Phase transformations in battery cathode materials during electrochemical-insertion reactions lead to capacity fading and low cycle life. One solution is to keep the same phase of cathode materials during cation insertion-extraction processes. Here, we demonstrate a novel strategy to control the phase and composition of Mn-based spinel oxides for magnesium-ion battery applications through the growth of thin films on lattice-matched substrates using pulsed laser deposition. Materials at two extreme conditions are considered: fully discharged cathode MgMn2O4 and fully charged cathode Mn2O4. The tetragonal MgMn2O4 (MMO) phase is obtained on MgAl2O4 substrates, while the cubic MMO phase is obtained on MgO substrates. Similarly, growth of the empty Mn2O4 spinel in the cubic phase is obtained on an MgO substrate. These results demonstrate the ability to control separately the phase of spinel thin films (e.g., tetragonal vs. cubic MMO) at nominally fixed composition, and to maintain a fixed (cubic) phase while varying its composition (MgxMn2O4, for x = 0, 1). This capability provides a novel route to gain insights into the operation of battery electrodes for energy storage applications.
I. INTRODUCTION
Spinel oxides, AB2O4, are widely used as catalysts,1,2 magnetic and ferroelectric devices,3,4 and as cathodes for lithium ion (e.g., LixMn2O4)5 and multivalent ion batteries (Mg2+, Al3+, etc., compared to Li+).6,7 Attractive properties include their low cost, low toxicity, and thermal stability.8 Li-based spinel cathodes operate only within a limited lithium composition range: λ-Mn2O4 ↔ LixMn2O4 for 0 ≤ x ≤ 1, corresponding to a maximum charge capacity of ∼140 mA h/g.5 This capacity limitation is associated with the occurrence of a structural phase change that occurs for x > 1, at which point LixMn2O4 (LMO) begins to transform from a cubic to a tetragonal phase.9 This structural change is associated with instabilities in LMO cathode, leading to capacity loss upon extended cycling. A similar issue was found for spinel-based Mg-cathodes: the insertion of Mg2+ into λ-MnO2 leads to a phase transformation from the cubic to tetragonal phase of MgMn2O4 (MMO).7 It is therefore expected that the phase stabilization of the cathode during cation insertion/extraction may be beneficial in the development of battery cathodes with high energy density and long cycle life.
It is difficult to maintain phase stability of powder materials under electrochemical charging/discharging conditions. For example, the empty (i.e., fully charged) cathode MnO2 has several structural phases, such as α-MnO2 (tetragonal I4/m; a = b = 9.777 Å and c = 2.855 Å),10 β-MnO2 (tetragonal P42/mnm; a = b = 4.398 Å and c = 2.873 Å),11,12 ε-MnO2 (hexagonal P63/mmc; a = b = 2.80 Å and c = 4.45 Å),13 γ-MnO2 (orthorhombic Pnma; a = 4.45 Å, b = 9.305 Å, and c = 2.85 Å),14 and λ-MnO2 (or spinel Mn2O4, cubic ; a = 8.029 Å).15 Although at room temperature the magnesiated (i.e., discharged) cathode, MMO, adopts a tetragonal spinel structure in space group I41/amd with a = b = 5.727 Å and c = 9.284 Å,16 it adopts a cubic phase in space group with a = 8.600 Å at high temperature (>950 °C)16,17 or high pressure (>15.6 GPa).18 To minimize the complexity associated with phase changes in powder-based electrodes, we chose, instead, to study epitaxial thin films, as many parameters such as geometry, orientation, and surface termination can be controlled rationally. Epitaxial thin films often follow the structures of the underlying substrates, and the relative stability of different phases can be controlled through the use of lattice strain.19,20
Herein, we demonstrate the growth and phase-control of the spinel MMO as epitaxial thin films for advanced battery cathode applications using pulsed laser deposition (PLD). PLD enables the growth of hetero-structured thin films with sharp interfaces and atomically smooth surfaces20 and enables the growth of metastable crystalline structures,21,22 allowing for the exploration of new materials with different properties and novel functionalities. Specifically, we demonstrate the phase-selective growth of high quality MMO thin films with either cubic or tetragonal phases by choosing different substrates, namely, MgO (cubic , a = 4.212 Å) vs. MgAl2O4 (MAO, cubic , a = 8.083 Å) and growth conditions (e.g., substrate temperature, gas pressure, and laser power). In addition, we also demonstrate the ability to control the composition of the cubic spinel by growing the empty spinel phase, Mn2O4, on an MgO substrate. This independent control of spinel phase (cubic vs. tetragonal) and composition (empty vs. Mg-containing) at room temperature and ambient condition introduces a new potential route for synthesizing battery cathodes with intrinsically higher stability.
