Next Article in Journal
A Design of High-Efficiency: Vertical Accumulation Modulators Based on Silicon Photonics
Previous Article in Journal
Upscale Synthesis of Magnetic Mesoporous Silica Nanoparticles and Application to Metal Ion Separation: Nanosafety Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Scalable Precursor-Assisted Synthesis of a High Voltage LiNiyCo1−yPO4 Cathode for Li-Ion Batteries

1
Department of Energy & Materials Engineering, Dongguk University, Seoul 04620, Republic of Korea
2
U.S.-Pakistan Center for Advanced Studies in Energy, National University of Sciences and Technology (NUST), Islamabad 44000, Pakistan
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(24), 3156; https://doi.org/10.3390/nano13243156
Submission received: 14 November 2023 / Revised: 11 December 2023 / Accepted: 14 December 2023 / Published: 16 December 2023

Abstract

:
A solid-solution cathode of LiCoPO4-LiNiPO4 was investigated as a potential candidate for use with the Li4Ti5O12 (LTO) anode in Li-ion batteries. A pre-synthesized nickel–cobalt hydroxide precursor is mixed with lithium and phosphate sources by wet ball milling, which results in the final product, LiNiyCo1−yPO4 (LNCP) by subsequent heat treatment. Crystal structure and morphology of the product were analyzed by X-ray powder diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and scanning electron microscopy (SEM). Its XRD patterns show that LNCP is primarily a single-phase compound and has olivine-type XRD patterns similar to its parent compounds, LiCoPO4 and LiNiPO4. Synchrotron X-ray absorption spectroscopy (XAS) analysis, however, indicates that Ni doping in LiCoPO4 is unfavorable because Ni2+ is not actively involved in the electrochemical reaction. Consequently, it reduces the charge storage capability of the LNCP cathode. Additionally, ex situ XRD analysis of cycled electrodes confirms the formation of the electrochemically inactive rock salt-type NiO phase. The discharge capacity of the LNCP cathode is entirely associated with the Co3+/Co2+ redox couple. The electrochemical evaluation demonstrated that the LNCP cathode paired with the LTO anode produced a 3.12 V battery with an energy density of 184 Wh kg−1 based on the cathode mass.

