitecture is beneficial to gaining elec-
tron from the Li atom.[59] As compared to NGs and PGs, these
defective structures demonstrate several folds of improvement
on the binding energy, i.e., 3.61 eV for NDGs, and 3.17 eV
Adv. Funct. Mater. 2018, 1706294
Figure 4. In-situ Raman spectroscopy test. In situ Raman spectra and the corresponding discharge-charge profiles at the first cycle of N-GCNs for LIBs
a,b) and SIBs c,d) with the current density of 0.25 A g1, respectively.
www.afm-journal.dewww.advancedsciencenews.com
1706294 (7 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
for DGs, suggesting that Li adsorption capability of nanogra-
phitic domains can be significantly enhanced by introducing
nanopore defects. On the other hand, the more negative EB
of NDGs than DGs provides direct evidence of the synergistic
effect between N-doping and nanopore defects for Li storage.[43]
To get further insight into the electronic structure during
the lithiation process, the charge density difference (yellow/
light blue color represents positive/negative charge differences,
respectively) between the charge densities of the adsorp-
tion system and those of the separated substrate/adatom (at
optimized positions) (Equation (2))[60] was also calculated
(Figure 5e–h; Figure S7a,b, Supporting Information)
ρρ ρρ
∆= +−[Lisubstrate][Li][substrate]
(2)
Figure 5e,f illustrates the large degree of charge transfer
between Li and the adjacent C atoms in PGs, whereby Li is
obviously oxidized by transferring the valence electrons of 2s to
carbon.[61] Whereas the NGs exhibits an electron rich character
which leads to the depressed adsorption ability for Li atoms
and attenuated charge transfer between Li and the adjacent C
atoms (Figure S7a,b, Supporting Information). Interestingly,
the introduction of defects could cause electron deficiency
and endow the graphene with an electron accepting tendency.
Therefore, the charge transfer ability between the lithium
and the carbon atoms are significantly enhanced for NDGs
(Figure 5g,h; Figure S7c,d, Supporting Information), which
would favor improved charge transfer kinetics when applied as
battery electrode.[31]
The electronic DOS for the above systems before and after
lithiation were also calculated (Figure 5; Figure S7, Supporting
Information). The PGs exhibits semiconductor characteristic
with the Dirac point at Fermi level (where the DOS is zero)
clearly visible before lithiation (Figure 5i). The Fermi level is
increased and shifted to the right after lithiation (Figure 5j),
which should be attributed to that the absorbed Li atom trans-
fers part of electrons to the conduction band of the PGs, and
the DOS near Fermi level is improved but bring little effect.
Hence, the pure graphene may not meet the requirements of
the faster electron transport kinetics. As compared to PGs,
the NDGs demonstrate prominent increment of electronic
states near the Fermi level both before and after lithiation
(Figure 5k,l). This suggests the enhanced electron mobility
within NDGs, which can greatly enhance the rate performance
of the electrode material during charge–discharge process.[31,61]
Thus, the DFT calculations evidenced the coupling effect
between N-doping and defects in the nanographitic domains of
N-GCNs for Li storage, which demonstrates improved binding
energy and charge transfer, as well as increased DOS around
Fermi level.
Adv. Funct. Mater. 2018, 1706294
Figure 5. The first-principles calculations. Models a–d) of the PGs, NGs, DGs, and NDGs, respectively. Charge density difference e,f) of an Li atom
adsorbed on the pristine graphene sheets, charge density difference g,h) of an Li atom adsorbed on the N-doped defective graphene sheets (isosur-
faces level = 0.0015e bohr3; yellow and light blue areas represent positive and negative charge differences). i,j) Corresponding DOS illustrations of the
pristine graphene sheets before and after lithiation, respectively. k,l) Corresponding DOS illustrations of the N-doped defects graphene sheets before
and after lithiation, respectively. The dashed line at 0 eV indicates the Fermi level.
www.afm-journal.dewww.advancedsciencenews.com
1706294 (8 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
Adv. Funct. Mater. 2018, 1706294
2.5. Charge Storage Mechanism of N-GCNs
Based on the above analysis, the as-prepared N-GCNs offer
obvious structural advantages in improving the Li/Na storage
capability (Figure 6): (i) The interconnected graphitic nanoshells
(diameter sizes from 200 to 600 nm) with small wall thickness of
20–50 nm can not only make certain fast and successive trans-
portation of electrons in N-GCNs, but also greatly reduce the
solid-state transport pathway for Li+/Na+ diffusion (Figure 6a)
because the electrolyte ions can be stored in both inside and out-
side surface of the ultrathin nanoshell, and the unique nanoshell
structure provides additional open hollow space in sub-micro-
scale, which can buffer volume expansion well during Li+/Na+
insertion and extraction (especially at high rates), assuring the
long cycle stability. (ii) The graphitic domains in the nanoshells
such as intermittent graphitic stripes (with short segments of
graphene layers) (Figure 6b) and crisscrossed graphene layers
(with different directions of ion access points) (Figure 6c) can
greatly increase the electronic transfer and ion diffusion coef-
ficient because of its high conductivity and short ion diffusion
pathway. (iii) The nanographitic domains (Figure 6d) can give
rise to defects such as edges and pores, which will provide suf-
ficient active sites to absorb Li+/Na+ storage, and the expanded
interlayer distance of 3.47–4.0 Å and the amorphous carbon with
nanopores (Figure 6e) can also favor the intercalation/interfacial
storage of Li+/Na+ ions, enabling high reversible capacity of the
N-GCNs, which is evidenced by in situ Raman study. (iv) The
coupled N-doping and nanopore defects in the nanographitic
domains can significantly improve the binding ability for
Li+/Na+, charge transfer ability, increase the DOS around Fermi
level before/after lithiation, thus enhance the charge transport
kinetics of the HCs with improved rate capability. Besides, the
surface oxygen groups introduced by the acidic treatment would
also contribute capacity through Faradaic reactions.