II. EXPERIMENTAL DETAILS
Epitaxial MgxMn2O4 (x ≈ 0, 1) films were prepared by PLD on polished single crystal substrates, having a size of 5 × 5 × 1 mm3 MgO(001) (from MTI Corporation) or MAO(001) (from CrysTec). The MMO target with 25 mm diameter was synthesized using solid-state methods from stoichiometric mixtures of MgO and Mn2O3 (Alfa Aesar, USA) powders and was calcined at 1100 °C in air for 72 h. At room temperature (25 °C) and ambient pressure (1 atm or 101 kPa), the MMO target exhibits the expected tetragonal phase (Figure S123). The MnO2 target (Figure S223) was prepared using β-MnO2 sintered at 1100 °C in air for 72 h. The MgO substrates were annealed in air at 1000 °C for 10 h to obtain atomically flat surfaces (Figure S323). PLD was performed using a KrF excimer laser (λ = 248 nm) at 500 °C under 10 mTorr O2 for 2500 pulses (∼10 nm) and 7500 pulses (∼30 nm) on MgO(001) and 1500 pulses (∼5 nm) on MgAl2O4(001). In addition, 10 000 pulses (∼40 nm) of Mn2O4 were grown on MgO(001) at 650 °C under 10 mTorr O2. Reflection high-energy electron diffraction (RHEED) was utilized for diagnostic in situ monitoring of the film growth. RHEED oscillations (Figure S423) were observed during MMO growth, indicating the layer-by-layer growth. The surface morphology of MMO films was examined by atomic-force microscopy (AFM, Veeco Nanoscope IV).
Thin film X-ray diffraction (XRD) was performed using a four-circle diffractometer (X'Pert PRO, PANalytical) in specular and off-specular configurations with Cu Kα1 (λ = 1.5406 Å) radiation. Further characterization was performed with X-ray reflectivity (XRR) at sector 33 BM of the Advanced Photon Source (APS) in Argonne National Laboratory (ANL) using 20 keV X-rays with a beam size of 2.0 mm horizontally and 0.1 mm vertically and an incident flux of ∼1010 photons per second. Scattered X-rays were detected using a pixel array detector (Dectris PILATUS 100K). The sample was placed in a custom-designed chamber and purged with helium during the XRR measurement to avoid oxidization related to ozone generation from intense X-ray beam. Low-angle XRR and specular (00L) crystal truncation rod (CTR) measurements were performed on both MMO/MgO and MMO/MAO films, and off-specular (11L) CTRs were carried out on MMO/MAO films only. The background and diffuse X-ray scattering contributions in CTR were subtracted by selection of regions of interest of the area detector images. The final intensities were normalized and corrected taking into account beam polarization and Lorentz factors.
X-ray photoelectron spectroscopy (XPS) spectra (Thermo Scientific ESCALAB 250Xi) were collected at the Keck II facility of NUANCE at Northwestern University using monochromated Al Kα X-rays. A low-energy electron flood gun was used to compensate for XPS induced surface charging. The carbon 1s line (284.8 eV) was used as a reference to calibrate the XPS spectra. The depth-dependent composition was obtained after the samples were Ar-etched for depths up to 4–5 nm and analyzed by XPS.
III. RESULTS AND DISCUSSION
AFM analysis (Figure 1) reveals that MMO films on both MAO(001) and MgO(001), and Mn2O4 film on MgO(001) have atomically smooth surfaces, with root-mean-square (RMS) roughness ∼2 Å. The original terraces from MgO substrates (Figure S2) are still visible after MMO and Mn2O4 growth, indicating a layer-by-layer growth mode, which is further supported by RHEED data collected during film growth (Figure S323).
1 × 1 μm2 AFM image of (a) 5 nm MgMn2O4 film on MgAl2O4(001) (MMO/MAO) (001), (b) 10 nm MMO film on MgO(001), and (c) 30 nm Mn2O4 film on MgO(001). The atomic terraces originally from MgO(001) substrate are still visible after the film growth, indicating conformal layer-by-layer growth mode from PLD.
1 × 1 μm2 AFM image of (a) 5 nm MgMn2O4 film on MgAl2O4(001) (MMO/MAO) (001), (b) 10 nm MMO film on MgO(001), and (c) 30 nm Mn2O4 film on MgO(001). The atomic terraces originally from MgO(001) substrate are still visible after the film growth, indicating conformal layer-by-layer growth mode from PLD.
The thin film epitaxial relationship can be seen from in situ RHEED pattern (Figure 2). The MMO thin film RHEED patterns along the MgO[100] (Figure 2(c)) and MgO[110] (Figure 2(d)) directions are aligned with original MgO patterns (Figures 2(a) and 2(b)), indicating epitaxial growth. In addition, the RHEED patterns after MMO growth appear as rods of intensity instead of as spots, suggesting very smooth (RMS around several Å) thin film surfaces, consistent with AFM measurements. Note that MMO RHEED patterns have double the truncation rod density compared to MgO patterns, indicating that the MMO in-plane lattice constant is twice as large as that of MgO. Similar RHEED patterns were observed for the epitaxial MMO/MAO (Figure S523) and Mn2O4/MgO (Figure S623) films.
Reflection high-energy electron diffraction (RHEED) patterns taken along the [100] and [110] azimuthal directions for (a) MgO [100], (b) MgO [110], (c) MgMn2O4 (MMO) [100], and (d) MMO [110]. Spinel patterns with in-plane lattice constants that are twice as large as those of the MgO substrate can be seen after MMO growth.
Reflection high-energy electron diffraction (RHEED) patterns taken along the [100] and [110] azimuthal directions for (a) MgO [100], (b) MgO [110], (c) MgMn2O4 (MMO) [100], and (d) MMO [110]. Spinel patterns with in-plane lattice constants that are twice as large as those of the MgO substrate can be seen after MMO growth.
The film structure and morphology was obtained with XRR. The XRR experimental data and model fits (Figure 3) are shown for samples with 1500 PLD pulses of MMO/MAO(001) and 2500 pulses of MMO/MgO(001) films. The data are plotted as a function of the vertical momentum transfer vector magnitude, q = 4π sin(2θ/2)/λ, where 2θ is the scattering angle and λ is the X-ray wavelength. The presence of oscillations extending to q > 0.8 Å−1 indicates smooth, uniform film layers, which is consistent with AFM and in situ RHEED analysis. The XRR data were fitted by a three-layer model, which consists of a semi-infinite substrate crystal (MAO or MgO), an MMO bulk layer, a transition layer, and a lower density surface overlayer (Figure 3, inset). This low-density surface overlayer could be due to incomplete coverage from PLD growth. This analysis reveals film thicknesses of 54 ± 9 Å for MMO/MAO(001), and 102 ± 9 Å for MMO/MgO(001) with electron densities that are within 5% of the expected value. Comparing the laser pulses used for each film, these results suggest that MMO was deposited at approximately the same rate, 0.04 Å/pulse, on both MAO and MgO substrates. The surface roughness of the MMO film from XRR analysis is ∼2 Å, which is consistent with AFM results. Furthermore, the derived electron density profiles show that MMO films have a bulk-like density on both substrates, but with an additional ∼20 Å-thick surface overlayer having a density that is ∼25%–45% less than that expected for the MMO film (see Table S1 for detailed fitting values23). This might be due to the partially covered surface during PLD growth. Further analysis by depth-dependent XPS shows that the surface (∼1.5 nm in depth) is rich in carbon and oxygen, thus reducing the average film density at the top, consistent with the formation of adventitious carbon layer. In contrast, the “bulk” (deeper than 1.5 nm) compositions are roughly close to the nominal values, namely, Mg1.05Mn2O3.98/MAO and Mg0.94Mn2O4.05/MgO. This slight off-stoichiometry could introduce point defects to alter the film lattice constants.
X-ray reflectivity data (black dots) and fits (red lines) for (a) 1500 PLD pulses of MMO on MAO(001) and (b) 2500 PLD pulses of MMO on MgO(001). Insets in each figure show the electron density depth profiles obtained from XRR fits normalized to the electron density of the substrate. The corresponding model parameters are listed in Table S1 in supplementary material.
X-ray reflectivity data (black dots) and fits (red lines) for (a) 1500 PLD pulses of MMO on MAO(001) and (b) 2500 PLD pulses of MMO on MgO(001). Insets in each figure show the electron density depth profiles obtained from XRR fits normalized to the electron density of the substrate. The corresponding model parameters are listed in Table S1 in supplementary material.
Epitaxial relationships between the MMO film and the MAO(001) substrate are further revealed in Figure 4. Specular (00L) CTR data (Figure 4(a)) for tetragonal MMO (MMOT) grown on MAO(001) show that the film has its c axis aligned with the substrate c axis. The off-specular (11L) CTR in Figure 4(b) indicates the following alignment: MMOT[101]//MAO[111]. This also implies that the MMOT film unit cell has a 45° in-plane rotation with respect to MAO substrate unit cell, which is further supported by in-plane ϕ scan (Figure S723). A crystal model is drawn in Figure 3(c) to better illustrate the epitaxial relationship between MMOT and MAO. This is not surprising as bulk MMOT has an in-plane lattice constant aMMO = 5.727 Å. To keep an epitaxial growth with minimum strain, a 45° in-plane rotation is necessary to match the substrate lattice constant . From the XRD data, the average unit cell parameters of the MMO film were determined to be c = 9.167 Å and a = 5.675 Å. Thus, there is compressive in-plane strain (εxx = −0.91%) with respect to the bulk MMOT lattice (a = b = 5.727 Å and c = 9.284 Å (Ref. 16)).
Crystal truncation rod data along (a) specular (00L) and (b) off-specular (11L) for MMOT/MAO(001). (d) Specular (00L) and (b) off-specular (20L) for MMOC/MgO(001). (c) and (d) Schematic crystal models showing the observed epitaxial relationships for MMOT/MAO(001) and MMOC/MgO(001).
Crystal truncation rod data along (a) specular (00L) and (b) off-specular (11L) for MMOT/MAO(001). (d) Specular (00L) and (b) off-specular (20L) for MMOC/MgO(001). (c) and (d) Schematic crystal models showing the observed epitaxial relationships for MMOT/MAO(001) and MMOC/MgO(001).
A surprising finding was the observation of the cubic phase of MMO (MMOC) grown on MgO(001). Specular (00L) and off-specular (20L) CTR scans (Figures 4(d) and 4(e), respectively) of a 10 nm MMO film show clear thickness fringes around the MgO (002) and MgO (202) Bragg peaks, indicating the high quality and uniformity of the MMOC thin films grown from PLD. These also reveal the following cube-on-cube epitaxial relationships: MMOC[001]//MgO[001] and MMOC[404]//MgO[202]. Further evidence is from additional measurements at the substrate half-order peak locations. These half-order peaks are forbidden for MgO cubic phase but are allowed for the spinel cubic phase as the lattice constant of cubic spinel is approximately as twice as that of MgO. The representative MMOC only peaks (i.e., some substrate half-order peaks) can be found in Figure S8. Note that a long-range specular θ–2θ scan (Figure S923) was also performed to check the potential presence of other phases. Neither peaks from MMO tetragonal phase nor other phases such as Mn2O3 (e.g., from the decomposition of MMO) were found. Similar experiments were conducted on 30 nm thick MMO films and the same cubic phase was found (Figure S1023). All these suggest that MMO films are preferentially grown in the cubic phase when grown on MgO(001) substrate via PLD (even using a MMOT target that has the tetragonal phase; Figure S1). Average lattice constants were c = 8.384 Å and a = 8.345 Å for 30 nm MMOC film, indicating a larger deviation from the bulk MMOC lattices (a = c = 8.600 Å (Ref. 24)). The 10 nm MMOC film is fully epitaxial with respect to the MgO substrate, namely, a = 2*aMgO = 8.424 Å, indicating that the strain imposed from MgO substrate brings a closer match for the ultra thin film to the MMOC bulk lattice spacing.
For MMO powders, the cubic phase is realized at temperatures above 950 °C (Refs. 16 and 17) or under a higher pressure ∼15.6 GPa.18 A possible cubic phase of an MMO thin film was recently suggested by sputtering deposition growth that showed a vertical coherence for the MMO thin film on an Al2O3 substrate.25 The formation of highly epitaxial MMOC/MgO and MMOT/MAO could be related to strains imposed by the substrates. As shown in Table I, although MgO and MAO are both cubic phases, MgO has a lattice constant 4.212 Å, while MAO has a lattice constant 8.083 Å. Compared to the bulk lattice constants of 8.600 Å (Ref. 24) for MMOC and 5.727 Å for MMOT, the in-plane strain induced by MgO, , is less than the strain if MMOT were grown on MgO, . However, to grow MMO on MAO, the tetragonal phase has a great advantage due to the smaller in-plane strain, , as compared to that of MMOC on MAO, . Therefore, we conclude that strain is a dominant factor for controlling and stabilizing the different MMO phases as thin films. In particular, MgO substrate may provide the interaction that stabilizes the high T/P cubic MMO phase that has not previously been observed at ambient conditions.16–18
Epitaxial relationship as well as bulk and thin film lattice constants of materials in this work. The in-plane strains are calculated using , where afilm_bulk is the in-plane bulk lattice constant of the film material and asub is the substrate lattice constant of MgO or MgAl2O4 (MAO). For comparison, the last row lists the calculated in-plane strains if MMOC were grown on MAO and MMOT were grown on MgO. R45° denotes the in-plane 45° rotation. Here, we use 30 nm MMOC and 54 nm MMOT thin films as examples to illustrate the role of strain.
Materials . | MgO . | MMOC . | MAO . | MMOT . |
---|---|---|---|---|
aBulk/cBulk (Å) | 4.212 | 8.600 | 8.083 | 5.727/9.284 |
Epitaxy | ... | Cube-on-cube | ... | R45° in-plane |
cFilm (Å) | ... | 8.384 | ... | 9.167 |
aFilm (Å) | ... | 8.345 | ... | 5.675 |
... | 2.09% MMOC/MgO | ... | 0.20% MMOT/MAO | |
εxx | ... | −3.86% MMOT/MgO | ... | 6.40% MMOC/MAO |
Materials . | MgO . | MMOC . | MAO . | MMOT . |
---|---|---|---|---|
aBulk/cBulk (Å) | 4.212 | 8.600 | 8.083 | 5.727/9.284 |
Epitaxy | ... | Cube-on-cube | ... | R45° in-plane |
cFilm (Å) | ... | 8.384 | ... | 9.167 |
aFilm (Å) | ... | 8.345 | ... | 5.675 |
... | 2.09% MMOC/MgO | ... | 0.20% MMOT/MAO | |
εxx | ... | −3.86% MMOT/MgO | ... | 6.40% MMOC/MAO |
The cubic MMO film has quite different properties compared to its tetragonal phase. Previous work on powder materials has shown that MMOC has an electrical resistivity of only ∼10 Ω cm instead of ∼106 Ω cm for the tetragonal phase,17 which is beneficial for a battery electrode. In addition, tetragonal spinel oxides are known to have the distorted oxygen octahedra due to Mn(III) electronic configuration, t2g3-eg1, which induces the Jahn-Teller distortion.26 This is believed to be the primary driving force for capacity fading in spinel materials such as LiMn2O4 as a lithium ion battery cathode. In particular, the tetragonal-to-cubic transition in MMO leads to a redistribution of Mg2+ into both tetrahedral and octahedral sites and will change the inversion degree of MMOC from 0 to 0.45,17 while Mg2+ only occupies the tetragonal sites in the tetragonal phase.16,17
These comparisons of MMOC with respect to powder MMO indicate that MMOC may be a good candidate as a model battery cathode. Although powder MMOC cannot be obtained at room temperature and ambient pressure, the stabilized thin film MMOC in this work might be utilized as a cathode in a thin film battery.6,27 As cations in A site of spinel AB2O4 are extracted and inserted during battery cycling, it is also important to produce a stable A-site empty B2O4 (fully charged state) with the same phase as AB2O4 (discharged state). For this reason, Mn2O4 thin film was grown on MgO(001) and Figure 5 compares MMOC (∼30 nm) and Mn2O4 (∼40 nm) grown on MgO(001). Results clearly show that the epitaxial Mn2O4 thin film is also cubic phase (see Figure S11 for more evidence). Note the cubic Mn2O4 is grown using a non-cubic phase of MnO2 target (Figure S2). To the best of our knowledge, this is the first time where high quality epitaxial spinel Mn2O4 thin films were obtained in the cubic phase. The corresponding (004) and (404) peaks of Mn2O4 are at higher diffraction angles than those of MMOC, indicating lattice constants of c = 8.280 Å and a = 8.417 Å. When compared to the lattice spacings of MMOC (c = 8.384 Å and a = 8.345 Å), the c lattice spacing of Mn2O4 is smaller due to the shrinkage of the crystal lattice as a result of the empty A site, while a of Mn2O4 is closer to 2*aMgO ∼ 8.424 Å, suggesting better in-plane coherence.
Specular and off-specular X-ray diffraction data for thick (∼20 nm) MgMn2O4 (MMO) and Mn2O4 grown on MgO (001) in the vicinity of (a), (c) MgO(002) and (b), (d) MgO(202), respectively. These suggest both MMO and Mn2O4 films are oriented by cubic-on-cubic epitaxy on MgO(001).
Specular and off-specular X-ray diffraction data for thick (∼20 nm) MgMn2O4 (MMO) and Mn2O4 grown on MgO (001) in the vicinity of (a), (c) MgO(002) and (b), (d) MgO(202), respectively. These suggest both MMO and Mn2O4 films are oriented by cubic-on-cubic epitaxy on MgO(001).
IV. CONCLUSION
In summary, we demonstrated the ability to synthesize phase-pure polymorphs of MgMn2O4 (MMO) thin films at room temperature and ambient pressure by PLD. The phase-selected growth of MMO thin films is controlled by the choice of substrate, with growth in the cubic phase (MMOC) on MgO substrates and the tetragonal phase (MMOT) on MgAl2O4 substrates. The ability to growth the Mn2O4 (MgxMn2O4 with x = 0) in cubic phase on MgO(001) (corresponding to a charged cathode) has also been established. As the MMOC is a high temperature (>950 °C)16,17 or high pressure (>15.6 GPa)18 phase in the powder form, the formation of MMOC as thin film at room temperature (25 °C) and ambient condition (101 kPa) is likely due to strains induced by the substrate host structure.
The controllable growth of stabilized tetragonal MMOT phase as well as the cubic MMOC and Mn2O4 phases opens up the use of these epitaxial thin film systems to understand the molecular-scale response of these materials to electrochemical reactions, as model cathodes for a multivalent battery system.6 For example, the ability to grow the cubic phase of MgxMn2O4 at two compositional extremes (x = 0 and x = 1) using the substrate confinement method provides a novel system for understanding the relevant structures and interactions in model battery cathode interfaces. MMO can be used as a model for the discharged cathode relevant to a Mg battery,28 while Mn2O4 represents the charged cathode for many battery chemistries (e.g., Li-, Zn-, Ca-batteries). Our recent work on MMOC thin films successfully demonstrated reversible electrochemical insertion and removal in a Mg cathode using this strategy,6 and this approach can also be used to study similar reactions for other cations (Li+, Zn2+, Mg2+, Ca2+) due to the stabilization of a specific host phase.
ACKNOWLEDGMENTS
This work was supported as part of the Joint Center for Energy Storage Research, an Energy Innovation Hub funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences. The Advanced Photon Source is supported by the DOE under Contract No. DE-AC02-06CH11357. This work made use of the Northwestern University Central Facilities supported by the Materials Research Science and Engineering Center (MRSEC) through National Science Foundation (NSF) under Contract No. DMR-1121262. The authors thank the beamline technical support from Christian M. Schlepuetz and Jenia Karapetrova at Sector-33, of APS. Authors thank Chunjoong Kim, Tanhong Yi, and Jordi Cabana from the University of Illinois at Chicago for helping synthesizing PLD target materials, Michael D. Biegalski for PLD growth support at Oak Ridge National Laboratory. The PLD preparation and characterization was conducted at the Center for Nanophase Materials Sciences, which is sponsored at Oak Ridge National Laboratory by the Scientific User Facilities Division, Office of Basic Energy Sciences, U.S. Department of Energy.