3.1 Limitations of CuO cathode
The separation of the active materials in SO2-in-salt systems with self-activating Cu-based cathodes, which can occur during the self-activation reaction and cycling, is attributed to the significant changes in their microstructure, leading to capacity fading [11]. The conversion of the inactive CuO phase into the active CuCl2 phase via spontaneous chlorination is thermodynamically favorable. The chlorination reaction of CuO in SO2-in-salt electrolytes can be expressed as follows:
\(3CuO+2LiAlC{l}_{4}\to 3CuC{l}_{2}+2LiCl+A{l}_{2}{O}_{3}\) \(\varDelta {G}^{0}=-342.48 kJ mo{l}^{-1}\), (1)
where ΔG0 is the calculated Gibbs free energy change for the chlorination reaction.
As described in Supplementary Information, CuO spontaneously reacts with LiAlCl4 in the electrolyte to form thermodynamically stable CuCl2, LiCl, and Al2O3. Moreover, CuO with a monoclinic distorted tetragonal PtS structure transforms into CuCl2 with a CdI2-like layered structure via an anion exchange reaction (Figure S1).[26] The change in the crystal structure and volume expansion during the anion exchange reaction can cause unavoidable morphological changes and separation of the active materials.[11]
Moreover, separation of the active materials occurs during the conversion reactions associated with the charge-discharge process. Typically, when the active materials are optimized using nanostructuring strategies, it is expected that their original morphology will be maintained even if their crystal structure changes. Most conversion-type active materials (MaXb; M = transition metal, such as Mn, Fe, Co, Ni, and Cu; X = anion, such as O, S, Se, F, N, and P) store Li+ ions via two-step reactions.[27, 28] During the first lithiation step, Li+ ions are inserted into the crystal lattice of the active material to form intermediate ternary Li–M–X phases as follows:
$${{M}_{a}X}_{b}+(b\bullet n)Li\leftrightarrow L{i}_{b\bullet n}{{M}_{a}X}_{b}$$
2
,
where n is the formal oxidation state of x. During the second lithiation step, as the reaction proceeds gradually, a transition metal precipitates and lithium compounds are formed, as follows:
$$L{i}_{b\bullet n}{{M}_{a}X}_{b}\leftrightarrow aM+ {bLi}_{n}X$$
3
Owing to the formation of the Li–M–X intermediate phase, the shapes of the MaXb active material and the transition metal did not change significantly. In contrast, CuCl2 with a monoclinic structure reacts with Li in the SO2-in-salt electrolyte system to form CuCl with a zinc-blende structure and LiCl with a cubic structure, as follows:[9–11]
$$Li + Cu{Cl}_{2} \leftrightarrow CuCl+ LiCl$$
4
The ionic radii of anions are larger than those of cations; hence, they play a critical role in determining the shape of the crystal lattice. Therefore, when Cl− ions deviate from the original crystal lattice and form a new phase, the remaining atoms form a new crystal structure. Consequently, the morphology of the active materials changes continuously during the redox reaction between CuCl2 and CuCl, which occurs during cycling and is accompanied by separation of the active materials.
3.2 Effect and modification of carbon materials and CuO
As indicated in Figure S2c-f, KB, the 0-D conductive additive promotes the fast conversion of the evenly distributed nanosized CuO particles into CuCl2. However, the resulting rapid change in shape led to the separation of the active material and block the conductive path because insulating CuCl2 covered the conductive KB nanoparticles. In this study, we used CuO microspheres as the active material instead of uniformly distributed CuO nanoparticles. Field-emission scanning electron microscopy (FE-SEM) image of the as-prepared MYS-CuO microspheres is presented in Fig. 1a. The MYS-CuO microspheres exhibited a diameter of ~ 1 µm and a partially hollow internal structure comprising nanoparticle clusters (~ 50 nm in size) inside the cavities. The conversion of CuO into CuCl2, which was accelerated by KB, progressed inward from the outside of the MYS-CuO clusters. Therefore, the effect of conductive KB on the conversion reaction was negligible, even though KB and CuO nanoparticles were homogeneously mixed.
The carbon-based conductive agent surrounding the active material played a catalyst-like role, accelerating the self-activation reaction. The FE-SEM images of the carbon-free and several carbon-containing CuO cathodes after immersion in LiAlCl4·3SO2 are presented in Fig. 1b, c, g, and S2c–h. After immersion for 48 h, short CuCl2 needles grew on the surface of the spherical CuO particles of the carbon-free cathode (Fig. 1b and S2c), and the chemical composition of the needles was confirmed using X-ray diffraction (XRD) (Figure S3). Moreover, as the spherical CuO nanoparticles were consumed, wire-shaped CuCl2 particles were formed at the KB-containing cathode (Fig. 1c, S2e, and S3). Owing to its high surface area, 0-D KB provided abundant active sites. Furthermore, KB served as the catalyst and template for the recrystallization of CuCl2 nanowires with low surface energy.29 This reaction was similar to the vapor-liquid-solid growth of 1-D structures, thus facilitating the self-activation of the cathode material. However, such changes in the morphology of the active cathode material electrically isolated it from the conductive agent, leading to the deterioration of the electrochemical properties of the cathode. In addition, the large volume changes during repeated charging–discharging reduced the physical contact between the cathode components, and consequently, the cathode cracks.