3. Conclusions
In conclusion, N-GCNs with homogeneous defective nanogra-
phitic domains are fabricated through the prechelation
between Ni2+ and chitosan with simulta-
neous N-doping and KOH-activation. The
“adsorption mechanism” of N-GCNs for Na+
storage and the “adsorption–intercalation
mechanism” of N-GCNs for Li+ storage are
verified through in situ Raman analysis.
DFT calculations are indicates the coupled
effect of heteroatom-doping and nanopore
defects down to the atomic scale in the nano-
graphitic domains of the N-GCNs. The as-
prepared N-GCNs deliver a high reversible
lithium storage capacity of 1253 mA h g1,
and outstanding rate performance which out-
performs most state-of-the-art hard carbons.
Interestingly, a high reversible capacity and
long cycle stability for sodium storage is also
obtained. This excellent long cycling perfor-
mance combined with the high capacity and
rate capability indicates that the N-GCNs are
promising anodes materials for application in rechargeable
LIBs and SIBs.
4. Experimental Section
Materials Preparation: The Artemia cyst shells with the main
component of chitosan were pretreated as described in the previous
paper.[32] Specifically 2.4 g as pretreated precursor was dispersed
into 200 mL deionized water and kept stirring for 6 h. 100 mL
0.025 m Ni(AC)2·4H2O, was then added dropwise and stirred overnight.
Subsequently, 18 mL 6 m KOH was added dropwise into the above
mixture and heated at 100 °C until the whole mixture turned into a
paste. The as-obtained materials were then put into a tube furnace and
heated to 850 °C with the ramping speed of 2 °C min1 and held there for
4 h. The as-obtained product was treated with concentrated 8 m HNO3,
1 m HCl solution and flushed with deionized water, respectively. After
being cleaned and dried, the final product was obtained and denoted
as N-GCNs. For comparison, the control samples calcined at 750 and
950 °C were also prepared following the same procedure, denoted as
N-GCNs-750 and N-GCNs-950, respectively.
Characterization: Raman spectra were recorded on a Renishaw in Via
Raman microscope with an Ar ion laser at the excitation wavelength
of 532 nm, and in situ Raman spectra (Witech Alpha 300R, 532 nm
wavelength) were tested at the excitation wavelength of 633 nm. Powder
XRD patterns between 10 and 50° (2
θ
) were collected by Rigaku D/
MAX-2500 powder diffractometer with Cu-K
α
radiation (
λ
= 0.154 nm)
operated at 40 kV, 200 mA. Hitachi-XPS was measured by a VG ESCALAB
MKIIX-ray photoelectron spectrometer using Mg-K
α
as the exciting
source (1253.6 eV). Carl Zeiss SUPRA 55 SAPPHIRE field emission
scanning electron microscope (Germany, 15 kV), Hitachi-7650 TEM
(Japan, 80 kV), and HRTEM (JEOL JEM-3000F) were used to investigate
the morphology and microstructure.
Electrochemical Measurements: The working electrodes were prepared
by mixing 80 wt% synthesized active materials, 10 wt% acetylene carbon
black, 10 wt% polyvinylidene fluoride (PVDF) binder in N-methyl-2-
pyrrolidone onto a Cu foil and then dried in a vacuum furnace at 120 °C
for 12 h. The 25 µm microporous monolayer membrane (PP, Celgard
2400) and Li/Na metal were used as separator and counter electrodes,
respectively, when assembled into a stainless-steel coin cell (2032) in an
Ar-filled glovebox (both O2 and H2O levels below 0.1 ppm). The charge
storage performance was tested in 1 m LiPF6 solution in a mixture (1:1:1,
vol%) of dimethyl carbonate, ethylene carbonate (EC), and ethylmethyl
carbonate for LIBs, and 1 m NaClO4 in a mixture (1:1, vol%) of EC and
diethyl carbonate solution for SIBs. CV curves were measured at the
Figure 6. Schematic illustration of Li (Na) storage mechanism in N-GCNs. a–e) Schematic
illustration of Li (Na) storage mechanism in N-GCNs.
www.afm-journal.dewww.advancedsciencenews.com
1706294 (9 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
Adv. Funct. Mater. 2018, 1706294
scanning rate of 0.1 mV s1. Electrochemical impedance spectroscopy
(EIS) spectra were recorded at the frequency range of 0.01 to 105 Hz.