1. Introduction

Demand for lithium-ion batteries (LIBs) has been increasing rapidly in portable electronics, electric vehicles, and other industries [1,2,3]. The rapid progression of LIB technology spurred an immediate demand for novel high-energy-density systems. A great deal of work is being put into exploring novel materials that can be used as the cathode for high-voltage LIBs. The first commercialized cathode LiCoO2 has a high operating voltage (~3.9 V) [4]. However, LiCoO2 has been gradually replaced by other commercialized cathode materials, such as spinel LiMn2O4, olivine LiFePO4, and layered-type LiNi1/3Mn1/3Co1/3O2 (NCM111) during the continuous development of LIBs owing to the high cost of Co resources [1,5]. Introducing redox couples with high redox potentials, such as Co2+/Co3+, Ni2+/Ni4+ or V3+/V4+/V5+, can further boost the voltage and energy density [6]. For example, LiNi0.5Mn1.5O4 demonstrates dramatically increased redox potential from 4.0 to 4.7 V by a partial replacement of Mn with Ni in spinel LiMn2O4 lattice [5]. In the case of layered LiNixMnyCozO2 chemistry, where x:y:z = 5:3:2 (NMC532), 6:2:2 (NMC622), 8:1:1 (NMC811), the higher Ni fraction translates into the increased energy density, boosting research attention on Ni-rich layered oxides [7,8]. Alternatively, the olivine family LiMPO4 (M = Fe, Mn, Co) cathodes have become the subject of intensive research thanks to their remarkable thermal stability attributed to the strong covalent P–O bond within the polyanionic structure. Typically, oxide atoms are packed in a slightly distorted hexagonal close packing (HCP) in an olivine structure where half of the octahedral sites contain Li+ and M2+ cations, while 1/8 of the tetrahedral sites contain P5+ cations, which is a hexagonal analogue of spinel. The MO6 octahedrons share four corners in the bc plane and are cross-linked by PO4 groups along the ‘a’ axis. The Li ions reside inside rows of edge-shared LiO6 octahedra aligned along the ‘a’ axis. Li+ ions are occupied by the three-dimensional network of perpendicular channels along the [010]– and [001]– direction. Having this network plays a crucial role in lithium-ion mobility, which makes olivine a viable cathode material [9]. Among them, LiFePO4 is the pioneer, reported in 1997 by Goodenough and co-workers [10]. It has been extensively investigated and successfully commercialized due to its favorable electrochemical properties, exceptional safety, abundance of Fe resources, and environmental friendliness. In addition, LiCoPO4 and LiNiPO4 also emerged as promising cathode materials on the pathway to future high-voltage (5 V) batteries due to their high operating voltage ~4.8 V and ~5.1 V, respectively, which translate to theoretical energy densities of 801 and 862 Wh kg−1 [5]. Nevertheless, the sluggish kinetics of the electronic and lithium-ion transport for these cathodes pose a major constraint to their development. Numerous modifications have been adapted to improve the Li extraction/insertion kinetics and electronic conductivity, including morphology control [11,12], metal ion doping [13], metal oxide coating [14,15,16], and carbon coating [17,18]. Manickam Minakshi et al. have conducted several studies on olivine cathodes [19,20,21,22,23]. A LiCoPO4/C nanocomposite with controlled morphology was synthesized using solid-state fusion by introducing a second phase (Co2P) in an inert atmosphere and utilizing Super P for amorphous carbon coating [19]. A LiCo1/3Mn1/3Ni1/3PO4 cathode, synthesized by sol-gel (SG) and solid-state methods, demonstrated superior performance in aqueous batteries with even microparticle distribution aided by PVP in the SG process [20]. Beyond these approaches, the overall electrochemical performance of any olivine-type materials can be enhanced via solid solutions. Basically, among all the cationic redox couples, Ni+2/+3/+4 is the most significant contributor to boosting the energy density of cathode materials, as demonstrated in the Ni-rich layered oxides [8]. However, it is difficult to synthesize pure phase LiNiPO4 due to the segregation of the Li4P2O7 and Ni3P impurity phases [23]. Alternatively, Ni doping is often widely adopted to influence the structural modification and electrochemical activity of other phospho-olivine cathodes, LiMnPO4 (M = Fe, Mn, and Co). Y. Ge et al. [24] have attributed the Ni doping to reducing the particle size of LiFePO4. Y. Liu et al. [25] stated that doped Ni can enhance the crystallinity of LiFePO4, as evidenced by XRD analysis. Y. Lu et al. [26] stated that doping nickel in LiFePO4/C composites strengthens the P–O bond, makes the structure more stable, and lowers the cathode particles and charge transfer resistance. The theoretical calculation predicted that Ni doping in LiMnPO4 could enhance the insertion potential of Li and reduce activation barriers that inhibit Li diffusion [27]. M. Minakshi and S. Kandhasamy [28] demonstrated improved voltage and capacity in LiMnPO4 with Ni substitutions. D. Shanmukaraj et al. reported that substitution of Ni in LiCoPO4 strengthens the P–O and M–O bonds [29]. Therefore, LiNiPO4–LiCoPO4 solid solutions were also investigated computationally [30] and experimentally [31,32,33,34] as potential cathodes for use in Li-ion batteries. S. T. Rommel et al. [34] have reported the synthesis of the LiNi1−yCoyPO4 (y = 0.25, 0.33, 0.66, 1.0) and observed a stabilization effect of cycle life at an optimum Ni content of 33%.
Various methods have been developed to synthesize those olivine-type LIB cathodes, including co-precipitation [29], sol-gel [32], hydrothermal [35], microwave [36], solid-state [16], supercritical fluid [11,37], polyol [38], and spray pyrolysis [33]. Since many of the available Co and Ni precursors are susceptible to form impurities, such as Co metal, CoO, Ni3P, and Li3PO4, most of the synthesis routes reported cannot be scaled up or require complicated heat-treatment steps to achieve pure stoichiometric LiMPO4 (M = Ni, Co) [16,39]. For example, NH4CoPO4 nanoplates can be used for LiCoPO4 preparation [40]. However, multiple heat treatments in both an air and an inert atmosphere were needed to make sure that the LiCoPO4 was stoichiometric because there is a possibility of cobalt metal impurities being formed by the decomposition of NH4CoPO4. In general, for synthesizing Ni-containing layered cathode materials that are successfully commercialized, the routine procedure entails two steps: precursors are synthesized through co-precipitation reactions, followed by their lithiation through solid-state reactions at high temperatures [41]. In the battery industry, co-precipitation is advantageous because it can incorporate various raw materials, such as chlorides, sulfates, nitrate salts, and organic anions. Carbonate and hydroxide co-precipitations are the most popular methods to prepare the precursors for the layered-type cathodes. Nevertheless, the precursor-assisted synthesis received much less attention for preparing olivine-type LiMPO4 (M = Ni, Co) cathodes [42]. In addition, a post-mortem study of the cycled electrodes is essential to reveal the charge storage mechanism of electrode materials. By analyzing dQ/dV (differential capacity) vs. voltage curves, only one study attempted to unravel the electrochemical behavior of LiNiPO4–LiCoPO4 solid-solution cathodes [34]. X-ray absorption spectroscopy (XAS) is considered one of the best tools in the scientific community for studying the electrochemical processes in battery materials. Its main important characteristics include (i) its element specificity, enabling a particular element to be studied by concentrating on its K (or, in some cases, L) absorption edge; (ii) the flexibility to customize it to different sites (such as Co and P in LiCoPO4), providing complementary information on the same compound. The XAS technique can track changes in oxidation states and local environmental conditions of metal atoms in battery materials. When X-rays are exposed to the target sample, a 1s photoelectron is excited into low-lying unoccupied states of the central atom at the K-edge, resulting in a normalized absorbance at specific energy [43].
On the anode side, a safety concern associated with Li dendrite formation in conventional graphitic anodes has sparked researchers to look at alternatives, such as Li4Ti5O12 (LTO). Lithium dendrite growth risk and solid electrolyte interface (SEI) formation can ultimately be elevated in LTO due to its lower-lying energy states of Ti3+/4+ redox couple than the LUMO (lowest unoccupied molecular orbital) level of electrolytes, leading to improved safety [44]. However, the redox potential (1.55 V vs. Li/Li+) of LTO reduces the overall working voltage in practical cell configurations. LiFePO4/LTO and Li(Ni1/3Co1/3Mn1/3)O2 (NCM111)/LTO full-cell batteries, for example, have offered cell voltages of only 1.85 and 2.2 V, respectively [45,46]. For the LTO anode, therefore, it may be appropriate to use “5 V-class cathodes”—Li2CoPO4F, LiCoPO4, LiNiPO4, and a LiCoPO4-LiNiPO4 solid solution—to secure a high output potential.
In this paper, LiNiyCo1−yPO4 (y = 0.33) has been synthesized by a low-energy ball mill using a pre-synthesized nickel–cobalt hydroxide precursor. The precursor was prepared by a hydroxide co-precipitation method. In addition to offering the advantages of mass production, the proposed synthesis strategy provided better homogeneity in the product. Moreover, the charge compensation mechanism of LiNiyCo1−yPO4 (y = 0.33) solid solution is revealed using synchrotron X-ray absorption spectroscopy (XAS). To date, no publication has reported the precursor-based synthesis of the LiCoPO4-LiNiPO4 solid-solution cathode, including the electrochemical performance test in a full-cell configuration in combination with the charge compensation mechanism revelation in a LiPF6-containing electrolyte. Our study reveals that incorporation of Ni into the LiCoPO4 olivine system is not worthy, as the Ni2+/Ni3+ redox couple was not activated during electrochemical cycling, which adversely affected the cathode-specific capacity.

2. Experimental Section

2.1. Synthesis

The precursor, [Ni0.33Co0.67](OH)2, was synthesized by co-precipitation method. The starting materials, NiSO4∙6H2O (Kanto Co., Tokyo, Japan) and CoSO4∙6H2O (Kanto Co., Tokyo, Japan) aqueous solutions, were pumped into a continuously stirred tank reactor (CSTR, 4 L) under a N2 atmosphere. The reactor was also simultaneously filled with NaOH solution (aq) in addition to the desired amount of NH4OH solution (aq) to act as a chelate. We carefully controlled the pH (11–12), temperature (60 °C), and stirring speed (1000 rpm) of the mixture in the reactor. After washing and filtering with distilled water, the resultant precursor powders were dried in a vacuum oven overnight. Finally, the obtained pink precursor powders were ball milled (at 150 rpm) with the ZrO2 media in ethanol for 12 h with Li2CO3, and NH4H2PO4. After the slurry was collected, it was dried in an oven at 80 °C for 12 h. The mixture was transferred to a vacuum-controlled box furnace equipped with flowing inert gas and calcined at over 700 °C at a heating rate of 2 °C min−1 for 8 h in a reductive (Ar 96% + H2 4%) atmosphere to yield LiNi0.33Co0.67PO4, the final product. A solid-state approach was used to prepare the LTO anode. To achieve a uniform mixture, stoichiometric amounts of Li2CO3 (Sigma-Aldrich, Burlington, MA, USA, 99%) and TiO2 (Junsei Chemical, Tokyo, Japan, 98.5%) were mixed in a mortar with a pestle. We added 5% extra Li2CO3 (wt.%) to the mixture to compensate for the possibility of Li vaporization during the calcination process. For homogeneous mixing, the mixture was ground repeatedly and then pressed into pellets at 300 bars by using a hydraulic press (Daehatech, Busan, Republic of Korea, HP-1B). As-prepared pellets were placed in a box furnace (MTI corporation, Richmond, CA, USA) and calcined at two stages with different temperatures, 850 and 950 °C, for 12 h each in air with an intermediate grinding step. The crystal structure and phase purity of the as-synthesized LTO anode were confirmed by X-ray diffraction analysis. The morphology of the LTO sample (inspected by Field emission-scanning electron microscopy, Fe-SEM) was micrometer-sized particles with an average diameter (primary particle) of 1.81 μm, as published in our previous research [47,48].

2.2. Electrochemical Measurements

We made the slurry of LNCP cathode containing the synthesized active material, carbon black (Super P), and polyvinylidene fluoride (PVDF) binder (Sigma-Aldrich, Burlington, MA, USA) with a weight ratio of 8:1:1 in N-methyl-2-pyrrolidone (NMP). The mixed slurry was cast onto an Al foil current collector. The active material loading of cathode used for this study was 2.8 mg cm−2. Before the electrode was used, it was dried in a vacuum oven at 80 °C for 12 h. The prepared working electrode was assembled into coin-type half cells (CR2032) in combination with a Li counter electrode, a polypropylene separator, and an electrolyte composed of 1 M LiPF6 in ethylene carbonate and ethyl methyl carbonate (3:7 volume ratio).
More details about structural characterization, LTO slurry making, full-cell assembly, electrochemical testing, ex situ sample preparation, and XAS measurement are described in Supplementary Materials.

3. Results and Discussion

The XRD result of LiNi0.33Co0.67PO4 (LNCP) material in Figure 1 exhibits well-defined diffraction peaks, matching with the standard data (ICSDS 00-000-1809) having the orthorhombic olivine structure with a Pnma space group. The sharp peaks reveal the high crystallinity of the material. A previous study observed lithium phosphate (Li3PO4) impurity phase formation when Mn is doped at the Co site [14]. Unlike this, no impurity peak was observed after the incorporation of the Ni at the Co site in the LiCoPO4 structure. The Ni3P is readily formed as a side phase during the synthesis of the LiNiPO4 material [32]. Herein, by maintaining a reductive atmosphere during heat treatment, Ni2+ and Co2+ are prevented from oxidizing into Ni3+ and Co3+, respectively, assuring that samples are impurity-free. Therefore, the XRD pattern signifies the formation of single-phase material.
The surface composition and valence state of the elements in the LNCP cathode were examined by X-ray photoelectron spectrometry (XPS). In Figure 2a, the Co 2p spectrum displays a typical spin-orbital split (2p3/2 and 2p1/2) with two shake-up satellite peaks at ~787.62 and 805.41 eV; the main peaks at 783.1 and 797.5 eV are attributed to Co2+ [49,50]. The Ni 2p region in Figure 2b exhibits two shake-up satellite peaks; the fitting peaks are located at the binding energies of 859 and 876.5 eV, indicating the oxidation states of Ni2+ [51]. The XPS spectrum for P 2p (Figure 2c) displays that the single binding energies peaked at 135.5 eV, which is assigned to the pentavalent P5+; this corresponds to the presence of the PO43− anion of LNCP [19]. Two peaks in the O 1s spectrum (Figure 2d) at ~530.13 and ~531.26 eV could be associated with metal phosphate and oxygen in chemically or physically adsorbed water molecules. The XPS spectrum of the Li 1s peak with binding energy 55.97 eV is shown in Figure 2e, confirming the presence of Li+ in LNCP.
The sample morphology was observed by FE-SEM, showing randomly distributed microparticles. The particles are loosely agglomerated without noticeable variation in their shape (Figure 3a). However, a closer inspection (Figure 3b) reveals two size distributions: small particles with a diameter range of about 200–400 nm appear on the surface of larger particles with an average size of 1–1.2 µm. It is evident that well-crystallized particles with a noticeable grain boundary and smooth surfaces are formed after calcination. During the electrochemical reaction, the loosely agglomerated particles allow the electrolyte to infiltrate into interparticle voids, ensuring proper utilization of the active material. Energy-dispersive X-ray spectroscopy (EDS) coupled with SEM was conducted to check the elemental distribution of the as-prepared LNCP cathode. In Figure 3c, SEM-EDS mapping results show that all elements, including Ni and Co, are homogeneously distributed within the sample. It is found that the atomic ratio of Ni and Co is 1:2 (Table S1).
Figure 4a shows the electrochemical response of the LNCP cathode in the cyclic voltammetry (CV) test at a scan rate of 0.1 mV s−1 with a lithium foil counter electrode. The first anodic and cathodic sweep demonstrates the oxidation peaks at 5.19 V and 5.14 V and the reduction peaks at 4.66 V and 4.53 V, which were moved and merged on successive cycles at 5.05 V and 4.65 V, respectively. From the second cycle, the CV curve features almost the shape of the pure LiCoPO4 sample, as observed in a previous report [15,31]. Otherwise, the oxidation and reduction peaks of Co2+/Co3+ and Ni2+/Ni3+ might be so close to each other such that only single oxidation and reduction peaks appear in the following cycles. The respective charge and discharge curves of the LNCP cathode (Figure 4b) between 5.2 to 3.5 V over the first five cycles are coherent with the CV curves. As expected, divergent voltage plateaus emerged at variable cell potentials during the first and subsequent cycles. For lithium extraction, LiCoPO4 follows a two-phase mechanism for both aqueous [21] and non-aqueous [52] electrolytes. Our previous study demonstrated two distinguishable flat plateaus during the galvanostatic charge of LiCoPO4 indicative of a two-phase mechanism [53]. It is difficult to distinguish such plateaus from the second charge curve of the LNCP cathode here. Nevertheless, two anodic peaks in the dQ/dV plot (Figure S1) of the second cycle support a two-phase lithium extraction mechanism (Note: The extensive decomposition of the electrolyte associated with SEI layer formation on the electrode surface makes it difficult to discern anodic peaks on first cycle dQ/dV plot). The second-cycle charging and discharging capacities of the LCNP at 0.1 C (1 C = 167 mAh g−1) are 100.4 and 82 mAh g−1, respectively. The electrolyte decomposition and SEI layer formation at high voltages likely result in a substantial irreversible capacity loss and low (~51%) coulombic efficiency (CE) in the first cycle. However, there is a significant improvement in CE (i.e., 81.7, 83.1, 86.9, 88%) from the second to fifth cycles. After an initial activation period lasting ≈ 20 cycles, the CE reached over 95%, and the reversibility of the electrode is excellent for at least 50 cycles, highlighting that the target material can sustain a stable lifespan (Figure 4c). After 50 cycles, the discharge capacity values maintain ~48 mAh g−1. The electrolyte composition likely plays a significant role in LCNP capacity fading, which remains for further investigations. The oxidative decomposition of LiPF6-containing electrolytes was reported for the LiCoPO4 and other high-voltage materials [54,55]. Applying a sulfone-based electrolyte [56] or incorporating HF scavenger separators [57] into this high-voltage cathode would be favorable from the electrochemical performance standpoint. Generally, a LiNiPO4 cathode exhibits a flat plateau at approximately 5 V, corresponding to the Ni3+/Ni2+ redox couple [58]. The LNCP cathode shows a discharge voltage plateau at about 4.8 V, which is slightly lower considering the expected Ni3+/Ni2+ redox couple but is in good agreement with previous publications for the Co3+/Co2+ redox couple in LiCoPO4 [11,13]. When an LNCP cathode is combined with an LTO anode, this full cell displays a discharge plateau at 3.12 V in the voltage profile during the galvanostatic cycling test (Figure 4d). For the full cell, the described potential is not vs. Li/Li+; it is instead the potential difference between an LNCP cathode and an LTO anode. The practical capacity of as-synthesized LTO anode was 136 mAh g−1 at 0.1 C (1 C = 175 mAh g−1) in a half cell [48]. By considering the anode-to-cathode capacity ratio, the active mass of the assembled full-cell electrodes was regulated to ensure that the cathode material was fully utilized. The full cell shows a maximum specific discharge capacity of 64 mAh g−1 at 0.1 C (16.7 mA g−1), which corresponds to a specific energy of 184 Wh kg−1. These values are based on the weight of the cathodic active material rather than the entire device; the purpose is to highlight the material performance of LNCP. The energy density is determined from the integrals of the second discharge curve. The full cell retains a specific capacity value of 43 mAh g−1 after 50 cycles, as shown in Figure 4d (inset). It is important to note that the capacity retention of LNCP in a full cell is shown to be better than that of a half cell.
Despite the high working potential of the presented cathodic material, its low specific capacity results in relatively low specific energy compared to the cathodic materials (LiCoPO4) used in pairing with LTO [59,60]. To understand the causes of the low specific capacity, the charge compensation mechanism of the LNCP cathode was investigated using ex situ X-ray absorption spectroscopy (XAS). We examined the changes in the oxidation state of nickel and cobalt atoms throughout the electrochemical reactions using XANES. The ATHENA software package (version 0.9.26) was used to handle and process the XAS data [61]. There is an absorption edge that results from a dipole-allowed 1s → 4p transition, which is commonly called the white line, and the oxidation state change of the absorbing atom is described by its relative energy position shift. Figure 5a,b illustrates X-ray absorption near edge structure (XANES) spectra at the Co and Ni K-edges during charging to 5.2 V (full CC) and discharging to 3.5 V (full DCh). Delithiation (charge) leads to the entire edge shift of the Co K-edge toward higher energy, suggesting that Co2+ is being oxidized to Co3+. During lithiation (discharge), the Co K-edge XANES spectrum shows reversible edge shifts back to its pristine state, suggesting quite reversible Co2+/3+ redox reaction in the LNCP during charging and discharging. As opposed to this, the position of the Ni K-edge (Figure 5b) remains unchanged during charging and discharging. In light of this finding, it can be assured that the capacity contribution to the LNCP cathode is due to cobalt oxidation and reduction. The inactivity of the Ni redox couple explains the lower specific capacity of the target material. Figure 5c,d presents Fourier-transformed intensity of extended X-ray absorption fine structure (FT-EXAFS) spectra of LNCP showing interatomic distance changing behavior at the Co and Ni K edges. Typically, an FT peak represents the bond distance between the absorbing and backscattering atoms. At R = 1.5 Å, the first peak is attributed to the scattering of the nearest oxygen atom in the first coordination shell of the TM-O6 octahedron. At R = 2.45 Å, the second peak is associated with the scattering of the second nearest TM cation in the second coordination shell of the Co/Ni-TM6 hexagon. For the EXAFS process, the k-range was 3.0~11 Å−1 with a k-weight factor of 3. After complete charge to 5.2 V, the interatomic Co–O, and Co–TM distances decreased. Upon subsequent discharge, cobalt interatomic distances and peak intensities were completely restored to their original pristine (OCV) state, indicating the highly reversible local structural changes of cobalt in LNCP. Conversely, the Ni K-edge EXAFS spectra conform well to the NiO rock salt structure [62]. The Ni–O and Ni–TM interatomic distances remain unchanged, indicating that the inactive NiO rock salt phase adversely affects the charge storage capability in the LNCP cathode. Thus, as determined by XANES, the Ni2+ ions do not participate in electrochemical reactions, which is consistent with the interpretation of the FT-EXAFS results.
After the charge/discharge cycle, ex situ X-ray diffraction (XRD) data were collected to obtain clues about the structural changes in the LNCP cathode as presented in Figure 6. It shows that all the main reflection peaks are maintained during charge–discharge cycles. Intense peaks are seen, which indicates suppression of amorphization of the LNCP cathode [63]. However, there is an additional peak at 2θ = 43.8° related to the NiO phase formation besides those for the LNCP material. This result confirms the detrimental rock salt phase (NiO) formation in the LNCP cathode [64]. Notably, S. M. Rommel et al. [34] also reported the inactivity of a Ni2+/Ni3+ redox couple in a LiCoPO4-LiNiPO4 solid solution cathode. However, the cause of the Ni content inactivity has not yet been identified. The present study has utilized a combination of ex situ XRD and synchrotron XAS techniques to reveal the underlying cause of this phenomenon. There is no doubt that the common polyanionic materials (LiMPO4) have structures that are generally stable with larger M cations (Fe2+, Co2+, Ni2+, and Mn2+, whose radii are 0.78, 0.72, 0.78, and 0.64 Å, respectively). The extremely small radius of the Ni3+ cation (0.56 Å) probably does not fit into the olivine structure’s octahedral environment or may cause strong structural distortion. Therefore, the electrochemical oxidation of the Ni2+ in LiNiyCo1−yPO4 material is most likely extremely difficult, at least under the standard conditions applied in this study.

4. Conclusions

A LiCoPO4-LiNiPO4 solid-solution cathode was synthesized using a pre-synthesized metal hydroxide precursor followed by wet ball milling with lithium (Li2CO3) and phosphate (NH4H2PO4) sources and subsequent heat treatment. The XRD result initially suggested that LiNiyCo1−yPO4 (y = 0.33) solid solutions were formed. The XPS results confirmed that in the pristine sample, the oxidation states of Co and Ni are +2. However, the results of synchrotron X-ray absorption spectroscopy (XAS) analysis of cycled electrodes indicate that Ni was not entirely doped into the structure of LiCoPO4. The cathode delivered low discharge capacity because the electrochemical activity of the Ni2+/Ni3+ redox couple was not activated due to the formation of electrochemically inactive rock salt-type NiO. Both cyclic voltammetry and discharge curves indicate that the LNCP cathode is driven by one redox couple, Co3+/Co2+, as confirmed by the XANES and EXAFS results. Therefore, substitution of Co2+ in LiCoPO4 by Ni2+ ion, in fact, negatively affects the cathode performance.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano13243156/s1, Figure S1. The dQ/dV curves of LiNiyCo1−yPO4 cathode in a half cell at (a) initial and (b) second cycles. Figure S2. Voltage profile of the LNCP/graphite full cells: (a) without pre-lithiation and (b) with pre-lithiated anode. Table S1. SEM-EDS analysis result for the LNCP cathode.

Author Contributions

Conceptualization, M.I.; data curation, G.A., S.Y. and H.A.; formal analysis, G.A., D.H. and M.F.; funding acquisition, K.-W.N.; investigation, M.I., D.H., M.F., S.Y. and B.A.; resources, K.-W.N.; supervision, K.-W.N.; methodology, B.A.; validation, H.A.; visualization, M.I.; writing—original draft, M.I.; writing—review & editing, K.-W.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Nano & Material Technology Development Program through the National Research Foundation of Korea (NRF) funded by Ministry of Science and ICT (RS-2023-00282389 and 2022R1A2C2009459).

Data Availability Statement

All the data analyzed in this study are available on request.

