Next Article in Journal
Evaluation of CO2 Hydrogenation in a Modular Fixed-Bed Reactor Prototype
Next Article in Special Issue
Energy Storage and CO2 Reduction Performances of Co/Co2C/C Prepared by an Anaerobic Ethanol Oxidation Reaction Using Sacrificial SnO2
Previous Article in Journal
Electrochemical Studies of Pd-Based Anode Catalysts in Alkaline Medium for Direct Glycerol Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrode Materials for Supercapacitors: A Review of Recent Advances

by
Parnia Forouzandeh
*,
Vignesh Kumaravel
and
Suresh C. Pillai
*
Nanotechnology and Bio-Engineering Research Group, Department of Environmental Science, Institute of Technology Sligo, Ash Lane, F91 YW50 Co. Sligo, Ireland
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(9), 969; https://doi.org/10.3390/catal10090969
Submission received: 10 July 2020 / Revised: 11 August 2020 / Accepted: 19 August 2020 / Published: 26 August 2020
(This article belongs to the Special Issue Electrocatalysis and Electrode Materials for Energy Production)

Abstract

:
The advanced electrochemical properties, such as high energy density, fast charge–discharge rates, excellent cyclic stability, and specific capacitance, make supercapacitor a fascinating electronic device. During recent decades, a significant amount of research has been dedicated to enhancing the electrochemical performance of the supercapacitors through the development of novel electrode materials. In addition to highlighting the charge storage mechanism of the three main categories of supercapacitors, including the electric double-layer capacitors (EDLCs), pseudocapacitors, and the hybrid supercapacitors, this review describes the insights of the recent electrode materials (including, carbon-based materials, metal oxide/hydroxide-based materials, and conducting polymer-based materials, 2D materials). The nanocomposites offer larger SSA, shorter ion/electron diffusion paths, thus improving the specific capacitance of supercapacitors (SCs). Besides, the incorporation of the redox-active small molecules and bio-derived functional groups displayed a significant effect on the electrochemical properties of electrode materials. These advanced properties provide a vast range of potential for the electrode materials to be utilized in different applications such as in wearable/portable/electronic devices such as all-solid-state supercapacitors, transparent/flexible supercapacitors, and asymmetric hybrid supercapacitors.

1. Introduction

Recently, due to the need for preserving natural resources and regulating global energy consumption, there has been a crucial demand for the development of clean energy sources. Among various unconventional electric-power devices, such as supercapacitors (SCs), batteries, and fuel cells, (SCs), which have been used in various fields, (e.g., hybrid cars and electric mass transit vehicles [1], and wearable/portable electronic devices [2], etc.,), SCs have inspired a vast range of research attention owing to their excellent electrochemical performances, including high specific power, superior cycling life and a rapid charging–discharging rate [3]. SCs are a type of capacitor that are also known as ultracapacitors, electrochemical capacitors (ECs) [4], gold capacitors, electrical double-layer capacitors (EDLCs), pseudocapacitors, or power coaches [5,6]. The main combination of SC configuration consists of an electrolyte, current collector, and electrodes with large specific surface area (SSA), which enhances the capacitance 10,000 times more than conventional capacitors [7,8] and thin dielectric separators, which provide higher capacitance for SC compared to conventional batteries [9].
The energy storage in SCs is based on the charge–discharge mechanism at the electrode–electrolyte interface [10] in which the principle is similar to conventional capacitors; however, the charge–discharge process is much faster in SCs [11]. The charge storage range for conventional capacitors is between micro to millifarads while for SCs is between 100–1000 F for each device retaining low equivalent series resistance (ESR) and specific power [12]. SCs could cover both the specific energy and power density (Pd) by several orders of magnitude compared to batteries exploiting proper design and efficient materials, which make them as a flexible energy storage option [12]. The specific energy (energy density) determines how long the energy storage device can be utilized, and the specific power (power density) signifies how rapidly a device can deliver the energy.
The segregating feature of the SC to the conventional capacitor is the absence of dielectric material in the SCs [13]. In comparison to SCs, conventional capacitors store and deliver very high energy density (Ed), but the charge and discharge time of batteries is much slower. SCs also show more safety, a low level of heating, device stability, light-weight, and flexible packaging [14,15,16]. Figure 1, [17], displays the Ragone plot of comparison between conventional capacitor, fuel cells, batteries, and hybrid supercapacitors (HSCs), as various energy storage devices, in terms of power density (Pd) and Ed in which the HSCs show a significant feature of Pd in comparison with batteries and fuel cells but considerably lower Pd than conventional capacitors.
According to the charge storage principle, SCs can be categorized into three groups, including (a) EDLCs that exploit double layer electrodes to store the charge in through non-faradic interactions (electrostatically) [18], (b) pseudocapacitors in which the charge storage mechanism is faradaically through oxidation-reduction reactions associated with their various potential [19], and (c) asymmetric supercapacitors (ASSC), in which, the charge storage mechanism of the two electrodes is not similar, and the storage principle is based on both Faradaic and non-Faradic behaviour of used material [13]. Generally, the “asymmetric” word refers to various charge storage mechanisms corresponding to the positive and negative electrodes in the SC device. According to this, the HSC has been categorized as one of the specific class of the ASSC. The energy storage mechanism of a device with supercapacitive materials as the cathode and a battery material as the anode is called HSC. ASSCs provide a high energy density and power density due to the extended operating voltage window (up to about 2.0 V in the case of aqueous electrolytes, and about 2.7 V in organic electrolytes, and even more than 4.0 V in ionic liquids) by taking the advantages of the two different electrode materials, which could operate at different potential windows. Figure 2 represents the schematic illustration of the EDLCs, pseudocapacitors, and the HSC.
The main characteristics for these energy storage systems, which provide the ability to distinguish one system from others are including energy density, power density, capacitance/capacity, cell voltage, kinetic storage mechanism, i–v relationship, operating temperature, cycle life, self-discharge, and used materials as an electrode, electrolyte, etc. Compared to EDLCs and pseudocapacitors, the HSCs displayed significant advantages such as higher energy density [20], and long cycle life (>100,000) [21], that has been obtained in a higher cell voltage (maximum 0.5) [22]. Lower self-discharge (energy dissipation) is another advantage of HSCs, in comparison with the EDLCs and pseudocapacitor devices [23]. Generally, self-discharge mechanisms, which are associated with various chemical and physical mechanisms, are mainly categorized into three classes: (a) parasitic faradaic reactions, (b) charge redistribution, and (c) ohmic leakage of current between the electrodes [24,25,26]. Accordingly, self-discharge occurs, if one or the combination of these processes is available.
Selecting a standard lifetime as one of the main parameters of the rechargeable energy storage devices is entirely subjected to the nature of the device, and it is highly linked to the operating conditions such as temperature, voltage, charge–discharge current, the depth of the discharge, and at some stages, the humidity [17,27]. It is essential to select the appropriate materials with high purity for both electrolyte and electrodes as it influences the cycling life and the current leakage. Cyclic stability can be estimated through GCD measurements by repeating the cycles continuously at a constant current density [28].
The cycling lifetime of the hybrid system is approximately 5000 cycles, which could vary with respect to the batteries (for non-Li-ion batteries around 500 cycles, and for Na- and Li-ion batteries around 2000 cycles) [29], and the other types of supercapacitor devices (around 10,000 cycles) [30]. The description of the life cycle is highly relient on the application of the corresponding device, where it can be defined as the number of full charge–discharges before the device capacitance/capacity reduces below 80% of the start-of-life value [31,32,33].
In the case of the death state of a battery intended for a hybrid or an ordinary electric vehicle, the life cycle is described as 80% of the corresponding initial capacity [34]. In the following sections, the cycle lifetime results for various energy storage devices are reported as the percentage of the initial capacity/capacitance of the device that remained after a specified number of cycles. In this review, a detailed description of the features of various electrode materials has been given. The electrochemical properties, preparation methods, and the applications of the materials have been discussed in detail. The following features have reviewed comprehensively: (i) the difference between the “pseudocapacitive” and “battery-type” electrode material, (ii) the importance on the selection of the electrode material, and their properties influencing the SCs performance, and (iii) the major categories of materials (carbonaceous materials, TMOs/TM(OH)s, and CPs), and their composites. The advantages and disadvantages of the electrode materials are also briefly compared. The charge storage behaviour of different components has been introduced through recent advances. Then, the importance of the utilization of 2D material (either individually or in combination with other materials) has been comprehensively explained. In the end, an outlook on the significance of the bioderived functional groups into the electrode material, their advantages, and application in various SCs such as transparent SCs, FSCs, and HSCs is given.

2. The Distinction between Terminologies Used: Pseudocapacitive and Battery-Like Electrode Materials

There are various materials, which can be utilized to provide a favorable charge storage mechanism. Generally, the carbonaceous materials are used as capacitive-type electrodes, while transition metal (TM) and conducting polymer-based material can be used as the battery type-electrodes. The battery-type materials can be classified into two groups, including, intercalation-type (intercalation-deintercalation), and conversion-type (two-phase conversion). This classification doesn’t mean that charge storage occurs only in one type of mechanism for the electrode materials. The charge storage process of the battery-type electrode materials is associated with both conversion and intercalation-deintercalation behaviour. There are also different electrode materials that exhibit either capacitive or battery-type behavior, depending upon various potential windows, the electrolyte selection, etc. Thus, to simplify the concept, the electrochemical behaviour of the redox-active materials (whether two-phase conversion or intercalation-type) can be described as “pseudocapacitive behaviour”.
In the field of supercapacitors, the term “pseudocapacitance” is exploited to define the behavior of electrode materials (e.g., RuO2 [35,36], MnO2 [37,38]) that exhibit the electrochemical characteristics of a capacitive electrode such as a linear dependency between the stored charge and the potential variation within the certain potential window, however, in this case, the charge storage should be originated from the electron-transfer mechanisms, instead of only relying on the ions accumulation in the electrochemical double layer (EDL), similar to that with activated carbons (AC) [39].
Several types of transition metal oxides (TMOs) have been considered as pseudocapacitive materials in aqueous electrolytes. The charge storage behavior of these materials is associated with their structural and hydration properties as well as the nature of the electrolyte. The structural water provides a large inner surface, which is accessible for ion absorption and leads to improved ion diffusion in the inner surface (e.g., RuO2·nH2O). Thus, the influence of surface-bound water is one of the significant issues, which makes a difference between the aqueous and non-aqueous pseudocapacitive materials. For instance, MnO2 is present in several polymorphs and mainly studied for the pseudocapacitive behavior in a neutral aqueous electrolyte, while it is considered as a battery-like material in alkaline electrolytes. The battery-type behavior in an alkaline electrolyte (e.g., KOH) is attributed to the formation of oxyhydroxides during the charging process [40]. Besides, the spinel compounds (e.g., Mn3O4 [41], Fe3O4 [42], and Co3O4 [43]), display both capacitive and battery-like behavior owing to the phase variations, which occurs during the cycling mechanism.
In various studies, the term “pseudocapacitive” has been used for battery-like material such as Ni(OH)2 [44,45] or for other materials that display faradic behavior (even for the materials, that are electrochemically irreversible) [46,47]. This may confuse the reader, since the term “capacitance” with the unit of the farad (F), cannot be applied for the material that their charge storage mechanism is only based on the faradic behavior. For these classes of material, the concept of “capacity” (with the unit of C (coulomb), or mAh) is a more significant and suitable term [39]. Figure 3 illustrates the general classification of the pseudocapacitive charge storage mechanism [17].
Among these various processes, the intercalation pseudocapacitance is hardly observed, since the charge storage for the corresponding materials is mostly hybrid, which can be defined as a combination of both “battery-type”, and “capacitive” behaviors [48]. In order to distinguish between the contributions from battery-like energy storage process and intercalative pseudocapacitance mechanism, the corresponding signatures from the electrochemical measurements (e.g., CVs and GCD), as well as the quantitative kinetics analysis, should be considered, [17,40,49], which are illustrated in Figure 4, [49]. According to the Equation (1) [50], the capacitive contribution and the diffusion-controlled contribution can be examined by investigating the CV data at different sweep rates, in which the calculated current “i” follows a power-law relationship with the sweep rate V. The “a” and “b” parameters are changeable, and the b values can be estimated from the slope of the plot of log i vs. log V [49,50].
i ( V ) = av b
As is shown in Figure 4, the “b” value corresponding to EDLC materials is always equal to 1, the CVs is rectangular, and in the case of their GCDs, the potential has a linear dependency on time. Pseudocapacitive materials exhibit the CVs in an almost rectangular shape, and the potential of GCDs is practically linear with respect to time. The corresponding “b” value is close to 1, and as it was mentioned before, the charge storage capability is represented using the F unit. In comparison with EDLCs and pseudocapacitive materials, the battery-type materials show the typical CVs with noticeable redox peaks, and the obvious plateaus are obtained for their corresponding GCDs. The “b” value for the battery-type material, depends on the nature of material, and it can be 0.5 and higher than 0.5 for conventional bulk battery-type materials or those with specific structural design electrode engineering, and for nanomaterial electrodes, respectively [49].
There are several studies, in which the presence of structural water in crystalline oxides resulted in intercalation pseudocapacitance-non-diffusion-controlled proton insertion into the interior of bulk-like oxides. For instance, structural water in layered-hydrated tungsten oxide (WO3·2H2O) induced the transition from battery-like electrochemical behavior to pseudocapacitive behavior for proton intercalation in comparison with the anhydrous WO3 [51].
The substantial difference between battery-like and pseudocapacitive electrode materials is that the charging and discharging time for the pseudocapacitive materials is in the range of seconds—minutes. Therefore, it is worth focusing on emerging pseudocapacitance in order to obtain high energy and high power densities from the same material [40].

3. Electrode Materials

The selection of electrode materials and their fabrication play a crucial role in enhancing the capacitive performance of SCs [52]. Electrodes of SCs must provide thermal stability, high SSA, corrosion resistance, high electrical conductivity, appropriate chemical stability, and suitable surface wettability. They should also be low-cost and environmentally benign. Besides, their capability of transferring the faradic charge is important to enhance the capacitance performance [53,54].
In fact, the specific capacitance factor is not only affected by surface area, but also there are other essential parameters including the simple morphology manipulation, including pore size distributions, as well as the pore shape, pore sizes, and their availability for the electrolyte [11,55,56].
Thus, two of the most important requirements for designing an SC device are (i) improving the electrochemically active sites by selecting the electrode materials with a high SSA, and (ii) tuning the pore size and pore shape (e.g., circle, vertical rectangle, horizontal rectangle, and square in the case of graphene nanopore [57], and cylindrical, spherical, and slit in the case of the biomass-derived carbon [58]) of the electrode material to make them suitable candidates which ease the transportation of electrolyte ions [11,59].
The electrode material having smaller pores offer higher capacitance and, thus, higher Ed. However, the smaller the pore size leads to having a higher the equivalent series resistance (ESR), therefore, decreasing the power density. Based on this, the application of the electrode material affects the selection of the materials. For example, for the applications in which having more peak currents is important, the electrode materials should satisfy the criteria of having greater pores, while for the applications that higher Ed is required, the electrode materials should have smaller pores [60,61].
Furthermore, the suitable size distribution can improve the retention capability, which is the indication of high Pd in an SC device. The effective distribution of micro/mesopores of the electrode material can provide a fast mass and ion transport through a continuous pathway, thus enhance the accessibility of the electrolyte, and make the material an appropriate choice for SC application [62].
The electrode materials are categorized based on three main classes, including carbonaceous materials, TMOs, and conducting polymers (CPs) [63].
Figure 5 shows the classification of the electrode materials for SC application from zero-dimensional (0D) to three-dimensional (3D) structure that is explained in detail through the following sections.

3.1. Nanostructured Carbon-Based Materials

The EDLC devices, capacitance mainly derives from the collected charges at the interface between electrodes and the electrolyte. Consequently, the capacitive performance of an EDLC device is extremely affected by the pore sizes, pore distribution and SSAs of the electrode materials [64,65]. Becker developed the first EDLC utilizing carbon-based materials in 1975 [66].
Mainly, the carbonaceous materials are utilized as electrode material for EDLCs owing to their improved properties such as high surface area, low cost, eco-friendliness, and their facile synthesis methods [67]. The physicochemical characteristics of the carbon-based material depend on the pore size that is approximately less than 1 nm [68].
Furthermore, their thermal stability, chemical and electrochemical stability (in various solutions from a strong acidic to basic media), high electrical conductivity, as well as the symmetrical galvanostatic charge–discharge profile and the good rectangular shape of CV curves propose that the carbon-based materials are the useful capacitive materials [54,69,70,71]. The capacitance of the electrodes extremely relies upon the SSA. Consequently, owing to the large SSA and the high porosity of the carbonaceous materials, they have displayed enhanced capacitance [72].
The charge storage mechanism of the carbon-based electrode mainly depends on the formation of a thin layer known as the Helmholtz layer on the surface of electrode material, which is in contact with the electrolyte ions (both non-dissolved and very weakly-dissolved ions). Secondly, the charge storage derives from the formation of a thicker outer Helmholtz layer, which is made through the existence of the solvated ions conducted by strong electrostatic forces. The indicated layer can be prolonged to a thicker layer, the so-called Gouy–Chapman diffuse layer, through solvated ions and the thermal motion. The degree of charge storage in an EDLC device can be described as the potential difference between the thin inner layer and outer thicker layer (the Helmholtz layers and the Gouy–Chapman diffuse layer), which is known as zeta potential (ξ-potential) [73]. Besides, there is a huge necessity to enhance the capacitance of the carbon-based material to obtain higher Ed. Introducing the redox-active moieties has provided the pseudocapacitive carbon-based material through different strategies such as utilizing redox-active functional groups [74], chemical doping of the carbon-based material using redox-active small molecules [75] and combining the aromatic isomers adsorbed onto the structure of carbon material [76].
Carbon-based materials could be used in various forms such as powder, fiber, monoliths, and foils in an SC device. In addition to the pore shape and SSA, there are other influential items such as the structure of the material, and pore size distribution, and electrical conductivity that have an important consequence on the electrochemical performance of an SC device [77,78,79]. There are several common examples of carbon-based material such as, carbon aerogels [80], graphene [81], AC [82], activated carbon fibres (ACF) [82], carbon nanotube (CNT) [83], and carbon cloth(CC) [84] as well as various carbon-based composites, which are suitable to be used as the electrode materials in SC devices, and are explained in detail through the following sections.

3.1.1. Activated Carbon Materials (ACs)

Activated carbons are the most commonly used material for EDLC electrodes with moderate cost [85] that show a complex porous structure comprising micropores (smaller than 2 nm), mesopores (in the range of 2–50 nm), and macropores larger than 50 nm that lead to having a high SSA [11,54,69,86]. Furthermore, although their good electrical conductivitie is lower (1250–2500 Sm−1), still, it is suitable for SCs. According to those properties, activated carbons are considered as useful electrode materials for SCs [82,87,88,89,90,91,92]. Activated carbon can be produced in two steps and involving either chemical or physical activation from various types of carbon-rich organic precursors such as fossil fuels, coke or synthetic polymers, wood, coconut shells [11].
The carbonization step involves the treatment of precursors in the exposure of oxidizing gases, such as CO2, steam, and air, at high temperatures (700–1200 °C) in an inert atmosphere to obtain the amorphous carbon. In the activation step, the applied temperature is lower (400 to 700 °C). The chemical activation process is carried out in the presence of the activating agents like carbonates, alkalis, chlorides, or acids such as NaOH, KOH, ZnCl2, and H3PO4 that lead to having a porous network in the carbon particles bulk [87,93].
Activated carbon shows plenty of physicochemical properties based on the carbon precursors used and activation methods that have a strong influence on its surface area, porous structure, and the pore size distribution [93].
As mentioned, activated carbon can be obtained from various carbon-rich organic precursors. The obtained coconut shell-based activated carbon exhibits supplementary micropores than the one which is made from charcoal. The research shows that a slighter pore size distribution can be conveniently measured in the case of activated carbon fibre made from activated carbon with the surface area around 2500 m2·g−1 [15]. Since the electrolyte ions with large size cannot inter to the smaller micropores, for activated carbon, the whole SSA does not contribute to the charge storage mechanism [70,71].
Here, a maximum obtainable surface area was estimated by 3000 m2·g−1; however, the effective surface area is obtained between 1000 to 2000 m2·g−1 [93]. Despite those enhancements in the case of specific area and porosity of the activated carbon, there are some reports indicating electrolyte decomposition and dangling bond positions derived by the increase in SSA [94].
The specific capacitance values for the activated carbon electrodes in aqueous electrolytes (between 100 to 300 F·g−1) is higher than in organic solution (<150 F·g−1), since the size of the ions for the organic electrode is larger [11,95] Wang et al., shown that by increasing the SSA of AC from 621 m2·g−1 to 2685 m2·g−1, the specific capacitance of the electrodes was improved from 17.68 F·g−1 to 171.2 F·g−1 [72].
As a result, estimating the pore size suitable for a specified ion size is one of the items that should be considered. Choosing an appropriate method is essential to improve the pore size distribution in the assembly process. According to previous reports, the pore size of 0.8 nm can be suitable in the case of organic electrolytes, while the pore size of 0.4 or 0.7 nm can be useful for the aqueous electrolytes [70,71]. Furthermore, the surface functionalization of the carbon material results in an increase faradaic redox reactions and, consequently, enhancement of about 5 to 10% in capacitance [96]. Achieving a maximum specific capacitance highly depends upon the coordination between the pore size and the ion size [97,98].
One of the significant disadvantages of activated carbon is its low density. Moreno-Fernandez et al. [82] showed that a high specific capacitance value (200 F·g−1) could be attained through the synthesis of the activated carbon fibre monolith that leads to a higher density by three times more. In another study, Li et al. [99] have fabricated activated carbon made by lignite at activation temperatures of 650 °C, which displayed cyclic stability over 3000 cycles, a specific capacitance of 207.5 F·g−1 as an electrode material for SC at a current density of 0.5 A·g−1.
Recently, hybrid lithium-ion SCs have been fabricated using the agar-derived AC as the cathode material, which stores the charge through the electric double-layer capacitance behaviour. The HSC provides both properties of high Pd (like capacitors), and high Ed (like lithium batteries). AC showed an appropriate pore volume and a narrow pore-size distribution that leads to having a large SSA more than 1672 m2·g−1, the high specific capacitance of 210.4 F·g·AC−1. The HSC provided a high Ed of 308.3 Wh·kg·AC−1 at a specific power of 0.7 kW·kg·AC−1 [100].
The electrochemical performance of the ACs could be improved through different strategies such as the Fermi level position modification of the ACs using ultrasonic radiation [101], modification of the AC surface through oxidation process [102], doping sulfur and oxygen functionalities on the porous structure of ACs [103], fabricating a composite using ACs and other carbon nanomaterials (e.g., carbon nanotubes (CNTs) and carbon nanofibers (CNFs), etc.) [104], and inserting polymers into the carbon substrate to form composite electrodes [105,106].
For instance, Cheng et al. [104] developed a new class nanocomposite electrode through the combination of the conventional ACs and carbon black (CB) with CNTs and CNFs. The resultant electrodes showed excellent electrochemical performance in comparison with that electrode of pure AC due to the synergistic effects of combined species. The quaternary nanocomposite electrode (Figure 6a–d) exhibited high capacitance retention in which 91.4% of capacitance remained after 30,000 charging–discharging cycles at 10 A·g−1 (Figure 6e). Compared to the AC electrode, the nanocomposite showed enhanced Ed of 29.6 Wh·L−1 at a Pd of 101.7 kW·L−1 (Figure 6f) as well as a high capacitance of 66.1 F·cm−3.

3.1.2. Carbon Nanotubes (CNTs)

Electrodes of CNTs have been developed due to their unique properties, such as their accessible external structure with high SSA and the specific internal network of mesopores, chemical and thermal stability, superior electrical conductivity, and low mass densities.
Furthermore, the mechanical flexibility of the CNTs and their open network provides them with a link to active material, higher interconnection, and to form a continuous distribution that leads to the more utilized SSA [11,107].
The CNT electrodes show lower ESR, and consequently higher power than ACs since the mesoporous structure of CNT networks allows the ions of the electrolyte to diffuse easier and faster [108] (Figure 7a,b). Du et al. decreased the ESR of electrodes outstandingly, through the adopting thin films fabricated by means of a highly concentrated colloidal suspension of CNTs [109]. The CNTs could be fabricated through different methods. Shi et al. [110] stated the importance of CNTs synthesize method on the performance of the electrode. CNTs were grown on 3D porous nickel (Ni) foam by means of the single-step CVD method, in which the resultant electrode exhibited a superior capacitive performance at higher current densities, less ESR, and a capacitance about 127% greater than the CNT electrode fabricated through the conventional transfer technique. However, the low SSA is the limiting factor for CNT application in high-performance SCs.
CNTs could be used as electrode material for SCs in the form of composite along with other materials such as graphene, metal oxides (MOs), polymer. Q. Cheng [108] studied the electrochemical properties of the graphene, CNTs, and their composite electrode material for SC in various electrolytes. Graphene/CNTs electrodes showed superior electrochemical properties (Figure 7c,d) Ed of 155.6 Wh·kg−1, and a maximum Pd of 263.2 kW·kg−1 in ionic liquid as well as a specific capacitance of 290.4 F·g−1. As is shown in Figure 7c, the graphene/CNTs composite exhibited the most rectangular geometry, which is an indication for an outstanding charge dispersion in the electrodes.
Furthermore, they have reported an interesting phenomenon known as “electro-activation”, in which the specific capacitance of graphene and graphene/CNTs not only did not reduce after a long time cycling in the 1 M KCl electrolyte but also have shown a high increase. It has been reported that the ions result in the larger spacing amid the graphene sheets, thus providing a larger surface area, which presents higher available double-layer capacitance (Figure 7e,f).
CNTs can be classified into two main categories such as SWCNTs and multi-walled carbon nanotubes (MWCNTs) [111]. Generally, CNTs can be synthesized via catalytic decomposition of certain hydrocarbons. Several parameters, such as crystalline orders, can be manipulated to obtain a different conformation like SWCNTs and MWCNTs [87]. In the case of the SWCNTs, they usually form bundles, which limits their surface area and only the outermost surface can be available for electrolyte ions [108].
The performance of the single-walled CNTs was profoundly affected by the accessibility of electrode materials and the size of the electrolyte ions [112]. The theoretical SSA for SWCNTs was reported at about 1315 m2·g−1, while the SSA for MWCNTs was smaller than ACs. This smaller SSA leads to a lower Ed for CNT in comparison with AC [87]. However, the MWCNTs mesopores structure provides easier ions movement across the electrode-electrolyte interface, their mechanical stability decreases because of the volume change derived from the frequent intercalation and depletion [113].
The CNT materials can go through the chemical activation using potassium hydroxide to improve the specific capacitance, which leads to enhancing the surface area of CNT to be two or three times higher while the morphology of the CNT is still defined [87].
The morphology and the purity and of the final CNT material are two crucial factors that affect the specific capacitance of the electrodes [87]. For instance, the MWCNT sheet electrodes were synthesized via catalytically-grown CNTs, which had a diameter of 80 Å and a higher SSA of 430 m2·g−1 than another CNT with a maximum specific capacitance of 104 F·g−1 derived from different methods [114]. In another report, the effect of the acid treatment on the specific capacitance of MWCNTs was studied. In comparison with the specific capacitance of the untreated CNTs (~10 F·g−1), the treated MWCNTs provided a specific capacitance of about 50 F·g−1. According to the FTIR spectrum, the obvious peak corresponds to the C–H bond in pristine CNTs became negligible in the acid-treated MWCNTs, which indicated that the acid treatment could chemically modify the C–H defects to introduce more oxygen-containing groups such as carboxyl, carbonyl, and hydroxyl groups to the surface of MWCNTs [115].
Xie et al. [53] synthesized hierarchical porous carbon microtubes (HPNCTs) through the simple carbonization and following KOH activation method utilizing willow catkins. The final product showed a tubular structure, derived from the willow catkins as well as a highly porous morphology as a result of the activation process. The resulting material showed a gravimetric capacitance of 292 F·g−1, the SSA of 1775.7 m2·g−1.
CNT materials could be deposited directly onto the conductive substrates. In this case, the CNTs act as mechanical support, which increases electron transport, thus decreasing the contact resistance between the current collectors and active materials [116].
Furthermore, various classes of materials and composites were used to coat the CNTs to enhance the electrochemical properties of the electrode. The CNT current collectors provide high cyclic stability, high rate capability, as well as enhanced Ed [117].
For instance, Liu et al. [117] fabricated flexible supercapacitors (FSCs) containing high-performance fibrous electrodes using the composite of MnO2 nanorods, which have been loaded on MXene nanosheets (NSs), and coated on carbon nanotube fibres (CNTFs). In this study, the resultant MnO2@MXene/CNTF displayed high capacitance retention after 5000 cycles, excellent specific capacitance, and superior flexibility. The CNTFs were applied as a current collector with excellent mechanical and electrical properties.
CNTs could be used as an electrode material along with novel MXene materials [118], TMOs [119], and their composites [117], which are the most common electrode materials for the SC.

3.1.3. Graphene

Graphene is a one-atom-thick layered 2D structure with sp2-bonded carbon atoms, which are arranged in a honeycomb-designed crystal structure [120,121,122,123,124]. Graphene, as a unique carbon material, shows excellent features including superior mechanical, morphological characteristics, and electrical conductivity, as well as high carrier mobility, large surface area, and high chemical stability [125,126,127]. Furthermore, the 2D structure of graphene decreases the thickness of the electrode that leads to obtaining a larger potential window, greater flexibility, chemical, and thermal stability [124]. Due to these properties, graphene is a powerful candidate for energy storage applications [123].
Accordingly, graphene could be utilized as electrode material in the application of high-performance SC, since, in comparison with other carbon-based materials such as ACs and CNTs, graphene-based electrodes do not depend on the distribution of pores in the solid-state [128,129]. Another advantage of utilizing the graphene as an electrode material is the accessibility of both exterior surfaces of the graphene sheet by the electrolyte [68,130,131].
A specific capacitance of 205 F·g−1 has been obtained for the SC fabricated using graphene-based electrodes. The resultant SC showed an Ed of 28.5 Wh·kg−1 and a Pd of 10 W·g−1 [132]. In another study, the flexible thin films of carbon showed that the final graphene material exhibited an electrical conductivity of 5880 S m−1 and a high SSA of 2400 m2·g−1. The obtained specific capacitance and the Ed of the graphene material were reported about 120 F·g−1 and 26 Wh·kg−1, respectively [133].
There are plenty of methods to produce various types of graphene, such as micromechanical exfoliation, epitaxial growth, chemical vapour deposition, unzipping of CNTs, electrochemical and chemical approaches, arch discharge method and intercalation techniques in graphite [68,130,131], Figure 8 illustrates the synthesis of N-doped porous graphene (N-pGr), in which the pGr was obtained through the oxidation of epoxy groups on the surface of reduced graphene oxide(rGO) using H2O2, in addition to the hydrolysis of hydroxyl functionalities through the parallel breaking of graphene C–C bonds [134,135].
Nanostructured graphene oxide (GO) could be synthesized with different morphology such as single layers, few layers, and 3D porous structures through the hydrothermal process. The morphology could be controlled through the process, adding various quantities of Ni2+. Among the various morphologies, the 3D porous structure displayed an optimum specific capacitance of 352 F·g−1 at a scan rate of 5 mV·s−1 [137].
Ogata et al. [138] prepared the rGO/GO/rGO device, which showed both SC and battery behaviour. The device behaved as a proton-type SC until the operating voltage about 1.2 V, and for voltages higher than 1.5 V, the device provided the redox reaction through the interaction between the oxygenated functional moieties on the rGO and behaving as a battery.
Moreover, Liu et al. [125] solved the problems of water-based electrolytes to function at broader voltage windows, synthesizing a surface compatible and pore accessible graphene for ionic liquid electrolytes. They prepared a mesoporous graphene network which was accessible to the ionic electrolyte and was able to operate a voltage up to 4 V. The synthesized material showed the Ed of 136 Wh·kg−1, and 90 Wh·kg−1 at 80 °C, and at room temperature, respectively.
In another study, a symmetrical SC was assembled utilizing various electrode materials made of GO, polyvinylidene fluoride (PVDF), and KOH with a different ratio. Among all the samples in this study, G1p4K6-2, and G1p9K6-4 showed better electrochemical performance, in which, for the sample G1p9K6-4, the maximum Ed of 78 Wh·kg−1 at Pd of 875 Wk·g−1 and high specific capacitance of 185 F·g−1 at 0.5 A·g−1 was obtained [139]. The samples were named as Gx1px2Kx3-x4, in which G stands for GO, p represents PVDF, K is for KOH, x1 and x2 signify the relative content ratio of PVDF to GO, x3 is the relative content ratio of KOH: (GO + PVDF), x4 is the reaction time, and “t” is the activation time. It was reported that extending the activation time led to having an increase in the effective crystallite size due to the loss of some amorphous carbon in the ignition during the time after completing the activation process. SEM image (Figure 9a) shows the highly porous structures for the sample G1p9K6-4 in comparison with other samples such as G1p4K6-2 (Figure 9b), which is the indication of higher SSA mesopore volume, thus having superior capacitive behaviour of that sample. Figure 9c represents the CV curves of the symmetric SC. The capacitive performance of the HSG device is presented in different galvanostatic charge–discharge current densities and various electrolytes through Figure 9d. The sample G1p9K6-4 having more SSA, revealed a higher Pd and Ed (Figure 9e), and a larger pore volume (Figure 9f) of 2.05 cm3 gr−1 in nano carbons, which ease the ion transport through the mesopores [139].
The graphene nanosheets (GNSs) could be used as electrode material for SC in 2D (also known as graphene paper), or 3D structure. In the case of 2D graphene, the layers, stacking could occur owing to the interplanar π π   interactions and van der Waals forces between the graphene sheets, which leads to a decrease in SSA, limiting the ion diffusion from the electrolyte, thus decrease in the electrochemical performance. There are various strategies to address this problem such as template-assisted growth [140], adding spacers [141,142], and crumpling the graphene layers [143].
For instance, Wang et al. [141] increased the distance between the graphene layers by incorporating a small quantity of the pillared carbon black (CB) nanoparticles, which are used as a spacer. Therefore, the electrochemical properties of the flexible graphene papers improved owing to the resultant open structure, which eases the charge storage and ion transfer.
Besides, graphene-based hydrogels are another class of electrode material that exhibited excellent electrochemical performance for SC owing to their highly porous structure and abundant active sites. Recently, an FSC device was fabricated using graphene-based composite hydrogel. The composite was obtained through the modification of the functionalized GQDs. The resultant electrode provided an excellent capacitance value of 451.7 F·g−1 and good mechanical flexibility.After 10,000 cycles, 89.0% of capacitance remained [144].
Besides, a redox SC was fabricated using the graphene-based aerogel material as the negative electrode (Nile Blue-conjugated graphene aerogel (NB-GA)), which exhibited excellent pseudocapacitive performance in acidic, neutral and alkaline electrolyte, as well as a high specific capacitance, more than 483 F·g−1 [142]. Incorporating the pseudocapacitive material between the GNSs prevented the agglomeration of the graphene layers, thus increasing the SSA and the supercapacitive performance. Figure 10 [143] shows that the Nile Blue material fulfils the requirements to be suitable as redox-active material including (a) well physical and chemical conjugation with the graphene surface, (b) having electrochemical behaviour, (c) staying stable over the whole range of PH, etc.

3.2. Transition Metal Oxides/Hydroxides-Based Materials

TM-based compounds are the most common redox-active materials, which are used as electrode materials for SCs, mostly HSCs, specifically in oxides and hydroxides form. Generally, there are various TMs, which are mainly used in the structure of TMOs and transition metal hydroxides (TM(OHs)) such as Ni, Fe, Co, Ti, Mo, V and Nb [145].
TMOs are one of the alternative materials for SCs electrodes with high energy and power owing to their higher specific capacitance and lower resistance in comparison with carbon material, as well as their high conductivity [146]. TMOs and CPs are the most common redox pseudocapacitive materials for owing to one of their properties of showing fast-reversible redox reactions, which lead to high specific capacitances and their long operation cycles [9]. There are different TMOs, as B.E. Conway defined, such as RuO2, Fe3O4, MnO2, NiO, Co3O4, IrO2 [147] as well as V2O5 and CuO, [146] in which their charge storage mechanism is based on the faradic behaviour.
Furthermore, there are several examples of various mixed-MOs such as CuO@NiO [148], CuO/Cu2O [149], Cu/Cu2O@TiO2 [150], Fe2O3@TiO2 [151], ZnO@Co3O4 [152], Fe2O3/Co3O4 [153], Fe2O3/NiCo2O4 [152], Co3O4/ZnFe2O4 [154], Co3O4/NiCo2O4 [155], ZnO/ZnFe2O4 [156,157,158], Cr2O3@TiO2 [159], NiFe2O4/Fe2O3 [160], NiO@ZnO [161], and ZnO@ZnCo2O4 [162], in which by developing a high surface area and the involving the multiple functionalities in the metal organic framework (MOF), they could be fabricated as electrodes for SCs to provide a high performance. Figure 11 shows various precursors that could be utilized to derive the porous MOs along with the corresponding surface area [163].
Besides, the layered transition metal hydroxide (TM(OH)s) such as Ni(OH)2, FeOOH, and Co(OH)2, with a large space between their layers provide suitable the redox reactions, enhancing the capacitance value of the electrodes in SCs. However, the TM(OH)s could offer more charge storage in comparison with their oxide form, the electronic conductivity of their corresponding oxides is higher. TM(OH)s have exhibited lower structure stability, and consequently, lower rate capability or less cycling performance compared to their corresponding oxides [197]. However, there are several strategies to address this limitation and improving the cycling stability of the material, such as layered engineering of the TM(OH)s through the anion and cation doping process. For instance, the cycle life of the α-Ni(OH)2 can be improved through the partial substitution of other TM cations such as Co2+, Ni3+, etc., to obtain bimetallic complexes [197].
These species attract the anions to insert into the layer spacing of α-Ni(OH)2 [198]. The anion insertion improves the cycle stability of the material by altering the structure of the primary α-Ni(OH)2 that makes the material to act as a pillar to keep structural stability. The intrinsic pillar effect was stated for the first time by Y. Chen et al. [197]. They fabricated an ASSC using MNC and graphene as electrode material (Figure 12a). It was reported that the structural stability of α-(Ni/Co)(OH)2 was improved owing to the pillar effect of the metaborate groups, which, unlike the external pillar groups in the previous works (such as SO42−, NO3−, and Cl), provides a strong connection between (Ni0.5Co0.5)(OH)2 layers through the strong B-O bonds. (Figure 12b) Besides the enhanced stability, the highly pGr with excellent conductivity led to increasing the Ed and cycling performance (Figure 12c) of the resultant electrode [197].
As it was mentioned in the introduction section, the battery-type TM can be classified into two main groups depending on their electrochemical reaction in HSC device, including intercalation-type and conversion-type (Figure 13) [199]. However, it is worth mentioning that, although TM-based materials are mainly categorized as battery-type materials, they also exhibit capacitive behaviour. For instance, Feng et al. fabricated the MnOx electrode materials by a simple electrodeposition method, which exhibited excellent capacitive performance through tuning the concentration of complexing agents in the plating electrolyte. They reported that the liquid-phase and solid-phase diffusion were facilitated due to the formation of the highly porous nanostructure and enhancing the contents of hydrous states such as H−O−H, and Mn−O−H for the thick oxide layer [200].

3.2.1. Ruthenium Oxide (RuO2)

Ruthenium oxide (RuO2) is one of the most studied TMOs for SC electrodes due to its superior chemical and thermal stability, high metallic type conductivity, broad potential window (1.2 V), highly reversible redox reaction, long cycle life, high theoretical capacitance (1200–2200 F·g−1), and good rate capability [201].
Furthermore, the superior properties of RuO2 such as its metallic conductivity, field emission behaviour, catalytic activities electrochemical reduction-oxidation properties in both crystalline and amorphous structures, make the RuO2 to be useful for various practical and theoretical applications, including a thick or thin resistor, ferroelectric films and integrated circuit development, and electronic applications, and the most recent in SCs [202].
In addition, such parameters as annealing temperature, particle size, and the crystallinity affect the pseudocapacitive performance of RuO2, [115,203], whereas the higher crystallinity of RuO2 leads to lower specific capacitances because of the compact structure of RuO2 that limits the insertion/depletion of ions and electrons, resulting to an improved electrochemical impedance [36]. Accordingly, most researches were focused on the amorphous RuO2 since the faradic redox reactions on both the exterior surface and the bulk of the material deliver a higher specific capacitance [36].
RuO2 materials are expensive, which could limit their large-scale applications. In order to develop cost-effective materials, the RuO2 hybrids were coupled with other pseudocapacitive or carbonaceous materials to reduce the amount of RuO2 without influencing the activity [204]. For instance, a nanosized RuO2/polyaniline (PANI)/carbon double-shelled hollow spheres nanocomposite was designed through an in situ electro-polymerization process [205].
RuO2 has three oxidation states as Ru4+, Ru3+, and Ru2+ [125,203,206]. The rapid-reversible redox reaction of RuO2 is shown by Eqquation (2) [207]:
RuOx(OH)y + δH+ + eH+ → RuOx-δ (OH)y+δ⋯(0˂δ˃2)
There are different methods to synthesis the RuO2 for SC electrode materials. For instance, Zheng et al. [208] synthesized a hydrous RuO2 through a sol-gel technique, which provided a specific capacitance of 720 F·g−1 at the annealing temperature of 150 °C.
In another research, RuO2 was prepared through electrodeposition technique for SC application resulting in a stable electrode with a large number of cycles and a specific capacitance of 498 F·g−1 at a scan rate of 5 mV·s−1 [209]. Kim et al. [210] also synthesized RuO2 loading on carbon via the colloidal method resulting in high specific capacitance of 407 F·g−1 obtained at a 40% RuO2 loading. The charge storage mechanism of RuO2 is based on four steps, as the electron hopping (a) within RuOx·nH2O, (b) between RuOx·nH2O particles, (c) between electrodes and the current collectors, and d) proton diffusion without RuOx·nH2O particles [36,211].
CV curve for RuO2 was almost rectangular, indicating a suitable capacitive behavior, and compared to another electrode material, RuO2 displayed lesser ESR, and consequently, higher Ed and Pd than conductive polymer-based SCs and EDLCs. However, their high cost and poor performance at high current densities [104] limit their application, whereas the researchers focused on reducing the loading of RuO2 via doping with other materials.
The RuO2-GC electrode was fabricated using the GNSs and CNTs, which were deposited onto CNFs. The combination of the RuO2 and carbon-based material provided a specific capacitance of 480.3 F·g−1 at 0.6 A·g−1. Furthermore, the composite-based SC showed a high energy density of 30.9 Wh·kg−1 at a power density of 14,000 W·kg−1 [212].

3.2.2. Manganese Dioxide (MnO2)

MnO2 is one of the most used electrode materials for SCs due to its superior properties such as outstanding capacitive performance in the aqueous electrolytes, low cost, and low environmental impact, which also make it useful for a broad range of applications such as in biosensor, molecular adsorption, energy storage, ion exchange, and catalysis [213,214].
The lesser cost and higher availability of MnO2 in comparison with RuO2-based materials [215,216,217], as well as another superior of properties such as a broad electrochemical potential window, low toxicity and a high theoretical specific capacitance (1370 F·g−1), makes it a promising alternative to replace as electrode materials [218,219,220].
The reversible redox reaction and the existence of various oxidation states for MnO2 lead to a high pseudocapacitance charge storage mechanism which takes place through two reversible mechanisms involving insertion/reinsertion of protons into the bulk of the electrode and secondly, adsorption/desorption of electrolyte cations onto the electrode surface that occur as the transition between III and IV oxidation states of MnO2 [219]. Equation (3) represents the charge storage mechanism as follows:
MnO2 + C+ + e → MnOOC
The chemical and physical factors, such as the morphology of the MnO2-based material, play an essential role in the electrochemical performance of MnO2, such as cyclic stability, which is highly dependent on the crystallinity of the porous MnO2. The nanocrystalline and the amorphous MnO2 provide easy accessibility to ion and cation diffusion owing to the porous nature of these compounds [221]. It has been reported that among all crystalline phases, α and δ show the highest specific capacitances [222].
For instance, Godbane et al. [223] considered the effect of the various structure on the charge storage performance of MnO2, which was prepared in a different allotropic phase such as 1D channels, 2D layers, and 3D interconnected tunnels and reported an optimal specific capacitance 241 F·g−1 for 3D interconnected tunnels and the lowest specific capacitance of 28 F·g−1 for 1D channels. It was also reported that the electrochemical performance was not affected by SSA.
In another study, Sun et al. [224] showed that the energy storage mechanism relied on the number of manganese (Mn) centres as active sites and reported that the enhancement in the percentage of active Mn centres by the decrease in crystal size leads to a higher specific capacitance and in contrast with the previous assumptions, the results in this research showed that the morphological surface to volume ratios, (or SSA) affects the electrochemical performance of MnO2-based SCs. Furthermore, the physical and chemical factors affect the electrochemical behaviour of the MnO2, as the hydration state of MnO2 has a significant influence on the specific capacitance, and the cyclic stability is being dependent on the morphology of the resultant material [225].
Among all various types of MnO2 nanostructure, the highest Ed and Pd were reported for tubular MnO2 nanostructure of about 21.1 Wh·kg−1 and 13.33 W·g−1, respectively. Despite these superior properties, SC, including MnO2-based electrodes, suffers from low specific capacitance values at high charge–discharge rates and insufficient cycling stability owing to MnO2 dissolution, low conductivity, broad band gaps, and low ionic diffusion constants [226,227].
According to these issues, the flower-like MnO2/nitrogen-doped graphene nanocomposite was synthesized via a hydrothermal technique, which offered an average specific capacitance of 220 F·g−1 at of 0.5 A·g−1 as well as capacitance retention of 98.3% after 3000 cycles [228].

3.2.3. Nickel Oxide (NiO)

Micro/nanomaterials such as nickel oxide have fascinated a lot of attention by the researcher as electrode material for electrochemical capacitors because of the shortened diffusion paths, fast redox reactions, and a large SSA in the solid phase [229,230,231,232,233].
NiO is one of the other promising pseudocapacitive materials for SCs electrodes owing to its properties such as, the high theoretical specific capacitance of 2584 F·g−1 [234,235,236], thermal stability, chemical stability, facile synthesis, reasonable cost, abundance, and being environmentally friendly [237]. The pseudocapacitive nature of NiO and Ni(OH)2 are debatable based on previous literature reports [39]. The NiO material can exhibit pseudocapacitive behavior. For instance, the NiO nanoflakes synthesized through the microwave-assisted method displayed a high pseudocapacitive performance as electrode material for supercapacitor, as well as a high conductivity of 33.87 Scm−1 at room temperature, and surface area as high as 206 m2g−1 [238]. In another study, a Co–Ni/Co3O4–NiO nano-heterostructures demonstrated a remarkable pseudocapacitance [44].
NiO, as an electrode material, has a lot of benefits in terms of electrochemical strategy, including accuracy, reliability, excellent electrochemical reaction reactivity, and simplicity [237].
NiO was obtained through the transformation of nickel hydroxide using the electrochemical strategy. NiO electrode resulting in this method showed an excellent specific capacitance of 1541 F·g−1 in 1M KOH (at the scan rate of 5 mV·s−1). After 3000 cycles, 91.8% of the capacity retention remained [239]. NiO could be utilized as electrode materials for SCs in the various forms such as nanowires, nanotubes [240], nanoparticles [241], nanoflowers [242], hierarchically nanoporous films [243], nanorods, and hollow nanofibers.
As other electroactive material, the electrochemical performance of the in the NiO material also highly depends on the morphology of the nanostructure while the porous structure provides a large active surface area that corresponds to the further faradic active sites. This structure has also been shown a very short pathway for ions. Besides, the NiO nanorod arrays offer high cycling stability as well as an ultra-high capacitance of 1536 F·g−1 and 2018 F·g−1 at high and low current density [244], respectively, which may be associated with the well-aligned arrays, narrow nanorod and the stable chemical bonding of NiO nanorods on the Ni substrate [237]. Furthermore, NiO nanofibers/Ni [245] provided almost 100% capacitive retention after 8000 cylces, also in the case of NiO nanoflowers [246], 99.7% of the capacitance remained after 1000 cycles.
In another research, a hierarchical shell-core structure was prepared using NiO nanospines and carbon as an inner core through hydrothermal reaction. NiO nanospines offered unimpeded diffusion paths for ion transport that increase the affinity of the electroactive materials to contribute to the redox reactions, whereas the inner carbon core provides physical support to prevent the volumetric variations of NiO nanospines which is associated with its good mechanical property. The obtained electrode showed a high specific capacitance of 1161 F·g−1, a current density of 2 A·g−1, and even high capacitance retention of 92.4% was obtained after 3000 cycles at a current density of 2 A·g−1 [247]. A sandwich-like nickel oxide/nanoporous nickel/metallic glass (NiO/np-Ni/MG) composite electrode was provided through the free-dealloying method in O2-rich HF solution. The free-standing electrode showed a high volume-specific capacitance of 745.3 F·cm−3 at 0.5 A·cm−3 as well as the capacitance retention about 92% after 6000 cycles, which implies on a relatively excellent cyclic stability [248].

3.2.4. Nickel Hydroxide (Ni (OH)2)

Among TM(OH)s, Nickel hydroxide has recently attracted attention from researchers owing to its excellent properties such as a high theoretical specific capacitance of 2082 F·g−1, good rate capability, ready availability, lower cost, environmentally benign nature, and its excellent thermal stability and high chemical stability in strong alkaline electrolytes [249]. Accordingly, α-Ni(OH)2 is the most powerful battery-type material for SC application, which is not only used as positive electrode material in nickel-based batteries, but it is also suitable for electrochemical SCs [250,251,252]. The electrochemical performance of nickel-based pseudocapacitor electrode materials is estimated based on the redox reaction of Ni(OH)2 or NiO in alkaline electrolytes, and it can be shown in two general equations for NiO and Ni(OH)2 in Equations (4) and (5), respectively [237].
NiO+ OH ↔ NiOOH+ e
Ni(OH)2+ OH ↔ NiOOH+ H2O + e
The supercapacitive performance of the nickel-based materials highly depends on its loading percentage on the electrode, the electrolyte, the morphology and porosity of the material as well as its conductivity and the surface area, whereas the more enhanced depth of the electrochemical process and larger SSA can provide a higher specific capacitance [253,254].
Furthermore, the performance of the Ni(OH)2 depends on several parameters such as the type of microstructure, the precipitation condition, structural defects, and crystallinity [208,255,256,257], whereas the crystal structure and crystallinity of Ni(OH)2 have a strong influence on the H+ mobility and conducting the charge–discharge rate [258]. For instance, Lang et al. [259] prepared a roughly packed Ni(OH)2, in which the poor crystallinity nanoflakes were shaped through the chemical precipitation method, and the synthesized sample exhibited a high specific capacitance of 2055 F·g−1. The enhanced capacitance and diffusion rates are due to the availability of OH ions in the electrolyte for the Ni(OH)2 nanoflakes.
Furthermore, nanoflower-like Ni(OH)2 was deposited onto a Ni foam to be used as an electrode for SCs, and the obtained electrode showed an outstanding rate capability with a specific capacitance of 1065 F·g−1 [260]. In another example, amorphous Ni(OH)2 nano-spheres deposited on a graphite rod and cycled at current densities of 20–70 A·g−1 in 1 M KOH, exhibited a high specific capacitance in the range of 1868–1330 F·g−1 that arises from the amorphous structure and the uneven surfaces [261].
Such as other MOs, besides all the advantages, Ni(OH)2, also has several drawbacks, like poor conductivity, low stability, and poor cycle performance that occur because of the massive volume alteration during the charging-discharging mechanism. Heteroatom doping of cobalt (Co) and zinc (Zn) could be utilized to improve the electrochemical performance of Ni(OH)2 [262,263].
Ni(OH)2 could restrict the rate capabilities due to its semiconductor nature, and the aggregation of Ni(OH)2 particles leads to decreasing the electrochemical performances. Different nanocomposites have been produced by highly conductive materials such as graphene and activated carbon to address these issues [264]. For instance, the Ni(OH)2/graphene composites were prepared using a chemical precipitation method in which the loosely 3D structure leads to more active sites and, consequently, more redox reactions. The coupling of graphene leads to enhanced stability and the conductivity of Ni(OH)2 [265].
The graphene sheet with good conductivity provides a larger surface area, whereas the increased interfacial contact with the electrolyte solution governs the electron transfer and the diffusion rate to the enhanced electrochemical performance. The obtained GNSs/Ni(OH)2 composite exhibited a specific capacitance of 2053 F·g−1, and 97% of the capacitance remained after 1000 cycles.
Ni(OH)2 composites could also be coupled with other oxides to increase the electrochemical properties. For example, Co(OH)2-Ni(OH)2 composites were synthesized to achieve a higher specific capacitance in comparison with the Ni(OH)2 or NiO in which this enhancement is attributed to the mesoporous structure and faster ion diffusion in this regular nanostructure [266,267]. After 1000 continuous charge–discharge cycles, the capacitance retention increased from 41.6% for the pure Ni(OH)2 electrode to 88.2% for the composite-based electrode with the weight ratio of 3:2. The weight ratio of Co(OH)2/Ni(OH)2 profoundly affects the SC behaviour [267].
As it was mentioned in Section 2, the phase characteristics of the electrode material could be modulated through the pillar effect.
Recently, Xiao et al. [268] fabricated the pillar-coordinated α-Ni(OH)2 (Ni-(BO2−)-LDH) by exploiting the pillar-coordinated strategy through the etching process of the Ni-MOF precursor, in which the organic ligand was substituted by OH and BO2− anions. The maintained MOFs structure provided a highly porous frame and abundant channels, which improved the electrolyte diffusion. They have described that the structure of α-Ni(OH)2 was supported owing to the pillar coordination, which offered a more stable system with a durable redox reaction in strong alkaline electrolyte. The enhanced Ni(BO2−)-LDH-4 exhibited a specific capacitance of 1760 F·g−1 at 1 A·g−1, and capacitance retention of 81% after 10,000 charging–discharging cycles. The HSC fabricated using the resultant Ni(BO2−)-LDH-4 and 3D-rGO as electrodes, has exhibited a high Ed of 56.5 Wh·kg−1 at the Pd of 0.111 kW·kg−1 (Figure 14) [268].
Besides, the electrochemical performance of the Ni(OH)2 (NSs) could be improved through growing the material on the carbon-based substrate to take advantage of the highly porous structure of these materials. For instance, Ni(OH)2 NSs were deposited on the electrochemically activated carbon cloth (ECC) to fabricate the Ni(OH)2/ECC (HT) composite (ECN-3), which was fabricated through the one-step ethanol-assisted supercritical deposition process (supercritical fluid deposition). Due to the high permeability, good solvent capability, and low surface tension of the supercritical ethanol fluid, it has the ability to propagate the reactant onto the porous structure of CC; thus the Ni(OH)2 NSs were distributed perfectly on the CC. This provided significant electrochemical performance for the Ni(OH)2 NSs/ECC flexible electrode. The accessibility of the redox sites to electrolyte ions increased owing to the resultant porous structure and obtained channels. The fabricated flexible solid-state SC (including ECC as the anode, and ECN-3 as the cathode), showed an Ed of 8.3 mWh·cm−3 at the Pd of 20 mW·cm−3, as well as the capacity retention of 87% after 10,000 cycles at 50 mA·cm−2 [84] (Figure 15).

3.2.5. Vanadium Pentoxide (V2O5)

Vanadium (V) oxide or vanadia is an inorganic-intercalation compound with the formula V2O5, which is generally known as vanadium pentoxide. V2O5 has attracted a lot of interest from researchers as a suitable electrode for pseudocapacitors in aqueous and organic electrolytes because of its natural abundance, high Ed, low cost, low toxicity, its mixed oxidation states (+2 to +5), ease of fabrication, and high capacitance [269,270]. The specific capacitance of 350 F·g−1 was reported for amorphous V2O5 in KCl aqueous electrolyte [271]. V2O5 shows an ultra-high theoretical capacitance of 2120 F·g−1 (under the wide potential window of 1 V), that is associated with the higher oxidation state of vanadium, which leads to transfer a higher number of electrons [40]. Nevertheless, low SSA and the low solubility of V2O5 limits the specific capacitance improvement. Furthermore, V2O5 is diffusely recommended as electrode material for SCs due to its controllable morphology. There are different methods to obtain the various morphology and structure of V2O5, such as electrospinning [272], the hydrothermal method [273], template-assisted growth [274], the sol-gel method [275], and so on.
For instance, the V2O5 nanofibers were synthesized using the electrospinning method, that provided the specific capacitance of 190 F·g−1 and 250 F·g−1 in 2 M KCl electrolyte and 1 M LiClO4 electrolyte, respectively [276]. Besides, the V2O5 nano-waxberries, nano-wires, and nanoflowers were prepared through the adjustment of the concentration of surfactant using the hydrothermal method [277]. In another study, the V2O5 nanoporous were synthesized by a facile capping-agent-assisted precipitation process, which resulted in an interconnected nanoporous network. The resultant nanochannels facilitated the accessibility of ions and improved ion diffusion. The achieved V2O5 showed the maximum specific capacitance of 316 F·g−1 and 76% of capacitance retained after 600 cycles [278].
The annealing temperatures have a significant effect on the nanomaterials structure and the electrochemical property of the prepared electrodes using such materials. For instance, the electrochemical performance of the different V2O5 films was investigated at different temperature. The specific capacitance of 346 F·g−1 at a scan rate of 5 mV·s−1 was reported for the V2O5-based material annealed for an hour at 300 °C which was much higher than those which were annealed at 400 °C (177 F·g−1), 350 °C (226 F·g−1), and 250 °C (241 F·g−1) [279]. In another study, V2O5 nanotubes were synthesized through the hydrothermal method, that showed a high specific capacitance of 417.3 mF·cm−2 at a scan rate of 5 mV·s−1 in 0.5 M Na2SO4 and suitable cycling stability with capacitance retention of 80% after 3000 cycles [280].
As it is shown in the previous examples, V2O5-based materials suffer from poor cycling performance and slow rate stability, which are due to the structural instability of V2O5 during the charge–discharge mechanism, its low conductivity (10−2 to 10−3 S cm−1), and slow electrochemical kinetics. To solve the problem associated with the low electrical conductivity of the V2O5, which is the most critical complication, researchers focused on shortening the diffusion length and electrochemical kinetics improvement of the electrolyte [281,282]. Furthermore, carbonaceous materials composites and metal fibres have been prepared to enhance the electronic conductivity of V2O5 [283].
In recent decades, to enhance the physicochemical performance of V2O5, researchers have focused on the development of the V2O5-based nanocomposite materials that can deliver supreme advantages of pure material [284]. Generally, the V2O5-based nanocomposites can be categorized into two main groups, including V2O5-carbon nanocomposites as well as V2O5-metal element nanocomposites such as V2O5/CNT [285], V2O5/rGO [286], V2O5/polypyrrole [287], and V2O5/SnO2 [288].
The Ed and the perdurability of the material could be improved through the synthesis of the composite materials. Heterogeneous structures play a significant role in composites [289].
Recently, the rGO/VO2 heterostructure was fabricated through a facile hydrothermal process. The resultant material showed a superior characteristic, (in the voltage range between 0.01–3 V (vs. Li/Li+)), to be used as the anode material, due to its high specific capacity, favourable rate capability, which arises from the mixed valency of V ions (V3+ and V4+) within the as-prepared VO2 matrix that eases the faradic reactions at a low potential. The battery−supercapacitor hybrid (BSH) device was assembled using the rGO@VO2 as the anode, and the AC@CC as the cathode (Figure 16a). The resultant BSH device provided a high Ed of 126.7 Wh·kg−1 as well as a supreme gravimetric Pd of about 10,000 W·kg−1 (Figure 16b), in the working voltage range between 1 V to 4 V. Figure 16c designates that the 2D VO2 nanoribbons are tightly attached to the large 2D rGO nanosheet [80].
The graphene-decorated V2O5 nanobelts were synthesized using the hydrothermal method, which revealed a high capacitance of 288 F·g−1 at a scan rate of 10 mV·s−1 and 82% of the capacitance remained even after 5000 cycles [290]. Chen et al. [291]. synthesized the V2O5-based composites in the nanowire structure, interpenetrating V2O5 into conductive porous CNT frames using an in-situ hydrothermal method. The prepared nanocomposite-based electrode exhibited a specific capacitance up to 440 F·g−1 at 0.25 A·g−1. The ASC device developed based on the V2O5/CNT hybrid composites showed a high Ed of 16 Wh·kg−1 at a Pd of 75 W·kg−1.
Furthermore, the graphene/V2O5-xerogel nanocomposites were fabricated utilizing a hydrothermal approach. The prepared electrode provided a good specific capacitance of 195.4 F·g−1 at a current density of 1 A·g−1. The electrical conductivity improvement was due to the presence of the graphene, which also improved the SSA of the V2O5-xerogels approximately twice in comparison with the V2O5-only xerogels [292]. Besides, Zhang et al. [293] synthesized the 3D porous V2O5 hierarchical spheres consisting of stacked platelets through the precipitation of precursor microspheres in aqueous solution at room temperature and the subsequent calcination. The as-obtained V2O5 hierarchical spheres showed an excellent high-rate capability. The porous V2O5 hierarchical spheres exhibited a high capacity of up to 473 C·g−1 at 0.2 A·g−1 as a battery-typed electrode.
Furthermore, crystalline α-V2O5 powders were successfully exfoliated into wrinkled V2O5 nanoribbons through a novel processing method based on specific microwave heating and were used to form a strongly bound V2O5/rGO hybrid material using a simple cation-mediated self-assembly mechanism and further thermal annealing treatment. The yield nanoribbons showed a specific capacitance of 423.6 F·g−1 at 1 A·g−1. The SSC developed using the hybrid material provides a high Ed of 58.8 Wh·kg−1 at and an average Pd of 500 W·kg−1 as well as an excellent high-rate capability. The specific capacitance remained nearly constant after up to 1800 charge–discharge cycles [294].
In another study, the reduced GO/vanadium pentoxide (V2O5)/PANI (GVP) nanocomposite was fabricated through in situ single-step chemical process, and the as-prepared material showed a high Ed of 54.62 Wh·kg−1, the specific capacitance of 273 F·g−1 (specific capacity of 328.5 C·g−1), and Pd of 1636.5 W·kg−1 at a current density of 1 A·g−1. The symmetric and asymmetric SCs using V2O5-based composites as electrode displayed high Ed and a wide voltage window that could compete with batteries [295].
Table 1 summarizes the electrochemical measurement results of different metal oxide-based electrode materials and their corresponding fabricated SCs using various types of electrolytes (e.g., liquid, solid/quasi solid-state, and redox-active electrolytes). As it is shown in Table 1, various types of supercapacitors, including SSCs, ASSCs, FSCs, and HSCs can be fabricated utilizing different TMOs and TM(OH)s composited with highly porous and conductive materials such as carbonaceous (e.g., CNTs, CCF, CB, etc.) and CPs, which provide rapid ionic diffusion as well as ion transport owing to their inter- and intraplanar pores. In several examples, the electrical conductivity of the TM(OH)s and TM(OH)s composite have been increased by utilizing the 2D nanomaterials (graphene, GPCNs, etc.), which leads to faster charge transfer owing to the 2D structure. The capacitive performance of the composite-based electrode materials and their corresponding supercapacitor was improved by introducing TMOs and TM(OH)s to carbonaceous materials which increased the hydrophilicity and the conductivity of carbon-based materials offering both EDLC and faradic behaviour to the resultant composite. Besides, in comparison with different mixed TMOs, the ternary metal-based nanomaterials (composed of three various MOs), exhibited excellent electrochemical performance owing to the combination of advantages between three components that can display extraordinary structural stability and synergies.

3.3. Conducting Polymer-Based Material (CPs)

Different CPs are another pseudocapacitive electrode material that has been widely researched because of their relatively high capacitance, high Ed, adjustable redox activity through chemical modification, good conductivity in doped states, high voltage windows as well as facile production, low environmental impact, and low cost [327,328,329].
CPs can store the charge in its bulk, since no structural alterations such as phase changes occur during the charge/discharge mechanism. Due to this, CPs can offer a higher capacitance because of the larger surface areas and redox storage capabilities [330,331,332,333].
In comparison with other carbon-based electrode materials, CPs have higher conductivity, capacitance, and low equivalent series resistance [334,335]. The specific capacitances in CPs derive from the fast-reversible redox reactions resulting from the π-conjugated polymer chains [90], in which the ions move to the polymer backbone through the oxidation process, also known as doping, and during the reduction process or de-doping, they are released back into the electrolyte solution. However, despite all the advantages, the reduction-oxidation in CPs causes mechanical stress and consequently limiting the stability and degradation through several charge–discharge cycles [13,336]. One of the other disadvantages CPs which hinders their performance is the low power densities, arise from the slow ion diffusion rates in bulk [337].
Sharma et al. [338] perceived a 50% decline in initial capacitance of the PPy-based SC electrode, just after 1000 cycles. Accordingly, the loss of Cs of PANI nanorods was about 29.5% after 1000 cycles. Therefore, the main concern related to CP-based electrodes is low cycling stability.
There are various methods for the synthesis of CPs, and among all, the oxidation of monomers via chemical or electrochemical approaches is the most commonly used approach [201].
The synthesized polymers arise in either two or three general states. As-formed polymers tend to be in the oxidized or “p-doped” state rather than the reduced or “n-doped” state. In the p-doped state, the polymer backbone is positively charged and exhibits a high electronic conductivity (between l to 100 Scm−1) [339]. Currently, p-dopable polymers are attracting more attention from researchers in comparison with non-dopable polymers due to their stable performances [330].
The CP-based electrodes for SC exist in three various configurations including (a) the p–p configuration in which both electrode materials are the same p-doped polymer (b) the SCs n-p type in which both electrode materials made from the same polymers, but one is positively charged (p-doped) and the other one is negatively charged (n-doped) electrodes and (c) p-doped and p–p′ SCs containing two various p-doped CPs with various oxidation and reduction electroactivities [334,339,340,341,342,343,344,345,346,347]. Notably, the n-p type polymer conformation has massive potential for high Ed and Pd [335].
The CPs electrode materials have limited pseudocapacitors from reaching their potential due to a lack of proficient n-doped conducting material and low cycling stability [334]. Various types of CPs commonly used for SC electrodes, including PANI, polythiophene (PTh), PPy, polyindole (PIND), polyfurane, poly(p-phenylene), poly(p-phenylenevinylene), poly 3,4-ethylenedioxythiophene (PEDOT) besides different derivatives, in which PTh is either p-doped or n-doped, while PANI and PPy can be only p-dopable, and their n-doping potential is much lesser than the reduction potential of the common electrolyte [346,347,348].
Conducting polymer hydrogels are successful candidates for FSC electrodes [349]. There are different strategies to synthesing the polymer hydrogel including (a) in situ polymerization of monomers into the insulating polymer matrix [349,350], (b) introducing crosslinkers through the polymerization process in order to develop the gelation [351]. However, the resultant CP hydrogels fabricated by the first method often provide a favourable mechanical performance; their electrochemical performance is not satisfying. The CP hydrogels synthesized through the second method have shown good electrochemical behaviour. Nevertheless, the processability and fabrication of the self-standing hydrogel with a favorable mechanical performance is difficult. Besides, synthesizing a polymer hydrogel electrolyte, which has both high electronic conductivity, and good electrochemical performance is challenging.
For instance, Yang et al. [352] exploited a simple macromolecular interaction assisted route to synthesize a self-standing CP hydrogel composed by PANI and PEDOT. The resultant hydrogel showed a high mechanical performance, good electronic conductivity, and good electrochemical properties simultaneously, owing to the synergic effect between PANI and PEDOT. The assembled all-hydrogel-state SC device exhibited good areal capacitance of 808.2 mF·cm−2, as well as a high Ed of 0.63 mWh·cm−3 at a high Pd of 28.42 mW·cm−3 (Figure 17).
Conducting polymer hydrogels could also be used as the electrolyte in FSCs due to their resistance against the deformation and physical damage. For instance, an FSC was fabricated by inserting the PEI−PVA−Bn-LiCl hydrogel electrolyte, (polyethyleneimine/polyvinyl alcohol/4-formylphe-nylboronic acid/Lithium chloride), within the two MWCNT electrodes, which showed a specific capacitance of 16.7 mF·cm−2, an enlarged operating potential window of 1.4 V, as well as high cycling stability, and superior mechanical stability [353] (Figure 18).

3.3.1. Polyaniline (PANI)

PANI is one of the most common CPs electrode material owing to its high conductivity, light-weight, superior capacitive performance for energy storage, facile synthesis, and low cost [354]. Furthermore, PANI is mechanically flexible, environment-friendly [355].
Besides these advantages, PANI is susceptible to rapid degradation in performance because of the repetitive charge–discharge cycles, (swelling and shrinkage). To address this problem the combination of PANI with carbon materials was utilized as the PANI layer, and it was coated on MOs/carbon composite to form a ternary hybrid composite (PANI/MOs/carbon) [355], which provided the reinforced cyclic stability for PANI and improved the capacitance as well [356].
Furthermore, PANI shows a wide range of electrochromic properties, such as exhibiting various colours, which is due to its many oxidation and protonation forms. These characteristics make it a useful candidate for the fabrication of electrochromic SCs [335].
Here, Wang et al. [342] synthesized PANI as the active material for an SC through an electrochemical polymerization method. It provided a high specific capacitance of 950 F·g−1 at 1 A·g−1. Here, it was reported that the array in unique nanowire structure leads to the reduction of diffusion paths, also the charge transfer resistance, which allows the material to have a high specific capacitance at high current densities.
The pathway of the polymerization (solution (Sol), emulsion (Eml), interfacial (Int)) affects the capacitance value of the CPs, whereas PANI-Int nanosphere, PANI-Eml nano-petal, and PANI-Sol nanocrystal, [357] were exhibited the Cs of, 300, 424 and 460 F·g−1, the Ed of 15, 21.2, and 23 Wh·kg−1, respectively at Pd of 200 W·kg−1.

3.3.2. Polypyrrole (PPy)

The requirement of developing energy storage devices for high-performance electrode materials can be derived from the structure-controlled conductive polymer hydrogels with mechanical flexibility and adjustable electrochemical properties. PPy is one of these conductive polymers that attract too much attention from the researchers. It has a higher density and greater flexibility compared to other CPs. It can endure a fast redox reaction for the charge storage, as well as showing a high electrical conductivity value between 10 to 500 S·cm−1 [358].
For example, a conductive polypyrrole hydrogels nanostructure was synthesized through an interfacial polymerization technique and obtained the 3D porous PPy network that showed excellent rate capabilities and high specific capacitance of 380 F·g−1 [345].
In another example, by coating the lignin with PPy, the Lignin-PPy composite was prepared through the polymerization of PPy in a process with and without the existence of methyl orange that is resulting in the development of PPy films of nanotubular and globular morphology. The experiment was conducted with the composite of carbon-based materials in nitrogen atoms by pyrolysis in an N2 atmosphere. It was reported that the SSA of the synthesized materials improved by 10 times more than that of pure carbon materials [359].

3.4. Nanocomposite-Based Materials

Nanocomposite-based electrodes integrated carbon-based material with either CPs or MOs and put them together to achieve a faradaic (chemical) and a non-faradaic (physical) charge storage mechanism in a single electrode.
Pseudocapacitive materials increase the capacitance of the nanocomposite while the SSA increases owing to the presence of the carbonaceous materials. There are various types of nanocomposites for the SC electrodes such as carbon–carbon composites, carbon-MOs composites, carbon-CPs composites and MOs-CPs composites that in the last case CNTs can be used as a supporting backbone for homogeneous distribution of CPs or MOs which leads to producing a high electric double-layer capacitance and pseudocapacitance. The specific capacitance of such categories of electrodes is higher than individual carbon, CPs, or MOs electrodes [63,360].

3.4.1. Carbon–Carbon Composites

The high capacitance in the carbon material arises from the effective SSA, which is reachable to the electrolyte ions. Therefore, improving the SSA of the carbon-based materials leads to having higher energy and Pd in SC. There are various types of carbon materials which can be used to build the carbon-based composites. In the case of graphene, the dispersion is problematic and must be improved. In order to find a solution, non-covalent functionalized graphene was synthesized [81]. Other drawbacks are associated with GO in which the material tends to restack and aggregate due to the presence of the van der Waals forces through the chemical reduction process, which leads to having a less accessible surface area to the electrolyte ions. The restacking issue can be solved by providing the intrapore for the solution using the SWCNTs [67].
The electrode fabricated using the SWCNTs inserted in graphene aerogels provided a high specific capacitance of 245.5 F·g−1 in 5M KOH. The graphene aerogel SC showed excellent cycle stability and a high rate capability [361].

3.4.2. Carbon-MOs Composites

As it was mentioned before, TMOs have low specific capacitance, poor electronic conductivity, and low electrochemical stability. The nanocomposite made of the combination of TMOs and the carbon material can improve the performance of the SCs. Therefore, in addition to the MOs and mixed MOs, there are different composites of MOs and carbon-based material, which has exhibited a higher surface area than those of pure metal oxide and carbon material [362].
One of the most important benefits of CNTs is that they can go through the surface functionalization process by various materials such as CPs and MOs, or they can be oxidized to solve the problem with the presence of some impurities. Accordingly, there is a lot of research carried out to compare the performance of the nanocomposite with pure materials. For instance, the composite of MnO2/CNT was prepared through a facile hydrothermal method, and the MnO2/CNT nanocomposite electrode showed a higher rate capability and specific capacitance in comparison with the pure MnO2 and CNT electrodes. It was reported that the high specific capacitance for the prepared nanocomposite could be due to the high SSA of the MnO2 and the highly porous structure [363].
In another study, MnO2/CNS exhibited ultra-high stability, with 96.1% of the capacitance retention at a current density of 5 A·g−1 even after 5000 cycles. It was reported that the superior capacity and cycling stability of MnO2/CNS were mainly because of the strong coupling between the carbon NSs and the in situ formed MnO2 nanosheet arrays (NSAs) with large SSA [364]. Furthermore, various hybrid materials such as Co3O4/CNTs [365], Co3O4/grapheme [366,367] and Co3O4/CNFs [368] were synthesized with a huge surface area and consequently an improved electrical conductivity.
Besides, an ASC device was fabricated using MnO2 and graphene, whereas the pure graphene was used as the anode and the composite of MnO2-coated/graphene was used as the cathode. The pure graphene electrodes showed capacitance of 245 F·g−1 at a current density of 1 mA while after depositing the MnO2, the capacitance increased up to 328 F·g−1 at the same current density [337].
A hybrid NiO/GF(graphene foam) nanocomposite was synthesized for the fabrication of a hybrid electrode for ASSCs that displayed a high specific capacitance of 1225 F·g−1 at 2 A·g−1. An ASSC was fabricated using HPNCNTs (hierarchical porous nitrogen-doped CNTs) as the negative electrode and NiO/GF as the positive electrode as well as in the aqueous KOH as the electrolyte, which showed a superior electrochemical performance due to the excellent capacitance of individual NiO/GF and HPNCNTs, and the synergetic effect between the two electrodes. For the resultant device, a high Ed of 32 Wh·kg−1 at a Pd of 700 W·kg−1 has been reported. Even after 2000 cycles, 94% of the capacitance remained [369].
Furthermore, the synthesized CoNi2S4/graphene (CNS/GR) nanocomposite exhibited a specific capacitance of 2009.1 F·g−1 at 1 A·g−1, and the capacitance could be reserved at 755.4 F·g−1 at 4 A·g−1 after 2000 cycles [370]. In another study, Perera et al. [371] fabricated a coin cell type ASC exploiting carbon fiber as a cathode and V2O5-CNT as an anode that showed Pd of 5.26 kW·kg−1 and an Ed of around 46.3Wh·kg−1.
Besides, the synthesized CNTs/NCSMnO2 composites showed a specific capacitance of 312.5 F·g−1 at 1 A·g−1 with capacitance retention of 92.7% after 4000 cycles. The ASC was fabricated using ACs (-ve electrode) and the CNTs/NCS/MnO2 composite (+ve electrode), which displayed a high specific capacitance and a stable potential window of 1.8 V with Ed of 27.3 Wh·kg−1 at Pd of 4500 W·kg−1 [165].
Another alternative is the RuO2-based composites in which the composition of RuO2 and the carbonaceous materials such as AC, CNTs, carbon aerosols, and graphene provides higher capacitive properties and lesser cost in comparison with the pure RuO2 based electrodes [372,373,374]. The MWCNT/RuO2 composite was assembled by inserting the RuO2 into MWCNTs in an acidic solution showing a maximum specific capacitance as high as 628 F·g−1 for different loading of the RuO2 films [375].
As it was mentioned before, CNTs could be utilized to support the hollow-structured nanoparticles. For instance, one of the main problems of the RuO2-based electrodes are cracking the RuO2 structure leading to poor long-term stability, which is due to the developed strain during the charge–discharge cycles [376,377]. In order to address this issue, Wang et al. [378] utilized a novel core–shell-templated method to assemble a CNT-supported hollow-structured RuO2 nanoparticle (hRuO2/CNT), in which there was no crack formation reported for the nanocomposite electrode. The fabricated electrode also has provided a high specific capacitance of 655.0 F·g−1 at a current density of 5 A·g−1. Guan et al. [379] fabricated the composite using needle-like Co3O4 deposited on graphene as electrode materials for SC. The resultant electrodes show a specific capacitance of 157 F·g−1 at a current density of 0.1 A·g−1.
Furthermore, there many examples of nanocomposites, or the single-phase materials of NixCo3-xS4 were synthesized with excellent energy storage properties. For instance, nanowires [380], nanoneedles [381], and (NSs) [382], of NiCo2S4/(GO) [383], Ni0.3Co2.7S4, NiCo2S4 hierarchical structures [384] and NiCo2S4/(CNTs) [385], etc. For instance, NiCo2S4 nanotubes were synthesized through the sacrificial templates, which showed the capacitance of 933 F·g−1 at 1 A·g−1 [386].
Ternary NiCo2S4 provides more redox reactions due to having a higher electronic conductivity in comparison with NixCo3-xS4 that decrease the charge transfer resistance, and consequently, a slighter interior resistance (IR) loss at higher current density. Therefore, a grater Pd and larger rate capability and can be obtained [376]. Besides, Ni–Co–Mn hydroxide nanoneedles were synthesized through a facile hydrothermal technique, which provided a specific capacitance of about 1400 F·g−1, higher Ed and Pd (30 Wh·kg−1 and 39 kW·kg−1, respectively), in comparison with the Ni–Co–S electrodes. There was no report of the capacitance loss even after 3000 cycles [387].
Carbon-based materials and its composites that are used for the SC electrodes can be compared in terms of various parameters, including surface area, pore size, specific capacitance, stability, rate capability, and cost. The composites consist of the MOs, and the carbon materials are one of the powerful candidates for the electrode materials, which can provide an excellent specific capacitance, good performance, and good rate capability as well as stability and less cost [376].
For instance, a ternary nanocomposite has been fabricated using the MnO2 nanorods and rGO NSs, which were reinforced on the poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) (PEDOT: PSS) (Figure 19a–e). The resultant composite (MGP) was used as electrode material for SC application and showed an improved specific capacitance of 633 F·g−1 at a current density of 0.5 A·g−1. The device remained stable 100% even after and 5000 cycles. The capacitance of the material is associated with both double-layer and pseudocapacitive mechanisms. Figure 19f,g shows asymmetrical CV curves for all different electrode materials (MnO2, MG(MnO2/rGO), and MGP), in which a larger area under the CV, dedicated to MGP, is an indication for a higher supercapacitance behaviour MGP compared to MnO2 and MG [388].
Besides, W. Peng fabricated the NiCo2S4/rGO nanospheres composite through a one-step hydrothermal method. The resultant composites showed an excellent specific capacitance of 1406 F·g−1 and outstanding cycle stability of 82.36% after 2000 charge–discharge cycles at the current density of 1 A·g−1, that were higher than the pure NiCo2S4 material (792 F·g−1 and 64.77% after the same number of cycles) [389].
In another study, iron cyclotetraphosphate (Fe2P4O12)–carbon composite was utilized as an electrode material for SCs. They reported that the phosphorus doping in the porous carbon structure and the presence of the functional groups in the phytic acid-derived carbon had increased the redox couple, thus improving the pseudocapacitive performance of the composite, which attributed to the highly reversible redox reaction. However, capacitive behaviour has been reported as the predominant charge storage mechanism, which mainly occurred in the acidic medium. They have displayed that the charge storage behaviour and the stability of the composite do not only depend on the electrode material but also are highly dependent on the electrolyte selection. Fe2P4O12 composite exhibited high stability in the 0.5 M H2SO4, a capacitance of 251 F·g−1 at a current density of 1 A·g−1, as well as a constant capacitance without any loss even after 9000 cycles [390].

3.4.3. Carbon-CPs Composites

There are different carbonaceous materials that can be coupled with CPs to form a nanocomposite. For instance, the nanocomposite using the activated carbon as the cathode and CPs as the anode can provide a higher Ed and Pd in comparison with EDLCs and an enhanced cycling performance compared to pseudocapacitors [360].
As an example, a composite electrode was fabricated through the colloidal self-assembly method using graphene-MWCNT-PPY nanofibers [83,391,392,393,394,395].
The freestanding PEDOT-PSS/SWCNTs composites exhibited the specific capacitance of 104 F·g−1 at 0.2 A·g−1, Pd of 825 W·kg−1, and Ed of 7 Wh·kg−1. Furthermore, 90% of the capacitance left after 1000 cycles [396].
In another study, PEDOT/PSS and MWNT composites showed a specific capacitance of 100 F·g−1 [397].
There are various types of electrodes fabricated using PANI. However, the low rate capability and relatively poor cycle stability limit their applications; still, single-phase PANI materials have been extensively considered as promising electrodes for SCs. There are various nanocomposites formed through the coupling of PANI nanostructures and carbon-based materials that show a better performance than that of PANI or carbon-based electrodes in which carbon material plays an important role because they can be considered as the efficient stabilizer for PANI materials with high availability in different morphologies [398].
For example, rGO aerogel-PANI composite was synthesized through the electro-deposition method in which the PANI arrays deposited on the rGO aerogel. Generally, the poor cyclability, Ed, and flexibility of rGO aerogels limit its application. The prepared composite has shown a rich open pore network and high capacitance due to the unique capacitance of the PANI as well as a high conductivity which arise from the cross-linked framework of 3D aerogel [399].
The specific capacitance of the resultant composite was reported of 432 F·g−1 at a current density of 1 A·g−1, and the stability of the materials in various bending conditions was outstanding. It provided an Ed of 25 Wh·kg−1, and 85% of the capacitance remained after 10,000 cycles [399]. In another study, an FSC, was fabricated using the prepared composite of PANI/etched carbon fiber cloth (eCFC), which exhibited 88% capacity retention and a higher specific capacitance (1035 F·g−1 at 1 A·g−1), compared to that of PANI nanowires at all current densities [400].
The graphene/PANI nanofiber composite was synthesized through the one-step in-situ polymerization method showing the specific capacitance of 526 F·g−1 at 0.2 A·g−1 current density [401]. The fabricated GNSs/PANI composite showed capacitance of 532.3 F·g−1 at 2 mV·s−1, and 99.6% capacitance has remained at 50 mV·s−1 [402]. Besides, a CNT hydrogel coupled with PANI has shown a specific capacitance of 680 mF·cm−2 at 1 mA·cm−2 [403]. In another study, the synthesized PANI nanorods/graphite NS composite showed a specific capacitance of 1665 F·g−1 at a current density of 1 A·g−1 [404].
Han et al. [405] fabricated electrodes using the composite of PEDOT/PSS and GO in 1M H2SO4, showing a specific capacitance of 108 F·g−1 and capacity retention of 78% after 1200 cycles. Furthermore, the symmetric (PPy/PPy) and asymmetric (PPy/AC) SCs have been fabricated utilizing Cladophora algae-derived cellulose as a binder which has exhibited the capacitance in the range of 0.45 F to 3.8 F [406].
In another study, three different composites have been synthesized through in-situ polymerization using GNSs, CNTs, and PANI. The GNSs/PANI composite exhibited a higher specific capacitance of 1046 F·g−1. The capacitance for the GNSs/CNT/PANI and CNT/PANI composites were reported of 1035 F·g−1 and 780 F·g−1. The capacitance reduced only 6% for GNSs/CNT/PANI, 52% GNSs/PANI, and 67% for CNT/PANI after 1000 cycles [407].
Polyphenylene diamine and hydrogen exfoliated GNSs have been prepared, which exhibited a specific capacitance of 248 F·g−1 at 2 A·g−1, and the assembled electrode showed an Ed of 8.6 Wh·kg−1 and a Pd of 0.5 kW·kg−1 [408].
In another study, the prepared graphene-based PVA composites showed 10 times enhancement in Young’s modulus with a tensile strength 150% better as well as the graphene loading of 1.8 vol % [409]. The PEDOT/PSS CNTs composite showed a specific capacitance in the range of 85 F·g−1 to 150 F·g−1. The Ed was reported up to 0.92 Wh·kg−1, and Pd varied from 100 W·kg−1 to 3000 W·kg−1 [335].
A hybrid 2D MoS2 NSs and 1D PANI nanowires were used as the SC electrode. The MoS2/PANI hybrid electrode showed capacitance of about 485 F·g−1 at a current density of 1 mAcm2. The higher performance was reported about 812 F·g−1, which was for the 1:2 ratio of MoS2 and PANI that can be associated with the synergistic interactions between conductive 2D MoS2 NSs with a large surface area and 1D PANI. The related Ed to that ratio and also the Pd was reported as high as 112 W·kg−1 and 0.6 kW·kg−1, respectively [410].

3.4.4. MOs-CPs Composites

The MOs/CPs-based material is another alternative composite for electrode material which provides an enhanced electrochemical characteristic in SCs resulting from the compatibility between the MOs and CPs [23,411].
The specific capacitance, rate capability, and cyclic stability can be increased, in comparison with pure MOs and pure CPs material for electrodes, by developing the composites of MOs coupled with CPs, which leads to improved conductivity of the electrodes.
For instance, the PPy coated MoO3 was synthesized through the in-situ polymerization method, which exhibited the specific capacitance of 110 F·g−1 at a current density of 100 m A·g−1 as well as an Ed of 20 Wh·kg−1 at Pd of 75 W·kg−1. ASC was assembled using the AC as the –ve electrode and the PPy/MoO3 as the +ve electrode and in 0.5 M K2SO4 aqueous solution that showed an Ed of 12 Wh·kg−1 at Pd of 3 kW·kg−1 [412].
The PANI/MnO2 nanocomposite was [413] prepared through the exchange reaction between PANI, and n-octadecyltrimethylammonium-intercalated MnO2 in N-methyl-2-pyrrolidone solvent in which the Fourier transform infrared (FTIR) spectroscopy results indicated the desirable interactions between MnO2 layers and intercalated PANI. It also has been found that the PANI still shows a proper conductivity after insertion into the layered MnO2. The composite exhibited a specific capacitance of 330 F·g−1 at the current density of 1 A·g−1 that was 187 F·g−1 and 208 F·g−1 higher than pristine PANI and MnO2, respectively, which is the result of the synergistic effect of PANI and MnO2. After 1000 cycles, 94% of the initial capacitance remained.
The MOs-CPs composite could be used extensively as an electrode material for both flexible and non-flexible planar SCs to improve different parameters such as the capacitance, Ed, and power density.
As an example, Raj et al. [414] synthesized Co3O4-PIND, (cobalt oxide-conducting polyindole), as electrode material for a non-flexible SC, which was fabricated in a free-binder system and showed a specific capacitance of 1805 F·g−1 and 1625 F·g−1 at the current density of 2 A·g−1 at 25 A·g−1, respectively.
There is also various electrode material composite, fabricated using carbon-based material, MOs and CPs, in which the ternary nanocomposite has exhibited improved electrochemical performance compared to those of corresponding individual components as electrode materials for SC.

4. Two-Dimensional (2D) Electrode Materials and Their Modification Methods

Since the tremendous discovery of exfoliated monocrystalline graphitic films in 2004 [415], the 2D material with a layered structure has appealed much interest from scientists due to their fantastic properties.
During 16 years after the successful mechanical exfoliation of graphene from three-dimensional (3D) graphite, various 2D nanomaterial has been obtained through various preparation and exfoliation methods, from different layered precursor materials such as hexagonal boron nitride (hBN) [416], metal nitrides (MNs) [416], TMOs [417], black phosphorus (BP) [418], transition metal dichalcogenides (TMDs) [419], metal-organic frameworks (MOFs) [420], covalent organic frameworks (COFs) [421], layered double hydroxides (LDHs) [422], graphitic carbon nitride (g-C3N4) [423], and different MAX phase (e.g., Ti2AlC) [424], as well as other precursors, which are utilized to obtain the 2D transition metal carbides, carbonitrides, and nitrides called “MXene” [425].
The structure and electrochemical properties (including the energy storage mechanism and etc.,) of various 2D materials, as well as their application as the electrode in SC devices, was explained in detail in our previous study [426] (Figure 20).
As it was mentioned in the previous sections, the efficiency of the SC device could be enhanced mainly by improving the electrode materials and modifying the electrolyte material. Here, we have summarized different modification methods, which provide more interesting outlooks on the study of 2D materials, in addition to their exclusive intrinsic properties. There are some common modification techniques to enlarge the surface area, which lead to enhancing the charge storage ability of the electrode materials, thus improving the properties of 2D materials. These techniques can be categorized into several groups including alloying [427], hybridization [428] intercalation [429], nanosizing, the top-down method [430], vertical/lateral heterojunctions [431], external field modification [432], strain engineering [433], and stacking order of the atomic layers [434].
There are different studies with regards to the enhancement of the capacitive performance of the 2D material through a core-shell formation strategy.
For instance, the hierarchical NiCo2O4@polypyrrole core-shell nanowires were embedded onto the hemp-derived carbon (HDC) microfiber as fiber-shaped electrode material for flexible planar SC, for which the as-fabricated electrode (HDC@NiCo2O4@PPy) exhibited an excellent rate performance as a specific capacitance of 2055 F·g−1 was attained at 1 A·g−1 that was higher than that of the HDC@NiCo2O4 composites which exhibited 1506 F·g−1 under the same conditions. The specific capacitance (742 F·g−1) was remained still higher than that of the HDC@NiCo2O4 composites (402 F·g−1 at 50 A·g−1), even after increasing the current density up to 50 A·g−1. After 5000 cycles, 90% of the initial capacity was left. The fabricated fiber-shaped all-solid-state SSC showed high physical flexibility, good stability, and an Ed of 17.5 Wh·kg−1 at Pd of 500 W·kg−1 [435]. The improvements in the pseudocapacitive performance arise from the high electrical conductivity of carbon microfiber and the PPY.
Yim et al. developed the nanospheres in core–shell structure through the electrodeposition method utilizing the 2D arrayed of polystyrene (PS) monolayer as the core and the RuO2 nanoparticles as the shell, in which the resultant nanospheres showed an excellent capacitive performance including a superior areal capacitance, well bonded to the current collector materials compared to that of planar RuO2-based electrode (Figure 21) [436].
There are several examples of different 2D materials that have been utilized for SC applications in addition to the mentioned examples in this review so far, and in our previous review, [426] in which we have described and highlighted the development of the 2D electrode materials such as MXene, MOFs, TMDs, COFs, MNs, BP, and perovskite (PSK).
Recently, three new β-ketoenamine-linked COFs have been fabricated using Schiff-base [3 + 3] polycondensations of 1,3,5-triformylphloroglucinol (TFP-3OHCHO) along with three tris(aminophenyl)-presenting derivatives-featuring amino, carbazole, and pyridine linkages, respectively under the solvothermal conditions [437].
β-ketoenamine connected naphthalene based porous covalent organic framework (TFP-NDA-COF) has been developed through a simple Schiff base condensation polymerization reaction between triformylphloroglucinol (TFP) and 1,5-napthalenediamine (NDA). Further, this material was utilized for the SC application in which the resultant electrode showed a specific capacitance up to 379 F·g−1 at 2 mV·s−1 scan rate and 348 F·g−1 at the current density of 0.5 A·g−1. It has shown an excellent supercapacitive behaviour and high chemical stability (75%) even after 8000 cycles which could be associated with the high SSA, high-porosity in the polymer network, the presence of naphthalene moiety for facile π conjugation as well as nitrogen as a hetero atom for redox activity [438].
The synthesized TFP-COFs were used in the SC and for gas storage applications and showed enhanced performance such as excellent crystallinity and superior electrochemical specific capacitance of 291.1 F·g−1 as well as a high BET SSA up to 686 m2·g−1. The improved performance can be associated with the conjugated enamine structures with redox-active triphenyl amino, carbazole, and pyridine units. The presence of more highly planar monomer units in the hexagonal structures of the synthesized COFs provided a stronger quadrupolar interaction with the included CO2 molecules [437].
It has been reported that the PEDOT-modified DAAQ-TFP COF films exhibited excellent high-rate performance and a better current response in the CV compared to that of pristine DAAQ-TFP COF film that can be associated with the wiring effect of PEDOT chains which ease the electron transport between electrodes and the redox-active sites of the DAAQ-TFP COF. Though, the ion transfer is enhanced through the vertical pore channels by shortening the diffusion length [439].Romero et al. [440] developed the COFs of polyimine with many metal ions, including Fe3+, Co2+, and Ni2+.
There are several disadvantages for COFs, such as less stability and electrical conductivity, which limits their application. A strategy has been exploited to eliminate these restrictions in which the CPs were confined within the porous frameworks that enhanced the cycle stability along with maintaining the mechanical stability and electrical conductivity [364]. For instance, Dichtel et al. [441] obtained a very high-rate charging (10−1600 °C) with higher Ed by exploiting this procedure via a simple electro-polymerization of EDOT into the COF pores.
Chandra et al. developed the COFs [TpPa-(OH)2, TpBD-(OH)2] as electrode materials for SCs. Figure 22 [442] illustrates the XRD pattern, the N2 adsorption curves, and the pore size distribution along with the TEM images of COFs. The resultant TpPa-(OH)2 presented a specific capacitance of 416 F·g−1 at a current density of 0.5 A·g−1 with 66% Cs retention after 10,000 cycles. The high specific capacitance is ascribed to the specific molecular control of redox-active units in the COF.
Recently, a COF complex was prepared as electrode material for flexible and extensible SCs using a template of hydroxyl-ended hyperbranched polymer (OHP) through a simple solid-state mechanical mixing process. Further, the COF@OHP complex was impregnated on a microporous carbon nanotube film (CNTF) to build a composite membrane (CHCM). The electrode showed superior electrochemical performance, including a high gravimetric capacitance of 249 F·g−1 as well as a superior lifespan in which 80% of initial capacitance remained after 10,000 charge–discharge cycles [443].
Furthermore, to obtain a high power/energy density SC device, the advance electrode materials are required to provide desirable properties depending upon the specific application. There are various approaches to improving the electrochemical properties of the “old” pseudocapacitive material such as MnO2 and RuOx to obtain the “advanced” pseudocapacitive material such as TMDs (such as MoS2, etc.,) and MOFs [426,444] For example, the functionalized of hybridization of 2D-TMDs with various carbon-based supporting materials to increase more active sites, and ease the charge transferability has been reported in several studies [418,445,446,447], (Figure 23) [428].
TMDs have fascinated a lot of interests owing to their excellent physicochemical properties including high surface area, semiconducting ability, low cost, high chemical stability, high mechanical performance, facile synthesis by means of different fabrication techniques such as mechanical exfoliation, direct sonication in solvents, chemical vapour deposition (CVD) and, etc. [419,448,449,450].
For instance, an ASC device was fabricated by using flowerlike MoS2/GNSs as the negative and MnO2/GNSs as the positive electrodes in which the MoS2/GNSs electrode displayed excellent electrochemical performance, such as high specific capacitance of 320 F·g−1 at 2 A·g−1. In addition, the resultant all-solid-state ASC achieved a remarkable Ed of 78.9 Wh·kg−1 at a Pd of 284.1 W·kg−1 [451].
Besides all the advantages, the interlayer interactions and high surface energy of the 2D-TMDs layers result in an increase in the restacking possibility of the layers, thus decreasing the number of active sites [452]. One of the efficient solutions to overcome the above disadvantages as well as increase the TMD activity for real advanced applications is to fabricate the new functional hybrid materials by coupling the 2D-TMDs with a highly conductive supporting material [419,445,453].
Furthermore, boron nitride (BN) is another excellent member of the 2D material, which has exhibited a superior dielectric property, high thermal conductivity. The BN layered material has shown superior stability and excellent dispersion in the organic electrolyte. With all these properties, BN material cannot be utilized alone as the SC electrode material, since its bandgap is large (between 4 eV to 5 eV). However, this drawback could be addressed through the bandgap engineering of the BN materials [454].
Recently, the PANI-coated CNTs has been grown on the exfoliated BN sheets to improve the electrochemical properties of the BN material. The resultant nanocomposite exhibited a specific capacitance of 515 F·g−1 as well as an Ed of 46 Wh·kg−1 in a three-electrode system.After 10,000 cycles, 98% of the capacitance remained. The nanohybrid ternary composite has also been used as electrode material in an ASSC, which provides an excellent Ed of 62 Wh·kg−1 and capacitance retention about 91% after 5000 cycles [455]. In another study, a ternary nanocomposite was synthesized through the hydrothermal method using rGO, exfoliated BN NSs, and Ni(OH)2. The resultant nanocomposite (rGO/BN/Ni(OH)2) exhibited a maximum specific capacitance of 349 F·g−1 as well as a high Ed of 48 Wh·kg−1, and a Pd of 118 W·kg−1 in 0.5 M KOH at a current density of 0.5 mA·cm−2 [456].
BP is another strong candidate that is comprised of phosphorene layers connected by the Van der Waals interaction [457]. One of the attractive characteristics of the BP material is the interlayer distance (0.55 nm), which is larger than that of graphene. This criterion makes the BP material to be suitable as electrode materials, owing to their capability of enhancing the ion intercalation, thus improving the electrochemical properties of the corresponding SC device [458].
Recently, the R-BP/SPC (red-black phosphorus/sulfonated porous carbon) composite was synthesized via the sonochemical method. The hybrid composite was utilized as the electrode in a SSC, which provided a high specific capacitance of 364.5 F·g−1 at a current density of 0.5 A·g−1. They reported superior capacitance retention of 89% after 10,000 cycles [459].
The oxygen vacancy is a significant factor that affects the ionic diffusion and conductivity of electrode materials, where the enhanced number of oxygen vacancies can decline the diffusion resistance [460] as a result, the porous networks can deliver higher SSA and increased the number of oxygen vacancies, thus improving the specific capacity/capacitance.
Similar PSKs exhibit different electrochemical behaviours in various types of electrolytes (pseudocapacitive behaviour and battery-like behavior). The charge storage performance of the PSKs with battery-like behaviour should be calculated by specific capacity (mAh·g−1 or C·g−1) rather than specific capacitance (F·g−1) [461].
The electrode materials with ultrathin nanosheet structures can achieve high specific capacity/capacitance and long-term stability because such structure decreases the strain resulted from the volume change during the charge–discharge process, thus providing more contact between the active material and electrolyte due to the high SSA [462].
PSK material is one of the novel alternative materials, which have attracted a lot of attention as electrode materials for SCs. PSK could be utilized as the anode material for SCs to provide an anion intercalation process. PSK oxides are promising multi-functional materials due to their enhanced physicochemical properties and the controllable oxygen vacancy number, which could be tuned through various techniques to improve the properties of the PSK oxides material as anodes for anion-intercalation SC. The number of oxygen vacancies and the high SSA of PSK oxides could enhance the electrochemical performance of SCs.
The electrical properties of PSKs could be changed through altering the temperature, in which most PSKs achieve a high conductivity at high temperature due to their structure with high oxygen diffusion rate and stability, which are different from that of those metals in the structure [463,464].
For example, the sphere La0.85Ca0.15MnO3 (LCM) and LaMnO3 (LMO) have been compared in terms of the morphology, chemical properties, specific capacitance and intrinsic resistance of the electroactive material itself. The specific capacitance was reported of 198 F·g−1 and 187 F·g−1 at the current density of 0.5 A·g−1 for La0.85Sr0.15MnO3 and LaMnO3, respectively [465].
Recently, the effect of A-site substitution on the crystalline structure of LMO PSK by using Ca or Sr substituentswas studied. The LCM and La0.85Sr0.15MnO3 (LSM) electrodes showed a specific capacitance of 140.5 mF·cm−2 and 129.0 mF·cm−2 at 0.5 mA·cm−2 in an aqueous electrolyte, respectively, which are four times higher than that of LMO (Figure 24) [466].
Sharma et al. prepared strontium titanate as the electrode material for SSC through the sol-gel method. The resultant electrode displayed a high capacitance of 592 F·g−1 at 5 mV·s−1 because of the mesoporous structure, and large SSA, of the electrode materials as well as a high mass transfer rate of electrolytic ions. The fabricated SSC using strontium titanate electrodes and aqueous electrolyte exhibited a high specific capacitance of 212.5 F·g−1 at 0.63 A·g−1, high Ed of 27.8 Wh·kg−1 with a Pd of 1921 W·kg−1. Furthermore, the device showed high cycling stability of 99% after 5000 cycles, and good capacitance retention of 74.5% at 2.5 A·g−1. The flexible and solid-state symmetric cell also exhibited high energy and Pd of 3.62 Wh·kg−1 and 965W·kg−1, respectively [467].
PSK materials are also a powerful candidate for flexible energy storage/conversion, such as flexible PSK solar cells and FSCs [2,468]. Table 2 summarizes the recent advances on the 2D nanomaterials as the electrode for supercapacitor devices as well as their reported charge storage mechanisms. As is shown in Table 2, the electrochemical performance of the supercapacitor can be improved using nanocomposites made by materials, which display different charge storage mechanism to obtain a battery-type nanocomposite. 2D material such as BN NSs and MXene flakes exhibit EDLC behaviour, while MOFs, COFs, and PSK can provide pseudocapacitance. Moreover, the EDLC behaviour of the 2D-based electrode materials can be improved through the incorporation of the carbonaceous materials, as the extra pseudocapacitance can be obtained by the utilization of various TMOs and CPs.

5. Applications and Future Perspectives

Recent studies on the development of various electrode materials, the investigation of the SC operation and their charge storage mechanisms have enabled novel perspectives on the utilization of advanced electrode materials in the asymmetric flexible hybrid supercapacitor (ASFHSC) increasingly in the future. The configuration of hybrid-asymmetric superconductor provides the advantage of having a wide operating potential window, which results from the unique potential range of cathodes and anodes [475]. However, the HSCs exploit the advantages of both battery and supercapacitor electrodes, to provide high energy density and power density, respectively. Still, there are several challenges with regard to the fabrication of such systems, including the cost-efficient and scalable fabrication methods, developing the effective procedures to provide more control on the surface features (e.g., porosity, morphology, etc.) of the materials.
FSCs have commanded a lot of concentration as one of the most important energy storage research focuses owing to their extraordinary capacitive behaviour, higher power density, higher rate capability, and higher cyclic stability in comparison with batteries and higher Ed compared to the traditional capacitors [476], which provide them with a vast range of potential to be utilized in various applications. Each category of SC has its own specific features and applications, such as in portable/wearable electronic devices as well as in computer and mobile phones [2,477], high power sources for heavy lift trucks or cranes [478,479], the self-powered SCs in medical applications (e.g., flexible artificial skin, artificial intelligence, portable or implantable medical/biomedical devices) [480,481], solar applications [482,483,484]. There are several SC companies such as ELTON (Russia), Nesscap (Korea), CAP-XX (Australia), and Nippon Chemicon (Japan), which are developing various kind of SCs for different applications [485].
Generally, the selection of the electrode material is one of the intense issues in the fabrication of an FSCs to obtain the appropriate mechanical strength and a high capacitance using the free-standing and binder-free electrodes, which are the main constituents of FSCs [476]. The electrode material for FSCs, as other classes of SCs, should fulfil the requirements of having a superior electrical conductivity, high tensile strain, large SSA, favourable chemical and thermal stability, high flexibility, wide potential window, and having enormous surface functional groups. For instance, among the carbon-based materials, graphene is one of the most powerful candidates for electrode material in FSCs owing to its large specific surface and high flexibility, which gives graphene the ability to be bent into large angles and provides a wide variety on the microstructures [486,487]. Recently, aerogels exhibited promising properties for 3D printed energy storage applications. For instance, graphene aerogels are commonly exploited owing to their high specific surface area, mechanical properties as well as tunable pore structures and porosity [488]. Therefore, as it has been reported in different studies [476,487] that carbonaceous materials could be used as a backbone, in the composite material along with redox-active materials such as MOs and CPs, or could be integrated with novel 2D materials such as MXene [489,490] to fabricate the appropriate flexible film as the electrode material for flexible/transparent SCs. Furthermore, incorporating the redox-active small molecules and bioderived functional groups into the pseudocapacitive material, as well as their utilization to increase the capacitance of the carbon-based material, have been provided more efficient, lower cost, non-toxic electrodes for SC application through eco-friendly synthesis [491,492].

6. Conclusions

The charge storage mechanisms, electrochemical behaviour and the principles of these materials have been investigated through several examples of recent advances. In this review, we have also described the distinction between the term “asymmetric” and “hybrid” as well as the difference between various electrochemical behaviour for the electrode materials including “battery-type” and “pseudocapacitive” behaviour. Besides, the main characteristic of the energy storage system including capacitance/capacity, operating temperature, energy density, power density, cell voltage, kinetic storage mechanism, I–V characteristics, cycling lifetime, self-discharge, and the nature of electrode materials have been briefly explained. The electrochemical signatures of EDLC, pseudocapacitive, and battery-like materials, including the CVs, GCDs, and i–v relationship, have been investigated. Three main groups of the electrode material include carbon-based material, metal oxides/hydroxides, and conductive polymers. Carbonaceous materials are promising electrodes to provide high specific surface area, favorable chemical and thermal stability as well as low electrical resistance, however their low energy density, which arises from their surface or quasi-surface energy storage, limits their application for large-scale. MOs such as RuO2, V2O5, and MnO2, exhibited an outstanding improvement as electrode materials (both the cathode and the anode) due to their variety, excellent specific capacitance, and environmental friendlyness. However, their low electrical conductivity is one of the drawbacks, which weakens their rate capability and consequently the Pd of the corresponding material. Conducting polymers have shown facile processability, improved conductivity, favourable flexibility, and advanced Ed owing to their redox-active heteroatoms; however, there are some drawbacks such as volume shrinking/expansion, which restrict the cyclic performance of the CPs during the charging–discharging process. Apart from the mentioned advantage of different materials, still, there are several drawbacks for each group of materials. For example, in the case of the pseudocapacitive materials, the mass loading and the thickness of the active material must be accurately considered, since when the mass loading/thickness of the active material is very low, or the charge–discharge mechanisms are directed at a low current density, the energy performance may be observed as high mistakenly.
The composite-based or hybrid electrodes designed employing 1D, 2D, and 3D nanomaterials exhibited higher SSA, offer shorter diffusion paths for electrons and ions, thus improving the pseudocapacitance. The recent examples of the developed 2D nanomaterial such as hBN, MNs, BP, TMOs, TMDs, MOFs, COFs, MXene, and PSK has been given. Furthermore, we have extensively reviewed the importance of different nanocomposite made of various carbonaceous materials, TMOs, TM(OH)s, CPs, and 2D material, which provided advanced properties for the electrode material such as larger SSA, higher conductivity (to provide a favourable charge transfer kinetics through the high availability for the electrolyte), facile manipulation of morphology (e.g., pore shape, particle distributions/dimensions), higher mechanical, chemical, and thermal stability as well as enhanced electrochemical properties such as higher specific capacitance, energy and Pd for the corresponding SC compared to that of the pure electrode material. In addition to the appropriate material selection, the electrochemical properties of the energy storage devices can be enhanced through the proper combination of the positive and negative electrodes, especially in the case of hybrid devices, where the battery-type and supercapacitor-type materials with various chemical natures are responsible for providing the high energy and high power densities, respectively.

Author Contributions

Conceptualization, S.C.P.; original draft preparation, P.F.; editing and review, V.K. and S.C.P.; supervision, S.C.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Presidents Bursary funding”, through the Institute of Technology Sligo.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

SCsSupercapacitors
EDLCsElectric double-layer capacitors
SSCsSymmetric supercapacitors
ASSCsAsymmetric supercapacitors
HSCsHybrid supercapacitors
FSCsFlexible supercapacitors
ASFHSCAsymmetric flexible hybrid supercapacitor
PdPower density
EdEnergy density
CVsCyclic voltammograms
SSASpecific surface area
0DZero-dimensional
1DOne-dimensional
2DTwo-dimensional
3DThree-dimensional
NSsNanosheets
NSAsNanosheet arrays
CBCarbon black
CCCarbon cloth
CCFCarbonized cotton fabric
CNTsCarbon nanotubes
CNFsCarbon nanofibers
MCNFsMulti-channel carbon nanofiber
ECCElectrochemically activated carbon
TPCTemplated porous carbon
SPCSulfurated porous carbon
HCNBsHollow carbon nano bowls
rGOReduced graphene oxide
GQDsGraphene quantum dot
PGrPorous graphene
GNSsGraphene nanosheets
GPCNsGraphene-like porous carbon nanosheets
MOsMetal oxides
TMsTransition metal
TMOsTransition metal oxide
TM(OH)sTransition metal hydroxide
TMDsTransition metal dichalcogenides
COFsCovalent organic frameworks
MOFsMetal organic framework
PANIPolyaniline
PThPolythiophene
PPyPolypyrrole
PINDPolyindole
PEDOTPoly 3,4-ethylenedioxythiophene
PVAPoly vinyl alcohol
PSPolystyrene
BNBoron nitrides
PNTsPolypyrrole nanotubes
hBNHexagonal boron nitride
BPBlack phosphorous
PSKPerovskite
MXene2D transition metal carbides, Carbonitrides, and nitrides

References

  1. Simon, P.; Gogotsi, Y.; Dunn, B. Where Do Batteries End and Supercapacitors Begin? Science 2014, 343, 1210–1211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Li, C.; Islam, M.; Moore, J.; Sleppy, J.; Morrison, C.; Konstantinov, K.; Dou, S.X.; Renduchintala, C.; Thomas, J. Wearable energy-smart ribbons for synchronous energy harvest and storage. Nat. Commun. 2016, 7, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Winter, M.; Brodd, R.J. What are batteries, fuel cells, and supercapacitors? Chem. Rev. 2004, 104, 4245–4269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kötz, R.; Carlen, M. Principles and applications of electrochemical capacitors. Electrochim. Acta 2000, 45, 2483–2498. [Google Scholar] [CrossRef]
  5. Lewandowski, A.; Galinski, M. Practical and theoretical limits for electrochemical double-layer capacitors. J. Power Sources 2007, 173, 822–828. [Google Scholar] [CrossRef]
  6. Bohlen, O.; Kowal, J.; Sauer, D.U. Ageing behaviour of electrochemical double layer capacitors. J. Power Sources 2007, 173, 626–632. [Google Scholar] [CrossRef]
  7. Huang, J.; Sumpter, B.G.; Meunier, V. Theoretical Model for Nanoporous Carbon Supercapacitors. Angew. Chem. Int. Ed. 2008, 47, 520–524. [Google Scholar] [CrossRef]
  8. Lu, Z.; Raad, R.; Safaei, F.; Xi, J.; Liu, Z.; Foroughi, J. Carbon Nanotube Based Fiber Supercapacitor as Wearable Energy Storage. Front. Mater. 2019, 6, 1–14. [Google Scholar] [CrossRef]
  9. Lokhande, P.E.; Chavan, U.S.; Pandey, A. Materials and Fabrication Methods for Electrochemical Supercapacitors: Overview. Electrochem. Energy Rev. 2020, 3, 155–186. [Google Scholar] [CrossRef]
  10. Kakaei, K.; Esrafili, M.D.; Ehsani, A. Graphene-Based Electrochemical Supercapacitors. Interface Sci. Technol. 2019, 27, 339–386. [Google Scholar]
  11. Zhang, L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 2520–2531. [Google Scholar] [CrossRef] [PubMed]
  12. Chmiola, J.; Yushin, G.; Gogotsi, Y.; Portet, C.; Simon, P.; Taberna, P.L. Anomalous increase in carbon at pore sizes less than 1 nanometer. Science 2006, 313, 1760–1763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Li, J.; Cheng, X.; Shashurin, A.; Keidar, M. Review of Electrochemical Capacitors Based on Carbon Nanotubes and Graphene. Graphene 2012, 1, 1–13. [Google Scholar] [CrossRef] [Green Version]
  15. Pan, H.; Li, J.; Feng, Y.P. Carbon nanotubes for supercapacitor. Nanoscale Res. Lett. 2010, 5, 654–668. [Google Scholar] [CrossRef] [Green Version]
  16. Forse, A.C.; Merlet, C.; Griffin, J.M.; Grey, C.P. New perspectives on the charging mechanisms of supercapacitors. J. Am. Chem. Soc. 2016, 138, 5731–5744. [Google Scholar] [CrossRef] [Green Version]
  17. Noori, A.; El-Kady, M.F.; Rahmanifar, M.S.; Kaner, R.B.; Mousavi, M.F. Towards establishing standard performance metrics for batteries, supercapacitors and beyond. Chem. Soc. Rev. 2019, 48, 1272–1341. [Google Scholar] [CrossRef]
  18. Woo, J.; Ahn, T.; Hun, J.; Myoun, J.; Kim, J. Electrochimica Acta Nanosheets based mesoporous NiO microspherical structures via facile and template-free method for high performance supercapacitors. Electrochim. Acta 2011, 56, 4849–4857. [Google Scholar] [CrossRef]
  19. Xing, Z.; Chu, Q.; Ren, X.; Ge, C.; Qusti, A.H.; Asiri, A.M.; Al-Youbi, A.O.; Sun, X. Ni3S2 coated ZnO array for high-performance supercapacitors. J. Power Sources 2014, 245, 463–467. [Google Scholar] [CrossRef]
  20. Xia, Q.; Yang, H.; Wang, M.; Yang, M.; Guo, Q.; Wan, L.; Xia, H.; Yu, Y. High Energy and High Power Lithium-Ion Capacitors Based on Boron and Nitrogen Dual-Doped 3D Carbon Nanofibers as Both Cathode and Anode. Adv. Energy Mater. 2017, 7, 1701336. [Google Scholar] [CrossRef]
  21. Ding, J.; Wang, H.; Li, Z.; Cui, K.; Karpuzov, D.; Tan, X.; Kohandehghan, A.; Mitlin, D. Peanut shell hybrid sodium ion capacitor with extreme energy–power rivals lithium ion capacitors. Energy Environ. Sci. 2015, 8, 941–955. [Google Scholar] [CrossRef]
  22. Khomenko, V.; Raymundo-Piñero, E.; Béguin, F. High-energy density graphite/AC capacitor in organic electrolyte. J. Power Sources 2008, 177, 643–651. [Google Scholar] [CrossRef]
  23. Chen, H.; Cong, T.N.; Yang, W.; Tan, C.; Li, Y.; Ding, Y. Progress in electrical energy storage system: A critical review. Prog. Nat. Sci. 2009, 19, 291–312. [Google Scholar] [CrossRef]
  24. Andreas, H.A. Self-Discharge in Electrochemical Capacitors: A Perspective Article. J. Electrochem. Soc. 2015, 162, A5047–A5053. [Google Scholar] [CrossRef]
  25. Black, J.; Andreas, H.A. Effects of charge redistribution on self-discharge of electrochemical capacitors. Electrochim. Acta 2009, 54, 3568–3574. [Google Scholar] [CrossRef]
  26. Abbas, Q.; Béguin, F. Influence of the iodide/iodine redox system on the self-discharge of AC/AC electrochemical capacitors in salt aqueous electrolyte. Prog. Nat. Sci. Mater. Int. 2015, 25, 622–630. [Google Scholar] [CrossRef] [Green Version]
  27. Kötz, R.; Hahn, M.; Gallay, R. Temperature behavior and impedance fundamentals of supercapacitors. J. Power Sources 2006, 154, 550–555. [Google Scholar] [CrossRef]
  28. Xiao, H.; Yao, S.; Liu, H.; Qu, F.; Zhang, X.; Wu, X. NiO nanosheet assembles for supercapacitor electrode materials. Prog. Nat. Sci. Mater. Int. 2016, 26, 271–275. [Google Scholar] [CrossRef] [Green Version]
  29. Peters, J.; Buchholz, D.; Passerini, S.; Weil, M. Life cycle assessment of sodium-ion batteries. Energy Environ. Sci. 2016, 9, 1744–1751. [Google Scholar] [CrossRef] [Green Version]
  30. Ding, J.; Hu, W.; Paek, E.; Mitlin, D. Review of Hybrid Ion Capacitors: From Aqueous to Lithium to Sodium. Chem. Rev. 2018, 118, 6457–6498. [Google Scholar] [CrossRef] [PubMed]
  31. Lu, Y.; Li, B.; Zheng, S.; Xu, Y.; Xue, H.; Pang, H. Syntheses and Energy Storage Applications of M x S y (M = Cu, Ag, Au) and Their Composites: Rechargeable Batteries and Supercapacitors. Adv. Funct. Mater. 2017, 27, 1703949. [Google Scholar] [CrossRef]
  32. Chemali, E.; Preindl, M.; Malysz, P.; Emadi, A. Electrochemical and Electrostatic Energy Storage and Management Systems for Electric Drive Vehicles: State-of-the-Art Review and Future Trends. IEEE J. Emerg. Sel. Top. Power Electron. 2016, 4, 1117–1134. [Google Scholar] [CrossRef]
  33. Talaie, E.; Bonnick, P.; Sun, X.; Pang, Q.; Liang, X.; Nazar, L.F. Methods and Protocols for Electrochemical Energy Storage Materials Research. Chem. Mater. 2017, 29, 90–105. [Google Scholar] [CrossRef]
  34. Liu, Q.; Li, Z.-F.; Liu, Y.; Zhang, H.; Ren, Y.; Sun, C.-J.; Lu, W.; Zhou, Y.; Stanciu, L.; Stach, E.; et al. Graphene-modified nanostructured vanadium pentoxide hybrids with extraordinary electrochemical performance for Li-ion batteries. Nat. Commun. 2015, 6, 6127. [Google Scholar] [CrossRef]
  35. Ardizzone, S.; Fregonara, G.; Trasatti, S. Inner and outer active surface of RuO2 electrodes. Electrochim. Acta 1990, 35, 263–267. [Google Scholar] [CrossRef]
  36. Zheng, J.P. Hydrous Ruthenium Oxide as an Electrode Material for Electrochemical Capacitors. J. Electrochem. Soc. 1995, 142, 2699. [Google Scholar] [CrossRef]
  37. Lee, H.Y.; Goodenough, J.B. Supercapacitor Behavior with KCl Electrolyte. J. Solid State Chem. 1999, 144, 220–223. [Google Scholar] [CrossRef]
  38. Brousse, T.; Toupin, M.; Dugas, R.; Athouël, L.; Crosnier, O.; Bélanger, D. Crystalline MnO2 as Possible Alternatives to Amorphous Compounds in Electrochemical Supercapacitors. J. Electrochem. Soc. 2006, 153, A2171. [Google Scholar] [CrossRef]
  39. Brousse, T.; Bélanger, D.; Long, J.W. To Be or Not To Be Pseudocapacitive? J. Electrochem. Soc. 2015, 162, A5185–A5189. [Google Scholar] [CrossRef] [Green Version]
  40. Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive oxide materials for high-rate electrochemical energy storage. Energy Environ. Sci. 2014, 7, 1597. [Google Scholar] [CrossRef] [Green Version]
  41. Lee, J.W.; Hall, A.S.; Kim, J.D.; Mallouk, T.E. A facile and template-free hydrothermal synthesis of Mn3O4 nanorods on graphene sheets for supercapacitor electrodes with long cycle stability. Chem. Mater. 2012, 24, 1158–1164. [Google Scholar] [CrossRef]
  42. Wang, S.-Y.; Ho, K.-C.; Kuo, S.-L.; Wu, N.-L. Investigation on Capacitance Mechanisms of Fe3O4 Electrochemical Capacitors. J. Electrochem. Soc. 2006, 153, 75–80. [Google Scholar] [CrossRef]
  43. Xia, X.-H.; Tu, J.-P.; Wang, X.-L.; Gu, C.-D.; Zhao, X.-B. Mesoporous Co3O4 monolayer hollow-sphere array as electrochemical pseudocapacitor material. Chem. Commun. 2011, 47, 5786. [Google Scholar] [CrossRef] [PubMed]
  44. Singh, A.K.; Sarkar, D.; Khan, G.G.; Mandal, K. Designing one dimensional Co-Ni/Co3O4-NiO core/shell nano-heterostructure electrodes for high-performance pseudocapacitor. Appl. Phys. Lett. 2014, 104, 133904. [Google Scholar] [CrossRef]
  45. Wang, H.; Yi, H.; Chen, X.; Wang, X. Facile synthesis of a nano-structured nickel oxide electrode with outstanding pseudocapacitive properties. Electrochim. Acta 2013, 105, 353–361. [Google Scholar] [CrossRef]
  46. Kandalkar, S.G.; Gunjakar, J.L.; Lokhande, C.D. Preparation of cobalt oxide thin films and its use in supercapacitor application. Appl. Surf. Sci. 2008, 254, 5540–5544. [Google Scholar] [CrossRef]
  47. Senthilkumar, B.; Sankar, K.V.; Vasylechko, L.; Lee, Y.-S.; Selvan, R.K. Synthesis and electrochemical performances of maricite-NaMPO4 (M = Ni, Co, Mn) electrodes for hybrid supercapacitors. RSC Adv. 2014, 4, 53192–53200. [Google Scholar] [CrossRef]
  48. Augustyn, V.; Come, J.; Lowe, M.A.; Kim, J.W.; Taberna, P.-L.; Tolbert, S.H.; Abruña, H.D.; Simon, P.; Dunn, B. High-rate electrochemical energy storage through Li+ intercalation pseudocapacitance. Nat. Mater. 2013, 12, 518–522. [Google Scholar] [CrossRef]
  49. Jiang, Y.; Liu, J. Definitions of Pseudocapacitive Materials: A Brief Review. Energy Environ. Mater. 2019, 2, 30–37. [Google Scholar] [CrossRef] [Green Version]
  50. Wang, J.; Polleux, J.; Lim, J.; Dunn, B. Pseudocapacitive Contributions to Electrochemical Energy Storage in TiO2 (Anatase) Nanoparticles. J. Phys. Chem. C 2007, 111, 14925–14931. [Google Scholar] [CrossRef]
  51. Mitchell, J.B.; Lo, W.C.; Genc, A.; LeBeau, J.; Augustyn, V. Transition from Battery to Pseudocapacitor Behavior via Structural Water in Tungsten Oxide. Chem. Mater. 2017, 29, 3928–3937. [Google Scholar] [CrossRef]
  52. Lai, L.; Yang, S.; Wang, L.; Teh, B.K.; Zhong, J.; Chou, H.; Chen, L.; Chen, W.; Shen, Z.; Ruoff, R.S.; et al. Preparation of supercapacitor electrodes through selection of graphene surface functionalities. ACS Nano 2012, 6, 5941–5951. [Google Scholar] [CrossRef] [PubMed]
  53. Xie, L.; Sun, G.-H.; Su, F.; Guo, X.; Kong, Q.; Li, X.; Huang, X.; Wan, L.; Song, W.; Li, K.; et al. Hierarchical porous carbon microtubes derived from willow catkins for supercapacitor applications. J. Mater. Chem. A 2016, 4, 1637–1646. [Google Scholar] [CrossRef]
  54. Frackowiak, E.; Béguin, F. Carbon materials for the electrochemical storage of energy in capacitors. Carbon 2001, 39, 937–950. [Google Scholar] [CrossRef]
  55. Kim, B.K.; Sy, S.; Yu, A.; Zhang, J. Electrochemical Supercapacitors for Energy Storage and Conversion. Handb. Clean Energy Syst. 2015, 1–25. [Google Scholar] [CrossRef]
  56. Sevilla, M.; Mokaya, R. Energy storage applications of activated carbons: Supercapacitors and hydrogen storage. Energy Environ. Sci. 2014, 7, 1250–1280. [Google Scholar] [CrossRef] [Green Version]
  57. Pongprayoon, P.; Chaimanatsakun, A. Revealing the Effects of Pore Size and Geometry on the Mechanical Properties of Graphene Nanopore Using the Atomistic Finite Element Method. Acta Mech. Solida Sin. 2019, 32, 81–92. [Google Scholar] [CrossRef]
  58. Dai, G.; Zhang, L.; Liao, Y.; Shi, Y.; Xie, J.; Lei, F.; Fan, L. Multi-Scale Model for Describing the Effect of Pore Structure on Carbon-Based Electric Double Layer. J. Phys. Chem. C 2020, 124, 3952–3961. [Google Scholar] [CrossRef]
  59. Yu, Z.; Tetard, L.; Zhai, L.; Thomas, J. Supercapacitor electrode materials: Nanostructures from 0 to 3 dimensions. Energy Environ. Sci. 2015, 8, 702–730. [Google Scholar] [CrossRef] [Green Version]
  60. Shang, T.; Xu, Y.; Li, P.; Han, J.; Wu, Z.; Tao, Y.; Yang, Q.-H. A bio-derived sheet-like porous carbon with thin-layer pore walls for ultrahigh-power supercapacitors. Nano Energy 2020, 70, 104531. [Google Scholar] [CrossRef]
  61. Miller, J.M. Deposition of Ruthenium Nanoparticles on Carbon Aerogels for High Energy Density Supercapacitor Electrodes. J. Electrochem. Soc. 1997, 144, 309. [Google Scholar] [CrossRef]
  62. Chen, L.; Zhang, X.-D.; Liang, H.-W.; Kong, M.; Guan, Q.-F.; Chen, P.; Wu, Z.-Y.; Yu, S.-H. Synthesis of Nitrogen-Doped Porous Carbon Nanofibers as an Efficient Electrode Material for Supercapacitors. ACS Nano 2012. [Google Scholar] [CrossRef] [PubMed]
  63. Vangari, M.; Pryor, T.; Jiang, L. Supercapacitors: Review of materials and fabrication methods. J. Energy Eng. 2013, 139, 72–79. [Google Scholar] [CrossRef]
  64. Qi, F.; Xia, Z.; Sun, R.; Sun, X.; Xu, X.; Wei, W.; Wang, S.; Sun, G. Graphitization induced by KOH etching for the fabrication of hierarchical porous graphitic carbon sheets for high performance supercapacitors. J. Mater. Chem. A 2018, 6, 14170–14177. [Google Scholar] [CrossRef]
  65. He, X.; Ma, H.; Wang, J.; Xie, Y.; Xiao, N.; Qiu, J. Porous carbon nanosheets from coal tar for high-performance supercapacitors. J. Power Sources 2017, 357, 41–46. [Google Scholar] [CrossRef]
  66. Becker, H.I.; General Electric Company. Low Voltage Electrolytic Capacitor. U.S. Patent 2,800,616, 1957. Available online: https://patentimages.storage.googleapis.com/a2/f8/a9/b7d5c04a415c8b/US2800616.pdf (accessed on 8 June 2020).
  67. Iro, Z.S.; Subramani, C.; Dash, S.S. A brief review on electrode materials for supercapacitor. Int. J. Electrochem. Sci. 2016, 11, 10628–10643. [Google Scholar] [CrossRef]
  68. Chen, S.M.; Ramachandran, R.; Mani, V.; Saraswathi, R. Recent Advancements in Electrode Materials for the Highperformance Electrochemical Supercapacitors—A Review. Int. J. Electrochem. Sci. 2014, 9, 4072–4085. [Google Scholar]
  69. Arbizzani, C.; Mastragostino, M.; Soavi, F. New trends in electrochemical supercapacitors. J. Power Sources 2001, 100, 164–170. [Google Scholar] [CrossRef]
  70. Raymundo-Piñero, E.; Kierzek, K.; Machnikowski, J.; Béguin, F. Relationship between the nanoporous texture of activated carbons and their capacitance properties in different electrolytes. Carbon N. Y. 2006, 44, 2498–2507. [Google Scholar] [CrossRef]
  71. Arenillas, A.; Menéndez, J.A.; Reichenauer, G.; Celzard, A.; Fierro, V.; Hodar, F.J.M.; Bailόn-Garcia, E.; Job, N. Organic and Carbon Gels: From Laboratory Synthesis to Applications; Springer: Berlin, Germany, 2019; p. 195. [Google Scholar] [CrossRef]
  72. Wang, C.-M.; Wen, C.-Y.; Chen, Y.-C.; Chang, J.-Y.; Ho, C.-W.; Kao, K.-S.; Shih, W.-C.; Chiu, C.-M.; Shen, Y.-A. The influence of specific surface area on the capacitance of the carbon electrodes supercapacitor. In Proceedings of the 2nd International Conference on Industrial Application Engineering 2015, Kitakyushu, Japan, 28–31 March 2015; pp. 439–442. [Google Scholar] [CrossRef] [Green Version]
  73. Varghese, J.; Wang, H.; Pilon, L. Simulating Electric Double Layer Capacitance of Mesoporous Electrodes with Cylindrical Pores. J. Electrochem. Soc. 2011, 158, A1106. [Google Scholar] [CrossRef] [Green Version]
  74. He, Y.-T.; Zhang, Y.; Li, X.; Lv, Z.; Wang, X.; Liu, Z.; Huang, X. Capacitive mechanism of oxygen functional groups on carbon surface in supercapacitors. Electrochim. Acta 2018, 282, 618–625. [Google Scholar] [CrossRef] [Green Version]
  75. Zhan, C.; Zhang, Y.; Cummings, P.T.; Jiang, D.E. Enhancing graphene capacitance by nitrogen: Effects of doping configuration and concentration. Phys. Chem. Chem. Phys. 2016, 18, 4668–4674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Zhao, Y.; Wang, X.; Wang, N.; Li, M.; Li, Q.; Liu, J. Unraveling Factors Leading to High Pseudocapacitance of Redox-Active Small Aromatics on Graphene. J. Phys. Chem. C 2019, 123, 994–1002. [Google Scholar] [CrossRef]
  77. Yang, H.; Kannappan, S.; Pandian, A.S.; Jang, J.H.; Lee, Y.S.; Lu, W. Graphene supercapacitor with both high power and energy density. Nanotechnology 2017, 28. [Google Scholar] [CrossRef] [PubMed]
  78. Boddula, R.; Ahamed, M.I.; Asiri, A.M. Inorganic Nanomaterials for Supercapacitor Design; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  79. Iro, Z.S.; Subramani, C.; Hafeez, H.Y. Study of asymmetric hybrid supercapacitor using carbon and metal oxides as electrode materials. Indian J. Sci. Technol. 2016, 9. [Google Scholar] [CrossRef]
  80. Sahoo, R.; Lee, T.H.; Pham, D.T.; Luu, T.H.T.; Lee, Y.H. Fast-Charging High-Energy Battery-Supercapacitor Hybrid: Anodic Reduced Graphene Oxide-Vanadium(IV) Oxide Sheet-on-Sheet Heterostructure. ACS Nano 2019, 13, 10776–10786. [Google Scholar] [CrossRef]
  81. Chen, Y.; Zhang, X.; Zhang, H.; Sun, X.; Zhang, D.; Ma, Y. High-performance supercapacitors based on a graphene-activated carbon composite prepared by chemical activation. RSC Adv. 2012, 2, 7747–7753. [Google Scholar] [CrossRef]
  82. Moreno-Fernandez, G.; Ibañez, J.; Rojo, J.M.; Kunowsky, M. Activated Carbon Fiber Monoliths as Supercapacitor Electrodes. Adv. Mater. Sci. Eng. 2017, 2017, 1–8. [Google Scholar] [CrossRef] [Green Version]
  83. Shi, K.; Zhitomirsky, I. Fabrication of polypyrrole-coated carbon nanotubes using oxidant-surfactant nanocrystals for supercapacitor electrodes with high mass loading and enhanced performance. ACS Appl. Mater. Interfaces 2013, 5, 13161–13170. [Google Scholar] [CrossRef]
  84. Zou, Q.; Khalafallah, D.; Wu, Z.; Chen, J.; Zhi, M.; Hong, Z. Supercritical ethanol deposition of Ni(OH)2 nanosheets on carbon cloth for flexible solid-state asymmetric supercapacitor electrode. J. Supercrit. Fluids 2020, 159, 104774. [Google Scholar] [CrossRef]
  85. Syarif, N.; Tribidasari, I.; Wibowo, W. Direct Synthesis Carbon/Metal Oxide Composites for Electrochemical Capacitors Electrode. Int. Trans. J. Eng. Manag. Appl. Sci. Technol. 2012, 3, 21–34. Available online: http://tuengr.com/V03/21-34.pdf (accessed on 3 January 2020).
  86. Bessaguet, C.; Dantras, E.; Lacabanne, C.; Chevalier, M.; Michon, G. Piezoelectric and mechanical behavior of NaNbO3/PEKK lead-free nanocomposites. J. Non. Cryst. Solids 2017, 459, 83–87. [Google Scholar] [CrossRef] [Green Version]
  87. Pandolfo, A.G.; Hollenkamp, A.F. Carbon properties and their role in supercapacitors. J. Power Sources 2006, 157, 11–27. [Google Scholar] [CrossRef]
  88. Simon, P.; Gogotsi, Y. Materials for Electrochemical Capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Salunkhe, R.R.; Lin, J.; Malgras, V.; Dou, S.X.; Kim, J.H.; Yamauchi, Y. Large-scale synthesis of coaxial carbon nanotube/Ni(OH)2 composites for asymmetric supercapacitor application. Nano Energy 2015, 11, 211–218. [Google Scholar] [CrossRef]
  90. Yan, J.; Wang, Q.; Wei, T.; Fan, Z. Recent Advances in Design and Fabrication of Electrochemical Supercapacitors with High Energy Densities. Adv. Energy Mater. 2014, 4, 1300816. [Google Scholar] [CrossRef]
  91. Choudhary, N.; Li, C.; Moore, J.; Nagaiah, N.; Zhai, L.; Jung, Y.; Thomas, J. Asymmetric Supercapacitor Electrodes and Devices. Adv. Mater. 2017, 29. [Google Scholar] [CrossRef]
  92. Yuan, C.; Wu, H.B.; Xie, Y.; Lou, X.W.D. Mixed Transition-Metal Oxides: Design, Synthesis, and Energy-Related Applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [Google Scholar] [CrossRef]
  93. Simon, P.; Burke, A. Nanostructured carbons: Double-layer capacitance and more. Electrochem. Soc. Interface 2008, 17, 38–43. [Google Scholar]
  94. Raymundo-Piñero, E.; Leroux, F.; Béguin, F. A high-performance carbon for supercapacitors obtained by carbonization of a seaweed biopolymer. Adv. Mater. 2006, 18, 1877–1882. [Google Scholar] [CrossRef]
  95. Ali, A.; Jamal, R.; shao, W.; Rahman, A.; Osman, Y.; Abdiryim, T. Structure and properties of solid-state synthesized poly(3,4-propylenedioxythiophene)/nano-ZnO composite. Prog. Nat. Sci. Mater. Int. 2013, 23, 524–531. [Google Scholar] [CrossRef] [Green Version]
  96. Zheng, H.; Wang, J.; Jia, Y.; Ma, C. In-situ synthetize multi-walled carbon nanotubes@MnO2 nanoflake core-shell structured materials for supercapacitors. J. Power Sources 2012, 216, 508–514. [Google Scholar] [CrossRef]
  97. Ania, C.O.; Khomenko, V.; Raymundo-Piñero, E.; Parra, J.B.; Béguin, F. The Large Electrochemical Capacitance of Microporous Doped Carbon Obtained by Using a Zeolite Template. Adv. Funct. Mater. 2007, 17, 1828–1836. [Google Scholar] [CrossRef] [Green Version]
  98. Zhong, C.; Deng, Y.; Hu, W.; Qiao, J.; Zhang, L.; Zhang, J. A review of electrolyte materials and compositions for electrochemical supercapacitors. Chem. Soc. Rev. 2015, 44, 7484–7539. [Google Scholar] [CrossRef]
  99. Li, L.; Wang, X.; Wang, S.; Cao, Z.; Wu, Z.; Wang, H.; Gao, Y.; Liu, J. Activated Carbon Prepared from Lignite as Supercapacitor Electrode Materials. Electroanalysis 2016, 28, 243–248. [Google Scholar] [CrossRef]
  100. Zhang, M.; Cheng, J.; Zhang, L.; Li, Y.; Chen, M.-S.; Chen, Y.; Shen, Z. Activated carbon by one-step calcination of deoxygenated agar for high voltage lithium ion supercapacitor. ACS Sustain. Chem. Eng. 2020, 8, 3637–3643. [Google Scholar] [CrossRef]
  101. Bordun, I.; Pohrebennyk, V.; Sadowa, M.; Ptashnyk, V.; Klos-Witkowska, A.; Martsenyuk, V. Impedance spectroscopy of supercapacitors on the basis on modified by the ultrasound activated carbon material. In Proceedings of the 2017 IEEE 9th International Conference on Intelligent Data Acquisition and Advanced Computing Systems: Technology and Applications, IDAACS 2017, Bucharest, Romania, 21–23 September 2017; pp. 86–90. [Google Scholar] [CrossRef]
  102. Grishchenko, L.M.; Tsapyuk, G.G.; Ricco, M.; Diyuk, V.E.; Boldyrieva, O.Y.; Mariychuk, R.; Matushko, I.P.; Pontiroli, D.; Lisnyak, V.V.; Scaravonati, S. Enhancing the Performance of Supercapacitor Activated Carbon Electrodes by Oxidation. In Proceedings of the 2020 IEEE 40th International Conference on Electronics and Nanotechnology, ELNANO, Kyiv, Ukraine, 22–24 April 2020; Igor Sikordky Kyiv Polytechnic Instituite: Kyiv, Ukraine; pp. 173–177. [Google Scholar] [CrossRef]
  103. Maria, F.R.; Raj, S.; Jaya, N.V.; Boopathi, G.; Kalpana, D.; Pandurangan, A. S-doped activated mesoporous carbon derived from the Borassus flabellifer flower as active electrodes for supercapacitors. Mater. Chem. Phys. 2019, 240, 122151. [Google Scholar] [CrossRef]
  104. Cheng, F.; Yang, X.; Zhang, S.; Lu, W. Boosting the supercapacitor performances of activated carbon with carbon nanomaterials. J. Power Sources 2019, 450, 227678. [Google Scholar] [CrossRef]
  105. Chen, W.C.; Wen, T.C.; Teng, H. Polyaniline-deposited porous carbon electrode for supercapacitor. Electrochim. Acta 2003, 48, 641–649. [Google Scholar] [CrossRef]
  106. Inal, I.I.G.; Gokce, Y.; Aktas, Z. Waste tea derived activated carbon/polyaniline composites as supercapacitor electrodes. In Proceedings of the 2016 IEEE International Conference on Renewable Energy Research and Applications, ICRERA 2016, Birmingham, UK, 20–23 November 2016; pp. 458–462. [Google Scholar] [CrossRef]
  107. Frackowiak, E.; Metenier, K.; Bertagna, V.; Beguin, F. Supercapacitor electrodes from multiwalled carbon nanotubes. Appl. Phys. Lett. 2000, 77, 2421–2423. [Google Scholar] [CrossRef]
  108. Cheng, Q.; Tang, J.; Ma, J.; Zhang, H.; Shinya, N.; Qin, L.C. Graphene and carbon nanotube composite electrodes for supercapacitors with ultra-high energy density. Phys. Chem. Chem. Phys. 2011, 13, 17615–17624. [Google Scholar] [CrossRef] [PubMed]
  109. Du, C.; Yeh, J.; Pan, N. High power density supercapacitors using locally aligned carbon nanotube electrodes. Nanotechnology 2005, 16, 350–353. [Google Scholar] [CrossRef]
  110. Shi, R.; Jiang, L.; Pan, C. A Single-Step Process for Preparing Supercapacitor Electrodes from Carbon Nanotubes. Soft Nanosci. Lett. 2011, 1, 11–15. [Google Scholar] [CrossRef] [Green Version]
  111. Rajender, I.; Mohammad, B.; Ahmer, F.; Asiri, A.M. Morphology Design Paradigms for Supercapacitors; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  112. Li, X.; Rong, J.; Wei, B. Electrochemical Behavior of Single-Walled Carbon Nanotube Supercapacitors under Compressive Stress. ACS Nano 2010, 4, 6039–6049. [Google Scholar] [CrossRef]
  113. Dillon, A.C. Carbon Nanotubes for Photoconversion and Electrical Energy Storage. Chem. Rev. 2010, 110, 6856–6872. [Google Scholar] [CrossRef]
  114. Niu, C.; Sichel, E.K.; Hoch, R.; Moy, D.; Tennent, H. High power electrochemical capacitors based on carbon nanotube electrodes. Appl. Phys. Lett. 1997, 70, 1480–1482. [Google Scholar] [CrossRef]
  115. Peng, C.; Jin, J.; Chen, G.Z. A comparative study on electrochemical co-deposition and capacitance of composite films of conducting polymers and carbon nanotubes. Electrochim. Acta 2007, 53, 525–537. [Google Scholar] [CrossRef]
  116. Wang, G.; Hou, C.; Wang, H. (Eds.) Flexible and Wearable Electronics for Smart Clothing; Wiley: Hoboken, NJ, USA, 2020. [Google Scholar]
  117. Liu, Q.; Yang, J.; Luo, X.; Miao, Y.; Zhang, Y.; Xu, W.; Yang, L.; Liang, Y.; Weng, W.; Zhu, M. Fabrication of a fibrous MnO2@MXene/CNT electrode for high-performance flexible supercapacitor. Ceram. Int. 2020, 46, 11874–11881. [Google Scholar] [CrossRef]
  118. Yang, L.; Zheng, W.; Zhang, P.; Chen, J.; Tian, W.; Zhang, Y.; Sun, Z. MXene/CNTs films prepared by electrophoretic deposition for supercapacitor electrodes. J. Electroanal. Chem. 2018, 830–831, 1–6. [Google Scholar] [CrossRef]
  119. Shimamoto, K.; Tadanaga, K.; Tatsumisago, M. All-solid-state electrochemical capacitors using MnO2/carbonnanotube composite electrode. Electrochim. Acta 2013, 109, 651–655. [Google Scholar] [CrossRef] [Green Version]
  120. Alam, S.N.; Sharma, N.; Kumar, L. Synthesis of Graphene Oxide (GO) by Modified Hummers Method and Its Thermal Reduction to Obtain Reduced Graphene Oxide (rGO)*. Graphene 2017, 6, 1–18. [Google Scholar] [CrossRef] [Green Version]
  121. Geim, A.K. Graphene: Status and prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef] [PubMed]
  123. Pumera, M. Graphene-based nanomaterials and their electrochemistry. Chem. Soc. Rev. 2010, 39, 4146–4157. [Google Scholar] [CrossRef]
  124. Tan, Y.B.; Lee, J.M. Graphene for supercapacitor applications. J. Mater. Chem. A 2013, 1, 14814–14843. [Google Scholar] [CrossRef]
  125. Liu, C.; Yu, Z.; Neff, D.; Zhamu, A.; Jang, B.Z. Graphene-Based Supercapacitor with an Ultrahigh Energy Density. Nano Lett. 2010, 10, 4863–4868. [Google Scholar] [CrossRef] [PubMed]
  126. Carbon Nanotube Fibres and Yarns: Production, Properties and Applications in Google Books. Available online: https://books.google.ie/books?id=uI6vDwAAQBAJ&pg=PA268&lpg=PA268&dq=H.+M.+Cheng,+IEEE,+(2010)+49&source=bl&ots=u8jvoQ_eMN&sig=ACfU3U2SIRQFlU2dHpKz1sdHXKK0VnyyMg&hl=en&sa=X&ved=2ahUKEwjv5suTz-zmAhU-XRUIHX_jAz0Q6AEwAXoECAkQAQ#v=onepage&q=H. M. Cheng%2C IEEE (accessed on 24 June 2020).
  127. Wang, H.; Hao, Q.; Yang, X.; Lu, L.; Wang, X. A nanostructured graphene/polyaniline hybrid material for supercapacitors. Nanoscale 2010, 2, 2164–2170. [Google Scholar] [CrossRef]
  128. Karthika, P.; Rajalakshmi, N.; Dhathathreyan, K.S. Functionalized Exfoliated Graphene Oxide as Supercapacitor Electrodes. Soft Nanosci. Lett. 2012, 2, 59–66. [Google Scholar] [CrossRef] [Green Version]
  129. Kim, T.; Jung, G.; Yoo, S.; Suh, K.S.; Ruoff, R.S. Activated Graphene-Based Carbons as Supercapacitor Electrodes with Macro- and Mesopores. ACS Nano 2013, 7, 6899–6905. [Google Scholar] [CrossRef]
  130. Sinniah, G.K.; Shah, M.Z.; Siong, H.C. The needs for changes in travel behaviour towards a Low Carbon Society. In Proceedings of the 26th International Business Information Management Association Conference—Innovation Management and Sustainable Economic Competitive Advantage: From Regional Development to Global Growth, IBIMA 2015, Madrid, Spain, 11–12 November 2015; pp. 3868–3876. [Google Scholar]
  131. Kuilla, T.; Bhadra, S.; Yao, D.; Kim, N.H.; Bose, S.; Lee, J.H. Recent advances in graphene based polymer composites. Prog. Polym. Sci. 2010, 35, 1350–1375. [Google Scholar] [CrossRef]
  132. Wang, Y.; Shi, Z.; Huang, Y.; Ma, Y.; Wang, C.; Chen, M.; Chen, Y. Supercapacitor devices based on graphene materials. J. Phys. Chem. C 2009, 113, 13103–13107. [Google Scholar] [CrossRef]
  133. Zhang, L.L.; Zhao, X.; Stoller, M.D.; Zhu, Y.; Ji, H.; Murali, S.; Wu, Y.; Perales, S.; Clevenger, B.; Ruoff, R.S. Highly conductive and porous activated reduced graphene oxide films for high-power supercapacitors. Nano Lett. 2012, 12, 1806–1812. [Google Scholar] [CrossRef]
  134. Paek, E.; Pak, A.J.; Kweon, K.E.; Hwang, G.S. On the origin of the enhanced supercapacitor performance of nitrogen-doped graphene. J. Phys. Chem. C 2013, 117, 5610–5616. [Google Scholar] [CrossRef]
  135. Palaniselvam, T.; Kashyap, V.; Bhange, S.N.; Baek, J.B.; Kurungot, S. Nanoporous Graphene Enriched with Fe/Co-N Active Sites as a Promising Oxygen Reduction Electrocatalyst for Anion Exchange Membrane Fuel Cells. Adv. Funct. Mater. 2016, 26, 2150–2162. [Google Scholar] [CrossRef]
  136. Singh, S.K.; Dhavale, V.M.; Boukherroub, R.; Kurungot, S.; Szunerits, S. N-doped porous reduced graphene oxide as an efficient electrode material for high performance flexible solid-state supercapacitor. Appl. Mater. Today 2017, 8, 141–149. [Google Scholar] [CrossRef]
  137. Li, N.; Tang, S.; Dai, Y.; Meng, X. The synthesis of graphene oxide nanostructures for supercapacitors: A simple route. J. Mater. Sci. 2014, 49, 2802–2809. [Google Scholar] [CrossRef]
  138. Ogata, C.; Kurogi, R.; Hatakeyama, K.; Taniguchi, T.; Koinuma, M.; Matsumoto, Y. All-graphene oxide device with tunable supercapacitor and battery behaviour by the working voltage. Chem. Commun. 2016, 52, 3919–3922. [Google Scholar] [CrossRef] [PubMed]
  139. Zhang, H.; Wang, K.; Zhang, X.; Lin, H.; Sun, X.; Li, C.; Ma, Y. Self-generating graphene and porous nanocarbon composites for capacitive energy storage. J. Mater. Chem. A 2015, 3, 11277–11286. [Google Scholar] [CrossRef]
  140. Karnan, M.; Subramani, K.; Sathish, M. Template Assisted Synthesis of Nitrogen doped 3D-Graphene for Supercapacitor Applications. Mater. Today Proc. 2017, 4, 12144–12151. [Google Scholar] [CrossRef]
  141. Wang, G.; Sun, X.; Lu, F.; Sun, H.; Yu, M.; Jiang, W.; Liu, C.; Lian, J. Flexible Pillared Graphene-Paper Electrodes for High-Performance Electrochemical Supercapacitors. Small 2012, 8, 452–459. [Google Scholar] [CrossRef]
  142. Shabangoli, Y.; Rahmanifar, M.S.; Noori, A.; El-Kady, M.F.; Kaner, R.B.; Mousavi, M.F. Nile Blue Functionalized Graphene Aerogel as a Pseudocapacitive Negative Electrode Material across the Full pH Range. ACS Nano 2019, 13, 12567–12576. [Google Scholar] [CrossRef] [PubMed]
  143. Casiraghi, C.; Hartschuh, A.; Lidorikis, E.; Qian, H.; Harutyunyan, H.; Gokus, T.; Novoselov, K.S.; Ferrari, A.C. Rayleigh imaging of graphene and graphene layers. Nano Lett. 2007, 7, 2711–2717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Xu, L.; Cheng, C.; Yao, C.; Jin, X. Flexible supercapacitor electrode based on lignosulfonate-derived graphene quantum dots/graphene hydrogel. Org. Electron. 2020, 78, 105407. [Google Scholar] [CrossRef]
  145. Nguyen, T.; Montemor, M.F. Redox active materials for metal compound based hybrid electrochemical energy storage: A perspective view. Appl. Surf. Sci. 2017, 422, 492–497. [Google Scholar] [CrossRef]
  146. Conway, B.E. Transition from Supercapacitor to Battery Behavior in Electrochemical Energy Storage. J. Electrochem. Soc. 1991, 138, 1539–1548. [Google Scholar] [CrossRef]
  147. Zhang, Y.; Ma, Q.; Feng, K.; Guo, J.; Wei, X.; Shao, Y.; Zhuang, J.; Lin, T. Effects of microstructure and electrochemical properties of Ti/IrO2–SnO2–Ta2O5 as anodes on binder-free asymmetric supercapacitors with Ti/RuO2–NiO as cathodes. Ceram. Int. 2020, 46, 17640–17650. [Google Scholar] [CrossRef]
  148. Guo, W.; Sun, W.; Wang, Y. Multilayer CuO@NiO Hollow Spheres: Microwave-Assisted Metal-Organic-Framework Derivation and Highly Reversible Structure-Matched Stepwise Lithium Storage. ACS Nano 2015, 9, 11462–11471. [Google Scholar] [CrossRef]
  149. Hu, L.; Huang, Y.; Zhang, F.; Chen, Q. CuO/Cu2O composite hollow polyhedrons fabricated from metal-organic framework templates for lithium-ion battery anodes with a long cycling life. Nanoscale 2013, 5, 4186–4190. [Google Scholar] [CrossRef]
  150. Mondal, I.; Pal, U. Synthesis of MOF templated Cu/CuO@TiO2 nanocomposites for synergistic hydrogen production. Phys. Chem. Chem. Phys. 2016, 18, 4780–4788. [Google Scholar] [CrossRef]
  151. Dekrafft, K.E.; Wang, C.; Lin, W. Metal-organic framework templated synthesis of Fe2O3/TiO2 nanocomposite for hydrogen production. Adv. Mater. 2012, 24, 2014–2018. [Google Scholar] [CrossRef]
  152. Huang, G.; Zhang, L.; Zhang, F.; Wang, L. Metal-organic framework derived Fe2O3@NiCo2O4 porous nanocages as anode materials for Li-ion batteries. Nanoscale 2014, 6, 5509–5515. [Google Scholar] [CrossRef] [PubMed]
  153. Kaneti, Y.V.; Zakaria, Q.M.D.; Zhang, Z.; Chen, C.; Yue, J.; Liu, M.; Jiang, X.; Yu, A. Solvothermal synthesis of ZnO-decorated α-Fe2O3 nanorods with highly enhanced gas-sensing performance toward n-butanol. J. Mater. Chem. A 2014, 2, 13283–13292. [Google Scholar] [CrossRef]
  154. Zhou, X.; Li, X.; Sun, H.; Sun, P.; Liang, X.; Liu, F.; Hu, X.; Lu, G. Nanosheet-Assembled ZnFe2O4 Hollow Microspheres for High-Sensitive Acetone Sensor. ACS Appl. Mater. Interfaces 2015, 7, 15414–15421. [Google Scholar] [CrossRef] [PubMed]
  155. Hu, H.; Guan, B.; Xia, B.; Lou, X.W. Designed formation of Co3O4/NiCo2O4 double-shelled nanocages with enhanced pseudocapacitive and electrocatalytic properties. J. Am. Chem. Soc. 2015, 137, 5590–5595. [Google Scholar] [CrossRef] [PubMed]
  156. Yang, X.; Xue, H.; Yang, Q.; Yuan, R.; Kang, W.; Lee, C.S. Preparation of porous ZnO/ZnFe2O4 composite from metal organic frameworks and its applications for lithium ion batteries. Chem. Eng. J. 2017, 308, 340–346. [Google Scholar] [CrossRef]
  157. Hou, L.; Lian, L.; Zhang, L.; Pang, G.; Yuan, C.; Zhang, X. Self-sacrifice template fabrication of hierarchical mesoporous bi-component-active ZnO/ZnFe2O4 sub-microcubes as superior anode towards high-performance lithium-ion battery. Adv. Funct. Mater. 2015, 25, 238–246. [Google Scholar] [CrossRef]
  158. Cao, H.; Zhu, S.; Yang, C.; Bao, R.; Tong, L.; Hou, L.; Zhang, X.; Yuan, C. Metal-organic-framework-derived two-dimensional ultrathin mesoporous hetero-ZnFe2O4/ZnO nanosheets with enhanced lithium storage properties for Li-ion batteries. Nanotechnology 2016, 27, 465402. [Google Scholar] [CrossRef]
  159. Zhang, W.; Liu, Y.; Lu, G.; Wang, Y.; Li, S.; Cui, C.; Wu, J.; Xu, Z.; Tian, D.; Huang, W.; et al. Mesoporous metal-organic frameworks with size-, shape-, and space-distribution-controlled pore structure. Adv. Mater. 2015, 27, 2923–2929. [Google Scholar] [CrossRef]
  160. Huang, G.; Zhang, F.; Zhang, L.; Du, X.; Wang, J.; Wang, L. Hierarchical NiFe2O4/Fe2O3 nanotubes derived from metal organic frameworks for superior lithium ion battery anodes. J. Mater. Chem. A 2014, 2, 8048–8053. [Google Scholar] [CrossRef]
  161. Li, G.C.; Zhu, S.-R.; Wu, M.-K.; Liu, P.-F.; Tao, K.; Yi, F.-Y.; Han, L. MOF-derived hierarchical double-shelled NiO/ZnO hollow spheres for high-performance supercapacitors. Dalton Trans. 2016, 45, 13311–13316. [Google Scholar] [CrossRef]
  162. Xu, X.; Cao, K.; Wang, Y.; Jiao, L. 3D hierarchical porous ZnO/ZnCo2O4 nanosheets as high-rate anode material for lithium-ion batteries. J. Mater. Chem. A 2016, 4, 6042–6047. [Google Scholar] [CrossRef]
  163. Salunkhe, R.R.; Kaneti, Y.V.; Yamauchi, Y. Metal-Organic Framework-Derived Nanoporous Metal Oxides toward Supercapacitor Applications: Progress and Prospects. ACS Nano 2017, 11, 5293–5308. [Google Scholar] [CrossRef]
  164. Li, T.T.; Qian, J.; Zheng, Y.Q. Facile synthesis of porous CuO polyhedron from Cu-based metal organic framework (MOF-199) for electrocatalytic water oxidation. RSC Adv. 2016, 6, 77358–77365. [Google Scholar] [CrossRef]
  165. Hu, L.; Yan, N.; Chen, Q.W.; Zhang, P.; Zhong, H.; Zheng, X.; Li, Y.; Hu, X. Fabrication Based on the Kirkendall Effect of Co3O4 Porous Nanocages with Extraordinarily High Capacity for Lithium Storage. Chem. A Eur. J. 2012, 18, 8971–8977. [Google Scholar] [CrossRef]
  166. Salunkhe, R.R.; Tang, J.; Kamachi, Y.; Nakato, T.; Kim, J.H.; Yamauchi, Y. Asymmetric Supercapacitors Using 3D Nanoporous Carbon and Cobalt Oxide Electrodes Synthesized from a Single Metal-Organic Framework. ACS Nano 2015, 9, 6288–6296. [Google Scholar] [CrossRef] [PubMed]
  167. Wu, R.; Qian, X.; Ruia, X.; Liu, H.; Yadian, B.; Zhou, K.; Wei, J.; Yan, Q.; Feng, X.-Q.; Long, Y.; et al. Zeolitic imidazolate framework 67-derived high symmetric porous Co3O4 hollow dodecahedra with highly enhanced lithium storage capability. Small 2014, 10, 1932–1938. [Google Scholar] [CrossRef]
  168. Wang, Y.; Wang, B.-F.; Xiao, F.; Huang, Z.-G.; Wang, Y.; Richardson, C.; Chen, Z.; Jiao, L.; Yuan, H. Facile synthesis of nanocage Co3O4 for advanced lithium-ion batteries. J. Power Sources 2015, 298, 203–208. [Google Scholar] [CrossRef]
  169. Liu, H.; Xia, G.; Zhang, R.; Jiang, P.; Chen, J.; Chen, Q. MOF-derived RuO2/Co3O4 heterojunctions as highly efficient bifunctional electrocatalysts for HER and OER in alkaline solutions. RSC Adv. 2017, 7, 3686–3694. [Google Scholar] [CrossRef] [Green Version]
  170. Han, Y.; Zhao, M.; Dong, L.; Feng, J.; Wang, Y.; Li, D.-J.; Li, X.-F. MOF-derived porous hollow Co3O4 parallelepipeds for building high-performance Li-ion batteries. J. Mater. Chem. A 2015, 3, 22542–22546. [Google Scholar] [CrossRef]
  171. Tian, D.; Zhou, X.L.; Zhang, Y.H.; Zhou, Z.; Bu, X.H. MOF-Derived Porous Co3O4 Hollow Tetrahedra with Excellent Performance as Anode Materials for Lithium-Ion Batteries. Inorg. Chem. 2015, 54, 8159–8161. [Google Scholar] [CrossRef] [PubMed]
  172. Li, C.; Chen, T.; Xu, W.; Lou, X.; Pan, L.; Chen, Q.; Hu, B. Mesoporous nanostructured Co3O4 derived from MOF template: A high-performance anode material for lithium-ion batteries. J. Mater. Chem. A 2015, 3, 5585–5591. [Google Scholar] [CrossRef]
  173. Soundharrajan, V.; Sambandam, B.; Song, J.; Kim, S.; Jo, J.; Kim, S.; Lee, S.; Mathew, V.; Kim, J. Co3V2O8 Sponge Network Morphology Derived from Metal-Organic Framework as an Excellent Lithium Storage Anode Material. ACS Appl. Mater. Interfaces 2016, 8, 8546–8553. [Google Scholar] [CrossRef] [PubMed]
  174. Zeng, G.; Chen, Y.; Chen, L.; Xiong, P.; Wei, M. Hierarchical cerium oxide derived from metal-organic frameworks for high performance supercapacitor electrodes. Electrochim. Acta 2016, 222, 773–780. [Google Scholar] [CrossRef]
  175. Zhang, L.; Wu, H.B.; Madhavi, S.; Hng, H.H.; Lou, X.W. Formation of Fe2O3 microboxes with hierarchical shell structures from metal-organic frameworks and their lithium storage properties. J. Am. Chem. Soc. 2012, 134, 17388–17391. [Google Scholar] [CrossRef]
  176. Zhang, F.; Wei, Y.; Wu, X.; Jiang, H.; Wang, W.; Li, H. Hollow zeolitic imidazolate framework nanospheres as highly efficient cooperative catalysts for [3 + 3] cycloaddition reactions. J. Am. Chem. Soc. 2014, 136, 13963–13966. [Google Scholar] [CrossRef]
  177. Xu, X.; Cao, R.; Jeong, S.; Cho, J. Spindle-like mesoporous α-Fe2O3 anode material prepared from MOF template for high-rate lithium batteries. Nano Lett. 2012, 12, 4988–4991. [Google Scholar] [CrossRef]
  178. Kim, T.K.; Lee, K.J.; Cheon, J.Y.; Lee, J.H.; Joo, S.H.; Moon, H.R. Nanoporous metal oxides with tunable and nanocrystalline frameworks via conversion of metal-organic frameworks. J. Am. Chem. Soc. 2013, 135, 8940–8946. [Google Scholar] [CrossRef]
  179. Zheng, F.; Xu, S.; Yin, Z.; Zhang, Y.; Lu, L. Facile synthesis of MOF-derived Mn2O3 hollow microspheres as anode materials for lithium-ion batteries. RSC Adv. 2016, 6, 93532–93538. [Google Scholar] [CrossRef]
  180. Maiti, S.; Pramanik, A.; Mahanty, S. Electrochemical energy storage in Mn2O3 porous nanobars derived from morphology-conserved transformation of benzenetricarboxylate-bridged metal-organic framework. CrystEngComm 2016, 18, 450–461. [Google Scholar] [CrossRef]
  181. Zhang, B.; Hao, S.; Xiao, D.; Wu, J.; Huang, Y. Templated formation of porous Mn2O3 octahedra from Mn-MIL-100 for lithium-ion battery anode materials. Mater. Des. 2016, 98, 319–323. [Google Scholar] [CrossRef]
  182. Kong, S.; Dai, R.; Li, H.; Sun, W.; Wang, Y. Microwave Hydrothermal Synthesis of Ni-based Metal-Organic Frameworks and Their Derived Yolk-Shell NiO for Li-Ion Storage and Supported Ammonia Borane for Hydrogen Desorption. ACS Sustain. Chem. Eng. 2015, 3, 1830–1838. [Google Scholar] [CrossRef]
  183. Gou, L.; Ma, L.; Zhao, M.-J.; Liu, P.-G.; Wang, X.-D.; Fan, X.-Y.; Li, D.-L. Co-based metal–organic framework and its derivatives as high-performance anode materials for lithium-ion batteries. J. Mater. Sci. 2019, 54, 1529–1538. [Google Scholar] [CrossRef]
  184. Han, Y.; Zhang, S.; Shen, N.; Li, D.; Li, X. MOF-derived porous NiO nanoparticle architecture for high performance supercapacitors. Mater. Lett. 2017, 188, 1–4. [Google Scholar] [CrossRef]
  185. Xiu, Z.; Alfaruqi, M.H.; Gim, J.; Song, J.; Kim, S.; Thi, T.V.; Duong, P.T.; Baboo, J.P.; Mathew, V.; Kim, J. Hierarchical porous anatase TiO2 derived from a titanium metal-organic framework as a superior anode material for lithium ion batteries. Chem. Commun. 2015, 51, 12274–12277. [Google Scholar] [CrossRef] [Green Version]
  186. Zheng, Y.; Zheng, S.; Xue, H.; Pang, H. Metal-organic frameworks for lithium-sulfur batteries. J. Mater. Chem. A 2019, 7, 3469–3491. [Google Scholar] [CrossRef]
  187. Zhang, X.; Wang, M.; Zhu, G.; Li, D.; Yan, D.; Lu, T.; Pan, L. Porous cake-like TiO2 derived from metal-organic frameworks as superior anode material for sodium ion batteries. Ceram. Int. 2017, 43, 2398–2402. [Google Scholar] [CrossRef]
  188. Kimitsuka, Y.; Hosono, E.; Ueno, S.; Zhou, H.; Fujihara, S. Fabrication of porous cubic architecture of ZnO using Zn-terephthalate MOFs with characteristic microstructures. Inorg. Chem. 2013, 52, 14028–14033. [Google Scholar] [CrossRef]
  189. Wang, T.; Shi, L.; Tang, J.; Malgras, V.; Asahina, S.; Liu, G.; Zhang, H.; Meng, X.; Chang, K.; He, J.; et al. A Co3O4-embedded porous ZnO rhombic dodecahedron prepared using zeolitic imidazolate frameworks as precursors for CO2 photoreduction. Nanoscale 2016, 8, 6712–6720. [Google Scholar] [CrossRef]
  190. Feng, Y.; Lu, H.; Gu, X.; Qiu, J.; Jia, M.; Huang, C.; Yao, J. ZIF-8 derived porous N-doped ZnO with enhanced visible light-driven photocatalytic activity. J. Phys. Chem. Solids 2017, 102, 110–114. [Google Scholar] [CrossRef]
  191. Guo, H.; Li, T.; Chen, W.; Liu, L.; Yang, X.-J.; Wang, Y.; Guo, Y. General design of hollow porous CoFe2O4 nanocubes from metal-organic frameworks with extraordinary lithium storage. Nanoscale 2014, 6, 15168–15174. [Google Scholar] [CrossRef]
  192. Liu, F.; Zhuang, Y.; Guo, M.; Chen, Y.; Tu, J.; Ding, L. 3D Copper Foam-Supported CuCo2O4 Nanosheet Arrays as Electrode for Enhanced Non-Enzymatic Glucose Sensing. Sensors 2018, 18, 1131. [Google Scholar] [CrossRef] [Green Version]
  193. Zheng, F.; Zhu, D.; Shi, X.; Chen, Q. Metal-organic framework-derived porous Mn1.8Fe1.2O4 nanocubes with an interconnected channel structure as high-performance anodes for lithium ion batteries. J. Mater. Chem. A 2015, 3, 2815–2824. [Google Scholar] [CrossRef]
  194. Du, M.; He, D.; Lou, Y.; Chen, J. Porous nanostructured ZnCo2O4 derived from MOF-74: High-performance anode materials for lithium ion batteries. J. Energy Chem. 2017, 26, 673–680. [Google Scholar] [CrossRef] [Green Version]
  195. Liang, K.; Cheang, T.; Wen, T.; Xie, X.; Zhou, X.; Zhao, Z.; Shen, C.; Jiang, N.; Wangc, Y. Facile Preparation of Porous Mn2SnO4/Sn/C Composite Cubes as High Performance Anode Material for Lithium-Ion Batteries. J. Phys. Chem. C 2016, 120, 3669–3676. [Google Scholar] [CrossRef]
  196. Zhao, J.; Wang, F.; Su, P.; Li, M.; Chen, J.; Yang, Q.; Li, C. Spinel ZnMn2O4 nanoplate assemblies fabricated via ‘escape-by-crafty-scheme’ strategy. J. Mater. Chem. 2012, 22, 13328–13333. [Google Scholar] [CrossRef]
  197. Chen, Y.; Pang, W.K.; Bai, H.; Zhou, T.; Liu, Y.; Li, S.; Guo, Z. Enhanced Structural Stability of Nickel-Cobalt Hydroxide via Intrinsic Pillar Effect of Metaborate for High-Power and Long-Life Supercapacitor Electrodes. Nano Lett. 2017, 17, 429–436. [Google Scholar] [CrossRef]
  198. Liang, J.; Renzhi, M.; Iyi, N.; Ebina, Y.; Takada, K.; Sasaki, T. Topochemical synthesis, anion exchange, and exfoliation of Co-Ni layered double hydroxides: A route to positively charged Co-Ni hydroxide nanosheets with tunable composition. Chem. Mater. 2010, 22, 371–378. [Google Scholar] [CrossRef]
  199. Liu, H.; Liu, X.; Wang, S.; Liu, H.K.; Li, L. Transition metal based battery-type electrodes in hybrid supercapacitors: A review. Energy Storage Mater. 2020, 28, 122–145. [Google Scholar] [CrossRef]
  200. Feng, D.Y.; Sun, Z.; Huang, Z.H.; Cai, X.; Song, Y.; Liu, X.X. Highly loaded manganese oxide with high rate capability for capacitive applications. J. Power Sources 2018, 396, 238–245. [Google Scholar] [CrossRef]
  201. Lokhande, V.C.; Lokhande, A.C.; Lokhande, C.D.; Kim, J.H.; Ji, T. Supercapacitive composite metal oxide electrodes formed with carbon, metal oxides and conducting polymers. J. Alloys Compd. 2016, 682, 381–403. [Google Scholar] [CrossRef]
  202. Zhao, C.; Zheng, W. A Review for Aqueous Electrochemical Supercapacitors. Front. Energy Res. 2015, 3. [Google Scholar] [CrossRef] [Green Version]
  203. Krishnan, D.; Raidongia, K.; Shao, J.; Huang, J. Graphene Oxide Assisted Hydrothermal Carbonization of Carbon Hydrates. ACS Nano 2014, 8, 449–457. [Google Scholar] [CrossRef] [PubMed]
  204. Hwang, J.Y.; El-Kady, M.F.; Wang, Y.; Wang, L.; Shao, Y.; Marsh, K.; Ko, J.M.; Kaner, R.B. Direct preparation and processing of graphene/RuO2 nanocomposite electrodes for high-performance capacitive energy storage. Nano Energy 2015, 18, 57–70. [Google Scholar] [CrossRef]
  205. Zhao, D.; Guo, X.; Gao, Y.; Gao, F. An Electrochemical Capacitor Electrode Based on Porous Carbon Spheres Hybrided with Polyaniline and Nanoscale Ruthenium Oxide. ACS Appl. Mater. Interfaces 2012, 4, 5583–5589. [Google Scholar] [CrossRef] [PubMed]
  206. Wang, H.; Zhang, D.; Yan, T.; Wen, X.; Shi, L.; Zhang, J. Graphene prepared via a novel pyridine-thermal strategy for capacitive deionization. J. Mater. Chem. 2012, 22, 23745–23748. [Google Scholar] [CrossRef]
  207. Long, J.W.; Swider, K.E.; Merzbacher, C.I.; Rolison, D.R. Voltammetric Characterization of Ruthenium Oxide-Based Aerogels and Other RuO2 Solids: The Nature of Capacitance in Nanostructured Materials. Langmuir 1999, 15, 780–784. [Google Scholar] [CrossRef]
  208. Ma, H.; Kong, D.; Xu, Y.; Xie, X.; Tao, Y.; Xiao, Z.; Lv, W.; Jang, H.D.; Huang, J.; Yang, Q.-H. Disassembly—Reassembly Approach to RuO2/Graphene Composites for Ultrahigh Volumetric Capacitance Supercapacitor. Small 2017, 13, 1701026. [Google Scholar] [CrossRef]
  209. Gujar, T.P.; Kim, W.Y.; Puspitasari, I.; Jung, K.D.; Joo, O.S. Electrochemically deposited nanograin ruthenium oxide as a pseudocapacitive electrode. Int. J. Electrochem. Sci. 2007, 2, 666–673. [Google Scholar]
  210. Kim, H.; Popov, B.N. Characterization of hydrous ruthenium oxide/carbon nanocomposite supercapacitors prepared by a colloidal method. J. Power Sources 2002, 104, 52–61. [Google Scholar] [CrossRef]
  211. Kate, R.S.; Khalate, S.A.; Deokate, R.J. Overview of nanostructured metal oxides and pure nickel oxide (NiO) electrodes for supercapacitors: A review. J. Alloys Compd. 2018, 734, 89–111. [Google Scholar] [CrossRef]
  212. Kong, S.; Cheng, K.; Ouyang, T.; Gao, Y.; Ye, K.; Wang, G.; Cao, D. Facile electrodepositing processed of RuO2-graphene nanosheets-CNT composites as a binder-free electrode for electrochemical supercapacitors. Electrochim. Acta 2017, 246, 433–442. [Google Scholar] [CrossRef]
  213. Wang, J.-C.; Yao, H.-C.; Fan, Z.-Y.; Zhang, L.; Wang, J.-S.; Zang, S.-Q.; Li, Z. Indirect Z-Scheme BiOI/g-C3N4 Photocatalysts with Enhanced Photoreduction CO2 Activity under Visible Light Irradiation. ACS Appl. Mater. Interfaces 2016, 8, 3765–3775. [Google Scholar] [CrossRef] [PubMed]
  214. Acznik, I.; Lota, K.; Sierczynska, A.; Lota, G. Carbon-supported manganese dioxide as electrode material for asymmetric electrochemical capacitors. Int. J. Electrochem. Sci. 2014, 9, 2518–2534. [Google Scholar]
  215. Chen, S.; Zhu, J.; Wu, X.; Han, Q.; Wang, X. Graphene oxide-Mno2 nanocomposites for supercapacitors. ACS Nano 2010, 4, 2822–2830. [Google Scholar] [CrossRef]
  216. Zhang, H.; Wang, Z.; Liu, D. A comprehensive review of stability analysis of continuous-time recurrent neural networks. IEEE Trans. Neural Netw. Learn. Syst. 2014, 25, 1229–1262. [Google Scholar] [CrossRef]
  217. Toupin, M.; Brousse, T.; Bélanger, D. Charge storage mechanism of MnO2 electrode used in aqueous electrochemical capacitor. Chem. Mater. 2004, 16, 3184–3190. [Google Scholar] [CrossRef]
  218. Wang, J.G.; Kang, F.; Wei, B. Engineering of MnO2-based nanocomposites for high-performance supercapacitors. Prog. Mater. Sci. 2015, 74, 51–124. [Google Scholar] [CrossRef]
  219. Zhu, J.; Shi, W.; Xiao, N.; Ruia, X.; Tan, H.; Lu, X.; Hng, H.H.; Ma, J.; Yan, Q. Oxidation-Etching Preparation of MnO2 Tubular Nanostructures for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2012, 4, 2769–2774. [Google Scholar] [CrossRef]
  220. Gupta, V.; Kusahara, T.; Toyama, H.; Gupta, S.; Miura, N. Potentiostatically deposited nanostructured α-Co(OH)2: A high performance electrode material for redox-capacitors. Electrochem. Commun. 2007, 9, 2315–2319. [Google Scholar] [CrossRef]
  221. Subramanian, V.; Zhu, H.; Vajtai, R.; Ajayan, P.M.; Wei, B. Hydrothermal synthesis and pseudocapacitance properties of MnO2 nanostructures. J. Phys. Chem. B 2005, 109, 20207–20214. [Google Scholar] [CrossRef]
  222. Devaraj, S.; Munichandraiah, N. Effect of Crystallographic Structure of MnO2 on Its Electrochemical Capacitance Properties. J. Phys. Chem. C 2008, 112, 4406–4417. [Google Scholar] [CrossRef]
  223. Ghodbane, O.; Pascal, J.-L.; Favier, F. Microstructural Effects on Charge-Storage Properties in MnO2 -Based Electrochemical Supercapacitors. ACS Appl. Mater. Interfaces 2009, 1, 1130–1139. [Google Scholar] [CrossRef] [PubMed]
  224. Sun, C.; Zhang, Y.; Song, S.; Xue, D. Tunnel-dependent supercapacitance of MnO2: Effects of crystal structure. J. Appl. Crystallogr. 2013, 46, 1128–1135. [Google Scholar] [CrossRef]
  225. Chang, J.K.; Huang, C.H.; Lee, M.T.; Tsai, W.T.; Deng, M.J.; Sun, I.W. Physicochemical factors that affect the pseudocapacitance and cyclic stability of Mn oxide electrodes. Electrochim. Acta 2009, 54, 3278–3284. [Google Scholar] [CrossRef]
  226. Ghodbane, O.; Pascal, J.-L.; Fraisse, B.; Favier, F. Structural in Situ Study of the Thermal Behavior of Manganese Dioxide Materials: Toward Selected Electrode Materials for Supercapacitors. ACS Appl. Mater. Interfaces 2010, 2, 3493–3505. [Google Scholar] [CrossRef] [PubMed]
  227. Wei, J.; Nagarajan, N.; Zhitomirsky, I. Manganese oxide films for electrochemical supercapacitors. J. Mater. Process. Technol. 2007, 186, 356–361. [Google Scholar] [CrossRef]
  228. Dong, J.; Lu, G.; Wu, F.; Xu, C.; Kang, X.; Cheng, Z. Facile synthesis of a nitrogen-doped graphene flower-like MnO2 nanocomposite and its application in supercapacitors. Appl. Surf. Sci. 2018, 427, 986–993. [Google Scholar] [CrossRef]
  229. Zhao, D.D.; Xu, M.W.; Zhou, W.J.; Zhang, J.; Li, H.L. Preparation of ordered mesoporous nickel oxide film electrodes via lyotropic liquid crystal templated electrodeposition route. Electrochim. Acta 2008, 53, 2699–2705. [Google Scholar] [CrossRef]
  230. Wang, Y.G.; Xia, Y.Y. Electrochemical capacitance characterization of NiO with ordered mesoporous structure synthesized by template SBA-15. Electrochim. Acta 2006, 51, 3223–3227. [Google Scholar] [CrossRef]
  231. Sun, C. (Ed.) Advanced Battery Materials; Wiley: Hoboken, NJ, USA, 2019. [Google Scholar]
  232. Xiong, S.; Yuan, C.; Zhang, X.; Xi, B.; Qian, Y. Controllable Synthesis of Mesoporous Co3O4 Nanostructures with Tunable Morphology for Application in Supercapacitors. Chem. A Eur. J. 2009, 15, 5320–5326. [Google Scholar] [CrossRef]
  233. Bae, J.; Song, M.K.; Park, Y.J.; Kim, J.M.; Liu, M.; Wang, Z.L. Fiber supercapacitors made of nanowire-fiber hybrid structures for wearable/flexible energy storage. Angew. Chem. Int. Ed. 2011, 50, 1683–1687. [Google Scholar] [CrossRef] [PubMed]
  234. Pang, H.; Ma, Y.; Li, G.; Chen, J.; Zhang, J.; Zheng, H.; Du, W. Facile synthesis of porous ZnO-NiO composite micropolyhedrons and their application for high power supercapacitor electrode materials. Dalton Trans. 2012, 41, 13284–13291. [Google Scholar] [CrossRef] [PubMed]
  235. Ghosh, D.; Giri, S.; Mandal, M.; Das, C.K. High performance supercapacitor electrode material based on vertically aligned PANI grown on reduced graphene oxide/Ni(OH)2 hybrid composite. RSC Adv. 2014, 4, 26094–26101. [Google Scholar] [CrossRef]
  236. Ci, S.; Wen, Z.; Qian, Y.; Mao, S.; Cui, S.; Chen, J. NiO-Microflower Formed by Nanowire-weaving Nanosheets with Interconnected Ni-network Decoration as Supercapacitor Electrode. Sci. Rep. 2015, 5, 11919. [Google Scholar] [CrossRef] [PubMed]
  237. Li, B.; Zheng, M.; Xue, H.; Pang, H. High performance electrochemical capacitor materials focusing on nickel based materials. Inorg. Chem. Front. 2016, 3, 175–202. [Google Scholar] [CrossRef]
  238. Goel, S.; Tomar, A.K.; Sharma, R.K.; Singh, G. Highly Pseudocapacitive NiO Nanoflakes through Surfactant-Free Facile Microwave-Assisted Route. ACS Appl. Energy Mater. 2018, 1, 1540–1548. [Google Scholar] [CrossRef]
  239. Aghazadeh, M.; Rashidi, A.; Ganjali, M.R.; Maragheh, M.G. Nickel oxide nano-rods/plates as a high performance electrode materials for supercapacitors; electrosynthesis and evolution of charge storage ability. Int. J. Electrochem. Sci. 2016, 11, 11002–11015. [Google Scholar] [CrossRef]
  240. Pang, H.; Lu, Q.; Li, Y.; Gao, F. Facile synthesis of nickel oxide nanotubes and their antibacterial, electrochemical and magnetic properties. Chem. Commun. 2009, 7542–7544. [Google Scholar] [CrossRef]
  241. Kim, S.I.; Lee, J.S.; Ahn, H.J.; Song, H.K.; Jang, J.H. Facile route to an efficient nio supercapacitor with a three-dimensional nanonetwork morphology. ACS Appl. Mater. Interfaces 2013, 5, 1596–1603. [Google Scholar] [CrossRef]
  242. Cheng, S.; Yang, L.; Liu, Y.; Lin, W.; Huang, L.; Chen, D.; Wong, C.P.; Liu, M. Carbon fiber paper supported hybrid nanonet/nanoflower nickel oxide electrodes for high-performance pseudo-capacitors. J. Mater. Chem. A 2013, 1, 7709–7716. [Google Scholar] [CrossRef]
  243. Soin, N.; Roy, S.S.; Mitra, S.K.; Thundat, T.; McLaughlin, J.A. Nanocrystalline ruthenium oxide dispersed Few Layered Graphene (FLG) nanoflakes as supercapacitor electrodes. J. Mater. Chem. 2012, 22, 14944–14950. [Google Scholar] [CrossRef]
  244. Lu, Z.; Chang, Z.; Liu, J.; Sun, X. Stable ultrahigh specific capacitance of NiO nanorod arrays. Nano Res. 2011, 4, 658–665. [Google Scholar] [CrossRef]
  245. Kundu, M.; Liu, L. Binder-free electrodes consisting of porous NiO nanofibers directly electrospun on nickel foam for high-rate supercapacitors. Mater. Lett. 2015, 144, 114–118. [Google Scholar] [CrossRef]
  246. Advanced Biosensors for Health Care Applications; Elsevier: Amsterdam, The Netherlands, 2019.
  247. Liu, P.; Yang, M.; Zhou, S.; Huang, Y.; Zhu, Y. Hierarchical shell-core structures of concave spherical NiO nanospines@carbon for high performance supercapacitor electrodes. Electrochim. Acta 2019, 294, 383–390. [Google Scholar] [CrossRef]
  248. Zheng, D.; Zhao, F.; Li, Y.; Qin, C.; Zhu, J.; Hu, Q.; Wang, Z.; Inoue, A. Flexible NiO micro-rods/nanoporous Ni/metallic glass electrode with sandwich structure for high performance supercapacitors. Electrochim. Acta 2019, 297, 767–777. [Google Scholar] [CrossRef]
  249. Ede, S.R.; Anantharaj, S.; Kumaran, K.T.; Mishra, S.; Kundu, S. One step synthesis of Ni/Ni(OH)2 nano sheets (NSs) and their application in asymmetric supercapacitors. RSC Adv. 2017, 7, 5898–5911. [Google Scholar] [CrossRef] [Green Version]
  250. Wang, H.; Casalongue, H.S.; Liang, Y.; Dai, H. Ni(OH)2 Nanoplates Grown on Graphene as Advanced Electrochemical Pseudocapacitor Materials. J. Am. Chem. Soc. 2010, 132, 7472–7477. [Google Scholar] [CrossRef] [Green Version]
  251. Gomez, I.J.; Arnaiz, B.; Cacioppo, M.; Arcudi, F.; Prato, M. Nitrogen-doped Carbon Nanodots for bioimaging and delivery of paclitaxel. J. Mater. Chem. B 2018, 6, 1–3. [Google Scholar] [CrossRef] [Green Version]
  252. Xu, L.; Ding, Y.-S.; Chen, C.-H.; Zhao, L.; Rimkus, C.; Joesten, R.; Suib, S.L. 3D Flowerlike α-Nickel Hydroxide with Enhanced Electrochemical Activity Synthesized by Microwave-Assisted Hydrothermal Method. Chem. Mater. 2008, 20, 308–316. [Google Scholar] [CrossRef]
  253. Oesten, R.; Wohlfahrt-Mehrens, M.; Ströbele, S.; Kasper, M.; Huggins, R.A. Structural aspects of undoped and doped nickel hydroxides. Ionics 1996, 2, 293–301. [Google Scholar] [CrossRef]
  254. Zhao, D.D.; Bao, S.J.; Zhou, W.J.; Li, H.L. Preparation of hexagonal nanoporous nickel hydroxide film and its application for electrochemical capacitor. Electrochem. Commun. 2007, 9, 869–874. [Google Scholar] [CrossRef]
  255. Ramani, M.; Haran, B.S.; White, R.E.; Popov, B.N. Synthesis and Characterization of Hydrous Ruthenium Oxide-Carbon Supercapacitors. J. Electrochem. Soc. 2001, 148, A374. [Google Scholar] [CrossRef]
  256. Kuratani, K.; Kiyobayashi, T.; Kuriyama, N. Influence of the mesoporous structure on capacitance of the RuO2 electrode. J. Power Sources 2009, 189, 1284–1291. [Google Scholar] [CrossRef]
  257. Mohajernia, S.; Hejazi, S.; Mazare, A.; Nguyen, N.T.; Hwang, I.; Kment, Š.; Zoppellaro, G.; Tomanec, O.; Zbořil, R.; Schmuki, P. Semimetallic core-shell TiO2 nanotubes as a high conductivity scaffold and use in efficient 3D-RuO2 supercapacitors. Mater. Today Energy 2017, 6, 46–52. [Google Scholar] [CrossRef]
  258. Kovalenko, V.; Kotok, V.; Bolotin, A. Definition of factors influencing on Ni(OH)2 electrochemical characteristics for supercapacitors. East. Eur. J. Enterp. Technol. 2016, 5, 17–22. [Google Scholar] [CrossRef] [Green Version]
  259. Lang, J.W.; Kong, L.B.; Wu, W.J.; Liu, M.; Luo, Y.C.; Kang, L. A facile approach to the preparation of loose-packed Ni(OH)2 nanoflake materials for electrochemical capacitors. J. Solid State Electrochem. 2009, 13, 333–340. [Google Scholar] [CrossRef]
  260. Lokhande, P.E.; Chavan, U.S. Nanoflower-like Ni(OH)2 synthesis with chemical bath deposition method for high performance electrochemical applications. Mater. Lett. 2018, 218, 225–228. [Google Scholar] [CrossRef]
  261. Li, H.B.; Yu, M.; Wang, F.X.; Liu, P.; Liang, Y.; Xiao, J.; Tong, Y.X.; Yang, G.W.; Wang, C.X. Amorphous nickel hydroxide nanospheres with ultrahigh capacitance and energy density as electrochemical pseudocapacitor materials. Nat. Commun. 2013, 4, 1–7. [Google Scholar] [CrossRef] [Green Version]
  262. Oshitani, M.; Yufu, H.; Takashima, K.; Tsuji, S.; Matsumaru, Y. Development of a Pasted Nickel Electrode with High Active Material Utilization. J. Electrochem. Soc. 1989, 136, 1590–1593. [Google Scholar] [CrossRef]
  263. Park, J.H.; Park, O.O.; Shin, K.H.; Jin, C.S.; Kim, J.H. An Electrochemical Capacitor Based on a Ni(OH)2/Activated Carbon Composite Electrode. Electrochem. Solid State Lett. 2002, 5, H7. [Google Scholar] [CrossRef]
  264. Zang, X.; Sun, C.; Dai, Z.; Yang, J.; Dong, X. Nickel hydroxide nanosheets supported on reduced graphene oxide for high-performance supercapacitors. J. Alloys Compd. 2017, 691, 144–150. [Google Scholar] [CrossRef]
  265. Chai, H.; Peng, X.; Liu, T.; Su, X.; Jia, D.; Zhou, W. High-performance supercapacitors based on conductive graphene combined with Ni(OH)2 nanoflakes. RSC Adv. 2017, 7, 36617–36622. [Google Scholar] [CrossRef] [Green Version]
  266. Li, M.; Xu, S.; Zhu, Y.; Yang, P.; Wang, L.; Chu, P.K. Heterostructured Ni(OH)2-Co(OH)2 composites on 3D ordered Ni-Co nanoparticles fabricated on microchannel plates for advanced miniature supercapacitor. J. Alloys Compd. 2014, 589, 364–371. [Google Scholar] [CrossRef]
  267. Su, L.; Gong, L.; Gao, J. The supercapacitive performances of Co(OH)2/Ni(OH)2 composites in lithium hydroxide solution: Selection of electrolyte and effect of weight ratio. J. Power Sources 2012, 209, 141–146. [Google Scholar] [CrossRef]
  268. Xiao, Z.; Liu, P.; Zhang, J.; Qi, H.; Liu, J.; Li, B.; Sun, X.; Zhang, Q.; Wei, C.; Wang, L. Pillar-Coordinated Strategy to Modulate Phase Transfer of α-Ni(OH)2 for Enhanced Supercapacitor Application. ACS Appl. Energy Mater. 2020, 3, 5628–5636. [Google Scholar] [CrossRef]
  269. Li, J.M.; Chang, K.H.; Hu, C.C. A novel vanadium oxide deposit for the cathode of asymmetric lithium-ion supercapacitors. Electrochem. Commun. 2010, 12, 1800–1803. [Google Scholar] [CrossRef]
  270. Qu, Q.T.; Liu, L.L.; Wu, Y.P.; Holze, R. Electrochemical behavior of V2O5·0.6H2O nanoribbons in neutral aqueous electrolyte solution. Electrochim. Acta 2013, 96, 8–12. [Google Scholar] [CrossRef]
  271. Lee, H.Y.; Goodenough, J.B. Ideal Supercapacitor Behavior of Amorphous V2O5·nH2O in Potassium Chloride (KCl) Aqueous Solution. J. Solid State Chem. 1999, 148, 81–84. [Google Scholar] [CrossRef]
  272. Yu, D.; Chen, C.; Xie, S.; Liu, Y.; Park, K.; Zhou, X.-Y.; Zhang, Q.; Li, J.-Y.; Cao, G. Mesoporous vanadium pentoxide nanofibers with significantly enhanced Li-ion storage properties by electrospinning. Energy Environ. Sci. 2011, 4, 858–861. [Google Scholar] [CrossRef]
  273. Xiong, C.; Aliev, A.E.; Gnade, B.; Balkus, K.J. Fabrication of Silver Vanadium Oxide and V2O5 Nanowires for Electrochromics. ACS Nano 2008, 2, 293–301. [Google Scholar] [CrossRef]
  274. Chen, W.; Zhou, C.; Mai, L.; Liu, Y.; Qi, Y.; Dai, Y. Field Emission from V2O5 n H2O Nanorod Arrays. J. Phys. Chem. C 2008, 112, 2262–2265. [Google Scholar] [CrossRef]
  275. Li, M.; Kong, F.; Wang, H.; Li, G. Synthesis of vanadium pentoxide (V2O5) ultralong nanobelts via an oriented attachment growth mechanism. CrystEngComm 2011, 13, 5317–5320. [Google Scholar] [CrossRef]
  276. Wee, G.; Soh, H.Z.; Cheah, Y.L.; Mhaisalkar, S.G.; Srinivasan, M. Synthesis and electrochemical properties of electrospun V2O5 nanofibers as supercapacitor electrodes. J. Mater. Chem. 2010, 20, 6720–6725. [Google Scholar] [CrossRef]
  277. Zeng, M.; Yin, H.; Yu, K. Synthesis of V2O5 nanostructures with various morphologies and their electrochemical and field-emission properties. Chem. Eng. J. 2012, 188, 64–70. [Google Scholar] [CrossRef]
  278. Saravanakumar, B.; Purushothaman, K.K.; Muralidharan, G. Interconnected V2O5 Nanoporous Network for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2012, 4, 4484–4490. [Google Scholar] [CrossRef]
  279. Shaikh, J.S.; Pawar, R.C.; Mali, S.S.; Moholkar, A.V.; Kim, J.H.; Patil, P.S. Effect of annealing on the supercapacitor performance of CuO-PAA/CNT films. J. Solid State Electrochem. 2012, 16, 25–33. [Google Scholar] [CrossRef]
  280. Balamuralitharan, B.; Cho, I.H.; Bak, J.S.; Kim, H.J. V2O5 nanorod electrode material for enhanced electrochemical properties by a facile hydrothermal method for supercapacitor applications. New J. Chem. 2018, 42, 11862–11868. [Google Scholar] [CrossRef]
  281. Zhai, T.; Liu, H.; Li, H.; Fang, X.; Liao, M.; Li, L.; Zhou, H.; Koide, Y.; Bando, Y.; Golberg, D. Centimeter-long V2O5 nanowires: From synthesis to field-emission, electrochemical, electrical transport, and photoconductive properties. Adv. Mater. 2010, 22, 2547–2552. [Google Scholar] [CrossRef]
  282. Devan, R.S.; Ma, Y.R.; Patil, R.A.; Lukas, S.M. Highly stable supercapacitive performance of one-dimensional (1D) brookite TiO2 nanoneedles. RSC Adv. 2016, 6, 6218–6225. [Google Scholar] [CrossRef] [Green Version]
  283. Cottineau, T.; Toupin, M.; Delahaye, T.; Brousse, T.; Bélanger, D. Nanostructured transition metal oxides for aqueous hybrid electrochemical supercapacitors. Appl. Phys. A Mater. Sci. Process. 2006, 82, 599–606. [Google Scholar] [CrossRef]
  284. Liang, B.; Chen, Y.; He, J.; Chen, C.; Liu, W.; He, Y.; Liu, X.; Zhang, N.; Roy, V.A.L. Controllable Fabrication and Tuned Electrochemical Performance of Potassium Co–Ni Phosphate Microplates as Electrodes in Supercapacitors. ACS Appl. Mater. Interfaces 2018, 10, 3506–3514. [Google Scholar] [CrossRef]
  285. Chen, Z.; Augustyn, V.; Wen, J.; Zhang, Y.; Shen, M.; Dunn, B.; Lu, Y. High-performance supercapacitors based on intertwined CNT/V2O5 nanowire nanocomposites. Adv. Mater. 2011, 23, 791–795. [Google Scholar] [CrossRef] [PubMed]
  286. Foo, C.Y.; Sumboja, A.; Tan, D.J.H.; Wang, J.; Lee, P.S. Flexible and Highly Scalable V2O5-rGO Electrodes in an Organic Electrolyte for Supercapacitor Devices. Adv. Energy Mater. 2014, 4, 1400236. [Google Scholar] [CrossRef]
  287. Qian, T.; Xu, N.; Zhou, J.; Yang, T.; Liu, X.; Shen, X.; Liang, J.; Yan, C. Interconnected three-dimensional V2O5/polypyrrole network nanostructures for high performance solid-state supercapacitors. J. Mater. Chem. A 2015, 3, 488–493. [Google Scholar] [CrossRef]
  288. Jayalakshmi, M.; Rao, M.M.; Venugopal, N.; Kim, K.-B. Hydrothermal synthesis of SnO2–V2O5 mixed oxide and electrochemical screening of carbon nano-tubes (CNT), V2O5, V2O5–CNT, and SnO2–V2O5–CNT electrodes for supercapacitor applications. J. Power Sources 2007, 166, 578–583. [Google Scholar] [CrossRef]
  289. Borenstein, A.; Hanna, O.; Attias, R.; Luski, S.; Brousse, T.; Aurbach, D. Carbon-based composite materials for supercapacitor electrodes: A review. J. Mater. Chem. A 2007, 5, 12653–12672. [Google Scholar] [CrossRef]
  290. Lee, M.; Balasingam, S.K.; Jeong, H.Y.; Hong, W.G.; Lee, H.-B.-R.; Kim, B.H.; Jun, Y. One-step hydrothermal synthesis of Graphene decorated V2O5 nanobelts for enhanced electrochemical energy storage. Sci. Rep. 2015, 5, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Chen, Z.; Qin, Y.; Weng, D.; Xiao, Q.; Peng, Y.; Wang, X.; Li, H.; Wei, F.; Lu, Y. Design and Synthesis of Hierarchical Nanowire Composites for Electrochemical Energy Storage. Adv. Funct. Mater. 2009, 19, 3420–3426. [Google Scholar] [CrossRef]
  292. Lu, S.; Xue, Z.; Chen, Y.; Wan, F.; Zhu, Q.; Zakharova, G.S. Researches on the structure and electrochemical properties of MoxV2-xO5+y nanosheets. Ionics 2017, 23, 2855–2862. [Google Scholar] [CrossRef]
  293. Zhang, Y.; Jing, X.; Wang, Q.; Zheng, J.; Jiang, H.; Meng, C. Three-dimensional porous V2O5 hierarchical spheres as a battery-type electrode for a hybrid supercapacitor with excellent charge storage performance. Dalton Trans. 2017, 46, 15048–15058. [Google Scholar] [CrossRef]
  294. Chen, Y.; Muthukumar, K.; Leban, L.; Li, J. Microwave-assisted high-yield exfoliation of vanadium pentoxide nanoribbons for supercapacitor applications. Electrochim. Acta 2020, 330, 135200. [Google Scholar] [CrossRef]
  295. Chen, D.; Li, J.; Wu, Q. Review of V2O5-based nanomaterials as electrode for supercapacitor. J. Nanoparticle Res. 2019, 21, 9. [Google Scholar] [CrossRef]
  296. Liu, B.; Liu, Y.; Chen, H.; Yang, M.; Li, H. MnO2 Nanostructures Deposited on Graphene-Like Porous Carbon Nanosheets for High-Rate Performance and High-Energy Density Asymmetric Supercapacitors. ACS Sustain. Chem. Eng. 2019, 7, 3101–3110. [Google Scholar] [CrossRef]
  297. Feng, X.; Huang, Y.; Li, C.; Xiao, Y.; Chen, X.; Gao, X.; Chen, C. Construction of carnations-like Mn3O4@NiCo2O4@NiO hierarchical nanostructures for high-performance supercapacitors. Electrochim. Acta 2019, 308, 142–149. [Google Scholar] [CrossRef]
  298. Sadak, O.; Wang, W.; Guan, J.; Sundramoorthy, A.K.; Gunasekaran, S. MnO2 Nanoflowers Deposited on Graphene Paper as Electrode Materials for Supercapacitors. ACS Appl. Nano Mater. 2019, 2, 4386–4394. [Google Scholar] [CrossRef]
  299. Qi, W.; Li, X.; Wu, Y.; Zeng, H.; Kuang, C.; Zhou, S.; Huang, S.; Yang, Z. Flexible electrodes of MnO2/CNTs composite for enhanced performance on supercapacitors. Surf. Coat. Technol. 2017, 320, 624–629. [Google Scholar] [CrossRef]
  300. Jia, H.; Cai, Y.; Zheng, X.; Lin, J.; Liang, H.; Qi, J.; Cao, J.; Feng, J.; Fei, W. Mesostructured Carbon Nanotube-on-MnO2 Nanosheet Composite for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2018, 10, 38963–38969. [Google Scholar] [CrossRef]
  301. Jeong, J.H.; Park, J.W.; Lee, D.W.; Baughman, R.H.; Kim, S.J. Electrodeposition of α-MnO2/γ-MnO2 on Carbon Nanotube for Yarn Supercapacitor. Sci. Rep. 2019, 9, 1–8. [Google Scholar] [CrossRef]
  302. Li, L.; Hu, Z.A.; An, N.; Yang, Y.Y.; Li, Z.M.; Wu, H.Y. Facile synthesis of MnO2/CNTs composite for supercapacitor electrodes with long cycle stability. J. Phys. Chem. C 2014, 118, 22865–22872. [Google Scholar] [CrossRef]
  303. Ko, W.Y.; Liu, Y.C.; Lai, J.Y.; Chung, C.C.; Lin, K.J. Vertically Standing MnO2 Nanowalls Grown on AgCNT-Modified Carbon Fibers for High-Performance Supercapacitors. ACS Sustain. Chem. Eng. 2019, 7, 669–678. [Google Scholar] [CrossRef]
  304. Sun, X.; Xu, T.; Bai, J.; Li, C. MnO2 Nanosheets Grown on Multichannel Carbon Nanofibers Containing Amorphous Cobalt Oxide as a Flexible Electrode for Supercapacitors. ACS Appl. Energy Mater. 2019, 2, 8675–8684. [Google Scholar] [CrossRef]
  305. Kumar, K.S.; Cherusseri, J.; Thomas, J. Two-Dimensional Mn3O4 Nanowalls Grown on Carbon Fibers as Electrodes for Flexible Supercapacitors. ACS Omega 2019, 4, 4472–4480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. He, Q.Q.; Wang, H.-Y.; Lun, N.; Qi, Y.-X.; Liu, J.; Feng, J.; Qiu, J.; Bai, Y.-J. Fabricating a Mn3O4/Ni(OH)2 Nanocomposite by Water-Boiling Treatment for Use in Asymmetric Supercapacitors as an Electrode Material. ACS Sustain. Chem. Eng. 2018, 6, 15688–15696. [Google Scholar] [CrossRef]
  307. Wang, T.; Le, Q.; Guo, X.; Huang, M.; Liu, X.; Dong, F.; Zhang, J.; Zhang, Y. Preparation of Porous Graphene@Mn3O4 and Its Application in the Oxygen Reduction Reaction and Supercapacitor. ACS Sustain. Chem. Eng. 2019, 7, 831–837. [Google Scholar] [CrossRef]
  308. Samuel, E.; Kim, T.G.; Park, C.W.; Joshi, B.; Swihart, M.T.; Yoon, S.S. Supersonically Sprayed Zn2SnO4/SnO2/CNT Nanocomposites for High-Performance Supercapacitor Electrodes. ACS Sustain. Chem. Eng. 2019, 7, 14031–14040. [Google Scholar] [CrossRef]
  309. Chakraborty, I.; Chakrabarty, N.; Senapati, A.; Chakraborty, A.K. CuO@NiO/Polyaniline/MWCNT Nanocomposite as High-Performance Electrode for Supercapacitor. J. Phys. Chem. C 2018, 122, 27180–27190. [Google Scholar] [CrossRef]
  310. Das, D.; Borthakur, L.J.; Nath, B.C.; Saikia, B.J.; Mohan, K.J.; Dolui, S.K. Designing hierarchical NiO/PAni-MWCNT core-shell nanocomposites for high performance super capacitor electrodes. RSC Adv. 2016, 6, 44878–44887. [Google Scholar] [CrossRef]
  311. Samuel, E.; Joshi, B.; Kim, Y.I.; Aldalbahi, A.; Rahaman, M.; Yoon, S.S. ZnO/MnOx Nanoflowers for High-Performance Supercapacitor Electrodes. ACS Sustain. Chem. Eng. 2020, 8, 3697–3708. [Google Scholar] [CrossRef]
  312. Jeon, S.; Jeong, J.H.; Yoo, H.; Yu, H.K.; Kim, B.H.; Kim, M.H. RuO2 Nanorods on Electrospun Carbon Nanofibers for Supercapacitors. ACS Appl. Nano Mater. 2020, 3, 3847–3858. [Google Scholar] [CrossRef]
  313. Ryu, I.; Yang, M.; Kwon, H.; Park, H.K.; Do, Y.R.; Lee, S.-B.; Yim, S. Coaxial RuO2-ITO nanopillars for transparent supercapacitor application. Langmuir 2014, 30, 1704–1709. [Google Scholar] [CrossRef]
  314. Xu, J.; Wang, Q.; Wang, X.; Xiang, Q.; Liang, B.; Chen, D.; Shen, G. Flexible Asymmetric Supercapacitors Based upon Co9S8 Nanorod//Co3O4 @RuO2 Nanosheet Arrays on Carbon Cloth. ACS Nano 2013, 7, 5453–5462. [Google Scholar] [CrossRef]
  315. Kaipannan, S.; Govindarajan, K.; Sundaramoorthy, S.; Marappan, S. Waste Toner-Derived Carbon/Fe3O4 Nanocomposite for High-Performance Supercapacitor. ACS Omega 2019, 4, 15798–15805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Tian, Y.; Hu, X.; Wang, Y.; Li, C.; Wu, X. Fe2O3 Nanoparticles Decorated on Graphene-Carbon Nanotubes Conductive Networks for Boosting the Energy Density of All-Solid-State Asymmetric Supercapacitor. ACS Sustain. Chem. Eng. 2019, 7, 9211–9219. [Google Scholar] [CrossRef]
  317. Gupta, A.; Sardana, S.; Dalal, J.; Lather, S.; Maan, A.S.; Tripathi, R.; Punia, R.; Singh, K.; Ohlan, A. Nanostructured Polyaniline/Graphene/Fe2O3 Composites Hydrogel as a High-Performance Flexible Supercapacitor Electrode Material. ACS Appl. Energy Mater. 2020, 3, 6434–6446. [Google Scholar] [CrossRef]
  318. Kumar, A.; Sarkar, D.; Mukherjee, S.; Patil, S.; Sarma, D.D.; Shukla, A. Realizing an Asymmetric Supercapacitor Employing Carbon Nanotubes Anchored to Mn3O4 Cathode and Fe3O4 Anode. ACS Appl. Mater. Interfaces 2018, 10, 42484–42493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Li, L.; Gao, J.; Cecen, V.; Fan, J.; Shi, P.; Xu, Q.; Min, Y. Hierarchical WS2@NiCo2O4 Core-shell Heterostructure Arrays Supported on Carbon Cloth as High-Performance Electrodes for Symmetric Flexible Supercapacitors. ACS Omega 2020, 5, 4657–4667. [Google Scholar] [CrossRef] [Green Version]
  320. Zhang, A.; Yue, L.; Jia, D.; Cui, L.; Wei, D.; Huang, W.; Liu, R.; Liu, Y.; Yang, W.; Liu, J. Cobalt/Nickel Ions-Assisted Synthesis of Laminated CuO Nanospheres Based on Cu(OH)2 Nanorod Arrays for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2020, 12, 2591–2600. [Google Scholar] [CrossRef]
  321. Kwak, C.S.; Ko, T.H.; Lee, J.H.; Kim, H.Y.; Kim, B.S. Flexible Transparent Symmetric Solid-State Supercapacitors Based on NiO-Decorated Nanofiber-Based Composite Electrodes with Excellent Mechanical Flexibility and Cyclability. ACS Appl. Energy Mater. 2020, 3, 2394–2403. [Google Scholar] [CrossRef]
  322. Xia, T.; Zhang, X.; Zhao, J.; Li, Q.; Ao, C.; Hu, R.; Zheng, Z.; Zhang, W.; Lu, C.; Deng, Y. Flexible and Conductive Carbonized Cotton Fabrics Coupled with a Nanostructured Ni(OH)2 Coating for High Performance Aqueous Symmetric Supercapacitors. ACS Sustain. Chem. Eng. 2019, 7, 5231–5239. [Google Scholar] [CrossRef]
  323. Ray, A.; Roy, A.; Saha, S.; Ghosh, M.; Chowdhury, S.R.; Maiyalagan, T.; Bhattacharya, S.K.; Das, S. Electrochemical Energy Storage Properties of Ni-Mn-Oxide Electrodes for Advance Asymmetric Supercapacitor Application. Langmuir 2019, 35, 8257–8267. [Google Scholar] [CrossRef] [Green Version]
  324. Deng, M.J.; Yeh, L.H.; Lin, Y.H.; Chen, J.M.; Chou, T.H. 3D Network V2O5 Electrodes in a Gel Electrolyte for High-Voltage Wearable Symmetric Pseudocapacitors. ACS Appl. Mater. Interfaces 2019, 11, 29838–29848. [Google Scholar] [CrossRef] [PubMed]
  325. Chen, H.C.; Lin, Y.C.; Chen, Y.L.; Chen, C.J. Facile fabrication of three-dimensional hierarchical nanoarchitectures of VO2/Graphene@NiS2 hybrid aerogel for high-performance all-solid-state asymmetric supercapacitors with ultrahigh energy density. ACS Appl. Energy Mater. 2019, 2, 459–467. [Google Scholar] [CrossRef]
  326. Bi, W.; Wu, Y.; Liu, C.; Wang, J.; Du, Y.; Gao, Z.; Wu, G.; Cao, G. Gradient oxygen vacancies in V2O5/PEDOT nanocables for high-performance supercapacitors. ACS Appl. Energy Mater. 2019, 2, 668–677. [Google Scholar] [CrossRef]
  327. Wang, D.W.; Li, F.; Zhao, J.; Ren, W.; Chen, Z.; Tan, J.; Wu, Z.-S.; Gentle, I.; Lu, G.; Cheng, H.-M. Fabrication of graphene/polyaniline composite paper via in situ anodic electropolymerization for high-performance flexible electrode. ACS Nano 2009, 3, 1745–1752. [Google Scholar] [CrossRef] [PubMed]
  328. Kalaji, M.; Murphy, P.J.; Williams, G.O. Study of conducting polymers for use as redox supercapacitors. Synth. Met. 1999, 102, 1360–1361. [Google Scholar] [CrossRef]
  329. Gupta, V.; Miura, N. High performance electrochemical supercapacitor from electrochemically synthesized nanostructured polyaniline. Mater. Lett. 2006, 60, 1466–1469. [Google Scholar] [CrossRef]
  330. Ryu, K.S.; Wu, X.; Lee, Y.-G.; Chang, S.H. Electrochemical capacitor composed of doped polyaniline and polymer electrolyte membrane. J. Appl. Polym. Sci. 2003, 89, 1300–1304. [Google Scholar] [CrossRef]
  331. Ammonia, L. The Acid Properties of Ammonium Saltsin. Aust. J. Chem. 1924, 822, 5–10. [Google Scholar]
  332. Shirakawa, H.; Louis, E.J.; MacDiarmid, A.G.; Chiang, C.K.; Heeger, A.J. Synthesis of electrically conducting organic polymers: Halogen derivatives of polyacetylene, (CH)x. J. Chem. Soc. Chem. Commun. 1977, 16, 578–580. [Google Scholar] [CrossRef]
  333. Ramya, R.; Sivasubramanian, R.; Sangaranarayanan, M.V. Conducting polymers-based electrochemical supercapacitors—Progress and prospects. Electrochim. Acta 2013, 101, 109–129. [Google Scholar] [CrossRef]
  334. Ellenbogen, J.C. Supercapacitors: A Brief Overview. 2006. Available online: https://www.mitre.org/sites/default/files/pdf/06_0667.pdf (accessed on 4 June 2020).
  335. Snook, G.A.; Kao, P.; Best, A.S. Conducting-polymer-based supercapacitor devices and electrodes. J. Power Sources 2011, 196. [Google Scholar] [CrossRef]
  336. Sharma, P.; Bhatti, T.S. A review on electrochemical double-layer capacitors. Energy Convers. Manag. 2010, 51, 2901–2912. [Google Scholar] [CrossRef]
  337. Cheng, Q.; Tang, J.; Ma, J.; Zhang, H.; Shinya, N.; Qin, L.C. Graphene and nanostructured MnO2 composite electrodes for supercapacitors. Carbon N. Y. 2011, 49, 2917–2925. [Google Scholar] [CrossRef]
  338. Sharma, R.K.; Rastogi, A.C.; Desu, S.B. Manganese oxide embedded polypyrrole nanocomposites for electrochemical supercapacitor. Electrochim. Acta 2008, 53, 7690–7695. [Google Scholar] [CrossRef]
  339. Rudge, A.; Davey, J.; Raistrick, I.; Gottesfeld, S.; Ferraris, J.P. Conducting polymers as active materials in electrochemical capacitors. J. Power Sources 1994, 47, 89–107. [Google Scholar] [CrossRef]
  340. Giannelis, E.P. Polymer Layered Silicate Nanocomposites. Adv. Mater. 1996, 8, 29–35. [Google Scholar] [CrossRef]
  341. Sivaraman, P.; Thakur, A.; Kushwaha, R.K.; Ratna, D.; Samui, A.B. Poly(3-methyl thiophene)-activated carbon hybrid supercapacitor based on gel polymer electrolyte. Electrochem. Solid State Lett. 2006, 9, 9. [Google Scholar] [CrossRef]
  342. Wang, K.; Huang, J.; Wei, Z. Conducting Polyaniline Nanowire Arrays for High Performance Supercapacitors. J. Phys. Chem. C 2010, 114, 8062–8067. [Google Scholar] [CrossRef]
  343. Cameron, C.G. Electrochemical capacitors. Springer Handb. 2017, 50, 563–589. [Google Scholar]
  344. Hashmi, S.A.; Upadhyaya, H.M. Polypyrrole and poly(3-methyl thiophene)-based solid state redox supercapacitors using ion conducting polymer electrolyte. Solid State Ion. 2002, 152–153, 883–889. [Google Scholar] [CrossRef]
  345. Shi, Y.; Pan, L.; Liu, B.; Wang, Y.; Cui, Y.; Bao, Z.; Yu, G. Nanostructured conductive polypyrrole hydrogels as high-performance, flexible supercapacitor electrodes. J. Mater. Chem. A 2014, 2, 6086–6091. [Google Scholar] [CrossRef]
  346. Ryu, K.S.; Kim, K.M.; Park, Y.J.; Park, N.G.; Kang, M.G.; Chang, S.H. Redox supercapacitor using polyaniline doped with Li salt as electrode. Solid State Ion. 2002, 152–153, 861–866. [Google Scholar] [CrossRef]
  347. Naoi, K.; Suematsu, S.; Manago, A. Electrochemistry of poly(1,5-diaminoanthraquinone) and its application in electrochemical capacitor materials. J. Electrochem. Soc. 2000, 147, 420–426. [Google Scholar] [CrossRef]
  348. Meng, Q.; Cai, K.; Chen, Y.; Chen, L. Research progress on conducting polymer based supercapacitor electrode materials. Nano Energy 2017, 36, 268–285. [Google Scholar] [CrossRef]
  349. Hao, G.P.; Hippauf, F.; Oschatz, M.; Wisser, F.M.; Leifert, A.; Nickel, W.; Mohamed-Noriega, N.; Zheng, Z.; Kaskel, S. Stretchable and semitransparent conductive hybrid hydrogels for flexible supercapacitors. ACS Nano 2014, 8, 7138–7146. [Google Scholar] [CrossRef] [PubMed]
  350. Li, W.; Gao, F.; Wang, X.; Zhang, N.; Ma, M. Strong and Robust Polyaniline-Based Supramolecular Hydrogels for Flexible Supercapacitors. Angew. Chem. Int. Ed. 2016, 55, 9196–9201. [Google Scholar] [CrossRef] [PubMed]
  351. Wang, Y.; Shi, Y.; Pan, L.; Ding, Y.; Zhao, Y.; Li, Y.; Shi, Y.; Yu, G. Dopant-Enabled Supramolecular Approach for Controlled Synthesis of Nanostructured Conductive Polymer Hydrogels. Nano Lett. 2015, 15, 7736–7741. [Google Scholar] [CrossRef] [PubMed]
  352. Yang, Z.; Shi, D.; Dong, W.; Chen, M. Self-Standing Hydrogels Composed of Conducting Polymers for All-Hydrogel-State Supercapacitors. Chem. A Eur. J. 2020, 26, 1846–1855. [Google Scholar] [CrossRef]
  353. Liu, J.; Huang, J.; Cai, Q.; Yang, Y.; Luo, W.; Zeng, B.; Xu, Y.; Yuan, C.; Dai, L. Design of Slidable Polymer Networks: A Rational Strategy to Stretchable, Rapid Self-Healing Hydrogel Electrolytes for Flexible Supercapacitors. ACS Appl. Mater. Interfaces 2020, 12, 20479–20489. [Google Scholar] [CrossRef]
  354. Cheng, Q.; Tang, J.; Shinya, N.; Qin, L.C. Polyaniline modified graphene and carbon nanotube composite electrode for asymmetric supercapacitors of high energy density. J. Power Sources 2013, 241, 423–428. [Google Scholar] [CrossRef]
  355. Han, G.; Liu, Y.; Zhang, L.; Kan, E.; Zhang, S.; Tang, J.; Tang, W. MnO2 Nanorods Intercalating Graphene Oxide/Polyaniline Ternary Composites for Robust High-Performance Supercapacitors. Sci. Rep. 2015, 4, 4824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  356. Li, J.; Xie, H.; Li, Y.; Liu, J.; Li, Z. Electrochemical properties of graphene nanosheets/polyaniline nanofibers composites as electrode for supercapacitors. J. Power Sources 2011, 196, 10775–10781. [Google Scholar] [CrossRef]
  357. Hu, C.C.; Chang, K.H.; Lin, M.C.; Wu, Y.T. Design and tailoring of the nanotubular arrayed architecture of hydrous RuO2 for next generation supercapacitors. Nano Lett. 2006, 6, 2690–2695. [Google Scholar] [CrossRef] [PubMed]
  358. Zhou, C.; Zhang, Y.; Li, Y.; Liu, J. Construction of High-Capacitance 3D CoO@Polypyrrole Nanowire Array Electrode for Aqueous Asymmetric Supercapacitor. Nano Lett. 2013, 13, 2078–2085. [Google Scholar] [CrossRef] [PubMed]
  359. Bober, P.; Gavrilov, N.; Kovalcik, A.; Micusik, M.; Unterweger, C.; Pašti, I.A.; Šeděnková, I.; Acharya, U.; Pfleger, J.; Filippov, S.K.; et al. Electrochemical properties of lignin/polypyrrole composites and their carbonized analogues. Mater. Chem. Phys. 2018, 213, 352–361. [Google Scholar] [CrossRef]
  360. Mastragostino, M.; Arbizzani, C.; Soavi, F. Conducting polymers as electrode materials in supercapacitors. Solid State Ion. 2002, 148, 493–498. [Google Scholar] [CrossRef]
  361. Shao, Q.; Tang, J.; Lin, Y.; Li, J.; Qin, F.; Yuan, J.; Qin, L.-C. Carbon nanotube spaced graphene aerogels with enhanced capacitance in aqueous and ionic liquid electrolytes. J. Power Sources 2015, 278, 751–759. [Google Scholar] [CrossRef] [Green Version]
  362. Poonam; Sharma, K.; Arora, A.; Tripathi, S.K. Review of supercapacitors: Materials and devices. J. Energy Storage 2019, 21, 801–825. [Google Scholar] [CrossRef]
  363. Liu, F.; Song, S.; Xue, D.; Zhang, H. Selective crystallization with preferred lithium-ion storage capability of inorganic materials. Nanoscale Res. Lett. 2012, 7, 1–17. [Google Scholar] [CrossRef] [Green Version]
  364. Wang, L.; Feng, X.; Ren, L.; Piao, Q.; Zhong, J.; Wang, Y.; Li, H.; Chen, Y.; Wang, B. Flexible solid-state supercapacitor based on a metal-organic framework interwoven by electrochemically-deposited PANI. J. Am. Chem. Soc. 2015, 137, 4920–4923. [Google Scholar] [CrossRef]
  365. Shan, Y.; Gao, L. Formation and characterization of multi-walled carbon nanotubes/Co3O4 nanocomposites for supercapacitors. Mater. Chem. Phys. 2007, 103, 206–210. [Google Scholar] [CrossRef]
  366. Xiang, C.; Li, M.; Zhi, M.; Manivannan, A.; Wu, N. A reduced graphene oxide/Co3O4 composite for supercapacitor electrode. J. Power Sources 2013, 226, 65–70. [Google Scholar] [CrossRef] [Green Version]
  367. He, G.; Li, J.; He, G.; Shi, J.; Sun, X.-Q.; Chen, S.; Wang, X. Hydrothermal preparation of Co3O4@graphene nanocomposite for supercapacitor with enhanced capacitive performance. Mater. Lett. 2012, 82, 61–63. [Google Scholar] [CrossRef]
  368. Abouali, S.; Garakani, M.A.; Zhang, B.; Xu, Z.-L.; Heidari, E.K.; Huang, J.-Q.; Huang, J.; Kim, J.-K. Electrospun Carbon Nanofibers with in Situ Encapsulated Co3O4 Nanoparticles as Electrodes for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 13503–13511. [Google Scholar] [CrossRef]
  369. Wang, H.; Yi, H.; Chen, X.; Wang, X. Asymmetric supercapacitors based on nano-architectured nickel oxide/graphene foam and hierarchical porous nitrogen-doped carbon nanotubes with ultrahigh-rate performance. J. Mater. Chem. A 2014, 2, 3223–3230. [Google Scholar] [CrossRef]
  370. Du, W.; Wang, Z.; Zhu, Z.; Hu, S.; Zhu, X.; Shi, Y.-F.; Panga, H.; Qian, X. Facile synthesis and superior electrochemical performances of CoNi2S4/graphene nanocomposite suitable for supercapacitor electrodes. J. Mater. Chem. A 2014, 2, 9613–9619. [Google Scholar] [CrossRef]
  371. Perera, S.D.; Patel, B.B.; Nijem, N.; Roodenko, K.; Seitz, O.; Ferraris, J.P.; Chabal, Y.J.; Balkus, K.J., Jr. Vanadium Oxide Nanowire-Carbon Nanotube Binder-Free Flexible Electrodes for Supercapacitors. Adv. Energy Mater. 2011, 1, 936–945. [Google Scholar] [CrossRef]
  372. Yan, S.; Wang, H.; Qu, P.; Zhang, Y.; Xiao, Z. RuO2/carbon nanotubes composites synthesized by microwave-assisted method for electrochemical supercapacitor. Synth. Met. 2009, 159, 158–161. [Google Scholar] [CrossRef]
  373. Zhang, C.; Xie, Y.; Zhao, M.; Pentecost, A.E.; Ling, Z.; Wang, J.; Long, D.; Ling, L.; Qiao, W. Enhanced electrochemical performance of hydrous RuO2/mesoporous carbon nanocomposites via nitrogen doping. ACS Appl. Mater. Interfaces 2014, 6, 9751–9759. [Google Scholar] [CrossRef]
  374. Lee, Y.; Kim, B.; Jung, H.J.; Shim, J.H.; Lee, Y.; Lee, C.; Baik, J.M.; Kim, W.; Kim, M.H. Hierarchically grown single crystalline RuO2 nanorods on vertically aligned few-walled carbon nanotubes. Mater. Lett. 2012, 89, 115–117. [Google Scholar] [CrossRef]
  375. Lee, J.K.; Pathan, H.M.; Jung, K.D.; Joo, O.S. Electrochemical capacitance of nanocomposite films formed by loading carbon nanotubes with ruthenium oxide. J. Power Sources 2006, 159, 1527–1531. [Google Scholar] [CrossRef]
  376. Zhi, M.; Xiang, C.; Li, J.; Li, M.; Wu, N. Nanostructured carbon-metal oxide composite electrodes for supercapacitors: A review. Nanoscale 2013, 5, 72–88. [Google Scholar] [CrossRef] [PubMed]
  377. Ghosh, A.; Lee, Y.H. Carbon-Based Electrochemical Capacitors. ChemSusChem 2012, 5, 480–499. [Google Scholar] [CrossRef] [PubMed]
  378. Wang, P.; Xu, Y.; Liu, H.; Chen, Y.; Yang, J.; Tan, Q. Carbon/carbon nanotube-supported RuO2 nanoparticles with a hollow interior as excellent electrode materials for supercapacitors. Nano Energy 2015, 15, 116–124. [Google Scholar] [CrossRef]
  379. Guan, Q.; Cheng, J.; Wang, B.; Ni, W.; Gu, G.; Li, X.; Huang, L.; Yang, G.; Nie, F. Needle-like Co3O4 anchored on the graphene with enhanced electrochemical performance for aqueous supercapacitors. ACS Appl. Mater. Interfaces 2014, 6, 7626–7632. [Google Scholar] [CrossRef] [PubMed]
  380. Yuan, C.; Li, J.; Hou, L.; Yang, L.; Shen, L.; Zhang, X. Facile template-free synthesis of ultralayered mesoporous nickel cobaltite nanowires towards high-performance electrochemical capacitors. J. Mater. Chem. 2012, 22, 16084–16090. [Google Scholar] [CrossRef]
  381. Zhang, G.Q.; Wu, H.B.; Hoster, H.E.; Chan-Park, M.B.; Lou, X.W. Single-crystalline NiCo2O4 nanoneedle arrays grown on conductive substrates as binder-free electrodes for high-performance supercapacitors. Energy Environ. Sci. 2012, 5, 9453–9456. [Google Scholar] [CrossRef]
  382. Zhang, G.; Lou, X.W.D. General Solution Growth of Mesoporous NiCo2O4 Nanosheets on Various Conductive Substrates as High-Performance Electrodes for Supercapacitors. Adv. Mater. 2013, 25, 976–979. [Google Scholar] [CrossRef]
  383. Umeshbabu, E.; Rajeshkhanna, G.; Justin, P.; Rao, G.R. Synthesis of mesoporous NiCo2O4-rGO by a solvothermal method for charge storage applications. RSC Adv. 2015, 5, 66657–66666. [Google Scholar] [CrossRef]
  384. Wu, H.B.; Pang, H.; Lou, X.W. Facile synthesis of mesoporous Ni0.3Co2.7O4 hierarchical structures for high-performance supercapacitors. Energy Environ. Sci. 2013, 6, 3619–3626. [Google Scholar] [CrossRef]
  385. Wang, X.; Han, X.; Lim, M.; Singh, N.; Gan, C.L.; Jan, M.; Lee, P.S. Nickel Cobalt Oxide-Single Wall Carbon Nanotube Composite Material for Superior Cycling Stability and High-Performance Supercapacitor Application. J. Phys. Chem. C 2012, 116, 12448–12454. [Google Scholar] [CrossRef]
  386. Wan, H.; Jiang, J.; Yu, J.; Xu, K.; Miao, L.; Zhang, L.; Chen, H.C.; Ruan, Y. NiCo2S4 porous nanotubes synthesis via sacrificial templates: High-performance electrode materials of supercapacitors. CrystEngComm 2013, 15, 7649–7651. [Google Scholar] [CrossRef]
  387. Xiong, G.; He, P.; Liu, L.; Chen, T.; Fisher, T.S. Plasma-grown graphene petals templating Ni-Co-Mn hydroxide nanoneedles for high-rate and long-cycle-life pseudocapacitive electrodes. J. Mater. Chem. A 2015, 3, 22940–22948. [Google Scholar] [CrossRef]
  388. Hareesh, K.; Shateesh, B.; Joshi, R.P.; Williams, J.F.; Phase, D.M.; Haram, S.K.; Dhole, S.D. Ultra high stable supercapacitance performance of conducting polymer coated MnO2 nanorods/rGO nanocomposites. RSC Adv. 2017, 7, 20027–20036. [Google Scholar] [CrossRef] [Green Version]
  389. Peng, W.; Chen, H.; Wang, W.; Huang, Y.; Han, G. Synthesis of NiCo2S4 nanospheres/reduced graphene oxide composite as electrode material for supercapacitor. Curr. Appl. Phys. 2020, 20, 304–309. [Google Scholar] [CrossRef]
  390. Soni, R.; Kurungot, S. Fe2P4O12-carbon composite as a highly stable electrode material for electrochemical capacitors. New J. Chem. 2019, 43, 399–406. [Google Scholar] [CrossRef]
  391. Shi, K.; Zhitomirsky, I. Electrophoretic nanotechnology of graphene-carbon nanotube and graphene-polypyrrole nanofiber composites for electrochemical supercapacitors. J. Colloid Interface Sci. 2013, 407, 474–481. [Google Scholar] [CrossRef]
  392. Shi, K.; Yang, X.; Cranston, E.D.; Zhitomirsky, I. Efficient Lightweight Supercapacitor with Compression Stability. Adv. Funct. Mater. 2016, 26, 6437–6445. [Google Scholar] [CrossRef]
  393. Liu, Y.; Shi, K.; Zhitomirsky, I. New colloidal route for electrostatic assembly of oxide nanoparticle—Carbon nanotube composites. Colloids Surfaces A Physicochem. Eng. Asp. 2014, 446, 15–22. [Google Scholar] [CrossRef]
  394. Liu, Y.; Ata, M.S.; Shi, K.; Zhu, G.Z.; Botton, G.A.; Zhitomirsky, I. Surface modification and cathodic electrophoretic deposition of ceramic materials and composites using celestine blue dye. RSC Adv. 2014, 4, 29652–29659. [Google Scholar] [CrossRef]
  395. Shi, K.; Giapis, K.P. Scalable Fabrication of Supercapacitors by Nozzle-Free Electrospinning. ACS Appl. Energy Mater. 2018, 1, 296–300. [Google Scholar] [CrossRef]
  396. Antiohos, D.; Folkes, G.; Sherrell, P.C.; Ashraf, S.; Wallace, G.G.; Aitchison, P.; Harris, A.T.; Chen, J.; Minett, A.I. Compositional effects of PEDOT-PSS/single walled carbon nanotube films on supercapacitor device performance. J. Mater. Chem. 2011, 21, 15987–15994. [Google Scholar] [CrossRef]
  397. Frackowiak, E.; Khomenko, V.; Jurewicz, K.; Lota, K.; Béguin, F. Supercapacitors based on conducting polymers/nanotubes composites. J. Power Sources 2006, 153, 413–418. [Google Scholar] [CrossRef]
  398. Liu, P.; Yan, J.; Guang, Z.; Huang, Y.; Li, X.; Huang, W. Recent advancements of polyaniline-based nanocomposites for supercapacitors. J. Power Sources 2019, 424, 108–130. [Google Scholar] [CrossRef]
  399. Yang, Y.; Xi, Y.; Li, J.; Wei, G.; Klyui, N.I.; Han, W. Flexible Supercapacitors Based on Polyaniline Arrays Coated Graphene Aerogel Electrodes. Nanoscale Res. Lett. 2017, 12, 1931–7573. [Google Scholar] [CrossRef]
  400. Yu, P.; Li, Y.; Yu, X.; Zhao, X.; Wu, L.; Zhang, Q. Polyaniline nanowire arrays aligned on nitrogen-doped carbon fabric for high-performance flexible supercapacitors. Langmuir 2013, 29, 12051–12058. [Google Scholar] [CrossRef]
  401. Mao, L.; Zhang, K.; Chan, H.S.O.; Wu, J. Surfactant-stabilized graphene/polyaniline nanofiber composites for high performance supercapacitor electrode. J. Mater. Chem. 2012, 22, 80–85. [Google Scholar] [CrossRef]
  402. Wang, R.; Han, M.; Zhao, Q.; Ren, Z.; Guo, X.; Xu, C.; Hu, N.; Lu, L. Hydrothermal synthesis of nanostructured graphene/polyaniline composites as high-capacitance electrode materials for supercapacitors. Sci. Rep. 2017, 7, 44562. [Google Scholar] [CrossRef] [Green Version]
  403. Zeng, S.; Chen, H.; Cai, F.; Kang, Y.; Chen, M.; Li, Q. Electrochemical fabrication of carbon nanotube/polyaniline hydrogel film for all-solid-state flexible supercapacitor with high areal capacitance. J. Mater. Chem. A 2015, 3, 23864–23870. [Google Scholar] [CrossRef]
  404. Li, Y.; Zhao, X.; Yu, P.; Zhang, Q. Oriented Arrays of Polyaniline Nanorods Grown on Graphite Nanosheets for an Electrochemical Supercapacitor. Langmuir 2013, 29, 493–500. [Google Scholar] [CrossRef]
  405. Han, Y.; Ding, B.; Tong, H.; Zhang, X. Capacitance properties of graphite oxide/poly(3,4-ethylene dioxythiophene) composites. J. Appl. Polym. Sci. 2011, 121, 892–898. [Google Scholar] [CrossRef]
  406. Keskinen, J.; Tuurala, S.; Sjödin, M.; Kiri, K.; Nyholm, L.; Flyktman, T.; Strømme, M.; Smolander, M. Asymmetric and symmetric supercapacitors based on polypyrrole and activated carbon electrodes. Synth. Met. 2015, 203, 192–199. [Google Scholar] [CrossRef]
  407. Yan, J.; Wei, T.; Fan, Z.; Qian, W.; Zhang, M.; Shen, X.; Wei, F. Preparation of graphene nanosheet/carbon nanotube/polyaniline composite as electrode material for supercapacitors. J. Power Sources 2010, 195, 3041–3045. [Google Scholar] [CrossRef]
  408. Jaidev; Ramaprabhu, S. Poly(p-phenylenediamine)/graphene nanocomposites for supercapacitor applications. J. Mater. Chem. 2012, 22, 18775–18783. [Google Scholar] [CrossRef]
  409. Zhao, X.; Zhang, Q.; Hao, Y.; Li, Y.; Fang, Y.; Chen, D. Alternate Multilayer Films of Poly(vinyl alcohol) and Exfoliated Graphene Oxide Fabricated via a Facial Layer-by-Layer Assembly. Macromolecules 2010, 43, 9411–9416. [Google Scholar] [CrossRef]
  410. Nam, M.S.; Patil, U.; Park, B.; Sim, H.B.; Jun, S.C. A binder free synthesis of 1D PANI and 2D MoS2 nanostructured hybrid composite electrodes by the electrophoretic deposition (EPD) method for supercapacitor application. RSC Adv. 2016, 6, 101592–101601. [Google Scholar] [CrossRef]
  411. Xie, K.; Wei, B. Materials and structures for stretchable energy storage and conversion devices. Adv. Mater. 2014, 26, 3592–3617. [Google Scholar] [CrossRef]
  412. Liu, Y.; Zhang, B.; Yang, Y.; Chang, Z.; Wen, Z.; Wu, Y. Polypyrrole-coated α-MoO3 nanobelts with good electrochemical performance as anode materials for aqueous supercapacitors. J. Mater. Chem. A 2013, 1, 13582–13587. [Google Scholar] [CrossRef]
  413. Zhang, X.; Ji, L.; Zhang, S.; Yang, W. Synthesis of a novel polyaniline-intercalated layered manganese oxide nanocomposite as electrode material for electrochemical capacitor. J. Power Sources 2007, 173, 1017–1023. [Google Scholar] [CrossRef]
  414. Raj, R.P.; Ragupathy, P.; Mohan, S. Remarkable capacitive behavior of a Co3O4-polyindole composite as electrode material for supercapacitor applications. J. Mater. Chem. A 2015, 3, 24338–24348. [Google Scholar] [CrossRef]
  415. Balbuena, J.O.; de Ureta, P.T.; Ruiz, E.R.; Pita, S.M. Electric Field Effect in Atomically Thin Carbon Films. Med. Clin. 2004, 146, 93–94. [Google Scholar] [CrossRef]
  416. Xiao, X.; Yu, H.; Jin, H.; Wu, M.; Fang, Y.; Sun, J.; Hu, Z.; Li, T.; Wu, J.; Huang, L.; et al. Salt-Templated Synthesis of 2D Metallic MoN and Other Nitrides. ACS Nano 2017, 11, 2180–2186. [Google Scholar] [CrossRef] [PubMed]
  417. Luo, S.; Zhao, J.; Zou, J.; He, Z.; Xu, C.; Liu, F.; Huang, Y.; Dong, L.; Wang, L.; Zhang, H. Self-Standing Polypyrrole/Black Phosphorus Laminated Film: Promising Electrode for Flexible Supercapacitor with Enhanced Capacitance and Cycling Stability. ACS Appl. Mater. Interfaces 2018, 10, 3538–3548. [Google Scholar] [CrossRef] [PubMed]
  418. Wang, D.Y.; Chang, C.L.; Ho, W.Y. Microstructure analysis of MoS2 deposited on diamond-like carbon films for wear improvement. Surf. Coat. Technol. 1999, 111, 123–127. [Google Scholar] [CrossRef]
  419. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  420. Yaghi, O.M.; Li, H. Hydrothermal Synthesis of a Metal-Organic Framework Containing Large Rectangular Channels. J. Am. Chem. Soc. 1995, 117, 10401–10402. [Google Scholar] [CrossRef]
  421. Alahakoon, S.B.; Thompson, C.M.; Occhialini, G.; Smaldone, R.A. Design Principles for Covalent Organic Frameworks in Energy Storage Applications. ChemSusChem 2017, 10, 2116–2129. [Google Scholar] [CrossRef]
  422. Yang, W.; Gao, Z.; Wang, J.; Ma, J.; Zhang, M.; Liu, L. Solvothermal one-step synthesis of Ni-Al layered double hydroxide/carbon nanotube/reduced graphene oxide sheet ternary nanocomposite with ultrahigh capacitance for supercapacitors. ACS Appl. Mater. Interfaces 2013, 5, 5443–5454. [Google Scholar] [CrossRef]
  423. Ong, W.J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer to Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  424. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for Synthesis and Processing of Two-Dimensional Titanium Carbide (Ti3C2Tx MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  425. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  426. Forouzandeh, P.; Pillai, S.C. Two-dimensional (2D) electrode materials for supercapacitors. Mater. Today Proc. 2020. [Google Scholar] [CrossRef]
  427. Xie, L.M. Two-dimensional transition metal dichalcogenide alloys: Preparation, characterization and applications. Nanoscale 2015, 7, 18392–18401. [Google Scholar] [CrossRef] [PubMed]
  428. Guan, G.; Han, M.Y. Functionalized Hybridization of 2D Nanomaterials. Adv. Sci. 2019, 6, 23. [Google Scholar] [CrossRef] [Green Version]
  429. Wan, J.; Lacey, S.D.; Dai, J.; Bao, W.; Fuhrer, M.S.; Hu, L. Tuning two-dimensional nanomaterials by intercalation: Materials, properties and applications. Chem. Soc. Rev. 2016, 45, 6742–6765. [Google Scholar] [CrossRef]
  430. Zhang, X.; Lai, Z.; Liu, Z.; Tan, C.; Huang, Y.; Li, B.; Zhao, M.; Xie, L.; Huang, W.; Zhang, H. A facile and universal top-down method for preparation of monodisperse transition-metal dichalcogenide nanodots. Angew. Chem. Int. Ed. 2015, 54, 5425–5428. [Google Scholar] [CrossRef]
  431. Pant, A.; Mutlu, Z.; Wickramaratne, D.; Cai, H.; Lake, R.K.; Ozkan, C.S.; Tongay, S. Fundamentals of lateral and vertical heterojunctions of atomically thin materials. Nanoscale 2016, 8, 3870–3887. [Google Scholar] [CrossRef]
  432. Sangwan, V.K.; Hersam, M.C. Electronic Transport in Two-Dimensional Materials. Annu. Rev. Phys. Chem. 2018, 69, 299–325. [Google Scholar] [CrossRef] [Green Version]
  433. Dai, Z.; Liu, L.; Zhang, Z. Strain Engineering of 2D Materials: Issues and Opportunities at the Interface. Adv. Mater. 2019, 31, 1–11. [Google Scholar] [CrossRef]
  434. Bao, W.; Jing, L.; Velasco, J.; Lee, Y.; Liu, G.; Tran, D.; Standley, B.; Aykol, M.; Cronin, S.B.; Smirnov, D.; et al. Stacking-dependent band gap and quantum transport in trilayer graphene. Nat. Phys 2011, 7, 948–952. [Google Scholar] [CrossRef] [Green Version]
  435. Xiong, W.; Hu, X.; Wu, X.; Zeng, Y.; Wang, B.; He, G.; Zhu, Z. A flexible fiber-shaped supercapacitor utilizing hierarchical NiCo2O4@polypyrrole core-shell nanowires on hemp-derived carbon. J. Mater. Chem. A 2015, 3, 17209–17216. [Google Scholar] [CrossRef]
  436. Hong, D.; Yim, S. RuO2 Thin Films Electrodeposited on Polystyrene Nanosphere Arrays: Growth Mechanism and Application to Supercapacitor Electrodes. Langmuir 2018, 34, 4249–4254. [Google Scholar] [CrossRef]
  437. L-Mahdy, A.F.M.E.; Hung, Y.-H.; Mansoure, T.H.; Yu, H.-H.; Hsu, Y.-S.; Wu, K.C.-W.; Kuo, S.-W. Synthesis of [3 + 3] β-ketoenamine-tethered covalent organic frameworks (COFs) for high-performance supercapacitance and CO2 storage. J. Taiwan Inst. Chem. Eng. 2019, 103, 199–208. [Google Scholar] [CrossRef]
  438. Das, S.K.; Bhunia, K.; Mallick, A.; Pradhan, A.; Pradhan, D.; Bhaumik, A. A new electrochemically responsive 2D Π-conjugated covalent organic framework as a high performance supercapacitor. Microporous Mater. 2018, 266, 109–116. [Google Scholar] [CrossRef] [Green Version]
  439. Xu, Q.; Dalapati, S.; Jiang, D. Charge up in wired covalent organic frameworks. ACS Cent. Sci. 2016, 2, 586–587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  440. Romero, J.; Rodríguez-San-Miguel, D.; Ribera, A.; Mas-Ballesté, R.; Otero, T.F.; Manet, I.; Licio, F.; Abellán, G.; Zamora, F.; Coronado, E. Metal-functionalized covalent organic frameworks as precursors of supercapacitive porous N-doped graphene. J. Mater. Chem. A 2017, 5, 4343–4351. [Google Scholar] [CrossRef] [Green Version]
  441. Mulzer, C.R.; Shen, L.; Bisbey, R.P.; McKone, J.R.; Zhang, N.; Abruña, H.D.; Dichtel, W.R. Superior charge storage and power density of a conducting polymer-modified covalent organic framework. ACS Cent. Sci. 2016, 2, 667–673. [Google Scholar] [CrossRef]
  442. Chandra, S.; Chowdhury, D.R.; Addicoat, M.; Heine, T.; Paul, A.; Banerjee, R. Molecular Level Control of the Capacitance of Two-Dimensional Covalent Organic Frameworks: Role of Hydrogen Bonding in Energy Storage Materials. Chem. Mater. 2017, 29, 2074–2080. [Google Scholar] [CrossRef]
  443. Xu, Z.; Chowdhury, D.R.; Addicoat, M.; Heine, T.; Paul, A.; Banerjee, R. Construction of extensible and flexible supercapacitors from covalent organic framework composite membrane electrode. Chem. Eng. J. 2020, 387, 124071. [Google Scholar] [CrossRef]
  444. Choi, K.M.; Jeong, H.M.; Park, J.H.; Zhang, Y.B.; Kang, J.K.; Yaghi, O.M. Supercapacitors of nanocrystalline metal-organic frameworks. ACS Nano 2014, 8, 7451–7457. [Google Scholar] [CrossRef]
  445. Jiang, J.W.; Park, H.S. Mechanical properties of MoS2/graphene heterostructures. Appl. Phys. Lett. 2014, 105, 033108. [Google Scholar] [CrossRef] [Green Version]
  446. Toth, P.S.; Velický, M.; Bissett, M.A.; Slater, T.J.A.; Savjani, N.; Rabiu, A.; Rakowski, A.M.; Brent, J.R.; Haigh, S.J.; O’Brien, P.; et al. Asymmetric MoS2/Graphene/Metal Sandwiches: Preparation, Characterization, and Application. Adv. Mater. 2016, 28, 8256–8264. [Google Scholar] [CrossRef] [Green Version]
  447. Zhang, X.; Selkirk, A.; Zhang, S.; Huang, J.; Li, Y.; Xie, Y.; Dong, N.; Cui, Y.; Zhang, L.; Blau, W.J.; et al. MoS2/Carbon Nanotube Core–Shell Nanocomposites for Enhanced Nonlinear Optical Performance. Chem. A Eur. J. 2017, 23, 3321–3327. [Google Scholar] [CrossRef] [PubMed]
  448. Nejati, K.; Akbari, A.R.; Davari, S.; Asadpour-Zeynali, K.; Rezvani, Z. Zn–Fe-layered double hydroxide intercalated with vanadate and molybdate anions for electrocatalytic water oxidation. New J. Chem. 2018, 42, 2889–2895. [Google Scholar] [CrossRef]
  449. Li, M.Y.; Chen, C.H.; Shi, Y.; Li, L.J. Heterostructures based on two-dimensional layered materials and their potential applications. Mater. Today 2016, 19, 322–335. [Google Scholar] [CrossRef] [Green Version]
  450. Huang, X.; Zeng, Z.; Zhang, H. Metal dichalcogenide nanosheets: Preparation, properties and applications. Chem. Soc. Rev. 2013, 42, 1934–1946. [Google Scholar] [CrossRef] [PubMed]
  451. Yang, X.; Niu, H.; Jiang, H.; Wang, Q.; Qu, F. A high energy density all-solid-state asymmetric supercapacitor based on MoS2/graphene nanosheets and MnO2/graphene hybrid electrodes. J. Mater. Chem. A 2016, 4, 11264–11275. [Google Scholar] [CrossRef]
  452. Leng, K.; Chen, Z.; Zhao, X.; Tang, W.; Tian, B.; Nai, C.T.; Zhou, W.; Loh, K.P. Phase Restructuring in Transition Metal Dichalcogenides for Highly Stable Energy Storage. ACS Nano 2016, 10, 9208–9215. [Google Scholar] [CrossRef] [Green Version]
  453. Yousaf, M.; Mahmood, A.; Wang, Y.; Chen, Y.; Ma, Z.; Han, R.P.S. Advancement in Layered Transition Metal Dichalcogenide Composites for Lithium and Sodium Ion Batteries. J. Electr. Eng. 2016, 4, 58–74. [Google Scholar] [CrossRef]
  454. Saha, S.; Jana, M.; Khanra, P.; Samanta, P.; Koo, H.; Murmu, N.C.; Kuila, T. Band gap engineering of boron nitride by graphene and its application as positive electrode material in asymmetric supercapacitor device. ACS Appl. Mater. Interfaces 2015, 7, 14211–14222. [Google Scholar] [CrossRef]
  455. Maity, C.K.; Sahoo, S.; Verma, K.; Behera, A.K.; Nayak, G.C. Facile functionalization of boron nitride (BN) for the development of high-performance asymmetric supercapacitors. New J. Chem. 2020, 44, 8106–8119. [Google Scholar] [CrossRef]
  456. Gomez, I.J.; Arnaiz, B.; Cacioppo, M.; Arcudi, F.; Prato, M. Synthesis of 3D Marigold Flower-Like rGO/BN/Ni(OH)2 Ternary Nanocomposites for Supercapacitor Applications. J. Mater. Chem. B 2018, 35. [Google Scholar] [CrossRef]
  457. Zhao, G.; Wang, T.; Shao, Y.; Wu, Y.; Huang, B.; Hao, X. A Novel Mild Phase-Transition to Prepare Black Phosphorus Nanosheets with Excellent Energy Applications. Small 2017, 13, 1602243. [Google Scholar] [CrossRef] [PubMed]
  458. Yu, X.; Ushiyama, H.; Yamashita, K. Comparative Study of Sodium and Lithium Intercalation and Diffusion Mechanism in Black Phosphorus from First-principles Simulation. Chem. Lett. 2014, 43, 1940–1942. [Google Scholar] [CrossRef]
  459. Gopalakrishnan, A.; Badhulika, S. Facile sonochemical assisted synthesis of a hybrid red–black phosphorus/sulfonated porous carbon composite for high-performance supercapacitors. Chem. Commun. 2020, 56, 7096–7099. [Google Scholar] [CrossRef] [PubMed]
  460. Kim, G.; Wang, S.; Jacobson, A.J.; Reimus, L.; Brodersen, P.; Mims, C.A. Rapid oxygen ion diffusion and surface exchange kinetics in PrBaCo2O5+x with a perovskite related structure and ordered a cations. J. Mater. Chem. 2007, 17, 2500–2505. [Google Scholar] [CrossRef]
  461. Mishra, M. (Ed.) Encyclopedia of Polymer Applications, 3 Volume Set; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  462. Yan, X.; Tong, X.; Wang, J.; Gong, C.; Zhang, M.; Liang, L. Synthesis of mesoporous NiO nanoflake array and its enhanced electrochemical performance for supercapacitor application. J. Alloys Compd. 2014, 593, 184–189. [Google Scholar] [CrossRef]
  463. Lu, K.; Shen, F. Long term behaviors of La0.8Sr0.2MnO3 and La0.6Sr0.4Co0.2Fe0.8O3 as cathodes for solid oxide fuel cells. Int. J. Hydrogen Energy 2014, 39, 7963–7971. [Google Scholar] [CrossRef]
  464. Zhang, L.; Zhou, Q.; He, Q.; He, T. Double-perovskites A2FeMoO6-δ (A = Ca, Sr, Ba) as anodes for solid oxide fuel cells. J. Power Sources 2010, 195, 6356–6366. [Google Scholar] [CrossRef]
  465. Wang, X.W.; Zhu, Q.Q.; Wang, X.E.; Zhang, H.C.; Zhang, J.J.; Wang, L.F. Structural and electrochemical properties of La0.85Sr0.15MnO3 powder as an electrode material for supercapacitor. J. Alloys Compd. 2016, 675, 195–200. [Google Scholar] [CrossRef]
  466. Ma, P.P.; Lu, Q.L.; Lei, N.; Liu, Y.K.; Yu, B.; Dai, J.M.; Li, S.H.; Jiang, G. Effect of A-site substitution by Ca or Sr on the structure and electrochemical performance of LaMnO3 perovskite. Electrochim. Acta 2020, 332, 135489. [Google Scholar] [CrossRef]
  467. Tomar, A.K.; Singh, G.; Sharma, R.K. Charge storage characteristics of mesoporous strontium titanate perovskite aqueous as well as flexible solid-state supercapacitor cell. J. Power Sources 2019, 426, 223–232. [Google Scholar] [CrossRef]
  468. Du, P.; Hu, X.; Yi, C.; Liu, H.C.; Liu, P.; Zhang, H.-L.; Gong, X. Self-Powered Electronics by Integration of Flexible Solid-State Graphene-Based Supercapacitors with High Performance Perovskite Hybrid Solar Cells. Adv. Funct. Mater. 2015, 25, 16–2420. [Google Scholar] [CrossRef]
  469. Ayman, I.; Rasheed, A.; Ajmal, S.; Rehman, A.; Ali, A.; Shakir, I.; Warsi, M.F. CoFe2O4 Nanoparticle-Decorated 2D MXene: A Novel Hybrid Material for Supercapacitor Applications. Energy Fuels 2020, 34, 7622–7630. [Google Scholar] [CrossRef]
  470. Yang, J.; Pan, Z.; Yu, Q.; Zhang, Q.; Ding, X.; Shi, X.; Qiu, Y.-C.; Zhang, K.; Wang, J.; Zhang, Y. Free-Standing Black Phosphorus Thin Films for Flexible Quasi-Solid-State Micro-Supercapacitors with High Volumetric Power and Energy Density. ACS Appl. Mater. Interfaces 2019, 11, 5938–5946. [Google Scholar] [CrossRef]
  471. Wang, L.; Liu, F.; Zhao, B.; Ning, Y.; Zhang, L.; Bradley, R.H.; Wu, W.; Lijian, W. Carbon Nanobowls Filled with MoS2 Nanosheets as Electrode Materials for Supercapacitors. ACS Appl. Nano Mater. 2020, 3, 6448–6459. [Google Scholar] [CrossRef]
  472. Kitchamsetti, N.; Choudhary, R.J.; Phase, D.M.; Devan, R.S. Structural correlation of a nanoparticle-embedded mesoporous CoTiO3 perovskite for an efficient electrochemical supercapacitor. RSC Adv. 2020, 10, 23446–23456. [Google Scholar] [CrossRef]
  473. Haldar, S.; Kushwaha, R.; Maity, R.; Vaidhyanathan, R. Pyridine-Rich Covalent Organic Frameworks as High-Performance Solid-State Supercapacitors. ACS Mater. Lett. 2019, 1, 490–497. [Google Scholar] [CrossRef]
  474. Liu, Y.; Wang, Y.; Chen, Y.; Wang, C.; Guo, L. NiCo-MOF nanosheets wrapping polypyrrole nanotubes for high-performance supercapacitors. Appl. Surf. Sci. 2020, 507, 145089. [Google Scholar] [CrossRef]
  475. Shao, Y.; El-Kady, M.F.; Sun, J.; Li, Y.; Zhang, Q.; Zhu, M.; Wang, H.; Dunn, B.; Kaner, R.B. Design and Mechanisms of Asymmetric Supercapacitors. Chem. Rev. 2018, 118, 9233–9280. [Google Scholar] [CrossRef]
  476. He, Y.; Chen, W.; Li, X.; Zhang, Z.; Fu, J.; Zhao, C.; Xie, E. Freestanding three-dimensional graphene/Mno2 composite networks as ultralight and flexible supercapacitor electrodes. ACS Nano 2013, 7, 174–182. [Google Scholar] [CrossRef] [PubMed]
  477. Lu, X.; Xia, Y. Electronic materials: Buckling down for flexible electronics. Nat. Nanotechnol. 2006, 1, 163–164. [Google Scholar] [CrossRef]
  478. Miller, J.R. Reliability of Electrochemical Capacitors. In Supercapacitors; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; pp. 473–507. [Google Scholar]
  479. Maggetto, G. Electric vehicles, hybrid electric vehicles and fuel cell electric vehicles: State of the art and perspectives. Ann. Chim. Sci. des Matà riaux 2001, 26, 9–26. [Google Scholar] [CrossRef]
  480. He, Z.; Gao, B.; Li, T.; Liao, J.; Liu, B.; Liu, X.; Wang, C.; Feng, Z.-Q.; Gu, Z. Piezoelectric-Driven Self-Powered Patterned Electrochromic Supercapacitor for Human Motion Energy Harvesting. ACS Sustain. Chem. Eng. 2019, 7, 1745–1752. [Google Scholar] [CrossRef]
  481. Hsu, H.H.; Khosrozadeh, A.; Li, B.; Luo, G.; Xing, M.; Zhong, W. An Eco-Friendly, Nanocellulose/RGO/in Situ Formed Polyaniline for Flexible and Free-Standing Supercapacitors. ACS Sustain. Chem. Eng. 2019, 7, 4766–4776. [Google Scholar] [CrossRef]
  482. Chia, Y.Y.; Lee, L.H.; Shafiabady, N.; Isa, D. A load predictive energy management system for supercapacitor-battery hybrid energy storage system in solar application using the Support Vector Machine. Appl. Energy 2015, 137, 588–602. [Google Scholar] [CrossRef]
  483. Habibzadeh, M.; Hassanalieragh, M.; Ishikawa, A.; Soyata, T.; Sharma, G. Hybrid Solar-Wind Energy Harvesting for Embedded Applications: Supercapacitor-Based System Architectures and Design Tradeoffs. IEEE Circuits Syst. Mag. 2017, 17, 29–63. [Google Scholar] [CrossRef]
  484. Zhang, Y.; Hao, N.; Lin, X.; Nie, S. Emerging challenges in the thermal management of cellulose nanofibril-based supercapacitors, lithium-ion batteries and solar cells: A review. Carbohydr. Polym. 2020, 234, 115888. [Google Scholar] [CrossRef]
  485. Miller, J.R. Market and Applications of Electrochemical Capacitors. In Supercapacitors; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; pp. 509–526. [Google Scholar]
  486. Li, Z.; Huang, T.; Gao, W.; Xu, Z.; Chang, D.; Zhang, C.; Gao, C. Hydrothermally Activated Graphene Fiber Fabrics for Textile Electrodes of Supercapacitors. ACS Nano 2017, 11, 11056–11065. [Google Scholar] [CrossRef]
  487. Wu, Q.; Xu, Y.; Yao, Z.; Liu, A.; Shi, G. Supercapacitors based on flexible graphene/polyaniline nanofiber composite films. ACS Nano 2010, 4, 1963–1970. [Google Scholar] [CrossRef]
  488. Zhu, C.; Liu, T.; Qian, F.; Han, T.Y.-J.; Duoss, E.B.; Kuntz, J.D.; Spadaccini, C.M.; Worsley, M.A.; Li, Y. Supercapacitors Based on Three-Dimensional Hierarchical Graphene Aerogels with Periodic Macropores. Nano Lett. 2016, 16, 3448–3456. [Google Scholar] [CrossRef] [PubMed]
  489. Couly, C.; Alhabeb, M.; Van Aken, K.L.; Kurra, N.; Gomes, L.; Anasori, B.; Alshareef, H.N.; Gogotsi, Y.; Navarro-Suárez, A.M. Asymmetric Flexible MXene-Reduced Graphene Oxide Micro-Supercapacitor. Adv. Electron. Mater. 2018, 4, 1700339. [Google Scholar] [CrossRef] [Green Version]
  490. Zhang, C.J.; Anasori, B.; Seral-Ascaso, A.; Park, S.-H.; McEvoy, N.; Shmeliov, A.; Duesberg, G.S.; Coleman, J.N.; Gogotsi, Y.; Nicolosi, V. Transparent, Flexible, and Conductive 2D Titanium Carbide (MXene) Films with High Volumetric Capacitance. Adv. Mater. 2017, 29, 1–9. [Google Scholar] [CrossRef] [PubMed]
  491. Fleischmann, S.; Mitchell, J.B.; Wang, R.; Zhan, C.; Jiang, D.; Presser, V.; Augustyn, V. Pseudocapacitance: From Fundamental Understanding to High Power Energy Storage Materials. Chem. Rev. 2020, 120, 6738–6782. [Google Scholar] [CrossRef]
  492. Miroshnikov, M.; Mahankali, K.; Thangavel, N.K.; Satapathy, S.; Arava, L.M.R.; Ajayan, P.M.; John, G. Bioderived Molecular Electrodes for Next-Generation Energy-Storage Materials. ChemSusChem 2020, 1–20. [Google Scholar] [CrossRef]
Figure 1. Ragone plot of the power-energy density range for different electrochemical energy storage devices.
Figure 1. Ragone plot of the power-energy density range for different electrochemical energy storage devices.
Catalysts 10 00969 g001
Figure 2. Schematic illustration of supercapacitors: (a) electrical double-layer capacitors (EDLC), (b) Pseudocapacitors, and (c) hybrid supercapacitors (HSCs).
Figure 2. Schematic illustration of supercapacitors: (a) electrical double-layer capacitors (EDLC), (b) Pseudocapacitors, and (c) hybrid supercapacitors (HSCs).
Catalysts 10 00969 g002
Figure 3. (a) Redox pseudocapacitance, as in RuO2-xH2O. (b) Intercalation pseudocapacitance, as in Nb2O5. (c) Doping pseudocapacitance, as in polypyrrole, and (d) under potential deposition based pseudocapacitance, such as Pb deposition on the surface of Au electrode. Reprinted with permission from [17], Copyright 2019, American Chemical Society.
Figure 3. (a) Redox pseudocapacitance, as in RuO2-xH2O. (b) Intercalation pseudocapacitance, as in Nb2O5. (c) Doping pseudocapacitance, as in polypyrrole, and (d) under potential deposition based pseudocapacitance, such as Pb deposition on the surface of Au electrode. Reprinted with permission from [17], Copyright 2019, American Chemical Society.
Catalysts 10 00969 g003
Figure 4. The electrochemical signatures for various type of charge storage mechanisms as EDLC, pseudocapacitive, and battery-type: (a) CVs, and (b) GCDs, along with their (c) quantitative kinetics analysis through the calculation of b in formula i(V) = avb. Reprinted with permission from [49], Copyright 2019, Wiley-VCH.
Figure 4. The electrochemical signatures for various type of charge storage mechanisms as EDLC, pseudocapacitive, and battery-type: (a) CVs, and (b) GCDs, along with their (c) quantitative kinetics analysis through the calculation of b in formula i(V) = avb. Reprinted with permission from [49], Copyright 2019, Wiley-VCH.
Catalysts 10 00969 g004
Figure 5. Schematic illustration of the most common supercapacitor electrode materials from 0D to the 3D structure.
Figure 5. Schematic illustration of the most common supercapacitor electrode materials from 0D to the 3D structure.
Catalysts 10 00969 g005
Figure 6. SEM images of (a) binary ACE (activated carbon electrode) composite electrode, with the weight percentage of 90 to 5 for activated carbons (AC)/carbon black (CB). (b) Ternary NCE-1 composite electrode (with the composition of AC/CB/carbon nanotube (CNT) (90/1.25/3.75)). (c) Ternary NCE-2 composite electrode (with the composition of (AC/CB/CNT/carbon nanofibers (CNF) (90/1.25/3.125/0.625)). (d) Quaternary NCE-3 composite electrode (with the composition of AC/CB/CNT/CNF (88.75/1.25/3.75/1.25)). (e) Capacitance retentions of various composite electrodes. (f) Ragone plots of the performance of various composite electrodes according to the total volume of the coating films of the two electrodes of the device. Reprinted with permission from [104], Copyright 2019, Elsevier.
Figure 6. SEM images of (a) binary ACE (activated carbon electrode) composite electrode, with the weight percentage of 90 to 5 for activated carbons (AC)/carbon black (CB). (b) Ternary NCE-1 composite electrode (with the composition of AC/CB/carbon nanotube (CNT) (90/1.25/3.75)). (c) Ternary NCE-2 composite electrode (with the composition of (AC/CB/CNT/carbon nanofibers (CNF) (90/1.25/3.125/0.625)). (d) Quaternary NCE-3 composite electrode (with the composition of AC/CB/CNT/CNF (88.75/1.25/3.75/1.25)). (e) Capacitance retentions of various composite electrodes. (f) Ragone plots of the performance of various composite electrodes according to the total volume of the coating films of the two electrodes of the device. Reprinted with permission from [104], Copyright 2019, Elsevier.
Catalysts 10 00969 g006
Figure 7. AC with high surface area; however, most of the micropores surface is not available for the electrolyte ions. (a) Single-walled carbon nanotube (SWCNT). (b) Cyclic voltammograms (CVs) for different electrode materials made of CNTs, graphene, and their composite in the aqueous electrolyte at a scan rate of 10 mV·s−1 (c), CVs for the same material as part (c), but in a different electrolyte (1 M TEABF4/PC electrolyte) at the same scan rate of 10 mV·s−1. (d) The illustration of the electro-activation phenomenon to improve the electrode surface area after charging–discharging cycles: Graphene layers before cycling, which are tending to restack. (e) After a long cycle life, the intercalated ions separated the graphene sheets, thus increasing the surface area accessible for electrolyte ions, and consequently an enhancement in the capacitive property of the material. (f) Reprinted with permission from [108], Copyright 2011, Royal Society of Chemistry.
Figure 7. AC with high surface area; however, most of the micropores surface is not available for the electrolyte ions. (a) Single-walled carbon nanotube (SWCNT). (b) Cyclic voltammograms (CVs) for different electrode materials made of CNTs, graphene, and their composite in the aqueous electrolyte at a scan rate of 10 mV·s−1 (c), CVs for the same material as part (c), but in a different electrolyte (1 M TEABF4/PC electrolyte) at the same scan rate of 10 mV·s−1. (d) The illustration of the electro-activation phenomenon to improve the electrode surface area after charging–discharging cycles: Graphene layers before cycling, which are tending to restack. (e) After a long cycle life, the intercalated ions separated the graphene sheets, thus increasing the surface area accessible for electrolyte ions, and consequently an enhancement in the capacitive property of the material. (f) Reprinted with permission from [108], Copyright 2011, Royal Society of Chemistry.
Catalysts 10 00969 g007
Figure 8. Graphical representation of the synthesis of N-pGr and graphene quantum dots (GQDs) and their potential application for the fabrication of the asymmetric supercapacitors (ASSC) device. Reprinted with permission from [136], Copyright 2017, Elsevier.
Figure 8. Graphical representation of the synthesis of N-pGr and graphene quantum dots (GQDs) and their potential application for the fabrication of the asymmetric supercapacitors (ASSC) device. Reprinted with permission from [136], Copyright 2017, Elsevier.
Catalysts 10 00969 g008
Figure 9. SEM images of the samples (a) G1p4K4-2, and (b) G1p9K6-4, (c)The CV curves illustration for the HSG device electrodes at a scan rate of 20 mV·s−1. (d) The specific capacitance of the various samples in different densities, for the HSG device with the TEABF4 as the electrolyte. (e) Ragone plot for the HSG device including the propylene carbonate + TEABF4, as electrolyte material. (f) The pore size distribution of the HSG electrode material samples. Reprinted with permission from [139], Copyright 2015, Royal Society of Chemistry.
Figure 9. SEM images of the samples (a) G1p4K4-2, and (b) G1p9K6-4, (c)The CV curves illustration for the HSG device electrodes at a scan rate of 20 mV·s−1. (d) The specific capacitance of the various samples in different densities, for the HSG device with the TEABF4 as the electrolyte. (e) Ragone plot for the HSG device including the propylene carbonate + TEABF4, as electrolyte material. (f) The pore size distribution of the HSG electrode material samples. Reprinted with permission from [139], Copyright 2015, Royal Society of Chemistry.
Catalysts 10 00969 g009
Figure 10. Illustration of the physically-chemically conjugation of Nile Blue. (a) The 3D porous structure of graphene aerogel (b), The possible faradic reactions of Nile Blue in strongly acidic and alkaline electrolyte, (c) The potential windows of the Nile Blue-conjugated graphene aerogel (NB-GA) composite-based electrodes in acidic, neutral, and alkaline media, (d) Potential windows of the NB-GA electrodes in acidic, neutral, and alkaline solutions. Reprinted with permission from [142], Copyright 2019, American Chemical Society.
Figure 10. Illustration of the physically-chemically conjugation of Nile Blue. (a) The 3D porous structure of graphene aerogel (b), The possible faradic reactions of Nile Blue in strongly acidic and alkaline electrolyte, (c) The potential windows of the Nile Blue-conjugated graphene aerogel (NB-GA) composite-based electrodes in acidic, neutral, and alkaline media, (d) Potential windows of the NB-GA electrodes in acidic, neutral, and alkaline solutions. Reprinted with permission from [142], Copyright 2019, American Chemical Society.
Catalysts 10 00969 g010
Figure 11. The development of the metal organic framework (MOF) synthesis derived from porous metal oxides (MOs). Bar graphs presentating a variation of MOF precursors, which have been utilized to arise MOs nanostructures (CuO [164], Co3O4 [165,166,167,168,169,170,171,172] Co3V2O8 [173], CeO2 [174], Fe2O3 [175,176,177], MgO [178], Mn2O3 [179,180,181], NiO [182,183,184], TiO2 [185,186,187] and ZnO [188,189,190]) and nanocomposites (CoFe2O4 [191], CuCo2O4 [192] Mn1.8Fe1.2O4 [193], ZnCo2O4 [194], ZnxCo3-XO4 [195], ZnMn2O4 [196], Co3O4@NiCo2O4 [155], Co3O4/ZnFe2O4 [154], CuO/Cu2O [149], CuO@NiO [148], Cu/Cu2O@TiO2 [150], Cr2O3@TiO2 [159], Fe2O3@TiO2 [151] and Fe2O3/NiCo2O4 [152]). The corresponding surface area values of the attained metal oxide nanostructures and nanocomposites. (Note: PB, Prussian Blue; BTC, benzene-1,3,5-tricarboxylic acid; ptcda, perylene-3,4,9,10-tetracarboxylic dianhydride; Cu BDC:[Cu(tpa)(dmf)]; tpa, terephthalic acid; DMF, N,N-dimethylformamide; Mg aph,Mg4(adipate)4(DMA)-(H2O)]·5DMA·2MeOH·4H2O; DMA, N,N-dimethylacetamide; MeOH, methanol). Reprinted with permission from [163], Copyright 2017, American chemical Society.
Figure 11. The development of the metal organic framework (MOF) synthesis derived from porous metal oxides (MOs). Bar graphs presentating a variation of MOF precursors, which have been utilized to arise MOs nanostructures (CuO [164], Co3O4 [165,166,167,168,169,170,171,172] Co3V2O8 [173], CeO2 [174], Fe2O3 [175,176,177], MgO [178], Mn2O3 [179,180,181], NiO [182,183,184], TiO2 [185,186,187] and ZnO [188,189,190]) and nanocomposites (CoFe2O4 [191], CuCo2O4 [192] Mn1.8Fe1.2O4 [193], ZnCo2O4 [194], ZnxCo3-XO4 [195], ZnMn2O4 [196], Co3O4@NiCo2O4 [155], Co3O4/ZnFe2O4 [154], CuO/Cu2O [149], CuO@NiO [148], Cu/Cu2O@TiO2 [150], Cr2O3@TiO2 [159], Fe2O3@TiO2 [151] and Fe2O3/NiCo2O4 [152]). The corresponding surface area values of the attained metal oxide nanostructures and nanocomposites. (Note: PB, Prussian Blue; BTC, benzene-1,3,5-tricarboxylic acid; ptcda, perylene-3,4,9,10-tetracarboxylic dianhydride; Cu BDC:[Cu(tpa)(dmf)]; tpa, terephthalic acid; DMF, N,N-dimethylformamide; Mg aph,Mg4(adipate)4(DMA)-(H2O)]·5DMA·2MeOH·4H2O; DMA, N,N-dimethylacetamide; MeOH, methanol). Reprinted with permission from [163], Copyright 2017, American chemical Society.
Catalysts 10 00969 g011
Figure 12. Graphical presentation of PMNC/G-2. (a) The layered structure of Ni0.5 Co0.5(BO2)y (OH)2−y xH2O. (b,c) Cycling performances of PMNC//AC and PMNC/G-2//AC ASSCs. Reprinted with permission from [197], Copyright 2017, American Chemical Society.
Figure 12. Graphical presentation of PMNC/G-2. (a) The layered structure of Ni0.5 Co0.5(BO2)y (OH)2−y xH2O. (b,c) Cycling performances of PMNC//AC and PMNC/G-2//AC ASSCs. Reprinted with permission from [197], Copyright 2017, American Chemical Society.
Catalysts 10 00969 g012
Figure 13. The graphical representation of the battery-type materials according to two electrochemical reaction mechanisms (intercalation and conversion-type battery), utilized in HSCs. Reprinted with permission from [199], Copyright 2020, Elsevier.
Figure 13. The graphical representation of the battery-type materials according to two electrochemical reaction mechanisms (intercalation and conversion-type battery), utilized in HSCs. Reprinted with permission from [199], Copyright 2020, Elsevier.
Catalysts 10 00969 g013
Figure 14. Graphical presentation of the fabrication procedure of the Ni(BO2−)-LDH-X. Reprinted with permission from [268], Copyright 2020, American Chemical society.
Figure 14. Graphical presentation of the fabrication procedure of the Ni(BO2−)-LDH-X. Reprinted with permission from [268], Copyright 2020, American Chemical society.
Catalysts 10 00969 g014
Figure 15. SEM image of the porous surface of electrochemically activated carbon cloth (ECC). (a), SEM images of ECN-3 composite electrode material (the top inset signifies the top-view image of the free-standing electrode at low-resolution mode). (b), SEM image of the compact gown nanosheets (NSs) with a distinct open porous structure. (c), the graphical illustration of the flexible ASC using ECN-3 and ECC electrodes. (d), CVs of various Ni(OH)2 NSs/ECC electrodes at 20 mV·s−1. (The samples were fabricated with various amount of nickel nitrate hexahydrate (0.1 g, 0.2 g, 0.3 g, and 0.4 g), and named ECN-1, ECN-2, ECN-3, and ECN-4, respectively. (e) Specific capacities of the ECN electrodes at various current densities. (blue line: ECN-1, green line: ECN-2, black line: ECN-3, red line: ECN-4) (f). Reprinted with permission from [84], Copyright 2020, Elsevier.
Figure 15. SEM image of the porous surface of electrochemically activated carbon cloth (ECC). (a), SEM images of ECN-3 composite electrode material (the top inset signifies the top-view image of the free-standing electrode at low-resolution mode). (b), SEM image of the compact gown nanosheets (NSs) with a distinct open porous structure. (c), the graphical illustration of the flexible ASC using ECN-3 and ECC electrodes. (d), CVs of various Ni(OH)2 NSs/ECC electrodes at 20 mV·s−1. (The samples were fabricated with various amount of nickel nitrate hexahydrate (0.1 g, 0.2 g, 0.3 g, and 0.4 g), and named ECN-1, ECN-2, ECN-3, and ECN-4, respectively. (e) Specific capacities of the ECN electrodes at various current densities. (blue line: ECN-1, green line: ECN-2, black line: ECN-3, red line: ECN-4) (f). Reprinted with permission from [84], Copyright 2020, Elsevier.
Catalysts 10 00969 g015
Figure 16. Schematic illustration of the as-fabricated battery−supercapacitor hybrid (BSH) device. (a) Comparative Ragone plot of the various stated hybrid storage devices with the as-fabricated BSH. (b) The FESEM of the rGO@VO2 composite. (c) Reprinted with permission from [80], Copyright 2019, American Chemical Society.
Figure 16. Schematic illustration of the as-fabricated battery−supercapacitor hybrid (BSH) device. (a) Comparative Ragone plot of the various stated hybrid storage devices with the as-fabricated BSH. (b) The FESEM of the rGO@VO2 composite. (c) Reprinted with permission from [80], Copyright 2019, American Chemical Society.
Catalysts 10 00969 g016
Figure 17. Graphical presentation of PEDOT/PANI hydrogels fabrication through two-step mechanism(a), Graphical presentation of the formation process of the PEDOT/PANI hydrogels. (b) The comparative CVs of PEDOT and PEDOT/PANI at 5 mV·s−1. (c), Compressive stress-strain curve of PEDOT/PANI hydrogels. (d) Reprinted with permission from [352], Copyright 2020, American Chemical Society.
Figure 17. Graphical presentation of PEDOT/PANI hydrogels fabrication through two-step mechanism(a), Graphical presentation of the formation process of the PEDOT/PANI hydrogels. (b) The comparative CVs of PEDOT and PEDOT/PANI at 5 mV·s−1. (c), Compressive stress-strain curve of PEDOT/PANI hydrogels. (d) Reprinted with permission from [352], Copyright 2020, American Chemical Society.
Catalysts 10 00969 g017
Figure 18. (a) The structure of the PEI−PVA−Bn hydrogel network (i) and the sliding development of the dynamic cross-linking sites. (ii) Graphical presentation of the SC consisting the PEI−PVA−Bn-LiCl-1 hydrogel electrolyte and MWCNT electrode. (b) Reprinted with permission from [353], Copyright 2020, American Chemical Society.
Figure 18. (a) The structure of the PEI−PVA−Bn hydrogel network (i) and the sliding development of the dynamic cross-linking sites. (ii) Graphical presentation of the SC consisting the PEI−PVA−Bn-LiCl-1 hydrogel electrolyte and MWCNT electrode. (b) Reprinted with permission from [353], Copyright 2020, American Chemical Society.
Catalysts 10 00969 g018
Figure 19. (a) The presentation of the MGP nanocomposite fabrication. The FESEM images of (b) GO, (c) MnO2 nanorods, (d) MG nanocomposite and (e) MGP nanocomposite. The CVs of (f) all the electrode material samples at a scan rate of 100 mV·s−1 and (g) the MGP nanocomposite at various scan rates. Reprinted with permission from [388], Copyright 2017, Royal Society of Chemistry.
Figure 19. (a) The presentation of the MGP nanocomposite fabrication. The FESEM images of (b) GO, (c) MnO2 nanorods, (d) MG nanocomposite and (e) MGP nanocomposite. The CVs of (f) all the electrode material samples at a scan rate of 100 mV·s−1 and (g) the MGP nanocomposite at various scan rates. Reprinted with permission from [388], Copyright 2017, Royal Society of Chemistry.
Catalysts 10 00969 g019
Figure 20. A graphical presentation of different types of ultrathin 2D nanomaterials.
Figure 20. A graphical presentation of different types of ultrathin 2D nanomaterials.
Catalysts 10 00969 g020
Figure 21. Illustration of the preparation process of the PS/RuO2 core-shells, including the surface modification of PS through the formation of the Ru−Cl bonds on Cl, terminated groups before the electrodeposition step. Reprinted with permission from [436], Copyright. 2018, American Chemical Society.
Figure 21. Illustration of the preparation process of the PS/RuO2 core-shells, including the surface modification of PS through the formation of the Ru−Cl bonds on Cl, terminated groups before the electrodeposition step. Reprinted with permission from [436], Copyright. 2018, American Chemical Society.
Catalysts 10 00969 g021
Figure 22. Evaluation of the observed (blue), predicted (green), and refined (red) PXRD patterns with their alteration (black) for (a) TpPa-(OH)2, (b) TpBD-(OH)2, and (c) TpPa-(OMe)2. Inset figures display the pore structure and π−π stacking distance for each of the COFs. Comparison of N2 adsorption isotherms at 77 K for (d) TpPa series [TpPa-(OH)2 (blue), TpPa-(OMe)2 (green), and TpPa-1 (brown)] and (e) TpBD series [TpBD-(OH)2 (blue), TpBD-(OMe)2 (green), and TpBD (brown)]. (f) Pore size distribution and (g) TEM images for TpPa-(OH)2, TpBD-(OH)2, TpPa-(OMe)2, and TpBD-(OMe)2. Reprinted with permission from [442], Copyright 2017, American Chemical Society.
Figure 22. Evaluation of the observed (blue), predicted (green), and refined (red) PXRD patterns with their alteration (black) for (a) TpPa-(OH)2, (b) TpBD-(OH)2, and (c) TpPa-(OMe)2. Inset figures display the pore structure and π−π stacking distance for each of the COFs. Comparison of N2 adsorption isotherms at 77 K for (d) TpPa series [TpPa-(OH)2 (blue), TpPa-(OMe)2 (green), and TpPa-1 (brown)] and (e) TpBD series [TpBD-(OH)2 (blue), TpBD-(OMe)2 (green), and TpBD (brown)]. (f) Pore size distribution and (g) TEM images for TpPa-(OH)2, TpBD-(OH)2, TpPa-(OMe)2, and TpBD-(OMe)2. Reprinted with permission from [442], Copyright 2017, American Chemical Society.
Catalysts 10 00969 g022
Figure 23. Functionalized hybridization of different 2D-materials through diverse strategies to combine various types of materials. Reprinted with permission from [428], Copyright 2019, Wiley-VCH.
Figure 23. Functionalized hybridization of different 2D-materials through diverse strategies to combine various types of materials. Reprinted with permission from [428], Copyright 2019, Wiley-VCH.
Catalysts 10 00969 g023
Figure 24. Crystal structures and the corresponding oxygen intercalation processes of A-site substituted compositions: (a) LCM; (b) LSM. Reprinted with permission from [466], Copyright 2020, Elsevier.
Figure 24. Crystal structures and the corresponding oxygen intercalation processes of A-site substituted compositions: (a) LCM; (b) LSM. Reprinted with permission from [466], Copyright 2020, Elsevier.
Catalysts 10 00969 g024
Table 1. A comparison of the recent metal oxide-based electrode materials performance utilized in various types of SC devices. (Cs: specific capacitance).
Table 1. A comparison of the recent metal oxide-based electrode materials performance utilized in various types of SC devices. (Cs: specific capacitance).
NanocompositeCs of the Electrode Material at Current Density (A·g−1) or at Scan Rate (mV·s−1)Cs of the SC at Current Density (A·g−1) or at Scan Rate (mV·s−1)Ed of the SCPd of the SCCapacitance Retention (%)/Number of Cycles/Current Density (A·g−1) or at Scan Rate (mV·s−1)ElectrolyteType of the SC DeviceRef.
MnO2/GPCNs438 F·g−1 at 0.5 A·g−172.4 F·g−150.2 Wh·kg−1516 W·kg−167.8/-/50 A·g−1 for the electrode, and 77.8/-/0.5 to 50 A·g−1 for the deviceAqueous 1.0 M Na2SO4ASSC (MnO2/GPCN-SS//GPCN-SS)[296]
Mn3O4@NiCo2O4@NiO1905 F·g−1 at 1 A·g−1216 F·g−1 at 1 A·g−176.8 Wh·kg−1800 W·kg−192/10,000Aqueous 2 M KOHASSC (MNN//AC)[297]
MnO2-coated graphene paper385.2 F·g−1 at 1 mV·s−176.8 mF·cm−2 at 0.05 mA·cm−26.14 mWh·cm−236 mW·cm−282.2/5000/-Aqueous 0.1 M Na2SO4Flexible solid-state ASSC (MnO2-coated graphene paper//graphene paper)[298]
MnO2/CNTs1980 Fm−2 at 10 A.m−2-177 Wh.m−2250 W.m−287/3000/10 A.m−2Aqueous 1.0 M Na2SO4SSC[299]
CNTs/MnO21229 F·g−1 at 1 A·g−1152 F·g−1 at 0.3 1 A·g−184.6 Wh·kg−1190 W·kg−194.4/10,000/- for the electrode, and 91/10,000/- for the deviceAqueous 1 M Na2SO4ASSC (CNTs/MnO2//AC)[300]
Mno2/CNTs-603.54 Fcm−2 at 1 mA·cm−2-93.8 μWh.cm−2193 μW.cm−2>98/1000/- for the electrodePolymer gel PVA-LiCl (6 M)SSC[301]
MnO2/CNTs201 F·g−1 at 20 A·g−1-13.3 Wh·kg−1600 W·kg−170/10,000/1 A·g−1 for the electrodeAqueous 1 M Na2SO4ASSC (AC//MnO2/CNTs)[302]
MnO2-AgCNT-CCup to 325 F·g−1 at 1 A·g−1----Aqueous 0.5 M Na2SO4FSC[303]
CoMCNFs@MnO2265 F·g−1 at 0.5 A·g−166 F·g−1 at 0.3 A·g−119.27 Wh·kg−1217.51 W·kg−198.7/10,000/for the electrode, and 94.9/10,000/for the deviceAqueous 3 M KOHASSC (CoMCNFs@MnO2//CNFs)[304]
Mn3O4/CNFs300.7 F·g−1 at 5 mV·s−1-79.4 mWh·cm−2-100/7500/-Aqueous 1 M Na2SO4SSC[305]
Mn3O4/Ni(OH)2742 F·g−1 at 1 A·g−143 F·g−1 at 0.2 A·g−115.3 Wh·kg−1168.8 W·kg−1-Aqueous 1 M KOHASSC (Mn3O4/Ni(OH)2//AC)[306]
Porous graphene@Mn3O4208.3 F·g−160.1 F·g−1 at 0.5 A·g−130.1 Wh·kg−1475 W·kg−186/2000/-Aqueous 1 M Na2SO4ASSC (porous graphene@Mn3O4//AG)[307]
Zn2SnO4/SnO2/CNT702 F·g−1 at 1 A·g−1-98 Wh·kg−11000 W·kg−193/15,000Aqueous 6 M KOHSSC[308]
CuO@NiO/PANI/MWCNT1372 F·g−1 at 5 mV·s−1---83/1500/-Aqueous 3 M NaOH-[309]
NiO/PANI/CNT356 F·g−1 at 5 mV·s−1---93.4/-/-Aqueous 0.1 M, 10 mL KCl in acetonitrile-[310]
ZnO/MnOx556 F·g−1 at 1 A·g−1-16 Wh·kg−1225 W·kg−196/10,000/-Aqueous 1 M Na2SO4SSC[311]
RuO2-CNFs188 F·g−1 at 1 mA·cm−2---93/3000/-for the electrodeAqueous 6 M KOH-[312]
RuO2−ITO nanopillars1235 F·g−1 at 50 mV·s−1---75/4000/-for the electrodeAqueous 0.5 M H2SO4-[313]
Co3O4@RuO2 nanosheet-3.42 F·cm−31.21 mWh·cm−313.29-Aqueous 3 M KOHASSC (Co3O4@RuO2 nanosheet/Co9S8)[314]
Co3O4@RuO2 nanosheet-4.28 F·cm−31.44 mWh·cm−30.89 W.cm−3-Solid-state PVA/KOH gelASSC (Co3O4@RuO2 nanosheet/Co9S8)[314]
carbon/Fe3O4536 F·g−1 at 3 A·g−1368 F·g−1 at 1 A·g−142 Wh·kg−114.5 k W·kg−177/20,000/20 A·g−1Aqueous 3.5 M KOHSSC[315]
Fe2O3/GNSs/CNTs675.7 F·g−1 at 1 A·g−1169.5 F·g−1 at 0.5 A·g−160.3 Wh·kg−1-82.2/5000/5 A·g−1Aqueous 6 M KOHAll-solid-state ASSC[316]
PANI/rGO/Fe2O3 @ CC1124 F·g−1 at 0.25 A·g−1---82/10,000/-Aqueous 1 M H2SO4-[317]
CNT−Fe3O4373 F·g−1 at 10 mV·s−1135.2 F·g−1 at 10 mV·s−137 Wh·kg−110 k W·kg−1100/15,000/-Aqueous 1 M Na2SO4HASSC (CNT−Fe3O4//CNT−Mn3O4)[318]
CNT−Mn3O4453 F·g−1 at 10 mV·s−1135.2 F·g−1 at 10 mV·s−137 Wh·kg−110 k W·kg−1100/15,000/-Aqueous 1 M Na2SO4HASSC (CNT−Fe3O4//CNT−Mn3O4)[276]
(Fe,Cr)2O3 oxide layer45.92 mF·cm−2 at 5 mV·s−116.88 mF·cm−2 at 5 mV·s−10.57 mWh·cm−2200 mW·cm−297/10 000/-for the electrode, and 90/10 000/-for the devicePolymer gel LiClO4−PCFlexible solid-state SSC[306]
WS2@NiCo2O4/CC2449.9 mF·cm−2 at 1 mA·cm−2-45.67 Wh·kg−1992.83 W·kg−185.59/5000/2 mA·cm−2Aqueous 3 M KOHFlexible solid-state SSC[319]
CoO/NiO−Cu@CuO2035 mF·cm−2 at 2 mA·cm−2-69.3 Wh·kg−11080 W·kg−190.3/5000/Solid-state PVA/KOH hydrogelSolid-state aASSC[320]
NiO@Nanofiber-based composite-161 at 1.0 A·g−122 Wh·kg−1491.9 W·kg−1∼96.0/5000/10 A·g−1Li gelFlexible transparent SSC[321]
Ni(OH)2@CCF502.00 at 1 A·g−1131.43 F·g−1 at 0.25 A·g−135.78 Wh·kg−10.35 k W·kg−1102.61 even at a bending angle of 180°Aqueous 6 M KOHFlexible and wearable SSC[322]
Ni-Mn-oxides1215.5 F·g−1 at 2 mV·s−1196.5 F·g−1 at 1.5 A·g−1132.3 Wh·kg−11651 W·kg−197/3000/-Aqueous 1 M Na2SO4ASSC (Ni-Mn-oxide//AC)[323]
V2O5/Pt/conductive printing paper343 F·g−1 at 10 mV·s−1160 F·g−1355 Wh·kg−10.2 kW·kg−191/5000/-Urea−LiClO4−PVA (ULP)Wearable SSC pseudocapacitor[324]
VO2/graphene@NiS2 hybrid aerogel1280.0 F·g−1 at 1 A·g−1222.2 F·g−1 at 2 mV·s−160.2 Wh·kg−1350.0 W·kg−186.2/10,000/-PVA/KOH gelAll-solid state ASSC (VO2/graphene@NiS2 hybrid aerogel//graphene aerogel)[325]
Gradient oxygen-vacant V2O5/PEDOT-614 F·g−1 at 0.5 A·g−185 Wh·kg−1250 W·kg−1122/50 000Aqueous 1 M Na2SO4SSC[326]
Table 2. A comparison of various 2D nanomaterials utilized as the electrode in SC devices.
Table 2. A comparison of various 2D nanomaterials utilized as the electrode in SC devices.
2D NanomaterialType of the Corresponding Composite-Based Electrode MaterialCharge Storage Behavior of the Resultant Composite-Based Electrode MaterialCs of the Electrode Material at Current Density (A·g−1) or at Scan Rate (mV·s−1)Capacitance Retention (%)/Number of CyclesRef.
MXeneCoF/MXeneThe faradaic charge storage mechanism of CoF NPs and the EDLC behaviour of MXene flakes resulted in a battery-like charge−discharge behaviour for the resultant composite.1268.75 F·g−1 at 1 A·g−197/5000[469]
BNTernary nanohybrids (mK-BN/CNT/PANI)The combination of EDLC and pseudocapacitive behaviour. (Pure BN and modified BN exhibited EDLC behaviour mainly. mK-BN showed the maximum pseudocapacitive involvement (28%) with respect to BN and mH-BN. The maximum pseudocapacitive contribution has been reported for the ternary composites.)515 F·g−198/10,000[455]
h-BN(h-BN)/rGOA prominent pseudo nature combined with EDLC was observed due to the EDLC behaviour of graphene and the intrinsic pseudocapacitive nature of h-BN heteroatoms.∼824 F·g−1 at 4 A·g−1100/4000[454]
BPBP thin filmEDLC behaviour 94.3/50,000[470]
TMDsMoS2@HCNBsThe domination of EDLC with slight faradaic pseudocapacitance, which arises from the faradaic charge transfer Mo in MoS2 structure560 F·g−1 at 0.2 A·g−194.4/5000[471]
PSKMesoporous CoTiO3 PSKPseudocapacitive behavior of the CoTiO3 microrods608.4 F·g−1 at 5 mV·s−182.3/1950[472]
LDHsNiCo-LDHsThe EDLC behaviour and faradic behaviour contribution1272 C·g−1 at 2 A·g−1103.9/5000
COFspyridyl-hydroxyl functionalized COFsThe domination of EDLC due to the high surface area of COFs (through the formation of the EDLC between the COF surface and the H+ ions) along with a pseudocapacitance, which increased by increasing the number of OH groups546 F·g−1 at 0.5 A·g−1100/10,000[473]
MOFsNiCo-MOF@PNTsThe pseudocapacitive behaviour, which was derived from the faradaic reactions of Ni2+/Ni3+ and Co2+/Co3+, as well as an extra pseudocapacitance due to the presence of the PNTs.1109 F·g−1 at 0.5 A·g−179.1/10,000[474]

Share and Cite

MDPI and ACS Style

Forouzandeh, P.; Kumaravel, V.; Pillai, S.C. Electrode Materials for Supercapacitors: A Review of Recent Advances. Catalysts 2020, 10, 969. https://doi.org/10.3390/catal10090969

AMA Style

Forouzandeh P, Kumaravel V, Pillai SC. Electrode Materials for Supercapacitors: A Review of Recent Advances. Catalysts. 2020; 10(9):969. https://doi.org/10.3390/catal10090969

Chicago/Turabian Style

Forouzandeh, Parnia, Vignesh Kumaravel, and Suresh C. Pillai. 2020. "Electrode Materials for Supercapacitors: A Review of Recent Advances" Catalysts 10, no. 9: 969. https://doi.org/10.3390/catal10090969

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