Brought to you by:

Structural and Electrochemical Investigation of Li ( Ni0.4Co0.15Al0.05Mn0.4 ) O2 Cathode Material

, , , , and

Published 6 October 2010 © 2010 ECS - The Electrochemical Society
, , Citation C. Rumble et al 2010 J. Electrochem. Soc. 157 A1317 DOI 10.1149/1.3494211

1945-7111/157/12/A1317

Abstract

was investigated to understand the effect of replacement of the cobalt by aluminum on the structural and electrochemical properties. In situ X-ray absorption spectroscopy (XAS) was performed, utilizing a novel in situ electrochemical cell, specifically designed for long-term X-ray experiments. The cell was cycled at a moderate rate through a typical Li-ion battery operating voltage range. XAS measurements were performed at different states of charge (SOC) during cycling, at the Ni, Co, and the Mn edges, revealing details about the response of the cathode to Li insertion and extraction processes. The extended X-ray absorption fine structure (EXAFS) region of the spectra revealed the changes of bond distance and coordination number of Ni, Co, and Mn absorbers as a function of the SOC of the material. The oxidation states of the transition metals in the system are , , and in the as-made material (fully discharged), while during charging the is oxidized to through an intermediate stage of , is oxidized toward , and Mn was found to be electrochemically inactive and remained as . The EXAFS results during cycling show that the Ni–O changes the most, followed by Co–O, and Mn–O varies the least. These measurements on this cathode material confirmed that the material retains its symmetry and good structural short-range order leading to the superior cycling reported earlier.

Export citation and abstract BibTeX RIS

Environmentally friendly and economical lithium battery electrode materials with higher capacity and good cycling stability are in high demand. With the emergence of portable telecommunications, computer equipment, and, ultimately, hybrid electric vehicles, there has been a great demand for less expensive batteries with longer lifetimes and smaller sizes and weights. has been the backbone for commercial use in Li-ion batteries since their inception by Sony due to its ease of production, stable electrochemical cycling, and acceptable specific capacity. The relatively high cost and toxicity of cobalt and the quest for materials with higher capacities has led to the study of possible alternatives, though many have limitations precluding their widespread use. is thermally unstable at high states of charge.1 Variants of exhibit good structural stability on overcharge but have low capacity and exhibit high solubility of in the electrolyte at elevated temperatures above .2 Layered systems such as have attracted attention as a cathode for lithium-ion batteries especially for hybrid electric vehicle applications.2 Compared to , this system proved to possess enhanced electrochemical performance at a lower projected cost.3 Other layered materials under consideration as cathode materials for Li-ion batteries include NMCs, which have the general formula (Ref. 3), where y is most commonly equal to . There have recently been attempts to further reduce the Co content (for example, ).4 This has met with some success, but the rate capability of the electrodes tends to decrease as the Ni content is raised due to increased antisite mixing. However, there is a great deal of interest in the synthesis and characterization of layered compounds with reduced Co content.47

Recently, , (Ref. 8), and , (Ref. 9), compounds were studied as cathode materials. For low values of y, aluminum substitution appears to improve the rate capability and cycling stability compared to the baseline materials. Other studies on Al substitution of layered oxide structures suggest that the thermal abuse tolerance is improved.10 These results motivated us to investigate specific Al-substituted materials using in situ X-ray absorption spectroscopy (XAS) in order to fully characterize the structural properties of this material. An ultimate goal is to allow design of low Co content layered oxides with improved electrochemical properties.

There has been a continuous effort to study the detailed changes that occur in various electrode materials during the charge–discharge process. Real-time tracking of the structure and valence changes during delithiation, relithiation, and prolonged cycling of layered cathode materials may have predictive value for determining the intrinsic electrochemical characteristics and can be achieved by employing an in situ electrochemical cell. XAS, accompanied by simultaneous electrochemical measurements, provides valuable and important information about the relationship between the structure and the electrochemical properties of the electrode materials,11 information that is not always accessible utilizing ex situ experimental measurements.

In this investigation, we present a comprehensive transmission XAS study to investigate electrodes (as used in Li-ion batteries) at various states of charge. Information about the oxidation state of the investigated transition metal elements (in this case, Mn, Co, and Ni) and their electronic configuration was obtained from the X-ray absorption near-edge spectroscopy (XANES) region of the K-absorption edges, whereas the extended X-ray absorption fine structure (EXAFS) region was used to probe the structure around the X-ray absorbing atoms.

Experimental Procedures

powder was synthesized and electrodes were prepared as previously described.9 The specifications of the electrode are given in Table I, where the specific capacities were calculated based on the valence changes and the amounts of the elements in the compound. Phase purity was assessed by X-ray powder diffraction using a Phillips X'Pert diffractometer with an X'celerator detector and Cu radiation. The experiments were performed using a complete half-cell. A special electrochemical in situ XAS cell was used for these experimental investigations; for a complete detailed design and assembly, the reader is referred to Deb et al.12 A sheet of porous polypropylene membrane (Celgard 3400) was utilized as a separator and 1M in ethylene carbonate (1:1 volume, Merck, LP40) as the electrolyte. A Li metal foil cut into a circular disk (diameter ) was used as the counter electrode. The charge–discharge cycling was performed at room temperature at the beam line with a potentiostat/galvanostat system (Princeton Applied Research, Model Versa) under constant current control. In situ XAS measurements performed during charging and discharging were carried out at various states of charge of the electrode. Charging and discharging were performed at a constant current density of between 1.0 and , which corresponds to about the C/25 rate, assuming that all of the Li could be extracted. The measurements performed at different points along the charging cycle are labeled in Fig. 1. The charging was paused during the XAS measurements to allow the electrode systems to stabilize. The spectra presented here were obtained during the charging process at different states of charge (to versus Li). For the measurements with the model compounds shown here for comparison, the samples were diluted with boron nitride (BN) (samples were mixed in a 1:10 ratio with BN using a mortar and pestle) and pellets were prepared. The sample pellets were then finally loaded on aluminum holders using Kapton adhesive foil on both sides of the sample.

Table I. Electrode properties and experimental conditions.

Formula weight: 94.5815
to specific capacity
to specific capacity
Total specific capacity
Active material weight
Electrode area
Active material loading
Total current
Figure 1.

Figure 1. Charging curve of the electrode. Labels shown A to O are the points where the measurements were performed during the charging cycle corresponding to (A), 0.96 (B), 0.92 (C), 0.88 (D), 0.84 (E), 0.80 (F), 0.76 (G), 0.67 (H), 0.55 (I), 0.49 (J), 0.38 (K), 0.31 (L), 0.27 (M), 0.23 (N) and 0.19 (O).

The XAS measurements were performed in the transmission mode at the bending magnet beamline station D of the DND-CAT (Sector 5), at the Advanced Photon Source, using a water cooled Si(111) double crystal monochromator, and the energy resolution of the monochromatic beam was determined to be . A beam size of about was used for the beam to pass easily through the in situ cell X-ray window resulting in an incident photon flux of . For both Mn and Ni K-absorption edges, the monochromator was scanned from 200 below to above the K-absorption edge, while the Co K-absorption edge data range was limited to a maximum of at the onset of the Ni K edge. The EXAFS data presented were analyzed using a combination of EXAFSPAK software package13 and ATHENA.14 The resulting function was weighted with to account for the damping of oscillations with increasing k. The radial structure functions presented here were obtained by Fourier transformation of using a k range of for Mn and Ni and for Co.

Results and Discussion

crystallizes in the R-3m space group and exhibits a well defined structure as does the parent material .9 Al substitution causes the a unit cell parameter to decrease by 0.1% and the c unit cell parameter to increase by 0.2% compared to the parent material. There is also an increase in the degree of antisite mixing (Ni in the 3a or lithium sites) from 6.6% to 7.3% and a slight increase in the lithium slab spacing, which results in improved lamellarity upon substitution with Al.9

In general, the shape of the K-edge XANES of the transition metal oxides provides unique information about the site symmetry and the nature of the bonding with surrounding ligands, while the threshold energy position of the absorption edge provides information about the oxidation state of the probed atom. To extract the information of the initial oxidation state of the transition metals in this material and elucidate the charge compensation mechanism in this system, Mn, Co, and Ni K-edge XAS experiments were carried out during the charge/discharge cycle at a constant current density of . The specific capacity was calculated from the elapsed time, current, and the mass of the active material in the cathode, assuming that all the current passed was due to the Li deintercalation. The comparison of the XANES region before charging [ (point A) in Fig. 1] with that of the model compounds is shown in Fig. 2. The comparison [Fig. 2a] reveals that the edge energy is essentially the same as that of , indicating that the oxidation state of Mn in this material at the start of the charging cycle is nearly all tetravalent . Figure 2b indicates that the XANES spectrum for the Co K edge of the sample is identical to that for , indicating that Co in the sample at the beginning of the charging is . Figure 2c shows the comparison of Ni K-edge XANES spectrum with that of the model compounds of nickel (II) oxide and layered . At , the XANES spectrum is similar to that of the nickel(II) oxide, indicating that the Ni in this system is divalent , whereas the XANES spectrum of the layered is similar to that observed at , confirming that Ni is in the trivalent state at .

Figure 2.

Figure 2. Normalized XANES comparison for of (a) Mn K edge at with the model compound and ; (b) Co K edge observed at with the model compound , and (c) Ni K edge at and with the model compounds nickel (II) oxide and , respectively.

XANES spectra for selected points during the charging process for Mn, Co, and Ni edges are shown in Fig. 3. For the Mn K-edge XANES [Fig. 3a], during charge, the edge position of the scans at different states of charge did not exhibit any significant edge shift to higher energies, though it does exhibit some changes in regard to the shape of the edge due to the change in the local environment of the Mn in the system as Li is deintercalated. The edge does not exhibit any rigid shift to higher energies, however, suggesting that the Mn oxidation state remains unchanged during the charging procedure; i.e., the atom is electrochemically inactive. The Co K edge [Fig. 3b] shows a progression of the entire pattern from lower energy to higher energy as a function of the decreased Li content (at the different charge states), indicating the oxidation of toward .

Figure 3.

Figure 3. Normalized XANES spectra at different states of charge of (a) Mn K edge, (b) Co K edge, and (c) Ni K edge, respectively. The insets show the nature of the pre-edge peaks of the respective XAS spectra; (d) Comparison of the linear combination (LC) best fit of the and spectra (with a composition of 62%, of and 38%, of ) with the spectra.

XANES spectra for the Ni K edge shown in Fig. 3c indicate that the Ni ion undergoes a two-stage change during the charging cycle, where the first stage ranges from to and the second stage is from to . Beyond (Fig. 4), the Ni K-edge XANES did not move significantly to a higher energy. This two-stage reaction observation may be attributed to the two one-electron reactions of the Ni ions ( and ) during charging. Similar observations of the redox reaction have been reported previously by Koyama et al.15 using ab initio calculations for consisting of , , and in the ranges of , , and , respectively. During discharge, these processes are reversed electrochemically: Li is inserted in the lattice, is reduced to , is reduced to , and M–O (metal–oxygen) and M–M (metal–metal) bond lengths return to their original values. For the purposes of this discussion, the results during charge will be discussed hereafter.

Figure 4.

Figure 4. Plot of the white line energy shift vs the x in , for the Ni K edge. The filled symbol (●) represents the data during the charge, while the empty symbol (o) represents the data during discharge.

Figure 5.

Figure 5. Mn-K edge -weighted Fourier transform at selected Li content (x) during the charge cycle (k ). M represents the transition metal atom.

Figure 6.

Figure 6. Co–K edge -weighted Fourier transform at selected Li content (x) during the charge cycle (k ). M represents the transition metal atom.

Figure 7.

Figure 7. Ni–K edge -weighted Fourier transform at selected Li content (x) during the charge cycle (k ). M represents the transition metal atom.

To make sure of the presence of the species at , we see that the absorption edges for Ni in the various states of charge cannot be represented by linear combinations of only and spectra. Figure 3d shows a comparison of the spectrum calculated as a linear combination of the two species of and spectra and the spectrum for . The important areas of mismatch are the region near (just before the absorption maximum), the absorption maximum, and regions near 8358 and (after the absorption maximum). The mismatch in the absorption maximum is an indication of the presence of instead of a combination of and . This evidence indicates that is an intermediate state for the Ni during oxidation/reduction cycling.

The preedge peaks of the respective XAS spectra are also shown in Fig. 3. These give additional information on the nature of the electronic states. For most transition metal elements, the preedge peaks that occur well below the main edge ( below) are assigned to transitions to empty states with d-like character transitions,16, 17 where n represents the initial number of d electrons and includes the excited electron in the final state, which includes the effect of the core hole. These transitions are directly allowed through a very weak quadrupole transition18 or allowed via an admixture of 3d and 4p states.19 The weak preedge in the Mn-absorption (Fig. 3a, peaks A and near 6542 and ) is the formally electric dipole-forbidden transition of a 1s electron to an unoccupied 3d orbital of a high-spin ion, which is partially allowed because of the pure electric quadrupole coupling and/or the 3d–4p (or Mn 3d–O 2p) orbital mixing. The presence of weak preedge intensity is indicative of octahedral coordination as opposed to tetrahedral coordination which would have resulted in a stronger preedge intensity.20 The two peaks (in the preedge region) for Mn are discernable, as Mn is in the tetravalent state , and a single peak structure in the preedge region is characteristic of trivalent Mn compounds. The latter single peak characteristic for trivalent manganese systems can be interpreted as splitting of the and energy levels modified by the Jahn–Teller distortion, though this is not observed in the studied system.21

For the Co preedge (Fig. 3b), the Co transitions are not well structured here, but two peaks (C and ) are still resolvable, as well as a weak feature (C and ) above with indication of an energy splitting in between C and is assigned to the transition, which is dipole and symmetry forbidden, but occurs, albeit very weakly, due to the mixing with the oxygen p orbitals and quadrupole transitions and transition which is dipole forbidden, and hence much weaker than the symmetry- and dipole-allowed main transition. The two transitions present here indicate at least a partial high-spin configuration, since the peaks are not clearly split, which may be due to the delocalized electrons in the band structure or an overlapping mixture of low-spin , intermediate , and high-spin states. The shoulder (B) around is the shake-down process, which occurs at a lower energy, since the binding energy decreases due to the core-hole screening effect in Co by an electron transferred in from the ligand.22, 23

Since the intensity of the preedge peak (D, in Fig. 3e) can be used as an indicator for the geometry of octahedral complexes, the presence of very weak Ni preedge during the cycling signifies that Ni atoms are not highly symmetrical, and the weak preedge arises due to the possible distortion of the octahedral 3b site in the rhombohedral R-3m space group, which results in the removal of the inversion symmetry.

Figure 4 is a quantitative picture of how the redox process progresses. It shows a plot of the relative white line peak position as a function of amount of Li present (x), in contrast to the edge position, as defined by the energy at half of the edge step for the K edge, observed during charge and discharge. The Co and the Ni peak positions change by and , respectively, as the lithium content changes from the fully discharged state to the charged ( versus Li) state. Though there is still a controversy regarding charge compensation via oxygen in these systems, the results here show clearly the contribution of the Ni and the Co ion to the charge compensation process. In a similar system, Yoon et al.24 reported that oxygen oxidation was part of the charge compensation process. In contrast with Yoon et al. , who reported only small shifts in metal XANES above 50% state of charge, we observe continuous shifting of both the metal XANES and the metal–oxygen bond lengths. There is thus no need to invoke oxygen oxidation; however, our data cannot rule out the possibility that a minor component of the charge compensation results from oxygen oxidation.

The local structure of the Mn, Co, and Ni atoms in the system was obtained from EXAFS measurements. The backgrounds were subtracted by extrapolating a Victoreen-type function from the preedge region, and EXAFS oscillations were extracted using cubic spline baseline functions. In all cases, the Li contribution to the EXAFS was ignored because of the low backscattering ability of the lithium. The FT (Fourier transform) of the data at selected SOCs is shown in Fig. 5, 6 and 7. The first peak (, labeled S) in the FT is attributed to the M–O interactions, while the second peak (, labeled T) is due to the M–M interactions. During the cycling process, the Mn–O distances remain nearly unchanged in contrast to the gradual dwindling of the Co–O distances accompanied by significant decreases in the Ni–O distances (Fig. 8). This change can be related to the change in the valence state in the transition metals where in an octahedral crystal field, the d orbitals split into a triply degenerate orbital set and a higher-energy doubly degenerate orbital set , which leads to an outermost electronic configuration for as . On the other hand, the outermost configuration for can be represented as . In a simple molecular-orbital picture, the orbitals are metal–ligand antibonding while the orbitals are metal–ligand nonbonding. As a consequence, changes in occupancy have a much larger effect on M–O distance than do changes in occupancy. Hence during the oxidation reaction of the charging cycle, the energy change for (or ) would be larger (than for Co) as the change takes place between the lower and the higher sets, inducing a larger change in ionic radius from to .

Figure 8.

Figure 8. Metal-oxygen first shell coordination bond length changes during cell cycling. to 0.19 represent first charging cycle and to 1.0 represent first discharge cycle.