Both of EIS and CV were test on CHI650e electrochemical workstation
(Chenhua, China). The constant current charge–discharge tests were
performed with a computer controlled cycling equipment (Land
CT2001A, China) in the potential rage of 0.01–3 V.
Supporting Information
Supporting Information is available from the Wiley Online Library.
Acknowledgements
The authors thank the financial supports from the National Natural
Science Foundation of China (51774251), Hebei Natural Science
Foundation for Distinguished Young Scholars (B2017203313 and
B2015203096), Hundred Excellent Innovative Talents Support Program
in Hebei Province (SLRC2017057), State Key Project of Research and
Development of China (2016YFA0200102), NSFC-RGC Joint Research
Scheme (51361165201), Scientific Research Foundation for the Returned
Overseas Chinese Scholars (CG2014003002), and the open funding from
State Key Laboratory of Advanced Technology for Materials Synthesis
and Processing (Wuhan University of Technology, 2017-KF-14).
Conflict of Interest
The authors declare no conflict of interest.
Keywords
hard carbon nanoshells, lithium/sodium storage, N-doping,
nanographitic domains
Received: October 30, 2017
Revised: December 1, 2017
Published online:
[1] Z. L. Yu, S. Xin, Y. You, L. Yu, Y. Lin, D. Xu, C. Qiao, Z. Huang,
N. Yang, S. Yu, J. B. Goodenough, J. Am. Chem. Soc. 2016, 138,
14915.
[2] J. B. Goodenough, K. S. Park, J. Am. Chem. Soc. 2013, 135, 1167.
[3] Z. Jian, Z. Xing, C. Bommier, Z. Li, X. Ji, Adv. Energy Mater. 2016, 6,
1501874.
[4] M. N. Tahir, B. Oschmann, D. Buchholz, X. Dou, I. Lieberwirth,
M. Panthofer, W. Tremel, R. Zentel, S. Passerini, Adv. Energy Mater.
2016, 6, 1501489.
[5] Z. Jian, Z. Xing, C. Bommier, Z. Li, X. Ji, Adv. Energy Mater. 2016, 6,
1501874.
[6] F. A. Soto, P. Yan, M. H. Engelhard, A. Marzouk, C. Wang, G. Xu,
Z. Chen, K. Amine, J. Liu, V. L. Sprenkle, Adv. Mater. 2017, 29,
1606860.
[7] L. Xiao, Y. Cao, W. A. Henderson, M. L. Sushko, Y. Shao, J. Xiao,
W. Wang, M. H. Engelhard, Z. Nie, J. Liu, Nano Energy 2016, 19,
279.
[8] Y. Cao, L. Xiao, M. L. Sushko, W. Wang, B. Schwenzer, J. Xiao,
Z. Nie, L. V. Saraf, Z. Yang, J. Liu, Nano Lett. 2012, 12, 3783.
[9] S. Qiu, L. Xiao, L. M. L. Sushko, K. S. Han, Y. Shao, M. Yan,
X. Liang, L. Mai, J. Feng, Y. Cao, X. Ai, H. Yang, J. LiuAdv. Energy
Mater. 2017, 7, 1700403.
[10] S. Wang, C. Xiao, Y. Xing, H. Xu, S. Zhang, J. Mater. Chem. A 2015,
3, 6742.
[11] A. Piotrowska, K. Kierzek, P. Rutkowski, J. Machnikowski, J. Anal.
Appl. Pyrolysis 2013, 102, 1.
[12] F. Bonino, S. Brutti, P. Reale, B. Scrosati, L. Gherghel, J. Wu,
K. Müllen, Adv. Mater. 2005, 17, 743.
[13] K. Hong, Q. Long, R. Zeng, Z. Yi, W. Zhang, D. Wang, W. Yin,
C. Wu, Q. Fan, W. Zhang, J. Mater. Chem. A 2014, 2, 12733.
[14] Y. Li, Y. Hu, H. Li, L. Chen, X. Huang, J. Mater. Chem. A 2016, 4, 96.
[15] D. A. Stevens, J. R. Dahn, J. Electrochem. Soc. 2000, 147, 1271.
[16] Y. Xu, Z. Lin, X. Zhong, B. Papandrea, Y. Huang, X. Duan, Angew.
Chem. 2015, 54, 5345.
[17] Y. Fang, Y. Lv, R. Che, H. Wu, X. Zhang, D. Gu, G. Zheng, D. Zhao,
J. Am. Chem. Soc. 2013, 135, 1524.
[18] X. Huang, H. Yu, J. Chen, Z. Lu, R. Yazami, H. H. Hng, Adv. Mater.
2014, 26, 1296.
[19] Y. Fang, Y. Lv, R. Che, H. Wu, X. Zhang, D. Gu, G. Zheng, D. Zhao,
J. Am. Chem. Soc. 2013, 135, 1524.
[20] X. Yang, L. Zhang, F. Zhang, Y. Huang, Y. Chen, ACS Nano 2014, 8,
5208.
[21] H. Zhou, S. Zhu, M. Hibino, I. Honma, M. Ichihara,
J. Cheminformatics 2004, 35, 2107.
[22] Z. Lyu, L. Yang, D. Xu, J. Zhao, H. Lai, Y. Jiang, Q. Wu, Y. Li,
X. Wang, Z. Hu, Nano Res. 2015, 8, 3535.
[23] L. Chen, Z. Wang, C. He, N. Zhao, C. Shi, E. Liu, J. Li, ACS Appl.
Mater. Interfaces 2013, 5, 9537.
[24] F. D. Han, Y. J. Bai, R. Liu, B. Yao, Y. X. Qi, N. Lun, J. X. Zhang, Adv.
Energy Mater. 2011, 1, 798.
[25] X. Liu, J. Zhang, S. Guo, N. Pinna, J. Mater. Chem. A 2016, 4, 1423.
[26] D. Xie, X. Xia, Y. Zhong, Y. Wang, D. Wang, X. Wang, J. Tu, Adv.
Energy Mater. 2017, 7, 1601804.
[27] A. L. M. Reddy, A. Srivastava, S. R. Gowda, H. Gullapalli, M. Dubey,
P. M. Ajayan, ACS Nano 2010, 4, 6337.
[28] J. Dong, Y. Xue, C. Zhang, Q. Weng, P. Dai, Y. Yang, M. Zhou, C. Li,
Q. Cui, X. Kang, Adv. Mater. 2017, 29, 1603692.
[29] Z. Pei, H. Li, Y. Huang, Q. Xue, Y. Huang, M. Zhu, Z. Wang, C. Zhi,
Energy Environ. Sci. 2017, 10, 742.
[30] J. Xu, M. Wang, N. P. Wickramaratne, M. Jaroniec, S. Dou, L. Dai,
Adv. Mater. 2015, 27, 2042.
[31] Y. Yang, D. M. Tang, C. Zhang, Y. Zhang, Q. Liang, S. Chen,
Q. Weng, M. Zhou, Y. Xue, J. Liu, Energy Environ. Sci. 2017, 10,
979.
[32] Y. Zhao, S. Huang, M. Xia, S. Rehman, S. Mu, Z. Kou, Z. Zhang,
Z. Chen, F. Gao, Y. Hou, Nano Energy 2016, 28, 346.
[33] Z. Zhang, S. Mu, B. Zhang, L. Tao, S. Huang, Y. Huang, F. Gao,
Y. Zhao, J. Mater. Chem. A 2016, 4, 2137.
[34] J. Wang, S. Kaskel, J. Mater. Chem. 2012, 22, 23710.
[35] A. Ōya, H. Marsh, J. Mater. Sci. 1982, 17, 309.
[36] J. Gillot, B. Lux, P. Cournault, F. du Chaffaut, Ber Deut. Keram. Ges.
1968, 45, 224.
[37] A. Oberlin, M. Endo, T. Koyama, J. Cryst. Growth 1976, 32, 335.
[38] R. P. Vidano, D. B. Fischbach, L. J. Willis, T. M. Loehr, Solid State
Commun. 1981, 39, 341.
[39] A. C. Ferrari, J. C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri,
F. Mauri, S. Piscanec, D. Jiang, K. S. Novoselov, S. Roth, Phys. Rev.
Lett. 2006, 97, 187401.
[40] Y. Hao, Y. Wang, L. Wang, Z. Ni, Z. Wang, R. Wang, C. K. Koo,
Z. Shen, J. T. Thong, Small 2010, 6, 195.
[41] W. Yang, H. Kai, Y. Zhu, F. Han, Y. Xu, I. Matsuda, Y. Ishii,
J. Cumings, C. Wang, Nat. Commun. 2014, 5, 4033.
[42] U. Sim, J. Moon, J. An, J. H. Kang, S. E. Jerng, J. Moon, S. P. Cho,
B. H. Hong, K. T. Nam, Energy Environ. Sci. 2015, 8, 1626.
[43] C. Ma, X. Shao, D. Cao, J. Mater. Chem. 2012, 22, 8911.
[44] M. Lu, W. Yu, J. Shi, W. Liu, S. Chen, X. Wang, H. Wang, Electrochim.
Acta 2017, 251, 369.
www.afm-journal.dewww.advancedsciencenews.com
1706294 (10 of 10) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
Adv. Funct. Mater. 2018, 1706294
[45] S. W. Lee, N. Yabuuchi, B. M. Gallant, S. Chen, B. S. Kim,
P. T. Hammond, S.-H. Yang, Nat. Nanotechnol. 2010, 5, 531.
[46] Y. Shao, J. Xiao, W. Wang, M. Engelhard, X. Chen, Z. Nie, M. Gu,
V. S. Laxmikant, E. Gregory, J. Zhang, L. Jun, Nano Lett. 2013, 13,
3909.
[47] J. Niu, R. Shao, J. Liang, M. Dou, Z. Li, Y. Huang, F. Wang, Nano
Energy 2017, 36, 322.
[48] J. Zhang, H. Zhu, P. Wu, C. Ge, D. Sun, L. Xu, Y. Tang, Y. Zhou,
Nanoscale 2015, 7, 18211.
[49] W. Tang, B. M. Goh, M. Y. Hu, C. Wan, B. Tian, X. Deng, C. Peng,
M. Lin, J. Z. Hu, K. P. Loh, J. Phys. Chem. C 2016, 120, 2600.
[50] Y. Li, Y. S. Hu, M. M. Titirici, L. Chen, X. Huang, Adv. Energy Mater.
2016, 6, 1600659.
[51] H. Song, G. Yang, C. Wang, ACS Appl. Mater. Interfaces 2014, 6,
21661.
[52] Y. Zheng, Y. Wang, Y. Lu, Y. S. Hu, J. Li, Nano Energy 2017, 39,
489.
[53] X. Liu, D. Chao, D. Su, S. Liu, L. Chen, C. Chi, J. Lin, Z. Shen,
J. Zhao, L. Mai, Y. Li, Nano Energy 2017, 37, 108.
[54] L. J. Hardwick, H. Buqa, P. Novák, Solid State Ionics 2006, 177,
2801.
[55] L. J. Hardwick, Ph.D. Thesis, Eidgenössische Technische Hochschule
ETH Zürich, 2007.
[56] L. J. Hardwick, M. Hahn, P. Ruch, M. Holzapfel, W. Scheifele,
H. Buqa, F. Krumeich, P. Novák, R. Kötz, Electrochem. Acta 2006,
52, 680.
[57] L. J. Hardwick, P. W. Ruch, M. Hahn, W. Scheifele, R. Kötz,
P. Novák, J. Phys. Chem. Solids 2008, 69, 1232.
[58] C. Sole, N. E. Drewett, L. J. Hardwick, Faraday Discuss. 2014, 172,
223.
[59] G. Mpourmpakis, G. Froudakis, J. Chem. Phys. 2006, 125, 56.
[60] F. Valencia, A. H. Romero, F. Ancilotto, P. L. Silvestrelli, J. Phys.
Chem. B 2006, 110, 14832.
[61] K. Shiota, T. Kawai, Jpn. J. Appl. Phys. 2017, 56, 06GE11.
... Carbonaceous materials with low intercalation potential, such as graphite (Gr), carbon nanotubes, and hard carbon (HC), are favorable LIC anode materials as they can provide broad operational voltage range and corresponding high energy density [2,4,[10][11][12][13]. Gr has been extensively used as the anode material in both LIBs and LICs owing to its high specific capacity (320-370 mAh g − 1 ), low lithiation/delithiation potential, and cost-effectiveness [10,11,[13][14][15][16][17][18][19]. However, its performance deteriorates in fast charging conditions, which restricts its largescale applications in electric vehicles and utility-scale energy storage. ...
... Fig. 3(d) shows that a higher Gr content in the anode corresponded to a higher ICE. HC exhibited the lowest ICE of approximately 52.7 %, similar to the value reported by Huang et al. (53.4 %) [17], followed by HC75Gr25, HC75Gr25DP, HC25Gr75, and Gr100, with ICEs of 62.3 %, 60.9 %, 76.3 %, and 82.0 %, respectively. The low ICE of HC was attributable to the excessive formation of SEI layers and Li passivation caused by its nanoporous structure and high specific surface area [10,20,21,25]. ...
Article
Full-text available
Li-ion capacitors (LICs), characterized by high energy and power densities, are widely regarded as promising electrochemical energy-storage devices. However, the application of these devices is limited by the poor stability of anode materials and inferior electrochemical performance owing to asymmetrical kinetics between the anode and cathode. To address these concerns, this study aimed at developing a highly stable composite anode material by tuning the mass ratio between graphite (Gr) and hard carbon (HC). When subjected to thrice-repeated deep prelithiation, the composite anode exhibited enhanced cycling performance without significant deterioration in the capacity. The incorporation of HC in the anode effectively improved the cyclability of Gr, and the thrice-repeated deep prelithiation strategy facilitated the increase in LIC capacity. The LIC using the composite anode exhibited an energy density of 48.5 Wh kg − 1 at a power density of 2170 W kg − 1 , surpassing those of Gr (33.8 Wh kg − 1 at 2030 W kg − 1) and HC (35.7 Wh kg − 1 at 2230 W kg − 1) LICs. The enhanced LIC retained 80.5 % of its initial energy density after 10,000 cycles, a significantly higher value than those of LICs based on Gr (26.1 %) and HC (73.3 %). In summary, deep prelithiation can enhance the overall performance of LICs to achieve improved cyclability and capacity compared with its semi-shallow prelithiated counterpart. This study provides a cost-effective solution for the design of LIC anode materials with high rate and long cycling performances.
... Other advantages of LTO include its high cycle life, thermal stability, and capacity. Even though LTO has (9) nLi + + C 6 + ne − ⇌ Li n C 6 challenges like reduced cell voltage and high cost due to the presence of titanium, it has still been commercialized [52][53][54]. ...
... Non-graphitised carbon is another intercalation material that has been studied due to its greater layer spacing [103]. However, it suffers from issues such as voltage hysteresis and high initial irreversible capacity [53,54]. Other carbon-based materials such as graphene [104,105], nanofibers [106][107][108], and nanotubes [109][110][111] are also being explored as potential materials for LIB anodes. ...
Article
Full-text available
This review paper presents a comprehensive analysis of the electrode materials used for Li-ion batteries. Key electrode materials for Li-ion batteries have been explored and the associated challenges and advancements have been discussed. Through an extensive literature review, the current state of research and future developments related to Li-ion battery electrodes were identified. The study also covers a wide range of subtopics, including the theoretical aspects of the basic functioning of lithium-ion batteries and the crystal structures of different electrode materials. Additionally, emerging trends and future directions in the development of high-performance commercial battery electrodes have been revealed, providing insights into promising avenues for further research. By synthesizing existing knowledge and analyzing the latest research, this review aims to provide a valuable resource for researchers, practitioners, and stakeholders interested in developing state-of-the-art high-performance Li-ion batteries. The findings and perspectives presented in this paper contribute to a deeper understanding of electrode materials for Li-ion batteries and their advantages and disadvantages, ultimately fostering advancements and innovations in commercial lithium-ion battery (LiB) electrode technology.
... It can be seen that D-band peak gradually become broaden and decreased from 2.20 to 0 V, but the position of this peak keeps unchanged, which can be attributed to the suppression of the breathing vibration of carbon rings at the defects and edges caused by the adsorption of Na + at defective sites and surface of pores. [67] As for the G-band peak, the position remains unchanged from 2.20 to 0.10 V and gradually become red shift from 0.10 to 0 V in the discharging process and then gradually shifts back from 0 to 0.1 V in the discharging process, corresponding to the weakening and lengthening of the CÀ C bond due to the transfer of electrons to π* antibonding-band with the insertion of Na + into the carbon layers. [15] Meanwhile, the width of G-peak increased from 0.10 to 0 V, demonstrating the diffusion process of Na + to pore structures and interact with sp2 carbon atoms and resulting in the change of E 2g stretching mode for the G-peak. ...
Article
Full-text available
Given the merits of abundant resource, low cost and high electrochemical activity, hard carbons have been regarded as one of the most commercializable anode material for sodium‐ion batteries (SIBs). However, poor rate capability is one of the main obstacles that severely hinder its further development. In addition, the relationships between preparation method, material structure and electrochemical performance have not been clearly elaborated. Herein, a simple but effective strategy is proposed to accurately construct the multiple structural features in hard carbon via adjusting the components of precursors. Through detailed physical characterization of the hard carbons derived from different regulation steps, and further combined with in‐situ Raman and galvanostatic intermittent titration technique (GITT) analysis, the network of multiple relationships between preparation method, microstructure, sodium storage behavior and electrochemical performance have been successfully established. Simultaneously, exceptional rate capability about 108.8 mAh g⁻¹ at 8 A g⁻¹ have been achieved from RHC sample with high reversible capacity and desirable initial Coulombic efficiency (ICE). Additionally, the practical applications can be extended to cylindrical battery with excellent cycle behaviors. Such facile approach can provide guidance for large‐scale production of high‐performance hard carbons and provides the possibility of building practical SIBs with high energy density and durability.
Article
Full-text available
Hard carbon materials have shown promising potential for sodium‐ion storage due to accommodating larger sodium ions. However, as for lithium‐ion storage, the challenge lies in tuning the high lithiation plateau capacities, which impacts the overall energy density. Here, hard carbon microspheres (HCM) are prepared by tailoring the cross‐linked polysaccharide, establishing a comprehensive methodology to obtain high‐performance lithium‐ion batteries (LIBs) with long plateau capacities. The “adsorption–intercalation mechanism” for lithium storage is revealed combining in situ Raman characterization and ex situ nuclear magnetic resonance spectroscopy. The optimized HCM possesses reduced defect content, enriched graphitic microcrystalline, and low specific surface area, which is beneficial for fast lithium storage. Therefore, HCM demonstrates a high reversible capacity of 537 mAh g⁻¹ with a significant low‐voltage plateau capacity ratio of 55%, high initial Coulombic efficiency, and outstanding rate performance (152 mAh g⁻¹ at 10 A g⁻¹). Moreover, the full cell (HCM||LiCoO2) delivers outstanding fast‐charging capability (4 min charge to 80% at 10 C) and impressive energy density of 393 Wh kg⁻¹. Additionally, 80% reversible capacity can be delivered under −40 °C with competitive cycling stability. This work provides in‐depth insights into the rational design of hard carbon structures with extended low‐voltage plateau capacity for high energy LIBs.
Article
Hollow carbon nanosphere anode materials were manipulated by a pyrolysis temperature-mediated microstructure strategy to simultaneously achieve high ICE and superior rate/cycling stability.
Article
Li-ion capacitors (LICs) are promising to simultaneously achieve battery-level energy density and supercapacitor-level power density, but the slow kinetics of diffusion-controlled battery anodes lead to unmatched two-electrode kinetics at the...
