Next Article in Journal
Interaction of Bis-(sodium-sulfopropyl)-Disulfide and Polyethylene Glycol on the Copper Electrodeposited Layer by Time-of-Flight Secondary-Ion Mass Spectrometry
Next Article in Special Issue
Synthesis of Triblock Polycarboxylate Superplasticizers with Well-Defined Structure and Its Dispersing Performance in β-Hemihydrate Gypsum
Previous Article in Journal
Recent Advances in g-C3N4-Based Materials and Their Application in Energy and Environmental Sustainability
Previous Article in Special Issue
In Situ Synthesis of C-N@NiFe2O4@MXene/Ni Nanocomposites for Efficient Electromagnetic Wave Absorption at an Ultralow Thickness Level
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polypyrrole-Coated Low-Crystallinity Iron Oxide Grown on Carbon Cloth Enabling Enhanced Electrochemical Supercapacitor Performance

Low-Carbon Technology & Chemical Reaction Engineering Lab, College of Chemical Engineering, Sichuan University, Chengdu 610065, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(1), 434; https://doi.org/10.3390/molecules28010434
Submission received: 7 December 2022 / Revised: 22 December 2022 / Accepted: 30 December 2022 / Published: 3 January 2023
(This article belongs to the Special Issue Feature Papers in Applied Chemistry)

Abstract

:
It is highly attractive to design pseudocapacitive metal oxides as anodes for supercapacitors (SCs). However, as they have poor conductivity and lack active sites, they generally exhibit an unsatisfied capacitance under high current density. Herein, polypyrrole-coated low-crystallinity Fe2O3 supported on carbon cloth (D-Fe2O3@PPy/CC) was prepared by chemical reduction and electrodeposition methods. The low-crystallinity Fe2O3 nanorod achieved using a NaBH4 treatment offered more active sites and enhanced the Faradaic reaction in surface or near-surface regions. The construction of a PPy layer gave more charge storage at the Fe2O3/PPy interface, favoring the limitation of the volume effect derived from Na+ transfer in the bulk phase. Consequently, D-Fe2O3@PPy/CC displayed enhanced capacitance and stability. In 1 M Na2SO4, it showed a specific capacitance of 615 mF cm−2 (640 F g−1) at 1 mA cm−2 and still retained 79.3% of its initial capacitance at 10 mA cm−2 after 5000 cycles. The design of low-crystallinity metal oxides and polymer nanocomposites is expected to be widely applicable for the development of state-of-the-art electrodes, thus opening new avenues for energy storage.

1. Introduction

The energy crisis and the environmental pollution problem caused by the excessive use of fossil energy have resulted in higher demand for developing sustainable energy around the world [1,2]. It is highly attractive to develop electrochemical energy storage devices by utilizing renewable energy sources (e.g., tidal, wind, and solar energy) [3,4]. Among the diverse energy storage devices, supercapacitors (SCs) have attracted considerable attention due to their high power density (1–10 kW kg−1), fast charge/discharge rate, safety, and long lifespan (>100,000 cycles) [5,6]. Electric double-layer capacitor (EDLC) materials are widely employed in SCs because of their high electrical conductivity, large specific surface area, and low cost [7]. However, the energy storage mechanism of EDLCs occurs at the surface of the electrode, which is responsible for the EDLCs’ poor capacity compared with batteries [8,9]. In addition, according to the equation E = 1/2CV2, constructing asymmetric supercapacitor devices (ASCs) can widen the operating voltage window and enhance the energy density [10]. However, the capacitance of ASCs is mainly limited by the inferior capacitance of negative materials.
Metal oxides, such as the pseudocapacitor materials (e.g., Co3O4, MnO2, and Bi2O3), can offer a higher specific capacitances than EDLC materials because of the charge storage mechanism of surface/near-surface redox reactions [11,12,13,14,15]. Among these metal oxides, Fe2O3 is one of the most interesting negative materials due to its low cost, environmental benignity, wide operating voltage window (0 to −0.1 V vs. Ag/AgCl), and so on [16,17]. However, it has usually exhibited an insufficient ionic diffusion rate, inferior electronic conductivity, and a volume effect, leading to poor specific capacitance and stability [18,19]. Generally, designing nanomaterials such as nanorods [20], nanosheets [21], and quantum dots [22] can give a sufficient surface area to make good use of active materials. In addition, low-crystallinity nanomaterials are capable of achieving better electrochemical performances than their high-crystallinity counterparts due to their increased structural disorder and defects [23,24]. Xia et al. [19] fabricated a crystalline/amorphous Fe2O3 nanocomposite to boost the utilization of active materials and enhance the Faradaic reaction, and it delivered a capacitance of 350 mF cm−2 at 1 mA cm−2. A defective Fe2O3 nanorod prepared using annealing under an inert atmosphere was reported by Wang et al. [25], and it displayed a capacitance of 381 mF cm−2 at 0.5 mA cm−2. However, the enhancement of capacitance derived from the boosted valence state change of metal oxide usually causes irreversible conversion in the bulk phase. Therefore, apart from the design of the nanostructure, it is imperative to develop strategies that can transform “bulk” redox reactions into “surface/near-surface” levels of redox reaction, which can simultaneously achieve the capacitance and electrochemical stability.
The construction of nanocomposites by introducing advanced carbon materials or conductive polymers not only favors the electron transfer of electrodes but also hinders the volume effect to improve the cycling stability [26,27]. Polypyrrole (PPy), a heterocyclic conjugate conductive polymer, has attracted considerable attention in energy storage devices [28,29]. Yang et al. [30] prepared a flexible PPy@Fe2O3@stainless steel yarn composite electrode and enhanced the capacitance compared with the Fe2O3@stainless steel yarn. Obaidat et al. [31] reposted a PPy-assisted Ag-doping strategy to improve the electrochemical performance of Co(OH)2, and the cycling stability was obviously improved after introducing PPy. Compared with the introduction of PPy by chemical oxidation, the electrodeposition strategy is considered to be a promising route for boosting energy storage. PPy prepared using electrochemical deposition methods was reported, and it possessed a molecular ordering formation assisted by an electric field [32]. By depositing a polymer on the surface of electrode to form the core–shell nanostructure, the formative interface between the polymer and the metal oxide can serve as the reservoir to achieve energy storage [33]. Moreover, this interface can efficiently limit ion transfer in the bulk phase, thus hindering the volume effect and further contributing to the electrochemical stability [33]. A low-crystallinity metal oxide with high disorder can usually possess a high charge storage due to the higher active area for enhancing EDLCs and Faradaic pseudocapacitance. However, they usually deliver low cycling stability. A conductive polymer coating layer can serve as the buffer layer to limit the volume expansion of active material and form an active interface to hinder the electrolyte ion transfer in the bulk phase. This could be an efficient strategy to boost the cycling stability. Therefore, we anticipate the construction of low-crystallinity Fe2O3 coated with a PPy layer, which can simultaneously boost capacitance and electrochemical stability.
In this work, we first constructed low-crystallinity Fe2O3 nanorod arrays on carbon cloth (D-Fe2O3/CC) using a simple chemical reduction reaction to enlarge the interface of the electrolyte and electrode. Then, PPy was deposited on the surface of low-crystallinity Fe2O3 using an electrochemical deposition method to construct a metal oxide/polymer core–shell nanocomposite (D-Fe2O3@PPy/CC). The NaBH4 treatment provided a rougher surface for D-Fe2O3/CC, favoring the Faradaic reaction derived from the ion diffusion in the bulk phase. The formation of a PPy layer on the surface of D-Fe2O3 enhanced the charge storage on the surface or near the surface of the electrode. Compared with the pristine Fe2O3/CC electrode with a capacitance of 114 mF cm−2 (123 F g−1) at 1 mA cm−2, this D-Fe2O3@PPy/CC electrode delivered an enhanced capacitance of 615 mF cm−2 (640 F g−1) at 1 mA cm−2. Notably, the D-Fe2O3@PPy/CC electrode also exhibited superior cycling stability, with an initial capacity retention of 79% at a current density of 10 mA cm−2 after 5000 cycles, which is higher than the initial capacity retention values of Fe2O3/CC (42%) and D-Fe2O3/CC (24%) electrodes.

2. Results and Discussion

The scheme of the fabrication of D-Fe2O3@PPy/CC is displayed in Figure 1a. First, Fe2O3 nanorod arrays were prepared on carbon cloth using a simple hydrothermal reaction and then immersed in 1 M NaBH4 to construct low-crystallinity Fe2O3 nanoarrays. To investigate the relationship between the reducing process and the electrochemical performance, D-Fe2O3/CC samples were prepared using various immersion times. After being treated with NaBH4, the D-Fe2O3/CC was coated with PPy to further improve the electrochemical stability. The mount of PPy loading could easily be regulated by changing the electrodeposition time.
Figure 1b displays the Raman spectra of D-Fe2O3/CC electrodes obtained by changing the reduction time. Clearly, three distinct peaks at 408, 291, and 223 cm−2 are observed for the pristine Fe2O3, D-Fe2O3-0.5 h, and D-Fe2O3-1 h, which imply the successful preparation of Fe2O3 [33,34]. No obvious characteristic peak can be observed in the Raman spectrum of D-Fe2O3-4 h, which indicates the breakdown of crystallinity on the Fe2O3 surface by the NaBH4 treatment. The Raman spectra of D-Fe2O3@PPy/CC obtained by varying the electrodeposition time (Figure 1c) also show the characteristic peaks of Fe2O3, but decreased intensity is observed for D-Fe2O3@PPy-120 s and D-Fe2O3@PPy-180 s, which is due to the PPy coating on the surface of Fe2O3. Notably, the obvious characteristic peaks at 1040, 954, and 916 cm−1 are observed, indicating the formation of PPy on the surface of Fe2O3 [33]. Similar results can be also observed in the XRD patterns of the D-Fe2O3/CC electrodes. As shown in Figure 1d, the characteristic peaks at 33.0°, 35.7°, 41.06°, 54.4°, and 64.3° are attributed to the (104), (110), (113), (116), and (300) planes of Fe2O3 (JCPDS#84–0308). The intensity of these characteristic peaks decreased by increasing the reduction time, which indicates the low crystallinity of Fe2O3 after the NaBH4 treatment. The NaBH4 treatment led to a reduction in crystal size, thus favoring Na+ diffusion near the surface. After being coated with PPy, no obvious change was observed (Figure 1e). This result indicates that the electrodeposition of PPy cannot influence the crystal structure of Fe2O3.
The surface chemistry of Fe2O3/CC, D-Fe2O3/CC, and D-Fe2O3@PPy/CC were investigated by XPS measurement, as shown in Figure 2. The high-resolution Fe 2p spectrum of Fe2O3/CC (Figure 2a) exhibits two peaks with binding energies (BEs) of 711.5 and 725.1 eV, corresponding to Fe 2p3/2 and Fe 2p1/2, respectively. The Fe 2p3/2 and Fe 2p1/2 peaks with 13.6 eV peak separation are attributed to the typical doublet peaks of Fe3+, implying the existence of an Fe3+ species [10]. This can be further confirmed by the satellite peak of Fe 2p3/2 with BEs at 719.1 eV [17,35]. After the NaBH4 treatment, the Fe 2p core-level XPS spectrum of D-Fe2O3/CC (Figure 2c) exhibited a negative shift compared with that of the pristine Fe2O3/CC, implying that the NaBH4 treatment influences the local chemical states of Fe3+. This result indicates the formation of oxygen vacancy on Fe2O3 [25,36]. The low-crystalline and high-disorder Fe2O3 can provide more ion diffusion pathways, which further enhance the Faradaic reaction in the bulk phase. The high-resolution O 1s spectrum of Fe2O3/CC (Figure 2b) shows three peaks. The peaks with BEs at around 530.4, 531.9, and 533.2 eV are assigned to lattice oxygen, chemisorbed oxygen, and physically adsorbed oxygen, respectively [10,37]. Among them, the chemisorbed oxygen, such as O, O2−, and O22− on the surfaces of materials, is attributed to the adsorption oxygen on the unpaired electrons generated by the surface defects. The relative contents of these O species can be determined from the proportional areas of the corresponding peaks. After the NaBH4 treatment, the O 1s spectrum of D-Fe2O3/CC (Figure 2d) showed four peaks with BEs at 529.6, 531.5, 533.1, and 535.5 eV, corresponding to the lattice oxygen, chemisorbed oxygen, physically adsorbed oxygen, and water [38,39]. As expected, the content of chemisorbed oxygen presented by D-Fe2O3/CC was calculated to be 36.8%, which was higher than that of pristine Fe2O3/CC (27.1%). This further confirmed the higher defect presentation after the chemical reduction procedure. After PPy coating, the wide-scan XPS spectrum of D-Fe2O3@PPy/CC (Supplementary Materials: Figure S1) exhibited the existence of N, Fe, O, C elements on the surface. The Fe 2p3/2 peak at 711.3 eV and the Fe 2p1/2 peak at 724.9 eV of D-Fe2O3@PPy/CC (Figure 2e) showed positive shifts compared with that of D-Fe2O3/CC. In addition, the high-resolution O 1s spectrum (Figure 2f) displayed two main peaks that can be deconvoluted into four peaks for 530.3, 531.7, 532.8, and 535.7 eV. The content of chemisorbed oxygen was determined to be 28%, which was due to the cover of PPy on the surface of Fe2O3. Figure S2 displays the high-resolution N 1s spectrum, and three deconvoluted peaks are observed at 397.7, 399.8, and 402.1 eV, which are ascribed to the Pyridinic-N, Pyrrolic-N, and nitrogen oxide, respectively [40,41,42]. This further indicates the successful formation of PPy on the surface of Fe2O3.
The morphology of all electrodes was investigated by SEM measurement. Clearly, the SEM image of Fe2O3/CC (Figure 3a) displays a nanorod structure, which uniformly covers the CC fiber (Figure 3a inset). Notably, by increasing the NaBH4 treatment time the morphology of the nanorod changed. When the reduction time was 0.5 h, some nanosheet structure between the nanorod was observed, as shown in Figure 3b. When the immersion time was more than 1 h, the SEM images of D-Fe2O3-1 h and D-Fe2O3-4 h (Figure 3c,d) mainly displayed nanosheet morphology. This could be ascribed to structure conversion by the NaBH4 treatment, which provides a strong reduction reaction on the surface of Fe2O3. A part of an oxygen atom is removed in the reduction reaction, giving a rougher surface for ion transfer.
Figure 4 shows the SEM images of D-Fe2O3@PPy/CC electrodes prepared using different electrodeposition times. No obvious change could be seen when conducting PPy electrodeposition for 30 s (Figure 4a). As the electrodeposition time increased to 60 s, some nonuniform particles were embedded on the surface of electrode and the nanostructure covered the CC fiber well. When conducting PPy electrodeposition for 120 s, the nanostructure was uniformly coated with PPy (Figure 4c). However, when the electrodeposition increased to 180 s, the nanorod or nanosheet structure could not be found (Figure 4d) and the PPy aggregated on the surface of the electrode (Figure 4d inset). Excessive PPy could obstruct the ion diffusion pathway, thus hindering ion transfer and limiting the utilization of the active material. The TEM image of D-Fe2O3@PPy (Figure S3) shows the rough surface, which is ascribed to the result of the SEM image. In addition, the selected area electron diffraction (SAED) pattern (Figure S3 inset) shows clear diffraction rings, indicating the disordered and polycrystalline features of Fe2O3.
To investigate the electrochemical performances of all electrodes, a three-electrode system was used in a 1 M Na2SO4 aqueous electrolyte. Figure S4a–d display the CV curves of D-Fe2O3/CC electrodes at different scan rates. Clearly, by increasing the NaBH4 treatment time, the area of the CV curve increased gradually. The corresponding GCD curves of D-Fe2O3/CC electrodes are illustrated in Figure S5a–d. The pristine Fe2O3, D-Fe2O3-0.5 h, D-Fe2O3-1 h, and D-Fe2O3-4 h electrodes delivered capacitances of 113, 701, 960, and 1220 mF cm−2 at a current density of 1 mA cm−2, respectively. Notably, the capacitance of Fe2O3/CC can be improved considerably by this chemical reduction treatment. This is ascribed to the formation of a rougher surface, which boosts the ion transfer and the utilization of the active materials. Notably, a weak charge/discharge platform at −0.6 V can be found in the GCD curves for the D-Fe2O3-0.5 h, D-Fe2O3-1 h, and D-Fe2O3-4 h electrodes. This implies enhanced ion transfer and Faradaic reaction in the bulk phase, which could provide additional battery-like charge storage. Figure 5a and Figure S6a–d show the CV curves of D-Fe2O3@PPy/CC electrodes. All CV curves show relatively quasi-rectangular shapes, implying that they exhibited more surface Faradaic reactions. The corresponding GCD curves (Figure 5b and Figure S7) present relatively asymmetric triangle shapes, further indicating the capacitive behavior of D-Fe2O3@PPy/CC electrodes. However, as the PPy electrodeposition time increased, the capacitance of the D-Fe2O3@PPy/CC electrode presented a decreased capacity. This was due to the limitation of Na+ transfer in the bulk phase of Fe2O3 due to the construction of the Fe2O3/PPy interface, which could favor electrochemical stability during the charge/discharge process. The rate performance of D-Fe2O3/CC is shown in Figure 5c. Clearly, the pristine Fe2O3/CC only delivered capacitances of 114, 103, 89, 79, and 69 mF cm−2 (123, 111, 96, 85, and 74 F g−1) at current densities of 1, 2, 5, 10, and 20 mA cm−2, respectively. After PPy coating, the D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) electrode (Figure 5d) delivered capacitances of 615, 525, 389, 298, and 208 mF cm−2 (640, 547, 405, 310, and 217 F g−1) at 1, 2, 5, 10, and 20 mA cm−2, respectively. This indicates the enhanced electrochemical performance of the Fe2O3 electrode using chemical reduction and PPy electrodeposition methods. In fact, the NaBH4 treatment can provide more surface roughness for Fe2O3 and give ion-insertion-like charge storage in the bulk phase, which can enhance the specific capacitance but deteriorate the volume effect. The construction of an Fe2O3/PPy interface could confine the Na+ and limit diffusion in the bulk phase, thus providing more surface Faradaic reactions for charge storage. Table S1 lists the detailed comparisons of Fe2O3-based and D-Fe2O3@PPy/CC anodes for supercapacitors [18,19,33,34,43,44,45,46,47,48,49]. The results further indicate the outstanding electrochemical performance of this D-Fe2O3@PPy/CC electrode.
To further investigate the charge storage mechanism, the surface capacitive behavior and diffusion-controlled contributions were divided using Dunn’s method to analyze the CV data according to the equation i(V) = k1v1/2 + k2v, where k1v1/2 and k2v represent the contributions from diffusion-controlled (a slow Faradaic reaction) and surface capacitive contributions (including EDLCs and fast surface Faradaic reactions). Usually, the enlarged specific area provided by the nanoarray structure and surface roughness can offer more EDLC capacitance (ion adsorption/desorption). The active sites of the electrode surface can give more Faradaic reactions from the increased chemisorbed redox sites on the surface of electrode. These two contributions compose the surface capacitive behavior. In addition, ion insertion/deinsertion in the near-surface region offers another redox reaction from the diffusion-controlled contribution. Figure 5e compares the diffusion-controlled and capacitive contribution for the pristine Fe2O3/CC, D-Fe2O3/CC (D-Fe2O3-1 h), and D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) electrodes. Notably, the diffusion-controlled capacity gradually decreased with the increased scan rate, implying that the entry of Na+ into the Fe2O3 bulk phase was limited at a high scan rate because of the electrochemical kinetic principle [50]. Owing to the “etching” process derived from the NaBH4 treatment, Fe2O3 could react with NaBH4 and generate structure defects on the surface of Fe2O3, offering more active sites and ion diffusion pathways. The diffusion-controlled and capacitive behavior contributions were improved considerably for the D-Fe2O3/CC electrode. Notably, capacitive contributions of 64%, 69%, 80%, and 87% were achieved for Fe2O3/CC electrode at scan rates of 10, 20, 60, and 100 mV s−1, respectively. After the NaBH4 treatment, the D-Fe2O3/CC electrode delivered decreased capacitive contributions (41%, 46%, 63%, and 76%) at different scan rates. This result confirms the enhanced ion-insertion-like charge storage. After PPy coating, the capacitive contribution further decreased, which was due to the PPy cover on the rough surface of Fe2O3, thus leading to the limitation of EDLCs. Notably, the electrochemical stability was promoted considerably. As shown in Figure 5f, the D-Fe2O3@PPy/CC electrode displayed enhanced cycling stability with a capacitance retention of 79.3% at a current density of 10 mA cm−2 after 5000 cycles. However, the D-Fe2O3/CC electrode with a high specific capacitance delivered poor cycling stability, and it only retained 23.9% of its initial capacitance. The pristine Fe2O3/CC also exhibited an unsatisfied capacitance retention of 42.5%. To further confirm the enhanced stability, the SEM images of Fe2O3/CC and D-Fe2O3@PPy/CC after the cycling test are displayed in Figure S8. Clearly, the nanostructure could not be maintained for the Fe2O3/CC electrode, and the collapse of the nanostructure was observed (Figure S7a). Notably, the SEM images of the D-Fe2O3@PPy/CC electrode show an obvious nanosheet structure (Figure S7b), and it covered the CC fiber well (Figure S8b inset).

3. Materials and Methods

3.1. Materials

Polypyrrole (C4H5N, >99%) was purchased from Aladdin Chemical Co., Ltd. (Shanghai, China). Nitric acid (HNO3, >68%), ethanol (C2H5OH, >99%), ferric chloride (FeCl3·6H2O, >99%), sodium sulfate (Na2SO4, >99%), sodium borohydride (NaHB4, >99%), and other chemical reagents were purchased from Chengdu Kelong Chemical Reagent Factory, China. The carbon cloth was purchased from Hongshan District, Wuhan Instrument Surgical Instruments (Wuhan, China). All chemical reagents were used as received without further purification.

3.2. Synthesis of Fe2O3/CC

The carbon cloth (CC) was cut into a number of 2 × 3 cm2 rectangles, immersed in concentrated HNO3, and heated in an oven at 110 °C for 2 h to attach hydroxyl on the surface of CC and remove the impurity on the surface. After the oven cooled to room temperature, the CC was washed with deionized water and ethanol for 30 min under ultrasonic conditions. Subsequently, 0.4 g of FeCl3·6H2O and 0.24 g of Na2SO4 were added to 35 mL of deionized water for dissolution. Then, the mixed solution was transferred into a 40 mL Teflon-lined stainless-steel autoclave, and a piece of CC (2 × 3 cm2) was immersed in the above solution and treated at 120 °C for 6 h to obtain the FeOOH precursor nanorod on CC. The Fe2O3 nanorod grown on CC (Fe2O3/CC) was obtained after heating the precursor at 450 °C in air for 3 h at a heating rate of 2 °C min−1. The mass loading of Fe2O3 on CC was measured to be 0.93 mg cm−2.

3.3. Synthesis of D-Fe2O3/CC

The chemical reduction process was conducted using a 1 M NaBH4 solution. First, 1.89 g of NaBH4 was added to 50 mL of deionized water to form a NaBH4 solution. Pieces of Fe2O3/CC were immersed in the NaBH4 solution for different times (0.5, 1, and 4 h) and then washed with water several times. The NaBH4 has strong reducibility and can serve as an “activator” to etch the crystallinity of Fe2O3. By changing the immersion time, the degree of reduction could be adjusted. The obtained low-crystallinity Fe2O3 on CC was named D-Fe2O3-xh, where x represents the 0.5, 1, and 4 h reduction times. The mass loading of D-Fe2O3-1 h was measured to be 0.72 mg cm−2.

3.4. Synthesis of D-Fe2O3@PPy/CC

The constant-voltage electrochemical deposition of PPy on low-crystallinity Fe2O3 on CC was conducted in a standard three-electrode system using D-Fe2O3/CC, Pt foil, and Ag/AgCl as the work, counter, and reference electrode, respectively. The electrolyte was prepared by adding 0.69 mL of a pyrrole monomer and 2.45 g of NaClO4 to 100 mL of deionized water. The mass loading of PPy can be adjusted by changing the electrodeposition time. The applied potential was set to 0.8 V with operation times of 30 s, 60 s, 120 s, and 180 s. Finally, the D-Fe2O3@PPy/CC electrodes were obtained after washing with deionized water and drying at 60 °C for 8 h. The obtained D-Fe2O3@PPy/CC electrodes with different electrodeposition times were named D-Fe2O3@PPy-30 s, D-Fe2O3@PPy-60 s, D-Fe2O3@PPy-120 s, and D-Fe2O3@PPy-180 s. The mass loading of D-Fe2O3@PPy-120 s was measured to be 0.96 mg cm−2.

3.5. Material Characterization

A Raman microscope (λ = 455 nm; Thermo Fisher Scientific Inc., Waltham, MA, USA), was used to obtain the Raman dates of all samples. A Philips PW1730 X-ray diffractometer was used to collect X-ray diffraction (XRD) patterns. The X-ray photoelectron spectroscopy (XPS) spectra of all samples were collected using an X-ray photoelectron spectrometer (AXIS NOVA, Kratos, UK). The scanning electron microscopy (SEM) images were characterized by an emission scanning electron microscope (Regulus 8230 field, Hitachi, Tokyo, Japan).

3.6. Electrochemical Measurement

A CHI660E electrochemical workstation was used to evaluate the electrochemical performance, including cyclic voltammetry (CV) and galvanostatic charge–discharge (GCD). All electrodes were measured in a three-electrode system using Pt foil as the counter electrode, Ag/AgCl as the reference electrode, and 1 M Na2SO4 as the electrolyte. All electrochemical performances were tested at 25 °C and ambient pressure. The specific capacitance (C, mF cm−2) was collected by the GCD and CV data according to the equations C = I × Δ t /(S × Δ V) and C = I d V /(2 × Δ V × v × S), respectively, where I (mA), I d V , Δ V (V), Δ t (s), and v (V s−1) represent the discharging current, the area of the CV curve, the potential window, the discharge time, and the scan rate, respectively. S (cm2) represents the geometric area of the working electrode (in this work, S = 0.5 cm2).

4. Conclusions

In summary, we demonstrated polypyrrole-coated low-crystallinity Fe2O3 supported on carbon cloth as the self-support anode for SCs. The chemical reduction treatment offered more surface roughness for D-Fe2O3/CC, favoring the Faradaic reaction derived from the ion diffusion in the bulk phase. The construction of a Fe2O3/PPy interface further enhanced the charge storage on the surface or near the surface of the electrode, thus efficiently hindering the volume effect of Fe2O3 during the charge/discharge process. Compared with the pristine Fe2O3/CC electrode, the enlarged surface roughness and Fe2O3/PPy interface gave more active sites for both Faradaic reactions and ion transfer, and the capacity contributions from diffusion-controlled and capacitive behaviors were simultaneously improved. The D-Fe2O3@PPy/CC electrode showed a specific capacitance of 615 mF cm−2 at 1 mA cm−2 and still retained 79.3% of its initial capacitance at 10 mA cm−2 after 5000 cycles.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules28010434/s1, Figure S1: Wide-scan XPS spectrum of D-Fe2O3@PPy/CC electrode; Figure S2: N 1s XPS spectrum of D-Fe2O3@PPy/CC electrode; Figure S3: TEM image of D-Fe2O3@PPy and the related SAED pattern; Figure S4: CV curves of D-Fe2O3/CC electrodes: (a) pristine Fe2O3, (b) D-Fe2O3-0.5 h, (c) D- Fe2O3-1 h, and (d) D-Fe2O3-4 h; Figure S5: GCD curves of D-Fe2O3/CC electrodes: (a) pristine Fe2O3, (b) D-Fe2O3-0.5 h, (c) D-Fe2O3-1 h, and (d) D-Fe2O3-4 h; Figure S6: CV curves of D-Fe2O3@PPy/CC electrodes: (a) D-Fe2O3@PPy-30 s, (b) D-Fe2O3@PPy-60 s, (c) D-Fe2O3@PPy-120 s, and (d) D-Fe2O3@PPy-180 s; Figure S7: GCD curves of D-Fe2O3@PPy/CC electrodes: (a) D-Fe2O3@PPy-30 s, (b) D-Fe2O3@PPy-60 s, (c) D-Fe2O3@PPy-120 s, and (d) D-Fe2O3@PPy-180 s; Figure S8: SEM images of (a) Fe2O3/CC and (b) D-Fe2O3@PPy/CC electrodes after long cycling test; Table S1: Electrochemical performances of Fe2O3-based and D-Fe2O3@PPy/CC anodes for supercapacitors. Citation: [18,19,33,34,43,44,45,46,47,48,49].

Author Contributions

Conceptualization, S.Y. and Y.W.; methodology, S.Y.; software, Y.W.; validation, S.Y., Y.W. and C.W.; formal analysis, C.W.; investigation, C.W.; resources, S.Y.; data curation, M.L.; writing—original draft preparation, C.W., M.L., Z.P., D.H. and Y.W.; writing—review and editing, Y.W. and S.Y.; visualization, Y.W.; supervision, S.Y.; project administration, S.Y.; funding acquisition, S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 2197812).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article.

Acknowledgments

We acknowledge Yingming Zhu from the Institute of New Energy and Low-Carbon Technology of Sichuan University for his help with SEM characterization and Panpan Li from the Shiyanjia Lab for XPS characterization (http//:www.shiyanjia.com) accessed on 21 December 2022. We also appreciate Xiang Lin, Wen Tian, Jie Wei, and Ji Li from the Engineering Teaching Center, School of Chemical Engineering of Sichuan University, for Raman and FTIR measurements.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

The samples are available from the authors.

References

  1. Sun, H.; Mei, L.; Liang, J.; Zhao, Z.; Lee, C.; Fei, H.; Ding, M.; Lau, J.; Li, M.; Wang, C.; et al. Three-dimensional holey-graphene/niobia composite architectures for ultrahigh-rate energy storage. Science 2017, 356, 599–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Xiao, X.; Song, H.; Lin, S.; Zhou, Y.; Zhan, X.; Hu, Z.; Zhang, Q.; Sun, J.; Yang, B.; Li, T.; et al. Scalable salt-templated synthesis of two-dimensional transition metal oxides. Nat. Commun. 2016, 7, 11296. [Google Scholar] [CrossRef] [Green Version]
  3. Wang, Y.; Wang, H.; Zhang, T.C.; Yuan, S.; Liang, B. N-doped porous carbon derived from rGO-Incorporated polyphenylenediamine composites for CO2 adsorption and supercapacitors. J. Power Sources 2020, 472, 228610. [Google Scholar] [CrossRef]
  4. Wang, Y.; Xiao, J.; Wang, H.; Zhang, T.C.; Yuan, S. N-doped porous carbon derived from solvent-free synthesis of cross-linked triazine polymers for simultaneously achieving CO2 capture and supercapacitors. Chem. Eur. J. 2021, 27, 7908–7914. [Google Scholar] [CrossRef]
  5. Mourad, E.; Coustan, L.; Lannelongue, P.; Zigah, D.; Mehdi, A.; Vioux, A.; Freunberger, S.A.; Favier, F.; Fontaine, O. Biredox ionic liquids with solid-like redox density in the liquid state for high-energy supercapacitors. Nat. Mater. 2017, 16, 446–453. [Google Scholar] [CrossRef]
  6. Sun, B.; Xiong, P.; Maitra, U.; Langsdorf, D.; Yan, K.; Wang, C.; Janek, J.; Schröder, D.; Wang, G. Design strategies to enable the efficient use of sodium metal anodes in high-energy batteries. Adv. Mater. 2020, 32, 1903891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Wang, Y.; Xiao, J.; Wang, H.; Zhang, T.C.; Yuan, S. Binary doping of nitrogen and phosphorus into porous carbon: A novel di-functional material for enhancing CO2 capture and super-capacitance. J. Mater. Sci. Tech. 2022, 99, 73–81. [Google Scholar] [CrossRef]
  8. Zhi, J.; Zhou, M.; Zhang, Z.; Reiser, O.; Huang, F. Interstitial boron-doped mesoporous semiconductor oxides for ultratransparent energy storage. Nat. Commun. 2021, 12, 445. [Google Scholar] [CrossRef]
  9. Zhu, Z.; Liu, Z.; Yin, Y.; Yuan, Y.; Meng, Y.; Jiang, T.; Peng, Q.; Wang, W.; Chen, W. Production of a hybrid capacitive storage device via hydrogen gas and carbon electrodes coupling. Nat. Commun. 2022, 13, 2805. [Google Scholar] [CrossRef]
  10. Wang, Y.; Zhang, T.; Xiao, J.; Tian, X.; Yuan, S. Enhancing electrochemical performance of ultrasmall Fe2O3-embedded carbon nanotubes via combusting-induced high-valence dopants. J. Mater. Sci. Technol. 2023, 134, 142–150. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Hu, Y.; Wang, Z.; Lin, T.; Zhu, X.; Luo, B.; Hu, H.; Xing, W.; Yan, Z.; Wang, L. Lithiation-induced vacancy engineering of Co3O4 with improved faradic reactivity for high-performance supercapacitor. Adv. Funct. Mater. 2020, 30, 2004172. [Google Scholar] [CrossRef]
  12. Mane, S.A.; Kashale, A.A.; Kamble, G.P.; Kolekar, S.S.; Dhas, S.D.; Patil, M.D.; Moholkar, A.V.; Sathe, B.R.; Ghule, A.V. Facile synthesis of flower-like Bi2O3 as an efficient electrode for high performance asymmetric supercapacitor. J. Alloys Compd. 2022, 926, 166722. [Google Scholar] [CrossRef]
  13. Huang, Z.; Song, Y.; Feng, D.; Sun, Z.; Sun, X.; Liu, X. High mass loading MnO2 with hierarchical nanostructures for supercapacitors. ACS Nano 2018, 12, 3557–3567. [Google Scholar] [CrossRef] [PubMed]
  14. Kumar, Y.A.; Das, H.T.; Guddeti, P.R.; Nallapureddy, R.R.; Pallavolu, M.R.; Alzahmi, S.; Obaidat, I.M. Self-supported Co3O4@Mo-Co3O4 needle-like nanosheet heterostructured architectures of battery-type electrodes for high-performance asymmetric supercapacitors. Nanomaterials 2022, 12, 2330. [Google Scholar] [CrossRef] [PubMed]
  15. Moniruzzaman, M.; Anil Kumar, Y.; Pallavolu, M.R.; Arbi, H.M.; Alzahmi, S.; Obaidat, I.M. Two-dimensional core-shell structure of cobalt-doped@MnO2 nanosheets grown on nickel foam as a binder-free battery-type electrode for supercapacitor application. Nanomaterials 2022, 12, 3187. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, H.; Xu, Z.; Yi, H.; Wei, H.; Guo, Z.; Wang, X. One-step preparation of single-crystalline Fe2O3 particles/graphene composite hydrogels as high performance anode materials for supercapacitors. Nano Energy 2014, 7, 86–96. [Google Scholar] [CrossRef]
  17. Zeng, Y.; Han, Y.; Zhao, Y.; Zeng, Y.; Yu, M.; Liu, Y.; Tang, H.; Tong, Y.; Lu, X. Advanced Ti-Doped Fe2O3@PEDOT core/shell anode for high-energy asymmetric supercapacitors. Adv. Energy Mater. 2015, 5, 1402176. [Google Scholar] [CrossRef]
  18. Wang, Y.; Xiao, J.; Zhang, T.; Ouyang, L.; Yuan, S. Single-step preparation of ultrasmall iron oxide-embedded carbon nanotubes on carbon cloth with excellent superhydrophilicity and enhanced supercapacitor performance. ACS Appl. Mater. Interfaces 2021, 13, 45670–45678. [Google Scholar] [CrossRef]
  19. Sun, S.; Zhai, T.; Liang, C.; Savilov, S.V.; Xia, H. Boosted crystalline/amorphous Fe2O3-δ core/shell heterostructure for flexible solid-state pseudocapacitors in large scale. Nano Energy 2018, 45, 390–397. [Google Scholar] [CrossRef]
  20. Wang, Y.; Wu, C.; Wu, Z.; Cui, G.; Xie, F.; Guo, X.; Sun, X. FeP nanorod arrays on carbon cloth: A high-performance anode for sodium-ion batteries. Chem. Commun. 2018, 54, 9341–9344. [Google Scholar] [CrossRef]
  21. Wang, Y.; Pan, Q.; Jia, K.; Wang, H.; Gao, J.; Xu, C.; Zhong, Y.; Alshehri, A.A.; Alzahrani, K.A.; Guo, X.; et al. Ni2P nanosheets on carbon cloth: An efficient flexible electrode for sodium-ion batteries. Inorg. Chem. 2019, 58, 6579–6583. [Google Scholar] [CrossRef] [PubMed]
  22. Zhai, T.; Sun, S.; Liu, X.; Liang, C.; Wang, G.; Xia, H. Achieving insertion-like capacity at ultrahigh rate via tunable surface pseudocapacitance. Adv. Mater. 2018, 30, 1706640. [Google Scholar] [CrossRef] [PubMed]
  23. Tang, J.; Yuan, H.; Duan, Q.; Liu, Y.; Wang, Y.; Yuan, S. Phosphorus-functionalized low-crystallinity transition-metal oxide nanorod arrays grown on carbon cloth for high-performance asymmetric supercapacitors. Colloids Surf. A 2022, 654, 130189. [Google Scholar] [CrossRef]
  24. Owusu, K.A.; Qu, L.; Li, J.; Wang, Z.; Zhao, K.; Yang, C.; Hercule, K.M.; Lin, C.; Shi, C.; Wei, Q.; et al. Low-crystalline iron oxide hydroxide nanoparticle anode for high-performance supercapacitors. Nat. Commun. 2017, 8, 14264. [Google Scholar] [CrossRef] [Green Version]
  25. Lu, X.; Zeng, Y.; Yu, M.; Zhai, T.; Liang, C.; Xie, S.; Balogun, M.; Tong, Y. Oxygen-deficient hematite nanorods as high-performance and novel negative electrodes for flexible asymmetric supercapacitors. Adv. Mater. 2014, 26, 3148–3155. [Google Scholar] [CrossRef]
  26. Wu, L.; Zheng, J.; Wang, L.; Xiong, X.; Shao, Y.; Wang, G.; Wang, J.; Zhong, S.; Wu, M. PPy-encapsulated SnS2 nanosheets stabilized by defects on a TiO2 support as a durable anode material for lithium-ion batteries. Angew. Chem. Int. Ed. 2019, 58, 811–815. [Google Scholar] [CrossRef]
  27. Shimoga, G.; Palem, R.; Choi, D.; Shin, E.; Ganesh, P.; Saratale, G.; Saratale, R.; Lee, S.; Kim, S. Polypyrrole-based metal nanocomposite electrode materials for high-performance supercapacitors. Metals 2021, 11, 905. [Google Scholar] [CrossRef]
  28. Yuan, L.; Yao, B.; Hu, B.; Huo, K.; Chen, W.; Zhou, J. Polypyrrole-coated paper for flexible solid-state energy storage. Energy Environ. Sci. 2013, 6, 470–476. [Google Scholar] [CrossRef]
  29. Liu, J.; Zhou, W.; Lai, L.; Yang, H.; Hua Lim, S.; Zhen, Y.; Yu, T.; Shen, Z.; Lin, J. Three dimensionals α-Fe2O3/polypyrrole (Ppy) nanoarray as anode for micro lithium ion batteries. Nano Energy 2013, 2, 726–732. [Google Scholar] [CrossRef]
  30. Zuo, W.; Zang, L.; Wang, X.; Liu, Q.; Qiu, J.; Liang, C.; Liu, X.; Yang, C. Flexible Polypyrrole@Fe2O3@stainless steel yarn composite electrode for symmetric thread-like supercapacitor with extended operating voltage window in Li2SO4-based aqueous electrolyte. Adv. Sustain. Syst. 2020, 4, 2000173. [Google Scholar] [CrossRef]
  31. Arbi, H.; Yadav, A.; Anil Kumar, Y.; Moniruzzaman, M.; Alzahmi, S.; Obaidat, I. Polypyrrole-assisted Ag doping strategy to boost Co(OH)2 nanosheets on Ni foam as a novel electrode for high-performance hybrid supercapacitors. Nanomaterials 2022, 12, 3982. [Google Scholar] [CrossRef] [PubMed]
  32. Huang, Y.; Zhu, M.; Pei, Z.; Huang, Y.; Geng, H.; Zhi, C. Extremely stable polypyrrole achieved via molecular ordering for highly flexible supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 2435–2440. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, Y.; Du, Z.; Xiao, J.; Cen, W.; Yuan, S. Polypyrrole-encapsulated Fe2O3 nanotube arrays on a carbon cloth support: Achieving synergistic effect for enhanced supercapacitor performance. Electrochim. Acta 2021, 386, 138486. [Google Scholar] [CrossRef]
  34. Wang, L.; Yang, H.; Liu, X.; Zeng, R.; Li, M.; Huang, Y.; Hu, X. Constructing hierarchical tectorum-like α-Fe2O3/PPy nanoarrays on carbon cloth for solid-state asymmetric supercapacitors. Angew. Chem. Int. Ed. 2017, 56, 1105–1110. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Y.; Zhang, Y.; Zhang, T.C.; Xiang, G.; Wang, X.; Yuan, S. Removal of trace arsenite through simultaneous photocatalytic oxidation and adsorption by magnetic Fe3O4@PpPDA@TiO2 core–shell nanoparticles. ACS Appl. Nano Mater. 2020, 3, 8495–8504. [Google Scholar] [CrossRef]
  36. Wu, X.; Liu, R.; Zhao, J.; Fan, Z. Advanced carbon materials with different spatial dimensions for supercapacitors. Nano Mater. Sci. 2021, 3, 241–267. [Google Scholar] [CrossRef]
  37. Liu, H.; Huo, W.; Zhang, T.C.; Ouyang, L.; Yuan, S. Photocatalytic removal of tetracycline by a Z-scheme heterojunction of bismuth oxyiodide/exfoliated g-C3N4: Performance, mechanism, and degradation pathway. Mater. Today Chem. 2022, 23, 100729. [Google Scholar] [CrossRef]
  38. Wang, Y.; Zhang, P.; Zhang, T.C.; Xiang, G.; Wang, X.; Pehkonen, S.; Yuan, S. A magnetic γ-Fe2O3@PANI@TiO2 core–shell nanocomposite for arsenic removal via a coupled visible-light-induced photocatalytic oxidation–adsorption process. Nanoscale Adv. 2020, 2, 2018–2024. [Google Scholar] [CrossRef] [Green Version]
  39. Wang, Y.; Liu, X.; Chen, Q.; Zhang, T.C.; Ouyang, L.; Yuan, S. Simultaneous photocatalytic oxidation and adsorption for efficient As(III) removal by magnetic BiOI/γ-Fe2O3 core–shell nanoparticles. Mater. Today Chem. 2022, 24, 100823. [Google Scholar] [CrossRef]
  40. Ji, W.; Xiong, Y.; Wang, Y.; Zhang, T.C.; Yuan, S. Multilayered TNAs/SnO2/PPy/β-PbO2 anode achieving boosted electrocatalytic oxidation of As(III). J. Hazard. Mater. 2022, 430, 128449. [Google Scholar] [CrossRef]
  41. Xiao, J.; Wang, Y.; Zhang, T.C.; Yuan, S. N,S-containing polycondensate-derived porous carbon materials for superior CO2 adsorption and supercapacitor. Appl. Surf. Sci. 2021, 562, 150128. [Google Scholar] [CrossRef]
  42. Xiao, J.; Wang, Y.; Zhang, T.C.; Ouyang, L.; Yuan, S. Phytic acid-induced self-assembled chitosan gel-derived N, P–co-doped porous carbon for high-performance CO2 capture and supercapacitor. J. Power Sources 2022, 517, 230727. [Google Scholar] [CrossRef]
  43. Sui, S.; Sha, J.; Deng, X.; Zhu, S.; Ma, L.; He, C.; Liu, E.; He, F.; Shi, C.; Zhao, N. Boosting the charge transfer efficiency of metal oxides/carbon nanotubes composites through interfaces control. J. Power Sources 2021, 489, 229501. [Google Scholar] [CrossRef]
  44. Cai, D.; Du, J.; Zhu, C.; Cao, Q.; Huang, L.; Wu, J.; Zhou, D.; Xia, Q.; Chen, T.; Guan, C.; et al. Iron oxide nanoneedles anchored on N-doped carbon nanoarrays as an electrode for high-performance hybrid supercapacitor. ACS Appl. Energy Mater. 2020, 3, 12162–12171. [Google Scholar] [CrossRef]
  45. Saravanakumar, B.; Ko, T.; Kim, B. Rational design of binder-free ZnCo2O4 and Fe2O3 decorated porous 3D Ni as high-performance electrodes for asymmetric supercapacitor. Ceram. Int. 2018, 44, 10635–10645. [Google Scholar] [CrossRef]
  46. Liang, H.; Xia, C.; Emwas, A.; Anjum, D.; Miao, X.; Alshareef, H. Phosphine plasma activation of α-Fe2O3 for high energy asymmetric supercapacitors. Nano Energy 2018, 49, 155–162. [Google Scholar] [CrossRef]
  47. Sun, J.; Huang, Y.; Fu, C.; Huang, Y.; Zhu, M.; Tao, X.; Zhi, C.; Hu, H. A high performance fiber-shaped PEDOT@MnO2//C@Fe3O4 asymmetric supercapacitor for wearable electronics. J. Mater. Chem. A 2016, 4, 14877–14883. [Google Scholar] [CrossRef]
  48. Lu, X.; Chen, X.; Zhou, W.; Tong, Y.; Li, G. α-Fe2O3@PANI Core–shell nanowire arrays as negative electrodes for asymmetric supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 14843–14850. [Google Scholar] [CrossRef]
  49. Yang, P.; Ding, Y.; Lin, Z.; Chen, Z.; Li, Y.; Qiang, P.; Ebrahimi, M.; Mai, W.; Wong, C.; Wang, Z. Low-cost high-performance solid-state asymmetric supercapacitors based on MnO2 nanowires and Fe2O3 nanotubes. Nano Let. 2014, 14, 731–736. [Google Scholar] [CrossRef]
  50. Shao, Y.; El-Kady, M.F.; Sun, J.; Li, Y.; Zhang, Q.; Zhu, M.; Wang, H.; Dunn, B.; Kaner, R.B. Design and mechanisms of asymmetric supercapacitors. Chem. Rev. 2018, 118, 9233–9280. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic illustration of the fabrication process of a D-Fe2O3@PPy/CC electrode. Raman spectra of (b) D-Fe2O3-xh (x = 0.5, 1, and 4) and (c) D-Fe2O3@PPy-ys (y = 30, 60, 120, and 180) electrodes. The corresponding XRD patterns of (d) D-Fe2O3-xh and (e) D-Fe2O3@PPy-ys electrodes.
Figure 1. (a) Schematic illustration of the fabrication process of a D-Fe2O3@PPy/CC electrode. Raman spectra of (b) D-Fe2O3-xh (x = 0.5, 1, and 4) and (c) D-Fe2O3@PPy-ys (y = 30, 60, 120, and 180) electrodes. The corresponding XRD patterns of (d) D-Fe2O3-xh and (e) D-Fe2O3@PPy-ys electrodes.
Molecules 28 00434 g001
Figure 2. The XPS spectra of Fe2O3/CC for (a) Fe 2p and (b) O 1s regions. The XPS spectra of D-Fe2O3/CC for (c) Fe 2p and (d) O 1s regions. The XPS spectra of D-Fe2O3@PPy/CC for (e) Fe 2p and (f) O 1s regions.
Figure 2. The XPS spectra of Fe2O3/CC for (a) Fe 2p and (b) O 1s regions. The XPS spectra of D-Fe2O3/CC for (c) Fe 2p and (d) O 1s regions. The XPS spectra of D-Fe2O3@PPy/CC for (e) Fe 2p and (f) O 1s regions.
Molecules 28 00434 g002
Figure 3. SEM images of D-Fe2O3/CC electrode prepared using different immersion times: (a) pristine Fe2O3, (b) D-Fe2O3-0.5 h, (c) D-Fe2O3-1 h, and (d) D-Fe2O3-4 h.
Figure 3. SEM images of D-Fe2O3/CC electrode prepared using different immersion times: (a) pristine Fe2O3, (b) D-Fe2O3-0.5 h, (c) D-Fe2O3-1 h, and (d) D-Fe2O3-4 h.
Molecules 28 00434 g003
Figure 4. SEM images of D-Fe2O3@PPy/CC electrode prepared using different electrodeposition times: (a) D-Fe2O3@PPy-30 s, (b) D-Fe2O3@PPy-60 s, (c) D-Fe2O3@PPy-120 s, and (d) D-Fe2O3@PPy-180 s.
Figure 4. SEM images of D-Fe2O3@PPy/CC electrode prepared using different electrodeposition times: (a) D-Fe2O3@PPy-30 s, (b) D-Fe2O3@PPy-60 s, (c) D-Fe2O3@PPy-120 s, and (d) D-Fe2O3@PPy-180 s.
Molecules 28 00434 g004
Figure 5. (a) CV curves of D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) at different scan rates. (b) GCD curves of D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) at different current densities. (c) Rate performances of D-Fe2O3/CC electrodes. (d) Rate performances of D-Fe2O3@PPy/CC electrodes. (e) Diffusion and capacitive contributions of Fe2O3/CC, D-Fe2O3/CC (D-Fe2O3-1 h), and D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) electrodes at scan rates of 10, 20, 60, 100 mV s−1. (f) Long-cycling performances at current density of 10 mA cm−1 for 5000 cycles.
Figure 5. (a) CV curves of D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) at different scan rates. (b) GCD curves of D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) at different current densities. (c) Rate performances of D-Fe2O3/CC electrodes. (d) Rate performances of D-Fe2O3@PPy/CC electrodes. (e) Diffusion and capacitive contributions of Fe2O3/CC, D-Fe2O3/CC (D-Fe2O3-1 h), and D-Fe2O3@PPy/CC (D-Fe2O3@PPy-120 s) electrodes at scan rates of 10, 20, 60, 100 mV s−1. (f) Long-cycling performances at current density of 10 mA cm−1 for 5000 cycles.
Molecules 28 00434 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wu, C.; Pei, Z.; Lv, M.; Huang, D.; Wang, Y.; Yuan, S. Polypyrrole-Coated Low-Crystallinity Iron Oxide Grown on Carbon Cloth Enabling Enhanced Electrochemical Supercapacitor Performance. Molecules 2023, 28, 434. https://doi.org/10.3390/molecules28010434

AMA Style

Wu C, Pei Z, Lv M, Huang D, Wang Y, Yuan S. Polypyrrole-Coated Low-Crystallinity Iron Oxide Grown on Carbon Cloth Enabling Enhanced Electrochemical Supercapacitor Performance. Molecules. 2023; 28(1):434. https://doi.org/10.3390/molecules28010434

Chicago/Turabian Style

Wu, Chunhui, Zifan Pei, Menglin Lv, Duchen Huang, Yuan Wang, and Shaojun Yuan. 2023. "Polypyrrole-Coated Low-Crystallinity Iron Oxide Grown on Carbon Cloth Enabling Enhanced Electrochemical Supercapacitor Performance" Molecules 28, no. 1: 434. https://doi.org/10.3390/molecules28010434

Article Metrics

Back to TopTop