Acknowledgments

We acknowledge the kind technical support received from Hun Gi Jung, Center for Energy Storage Research, Korea Institute of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, C.; Dai, Q.; Gaines, L.; Hu, M.; Tukker, A.; Steubing, B. Future material demand for automotive lithium-based batteries. Commun. Mater. 2020, 1, 99. [Google Scholar] [CrossRef]
  2. Olivetti, E.A.; Ceder, G.; Gaustad, G.G.; Fu, X. Lithium-ion battery supply chain considerations: Analysis of potential bottlenecks in critical metals. Joule 2017, 1, 229–243. [Google Scholar] [CrossRef]
  3. Liang, Y.; Zhao, C.-Z.; Yuan, H.; Chen, Y.; Zhang, W.; Huang, J.-Q.; Yu, D.; Liu, Y.; Titirici, M.-M.; Chueh, Y.-L.; et al. A review of rechargeable batteries for portable electronic devices. InfoMat 2019, 1, 6–32. [Google Scholar] [CrossRef]
  4. Mizushima, K.; Jones, P.C.; Wiseman, P.J.; Goodenough, J.B. LixCoO2 (0 < x < −1): A new cathode material for batteries of high energy density, Mater. Res. Bull. 1980, 15, 783–789. [Google Scholar]
  5. Li, W.; Song, B.; Manthiram, A. High-voltage positive electrode materials for lithium-ion batteries. Chem. Soc. Rev. 2017, 46, 3006–3059. [Google Scholar] [CrossRef] [PubMed]
  6. Jeong, G.; Kim, Y.-U.; Kim, H.; Kim, Y.-J.; Sohn, H.-J. Prospective materials and applications for Li secondary batteries. Energy Environ. Sci. 2011, 4, 1986–2002. [Google Scholar] [CrossRef]
  7. Li, W.; Lee, S.; Manthiram, A. High-Nickel NMA: A Cobalt-free alternative to NMC and NCA Cathodes for Lithium-ion Batteries. Adv. Mater. 2020, 32, 2002718. [Google Scholar] [CrossRef] [PubMed]
  8. Mohamed, N.; Allam, N.K. Recent advances in the design of cathode materials for Li-ion batteries. RSC Adv. 2020, 10, 21662–21685. [Google Scholar] [CrossRef]
  9. Fisher, C.A.J.; Hart Prieto, V.M.; Islam, M.S. Lithium battery materials LiMPO4 (M = Mn, Fe, Co, and Ni): Insights into defect association, transport mechanisms, and doping behavior. Chem. Mater. 2008, 20, 5907–5915. [Google Scholar] [CrossRef]
  10. Padhi, A.K.; Nanjundaswamy, K.S.; Goodenough, J.B. Phospho-olivines as positive-electrode materials for rechargeable lithium batteries. J. Electrochem. Soc. 1997, 144, 1188–1194. [Google Scholar] [CrossRef]
  11. Truong, Q.D.; Devaraju, M.K.; Honma, I. Benzylamine-directed growth of olivine-type LiMPO4 nanoplates by a supercritical ethanol process for lithium-ion batteries. J. Mater. Chem. A 2014, 2, 17400–17407. [Google Scholar] [CrossRef]
  12. Manzi, J.; Curcio, M.; Brutti, S. Structural and morphological tuning of LiCoPO4 materials synthesized by solvo-thermal methods for Li-cell applications. Nanomaterials 2015, 5, 2212–2230. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Y.; Qiu, J.; Li, M.; Zhu, X.; Wen, Y.; Li, B. One-Step Synthesis of LiCo1−1.5xYxPO4@C cathode material for high-energy lithium-ion batteries. Materials 2022, 15, 7325. [Google Scholar] [CrossRef] [PubMed]
  14. Örnek, A.; Can, M.; Yeşildağ, A. Improving the cycle stability of LiCoPO4 nanocomposites as 4.8 V cathode: Stepwise or synchronous surface coating and Mn substitution. Mater. Charact. 2016, 116, 76–83. [Google Scholar] [CrossRef]
  15. Eftekhari, A. Surface modification of thin-film based LiCoPO4 5 V cathode with metal oxide. J. Electrochem. Soc. 2004, 151, A1456. [Google Scholar] [CrossRef]
  16. Tolganbek, N.; Yerkinbekova, Y.; Kalybekkyzy, S.; Bakenov, Z.; Mentbayeva, A. Current state of high voltage olivine structured LiMPO4 cathode materials for energy storage applications: A review. J. Alloys Compd. 2021, 882, 160774. [Google Scholar] [CrossRef]
  17. Kumar, P.R.; Rao, V.M.; Nageswararao, B.; Venkateswarlu, M.; Satyanarayana, N. Enhanced electrochemical performance of carbon-coated LiMPO4 (M = Co and Ni) nanoparticles as cathodes for high-voltage lithium-ion battery. J. Solid State Electrochem. 2016, 20, 1855–1863. [Google Scholar] [CrossRef]
  18. Örnek, A.; Can, M.; Yeşildağ, A. Analysis of the effects of different carbon coating strategies on structure and electrochemical behavior of LiCoPO4 material as a high-voltage cathode electrode for lithium-ion batteries. Electrochim. Acta 2018, 279, 108–117. [Google Scholar]
  19. Gangulibabu; Nallathamby, K.; Meyrick, D.; Minakshi, M. Carbonate anion anion-controlled growth of LiCoPO4/C nanorods and its improved electrochemical behavior. Electrochim. Acta 2013, 101, 18–26. [Google Scholar] [CrossRef]
  20. Kandhasamy, S.; Pandey, A.; Minakshi, M. Polyvinylpyrrolidone assisted sol-gel route LiCo1/3Mn1/3Ni1/3PO4 composite cathode for aqueous rechargeable battery. Electrochim. Acta 2012, 60, 170–176. [Google Scholar] [CrossRef]
  21. Minakshi, M.; Singh, P.; Sharma, N.; Blackford, M.; Ionescu, M. Lithium extraction−insertion from/into LiCoPO4 in aqueous batteries. Ind. Eng. Chem. Res. 2011, 50, 1899–1905. [Google Scholar] [CrossRef]
  22. Minakshi, M.; Singh, P.; Ralph, D.; Appadoo, D.; Blackford, M.; Ionescu, M. Structural characteristics of olivine Li(Mg0.5Ni0.5)PO4 via TEM analysis. Ionics 2012, 18, 583–590. [Google Scholar] [CrossRef]
  23. Minakshi, M.; Singh, P.; Appadoo, D.; Martin, D.E. Synthesis, and characterization of olivine LiNiPO4 for aqueous rechargeable battery. Electrochim. Acta 2011, 56, 4356–4360. [Google Scholar] [CrossRef]
  24. Ge, Y.; Yan, X.; Liu, J.; Zhang, X.; Wang, J.; He, X.; Wang, R.; Xie, H. An optimized Ni doped LiFePO4/C nanocomposite with excellent rate performance. Electrochim. Acta 2010, 55, 5886–5890. [Google Scholar] [CrossRef]
  25. Liu, Y.; Gu, Y.-J.; Luo, G.-Y.; Chen, Z.-L.; Wu, F.-Z.; Dai, X.-Y.; Mai, Y.; Li, J.-Q. Ni-doped LiFePO4/C as high-performance cathode composites for Li-ion batteries. Ceram. Int. 2020, 46, 14857–14863. [Google Scholar] [CrossRef]
  26. Lu, Y.; Shi, J.; Guo, Z.; Tong, Q.; Huang, W.; Li, B. Synthesis of LiFe1−xNixPO4/C composites and their electrochemical performance. J. Power Sources 2009, 194, 786–793. [Google Scholar] [CrossRef]
  27. Sgroi, M.F.; Lazzaroni, R.; Beljonne, D.; Pullini, D. Doping LiMnPO4 with cobalt and nickel: A first principle study. Batteries 2017, 3, 11. [Google Scholar] [CrossRef]
  28. Minakshi, M.; Kandhasamy, S. Utilizing active multiple dopants (Co and Ni) in olivine LiMnPO4. Curr. Opin. Solid State Mater. Sci. 2012, 16, 163–167. [Google Scholar] [CrossRef]
  29. Shanmukaraj, D.; Murugan, R. Synthesis, and characterization of LiNiyCo1−yPO4 (y = 0–1) cathode materials for lithium secondary batteries. Ionics 2004, 10, 88–92. [Google Scholar] [CrossRef]
  30. Ping, L.Z.; Jun, Z.Y.; Ming, Z.Y. Li-site and metal-site ion doping in phosphate-olivine LiCoPO4 by first-principles calculation, Chin. Phys. Lett. 2009, 26, 038202. [Google Scholar]
  31. Wolfenstine, J.; Allen, J. LiNiPO4–LiCoPO4 solid solutions as cathodes. J. Power Sources 2004, 136, 150–153. [Google Scholar] [CrossRef]
  32. Örnek, A.; Bulut, E.; Can, M. Influence of gradual cobalt substitution on lithium nickel phosphate nano-scale composites for high voltage applications. Mater. Charact. 2015, 106, 152–162. [Google Scholar] [CrossRef]
  33. Li, Y.; Taniguchi, I. Synthesis of LiNi1-xCoxPO4/C nanocomposite cathode for lithium-ion batteries by a combination of aerosol and powder technologies. Adv. Powder Technol. 2019, 30, 180–189. [Google Scholar] [CrossRef]
  34. Rommel, S.M.; Rothballer, J.; Schall, N.; Brünig, C.; Weihrich, R. Characterization of the carbon-coated LiNi1−y CoyPO4 solid solution synthesized by a non-aqueous sol-gel route. Ionics 2015, 21, 325–333. [Google Scholar] [CrossRef]
  35. Neef, C.; Meyer, H.-P.; Klingeler, R. Morphology-controlled two-step synthesis and electrochemical studies on hierarchically structured LiCoPO4. Solid State Sci. 2015, 48, 270–277. [Google Scholar] [CrossRef]
  36. Murugan, A.V.; Muraliganth, T.; Ferreira, P.J.; Manthiram, A. Dimensionally modulated, single-crystalline LiMPO4 (M = Mn, Fe, Co, and Ni) with nano-thumblike shapes for high-power energy storage. Inorg. Chem. 2009, 48, 946–952. [Google Scholar] [CrossRef] [PubMed]
  37. Devaraju, M.K.; Truong, Q.D.; Hyodo, H.; Sasaki, Y.; Honma, I. Synthesis, characterization, and observation of antisite defects in LiNiPO4 nanomaterials. Sci. Rep. 2015, 5, 11041. [Google Scholar] [CrossRef]
  38. Karthickprabhu, S.; Hirankumar, G.; Maheswaran, A.; Sanjeeviraja, C.; Bella, R.D. Structural and conductivity studies on LiNiPO4 synthesized by the polyol method. J. Alloy. Compd. 2013, 548, 65–69. [Google Scholar] [CrossRef]
  39. Rommel, S.M.; Schall, N.; Brünig, C.; Weihrich, R. Challenges in the synthesis of high voltage electrode materials for lithium-ion batteries: A review on LiNiPO4. Monatsh Chem. 2014, 145, 385–404. [Google Scholar] [CrossRef]
  40. Oh, S.-M.; Myung, S.-T.; Sun, Y.-K. Olivine LiCoPO4-carbon composite showing high rechargeable capacity. J. Mater. Chem. 2012, 22, 14932–14937. [Google Scholar] [CrossRef]
  41. Wang, D.; Belharouak, I.; Ortega, L.H.; Zhang, X.; Xu, R.; Zhou, D.; Zhou, G.; Amine, K. Synthesis of high-capacity cathodes for lithium-ion batteries by morphology-tailored hydroxide co-precipitation. J. Power Sources 2015, 274, 451–457. [Google Scholar] [CrossRef]
  42. Koleva, V.; Zhecheva, E.; Stoyanova, R. Ordered olivine-type lithium–cobalt and lithium–nickel phosphates prepared by a new precursor method. Eur. J. Inorg. Chem. 2010, 2010, 4091–4099. [Google Scholar] [CrossRef]
  43. Giorgetti, M.; Stievano, L. X-ray absorption spectroscopy study of battery materials. In X-ray Characterization of Nanostructured Energy Materials by Synchrotron Radiation, 1st ed.; Khodaei, M., Petaccia, L., Eds.; IntechOpen: London, UK, 2017; Chapter 4. [Google Scholar] [CrossRef]
  44. Robertson, A.; Trevino, L.; Tukamoto, H.; Irvine, J. New inorganic spinel oxides for use as negative electrode materials in future lithium-ion batteries. J. Power Sources 1999, 81, 352–357. [Google Scholar] [CrossRef]
  45. Yang, C.-C.; Hu, H.-C.; Lin, S.J.; Chien, W.-C. Electrochemical performance of V-doped spinel Li4Ti5O12/C composite anode in Li-half and Li4Ti5O12/LiFePO4-full cell. J. Power Sources 2014, 258, 424–433. [Google Scholar] [CrossRef]
  46. Guo, X.; Xiang, H.F.; Zhou, T.P.; Li, W.H.; Wang, X.W.; Zhou, J.X.; Yu, Y. Solid-state synthesis, and electrochemical performance of Li4Ti5O12/graphene composite for lithium-ion batteries. Electrochim. Acta 2013, 109, 33–38. [Google Scholar] [CrossRef]
  47. Ali, B.; Muhammad, R.; Islam, M.; Anang, D.A.; Han, D.-S.; Moeez, I.; Chung, K.Y.; Cho, M.K.; Kim, J.-Y.; Kim, M.-G.; et al. Cd-Doped Li4–xCdxTi5O12 (x = 0.20) as a high rate capable and stable anode material for lithium-ion batteries. ACS Appl. Energy Mater. 2023, 6, 4198–4210. [Google Scholar] [CrossRef]
  48. Ali, B.; Muhammad, R.; Anang, D.A.; Cho, M.K.; Kim, J.-Y.; Nam, K.-W. Ge-doped Li4Ti 5-xGexO12 (x= 0.05) as a fast-charging, long-life bi-functional anode material for lithium-and sodium-ion batteries. Ceram. Int. 2020, 46, 16556–16563. [Google Scholar] [CrossRef]
  49. Belgibayeva, A.; Nagashima, T.; Taniguchi, I. Synthesis of LiCoPO4/C nanocomposite fiber mats as free-standing cathode materials for lithium-ion batteries with improved electrochemical properties. Electrochim. Acta 2022, 419, 140400. [Google Scholar] [CrossRef]
  50. Alex, C.; Sarma, S.C.; . Peter, S.C.; John, N.S. Competing effect of Co3+ reducibility and oxygen-deficient defects toward high oxygen evolution activity in Co3O4 systems in alkaline medium. ACS Appl. Energy Mater. 2020, 3, 5439–5447. [Google Scholar] [CrossRef]
  51. Mirghni, A.A.; Madito, M.J.; Oyedotun, K.O.; Masikhwa, T.M.; Ndiaye, N.M.; Ray, S.J.; Manyala, N. A high energy density asymmetric supercapacitor utilizing a nickel phosphate/graphene foam composite as the cathode and carbonized iron cations adsorbed onto polyaniline as the anode. RSC Adv. 2018, 8, 11608–11621. [Google Scholar] [CrossRef]
  52. Kaus, M.; Issac, I.; Heinzmann, R.; Doyle, S.; Mangold, S.; Hahn, H.; Chakravadhanula, V.S.K.; Kübel, C.; Ehrenberg, H.; Indris, S. Electrochemical delithiation/relithiation of LiCoPO4: A two-step reaction mechanism investigated by in situ X-ray diffraction, in situ X-ray absorption spectroscopy, and ex situ 7Li/31P NMR spectroscopy. J. Phys. Chem. C 2014, 118, 17279–17290. [Google Scholar] [CrossRef]
  53. Islam, M.; Akbar, M.; Ali, G.; Nam, K.W.; Chung, K.Y.; Jung, H.-G. A high voltage Li-ion full-cell battery with MnCo2O4/LiCoPO4 electrodes. Ceram. Int. 2020, 46, 26147–26155. [Google Scholar] [CrossRef]
  54. Markevich, E.; Sharabi, R.; Gottlieb, H.; Borgel, V.; Fridman, K.; Salitra, G.; Aurbach, D.; Semrau, G.; Schmidt, M.A.; Schall, N.; et al. Reasons for capacity fading of LiCoPO4 cathodes in LiPF6 containing electrolyte solutions. Electrochem. Commun. 2012, 15, 22–25. [Google Scholar] [CrossRef]
  55. Manzi, J.; Brutti, S. Surface chemistry on LiCoPO4 electrodes in lithium cells: SEI formation and self-discharge. Electrochim. Acta 2016, 222, 1839–1846. [Google Scholar] [CrossRef]
  56. Xue, W.; Huang, M.; Li, Y.; Zhu, Y.G.; Gao, R.; Xiao, X.; Zhang, W.; Li, S.; Xu, G.; Yu, Y.; et al. Ultra-high-voltage Ni-rich layered cathodes in practical Li metal batteries enabled by a sulfonamide-based electrolyte. Nat. Energy 2021, 6, 495–505. [Google Scholar] [CrossRef]
  57. Li, P.; Wang, Y.; Liu, Z.; Hu, X. Acid-scavenging separators promise long-term cycling stability of lithium-ion batteries. Mater. Chem. Front. 2023, 7, 6318–6344. [Google Scholar] [CrossRef]
  58. Örnek, A. Optimization of dielectric heating parameters in the production of high-voltage LiNiPO4-core and carbon-shell ceramics. J. Am. Ceram. Soc. 2017, 100, 5668–5680. [Google Scholar] [CrossRef]
  59. Ni, J.; Liu, W.; Liu, J.; Gao, L.; Chen, J. Investigation on a 3.2 V LiCoPO4/Li4Ti5O12 full battery. Electrochem. Commun. 2013, 35, 1–4. [Google Scholar] [CrossRef]
  60. Du, C.Q.; Tang, Z.Y.; Wu, J.W.; Tang, H.Q.; Zhang, X.H. A three-volt lithium-ion battery with LiCoPO4 and zero-strain Li4Ti5O12 as insertion material. Electrochim. Acta 2014, 125, 58–64. [Google Scholar] [CrossRef]
  61. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data Analysis for X-Ray Absorption Spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef]
  62. Abbott, D.F.; Fabbri, E.; Borlaf, M.; Bozza, F.; Schäublin, R.; Nachtegaal, M.; Graule, T.; Schmidt, T.J. Operando X-ray absorption investigations into the role of Fe in the electrochemical stability and oxygen evolution activity of Ni1-xFexOy nanoparticles. J. Mater. Chem. A 2018, 6, 24534–24549. [Google Scholar] [CrossRef]
  63. Sharabi, R.; Markevich, E.; Borgel, V.; Salitra, G.; Gershinsky, G.; Aurbach, D.; Semrau, G.; Schmidt, M.A.; Schall, N.; Stinner, C. Raman study of structural stability of LiCoPO4 cathodes in LiPF6 containing electrolytes. J. Power Sources 2012, 203, 109–114. [Google Scholar] [CrossRef]
  64. Wu, F.; Dong, J.; Chen, L.; Chen, G.; Shi, Q.; Nie, Y.; Lu, Y.; Bao, L.; Li, N.; Song, T.; et al. Removing the intrinsic NiO phase and residual lithium for high-performance nickel-rich materials. Energy Mater. Adv. 2023, 4, 0007. [Google Scholar] [CrossRef]
Figure 1. XRD pattern for LNCP powder after calcined at 700 °C.
Figure 1. XRD pattern for LNCP powder after calcined at 700 °C.
Nanomaterials 13 03156 g001
Figure 2. The high-resolution XPS spectra of (a) Co 2p, (b) Ni 2p, (c) P 2p, (d) O 1s, and (e) Li 1s in the LNCP powder.
Figure 2. The high-resolution XPS spectra of (a) Co 2p, (b) Ni 2p, (c) P 2p, (d) O 1s, and (e) Li 1s in the LNCP powder.
Nanomaterials 13 03156 g002
Figure 3. SEM micrographs at (a) low and (b) high resolutions; (c) SEM-EDS elemental mapping for LNCP.
Figure 3. SEM micrographs at (a) low and (b) high resolutions; (c) SEM-EDS elemental mapping for LNCP.
Nanomaterials 13 03156 g003
Figure 4. (a) Cyclic voltammetry at a scan rate of 0.1 mV s−1, (b) Galvanostatic voltage profiles, and (c) cycling performance of LNCP cathode material in half cell at 0.1 C-rate (1 C = 167 mAh g−1). (d) Voltage profile and cycling performance of LNCP–LTO full cell.
Figure 4. (a) Cyclic voltammetry at a scan rate of 0.1 mV s−1, (b) Galvanostatic voltage profiles, and (c) cycling performance of LNCP cathode material in half cell at 0.1 C-rate (1 C = 167 mAh g−1). (d) Voltage profile and cycling performance of LNCP–LTO full cell.
Nanomaterials 13 03156 g004
Figure 5. Ex situ analysis of LNCP electrodes during the first charge and discharge electrochemical cycle: (a) XANES and (c) corresponding EXAFS spectra at Co K-edge, and (b) XANES and (d) corresponding EXAFS spectra at Ni K-edge.
Figure 5. Ex situ analysis of LNCP electrodes during the first charge and discharge electrochemical cycle: (a) XANES and (c) corresponding EXAFS spectra at Co K-edge, and (b) XANES and (d) corresponding EXAFS spectra at Ni K-edge.
Nanomaterials 13 03156 g005
Figure 6. Ex situ XRD for fully charged and discharged LNCP electrodes in a 3.5–5.2 V voltage range.
Figure 6. Ex situ XRD for fully charged and discharged LNCP electrodes in a 3.5–5.2 V voltage range.
Nanomaterials 13 03156 g006
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.

Share and Cite

MDPI and ACS Style

Islam, M.; Ali, G.; Faizan, M.; Han, D.; Ali, B.; Yun, S.; Ahmad, H.; Nam, K.-W. Scalable Precursor-Assisted Synthesis of a High Voltage LiNiyCo1−yPO4 Cathode for Li-Ion Batteries. Nanomaterials 2023, 13, 3156. https://doi.org/10.3390/nano13243156

AMA Style

Islam M, Ali G, Faizan M, Han D, Ali B, Yun S, Ahmad H, Nam K-W. Scalable Precursor-Assisted Synthesis of a High Voltage LiNiyCo1−yPO4 Cathode for Li-Ion Batteries. Nanomaterials. 2023; 13(24):3156. https://doi.org/10.3390/nano13243156

Chicago/Turabian Style

Islam, Mobinul, Ghulam Ali, Muhammad Faizan, Daseul Han, Basit Ali, Sua Yun, Haseeb Ahmad, and Kyung-Wan Nam. 2023. "Scalable Precursor-Assisted Synthesis of a High Voltage LiNiyCo1−yPO4 Cathode for Li-Ion Batteries" Nanomaterials 13, no. 24: 3156. https://doi.org/10.3390/nano13243156

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop