Muxuan
Yang
,
Pratik
Kasbe
,
Jinyu
Bu
and
Weinan
Xu
*
School of Polymer Science and Polymer Engineering, The University of Akron, Akron, OH 44325, USA. E-mail: weinanxu@uarkon.edu
First published on 27th March 2024
Two-dimensional metal oxide (MO) nanostructures have unique properties compared with their bulk or 0D and 1D (nanoparticle and nanowire) counterparts. Their abundant surface area and atomically thin 2D structure are advantageous for their applications in catalysis and energy, as well as integration with 2D layered materials such as graphene and reduced graphene oxide (rGO). However, fast and scalable synthesis of 2D MOs and their nanocomposites remains challenging. Here, we developed a microwave-assisted solid-state synthesis method for the scalable generation of 2D MOs and 2D MO/rGO nanocomposites with tunable structure and composition. The structures and properties of 2D Fe2O3 and 2D ZnO as well as their nanocomposites with rGO were systematically investigated. The excellent electrochemical properties of such 2D MO/rGO nanocomposites also enable us to use them as electrode materials to fabricate microsupercapacitors. This work provides new insights into the scalable and solid-state synthesis of 2D nanocomposites and their potential applications in catalysis, energy conversion and storage.
Based on the energy storage mechanisms of supercapacitors, they can be classified into two types: electrochemical double-layer capacitors (EDLCs) and pseudocapacitive capacitors. The difference mainly comes from their behavior at the electrode/electrolyte interface. EDLCs are mainly based on carbonaceous electrode materials, and they achieve separation of charges in a Helmholtz double layer at the electrode/electrolyte interface. Pseudocapacitance is achieved by faradaic electron charge transfer with redox reactions, intercalation or electrosorption. The corresponding electrode materials (metal oxides/hydroxides or conductive polymers) usually have the ability to provide higher specific capacitance.9,10
Transition metal oxides (MO) have emerged as important electrode materials for supercapacitors due to their tunable valency, high theoretical capacitance, abundance in nature, resistance to corrosion and good thermal stability.11,12 The combination of transition metal oxides with carbon nanomaterials (especially graphene and reduced graphene oxide (rGO)) to form hybrids or nanocomposites is a promising strategy to further enhance their energy storage performance due to the synergy between EDLCs and pseudocapacitance mechanisms.13–17 Most of the previous reports involve physical mixing or blending of MO nanostructures with graphene. The MO nanostructures are usually in the form of 0D nanoparticles or 1D nanorods and nanowires. Their synthesis is also primarily based on sol–gel, hydrothermal/solvothermal, and vapor phase deposition,18–21 which are time-consuming, expensive, and not scalable.22–25
Because of the 2D atomically thin nature of graphene, it is expected to have stronger and more intimate interactions with MOs if the MOs also have 2D atomically thin structures. The abundant contact area and strong van der Waals interaction between graphene and 2D MOs will lead to synergistic property enhancement.26–29 There are several pioneering works on the synthesis and characterization of 2D MOs. For instance, 2D iron oxide (Fe2O3) was synthesized by liquid exfoliation in dimethylformamide from natural ore hematite (α-Fe2O3) and named hematene.30 Exfoliation of 2D Fe2O3 in melamine aqueous solution under mild sonication was also reported.31 Chahal et al. developed a microwave-assisted synthesis of 2D MOs in dimethylformamide or isopropanol solvent using metal chlorides as precursors.32 These 2D MOs have already been investigated for applications including catalysis,30–33 optics,34,35 electronics,36,37 and sensing.38,39 However, scalable and solid-state synthesis of 2D MO/rGO nanocomposites and their utilization in energy storage have not been demonstrated before.
To fill this knowledge gap, in this work, we developed a microwave-assisted solid-state synthesis approach for 2D MOs and 2D MO/rGO nanocomposites. Our approach is simple, fast, and scalable. Two different 2D MOs (Fe2O3 and ZnO) and their rGO nanocomposites were prepared and systematically investigated. The electrochemical properties of 2D MO/rGO nanocomposites were studied and they showed excellent performance. We also used the 2D Fe2O3/rGO nanocomposites as the main electrode materials to fabricate symmetric and asymmetric microsupercapacitors (MSCs). Our research provides a versatile and important platform for the scalable synthesis of 2D MO/rGO nanocomposites and their use in energy storage.
For the synthesis of 2D Fe2O3/rGO nanocomposites, FeCl3 powder was first mixed with graphene oxide in a calculated mass ratio, and then the mixture was used for the same microwave-assisted synthesis and washing processes as described above. The synthesis of 2D ZnO/rGO nanocomposites follows a similar procedure by using ZnCl2 as the metal precursor. The weight ratio of metal chloride and GO in the mixture can be tuned to control the final composition of the 2D nanocomposites.
The second approach for MSC fabrication is based on laser cutting and spray coating, which allows the fabrication of asymmetric MSCs. Laser cutting was used to create a polyester shadow mask for spray coating of the electrode materials with a commercial handheld airbrush. A controlled amount of the ethanol suspension of 2D Fe2O3/rGO nanocomposites was spray-coated through the shadow mask on a preheated (130 °C) polyimide substrate and a calculated amount of EG was added to the ink as a conductive additive. Then another shadow mask was used for the spray-coating of EG on the polyimide substrate as the second electrode. The ink for the positive electrode contains 13.5 mg of 2D Fe2O3/rGO nanocomposites and 3.5 mg of EG. The ink for the negative electrode contains 17.5 mg of EG. Subsequently, the KOH/PVA gel electrolyte was applied to the electrodes to complete the device fabrication.
Electrochemical measurements were performed in both 3-electrode and 2-electrode systems. For the 3-electrode test, the working electrode was prepared by mixing the active material, carbon black and polyvinylidene fluoride (PVDF) in a mass ratio of 90:5:5 to make a slurry, then it was coated onto Ni foam, followed by drying in a vacuum oven. KOH solution (1M) was used as the electrolyte, Ag/AgCl was used as the reference electrode, and a platinum wire was used as the counter electrode.
All the measurements were carried out using a CHI 660D electrochemical workstation. Cyclic voltammetry (CV) was performed at different scan rates ranging from 1 mV s−1 to 100 mV s−1. Galvanostatic charge and discharge (GCD) curves were recorded at different current densities from 1 mA cm−2 to 10 mA cm−2. Electrochemical impedance spectroscopy (EIS) was performed between a frequency range of 0.1 Hz and 1 MHz. The capacitance calculation equation used can be seen in the ESI.†
Microwave irradiation contains both the electric field and magnetic field that act normal to each other. The metal chloride precursors absorb electromagnetic energy and convert it into thermal energy, which can be further enhanced with the incorporation of GO. Such a local high thermal energy can rapidly break the metal chloride bonds when the thermal energy exceeds the bond dissociation energy.32 The oxygen molecules from air are also excited by microwave irradiation and generate radicals, which react with activated metal atoms to form metal oxides.32,41 The small amount of moisture in air can also participate in the reaction.42 The probable reaction equations are: 4FeCl3 + 3O2 → 2Fe2O3 + 6Cl2; FeCl3 + H2O → FeOCl + 2HCl; and 2FeOCl + H2O → Fe2O3 + 2HCl (for 2D Fe2O3). And 2ZnCl2 + O2 → 2ZnO + 2Cl2; ZnCl2 + H2O → Zn(OH)Cl + HCl; and Zn(OH)Cl → ZnO + HCl (for 2D ZnO).
The exact mechanism for the formation of 2D metal oxides under such a condition requires further investigation. Our hypothesis is that the electric field from microwave irradiation acts in a static plane (with sinusoidal magnitude over time) facilitates the atom rearrangement in a 2D plane. In addition, the incorporation of 2D GO nanosheets can act as an atomically thin 2D template to further promote 2D MO formation.
We estimated the yield of such a solid-state synthesis of 2D Fe2O3 and ZnO by measuring the ratio of the actual product weight and the theoretical amount. Without the incorporation of GO, the yield for 2D Fe2O3 is about 10.2%, and the yield increases to 36.0% when 10 wt% of GO is incorporated into the precursors (GO has a 7 wt% weight loss during this reduction process, which was taken into consideration during the yield calculation). The yield for 2D ZnO under the same reaction conditions is lower (7.9%) due to the lower microwave absorption capability. Therefore, in the following discussion, we will focus on the 2D Fe2O3 and 2D Fe2O3 nanocomposites. We also varied the weight fraction of GO in the metal chloride precursors for the solid-state synthesis, for instance, the 2D Fe2O3/rGO (10/1) sample has the amount of GO equals to 10 wt% of iron chloride in the precursor and the 2D Fe2O3/rGO (20/1) sample has the amount of GO equals to 5 wt% of iron chloride in the precursor.
