Next Article in Journal
Functional Polymer Nanocomposites with Increased Anticorrosion Properties and Wear Resistance for Water Transport
Next Article in Special Issue
Research Status of and Prospects for 3D Printing for Continuous Fiber-Reinforced Thermoplastic Composites
Previous Article in Journal
Effect of Fibres on Physico-Mechanical Properties of Bulk-Fill Resin Composites
Previous Article in Special Issue
Flexible Wet-Spun PEDOT:PSS Microfibers Integrating Thermal-Sensing and Joule Heating Functions for Smart Textiles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Highly Conductive Porous Fe3O4@RGO/PEDOT:PSS Composite Films via Acid Post-Treatment and Their Applications as Electrochemical Supercapacitor and Thermoelectric Material

1
College of Physics and Electronic Engineering, Xinyang Normal University, Xinyang 464000, China
2
Key Laboratory of Advanced Micro/Nano Functional Materials of Henan Province, Xinyang Normal University, Xinyang 464000, China
3
Energy-Saving Building Materials Innovative Collaboration Center of Henan Province, Xinyang Normal University, Xinyang 464000, China
*
Author to whom correspondence should be addressed.
Submission received: 4 July 2023 / Revised: 16 August 2023 / Accepted: 16 August 2023 / Published: 18 August 2023
(This article belongs to the Special Issue Polymer-Based Flexible Materials)

Abstract

:
As a remarkable multifunctional material, ferroferric oxide (Fe3O4) exhibits considerable potential for applications in many fields, such as energy storage and conversion technologies. However, the poor electronic and ionic conductivities of classical Fe3O4 restricts its application. To address this challenge, Fe3O4 nanoparticles are combined with graphene oxide (GO) via a typical hydrothermal method, followed by a conductive wrapping using poly(3,4-ethylenedioxythiophene):poly(styrene sulfonic sulfonate) (PEDOT:PSS) for the fabrication of composite films. Upon acid treatment, a highly conductive porous Fe3O4@RGO/PEDOT:PSS hybrid is successfully constructed, and each component exerts its action that effectively facilitates the electron transfer and subsequent performance improvement. Specifically, the Fe3O4@RGO/PEDOT:PSS porous film achieves a high specific capacitance of 244.7 F g−1 at a current of 1 A g−1. Furthermore, due to the facial fabrication of the highly conductive networks, the free-standing film exhibits potential advantages in flexible thermoelectric (TE) materials. Notably, such a hybrid film shows a high electric conductivity (σ) of 507.56 S cm−1, a three times greater value than the Fe3O4@RGO component, and achieves an optimized Seebeck coefficient (S) of 13.29 μV K−1 at room temperature. This work provides a novel route for the synthesis of Fe3O4@RGO/PEDOT:PSS multifunctional films that possess promising applications in energy storage and conversion.

1. Introduction

Considering the rapid growth of the world’s economy and the continual consumption of fossil fuels, green energy (such as wind, hydropower, thermoelectric, etc.) and electrical energy storage devices are in urgent need for many applications such as portable electronic devices and electric vehicles [1,2,3,4,5,6]. Among the variety of energy storage devices, supercapacitors (SCs) have attracted widespread attention for their high power density, reliable safety, outstanding cycling stability, and low cost [2,7,8,9,10]. Generally, there are two kinds of supercapacitors: electrochemical double-layer capacitors (EDLCs) and pseudocapacitors. While EDLCs store electricity through the double-layer effect, the pseudocapacitor works through a fast redox reaction, which is essential for harvesting outstanding capacitive ability [11,12,13].
To explore desired electrode materials for supercapacitors, many efforts have been put into researching transition metal oxides, such as Fe3O4, Fe2O3, Co3O4, RuO2, MnO2, etc. [10,14,15,16,17,18]. Among them, iron ferrite of Fe3O4 has been proposed as a potential supercapacitor material because of its high specific capacitance, easy redox reaction, rich natural storage, and environmental friendliness [10]. Nevertheless, Fe3O4 has a low conductivity in nature, which limits its electrochemical performance. Furthermore, it remains a challenge to avoid nanoparticle agglomeration during the preparation of electrode materials. In order to solve the above problems, Fe3O4 is commonly combined with carbon-based materials, especially graphene and CNTs. For instance, through a layer-by-layer method, the obtained Fe3O4/RGO multilayer electrodes exhibited a specific capacitance of 151 F g−1 when a current density of 0.9 A g−1 was used, and after 1000 cycles, the capacitance retained 85% of its original value, indicating good cycling stability [19]. A sandwich-like Fe3O4/MnO2/RGO nanocomposite was explored and the value of specific capacitance reached 77.5 F g−1 at 0.5 A g−1 and kept 35 F g−1 at 20 A g−1 in 1 M Na2SO4 [20]. CNT/Fe3O4 nanocomposites synthesized through the hydrothermal method also achieved a specific capacitance of 117.2 F g−1 at 10 mA cm−2 in a 6 M KOH electrolyte [21]. A novel BRGO/Fe3O4-MWCNT hybrid nanocomposite was successfully fabricated and possessed good supercapacitance performance (165 F g−1 at a current density of 2 A g−1) [22]. Not only that, the obtained BRGO/Fe3O4-MWCNT composites also possessed a high photo degradation efficiency. As conductive frameworks, carbon materials in these strategies effectively avoid the collapse of the nano-Fe3O4 particles and thus improve their electrochemical properties. However, to realize large capacitance and practical applications requires high mass loading of active Fe3O4, which in turn increases the electrode resistance and thus limits the performance characteristics of the composite electrodes. In addition, binder materials like polytetrafluoroethylene (PTFE) and polyvinylidene fluoride (PVDF) are frequently used during the preparation of metal-oxide-based nanocomposite films for the preparation of flexible composite materials [23,24]. However, these binders are nonconductive and decrease the electrical conductivity of the electrodes.
To address these issues, one promising strategy is to incorporate Fe3O4 nanoparticles onto carbon-based frameworks and coat them with conducting polymers to form highly interconnected networks for charge transformation. Thus, there is an urgent demand for highly conductive binders that can further disperse the packed Fe3O4/carbon nanostructures. Among various conducting polymers, PEDOT:PSS is water soluble and can be used as a binder that is capable of dispersing carbon-based materials and/or other kinds of nanomaterials in water. Additionally, PEDOT:PSS can achieve high conductivities via incorporating additives (such as organic solvents [25,26], ionic liquids [27], and inorganic salts [28], etc.) and post-treatment through polar solvents (e.g., DMSO [29,30], EG [31,32], etc.) or acids [33,34]. It is believed that the Fe3O4/carbon/PEDOT:PSS composite can be employed as an excellent capacity electrode material that possesses considerable potential application in energy storage devices. Furthermore, such a unique structure can effectively increase the conductivity of the composite material, and its potential application in energy conversion technologies, such as thermoelectric, photoelectric, and thermal sensor, can also be expected.
Herein, we construct a ternary system based on Fe3O4@RGO/PEDOT:PSS; the graphene oxide (GO) in the compound acts as a base supporting material, while the PEDOT:PSS serves as a highly conductive wrapping material. Notably, after acid treatment, the as-prepared hybrid composite is easily stripped from the glass substrate and forms porous, highly conductive, and flexible electrode films. Benefiting from the large-scale construction of porous structures and the highly connected conducting networks, the Fe3O4@RGO/PEDOT:PSS electrodes exhibit a high specific capacitance of 244.7 F g−1 at 1 A g−1, and a good rate capability that remains 146.0 F g−1 at 10 A g−1. Except for energy storage, the constructed hybrid films can also be used as thermoelectric (TE) materials, which are capable of converting low-grade and/or waste heat into electricity, making them an important source of green energy. A dimensionless figure of merit, ZT = S2σT/κ, is usually applied to evaluate the TE materials’ conversion efficiency, in which σ is the electrical conductivity of the material, S stands for the Seebeck coefficient, and T and κ represent absolute temperature and thermal conductivity, respectively. For polymers and their composites, the thermal conductivity is relatively low (lower than that of the inorganic TE materials by almost one to three orders of magnitude). Therefore, the power factor S2σ is a good approximation for comparing organic and hybrid thermoelectric materials. The experiment results reveal that the Fe3O4@RGO/PEDOT:PSS hybrid films possess better TE properties than those of their single components. The related mechanism is also discussed in detail.

2. Materials and Methods

2.1. Materials

PEDOT:PSS (PH1000, Mw = 326.388) was purchased from Heraeus Company (Hanau, Germany); both ferroferric oxide nanoparticles (Fe3O4, Mw = 231.54 g/mol, around 20 nm particle size) and lithium sulfate (Li2SO4, Mw = 109.94 g/mol) from Beijing InnoChem Science & Technology Co., Ltd. (Beijing, China); GO aqueous solution (5 mg/mL) from Suzhou Tanfeng Graphene Technology Co., Ltd. (Suzhou, China); perchloric acid (HClO4, 70~72%, ~1.76 g/mL) and hydroiodic acid (HI, 57%, 5.23 g/mL) were obtained from Tianjin DaMao Chemical Reagent Factory (Tianjin, China) and Shanghai Mclean Biochemical Technology Co., Ltd. (Shanghai, China), respectively. Water used in this work was all deionized (DI) water (its resistance is around 18.2 MΩ cm). All the agents employed in the experiments were utilized directly.

2.2. Preparation of Fe3O4@GO and Fe3O4@RGO

The Fe3O4@GO composites were prepared according to a typical hydrothermal method. Firstly, Fe3O4 nanoparticles (12.5 mg) were added to 12.5 mL of GO aqueous solution (2 mg mL−1). Then, 12.5 mL of deionized water was added and the mixture was shaken well. After 30 min of sonication, the homogenous solution obtained was transferred to a 50 mL Teflon-lined steel autoclave and heated to 180 °C for 12 h. Subsequently, the steel autoclave was taken out and cooled to room temperature. The black sediment was collected and allowed to freeze-dry for 24 h. Finally, ~20.3 mg of black-brown fluffy product was obtained. For comparison, some Fe3O4@GO powder was also treated using HI for preparation of Fe3O4@RGO.

2.3. Preparation of Fe3O4@GO/PEDOT:PSS Composite Films

The Fe3O4@GO/PEDOT:PSS composite films were prepared via a drop-coating method; 25 mg of Fe3O4@GO lyophilized powder was dispersed in 10 mL PEDOT:PSS solution (1 mg mL−1). The obtained suspension was sonicated for about 1 h at room temperature. Thereafter, the as-prepared solution was allowed to drop onto a precleaned glass substrate, followed by a drying process at room temperature.

2.4. Preparation of Fe3O4@RGO/PEDOT:PSS Free-Standing Films

For fabrication of the free-standing Fe3O4@RGO/PEDOT:PSS films, an acid post-treatment was applied. The as-prepared Fe3O4@GO/PEDOT:PSS hybrid films were first immersed in HClO4 for 24 h. After that, the films were washed using DI water, followed by an air-drying process. To ensure good conductivities of the electrodes, the samples were further treated using HI with a similar procedure as described above. During immersion using HI, a chemical reductant, GO, can be effectively reduced to RGO. Finally, the free-standing Fe3O4@RGO/PEDOT:PSS hybrid films were successfully fabricated.

2.5. Characterization and Measurements

A Hitachi 4800 field emission scanning electron microscope (FE-SEM) was employed to analyze the morphologies of the films (Hitachi Limited, Tokyo, Japan). A Thermo K-Alpha X-ray photoelectron spectroscope (XPS) was used to determine the electron-binding energies of the samples (Thermo Fisher Scientific, Shanghai, China). Raman spectroscopy was performed using a LabRAM HR Evolution instrument with a 532 nm laser (HORIBA, Shanghai, China). An SDT Q600 (TA Instruments, New Castle, DE, USA) was applied for thermogravimetric analysis (TGA). An electrochemical performance analysis of the samples was conducted using a CHI 660E electrochemical workstation (Shanghai CH instruments Co., Shanghai, China). The thermoelectric properties were measured by employing a thin-film thermoelectric test system (MRS-3 M, Wuhan Joule Yacht Science &Technology Co., Ltd., Wuhan, China).
In this study, cyclic voltammogram (CV) curves, galvanostatic charge–discharge (GCD) curves, and electrochemical impedance spectroscopy (EIS) were collected via a three-electrode system in which a platinum mesh, saturated calomel electrode, and the as-prepared sample were used as the counter electrode, reference electrode, and working electrode, respectively. A 1 M Li2SO4 aqueous solution was used as the electrolyte. The specific capacitance (Cg, F g−1) of the samples was calculated with the formula Cg = IΔt/mΔV, where I stands for constant discharge current (A), Δt for discharge time (s), and m and ΔV for the mass of the film sample and the potential window, respectively.

3. Results and Discussion

3.1. Fabrication of Fe3O4@RGO/PEDOT:PSS Free-Standing Film

Figure 1 provides a schematic illustration of how Fe3O4@RGO/PEDOT:PSS free-standing films are fabricated. First, Fe3O4 nanoparticles are combined with GO nanosheets for preparation of the Fe3O4@GO nanocomposites. After washing and freeze-drying, the Fe3O4@GO framework is successfully fabricated. For further improvement of the electric conductivity, a PEDOT:PSS conducting polymer is incorporated for conductive wrapping. With a subsequent HClO4 treatment, the conductivity of the PEDOT:PSS is dramatically improved, owing to the removal of nonconductive PSS and the formation of ordered molecular packing. Furthermore, a secondary acid treatment is conducted using HI, an efficient reducing agent that can eliminate the attached oxygen-containing functional groups and enhance the electrical conductivity of GO. Moreover, the Fe3O4 component is partly etched during the acid treatment, which contributes to the construction of interconnected porous nanostructures. Based on the above analysis, we infer that the obtained flexible, free-standing, and porous film should possess good electrochemical properties, which will be discussed latterly.

3.2. Structure Characterization and Analysis

First, SEM was applied to examine the surface morphology of the obtained precursors and the final products. The as-prepared Fe3O4@GO exhibits an interconnected, highly porous microstructure, as shown in Figure 2a,b. First, the ultrasonic process results in homogeneous dispersion of the GO aqueous solution and Fe3O4 nanoparticles. And the subsequent liquid interfacial polymerization, under high pressure and temperature, allows the Fe3O4 structures to grow uniformly on the surface of the GO nanosheets. Finally, the freeze-drying procedure maintains the lamellar structure of the GO, and a homogeneous porous nanostructure is successfully constructed. With the introduction of the conductive PEDOT:PSS polymer, the gained composite films no longer possess a porous architecture. As is well known, PEDOT:PSS is water soluble and can wrap around the surface of the Fe3O4@GO precursor, forming an electrically conductive polymer shell. As depicted in Figure 2c,d, it can be observed visually that the Fe3O4@GO/PEDOT:PSS film shows a highly crumpled surface, which should be derived from the Fe3O4@GO core and PEDOT:PSS shell.
Notably, the superficial structure of the hybrid has changed significantly after the acid treatment. As exhibited in Figure 2e,f, the porous nanostructure is rediscovered in the complex architecture of Fe3O4@RGO/PEDOT:PSS, which can be ascribed to the acid treatment. As those meso-/macropores would provide a large surface area, electrolyte transport and access to active sites could be enhanced during the charging/discharging process. EDS element mapping was also conducted, and the results (Figure 2g,h) clearly reveal that the Fe element is evenly distributed in the Fe3O4@RGO/PEDOT:PSS sample. All these features would be beneficial for enhancing the supercapacitive performance of the hybrid films. On the other hand, when it is used as a thermoelectric material, the porous nanostructure is of importance in suppressing the thermal conductivity, whereas the large number of holes is not conducive to the formation of conductive networks and thus leads to a reduction in electrical conductivity and a decrease in mechanical properties. And thus, the presence of a conductive binder becomes very important. Furthermore, the introduction of conductive polymers is conducive to achieving relatively low thermal conductivity (≤1 W·m−1·K−1 in general), and therefore exhibits enormous potential in TE applications.
In order to account for the structural variation in Fe3O4@RGO/PEDOT:PSS, a Raman spectroscopy experiment was performed. As is depicted in Figure 3a, the pristine Fe3O4@GO/PEDOT:PSS presents only two main characteristic peaks at 1341 and 1591 cm−1, which are associated with the D and G bands of GO, respectively [8,33,35]. Furthermore, two small peaks (centered at 218 and 284 cm−1), corresponding to Fe3O4, also appear in the Raman spectrum [21]. After acid treatment, the most obvious difference is that some novel peaks associated with PEDOT appear. As exhibited in Figure S1, for acid-treated PEDOT:PSS, peaks at 1561 cm−1 and 1504 cm−1 are assigned to the asymmetric Cα=Cβ stretching, while the peak position 1430 cm−1 corresponds to the symmetric Cα=Cβ(–O) stretching in the five-membered ring, 1366 cm−1 to the Cβ–Cβ stretching, 1254 cm−1 to the inter-ring Cα–Cα stretching, 1095 cm−1 to the C–O–C deformation, 990 cm−1 and 576 cm−1 to oxyethylene ring deformation, 857 cm−1 and 699 cm−1 to C–S bonds, and 437 cm−1 to SO2 bending [36]. With the incorporation of Fe3O4, a shift of some characteristic peaks is noticeable (see Figure 3b and Figure S1), indicating the interaction between the PEDOT:PSS and Fe3O4 filler. Furthermore, it is of importance to note that the peak at 1430 cm−1 (symmetric Cα=Cβ(–O) stretching) shifts to 1429 cm−1 in the composite sample. In general, the shift of symmetric Cα=Cβ(–O) stretching vibration is mainly related to the ratio between the benzoid and quinoid conformations. Due to the lack of conjugated π-electrons in Cα–Cβ, the band red-shift of the symmetric Cα=Cβ(–O) indicates that more quinoid conformations generate. Namely, with the introduction of Fe3O4@RGO nanoparticles, there is a conformation transition of the PEDOT molecules from the coiled benzoid to the extended quinoid structure, which facilitates the carrier transformation and is beneficial for electrical performance. Notably, with the acid treatment, the iron oxide characteristic peaks gradually weaken, indicating a decrease in the Fe3O4 component. This change in Fe3O4@RGO/PEDOT:PSS would favor the formation of porous nanostructures. When compared with the acid-treated PEDOT:PSS, some PEDOT characteristic peaks become indistinct in the composite film because of the decrease in the PEDOT proportion induced by the addition of the Fe3O4 filler. All these phenomena indicate the presence of residual Fe3O4 after acid treatment, and the nanostructures introduced certainly have an influence on the energy performance, which will be further discussed in the subsequent sections.
In order to further investigate the elemental composition and chemical state of the obtained composite films, X-ray electron spectroscopy (XPS) analysis was conducted, and the results are shown in Figure 4 and Figure S2. As depicted in the XPS survey spectrum (see Figure S2), after acid treatment the porous Fe3O4@RGO/PEDOT:PSS contains elements of Fe, C, O, and S. The high resolution of the Fe2p spectra was decomposed, and the result is shown in Figure 4a. As presented, Fe2+ is predominantly correlated with the peaks at 711.2 (Fe2p3/2) and 724.5 eV (Fe2p1/2), while Fe3+ is mainly associated with the peaks at 713.8 (Fe2p3/2) and 726.7 eV (Fe2p1/2) [37,38]. Additionally, the satellite peaks at 719.1 and 731.8 eV belong to Fe2+ and Fe3+, respectively [37]. All these phenomena illustrate the presence of residual Fe3O4 even after a relatively lengthy treatment with HClO4 and HI, which serve an important role in electrochemical performance.
To deepen understanding of the influence of acid treatment on structural variations in the composite films, the high-resolution S2p spectra for Fe3O4@GO/PEDOT:PSS (as prepared) and flexible Fe3O4@RGO/PEDOT:PSS (after acid treatment) were analyzed, and the results are displayed in Figure 4b,c. As can be seen, the as-prepared Fe3O4@GO/PEDOT:PSS exhibits two distinct types of S element, because the S2p binding energy of the thiophene unit in PEDOT is very different from that of the sulfonate group in PSS. Namely, there are two types of S; one is related to PSS (high-binding-energy region, 171.8~166.5 eV), and the other is associated with PEDOT (low-binding-energy region, 166.5~162.5 eV). For the S element in PSS, S2p peaks can be divided into S2p3/2 and S2p1/2 peaks, showing peak positions centered at 168.1 eV and 169.2 eV, respectively. Meanwhile, for the S element in PEDOT, S2p peaks can also be classified into S2p3/2 and S2p1/2 peaks, centered at 164.1 eV and 165.5 eV, respectively. Comparing Figure 4b with Figure 4c, it can be found that the relative content of PSS to PEDOT decreases distinctly. This ratio decrease can be attributed to conformational variations caused by PSS removal during acid treatment. PSS extraction can be generally quantified by figuring the integral area ratio of the characteristic peaks. The calculation results are shown in Figure S3 (see Supporting Information). A PSS/PEDOT surface element ratio of 2.06 can be achieved for the as-obtained Fe3O4@GO/PEDOT:PSS composite, whereas the post-treatment using perchloric acid and hydroiodic acid reduces the PSS/PEDOT ratio to 1.48 (for the Fe3O4@RGO/PEDOT:PSS sample), confirming the PSS removal effect of the acids’ treatment. Because PSS itself is not conductive in nature, its reduction in content can effectively improve the conductive characteristics of the hybrid films, which would not only benefit the enhancement of the electrochemical performance, but also potentially improve the TE properties of the composite electrodes.

3.3. Electrochemical Properties of Fe3O4@RGO/PEDOT:PSS Free-Standing Films

CV and GCD were conducted to estimate the electrochemical properties of the electrodes, and the results are presented in Figure 5a,b. It is suggested that the GO skeleton is partially reduced to RGO via HI treatment. Combined with the wrapping effect of the highly conductive PEDOT:PSS, a continuous conductive network is formed, which provides good paths for the transport of ions and rapid redox reactions. This flexible film is expected to be employed as a high-performance film electrode for SCs, working simultaneously as an electrically conducting current collector and active electrode material. As can be observed, the Fe3O4@RGO/PEDOT:PSS self-supporting film exhibits a quasi-rectangular CV curve at a scan rate of 50 mV s−1 (see Figure 5a). For the Fe3O4@RGO electrode, nearly no redox peak can be seen, owing to its relatively low electric conductivity, whereas for the Fe3O4@RGO/PEDOT:PSS film, the corresponding redox couples become more clear, which is related to the surface redox reactions between Fe2+ and Fe3+ [21]. Meanwhile, the area enclosed by the CV curve of the Fe3O4@RGO/PEDOT:PSS film triples that of Fe3O4@RGO, representing a larger capacitance. Nevertheless, the pristine PEDOT:PSS itself possesses a smaller CV curve area due to its poor electric conductivity. Promisingly, the nonconductive PSS can be partly removed via acid treatment, which is beneficial for improving conductivity. Meanwhile, Fe3O4 nanoparticles would be partially etched away via acid treatment, and some macroporous microstructures formed where the reactions take place, benefiting electrolyte transport and the electrochemical redox reactions at the electrolyte/electrode interfaces. The influence of the PEDOT:PSS wrapping effect on the electrochemical performance was further evaluated via GCD measurement, shown in Figure 5b. The specific capacitance of Fe3O4@RGO is 71.68 F g−1, which is comparable with a previous report [10], and can be substantially improved with the introduction of PEDOT:PSS and the subsequent acid treatment. As is shown in Figure 5b, the specific capacitance of Fe3O4@RGO/PEDOT:PSS can reach a high value of 244.7 F g−1 at 1 A g−1. Notably, the shape of the Fe3O4@RGO/PEDOT:PSS curve deviates from the ideal triangular, implying the pseudocapacitive performance is contributed from the Fe3O4 component. These results are highly consistent with the CV analysis. Therefore, we infer that the introduction of PEDOT:PSS and the subsequent acid treatment contribute to the pseudocapacitance storage of Fe3O4, resulting in better SC performance.
The effect of the PEDOT:PSS content on the electrochemical performance was also researched, and the results are given in Figure S4. After PEDOT:PSS addition and the following acid treatment, the PEDOT:PSS molecules wrap intimately around the Fe3O4@RGO framework via strong interactions, generating a continuous and conductive network. Combined with the high electric conductivity of PEDOT:PSS, the porous architecture of Fe3O4@RGO/PEDOT:PSS facilitates ion transport and fast redox reactions, leading to an enhanced capacitance. Nevertheless, when the addition amount exceeds 28.6 wt% (which becomes 34.06 wt% after acid treatment), the presence of excessive PEDOT:PSS may disrupt the interfacial contact between Fe3O4@RGO and the electrolyte, causing a significant reduction in redox activity sites. As a result, the capacitive performance of the flexible film is deteriorated. Specifically, when 71.4 wt% (51.08 wt% was left after acid treatment) Fe3O4@RGO was used, the specific capacity possessed a maximum value. As is exhibited in Figure 5c, further CV curves in a wide range of scan rates (from 10 to 100 mV s−1) indicate a good reversibility of the redox reactions. GCD tests of this sample with different current densities were also performed (Figure 5d). They show high specific capacitance of 244.7, 205.0, 181.0, and 146.0 F g−1 at 1, 2.5, 5, and 10 A g−1, respectively, which is in accordance with the CV curves. From the relative contents of each component (Figure S5), the quantitative contribution of the Fe3O4@RGO and PEDOT:PSS can be determined. In our previous report, an acid-treated PEDOT:PSS film achieved a specific capacity of around 43.2 F g−1 at 1 A g−1 [33]. According to this value, the specific capacitance of Fe3O4@RGO based on its own weight is around 379.0 F g−1. These results are comparable or even better when compared with other Fe3O4-based electrodes [18,19,20,21,22,39,40,41]. The excellent electrochemical performance of the porous Fe3O4@RGO/PEDOT:PSS could be ascribed mainly to the acid treatment. First, the nonconductive PSS was effectively removed via HClO4 and HI immersion. Secondly, for the GO component, some functional groups like oxygen-containing groups would favor rapid ion transfer between the film surface and interior [42]. However, the existence of these functional groups is not beneficial for electron transport, resulting in a low electric conductivity. Via HI treatment, the balance between the oxygen-containing functional groups and conductivity could be properly adjusted for enhancement of the electrochemical properties. Last but not least, the interconnected conductive framework was successfully fabricated via acid treatment (Figure 2e–g), ensuring the fast ion and electron transport.
In addition, EIS measurement was conducted to investigate the ion diffusion and electron transfer resistance of the prepared electrodes. As can be seen from Figure 5e, the Nyquist curve mainly consists of two parts: the high-frequency Nyquist curve is composed of a semicircle, while the low-frequency curve is made up of a nearly straight line. Generally, the semicircle diameter at high frequency is related to the charge transfer resistance (Rct). To be specific, the Fe3O4@RGO possesses a charge transfer resistance of around 50 Ω. It is worthy to note that, after being wrapped with PEDOT:PSS, the electrode shows a marked decline in Rct value. Furthermore, the slope of the Fe3O4@RGO/PEDOT:PSS electrode at lower frequencies is very similar to that of the Fe3O4@RGO, indicating that good diffusive behavior of the electrolyte ions is maintained after PEDOT:PSS addition. All results of these EIS analyses are consistent with those collected from the CV and GCD studies, illustrating that the PEDOT:PSS wrapping indeed facilitates efficient ion transport and consequently enhances its capacitance performance. Moreover, a further cyclic stability test was carried out, and the result reveals that even at a very high current density (20 A g−1), the Fe3O4@RGO/PEDOT:PSS integrated electrode film still retains nearly 70% of its original capacitance when it is charged and discharged for 800 cycles (see Figure 5f).

3.4. Thermoelectric Performance of Fe3O4@RGO/PEDOT:PSS Free-Standing Films

The excellent performance of the hybrid Fe3O4@RGO/PEDOT:PSS electrodes can be attributed to their unique architecture, which may endow this free-standing film with wider application areas. Herein, we also investigate the thermoelectric properties of this Fe3O4@RGO/PEDOT:PSS integrated film. For comparison, a Fe3O4@RGO sample with a similar thickness (Figure S6) was also prepared. Figure 6 shows the variation in the TE parameters of the composite films as a function of absolute temperature. As can be seen, the σ of the Fe3O4@RGO sample reveals a value of 150 S cm−1 and keeps almost constant in the tested temperature interval (Figure 6a). With the incorporation of PEDOT:PSS and the following acid treatment, the obtained Fe3O4@RGO/PEDOT:PSS integrated film shows a much higher σ value (507.56 S cm−1). With increasing temperature, the σ value just shows a slight decrease, and this variation trend is very similar to those reported in the literature [43,44]. Namely, the electric conductivity is not sensitive to the change in temperature. As shown in Figure 6b, all the composite films exhibit positive Seebeck coefficients, illustrating that the predominant charge carriers are holes in the above obtained samples. The Fe3O4@RGO film shows a Seebeck coefficient of 18.96 μV K−1 at room temperature. When the test temperature increases, the S demonstrates a slight downward trend and decreases to 15.07 μV K−1 at 380 K. In contrast, the free-standing Fe3O4@RGO/PEDOT:PSS film possesses an S value of only 13.29 μV K−1 at room temperature. Upon increasing the testing temperature from 300 K to 380 K, the Seebeck coefficient presents a visible increasing trend, and achieves an S value of 15.07 μV K−1 (the same result with the Fe3O4@RGO sample) at 380 K. Taking into account the big variation in electric conductivity, the Fe3O4@RGO/PEDOT:PSS film exhibits a much higher power factor when compared with the Fe3O4@RGO sample. To be specific (shown in Figure 6c), a maximum PF value of 11.06 μW⋅m−1⋅K−2 is achieved at 380 K, which is much higher than that of the Fe3O4@RGO, by nearly four times.

4. Conclusions

In summary, we have demonstrated the creation of highly conductive Fe3O4@RGO/PEDOT:PSS porous films via a facile but efficient method. Benefiting from the supporting effect of the GO framework and the highly conductive networks arising from PEDOT:PSS wrapping and subsequent acid treatment, the ternary Fe3O4@RGO/PEDOT:PSS composites exhibit excellent electrochemical properties: a high specific capacitance of 244.7 F g−1 can be achieved at a current density of 1 A g−1; meanwhile, a high level of cycling stability of ~70% is maintained after 800 cycles. The superior electrochemical performance of Fe3O4@RGO/PEDOT:PSS hybrid films greatly benefits from their unique structures. Notably, the free-standing flexible hybrid films also show relatively high thermoelectric properties, and a high power factor of 11.06 μW⋅m−1⋅K−2 is reached at 380 K. The novel method proposed in this paper paves an effective way to fabricate and design highly conductive and porous PEDOT:PSS-based composites for electrochemical energy storage and thermoelectric applications.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/polym15163453/s1, Figure S1: Raman spectra of the PEDOT:PSS film after acid treatment via HClO4; Figure S2: XPS survey spectrum of the porous Fe3O4@RGO/PEDOT:PSS; Figure S3: PSS/PEDOT ratio calculated for Fe3O4@GO/PEDOT:PSS samples before acid treatment and Fe3O4@RGO/PEDOT:PSS after acid treatment.; Figure S4: CV curves of Fe3O4@RGO/PEDOT:PSS with different PEDOT:PSS content at a scan rate of 50 mV s−1: a, the content of PEDOT:PSS is 16.7 wt%; b, the content of PEDOT:PSS is 28.6 wt%; c, the content of PEDOT:PSS is 50.0 wt%; Figure S5: TG curves for Fe3O4@RGO/PEDOT:PSS flexible film; Figure S6: The thickness of Fe3O4@RGO and Fe3O4@RGO/PEDOT:PSS films.

Author Contributions

Conceptualization, F.L. (Fuwei Liu) and L.G.; methodology, investigation, and validation, L.G., Q.W., Z.C., J.D. and F.L. (Fuqun Li); writing—original draft preparation, L.G.; writing—review and editing, F.L. (Fuwei Liu); discussion of experiments, H.L., R.L., M.W., J.L. and L.W.; supervision, F.L. (Fuwei Liu). 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 (Grant No. U2004174) and the Nanhu Scholars Program for Young Scholars of XYNU.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

All data are contained within the article, and Supplementary Materials are available upon request from the authors.

Acknowledgments

The project was supported by the National Natural Science Foundation of China (Grant No. U2004174) and the Nanhu Scholars Program for Young Scholars of XYNU. We thank the Analysis Testing Center of Xinyang Normal University for the help with characterization analyses. We also appreciate the support and help from Shichao Wang.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, X.; Xiao, F.; Lei, Y.; Lu, H.; Zhang, J.; Yan, M.; Xu, J. A novel approach for synthesis of expanded graphite and its enhanced lithium storage properties. J. Energy Chem. 2021, 59, 292–298. [Google Scholar] [CrossRef]
  2. Zhou, Y.; Qi, H.; Yang, J.; Bo, Z.; Huang, F.; Islam, M.S.; Lu, X.; Dai, L.; Amal, R.; Wang, C.H.; et al. Two-birds-one-stone: Multifunctional supercapacitors beyond traditional energy storage. Energy Environ. Sci. 2021, 14, 1854–1896. [Google Scholar] [CrossRef]
  3. Simon, P.; Gogotsi, Y.; Dunn, B. Where do batteries end and supercapacitors begin? Science 2014, 343, 1210–1211. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, M.; Lu, Y.; Yang, Y.; Zhang, M.; Yue, Z.; Zhang, N.; Peng, T.; Liu, X.; Luo, Y. A vanadium-based oxide-nitride heterostructure as a multifunctional sulfur host for advanced Li-S batteries. Nanoscale 2021, 13, 13085–13094. [Google Scholar] [CrossRef]
  5. Xie, W.; Wang, Q.; Wang, W.; Xu, Z.; Li, N.; Li, M.; Jia, L.; Zhu, W.; Cao, Z.; Xu, J. Carbon framework microbelt supporting SnO(x) as a high performance electrode for lithium ion batteries. Nanotechnology 2019, 30, 325405. [Google Scholar] [CrossRef]
  6. Peng, T.; Guo, Y.; Zhang, Y.; Wang, Y.; Zhang, D.; Yang, Y.; Lu, Y.; Liu, X.; Chu, P.K.; Luo, Y. Uniform cobalt nanoparticles-decorated biscuit-like VN nanosheets by in situ segregation for Li-ion batteries and oxygen evolution reaction. Appl. Surf. Sci. 2021, 536, 147982. [Google Scholar] [CrossRef]
  7. Jiang, Y.; Ou, J.; Luo, Z.; Chen, Y.; Wu, Z.; Wu, H.; Fu, X.; Luo, S.; Huang, Y. High Capacitive Antimonene/CNT/PANI Free-Standing Electrodes for Flexible Supercapacitor Engaged with Self-Healing Function. Small 2022, 18, 2201377. [Google Scholar] [CrossRef]
  8. Zhu, Q.; Zhao, D.; Cheng, M.; Zhou, J.; Owusu, K.A.; Mai, L.; Yu, Y. A New View of Supercapacitors: Integrated Supercapacitors. Adv. Energy Mater. 2019, 9, 1901081. [Google Scholar] [CrossRef]
  9. Bertana, V.; Scordo, G.; Camilli, E.; Ge, L.; Zaccagnini, P.; Lamberti, A.; Marasso, S.L.; Scaltrito, L. 3D Printed Supercapacitor Exploiting PEDOT-Based Resin and Polymer Gel Electrolyte. Polymers 2023, 15, 2657. [Google Scholar] [CrossRef]
  10. Chang, J.; Adhikari, S.; Lee, T.H.; Li, B.; Yao, F.; Pham, D.T.; Le, V.T.; Lee, Y.H. Leaf vein-inspired nanochanneled graphene film for highly efficient micro-supercapacitors. Adv. Energy Mater. 2015, 5, 1500003. [Google Scholar] [CrossRef]
  11. Huang, Y.; Zhu, M.; Meng, W.; Fu, Y.; Wang, Z.; Huang, Y.; Pei, Z.; Zhi, C. Robust reduced graphene oxide paper fabricated with a household non-stick frying pan: A large-area freestanding flexible substrate for supercapacitors. RSC Adv. 2015, 5, 33981–33989. [Google Scholar] [CrossRef]
  12. Li, X.; Wu, H.; Guan, C.; Elshahawy, A.M.; Dong, Y.; Pennycook, S.J.; Wang, J. (Ni,Co)Se2/NiCo-LDH Core/Shell Structural Electrode with the Cactus-Like (Ni,Co)Se2Core for Asymmetric Supercapacitors. Small 2018, 15, 1803895. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, H.; Qiu, F.; Lu, C.; Zhu, J.; Ke, C.; Han, S.; Zhuang, X. A Terpyridine-Fe(2+)-Based Coordination Polymer Film for On-Chip Micro-Supercapacitor with AC Line-Filtering Performance. Polymers 2021, 13, 1002. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, Z.-S.; Wang, D.-W.; Ren, W.; Zhao, J.; Zhou, G.; Li, F.; Cheng, H.-M. Anchoring Hydrous RuO2 on Graphene Sheets for High-Performance Electrochemical Capacitors. Adv. Funct. Mater. 2010, 20, 3595–3602. [Google Scholar] [CrossRef]
  15. Yu, G.; Hu, L.; Liu, N.; Wang, H.; Vosgueritchian, M.; Yang, Y.; Cui, Y.; Bao, Z. Enhancing the supercapacitor performance of graphene/MnO2 nanostructured electrodes by conductive wrapping. Nano Lett. 2011, 11, 4438–4442. [Google Scholar] [CrossRef]
  16. Zhang, R.; Wang, X.; Cai, S.; Tao, K.; Xu, Y. A Solid-State Wire-Shaped Supercapacitor Based on Nylon/Ag/Polypyrrole and Nylon/Ag/MnO2 Electrodes. Polymers 2023, 15, 1627. [Google Scholar] [CrossRef]
  17. Fang, L.; Lan, M.; Liu, B.; Cao, Y. Synthesis and Electrochemical Performance of Flower-like Zn0.76Co0.24S. J. Xinyang Norm. Univ. (Nat. Sci. Ed.) 2022, 35, 615–620. [Google Scholar]
  18. Liu, D.; Wang, X.; Wang, X.; Tian, W.; Liu, J.; Zhi, C.; He, D.; Bando, Y.; Golberg, D. Ultrathin nanoporous Fe3O4–carbon nanosheets with enhanced supercapacitor performance. J. Mater. Chem. A 2013, 1, 1952–1955. [Google Scholar] [CrossRef]
  19. Khoh, W.-H.; Hong, J.-D. Layer-by-layer self-assembly of ultrathin multilayer films composed of magnetite/reduced graphene oxide bilayers for supercapacitor application. Colloids Surf. A 2013, 436, 104–112. [Google Scholar] [CrossRef]
  20. Li, J.; Chen, Y.; Wu, Q.; Xu, H. Synthesis and electrochemical properties of Fe3O4/MnO2/RGOs sandwich-like nano-superstructures. J. Alloy. Compd. 2017, 693, 373–380. [Google Scholar] [CrossRef]
  21. Guan, D.; Gao, Z.; Yang, W.; Wang, J.; Yuan, Y.; Wang, B.; Zhang, M.; Liu, L. Hydrothermal synthesis of carbon nanotube/cubic Fe3O4 nanocomposite for enhanced performance supercapacitor electrode material. Mater. Sci. Eng. B 2013, 178, 736–743. [Google Scholar] [CrossRef]
  22. Ramanathan, S.; SasiKumar, M.; Radhika, N.; Obadiah, A.; Durairaj, A.; Helen Swetha, G.; Santhoshkumar, P.; Sharmila Lydia, I.; Vasanthkumar, S. Musa paradisiaca reduced graphene oxide (BRGO) /MWCNT-Fe3O4 nanocomposite for supercapacitor and photocatalytic applications. Mater. Today Proc. 2021, 47, 843–852. [Google Scholar] [CrossRef]
  23. Fang, L.; Qiu, Y.; Zhai, T.; Wang, F.; Zhou, H. Synthesis of Ni(OH)2-VS2 Nanocomposite and Their Application in Supercapacitors. J. Xinyang Norm. Univ. (Nat. Sci. Ed.) 2017, 30, 109–113. [Google Scholar]
  24. Mondal, S.; Thakur, S.; Maiti, S.; Bhattacharjee, S.; Chattopadhyay, K.K. Self-Charging Piezo-Supercapacitor: One-Step Mechanical Energy Conversion and Storage. ACS Appl. Mater. Interfaces 2023, 15, 8446–8461. [Google Scholar] [CrossRef] [PubMed]
  25. Ouyang, J.; Chu, C.W.; Chen, F.C.; Xu, Q.; Yang, Y. High-Conductivity Poly(3,4-ethylenedioxythiophene):Poly(styrene sulfonate) Film and Its Application in Polymer Optoelectronic Devices. Adv. Funct. Mater. 2005, 15, 203–208. [Google Scholar] [CrossRef]
  26. Ouyang, J.; Xu, Q.; Chu, C.-W.; Yang, Y.; Li, G.; Shinar, J. On the mechanism of conductivity enhancement in poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) film through solvent treatment. Polymer 2004, 45, 8443–8450. [Google Scholar] [CrossRef]
  27. Badre, C.; Marquant, L.; Alsayed, A.M.; Hough, L.A. Highly Conductive Poly(3,4-ethylenedioxythiophene):Poly (styrenesulfonate) Films Using 1-Ethyl-3-methylimidazolium Tetracyanoborate Ionic Liquid. Adv. Funct. Mater. 2012, 22, 2723–2727. [Google Scholar] [CrossRef]
  28. Fan, Z.; Du, D.; Yu, Z.; Li, P.; Xia, Y.; Ouyang, J. Significant Enhancement in the Thermoelectric Properties of PEDOT:PSS Films through a Treatment with Organic Solutions of Inorganic Salts. ACS Appl. Mater. Int. 2016, 8, 23204–23211. [Google Scholar] [CrossRef]
  29. Kim, G.H.; Shao, L.; Zhang, K.; Pipe, K.P. Engineered doping of organic semiconductors for enhanced thermoelectric efficiency. Nat. Mater. 2013, 12, 719–723. [Google Scholar] [CrossRef]
  30. Lee, S.H.; Park, H.; Kim, S.; Son, W.; Cheong, I.W.; Kim, J.H. Transparent and flexible organic semiconductor nanofilms with enhanced thermoelectric efficiency. J. Mater. Chem. A 2014, 2, 7288–7294. [Google Scholar] [CrossRef]
  31. Takano, T.; Masunaga, H.; Fujiwara, A.; Okuzaki, H.; Sasaki, T. PEDOT Nanocrystal in Highly Conductive PEDOT:PSS Polymer Films. Macromolecules 2012, 45, 3859–3865. [Google Scholar] [CrossRef]
  32. Crispin, X.; Jakobsson, F.L.E.; Crispin, A.; Grim, P.C.M.; Andersson, P.; Volodin, A.; van Haesendonck, C.; Van der Auweraer, M.; Salaneck, W.R.; Berggren, M. The Origin of the High Conductivity of Poly(3,4-ethylenedioxythiophene)-Poly(styrenesulfonate) (PEDOT-PSS) Plastic Electrodes. Chem. Mater. 2006, 18, 4354–4360. [Google Scholar] [CrossRef]
  33. Liu, F.; Xie, L.; Wang, L.; Chen, W.; Wei, W.; Chen, X.; Luo, S.; Dong, L.; Dai, Q.; Huang, Y.; et al. Hierarchical Porous RGO/PEDOT/PANI Hybrid for Planar/Linear Supercapacitor with Outstanding Flexibility and Stability. Nano Micro Lett. 2020, 12, 17. [Google Scholar] [CrossRef] [PubMed]
  34. Bießmann, L.; Saxena, N.; Hohn, N.; Hossain, M.A.; Veinot, J.G.C.; Müller-Buschbaum, P. Highly Conducting, Transparent PEDOT:PSS Polymer Electrodes from Post-Treatment with Weak and Strong Acids. Adv. Electron. Mater. 2019, 5, 1800654. [Google Scholar] [CrossRef]
  35. Song, H.; Liu, C.; Xu, J.; Jiang, Q.; Shi, H. Fabrication of a layered nanostructure PEDOT:PSS/SWCNTs composite and its thermoelectric performance. RSC Adv. 2013, 3, 22065–22071. [Google Scholar] [CrossRef]
  36. Ely, F.; Matsumoto, A.; Zoetebier, B.; Peressinotto, V.S.; Hirata, M.K.; de Sousa, D.A.; Maciel, R. Handheld and automated ultrasonic spray deposition of conductive PEDOT:PSS films and their application in AC EL devices. Org. Electron. 2014, 15, 1062–1070. [Google Scholar] [CrossRef]
  37. Wang, L.; Liu, F.; Pal, A.; Ning, Y.; Wang, Z.; Zhao, B.; Bradley, R.; Wu, W. Ultra-small Fe3O4 nanoparticles encapsulated in hollow porous carbon nanocapsules for high performance supercapacitors. Carbon 2021, 179, 327–336. [Google Scholar] [CrossRef]
  38. Li, X.; Xu, Y.; Wu, H.; Qian, X.; Chen, L.; Dan, Y.; Yu, Q. Porous Fe3O4/C nanoaggregates by the carbon polyhedrons as templates derived from metal organic framework as battery-type materials for supercapacitors. Electrochim. Acta 2020, 337, 135818. [Google Scholar] [CrossRef]
  39. Du, X.; Wang, C.; Chen, M.; Jiao, Y.; Wang, J. Electrochemical Performances of Nanoparticle Fe3O4/Activated Carbon Supercapacitor Using KOH Electrolyte Solution. J. Phys. Chem. C 2009, 113, 2643–2646. [Google Scholar] [CrossRef]
  40. Wang, X.; Jiang, D.; Jing, C.; Liu, X.; Li, K.; Yu, M.; Qi, S.; Zhang, Y. Biotemplate Synthesis of Fe3O4/Polyaniline for Supercapacitor. J. Energy Storage 2020, 30, 101554. [Google Scholar] [CrossRef]
  41. Li, J.; Lu, W.; Yan, Y.; Chou, T.-W. High performance solid-state flexible supercapacitor based on Fe3O4/carbon nanotube/polyaniline ternary films. J. Mater. Chem. A 2017, 5, 11271–11277. [Google Scholar] [CrossRef]
  42. Qiu, C.; Jiang, L.; Gao, Y.; Sheng, L. Effects of oxygen-containing functional groups on carbon materials in supercapacitors: A review. Mater. Des. 2023, 230, 111952. [Google Scholar] [CrossRef]
  43. Du, Y.; Shi, Y.; Meng, Q.; Shen, S.Z. Preparation and thermoelectric properties of flexible SWCNT/PEDOT:PSS composite film. Synth. Met. 2020, 261, 116318. [Google Scholar] [CrossRef]
  44. Wei, S.; Huang, X.; Deng, L.; Yan, Z.-C.; Chen, G. Facile preparations of layer-like and honeycomb-like films of poly(3,4-ethylenedioxythiophene)/carbon nanotube composites for thermoelectric application. Compos. Sci. Technol. 2021, 208, 108759. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the fabrication process of the free-standing Fe3O4@RGO/PEDOT:PSS film.
Figure 1. Schematic illustration of the fabrication process of the free-standing Fe3O4@RGO/PEDOT:PSS film.
Polymers 15 03453 g001
Figure 2. SEM images of (a,b) the Fe3O4@GO composites; (c,d) Fe3O4@GO/PEDOT:PSS hybrid films as prepared; (eg) Fe3O4@RGO/PEDOT:PSS flexible free-standing films after acid treatment; (h) the corresponding EDS mapping image of (g) Fe3O4@RGO/PEDOT:PSS.
Figure 2. SEM images of (a,b) the Fe3O4@GO composites; (c,d) Fe3O4@GO/PEDOT:PSS hybrid films as prepared; (eg) Fe3O4@RGO/PEDOT:PSS flexible free-standing films after acid treatment; (h) the corresponding EDS mapping image of (g) Fe3O4@RGO/PEDOT:PSS.
Polymers 15 03453 g002
Figure 3. Raman spectra of the obtained hybrid films: (a) Fe3O4@GO/PEDOT:PSS films before acid treatment and (b) Fe3O4@RGO/PEDOT:PSS films after acid treatment.
Figure 3. Raman spectra of the obtained hybrid films: (a) Fe3O4@GO/PEDOT:PSS films before acid treatment and (b) Fe3O4@RGO/PEDOT:PSS films after acid treatment.
Polymers 15 03453 g003
Figure 4. The Fe2p and S2p high-resolution XPS spectra of composite films including fits for the components: (a) Fe2p of Fe3O4@RGO/PEDOT:PSS after acid treatment; (b) S2p of Fe3O4@GO/PEDOT:PSS before acid treatment; (c) S2p of Fe3O4@RGO/PEDOT:PSS after acid treatment.
Figure 4. The Fe2p and S2p high-resolution XPS spectra of composite films including fits for the components: (a) Fe2p of Fe3O4@RGO/PEDOT:PSS after acid treatment; (b) S2p of Fe3O4@GO/PEDOT:PSS before acid treatment; (c) S2p of Fe3O4@RGO/PEDOT:PSS after acid treatment.
Polymers 15 03453 g004
Figure 5. Electrochemical properties of PEDOT:PSS, Fe3O4@RGO, and Fe3O4@RGO/PEDOT:PSS electrodes: (a) CV curves of samples at a scan rate of 50 mV s−1; (b) GCD curves of samples at a current density of 1 A g−1; (c) CV curves of Fe3O4@RGO/PEDOT:PSS at different scan rates; (d) CD curves of Fe3O4@RGO/PEDOT:PSS at different current densities; (e) Nyquist plots of Fe3O4@RGO and Fe3O4@RGO/PEDOT:PSS composite samples; and (f) cycle stabilities of Fe3O4@RGO/PEDOT:PSS during the long-term charging/discharging process at a current density of 20 A g−1.
Figure 5. Electrochemical properties of PEDOT:PSS, Fe3O4@RGO, and Fe3O4@RGO/PEDOT:PSS electrodes: (a) CV curves of samples at a scan rate of 50 mV s−1; (b) GCD curves of samples at a current density of 1 A g−1; (c) CV curves of Fe3O4@RGO/PEDOT:PSS at different scan rates; (d) CD curves of Fe3O4@RGO/PEDOT:PSS at different current densities; (e) Nyquist plots of Fe3O4@RGO and Fe3O4@RGO/PEDOT:PSS composite samples; and (f) cycle stabilities of Fe3O4@RGO/PEDOT:PSS during the long-term charging/discharging process at a current density of 20 A g−1.
Polymers 15 03453 g005
Figure 6. The electrical conductivity (a), Seebeck coefficient (b), and power factor (c) of Fe3O4@RGO and Fe3O4@RGO/PEDOT:PSS composite films from 300 to 380 K.
Figure 6. The electrical conductivity (a), Seebeck coefficient (b), and power factor (c) of Fe3O4@RGO and Fe3O4@RGO/PEDOT:PSS composite films from 300 to 380 K.
Polymers 15 03453 g006
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

Gao, L.; Liu, F.; Wei, Q.; Cai, Z.; Duan, J.; Li, F.; Li, H.; Lv, R.; Wang, M.; Li, J.; et al. Fabrication of Highly Conductive Porous Fe3O4@RGO/PEDOT:PSS Composite Films via Acid Post-Treatment and Their Applications as Electrochemical Supercapacitor and Thermoelectric Material. Polymers 2023, 15, 3453. https://0-doi-org.brum.beds.ac.uk/10.3390/polym15163453

AMA Style

Gao L, Liu F, Wei Q, Cai Z, Duan J, Li F, Li H, Lv R, Wang M, Li J, et al. Fabrication of Highly Conductive Porous Fe3O4@RGO/PEDOT:PSS Composite Films via Acid Post-Treatment and Their Applications as Electrochemical Supercapacitor and Thermoelectric Material. Polymers. 2023; 15(16):3453. https://0-doi-org.brum.beds.ac.uk/10.3390/polym15163453

Chicago/Turabian Style

Gao, Luyao, Fuwei Liu, Qinru Wei, Zhiwei Cai, Jiajia Duan, Fuqun Li, Huiying Li, Ruotong Lv, Mengke Wang, Jingxian Li, and et al. 2023. "Fabrication of Highly Conductive Porous Fe3O4@RGO/PEDOT:PSS Composite Films via Acid Post-Treatment and Their Applications as Electrochemical Supercapacitor and Thermoelectric Material" Polymers 15, no. 16: 3453. https://0-doi-org.brum.beds.ac.uk/10.3390/polym15163453

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop