10 1016@j Cej 2019 123576

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Journal Pre-proofs

Extended plateau capacity of phosphorus-doped hard carbon used as an anode


in Na- and K-ion batteries

Stevanus Alvin, Christian Chandra, Jaehoon Kim

PII: S1385-8947(19)32991-2
DOI: https://doi.org/10.1016/j.cej.2019.123576
Reference: CEJ 123576

To appear in: Chemical Engineering Journal

Received Date: 1 October 2019


Revised Date: 18 November 2019
Accepted Date: 20 November 2019

Please cite this article as: S. Alvin, C. Chandra, J. Kim, Extended plateau capacity of phosphorus-doped hard carbon
used as an anode in Na- and K-ion batteries, Chemical Engineering Journal (2019), doi: https://doi.org/10.1016/
j.cej.2019.123576

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will
undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing
this version to give early visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier B.V.


Extended plateau capacity of phosphorus-doped hard carbon

used as an anode in Na- and K-ion batteries

Stevanus Alvina, Christian Chandrab, Jaehoon Kima,b,c

a SKKU Advanced Institute of Nano Technology (SAINT), Sungkyunkwan University, 2066, Seobu-

ro, Jangan-gu, Suwon, Gyeonggi-do 16419, Republic of Korea


b School of Mechanical Engineering, Sungkyunkwan University, 2066, Seobu-ro, Jangan-gu, Suwon,

Gyeonggi-do 16419, Republic of Korea


c School of Chemical Engineering, Sungkyunkwan University, 2066, Seobu-ro, Jangan-gu, Suwon,

Gyeonggi-do 16419, Republic of Korea

*To whom correspondence should be addressed: Tel: +82-31-299-4843, Fax: +82-31-290-5889,

e-mail: jaehoonkim@skku.edu (Prof. Jaehoon Kim)

0
ABSTRACT

Hard carbon is one of most promising anode materials used in sodium-ion batteries (SIBs)

because of its high low-voltage plateau capacity. Heteroatom doping into the carbon structure is

considered an effective method to enhance the Na+-ion uptake. However, heteroatom doping is not

utilized to increase the low-voltage plateau capacity because the carbonization temperatures are

limited to low values (600–1100 °C). In addition, the formation of excess defect sites, which is caused

by heteroatom doping leads to lower initial Coulombic efficiency (ICE). Herein, to increase the low-

voltage plateau capacity and to maintain high ICE, combination of high-temperature carbonization

and low-level heteroatom doping is investigated. The P-doped hard carbon synthesized at 1300 °C

with doping level of 1.1 at.% exhibits enhanced reversible capacity of 328 mAh g−1 at 50 mA g−1, and

high ICE of 72% in SIBs. After the P-doping, the low-voltage plateau capacity increases, while the

high-voltage sloping capacity does not change significantly. This is attributed to the enlargement of

the interlayer spacing between the graphitic layers, which enhances Na+-ion intercalation. The P-

doped hard carbon delivers a high reversible capacity of 302 mAh g−1 in potassium-ion batteries

(KIBs); this value is 23% larger than that of undoped hard carbon.

Keywords: hard carbon, doping, sodium-ion battery, potassium-ion battery, phosphorus, interlayer

space, intercalation

1
1. INTRODUCTION

Sodium-ion batteries (SIBs) have recently received considerable attention in large-scale energy

storage fields, such as in energy storage systems, because of the low-cost, global distribution and

abundance of sodium in sharp contrast to the Li counterpart [1, 2]. Although graphite is commercially

used as an anode material in lithium-ion batteries, the Na+-ion uptake in graphite is extremely small

because of the unfavorable formation of sodium-graphite intercalation compounds [3-5]; this makes it

difficult to fabricate high-energy-density electrode materials using graphite for use as an anode in

SIBs [6-8]. Therefore, numerous studies have been carried out to develop potential anode materials for

LIBs and SIBs, which include carbon-based, alloying-based, and metal oxide-based materials [1, 2, 9-

13]. Among these, hard carbon, which consists of irregularly-oriented, few graphitic layers stacked in

an approximately parallel fashion and amorphous carbon domains, has been attracting attention as one

of the most promising anode materials for SIBs because of its low-cost, ease of synthesis, its high

stability during Na+-ion insertion and deinsertion, and the possibility of using renewable resources

(e.g., wood, macro-algae, fruit shell) for producing it [14-27]. In addition, the large interlayer spacing

between the basal plane of the graphitic layers (0.36–0.40 nm) and defects, and micropores present in

hard carbon can provide favorable sites for Na+-ion uptake, resulting in high reversible capacities of

300–350 mAh g−1. More importantly, the low-voltage plateau capacity below 0.1 V of hard carbon,

which is not found in other types of carbon (e.g., soft carbon [28, 29]) renders it highly promising for

developing high-energy-density electrodes for full cells. Besides the advantages of low-voltage

plateau capacity to energy density, the Na metal plating could be a great challenge at operation of low

temperature and/or elevated charge rates [30]; however, such problem can be overcome by decreasing

the surface area of hard carbon.

Despite significant progress in the development of hard carbon as a potential anode for SIBs, the

complex nature of hard carbon, including a wide range of interlayer spacing, graphitization degree,

number of stacked layers, defects, micropores, and residual chemical functionalities, makes it difficult

2
to understand the property–performance correlation. For example, the origin of the low-voltage

plateau capacity is still a topic of debate; previously, filling of the micropore with Na+ ion was

considered to be responsible for the low-voltage plateau capacity [31-35]. Recently, however,

intercalation of Na+-ion into the graphitic layers is regarded as the origin for the low-voltage plateau

capacity [15, 16, 19, 22, 36-41]. Depending on the Na+-ion storage mechanism, strategies to extend

the low-voltage plateau region should be developed. Based on ex situ X-ray diffraction (XRD), 23Na

solid-state nuclear magnetic resonance (NMR) spectroscopy, and space-filling model analyses, the

low-voltage plateau capacity of lignin-derived hard carbon is speculated to originate from the Na+ ion

intercalation into the graphitic layers [25].

The physicochemical properties of hard carbon depend on many factors including the type of

precursor [42, 43] and carbonization conditions (e.g., temperature [15, 16, 44], sweep gas flow rate

[45], heating rate [40]). Other than controlling the process parameters and sources of carbon, doping

of heteroatoms (e.g., phosphorous, nitrogen, sulfur, boron, fluorine) into the carbon structure has been

considered a promising method to control the physicochemical and textural properties of hard carbon

[17, 46-65]. Overall, the heteroatom doping changes the local structure of hard carbon, leading to

enlargement of the interlayer spacing, increase in the micropore volume, creation of defect sites, and

increase in the electronic conductivity. These beneficial features due to heteroatom doping can lead to

more active sites for Na+-ion uptake, reduced diffusion barrier, and enhanced charge transfer in the

composite electrode. As summarized in Table 1, numerous heteroatom doping approaches have been

explored, resulting in increased discharge–charge capacities and high-rate performance. For example,

Wang et al. reported that N-doped hard carbon prepared using polypyrrole as a precursor delivered a

specific capacity of 134.2 mAh g−1 at a high specific current of 200 mA g−1 after 200 cycles [46]. The

excellent high-rate performance was attributed to N-doping, which enhanced the charge transfer

kinetics in the hard carbon layers. Qie et al. prepared S-doped hard carbon and observed a reversible

capacity of 300 mAh g−1 at a high specific current of 0.5 A g−1. S-doping enlarged the interlayer

3
spacing of the graphitic layer, which can facilitate sodium ions insertion between the graphene sheets

[48]. Phosphorus has also been regarded as a promising dopant because it can improve the

electrochemical performance by increasing the defect sites and enlarging the interlayer spacing of the

graphitic layer [66]. Hou et al. reported large-area carbon nanosheets doped with phosphorus

containing highly dilated interlayers with a spacing of 0.42 nm [54]. The P-doped hard carbon

delivered a high reversible capacity of 328 mAh g−1 at 0.1 A g−1.

Although plausible reasons for the improved electrochemical performance of heteroatom-doped

hard carbon have been discussed well by others, there are still important issues to be carefully

addressed for developing practical SIBs. First, most of the previous heteroatom doping has been

performed at relatively lower temperatures of 600–1100 °C [46-51, 53-56, 58-65, 67, 68]; at this low

carbonization temperature range, the development of the graphitic layer was not activated, resulting in

small low-voltage plateau capacities. A well-developed hexagonal structure in the graphitic layer can

facilitate Na+-ion diffusion, leading to an extended plateau capacity. The increment in the high-

voltage sloping capacity without improving the low-voltage plateau capacity would not be beneficial

in developing a practical, high-energy-density anode for SIBs. Second, increasing the doping level

sometimes causes an increase in the defect sites in the graphitic layers, leading to the lowering of the

initial Coulombic efficiency (ICE) [46-51, 53-56, 58-65, 67, 68]; as listed in Table 1, the ICE values

of the doped hard carbon are in the range of 18–60%. The large loss of the initial discharge capacity is

typically caused by the undesirable electrolyte decomposition and irreversible Na+-ion trapping on the

surface of the composite electrode. Therefore, it is imperative to develop heteroatom-doped hard

carbon that has an extended voltage plateau and high ICE.

To increase the low-voltage plateau capacity while simultaneously maintaining high ICE, new

doping strategies should be developed. Different from the previous doping approaches, the doping

strategy in this study involves the use a high carbonization temperature and suppression of excessive

activation of carbon by maintaining a low heteroatom-doping level. Herein, P-doped hard carbon was

4
synthesized using lignin as the carbon source, at a high carbonization temperature of 1300 °C. When

tested as an anode in SIBs, the hard carbon with 1.1 at.% P-doping delivered a total reversible

capacity of 328 mAh g−1 with a high low-voltage plateau capacity of 223 mAh g−1 and maintained

high ICE of 72%. In addition, potassium-ion batteries (KIBs) have been investigated with several

potential anode materials [69, 70]. The P-doped hard carbon exhibited excellent electrochemical

performance as an anode in KIB; a high reversible capacity of 302 mAh g−1 was achieved, which is 23%

larger than that of undoped hard carbon. As listed in Table 1, the low-voltage plateau capacity of the

P-doped hard carbon synthesized in this study is higher than those of most of the previous heteroatom-

doped hard carbon materials. For comparison, low-temperature P-doping at 900 °C was also

performed, as has been done in most of the previous studies, in order to gain insights into the

correlation between the physicochemical properties of P-doped hard carbon and the electrochemical

performance. In addition, to investigate the possibility of using the high-temperature doping strategy

for doping other heteroatoms, N-doped and S-doped hard carbons were also prepared at 1300 °C, and

their electrochemical performance was evaluated.

5
Table 1. Comparison of the surface area, ICE, total capacity, and plateau capacity of heteroatom-doped carbons in SIB

Plateau capacity Sloping capacity


Pyrolysis Surface Doping Total reversible
(mAh g−1) (mAh g−1) ICE
Doping Materials Temperature Area level capacity Refs.
(% of the total (% of the total (%)
(°C) (m2 g−1) (%) (mAh g−1)
capacity) capacity)
N-doped carbon 271 at 50 mA g−1 Li et al., Chem
800 122.6 2.5 at.% 271 at 50 mA g−1 None 39
nanotubes (100%) Comm, 2015 [71]
N-doped porous 284 at 75 mA g−1 Chen et al., Chem,
700 438 6.0 wt% 284 at 75 mA g−1 None 47
carbon tubule (100%) 2017 [72]

N-doped carbon 350 at 50 mA g−1 Wang et al.,


800 1477 N.A.a 350 at 50 mA g−1 None 35
sheets (100%) ChemSusChem, 2013
[46]
N-doped
catalytic 227 at 50 mA g−1 Wang et al., Sci.
600 1008 1.32 at.% 227 at 50 mA g−1 None 40
graphitized hard (100%) Rep., 2018 [61]
carbon
N-rich Liu et al., ACS Appl.
70 at 30 mA g−1 268 at 30 mA g−1
N-doping mesoporous 700 113 7.78 at.% 338 at 30 mA g−1 54 Mater. Inter., 2015
(21%) (79%)
carbon [67]
Free-standing
564 at 100 mA g−1 Wang et al., Adv.
N-doped carbon 650 564.4 7.15 wt% 564 at 100 mA g−1 None 36
(100%) Energy Mater., 2015
nanofiber films
[52]
N-doping and
defective
Huang et al., Adv.
nanographitic 1236 at 100 mA g−1
950 1490 1.3 at.% 1236 at 100 mA g−1 None 53 Funct. Mater., 2018
domain-coupled (100%)
[60]
hard carbon
nanoshells
N-doped carbon 140 at 56 mA g−1 200 at 56 mA g−1 Yang et al., Adv.
650 N.A. 9.89 at.% 340 at 56 mA g−1 64
sheets (41%) (59%) Mater., 2016 [51]
S-doped hard 190 at 50 mA g−1 138 at 50 mA g−1 Li et al., Adv. Energy
1100 5.2 0.1 wt% 328 at 50 mA g−1 80
carbon (58%) (42%) Mater., 2017 [66]
S-doping Tang et al.,
S-doped carbon 443 at 50 mA g−1
600 1006.7 11.5 at.% 443 at 50 mA g−1 None 47 Electrochim. Acta
spheres (100%)
[55]
6
S-doped
322 at 50 mA g−1 Zou et al., Small,
graphitic carbon 800 898.8 2.12 at.% 322 at 100 mA g−1 None 60
(38%) 2017 [53]
nanosheets
S-doped Li et al., Energy
516 at 20 mA g−1
disordered 500 117.3 26.9 wt% 516 at 20 mA g−1 None 63 Environ. Sci., 2015
(100%)
carbon [47]
482 at 20 mA g−1 Qie et al., Adv. Sci.,
S-doped carbon 700 39.8 15.17 wt% 482 at 100 mA g−1 None 74
(100%) 2015 [48]
S-doped hard 320 at 100 mA g−1 Hong et al., Adv.
600 276.9 6.3 at.% 320 at 100 mA g−1 None 56
carbon (100%) Mater., 2018 [64]
S-doped
259 at 100 mA g−1 Shi et al., Carbon,
mesoporous 950 573 2.5 at.% 259 at 100 mA g−1 None 18
(100%) 2017 [56]
carbon
P-doped hard 205 at 50 mA g−1 154 at 50 mA g−1 Li et al., Adv. Energy
1100 7.3 3 wt% 359 at 50 mA g−1 74
carbon (57%) (43%) Mater., 2017 [66]
P-doped hard Wu et al., ACS Appl.
54 at 50 mA g−1 234 at 50 mA g−1
carbon 1000 50.3 1.4 wt% 288 at 50 mA g−1 56 Mater. Inter., 2018
(19%) (81%)
nanofibers [62]
Large-area
carbon
328 at 100 mA g−1 Hou et al., Adv. Sci.,
P-doping nanosheets 900 549.8 1.39 at.% 328 at 100 mA g−1 None 47
(100%) 2017 [54]
doped with
phosphorus
Flexible P- Lü et al., ACS Appl.
243 at 50 mA g−1
doped carbon 700 345.9 5.2 at.% 243 at 50 mA g−1 None 72 Mater. Inter., 2017
(100%)
cloth [55]
Yu et al., Adv.
P-functionalized 65 at 50 mA g−1 328 at 50 mA g−1
1000 22 N.A. 393 at 50 mA g−1 59 Energy Mater., 2018
hard carbon (17%) (83%)
[63]
F-doped carbon 110 at 50 mA g−1 118 at 50 mA g−1 Wang et al., J. Phys.
1100 46.4 1.1 at.% 228 at 50 mA g−1 52
particles (48%) (52%) Chem. C, 2015 [49]
Other
B-doped hard 77 at 20 mA g−1 70 at 20 mA g−1 Li et al., Adv. Energy
heteroatom 1100 8.0 3.0 wt% 147 at 20 mA g−1 37
carbon (52%) (48%) Mater., 2017 [66]
doping
Wu et al., ACS Appl.
K+-doped hard 196 at 50 mA g−1 118 at 50 mA g−1
1100 78.97 0.6 at.% 314 at 50 mA g−1 69 Mater. Inter., 2018
carbon (62%) (38%)
[63]

7
N and S co-
7.01 at.%
doped hollow 452 at 50 mA g−1 Liao et al., Nanoscale
800 397.7 (N), 3.15 452 at 50 mA g−1 None 35
carbon (38%) Adv., 2018 [65]
at.% (S)
nanofiber
N and S dual- 2.2 at.%
180 at 200 mA g−1 Xu et al., Green
doped porous 800 10.2 (N), 0.97 180 at 200 mA g−1 None 35
(100%) Chem, 2014 [73]
carbon at.% (S)

S-doped N-rich 20.0 at.%


350 at 50 mA g−1 Yang et al., Adv.
Two carbon 650 379.4 (N), 9.2 350 at 50 mA g−1 None 44
(100%) Mater., 2016 [59]
heteroatom nanosheets at% (S)
co-doping
B and N dual-
doped 3D
2.51 at.%
interconnected 536 at 50 mA g−1 Wang et al., Adv.
1000 1585 (N), 10.1 536 at 50 mA g−1 None 37
carbon (100%) Sci., 2017 [58]
at.% (B)
nanofiber thin
film
N and P co- 1.00 at.%
110 at 50 mA g−1 225 at 50 mA g−1 Li et al., Carbon 2016
doped carbon 800 36.8 (N), 1.19 335 at 50 mA g−1 47
(33%) (67%) [51]
microspheres at.% (P)
P-doped hard
223 at 50 mA g−1 105 at 50 mA g−1
carbon derived 1300 24.5 1.10 at.% 328 at 50 mA g−1 72 This work
(68%) (32%)
from lignin
aN.A.: Not available

8
2. EXPERIMENTAL

2.1 Materials

The lignin sample was obtained from a pilot-scale delignification apparatus installed in South

Korea. Lignin was produced from oak hardwood using a concentrated sulfuric acid hydrolysis process.

Detailed information on the lignin sample used in this study is given elsewhere [74]. Phosphoric acid

(purity, >85%) was purchased from J.T. Baker Company (USA). High-purity nitrogen gas (>99.99%)

was purchased from the JC Gas Company (South Korea). Deionized and distilled water was obtained

using an AQUAMaxTM-Basic 363 water purification system equipped with a 0.22-µm filter (Young

Lin Instrument Co., Ltd., South Korea). Polyvinylidene fluoride (PVDF) used as a binder was

purchased from Kureha Chemical Industry Co., Japan. Acetylene black used as a conducting agent

was purchased from DENKA Co. Ltd (Japan). N-methyl-2-pyrrolidinone (NMP, >98%) was

purchased from Alfa Aesar (USA).

2.2 Preparation of undoped hard carbon

The schematic of the preparation of undoped and P-doped hard carbons is shown in Scheme S1

(see Supplementary data). Before carbonization, the lignin sample was purified by washing thrice with

an aqueous KOH solution (20 wt%) at 70 °C, followed by washing with 1 M HCl at 60 °C for 15 h to

remove inorganic species. Five grams of the purified lignin sample was transferred into a model STF

16/180 tubular furnace (Carbolite Gero, UK). The tube was purged with N2 gas for 30 min at room

temperature to remove oxygen or other reactive gases in the tube. The temperature of the tube was

then either increased to 900 or 1300 °C at a ramping rate of 5 °C min−1. When the temperature

reached 900 or 1300 °C, the carbonization was carried out at the final constant temperature for 6 h

under N2 gas flowing at 100 mL min−1. The tube was then cooled to room temperature at a cooling

rate of 5 °C min−1 and the samples carbonized at 900 and 1300 °C were collected; these are referred to

as HC-900 and HC-1300, respectively.

9
2.3 Preparation of doped hard carbon

To produce doped hard carbon, a doping agent (H3PO4 for P-doping, H2SO4 for S-doping, or urea

for N-doping) was added to purified lignin at a mass ratio of dopant:lignin = 0.05:1, 0.1:1, and 0.2:1.

The doping solutions were prepared by dissolving a desired amount of H3PO4, H2SO4, or urea in water.

Then the purified lignin powder was added to the doping solution. The whole mixture was mixed for 2

h by mild stirring. The sample was then slowly dried in a convection oven at 80 °C for 48 h. Then, the

dried sample was calcined under the same conditions described above for producing the undoped hard

carbon. Note that the presence of an excess amount of the dopant during the carbonization led to the

activation of the carbon sheets, resulting in the formation of highly defective, microporous hard

carbon. The doped hard carbon was designated as HC-T-XC-A, T, X, and C represent the

carbonization temperature, dopant species, and dopant concentration, respectively, and A represents

activation of the hard carbon structure.

As discussed in the introduction, highly defective hard carbons typically exhibit low ICE when

tested as an anode in SIBs [40, 60, 75]. To suppress the activation of carbon during the heteroatom

doping, the mixture of the purified lignin sample and dopant solution was filtrated to remove excess

dopants that did not directly interact with the lignin structure. The filtered lignin sample was then

carbonized under the same condition used to produce the undoped hard carbon sample. The produced

hard carbon with suppressed activation is denoted as HC-T-XC, where T, X, and C represent the

carbonization temperature, dopant species, and dopant concentration, respectively.

2.4 Material characterization

An X-ray diffractometer (XRD, D/Max-2500V/PC Rigaku, Japan) equipped with Cu-Kα radiation

source was used to characterize the phase structure. XRD was performed at diffraction angles (2θ)

between 3 and 90° at a scanning rate of 5° min−1. The carbon structure was analyzed using a LabRAM

10
HR800 confocal Raman microscope (HORIBA Jobin-Yvon, USA). The surface elemental

composition was determined by X-ray photoelectron spectroscopy (XPS, model ESCALAB 250Xi,

UK). The morphology of the samples was examined using a JEOL JSM-7500F field-emission

scanning electron microscope (FE-SEM, JEOL Inc., USA). High-resolution transmission electron

microscopy (HR-TEM) images were captured using a JEM ARM 200F instrument (JEOL, USA, Inc.)

operated at 200 kV. The Brunauer-Emmett-Teller (BET) surface area was calculated and the

micropore characteristics were determined using nitrogen adsorption–desorption isotherms acquired

with a BELSORP-MAX instrument (BEL Japan Inc., Japan).

2.5 Electrochemical measurements

The electrochemical properties of the doped and undoped hard carbon were evaluated by the

galvanostatic cycling of a CR2032 coin cell with these materials used as the anode. The coin cell

consisted of metallic sodium (Wellcos Corp., South Korea) or potassium film (Sigma-Aldrich) as both

a counter electrode and a reference electrode, a mixture of 70 wt% hard carbon, 10 wt% acetylene

black, and 20 wt% PVDF as a working electrode, a glass microfiber (GF/B, Whatman, UK) as a

separator, 1 M NaClO4 dissolved in ethylene carbonate (EC)/dimethyl carbonate (DMC)/propylene

carbonate (PC) solvent (volume ratio of EC:DMC:PC = 9:2:9) as the sodium electrolyte, and 0.8 M

KPF6 in EC/diethyl carbonate (DEC) solvent (volume ratio of EC:DEC = 1:1) as the potassium

electrolyte. To prepare the electrode, known amounts of hard carbon, acetylene black, and PVDF were

dispersed in NMP to obtain a slurry. The slurry was then coated onto copper foil (Wellcos Corp.,

South Korea), followed by drying in a vacuum oven at 80 °C overnight to evaporate NMP. The slurry

film was pressed using a twin heat roller (model WCRP-1015HG, Wellcos Corp.) at 100 °C for 10s

for four times to ensure close contact between the Cu foil and electrode materials. Electrode disks

with diameter of 14 mm with an area of 1.54 cm2 were then punched out and weighed. The loading of

the active material was adjusted in the range of 1.0–1.2 mg cm-2. The coin cells were assembled in a

11
glove box filled with ultra-high purity Ar gas (>99.9999%). Oxygen and water contents in the glove

box were controlled below 0.1 ppm. The SIB coin cells were galvanostatically discharged to 0.005 V

(to prevent deposition of sodium metal on the electrode surface) and charged to 2.5 V (vs. Na+/Na)

using a WBCS 3000 model battery test equipment (WonATech Corp., South Korea) at room

temperature. Cyclic voltammetry (CV) of the coin cells was carried out using a model ZIVE MP1

potentiostat analyzer (WonATech Corp., South Korea) at a scanning rate of 0.1 mV s−1 within the

voltage range of 0.005–2.5 V. The rate capability of the hard carbon samples was evaluated by

varying the specific currents from 50 mA g−1 to 5 A g−1 in the voltage range of 0.005–2.5 V.

Galvanostatic intermittent titration technique (GITT) was used to evaluate the Na+-ion diffusion

kinetics in the electrodes at a specific current of 50 mA g−1 and a current pulse time of 10 min at 60

min intervals. The KIB coin cells were galvanostatically discharged to 0.005 V and charged to 2.0 V

(vs. K+/K) using the WBCS 3000 model battery test equipment at room temperature.

3. RESULTS AND DISCUSSION

One of the most important parameters determining the electrochemical performance of hard

carbon is the carbonization temperature. As previously reported, when carbonized in the temperature

range of 1200–1600 °C, the low-voltage plateau below 0.1 V of the hard carbon was enhanced as

compared to those of the hard carbons synthesized at lower temperatures of <1100 °C [46-51, 53-56,

58-65, 67]. With different types of carbon precursors, different carbonization temperatures are

required for producing hard carbons with an extended low-voltage plateau, because of the different

chemical compositions and structures of the sources [42, 43]. In case of lignin-derived hard carbon,

we previously observed that the most favorable Na+ ion intercalation into the graphitic layer occurred

in case of the hard carbon synthesized at 1300 °C [25]. However, a high carbonization temperature is

not favorable for achieving a high level of heteroatom-doping. As discussed in previous reports, the

C–N and C–S bonds decompose quickly at temperatures above 900 °C [52, 76-78]. Similarly, the C–P,

12
–P=O, –P–O bonds form at around 700 °C as the proportion of oxygen starts to bound to phosphorus

[79] and phosphorus initiates to interact with aromatic rings [80], while at high temperatures

(>800 °C), the elimination of –P=O bond begins, leading to low P-doing levels [81]. To first test the

possibility of increasing the doping level while suppressing excess activation, the lignin sample was

charged into the aqueous H3PO4 solution at varying H3PO4:lignin weight ratios ranging from 0.05:1 to

1:1. After the mixture was mildly stirred for 2 h, it was filtered and the amount of H3PO4 loaded in

lignin was measured using a high accuracy balance (model EX125 , OHAUS Co. Ltd., USA); the

results are shown in Fig. S1. When the weight ratio of H3PO4:lignin was increased from 0.05:1 to

0.1:1, the amount of PO43– ions in the lignin increased from 1.3 to 2.7 wt%. A further increase in the

weight ratio by an order of magnitude, from 0.1:1 to 1:1, did not apparently increase the PO43– ion

uptake by lignin. This indicates that the dopant uptake sites in the lignin are highly limited; once all

the uptake sites in lignin interact with PO43− ions at more than a critical concentration (which is a ratio

of H3PO4:lignin = 0.1:1), further increase in the dopant concentration in the aqueous solution does not

increase the uptake of PO43–. Because of the complexity of the lignin structure with different types of

bonding (e.g., –O–4, –O–4, 4–O–5) and oxygenated functionalities (e.g., methoxy group, hydroxyl

group), it is difficult to determine what kind of functionalities in lignin are favorable for the

adsorption of PO43– ions. A plausible site for the PO43– ion uptake could be the -hydroxyl group, as

shown in Fig. S2. Therefore, once all the possible adsorption sites are occupied, an increase in the

dopant concentration does not lead to increased doping of the hard carbon, as listed in Table 2. The

level of P-doping, estimated using the survey XPS scan profiles (Fig. S3), increased slightly from 0.72

to 1.10 at.% with an increase in the dopant:lignin ratio from 0.05:1 to 0.2:1.

13
Table 2. Structural and physical properties of the undoped and heteroatom-doped hard carbons

Doping BET surface Total pore Micropore Mesopore


d002b La b Lc b Lc/d002 + 1 b ID3/IG
Material level aread volumed volumed volumed
(nm) (nm) (nm) ratioc
(at.%)a 2
(m g )−1 3 −1
(cm g ) (%) (%)

HC-1300 None 0.3748 4.17 1.131 4.02 0.72 48.3 0.112 65.7 34.3

HC-1300-P0.72 0.72 0.3797 4.09 1.107 3.93 0.74 50.9 0.089 73.3 26.7

HC-1300-P0.96 0.96 0.3831 4.07 1.097 3.87 0.77 12.1 0.033 34.1 65.9

HC-1300-P1.10 1.10 0.3868 4.03 1.109 3.85 0.79 24.5 0.059 44.2 55.8

HC-1300-P1.12-A 1.12 N.A.e N.A. N.A. N.A. N.A. 1338 2.142 94.2 5.8

HC-900 None 0.3969 2.89 0.765 2.92 1.79 626.9 0.900 95.1 4.9

HC-900-P0.98 0.98 0.3870 3.02 0.912 3.36 1.81 504.1 0.728 94.3 5.7

HC-1300-N0.84 0.84 0.3751 4.03 1.016 3.71 N.A. 15.1 0.031 33.7 66.3

HC-1300-N1.66-A 1.66 0.3750 4.18 1.121 3.98 N.A. 28.3 0.046 76.5 23.5

HC-1300-S0.04 0.04 0.3780 4.01 0.857 3.27 N.A. 68.3 0.123 78.8 21.2

HC-1300-S0.17-A 0.17 0.3849 4.07 0.939 3.44 N.A. 267.6 0.488 92.1 7.9
a Determined using the XPS survey scan
b Determined by XRD analysis
c Estimated from Raman spectra
d Estimated using N adsorption–desorption isotherms
2
e N.A. = not available

14
One can speculate that a potential strategy to increase the doping level is to simply embed excess

PO43− ions in the carbon precursor by slowly evaporating water from an aqueous mixture of dopant

and carbon precursor without filtering the mixture. In fact, this non-filtration approach was employed

in most of the previous works, as listed in Table 1. In case of lignin-H3PO4, the excess PO43– ions that

remained in the lignin matrix upon evaporating water (without filtration) did not increase the P-doping

level. As listed Table 2, the doping level of HC-1300-P1.12-A (without filtration) was very similar to

that of HC-1300-P1.10 (with filtration). This indicates that the excess PO43– ions that were loosely

attached to the surface of hard carbon decomposed during the carbonization of lignin. On the other

hand, the BET surface area of HC-1300-P1.12-A was found to be two orders of magnitude higher than

that of HC-1300-P1.10 (Table 2). This indicates that the presence of excess amounts of dopants

caused undesirable activation of the hard carbon structure, resulting in the generation of a large

number of defects and micropores. The altered textural properties of hard carbon adversely affect the

electrochemical performance of SIBs; the defect sites in hard carbon contribute to an increase in the

high-voltage capacitance-like sloping capacity above 0.1 V [19, 40, 41], whereas excessive carbon

activation causes a decrease in the graphitic domains in hard carbon, thereby leading to lowering of

the plateau capacity [23, 24, 82]. In addition, the high degree of defect formation causes a decrease in

ICE [40]. To overcome the excess activation of carbon, the excess amount of H3PO4 that did not

strongly interact with lignin should be removed before carbonization.

Figs. 1a presents the XRD patterns of the undoped and P-doped hard carbons. Two main broad

reflections are located at diffraction angles (2θ) in the range of 22–24° and 43–44°, which correspond

to the (002) and the (100) crystallographic planes of the graphitic layers in hard carbon, respectively.

Using the Scherrer equation (2dsin θ = nλ), with the 2θ value corresponding to the (002) plane, the

interlayer spacings of the undoped and P-doped hard carbons were calculated (see Fig. 1b). As the P-

doping level increased from 0.72 to 1.10 at.%, the interlayer spacing between the graphene sheets in

the graphitic domains increased from 0.375 to 0.387 nm, indicating that the P-doping increased the

15
interlayer spacing. This trend agrees well with that reported previously [17, 63, 83]; the dilated d-

spacing with P-doping is caused by the larger size of the P atom than the carbon atom size, and it

weakens the van der Waals interactions between the graphene layers. The apparent carbon domain

length parallel to the a-axis (La) and the apparent carbon domain thickness parallel to the c-axis (Lc)

were calculated from the full-width at half maximum (FWHM) of the XRD reflection, using Eq. 1.
𝐾λ
𝐿(𝑛𝑚) = β cosθ (1)

where 𝜆 is the wavelength of the X-ray used (0.154 nm) and 𝛽 is the FWHM of the XRD reflection.

With the 𝐾 values of 0.9 and 1.84 [84] for (002) and (100) reflections of the carbon structure,

respectively, La and Lc of the undoped and P-doped hard carbons were calculated, which are found to

be similar in the range of 4.03–4.17 and 1.109–1.131 nm, respectively (Table 2). This indicates that

the low-level P-doping did not affect the number of stacked graphitic layers and the length of the

graphene sheets, but only enlarged the interlayer spacing. On the other hand, HC-1300-P1.12-A

exhibited a significantly enlarged interlayer spacing of 0.4091 nm, while showing much smaller La

and Lc values as compared to those of HC-1300-P1.10. This is probably caused by the enhanced

formation of defects and micropores, which hinders the ordering of the hexagonal structure of carbon

atoms and reduces the van der Waals interactions between the graphene layers. The enhanced

formation of defects and disordered regions due to the activation of carbon also decreases the

graphitic domains in the HC-1300-P1.12-A sample.

16
Fig. 1. (a) XRD patterns and (b) interlayer spacing of the undoped and P-doped hard carbons.

The chemical states of the doped phosphorus species in the hard carbon were analyzed using the

high-resolution XPS spectra of the P 2p region, as shown in Fig. 2. The P 2p peak was deconvoluted

into four peaks at 130.8, 132.8, 134.2, and 135.6 eV, which can be assigned to –P=O, –P–O, –P–C

bonds [63, 83] and P2O5 (P5+) [85], respectively. The area of each peak and the percentages of various

bonding states are listed in Table S1. As the P-doping level increased from 0.72 to 1.10 at.%, the

percentage of –P–O bonds increased significantly from 0.17 to 0.71 at.%, while those of –P=O and –

P–C bonds did not change significantly. Only the HC-1300-P1.12 sample contains P2O5, suggesting

that excess PO43– ions that were loosely attached to the surface of lignin transformed into P2O5 during

the calcination [86]. On the other hand, the filtering step effectively removed the loosely attached

PO43– ions, and thus the HC-1300-P0.72, HC-1300-P0.96, and HC-1300-P1.10 samples did not exhibit

17
P2O5 species. Considering the fact that Na+ ion adsorption energies are higher for –P=O and –P–C

bonds than those for the –P–O bond and carbon [63], the dominant –P–O bond in HC-1300-P1.10

could increase the interlayer spacing of the graphitic layers, while the Na+-ion adsorption can be

suppressed.

Fig. 2. High-resolution XPS spectra of the P 2p region of the P-doped hard carbons.

18
The morphologies of the doped and undoped hard carbon samples were investigated by SEM

and HR-TEM (Fig. 3). Both the P-doped and undoped hard carbons without activation show quite

similar macroscopic and microscopic morphologies; irregularly shaped particles in the range of 1–5

m are loosely aggregated to form large particles in the range of 15–20 m. Three to four graphitic

layers are stacked together and embedded in amorphous carbon domains. Further, micropores are

observed between the stacked graphitic regions. As the doping concentration increased, the interlayer

spacing increased, which agrees well with the XRD results. In case of HC-1300-P1.12-A, it was very

difficult to identify the stacked graphene layers, which indicates the poor degree of graphitization.

19
Fig. 3. (a)–(e) FE-SEM images, (f)–(j) low-magnification HR images, and (k)–(o) HR-TEM images of
HC-1300, HC-1300-P0.72, HC-1300-P0.96, HC-1300-P1.10, and HC-1300-P1.12-A.

The micropores in hard carbon play an important role in determining its electrochemical

performance, although their effect on the sloping and plateau voltage capacities is not yet clear [31-

35]. Fig. 4a shows the distribution of the micropores in the undoped and P-doped hard carbons, as

20
analyzed by the nonlocal density functional theory (NLDFT), and Fig. 4b shows the N2 adsorption–

desorption isotherms of the hard carbon samples. HC-1300-P1.12-A shows high evolution of micro-

to-meso pores in the range of 0.2–4 nm because of the excess activation, while the micropore

evolution of HC-1300-P0.72, HC-1300-P0.96 and HC-1300-P1.10 was highly suppressed. Both hard

carbon samples exhibited Type 1 isotherms with very steep N2 uptake at low p/po, indicating that they

had a microporous structure [87]. The BET surface area and the micropore volume of the activation-

suppressed P-doped hard carbon decreased as the doping level increased (Table 2); at high P-doping,

the BET surface areas are in the range of 12–25 m2 g−1 and the micropore volumes are in the range of

0.01–0.03 cm3 g−1, which are much smaller than those of the undoped hard carbon (48 m2 g−1, 0.07

cm3 g−1). On the other hand, the HC-1300-P1.12-A sample exhibited two orders of magnitude larger

BET surface area of 1338 cm3 g−1 and micropore volume of 2.02 cm3 g−1 as compared to those of the

undoped and activation-suppressed P-doped hard carbons. This indicates that the direct carbonization

of H3PO4-loaded wood without filtering affect the pore size of the hard carbons significantly. As listed

in Table 1, most of the previous heteroatom-doped hard carbons synthesized at temperatures less than

1000 °C exhibited extremely high surface areas (500–1500 m2 g−1), owing to incomplete formation of

graphitic layers and a high degree of defects; this has a detrimental effect on the ICE (18–60%) [46,

47, 49-51, 53-56, 58-65, 67]. In addition, the major reversible capacity of the hard carbon produced at

a low temperature originates from the sloping voltage due to the Na+-ion adsorption mechanism [15-

17, 19, 32, 41, 44, 75]. On the contrary, in this study, the high carbonization temperature (1300 °C)

increased the ordering of the amorphous carbon domain in the hexagonal structure of graphitic carbon

and caused merging of some of the micropores to form mesopores, resulting in a lower specific

surface area.

21
Fig. 4. (a) Micropore distribution determined by the NLDFT method and (b) N2 adsorption–
desorption isotherms of the undoped and P-doped hard carbons (right panels: zoom-in image).

To investigate the structures of hard carbon, Raman spectroscopy was employed, and the spectra

are shown in Fig. S4. The spectra in the range of 800–2000 cm−1 were deconvoluted into four

Gaussian peaks [88-90]; the presence of the G band at 1594 cm−1 is ascribed to the ideal graphitic

domain with E2g symmetry, while the D1 band at 1341 cm−1 is assigned to the disordered graphitic

domains like the edges of hexagonal planes with A1g symmetry. The D4 and D3 bands at 1200 and

1500 cm−1, respectively, could be attributed to the short-range vibration of sp3 coordinated carbon of

the amorphous carbon structure and the disordered graphitic domains, respectively [89]. The degree of

graphitization could be calculated by identifying the area percentage ratio of the D3 to G peak (ID3/IG)

[50, 89]. As listed in Table 2, the ID3/IG increased slightly from 0.72 (HC-1300) to 0.79 (HC-1300-

P1.10) as the P-doping level increased from 0 to 1.10 at.%. The presence of phosphorus species on the

surface of graphitic carbon increased the degree of disorder in the hard carbon to some extent, which

can be caused by the disturbance of the hexagonal structure by the doped P. However, the increase in

the disorder in the P-doped hard carbon is not significant because of the low doping level. The Raman
22
and N2 adsorption–desorption isotherm analyses indicate that the P-doping of the hard carbon after the

removal of excess dopants that did not directly interact with the lignin structure led to a lower number

of defects and low micropore volume, which are beneficial for achieving high ICE and suppressing

the high-voltage sloping capacity.

As discussed thus far, the high-temperature carbonization and low-level P-doping increased the

interlayer spacing between the graphene sheets, but the other properties of the hard carbon such as the

La, Lc, BET surface area, micropore volume, and degree of graphitization did not change significantly.

To investigate the effect of P-doping on the electrochemical performance of hard carbon,

galvanostatic discharge–charge profiles of the undoped and the P-doped hard carbons were collected

in the voltage range of 0.005–2.5 V at a specific current of 50 mA g−1. As shown in Fig. 5a, the

undoped hard carbon delivered an initial discharge capacity of 378 mAh g−1 and initial charge

capacity of 261 mAh g−1 with ICE of 69%. As the P-doping level increased from 0.72 to 1.10, the

initial discharge capacities of the P-doped hard carbons increased from 424 to 443 mAh g−1 and the

initial charge capacities increased from 301 to 319 mAh g−1 with the ICE being in the range of 71–

73%. Table S2 shows the relationship between surface area and ICE values. The high ICE values are

attributed to the very low surface area of the hard carbons (below 50 m2 g−1). In sharp contrast, the

HC-1300-P1.12-A sample delivered initial discharge and charge capacities of 387 and 42 mAh g−1,

resulting in extremely low ICE of 11%. The HC-1300-P1.12-A sample rich in defects and micropores

provides potential sites for the undesirable solid-electrolyte interphase (SEI) formation during the

initial sodiation. Therefore, the effective suppression of the activation of carbon during the P-doping

led to a low BET surface area and thus suppressed the SEI layer formation, which is important in the

practical application of the anode [91, 92]. To further investigate the effect of P-doping on the Na+-ion

storage mechanism, the 2nd and 10th discharge capacities of the undoped and P-doped hard carbons

were divided into a sloping voltage region between 1.0 and 0.1 V and a flat voltage region below 0.1

V, as shown in Figs. 5b–c. The sloping capacities of the undoped and P-doped hard carbons were

23
found to be quite similar, whereas the plateau capacities increased from 183 (HC-1300) to 223 mAh

g−1 (HC-1300-P1.10) as the doping level increased from 0 to 1.10 at.%. The contribution of the low-

voltage plateau capacity to the total capacity increased from 64% (HC-1300) to 68% (HC-1300-P1.10).

This indicates that the P-doped hard carbon produced at the high carbonization temperature of

1300 °C can enhance the plateau capacity without affecting the sloping capacity, which is beneficial

for fabricating high-energy-density batteries in full cell configuration. Although the origin of the

plateau capacity below 0.1 V is a topic of debate (micropore volume filling [31-35] vs. intercalation

into the graphitic layers [15, 16, 19, 22, 36-41]), our recent investigation using ex situ XRD and solid

state 23Na NMR suggests that intercalation of Na+ ion into the graphitic layer of the lignin-derived

hard carbon could be responsible for the low-voltage plateau capacity, followed by filling Na+ ions

into micropores below 0.03 V [25]; most of the low potential capacity was originated from the Na+

ions intercalation into the graphitic layer in the voltage of 0.1–0.03 V, followed by a small capacity

contribution (approximately 40 mAh g−1) that could be originated from micropore filling in the

voltage of 0.03–0.005 V. By increasing the interlayer spacing with the P-doping, the low potential

capacity can be enhanced somewhat significantly from 183 mAh g−1 to 223 mAh g−1. As listed in

Table 2, the micropore volume of the P-doped hard carbon is slightly lower than that of undoped hard

carbon. In fact, the capacities of undoped and P-doped hard carbon at the voltage of 0.03–0.005 V

were very similar (39–41 mAh g−1), indicating that the undoped and P-doped hard carbons have

similar micropore-filling capacities at the low voltages. Therefore, the phosphorus species embedded

in the hard carbon structure play an important role in facilitating Na+-ion diffusion into the interlayer

spacing between the graphene sheets without affecting the sloping capacity and ICE values. The

enlarged interlayer spacing between the hexagonal planes could provide increased active sites for Na+-

ion uptake [66]. On the other hand, the HC-1300-P1.12-A, which was enriched with micropores,

delivered much smaller flat voltage capacity of 24 mAh g−1 and total capacity of 62 mAh g−1 as

compared to those of the undoped hard carbon and the activation-suppressed P-doped hard carbon.

24
Therefore, the excess defects and micropores formed due to the activation of carbon have a

detrimental effect on the Na+-ion uptake in both the sloping voltage and plateau voltage regions.

Although it is not clear what causes this, the excessive formation of the SEI layer on the surface of the

HC-1300-P1.12-A electrode and strong interaction between the pre-adsorbed Na+ ions in the defect

sites could retard the subsequent Na+-ion diffusion into both the free defect sites, the graphitic region,

and the micropores.

Fig. 5. (a) Discharge–charge profiles and capacities from the sloping voltage above 0.1 V and from
the plateau voltage below 0.1 V in the (b) 2nd cycle and (c) 10th cycle of the undoped and P-doped
hard carbons.

25
To further investigate the effect of P-doping on the Na+ ion storage mechanism, Na+-ion

diffusivity was analyzed using GITT. The GITT profiles of the undoped and the P-doped hard carbons

were collected in the second discharge–charge cycle from 0.005 V to 2.5 V with a pulse current of 50

mA g−1 for 10 min between the rest intervals of 60 min (Figs. S5a–b). Based on the Fick’s second law

diffusion, the diffusivity of Na+ ions (𝐷𝑁𝑎 + ) can be calculated using Eq. 2 [92].

2
𝐷𝑁𝑎 + =
4 𝑚𝑐 𝑉𝑐 2
𝜋 ( )
𝑀𝑐 𝑆 ( )
∆𝐸𝑠
𝑑𝐸𝑡
𝜏 (𝑑 𝜏)
(2)

where 𝑉𝑐 is the molar volume of the carbon material (8 cm3 mol−1), 𝑚𝑐 is the mass of the hard carbon

(0.00154–0.00168 g), 𝑀𝑐 is the molar mass of the carbon material (12 g mol−1), 𝑆 is the surface area of

the electrode disk (1.54 cm2), 𝜏 is the pulse duration (10 min), ∆Es is the difference between two

sequential open-circuit potentials (V), and ∆𝐸𝑡 is the voltage shift when the current pulse is applied

(V). A detailed illustration of ∆𝐸𝑠 and ∆𝐸𝑡 in each current step is shown in Fig. S5d. Eq. 2 can be

simplified using the linear relationship between Et and √t, as presented in Fig. S5c.

𝑚𝑐 𝑉𝑐 2 ∆𝐸𝑠 2
𝐷𝑁𝑎 + =
4
𝜋𝜏 ( )( )
𝑀𝑐 𝑆 ∆𝐸𝑡
(3)

The 𝐷𝑁𝑎 + values calculated using Eq. 3 during the sodiation and desodiation processes are

shown in Fig. 6a and 6b, respectively. At the beginning of sodiation, the apparent 𝐷𝑁𝑎 + decreased

gradually to 1.3 × 10–9 cm2 s−1 as the voltage decreased to 0.2 V. In the voltage range of 0.2–0.1 V,

the 𝐷𝑁𝑎 + increased slightly, followed by a sharp decrease in 𝐷𝑁𝑎 + to 2 × 10–10 cm2 s−1 when the

voltage decreased to 0.04 V. A further decrease in the voltage to 0.005 V led to an increase in 𝐷𝑁𝑎 + to

8 × 10–10 cm2 s−1. During desodiation, the 𝐷𝑁𝑎 + exhibited an opposite behavior to that of the sodiation

process, indicating reversible insertion and extraction of Na+ ions into the hard carbon. The larger

𝐷𝑁𝑎 + values in the in the sloping region (> 0.2 V) than those in the plateau region (<0.1 V) is possibly

because of the thermodynamically favorable adsorption of Na+ ions onto the defect sites [19, 41]; once

the surface active sites are occupied with the Na+ ions, it is difficult for the subsequent Na+ ions to

26
diffuse into the microdomain consisting of graphite-like carbon layers. The sharp decrease in 𝐷𝑁𝑎 + in

the voltages between 0.1 and 0.04 V may be because more energy is required to overcome the

repulsive forces generated from the strongly interacting adsorbed Na+ ions and defect sites on the

surfaces before the Na+ ions intercalate into the microdomain of graphitic layers [32, 41]. The U-turn

behavior at 0.2 V and 0.03 V could be ascribed to Na+ ions adsorbing onto or filling into the

micropores, as discussed previously using ex situ XPS and space-filling model [25]. The Na+-ion

diffusivity behavior of the undoped hard carbon was very similar to those of the activation-suppressed

P-doped hard carbons, suggesting that the P-doping did not change the sodiation and desodiation

mechanisms. In case of the HC-1300-P1.12-A sample, the rapid decay of 𝐷𝑁𝑎 + in the sloping region is

probably due to the formation of a thick SEI layer and strong interaction between the defect sites and

previously adsorbed Na+ ions, which make it difficult for subsequent Na+ ion to adsorb on the

unoccupied defect sites and to diffuse into the graphitic region. In addition, the presence of excess

micropores could hinder the intercalation of Na+ ions into the graphitic layer because the energy

required for the ions to diffuse into the graphitic region is reduced by the accumulation of Na+ ions

that are pre-adsorbed on the micropore surfaces.

27
Fig. 6. Apparent Na+ ion diffusion coefficients calculated from the GITT profiles of the undoped and
P-doped hard carbons during (a) sodiation and (b) desodiation in the 2nd cycle.

To further investigate the role of P-doping as a redox active center or as an enlarger of the

interlayer spacing, CV profiles of the undoped and P-doped hard carbons were collected at a scanning

rate of 0.1 mV s−1, and the results are shown in Fig. S6. In the first cathodic scan, a broad irreversible

peak at ~0.5 V was observed for both undoped and the activation-suppressed P-doped hard carbons,

which can be attributed to the SEI formation [93, 94]. The CV profiles of the P-doped hard carbons

are very similar to that of the undoped hard carbon. The absence of a peak associated with the

reduction of POx species, which is typically caused by the adsorption of Na+ ions onto the active sites

of POx, indicates that the P-doping has a negligible effect on the direct uptake of Na+ ions. This could

be because of the low P-doping level and dominant –P–O bonds, which could suppress the Na+ ion

adsorption in the high voltage region. In case of the HC-1300-P1.12-A sample, the absence of sharp

redox peak pairs at 0.01/0.12 V indicates that the Na+ ion intercalation into the graphitic region was

negligible.

The rate performance and long-term cycling stability of the undoped and the P-doped hard

carbons were evaluated by varying the specific currents between 0.05–5.0 A g−1, and the results are

shown in Fig. 7. The P-doped hard carbon consistently delivered larger capacities than those of the

undoped hard carbon. As shown in the discharge–charge profiles in Fig. S7 and summarized in Fig. 7b,

at the high specific current of 0.5 A g−1, the HC-1300-P0.96 and HC-1300-P1.10 samples still

maintained their low-voltage plateau, while the HC-1300 and HC-1300-P0.72 samples did not exhibit

the low-voltage plateau. The contributions of the low-voltage plateau capacity to the total capacity of

the HC-1300-P0.96 and HC-1300-P1.10 samples are 36 and 42%, respectively. Because of the

presence of the low-voltage plateau for the highly P-doped hard carbon at 0.5 A g−1, the total capacity

increased with an increase in the P-doping level. Based on the “adsorption–intercalation” mechanism

[15, 16, 19, 22, 36-41], the presence of the low-voltage plateau at 0.5 A g−1 indicates that Na+-ion

penetration into the graphitic region of the high-level P-doping samples was facilitated, which could
28
be because of the more dilated graphitic layers as compared to those of the undoped and low-level P-

doped hard carbon samples. In addition, the phosphorus species embedded in the hard carbon could

promote charge transfer kinetics of Na+ ions [62, 63]. At the high specific currents of > 1.0 A g–1, the

difference in reversible capacities of the undoped and P-doped hard carbons became similar, which

could be because only the slopping capacities contribute to the total capacities, which are dependent

on the defect and edge sites. To examine the charge transfer kinetics, EIS data of HC-1300 and HC-

1300-P1.10 were collected, and the results are shown in Figs. S8a–b. The EIS data were fitted using

an equivalent circuit model (Fig. S8c), where Re is the electrolyte solution resistance, RSEI is the SEI

formation resistance, Rct is the charge transfer resistance associated with the Na/Na+ redox reaction at

the active material surface, CPE is the constant phase angle element associated with the double-layer

capacitance, and W is the Warburg impedance. As presented by the Bode plots (Figs. S9–S10), the

change in the slope of the high-frequency region indicates the presence of two overlapped semicircles

in the low-to-medium frequency range in the Nyquist plot. As listed in Table S3, the Rct values of HC-

1300-P1.10 were lower than those of HC-1300 during the 2nd sodiation and desodiation process,

suggesting that enhanced Na+-ion transport in the P-doped hard carbon electrode. The voltage plateau

capacity decreases at the high specific currents of 1–5 A g–1, because the Na+ ions do not have enough

time to diffuse into the graphitic layers under the fast discharge–charge conditions. When the specific

current was returned to 0.05 A g−1, each electrode recovered its initial capacity, thus indicating the

excellent integrity of cells with both the undoped and the P-doped hard carbon anodes. As shown in

Fig. 7c, the long-term cyclability of the hard carbon samples indicates a marginal capacity loss at a

high specific current of 0.3 A g−1. In addition, the low-voltage plateau capacities of the P-doped hard

carbons were maintained during 700 cycles (Fig. 7d). This indicates that as in the case of undoped

hard carbon, the P-doped hard carbon remained very stable under the fast discharge–charge condition.

29
Fig. 7. (a) Rate performance, (b) capacity from the sloping voltage above 0.1 V and from the plateau
voltage at 0.5 A g−1, (c) long-term cyclability at 0.3 A g−1 up to 700 cycles, and (d) capacities from the
sloping voltage above 0.1 V and from the plateau voltage at the 700th cycle with a specific current of
0.3 A g−1 of the undoped and P-doped hard carbons. In Figs. 7a and c, filled circles are discharge
capacities and open circles are charge capacities. The starting potentials of the plateau capacities at
current densities of 0.5 A g−1 and 0.3 A g−1 were 0.04 V and 0.05 V, respectively, because of the
difference in polarization depending on the current densities.

The correlation between the physicochemical properties and electrochemical properties of the

undoped and P-doped hard carbons discussed so far indicates that the superior capacity of the P-doped

hard carbon can be attributed to the dilated interlayer spacing between the basal planes of the graphitic

layers, which enhanced the Na+-ion diffusion kinetics and extended the low-voltage plateau capacity.

Although expanded hard carbon could be obtained by decreasing the carbonization temperature (900–

1100 °C), incomplete development of the graphitic region at the low carbonization temperatures leads

to shortening of the plateau capacity [25, 26]. Some of heteroatom-doped hard carbons synthesized at

30
low carbonization temperatures did not exhibit the low-voltage plateau capacity [46, 47, 51-55, 58, 60,

65, 72, 73], possibly because of the absence of well-developed graphitic domains. In contrast to the

previous low-temperature doping approaches (600–1100 °C) [17, 46-65, 67, 72, 73, 84], the increase

in the low-voltage plateau capacity of the P-doped hard carbon synthesized at the high calcination

temperature of 1300 °C in this study is clearly advantageous for fabricating high-energy-density

electrodes. In addition, the low doping level did not decrease the ICE. For example, as listed in Table

1, the low-voltage plateau capacity of HC-1300-P1.10 was found to be 224 mAh g−1 at 50 mA g−1,

with the ICE being 72%, values that are much higher than those reported for most of the previous

heteroatom-doped hard carbons. To investigate the possibility of increasing the low-voltage plateau

capacity of P-doped hard carbon synthesized at a low carbonization temperature, lignin-derived

undoped and P-doped hard carbons were synthesized at 900 °C, and the results are shown in Fig. S11.

After P-doping, the low-voltage plateau capacity increased from 15 to 56 mAh g−1 at 0.05 A g−1 after

10 cycles (Fig. S11b), while the high-voltage sloping capacities of the P-doped and undoped hard

carbons are quite similar. In addition, at a high specific current of 0.3 A g−1, the P-doped hard carbon

still exhibited the low-voltage plateau, while the undoped hard carbon did not exhibit it (Fig. S11c).

As shown in Figs. S11d–e, the high-rate capacities of HC-900-P0.98 are consistently larger than those

of HC-900. To investigate the beneficial effect of the P-doping at the low calcination temperature of

900 °C, the physicochemical properties of the P-doped and undoped hard carbons were analyzed, and

the results are shown in Figs. S11f–i and Table 2. The Raman spectra (shown in Fig. S11f) and the

ID3/IG values of HC-900-P0.98 and HC-900 are quite similar, which could be the reason for the similar

sloping capacity. In contrast, as shown in Fig. S11g, the (002) crystallographic plane of the graphitic

layers in HC-900-P0.98 was more developed as compared to that in HC-900. As a result, HC-900-

P0.98 exhibited longer and thicker graphitic layers than those of HC-900, which is probably due to the

swift graphitization around the P species. The presence of more ordered stacked graphene layers in

HC-900-P0.98 resulted in increased plateau capacity and high-rate performance. The textural

31
properties of HC-900 and HC-900-P0.98 were also investigated (Figs. S11h–i). The slightly lower

micropore volume of HC-900-P0.98 could be attributed to the enhanced ordering of the amorphous

carbon matrix and the collapse of the micropores. The lower amount of micropores in HC-900-P0.98

could be the reason of higher ICE value of 42% than that of HC-900 (31%).

Previously, it has been noted that sulfur and nitrogen could be promising dopants to improve the

electrochemical performance of hard carbon, as listed in Table 1. However, as in the case of P-doping,

almost all the S- and N-doping of hard carbon was conducted at low temperatures of 600–950 °C,

which are not enough to endow the hard carbon with the unique plateau capacity. In this section, we

investigated the possibility of maintaining the plateau capacity of S- and N-doped hard carbon

synthesized with a dopant-to-lignin weight ratio of 0.2:1 at the high temperature of 1300 °C using a

doping procedure identical to that of P-doping. As listed in Table 1, the N-doping levels are similar to

that of the P-doping levels, while the S-doping levels are extremely small. This indicates that instead

of being incorporated in the carbon host, the S dopant species might have decomposed to SOx during

the high-temperature calcination. Because of the extremely low doping level, HC-1300-S0.04

exhibited discharge–charge profiles and reversible capacities similar to those of undoped hard carbon

(Figs. S12a–d). To increase the S-doping level, the SO42−-adsorbed lignin was directly carbonized

without filtration of the precursor mixture. A slight increase in the S-doping level to 0.17 at.% was

observed. However, both the sloping and plateau capacities decreased significantly for the HC-1300-

S0.17-A electrode, as shown in Fig. S12c. The crystalline structure of HC-1300-S0.17-A was quite

similar to those of undoped hard carbon and HC-1300-S0.04, whereas the BET surface area and

micropore volume of HC-1300-S0.17-A were significantly increased. The presence of excess

micropores in HC-1300-S0.17-A, because of enhanced activation of carbon caused by sulfuric acid

treatment, is the reason for the lower ICE (49%) as compared to that of HC-1300 and HC-1300-S0.04

(69%), and for the decreased capacity, as in the case of HC-1300-P-1.12A.

32
In case of N-doping, the doping level without filtration was much higher (1.66 at.%) than that

with filtration (0.84 at.%). Although the N-doping levels were much higher than the S-doping levels,

the sloping/plateau capacities and the total reversible capacities of the N-doped hard carbons were

similar to those of undoped hard carbon (Figs. S12a–d). Analysis of the structural and textural

properties indicated that the undoped and N-doped hard carbons have similar characteristics (Figs.

S12e–d, Table 2). In contrast to the previous studies that demonstrated the effective contribution of N-

doping in increasing the Na+-ion uptake by the hard carbons produced at a low temperature [46, 47, 52,

59-61, 67, 72, 95], the N-doping at the high carbonization temperature did not result in a change in the

physicochemical properties and electrochemical performance.

Lastly, to examine the possibility of enhancing the electrochemical performance of the P-doped

carbon in KIBs, discharge–charge, rate-performance, and long-term cyclability of HC-1300 and HC-

1300-P1.10 in were studied using the potassium electrolyte (Fig. 8). The galvanostatic discharge–

charge profiles corresponding to the initial 10 cycles at a specific current of 50 mA g−1 show an

increase in the low-voltage capacity below 0.25 V from 153 to 192 mAh g−1 after P-doping, while the

high-voltage sloping capacities above 0.25 V are similar for both the undoped and P-doped hard

carbons (Figs. 8a–b). The voltage profile of hard carbon in KIBs is different from in SIBs because of

difference in the binding energy of K+ and Na+ ions into hard carbon, as discussed in a previous paper

[70]. As shown in Fig. 8a, small but clear plateau potentials at around 0.25 V in KIBs were observed.

As a result, the total reversible capacity of the P-doped hard carbon was found to be higher (302 mAh

g−1) than that of the undoped hard carbon (245 mAh g−1). The ICE values of HC-1300-P1.10 and HC-

1300 were found to be quite similar (46–47%). As in the case of SIBs, the HC-1300-P1.10 electrode

exhibited consistently enhanced high-rate performance at the specific currents of 0.10–1.00 A g−1 in a

KIB. At a high current density of 1.0 A g−1 (Fig. S13), HC-1300-P1.10 still maintained its low-voltage

plateau capacity, while the low-voltage plateau disappeared in the HC-1300 electrode at 1.0 A g−1.

This indicates that K+-ion penetration into the graphitic region of the P-doped hard carbon sample was

33
facilitated because of the dilated graphitic layers as compared to those of the undoped hard carbon

sample. In addition, the P-doped carbon delivered extremely stable capacities over 700 cycles with a

very small capacity decay of 0.02 mAh g−1 per cycle at 0.3 A g−1. Therefore, the high-temperature P-

doping of hard carbon can be a promising approach to enhance the low-voltage plateau capacity and

high-rate performance of both the SIBs and KIBs with the maintenance of the high ICE, which is very

important for developing high-energy-density cells.

Fig. 8. (a) Discharge–charge profiles, (b) sloping and plateau capacities at the 2nd cycle and 0.05 A g−1,
(c) rate performance, and (d) long-term cyclability of the undoped and P-doped hard carbons in KIBs.
In Fig. 8d, filled circles are discharge capacities and open circles are charge capacities.

34
4. CONCLUSIONS

In summary, P-doped hard carbon was synthesized by simply mixing lignin with phosphoric acid,

filtering the mixture to remove unabsorbed PO43– ions, and carbonizing the material at the high

temperature of 1300 °C. The activation of hard carbon, which is typically caused by the excess dopant,

was prevented, thus resulting in a low specific surface area (~25 m2 g−1) and low micropore volume

(0.026 cm3 g−1). The major chemical species in the basal plane of the P-doped hard carbon was found

to be the –P–O bond, which increases the interlayer spacing between the graphitic layers from 0.375

nm (undoped) to 0.387 nm (P-doped), while suppressing Na+ ion adsorption. The effective

suppression of defects and micropores and the dominant –P–O bonds in the P-doped hard carbon

enabled it to deliver much higher low-voltage plateau capacity of 223 mAh g−1 as compared to that of

undoped hard carbon (183 mAh g−1) at 50 mA g−1 when tested as an anode in SIBs. The increase in

the low-voltage plateau capacity of the P-doped hard carbon is the main reason for the enhancement of

the total reversible capacity to 328 mAh g−1 as compared to that of the undoped hard carbon (286

mAh g−1). In addition, a high ICE of 72% was maintained because of the suppression of defects and

micropores. The P-doping was also effective in enhancing the low-voltage plateau capacity of the hard

carbon synthesized at a low temperature (900 °C). Other types of dopant (S and N) did not enhance

the electrochemical performance of doped hard carbon because of the decomposition of the dopant at

the high carbonization temperature and negligible changes in the interlayer spacing. The P-doped hard

carbon was effective in enhancing the electrochemical performance of KIBs. The low-voltage

capacity below 0.25 V increased from 153 to 192 mAh g−1 and the total reversible capacity increased

from 245 to 302 mAh g−1 after P-doping. The results of this study indicate that low-level P-doping of

lignin-derived hard carbon is a very promising approach for developing high-energy-density anodes

with high ICE, which is important in the practical application of SIBs and KIBs.

35
ACKNOWLEDGMENTS

This research was supported by a National Research Foundation of Korea (NRF) grant funded by

the Korean government (MSIP) (grant number: NRF-2016R1A2B3008800). Additional support from

the Waste-to-Energy Technology Development Program of the Korea Environmental Industry &

Technology Institute, a financial resource grant from the Ministry of Environment, Republic of Korea

(No. 2018001580001).

REFERENCES

[1] M.D. Slater, D. Kim, E. Lee, C.S. Johnson, Sodium-Ion Batteries, Adv. Funct. Mater. 23 (2013)

947-958.

[2] N. Yabuuchi, K. Kubota, M. Dahbi, S. Komaba, Research Development on Sodium-Ion Batteries,

Chem. Rev. 114 (2014) 11636-11682.

[3] K. Nobuhara, H. Nakayama, M. Nose, S. Nakanishi, H. Iba, First-principles study of alkali metal-

graphite intercalation compounds, J. Power Sources 243 (2013) 585-587.

[4] H. Moriwake, A. Kuwabara, C.A.J. Fisher, Y. Ikuhara, Why is sodium-intercalated graphite

unstable?, RSC. Adv. 7 (2017) 36550-36554.

[5] Y. Liu, B.V. Merinov, W.A. Goddard, Origin of low sodium capacity in graphite and generally

weak substrate binding of Na and Mg among alkali and alkaline earth metals, Proc. Natl. Acad. Sci.

113 (2016) 3735.

[6] B. Jache, P. Adelhelm, Use of Graphite as a Highly Reversible Electrode with Superior Cycle Life

for Sodium-Ion Batteries by Making Use of Co-Intercalation Phenomena, Angew. Chem., Int. Ed. 53

(2014) 10169-10173.

[7] J. Sangster, C-Na (Carbon-Sodium) System, J. Phase Equilib. Diffus. 28 (2007) 571-579.

[8] K. Nobuhara, H. Nakayama, M. Nose, S. Nakanishi, H. Iba, First-principles study of alkali metal-

graphite intercalation compounds, J. Power Sources 243 (2013) 585-587.

36
[9] S.-W. Kim, D.-H. Seo, X. Ma, G. Ceder, K. Kang, Electrode Materials for Rechargeable Sodium-

Ion Batteries: Potential Alternatives to Current Lithium-Ion Batteries, Adv. Energy Mater. 2 (2012)

710-721.

[10] C. Wang, Y. Zhang, W. He, X. Zhang, G. Yang, Z. Wang, M. Ren, L. Wang, Na-Doped C70

Fullerene/N-Doped Graphene/Fe-Based Quantum Dot Nanocomposites for Sodium-Ion Batteries with

Ultrahigh Coulombic Efficiency, ChemElectroChem 5 (2018) 129-136.

[11] X. Yi, W. He, X. Zhang, G. Yang, Y. Wang, Hollow mesoporous MnO/MnS/SiC/S-CN

composites prepared from soda pulping black liquor for lithium-ion batteries, J. Alloys Compd. 735

(2018) 1306-1313.

[12] X. Yi, W. He, X. Zhang, Y. Yue, G. Yang, Z. Wang, M. Zhou, L. Wang, Graphene-like carbon

sheet/Fe3O4 nanocomposites derived from soda papermaking black liquor for high performance

lithium ion batteries, Electrochim. Acta 232 (2017) 550-560.

[13] W. Yang, W. He, X. Zhang, G. Yang, J. Ma, Y. Wang, C. Wang, Na3V2(PO4)3/N-doped Carbon

Nanocomposites with Sandwich Structure for Cheap, Ultrahigh-Rate, and Long-Life Sodium-Ion

Batteries, ChemElectroChem 6 (2019) 2020-2028.

[14] F. Shen, W. Luo, J. Dai, Y. Yao, M. Zhu, E. Hitz, Y. Tang, Y. Chen, V.L. Sprenkle, X. Li, L. Hu,

Ultra-Thick, Low-Tortuosity, and Mesoporous Wood Carbon Anode for High-Performance Sodium-

Ion Batteries, Adv. Energy Mater. 6 (2016) 1600377.

[15] J. Ding, H. Wang, Z. Li, A. Kohandehghan, K. Cui, Z. Xu, B. Zahiri, X. Tan, E.M. Lotfabad, B.C.

Olsen, D. Mitlin, Carbon Nanosheet Frameworks Derived from Peat Moss as High Performance

Sodium Ion Battery Anodes, ACS Nano 7 (2013) 11004-11015.

[16] E.M. Lotfabad, J. Ding, K. Cui, A. Kohandehghan, W.P. Kalisvaart, M. Hazelton, D. Mitlin,

High-density sodium and lithium ion battery anodes from banana peels, ACS Nano 8 (2014) 7115-

7129.

37
[17] Y. Miao, J. Zong, X. Liu, Phosphorus-doped pitch-derived soft carbon as an anode material for

sodium ion batteries, Mater. Lett. 188 (2017) 355-358.

[18] P. Liu, Y. Li, Y.-S. Hu, H. Li, L. Chen, X. Huang, A waste biomass derived hard carbon as a

high-performance anode material for sodium-ion batteries, J. Mater, Chem. A 4 (2016) 13046-13052.

[19] S. Qiu, L. Xiao, M.L. Sushko, K.S. Han, Y. Shao, M. Yan, X. Liang, L. Mai, J. Feng, Y. Cao, X.

Ai, H. Yang, J. Liu, Manipulating Adsorption–Insertion Mechanisms in Nanostructured Carbon

Materials for High-Efficiency Sodium Ion Storage, Adv. Energy Mater. 7 (2017) 1700403-n/a.

[20] Y. Yao, F. Wu, Naturally derived nanostructured materials from biomass for rechargeable

lithium/sodium batteries, Nano Energy 17 (2015) 91-103.

[21] V. Simone, A. Boulineau, A. de Geyer, D. Rouchon, L. Simonin, S. Martinet, Hard carbon

derived from cellulose as anode for sodium ion batteries: Dependence of electrochemical properties

on structure, J. Energy Chem. 25 (2016) 761-768.

[22] Q. Wang, X. Zhu, Y. Liu, Y. Fang, X. Zhou, J. Bao, Rice husk-derived hard carbons as high-

performance anode materials for sodium-ion batteries, Carbon 127 (2018) 658-666.

[23] H. Wang, W. Yu, J. Shi, N. Mao, S. Chen, W. Liu, Biomass derived hierarchical porous carbons

as high-performance anodes for sodium-ion batteries, Electrochim. Acta 188 (2016) 103-110.

[24] K.-l. Hong, L. Qie, R. Zeng, Z.-q. Yi, W. Zhang, D. Wang, W. Yin, C. Wu, Q.-j. Fan, W.-x.

Zhang, Y.-h. Huang, Biomass derived hard carbon used as a high performance anode material for

sodium ion batteries, J. Mater, Chem. A 2 (2014) 12733-12738.

[25] S. Alvin, D. Yoon, C. Chandra, H.S. Cahyadi, J.-H. Park, W. Chang, K.Y. Chung, J. Kim,

Revealing sodium ion storage mechanism in hard carbon, Carbon 145 (2019) 67-81.

[26] D. Yoon, J. Hwang, W. Chang, J. Kim, Carbon with Expanded and Well-Developed Graphene

Planes Derived Directly from Condensed Lignin as a High-Performance Anode for Sodium-Ion

Batteries, ACS Appl. Mater. Interfaces. 10 (2018) 569-581.

38
[27] Y. Qi, Y. Lu, F. Ding, Q. Zhang, H. Li, X. Huang, L. Chen, Y.-S. Hu, Slope-Dominated Carbon

Anode with High Specific Capacity and Superior Rate Capability for High Safety Na-Ion Batteries,

Angew. Chem. Int. Ed. 58 (2019) 4361−4365.

[28] W. Luo, Z. Jian, Z. Xing, W. Wang, C. Bommier, M.M. Lerner, X. Ji, Electrochemically

Expandable Soft Carbon as Anodes for Na-Ion Batteries, ACS Cent. Sci. 1 (2015) 516-522.

[29] Z. Jian, C. Bommier, L. Luo, Z. Li, W. Wang, C. Wang, P.A. Greaney, X. Ji, Insights on the

Mechanism of Na-Ion Storage in Soft Carbon Anode, Chem. Mater. 29 (2017) 2314-2320.

[30] M. Wahid, D. Puthusseri, Y. Gawli, N. Sharma, S. Ogale, Hard Carbons for Sodium-Ion Battery

Anodes: Synthetic Strategies, Material Properties, and Storage Mechanisms, ChemSusChem 11 (2018)

506-526.

[31] S. Komaba, W. Murata, T. Ishikawa, N. Yabuuchi, T. Ozeki, T. Nakayama, A. Ogata, K. Gotoh,

K. Fujiwara, Electrochemical Na Insertion and Solid Electrolyte Interphase for Hard-Carbon

Electrodes and Application to Na-Ion Batteries, Adv. Funct. Mater. 21 (2011) 3859-3867.

[32] P. Bai, Y. He, X. Zou, X. Zhao, P. Xiong, Y. Xu, Elucidation of the Sodium-Storage Mechanism

in Hard Carbons, Adv. Energy Mater. 8 (2018) 1703217.

[33] D.A. Stevens, J.R. Dahn, An In Situ Small‐Angle X‐Ray Scattering Study of Sodium

Insertion into a Nanoporous Carbon Anode Material within an Operating Electrochemical Cell, J.

Electrochem. Soc. 147 (2000) 4428-4431.

[34] D.A. Stevens, J.R. Dahn, The Mechanisms of Lithium and Sodium Insertion in Carbon Materials,

J. Electrochem. Soc. 148 (2001) A803-A811.

[35] J.M. Stratford, P.K. Allan, O. Pecher, P.A. Chater, C.P. Grey, Mechanistic insights into sodium

storage in hard carbon anodes using local structure probes, Chem. Commun. 52 (2016) 12430-12433.

[36] P.-c. Tsai, S.-C. Chung, S.-k. Lin, A. Yamada, Ab initio study of sodium intercalation into

disordered carbon, J. Mater, Chem. A 3 (2015) 9763-9768.

39
[37] H. Lu, F. Ai, Y. Jia, C. Tang, X. Zhang, Y. Huang, H. Yang, Y. Cao, Exploring Sodium-Ion

Storage Mechanism in Hard Carbons with Different Microstructure Prepared by Ball-Milling Method,

Small 14 (2018) 1802694.

[38] Z. Li, L. Ma, T.W. Surta, C. Bommier, Z. Jian, Z. Xing, W.F. Stickle, M. Dolgos, K. Amine, J.

Lu, T. Wu, X. Ji, High Capacity of Hard Carbon Anode in Na-Ion Batteries Unlocked by POx Doping,

ACS Energy Lett. 1 (2016) 395-401.

[39] P. Wang, X. Zhu, Q. Wang, X. Xu, X. Zhou, J. Bao, Kelp-derived hard carbons as advanced

anode materials for sodium-ion batteries, J. Mater, Chem. A 5 (2017) 5761-5769.

[40] L. Xiao, H. Lu, Y. Fang, M.L. Sushko, Y. Cao, X. Ai, H. Yang, J. Liu, Low-Defect and Low-

Porosity Hard Carbon with High Coulombic Efficiency and High Capacity for Practical Sodium Ion

Battery Anode, Adv. Energy Mater. 8 (2018) 1703238.

[41] C. Bommier, T.W. Surta, M. Dolgos, X. Ji, New Mechanistic Insights on Na-Ion Storage in

Nongraphitizable Carbon, Nano Lett. 15 (2015) 5888-5892.

[42] X. Dou, I. Hasa, M. Hekmatfar, T. Diemant, R.J. Behm, D. Buchholz, S. Passerini, Pectin,

Hemicellulose, or Lignin? Impact of the Biowaste Source on the Performance of Hard Carbons for

Sodium-Ion Batteries, ChemSusChem 10 (2017) 2668-2676.

[43] H. Yamamoto, S. Muratsubaki, K. Kubota, M. Fukunishi, H. Watanabe, J. Kim, S. Komaba,

Synthesizing higher-capacity hard-carbons from cellulose for Na- and K-ion batteries, J. Mater, Chem.

A 6 (2018) 16844-16848.

[44] N. Sun, H. Liu, B. Xu, Facile synthesis of high performance hard carbon anode materials for

sodium ion batteries, J. Mater, Chem. A 3 (2015) 20560-20566.

[45] E. Irisarri, A. Ponrouch, M.R. Palacin, Review—Hard Carbon Negative Electrode Materials for

Sodium-Ion Batteries, J. Electrochem. Soc. 162 (2015) A2476-A2482.

40
[46] H.-g. Wang, Z. Wu, F.-l. Meng, D.-l. Ma, X.-l. Huang, L.-m. Wang, X.-b. Zhang, Nitrogen-

Doped Porous Carbon Nanosheets as Low-Cost, High-Performance Anode Material for Sodium-Ion

Batteries, ChemSusChem 6 (2013) 56-60.

[47] W. Li, M. Zhou, H. Li, K. Wang, S. Cheng, K. Jiang, A high performance sulfur-doped

disordered carbon anode for sodium ion batteries, Energy Environ. Sci. 8 (2015) 2916-2921.

[48] L. Qie, W. Chen, X. Xiong, C. Hu, F. Zou, P. Hu, Y. Huang, Sulfur-Doped Carbon with Enlarged

Interlayer Distance as a High-Performance Anode Material for Sodium-Ion Batteries, Adv. Sci. 2

(2015) 1500195.

[49] P. Wang, B. Qiao, Y. Du, Y. Li, X. Zhou, Z. Dai, J. Bao, Fluorine-Doped Carbon Particles

Derived from Lotus Petioles as High-Performance Anode Materials for Sodium-Ion Batteries, J. Phys.

Chem. C 119 (2015) 21336-21344.

[50] J.W. Jeon, L. Zhang, J.L. Lutkenhaus, D.D. Laskar, J.P. Lemmon, D. Choi, M.I. Nandasiri, A.

Hashmi, J. Xu, R.K. Motkuri, C.A. Fernandez, J. Liu, M.P. Tucker, P.B. McGrail, B. Yang, S.K.

Nune, Controlling Porosity in Lignin‐Derived Nanoporous Carbon for Supercapacitor Applications,

ChemSusChem 8 (2015) 428-432.

[51] T. Yang, T. Qian, M. Wang, X. Shen, N. Xu, Z. Sun, C. Yan, A Sustainable Route from Biomass

Byproduct Okara to High Content Nitrogen-Doped Carbon Sheets for Efficient Sodium Ion Batteries,

Adv. Mater. 28 (2016) 539-545.

[52] S. Wang, L. Xia, L. Yu, L. Zhang, H. Wang, X.W. Lou, Free-Standing Nitrogen-Doped Carbon

Nanofiber Films: Integrated Electrodes for Sodium-Ion Batteries with Ultralong Cycle Life and

Superior Rate Capability, Adv. Energy Mater. 6 (2016) 1502217.

[53] G. Zou, C. Wang, H. Hou, C. Wang, X. Qiu, X. Ji, Controllable Interlayer Spacing of Sulfur-

Doped Graphitic Carbon Nanosheets for Fast Sodium-Ion Batteries, Small 13 (2017) 1700762.

41
[54] H. Hou, L. Shao, Y. Zhang, G. Zou, J. Chen, X. Ji, Large-Area Carbon Nanosheets Doped with

Phosphorus: A High-Performance Anode Material for Sodium-Ion Batteries, Adv. Sci. 4 (2017)

1600243.

[55] H. Tang, D. Yan, T. Lu, L. Pan, Sulfur-doped carbon spheres with hierarchical micro/mesopores

as anode materials for sodium-ion batteries, Electrochim. Acta 241 (2017) 63-72.

[56] X. Shi, Y. Chen, Y. Lai, K. Zhang, J. Li, Z. Zhang, Metal organic frameworks templated sulfur-

doped mesoporous carbons as anode materials for advanced sodium ion batteries, Carbon 123 (2017)

250-258.

[57] H.B. Wang, Q. Yu, J. Qu, Synthesis of phosphorus-doped soft carbon as anode materials for

lithium and sodium ion batteries, Russ. J. Phys. Chem. A 91 (2017) 1152-1155.

[58] M. Wang, Y. Yang, Z. Yang, L. Gu, Q. Chen, Y. Yu, Sodium-Ion Batteries: Improving the Rate

Capability of 3D Interconnected Carbon Nanofibers Thin Film by Boron, Nitrogen Dual-Doping, Adv.

Sci. 4 (2017) 1600468.

[59] J. Yang, X. Zhou, D. Wu, X. Zhao, Z. Zhou, S-Doped N-Rich Carbon Nanosheets with Expanded

Interlayer Distance as Anode Materials for Sodium-Ion Batteries, Adv. Mater. 29 (2017) 1604108.

[60] S. Huang, Z. Li, B. Wang, J. Zhang, Z. Peng, R. Qi, J. Wang, Y. Zhao, N-Doping and Defective

Nanographitic Domain Coupled Hard Carbon Nanoshells for High Performance Lithium/Sodium

Storage, Adv. Funct. Mater. 28 (2018) 1706294.

[61] N. Wang, Q. Liu, B. Sun, J. Gu, B. Yu, W. Zhang, D. Zhang, N-doped catalytic graphitized hard

carbon for high-performance lithium/sodium-ion batteries, Sci. Rep. 8 (2018) 9934.

[62] F. Wu, R. Dong, Y. Bai, Y. Li, G. Chen, Z. Wang, C. Wu, Phosphorus-Doped Hard Carbon

Nanofibers Prepared by Electrospinning as an Anode in Sodium-Ion Batteries, ACS Appl. Mater.

Interfaces. 10 (2018) 21335-21342.

42
[63] Y. Li, Y. Yuan, Y. Bai, Y. Liu, Z. Wang, L. Li, F. Wu, K. Amine, C. Wu, J. Lu, Insights into the

Na+ Storage Mechanism of Phosphorus-Functionalized Hard Carbon as Ultrahigh Capacity Anodes,

Adv. Energy Mater. 8 (2018) 1702781.

[64] Z. Hong, Y. Zhen, Y. Ruan, M. Kang, K. Zhou, J.-M. Zhang, Z. Huang, M. Wei, Rational Design

and General Synthesis of S-Doped Hard Carbon with Tunable Doping Sites toward Excellent Na-Ion

Storage Performance, Adv. Mater. 30 (2018) 1802035.

[65] K. Liao, H. Wang, L. Wang, D. Xu, M. Wu, R. Wang, B. He, Y. Gong, X. Hu, A high-energy

sodium-ion capacitor enabled by a nitrogen/sulfur co-doped hollow carbon nanofiber anode and an

activated carbon cathode, Nanoscale Adv. 1 (2019) 746-756.

[66] Z. Li, C. Bommier, Z.S. Chong, Z. Jian, T.W. Surta, X. Wang, Z. Xing, J.C. Neuefeind, W.F.

Stickle, M. Dolgos, P.A. Greaney, X. Ji, Mechanism of Na-Ion Storage in Hard Carbon Anodes

Revealed by Heteroatom Doping, Adv. Energy Mater. 7 (2017) 1602894.

[67] H. Liu, M. Jia, N. Sun, B. Cao, R. Chen, Q. Zhu, F. Wu, N. Qiao, B. Xu, Nitrogen-Rich

Mesoporous Carbon as Anode Material for High-Performance Sodium-Ion Batteries, ACS Appl.

Mater. Interfaces. 7 (2015) 27124-27130.

[68] Q. Wang, X. Ge, J. Xu, Y. Du, X. Zhao, L. Si, X. Zhou, Fabrication of Microporous Sulfur-

Doped Carbon Microtubes for High-Performance Sodium-Ion Batteries, ACS Appl. Energy Mater. 1

(2018) 6638-6645.

[69] X. Ge, S. Liu, M. Qiao, Y. Du, Y. Li, J. Bao, X. Zhou, Enabling Superior Electrochemical

Properties for Highly Efficient Potassium Storage by Impregnating Ultrafine Sb Nanocrystals within

Nanochannel-Containing Carbon Nanofibers, Angew. Chem. Int. Ed. 58 (2019) 14578−14583.

[70] Z. Jian, Z. Xing, C. Bommier, Z. Li, X. Ji, Hard Carbon Microspheres: Potassium-Ion Anode

Versus Sodium-Ion Anode, Adv. Energy Mater. 6 (2016) 1501874.

43
[71] D. Li, L. Zhang, H. Chen, L.-x. Ding, S. Wang, H. Wang, Nitrogen-doped bamboo-like carbon

nanotubes: promising anode materials for sodium-ion batteries, Chem. Commun. 51 (2015) 16045-

16048.

[72] Y. Chen, X. Li, K. Park, W. Lu, C. Wang, W. Xue, F. Yang, J. Zhou, L. Suo, T. Lin, H. Huang, J.

Li, J.B. Goodenough, Nitrogen-Doped Carbon for Sodium-Ion Battery Anode by Self-Etching and

Graphitization of Bimetallic MOF-Based Composite, Chem 3 (2017) 152-163.

[73] G. Xu, J. Han, B. Ding, P. Nie, J. Pan, H. Dou, H. Li, X. Zhang, Biomass-derived porous carbon

materials with sulfur and nitrogen dual-doping for energy storage, Green Chem. 17 (2015) 1668-1674.

[74] A. Riaz, C.S. Kim, Y. Kim, J. Kim, High-yield and high-calorific bio-oil production from

concentrated sulfuric acid hydrolysis lignin in supercritical ethanol, Fuel 172 (2016) 238-247.

[75] Z. Li, Y. Chen, Z. Jian, H. Jiang, J.J. Razink, W.F. Stickle, J.C. Neuefeind, X. Ji, Defective Hard

Carbon Anode for Na-Ion Batteries, Chem. Mater. 30 (2018) 4536-4542.

[76] J.P. Paraknowitsch, A. Thomas, J. Schmidt, Microporous sulfur-doped carbon from thienyl-based

polymer network precursors, Chem. Commun. 47 (2011) 8283-8285.

[77] S.-A. Wohlgemuth, R.J. White, M.-G. Willinger, M.-M. Titirici, M. Antonietti, A one-pot

hydrothermal synthesis of sulfur and nitrogen doped carbon aerogels with enhanced electrocatalytic

activity in the oxygen reduction reaction, Green Chem. 14 (2012) 1515-1523.

[78] X. Liu, L. Zhou, Y. Zhao, L. Bian, X. Feng, Q. Pu, Hollow, Spherical Nitrogen-Rich Porous

Carbon Shells Obtained from a Porous Organic Framework for the Supercapacitor, ACS Appl. Mater.

Interfaces. 5 (2013) 10280-10287.

[79] A.M. Puziy, O.I. Poddubnaya, A. Martínez-Alonso, F. Suárez-García, J.M.D. Tascón, Surface

chemistry of phosphorus-containing carbons of lignocellulosic origin, Carbon 43 (2005) 2857-2868.

[80] T.-H. Liou, Development of mesoporous structure and high adsorption capacity of biomass-based

activated carbon by phosphoric acid and zinc chloride activation, Chem. Eng. J. 158 (2010) 129-142.

44
[81] C. Wang, L. Sun, Y. Zhou, P. Wan, X. Zhang, J. Qiu, P/N co-doped microporous carbons from

H3PO4-doped polyaniline by in situ activation for supercapacitors, Carbon 59 (2013) 537-546.

[82] X.-F. Luo, C.-H. Yang, Y.-Y. Peng, N.-W. Pu, M.-D. Ger, C.-T. Hsieh, J.-K. Chang, Graphene

nanosheets, carbon nanotubes, graphite, and activated carbon as anode materials for sodium-ion

batteries, J. Mater, Chem. A 3 (2015) 10320-10326.

[83] Y. Liu, B.V. Merinov, W.A. Goddard, 3rd, Origin of low sodium capacity in graphite and

generally weak substrate binding of Na and Mg among alkali and alkaline earth metals, Proc. Natl.

Acad. Sci. U. S. A. 113 (2016) 3735-3739.

[84] J. Jin, B.-j. Yu, Z.-q. Shi, C.-y. Wang, C.-b. Chong, Lignin-based electrospun carbon nanofibrous

webs as free-standing and binder-free electrodes for sodium ion batteries, J. Power Sources 272 (2014)

800-807.

[85] R. Franke, T. Chassé, P. Streubel, A. Meisel, Auger parameters and relaxation energies of

phosphorus in solid compounds, J. Electron Spectrosc. Relat. Phenom. 56 (1991) 381-388.

[86] B. Jibril, O. Houache, R. Al-Maamari, B. Al-Rashidi, Effects of H3PO4 and KOH in

carbonization of lignocellulosic material, J. Anal. Appl. Pyrolysis 83 (2008) 151-156.

[87] G. Fagerlund, Determination of specific surface by the BET method, Mater. Struct. 6 (1973) 239-

245.

[88] A. Cuesta, P. Dhamelincourt, J. Laureyns, A. Martínez-Alonso, J.M.D. Tascón, Raman

microprobe studies on carbon materials, Carbon 32 (1994) 1523-1532.

[89] A. Sadezky, H. Muckenhuber, H. Grothe, R. Niessner, U. Pöschl, Raman microspectroscopy of

soot and related carbonaceous materials: Spectral analysis and structural information, Carbon 43

(2005) 1731-1742.

[90] J.D. Wilcox, M.M. Doeff, M. Marcinek, R. Kostecki, Factors Influencing the Quality of Carbon

Coatings on LiFePO4, J. Electrochem. Soc. 154 (2007) A389-A395.

45
[91] W. Luo, C. Bommier, Z. Jian, X. Li, R. Carter, S. Vail, Y. Lu, J.-J. Lee, X. Ji, Low-Surface-Area

Hard Carbon Anode for Na-Ion Batteries via Graphene Oxide as a Dehydration Agent, ACS Appl.

Mater. Interfaces. 7 (2015) 2626-2631.

[92] H. Hou, X. Qiu, W. Wei, Y. Zhang, X. Ji, Carbon Anode Materials for Advanced Sodium‐Ion

Batteries, Adv. Energy Mater. 7 (2017) 1602898.

[93] J. Jin, Z.-q. Shi, C.-y. Wang, Electrochemical Performance of Electrospun carbon nanofibers as

free-standing and binder-free anodes for Sodium-Ion and Lithium-Ion Batteries, Electrochim. Acta

141 (2014) 302-310.

[94] A. Ponrouch, A.R. Goñi, M.R. Palacín, High capacity hard carbon anodes for sodium ion

batteries in additive free electrolyte, Electrochem. Commun. 27 (2013) 85-88.

[95] L. Wang, C. Yang, S. Dou, S. Wang, J. Zhang, X. Gao, J. Ma, Y. Yu, Nitrogen-doped

hierarchically porous carbon networks: synthesis and applications in lithium-ion battery, sodium-ion

battery and zinc-air battery, Electrochim. Acta 219 (2016) 592-603.

Graphical Abstract (GA)

46
Highlights
 High-temperature carbonization with low-level heteroatom doping was developed.

 P-doped carbon exhibited enhanced low-voltage plateau capacity of 223 mAh g−1 in SIBs.

 High reversible capacity of 328 mAh g−1 with high ICE of 72% was achieved in SIBs.

 As an anode in KIBs, high reversible capacity of 302 mAh g−1 was achieved.

47

You might also like