Open Access Article
Beth I. J. Johnston
*abc,
Satish Bollojuab,
Stephen W. T. Price
bd,
Alexander G. Squirese,
Lavan Ganeshkumarbf,
Muhammad Ansab,
James A. Gotta,
Narayan Simrit Kaurc,
Innes McClelland
bg,
Samuel G. Booth
bc,
Andrew M. Bealed,
Simon D. M. Jacquesd,
Ashok S. Menon
ab,
David O. Scanlon
be,
Louis F. J. Piper
ab and
Serena A. Cussen
*bg
aWMG, University of Warwick, Coventry, CV4 7AL, UK. E-mail: beth.johnston@warwick.ac.uk
bThe Faraday Institution, Quad One, Harwell Campus, Didcot, OX11 0RA, UK
cSchool of Chemical, Materials and Biological Engineering, The University of Sheffield, Sheffield, S1 3DJ, UK
dFinden Limited, Building R71, Rutherford Appleton Laboratory, Harwell, Oxford, OX11 0QX, UK
eSchool of Chemistry, The University of Birmingham, Birmingham, B15 2TT, UK
fDepartment of Chemistry, University College London, London, WC1H 0AJ, UK
gSchool of Chemistry, University College Dublin, Belfield, Dublin 4, Ireland. E-mail: serena.cussen@ucd.ie
First published on 16th October 2025
LiNiO2 cathodes for lithium-ion batteries offer the prospect of high specific capacities; however, a plethora of structural and surface instabilities occur during cycling, which can limit their lifetime and impinge on their safety. Structural and surface modification strategies such as cation-doping have been shown to stabilise cycling performance and prolong cathode lifetimes, yet they often tackle either surface or bulk driven degradation processes. Here, we present a dual-cation substitution approach for the LiNiO2 cathode which produces a coat-doped cathode in a single step. Judicious selection of cation substituents enables the targeted stabilisation of both bulk- and surface-originating instabilities, in this case magnesium and tungsten, respectively. While the addition of tungsten as a sole substituent promotes a rock-salt surface layer which typically reduces the observable capacity, we demonstrate that the incorporation of Mg into W-containing compositions can mitigate these structural transformations. These coat-doped Mg/W-LiNiO2 cathodes exhibit superior cycling stabilities compared to unmodified LiNiO2 and singly-substituted Mg- or W-LiNiO2. X-ray diffraction computed tomography methods complement these findings, providing spatially resolved structural information on the location and heterogeneity of the coat-doped cathodes, guiding synthetic pathways to optimised materials that outperform undoped LiNiO2 even in high-mass loading cell environments.
In addition to these complex surface degradation effects, bulk structural instabilities also occur and are principally linked to a lattice collapse that occurs at high states of charge for TM layered oxides.21 These can be tracked through in situ and operando diffraction methods, which demonstrate several first-order phase transitions during cycling.22–26 For LNO, during charging, these transitions proceed via H1 (pristine LNO) ↔ M ↔ H2 ↔ H3 where H denotes a hexagonal phase and M denotes a monoclinic phase. This is accompanied by a gradual decrease in the a lattice parameter throughout charging, motivated by the Ni–O environment which experiences oxidation during delithiation. Meanwhile, the c lattice parameter gradually expands with lithium removal due to electrostatic repulsion between the [NiO2] layers until ≈80% delithiation is reached. On reaching this point, which occurs between 4.15 V and 4.2 V vs. Li/Li+, the first order transition from the H2 to H3 phase is accompanied by a rapid c-axis contraction and lattice collapse, believed to be a result of the now almost empty Li layer where steric effects prevail over electrostatic repulsions.27 The anisotropy between lattice parameter evolutions causes significant strain in the material.28 Such induced strain in Ni-rich materials can manifest on the macroscale through particle cracking, which exposes fresh, highly reactive surfaces and can also lead to areas that become electronically disconnected.29–31
Attempts to mitigate such degradation modes and extend the cycling lifetime of LNO have previously included chemical stabilisation, such as surface coatings32 or elemental substitution of Li or Ni with foreign cations. When small fractions of substituents are introduced (on the order of 0.5–5 mol%), we can consider this a substitution-type approach.9,33,34 Substitution has proved to be an effective method in stabilising electrochemical performances in Ni-rich layered oxides with a wide variety of substituent cations able to stabilise cycling behaviour such as Al3+,35–37 Mg2+,38–40 Ti4+,41–43 Ga3+,44,45 B3+,46,47 Zn4+,48,49 and Nb5+.50 The location of the substituent can also influence the mode of stabilisation. For example Mg2+ is predicted to preferentially reside on Li+ sites owing to the fact that it has a similar size to Li+ (rLi+ = 0.76 Å, rMg2+ = 0.72 Å) and a lower formation energy where it is thought to stabilise the [LiO2] interlayer against collapse through a pillaring effect.51,52 More recently, high valence substituents which are more insoluble in the LiMO2 structure, such as W6+, Nb5+, Ta5+, and Mo6+, have been investigated owing to their propensity to segregate to grain boundaries between particles where they may act to stabilise these surfaces during cycling.53,54 Tungsten has also been shown to inhibit primary particle growth, in addition to suppressing the magnitude of the H2 ↔ H3 transition and improving the mechanical properties of LNO as a result of LixWyOz surface phases which form.55–58 Given different dopants can differently affect structure and stability, doping with multiple cations is an appealing prospect to access a combination of benefits for LNO cathodes.59–61 Here, we consider the synergy between candidate substituent cations to deliver a coat-doped arrangement for LNO cathodes, combining the pillar-stabilising effects of magnesium within the particle bulk and the coating-stabilising effect of a tungsten-enriched surface phase. The efficacy of the dual-substitution strategy, utilising low substituent concentrations <2 mol% in total, was established through higher capacity retention, outperforming singly substituted LNO. While Mg-inclusion preserved the well-layered nature of the LNO structure, W incorporation was shown to promote a more rock-salt like structure and lower specific capacities. Through XRD-CT on charged samples, position-resolved crystallographic information revealed structural inhomogeneities across W-containing materials, prompting an improved synthetic strategy. Improved W-containing samples showed a more preserved layered oxide structure, leading to higher specific capacities. During cycling, the dual-substituted, coat-doped LNO consistently outperformed singly-substituted compositions across half- and full-cells, indicating promise in higher mass-loading environments, demonstrating the efficacy of this targeted, scalable compositional tuning approach for high energy density ultra-Ni-rich cathode materials.
:
TM ratio maintained at 2
:
1) and an NH4OH chelating agent solution (NH4OH
:
TM ratio = 5
:
1) were pumped separately into the reactor. The reaction was stirred overnight (1000 rpm) and a Ni(OH)2 precipitate was obtained after washing. The powder was dried at 80 °C overnight.
The obtained Ni(OH)2 powder was thoroughly mixed by grinding with LiOH·H2O (99.995%, Thermo Fisher Scientific), WO3 (99.9%, Merck) and Mg(CH3COO)2 (98%, Merck) in suitable quantities to obtain a total of six samples of target stoichiometries as follows: parent, unmodified LiNiO2 (LNO), Li0.99Mg0.01NiO2 (1Mg-LNO), Li0.98Mg0.02NiO2 (2Mg-LNO), LiNi0.99W0.01O2 (1W-LNO), LiNi0.98W0.02O2 (2W-LNO) and Li0.99Mg0.01Ni0.99W0.01O2 (1Mg1W-LNO). A molar excess of 1% LiOH·H2O was used for synthesis initially, increasing to 5% for W-containing LNOs in later experiments (as described in the discussion). Ground samples were transferred to a tube furnace and heated at 480 °C for 5 hours followed by a final calcination at 770 °C (reduced to 710 °C for the parent LNO) for 15 hours under a flow of oxygen. After the reaction, samples were cooled to 200 °C and quickly removed to an Ar filled glovebox to avoid moisture and CO2 uptake.
Particle size and morphology were examined here using scanning electron microscopy (SEM). Samples were prepared on adhesive carbon tabs and coated with Au using a sputter coater to avoid charging feedback. SEM images were taken using an Inspect F50 electron microscope operating at 20 kV. Cross sectional electrode images (SEM and SEM-EDS) were acquired using a Thermo Fisher Scientific Scios Dualbeam system (operating at 2 kV and 0.1 nA), equipped with a CleanConnect inert gas transfer system. The method for ion milling and image acquisition is identical to that reported in previous work.63
:
5
:
5 in NMP (N-methyl-2-pyrrolidone) solvent (Merck). The obtained slurry was cast onto carbon-coated Al foil using a doctor blade set at a height of 150 μm. The slurry was dried at 100 °C to evaporate the NMP solvent before further drying at 80 °C overnight in a vacuum oven. Electrodes of 10 mm diameter were then punched from the dried film and transferred into an Ar-filled glovebox. Active material loadings were 4–5 mg cm−2.CR2032 coin-type half-cells were constructed in an Ar-filled glovebox using the as-prepared LNO electrodes as the cathode, a 15.8 mm diameter Li metal anode (Cambridge Energy Solutions) and a Whatman glass microfibre separator. 100 μL of 1 M LiPF6 in ethylene carbonate (EC)
:
ethyl methyl carbonate (EMC) (EC
:
EMC 3
:
7 by weight%) with 2% vinylene carbonate (VC) additive (Solvionic) was used as the electrolyte. Galvanostatic charge and discharge tests were carried out in a temperature-controlled chamber at 25 °C using an Arbin LBT21084 series cycler. In all cases 1C was defined as 220 mA g−1. Initial slow cycles at C/10 or C/20 (as detailed in the text) were initially carried out using a constant-current constant-voltage (CCCV) protocol (current limits of C/20 and C/40 respectively), followed by longer term cycling at C/2 and/or C/3. Cells were cycled between 3.0 and 4.3 V versus Li/Li+ and 5 minute rest steps were implemented after each discharge. For ex situ XRD-CT tests, half-coin cells were cycled to defined upper cut-off potentials of 3.8, 4.2, 4.4 and 4.7 V vs. Li/Li+ at a rate of C/20. After reaching the desired cut-off voltage, cells were immediately disassembled in an argon filled glovebox, washed with dimethyl carbonate (DMC), dried under vacuum and stored under argon.
:
5
:
5 (wt%) in NMP. The slurries were applied onto aluminium foil substrates using a precision drawdown coater (Erichsen Coatmaster 510). The active material loading was 13–14 mg cm−2, and the electrodes were calendered to achieve a target density of 3.0 g cm−3. The anode electrode consisted of 95.25 wt% graphite (BTR V-H), 1 wt% carbon black (Imerys C45), 1.50% CMC (Ashland BVH8), and 2.25 wt% SBR (Zeon BM451) binder. The mass loading was 9.3 mg cm−2, and the electrode was calendered to a density of 1.23 g cm−3. The negative to positive electrode capacity ratio (N/P) of electrodes was 1.17. The electrolyte used was 1 M lithium hexafluorophosphate (LiPF6) in ethylene carbonate (EC) and ethyl methyl carbonate (EMC) (LP57) with 2% vinylene carbonate (VC) (Solvionic).
XRD-CT experiments were conducted at beamline P07 (EH2) of PETRA III, DESY. A monochromatic X-ray beam of 73.799 keV (0.168 Å) was focussed to 1 μm × 10 μm. Data was collected using a Pilatus3 X CdTe 2M detector and calibrated with a CeO2 standard (99.9% <5 μm). XRD-CT scans were measured at 250 Hz with a translation step size of 1 μm, and 4 interlaced65 rotational scans with steps of 0.72° within each scan, and steps of 0.18° between scans. The interlaced approach served firstly to assess for beam induced sample degradation, and secondly to mitigate the risk of unexpected failure during a very long duration scan.
Phase identification of the reconstructed XRD-CT datasets was carried out using X'Pert HighScore Plus, whilst quantification of the datasets was carried out using BeamStop68 (in-house high-volume data processing GUI) and Topas V7. For cathodes charged to higher voltages, two peaks corresponding to the LiNiO2 (003) reflection with different c-axis lengths were observed in some pixels, i.e. there were regions smaller than the resolution of the measurement in the sample that had differing degrees of (de)lithiation in close proximity. To account for this in the refinement, two separate R
m LiNiO2 models were used, with restrictions on the c-axis lattice parameter being above or below 14 Å. The analysis here focusses on probing the relative differences in the unit cell parameters per pixel, thus the LiNiO2 models used here did not account for off-stoichiometry (Li1−xNi1+xO2 with x Ni2+ residing on the Li sites). Once fitting was completed, the scale factor for each LiNiO2 phase was used to mask the respective lattice parameters and crystallite sizes, and then the results were combined to provide the average value per pixel presented in the figures below.
Initial trial dopant-defect structures within Jahn–Teller distorted LiNiO2 (space group P21/c) were generated using ShakeNBreak,77,78 with low-energy configurations further optimized and analyzed with doped.79 Defect relaxations used a 2 × 4 × 2 supercell of the primitive P21/c unit cell and were sampled with a 2 × 2 × 2 Γ-centered Monkhorst–Pack k-point mesh. Convergence criteria were 1 × 10−5 eV for total energies and 0.01 eV Å−1 for forces.
Defect formation energies were calculated as:
To investigate the effect of both Mg and W substituents on the layered LiNiO2 structure, samples were analysed by X-ray diffraction (XRD) and Rietveld refinement. Fig. 1(c) shows the powder XRD patterns obtained from LNO, 1Mg-LNO, 2Mg-LNO, 1W-LNO, 2W-LNO and 1Mg1W-LNO. The data have been normalised to the peak height of the (003) reflection for each sample. The observed diffraction peaks could be indexed to hexagonal type α-NaFeO2 with space group R
m typical of transition metal layered oxide cathodes, with Li and Ni occupying alternating layers amongst a cubic close packed oxygen arrangement (Fig. S6). Fig. 1(d) magnifies the 2θ region where reflections from common impurity phases often appear between 20 and 35°. In this region, 1W-LNO and 2W-LNO show evidence of an impurity phase (marked with *) identified as Li6WO6.56 These reflections in the 2W-LNO XRD pattern have a higher intensity indicating that the amount of Li6WO6 increases as the W substitution level increases. In contrast, compositions containing only Mg exhibit no evidence of Mg-containing impurities, suggesting Mg incorporation into the layered structure. For 1Mg1W-LNO, no Li6WO6 impurities are distinguishable from the background suggesting the ability of this dual cation strategy to mitigate the formation of the Li6WO6 phase. The mitigation of Li6WO6 formation during synthesis could be beneficial as such phases may leach lithium from LiNiO2, leading to more rock-salt like, lithium deficient structures along the NiO–LiNiO2 phase line, decreasing the practical capacities. For all samples here, no further impurities (e.g. Li2CO3, LiOH and Li2SO4) are detected in the XRD data.
Fig. 1(e) highlights the (006)/(102) and the (108)/(110) peak doublets where the degree of splitting is an indication of the degree of layering in the crystal structure of LiNiO2. These peaks coincide in the rock-salt structure, but as the structure begins to layer, they become increasingly separated.9 Here, the degree of splitting appears reduced by W-substitution across 1W-LNO, 2W-LNO and 1Mg1W-LNO compared to pristine LNO and Mg-LNOs. This correlates well with the reported ability of W-substitution to reduce the layered character of the structure in favour of a more rock-salt like character58,88 (although it is clear from the XRD data here that this structure remains predominantly layered in character). Vacancies in the Li layer likely promote a greater degree of off-stoichiometry in the samples and encourage a shift towards more rock-salt (LixNi1−xO) behaviour, reducing the layering in the structure. However, upon coat-doping, improvements in peak splitting are seen here for 1Mg1W-LNO compared to 1W-LNO and 2W-LNO. As observed previously, the appearance of the Li6WO6 impurity for 1W-LNO and 2W-LNO, and its absence in 1Mg1W-LNO, suggest that the Mg-doping in W-LNOs can mitigate against the formation of Li–W–O species and help to preserve the Li content in the active layered phase and preserve more of its layered character.
To investigate the effect of Mg and W substitution on the LiNiO2 structure in more detail, Rietveld refinements were performed as shown in Fig. 2(a) for LNO and (b) for 1Mg1W-LNO (Fig. S7(a) for 1Mg-LNO, (b) 2Mg-LNO, (c) 1W-LNO and (d) 2W-LNO) with structural parameters for all summarised in Table 1. Refinements undertaken assumed a hexagonal NaFeO2 layered structure with space group R
m where Li occupies the 3a site, Ni occupies mainly the 3b site and oxygen occupies the 6c site. A Li6WO6 phase was added to the refinements for 2W-LNO where the associated peaks are clearly distinguishable from the background. A small amount of nickel was allowed to reside on the 3a site (denoted herein as NiLi) to represent the off-stoichiometry (Li1−xNi1+xO2) where the total occupancy of the 3a site was equal to unity. Li occupancy in the Ni layer was not considered as this is reported only to occur in LNO where the off-stoichiometry is high (NiLi > 0.07).89 For Mg and W containing compositions, Mg was assumed to sit on the 3a site and W was assumed to fit in the 3b site and their occupancies were fixed at their theoretical value. Although W is predicted to accumulate at particle surfaces, a study by Bianchini et al. found that modelling the unit cell with W on 3b sites led to better quality fits.58 Besides, at least some W incorporation into 3b sites is not completely ruled out and as a bulk averaged probe, XRD is unable to resolve the W distribution in the sample and from the XRD patterns exhibited in Fig. 1, it is clear here that the process of W-incorporation yields changes to the bulk LNO structure.
![]() | ||
Fig. 2 Rietveld refinement plots of (a) LNO and (b) 1Mg-LNO showing the data (circles), fit (black/orange coloured line), difference (turquoise line) and the Bragg position ticks for R m LiNiO2 (red). | ||
| Sample | a (Å) | c (Å) | V (Å3) | c/a | NiLi | wR (%) |
|---|---|---|---|---|---|---|
| LNO | 2.87741(4) | 14.2039(3) | 101.846(3) | 4.936 | 0.016(1) | 2.51 |
| 1Mg-LNO | 2.87886(4) | 14.2088(2) | 101.983(3) | 4.936 | 0.018(1) | 2.53 |
| 2Mg-LNO | 2.87837(4) | 14.2094(3) | 101.953(3) | 4.937 | 0.014(1) | 2.57 |
| 1W-LNO | 2.87907(6) | 14.2064(4) | 101.981(5) | 4.934 | 0.045(1) | 2.16 |
| 2W-LNO | 2.88076(7) | 14.2096(5) | 102.123(5) | 4.933 | 0.062(1) | 1.98 |
| 1Mg1W-LNO | 2.87916(5) | 14.2093(3) | 102.009(4) | 4.935 | 0.037(1) | 1.96 |
Compared to the undoped LNO, all doped samples display enlarged lattice parameters and thus enlarged unit cell volumes. Lattice parameter changes upon Mg and W-addition are compared in Fig. S8. Here, the laboratory X-ray measurements are not able to resolve the Mg positions for such a low doping level but its effect on the structural parameters can be established. Upon addition of Mg2+ to LNO, the a-lattice parameter is found to expand at 1 mol% Mg but shrink for 2 mol% Mg while the c-lattice parameter grows for both 1 mol% and 2 mol% Mg, however not linearly. This behaviour is similar to that reported by Bianchini et al.90 The reasons for such behaviour could lie in the distribution of Mg2+ over the 3a site (where it substitutes for Li) and the 3b site (where it substitutes for Ni). Here, laboratory X-ray measurements are not able to resolve the Mg positions and occupancies at low levels. For W-containing samples, both a- and c-lattice parameters and thus the unit cell volume increase in a linear fashion upon increasing W addition, consistent with previous reports and expected due to the larger ionic radius of W6+ compared to Ni3+. For coat-doped 1Mg1W-LNO, only small changes compared to 1Mg-LNO and 1W-LNO occur. A small increase in the a-lattice parameter occurs from 1Mg-LNO to 1Mg1W-LNO accompanied by a minor increase in the c-lattice parameter, attributed to the reduction of Ni3+ in the Ni layer and additional Ni2+ in the interlayer as the off-stoichiometry increases. Meanwhile, no change in the a-lattice parameter is seen between 1W-LNO and 1Mg1W-LNO but the c-parameter is increased, likely a result of the interlayer spacing increasing on Mg-doping.
The c/a ratio, which is a commonly reported metric relating to the degree of layering in transition metal layered oxides such as LNO, is not greatly affected by Mg-substitution and slightly lowered on W substitution. This is seemingly increased for 1Mg1W-LNO, suggesting that the presence of Mg in W-containing compositions could aid in stabilising the layered phase over the rock-salt phase, also consistent with the trends observed in the (108)/(110) peak splitting behaviour observed in the XRD patterns (Fig. 1(e)). These trends are also reflected in the degrees of off-stoichiometry obtained, represented by the amount of Ni2+ residing in the Li layer (NiLi) calculated from refinements. In general, as more Ni occupies the 3a sites the structure evolves from well-layered to one with more rock-salt like character. Minimal change is found between LNO, 1Mg-LNO and 2Mg-LNO where a small (<2%) fraction of NiLi is calculated. A much larger degree of off-stoichiometry is found for 1W-LNO (4.5%) and 2W-LNO (6.2%), consistent with previous reports of W-doping into LNO.57,58 For the coat-doped sample, 1Mg1W-LNO, the off-stoichiometry is decreased compared to 1W-LNO to 3.7%, further establishing the ability of dual-cation incorporation in reducing the extent of structural change affected by W-addition and helping to preserve more of the layered character of the LNO. While the total amount of non-Li atoms in the Li layer (MgLi + NiLi) is consistent across 1Mg1W-LNO (assuming here total incorporation of Mg into the Li sites), 1W-LNO and 2W-LNO, reducing the amount of NiLi defects is beneficial for mitigating against the possible local layer contractions reported from the reduction of NiLi during charging (since the Mg oxidation state will stay constant).51 So the benefits of dual substitution here are apparent in the suppression of unwanted NiLi defects. While singly-doped LNOs have been shown here to behave in line with the literature, it should be noted that the benefits of coat-doping with Mg and W are shown to reverse, to some extent, the structural distortions brought about by W-addition and help to preserve the well-layered character of LiNiO2, which will likely impact the electrochemical behaviour.
| Mg-LNO | ΔμLi (eV) | ΔμNi (eV) | ΔμO (eV) | ΔμMg (eV) |
|---|---|---|---|---|
| A | −3.26 | −2.93 | 0.00 | −5.79 |
| B | −3.04 | −3.15 | 0.00 | −5.73 |
| C | −2.74 | −2.85 | −0.30 | −5.43 |
| D | −2.66 | −2.60 | −0.47 | −5.28 |
| E | −2.72 | −2.39 | −0.54 | −5.25 |
| F | −2.66 | −2.45 | −0.54 | −5.23 |
| W-LNO | ΔμLi (eV) | ΔμNi (eV) | ΔμO (eV) | ΔμW (eV) |
|---|---|---|---|---|
| A | −3.26 | −2.93 | 0.00 | −9.79 |
| B | −3.04 | −3.15 | 0.00 | −10.65 |
| C | −2.74 | −2.85 | −0.30 | −10.35 |
| D | −2.66 | −2.60 | −0.47 | −9.49 |
| E | −2.72 | −2.39 | −0.54 | −9.52 |
| F | −2.66 | −2.45 | −0.54 | −9.86 |
For Mg-LNO, the predicted defect concentrations at these vertices and the interpolated paths between them are shown in Fig. 3(a). Two features of the plot stand out: firstly, the concentrations of the dopant defects are high relative to the intrinsic defect concentrations, suggesting that Mg incorporates into the bulk of LNO easily under most synthesis regimes, and secondly, that the preference for Mg substituting for Li and Ni is dependent on the chemical potentials of the cations which cannot be excluded as a potential contributor to the challenge of assigning the site preference of Mg dopants alongside instrument sensitivity. Repeating the same analysis for W-doped LNO (Fig. 3(b)) shows a stark difference from Mg-LNO. Here, there is a significant preference for WNi relative to WLi in contrast to Mg-LNO where the preferred dopant site is dependent on the chemical potential limit the concentrations are calculated at, but also the concentrations of the WNi dopant species are significantly lower than most intrinsic defects. The concentrations are on the order of the nickel vacancy, which has been shown to be highly unstable in first principles studies.91 From a thermodynamic perspective, bulk incorporation of W into layered LNO is unlikely. Any residual W content in the bulk of the material is likely to be due to out-of-equilibrium or kinetic effects; the system is likely to favour the segregation of W-rich phases. Thus, while the SEM-EDS mapping (Fig. S5) shows Mg and W distribution across the secondary particles, it is likely that while the Mg is doped into the bulk, W is segregated to the primary particle surfaces and grain boundaries (this scale is not resolvable in the EDS image).
On discharge, larger irreversible capacity losses are exhibited by all doped samples compared to the undoped LNO. Kinetic hindrances associated with greater off-stoichiometry are thought to govern first cycle losses in LNO.92 So it is no surprise that all doped samples have higher irreversible capacities and poorer coulombic efficiencies compared to the undoped LNO material. However, if these capacity losses purely stem from local layer contractions around oxidised NiLi at high states of charge,93 this does not account for the larger irreversible capacities (compared to undoped LNO) observed in the Mg-doped LNOs. Instead, the revised mechanism for kinetic hindrances in off-stoichiometric LNO reported by Sicolo et al. explains the observed behaviour.94 Here, positively charged NiLi defects do not oxidise during charging but rather act as a sink for Li vacancies, preventing Li reintercalation to these vacancies under standard cycling conditions. For Mg-doped samples, the positively charged MgLi should act in the same manner and induce Li vacancy (VLi) trapping around it. Indeed, calculating MgLi–VLi binding energies as compared to NiLi–VLi binding energies, Eb, according tow
Fig. 4 shows the voltage-capacity profiles and the corresponding differential capacity (dQ/dV) plots for all LNO compositions during the third cycle. Fig. 4(a) and (b) show the effects on these for all Mg-containing samples while Fig. 4(c) and (d) highlight the changes observed for W-containing samples. Again, all substituted compositions exhibit lower discharge capacities compared to the undoped LNO. It should be noted that some compositions appear to have increased capacity compared to the first cycle, however this is commonly observed due to cathode conditioning processes that may occur. The more sloping nature of the doped compositions (Fig. 4(a) and (c)) is clear. This is also reflected in the dQ/dV profiles in Fig. 4(b) and (d). Upon Mg and W incorporation, the peaks broaden, marking a shift from two-phase towards more solid-solution like behaviour. Suppressing the two-phase behaviour in favour of solid-solution behaviour is beneficial to the electrochemical performance and cycling stability as it can mitigate against pronounced volume changes at high states of charge which are known to induce particle cracking.96 In particular, the feature at 3.9 V vs. Li/Li+ in the monoclinic region of the dQ/dV plots (magnified in each inset) corresponds to the stable ordering of Li vacancies in LixNiO2 for x = 2/5 and is sensitive to structural modifications such as doping.97 The Li vacancy orderings are disrupted by foreign cations (NiLi and MgLi) in the Li layer and thus this feature is suppressed in the substituted samples here. Similarly, the 3.5 V plateau on discharge, which is observed if the discharge (relithiation) kinetics are good, is also suppressed on substitution as a result of the foreign cations. The fact that this feature is suppressed even in Mg only containing samples (despite the similar NiLi amounts between LNO, 1Mg-LNO and 2Mg-LNO) is a strong indicator that Mg is present in the Li layer.
Further changes are clear in the H2 ↔ H3 transition that occurs between 4.1 and 4.2 V vs. Li/Li+ where the peaks broaden upon doping while also experiencing a shift to higher voltages. Broadening the H2 ↔ H3 peak can alleviate the severe and abrupt nature of the phase transition and help to mitigate the deleterious processes that originate from it, e.g. the anisotropic lattice distortion leading to abrupt lattice collapse and particle cracking. It is clear from the magnification of the H2 ↔ H3 region in Fig. 4(b) and (d) that W-addition suppresses this phase transition to a greater extent than Mg-addition, although all doped samples show significant suppression of the H2 ↔ H3 peak compared to the undoped LNO. Fig. 4(b) highlights greater suppression for 1Mg1W-LNO than for 1Mg- and 2Mg-LNO, while 4(d) demonstrates increasing suppression on subsequent W-addition (LNO to 1W-LNO to 2W-LNO) while 1Mg1W-LNO is similar to 1W-LNO – suggesting that W-addition alters the H2 ↔ H3 transition characteristics more than Mg-addition (although both alter the behaviour compared to undoped LNO). The shift of the H2 ↔ H3 peaks to higher voltages on doping indicates a change in the thermodynamic potential of the transition during charge and discharge and is consistent with other reports of W and Mg containing LNOs.57,58,90
To assess the longer-term cycling behaviour of Mg and W containing LNOs, after 5 conditioning cycles at C/10, the current rate was increased to C/2. The specific capacities achieved across all compositions across 100 cycles are demonstrated in Fig. 4(e). Upon higher applied currents each composition experiences a capacity drop of approximately 10 to 15 mA h g−1 in line with kinetic expectations when moving to higher current values. After 100 cycles, undoped LNO, 2Mg-LNO and 1Mg1W-LNO all deliver specific capacities close to 160 mA h g−1, while 1W-LNO and 2W-LNO exhibit lower capacities close to 150 mA h g−1. However, by examining the normalised capacity retention (normalised to the specific discharge capacity achieved during the first cycle at C/2, Fig. 4(f)) the benefits of dual-substitution are clear. While pristine LNO retains approximately 77% of its initial capacity after 100 cycles, 1Mg1W-LNO displays superior retention of 85% after 100 cycles. Although the 2W-LNO sample achieved a similar normalised capacity retention to 1Mg1W-LNO here after 100 cycles, the disparity in specific capacity between these materials is evident. Overall, when considering both specific capacity and normalised capacity retention, the superior cycling behaviour of the 1Mg1W-LNO compared to the other compositions studied here becomes evident and the implications of the Mg/W coat-doping strategy of LNO are promising.
However, despite the advantages of the dual-doped 1Mg1W-LNO shown here, the W-containing materials take a larger hit on capacity compared to the Mg only doped LNOs. The significant structural changes examined by PXRD (Fig. 1) and refinement of the XRD data (Fig. 2, S7 and Table 1), including the large calculated off-stoichiometries and the less layered character, likely lead to the reduction in capacity observed. The capacity drop observed on W-incorporation warrants improvements to the W-containing compositions. While the PXRD data presented here (above) yield information on the uncharged state of the LNOs, probing the structural evolution of the W-containing materials as a function of state of charge becomes important in determining suitable improvement strategies towards higher performance materials.
The XRD-CT measurements were performed on electrodes recovered from coin cells (see the Experimental), with each pixel in the image representing an area of roughly 10 × 10 μm2 which is on the order of the secondary particle diameter. Thus, resolution on the secondary, but not primary particle level is expected. Only values extracted from the refinement of the R
m structure are shown, thus the Al current collector is not visible here. Each composition was charged up to an upper cut-off voltage between 3.8 V and 4.7 V vs. Li/Li+ at a slow C/20 rate to promote uniformity of Li removal through more kinetically favourable cycling conditions. The voltages chosen for analysis aim to provide a view of the cathode in the H1 phase (open circuit voltage, OCV) and the M phase (3.8 V), during the H2 → H3 transition (4.2 V), above the transition (4.4 V) and at extreme delithiation (4.7 V). Examining the evolution of the c lattice parameter in LNO at increasing states of charge allows for indirect monitoring of the homogeneity of the extent of delithiation in the LNO compositions across the electrodes. This is shown in Fig. 5 for undoped LNO, 1W-LNO and 1Mg1W-LNO at the voltages stated above. All samples show a uniform expansion of the c lattice parameter across the area of examination up to 4.2 V, characterised by the deeper red colour observed at 4.2 V as compared to OCV and 3.8 V. This initial expansion in the c direction is well known owing to the stronger repulsion between adjacent oxygen slabs as lithium is initially removed from the layer.9 Above 4.2 V the data show that all samples experience a c-axis contraction, visually represented through a shift to green/cyan. This occurs to a greater extent in undoped LNO compared to 1W-LNO and 1Mg1W-LNO, demonstrating the substituents' ability to suppress the phase transitions that induce structural instabilities (abrupt contraction of the c-axis during the H2 → H3 transition) that lead to interparticle cracking. Importantly however, the c-lattice parameter distribution is more spatially homogeneous in undoped LNO compared to the doped samples (seen also in the a-lattice parameter and the c/a ratio shown in Fig. S10(a) and (b)) which display a spread of colours across the electrode cross-section, representing a wider range of c-axis values within the sample. This is indicative of non-uniform delithiation characteristics across the doped cathodes. As the extent of delithiation ultimately determines the capacity, this non-uniformity could be the cause of the lower specific capacities observed across W-containing LNOs (1W-LNO, 2W-LNO, 1Mg1W-LNO). That some areas of the W-containing cathode powder respond differently under charging (deviating from the behaviour of the undoped LNO despite identical electrode and coin-cell fabrication protocols) suggests that inhomogeneity exists across the powder likely arising as a result of the synthetic protocol, advocating for a revised synthetic route. Inhomogeneity of W distribution is also observed in the particles, as examined on an OCV 1Mg1W-LNO electrode through X-ray fluorescence computed tomography (XRF-CT) shown in Fig. S11. It is also noted for undoped LNO charged up to 4.4 V (and persisting at 4.7 V), there exists a thin, distinct layer on the electrode surface of higher c-axis value (not observed in the a-axis values, Fig. S10(a)), indicative of a less oxidised, less delithiated surface99 with a smaller degree of c-axis collapse.
As before, the dominant peaks observed could be indexed to the R
m LNO structure. For all W-containing compositions the presence of Li6WO6 peaks is evident in the magnified impurity region (19° ≤ 2θ ≤ 35°) where these peaks are higher in intensity for the 2W-LNO sample. In the previous 1Mg1W-LNO sample, no impurity peaks were observed, indicating that adding an excess of Li during the calcination can promote the formation of LixWOy phases. It is likely that the formation of such lithium tungstates is favoured over W-incorporation into the layered structure. While all W-containing samples exhibit less peak splitting of the (108)/(110) compared to the baseline, the undoped LNO here (magnified in Fig. S12), no discernible differences in the (108)/(110) peak splitting are observed across the W-containing series, indicating that the layered structure formed is not vastly changed here as the dopant levels are changed and as dual cation substitution is exploited – in contrast to the changes that were observed in the previous samples.
To examine the structures in more detail, Rietveld refinements from the PXRD data were carried out in the same manner as discussed for the first sample set. The refinement plots for 1W-LNO and 1Mg1W-LNO are shown in Fig. 6(b) and (c), respectively, with the plots obtained from refinements of 2W-LNO and 2Mg1W-LNO being shown in Fig. S13 with structural parameters calculated from the refinements displayed in Table 3. The values for undoped LNO here are the same as those shown in Table 1 for the baseline LNO. As before, all doped samples possess larger unit cell parameters and volumes compared to the undoped LNO as expected for doped variants; however, this occurs to a lesser extent than for previously observed doped samples (Table 1), indicating less deviation away from the structural properties of undoped LNO when using a greater Li precursor excess during synthesis. This behaviour could indicate a more well-layered structure attained through relieving the Li deficiency that can occur on formation of LixWyOz phases. The structural changes brought about by this extra Li during synthesis are perhaps most clear through the NiLi values obtained from the refinements. While an increase in NiLi is observed for the 1W-LNO here at 2.8% this is much lower than the previous value of 4.5% obtained for 1W-LNO in the previous sample set and it is likely that the additional Li helps to mitigate against Ni migration into the Li layer. This is similarly observed for the 2W-LNO sample here compared to the previous 2W-LNO sample (1 mol% excess LiOH·H2O during synthesis); however, here, there is no increase in the NiLi value on doubling the W-content from 1W-LNO to 2W-LNO despite increasing lattice parameters. The slightly enlarged c-parameter in the absence of additional NiLi defects could be caused by a slight Li deficiency in the Li layer that does not promote further off-stoichiometry. The structural changes could originate from the removal of Li from the layered structure to accommodate the Li6WO6 phase – so, a more Li deficient layered phase for 2W-LNO compared to 1W-LNO which does not cause a detectable change in the off-stoichiometry. Consistent with the results demonstrated for the previous Mg/W-doped sample set, the coat-doping is shown to reduce the off-stoichiometry from 2.8% to 2.0%, further confirming the ability of the Mg-dopant to mitigate against Ni migration in high-valence doped LNO compositions. Doubling the Mg-content in the coat-doped samples further demonstrates this, where for 2Mg1W-LNO the NiLi quantity was calculated to be 1.4% from the refinement, which brings this value in line with that calculated for undoped LNO (although the total amount of non-Li atoms in the Li layer will be larger after accounting for MgLi).
| Sample | a (Å) | c (Å) | V (Å3) | c/a | NiLi | wR (%) |
|---|---|---|---|---|---|---|
| LNO | 2.87741(4) | 14.2039(3) | 101.846(3) | 4.936 | 0.016(1) | 2.51 |
| 1W-LNO | 2.87835(5) | 14.2052(3) | 101.922(4) | 4.935 | 0.028(1) | 2.14 |
| 2W-LNO | 2.87947(4) | 14.2098(1) | 102.033(3) | 4.935 | 0.028(1) | 2.15 |
| 1Mg1W-LNO | 2.87814(5) | 14.2046(3) | 101.902(4) | 4.935 | 0.020(1) | 2.15 |
| 2Mg1W-LNO | 2.87847(5) | 14.2095(3) | 101.960(4) | 4.936 | 0.014 (1) | 2.17 |
Electrochemical tests were conducted to examine two outcomes: (i) whether the improved structural properties of W-containing LNOs translated to improved electrochemical properties and (ii) to examine whether the enhanced stability imparted through Mg/W coat-doping was consistent across both sample sets. A clear increase in initial capacity for 1W-LNO is afforded when a greater LiOH·H2O excess is employed during the synthesis shown in Fig. S14, demonstrating the advantage of this in W-containing LNOs. Fig. 7(a) shows the first cycle voltage-capacity profiles for LNO, 1W-LNO, 2W-LNO, 1Mg1W-LNO and 2Mg1W-LNO cycled between 3.0 and 4.3 V at a current rate of C/20 (where 1C = 220 mA g−1) at 25 °C, with Table S2 summarising the first cycle characteristics. Upon charging, LNO and doped compositions containing 1 mol% W reach similar capacities in the region of ∼260 mA h g−1; however, that of 2W-LNO is reduced to ∼240 mA h g−1. Although the quantity of NiLi defects in 1W-LNO and 2W-LNO was calculated to be the same, we cannot infer the Li content in the samples from the NiLi concentration due to the poor visibility of Li to X-rays which renders refinements of Li contents non-sensical. While they share a common off-stoichiometry, there is a clear enlargement of the unit cell parameters and cell volume for 2W-LNO compared to 1W-LNO (Table 3), suggestive of a Li deficiency in 2W-LNO. A combination of excess LixWyOz phases and a Li deficiency in 2W-LNO is likely to account for the reduced capacity for 2W-LNO. Interestingly, however, Hua et al. explored how surface Li+ vacancies could improve the electrochemical properties in Li-rich layered oxide cathode materials, where creating a shell with Li+ vacancies was shown to drastically improve the first cycle coulombic efficiencies and the resulting cycle life through improving the bulk phase activity and redox kinetics.100 Here, while suggested Li+ deficiencies can lower the specific capacities of LNO, the effect on capacity stabilisation is clear. On discharge, poorer first cycle coulombic efficiencies/larger irreversible capacity losses are exhibited by all doped samples compared to the undoped LNO as observed in the first sample set. However, for all samples here, the discharge plateau at 3.5 V, indicative of good lithiation kinetics, is not suppressed in any W-containing samples, likely due to the improved off-stoichiometries afforded by the altered synthesis route.
Fig. 7(b) shows the voltage profiles during the 2nd cycle in coin half-cells where all materials now exhibit discharge capacities >200 mA h g−1 with the trend LNO > 1W-LNO > 2Mg1W-LNO > 2W-LNO ≈ 1Mg1W-LNO. The capacity penalty associated with the first cycle irreversible capacity loss determines this trend with the doped materials all displaying lower capacities compared to the undoped LNO as observed before. Higher initial specific capacities are obtained when doubling the Mg content from 1 mol% to 2 mol% here between 1Mg1W-LNO and 2Mg1W-LNO and higher overall specific capacities are achieved when increasing the LiOH·H2O excess during synthesis from 1 mol% to 5 mol% for W-containing LNOs attributed to the decrease in off-stoichiometry in the samples.
The differential capacity (dQ/dV) profiles for the 2nd cycle are shown in Fig. 7(c) where the LNO phase transitions during de/lithiation are observed for all samples. As before, on dopant incorporation, the peaks become less intense and broader as the phase transition behaviour becomes more solid-solution like. The minimum at ∼3.9 V vs. Li/Li+ associated with the stable lithium/vacancy ordering at x = 2/5 (in LixNiO2) also disappears here for all doped materials, mirroring the behaviour displayed for the doped LNO samples discussed above. The H2 ↔ H3 transition (Fig. 7(d)) also mirrors the previous samples' behaviour where the peak associated with the transition becomes broader, in agreement with expectations. Within this sample set, the suppression in both coat-doped samples (1Mg1W-LNO and 2Mg1W-LNO) lies between that of 1W-LNO and 2W-LNO, indicating that in these compositions W addition has more of a suppressive effect on the transition than Mg. While suppression generally leads to more stable cycling in these materials and is seen as a favourable outcome, it is often at the expense of capacity and a middle-ground must be sought. To assess the longer-term cycling behaviour of these doped LNOs, after two formation cycles at C/20, the current rate was increased to C/3 and the performance was assessed over 100 cycles. The specific capacities of LNO, 1W-LNO, 2W-LNO, 1Mg1W-LNO and 2Mg1W-LNO are shown in Fig. 7(e) with the normalised capacity retention in 7(f). Upon increase of current density, the capacity for each composition drops between 10 and 15 mA h g−1. After 100 cycles, the capacity delivered by the dual-doped 1Mg1W-LNO (170 mA h g−1) is higher than that of undoped LNO (152 mA h g−1) and the other doped compositions (165 and 161 mA h g−1 for 1W-LNO and 2W-LNO, respectively). When normalised, the enhancement provided to the capacity retention by the coat-doping is even more clear. After 100 cycles, 1Mg1W-LNO retained 91% of its starting capacity (at C/3) compared to 74% for undoped LNO and 84 and 88% for 1W-LNO and 2W-LNO, respectively. As before (under initial synthesis conditions employing 1 mol% of LiOH·H2O precursor), the coat-doped sample 1Mg1W-LNO exhibits improved cycling stability.
For the 3E cells cycled with a CCCV charging protocol above, electrochemical impedance spectroscopy (EIS) was performed to evaluate the evolution of the cell impedances over long term cycling across parent LNO and coat-doped 1Mg1W-LNO, yielding insights into surface degradation. The Nyquist plots generated using 200 cycles of EIS data obtained from the cathode component of the 3E cell for LNO and 1Mg1W-LNO are shown in Fig. 8(e). The Nyquist plots generated using the full EIS data (cathode, anode and full cell) after formation, after 100 cycles and after 200 cycles are shown in Fig. S16, where the majority of the cell impedance contribution comes from the cathode during aging. After formation (Fig. S16(a) and (d)), a lower frequency semicircle attributed to the charge transfer resistance dominates, followed by a Warburg type impedance tail. Here, the 1Mg1W-LNO cathode exhibits an initial higher charge-transfer resistance compared to the parent LNO, likely from the hindered kinetics caused through cation substitution. However, after 100 (Fig. S16(b), (c), (e) and (f)) and 200 cycles, the charge transfer resistance for the parent LNO increases significantly while that of the 1Mg1W-LNO sample undergoes a much smaller growth. The improved surface impedance characteristics on coat-doping signified decreased surface degradation leading to the increased capacity retention observed consistently across long-term cycling. The cross-sectional SEM images of LNO and 1Mg1W-LNO electrodes harvested from 3E cells after 200 cycles (CC only charging) are shown in Fig. 8(f) and (g), respectively. Significant particle pulverisation is seen across many secondary particles across the LNO electrode, arising from repeated microcracks originating during cycling. In contrast, no significant particle pulverisation is observed across the 1Mg1W-LNO electrode. Microcrack suppression here is likely caused by the smoother phase transition behaviour associated with the doped LNO materials. The altered primary particle morphology also plays a role, where the smaller, more rod-like particles are able to more effectively dissipate internal strains arising from the anisotropic lattice distortions.55 Relieving particle microcracking and pulverisation through this doping method thus minimises the proliferation of fresh surfaces vulnerable to electrolyte attack and reduces the extent of surface degradation, reflected across the improved cycling stability and impedance results reported here.
| This journal is © The Royal Society of Chemistry 2025 |