Next Article in Journal
Theoretical Prediction of Mechanical Strength and Desalination Performance of One-Atom-Thick Hydrocarbon Polymer in Pressure-Driven Separation
Previous Article in Journal
Effect of Graphene Oxide Coating on Natural Fiber Composite for Multilayered Ballistic Armor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Redox-Active Gel Electrolyte Combined with Branched Polyaniline Nanofibers Doped with Ferrous Ions for Ultra-High-Performance Flexible Supercapacitors

1
Institute of Advanced Wear & Corrosion Resistance and Functional Materials, Jinan University, Guangzhou 510632, China
2
School of Aerospace, Mechanical and Mechatronic Engineering J07, University of Sydney, Sydney, NSW 2006, Australia
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(8), 1357; https://doi.org/10.3390/polym11081357
Submission received: 21 July 2019 / Revised: 4 August 2019 / Accepted: 6 August 2019 / Published: 16 August 2019

Abstract

:
In this work, the effects of utilizing an Fe2+/Fe3+ redox-active electrolyte and Fe2+-doped polyaniline (PANI) electrode material on the performance of an assembled supercapacitor (SC) were studied. The concentration of the redox couple additive in the electrolyte of the SC was optimized to be 0.5 M. With the optimized concentration of 0.4 M Fe2+, the doped PANI branched nanofibers electropolymerized onto titanium mesh were much thinner, cleaner, and more branched than normal PANI. A specific capacitance (Cs) of 8468 F g−1 for the 0.4 M Fe2+/PANI electrode in the 1 M H2SO4 + 0.5 M Fe2+/Fe3+ gel electrolyte and an energy density of 218.1 Wh kg−1 at a power density of 1854.4 W kg−1 for the resultant SC were achieved, which were much higher than those of the conventional PANI electrode tested in a normal H2SO4 electrolyte (404 F g−1 and 24.9 Wh kg−1). These results are among the highest reported for PANI-based SCs in the literature so far and demonstrate the potential effectiveness of this strategy to improve the electrochemical performance of flexible SCs by modifying both the electrode and electrolyte.

Graphical Abstract

1. Introduction

Supercapacitors (SCs) are attractive because of their high power, long periodicity, and environmental friendliness [1,2]. The electrode materials used in SCs include carbon-based materials, conductive polymers, and transition metal oxides/hydroxide [2,3,4,5,6]. Polyaniline (PANI) is not only easy to prepare but also has good conductivity, excellent electrochemical activity, fast Faraday reactivity, facile fabrication, and so forth. [7]. Nonetheless, long-term charge and discharge easily cause volume expansion and contraction, which leads to a decrease in mechanical and electrochemical performance, thereby limiting its application in SCs [8]. So far, some strategies have been developed to promote its electrochemical behaviors, including combinations with other materials, such as PANI/carbon-based materials [9,10], PANI/metallic oxide [11,12], and PANI/metallic oxides/carbon materials [13], or the doping of PANI with certain materials/ions, including protonic acids/anions (HCl [14], H2SO4 [15], HClO4 [16], and H3PO4 [17]) and metallic cations (LiCl [18], Ni2+ [19], Co2+ [20], Fe3+ [21], Zn2+ [22,23], Cu2+ [24], and Mn2+ [25]).
The effective doping of electrodes could produce electroactive materials with certain physicochemical structures and provide enough and suitable active sites which would favor the electrochemical process in devices [26,27,28]. A common, well-investigated method is the anion/protonic acid doping of PANI. PANI nanofibers, nanorods, nanoparticles, and pyramid-like PANI have been deposited uniformly on carbonized kapok surfaces by using HCl, H2SO4, HClO4, HPO4, and PTSA as dopants, and their specific capacitances (Cs) were measured to be 241, 245, 580, 233, and 289 F g−1 at 1.0 A g−1, respectively [17]. Recently, transition metal ions and other cations have been used as dopants for PANI due to their unique electronic exchange properties [20]. Doping cations in PANI can not only be used as redox catalysts and corrosion inhibitors [29,30] but also to enhance the energy storage of the polymer in certain cases. PANI/Co2+, PANI/Ni2+, PANI/Cu2+, PANI/Zn2+, and PANI/Fe3+ films have been prepared by electrodeposition and they displayed a higher specific capacitance than that of pure PANI [19,20,21,22,23]. It was found that PANI/Mn2+ film has more reaction active centers than pure PANI film due to the existence of metal ion dopants [21]. Moreover, compared with pure PANI, the addition of Co2+ to PANI is believed to widen the localization of the charge on the macromolecular skeleton due to the transformation of the quinoid ring structures to benzenoid rings by protonation [20,31].
To improve the performance of SCs (e.g., their energy density), two approaches have been adopted. One is developing new electrode materials and their composites [32,33]. Another is achieved by adjusting the configuration of the devices or the design and utilization of specific electrolytes [34]. Generally, novel electrode materials and their assembled symmetric/asymmetric SCs with high electrochemical performance can be designed and fabricated, including the above-described doped PANI electrode materials. For the utilization of specific electrolytes, besides organic electrolytes (such as ionic liquids) with a widening working potential window, redox-active electrolytes were recently developed by adding redox additives to electrolytes [35]. These redox media can afford additional capacitive contributions through their reversible electrochemical reactions and can improve the electrochemical performance of assembled symmetric/asymmetric SCs. Electrolytes with a variety of redox-active additions have been studied, such as H2SO4 + FeBr3 and KCl + VOSO4 [36,37], H2SO4 + KI, H2SO4 + VO2+/VO2+ and KOH + K3Fe(CN)6 [38,39,40], H2SO4 + hydroquinone [41], and H2SO4 + Fe2+/Fe3+ [42,43,44]. Among these, active electrolytes containing Fe3+/Fe2+ redox couples are easily available and cost effective, and they have been shown to be promising electrolytes in SCs, where PANI has been found to exhibit an enhanced Cs of 1062 F g−1 at a current density of 2 A g−1 [42].
It is expected that advanced SCs with enhanced capacitive performance will be developed by combining the advances in both electrodes and the electrolyte. Although various polyaniline material electrodes and redox-active electrolytes have been used in capacitors, little information on the capacitive performance of Fe2+-doped PANI in an Fe3+/Fe2+ active electrolyte is available so far. In this work, flexible symmetric SCs were designed and fabricated by both PANI electrode materials doped with Fe2+ and utilizing a redox-active electrolyte containing an Fe2+/Fe3+ additive. The effects of the presence of Fe2+ on the physicochemical structure and performance of the modified PANI electrode materials were investigated and the electrode materials were optimized. Moreover, the influence of incorporating an Fe2+/Fe3+ redox-active additive into the electrolyte combined with the optimized electrode materials on the performance of SCs was studied in detail.

2. Experimental

2.1. Preparation of PANI/Fe2+ and PANI Materials

PANI materials were electrodeposited in a three-electrode system on a CHI 760e electrochemical work station (CH Instruments, Inc., Austin, TX, USA). A saturated calomel electrode (SCE), a platinum net, and titanium mesh with a working area of 1 cm2 were used as the reference electrode, the counter electrode, and the working electrode, respectively. Before depositing the polymer, the titanium mesh was washed in an ultrasonic bath of ethanol for several minutes and finally air-dried at 70 °C. The electropolymerization of PANI was prepared by conducting a cyclic voltammetry (CV) test in a mixture of 0.5 M HCl and 0.2 M aniline with 0.2, 0.4, and 0.8 M FeCl2 in solution at a scanning rate of 20 mV s−1, and the potential window ranged from 0 to 0.9 V. After electrodeposition, the titanium mesh coated with polymers was immersed in distilled water to remove the soluble monomers or oligomers and finally air-dried at 70 °C. The mass of active material deposited on the titanium mesh (around 1 mg) was controlled by adjusting the CV cycles and measured on a precise analytical balance. Also, pure PANI was electrosynthesized in the absence of FeCl2 according to the abovementioned process.

2.2. Preparation of H2SO4/Fe2+/3+/Poly(Vinyl Alcohol) (PVA) Gel Electrolyte

H2SO4/PVA was obtained by mixing and stirring 5 g of PVA and 50 mL of 1 M H2SO4 together until it became stable and clear at 90 °C [45]. Subsequently, a certain amount of FeSO4·7H2O and Fe2(SO4)3 was added and the resultant mixture was stirred until it dissolved. Finally, the H2SO4/Fe2+/3+/PVA gel electrolyte was cooled down to ambient temperature. H2SO4/PVA gel electrolyte was also prepared without adding the ferric and ferrous salts.

2.3. Characterization

XRD was performed using an X-ray diffractometer (UItima IV) (Rigaku Corporation, The Woodlands, TX, USA) equipped with Co-Kα radiation, with the scanning angle from 3° to 60° at a rate of 10° min−1. FTIR transmission spectra were taken on an FTIR spectrometer (iS50, EQUINOX 55). The morphology of the samples was observed on a field emission scanning electron microscope (FESEM, S3700N).

2.4. Electrochemical Measurements

The CV, galvanostatic charge–discharge (GCD), and electrochemical impedance spectroscopy (EIS) were tested and recorded in either a two- or three-electrode system on a CHI 760e electrochemical analyzer (CH Instruments, Inc., Austin, State of Texas, USA). The electrolytes in the two- and three-electrode systems were 1 M H2SO4 and 1 M H2SO4/Fe2+/3+ redox-active electrolytes with different concentrations of Fe2+/3+ (0.2, 0.5, and 0.8 M), respectively. In the three-electrode system, an SCE, a platinum net, and polymer-coated titanium mesh were used as the reference electrode, counter electrode, and working electrode, respectively. For the solid-state SCs, the devices were assembled using polymer-coated titanium mesh as the electrodes and a H2SO4/Fe2+/3+/PVA gel electrolyte as both separator and electrolyte [45]. EIS measurements were performed in the frequency range of 100 kHz to 10 mHz at an AC amplitude of 5 mV.
The total capacitance of the SC (Ctotal, F g−1), the specific capacitance of a single electrode (Cs, F g−1), the energy density (E, W h Kg−1), and the power density (P, W Kg−1) were calculated according to these equations:
C t o t a l = I × Δ t V × m
C s = 4 × C t o t a l
E = I × Vd t m
P = 3600 E Δ t
where I is the charge/discharge current (A), Δt is the discharge time (s), V is the potential drop in the discharge progress, ∫Vdt is the galvanostatic discharge current area, and m denotes the total mass of the PANI electrodeposited for both electrodes [42,43,46].

3. Results

Fe3+-doped polyaniline has previously been prepared by an electrochemical method [21,47]. However, it was found that the electroplating electrolyte containing Fe3+ and aniline were very unstable and PANI precipitate occurred in tens of minutes, possibly due to the chemical reaction between Fe3+ and the monomer, as FeCl3 was often used as an oxidant reagent for the chemical polymerization of PANI. So here, Fe2+ was utilized in the electrolyte for the electrodeposition of PANI. The structures of the electropolymerized samples were studied by XRD and FTIR, respectively. In the XRD pattern (Figure 1A), three evident peaks were observed at 2θ = 7.1°, 20.5°, and 24.6°. The peak at 2θ = 7.1° confirmed the short-range ordered configuration of the polymer chains [42] and the others indicated the periodic structures parallel and vertical to the polymer chain in PANI, implying lower crystallinity and a conductive emeraldine salt structure [20,23]. Compared with PANI, the 0.4 M Fe2+/PANI electrode only showed two evident broad peaks at 2θ = 7.1° and 24.6°, and the weaker peaks indicated its lower crystallinity compared with PANI [47].
The characteristic peaks of PANI in Figure 1B were observed as follows: two bands located at 1560.18 and 1482.28 cm−1 were assigned to the stretching vibration of N=Q=N and N–B–N, where Q and B represent the quinoid and the benzenoid units, respectively. The bands located at 1293.38 and 1240.08 cm−1 were attributed to C–N and C=N stretching modes, respectively, while the band at 1119.74 cm−1 was related to the in-plane bending vibrations and the band at 797.66 cm−1 was due to the C–H bonds’ in-plane bending vibration in the 1,4-disubstituted aromatic ring [20,31,42,48]. It was noted that the peaks at 1560.18, 1482.28, and 1119.74 cm−1 displayed obvious red shifts after the modification with Fe2+ and appeared at 1559.79, 1476.99, and 1109.42 cm−1, respectively. The red shift implied the conversion of a quinone-like ring to a benzene ring by a proton-induced spin mismatch mechanism [49] and the decreasing charge delocalization in the main chain of PANI backbone, and the comprehensive effect of the protonation and pseudoprotonation processes had a decisive influence on the final chemical state of the PANI [31,50]. The XRD and FTIR spectrum indicated that PANI and Fe2+-modified PANI having an emeraldine chemical structure were prepared successfully and the PANI prepared in the presence of Fe2+ displayed lower crystallinity and charge delocalization on the PANI backbone.
The morphology of PANI and the other electrode materials were characterized by FESEM, as shown in Figure 2. All of the products showed a coral-like morphology which was similar to branched PANI nanofibers synthesized by chemical oxidation polymerization and electropolymerization with FeCl3 as an oxidizer, although the diameter of most of the Fe2+/PANI nanofibers here (~70 nm) was much larger than that of branched nanofibers (~40 nm) produced in previous works [51,52]. Obviously, the presence of metal cations during the electropolymerization of PANI affected its microstructure. Compared with the Fe2+/PANI with a clean fiber surface and a fiber diameter of around 70 nm (Figure 2b–d), the PANI obtained in the absence of Fe2+ displayed a much rougher fiber surface and a larger fiber diameter of 150–400 nm (Figure 2a). Interestingly, the 0.4 M Fe2+/PANI fibers seemed to be more branched than both the 0.2 M Fe2+/PANI and 0.8 M Fe2+/PANI nanofibers, which was beneficial for their higher electrochemical activity. According to their EDS analysis (Table S1), the Fe content in 0.4 M Fe2+/PANI was 3.03%. Based on the results of EDS and FTIR, the Fe2+ in the PANI may have attached to nitrogen atoms, which acted as the contact site in the polymer backbone, leading to the formation of branch-like PANI [53]. Unlike the PANI doped with other transition metallic ions (Ni2+, Cu2+, or Fe3+) [19,21,24], due to the coexistence of Fe2+ and aniline in the electrolyte, an excessive electrochemical reaction of Fe2+ occurred during the electropolymerization of PANI and had some influence on the electrochemical synthesis process, thus affecting the morphology of the products. As the nanostructure of the electrode materials has great positive effects on their electrode properties [54], the resultant ferrous-ion-modified PANI branched nanofibers were expected to exhibit improved electrochemical capacitive performance.
The electrochemical performance of the electropolymerized PANI materials was first tested in the three-electrode system in 1 M H2SO4. In Figure 3A, the discharge times of all three Fe2+/PANI electrodes at 1 A g−1 were longer than pure PANI, indicating they had a higher Cs than that of PANI. The corresponding Cs was 328 F g−1 for 0.2 M Fe2+/PANI, 335 F g−1 for 0.4 M Fe2+/PANI, and 304 F g−1 for 0.8 M Fe2+/PANI, all of which were larger than the 298.8 F g−1 of pure PANI. Obviously, the polymer deposited in the presence of 0.4 M Fe2+ exhibited the highest Cs among these electrodes. Therefore, 0.4 M Fe2+/PANI was used to assemble SCs in the following work.
The PANI or Fe2+/PANI electrodes were used to assemble the symmetric SCs with the 1 M H2SO4/PVA gel electrolyte. The CV curves of 0.4 M Fe2+/PANI and PANI (Figure 3B) were similar and both of them showed a rectangular form with a little deviation. There was a pair of evident redox peaks in the curves between 0.2 and 0.4 V for both pure PANI and 0.4 M Fe2+/PANI, which was attributed to the conversion between the semiconducting (leucoemeraldine form) and conducting (polaronic emeraldine form) states. Another evident pair of redox peaks occurred between −0.2 and −0.1 V for the 0.4 M Fe2+/PANI electrode, which was close to that of a similar work on symmetric PANI SCs with a 0.5 M H2SO4 electrolyte containing an Fe2+/Fe3+ additive [42]. However, its peak difference (~70 mv) here was clearly much less than that which was contributed by the Fe2+/Fe3+ additive in the electrolyte [42], which could have been due to the active transition metal ion doping inside the PANI electrode materials in our work rather than dissolving in the electrolyte. These results indicate the successful doping of Fe2+ in PANI and the corresponding change of its electrochemical behavior. According to the GCD tests at a current density of 5 A g−1 (Figure 3C), the Cs of the 0.4 M Fe2+/PANI electrode was 452 F g−1, which was much higher than that of pure PANI (404 F g−1). Moreover, a better rate capability of the materials could be achieved when the electrodeposition of PANI was carried out in the presence of Fe2+. For pure PANI, 79.21% of Cs could be preserved when the GCD current was increased from 5 to 100 A g1, while 87.79% of Cs still remained for 0.4 M Fe2+/PANI, as shown in Figure 3C. Furthermore, the results of the cyclic stability tests also demonstrated the positive effect of the Fe2+ modification of PANI. As shown in Figure S1, after 1000 cycles at 20 A g1, the Cs of 0.4 M Fe2+/PANI displayed only a slight drop of 0.41%, which was much less than that of PANI (12.5%). The positive effect of Fe2+ doping may originate from the additional redox process and extra active sites relative to the metal ions in PANI, which stabilized and enhanced the performance of PANI.
To improve the performance of the capacitors further, Fe2+/Fe3+ redox couples were introduced into the electrolyte in the SCs. In contrast to the nearly rectangular CV curve in SCs with only the H2SO4 electrolyte (Figure 3B), it can be seen that there was one pair of strong redox peaks with good symmetry in the CV curves of both PANI and 0.4 M Fe2+/PANI electrodes in the redox-active electrolyte (Figure 4A), which clearly shows the contribution of the redox couple in the electrolyte [44]. The peak potential difference (|Epc-Epa|) increased with the increase of the concentration of the redox additive in the electrolyte. The redox peak current first increased as the concentration of Fe2+/Fe3+ increased from 0.2 to 0.5 M and then noticeably decreased after reaching 0.8 M. The results indicate that a proper ion concentration helps to enhance the electrochemical performance of SCs, although an excessive metal ion concentration may reduce water hydration, thus decreasing ion activity and mobility [44]. Also, the redox peak current of 0.4 M Fe2+/PANI increased slightly relative to that of PANI, which could have been due to the limited contribution of the metal ions doped in the electrode material to the whole electrochemical process in the redox-active electrolyte. The discharge time of SCs increased when the concentration of Fe2+/Fe3+ in the electrolyte increased from 0 to 0.5 M and then decreased when the concentration of the redox couple increased to 0.8 M, as shown in the GCD curves obtained at 5 A g1 (Figure 4B). Correspondingly, the calculated Cs of the 0.4 M Fe2+/PANI electrodes was enhanced from 452 F g1 in the H2SO4 electrolyte to 8468 F g1in H2SO4 + 0.5 M Fe2+/Fe3+. The increased Cs could have been due to the additional electrochemical reaction of the Fe2+/Fe3+ additive in the electrolyte on the active sites of the electrodes [42]. By adjusting the concentration of the active additive in the SCs and performing the relative GCD tests, its optimum concentration could be obtained. Also, the Cs values of the 0.4 M Fe2+/PANI electrodes in the electrolyte with the addition of 0.2 and 0.8 M Fe2+/Fe3+ was 3150 and 6160 F g1, respectively, indicating the optimum Fe2+/Fe3+ in the electrolyte was 0.5 M. These Cs values were much larger than others which have been reported in the literature (Table 1), and this could have been due to the advantage of modifying both the electrode material and the electrolyte with the redox-active additive in this work.
To study the rate capability of the SCs with the redox-active additive in the electrolyte, their GCD measurements were performed at different current densities. For both PANI and 0.4 M Fe2+/PANI electrode materials, a sharp decline of Cs with the increasing GCD current density could be observed, as shown in Figure 4C. This can be ascribed to the insufficiency of the redox reaction of Fe2+/Fe3+ at high current densities. The Cs of the 0.4 M Fe2+/PANI in 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte gradually decreased from 8468 to 6120, 3896, 1924, and 972 F g1 as the current increased from 5 to 10, 20, 50, and 100 A g1, respectively, which was calculated according to the GCD curves in Figure 5A. The Cs retention of 11.47% at 100 A g−1 could have been due to the inadequate redox reaction of Fe2+/Fe3+, the concentration polarization, and the reduced accessible area during the charge–discharge process with such a high current density. However, this value was still much larger than PANI in the normal H2SO4 electrolyte. To the best of our knowledge, this is the first report of such a remarkable Cs at high current densities for PANI-based symmetric SCs. Similarly, the CV curves at higher scan rates showed an increased peak potential difference (Figure 5B).
The energy density and power density of energy storage devices are very important for their practical application. As can be seen in Figure 4D, the 0.4 M Fe2+/PANI symmetric SC with 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte processed an energy density as high as 218.1 Wh kg−1 at a power density of 1854.4 W kg−1, which was not only better than that of the other samples, such as PANI in a 1 M H2SO4 electrolyte (24.9 Wh kg−1 at 4437.6 W kg−1), 0.4 M Fe2+/PANI in a 1 M H2SO4 electrolyte (27.35 Wh kg−1 at 4318.4 W kg−1) (Figure 3D), and PANI in a 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte (196 Wh kg−1 at 1834.6 W kg−1) (Figure 4D), but also higher than those of other reports on PANI-based electrode materials tested in a H2SO4 + Fe2+/Fe3+ electrolyte, including PANI (22.1 Wh kg−1 at 774.0 W kg−1) [42] and PANI/CNT (22.9 Wh kg−1 at 700.1 W kg−1) [43].
To further study the effect of both modifying the electrode materials and utilizing the redox-active gel electrolyte on the performance of SCs, the cycle stability of the assembled SCs was tested for 1000 charge–discharge cycles at 20 A g−1. As shown in Figure 6A, after testing for 1000 cycles in a 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte, 0.4 M Fe2+/PANI and PANI electrode materials showed a Cs retention of 66.09% and 50.23%, respectively. Meanwhile, 0.4 M Fe2+/PANI showed better Cs retention (99.59%) than that of PANI (87.5%) in 1 M H2SO4 (Figure S1), indicating that by modifying PANI with Fe2+, the cyclic stability of the electrode can be also enhanced greatly in the redox-active electrolyte. Similar cyclic stability improvements have been reported in the case of PANI films doped with Fe3+ in an H2SO4 electrolyte [21] and could be attributed to the synergistic effect of PANI and metal ion dopants and the ability to better adjust the volume change during the electrochemical process [47], originating from the electrochemical redox reaction of both PANI and transition metal ions in the electrode composite materials, as well as their thinner and highly branched fiber structure (Figure 2c), which provide more redox reaction active sites and room to accommodate the volume expansion and shrinkage of PANI during charge–discharge of SCs. It is also noted that the cyclic stability of PANI and 0.4 M Fe2+/PANI in the H2SO4 + Fe2+/Fe3+ electrolyte was inferior to that in the H2SO4 electrolyte, which could have been due to the extensive redox reaction caused by the redox additive in the electrolyte and the destruction of the polymer chain by the additional electrochemical process of the Fe2+/Fe3+ redox couple on the electrode materials [56]. The electrochemical behavior of all-solid-state flexible SCs were also characterized during the deformation of the SCs. In Figure S2, the GCD curves of 0.4 M Fe2+/PANI SCs with a 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte tested under bending and twisting changed little, indicating the good capacitive stability of the assembled flexible PANI-based SCs with high performance at different deformations.
Nyquist plots of PANI SCs and 0.4 M Fe2+/PANI SCs with different concentrations of Fe2+/Fe3+ in a 1 M H2SO4 electrolyte are shown in Figure 6B. In the high-frequency region, the initial nonzero intercept of the semicircle on the real axis represents the combined series resistance of the electrolyte, electrode, current collector, and contact resistance of the active material/current collector (Rs), and the minimum value of Rs means a greater conducting nature of the composite electrodes in the electrochemical system [20,57]. Another significant feature of the plots is the nearly vertical line to the abscissa axis in the low-frequency region, which corresponds to the resistance of transport and the diffusion process in the electrolyte [58,59,60]. The Rs value of 0.4 M Fe2+/PANI in the 1 M H2SO4 electrolyte was 0.872 Ω, which was much less than that in the 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte (1.218 Ω). Moreover, as Fe2+/Fe3+ in the electrolyte increased from 0.2 to 0.8 M, the Rs value increased significantly from 0.782 to 1.329 Ω, indicating the evident resistance contribution of the high concentration of Fe2+/Fe3+ in the electrolyte to the combined series resistance in the SCs. Also, both the Rs of PANI in the 1 M H2SO4 (Figure 6C) and the 1 M H2SO4 + 0.5 M Fe2+/Fe3+ (Figure 6B) electrolyte was less than that of 0.4 M Fe2+/PANI, which can be attributed to the amorphous configuration of PANI with lower charge delocalization and a thinner polymer nanofiber structure formed in the presence of Fe2+. The pseudocapacitance contributed by both Fe2+/PANI electrodes and the Fe2+/Fe3+ redox additive in the electrolyte resulted in the oblique line deviating from a perfect vertical line in the low-frequency region, the diffusion-controlled doping/undoping of PANI, and the Fe2+/Fe3+ redox reaction leading to the Warburg behaviors as shown in the plots [43,61].
In addition, the Nyquist plots of PANI and 0.4 M Fe2+/PANI with Fe2+/Fe3+ (0 and 0.5 M) in the electrolyte were also recorded before and after 1000 GCD cycles. Interestingly, before the first cycle, the Rs value of 0.4 M Fe2+/PANI in the 1 M H2SO4 electrolyte was 1.202 Ω, which was bigger than that of PANI in the same electrolyte (0.824 Ω), indicating that the Fe2+-modified PANI electrodes were less conductive. This is consistent with the aforementioned XRD and FTIR analyses. In Figure 6C,D, the diameter of the semicircle at the high-frequency region represents the charge-transfer resistance (Rct), which is related to the ion diffusion between the electrode and electrolyte interface [62,63]. After 1000 cycles, for both PANI and 0.4 M Fe2+/PANI in 1 M H2SO4 + Fe2+/Fe3+ (0 or 0.5 M) electrolytes, the Rct became larger than that before the first cycle. This implies the volume expansion of the polymer and/or the decrease of the active sites after 1000 cycles. Especially with the incorporation of Fe2+/Fe3+ in the electrolyte, the additional redox reaction of iron ions at the electrode/electrolyte interface may have caused higher charge-transfer resistance and harder ion diffusion in the SCs.

4. Conclusions

In summary, Fe2+-modified PANI materials were prepared and optimized by electropolymerization. The effects of Fe2+ modification on the physical and chemical properties of electropolymerized PANI were studied. The PANI prepared in the presence of Fe2+ displayed a thinner and more branched fiber structure than that synthesized in the absence of Fe2+. The Cs and capacitance retention of 0.4 M Fe2+/PANI electrode materials in the 1 M H2SO4/PVA electrolyte was 452 F g−1 and 99.59% after 1000 cycles at 5 A g−1, respectively, both of which were much higher than that of its counterpart prepared without the presence of Fe2+. An Fe2+/Fe3+ redox couple additive was also introduced into the electrolyte, and its concentration in the electrolyte was optimized to be 0.5 M. As a result, 0.4 M Fe2+/PANI electrodes tested in the 1 M H2SO4 + 0.5 M Fe2+/Fe3+ gel electrolyte exhibited a remarkable Cs of 8468 F g−1 at 5 A g−1. Moreover, the resultant SCs exhibited enhanced energy density, which was nearly 9 times that of SCs with PANI electrodes without the Fe2+ modification and the Fe2+/Fe3+ additive in the electrolyte and was much higher than that of other reports on PANI symmetric SCs with an H2SO4 + Fe2+/Fe3+ electrolyte. This work demonstrated an effective strategy to fabricate SCs with excellent electrochemical performance through the modification of conductive polymer electrodes with transition metal ions, as well as by adding a redox-active additive in the electrolyte.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/11/8/1357/s1, Table S1: EDS of PANI and 0.4 M Fe 2+/PANI (atomic concentration), Figure S1: Cycling performance of symmetric PANI and 0.4 M Fe2+/PANI SCs in 1 M H2SO4 electrolytes: capacitance retention at 20 A g−1 for 1000 cycles, Figure S2: GCD curves of all-solid-state device under normal, bending, and flexibility conditions.

Author Contributions

Y.M., Y.X., and W.M. performed the experiments; Y.M. wrote the paper; X.D. supervised the project and research work; and X.D. proposed the original idea.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities of China (Project No. 21618408).

Acknowledgments

We would like to acknowledge the financial support from the Fundamental Research Funds for the Central Universities of China (Project No. 21618408).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Senthilkumar, S.T.; Selvan, R.K.; Melo, J.S. Redox additive/active electrolytes: A novel approach to enhance the performance of supercapacitors. J. Mater. Chem. A 2013, 1, 12386–12394. [Google Scholar] [CrossRef]
  2. Zhang, L.L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 2520–2531. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, D.; Jing, H.; Wang, Z.; Li, J.; Hu, M.; Lv, R.; Zhang, R.; Chen, D. Coupled ultrasonication-milling synthesis of hierarchically porous carbon for high-performance supercapacitor. J. Colloid Interface Sci. 2018, 528, 208–224. [Google Scholar] [CrossRef] [PubMed]
  4. Snook, G.A.; Kao, P.; Best, A.S. Conducting-polymer-based supercapacitor devices and electrodes. J. Power Sources 2011, 196, 1–12. [Google Scholar] [CrossRef]
  5. Long, J.Y.; Yan, Z.S.; Gong, Y.; Lin, J.H. MOF-derived Cl/O-doped C/CoO and C nanoparticles for high performance supercapacitor. Appl. Surf. Sci. 2018, 448, 50–63. [Google Scholar] [CrossRef]
  6. Huang, Z.-H.; Sun, F.-F.; Batmunkh, M.; Li, W.-H.; Li, H.; Sun, Y.; Zhao, Q.; Liu, X.; Ma, T.-Y. Zinc–nickel–cobalt ternary hydroxide nanoarrays for high-performance supercapacitors. J. Mater. Chem. A 2019, 7, 11826–11835. [Google Scholar] [CrossRef]
  7. Meng, Y.; Wang, K.; Zhang, Y.; Wei, Z. Hierarchical Porous Graphene/Polyaniline Composite Film with Superior Rate Performance for Flexible Supercapacitors. Adv. Mater. 2013, 25, 6985–6990. [Google Scholar] [CrossRef]
  8. Ma, L.; Su, L.; Zhang, J.; Zhao, D.; Qin, C.; Jin, Z.; Zhao, K. A controllable morphology GO/PANI/metal hydroxide composite for supercapacitor. J. Electroanal. Chem. 2016, 777, 75–84. [Google Scholar] [CrossRef]
  9. He, H.; Ma, L.; Fu, S.; Gan, M.; Hu, L.; Zhang, H.; Xie, F.; Jiang, M. Fabrication of 3D ordered honeycomb-like nitrogen-doped carbon/PANI composite for high-performance supercapacitors. Appl. Surf. Sci. 2019, 484, 1288–1296. [Google Scholar] [CrossRef]
  10. Liu, D.; Liu, J.; Wang, Q.; Du, P.; Wei, W.; Liu, P. PANI coated microporous graphene fiber capable of subjecting to external mechanical deformation for high performance flexible supercapacitors. Carbon 2019, 143, 147–153. [Google Scholar] [CrossRef]
  11. Yang, Z.; Ma, J.; Araby, S.; Shi, D.; Dong, W.; Tang, T.; Chen, M. High-mass loading electrodes with exceptional areal capacitance and cycling performance through a hierarchical network of MnO2 nanoflakes and conducting polymer gel. J. Power Sources 2019, 412, 655–663. [Google Scholar] [CrossRef]
  12. Zhu, C.; He, Y.; Liu, Y.; Kazantseva, N.; Saha, P.; Cheng, Q. ZnO@MOF@PANI core-shell nanoarrays on carbon cloth for high-performance supercapacitor electrodes. J. Energy Chem. 2019, 35, 124–131. [Google Scholar] [CrossRef]
  13. Xia, X.; Hao, Q.; Lei, W.; Wang, W.; Sun, D.; Wang, X. Nanostructured ternary composites of graphene/Fe2O3/polyaniline for high-performance supercapacitors. J. Mater. Chem. 2012, 22, 16844–16850. [Google Scholar] [CrossRef]
  14. Geethalakshmi, D.; Muthukumarasamy, N.; Balasundaraprabhu, R. Effect of dopant concentration on the properties of HCl-doped PANI thin films prepared at different temperatures. Optik 2014, 125, 1307–1310. [Google Scholar] [CrossRef]
  15. Gizdavic-Nikolaidis, M.R.; Jevremovic, M.M.; Milenkovic, M.; Allison, M.C.; Stanisavljev, D.R.; Bowmaker, G.A.; Zujovic, Z.D. High yield and facile microwave-assisted synthesis of conductive H2SO4 doped polyanilines. Mater. Chem. Phys. 2016, 173, 255–261. [Google Scholar] [CrossRef]
  16. Murugesan, R.; Subramanian, E. Effect of organic dopants on electrodeposition and characteristics of polyaniline under the varying influence of H2SO4 and HClO4 electrolyte media. Mater. Chem. Phys. 2003, 80, 731–739. [Google Scholar] [CrossRef]
  17. Xu, W.; Mu, B.; Wang, A. Morphology control of polyaniline by dopant grown on hollow carbon fibers as high-performance supercapacitor electrodes. Cellulose 2017, 24, 5579–5592. [Google Scholar] [CrossRef]
  18. Dominic, J.; David, T.; Vanaja, A.; Muralidharan, G.; Maheswari, N.; Kumar, K.S. Supercapacitor performance study of lithium chloride doped polyaniline. Appl. Surf. Sci. 2018, 460, 40–47. [Google Scholar] [CrossRef]
  19. Xu, H.; Li, J.; Peng, Z.; Zhuang, J.; Zhang, J. Investigation of polyaniline films doped with Ni2+ as the electrode material for electrochemical supercapacitors. Electrochim. Acta 2013, 90, 393–399. [Google Scholar] [CrossRef]
  20. Xu, H.; Wu, J.-X.; Li, C.-X.; Zhang, J.-X.; Wang, X.-X. Investigation of polyaniline films doped with Co2+ as the electrode material for electrochemical supercapacitors. Ionics 2015, 21, 1163–1170. [Google Scholar] [CrossRef]
  21. Xu, H.; Wu, J.; Li, C.; Zhang, J.; Liu, J. Investigation of polyaniline films doped with Fe3+ as the electrode material for electrochemical supercapacitors. Electrochim. Acta 2015, 165, 14–21. [Google Scholar] [CrossRef]
  22. Xu, H.; Zhang, J.; Chen, Y.; Lu, H.; Zhuang, J. Electrochemical polymerization of polyaniline doped with Zn2+ as the electrode material for electrochemical supercapacitors. J. Solid State Electrochem. 2014, 18, 813–819. [Google Scholar] [CrossRef]
  23. Li, J.; Cui, M.; Lai, Y.; Zhang, Z.; Lu, H.; Fang, J.; Liu, Y. Investigation of polyaniline co-doped with Zn2+ and H+ as the electrode material for electrochemical supercapacitors. Synth. Met. 2010, 160, 1228–1233. [Google Scholar] [CrossRef]
  24. Xu, H.; Zhang, J.; Chen, Y.; Lu, H.; Zhuang, J. Electrochemical polymerization of polyaniline doped with Cu2+ as the electrode material for electrochemical supercapacitors. RSC Adv. 2014, 4, 5547–5552. [Google Scholar] [CrossRef]
  25. Xu, H.; Zhang, J.; Chen, Y.; Zhuang, J.; Lu, H. Electrochemical polymerization of polyaniline doped with Mn2+ as electrode material for electrochemical supercapacitors. Polym. Mater. Sci. Eng. 2014, 30, 23–27. [Google Scholar]
  26. Wang, X.; Xu, M.; Fu, Y.; Wang, S.; Yang, T.; Jiao, K. A Highly Conductive and Hierarchical PANI Micro/nanostructure and Its Supercapacitor Application. Electrochim. Acta 2016, 222, 701–708. [Google Scholar] [CrossRef]
  27. Wu, Z.-S.; Parvez, K.; Li, S.; Yang, S.; Liu, Z.; Liu, S.; Feng, X.; Müllen, K. Alternating Stacked Graphene-Conducting Polymer Compact Films with Ultrahigh Areal and Volumetric Capacitances for High-Energy Micro-Supercapacitors. Adv. Mater. 2015, 27, 4054–4061. [Google Scholar] [CrossRef] [PubMed]
  28. Jun, S.C.; Patil, U.M.; Lee, S.C.; Kulkarni, S.B.; Sohn, J.S.; Nam, M.S.; Han, S. Nanostructured pseudocapacitive materials decorated 3D graphene foam electrodes for next generation supercapacitors. Nanoscale 2015, 7, 6999–7021. [Google Scholar]
  29. Palaniappan, S.; John, A.; Amarnath, C.A.; Rao, V.J. Mannich-type reaction in solvent free condition using reusable polyaniline catalyst. J. Mol. Catal. A Chem. 2004, 218, 47–53. [Google Scholar] [CrossRef]
  30. Sathiyanarayanan, S.; Jeyaprabha, C.; Venkatachari, G. Influence of metal cations on the inhibitive effect of polyaniline for iron in 0.5M H2SO4. Mater. Chem. Phys. 2008, 107, 350–355. [Google Scholar] [CrossRef]
  31. Tao, S.; Hong, B.; Kerong, Z. An infrared and Raman spectroscopic study of polyanilines co-doped with metal ions and H+. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2007, 66, 1364–1368. [Google Scholar] [CrossRef] [PubMed]
  32. Lu, Q.; Chen, J.G.; Xiao, J.Q. Nanostructured Electrodes for High-Performance Pseudocapacitors. Angew. Chem. Int. Ed. 2013, 52, 1882–1889. [Google Scholar] [CrossRef] [PubMed]
  33. Krüner, B.; Tolosa, A.; Kim, D.; Choudhury, S.; Presser, V.; Lee, J.; Sathyamoorthi, S.; Seo, K.-H. Tin/vanadium redox electrolyte for battery-like energy storage capacity combined with supercapacitor-like power handling. Energy Environ. Sci. 2016, 9, 3392–3398. [Google Scholar] [Green Version]
  34. Zheng, C.; Yoshio, M.; Qi, L.; Wang, H. A 4 V-electrochemical capacitor using electrode and electrolyte materials free of metals. J. Power Sources 2014, 260, 19–26. [Google Scholar] [CrossRef]
  35. Fic, K.; Frackowiak, E.; Béguin, F. Unusual energy enhancement in carbon-based electrochemical capacitors. J. Mater. Chem. 2012, 22, 24213–24223. [Google Scholar] [CrossRef]
  36. Wang, Y.; Chang, Z.; Qian, M.; Zhang, Z.; Lin, J.; Huang, F. Enhanced specific capacitance by a new dual redox-active electrolyte in activated carbon-based supercapacitors. Carbon 2019, 143, 300–308. [Google Scholar] [CrossRef]
  37. Wang, A.-Y.; Chaudhary, M.; Lin, T.-W. Enhancing the stability and capacitance of vanadium oxide nanoribbons/3D-graphene binder-free electrode by using VOSO4 as redox-active electrolyte. Chem. Eng. J. 2019, 355, 830–839. [Google Scholar] [CrossRef]
  38. Senthilkumar, S.T.; Selvan, R.K.; Lee, Y.S.; Melo, J.S. Electric double layer capacitor and its improved specific capacitance using redox additive electrolyte. J. Mater. Chem. A 2013, 1, 1086–1095. [Google Scholar] [CrossRef]
  39. Senthilkumar, S.T.; Selvan, R.K.; Ponpandian, N.; Melo, J.S.; Lee, Y.S. Improved performance of electric double layer capacitor using redox additive (VO2+/VO2+) aqueous electrolyte. J. Mater. Chem. A 2013, 1, 7913–7919. [Google Scholar] [CrossRef]
  40. Wang, G.; Zhang, M.; Xu, H.; Lu, L.; Xiao, Z.; Liu, S. Synergistic interaction between redox-active electrolytes and functionalized carbon in increasing the performance of electric double-layer capacitors. J. Energy Chem. 2018, 27, 1219–1224. [Google Scholar] [CrossRef] [Green Version]
  41. Chen, W.; Xia, C.; Rakhi, R.; Alshareef, H.N. A general approach toward enhancement of pseudocapacitive performance of conducting polymers by redox-active electrolytes. J. Power Sources 2014, 267, 521–526. [Google Scholar] [CrossRef]
  42. Ren, L.; Zhang, G.; Yan, Z.; Kang, L.; Xu, H.; Shi, F.; Lei, Z.; Liu, Z.-H. High capacitive property for supercapacitor using Fe3+/Fe2+ redox couple additive electrolyte. Electrochim. Acta 2017, 231, 705–712. [Google Scholar] [CrossRef]
  43. Xia, C.; Leng, M.; Tao, W.; Wang, Q.; Gao, Y.; Zhang, Q. Polyaniline/carbon nanotube core–shell hybrid and redox active electrolyte for high-performance flexible supercapacitor. J. Mater. Sci. Mater. Electron. 2019, 30, 4427–4436. [Google Scholar] [CrossRef]
  44. Liu, T.T.; Zhu, Y.H.; Liu, E.H.; Luo, Z.Y.; Hu, T.T.; Li, Z.P.; Ding, R. Fe3+/Fe2+ redox electrolyte for high-performance polyaniline/SnO2 supercapacitors. Trans. Nonferrous Met. Soc. China 2015, 25, 2661–2665. [Google Scholar] [CrossRef]
  45. Bu, Y.; Cao, M.; Jiang, Y.; Gao, L.; Shi, Z.; Xiao, X.; Wang, M.; Yang, G.; Zhou, Y.; Shen, Y. Ultra-thin bacterial cellulose/poly(ethylenedioxythiophene) nanofibers paper electrodes for all-solid-state flexible supercapacitors. Electrochim. Acta 2018, 271, 624–631. [Google Scholar] [CrossRef]
  46. Mai, L.-Q.; Minhas-Khan, A.; Tian, X.; Hercule, K.M.; Zhao, Y.-L.; Lin, X.; Xu, X. Synergistic interaction between redox-active electrolyte and binder-free functionalized carbon for ultrahigh supercapacitor performance. Nat. Commun. 2013, 4, 2923. [Google Scholar] [CrossRef]
  47. Asen, P.; Shahrokhian, S.; Zad, A.I. Transition metal ions-doped polyaniline/graphene oxide nanostructure as high performance electrode for supercapacitor applications. J. Solid State Electrochem. 2018, 22, 983–996. [Google Scholar] [CrossRef]
  48. Chen, W.; Rakhi, R.; Alshareef, H.N. Facile synthesis of polyaniline nanotubes using reactive oxide templates for high energy density pseudocapacitors. J. Mater. Chem. A 2013, 1, 3315–3324. [Google Scholar] [CrossRef]
  49. Kim, Y.; Foster, C.; Chiang, J.; Heeger, A. Localized charged excitations in polyaniline: Infrared photoexcitation and protonation studies. Synth. Met. 1989, 29, 285–290. [Google Scholar] [CrossRef]
  50. Chiang, J.-C.; MacDiarmid, A.G. ‘Polyaniline’: Protonic acid doping of the emeraldine form to the metallic regime. Synth. Met. 1986, 13, 193–205. [Google Scholar] [CrossRef]
  51. Du, X.-S.; Zhou, C.-F.; Wang, G.-T.; Mai, Y.-W. Novel Solid-State and Template-Free Synthesis of Branched Polyaniline Nanofibers. Chem. Mater. 2008, 20, 3806–3808. [Google Scholar] [CrossRef]
  52. Du, X.-S.; Zhou, C.-F.; Mai, Y.-W. Facile Synthesis of Hierarchical Polyaniline Nanostructures with Dendritic Nanofibers as Scaffolds. J. Phys. Chem. C 2008, 112, 19836–19840. [Google Scholar] [CrossRef]
  53. Arulmani, S.; Wu, J.J.; Anandan, S. Ultrasound promoted transition metal doped polyaniline nanofibers: Enhanced electrode material for electrochemical energy storage applications. Ultrason. Sonochem. 2019, 51, 469–477. [Google Scholar] [CrossRef] [PubMed]
  54. Li, H.; Sun, Y.; Yuan, Z.-Y.; Zhu, Y.-P.; Ma, T.-Y. Titanium Phosphonate Based Metal-Organic Frameworks with Hierarchical Porosity for Enhanced Photocatalytic Hydrogen Evolution. Angew. Chem. 2018, 57, 3222–3227. [Google Scholar] [CrossRef] [PubMed]
  55. Alamin, A.A.; Elhamid, A.E.M.A.; Anis, W.R.; Attiya, A.M. Fabrication of symmetric supercapacitor based on relatively long lifetime polyaniline grown on reduced graphene oxide via Fe2+ oxidation sites. Diam. Relat. Mater. 2019, 96, 182–194. [Google Scholar] [CrossRef]
  56. Chen, W.; Rakhi, R.; Alshareef, H.N. Capacitance enhancement of polyaniline coated curved-graphene supercapacitors in a redox-active electrolyte. Nanoscale 2013, 5, 4134–4138. [Google Scholar] [CrossRef] [PubMed]
  57. Thakur, A.K.; Deshmukh, A.B.; Choudhary, R.B.; Karbhal, I.; Majumder, M.; Shelke, M.V. Facile synthesis and electrochemical evaluation of PANI/CNT/MoS 2 ternary composite as an electrode material for high performance supercapacitor. Mater. Sci. Eng. B 2017, 223, 24–34. [Google Scholar] [CrossRef]
  58. Zhang, H.; Zhang, L.; Deng, J.; Han, Y.; Li, X. Fabrication of porous Co3O4 with different nanostructures by solid-state thermolysis of metal-organic framework for supercapacitors. J. Mater. Sci. 2018, 53, 8474–8482. [Google Scholar] [CrossRef]
  59. Dai, Y.; Zhu, S.; Wang, C.; Cong, Y.; Zeng, Y.; Jiang, T.; Huang, H.; Huang, S.; Meng, X. In-situ fabrication of Co foam@Co3O4 porous nanosheet arrays for high performance supercapacitors. J. Alloy. Compd. 2018, 748, 291–297. [Google Scholar] [CrossRef]
  60. Wu, T.; Li, J.; Hou, L.; Yuan, C.; Yang, L.; Zhang, X. Uniform urchin-like nickel cobaltite microspherical superstructures constructed by one-dimension nanowires and their application for electrochemical capacitors. Electrochim. Acta 2012, 81, 172–178. [Google Scholar] [CrossRef]
  61. Liu, Y.; Yan, D.; Li, Y.; Wu, Z.; Zhuo, R.; Li, S.; Feng, J.; Wang, J.; Yan, P.; Geng, Z. Manganese dioxide nanosheet arrays grown on graphene oxide as an advanced electrode material for supercapacitors. Electrochim. Acta 2014, 117, 528–533. [Google Scholar] [CrossRef]
  62. Zhou, H.; Zhi, X.; Zhai, H.-J. A facile approach to improve the electrochemical properties of polyaniline-carbon nanotube composite electrodes for highly flexible solid-state supercapacitors. Int. J. Hydrog. Energy 2018, 43, 18339–18348. [Google Scholar] [CrossRef]
  63. Malik, R.; Zhang, L.; McConnell, C.; Schott, M.; Hsieh, Y.-Y.; Noga, R.; Alvarez, N.T.; Shanov, V. Three-dimensional, free-standing polyaniline/carbon nanotube composite-based electrode for high-performance supercapacitors. Carbon 2017, 116, 579–590. [Google Scholar] [CrossRef]
Figure 1. XRD pattern (A) and FTIR spectrum (B) of polyaniline (PANI) and 0.4 M Fe2+/PANI electrode material.
Figure 1. XRD pattern (A) and FTIR spectrum (B) of polyaniline (PANI) and 0.4 M Fe2+/PANI electrode material.
Polymers 11 01357 g001
Figure 2. Field emission scanning electron microscopy (FESEM) images of PANI (A,a), 0.2 M Fe2+/PANI (B,b), 0.4 M Fe2+/PANI (C,c), and 0.8 M Fe2+/PANI (D,d) electrode materials.
Figure 2. Field emission scanning electron microscopy (FESEM) images of PANI (A,a), 0.2 M Fe2+/PANI (B,b), 0.4 M Fe2+/PANI (C,c), and 0.8 M Fe2+/PANI (D,d) electrode materials.
Polymers 11 01357 g002
Figure 3. (A) Galvanostatic charge–discharge (GCD) curves of PANI and PANI doped with Fe2+ (0.2, 0.4, and 0.8 M) in a three-electrode system in 1 M H2SO4 at a current density of 1 A g−1, respectively. Electrochemical performance of symmetric PANI and 0.4 M Fe2+/PANI supercapacitors (SCs) with 1 M H2SO4 electrolyte: (B) cyclic voltammetry (CV) curves at a scan rate of 20 mV s−1, (C) specific capacitances at different current densities, and (D) Ragone plots.
Figure 3. (A) Galvanostatic charge–discharge (GCD) curves of PANI and PANI doped with Fe2+ (0.2, 0.4, and 0.8 M) in a three-electrode system in 1 M H2SO4 at a current density of 1 A g−1, respectively. Electrochemical performance of symmetric PANI and 0.4 M Fe2+/PANI supercapacitors (SCs) with 1 M H2SO4 electrolyte: (B) cyclic voltammetry (CV) curves at a scan rate of 20 mV s−1, (C) specific capacitances at different current densities, and (D) Ragone plots.
Polymers 11 01357 g003
Figure 4. Electrochemical performance of symmetric PANI and 0.4 M Fe2+/PANI SCs with 1 M H2SO4 + Fe2+/Fe3+ (0, 0.2, 0.5, and 0.8 M) electrolytes: (A) CV curves at a scan rate of 20 mv s−1, (B) GCD curves at a current density of 5 A g−1, (C) specific capacitances at different current densities, and (D) Ragone plots.
Figure 4. Electrochemical performance of symmetric PANI and 0.4 M Fe2+/PANI SCs with 1 M H2SO4 + Fe2+/Fe3+ (0, 0.2, 0.5, and 0.8 M) electrolytes: (A) CV curves at a scan rate of 20 mv s−1, (B) GCD curves at a current density of 5 A g−1, (C) specific capacitances at different current densities, and (D) Ragone plots.
Polymers 11 01357 g004
Figure 5. Electrochemical performance of symmetric 0.4 M Fe2+/PANI SCs in 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte: (A) GCD curves at different current densities and (B) CV curves at different scan rates.
Figure 5. Electrochemical performance of symmetric 0.4 M Fe2+/PANI SCs in 1 M H2SO4 + 0.5 M Fe2+/Fe3+ electrolyte: (A) GCD curves at different current densities and (B) CV curves at different scan rates.
Polymers 11 01357 g005
Figure 6. Cycling performance of symmetric PANI and 0.4 M Fe2+/PANI SCs with 1 M H2SO4 + Fe2+/Fe3+ (0 and 0.5 M) electrolytes: (A) capacitance retention at 20 A g−1 for 1000 cycles; Nyquist plots of (B) symmetric PANI and 0.4 M Fe2+/PANI SCs with 1 M H2SO4 + Fe2+/Fe3+ (0, 0.2, 0.5, and 0.8 M) electrolytes (inset: the close-up view of the high-frequency region); (C) symmetric PANI and 0.4 M Fe2+/PANI SCs with a 1 M H2SO4 electrolyte (inset: the close-up view of the high-frequency region); and (D) symmetric PANI and 0.4 M Fe2+/PANI SCs with a 1 M H2SO4 + 0.5 M Fe2+/Fe3+electrolyte before and after 1000 cycles.
Figure 6. Cycling performance of symmetric PANI and 0.4 M Fe2+/PANI SCs with 1 M H2SO4 + Fe2+/Fe3+ (0 and 0.5 M) electrolytes: (A) capacitance retention at 20 A g−1 for 1000 cycles; Nyquist plots of (B) symmetric PANI and 0.4 M Fe2+/PANI SCs with 1 M H2SO4 + Fe2+/Fe3+ (0, 0.2, 0.5, and 0.8 M) electrolytes (inset: the close-up view of the high-frequency region); (C) symmetric PANI and 0.4 M Fe2+/PANI SCs with a 1 M H2SO4 electrolyte (inset: the close-up view of the high-frequency region); and (D) symmetric PANI and 0.4 M Fe2+/PANI SCs with a 1 M H2SO4 + 0.5 M Fe2+/Fe3+electrolyte before and after 1000 cycles.
Polymers 11 01357 g006
Table 1. Comparison of specific capacitance (Cs) values of PANI-based electrode materials.
Table 1. Comparison of specific capacitance (Cs) values of PANI-based electrode materials.
Material Fabrication MethodCs/F·g−1ElectrolyteCitation
PANIChemical oxidative polymerization1062/2 A g−11 M H2SO4 + 0.8 M Fe3+/2+[42]
PANI/SnO2Chemical oxidative polymerization1172/1 A g−11 M H2SO4 + 0.4 M Fe3+/2+[44]
PANI/CNTElectrodeposition1128/5 A g−11 M H2SO4 + 0.02 M Fe3+/2+[43]
Fe3+-Zn2+-PANI/GOElectrodeposition1140/10 A g−10.5 M Na2SO4[47]
PANI/RGOChemical oxidative polymerization565/0.1 A g−16 M KOH[55]
Fe3+/PANIElectrodeposition602/3 mA cm−20.5 M H2SO4[21]
Fe2+/PANIElectrodeposition8468/5 A g−11 M H2SO4 + 0.5 M Fe3+/2+This work

Share and Cite

MDPI and ACS Style

Mo, Y.; Meng, W.; Xia, Y.; Du, X. Redox-Active Gel Electrolyte Combined with Branched Polyaniline Nanofibers Doped with Ferrous Ions for Ultra-High-Performance Flexible Supercapacitors. Polymers 2019, 11, 1357. https://doi.org/10.3390/polym11081357

AMA Style

Mo Y, Meng W, Xia Y, Du X. Redox-Active Gel Electrolyte Combined with Branched Polyaniline Nanofibers Doped with Ferrous Ions for Ultra-High-Performance Flexible Supercapacitors. Polymers. 2019; 11(8):1357. https://doi.org/10.3390/polym11081357

Chicago/Turabian Style

Mo, Youtian, Wei Meng, Yanlin Xia, and Xusheng Du. 2019. "Redox-Active Gel Electrolyte Combined with Branched Polyaniline Nanofibers Doped with Ferrous Ions for Ultra-High-Performance Flexible Supercapacitors" Polymers 11, no. 8: 1357. https://doi.org/10.3390/polym11081357

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