1. Introduction
Sodium-ion batteries (NIBs) are nowadays considered to be the most promising candidate to replace Lithium-ion batteries (LIBs) in the near future, mainly due to the chemical similarity of lithium and sodium ions and the high natural abundance of sodium [
1,
2]. The main limiting issue for NIBs application is related to their cathode counterpart, which underwent uncontrollable phase transitions and volume changes during repeated charge-discharge cycles, preventing short-term battery commercialization. For this reason, the main efforts of researchers have been devoted to developing high-performance cathodes [
3,
4]. The most intriguing of the identified materials are the transition layered metal oxides, Na
xMO
2 (M = Mn, Co, Ni, Fe), due to their high theoretical specific capacity [
5]. This is particularly true for Na
0.67MnO
2 (NMO) with a P2-type structure (following the current Delmas notation for this class of materials [
6]) and a capacity of > 175 mAh/g [
7]. The P2-type phase, with AA-BB octahedral layer stacking, allows the direct migration of sodium between adjacent prismatic sites, thus enabling rapid Na
+ ions transport and showing better applicability with respect to the O3-type phase. In addition, NMO has low-cost constituting elements and the environmental friendliness of manganese. Unfortunately, some challenges are related to the irreversible structural distortions caused by Mn
3+ (Jahn-Teller) ions, poor cycling stability due to severe volume changes and mechanical stress during intercalation/deintercalation, and an unsatisfactory energy density limited by the low redox potential of the Mn
3+/Mn
4+ redox couple [
5]. Attempts to solve these issues include cation doping or co-doping and nano-structuring, as well as the formation of composites with graphene, strategies used for many other electrode materials [
8,
9,
10,
11,
12,
13,
14,
15,
16,
17,
18]. As has been well demonstrated in the battery literature, the samples’ morphology plays an important role in the electrochemical performances. It was proven that, for layered materials, the performance can be optimized by adjusting the Na/Mn ratio, the sintering temperature, and the cooling rate [
13], which at the same time can be useful to control the Mn vacancies that were suggested to form on transition metal layers due to the uptake of oxygen during the cooling process, resulting in an increase in the manganese oxidation state and the formation of manganese-deficient phases [
11,
19]. It was suggested that the quenching process from high temperatures could avoid the Mn vacancies, stabilizing an orthorhombic P2 form with the
Cmcm space group [
7,
10]. However, a high Mn
3+ amount is still present in the sample, making unavoidable the multiphase transitions during cycling. A clear correlation between the cooling treatment and structure stabilization was not identified, particularly for the doped Na
0.67MnO
2 samples [
11,
15]. The doping was suggested to improve the structural stability during sodiation/desodiation and, indeed, the long-term cyclability. One of the most investigated doped compounds is Na
0.67Mn
0.667Ni
0.333O
2, with the substitution of 1/3 of the manganese in the original NMO with nickel: in this case, an increase in the redox potential and the improvement of cycling performance have been demonstrated [
8]. Starting from this stoichiometry, other compositions were obtained by substituting manganese with variable amounts of Fe, Co, Ti, Al, Mg, and Zn or a combination of them, achieving very complex stoichiometries [
9,
10,
11,
12,
13,
14,
15]. However, Ni and Co are toxic and expensive elements that should be avoided for the necessary transition toward the next generation of sustainable materials for NIBs. Concerning the other dopants, it seems that divalent cations may enhance the structural stability of the main phase by increasing the oxidation state of manganese and limiting the Jahn-Teller effect related to Mn
3+ ions. In addition, some of them, such as copper, a harmless element, which is less expensive with respect to Ni, can improve the air/water stability of the P2 phase. It was also proven that the Cu
2+/Cu
3+ redox couple is electrochemically active in P2-type structures [
20]. Iron, too, is a cheap and abundant ion, itself possibly contributing to the redox processes and helping to limit the phase transformations [
21].
The aim of the present paper was to verify the effect of both doping (with the substitution of manganese with harmless and inexpensive dopant ions) and cooling treatment on the physico-chemical properties of Na0.67MnO2. In particular, we performed a systematic study of the stabilized phases and morphology of undoped and Cu- or Fe-doped Na0.67MnO2 samples synthesized via the sol-gel route, with natural cooling or quenching from high temperature to room temperature. Wide use has been made of X-ray powder diffraction, with Rietveld structural refinements and the employment of Scanning electron microscopy with Energy-dispersive spectroscopy. The Mössbauer (for Fe-doped samples) and Electron paramagnetic resonance spectroscopies allowed us to determine the oxidation states of transition metal ions and the possible presence of magnetic impurities. The structural stability of the samples in the air was studied by X-ray powder diffraction. Preliminary cyclic voltammetry and charge-discharge measurements were performed to evaluate the effect of doping on capacity values and capacity retention; the results were discussed and interpreted on the basis of the structural and spectroscopic findings.
2. Materials and Methods
2.1. Synthesis
Na0.67Mn1-xMxO2 (M=Cu, Fe; x = 0 or 0.2) samples were synthesized via the sol-gel method, with two different cooling thermal treatments of slow cooling or quenching.
(CH3COO)Na (Merck, ACS grade), (CH3COO)2Mn 4H2O (Sigma-Aldrich, >99%), and (CH3COO)2Cu H2O (Merck, ACS grade) or (CH3COO)2Fe (Sigma-Aldrich, 95%), were weighed in the proper amounts (Na:Mn:M = 0.67:1−x:x; x = 0 or 0.2) and dissolved in about 100 mL of distilled water, together with citric acid (2:1 moles ratio with respect to the sum of the reagents). In order to compensate for the possible loss of sodium due to the annealing process, a 10 wt % excess (with respect to 0.67 moles of Na) of sodium acetate was added during the synthesis. The solution was stirred, heated to 65 °C, and maintained overnight until all the solvent had evaporated. The obtained white powder was heat-treated in an oven at 350 °C for 5 h to remove the organic component. Afterward, the powder was ground, pelletized, and heat-treated at 800 °C for 12 h following two different cooling treatments: in the first one, the pellets were allowed to slowly cool to room temperature in the oven and in the second one, the pellets were quenched via rapid extraction from the oven. The pellets were ground down and the resulting powders were preserved in a glovebox to avoid contact with the atmosphere until their subsequent use. A portion of the powders was also maintained in air for the structural stability study. In the following sections, the Na0.67MnO2, Na0.67Mn0.8Cu0.2O2, and Na0.67Mn0.8Fe0.2O2 samples will be named NMO, Cu02, and Fe02 for slowly cooled samples, with the same acronyms followed by Q for the quenched samples.
2.2. Instruments
X-ray powder diffraction (XRPD) measurements were performed using a Bruker D5005 diffractometer (Karlsruhe, Germany) with CuKα radiation (40 kV, 40 mA), graphite monochromator, and scintillation detector. The patterns were collected in the angular range of 10–120°, with a step size of 0.03° and a counting time of 12 s per step in a silicon sample holder with low background. Rietveld structural and profile refinement was carried out by means of TOPAS 3.0 Bruker software [
22], on the basis of the known crystal structure models of the layered polymorphs. During the refinement, the background coefficients, scale factor, zero or displacement error, lattice parameters, crystallite sizes, isotropic thermal factors, and atomic positions were varied. The occupancies of sodium and dopant ions were fixed to stoichiometric values, due to the complexity of the samples and to the similar X-ray scattering factors of Mn, Cu, and Fe ions. To evaluate the structural stability, the patterns were also collected on all the samples in the 10–70° 2θ range, step 0.03° and 2s/step of counting time after 7, 14, and 30 days of the maintenance in air.
A Tescan Mira 3 (Tescan USA Inc., Warrendale, PA, USA) scanning electron microscope (SEM) was used for the morphological study on gold-sputtered powder samples. Energy-dispersive X-ray spectroscopy (EDS) analysis was also performed with the same instrument on non-sputtered powder samples.
Element measurements were performed via inductively coupled plasma optical emission spectroscopy (ICP-OES iCAP 7400, Thermo Fisher Scientific, Waltham, MA USA) equipped with a concentric nebulizer, cyclonic spray chamber, and ceramic torch, according to the operating conditions suggested by the manufacturer. The linearity range of intensity vs. concentration was obtained using standard solutions (1–10 mg/L) prepared from a 1 mg/mL stock solution. Small amounts, exactly weighted, of each sample were treated with 0.5 mL of ultrapure 65% HNO3 and 1.5 mL of ultrapure 37% HCl, refluxed for 10 min, evaporated to a small volume, and then diluted to 50 mL with ultrapure water. Clear solutions were obtained and analyzed for the determination of element content.
The Mössbauer measurements were performed by means of a standard Mössbauer setup (home assembled) in transmission geometry. The spectra were collected by means of a Kr-CO
2 proportional counter, Fast
TM (LND Inc., New York, USA) electronics for gamma-ray spectroscopy, and a Wissel
TM spectrometer (Blieskastel, Germany), which was run in sinusoidal acceleration mode (v
max = 4.0 mm/s) and calibrated using a standard metal iron foil. The γ-ray source was a 25-mCi
57Co in rhodium matrix with the Lamb-Mössbauer factor
f = 0.63, as measured by applying the method described by Spina and Lantieri [
23]. Approximately 33 mg/cm
2 of each compound was used for the measurements. The Mössbauer spectra were interpreted by means of a fitting procedure, based on the evaluation of the transmission integral function, which takes into account the dependence of the Mössbauer spectra on the sample’s effective thickness. The complete expression used to fit the spectra was:
where
and
are the detected counts and the spectrum baseline, respectively, as a function of the transducer velocity
[
24]. Moreover,
is the reduced recoilless fraction of the source and
is the Voigt distribution (having
and
as center and FWHM, respectively) used to describe the source line shape. The Voigt profile has a Lorentzian component with natural line width, while that of the Gaussian one is suitable to reproduce the total linewidth of the source provided by the manufacturer (Γ
S = 0.114 mm/s). Finally,
is the absorption cross-section of the sample as a function of the energy
, expressed in mm/s, and
is the effective thickness of the sample. In the limit of “thin absorption approximation” (
< 1), each contribution to
is expressed as a Voigt doublet, having a Lorentzian component with the natural line width (Γ
n) and a Gaussian one with a broadening σ, describing a particular distribution of hyperfine parameters [
25]. Consequently, the total linewidth of each contribution is approximately given by Γ
tot = Γ
S + Γ
n + σ. For both samples, a rather good agreement between experimental and best-fit data was obtained with χ
2 ~ 1200 for the 512 points.
Electron paramagnetic resonance (EPR) measurements have been performed at about 9.46 GHz with a Bruker spectrometer (Karlsruhe, Germany). Particular care has been devoted to determining sample mass and position in the resonant cavity for the comparison of signal intensities (areas). The temperature dependence of the spectra has been investigated in the range of 115–370 K.
The slurries for the electrochemical measurements were prepared by mixing the active materials with carbon (Super C65) and PVdF binder (Solef 5130) at a weight ratio of 80:10:10 in N-methyl-2-pyrrolidone (NMP) (Sigma-Aldrich) and magnetically stirred for about one hour. Afterward, the slurries were coated onto an aluminum foil using a homemade doctor blade, maintained at room temperature until dried, heat-treated at 80 °C overnight in a vacuum oven, then hot-pressed at 200 psi at 100 °C for 5 min. The slurries were maintained in a glove box (MBraun, Garching bei München, Germany, O2 < 1 ppm, H2O < 1 ppm) with an Ar atmosphere to avoid contact with moisture.
The electrodes were cut in the form of discs (1 cm in diameter) with a mass loading of about 3 mg/cm2. Swagelok cells were assembled in the glove box, with the slurries acting as the working electrode, Na metal as the reference and counter electrode, and a Whatman GF/A disc as the separator. The chosen electrolyte was 1 M NaPF6 in EC:DEC (1:1 wt %).
Cyclic voltammetry (CV) was performed by using an Autolab PGSTAT30 (Eco Chemie, Metrohm, Utrecht, The Netherlands) at a scan rate of 0.1 mV/s for five cycles in the potential range of 1.7–4.4 V. For galvanostatic charge–discharge tests, the Swagelok cells were cycled on a Neware (Hong-Kong, China) battery tester in the same potential range for 10 cycles at C-rates of between C/5 and 5C, after a conditioning cycle at C/20 and C/10.
4. Discussion
The Na0.67MnO2 cathode material is currently attracting attention for its application in SIBs, mainly due to the high capacity and structural stability of the P2 phase, the main stable polymorph for this sodium composition. The wide range of possible cationic substitutions further allows for improving its functional properties. However, to develop electrochemically performing manganese-based P2 cathode materials, the elemental substitution necessitates the control of the Mn oxidation state. We achieved this goal by the combined use of structural and spectroscopic techniques.
Cu- and Fe-doped Na
0.67MnO
2 samples were successfully synthesized as mixtures of phases [
31,
32] due to the low temperature of thermal treatment, 800 °C, that was intentionally chosen to verify if the stabilization of the polymorph mixtures could allow to reach higher capacities or achieve better cycling stability, thanks to the favorable cushioning of structural transitions [
7,
11,
36].
As previously demonstrated by XRPD, P2 and P’2 were the main phases for all the samples, apart from NMO-Q. This is in line with the literature findings on Na
0.67Mn
1-xMg
xO
2 [
11], in which the quenching process favors, as in our case,
Cmcm stabilization for undoped and slightly doped samples, apart from x = 0.2, the same doping level used in the present work, for which the P2 phase is instead stabilized. In our case, in fact, the doping always allowed the stabilization of the P2 hexagonal polymorph as the main phase, independently of the cooling treatment, even if the amount of the P’2 phase increased in the quenched samples. This observation, based on the XRPD evidence (
Table 1), was also consistent with the EPR results. In fact, the P’2 phase is stabilized when a low number of Mn vacancies is formed, with a consequent reduction in the amount of Mn
4+ ions, which are, therefore, preferentially surrounded by Mn
3+ ions, thereby giving rise to broad EPR lines with g-values that are generally far from the ideal 1.996 value, which is typical of samples with Mn
4+ as the unique magnetic ion. For slowly cooled samples, the EPR spectra of the Cu-doped samples showed a lower intensity with respect to pure and Fe-doped samples. This evidence suggests a complex behavior that cannot be explained only on the basis of the possible variation of manganese oxidation states as a consequence of the dopant introduction or quenching treatment [
11].
O-type polymorphs were also present (
Table 1), apart from the Cu02 sample, which only contained P-type polymorphs. The intergrowth of the P- and O-type phases is not new for layered cathodes: similar composites were obtained experimentally but were also simulated by structure modeling to calculate the diffraction patterns, to be compared with the complex experimental patterns [
27,
31,
32,
37]. For slowly cooled samples, the O polymorphs’ amount is similar for NMO and Fe02, while the presence of Cu ions suppresses the gliding of planes, avoiding the segregation of O-type phases. For the quenched samples, the O-type phases are present in all the samples in a higher amount for NMO-Q, demonstrating that the cooling treatment is a prevailing factor with respect to doping.
We demonstrated that iron and copper can easily substitute for manganese on the octahedral sites of the layers, as suggested by the absence of phases only containing the dopant ions and by the changes in lattice parameters with respect to undoped NMO, which was particularly evident in the quenched samples. The Mössbauer data suggested that iron ions were present in both the P-type polymorphs, with a preference for the more regular octahedral sites of the P2 structure. The distribution of the dopants in both the structures can be predicted by the unit cell volume of P2 and P’2: after doping, they are about 2–3% and 4–6% higher with respect to the undoped sample. For the main P2 phase, both the
a and
c crystallographic axes increased (
Table 1). The enlargement of the
c-axis could, in principle, be useful for an easy Na
+ insertion/extraction during cell functioning, but, at the same time, it could favor the instability of the electrode structure [
33]. The increase in the
a-axis instead weakens the repulsion in the transition metal layers and, during structure changes upon cycling, particularly at high voltage, the long-range structure can change and the distortion effect will begin to have an impact on the structural reversibility, which can lead to poor capacity retention and rate performance. This scenario seems to suggest that, from the structural point of view, the dopant ions could have an unfavorable influence on the electrochemical properties. The lattice parameter variations can be attributed to the differences in the ionic radii of the stabilized species in octahedral coordination, Mn
3+/Mn
4+ (suggested by the EPR spectra analysis), and Cu
2+ and Fe
3+ ions (as determined by Mössbauer spectroscopy) [
38]. This is particularly true for the quenched samples, where no Mn vacancies should be present.
The crystallite sizes of the P2 phase were about twice those of the P’2 phase, for all the synthesized samples, independently of the cooling treatment. This is in line with the higher structural order degree of the P2 polymorph with respect to P’2, as is consistent with the line width of the relative EPR contributions. In particular, the copper ions heavily affected the crystallite size values, which are bigger than those of the pure and iron-doped samples for both the polymorphs (
Table 1), as well as the external morphology of the particles, with micron-sized dimensions and without clear platelet particles. The morphology of the samples also changed, passing from the slowly cooled to the quenched samples: the rapid cooling maintained the grains’ aggregation, which could be the basis of the worsening of electrochemical performances. It is well known that the downsizing of the grains, decreasing the diffusion paths, favors the Na
+ migration. In particular, the unfavorable morphology could justify the worst performances of the Cu-doped slowly cooled and quenched samples, (
Figure 9). The marked influence of the dopant ions was evident in terms of the structural stability as a function of time. The undoped sample had good stability for at least 7–14 days. This is due to the intrinsic characteristics of the P2 polymorph, considered one of the most stable between the P- and O-type layered phases, but this is also valid for
Cmcm, the main phase of the NMO-Q sample. The introduction of iron does not markedly change the behavior of NMO: only in the case of Fe02-Q, after 30 days, high degradation was evident, with a concomitant loss of crystallinity of the main P2 phase. The copper substitution, instead, helped to improve the stability because the formed phases remained practically unchanged for up to 30 days. This evidence is in agreement with the literature findings, suggesting copper substitution as a way to enhance the stability of layered electrode materials [
14].
We could try to relate the electrochemical performances (
Figure 8,
Figure 9, and
Figure S5) to the structural and morphological peculiarities of the samples. The CV curves were complex, with multiple events, due to the presence of complex mixtures of phases that were electrochemically active. The NMO sample, with the highest amount of P2 phase, had a well-defined CV, in which the redox peak of Mn
3+/Mn
4+ couple and the peak at about 4.39 V, due to P2-O2 transformation, can be recognized. However, the peak intensities decreased, and the positions shifted to higher potential in the anodic scan, passing from the first to the fifth cycle. These effects could be related to the marked structural changes occurring during cycling. The introduction of dopants limited the intensity decrease during cycling: both the doped samples have better reversibility. This may be due to the coexistence of a higher amount of structurally different phases. In fact, the doping and the quenching caused the increase in P’2 polymorph amount, as well as O-type phases, particularly in the quenched samples. The Fe-doped samples and, partly, the Cu samples (at least at lower potentials) had broader peaks, suggesting that the electrochemical phenomena occurred in a wide potential window, due to the increase in the amount of distorted polymorphs. These observations were in agreement with the line-broadening of EPR spectra.
In the slowly cooled samples, the NMO had the highest capacity values (about 162 mAh/g); however, these rapidly decreased by increasing the C-rates, even if the capacity at 2C was again acceptable. A similar trend was verified by Guo et al. [
37] on layered Na
0.66Li
0.18Mn
0.71Ni
0.21Co
0.08O
2+δ P2/O3 composite. The minor presence of the O3 phase, together with the main P2 phase, is responsible for the satisfactory rate capability. The O3 phase can supply more sodium ions, while the enlarged layered spacings of the P2 phase are beneficial for the easy diffusion of sodium ions upon the charging process. A good rate capability is shown by the Fe02 sample, also having some amount of O3 Na+ reservoir phase, but in this case, the enlargement of the lattice parameters of P2 phase and the stabilization of a larger amount of P’2 polymorph seem to negatively influence the electrochemical performance. Cu02 has the worst performance, although with a lower decrease in capacity during cycling at an increasing C-rate. The Cu-doped samples’ poor performance may mainly be due to the kinetic limitations related to the higher diffusion paths of sodium, because of the larger particle sizes, and due to the absence of O-type phases. So the NMO, containing a higher amount of P2 with respect to Fe02 and Cu02 and with smaller lattice parameters, could better buffer the structural changes during cycling. In some papers, it has been suggested that both the Cu
2+ and Fe
3+ in the P2 phase are active redox couples providing themselves a contribution to the overall capacity [
14,
21]. In fact, from the CV data of the Cu-doped samples (
Figure S5), the peak at about 4 V can be attributed to the Cu
2+/Cu
3+ redox couple [
14]. Despite the possible dopant contribution to the capacity, in the slowly cooled samples, the prevailing factor on the capacity values is the P2 phase prevalence, together with a small amount of O-type phases (particularly O3). The effect of dopant, as demonstrated by the EPR measurements, is different from what was expected. In fact, particularly in the case of the Cu-doped samples, the amount of Mn
4+ did not increase, as demonstrated by the decrease in EPR signal intensity with respect to NMO (
Figure 4), suggesting a possible decrease in oxygen absorption and, in turn, a lower number of manganese vacancies with respect to NMO. This evidence could also help to explain the poor electrochemical capacities. The performance of the quenched samples was lower with respect to the slowly cooled samples. The reason for this difference could be due to the different morphologies and to the different stabilizations of the phases. In fact, NMO-Q (
Figure 2) presented particles that seemed to melt together, forming large aggregates, which was different from NMO. In addition, the main phase of this sample was the
Cmcm polymorph, with a more distorted structure, limiting easy Na+ diffusion. In addition, a high percentage of O-type phases was stabilized, which may have worsened the electrochemical performance. In fact, the intervention of intermediate sites in the sodium migration needs to overcome a high-energy barrier for the O3-type structure; the inevitable complex phase transition and weak kinetics performance will directly influence the electrochemical properties. Only Fe02-Q, with a low degree of agglomeration and a low percentage of O-type polymorphs, together with the P2 phase as the main phase (similarly to NMO) with respect to NMO-Q, has a satisfying rate capability (
Figure 9). It also demonstrated high capacity values at higher C-rates with respect to NMO, and a less pronounced capacity decrease by increasing the C-rates. The Cu-doped sample also yields the worst performances, due to the same reasons previously given for Cu02. In the case of the quenched samples, however, the P2 amount was further decreased in the doped samples and, from the broad EPR signals, we could hypothesize a high value for the Mn
3+/Mn
4+ ratio, as also indicated by the EDS results (
Table 2), which can also justify the limited electrochemical results.
The deepening of the electrochemical results by performing long-cycling measurements in the search for the most competitive samples or ex situ XRPD measurements, to better clarify the mechanisms of Na storage, were outside the aim of the present work and will be the subject of future work, starting from these preliminary results and working on optimized samples.