The SEM image (Fig. 2g) of the 2D Fe2O3/rGO (10/1) nanocomposite shows that it has a nanosheet morphology with lateral sizes in the range of 100–800 nm. The TEM image (Fig. 2h) shows that domains of higher contrast which correspond to 2D Fe2O3 are on the surface of rGO nanosheets. The high-resolution TEM images (Fig. 2i–k) further show that these two domains have different crystalline structures: the one (Fig. 2k) with well-defined lattice fringes and a d spacing of 0.35 nm corresponds to 2D Fe2O3 and the other one with lower contrast and large spacing (0.44 nm) corresponds to the few-layer graphene domain. The SEM image of the 2D ZnO/rGO nanocomposites shows a similar morphology (Fig. S1†) with 2D ZnO nanosheets dispersed on the surface of rGO flakes. The EDX spectrum (Fig. S1†) further confirms the existence of 2D ZnO nanosheets in the nanocomposites.
To further investigate the structure and properties of the 2D MOs and their nanocomposites, several types of spectroscopy and scattering were conducted. Raman spectroscopy was used to characterize the structures of 2D Fe2O3 and the 2D Fe2O3/rGO nanocomposites (Fig. 3a). Characteristic peaks of the α-phase of Fe2O3 can be clearly observed, including the peak at 224 cm−1 which corresponds to the A1g mode and the peaks at 296, 410, and 613 cm−1 which correspond to the Eg mode, and the peak at 1315 cm−1 is attributed to two-magnon scattering.43,44 Moreover, a forbidden disorder-originated vibrational peak at 650 cm−1 confirms the formation of 2D crystals since there is no such disorder in conventional 3D crystals.45 In addition, the intensity ratio of A1g and Eg peaks at 224 and 296 cm−1 was calculated to be 0.77, which further confirms the 2D nature of the synthesized Fe2O3 nanosheets, because a such ratio is larger than 1.0 for bulk crystals.30 For the 2D Fe2O3/rGO nanocomposites, the major peaks for 2D Fe2O3 remain and the peaks corresponding to the G band and D band of rGO at 1570 and 1360 cm−1 can also be observed. It is noted that the peak at 1315 cm−1 has a substantially reduced intensity for the 2D Fe2O3/rGO nanocomposites, the probable reasons include the intimate contact between rGO and Fe2O3 leads to changes in the two-magnon interaction or the reduced FeOOH side product in the nanocomposites.43
We also characterized the 2D ZnO/rGO nanocomposites by Raman spectroscopy (Fig. S2†). The three characteristic graphene peaks, the D band (1360 cm−1), G band (1570 cm−1), and 2D band (2700 cm−1), are present in the Raman spectrum. In addition, two peaks at 94 cm−1 and ∼430 cm−1 are observed, which correspond to the E2 vibration mode of ZnO.46–48 The peaks from ZnO have relatively small intensity, which is due to the lower yield during the solid-state synthesis.
The crystalline structure was further investigated by XRD (Fig. 3b). 2D Fe2O3 shows all the characteristic peaks of α-Fe2O3 with a well-defined shape. For the 2D Fe2O3/rGO nanocomposite, besides all the peaks from α-Fe2O3, there is also an additional peak with a 2θ value of 26°, which corresponds to the (002) crystal plane of multilayered rGO.49,50 This peak intensity is relatively low because rGO is the minor component in the 2D Fe2O3/rGO (10/1) nanocomposite.
The UV-vis spectrum of 2D Fe2O3 shows absorption peaks at 246, 365, and 551 nm (Fig. 3c). The optical adsorption spectrum can be used to estimate the band gap energy of 2D Fe2O3 by using the Tauc plot (Fig. 3c inset). The calculated band gap for our 2D Fe2O3 is 2.86 eV, which is consistent with literature reports. Such a band gap is larger than that of bulk Fe2O3 crystals due to the quantum effect when the size reduces to the nanoscale.30,32,51 For the 2D Fe2O3/rGO nanocomposites (Fig. 3d), the peaks for 2D Fe2O3 at 246 and 365 nm can clearly be observed. There is also a strong peak at 221 nm, which corresponds to the absorption peak of rGO in the nanocomposites. The band gap calculation from the Tauc plot shows almost the same value (2.85 eV) as that of 2D Fe2O3.
XPS was conducted to confirm the structures and compositions of 2D Fe2O3 and the 2D Fe2O3/rGO nanocomposites. The survey scan (Fig. S3†) of the 2D nanocomposites shows characteristic peaks of oxygen, iron, and carbon. The high-resolution scan of the Fe 2p peak (Fig. 3e) shows two distinct peaks located at 709.3 eV and 722.6 eV, corresponding to the two spin states of iron: Fe 2p3/2 and Fe 2p1/2, respectively.32,52,53 The energy separation between the two peaks is 13.3 eV, which is consistent with previous reports on Fe2O3.54 In addition, a satellite peak at 717.1 eV appears, which is characteristic of Fe3+ ions in Fe2O3,55 this further confirms that the 2D iron oxide is primarily Fe2O3 rather than other forms such as Fe3O4.56 The high-resolution scan of the O 1s peak and its deconvolution (Fig. 3f) shows three sub-peaks, the major one at 530.3 eV is attributed to the lattice oxygen involved in the binding of α-Fe2O3 and two minor peaks at 531.6 eV and 533.3 eV primarily correspond to the residue surface oxygen groups including CO and C–O from rGO.57,58
CV scans of the 2D Fe2O3/rGO (10/1) nanocomposite at scan rates from 5 mV s−1 to 50 mV s−1 are shown in Fig. 4a. The curves display a pair of typical redox peaks corresponding to the valence state change of iron between Fe3+ and Fe2+, which indicates pseudocapacitive behavior. With the increase of scan rate, the CV curves maintain the overall shape, the anodic peaks shift toward larger potential values, and the cathodic peaks shift toward lower potential values. The shift can be explained by the Randles–Sevcik equation59 and the increased ionic diffusion resistance at a high scan rate.60–64
Galvanostatic charge–discharge (GCD) measurements of the 2D Fe2O3/rGO (10/1) nanocomposite at different current densities were also conducted (Fig. 4b). Charging to 0.35 V is completed in a few seconds, followed by a slower charging to 0.47 V. In the discharge curves, there is a plateau at around 0.25 V in low current density measurements, which is characteristic of pseudocapacitive behavior and matches with the reduction peak in CV measurements. The most pronounced plateaus for both 2D MOs were observed in the low current density curves, this is due to the sufficient time that ensure the electrolyte ions to interact with the electrode at a low charging/discharging rate. When the current density increased, the plateaus were obviously shortened since the insufficient time for the ions to reach the entire electrode surface area and the redox reaction become more restricted to the more easy accessible area, which limits the charge storage capability.65,66
Moreover, we also studied the electrochemical performance of pristine 2D Fe2O3 (without incorporation of graphene oxide during the synthesis) and 2D Fe2O3/rGO nanocomposites with different ratios between the two components. The CV scans of 2D Fe2O3 and 2D Fe2O3/rGO (20/1) are shown in Fig. S4.† Both samples show similar shapes and peak positions in the CV curves compared with 2D Fe2O3/G (10/1), but the current density and area within the CV curves are smaller. The calculated specific capacitance values for the three samples were compared and are presented in Fig. 4c. It can be seen that pristine 2D Fe2O3 has the lowest capacitance of 45.7 F g−1 (at a scan rate of 10 mV s−1), primarily due to the low electrical conductivity. The two 2D Fe2O3/rGO composites have substantially improved capacitance, especially for 2D Fe2O3/rGO (10/1), with a capacitance of 258.9 F g−1 at 10 mV s−1 and 331.4 F g−1 at 1 mV s−1, respectively.
The electrochemical capacitance of pristine 2D Fe2O3 is limited by its intrinsic low conductivity that limits charge transfer.13 After the incorporation of rGO, there are three factors that can contribute to the electrochemical performance of the 2D Fe2O3/rGO nanocomposites. First, the intercalated hybrid structures with smaller 2D Fe2O3 nanosheets on the surface or between rGO flakes increase the electrochemically active sites. The enhanced intercalation and surface accumulation of ions increase the electrochemical kinetics.14,67 Second, the high conductivity of rGO promotes charge transfer during the reversible charge storage–release process.65,66 Third, rGO also exhibits a certain extent of pseudocapacitive behavior due to the oxygen-containing groups that can contribute to the overall pseudocapacitive capacitance.57,68,69
Our solid-state synthesis approach is versatile and can be used to synthesize other types of 2D MO/rGO composites including 2D ZnO/rGO. The electrochemical properties of the 2D ZnO/rGO nanocomposites were also investigated. The CV curves of 2D ZnO/rGO (20:1) at different scan rates are shown in Fig. 4d. A pair of redox peaks located at around 0.25 V and 0.42 V can be observed, which primarily correspond to the intercalation and deintercalation of K+ from the electrolyte into ZnO (ZnO + K+ + e− ↔ ZnOK).70 GCD tests for 2D ZnO/rGO (20:1) at different current densities were also conducted (Fig. 4e). In the discharge curves, there is a plateau at around 0.30 V, which is characteristic of pseudocapacitive behavior and matches with the reduction peak in CV measurements.
We also varied the ratio of ZnO to rGO in the 2D ZnO/rGO nanocomposites. The specific capacitance values for two samples, 2D ZnO/rGO (20:1) and 2D ZnO/rGO (10:1), are summarized in Fig. 4f (see also Fig. S5†), which shows that the two samples have a capacitance of 247.3 F g−1 and 199.0 F g−1, respectively, at a scan rate of 10 mV s−1. Such performance is comparable to the 2D Fe2O3/rGO nanocomposite. Due to the higher yield of 2D the Fe2O3/rGO nanocomposites in the solid-state synthesis, we will focus the following discussion of supercapacitor devices to those with the 2D Fe2O3/rGO electrodes.
We compared the electrochemical performance of our 2D Fe2O3/rGO nanocomposites with literature reports on similar material systems composed of iron oxide and carbon nanostructures, including Fe2O3@N-doped porous carbon,71 Fe3O4 nanoparticles on rGO,72 α-Fe2O3 nanotube arrays,73 RGO-Fe3O4,69 Fe3O4/MWCNTs,74 hydrothermal Fe3O4 nanoparticles,75 and Fe2O3/3D graphene aerogels.76 The results are summarized in Fig. S6 and Table S1.† Our 2D Fe2O3/rGO nanocomposite has superior specific capacitance (230.1 F g−1 capacitance at 1 A g−1 from GCD data) compared with others. The excellent electrochemical performance in combination with scalable solid-state synthesis makes the 2D Fe2O3/rGO nanocomposites promising candidates as electrode materials for energy storage devices.
The fabricated MSC device is shown in Fig. 5a, the length of each interdigitated electrode is 4 mm, the width is 0.4 mm, and the gap between neighboring fingers is 0.2 mm. The overall device size is 5.4 × 7.7 mm. The optical microscope image shows the lithography-patterned electrode with a well-defined size and shape. The SEM image of the electrode surface shows a high density of 2D Fe2O3/rGO nanoflakes.
CV measurements of the MSC device with an electrode composed of a 2D Fe2O3/G and EG mixture (1:1 weight ratio) at different scan rates are shown in Fig. 5b. The data show characteristics of both pseudocapacitive and EDLC features originated from the 2D Fe2O3/rGO nanocomposites. The redox peaks become less pronounced at scan rates of 50 mV s−1 and above. This could be attributed to the limitation of the ion transport rate to the electroactive surface and the more pronounced double-layer charging at high scan rates. The GCD curves of the MSC device are shown in Fig. 5c. At a low current density (such as 0.6 mA cm−2), there is a plateau in the discharge process (potential range of 0.2–0.4 V), which can be attributed to redox process at this range and the rate of which is slower than double-layer discharging. At high current densities, the GCD curves have a symmetric triangle shape without any plateau.
We also studied the effect of incorporating EG into the 2D Fe2O3/rGO nanocomposites on the MSC device performance. The electrochemical performance of two types of devices, one using only 2D Fe2O3/rGO as the electrodes and the other using 2D Fe2O3/rGO mixed with EG (weight ratio 1:1) as the electrodes, was compared by plotting their specific capacitance at different scan rates (Fig. 5d). It can be seen that the incorporation of EG effectively increased the capacitance. For instance, the device with the mixture in the electrodes has a capacitance of 35.5 mF cm−2 at 5 mV s−1, while the MSC with only 2D Fe2O3/rGO in electrodes shows a capacitance of 10.9 mF cm−2 at the same scan rate. The Nyquist plots of the two devices from EIS measurements and their fitting are shown in the inset (more details are given in Fig. S8 and Table S2†). The MSC with mixture electrodes shows a larger intercept on the real axis and a substantially higher slope in the low-frequency region.57,77 This result indicates that the enhanced performance of MSC with mixture electrodes is primarily due to the enhanced ion adsorption and diffusion rate, and the increased contribution from the electric double-layer capacitance.
For the symmetric MSC devices discussed above, despite their high resolution fabrication and small form factor, the main limitation is the relatively narrow operation window (0.8 V) due to the symmetric electrodes. In order to expand the operation window, we used another approach, which is based on laser cutting and spray coating, to fabricate asymmetric MSCs. Two different suspensions can be used to spay-coat and fabricate the two different electrodes on each side.
For the asymmetric MSC device shown in Fig. 6a, the interdigital electrodes on the two sides are 2D Fe2O3/rGO (10/1) and EG, respectively. The SEM image shows the surface of the electrodes has a high density of loosely connected 2D nanosheets. The CV scans of the two individual electrodes in the half-cell configuration are shown in Fig. 6b, from which the specific capacitances (at a scan rate of 50 mV s−1) of 2D Fe2O3/rGO (10/1) and EG were calculated to be 150.8 F g−1 and 22.1 F g−1, respectively. The mass of each electrode to be deposited was calculated based on their capacitance to reach charge balance during operation.
CV scans of the asymmetric MSC at different scan rates are shown in Fig. 6c. The operation voltage window substantially increased to 1.5 V. The CV curves show both pseudocapacitive and EDLC features. The GCD data (Fig. 6d) show relatively fast charging and discharge. At a low discharge rate (0.02 mA cm−2), there is a plateau at around 0.25 V in the curve, which corresponds to the redox peak in the CV scan. The area-specific capacitance of the asymmetric MSC is lower than that of the symmetric MSC (for instance, 2.5 mF cm−2vs. 35.4 mF cm−2 at 10 mV s−1 scan rate). The main reason is the loose structure and smaller thickness of the electrodes from the spray coating method, as shown by the SEM image in Fig. 6a, which can lead to less continuity and lower conductivity compared with electrodes prepared by the vacuum filtration method.
To further confirm that such differences in the MSC performance is mainly due to the fabrication method, we fabricated symmetric MSC devices also by the spray coating method using the 2D Fe2O3/rGO (10/1) nanocomposites. The CV scan curves and GCD curves of the MSC are shown in Fig. S9.† Such a device has a much lower capacitance (0.59 mF cm−2 at 10 mV s−1) compared with the symmetric MSC fabricated by vacuum filtration (Fig. 5b and c), but close to that of the asymmetric MSC (2.50 mF cm−2) also fabricated by the spray coating method. Such a comparative study also confirms that the asymmetric MSC has higher capacitance due to the enlarged electrochemical operation window.
Systematic characterization of these 2D MOs and 2D MO/rGO nanocomposites was conducted using spectroscopies, electron microscopies and diffraction methods. The electrochemical properties of the 2D Fe2O3/rGO nanocomposites were investigated, which show an excellent specific capacitance of 331.4 F g−1 at a 1 mV s−1 scan rate. Such 2D Fe2O3/rGO nanocomposites were further used as the main electrode materials to fabricate MSCs. Both symmetric and asymmetric MSCs were fabricated and tested, and their energy storage capability was demonstrated. This work is of high importance to the fields of solid-state chemistry, electrochemistry, and energy storage and has the potential to be used in next-generation supercapacitors and batteries as high-performance electrodes. Our approach is also highly flexible and can be used for the synthesis of mixed MO nanostructures and their graphene nanocomposites. Such multifunctional nanocomposites with multiple types of MOs can have highly tunable band gaps and electrochemical activities for broad applications.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nr00587b |
This journal is © The Royal Society of Chemistry 2024 |