You are on page 1of 14

batteries

Article
Facile Synthesis of Nickel Phosphide @ N-Doped Carbon
Nanorods with Exceptional Cycling Stability as Li-Ion and
Na-Ion Battery Anode Material
Fang Fu 1,† , Qiuchen He 1,† , Xuan Zhang 2 , Julian Key 1, *, Peikang Shen 1 and Jinliang Zhu 1, *

1 Collaborative Innovation Center of Sustainable Energy Materials, Guangxi Key Laboratory of Electrochemical
Energy Materials, Guangxi University, Nanning 530004, China; 2015301022@st.gxu.edu.cn (F.F.);
heqiuchen0817@163.com (Q.H.)
2 Gansu Yinguang Chemical Industry Ltd., Baiyin 730900, China; zhangxy666@foxmail.com
* Correspondence: joolskey@yahoo.com (J.K.); jlzhu@gxu.edu.cn (J.Z.)
† These authors contributed equally to this work.

Abstract: Nickel phosphide (Ni2 P), as an anode material for both lithium- and sodium-ion batteries,
offers high theoretical specific and volumetric capacities. However, considerable challenges include
its limited rate capability and low cycle stability arising from its volume change and degradation
during cycling. To solve these issues, appropriate composite micro/nanoparticle designs can improve
conductivity and provide confinement. Herein, we report a simple pyrolysis method to synthesize
nitrogen-doped carbon-coated Ni2 P nanorod arrays (Ni2 P@NC) from nickel foam and an ionic resin
as a source of carbon, nitrogen and phosphorus. The N-doped open-ended carbon shells provide
Ni2 P containment, good electrical conductivity, efficient electrolyte access and the buffering of bulk
strain during cycling. Consequently, as a LIB anode material, Ni2 P@NC has impressive specific
capacity in long-term cycling (630 mAh g−1 for 150 cycles at 0.1 A g−1 ) and a high rate capability
of 170 mAh g−1 for 6000 cycles at 5 A g−1 . Similarly, as a SIB anode, Ni2 P@NC retains a sizable
288 mAh g−1 over 300 cycles at 0.1 A g−1 , and 150 mAh g−1 over 2000 cycles at 2 A g−1 . Furthermore,
Citation: Fu, F.; He, Q.; Zhang, X.; due to a sizable portion of its capacity coinciding with adequately low voltage, the material shows
Key, J.; Shen, P.; Zhu, J. Facile promise for high volumetric energy storage in full-cell format. Lastly, the simple synthesis method
Synthesis of Nickel Phosphide @
has the potential to produce other carbon-coated metal phosphides for electrochemical applications.
N-Doped Carbon Nanorods with
Exceptional Cycling Stability as
Keywords: nickel phosphide; nanorod; cycling stability; lithium-ion battery; sodium-ion battery
Li-Ion and Na-Ion Battery Anode
Material. Batteries 2023, 9, 267.
https://doi.org/10.3390/
batteries9050267
1. Introduction
Academic Editor: Hirotoshi
Present-day commercial lithium-ion batteries (LIBs) and sodium-ion batteries (SIBs)
Yamada
employ graphitic carbon and hard carbon (HC) respectively as anode materials [1,2]. For
Received: 29 March 2023 practical applications, the overall merits of graphite and HC remain unmatched. This is
Revised: 27 April 2023 due mainly to their combination of charge storage at very low voltages (at an average of
Accepted: 8 May 2023 0.15 vs. Li+ /Li for graphite and approximately 0.45 V vs. Na+ /Na for HC) [3,4], moderate
Published: 11 May 2023 specific capacities (372 mAh g−1 for graphite and up to ~250 mAh g−1 HC), low material
costs and adequate rate capability and cycling stability [1,3,5]. However, both materials
are in fact quite low in specific and volumetric capacities, the latter owing to their low
densities of ~2.26 and 1.5 g cm−3 , respectively [6,7]. Moreover, the high specific energy
Copyright: © 2023 by the authors.
contribution of graphite and HC comes mainly from their low voltage profiles rather than
Licensee MDPI, Basel, Switzerland.
their storage capacity contribution. Therefore, from the perspective of increasing the specific
This article is an open access article
and volumetric energy of LIBs and SIBs, alternative anode materials with particularly high
distributed under the terms and
specific and volumetric capacities are of investigative interest, provided that an adequately
conditions of the Creative Commons
Attribution (CC BY) license (https://
low anode voltage is also achievable [8–11].
creativecommons.org/licenses/by/
Ni2 P has been widely researched as an electrocatalyst material but has only recently
4.0/). received attention as a LIB and SIB anode material [12–14]. As an anode in both the

Batteries 2023, 9, 267. https://doi.org/10.3390/batteries9050267 https://www.mdpi.com/journal/batteries


Batteries 2023, 9, 267 2 of 14

LIB and SIB cell formats, Ni2 P stores charge via the reversible conversion reaction of
Ni2 P + 3Li/Na→Li3 P/Na3 P + 2Ni, which, as a three-electron process, offers a high theoret-
ical specific capacity of 542 mAh g−1 [15]. It also has favorable thermal stability (compared
to low-voltage storage in carbons) owing to its higher average voltage profile, within which
a sizable charge storage capacity can be practically accessed within an average voltage of
~0.5 V vs. Li+ /Li and Na+ /Na [15]. Furthermore, as a transition-metal phosphide with a
high crystal density of 7.55 g cm−3 [16], Ni2 P has almost three times the crystal density
of graphite and over five times that of HC [6,7], which is conducive to reaching higher
volumetric energy storage in LIBs and SIBs.
However, typical of all conversion reaction materials, Ni2 P undergoes large volume
changes during cycling that cause phase decomposition and resultant kinetic problems of
poor electron transfer conductivity [17,18]. Solutions to these problems include the particle
design of appropriate micro- and nanostructures and the use of carbon as a protective layer
to contain material breakage during cycling [19,20]. For example, Park’s team successfully
prepared amorphous and crystalline Ni2 P and self-assembled Ni2 P nanoparticles by ther-
mal injection using an Ar atmosphere and standard Schlenk-line technology. By varying the
quantity of the nickel acetylacetonate reactant, Ni2 P nanoparticles were produced with com-
parable shapes and sizes but with different crystalline states. Notably, the amorphous Ni2 P
had greatly improved initial coulombic efficiency (ICE) and capacity retention, achieving a
reversible capacity of 573.2 mA h g−1 at 0.5 C and, in successive cycles, a superb CE of 99.1%
in LIBs [21]. With the use of Ni2 P as a SIB anode material, Yin et al. successfully prepared
Ni2 P@Carbon/graphene with three-dimensionally interconnected porous structures. The
synergistic interaction between carbon layers and Ni2 P nanoparticles provided structural
stability during cycling, maintaining a specific capacity of 124.5 mA h g−1 at 1 A g−1 for
2000 cycles. Such a synthesis approach combines solvothermal reactions and in situ phos-
phorylation procedures [22]. Overall, the carbon covering prevented the nanoparticles
from clumping together, while it also improved the electron transfer rate and provided
containment for volume changes.
The reported morphologies of synthesized Ni2 P in the literature include nanoparticles,
hollow spheres, nanosheets and nanobelts [12,23–28]. The two main methods commonly
used to synthesize Ni2 P are (1) the solvothermal/hydrothermal method (where it is not easy
to manipulate the morphology or to remove the excess solvent [29,30]), and (2) chemical
vapor deposition using hypophosphite as a phosphorus source with prepared Ni/NiO
precursors of the designed morphology as the nickel source. However, in the latter method,
the reaction releases toxic and flammable PH3 gas and is quite labor-intensive (Table S1).
Therefore, achieving an effective Ni2 P architecture via a convenient method with low
toxicity remains an important goal for its realization as an anode material.
Herein, we report a simple and safe pyrolysis method to synthesize nitrogen-doped
carbon-coated (NC) Ni2 P nanorods with structural features that enhance Ni2 P perfor-
mance as a LIB and SIB anode. The synthesis method involves the pyrolysis of a novel
nitrogen- and phosphorous-containing resin in the presence of nickel foam. The phosphati-
zation growth/carbon deposition reaction occurs on the nickel metal surface to form self-
assembling superstructure arrays comprising composite 100 nm diameter Ni2 P nanorods
that develop a carbon shell throughout the synthesis process. The unique nanostructure and
composition of Ni2 P@NC were determined by various physical characterization methods,
followed by electrochemical testing. In both high-rate and long-term cycling in half-cell for-
mat, Ni2 P@NC had excellent performance, which can be attributed to its unique structure.
For a comparison, a number of controls were also tested, which included a synthesized
Ni2 P, NC, and Ni2 P/NC mixture as well as commercial Ni2 P (see Materials and Meth-
ods). Finally, the possible specific and volumetric energy contributions the material might
facilitate in full-cell formats was calculated.
Batteries 2023, 9, 267 3 of 14