Article
Given the merits of abundant resource, low cost and high electrochemical activity, hard carbons have been regarded as one of the most commercializable anode material for sodium‐ion batteries (SIBs). However, poor rate capability is one of the main obstacles that severely hinder its further development. In addition, the relationships between preparation method, material structure and electrochemical performance have not been clearly elaborated. Herein, a simple but effective strategy is proposed to accurately construct the multiple structural features in hard carbon via adjusting the components of precursors. Through detailed physical characterization of the hard carbons derived from different regulation steps, and further combined with in‐situ Raman and galvanostatic intermittent titration technique (GITT) analysis, the network of multiple relationships between preparation method, microstructure, sodium storage behavior and electrochemical performance have been successfully established. Simultaneously, exceptional rate capability about 108.8 mAh g‐1 at 8 A g‐1 have been achieved from RHC sample with high reversible capacity and desirable initial Coulombic efficiency (ICE). Additionally, the practical applications can be extended to cylindrical battery with excellent cycle behaviors. Such facile approach can provide guidance for large‐scale production of high‐performance hard carbons and provides the possibility of building practical SIBs with high energy density and durability.
Article
Full-text available
The main challenge associated with sodium-ion battery (SIB) anodes is a search for novel candidate materials with high capacity and excellent rate capability. The most commonly used and effective route for graphene-based anode design is the introduction of in-plane "hole" defects via nitrogen-doping; this creates a spacious reservoir for storing more energy. Inspired by mountains in nature, herein, we propose another way-the introduction of blistering in graphene instead of making "holes"; this facilitates adsorbing/inserting more Na⁺ ions. In order to properly answer the key question: ""protrusions" or "holes" in graphene, which is better for sodium ion storage?", two types of anode materials with a similar doping level were designed: a phosphorus-doped graphene (GP, with protrusions) and a nitrogen-doped graphene (GN, with holes). As compared with GN, the GP anode perfectly satisfies all the desired criteria: it reveals an ultrahigh capacity (374 mA h g⁻¹ after 120 cycles at 25 mA g⁻¹) comparable to the best graphite anodes in a standard Li-ion battery (∼372 mA h g⁻¹), and exhibits an excellent rate capability (210 mA h g⁻¹ at 500 mA g⁻¹). In situ transmission electron microscopy (TEM) experiments and density functional theory (DFT) calculations were utilized to uncover the origin of the enhanced electrochemical activity of "protrusions" compared to "holes" in SIBs, down to the atomic scale. The introduction of protrusions through P-doping into graphene is envisaged to be a novel effective way to enhance the capacity and rate performance of SIBs.
Article
Nitrogen, oxygen and sulfur tridoped porous carbons have been successfully synthesized from natural biomass algae-Carrageen by using a simultaneous carbonization and activation procedure. The doped carbons with sponge-like interconnected architecture, partially ordered graphitic structure, and abundant heteroatom doping perform outstanding features for electrochemical energy storage. When tested as lithium-ion battery anodes, a high reversible capacity of 839 mAh g⁻¹ can be obtained at the current density of 0.1 A g⁻¹ after 100 cycles, while a high capacity of 228 mAh g⁻¹ can be maintained at 10 A g⁻¹. Tested against sodium, a high specific capacity of 227 can be delivered at 0.1 A g⁻¹ after 100 cycles, while a high capacity of 109 mAh g⁻¹ can be achieved at 10 A g⁻¹. These results turn out that the doped carbons would be potential anode materials for lithium- and sodium-ion batteries, which can be achieved by a one-step and large-scale synthesis route. Our observation indicates that heteroatom doping (especially sulfur) can significantly promote ion storage and reduce irreversible ion trapping to some extent. This work gives a general route for designing carbon nanostructures with heteroatom doping for efficient energy storage.
Article
Hard carbon anode materials for sodium-ion batteries (SIB) have usually been tested in half-cells by cycling between 0-2 V, and is believed to exhibit low rate capability. However, we find that the specific capacity, the rate performance, and the cycling performance may all be severely underestimated with the traditional half-cell cycling evaluation method, due to premature truncation of part II of the capacity (part I is “sloping”, part II is “plateauing”, while part III is Na metal deposition). Here we introduce a sodium-matched SIB full-cell architecture, with newly developed hard carbon derived from macadamia shell (MHC) as anode and (NCNFM) as the cathode material, with anode/cathode areal capacity ratio of 1.02-1.04. Our carefully balanced full-cells exhibit a cell-level theoretical specific energy of 215 Wh kg⁻¹ at C/10 and 186 Wh kg⁻¹ at 1C based on cathode-active and anode-active material weights, and an outstanding capacity retention of 70% after 1300 cycles ( ). Traditional half-cell test (THT) of MHC using superabundant Na metal counter electrode shows only 51.7 mAh g⁻¹ capacity at 1C, and appears to die in no more than 100 hours due to low open-circuit voltage slope and large polarization. A revised half-cell test (RHT) which shows much better agreements with full-cell test results, delivers a specific capacity of 314 mAh g⁻¹, with an initial Coulombic efficiency of , which is comparable to that of graphite in Lithium-ion batteries.
Article
We investigated the fundamental properties of the interaction between a Li atom and a graphene surface with various defect structures by first-principles electronic state calculations to improve the capacity and charge rate of a graphitic anode for Li ion battery applications. The adsorption energy tends to decrease as the number of deficit carbon atoms at a neighboring defect increases even for adsorption at a hexagonal ring (HR) away from defects, although the interaction between a Li adatom and an HR is similar independent of the defect structure. The reason for the change in adsorption energy is the electronic charge transfer from the Li 2s-like state to the defect-induced state near the Fermi level. We also found that a Li atom diffuses through a V6 defect without a diffusion barrier practically.
Article
Hard carbon is one of the most promising anode materials for sodium-ion batteries, but the low Coulombic efficiency is still a key barrier. In this paper, a series of nanostructured hard carbon materials with controlled architectures is synthesized. Using a combination of in situ X-ray diffraction mapping, ex situ nuclear magnetic resonance (NMR), electron paramagnetic resonance, electrochemical techniques, and simulations, an “adsorption–intercalation” mechanism is established for Na ion storage. During the initial stages of Na insertion, Na ions adsorb on the defect sites of hard carbon with a wide adsorption energy distribution, producing a sloping voltage profile. In the second stage, Na ions intercalate into graphitic layers with suitable spacing to form NaC x compounds similar to the Li ion intercalation process in graphite, producing a flat low voltage plateau. The cation intercalation with a flat voltage plateau should be enhanced and the sloping region should be avoided. Guided by this knowledge, nonporous hard carbon material has been developed which has achieved high reversible capacity and Coulombic efficiency to fulfill practical application.
Article
Porous carbons with high specific surface area and high defect density have been prepared through direct carbonization of cattle bones without any additional activators and templates. Benefiting from self-activation induced by hydroxyapatites within the cattle bones, the high-defect porous carbons obtained at 1100 °C (PC-1100) possess the high specific surface area (2096 m² g⁻¹), largest mesopore volume (1.829 cm³ g⁻¹), a narrow mesopore size distribution centered at approximately 4.0 nm and good electrical conductivity (5141 S m⁻¹). Due to the synergistic effect of the defects and pores, PC-1100 as the anode for Li-ion battery exhibits a high reversible capacity of 1488 mAh g⁻¹ after 250 cycles at 1 A g⁻¹ and 661 mAh g⁻¹ after 1500 cycles at 10 A g⁻¹. Even at 30 A g⁻¹, PC-1100 can still deliver a high reversible capacity of 281 mAh g⁻¹, showing superior lithium storage capability. Moreover, the symmetric supercapacitor based on the PC-1100 in neat EMIM-BF4 electrolyte delivers a high energy density of 109.9 Wh kg⁻¹ at a power density of 4.4 kW kg⁻¹, and maintains an energy density of 65.0 Wh kg⁻¹ even at an ultrahigh power density of 81.5 kW kg⁻¹, as well as a superior cycling performance (96.4% of the capacitance retention after 5000 cycles).
Article
Solid-electrolyte interphase (SEI) films with controllable properties are highly desirable for improving battery performance. In this paper, a combined experimental and theoretical approach is used to study SEI films formed on hard carbon in Li- and Na-ion batteries. It is shown that a stable SEI layer can be designed by precycling an electrode in a desired Li- or Na-based electrolyte, and that ionic transport can be kinetically controlled. Selective Li- and Na-based SEI membranes are produced using Li- or Na-based electrolytes, respectively. The Na-based SEI allows easy transport of Li ions, while the Li-based SEI shuts off Na-ion transport. Na-ion storage can be manipulated by tuning the SEI layer with film-forming electrolyte additives, or by preforming an SEI layer on the electrode surface. The Na specific capacity can be controlled to < 25 mAh g−1; ≈ 1/10 of the normal capacity (250 mAh g−1). Unusual selective/preferential transport of Li ions is demonstrated by preforming an SEI layer on the electrode surface and corroborated with a mixed electrolyte. This work may provide new guidance for preparing good ionselective conductors using electrochemical approaches.
Article
Facile yet rational design of an efficient reversible oxygen electrocatalyst is critical for many renewable energy conversion and storage technologies. Here we report a simple and general synthetic protocol for fabricating a hierarchically porous and heteroatom doped carbon catalyst, which exhibited outstanding oxygen reduction/evolution activities (with a metric potential difference of 0.72 V in 1 M KOH, the best value for metal-free catalysts reported to date) with good stability in different electrolytes. The excellent performances of the catalyst were primarily endowed by our synthetic protocol, which integrates good conductivity, abundant accessible dopant species and suitable porous architectures within an in situ pyrolysis reaction. As a result, the performances of rechargeable Zn-air batteries based on the optimized catalyst substantially outperform those afforded by a benchmark Pt/C catalyzer. Our work is expected to open up new avenues for developing other efficient catalysts in a facile and viable way.
Article
A novel carbon structure, highly branched homogeneous-N-doped graphitic (BNG) tubular foam, is designed via a novel N, N-dimethylformamide (DMF)-mediated chemical vapor deposition method. More structural defects are found at the branched portions as compared with the flat tube domains providing abundant active sites and spacious reservoirs for Li(+) storage. An individual BNG branch nanobattery is constructed and tested using in situ transmission electron microscopy and the lithiation process is directly visualized in real time.