To mitigate this volume change, we mixed a solution of PAN in N-methyl-2-pyrrolidone with MYS-CuO and KB to ensure robust conduction pathways. The mixture was subjected to cyclization at 280°C under ambient air conditions. The cross-sectional microstructure of the PAN-coated MYS-CuO cyclized at 280°C (P-280-MYS) was analyzed using focused ion beam microscopy and transmission electron microscopy, respectively, and the results are presented in Figs. 1d and e, respectively. The (100) plane of monoclinic CuO can be observed in the fast Fourier transform pattern of the region enclosed by the dashed red square in Fig. 1e that is illustrated in Fig. 1f. The FE-SEM images and XRD profile of the P-280-MYS cathode after immersion in the electrolyte (Fig. 1g and S2g and Figure S3, respectively) revealed the presence of mossy-like and sprout-shaped CuCl2 particles, indicating that cyclized PAN, which is a carbon material containing N-doped delocalized sp2 π-bonds, affected the formation of the CuCl2 phase via the conversion reaction at the cathode. A schematic of the structural transformation of carbon-containing CuO particles and MYS-CuO cathodes during immersion in the electrolyte is depicted in Fig. 1h.
Similar changes in morphology were observed for cathodes comprising hollow CuO nanocubes (CuO HNCs) as the active material with similar cathode compositions. (Figure S2 b, d, f, and h) However, the CuCl2 particles of the cyclized PAN-containing CuO HNC cathode were plate-shaped (Figure S2h) unlike the sprout-shaped CuCl2 particles of the P-280-MYS cathode, which was attributed to the initial shape of the CuO nanoparticles.
3.3 Characterization of oxidative cyclized PAN
The oxidative cyclization of PAN via heat treatment at 200–300°C under ambient air converted the polymer into pyrogenic carbon. During this process, the nitrile (C ≡ N) groups of the PAN chains were converted into C = N and C–N bonds, resulting in a cross-linking ring structure with pyridinic, pyrrolic, and quaternary N atoms. The extent of cyclization depended on the reaction temperature.
To analyze the relationship between the chemical structure and cyclization temperature of the PAN-containing MYS-CuO cathodes, we performed X-ray photoelectron spectroscopy (XPS) and Fourier-transform infrared (FTIR) spectroscopy experiments. The N 1s XPS profiles of the MYS-CuO cathodes with non-cyclized PAN (P-MYS) and PAN cyclized at 280°C (P-280-MYS) are illustrated in Fig. 2a–c and S4. The only peak at 400.0 eV in the N 1s XPS profile of the P-MYS cathode was attributed to the presence of C ≡ N groups. After cyclization, this peak was broadened by the evolution of three new peaks at 399.2, 400.6, and 401.6 eV, which were ascribed to pyridinic-, pyrrolic-, and quaternary-N, respectively. The intensities of these peaks increased with increasing cyclization temperature, whereas the intensity of the peak corresponding to the C ≡ N groups decreased (Fig. 2b and S4). The structures of C ≡ N and pyridinic-, pyrrolic-, and quaternary-N groups are shown in Fig. 2c. The peaks in the N 1s XPS profile were upshifted by 0.5 eV compared with the values reported in the literature.[30, 31] The shift was attributed to the strong interaction between the N species and Cu(II) sites of CuO.[32–34]
The FTIR spectra of the raw and cyclized PAN powders at temperatures between 150°C and 300°C are presented in Fig. 2d. As the oxidative cyclization temperature increased, the intensities of the peaks at 2241 cm-1 (C ≡ N group), 2935, and 2870 cm-1 (C–H stretching of the –CH2– groups of the PAN backbone), 1450 cm-1 (C–H bending vibrations of the –CH2– groups of the PAN backbone), and 1358 cm-1 (C–H bending vibrations of the –CH2– and –CH– groups of the PAN backbone) decreased. Three new bands at 1580 cm-1 (attributed to the stretching of the C = C bonds, cyclic C = N bonds, and in-plane bending of the N–H bonds), 1380 cm-1 (assigned to the bending vibration of the C–H and N–H bonds in the rings), and 810 cm-1 (ascribed to the C = CH bonds of the aromatic rings) emerged at 200°C. The intensities of these bands increased with increasing cyclization temperature, revealing that the C ≡ N groups were converted into a cyclic structure when the temperature exceeded 200°C. The extent of reaction (EOR) for the cyclization of PAN with respect to the heat treatment temperature (Fig. 2e) was calculated as follows:
$$EOR \left(\text{\%}\right)=\frac{{I}_{1580}}{({I}_{1580}+{I}_{2241})}\times 100$$
5
where \({I}_{1580}\) and \({I}_{2241}\) are the intensities of the bands at 1580 and 2241 cm-1, respectively. The EOR was approximately 100% at temperatures higher than 280°C. The differential scanning calorimetry profile of PAN revealed that it underwent pyrolysis, and exothermic peaks were observed at temperatures higher than 300°C (Figure S5).
The presence of O during the oxidative cyclization reaction led to the formation of numerous O-containing groups.(Fig. 2d) In particular, bands at 1720 cm-1 (ascribed to the stretching of the C = O bonds in ketones, aldehydes, and –COOH groups), 1662 cm-1 (attributed to the stretching of the C = O bonds in the highly conjugated acridone rings), 1161–1200 cm-1 (corresponding to the stretching of the C–O bonds), and 1250 cm-1 (attributed to the asymmetric stretching of the C–O–C bonds) were observed in the FTIR spectra of the PAN-containing MYS-CuO cathodes cyclized at temperatures higher than 200°C.
3.4 Characterization of PAN-coated MYS-CuO cathodes
The N-doping of carbon materials could increase their electron transport rate, electron shuttling efficiency, and electron carrier concentration. The resistance of the P-280-MYS cathode was as low as that of the KB-containing MYS-CuO cathode ( CuO: KB: PTFE = 6:2:2). The resistivity of the cathode containing non-cyclized PAN (P-MYS, CuO: KB: PAN: PTFE = 6:2:1:1) was high (5.31 Ω cm2; Fig. 3a). As the cyclization temperature was increased from 200°C (P-200-MYS) to 250°C (P-250-MYS), and 280°C (P-280-MYS), the cathode resistivity decreased to 5.24, 4.56, and 3.81 Ω cm2, respectively (Fig. 3a). These results demonstrated that the formation of N-doped carbon with delocalized sp2 π-bonds during PAN cyclization improved the electrical conductivity of the cathode.
To evaluate the charge transfer capability of the cathodes after resting, we performed electrochemical impedance spectroscopy measurements for symmetric cells featuring cathodes cyclized at different temperatures after resting for a week. The impedance of the P-280-MYS cathode after self-activation was the lowest, indicating that the N-doped carbon surrounding the active materials lowered the charge-transfer resistance (Rct) of the cathode (Fig. 3b and c). Moreover, the N-doped carbon particles captured Cu ions, preventing their migration outside the carbon coating, and acted as active sites for the recrystallization of CuCl2 and CuCl during charging-discharging.
3.5 Electrochemical performance
The electrochemical performances of the MYS, P-MYS, P-200-MYS, P-250-MYS, and P-280-MYS cathodes for Li metal dual-ion batteries with SO2-in-salt electrolytes are presented in Fig. 4. The discharge capacity and capacity retention of the batteries significantly depended on the cathode resistivity and Rct. The initial discharge specific capacity of the P-280-MYS cathode was the highest (315.9 mAh g-1) among all analyzed cathodes. This value was comparable to the theoretical capacity of 336.9 mAh g-1 at 0.2 C, which corresponded to an energy density of ~ 1295 Wh kgCuO-1. Moreover, the P-280-MYS cathode exhibited excellent cycling stability, with a capacity retention of ~ 84% (~ 266.8 mAh g-1) after 200 cycles. The discharge specific capacity of the MYS cathode faded faster than that of the P-280-MYS cathode, dropping from 284.2 mAh g-1 in the 1st cycle to 177.0 mAh g-1 in the 200th cycle, with a capacity retention rate of only 62.3%. The P-200-MYS and P-250-MYS cathodes exhibited good capacity retention in the 200th cycle, compared to their highest capacities of 87.9% and 83.6%, respectively (the highest specific capacity was 160.2 mAh g-1 in the 124th cycle for P-200-MYS and 235.7 mAh g-1 in the first cycle for P-250-MYS). We speculated that the O-containing groups formed by heat treatment in an oxygenated environment (found in the FT-IR analysis, as shown in Fig. 2d) may be responsible for the improved cycle retention of P-200-MYS, P-250-MYS, and P-280-MYS cathodes, as PAN functionalized with negatively charged groups was reported to be an adsorbent for metal ions that can suppress the separation of Cu ions from the cathode.35–37 Meanwhile, the O-containing groups and residual C ≡ N groups of P-200-MYS and P-250-MYS limited their electrical conductivity, which contributed to the large initial ΔV values of the cathodes, as illustrated in their voltage profiles (Fig. 4b). The ΔV values were calculated at a depth of discharge (DOD) of 50% and a state of charge (SOC) of 50%. In addition, C ≡ N groups with delocalized electron structures typically repel anions and attract cations, thereby hindering ion transport during charging-discharging. This effect was overcome as the cycles progressed (Fig. 4c).
The charge-discharge profiles of the MYS-CuO cathode during the 1st and 233rd cycles are presented in Figs. 3b and c, respectively. The cell voltage during discharging (Vdis) and charging (Vch) is expressed as follows:
$${V}_{dis}={V}_{emf}-Ir$$
6
and
$${V}_{ch}={V}_{emf}+Ir$$
7
where Vemf is the electromotive force of the battery, I is the current, and r is the internal resistance of the cell, which includes contact resistance, electrolyte resistance, activation polarization, and concentration polarization of the cathode.
The fluctuations in the ΔV value suggested that the r value changed during cycling. During heat treatment, the C ≡ N groups of PAN were converted into amide groups owing to cyclization or the so-called “stabilization” process. Cyclized PAN is an N-doped carbon material, and its electrical conductivity is higher than that of the raw PAN (P-MYS). Because all the C ≡ N groups of PAN heat-treated at 280°C were converted into amide groups, P-MYS-280 was the only cathode that retained a high discharge capacity over 200 cycles, and its ΔV value was constant, as presented in Fig. 4d.
The rate capabilities of the cathodes are shown in Fig. 4e. The charge and discharge rate capabilities were measured at several current rates between 0.1 and 1 C and in the voltage range of 3.2–3.7 V vs. Li+/Li. The charge and discharge rates were maintained constant. The capacity of P-280-MYS was the highest at all C-rates and remained above 250 mAh g-1 even at a high rate of 1 C, whereas the other cathodes exhibited relatively poor rate properties.
In the Ragone plot shown in Fig. 5, the rate performance of the MYS-CuO cathode (P-MYS-280) is compared with that of state-of-the-art cathode materials for Li/Na-SO2 batteries. The simple nanosizing improved the energy density as a result of the Li-Cu2O nanocrystals [9] and Li-CuO hollow nanocubes (HNCs) [11] shown in Fig. 5; however, as discussed earlier, the improvements in the power density are limited. The energy density of MYS-CuO, which is a nanocluster, was over 900 Wh kg-1 at any given power density. Moreover, the C-rate was maintained at 1 C during charging, and there was no significant deterioration in the discharge capacity (the energy density and power density at 1 C were 903 Wh kg-1 and 1020 W kg-1, respectively).
3.6 Morphology evolution of PAN-coated MYS-CuO during charge-discharge
Changes in the microstructure of the P-280-MYS cathode during charging and discharging were evaluated using FE-SEM to elucidate the reasons for its excellent electrochemical performance. The P-280-MYS cathodes were disassembled at DODs and SOCs of 25%, 50%, and 100% of the first cycle (Fig. 6a), and their corresponding FE-SEM images are presented in Fig. 6b–g and S6a. During the early discharge stage (DOD of 25%), the buds that sprouted from the seeds (MYS) were planted in the field (cathode) and convexly protruded on the cathode surface. Thereafter, the buds gradually swelled and opened as the DOD increased to 50% and turned into blooming flowers when the cathode was completely discharged at a DOD of 100% (Fig. 6b–d and S6a). The XRD patterns and energy-dispersive X-ray spectroscopy mappings of the cathodes revealed that the petals produced during discharging consisted of CuCl and LiCl (Figure. S7). During charging, the CuCl and LiCl petals at the center of the flower gradually disappeared (Fig. 6e and f). Finally, the active material returned to its initial shape at 100% SOC (Fig. 6g and S6a). The overall morphology evolution of the active material particles during the first charge–discharge cycle is illustrated in Fig. 6h. The MYS-CuO microspheres surrounded by cyclized PAN were converted to CuCl2 through a self-activation process. At this point, recrystallization caused a significant change in the volume of the initial yolk–shell microspheres, which split and resemble germinating seeds. Upon discharge, the Cl– ions originating from CuCl2 combined with the Li+ ions in the electrolyte to form LiCl, and the reduced monovalent Cu+ ions recrystallized with the residual Cl– ions to form CuCl. The initial CuCl2 particles presented curved shapes, with the convex part covered by the cyclized PAN and the concave part in direct contact with the electrolyte. As a result, Cl– ions easily diffused into the concave parts of the CuCl2 particles, and the byproducts generated via discharge grew inward and exhibited a bud-like morphology. When the generated discharge products grew larger than the space inside the buds, the discharge product particles bloomed outward to form a flower-like morphology. During charging, the byproducts disappeared from the center of the flower shape and returned to the sprout shape, similar to the process described above. It is possible to provide a variable space inside the concave area of the active material and ensure constant conductive pathways through the convex side.