2. Materials and Methods


Ni2 P@NC was synthesized as follows: 2.5 g of a purchased nitrogen- and phosphorus-
containing resin (Sunresin Materials Co., Ltd., Zhejiang, China) was first placed in a quartz
boat to form a uniform depth layer. A nickel foam sheet (6 × 3 × 0.15 cm) was then
positioned on top of the resin, and the assembly was heated for 60 min in a tube furnace at
1000 ◦ C under a N2 atmosphere at 10 ◦ C min−1 . After cooling to room temperature, the
Ni2 P@NC products were detached from the nickel foam surface via dry sonication. ICP
analysis revealed the Ni2 P content to be approximately 85 wt.%. Nitrogen-doped carbon
(NC) was made using the above procedure but with the omission of nickel foam. To make
nickel phosphide nanopowder or micropowder (Ni2 P), red phosphorus and nickel powder
were heated at 1000 ◦ C for 2 h. Ni2 P/NC was made by combining Ni2 P and NC in a ball
mill for two hours at a weight ratio of 85:15. Commercial Ni2 P (C-Ni2 P) was purchased
from (Aldrich Chemical Corporation, Milwaukee, WI, 53201, USA).
Crystal structure determination was carried out using X-ray diffractometry (XRD)
on a D/Max-III XRD diffractometer (Rigaku Co., Cu Kα, Akishima City, Tokyo, Japan).
Structural, compositional and particle size analysis employed scanning electron microscopy
(SEM) on a FESEM SU8220 (Hitachi Corp., Tokyo, Japan), and transmission electron mi-
croscopy (TEM) was performed using a Titan ETEM G2 80–300 (FEI Co., Hillsboro, OR,
USA). Surface-state and chemical environment analyses were carried out using X-ray
photoelectron spectroscopy (XPS) with Al K radiation on an ESCALAB 250 (Thermo-VG
Scientific, New York, NY, USA). Pore structure characterization and specific surface area
were determined by N2 adsorption–desorption measurements (Micromeritics, ASAP 2420
instrument, Atlanta, GA, USA). For this purpose, a 100 mg sample was placed in a 77 K
quartz glass tube and degassed in a Dewar bottle cooled with sufficient nitrogen, after
which the sample was transferred to the analytical station of the adsorber for adsorption
testing. The specific surface area and pore size distribution information of the material
were calculated using the Barrett, Joyner and Halenda (BJH) method.
Electrodes were made using the slurry cast method, with slurries comprising 80 wt.%
active materials (either Ni2 P@NC, Ni2 P, Ni2 P/NC, NC or C-Ni2 P), 10 wt.% acetylene black,
10 wt.% carboxymethyl cellulose and deionized water. The slurries were cast onto copper
foil followed by drying at 85 ◦ C for 12 h and cut into 14 mm diameter electrode discs. The
active material mass loading for each electrode was approximately 2.1 mg (when testing
the CV curve, the cell had an active material load of approximately 0.6 g). Half cells were
constructed with Li or Na metal foil counter electrodes, and glass microfiber separators
(Whatman GF/D) were used with an electrolyte of 1 M LiPF6 dissolved in EC/DMC/DEC
(1:1:1 (v:v:v)) or 1 M NaClO4 dissolved in PC/EC (1:1 (v:v)) with 5% fluoroethylene carbon-
ate (FEC) added, respectively. Cyclic voltammetry (CV) and electrochemical impedance
spectroscopy (EIS) were carried out on an electrochemical instrument (IM6, Zahner-Elektrik,
Kronach, Germany). Charge/discharge cell cycling was carried on the Neware cell test
system (Neware Battery, Shenzhen, China) set at a potential range of 0.01 to 3 V.

3. Results and Discussion


Figure 1 shows the XRD patterns of Ni2 P@NC, Ni2 P/NC, Ni2 P and NC powders. All
samples containing Ni2 P produced sharp 2θ diffraction peaks at 40.7, 44.6, 47.4, 54.2, 55.0
and 74.7◦ , relating, respectively, to the (111), (201), (210), (300), (211) and (400) crystal planes
of hexagonal Ni2 P (PDF#65-3544). Ni2 P@NC and Ni2 P/NC patterns show a poorly defined
peak at around 26◦ , analogous to the strongly pronounced peak seen in the NC sample.
SEM imaging of the Ni2 P@NC sample under low magnification revealed a uniform
growth distribution of Ni2 P@NC arrays formed on the nickel foam substrate. Under
high magnification, the arrays can be seen to consist of nanorods with a consistent width
and length, with no entanglement or agglomeration (Figure 2A–C). In contrast, Ni2 P
(Figure S2) formed aggregated structures containing a large number of nanoparticles. In
the Ni2 P/NC mixture (Figure S3), Ni2 P and NC particles appeared to be freely mixed
with disorderly contact, providing only the smallest stabilized structure. Furthermore,
Batteries 2023, 9, 267 4 of 14

the presence of irregular holes in NC suggests the loss of phosphorus during pyrolysis
(Figure S4). TEM imaging of an individual Ni2 P@NC nanorod (Figure 2D) shows that
the nanorods are roughly 100 nm in diameter and that carbon layers wrap uniformly
around them, and the corresponding Fourier transform image (inset) also confirms good
single-crystal characteristics. High-resolution TEM (HRTEM) shows that the carbon layer is
~12 nm in thickness, and the Ni2 P lattice spacing of 5.05 Å corresponds to the (100) crystal
plane (Figure 2E,F). In the sonicated samples (Figure 2D), the nanorods can be seen to have
open tips caused by their detachment from the nickel foam. The open-tip structure should
facilitate good electrolyte access to the battery material, and Figure 2E,F show areas of
close contact between the carbon and Ni2 P, as well as some void non-contact areas. These
features could, respectively, provide buffering expansion space and support good electrical
contact during cycling. Finally, elemental mapping (Figure 2G–L) shows a broad dispersion
Batteries 2023, 9, x FOR PEER REVIEW 4 of 16
of the four elements in a single Ni2 P@NC nanorod and a uniform dispersion of N and P
elements throughout the carbon layer (which may also promote electrochemical capacity).

Figure 1. XRD patterns of Ni22P@NC,


P@NC, Ni
Ni22P/NC,
P/NC,Ni2P P
Ni2 and
andNC.
NC.

Figure 3A shows
SEM imaging N22P@NC
of the Ni adsorption–desorption
sample under low curve for Ni2 P@NC,
magnification whichaisuniform
revealed clearly
characterized by a hysteresis
growth distribution of Ni2P@NC looparrays
and type IV curve
formed on the[31]. Thefoam
nickel tests substrate.
reveal thatUnder
Ni2 P@NC
high
has a specific surface area of 272.3 m 2 g−1 according to the Brunauer–Emmett–Teller (BET)
magnification, the arrays can be seen to consist of nanorods with a consistent width and
analysis, and no
length, with it isentanglement
clear from theor pore size distribution
agglomeration curve
(Figure that the
2A–C). In sample
contrast,predominantly
Ni2P (Figure
contains
S2) formed mesopores.
aggregated Thestructures
specific surface areas of
containing a Ni 2 P, Ni
large 2 P/NCof
number and NC were 18.9,
nanoparticles. In32.5
the
and
Ni2P/NC m2 g−1 , (Figure
308.9 mixture respectively (Figures
S3), Ni 2P andS5–S7), showing
NC particles an increasing
appeared trend with
to be freely mixed carbon
with
content.
disorderlyNotably,
contact, theproviding
NC sample hasthe
only a high specific
smallest surface area
stabilized resulting
structure. from both the
Furthermore, the
absence of Ni P and the formation of additional pores due to the removal
presence of irregular holes in NC suggests the loss of phosphorus during pyrolysis (Fig-
2 of phosphorus
from the TEM
ure S4). carbon during of
imaging pyrolysis.
an individual Ni2P@NC nanorod (Figure 2D) shows that the
nanorods are roughly 100 nm inand
The elemental composition bond types
diameter of Ni
and that 2 P@NC
carbon werewrap
layers determined
uniformlyusing XPS.
around
Figure 3B shows the detection of four major elements in the broad-scan
them, and the corresponding Fourier transform image (inset) also confirms good single- spectra. The Ni
2p spectra
crystal (Figure 3C) High-resolution
characteristics. show six clear peaksTEMassigned
(HRTEM)toshows Ni 2p3/2 that(853.2 eV) and
the carbon Ni is
layer 2p~12
1/2
(869.9 eV) in Ni-P; Ni 2p (856.9 eV) and Ni 2p (874.9 eV) in
nm in thickness, and the Ni2P lattice spacing of 5.05 Å corresponds to the (100) crystal
3/2 1/2 Ni-O; and also two
satellite peaks 2E,F).
plane (Figure at 863.1In eV
the and 880.7 eV
sonicated [8,32].(Figure
samples The P 2p 2D),spectra (Figure 3D)
the nanorods reveal
can be seenthe
to
bond types of P-Ni (129.6 eV), P-C (133.3 eV) and P-O (134.2 eV) [33,34].
have open tips caused by their detachment from the nickel foam. The open-tip structure The C 1s spectra
(Figure
should 3E) revealgood
facilitate C-C (284.6 eV), C-P
electrolyte (283.6
access eV),battery
to the C-N (285.7 eV), C-O
material, and(286.4
FigureeV) andshow
2E,F O-C=Oar-
(288.9 eV) [35,36]. The N 1s spectra (Figure 3F) reveal pyridinic N (398.9
eas of close contact between the carbon and Ni2P, as well as some void non-contact areas. eV), pyrrolic N
(399.9 eV) and graphitic N (401.1 eV) [37,38], which indicates the N-doping of carbon. The
These features could, respectively, provide buffering expansion space and support good
detection of C-P- and C-N-type bonds indicates the N and P doping of carbon, both of
electrical contact during cycling. Finally, elemental mapping (Figure 2G–L) shows a
which are known to have donor/acceptor properties, resulting in expanded lattice spacing
broad dispersion of the four elements in a single Ni2P@NC nanorod and a uniform dis-
and the formation of external defects, which can enhance Li+ /Na+ adsorption/desorption,
persion of N and P elements throughout the carbon layer (which may also promote elec-
storage capacity, cycling stabilization and rate performance [35,39,40].
trochemical capacity).
Batteries 2023,9,9,267
Batteries2023, x FOR PEER REVIEW 55ofof14
16

Figure2.2. SEM
Figure SEM (A–C)
(A–C) and
and TEM
TEM (D–F)
(D–F) images
images of
of Ni
Ni2P@NC
P@NC at
at different
different magnifications.
magnifications. (G–L)
(G–L) Ele-
El-
2
mental mapping images of Ni2P@NC in regard to the original TEM image showing elemental dis-
emental mapping images of Ni2 P@NC in regard to the original TEM image showing elemen-
tributions.
tal distributions.

FigureS1
Figure 3Ashows
showsSEM the images
N2 adsorption–desorption
of the timeline of the curve for Ni2P@NC,
progression which nanorod
of Ni2 P@NC is clearly
characterized by a hysteresis loop and type IV curve [31]. The tests
growth from the Ni foam surface. After 5 min, the surface of the Ni foam appears2P@NC reveal that Ni quite
has a (Figure
rough specific S1C),
surface area under
which, of 272.3 m2 gmagnification,
higher
−1 according to the Brunauer–Emmett–Teller
reveals the presence of short,
(BET) analysis,
finger-like and it is
projections of clear from thenanorods
the budding pore size (Figures
distributionS1Dcurve that the
and S8). Thesample pre-
core–shell
dominantly contains mesopores. The specific surface areas of Ni
growth pattern suggests that carbon (present in the gaseous hydrocarbons arising from 2 P, Ni 2P/NC and NC
were 18.9, 32.5 and 308.9 m 2 g−1, respectively (Figures S5–S7), showing an increasing
the pyrolyzed resin) first dissolves in the molten Ni2 P and is then deposited as a carbon
trendonwith
shell carbon
the Ni content.
2 P surface onceNotably, the NC
its saturation sample
point has a high
is reached. specific
Similar surface area re-
“vapor-liquid-solid”
sultingmechanisms
growth from both the absence
have of Ni2P and
been proposed the formation
for carbon deposition of on
additional pores due
transition-metal to the
carbides
removal
and metalofcarbides
phosphorus from
[41,42]. the carbon
Rod-like during
Ni2 P growth pyrolysis.
with carbon shell deposition is likely
The elemental
to involve composition
capillary forces that drawandmolten
bond types
nickel of
upNi the2P@NC were the
rods while determined using
lower-density
XPS. Figure
saturated 3B shows
carbon the detection
is pushed of four
to the surface major
of the rodselements
(FiguresinS1E,F
the broad-scan
and S9) to spectra.
obtain
The Ni 2p spectra
carbon-coated Ni2 P(Figure 3C) show
nanorods (FiguresixS1H).
clear Upon
peaks assigned
cooling from to Nithe
2p3/2 (853.2toeV)
liquid theand Ni
solid
2p1/2 (869.9
state, eV) generated
voids are in Ni-P; Nibetween
2p3/2 (856.9 eV) andshell
the carbon Ni 2p 1/2 (874.9
and eV) in
Ni2 P core due Ni-O; anddifferent
to their also two
cooling
satellitecoefficients.
peaks at 863.1 eV and 880.7 eV [8,32]. The P 2p spectra (Figure 3D) reveal the
bond types of P-Ni (129.6 eV), P-C (133.3 eV) and P-O (134.2 eV) [33,34]. The C 1s spectra
(Figure 3E) reveal C-C (284.6 eV), C-P (283.6 eV), C-N (285.7 eV), C-O (286.4 eV) and O-
C=O (288.9 eV) [35,36]. The N 1s spectra (Figure 3F) reveal pyridinic N (398.9 eV), pyr-
rolic N (399.9 eV) and graphitic N (401.1 eV) [37,38], which indicates the N-doping of
carbon. The detection of C-P- and C-N-type bonds indicates the N and P doping of car-
bon, both of which are known to have donor/acceptor properties, resulting in expanded
Batteries 2023, 9, 267 6 of 14
lattice spacing and the formation of external defects, which can enhance Li+/Na+ adsorp-
tion/desorption, storage capacity, cycling stabilization and rate performance [35,39,40].

Figure 3.
Figure 3. (A)
(A)Nitrogen
Nitrogen adsorption–desorption isotherms
adsorption–desorption and the
isotherms andpore
thesize distribution
pore of Ni2 P@NC.
size distribution of
(B)2P@NC.
Ni XPS spectra of Ni
(B) XPS P@NC.
spectra
2 of High-resolution
Ni 2 P@NC. XPS spectra
High-resolution and
XPS the fitting
spectra and results
the of (C)
fitting Ni 2p,
results (D)
of P
(C)
Ni
2p,2p,
(E)(D)
C 1sP and
2p, (E)
(F)CN1s1s.and (F) N 1s.

Galvanostatic
Figure S1 shows discharge/charge
SEM images of the cycling curves
timeline of Ni
of the 2 P@NC at of
progression 0.1Ni g−1 asnano-
A2P@NC a LIB
anode
rod (in half-cell
growth from theformat
Ni foamagainst
surface.Li metal)
After 5are shown
min, in Figure
the surface 4A.Ni
of the While
foamthe initial
appears
discharge
quite rough capacity
(Figure wasS1C), mAh g−
1058which, 1 and the initial charge capacity was 690 mAh g−1 (i.e.,
under higher magnification, reveals the presence of
65.28%finger-like
short, of ICE), a high CE wasof
projections quickly reached nanorods
the budding by the 2nd(Figures
and 3rd cycles
S1D and thatS8).
remained well
The core–
above 90% by the 150th cycle at 630 mAh g −1 . Figure 4B shows the CVs of Ni P@NC in the
shell growth pattern suggests that carbon (present in the gaseous hydrocarbons 2 arising
LIB electrolyte
from the pyrolyzed over resin)
the first fivedissolves
first cycles at in 0.2the s−1 . During
mVmolten Ni2P and cathodic
is then scanning,
deposited theas
first
a
scan shows an extended cathodic peak relating to SEI film formation,
carbon shell on the Ni2P surface once its saturation point is reached. Similar “vapor- which, by the 2nd
to 5th cycle, was
liquid-solid” completely
growth mechanismsgone, haveshowingbeenthat such anfor
proposed irreversible capacity component
carbon deposition on transi-
tion-metal carbides and metal carbides [41,42]. Rod-like Ni2P growthVwith
is restricted to the 1st cycle only. The anodic peaks at 1.08 V and 2.37 correspond
carbon to the
shell
oxidation of Ni to NiP and Ni
deposition is likely to involve capillary 2 P, respectively, and consistently appeared
forces that draw molten nickel up the rods while in subsequent
scans
the [43,44]. In contrast,
lower-density Ni2 carbon
saturated P powder is (Figure
pushedS10) in the
to the second
surface of and
the third
rods cycles
(Figures exhibited
S1E,F
a relatively destabilized anomalous curve
and S9) to obtain carbon-coated Ni2P nanorods (Figure S1H). shape. To further examine the electrochemical
Upon cooling from the liq-
behavior
uid to the andsolidstability of Niare
state,−voids 2 P@NC,
generatedCV atbetween
0.1 to 2 the s−1 was
mVcarbon carried
shell and outNi2Pafter
coreSEIduefilm
to
formation at 0.2 mV s 1 (shown earlier in Figure 4C). The subsequent curves over increasing
their different cooling coefficients.
rates of voltage change remained largely unchanged, indicating that Ni2 P@NC has a stable
Galvanostatic discharge/charge cycling curves of Ni2P@NC at 0.1 A g−1 as a LIB an-
reversible capacity as a LIB anode.
ode (in half-cell format against Li metal) are shown in Figure 4A. While the initial dis-
The rate capabilities of Ni2 P@NC as a LIB anode at 0.1 and 10 A g−1 (Figure 4D) were
charge capacity was−1058 mAh g and the initial charge capacity was 690 mAh g−1 (i.e.,
−1
750 and 130 mAh g 1 , respectively, and recovered a high capacity of 670 mAh g−1 when
65.28% of ICE), a high CE was quickly reached by the 2nd and 3rd cycles that remained
returned to 0.1 A g−1 , i.e., resuming at 89.3% retention of the original value. In comparison,
well above 90% by the 150th cycle at 630 mAh g−1. Figure 4B shows the−CVs of Ni2P@NC
the rate capabilities of the control materials (Figure 4D) at 0.1−1and 10 A g 1 were as follows:
in the LIB electrolyte over the first five cycles at 0.2 mV s . During cathodic scanning,
Ni2 P (407 and 13 mAh g−1 ); NC (267 and 35 mAh g−1 ); Ni2 P/NC (330 and 14 mAh g−1 );
the first scan shows an extended cathodic peak relating to SEI film formation, which, by
and the C-Ni2 P control (414 and 65 mAh g−1 , Figure S11). Therefore, Ni2 P@NC clearly
the 2nd to 5th cycle, was completely gone, showing that such an irreversible capacity
had a notably higher rate capability than the four control materials. Figure 4E shows the
component is restricted to the 1st cycle only. The anodic peaks at 1.08 V and 2.37 V cor-
long-term cycling tests of Ni2 P@NC as a LIB anode. Here, the initial first cycle’s irreversible
respond
capacityto thedue
loss oxidation
to the of Ni to NiP
formation of and
the SEINi2P,film
respectively,
is included andinconsistently
the plot [45]. appeared
During
in subsequent scans [43,44]. In contrast, Ni
subsequent charge/discharge cycles, the CE holds at >90% at 630 mAh g . Notably,and
2P powder (Figure S10) in the − 1 second this
third
valuecycles exhibited
considerably a relatively
exceeds the 542 destabilized anomalous
mAh g−1 theoretical curve shape.
capacity of Ni2 PToand further exam-
the isolated
ine
NCthe electrochemical
sample. However,behavior and stability
excess theoretical of Ni2P@NC,
capacities for Ni CVPatare 0.1commonly
to 2 mV s−1 reported,
was car-
2
particularly with combinations of carbon (see Table S2). In the present case, it is possible
that the N-doped carbon shell is able to greatly boost the ion storage capacity of Ni2 P and
that interfacial charge storage and the additional contribution of pseudo-capacitance also
result in an increase in the material’s capacity [41,46]. Impressively, Ni2 P@NC retains a
stabilized capacity of >600 mAh g−1 after 150 cycles, unlike the three controls Ni2 P, NC and
Ni2 P/NC, which dropped to 75.5, 103.7 and 84.9 mAh g−1 , respectively, by the 70th cycle.
Furthermore, Ni2 P@NC cycled at 5 A g−1 (Figure 4F) was also very stable, dropping from
Batteries 2023, 9, 267 7 of 14

