Graphene Oxide Modified Lini Co MN O As Cathode Material For Lithium Ion Batteries and Its Electrochemical Performances
Graphene Oxide Modified Lini Co MN O As Cathode Material For Lithium Ion Batteries and Its Electrochemical Performances
Graphene Oxide Modified Lini Co MN O As Cathode Material For Lithium Ion Batteries and Its Electrochemical Performances
17
International Journal of
ELECTROCHEMICAL
SCIENCE
www.electrochemsci.org
LiNi1/3Co1/3Mn1/3O2(NCM) was coated with graphene oxide(GO) by a simple chemical approach. The
prepared materials were characterised using various techniques such as X-ray diffraction (XRD),
thermogravimetric analysis, scanning electron microscopy, energy dispersive spectroscopy,
transmission electron microscopy and X-ray photoelectron spectroscopy. XRD revealed that the crystal
structure and crystallinity do not change after surface modification with GO. The 2.5 wt% GO-coated
NCM exhibits stable cycling stability with a capacity retention of 84.2% at 1 C after 100 cycles, which
is higher than the corresponding value of 74.3% for the pristine sample. The rate performance results
show that the GO coating increases the discharge capacity from 87.1 to 100.9 mAh g−1 at a high rate of
5 C. Therefore, GO-coated NCM is a highly promising cathode material for lithium-ion batteries
because it provides good cycling stability and high rate performance.
1. INTRODUCTION
Advanced energy storage devices, such as lithium-ion batteries [1], supercapacitors [2],
sodium-sulphur batteries [3] and solar cells [4], have been widely researched to address the global
energy crisis and climate change problems. Among various forms of energy storage equipment,
lithium-ion batteries have been regarded suitable power sources for battery electric vehicles, hybrid
electric vehicles (HEVs) and parallel HEVs because of their high energy density, good cycling
performance and long cycle life [5–7]. Lithium-ion batteries consist of cathodes, anodes and
Int. J. Electrochem. Sci., Vol. 13, 2018 10441
electrolytes. However, the bottleneck problem that hinders the development of lithium-ion batteries is
the limiting factor in high-performance cathode materials. LiCoO2 [8] has been used as a cathode
material in commercial LIBs because of its high charge/discharge voltage, good cycling ability and
easy preparation; however, its low thermal stability, high cost and toxicity restrict its further
development. LiNiO2 [9] has been considered for use in second-generation commercial LIBs because
of its high power and energy density, low cost and high discharge capacity; however, the chemical
instability of Ni3+ ions makes it difficult to synthesise. LiMn2O4 [10, 11] has the advantages of low
cost, high thermal stability and high voltage, but its reversible capacity fades rapidly at high current
rates. Compared with other cathode materials, such as LiCoO2, LiNiO2 and LiMn2O4, layered
LiNixCoyMn1-x-yO2 has higher theoretical capacity, lower cost, higher safety and higher working
voltage and environmentally friendly; it is also acknowledged as a promising candidate material for
commercial LIBs. Recently, a series of LiNixCoyMn1-x-yO2 was investigated including
LiNi1/3Co1/3Mn1/3O2 [12], LiNi0.5Co0.2Mn0.3O2 [13], LiNi0.6Co0.2Mn0.2O2 [14] and LiNi0.8Co0.1Mn0.1O2
[15], in which the Ni content imparts a high charge/discharge capacity because of its redox reactions.
However, the addition of Ni will cause cation mixing because the ionic radius of Li+(0.76 Å) is similar
to that of Ni2+(0.69 Å), thus resulting in poor electrochemical performance. Among the LiNixCoyMn1-x-
yO2 mixed oxide series, NCM (111) has attracted great attention from academic and industrial
engineers because of its high specific capacity, superior structural stability and environmental
benignancy.
However, several drawbacks limit its further application. The first is the degradation of its
reversible capacity during prolonged cycling because of an unstable surface layer between the
electrode and electrolyte. The second is its poor rate capability, which can be attributed to poor
electronic conductivity and side reactions with the cathode material. To overcome the shortcomings
mentioned above, surface modification has proven to be an effective technique for improving the
electrochemical performance of NCM. Thus, the most common coating materials including Al2O3 [16],
MgO [17], TiO2 [18], Nb2O5 [19], V2O5 [20], SrF2 [21], AlPO4 [22], Li2ZrO3 [23], Li3VO4 [24] and
Li2MnO3 [25] have contributed to improvements in thermal stability and rate capacity. A suitable
coating layer can act as a protective shell to prevent interfacial side reactions. However, most metal
oxides have polarised the electrode and decreased the reversible capacity because of their inactivity
with respect to lithium insertion/extraction. Therefore, it is essential to develop better conductive
material to improve the electrochemical performance of NCM. Carbon is highly abundant and is
associated with mature technologies. Therefore, its use in lithium-ion batteries should be pursued.
Graphene oxide (GO), with its superior electronic conductivity, large specific surface area, stable
chemical properties and mechanical strength, is considered an excellent conductive additive to
nanostructure materials for lithium-ion batteries [26, 27]. In this work, we have explored the possibility
of coating NCM particles with a layer of GO coating layer. GO was synthesised by a modified
Hummers method, and a simple chemical approach was used to wrap it around NCM particles to
improve their electrochemical performance. The influence of the GO coating on the cathode structure,
morphology and electrochemical performance is discussed in detail. The satisfactory performance of
GO-coated NCM (GO/NCM) renders it a promising cathode material that exhibits higher reversible
capability, longer cycling stability and better rate capability than NCM.
Int. J. Electrochem. Sci., Vol. 13, 2018 10442
2. EXPERIMENTAL
NCM was synthesised by a sol-gel citric acid assisted method. In a typical experiment, lithium
acetate, nickel acetate, cobalt acetate and manganese, with a stoichiometric ratio of
Li:(Ni+Co+Mn)=1.05:1, were selected as the starting materials and dissolved in ethyl alcohol. Excess
5% Li+ was added to offset its loss, which occurs at high temperature. Citric acid solution was added
dropwise to the above solution with vigorous stirring. The molar ratio of metal ions to citric acid was
1:1.2. The mixture was evaporated at 80°C with continuous stirring until a gel was formed, which was
subsequently frozen at -80°C in a cryogenic refrigerator for 12 h, followed by freeze-drying for 24 h to
acquire the precursor. The precursor was then heated at 450°C in air for 5 h with subsequent
calcinations at 900°C for 12 h to obtain NCM.
GO/NCM was prepared by a simple chemical approach. First, GO was produced using a
modified Hummers method [28, 29].Potassium persulphate (0.5 g), phosphorous pentoxide (0.5 g) and
concentrated sulphuric acid (20 ml) were added into a 100 ml flask kept at 80°C; 1 g graphite powder
was then placed in the flask, and the resultant dark blue mixture was kept at 80°C for 6 h. The pre-
oxidised product was diluted with deionised water, washed and then filtered using a circulating water
type vacuum pump until the syringe water pH=7. The pre-oxidised product was dried at 80°C for 8 h.
Thereafter, 1 g of pre-oxidised product and 0.5 g NaNO3 were dissolved in 23 ml of 98% H2SO4 at 0°C
and 3 g KMnO4 was slowly added, keeping the temperature of the suspension below 5°C with an ice
bath. The mixture was heated to 35°C and stirred for 3 h before adding distilled water (80 ml). The
mixture was then heated to 98°C and further stirred for 30 min before adding 3% H2O2 solution (10
ml) to remove the remaining KMnO4, after which the colour of the mixture changed to bright yellow.
The mixture was filtered and washed with 10% HCl solution to remove the residual metal ions and
acids and was maintained at pH=7. The material was frozen at -80°C in a cryogenic refrigerator for 12
h, followed by freeze-drying for 24 h. Thus, the GO sample was obtained.
GO (25 mg) was dispersed in 30 ml ethanol with ultrasonication for 3 h to ensure that it was
completely dissolved. Thereafter, 1 g of NCM powder was immersed in the mixture with vigorous
stirring, and the mixture was heated at 60°C until the alcohol completely evaporated. Finally, the
obtained material was placed in a freeze dryer at -50°C for 10 h.
The crystal structure was analysed by X-ray diffraction (XRD) (D8 advance, Bruker) with Cu-
K radiation (40 kV, 40 mA) in the 2 range 10°–80° at a scan rate of 10°/min. The particle site,
surface morphologies and elemental composition of the samples were performed by scanning electron
microscopy (SEM) (S-4800, Hitachi) equipped with energy dispersive spectroscopy (EDS).
Int. J. Electrochem. Sci., Vol. 13, 2018 10443
Thermogravimetric analysis (TGA) (AQ20, TA) was performed from room temperature to 900°C at a
heating rate of 10°/min in air. Transmission electron microscopy (TEM) (Talos F200S, FEI) was
employed to estimate the thickness of the coating on the surface of the cathode powders. X-ray
photoelectron spectroscopy (XPS) (Thermo Scientific ESCALAB 250Xi) was used to evaluate the ion
valence states in the metal oxide.
The electrochemical properties of pristine and GO/NCM were measured using a CR2032 coin
cell. The electrodes were prepared by mixing the active material with acetylene black and
polyvinylidene fluoride (80:10:10) in N-methyl-2-pyrrolidone to obtain a slurry, which was coated on
cleaned and polished aluminum foil and dried at 80°C for 10 h in a vacuum oven. The cells were
assembled in an argon-filled glove box, in which water and oxygen levels were maintained below
1ppm. LiPF6 (1 M) dissolved in an EC: DMC (1:1 vol.%) electrolyte. The cells consist of a prepared
cathode, with lithium metal as an anode and a polypropylene microporous film (Celgard 2400) as the
separator. The weight of active material in the electrode was 3 mg/cm2. A galvanostatic charge–
discharge test was performed at rates of 0.2, 0.5, 1, 2 and 5 C in the voltage range 2.5–4.3 V at room
temperature. Electrochemical impedance spectroscopy (EIS) was also conducted on an Autolab
PGSTAT302N in the frequency range 100–0.01 Hz with a potential perturbation of 5 mV.
The chemical compositions and crystal structures of NCM and GO/NCM were determined by
XRD. As shown in Fig.1(a), all the diffraction peaks are sharp and well-defined, thus indicating that
the synthesised materials are well-crystallised. All the materials can be indexed as layered oxide
structures based on a hexagonal -NaFeO2 structure with an ̅ m space group, with no impurity
phases detected [30]. The splitting of the (006)/(102) and (108)/(110) peaks at 38° and 65° is clearly
Int. J. Electrochem. Sci., Vol. 13, 2018 10444
identified, thus indicating that the materials have highly ordered layered structures [31]. The XRD
pattern of GO shown in Fig.1(b) has a sharp peak at 11°, which can be attributed to the complete
oxidation of graphite to graphite oxide [32]. However, no diffraction peaks of GO can be observed in
the XRD patterns of GO/NCM. This is probably because the strong diffraction peaks from the highly
crystalline NCM and the coated GO (2.5 wt%) are below the detection limit of XRD [33, 34]. The
calculated lattice parameters (a and c) for NCM (a=2.857 Å and c=14.205 Å) and GO/NCM (a=2.860
Å and c=14.230 Å) are in accordance with those reported in the literature[35, 36].The intensity ratio
I(003)/I(104) is a sensitive parameter with respect to the degree of cation mixing of the samples. In a
previous study, in which I(003)/I(104)>1.2, it was confirmed that the material possessed a good layer
structure with low cation mixing [37]. In the present study, the I(003)/I(104) values for the NCM and
GO/NCM powders were calculated as 1.28 and 1.32, respectively. All the results indicate that the
crystal structure and crystallinity do not change after surface modification with GO.
(a) (b)
GO
(c) (d)
GO
Figure 3. SEM images of bare (a and c) and GO/NCM cathodes (b and d).
Int. J. Electrochem. Sci., Vol. 13, 2018 10445
(a) (b)
Ni Ka1 Co Ka1
(c) (d)
Mn Ka1 C Ka1
(a) (b) GO
coating
layer
(003)
d=4.7 Å
5 nm 10 nm
TGA was used to quantify the amount of GO present in the GO/NCM composites. Fig.2 shows
the thermogravimetric (TG) curves of the GO/NCM when it was healed from 25°C to 900°C at a rate
of 5°/min. According to the TG curve, the weight loss of GO (0.5%) below 200°C resulted from the
Int. J. Electrochem. Sci., Vol. 13, 2018 10446
removal of absorbed water molecules from its surface. The weight loss from 200°C to 350°C was due
to the decomposition of oxygen-containing functional groups in the GO sheets, and the weight loss
above 350°C was due to the combustion of carbon skeleton [38]. The TG results confirm that the
content of GO in the NCM cathode material is 2.5%.
The apparent morphologies of NCM and GO/NCM were examined by SEM, as shown in Fig.3.
Both samples consisted of homogeneous primary particles with average particle sizes of 200–500 nm.
The images of the GO/NCM composite show NCM particles embedded in the GO sheets. The uniform
GO/NCM structures can be ascribed to the simple chemical mixing method. Furthermore, continuous
stirring and ultrasonication were effective in distributing the materials throughout the GO sheets. Thus,
the GO network provides conductive connections between the NCM particles. EDS was also used to
confirm the distribution of the GO coating layer on the surface of NCM. As illustrated in Fig.4,
elemental C originating from GO was uniformly distributed throughout the cathode material, thus
indicating that the cathode material is successfully and evenly coated with GO via this simple chemical
approach.
To further confirm the formation of the coated structure, high-resolution TEM (HRTEM) was
performed to evaluate the interplanar spacing and observe the microstructures of the coated materials.
The bare sample presents perfect crystallinity in Fig.5a, with clear lattice fringes extending to the
particle boundary. The interplanar spacing of the parallel lattice is approximately 4.7 Å, which
corresponds to the spacing of the (003) plane of the hexagonal layered structure [39]. Fig.5b shows the
GO coating layer on the surface of NCM with a thickness approximately 10–12 nm, and an obvious
gap was found between the GO and the cathode material. This finding indicated tight adhesion
between GO and NCM.
The modification of NCM with GO, can lead to the transition metals (Ni, Co, Mn) changing
their oxidation states and result in low discharge capacity and poor cycling performance. Thus, XPS
measurements were performed to detect the valence state of the transition metals in the GO/NCM
materials. The results are shown in Fig.6. The XPS wide spectra show that GO/NCM has Li1s, Ni2p,
Co2p, Mn2p, C1sand O1s peaks. The binding energies of Ni2p3/2, Co2p3/2 and Mn2p3/2 for GO/NCM
are 855.3, 780.1 and 624.6 eV, respectively, thus suggesting that Ni, Co, and Mn in our materials are
respectively divalent, trivalent and tetravalent, which is consistent with reported values [40]. Two
major peaks at about emerge in the Ni2p spectrum (Fig.6b); these peaks are assigned to the Ni2p3/2 and
Ni2p1/2 peaks. The presence of a satellite peak can be attributed to the multiple splitting of nickel oxide
energy levels [41]. The electron binding energies of Co2p3/2 and Co2p1/2 were recorded at 780.1 and
795.5 eV, thus implying that the existence of Co3+. The Mn2p3/2 energy at 646.2 eV confirms Mn4+
oxidation states in the samples [42]. The C1s spectra of the materials are deconvoluted into three main
peaks. The peaks with binding energies of 288.5, 285.2 and 284.5 eV are ascribed to carboxyl groups
(–O–C=O), hydroxyl groups (C–OH) and carbon networks (C–C/C=C), respectively [43]. The results
indicate that GO in the 2.5 wt% GO/NCM has no impact on the valence state of the transition metals.
Int. J. Electrochem. Sci., Vol. 13, 2018 10447
Figure 6. XPS spectra of GO/NCM for all elements(a), Ni2p(b), Co2p(c), Mn2p(d) and C1s(e).
Fig.7 shows the initial charge-discharge curves of the bare and 2.5 wt% GO/NCM samples at
0.1 C in the voltage range 2.5–4.3 V. All charge-discharge curves had a potential plateau at 3.6 V,
consistent with a typical-layer-structured NCM. At room temperature, the initial charge-discharge
capacities for GO/NCM and NCM are 195.8/177.2 and 190.7/172.6 mAh g−1, respectively. It is clear
that the GO plays an important role in improving the capacity of the cathode materials. The increased
discharge capacity of GO/NCM is mainly attributed to two factors. Firstly, coating GO, which has a
large surface area, on the surface of NCM increases the contact area between the cathode material and
electrolyte, further improving the performance of the active material. Secondly, lithium storage on the
GO sheets could also reduce the electrochemical polarisation during charge/discharge.
Figure 8. Cycling performances of NCM and GO-NCM samples at 0.2C (a) and 1C (b) after 100
cycles.
Fig.8(a) shows the cycling performances of the NCM and 2.5 wt% GO/NCM electrodes at a
constant current of 0.2 C between 2.5 and 4.3 V at room temperature. The NCM delivers initial
capacities of 165.8 mAh g−1 after 100 cycles, and it retains a capacity of 140.3 mAh g−1 with a capacity
retention of 84.6%. On the contrary, the 2.5 wt% GO/NCM materials deliver initial discharge
capacities of 152.4 mAh g−1 after 100 cycles, corresponding to a capacity retention of 89%. This
suggests that the GO coating layer on the NCM surface enhanced its cycling stability and capacity
retention.
To observe the effect of the coating on the high-rate material cycling performance, the cells
were measured at a current density of 1 C between 2.5 and 4.3 V. The corresponding results are shown
in Fig.8 (b). The discharge capacity of NCM decreased from 154.9 to 115.1 mAh g−1 after 100 cycles,
with a capacity retention of only 74.3%. On the contrary, 2.5 wt% GO/NCM delivered a discharge
capacity of 162.5 mAh g−1 in the first cycle and 136.8 mAhg−1 after 100 cycles, with a corresponding
capacity retention of 84.2%. Comparing the cycle performances at 0.2 C, GO/NCM performed slightly
better than the bare NCM; however, at 1 C, GO/NCM exhibited remarkable improvement compared
with bare NCM. The better cycling performance of GO/NCM can be ascribed to the protection by the
GO coating layer, which reduced the corrosion of the cathode material by the electrolyte and
suppressed side reactions between the active material and electrolyte during the prolonged cycling
[44].
Int. J. Electrochem. Sci., Vol. 13, 2018 10449
(b)
Re Rct Zw
CPE
Figure 10. (a) Nyquist plots of NCM and GO/NCM electrodes after 100cycles. (b) Equivalent circuit
model. Element Freedom Value Error Error %
Re Free(+) 11 1.1213 10.194
Rct bare and Free(+)
The rate performances of the GO/NCM cathodes601.3 are shown
27.734 in Fig.9.
4.6123
The cells were
+
charged and discharged between 2.5CPE-T
and 4.3 VFixed(X)
vs. Li/Li at 0current ratesN/A N/A2 and 5 C. The
of 0.2, 0.5, 1,
rate was increased from 0.2 to 5 CPE-P
C stepwise and Fixed(X) 1
finally returned to 0.2N/AC every fiveN/Acycles at each
Zw-R gradualFixed(X)
current rate. All the cathodes exhibited decreases in0 discharge capacity
N/A as theN/Acurrent density
increased. The 2.5 wt% GO/NCM Zw-T sample Fixed(X)
presented a0 distinctly improved
N/A N/A
discharge capacity
−1
compared with NCM (100.9 vs. 87.1 Zw-PmAh g )Fixed(X)
at a current 0.5 N/AHowever, N/A
density of 5 C. when the current
density was decreased to 0.2 C again, the capacity of GO/NCM almost recovered its initial value, thus
Chi-Squared: 0.39329
Weighted Sum of Squares: 38.542
implying that a high current density and rapid lithiation/delithiation do no effect the crystal structure.
By contrast, the NCM did not recover its initial capacity because of an irreversible capacity loss
resulting from an increase in its inner resistance. The GO introduced into the NCM material improved
the reaction kinetics by increasing the electrical conductivity via the GO 2D network.
EIS measurements were performed to further investigate the possible reasons why the GO-
coated NCM improves the electrochemical performance and reaction kinetics, particularly at high
current rates. Fig.10 displays the impedance spectrum obtained with a coin cell after the 100th charge–
discharge cycle. All these curves consist of a single semicircle in the high-to medium-frequency region
and a quasi-linear line in the low-frequency region [45]. In general, the semicircle in the high-to
medium-frequency region is related to the charge transfer resistance (Rct), and the quasi-linear line in
the low-frequency region is attributed to the Warburg impedance (Zw), which is ascribed to lithium-ion
diffusion in the solid phase of the electrode [46,47]. The equivalent electrical circuit is shown in
Fig.10(b), where Re represents the solution resistance, CPE stands for the double layer capacitance and
Rct represents the charge transfer resistance. The values of Re and Rct were obtained by fitting with
Zview software.
After the 100th cycle, Rct for the pristine samples was 362 Ω. However, for the 2.5 wt%
GO/NCM, Rct was only 247 Ω. This directly demonstrates that a minimum amount GO-coated on the
electrode improved the diffusion of Li ions, increased the electrical conductivity and suppressed the
increase in resistance. Furthermore, the GO-coating layer can inhibit side reactions between the
electrode and electrolyte during the high-rate charge-discharge process, thus slowing down the
increase in the charge transfer resistance (Rct). The large surface area and better electrical conductivity
of GO lead to a reduction in the charge transfer resistance of the coated cathode material, thus
enhancing the electrode kinetics and consequently resulted in a superior electrochemical performance.
To further study the electrical conductivity for the bare and GO/NCM, the diffusion coefficient
of lithium ions can be evaluated according to the following equation [48, 49]:
( )
where R is the gas constant (8.314 Jmol−1K−1), T is the absolute temperature (298.15 K), A is
the surface area of the electrode, n is the number of electrons per molecule during Li ion insertion, F is
the Faraday constant (96486 Cmol−1), C is the lithium-ion concentration and σ is the Warburg factor,
which can be obtained using the following equation:
( )
Re is the solution resistance; Rct is the charge transfer resistance and is the angular frequency.
Fig.11 shows the linear relationship between and −1/2 (square root of frequency) in the low-
frequency region. The lithium-ion diffusion coefficients of NCM and GO/NCM were calculated are
5.29×10−13 and 7.92×10−13 cm2/s.
Int. J. Electrochem. Sci., Vol. 13, 2018 10451
-1/2
Figure 11. The relationship between and (s-1/2) for NCM and GO/NCM.
The improved DLi for the GO/NCM particles due to the large surface area of GO sheets allows
the fabrication of a binder free cathode, which also enhances the electrical conductivity and facilitates
electrolyte through the cathode. Thus, the GO coating layer increases the lithium-ion diffusion
coefficient compared with the NCM samples.
4. CONCLUSIONS
In summary, NCM was coated with GO through a simple chemical approach. The 2.5 wt%
GO/NCM retained its original crystal structure and even exhibited better electrochemical performance
than the pristine material in terms of discharge capacity, high columbic efficiency, rate capacity,
cycling performance and lithium-ion diffusion coefficient. These performance improvements are
mainly ascribed to the GO coating layer, which has a large surface area and better electrical
conductivity. Thus, GO not only prevented the side reaction between the active material and
electrolyte but also stabilised the crystal structure at high current densities. GO-coated NCM is a
promising cathode material candidate for potential applications in electric vehicles and HEVs.
ACKNOWLEDGEMENT
This work was financially supported by the National Science Foundation of China (Grant
NO.51378183) and the Research Program of Hubei Provincial Department of Education (T201215),
and the Petroleum and Chemical Industry Federation Program of China (Grant NO.20170902).
References
1. P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J.M. Tarascon, Nature, 407 (2000) 496.
2. R.F. Service, Science, 313 (2006) 902.
3. C.W. Park, H.S. Ryu, K.W. Kim, J.H. Ahn, J.Y. Lee and H.J. Ahn, J. Power Sources, 165 (2007)
450.
Int. J. Electrochem. Sci., Vol. 13, 2018 10452
45. C.J. Jafta, K.I. Ozoemena, M.K. Mathe and W.D. Roos, Electrochim. Acta, 85 (2012) 411.
46. S.Y. Tan, L. Wang, B. Liang, J.B. Xu, W. Ren, P.F. Hu, et al. J. Power Sources, 277 (2015) 139.
47. M. Wang, M. Luo, Y. Chen, Y. Su, L. Chen and R. Zhang, J. Alloys Compd., 696 (2016) 907.
48. N. Takami, A. Satoh, M. Hara and T. Ohsaki, Cheminform, 26 (1995) .
49. A.J. Bard, L.R. Faulkner, Electrochemical Methods, second ed., Wiley, (2001) New Jersey,
American
© 2018 The Authors. Published by ESG (www.electrochemsci.org). This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/4.0/).