From Fig. 6 and 7, it can be seen that the amplitude of the Co–O peak remains almost unchanged while the Ni–O peak increases during charging. These observations can best be explained if we consider the simpler systems of and . For studied earlier,24 Ni is surrounded by oxygen atoms at two different distances and the distortion of the octahedral coordination is consistent with the presence of the Jahn–Teller effect due to the in the low-spin state (in addition to the distortion inherent in R-3m symmetry). On delithiation of , the increase of the Ni–O peak amplitude is ascribed to the change in the oxidation from to , and since is not a Jahn–Teller active ion, the gradual change of to during charging diminishes the effect of the Jahn–Teller distortion and results in an increased amplitude of the Ni–O peak. Here in this system based on the above observation, we can explain that during delithiation from to (when Ni is oxidized from to ), a diminution of the Jahn–Teller effect results in increased amplitude of the Ni–O peak. On the contrary, from to (when Ni is oxidized from to ), increase of the Jahn–Teller effect results in a decrease of amplitude of the Ni–O peak. For the delithiation there is no change in the amplitude of the Co–O peak due to the absence of the Jahn–Teller effect for . Thus similarly for this system, during charging, Co changes from to neither of which are Jahn–Teller active ions (except that low-spin will be Jahn–Teller active, albeit with a small distortion), and thus a change in amplitude of the Co–O peak is not observed.

Structural parameters of the Mn, Co, and Ni absorbers were analyzed based on a two shell model, where bond distances (R) and the Debye–Waller factors were left as free parameters and the coordination number for Mn–O, Co–O, and Ni–O was kept fixed to the crystallographic value of 6 (Fig. 8). During cycling, the change in the Ni–O bond length is the largest , followed by Co–O , and the Mn–O change is negligible . The Ni–O bond length at the start of charge is , while at it is , and finally at it is . These Ni–O distances are consistent with octahedrally coordinated systems where the bond length is about , exhibiting a static Jahn–Teller distortion shows an average bond at , and finally the bond length is .2529 Thus these fit results confirm that the average oxidation state of Ni ion at the start of the charge is , whereas at it is , and finally at the end of the charge, Ni ion is close to . The significant change in the Ni–O bond length can be explained by the change in oxidation state of since the ionic radius of is larger than that of . For Co, during charging, the observed changes in the Co–O distance are attributed to the oxidation of to , and the predicted decrease in Co–O distance between and is .29 Using Ref. 30, it can be estimated that between and 50% of the Co is oxidized during the charging cycle. Finally, for the metal–metal (Mn–M, Co–M, and Ni–M) interaction fits, we have taken two approaches. In one approach, we have fitted the experimental profile, taking Co as the scatterer (Fig. 9), but because this scattering atom can either be Mn, Co, or Ni (though not Al because it is too light), we also did fits with the Mn and Ni atoms. This was done to understand the sensitivity of the data, taking Co as the scatterer. In this approach we found that if instead of Co it was fitted with Ni or Mn, the corresponding bond distances shown in Fig. 9 during the cycling change by for Ni and for Mn. In the other approach, which actually gives similar result as shown in Fig. 9, we fitted the M–M peak by taking into consideration the ratio of Mn, Co, and Ni present initially at the beginning of the charging cycle. Regardless of the absolute accuracy of the apparent M–M distance, the changes in distance (Fig. 9) should be reliable.

Figure 9.

Figure 9. Metal-metal first shell coordination bond length changes during cell cycling. to 0.19 represent first charging cycle and to 1.0 represent first discharge cycle.

Conclusions

In situ XAS characterizations for Al-substituted have been performed during the first charge and discharge process, providing us with an excellent tool for analyzing the changes that occur when Li is cycled in and out of the layered lattice. During cycling, Mn is electrochemically inactive and remains tetravalent , whereas divalent nickel is oxidized mostly to tetravalent nickel , passing through an intermediate stage of trivalent nickel , and about of is converted to during charge. Bond length changes upon delithiation and relithiation of this system appear to be very reversible, which may explain the excellent cycling stability that has been previously observed.

Acknowledgments

Work was performed at the DND-CAT beam line, which is supported by the E. I. DuPont de Nemours and Co., the Dow Chemical Company, the U.S. National Science Foundation through Grant no. DMR-9304725 , and the State of Illinois through the Department of Commerce and the Board of Higher Education Grant no. IBHE HECA NWU 96. This work was supported by the Assistant Secretary for Energy Efficiency and Renewable Energy, Office of Vehicle Technologies of the U.S. Department of Energy under Contract no. DE-AC02-05CH11231 .

University of Michigan assisted in meeting the publication costs of this article.

Please wait… references are loading.
10.1149/1.3494211