226 to 170 mAh g−1 in merely 6000 cycles, equating to a 0.00413% average capacity drop per
cycle. Clearly, the unique structure and composition of Ni2 P@NC are conducive to steady
long-term cycling and high-rate cycling. Table S2 shows a significant increase in cycling
stabilization and the retention of the relative capacity of Ni2 P@NC over the previously
reported Ni2 P-based materials as LIB anodes. The particularly high capacity retention
Batteries 2023, 9, x FOR PEER REVIEW 8 of 16
of our Ni2 P@NC composites is evidently associated with the appropriate nanostructure
combined with a novel nanorod morphology.

Figure
Figure4.4.(A) Discharge–charge
(A) Discharge–chargecurves of Ni2P@NC
curves of Ni2against
P@NC Li metal in
against LiLIB electrolyte
metal in LIBatelectrolyte
0.1 A g−1. at 0.1 A
(B) Cyclic voltammetry (CV) curves of Ni 2P@NC. (C) CVs of Ni2P@NC at different scan rates. (D)
g−1 . (B) Cyclic voltammetry (CV) curves of Ni2 P@NC.
Rate testing between 0.1 and 10 A g−1. (E) Long-term cycling
(C) CVs of Ni P@NC at different scan
performance at 0.12 A g−1. (F) Long-
rates.cycling
(D) Rate testing between 0.1 at
and − 1 −1
term performance of Ni2P@NC 5 A10
g−1A
. g . (E) Long-term cycling performance at 0.1 A g .
(F) Long-term cycling performance of Ni2 P@NC at 5 A g−1 .
Figure 5A shows the galvanostatic discharge/charge cycling curves of Ni2P@NC at
0.1 A Figure 5Aashows
g−1 against thecounter
Na metal galvanostatic
electrodedischarge/charge cycling
in SIB half-cell format. curves
Ni2P@NC of Ni2 P@NC at
delivered
555.5 g−1 against
0.1 AmAh a Na
g−1 in the firstmetal counter
discharge and electrode
343.2 mAhingSIB−1 in half-cell format.for
the first charge Nian
2 ICE ofdelivered
P@NC
555.5 mAh
61.8%. −1 in the capacity
A highgreversible first discharge andg343.2
of 290 mAh −1 wasmAh g−1 inwith
maintained the nearly
first charge
100% CEforatan ICE of
the 300th cycle. Figure 5B shows CV curves obtained− 1at 0.2 mV s−1. In the cathodic scan,
61.8%. A high reversible capacity of 290 mAh g was maintained with nearly 100% CE at
the
the intense reduction
300th cycle. peak
Figure 5Batshows
1.0 V CVcorresponds to the formation
curves obtained at 0.2 mV −13.P.InInthe
of sNa thecathodic
first scan,
scan, between 0.01 and 1.5 V, the formation of the SEI film generated a higher current,
while in the following four scans, the tight overlap of the curves occurred. In the anodic
scans, peaks at 2.0 V correspond to the oxidation of Ni to Ni2P, and subsequent scans
Batteries 2023, 9, 267 8 of 14

the intense reduction peak at 1.0 V corresponds to the formation of Na3 P. In the first scan,
between 0.01 and 1.5 V, the formation of the SEI film generated a higher current, while
in the following four scans, the tight overlap of the curves occurred. In the anodic scans,
peaks at 2.0 V correspond to the oxidation of Ni to Ni2 P, and subsequent scans overlap
well. In contrast, Ni2 P powder (Figure S12) produced less overlap in the second and
third cycles, indicating lower stability. Over sweep rates of 0.1 to 1.5 mV s−1 (Figure 5C),
Ni2 P@NC generated constant curve shapes (showing little deviation in reduction–oxidation
peak locations), which indicates that the material facilitates rapid ion mobility. Moreover,
Ni2 P@NC as a SIB anode had a high rate performance between 0.1 and 5 A g−1 (Figure 5D)
of 343.2 and 115 mAh g−1 , respectively. On return to 0.1 A g−1 , the reversible capacity of
280.3 mAh g−1 was restored to 81.7%, indicating excellent rate performance. The percentage
recovery on return to 0.1 A g−1 followed the order NC > Ni2 P@NC > C-Ni2 P > Ni2 P/NC >
Ni2 P, which reveals the stability of NC and its stabilizing effect on cycling in both Ni2 P@NC
and Ni2 P/NC. Notably, as with LIB cycling, Ni2 P@NC as a SIB anode produced the highest
capacity at all rates tested, remaining 290 mAh g−1 over 300 cycles at 0.1 A g−1 , compared
to values < 100 mAh g−1 for the other control materials after 150 cycles (Figure 5E). The
related electrochemical properties of C-Ni2 P are shown in Figure S13. The long-term cycling
of Ni2 P@NC at 2 A g−1 (Figure 5F) had a minimal capacity loss from the 1st to 300th cycles
at 179.6 to 151 mAh g−1 , equating to a 0.008% loss per cycle. Therefore, Ni2 P@NC has
outstanding stable cycling performance as a SIB anode material at a high current density.
Table S3 compares the cycling performance of Ni2 P@NC in this study to earlier studies
on Ni2 P and Ni2 P-carbon combinations as a SIB anode. Similar to its behavior as a LIB
anode, Ni2 P@NC as a SIB anode demonstrates exceptional cycling stability and a seemingly
unmatched rate capability. This is clearly facilitated by the unique carbon-coated nanorod
structure that enhances electrical conductivity and shortens the ion transport pathway.
Electrochemical impedance spectroscopy (EIS) revealed that the samples had an order
of increasing electrode/electrolyte transfer resistance of Ni2 P > Ni2 P/NC > Ni2 P@NC
> NC (Figure S14). This suggests that close carbon contact between the NC shell and
Ni2 P in Ni2 P@NC is synergistic to charge transfer. This, in conjunction with its open
tubular morphology, appears to account for its improved cycling stability and higher rate
capability over those of Ni2 P and Ni2 P/NC. More specifically, the collective advantages of
the Ni2 P@NC structure appear to include (1) the buffering of volume expansion via carbon
shells that are open-ended with void spaces between the core–shell contact; (2) the N- and
P-doped widening of pores in carbon, allowing rapid ion movement and high capacity;
(3) a nanorod structure, facilitating fast ion and electron transfer.
Finally, we consider the possible energy storage contribution of Ni2 P@NC as a LIB and
SIB anode material. Of notable concern is the high-voltage slope tail region in the Ni2 P@NC
charge curve profiles (Figures 4A and 5A), which appears detrimental to energy storage.
However, it is also evident, owing to its high specific capacity, that a significant capacity still
resides in the lower-voltage region of the curves. Therefore, it was of interest to determine
whether restricting cycling to this region could theoretically produce higher cell specific
energy than using its full capacity capability. The calculation process used is described in
Section S2 of the SI (“Energy calculations for Ni2 P@NC as a LIB and SIB anode material”) in
Figures S15–S17. Firstly, the average voltages preceding all possible voltage/charge cutoff
points were calculated and plotted against the corresponding capacity–mass balancing
of a typical LIB or SIB cathode for the range of all possible anode capacity cutoff points
(Figures S15 and S16). This produced the range of all theoretical full-cell specific capacities,
which were then multiplied by the corresponding average cell voltages to obtain the cell
specific energy. Plotting the full range of specific energy data points revealed maxima
(Figure S17). Notably, the maxima positions do not correspond to the use of the full-
anode capacity at the 3 V cutoff. For example, the theoretical specific energy maximum
(active material only) for Ni2 P@NC in full-cell LIB format as a Ni2 P@NC/NMC 333 cell
was 435.4 Wh kg−1 (Figure 6). This resulted from an anode-to-cathode mass balance
ratio of 1:2.19 and a full-cell cutoff at 2.54 V. Here, the corresponding anode specific
Batteries 2023, 9, 267 9 of 14

capacity cuts off at 438.4 mAh g−1 and 1.2 V vs. Li+ /Li at an average anode voltage
of 0.63 V vs. Li+ /Li. However, the calculated specific energy, unfortunately, still falls
8.3% short of graphite/NMC 333 active materials at an optimal mass balance of 1:1.86.
Moreover, the result for the equivalent SIB full cell, compared to a hard carbon (HC)
combination with a high-capacity cathode (Table S5), was even lower at ~24%
Batteries 2023, 9, x FOR PEER REVIEW short
10 of 16 of the
hard carbon equivalent.

Figure 5. (A)
Figure (A) Discharge–charge
Discharge–charge curves
curves of Ni
of2P@NC against
Ni2 P@NC Na metal
against in SIB electrolyte
Na metal at−1.0.1 A g−1 .
at 0.1 A g
in SIB electrolyte
(B) Cyclic voltammetry (CV) curves of Ni 2P@NC. (C) CVs of Ni2P@NC at different scan rates. (D)
(B) Cyclic voltammetry (CV) curves of Ni2 P@NC. (C) CVs of Ni2 P@NC at different scan rates. (D) Rate
Rate testing between 0.1 and 5 A g−1. (E) Long-term cycling performance at 0.1 A g−1. (F) Long-term
testing between 0.1 and 5 A g−1 . (E) Long-term cycling performance at 0.1 A g−1 . (F) Long-term
cycling performance of Ni2P@NC at 2 A g−1.
cycling performance of Ni2 P@NC at 2 A g . − 1

Electrochemical impedance spectroscopy (EIS) revealed that the samples had an or-
der of increasing electrode/electrolyte transfer resistance of Ni2P > Ni2P/NC > Ni2P@NC >
NC (Figure S14). This suggests that close carbon contact between the NC shell and Ni2P
in Ni2P@NC is synergistic to charge transfer. This, in conjunction with its open tubular
morphology, appears to account for its improved cycling stability and higher rate capa-
anode-to-cathode mass balance ratio of 1:2.19 and a full-cell cutoff at 2.54 V. Here, the
corresponding anode specific capacity cuts off at 438.4 mAh g−1 and 1.2 V vs. Li+/Li at an
average anode voltage of 0.63 V vs. Li+/Li. However, the calculated specific energy, unfor-
tunately, still falls 8.3% short of graphite/NMC 333 active materials at an optimal mass
balance of 1:1.86. Moreover, the result for the equivalent SIB full cell, compared to a hard
Batteries 2023, 9, 267 carbon (HC) combination with a high-capacity cathode (Table S5), was even lower 10 ofat
14
~24% short of the hard carbon equivalent.

(A) (B)
Figure
Figure6.6.Calculation
Calculationofofthe
thetheoretical
theoreticalspecific
specificand
and volumetric
volumetricenergy
energyachievable with
achievable Ni2Ni
with P@NC
2 P@NC
vs. LIB and SIB cathodes. (A) Comparison to graphite in LIB format using NMC 333
vs. LIB and SIB cathodes. (A) Comparison to graphite in LIB format using NMC 333 cathodes. cathodes. (B)
Comparison to hard carbon in SIB format using Na 0.76Mn0.5Ni0.3Fe0.1Mg0.1O2 cathodes (MNFM).
(B) Comparison to hard carbon in SIB format using Na0 .76 Mn0 .5 Ni0 .3 Fe0 .1 Mg0 .1 O2 cathodes (MNFM).
See
SeeSupplementary
SupplementaryMaterials
MaterialsSection
Section S2
S2 for
for details.
details.

Interestingly,
Interestingly,while
while Ni
Ni22P@NC
P@NCclearly
clearlyfalls
fallsshort
shortofofgraphite
graphiteandandhard
hardcarbon
carboninin its
its
specific
specific energy contribution, its calculated volumetric energy contribution was found to to
energy contribution, its calculated volumetric energy contribution was found be
be considerably
considerably higher
higher thanthan
thatthat of graphite
of graphite andThe
and HC. HC.calculation
The calculation of volumetric
of volumetric en-
energy first
ergy first the
required required the calculation
calculation of the totalof the total
densities ofdensities of the
the materials in materials in the
the cells that cells that
produced the
produced the highest
highest specific energyspecific energy
(Table S4), (Table
followed byS4), followed by the
the multiplication multiplication
of cell of cell
density by specific
density
energy. by specific
Notably, energy.
these valuesNotably, these valueshigh
were impressively 993.2 Wh L−1high
wereatimpressively andat 993.2
402 Wh WhL−1 Lfor
−1

Ni2 P@NC in LIB and SIB formats, respectively, regarding active materials. This corresponds
to a ~25.2 and ~30.4% increase, respectively, compared to graphite and HC cells (Figure 6).
Therefore, despite the considerably higher average voltage profile of Ni2 P@NC compared
to both graphite and HC, the combination of its high specific capacity and high density with
optimized cathode balancing revealed, via calculations, that it could achieve much higher
volumetric energy than graphite and HC. Moreover, the above method provides a useful
first step for predicting the energy contribution of materials based on half-cell specific
capacity results prior to full-cell format testing. It should also be noted that previous
studies using other forms of ab initio calculations have been extremely useful in the
battery research field, e.g., for the prediction of a 5 V LIB employing Li2 CoMn3 O8 cathode
material in studies by Eglitis and Borstel [47,48]. Furthermore, from the perspective of
developing and evaluating alternative electrode materials with high specific capacities, the
calculated results show the importance of including maximum calculations in conjunction
with material density determination.

4. Conclusions
In this study, an environmentally friendly solid ionic resin provided a carbon and
phosphorus source to synthesize novel N-doped carbon-encapsulated Ni2 P nanorod ar-
rays (Ni2 P@NC) on nickel foam via “pyrolytic growth/carbon deposition”. Structural
and compositional analyses found that Ni2 P@NC comprises uniform N-doped carbon
shells encasing highly crystalline Ni2 P nanorods. These features significantly improve the
composite’s structural stability and electrical conductivity. Additionally, the open-ended
nanorod structure allows Ni2 P to be in direct contact with the electrolyte, thus simplifying
the ion transport path. Consequently, Ni2 P@NC cycled as a LIB anode material with an
impressively high capacity, a high rate capability and high cycling stability (retaining
630 mAh g−1 at 0.1 A g−1 after 150 cycles and 170 mAh g−1 at 5 A g−1 after 6000 cycles
with a capacity loss of only 0.00413% per cycle). Similarly, Ni2 P@NC as a SIB anode material
had a strong performance of 343.2 mAh g−1 at 0.1 A g−1 and 151 mAh g−1 at 2 A g−1 after
2000 cycles with only 0.00796% capacity loss per cycle. Furthermore, it was determined by
calculations that a sizable portion of the capacity coincides at a low enough average voltage
to offer a potentially significant increase in volumetric energy over standard graphite
Batteries 2023, 9, 267 11 of 14

and HC anodes. Lastly, the simple synthesis method has the potential to produce other
carbon-coated metal phosphides for electrochemical applications.

Supplementary Materials: The following supporting information can be downloaded at https://


www.mdpi.com/article/10.3390/batteries9050267/s1. Figure S1: SEM images of Ni2 P@NC nanorod
growth from Ni foam after different pyrolysis times: (A,B) 0 min (fresh nickel foam); (C,D) 5 min;
(E,F) 30 min and (G,H) 60 min; Figure S2: SEM images of Ni2 P; Figure S3: SEM images of Ni2 P/NC;
Figure S4: SEM images of NC; Figure S5: N2 adsorption–desorption isotherms of Ni2 P; Figure S6:
N2 adsorption–desorption isotherms of Ni2 P/NC; Figure S7: N2 adsorption–desorption isotherms
of NC; Figure S8: XRD pattern of Ni2 P@NC in 5 min; Figure S9: EDS analysis of nanorod array of
Ni2 P@NC in 30 min; Figure S10: CV curves of Ni2 P powder at 0.2 mV s–1 between 0.01 and 3.0 V
vs. Li+ /Li; Figure S11: (A) Discharge–charge curves of C-Ni2 P against Li metal in LIB electrolyte at
0.1 A g–1 . (B) Rate testing between 0.1 and 10 A g–1 . (C) Long-term cycling performance at 0.1 A g–1 ;
Figure S12: CV curves of Ni2 P powder at 0.2 mV s–1 between 0.01 and 3.0 V vs. Na+ /Na; Figure S13:
(A) Discharge–charge curves of C-Ni2 P against Na metal in SIB electrolyte at 0.1 A g–1 . (B) Rate testing
between 0.1 and 5 A g–1 . (C) Long-term cycling performance at 0.1 A g–1 ; Figure S14: EIS spectra
and equivalent circuit for Ni2 P@NC, Ni2 P/NC, Ni2 P and NC as SIB anode materials, where Re , Rct ,
CPE and Wo in the fitted equivalent circuit are electrolyte resistance, charge-transfer resistance at the
electrode/electrolyte interface; Figure S15: Average voltages of Ni2 P@NC in (A) LIBs and (B) SIBs.
(A) Comparison to graphite (0.15 V average) and capacity–mass-balanced LiNi0.3 Mn0.3 Co0.3 O2
(NMC 333: 3.8 V average). (B) Comparison to a typical hard carbon anode (HC: 0.45 V average)
and Na0.76 Mn0.5 Ni0.3 Fe0.1 Mg0.1 O2 (MNFM: 3.2 V average) [3,4,49,50]; Figure S16: (A and B) Average
voltages vs. cell specific capacity of Ni2 P@NC/cathodes in (A) LIBs and (B) SIBs. (C and D) Specific
and volumetric energy of Ni2 P@NC/cathodes in (A) LIBs and (B) SIBs with labeled maxima points;
Figure S17: Maximum energy density cutoff capacities and related voltages for Ni2 P@NC in LIBs
and SIBs demarked by dashed lines (see Table S4 for cutoff values); Table S1: Commonly reported
synthesis methods for nickel phosphides [12,24,26,28,51–66]; Table S2: Cycling performance of
reported Ni2 P as LIBs anode material [21,29,44,67–69]; Table S3: Cycling performance of reported
Ni2 P as SIB anode material [13,70,71]; Table S4: Density of active materials in cells yielding maximum
specific energy [72]; Table S5: Parameters for maximized energy of active materials.
Author Contributions: F.F.: Writing—Original Draft Preparation, Methodology, and Writing—
Reviewing and Editing. Q.H.: Software, Writing—Original Draft Preparation, and Methodology.
X.Z.: Software and Methodology. J.K.: Conceptualization and Writing—Reviewing and Editing. P.S.:
Funding Acquisition and Software. J.Z.: Conceptualization, Supervision, and Writing—Reviewing
and Editing. All authors have read and agreed to the published version of the manuscript.
Funding: This work was supported by the National Natural Science Foundation of China (51962002)
and the Natural Science Foundation of Guangxi (2022GXNSFAA035463).
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Asenbauer, J.; Eisenmann, T.; Kuenzel, M.; Kazzazi, A.; Chen, Z.; Bresser, D. The success story of graphite as a lithium-ion anode
material—Fundamentals, remaining challenges, and recent developments including silicon (oxide) composites. Sustain. Energy
Fuels 2020, 4, 5387–5416. [CrossRef]
2. Hwang, J.Y.; Myung, S.T.; Sun, Y.K. Sodium-ion batteries: Present and future. Chem. Soc. Rev. 2017, 46, 3529–3614. [CrossRef]
3. Mao, C.; Wood, M.; David, L.; Jin An, S.; Yangping Sheng, Y.; Du, Z.; Meyer, H.; Rose, E.; Ruther, R.E.; Wood, D.L. Selecting the
Best Graphite for Long-Life, High-Energy Li-Ion Batteries. J. Electrochem. Soc. 2018, 165, A1837. [CrossRef]
4. Moon, H.; Innocenti, A.; Liu, H.; Zhang, H.; Weil, M.; Zarrabeitia, M.; Passerini, S. Hard carbon cycling Bio-Waste-Derived Hard
Carbon Anodes Through a Sustainable and Cost-Effective Synthesis Process for Sodium-Ion Batteries. ChemSusChem 2023, 16,
e202201713.
5. Liu, L.; Tian, Y.; Abdussalam, A.; Gilani, M.R.H.S.; Zhang, W.; Xu, G. Hard Carbons as Anodes in Sodium-Ion Batteries: Sodium
Storage Mechanism and Optimi-zation Strategies. Molecules 2022, 27, 6516. [CrossRef] [PubMed]
6. Burchell, T.D.; McEnaney, B. Carbon Materials for Advanced Technologies (Structure and Bonding in Carbon Materials); Elsevier:
Amsterdam, The Netherlands, 1999; Volume 1, p. 2.
7. Pistoia, G. Lithium-Ion Batteries. Advances and Applications; Elsevier: Amsterdam, The Netherlands, 2014; Volume 2, p. 25.
Batteries 2023, 9, 267 12 of 14

8. Bates, A.M.; Preger, Y.; Torres-Castro, L.; Harrison, K.L.; Harris, S.J.; Hewson, J. Are solid-state batteries safer than lithium-ion
batteries? Joule 2022, 6, 742–755. [CrossRef]
9. Lin, L.; Zhang, C.; Huang, Y.; Zhuang, Y.; Fan, M.; Lin, J.; Wang, L.; Xie, Q.; Peng, D.-L. Challenge and Strategies in Room
Temperature Sodium–Sulfur Batteries: A Comparison with Lithium–Sulfur Batteries. Small 2022, 18, 2107368. [CrossRef]
10. Xu, X.; Feng, J.; Liu, J.; Lv, F.; Hu, R.; Fang, F.; Yang, L.; Ouyang, L.; Zhu, M. Robust spindle-structured FeP@C for high-
performance alkali-ion batteries anode. Electrochimi. Acta 2019, 312, 224–233. [CrossRef]
11. Sun, W.; Zhao, W.; Yuan, S.; Zhang, L.; Yang, Y.; Ge, P.; Ji, X. Designing Rational Interfacial Bonds for Hierarchical Mineral-Type
Trogtalite with Double Carbon towards Ultra-Fast Sodium-Ions Storage Properties. Adv. Funct. Mater. 2021, 31, 2100156.
[CrossRef]
12. Sun, H.; Xu, X.; Yan, Z.; Chen, X.; Cheng, F.; Weiss, P.S.; Chen, J. Porous Multishelled Ni2 P Hollow Microspheres as an Active
Electrocatalyst for Hydrogen and Oxygen Evolution. Chem. Mater. 2017, 29, 8539–8547. [CrossRef]
13. Liang, J.; Zhu, G.; Zhang, Y.; Liang, H.; Huang, W. Conversion of hydroxide into carbon-coated phosphide using plasma for
sodium ion batteries. Nano Res. 2022, 15, 2023–2029. [CrossRef]
14. Li, Y.; Zhang, J.; Li, D.; Ding, J.; Liu, Y.; Cai, Q. Fabrication of Core-Shell Ni2 P@N, P−Co-Doped Carbon/Reduced Graphene
Oxide Composite as Anode Material for Lithium- and Sodium-Ion Batteries. ChemElectroChem 2019, 6, 5492–5498. [CrossRef]
15. Li, Q.; Yang, D.; Chen, H.; Lv, X.; Jiang, Y.; Feng, Y.; Rui, X.; Yu, Y. Advances in metal phosphides for sodium-ion batteries. Sustain.
Mat. 2021, 1, 359–392.
16. Jain, A.; Ong, S.P.; Hautier, G.; Wei Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al.
Commentary: The Materials Project: A Materials Genome Approach to Accelerating Materials Innovation; APL Materials 1; AIP: Long
Island, NY, USA, 2013; p. 011002.
17. Zhou, S.; Huang, P.; Xiong, T.; Yang, F.; Yang, H.; Huang, Y.; Li, D.; Deng, J.; Balogun, M.S. Sub-Thick Electrodes with Enhanced
Transport Kinetics via In Situ Epitaxial Heterogeneous Interfaces for High Areal-Capacity Lithium Ion Batteries. Small 2021, 17,
2100778. [CrossRef]
18. Wang, Z.; Yu, C.; Zhao, C.; Guo, W.; Yu, J.; Qiu, J. Interface Inversion: A Promising Strategy to Configure Ultrafine Nanoparticles
over Graphene for Fast Sodium Storage. Small 2021, 17, 2005119. [CrossRef] [PubMed]
19. Liu, B.; Jiang, K.; Zhu, K.; Liu, X.; Ye, K.; Yan, J.; Wang, G.; Cao, D. Cable-like polyimide@carbon nanotubes composite as a
capable anode for lithium ion batteries. Chem. Eng. J. 2022, 446, 137208. [CrossRef]
20. Liu, S.; Niu, K.; Chen, S.; Sun, X.; Liu, L.; Jiang, B.; Chu, L.; Lv, X.; Li, M. TiO2 bunchy hierarchical structure with effective
enhancement in sodium storage behaviors. Carbon Energy 2022, 4, 645–653. [CrossRef]
21. Kim, C.; Kim, H.; Choi, Y.; Lee, H.A.; Jung, Y.S.; Park, J. Facile Method to Prepare for the Ni2 P Nanostructures with Controlled
Crystallinity and Morphology as Anode Materials of Lithium-Ion Batteries. ACS Omega 2018, 3, 7655–7662. [CrossRef]
22. Miao, X.; Yin, R.; Ge, X.; Li, Z.; Yin, L. Ni2 P@Carbon Core–Shell Nanoparticle-Arched 3D Interconnected Graphene Aerogel
Architectures as Anodes for High-Performance Sodium-Ion Batteries. Small 2017, 13, 1702138. [CrossRef]
23. Xiong, X.; You, C.; Cao, X.; Pang, L.; Kong, R.; Sun, X. Ni2 P nanosheets array as a novel electrochemical catalyst electrode for
non-enzymatic H2 O2 sensing. Electrochim. Acta 2017, 253, 517–521. [CrossRef]
24. Stern, L.-A.; Feng, L.; Song, F.; Hu, X. Ni2 P as a Janus catalyst for water splitting: The oxygen evolution activity of Ni2 P
nanoparticles. Energy Environ. Sci. 2015, 8, 2347–2351. [CrossRef]
25. Shi, F.; Xie, D.; Zhong, Y.; Wang, D.H.; Xia, X.H.; Gu, C.D.; Wang, X.L.; Tu, J.P. Facile synthesis of self-supported Ni2 P
nanosheet@Ni sponge composite for high-rate battery. J. Power Sources 2016, 328, 405–412. [CrossRef]
26. Jin, Y.; Zhao, C.; Wang, L.; Jiang, Q.; Ji, C.; He, X. Preparation of mesoporous Ni2 P nanobelts with high performance for
electrocatalytic hydrogen evolution and supercapacitor. Int. J. Hydrogen Energy 2018, 43, 3697–3704. [CrossRef]
27. Li, Q.; Li, X.; Gu, J.; Li, Y.; Tian, Z.; Pang, H. Porous rod-like Ni2 P/Ni assemblies for enhanced urea electrooxidation. Nano Res.
2021, 14, 1405–1412. [CrossRef]
28. Chen, T.; Liu, D.; Lu, W.; Wang, K.; Du, G.; Asiri, A.M.; Sun, X. Three-Dimensional Ni2 P Nanoarray: An Efficient Catalyst
Electrode for Sensitive and Selective Nonenzymatic Glucose Sensing with High Specificity. Anal. Chem. 2016, 88, 7885–7889.
[CrossRef]
29. Zheng, J.; Huang, X.; Pan, X.; Teng, C.; Wang, N. Yolk-shelled Ni2 P@carbon nanocomposite as high-performance anode material
for lithium and sodium ion batteries. Appl. Surf. Sci. 2019, 473, 699–705. [CrossRef]
30. Duan, C.W.; Hu, L.X.; Ma, J.L. Ionic liquids as an efficient medium for the mechanochemical synthesis of α-AlH3 nano-composites.
J. Mater. Chem A 2018, 6, 6309–6318. [CrossRef]
31. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases,
with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem.
2015, 87, 1051–1069. [CrossRef]
32. Lin, X.-M.; Chen, J.-H.; Fan, J.-J.; Ma, Y.; Radjenovic, P.; Xu, Q.-C.; Huang, L.; Passerini, S.; Tian, Z.-Q.; Li, J.-F. Synthesis and
Operando Sodiation Mechanistic Study of Nitrogen-Doped Porous Carbon Coated Bimetallic Sulfide Hollow Nanocubes as
Advanced Sodium Ion Battery Anode. Adv. Energy Mater. 2019, 9, 1902312. [CrossRef]
33. Liu, X.; Li, W.; Zhao, X.; Liu, Y.; Nan, C.-W.; Fan, L.-Z. Two Birds with One Stone: Metal–Organic Framework Derived Micro-
/Nanostructured Ni2 P/Ni Hybrids Embedded in Porous Carbon for Electrocatalysis and Energy Storage. Adv. Funct. Mater. 2019,
29, 1901510. [CrossRef]
Batteries 2023, 9, 267 13 of 14

34. Yuan, H.; Wu, M.; Zheng, J.; Chen, Z.-G.; Zhang, W.; Luo, J.; Jin, C.; Sheng, O.; Liang, C.; Gan, Y.; et al. Empowering Metal
Phosphides Anode with Catalytic Attribute toward Superior Cyclability for Lithium-Ion Storage. Adv. Funct. Mater. 2019, 29,
1809051. [CrossRef]
35. Yuan, W.; Jiang, T.; Fang, X.; Fan, Y.; Qian, S.; Gao, Y.; Cheng, N.; Xue, H.; Tian, J. Interface engineering of S-doped Co2 P@Ni2 P
core–shell heterostructures for efficient and energy-saving water splitting. Chem. Eng. J. 2022, 439, 135743. [CrossRef]
36. Li, D.; Li, Z.; Zou, R.; Shi, G.; Huang, Y.; Yang, W.; Yang, W.; Liu, C.; Peng, X. Coupling overall water splitting and biomass
oxidation via Fe-doped Ni2 P@C nanosheets at large current density. Appl. Catal. B 2022, 307, 121170. [CrossRef]
37. Wan, Y.; Song, K.; Chen, W.; Qin, C.; Zhang, X.; Zhang, J.; Dai, H.; Hu, Z.; Yan, P.; Liu, C.; et al. Ultra-High Initial Coulombic
Efficiency Induced by Interface Engineering Enables Rapid, Stable Sodium Storage. Angew. Chem. Int. Ed. 2021, 60, 11481–11486.
[CrossRef] [PubMed]
38. Zhu, J.; He, Q.; Liu, Y.; Key, J.; Nie, S.; Wu, M.; Shen, P.K. Three-dimensional, hetero-structured, Cu3 P@C nanosheets with
excellent cycling stability as Na-ion battery anode material. J. Mater. Chem A 2019, 7, 16999–17007. [CrossRef]
39. Zhu, J.; Wu, Q.; Key, J.; Wu, M.; Shen, P.K. Self-assembled superstructure of carbon-wrapped, single-crystalline Cu3 P porous
nanosheets: One-step synthesis and enhanced Li-ion battery anode performance. Energy Storage Mater. 2018, 15, 75–81. [CrossRef]
40. Jiang, J.; Ma, C.; Zhang, W.; He, Y.; Li, X.; Yuan, X. Controlled design for integration of FeP into 3D carbon frameworks for
superior Na storage. Chem. Eng. J. 2022, 429, 132271. [CrossRef]
41. Wei, P.; Zhu, J.; Qiu, Y.; Wang, G.; Xu, X.; Ma, S.; Shen, P.K.; Wu, X.-L.; Yamauchi, Y. One-dimensional core–shell motif nanowires
with chemically-bonded transition metal sulfide-carbon heterostructures for efficient sodium-ion storage. Chem. Sci. 2021, 12,
15054–15060. [CrossRef]
42. Abdullaeva, Z.; Omurzak, E.; Iwamoto, C.; Ganapathy, H.S.; Sulaimankulova, S.; Liliang, C.; Mashimo, T. Onion-like carbon-
encapsulated Co, Ni, and Fe magnetic nanoparticles with low cytotoxicity synthesized by a pulsed plasma in a liquid. Carbon
2012, 50, 1776–1785. [CrossRef]
43. Bai, Y.; Zhang, H.; Fang, L.; Liu, L.; Qiu, H.; Wang, Y. Novel peapod array of Ni2 P@graphitized carbon fiber composites growing
on Ti substrate: A superior material for Li-ion batteries and the hydrogen evolution reaction. J. Mater. Chem A 2015, 3, 5434–5441.
[CrossRef]
44. Bai, Y.; Zhang, H.; Li, X.; Liu, L.; Xu, H.; Qiu, H.; Wang, Y. Novel peapod-like Ni2 P nanoparticles with improved electrochemical
properties for hydrogen evolution and lithium storage. Nanoscale 2015, 7, 1446–1453. [CrossRef] [PubMed]
45. Shi, W.; Zhang, Y.; Tian, Z.Q.; Pan, Z.; Key, J.; Shen, P.K. Low temperature synthesis of polyhedral hollow porous carbon with high
rate capability and long-term cycling stability as Li-ion and Na-ion battery anode material. J. Power Sources 2018, 398, 149–158.
[CrossRef]
46. Sun, C.; Chen, S.; Li, Z. Controllable synthesis of Fe2 O3 -carbon fiber composites via a facile sol-gel route as anode materials for
lithium ion batteries. Appl. Surf. Sci. 2018, 427, 476–484. [CrossRef]
47. Eglitis, R.I.; Borstel, G. Towards a practical rechargeable 5 V Li ion battery. Phys. Stat. Sol. A. 2005, 202, R13–R15.
48. Eglitis, R.I. Theoretical prediction of the 5 V rechargeable Li ion battery using Li2 CoMn3 O8 as a cathode. Phys. Scr. 2015, 90, 9.
[CrossRef]
49. Shaju, K.M.; Subba Rao, G.V.; Chowdari, B.V.R. Performance of layered Li(Ni1/3 Co1/3 Mn1/3 ) O2 as cathode for Li-ion batteries.
Electrochim. Acta 2002, 48, 145–151. [CrossRef]
50. Keller, M.; Buchholz, D.; Passerini, S. Layered Na-Ion Cathodes with Outstanding Performance Resulting from the Synergetic
Effect of Mixed P- and O-Type Phases. Adv. Energy Mater. 2016, 6, 1501555.
51. Chen, Y.; Qin, Z. General Applicability of Nanocrystalline Ni2 P as a Noble-Metal-Free Cocatalyst to Boost Photocatalytic
Hydrogen Generation. Catal. Sci. Technol. 2016, 6, 8212–8221. [CrossRef]
52. Mai, Y.-J.; Xia, X.; Jie, X.-H. Cross-Linked Hierarchical Arrays of Ni2 P Nanoflakes Prepared Via Directional Phosphorization and
Their Applications for Advanced Alkaline Batteries. J. Power Sources 2017, 367, 116–121. [CrossRef]
53. Liu, G.; He, D.; Yao, R.; Zhao, Y.; Li, J. Enhancing the Water Oxidation Activity of Ni2 P Nanocatalysts by Iron-Doping and
Electrochemical Activation. Electrochimi. Acta 2017, 253, 498–505. [CrossRef]
54. Li, Q.; Ma, J.; Wang, H.; Yang, X.; Yuan, R.; Chai, Y. interconnected Ni2 P Nanorods Grown on Nickel Foam for Binder Free
Lithium Ion Batteries. Electrochim. Acta 2016, 213, 201–206. [CrossRef]
55. Cheng, J.; Zhao, D.; Fan, L.; Wu, X.; Wang, M.; Zhang, N.; Sun, K. Ultra-High Rate Li-S Batteries Based on A Novel Conductive
Ni2 P Yolk-Shell Material as the Host for the S Cathode. J. Mater. Chem. A 2017, 5, 14519–14524. [CrossRef]
56. Laursen, A.B.; Wexler, R.B.; Whitaker, M.J.; Izett, E.J.; Calvinho, K.U.D.; Hwang, S.; Rucker, R.; Wang, H.; Ji, J.; Garfunkel, E.; et al.
Climbing the Volcano of Electrocatalytic Activity while Avoiding Catalyst Corrosion: Ni3 P, a Hydrogen Evolution Electrocatalyst
Stable in Both Acid and Alkali. Acs Catal. 2018, 8, 4408–4419. [CrossRef]
57. Wang, X.D.; Cao, Y.; Teng, Y.; Chen, H.Y.; Xu, Y.F.; Kuang, D.B. Large-Area Synthesis of A Ni2 P Honeycomb Electrode for Highly
Efficient Water Splitting. ACS Appl. Mater. Interfaces 2017, 9, 32812–32819. [CrossRef]
58. Shi, Y.; Xu, Y.; Zhuo, S.; Zhang, J.; Zhang, B. Ni2 P Nanosheets/Ni Foam Composite Electrode for Long-Lived and Ph-Tolerable
Electrochemical Hydrogen Generation. ACS Appl. Mater. Interfaces 2015, 7, 2376–2384. [CrossRef]
59. Liu, S.; Feng, J.; Bian, X.; Liu, J.; Xu, H. Electroless Deposition of Ni3 P-Ni Arrays on 3-D Nickel Foam as A High Performance
Anode for Lithium-Ion Batteries. RSC Adv. 2015, 5, 60870–60875. [CrossRef]
Batteries 2023, 9, 267 14 of 14

60. Chen, G.-F.; Ma, T.Y.; Liu, Z.-Q.; Li, N.; Su, Y.-Z.; Davey, K.; Qiao, S.-Z. Efficient and Stable Bifunctional Electrocatalysts Ni/Nixmy
(M = P, S) for Overall Water Splitting. Adv. Funct. Mater. 2016, 26, 3314–3323. [CrossRef]
61. Wang, X.; Li, W.; Xiong, D.; Liu, L. Fast Fabrication of Self-Supported Porous Nickel Phosphide Foam for Efficient, Durable
Oxygen Evolution and Overall Water Splitting. J. Mater. Chem. A 2016, 4, 5639–5646. [CrossRef]
62. Shi, S.; Li, Z.; Sun, Y.; Wang, B.; Liu, Q.; Hou, Y.; Huang, S.; Huang, J.; Zhao, Y. A Covalent Heterostructure of Monodisperse
Ni2 P Immobilized on N, P-Co-Doped Carbon Nanosheets for High Performance Sodium/Lithium Storage. Nano Energy 2018, 48,
510–517. [CrossRef]
63. Sun, Y.; Zhang, T.; Li, X.; Bai, Y.; Lyu, X.; Liu, G.; Cai, W.; Li, Y. Bifunctional Hybrid Ni/Ni2 P Nanoparticles Encapsulated By
Graphitic Carbon Supported with N, S Modified 3D Carbon Framework for Highly Efficient Overall Water Splitting. Adv. Mater.
Interfaces 2018, 5, 1800473. [CrossRef]
64. Sun, T.; Dong, J.; Huang, Y.; Ran, W.; Chen, J.; Xu, L. Highly Active and Stable Electrocatalyst of Ni2 P Nanoparticles Supported
on 3D Ordered Macro-/Mesoporous Co-N-Doped Carbon for Acidic Hydrogen Evolution Reaction. J. Mater. Chem. A 2018, 6,
12751–12758. [CrossRef]
65. Li, Y.; Cai, P.; Ci, S.; Wen, Z. Strongly Coupled 3D Nanohybrids with Ni2 P/Carbon Nanosheets as Ph-Universal Hydrogen
Evolution Reaction Electrocatalysts. Chemelectrochem 2017, 4, 340–344. [CrossRef]
66. Feng, Y.; Zhang, H.; Mu, Y.; Li, W.; Sun, J.; Wu, K.; Wang, Y. Monodisperse Sandwich-Like Coupled Quasi-Graphene Sheets
Encapsulating Ni2 P Nanoparticles for Enhanced Lithium-Ion Batteries. Chemistry 2015, 21, 9229–9235. [CrossRef] [PubMed]
67. Yao, C.; Xu, J.; Shen, Y.; Xie, A. Synthesis and Excellent Performance of Porous Ni2 P@C/Cnts Nanocomposite Derived From
Ni-Mofs as an Anode for Lithium-Ion Batteries. Int. J. Energy Res. 2022, 46, 10875–10884. [CrossRef]
68. Zhang, R.Z.; Zhu, K.J.; Huang, J.D.; Yang, L.Y.; Li, S.T.; Wang, Z.Y.; Xie, J.R.; Wang, H.; Liu, J. Ultrafine Ni2 P Nanoparticles
Embedded in one-Dimensional Carbon Skeleton Derived From Metal-Organic Frameworks Template as a High-Performance
Anode for Lithium Ion Battery. J. Alloys Compd. 2019, 775, 490–497. [CrossRef]
69. Dong, C.; Guo, L.; He, Y.; Chen, C.; Qian, Y.; Chen, Y.; Xu, L. Sandwich-Like Ni2 P Nanoarray/Nitrogen-Doped Graphene
Nanoarchitecture as A High-Performance Anode for Sodium and Lithium Ion Batteries. Energy Storage Mater. 2018, 15, 234–241.
[CrossRef]
70. Zhou, D.; Xue, L.-P.; Wang, N. Robustly Immobilized Ni2 P Nanoparticles in Porous Carbon Networks Promotes High-Performance
Sodium-Ion Storage. J. Alloys Compd. 2019, 776, 912–918. [CrossRef]
71. Li, H.; Wang, X.; Zhao, Z.; Tian, Z.; Zhang, D.; Wu, Y. Ni2 P Nanoflake Array/Three Dimensional Graphene Architecture as
Integrated Free-Standing Anode for Boosting the Sodiation Capability and Stability. ChemElectroChem 2019, 6, 404–412. [CrossRef]
72. Yin, S.; Rho, Y.; Swainson, I.; Nazar, L. X-ray/Neutron Diffraction and Electrochemical Studies of Lithium De/Re-Intercalation in
Li1-x Co1/3 Ni1/3 Mn1/3 O2 (x = 0→1). Chem. Mater. 2006, 18, 1901–1910. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like