Next Article in Journal
Numerical Investigation of the Thermal Performance of Air-Cooling System for a Lithium-Ion Battery Module Combined with Epoxy Resin Boards
Previous Article in Journal
NaBH4-Poly(Ethylene Oxide) Composite Electrolyte for All-Solid-State Na-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Capacitive Materials for On-Chip Electronics Energy Storage Technologies

by
Bukola Jolayemi
1,2,
Gaetan Buvat
1,2,3,
Pascal Roussel
3,* and
Christophe Lethien
1,2,4,*
1
Institut d’Electronique, de Microélectronique et de Nanotechnologies, Université de Lille, CNRS, Université Polytechnique Hauts-de-France, UMR 8520-IEMN, F-59000 Lille, France
2
Réseau sur le Stockage Electrochimique de l’Energie (RS2E), CNRS FR 3459, 33 rue Saint Leu, 80039 Amiens CEDEX, France
3
Unité de Catalyse et de Chimie du Solide (UCCS), Université de Lille, CNRS, Centrale Lille, Université d’Artois, UMR 8181–UCCS, F-59000 Lille, France
4
Institut Universitaire de France (IUF), 92073 Paris, France
*
Authors to whom correspondence should be addressed.
Batteries 2024, 10(9), 317; https://doi.org/10.3390/batteries10090317
Submission received: 22 July 2024 / Revised: 20 August 2024 / Accepted: 23 August 2024 / Published: 7 September 2024
(This article belongs to the Section Supercapacitors)

Abstract

:
Miniaturized energy storage devices, such as electrostatic nanocapacitors and electrochemical micro-supercapacitors (MSCs), are important components in on-chip energy supply systems, facilitating the development of autonomous microelectronic devices with enhanced performance and efficiency. The performance of the on-chip energy storage devices heavily relies on the electrode materials, necessitating continuous advancements in material design and synthesis. This review provides an overview of recent developments in electrode materials for on-chip MSCs and electrostatic (micro-/nano-) capacitors, focusing on enhancing energy density, power density, and device stability. The review begins by discussing the fundamental requirements for electrode materials in MSCs, including high specific surface area, good conductivity, and excellent electrochemical stability. Subsequently, various categories of electrode materials are evaluated in terms of their charge storage mechanisms, electrochemical performance, and compatibility with on-chip fabrication processes. Furthermore, recent strategies to enhance the performance of electrode materials are discussed, including nanostructuring, doping, heteroatom incorporation, hybridization with other capacitive materials, and electrode configurations.

1. Introduction

The global consumer markets of artificial intelligence (AI) and internet of things (IoT) technology are expanding rapidly, thanks to wireless microsensor networks, wearable and implantable microelectronics, and other smart devices. These technologies find applications in home automation, medical control, space and defense surveillance, agriculture, environmental monitoring, unmanned aerial vehicles (UAVs), and industrial process tracking. However, the full autonomy of these wirelessly interconnected microelectronics is constrained by self-powering mechanisms, a major drawback of the technology. To overcome this limitation, energy scavenging/harvesting techniques such as solar cells [1,2] and nanogenerators [3,4] are used to harness and utilize ambient energy sources [5]. Intermittent and unstable energy generation sources necessitate complementary micro-energy storage technologies to capture the energy generated and deliver it on demand to mitigate discontinuity, periodicity, and indeterminacy [3,4,6,7,8,9]. Presently, the IoT industries are constrained by on-chip energy supply, which hinders their ability to meet the fast-growing consumer markets. Therefore, there is an urgent need for the expeditious implementation and improvement of on-chip electronics power sources to ensure self-powered and maintenance-free systems with minimal human intervention. By addressing these energy constraints through careful design, optimization, and management strategies, IoT industries can create devices that deliver the desired functionality while maximizing energy efficiency and minimizing operational costs.
In recent years, microelectronic systems have continued to evolve towards ultra-miniaturized, flexible, aesthetic, versatile, and integrable devices in ultra-low form factors; they require commensurate features with on-chip electronics energy storage/supply modules [8,10]. The ongoing efforts to meet the energy requirements are enabled by technological advancements in system-on-chip (SoC) and system-in-package (SiP). SoC integrates multiple components of an electronic system into a single millimeter-sized silicon chip, and SiP incorporates various SoC components, such as sensors, data management units, radiofrequency transceivers, and energy supply units, into a single module to optimize space utilization. These technologies demand stringent requirements on the form factor of the on-chip energy supply unit, leading to the miniaturization of traditional energy storage systems like electrostatic capacitors, electrochemical capacitors or supercapacitors, and Li-ion batteries.
Miniaturization of classical energy storage systems through advanced microfabrication technologies has been remarkably prioritized in the expedition to meet the on-chip energy needs for IoT applications. The on-chip energy storage devices are miniaturized power sources integrated directly and seamlessly onto semiconductor chips to provide localized power supply solutions for various electronic devices, including IoT devices. They are designed to be compact with a footprint close to several square millimeters and the footprint of the electrode in the range, depending on the power/energy storage capacity and form factor requirements [1,6,9,11,12]. Considering the ultra-low form factors, downsizing the components of the energy storage devices, particularly the active materials (i.e., electrode materials/dielectrics), will inevitably jeopardize the energy and power deliveries since the performance of the device strongly relies on these materials [6,13,14,15,16,17,18]. As such, the energy and power densities are critically dependent on the electrode materials or dielectric properties, such as specific capacity/capacitance, rate capability, cycling stability, dielectric constant, voltage breakdown (in the case of electrostatic capacitors), and also the topology or configuration of the device [6,18].
The recent cutting-edge on-chip energy storage microsystems technologies have been focusing on engineering and developing new functional materials, innovative electrode design, and advanced microfabrication processes to concomitantly enhance the energy and power densities, aiming to improve the system performance, efficiency, and reliability in order to surmount the constraints of on-chip energy supply restricting the IoT industries to meet the fast-growing consumer markets. Miniaturized electrostatic capacitors (micro-capacitors), micro-supercapacitors (MSCs), and microbatteries (MBs) have drawn considerable attention as on-chip electronics energy storage systems. Recently, several authors have conducted reviews on emerging functional materials such as anode, cathode, and solid electrolyte materials used in MBs and provided state-of-the-art technologies [19,20,21,22,23,24,25,26,27,28,29,30,31,32]. The present review focuses on emerging (pseudo)capacitive materials used in electrostatic (micro-) capacitors and MSCs.
Electrostatic Metal/Insulator/Metal (MIM) (nano-/micro-) capacitors in silicon technology are essential components of any power distribution network in microelectronics; not only do they provide power to the microdevices, but they also subdue voltage instabilities [10]. They are known for delivering ultrahigh power density and ultra-fast charge/discharge rates, making them suitable for many high-power microelectronic systems, including DC/AC converters [33] and dynamic random-access memory (DRAM) circuits [34,35,36,37]. Apart from correcting the power factor, they are widely utilized in bandpass filters [38,39], and they can also perform specific functions, such as timing and decoupling. Their aim is to respond instantaneously to current peaks in the electrical circuit to which they are connected when used as decoupling elements [40]. They, therefore, reduce variations in the supply voltage, providing (very quickly) the amount of current needed to keep the voltage constant. They can be directly integrated via (i) conventional (2D heterogeneous integration of chips next to each other), (ii) on silicon interposers (2.5D integration, coupling the concept of side-by-side and 3D chip integration), or (iii) in 3D integration (vertical stacking) for optimal integration densification. However, the energy density is extremely low compared with MBs and MSCs, as revealed in the Ragone plot in Figure 1; this impedes its use in many applications compared to MSCs and MBs microdevices [36]. Nevertheless, the advantages of the micro-capacitors could be leveraged when coupled with other energy storage micro-devices, especially the MBs, to improve the overall performance and efficiency of the energy storage and delivery system; this configuration could enable ultrafast energy storage and release, commonly desirable in applications requiring high pulse discharge or high peak power, which MBs alone might not efficiently offer due to their typically much slower charge/discharge rate, though maintaining a steady energy supply [10]. On the other hand, micro-capacitors could serve as a stabilizer for MB’s output (power supply) in applications that require consistent power delivery to smooth out fluctuations in voltage and current. Coupled with a Li-ion battery in a smartphone or a Li-ion micro-battery within a miniaturized connected object, a 30% gain in (micro-) battery life is expected [41]. Note that surface-mount capacitors (SMDs) are too bulky to cope with this need for miniaturization.
On the other hand, MSCs are miniaturized electrochemical capacitors that operate on the same principles as traditional supercapacitors; they offer higher capacitance and energy density but lower power density, rapid charge/discharge rates, and a longer cycle life compared with electrostatic (micro-/nano-) capacitors [42,43]. MSCs are a bridge between devices operating with short charge–discharge times (<0.1 s) and those with longer discharge times (>0.1 h) under and over which electrostatic (micro-/nano-) capacitors and microbatteries, respectively, find their niche [44]. The Ragone plot relates the energy density (E) and power density (P) for the micro-energy storage devices, using the numerical values to allow a performance comparison exhibited by different devices. The performance requirement of MSC depends on the intended applications. Nevertheless, it is important to take into account the collective fabrication of MSC at the wafer level and subsequently process compatibility with micro-/nanoelectromechanical systems (MEMS/NEMS) technology instead of lab prototyping in order to facilitate the technological readiness level of such MSC [6].
The material selection process is crucial to enhancing energy storage performance, especially for on-chip integration, considering factors such as energy density, power requirements, form factor requirements, and microfabrication techniques. Additionally, the properties of the functional capacitive materials (e.g., specific capacitance) are the main drivers of the energy and power densities of a capacitor (whether electrostatic or electrochemical). The energy stored (E) in a capacitor and the rate at which it delivers the energy (i.e., charge/discharge rates; power density) are important performance metrics; the energy is expressed as a function of both the capacitance (C) and the cell voltage (V) (Equation (1)) [40,45].
E = 1 2 C V 2
Another important performance metric to evaluate the charge/discharge rates of energy storage devices is the time constant (τ), which is the time required for the capacitor to charge/discharge in response to a step change in voltage [46,47]. The time constant is highly influenced by the material properties and is often determined to identify the materials that are capable of retaining high capacity at high charge/discharge rates by providing insights into the device’s response time. It is calculated as the product of the equivalent series resistance (ESR) and capacitance (C), as expressed in Equation (2).
τ = R ESR · C
Materials that are capable of retaining high capacity at high charge/discharge rates typically exhibit low ESR and high capacitance [48]. Low ESR minimizes energy losses and heat generation during charge/discharge cycles, while high capacitance gives rise to storing more amounts of energy. The relationship between the various devices’ energy and time constants is shown in Figure 1b. However, the overall performance of MSCs, for instance, is influenced by two main factors [47]—the choice of the active electrode materials, which define the capacitance of the device, and the electrolyte, which determines the operational voltage. When evaluating electrode materials (and/or electrolytes) for MSCs, it is essential to consider their properties, such as electrical conductivity, surface area, pore structure, and ion transport kinetics [47,48].
Materials with high electrical conductivity and a large surface area can facilitate rapid electron/ion transfer and, in return, faster charge/discharge rates. Also, materials with efficient ion transport properties enable swift diffusion of ions within the electrode/electrolyte interface, which enhances the device’s overall performance. The trade-off between the energy and the time constant (τ = RC) of the device is essential for practical applications; to achieve a significant increase in power capability, for instance, a substantial energy reduction may be necessary as a sacrifice in order to get a commensurate reduction in the time constant [47,48,49,50,51,52,53].
Moreover, realizing miniaturized on-chip electronic energy storage devices requires meticulous functional material design and deposition techniques compatible with advanced microfabrication processes. Most electrode preparation methods employed in traditional macro-energy storage devices cannot satisfy the requirements imposed by the miniaturization process. In addition, many of the recent electrode preparation techniques for the development of flexible microdevices for printable electronics largely rely on wet processing routes by using colloidal solutions or suspensions of particles, which are not fully compatible with semiconductor device manufacturing usually encountered in the electronics industry [54,55,56]. Nevertheless, developing various functional material layers may involve more than one growth mechanism and micro-processing techniques. The commonly used electrode film deposition and device microfabrication techniques are shown in Figure 2. The microfabrication process may further be classified as top-down etching techniques, bottom-up printing techniques; and emerging microfabrication techniques, the details on these techniques can be found in the literature [6,11,57,58]. Thin film deposition of functional materials compatible with MEMS/NEMS technology is prioritized to speed up the technological readiness level of efficient on-chip micro-energy storage devices by collective fabrication at the wafer level. Vacuum deposition techniques, such as physical vapor deposition (PVD) and atomic layer deposition (ALD), are favored for this task and are commonly employed in the electronics industry. Although other techniques, especially solution-based routes such as electrophoretic deposition/electrodeposition (Figure 2), have been used to produce high-performance porous electrode materials or high-k dielectric materials for on-chip energy storage applications—such deposition techniques are instead suitable for lab prototyping [6,59].
It is also essential to consider the selection of a suitable substrate for film development to accommodate different electrode microfabrication techniques. The substrate provides mechanical support to the functional materials, facilitating electron transport and influencing the device’s performance in certain extreme conditions. Substrates are usually classified as rigid (Si, Si/SiO2, glass, and quartz) and flexible (polyethylene terephthalate, polyimide, polyvinyl alcohol, polydimethylsiloxane, paper, and textile) [62,63]. However, silicon wafers are highly versatile substrates compatible with various semiconductor fabrication processes and technologies. They provide a reliable foundation for implementing various semiconductor technologies and are widely adopted across different segments of the electronics industry. Their uniformity, purity, scalability, and compatibility with semiconductor manufacturing processes make them indispensable in producing advanced electronic components that power modern technology. Several reviews have been conducted on on-chip energy storage devices, focusing more on design, fabrication, integration, applications, and performance metrics. This review provides an overview of capacitive materials for on-chip electrostatic (micro-/nano-) capacitors and electrochemical MSCs, with more emphasis on recent electrode developments, challenges, and future directions. Emphasis is given to the electrode materials, and the preparations are largely compatible with MEMS technology.

2. Materials for Electrostatic (Nano-/Micro-) Capacitors

Electrostatic capacitors are energy storage devices typically consisting of two closely spaced parallel conducting electrodes separated by an insulating material known as a dielectric (see Figure 3) [64]. The configuration is designated as metal–insulator–metal (MIM). The device obeys Coulomb’s law when a voltage (V) bias is applied to the electrodes; it initiates a polarization of the dielectric by the applied electric field, resulting in a charge accumulation on opposing electrode surfaces; this is known as the electrostatic charge storage process [65,66,67]. Capacitance (C) is the most important performance metric for a capacitor, as it determines its capability to store charge.
The ratio of the total accumulated charge (Q) and the electric potential (V) is used to express capacitance, which can be calculated using the following formula:
C = Q V
Q = C · V
The ability of a dielectric material to hold or store charge is an important intrinsic property known as dielectric permittivity (ε). However, in the absence of the dielectric, the space between the electrodes becomes empty; the ε is replaced by the permittivity of the free space or vacuum, ε 0 , a constant that is approximately equal to 8.854 × 10−12 Fm−1. The capacitance further defines the relationship between the surface area (S) of the electrodes, the distance (d) between the electrodes, and the dielectric permittivity (ε), as given in Equation (5). In order to measure how effectively the dielectric can store energy in an electric field relative to a vacuum, another parameter, known as relative permittivity or dielectric constant, κ , is introduced, as shown in Equation (6).
C = ε S / d
C = κ ε 0 S / d
where κ = ε r = ε ε 0 .
A distinctive feature of electrostatic capacitors is their high power due to their rapid charge/discharge rates, with the rates limited only by external circuit RCs. However, energy storage is limited as only a flat surface or parallel plate electrode configuration is utilized, unlike MSCs, where charges are stored in electric double layers or through fast redox reactions in porous electrodes, allowing larger energy density storage on the porous electrode surfaces [67]. The energy (E) is an important performance metric, which is a function of both the capacitance (C) and the voltage (V), i.e., E = 1 2 CV 2 (Equation (1)), derived from dE = VdQ [40]. Materials with a higher dielectric constant (known as high-k) and/or increased surface area of the electrodes can enhance charge storage, thus increasing the energy density. Applying a higher voltage can also increase the energy while taking the breakdown voltage of the dielectric material into account. However, it is noteworthy that other factors like polarization and applied electric field also play a significant role in determining the capacitor’s energy performance, which is detailed elsewhere [40,68,69].
Miniaturization of the capacitors can compromise their performance/efficiency as the size is significantly reduced along the footprint of the electrodes, leading to capacitance loss [70] To compensate for the loss of capacitance, the thickness of the dielectric films (d) needs to be reduced. Having an extremely thin dielectric can result in leakage current from direct electronic tunneling through the dielectric, leading to a complete breakdown of the dielectric when the leakage current becomes too high at a certain electric field Ebd [70,71]. The voltage at which the dielectric materials experience breakdown is known as breakdown voltage (Vbd) and is expressed as:
V bd = E bd · d
E bd = V bd d
where Ebd is known as the dielectric breakdown strength (DBS).
There is a trade-off between dielectric constant (k) and breakdown voltage (Ebd) when achieving high capacitance and voltage due to thickness, as expressed below [71,72]
E bd   orDBS ~ 20 κ   MV   cm 1
Hence, electrostatic capacitors are typically metal–insulator–metal (MIM) and designed to be integrated directly onto silicon substrates alongside other electronic components, enabling highly integrated systems-on-chip (SoCs). Physical vapor deposition (PVD) techniques such as sputtering deposition are popular in the electronics industry and can be used to deposit dielectric and electrode materials on top of the silicon substrate, but the issues of film conformality and uniformity limit these deposition techniques for 3D silicon-integrated capacitors [73,74,75]. A 3D nanostructured electrode configuration is introduced to achieve a higher surface area and, thus, a higher capacitance value. Hence, realizing extremely high aspect ratio structures such as deep reactive ion etched (DRIE) silicon [64] and anodized aluminum oxide (AAO) [67], electrostatic silicon-integrated capacitors are typically fabricated using semiconductor microfabrication techniques, including photolithography, etching, and deposition processes, particularly atomic layer deposition (ALD) (Figure 1). These microfabrication techniques ensure precise control over capacitor dimensions, dielectric materials, and electrode configurations.
The ALD technique is usually preferred due to its unique features for fabricating ultrathin films, nano-laminating layers, high degree of thickness control, and conformal surface coatings in high-aspect-ratio trenches [67,70]. The ALD technique has shown significant improvements in the areal capacitance of nearly 1 μF mm−2 [64,67,71]. However, some dielectric films prepared by ALD exhibit a low dielectric constant due to the low deposition temperature, inhibiting a high degree of crystallization [70,76]. Extensive screening and exploration of various high-k dielectric materials to increase the capacitance and energy density of silicon-integrated capacitors while maintaining a good breakdown voltage are on the radar of the researchers. The commonly used dielectrics are shown in Figure 4, featuring their bandgap versus the dielectric constant. It is essential that the k-value be high, over 12 or preferably 25–35, while the high bandgap is required to suppress leakage current [77,78]. There is a critical trade-off relationship between high-k dielectrics and bandgap that needs to be taken into account; when the dielectric constant is increased, the band gap is decreased [78]. As the dielectric constant increased, the conduction band offsets (i.e., the barrier height) between the metal electrode and dielectric decreased, leading to a decrease in dielectric strength [78]. The phase of the dielectric materials with a uniform crystal structure is expected to remain unchanged in the temperature range of operation to prevent grain boundaries and crystal dislocations favoring carrier recombination and potential leakage paths [70].
The summary of the emerging dielectric materials is given in Table 1. SiO2 deposited on Si wafers is the most commonly used substrate, suitable for on-chip capacitors; however, other types of substrates have been reported. The same electrodes are traditionally used for both bottom and top electrodes in a symmetric configuration, even though some may prefer to use different materials with high work functions to suppress leakage current. Nonetheless, the performance of micro-capacitors mainly depends on the dielectric material [79]. The electrode materials commonly used are usually the participating metals in the oxide dielectric compounds; other electrode materials with higher work functions, including Ni, Al, Pt, Ag, Au, Cu, TiN, and TaN, may be used to minimize the leakage current [79,80,81,82].

2.1. Binary Oxide Dielectric Materials

Al2O3, Si3N4 (nitride dielectric), and SiO2 exhibit good dielectric breakdown strength but suffer low dielectric constants. Efforts have been made to explore several high-k dielectric materials, such as HfO2 [83], Sm2O3 [84], Er2O3 [85], Ta2O5 [82,86], ZrO2 [87,88,89], and La2O3 [90,91]. However, the low breakdown field of these high-k insulators is a major concern for capacitor reliability [92]. To address the low breakdown field and improve leakage performance, binary oxide dielectrics are usually stacked. Examples are the high-k nanolaminated MIM capacitors [93,94,95]. The preparation of the stacked oxide dielectrics is usually carried out via vacuum deposition techniques such as sputtering, PLD, and ALD. Yu et al. [96] fabricated planar MIM capacitors with HfO2-Al2O3 laminate dielectric on SiO2 substrate using the ALD technique, demonstrating a capacitance density of 12.8 fF μm−2 from 10 kHz up to 20 GHz and a low leakage current of 3.2 × 10−8 A cm−2 at 3.3 V. Si3N4/laminated Al2O3–HfO2 was fabricated using back-end-of-line (BEOL) fabrication to demonstrate a capacitance density of 4.2 fFμm−2 [92]. Other stacked dielectrics have also been reported, including ZrO2/Ta2O5/ZrO2 [97], HfO2/Ta2O5/HfO2 [97,98], Al2O3/TiO2/HfO2 [99], and Al2O3/HfO2/Al2O3 [100]. Films have been exploited for the MIM capacitors to increase the capacitance density. In addition, the breakdown field can also be improved through doping, as reported for HfxZr1−xO2 thin films fabricated by the atomic layer deposition (ALD) technique on SiO2/Si substrate [101,102].
Table 1. Emerging dielectrics for micro-capacitors.
Table 1. Emerging dielectrics for micro-capacitors.
DielectricThickness (nm)Preparation TechniqueBreakdown Voltage (MVcm−1)ESD
(Jcm−3)
Reference
Al2O325ALD4.8-[67]
BT-BMZ230Sputtering7.987.26[73]
BSMT280Sputtering391[75]
BZT130Sputtering4.5640.6[37]
Sm-BFO-BTO650PLD3.5152[103]
BFO25-BTO75-2.5Mn200PLD380[104]
BST–BMN400PLD586[105]
BZT60PLD6.565.1[106]
BZT90PLD7.9489.9[74]
La:HZO10ALD450[102]
Si:HfO29ALD3.340[107]
Si:HZO10ALD4.553[108]
Al:HZO10ALD552[108]
HZO/Hf0.25Zr0.75O21/9ALD671.97[109]
TZT8ALD5.38114.5[110]
HfO2/ZrO22.2/6.6ALD449.9[111]
ESD = electrostatic discharge, PLD = pulsed laser deposition; ALD = atomic layer deposition, BST = barium strontium titanate; BT-BMZ = 0.85BaTiO3-0.15Bi(Mg0.5Zr0.5)O3, BSMT = Ba0.5Sr0.5(Ti0.97Mn0.03)O3, BZT = BaZr0.35Ti0.65O3, HZO = Hf0.5Zr0.5O2, TZT = TiO2/ZrO2/TiO2.
The use of highly regular three-dimensional (3D) architecture has been exploited and demonstrated both high energy and high-power density [67,112]. For instance, Banerjee et al. [67] fabricated 3D arrays of MIM nanocapacitors in self-assembly and self-aligning anodic aluminum oxide nanopores via ALD, employing TiN as the conducting electrodes and Al2O3 as the insulator (Figure 5). It was demonstrated that the conformality for the bottom electrode TiN and Al2O3 in a 20:1 aspect ratio pore was better than 95%. The SEM images show the nanotubular MIM structure, indicating the AAO barrier layer and three layers corresponding to the TiN bottom electrode (BE), Al2O3, and TiN top electrode (TE), and wafer-based production showing capacitors of different areas, with each dot capacitor being 125 µm wide and comprising about 106 nanocapacitors (or nanotubes). The nanocapacitors achieved an enhanced 100 µF cm−2, substantially exceeding previously reported values for nanostructured electrostatic capacitors [67]. In another study, Hourdakis et al. [113] fabricated 3D structured MIM capacitors using photolithography and plasma etching, HfO2 dielectric was deposited via ALD, and the device achieved 3.2 μF cm−2 with stable operation up to a frequency of 105 Hz.

2.2. Ternary Oxide Dielectric Materials

As the efforts to improve the performance of the dielectric capacitors intensified, ternary oxide dielectric materials with perovskite crystal structure, especially class II ceramic capacitor materials such as barium titanate (BaTiO3, BTO), strontium titanate (SrTiO3, STO), bismuth ferrite (BiFeO3, BFO), silver niobate (AgNbO3, ANO) [114], BaZrO3, SrTiO3, SrRuO3 [115], and LaGdO3 [116], are considered to be very promising due to the exceptionally high permittivity (>1000) obtainable in the perovskite crystal structure [78]. The permittivity of the materials can be extensively modified to vary the material composition, which could lead to different polar structures of perovskite dielectrics, resulting in the formation of linear dielectrics, ferroelectrics, relaxor-ferroelectrics, and anti-ferroelectrics [117]. However, the MIM capacitor-based on perovskite dielectrics also suffers from severe leakage current problems due to the trade-off relationship between the dielectric constant and bandgap; the main reason for severe leakage current problems is because of their narrow bandgaps [77,78,117,118,119]. Like binary compounds via dielectric stacking, doping, or elemental substitution [120]. The breakdown voltage of the perovskite materials could be significantly improved when their A-site or B-site are fully or partly substituted or are found in stacked systems [68,79]. Unfortunately, many substitutions are carried out via solution-based processes, which are not suitable for the on-chip microfabrication protocols. Nevertheless, some of the derivatives from the substitution/doping of BTO have been obtained via PLD, including La-doped Ba1−xLax(Zr0.25Ti0.75)O3 (BLZT) [74,121] and Mn-doped BiFeO3–BaTiO3 thin films [105], BiFeO3-BaTiO3-SrTiO3(0.55 − x)BFO−xBTO-0.45STO) were also realized by PLD, exhibiting high energy density with a low leakage voltage [73].

3. Materials for Micro-Supercapacitors

For convenience, it is often necessary to classify electrochemical micro-supercapacitors (MSCs) into two according to the charge storage mechanisms, namely, miniaturized electric double-layer capacitors (EDLCs) and pseudo(micro-) capacitors [6,45]. It is also important to discuss the mechanisms for the charge storage process in MSCs in terms of EDLC and pseudo-capacitance separately. Miniaturized EDLCs store charges principally on the adsorption/desorption (physisorption) of ions on the electrode surface, particularly carbon-based materials, where under an electric field, the charge carriers become polarized and accumulate on the parallel porous electrode surfaces to form electrochemical double layers (EDL); the physisorption is very fast and reversible, and they usually operate in an organic electrolyte or ionic liquid [45,122].
In contrast, the charge storage processes in pseudo(micro-) capacitors utilize non-polarized electrodes (known as pseudocapacitive electrodes) that operate based on chemical potential, undergoing fast reversible Faradic reactions at the surface or subsurface, thereby facilitating high power capabilities [12,66,122]. MSCs are distinctly different from microbatteries (MBs) that exhibit an in-depth and relatively sluggish bulky redox reaction mechanism involving Li-ions shuttling between the cathode and anode; MBs, unlike pseudo(micro-) capacitors, exhibit higher energy density but lower power density, poor rate capability, and a shorter lifespan [25,123,124,125,126,127].
MSCs are known for power delivery, capable of operating at a high rate but relatively low energy density. Since both energy and power densities are important for IoT applications, their energy delivery needs critical improvement. The energy density (E) of the MSCs can be maximized by either increasing the capacitance and/or the working voltage window, since E = 1 2 C · Δ V MSC 2 , where C is the capacitance and ΔVMSC is the cell voltage [45,128,129]. Both C and V are influenced by the nature or properties of the electrode materials and electrolyte [129]. Besides the electrolytes’ influence, increasing the surface area of the electrode films can effectively enhance the capacitance and, thus, energy density. The performance of the MSCs strongly depends on the properties of the electrode materials, such as the crystallographic structure and morphology of the active material (e.g., nanorods, nanowires, nanofeather, and porous nanoflakes), and other factors include electrical contact with current collectors, exposed surface area, amounts of cations, etc. [130,131].
In the meantime, increasing surface capacitance results in a higher time constant, according to this formula, τ = RC MSC , leading to lower power performance, necessitating a trade-off between capacitance and power capabilities [6]. A pertinent strategy consists of lowering the resistance of the electrode materials [53]. In addition, the energy and power densities can also be enhanced through the topology or configuration of the device (the arrangement or geometry of the positive and negative electrodes) [6,132]. The overview of the schematic topologies used in MSCs, including parallel plate, interdigitated, and three dimensional (3D) interdigitated configurations, is shown in Figure 6a. The interdigitated topology is the most suitable configuration for on-chip MSC applications, having both electrodes in the same plane. However, the amount of active material per electrode is reduced by half due to the inactive gap between the two interdigitated electrode fingers. As a result, the surface capacitance decreases drastically to less than one-fourth of the areal capacitance of a single electrode. In order to enhance the energy density of MSCs, 3D interdigitated configurations have been explored. Lethien et al. [6] provide comprehensive details on the different MSC topologies and configurations. Again, the energy storage capacity can be improved through a 3D electrode design that unblocks the “dead surface” of MSC electrodes while maintaining high mass loading within the device’s footprint area [6,132,133].
Further, increasing the cell voltage will also substantially enhance the energy to the tune of four-fold, according to Equation (1) [129]. The width of the safe electrochemical window of the electrode material, which defines the cell voltage, needs to be critically considered. Several electrode materials, especially pseudocapacitive materials, exhibit high capacitance in aqueous electrolytes, but their cell voltage is limited to around 1 V due to the water splitting at 1.23 V. However, the most efficient energy density delivered by these MSCs is around 100 μWh cm−2 [6,128]. Nevertheless, the safe electrochemical window can be extended when two different active materials are used in the negative and positive electrodes to have an asymmetric configuration (as shown in Figure 6b), combining the potential windows of both electrodes (i.e., with complementary electrochemical potential windows) without damaging the materials [6,45]. This is a major departure from the symmetric configuration, where the same (pseudo)capacitive material is used in the two electrodes and the cell voltage is restricted to about 1.2 V in the potential window.
Asymmetric MSCs could exhibit up to 2 V cell voltage in aqueous media [134,135,136,137,138,139,140,141,142,143,144,145]; the cell voltage depends on the positive electrode’s upper potential limit and the negative electrode’s lower potential limit [146]. The overpotentials at the positive electrode (for oxygen evolution) and the negative electrode (for hydrogen evolution) extend the cell voltage windows beyond the thermodynamic limit of 1.23 V of the water decomposition voltage [146]. An example of asymmetric configuration was demonstrated by Asbani et al. [145] (Figure 6c), combining sputtered vanadium nitride film (VN) and electrodeposited hydrous ruthenium oxide (hRuO2) film to fabricate an asymmetric micro-supercapacitor (A-MSC). Taking advantage of their complementary electrochemical potential windows in 1 M KOH electrolyte to achieve a cell voltage up to 1.15 V and high specific capacitance values for the device (≈100 mF cm−2). The effect of asymmetric configuration is evident in the galvanostatic charge and discharge plots shown in Figure 6c. Concerted efforts have been focused on exploring various active electrode materials, including carbon-based, conducting polymer, and metal-based materials (oxides, nitride, and MXene), to optimize the power and energy delivery of on-chip MSCs. Additionally, there have been attempts to combine battery-type materials with (pseudo) capacitive materials to form hybrid (asymmetric) MSCs, commonly referred to as Li-ion micro-supercapacitors [147,148,149,150]. However, these efforts have only been able to obtain prototyping MSCs [147,148]. An example of this was demonstrated by Zheng et al. [8], who fabricated a prototype hybrid MSCs featuring lithium titanate nanospheres and graphene in an MSC asymmetric configuration. Another area of interest is optimizing the existing solid electrolytes and developing new ones for better performance.
MSC performance is evaluated based on the surface area or volume of the device to normalize capacitance (F), energy (Wh), and power (W). The surface performance metrics are given as mF cm−2, mWhcm−2, and mW cm−2, which are normalized by the operated area of the active electrode [6,9,56,151]. Volumetric performance is derived by normalizing the surface metrics to the thickness of the active layer, giving rise to mF cm−3, µWhcm−3, and mW cm−3 for capacitance, energy, and power, respectively [56,151]. When comparing the performance of MSCs, volumetric performance metrics are appropriate, but it is important in that case not to take into account the thickness of the substrate, which is only used as a mechanical support in most cases with various thicknesses (from 200 µm up to 2 mm as an example). It is useless to divide the surface performance by this substrate thickness.
Figure 6. Representative electrode topologies/configurations for on-chip micro-supercapacitors. (a) Various electrode configurations. Reproduced with permission from [6]. Copyright (2019), Royal Society of Chemistry. (b) MSC based on interdigitated topologies with a symmetric configuration or an asymmetric configuration. Reproduced with permission from [45]. © American Chemical Society. CC BY 4.0. (c) Sketch up of the VN/hRuO2 asymmetric MSC in face-to-face topology operating in 0.5 M H2SO4 and 1 M KOH aqueous electrolytes, and galvanostatic charge and discharge plots of the symmetric and asymmetric MSCs, using hRuO2 as the positive and VN as the negative electrodes, respectively. Reproduced with permission from [145]. Copyright (2021), Elsevier.
Figure 6. Representative electrode topologies/configurations for on-chip micro-supercapacitors. (a) Various electrode configurations. Reproduced with permission from [6]. Copyright (2019), Royal Society of Chemistry. (b) MSC based on interdigitated topologies with a symmetric configuration or an asymmetric configuration. Reproduced with permission from [45]. © American Chemical Society. CC BY 4.0. (c) Sketch up of the VN/hRuO2 asymmetric MSC in face-to-face topology operating in 0.5 M H2SO4 and 1 M KOH aqueous electrolytes, and galvanostatic charge and discharge plots of the symmetric and asymmetric MSCs, using hRuO2 as the positive and VN as the negative electrodes, respectively. Reproduced with permission from [145]. Copyright (2021), Elsevier.
Batteries 10 00317 g006aBatteries 10 00317 g006b
This metric is also used by many researchers to exacerbate the performance of nanometer-thick film electrodes. In that case, the volumetric energy and power densities are exaggerated, and that does not reflect the real performance of the MSC. The most important thing for a microdevice is the footprint surface of the energy storage source, which is constrained by the miniaturization of the electronics. It is thus reasonable to evaluate the performance of the MSC based on the surface metric since the footprint area is a crucial factor to consider in on-chip applications [9,56,151]. The performance of the MSCs could be influenced by the following parameters:
  • MSC topology
  • Capacitance retention of the MSC
  • Material/electrolyte conductivity
  • Material surface area
  • Material/electrolyte ion transport properties.
  • Mass-loading of the active materials
  • The width of the safe electrochemical window

3.1. Electric Double-Layer Capacitor Materials

Electric double-layer capacitor (EDLC) materials are low-cost porous carbon-based electrodes, such as carbon nanotubes [152,153,154], activated carbons [155], carbide-derived carbons (CDC) [44,156,157], onion-like carbon [158], carbon foam [159], and graphene [160]. They exhibit a high specific surface area, good electrical conductivity, and chemical stability. They are usually operated in organic electrolytes, demonstrating true capacitive behavior and excellent chemical stability upon cycling [6,36,44,54,155,156,159,161,162]. However, carbon-based MSCs suffer from low energy density, limiting their practical applications in high energy microdevices [159]. To improve the energy performance, carbon-based materials are usually composited/doped with other (pseudo)capacitive materials. Table 2 provides a summary of the recently researched carbon-based MSCs electrodes for on-chip applications.

3.1.1. Graphene-Based Materials

Graphene-based materials are the most explored carbon-based electrodes for MSCs applications. They exhibit excellent electrical properties due to their high electron mobility (2.5 × 105 cm2 V−1 s−1) at room temperature [163]. They also possess a large theoretical surface area (2630 m2g−1) [43] and have the capability to offer ultrahigh power performance and excellent frequency response [160,164]. Graphene is a single layer of sp2-hybridized carbon atoms tightly bound in a two-dimensional (2D) honeycomb lattice with monoatomic layer thickness [151]; the intrinsic capacitance of single-layer graphene (SLG) of ~21 µF cm−2 sets the upper limit for EDL capacitance for all carbon-based materials [165,166,167]. They are typically synthesized by the chemical vapor deposition (CVD) method under a high temperature and with a catalytic metal substrate (mostly grown on copper or nickel foam), requiring transfer to a final substrate; the processing conditions, restacking due to the van der Waals interaction between platelets [168,169,170], and low throughput limit their widespread application in in-plane (on-chip) MSCs [167,171]. Therefore, other synthesis routes are being explored, including solution-processing techniques (especially the liquid-phase exfoliation method), reduction in graphene oxides (GO) [168], and laser induction.
Reduced GO (rGO) constitutes the majority of the reported graphene-based MSCs electrodes. They are obtained by removing the oxygen-containing functional groups from graphene oxides (GO) through various methods such as laser, thermal, chemical, or ultraviolet irradiation; however, the chemical reduction method is the most widely used [172,173,174] Generally, GO is a highly resistive material; hence, its electrical conductivity needs to be improved to meet the requirements of MSC applications, while on the other hand, rGO is a conducting material that is very close to pure graphene. For instance, Wu et al. [168] used the methane-plasma reduction technique to obtain rGO on Si substrate, followed by oxygen-plasma etching for interdigitated patterning of the film. The MSCs delivered low capacitance values (80.7 μF cm−2/17.9 F cm−3) with power and energy densities of 495 W cm−3 and 2.5 mWh cm−3, respectively, showing how maximizing volumetric performance while surface energy and power densities are very low. However, some of the conversion methods involve wet processing routes that could be detrimental to the microfabrication protocols, thereby impeding their implementation in on-chip MSC applications [170]. Additionally, many of the reduction processes, such as high temperature, plasma, or reductant treatment, are not suitable for scalable and on-chip MSC integrated applications [170].

Exfoliated Graphene

Exfoliated graphene (EG) is typically in powder form and is prepared through a process called liquid-phase exfoliation. To make an ink or slurry, it is dispersed in a solvent along with other additives such as surfactants, stabilizers, binders, and conductive agents. The resulting mixture can be deposited onto substrates or current collectors using microfabrication techniques (see Figure 2b) such as screen printing [175,176], mask-assisted filtration [177,178], inkjet printing [179], electrophoretic deposition [180], spray coating [181,182], or electrodeposition [181,182]. Additionally, EG can be combined with other electroactive materials like transition metal oxides and conducting polymers to create composite materials [151,183,184].
Table 2. EDLC materials for on-chip MSCs.
Table 2. EDLC materials for on-chip MSCs.
Electrode MaterialSubstrateSynthesis/FabricationThicknessElectrolyteCell Voltage/
Pot. Window
Areal Cap.Vol. Cap.Energy DensityPower DensityConfigurationReference
nmVmFcm−2Fcm−3mWhcm−3Wcm−3
GraphenePETCVD/DLW17PVA-H2SO4 hydrogel10.06336.85.11714Intedigitated[170]
GraphenePETCVD/DLW17EMITFSI/Fumed silica ionogel2.50.045147231860Intedigitated[107]
GrapheneSi/SiO2CVD/Lithography5PVA-H2SO4 hydrogel10.066131NANAIntedigitated[185]
GrapheneGlass/NiDrop casting/plasma etching1000PVA-H2SO4 hydrogel0.80.1201.52000Intedigitated[186]
GrapheneSi/SiO2CVD/Plasma etching5PVA-H2SO4 hydrogel10.230742.62000Intedigitated[187]
GrapheneSi/SiO2 and PETSpin coating/Inkjet printing750PSSH 10.79.31.30.6Intedigitated[187]
Graphene (Exfoliated)PET Electrochemical exfoliation/Mask-assisted filtration600PVA-H2SO4 hydrogel1.20.82213.72.74493Intedigitated[188]
Graphene (Exfoliated)PET Electrochemical exfoliation/Mask-assisted filtration600KTFSI-P14TFSI ionogel3.40.549.0314.52.6Intedigitated[188]
GraphenePolyimideLaser scribing25,000H2SO4141.60.450Intedigitated[189]
rGOPolyimideMask-assisted vacuum filtration1020PVA-H3PO4 hydrogel0.80.68456.70.375Intedigitated[190]
rGOPETSpin-coating/lithography15PVA-H2SO4 hydrogel10.0817.92.5495Intedigitated[168]
rGOSi/SiO2Spin coating/Laser scribing7600EMITFSI/Fumed silica ionogel2.52.322.322.1200Intedigitated[160]
rGO-CNTsSi/SiO2/TiAuElectrostatic spray deposition (ESD)/photolithography60003 M KCl15.64.90.6877Intedigitated[164]
Graphene-CNTsPET Wet-Jet Milling Exfoliation/screen printing27,000PVA-H3PO4 hydrogel1.81.320.490.22<1Intedigitated[191]
CNTs 3D Printing27,600PVA-H3PO4 hydrogel14.691.70.123.72Intedigitated[152]
CNTs Si/SiO2Injection/Plasma etching5000PVA-H3PO4 hydrogel0.82.755.50.40.19Intedigitated[153]
CNTsSi/SiO2/FePlasma enhanced CVD PVA–KCl hydrogel0.852.5--Single electrode[192]
CNTs-carbon nanosheet (CN)Si/SiO2CVD/photolithography 100,000PVA-H3PO4 hydrogel11101120.45Intedigitated[193]
Carbon nanowires (CNWs)Si/Si3N4/Cr/Pt Electrodeposition + CVD/direct laser writing12,000PVA-H3PO4-SiWA0.85.74.75NANASingle electrode [194]
Activated carbonSi/SiO2Photolithography/etching process/inkjet printing20,0001 M Et4NBF42.52.1---Intedigitated[155]
TiC-CDCSi/SiO2DC magnetron sputtering + Chlorintion14001 M H2SO40.94935010.12Intedigitated[54]
TiC-CDCSi/SiO2DC magnetron sputtering + Chlorintion41002 M EMI, BF4:CH3CN361.5150351parallel plates [54]
TiC-CDCSi/SiO2DC magnetron sputtering + Chlorintion22002 M EMI, BF4:CH3CN335.21603010parallel plates [54]
TiC-CDCSi/SiO2DC magnetron sputtering + Chlorination50001 M H2SO40.8205410NANASingle electrode [54]
TiC-CDCSi/SiO2DC magnetron sputtering + Chlorination32001 M H2SO40.9112350390.94parallel plates [161]
TiC-CDC Si/SiO2DC magnetron sputtering + Chlorination46001 M KOH1.171152NANASingle electrode [156]
TiC-CDC-Anthraquinone (AQ)Si/SiO2DC magnetron sputtering + Chlorination + Electrochemical grafting46001 M KOH1.144338NANASingle electrode [156]
CVD = Chemical Vapor Deposition; DLW = direct laser writing; PET = polyethylene terephthalate

Laser-Induced Graphene

Laser-induced graphene (LIG) or laser scribed graphene (LSG) has attracted much attention for its application as MSC electrodes since its discovery in 2014 as a 3D porous material prepared from various carbon materials by direct laser writing [151,195]. Its popularity is due to the fact that it can be prepared and patterned into highly porous 3D graphene in a single step without the need for wet chemical processes [151]. The high porosities created, due to the release of gas during the laser scribing process, result in a large surface area, thereby facilitating the electrolyte penetration into the active materials [195].
Some innovative strategies, including photochemical reduction in GO, are being explored to address the scalability challenges [54,196]. Another strategy is the introduction of spacers, such as carbon nanotubes (CNTs), between the graphene sheets to prevent their restacking [164]. Heteroatom doping can effectively tune the material structure, introduce pseudocapacitance, and facilitate electrolyte wettability, consequently enhancing charge storage capacity [151,197,198,199]. Such doping elements include S [199], F [197], P [198], and heteroatoms O/N/S [200]. In addition, surface modification is an effective approach to incorporating high-mass-loaded electroactive materials such as other carbon materials [181], such as other carbon materials layer [181], transition metal oxides/hydroxides [201,202,203], MXene [173], and conductive polymers [204]) to the porous surfaces to enhance the electrochemical performance [167,171,205,206,207].
The use or incorporation of the pseudocapacitive materials with a larger specific capacitance moves the carbon electrode from pure double-layer to pseudocapacitive charge storage, combining their synergistic/complementary properties (pseudocapacitance and EDLC) to offer better MSCs performance. However, this strategy has some drawbacks, as it can lead to reduced power capabilities, a shorter cycle life due to the kinetic limitations of the redox reactions, and low electronic conductivity [54,208].
An example is reported by Lin et al. [209] where interdigitated MSC electrodes based on a porous hierarchic 3D graphene-CNT composite were fabricated using conventional photolithography on a silicon substrate. The graphene-CNT composite was grown by CVD on a patterned Ni (a current collector and catalyst for the composite growth). The scheme of the device is shown in Figure 7a. The figure shows the SEM image of the fabricated MSC and a cross-sectional SEM image of the graphene-CNT structure grown on a Ni catalyst/current collector, respectively. The CV curves with a near-perfect rectangular shape show that the MSC based on the composite electrode (tested in an aqueous electrolyte of 1 M Na2SO4) shows improved performance over the one with only the graphene electrode. The enhanced capacitance was attributed to the ion absorption/desorption in the CNTs. The device achieved specific capacitances up to 2.16 mF cm−2 with a maximum power density of 115 W cm−3 in an aqueous electrolyte; the surface capacitance and volumetric energy density achieved in ionic liquids were 3.93 mF cm−2 and 2.42 mWh cm−3, respectively.
Beidaghi et al. [164] fabricated interdigitated graphene-CNT composite-based MSC, combining electrostatic spray deposition (ESD) and photolithography lift-off methods for the device fabrication. The device was tested in a 3 M KCl aqueous electrolyte, exhibiting a capacitance of 6.1 mF cm−2 at 10 mV s−1, an energy density of 0.68 mWh cm−3 and a power density of 77 W cm−3. Lin et al. [189] obtained patterned 3D networks (porous) graphene films from polyimide (PI) films using a CO2 infrared laser; the sp3-carbon atoms of the PI were photothermally converted to sp2-carbon atoms by pulsed laser irradiation, resulting in LIG (LI-rGO) formation. The interdigitated electrode exhibits a low specific capacitance of >4 mF·cm−2 and power densities of ~9 mW·cm−2. The enhanced capacitance, compared to other carbon-based electrodes, has been attributed to the 3D network of highly conductive graphene, providing easy access for the electrolyte to form a Helmholtz layer [189].
In another study, Gao et al. [203] prepared LSG and LSG-MnO nanocomposite films on PET as interdigital electrodes for MSCs by one-step picosecond laser direct-writing (LDW) in air, having thicknesses of 52 μm and 55 μm, respectively. The performance of the LSG and LSG-MnO electrodes showed a specific capacitance of 191 and 470 mF cm−2, respectively; the CV curves of the electrodes over 0.8 V showed that LSG-MnO exhibits a larger integrated area and greater current response than the LSG electrode, indicating higher capacitance and better electrochemical activity, which is attributed to the synergistic effects of the EDLC of LSG and the pseudocapacitance process of MnO films. Moreover, the maximum area-specific capacitance of LSG-MnO-based MSCs with PVA/H3PO4 gel electrolyte is 55 mF cm−2 and recorded a good capacitance retention of about 96% after 5000 cycles. The device delivers a maximum energy density with PVA/H3PO4 gel electrolyte of 4.89 μWh cm−2 and a maximum power density of 0.72 mW cm−2.

3.1.2. Carbon Nanotubes (CNTs)

Carbon nanotubes (CNTs) are considered promising electrode materials for on-chip MSC applications due to their high surface area, excellent electrical conductivity, mechanical flexibility, and chemical stability [152,153,210,211,212] CNTs are categorized into single-wall CNT (SWCNT) and multi-wall CNT (MWCNT) [154,210]. However, a major issue associated with the use of CNTs is their structural breakdown, which results in decreased electrical conductivity and mechanical strength [213]. This could be addressed by incorporating other functional materials inside their interconnected pores to prevent such structural breakdown [213]. In addition, CNT film preparations are mostly performed through solution routes, limiting their use for on-chip MSC applications. Several studies have explored alternative approaches to fabricating MSCs based on CNT electrodes, especially flexible MSCs [152]. Kim et al. [153] used selective wetting-induced micropatterning for MWCNTs-based MSCs. It has been demonstrated from recent studies that the performance of CNT-based electrode materials is highly dependent on nanotube alignment and packing density [154,210]. Highly aligned CNTs that are either horizontally (HACNTs) or vertically (VACNTs) aligned present good pathways for electron transport, which results in superior performance compared to randomly entangled CNT networks. HACNTs-sheet-based MSCs exhibit better electrochemical performance when compared with VACNT, as the latter exhibits lower areal capacitance and volumetric performance. This is because HACNTs have a long-range alignment interconnected network that facilitates charge transport, leading to better conductivity and a higher packing density, which enhances the energy density [152,210,211].
Cao et al. [192] prepared multiwall CNT-forest by plasma enhanced chemical vapor deposition (PECVD) on silicon wafer (before transferred onto elastomer) to achieve very low capacitance values (5 mF cm−2 and ≈ 2.5 F cm−3) and retain close to 90% of the initial capacitance after 10,000 cycles (see Figure 7b). Pseudocapacitive materials are usually employed to enhance the electrochemical performance of the CNT-based electrodes, combining the synergetic effects of EDLC and pseudocapacitance [214,215]. This was demonstrated by Dousti et al. in 2020 [154], who for the first time fabricated an interdigitated electrode with HACNT using a single oxygen plasma etching process for fabrication. The device with HACNT (~300 nm-thick) films achieved volumetric capacitance, energy, and power densities of 75.7 Fcm−3 (at 5 mV s−1), 10.52 mWhcm−3 and 19.33 Wcm−3, respectively, in 0.1 M Na2SO4, while the performance was significantly improved using the HACNT-MnO2 composite (~300 nm-thick) electrode, demonstrating a volumetric capacitance, energy, and power densities of 242 Fcm−3 (at a scan rate of 5 mV s−1), 33.7 mWhcm−3, and 31 Wcm−3, respectively, with an impressive long cycle life (>90% capacitance retention after 7000 cycles). It was reported that the interdigitated electrode HACNT-MnO2 composite sheet-based MSCs outperformed the existing planar thin-film MSCs at the time and could match the performance of 3D MSCs [154]. Ouldhamadouche et al. [214] prepared hierarchically composite electrodes consisting of porous and nanostructured VN grown on vertically aligned CNTs (VACNTs) for micro-supercapacitor applications. The CNTs were grown on a Si/SiO2 substrate through a distributed electron cyclotron resonance (ECR)–plasma enhanced chemical vapor deposition (PECVD) process, after which a direct deposition of vanadium nitride (VN) films over the as-prepared CNTs arrays was performed by means of reactive DC sputtering. The composite VACNTs-VN electrode demonstrated an areal capacitance of 37.5 mF cm−2 at a scan rate of 2 mV s−1 in a 0.5 M K2SO4 electrolyte solution between −0.1 and 0.4 V potential windows. Pitkänen et al. [216] fabricated a composite VACNTs-MnOx-electrode-based on-chip MSCs. The device achieved a specific areal capacitance of 37 mF cm−2, and an energy density of 6.7 μWh cm−2. Zhang et al. [217] fabricated interdigital 3DTiN-TiO2-VACNT-based MSC using area-selective atomic layer deposition by aerosol jet printing technique. The device achieved the composite TiN-TiO2-VACNT MSC with an areal capacitance of 6.55 mF cm−2. This achieved nearly two orders of magnitude improvement compared to the pure VACNT MSC (0.07 mF cm−2). The presence of the pseudocapacitive materials (TiN-TiO2) significantly improved the energy density (3.28 mWh cm−2) while maintaining a high power density (2.34 mW cm−2).
Figure 7. Carbon-based electrodes (a) Graphene-CNT grown on a Si substrate. Reproduced with permission from [209]. Copyright © 2024, American Chemical Society. (b) Vertically aligned carbon nanotube (VACNT) forests grown on a silicon wafer. Reproduced with permission from [192]. Copyright (2019), John Wiley & Sons. (c) Carbide-derived carbon (CDC) on a Si wafer. Reproduced with permission from [54]. Copyright (2016), American Association for the Advancement of Science.
Figure 7. Carbon-based electrodes (a) Graphene-CNT grown on a Si substrate. Reproduced with permission from [209]. Copyright © 2024, American Chemical Society. (b) Vertically aligned carbon nanotube (VACNT) forests grown on a silicon wafer. Reproduced with permission from [192]. Copyright (2019), John Wiley & Sons. (c) Carbide-derived carbon (CDC) on a Si wafer. Reproduced with permission from [54]. Copyright (2016), American Association for the Advancement of Science.
Batteries 10 00317 g007

3.1.3. Porous Carbide-Derived Carbons

Carbide-derived carbons (CDC) are a type of nanoporous carbon material that is obtained by subjecting a metal carbide, typically titanium carbide (TiC), to chlorination. This process results in the formation of a bilayer of TiC and porous carbon (TiC-CDC). TiC-CDC is a highly promising material for on-chip MSC applications due to its unique combination of high surface area, high capacitance, electrical conductivity, and chemical stability [54,157,161,218,219]. The use of TiC-CDC as electrode materials eliminates the presence of polymer binders between the substrate and the active materials, which is common to other carbon materials, ensuring a naturally strong interface between the current collector and the active material [157]. TiC films are typically deposited on Si wafers by reactive DC magnetron sputtering using a titanium target and acetylene (C2H2) gas as a carbon source. The reactive-sputtered TiC films are subsequently etched by chlorine gas to form CDC films [54,157,161,218,219]. The following equation expresses the reaction process of carbide with chlorine gas [157,220].
TiC s + 2 Cl 2 g TiCl 4 g + C s
A number of studies have fabricated TiC-CDC films as on-chip MSC electrodes at the wafer level [54,161,219]. For instance, Huang et al. [54] achieved MSC based on TiC-CDC electrodes via reactive DC magnetron sputtering followed by chlorination at the wafer level, as shown in Figure 7c. (A) The Si/TiC/CDC electrode with a 5 µm-thick CDC operated in 1 M H2SO4 electrolyte achieved 205 mF cm−2 (410 F cm−3) with no capacitance loss after 10,000 cycles. Also, the MSC recorded 350 mF cm−2 in 1 M H2SO4 electrolyte at 10 mV s−1.

3.2. Pseudocapacitive Electrodes

Pseudocapacitive materials are becoming the principal candidates for simultaneously delivering high energy and power densities as they occupy a middle ground between EDLCs and battery-type materials; they exhibit near-rectangular cyclic voltammetry (CV) profiles and almost linear galvanostatic charge/discharge (GCD) curves similar to EDLC [146,221]. The charge storage processes rely on a charge transfer mechanism instead of physically storing charge by electrostatic charge adsorption. The charge storage mechanism emanates from the electrosorption and/or fast Faradaic reversible redox reactions at the surface or subsurface of the redox-active materials; the surface Faradaic electron transfer is enabled by charge-compensating ions [45,222,223,224]. The Faradaic reaction in pseudocapacitive materials is ultrafast, which differs from battery-type materials where faradaic reactions occur at a constant potential, usually originating from the ideal Nernstian process [222,223]. Pseudocapacitance can arise from one or a combination of the following fast Faradic processes: redox, intercalation, and electrosorption [146,224,225].
From a technological point of view, it is highly desirable to prepare pseudocapacitive electrode films that are compatible with the cleanroom protocol, which is required by on-chip integration, and it is also important to consider microfabrication techniques capable of facilitating scale-up production of electrode materials at the wafer level. There are two main categories of pseudocapacitive materials: conducting polymers such as polyaniline (PANI) [226], poly(3,4-ethylene dioxythiophene) (PEDOT) [227], polypyrrole (PPy) polypyrrole (PPy) [227]), and metal-based pseudocapacitive compounds [228,229,230,231]. Metal-based materials are often preferred for on-chip applications due to their versatility and compatibility with microfabrication processes. Metal compound-based MSCs remain the exploration hotspot for achieving high power and energy densities in a limited footprint area and volume. The emerging metal-based electrode materials for on-chip MSC applications can be classified as follows:
i
Metal oxides/hydroxides
ii
Metalnitrides
iii
2D transition metal carbides/carbonitrides (MXene)
Other metal compounds, including metal dichalcogenides, such as MoS2, exhibit pseudocapacitance properties, high electrical conductivity, and a good hydrophilic nature. However, their implementation for on-chip implementation is restricted due to processing techniques, including hydrothermal [232,233], colloidal [234], chemical bath [235], chemical precipitation and ion-exchange process [140], and successive ionic layers [6,9,11,12,236].

3.2.1. Metal Oxides

Metal oxides are the centerpiece of (micro-)pseudocapacitive electrode material exploration; they exhibit multiple oxidation states, allowing for effective redox charge transfer [150]. Oxides or hydroxides of different metals, such as Ru, Mn, Ni, Co, Fe, Ti, V, Mo, W, and Nb, have shown promise as redox-active materials for MSC electrodes. The electrochemical behavior of these oxides in aqueous electrolytes explains their charge storage mechanisms, involving fast Faradaic and implicit non-Faradaic EDLC processes at the surface/subsurface [146]. The fast Faradaic redox process involves [146,237,238]:
  • valence changes (electron transfer) of the metallic species located near the oxide surface,
  • serving as redox-active centers through cation intercalation/deintercalation, surface adsorption/desorption, or
  • surface redox reactions with anions.
The valence changes in the oxide materials occur at different potential ranges, which are essentially determined by the nature of the electrolytes [237]; in aqueous electrolytes, for instance, the valence changes in Ni2+/Ni3+ and Co2+/Co3+ occur in the ranges of ≈0 and 0.6 V (0–0.6 V vs. Hg/HgO [239]), Mn3+/Mn4+ (≈0–0.8 V vs. Ag/AgCl [228]), and Fe2+/Fe3+ (−1–0 V vs. SCE) [240]. RuO2 and MnO2 are the most representative metal oxide electrode materials and are widely known for their high theoretical capacitance and high energy density capability [241,242]. Several other metal oxides, such as MoO2 [243], Co3O4, NiO [244], MnFe2O4 [236], IrO2 [245], and FeWO4 [128], are currently investigated for on-chip MSCs applications. In addition, the synergistic contributions from pseudocapacitance and EDLC processes are considered to enhance the energy storage capabilities of the MSCs by exploring metal oxide/carbon composites as electrode materials [18,221].
Besides on-chip compatibility issues, the electrochemical performance is also influenced by the preparation techniques. Notably, studies have shown that amorphous structures or hydrous forms of transition binary oxides (e.g., RuO2·nH2O, IrO2·nH2O, and WO3·nH2O) exhibit better pseudocapacitive behavior than those of anhydrous structures or crystalline phases [6,246]. Currently, chemical or solution routes are dominant in the preparation of oxide electrodes. These methods typically require the use of expensive precursor solutions (such as RuCl3·xH2O for RuO2 synthesis) and complex steps to prepare porous nanostructured oxide films with a large surface area [247,248,249]. Solution-phase processes have the potential to offer better capacitance performance compared to vacuum-deposited oxide thin films, such as physical vapor deposition (PVD) techniques (ex: sputtering deposition). However, the chemical routes are typically not fully compatible with microfabrication protocols, especially when interfacing with other electronic devices. Indeed, some chemical routes, like electrodeposition and electrophoretic deposition techniques, are widely used to produce high-performance porous oxide thin film electrodes, especially 3D oxide electrodes, taking advantage of the film conformality feature [145,250,251,252]. These techniques may be attractive and compatible with electronic microfabrication protocols, but they are only suitable for individual prototyping and lab-scale applications [59,241]. Table 3 summarizes the recently explored metal oxide electrodes for on-chip MSC applications.
Electrodeposition has successfully created mesoporous oxide MSC electrodes with ordered 3D architectures, such as RuO2 and MnO2 electrodes [145,250,251,252]. These 3D electrode materials allow good penetration of electrolytes and reactants into the entire electrode matrix and high mass loading of active materials [6,253]. Three-dimensional electrode materials are desirable where the footprint area is limited, such as on-chip MSC, to improve the surface-to-volume ratio, leading to significantly enhanced energy densities [6,253]. Additionally, the active layer is conformally deposited on a 3D scaffold with a high area enhancement factor (AEF), resulting in at least one order of magnitude improvement in the areal capacitance of 3D MSCs compared to planar geometry [6].
Physical vapor deposition (PVD) techniques are receiving more attention for the preparation of high-quality and contamination-free oxide thin film electrode materials for MSC applications [131,254]. These techniques are more eco-friendly than chemical routes. Different types of PVD techniques are available, such as molecular beam epitaxy, evaporation, and magnetron sputtering. The semiconductor industry commonly uses these techniques to deposit dense metallic or semiconducting films on large substrates [59,128,236,255]. Among these techniques, magnetron sputtering is considered the most fascinating tool due to its versatility and ability to produce tailored porous oxide film electrodes for on-chip MSCs [59]. Additionally, sputtering deposition leverages the ability to prepare thin films on large substrates or silicon wafers, making it a preferred choice for mass production of porous metal compound-based electrodes used in on-chip MSC applications. However, oxide electrode films produced through sputtering may exhibit lower capacitance compared to 3D electrodes produced through chemical routes.

Ruthenium Oxides

Ruthenium oxide (RuO2) is a model pseudocapacitive electrode material for MSC applications. It is widely known for its high specific pseudocapacitance owing to its ability to undergo rapid redox reactions with its multiple valence states in a reversible manner [256], excellent conductivity, and high energy density capability [241,242,257]. It has multiple valence states for electron transition (Ru2+, Ru3+, Ru4+, and Ru6+) accessible within a 1.2 V electrochemical potential window during redox reactions [20,222,249]. The charge storage process in RuO2 operated in acidic solutions is described in Equation (11) [6,18].
RuO 2 + xH + + x · e RuO 2 x ( OH ) x     ( 0 > x > 2 )
The equation explains the participation of protons during the fast redox reaction where RuO2 acts as a proton condenser; this reaction is electrochemically reversible [246]. The continuous change of ‘x’ during the fast redox reaction produces high specific capacitance [6]. Recent research has found that the specific capacitance of amorphous structures or hydrous forms of RuO2 (RuO2·nH2O) is superior to those of anhydrous structures or crystalline phase (i.e., rutile polymorph of RuO2); this is because there are more permeable sites in RuO2·nH2O for protons, and the proton conduction inside RuO2 is dominant compared to the electron conduction [6,246]. Hydrous regions are more permeable to protons, and as a result, proton conduction is the dominant form of conduction inside ruthenium dioxide instead of electron conduction. However, the hydrous forms of RuO2 are usually prepared by solution-phase processes. The high cost associated with bulk RuO2 has been controversial regarding its use for MSC applications. However, this concern has been arguably laid to rest as low mass loading (RuO2) electrodes are required for MSC applications, typically less than 1 mg cm−2 [6,241,258].
Pioneering work on MSCs began two decades ago with a RuO2 thin film electrodeposited by the reactive DC sputtering technique [256,259]. A solid electrolyte LIPON was sandwiched between two RuO2 electrodes, but the capacitance realized was found to be much lower than the chemically obtained bulk RuO2. The low capacitance was attributed to the dense morphology and anhydrous form of the thin film electrodes. Another issue identified was the low ionic conductivity of the solid electrolyte, limiting the rate capability due to the high ohmic drop [6,259]. Efforts have been made to improve the electrochemical performance through electrode topology/configurations. Hota et al. [247] explored different fractal electrode topologies, including Hilbert, Moore, and Peano, and the conventional interdigitated configuration of the sputtered RuO2-based MSCs (Figure 8a).
Among the different electrode designs, Moore-design MSC topology exhibited the best electrochemical performance, with a capacitance of 11 mF cm−2 (168 F cm−3) and an energy density of 23.2 mWh cm−3 at a power density of 769 mW cm−3. The conventional interdigitated electrode structure has a capacitance of 8 mF cm−2 (125 F cm−3) and delivers an energy density of 17.5 mWh cm−3 at the same power density. The MSC with the Moore design showed a 32% increase in energy density, which can be attributed to the active electrode surface being responsible for the improved performance and the increase in the electric lines of force coming from the edging effects in the fractal electrodes [247].
Another approach to improving the electrochemical performance of RuO2-based MSCs is the synergistic contributions of the unique properties of RuO2 with other (pseudo)capacitive materials in a composite electrode configuration. Han et al. [249] demonstrated an example of this by uniformly coating vertical graphene with a porous RuO2 film using a solution-free reactive magnetron sputtering technique. The composite electrode achieved an area capacitance of 15.3 mF cm−2, with 50 nm-thick RuO2 on 5 µm-thick graphene. Furthermore, the composite electrode retained excellent capacitance after 10,000 charging and discharging cycles. In another study featuring hydrous RuO2 prepared on vertically aligned carbon nanowalls (CNW) through an expensive solution-based electrodeposition method reported by Dinh et al. [194], a high areal capacitance of 1 F cm−2 was achieved. This method is mainly suitable for laboratory applications. Despite concerted efforts, the satisfactory electrochemical performance of RuO2-based MSCs via PVD techniques is yet to be achieved.
Several strategies have been proposed to scale up the collective fabrication of chemically prepared RuO2 electrode-based MSCs and other oxide electrodes. For instance, Patrice Simon’s group [241], demonstrated using laser writing techniques to integrate flexible MSCs on current collector-free polyimide foils. The microdevice delivered 27 mF cm−2 (540 F cm−3) in 1 M H2SO4 and retained 80% of the initial capacitance after 10,000 cycles. The authors alluded to the fact that the laser writing fabrication process is simple and scalable on a large scale [241]. However, despite all the recorded landmarks with 2D electrode materials, the electrochemical performance of oxides has yet to meet the energy requirements of the microdevices employed in IoT applications.
This has led to the development of a novel 3D electrode material configuration, usually prepared through solution-phase processes, to boost the device’s performance [251,260]. Asbani et al. [251] fabricated an efficient 3D RuO2 electrode (∼400 nm-thick), which was step-conformally electrodeposited on a 3D silicon microtube. The 3D silicon microtubes were fabricated using deep reactive ion etching (DRIE). The 3D RuO2 electrode exhibited remarkable areal capacitance of 4.5 Fcm−2 at 2 mVs−1, while maintaining more than 2 Fcm−2 at 100 mVs−1 (10 s charge/discharge time) and maintaining 90% of the initial capacitance value after 10,000 cycles.
Table 3. Metal oxide electrodes for on-chip MSCs.
Table 3. Metal oxide electrodes for on-chip MSCs.
Electrode MaterialSubstrateSynthesis/FabricationThickness (nm)ElectrolyteCell Vol.
/Pot. Window, V
Areal Cap.
mF cm−2
ConfigurationReference
RuO23D-Pt nanotubesElectrodeposition1300.5 M H2SO41.353203D Single electrode [261]
RuO2Si/SiO2/Ti/Au Electrodeposition/photolithography 0.5 M H2SO40.93 [262]
hRuO23D silicon microtubes scaffold + Al2O3/PtElectrodeposition/photolithography4270.5M H2SO4143003D-parallel plates [251]
RuO2-Au ElectrodepositionNAPVA-H3PO4-SiWA0.934733D-parallel plates [260]
hRuO2-carbon nanowires (CNWs) Si/Si3N4/Cr/Pt Electrodeposition + CVD/direct laser writing12,0000.5 M H2SO40.81094Single electrode [194]
RuO2-grapheneSi-SiO2PECVD + reactive sputtering550PVA-H3PO4 hydrogel115.3Single electrode [249]
RuOxNySzSiO2-Ti-Au-3D Porous AuElectrodeposition/photolithography200,0000.5 M H2SO40.8514,3003D Single electrode [257]
RuOxNySzSiO2-Ti-Au-3D Porous AuElectrodeposition/photolithography30,000PVA-SIWA1.17143D Interdigitated[257]
MnO23D silicon microtubes scaffoldElectrodeposition2005 M LiNO31185Single electrode [250]
MnO2Si-SiO2Electrodeposition/photolithigraphy30001 M Na2SO40.856.3Intedigitated[228]
MnO23D silicon microtubes scaffold + Al2O3/PtElectrodeposition/photolithography3500.5 M Na2SO40.86503D-interdigitated[263]
MnO23D siliconmicrotubes + Al2O3/PtElectrodeposition150.5 M Na2SO40.86703D-interdigitated[162]
MnO2nanotubesPET/3D polycarbonate (PC) membraneElectrodeposition/PDMS-assisted transfer/photolithigraphy8000PVA-Na2SO4 hydrogel113.23D-interdigitated[264]
MnO2-grapheneGraphene (LSG)Electrodeposition/direct laser writing15,0001 M Na2SO40.9852Interdigitated[201]
MnOx-AuPETE-beam evaporation/photolithography50PVA-H2SO4 hydrogel0.80.164Interdigitated[265]
Ni(OH)2PET-Nihydrothermal + spin coating/photolithigraphy600PVA-KOH hydrogel0.70.528Interdigitated[266]
Ni(OH)2Polyethylene naphthalate (PEN) sheetchemical bath deposition process (CBD)/photolithigraphy5001 M KOH 0.616Interdigitated[239]
NiFe2O4PET-Nielectrospinning/photolithography300PVA-KOH hydrogel0.80.067Interdigitated[267]
FeWO4Si/Al2O3/PtReactive DC magnetron sputtering 9005 M LiNO30.63.5Single electrode [128]
(Mn,Fe)3O4Si/Al2O3/PtReactive DC magnetron sputtering 38701 M Na2SO4180Single electrode [236]
IrOxSi/SiC/Ti/PtDC magnetron sputtering/lift-off photolithography300Phosphate-buffered saline (PBS) 112.753D Interdigitated[245]
S-doped grapheneSi/SiO2Spin coating/plama etching10PVA-H2SO4 hydrogel10.582Interdigitated[199]
F-doped graphenePETMask-assisted filtration1000EMIMBF4/PVDF-HFP ionogel3.517.4Intedigitated[197]
P-doped graphene Kevlar fabricLaser direct writing59,400PVA-H3PO4 hydrogel0.8125.35Intedigitated[198]
Cl-doped graphenePETMask-assisted filtration500EMIMBF4/PVDF-HFP ionogel3.58Intedigitated[268]
O-N-S-co-doped grapheneWoodSlurry coating/Laser direct writing 60,100PVA-H2SO4 hydrogel0.882.1Intedigitated[200]
Graphen-ThiophenePET Vacuum-filtering/Plasma etching105PVA-H2SO4 hydrogel13.9Intedigitated[269]
Graphene-Phosphorene PET Mask-assisted vacuum filtration2000BMIMPF639.8Intedigitated[177]
Graphene-PEDOTPET Pen lithography20,700PVA-H2SO4 hydrogel1.216Intedigitated[204]
Graphene-CNTs/Ag nanowiresSi/SiO2 and PETPlasma-jet based 3D printing20,000PVA-H3PO4 hydrogel0.821.6Intedigitated[270]
rGO-RuO2Free-standing fiber sheetmodified Hummers + hydrothermal42,5001 M H2SO414479.5Free-standing fiber sheet[202]
rGO-RuO2Free-standing fiber sheetmodified Hummers + hydrothermal42,500PVA-H3PO4 hydrogel1833.4253D-interdigitated[202]
rGO-MnO2NA (tansfarable to any substrate)Photomodulation/shaped femtosecond laser (SSFL)30000.5 M Na2SO421283D-interdigitated[271]
rGO-WO3highly oriented pyrolytic graphite (HOPG) Electrodeposition31,5000.158 M H2SO41178parallel plates[271]
rGO-MnO2Polyimide/AuDoctor blade + electrodeposition/Laser scribing15,0000.5 M Na2SO40.98523D-interdigitated[201]
rGO-ZnOPEThydrothermal reaction/Laser scribing11,000PVA-H2SO4 hydrogel14.3Intedigitated[272]
Graphene-V2O5PETElectrochemical exfoliation + hydrothermal method/Plasma etching300PVA-LiCl hydrogel13.9Intedigitated[273]
rGO-V8C7PETContinuous centrifugal coating and laser scribing13,000PVA-LiCl hydrogel0.849.5Intedigitated[274]
rGO-PPyFree-standing nanosheetsPolymerization + Mask-assisted filtration7000PVA-H2SO4 hydrogel0.881Intedigitated[275]
CNTs-NiPolycarbonate (PC) sheetMaskless laser-assisted dry transfer/direct laser writing50,350TMOS:FA:EMI TFSI ionogel30.43Intedigitated[211]
In addition, the introduction of heteroatoms like N and S in RuO2 has been reported to influence the oxide microstructure and surface significantly and provide better access to active sites in MSCs, as demonstrated by the group of David Pech [257]. The group fabricated interdigitated MSCs based on hydrous RuOxNySz electrodes (Figure 8b), and the conformally electrodeposited active material (RuOxNySz) on 3D porous current collectors exhibits an extremely high surface area with a nanodendritic network, demonstrating a high areal capacitance of 14.3 F cm−2 for the electrode (30 µm-thick films tested in 0.5 M H2SO4 at 5 mV s−1) in an electrochemical window of 0 to 1.1 V and 714 mF cm−2 for an all-solid-state MSC fabricated on silicon wafer, using a poly(vinyl alcohol) (PVA)-based electrolyte doped with silicotungstic acid (H4SiW12O40, SiWa), with stable performance (>80% retention after 5000 cycles). The device achieved a specific energy density of 120 µWh cm−2 and a maximum power density of 421 mW cm−2. The cell voltage of the device was extended up to 2.3 V when doped [EMIM] [TFSI] was used as the electrolyte. However, this came at the expense of a lower cell capacitance and higher ESR, while it achieved the energy and power density of 128 µWh cm−2 and 110 mW cm−2, respectively [257].

Manganese Oxides

Manganese oxides (MnOx), especially MnO2, are among the most appealing pseudocapacitive electrode materials due to their abundance, low cost, and high theoretical capacitance [18]. MnO2 has been a focus of exploration as pseudocapacitive electrodes for MSC applications since the publication of the pioneering work by the Goodenough research group in 1999 [276]. Their work highlighted the pseudocapacitive behavior of MnO2 in a neutral aqueous solution. However, achieving high capacitance performance with pristine MnO2-based electrodes in a planar ultrathin configuration is challenging. This is primarily due to the low mass loading, poor electrical conductivity (10−5 to 10−6 S cm−1), low structural stability and flexibility caused by volume expansion during charge and discharge, and electrochemical dissolution of active materials, leading to rapid capacitance degradation [146,253,265].
Consequently, several strategies have been undertaken to stabilize and improve the capacitance of the MnO2 through structural optimization (e.g., amorphous and α-, β- and λ-type crystalline MnO2 electrodes) [277,278], defect chemistry (e.g., mixed oxides, MnO2–polymer composite electrodes, MnO2-nanostructured carbon composites, noble metals) [131,279], tuning the morphologies (e.g., nanofilbers, nanorods, nanosheets) [130,280], and controllable incorporation of porosities [281], into the material to provide sufficient void nanospaces to accommodate volume expansions (i.e., better stress/strain accommodation). The porous morphology can provide a large accessible surface area, facilitate electrons/ions diffusion kinetics, stabilize the electrodes during the cycling process, and ensure good structural integrity for a long life cycle [131,282,283,284]. In addition, α-MnO2 is considered the most promising phase among the various crystallographic structures of MnO2; it has large tunnel sizes formed from double chains of MnO6 octahedra, which position the material to store more foreign cations while enabling the conversion of Mn4+ to Mn3+ ions for charge balance in reversible redox reactions, delivering it with the highest specific capacitance [131]. The pseudocapacitive behavior of MnO2 operated in neutral electrolytes can be explained by two fast Faradaic redox reactions: the intercalation process and the surface adsorption/desorption mechanism [285]. The first process involves the fast intercalation of protons (H+) and/or alkali cations (C+ = Na+, Li+, K+…) coming from the aqueous electrolyte during the redox reaction via fast Faradaic reaction in the bulk film of the MnO2, as described in Equation (12) [6,285].
MnO 2 + xH + + yC + + x + y · e MnOOH x C y
where x and y represent the number of moles of H+ and cations of C+ in the electrolyte intercalated in MnO2, respectively.
On the other hand, the adsorption/desorption mechanism involves fast and reversible surface redox reactions at the surface or near the surface of MnO2 rather than ion intercalation [253,285], as expressed in Equation (13).
( MnO 2 ) surface + C + + e ( MnOOC ) surface
In both Faradaic processes, a redox reaction occurs between the III and IV oxidation states of Mn ions [253,285]. The pseudocapacitive behavior contributes to the high-power density of MnO2-based electrochemical devices.
Thin-film electrodes based on MnO2 are usually prepared via chemical routes such as sol–gel coating [286], vacuum filtration [284], electrodeposition [287,288], and electrophoretic deposition [287,288]. MnO2 thin film electrode deposition via PVD techniques such as electron beam evaporation [265] and sputtering [131,289] has been reported. Si et al. [265] deposited MnO2 thin film electrodes on polyethylene terephthalate (PET) substrates using electron beam evaporation; the 50 nm-thick MnOx electrode exhibits a volumetric capacitance of 37.9 F cm−3 at a scan rate of 10 mV s−1. The low capacitance obtained was attributed to the poor electronic conductivity of the film; the conductivity was enhanced by incorporating very thin layers of gold into the MnOx layer as conductive additives to enable effective charge transport and electrode integrity. The 50 nm-thick MnOx/Au electrode exhibits a volumetric capacitance of 78.6 F cm−3 at a scan rate of 10 mV s−1, resulting in ~107% performance enhancement over the pristine MnOx electrode. The composite electrode stores charge through the synergetic contribution of the double-layer and pseudocapacitive processes of the active MnOx coupled with the highly conductive gold, contributing to the total capacitance [265]. Interdigitated MSCs fabricated with 50 nm-thick MnOx/Au composite film electrode fingers demonstrated a volumetric capacitance of 58.3 F cm−3, exhibiting a maximum volumetric energy density of 1.75 mW h cm−3 and a maximum power density of 3.44 W cm−3, which is adjudged to have performed better than many existing double layer MSCs [265].
Kumar et al. [131] deposited a high specific surface area and large aspect ratio a-MnO2 nanorod’s forest (~1 µm-thick) via reactive DC sputtering technique on silver (50 nm-thick) coated porous anodic aluminum oxide (AAO) substrate. The areal capacitance of the electrode tested in 1 M Na2 SO4 was found to be 207 mFcm−2 at a scan rate of 2 mVs−1. Broughton and Brett [289] used a double-step sputtering-electrochemical oxidation process to obtain manganese oxide thin films; the electrode thin films were synthesized by anodic oxidation of metallic films deposited by sputtering, starting from vacuum deposition of manganese metal onto Pt coated Si wafers in an Argon atmosphere and subsequent electrochemical conversion of the metal film into capacitive oxide with a porous and dendritic structure in a 1 M Na2SO4 electrolyte. The oxidation behavior was studied by chronopotentiometry (CP) under constant current and linear sweep voltammetry (LSV) techniques. The thin film electrode exhibits pseudocapacitive behavior in the neutral electrolyte, but with a relatively low capacitance value.
Manganese oxide (MnOx) composite electrodes have been reported for the MSC application [140,290,291], exhibiting higher stability of the oxidation states, which is crucial for long-term performance. The redox reaction of Mn ions in these oxidation states occurs spontaneously during the charge-discharge process and their relatively broad work potential window in aqueous electrolyte solution [292,293,294,295]. Li et al. [296] deposited Mn3O4 electrode thin films via electron beam evaporation on 300 nm-thick vertically porous nickel (VPN) sputtered on polyethylene terephthalate (PET) substrate. The Mn3O4 electrode achieved a volumetric capacitance of 533 F cm−3 at the scan rate of 2 mV s−1; the interdigitated MSCs based on the Mn3O4 electrode were fabricated by a conventional photolithography process and achieved a volumetric capacitance of 110 F cm−3 at the current density of 20 μA cm−2, retaining 95% of the initial capacitance after 5000 cycles under the current density of 20 μA cm−2 [296]. In another study, an asymmetric MSC comprising rGO//MnOx exhibited an energy density of 1.02 mWhcm−3 at a maximal power density of 3.44 Wcm−3 [296], and the Sun group [297] achieved a specific capacitance of 73.25 mF cm−2 with graphene/MnO-Mn3O4 composite-based interdigitated MSC fabricated by the laser direct-writing method. The device delivered a maximum power and energy density of 1.29 mW cm−2 and 14.65 μWh cm−2, respectively, retaining 90% of its initial capacitance after 5000 cycles.
Like 3D RuO2 electrode materials, 3D MnO2 electrode materials have been demonstrated by various groups [162,250,252,298]. For instance, Bounor et al. [250] demonstrated a collective fabrication of 3D MSCs integrated on silicon wafers (Figure 8c), using MnO2 as the active electrode material deposited by a pulsed electrodeposition method. The 3D electrodeposited MnO2 electrode films (1.24 µm-thick) exhibited 1.7 F cm−2 in 5 M LiNO3 measured at a scan rate of 2 mV s−1. The MSCs fabricated with 3D MnO2 (450 nm-thick) exhibited the areal capacitance of 0.75 F cm−2 at 2 mV s−1 in 5 M LiNO3 with the cell voltage of 1 V, delivering the energy density of 50–100 µWh cm−2 while keeping the power density high (>1 mW cm−2) and the cycling life, retaining > 82% of the initial capacitance after 10,000 cycles [250]. In a recent study by the David Pech group, 3D MSCs were fabricated based on highly electrodeposited porous scaffolds of Ni/MnO2. The electrodes exhibit high areal capacitance of over 4 F cm−2 and excellent cycling stability in the ionic liquid-based electrolyte using NaFSI/Pyr13FSI [299].

Molybdenum Oxides

Molybdenum oxides (MoOx, 2 ≤ x ≤ 3) exhibit (pseudo)capacitive behavior in neutral or slightly acidic aqueous solutions, making them highly promising as MSC electrode materials, especially as negative electrodes [144,300,301]. The pseudocapacitive behavior of molybdenum oxides primarily arises from fast and reversible surface redox reactions. Mo oxides are structurally diverse with multi-valence states, multielectron redox is expected, given that Mo6+, Mo5+, and Mo4+ are all accessible and stable oxidation states [243,302,303,304]. Mo oxides exist in various stoichiometries, ranging from the nonconductive full stoichiometry MoO3 with a large bandgap of about 3.0 eV to the more conductive reduced oxide MoO2 [246,303]. These oxides are composed of layered MoO6 octahedra structures formed by der Waal forces through edge or corner sharing; the structures are beneficial for the permeation of small ions such as H+, Na+, and K+ [225,303,305,306]. Various forms of Mo oxide nanostructures [307,308] and nanocomposites [309] have been widely investigated as pseudocapacitive electrodes in classical supercapacitors, exhibiting outstanding theoretical specific capacitance (2700 F g−1) [307,308]. However, there are limited reports on the use of Mo oxides as on-chip MSC electrodes [300,310].
The charge storage mechanism of MoOx in an aqueous electrolyte can be explained by the fast Faradaic redox reactions, involving ion adsorption at the surface or near the surface of MoOx accompanied by rapid and reversible electron transfer, as expressed in Equation (14) [225].
MoO x + zC + + ze MoOOC
where C+ and z represent surface-adsorbed alkali cations (C+ = Na+, Li+, K+, Zn2+…) coming from the aqueous electrolyte and the number of electron transfers, respectively.
Orthorhombic phaseα-MoO3 and MoO2 are attracting more attention due to their thermodynamic stability, the unique layered crystal structure of α-MoO3 with two interleaved sublayers, and the tunnel structure of MoO2 [144,225,243,307,308,311,312,313]. However, the widespread practical applications of MoO3-based electrodes in MSCs are limited due to poor electrical conductivity, low specific capacitance, and rapid deterioration owing to the lack of structural reliability during electrochemical reactions, resulting in poor electrochemical performance, which is attributed to the lack of electrochemically active centers and the low surface electrochemical activity of MoO3-based materials [243,309,314].
The electrical conductivity of MoO3 can be improved by (i) doping engineering, (2) defect engineering, such as introducing oxygen vacancies, MoO3−x (2 < x < 3), to act as shallow donors by increasing the carrier concentration, and (3) crystal engineering, such as reducing crystal size and converting into quantum dots [243,315,316]. Both doping engineering and crystal engineering involve complex material preparation processes, which can limit their applications in on-chip devices. Additionally, accurately controlling doping atoms can be challenging, while crystal engineering may face difficulties such as coalescence. To introduce oxygen vacancies, several techniques can be employed, including heat treatment in a specific atmosphere, high-energy particle bombardment, and chemical post-treatment, among others [225,243,315,317]. Non-stoichiometric oxygen-deficient MoO3−x (2 < x < 3) has been demonstrated to enhance the intrinsic electrical conductivity and produce an expanded interlayer distance in the MoO3, structure, thus improving the electrochemical performance [315,317,318]. Synthesis of composites and doped MoO3 electrode materials with metal ions can increase the number of electrochemically active centers at the surface, thus addressing the low specific capacities and rapid deterioration [309,313,319,320].
Zhang et al. [321] employed the screen printing technique to fabricate interdigitated MSCs on Au/polyimide (PI) substrates. The device delivered 41.7 mF cm−2 and 5.8 μWh cm−2 for the MSC with MoO3-x nanorod microelectrodes with PVA/H2SO4 gel electrolyte, retaining 81% of the initial capacitance after long-term charging-discharging cycling at 2.5 mA cm−2 for 8000 cycles; this is considered one of the highest for the MSCs with interdigital metal oxide electrodes [296,321,322,323]. For on-chip implementation, more attention is given to growing the oxide film electrodes directly onto the substrate by PVD, especially the sputtering techniques; for instance, the Zhang group [243,300] deposited MoOx (x = 2.3) thin film via RF magnetron sputtering, and the MoOx (~1 µm-thick) electrode achieved an areal capacitance of 31 mFcm−2 at 5 mVs−1 in 0.5 M Li2SO4 [243]. The group further investigated the performance of the sputtered films (860 nm-thick) in an asymmetric MSC configuration containing MoO2+x(−)//2 M Li2SO4//MnO2(+), demonstrating an energy density of 2.8 µWh cm−2 at areal power density of 0.35 mW cm−2, and good stability with no capacitance loss for 10000 cycles [243,300].
As earlier remarked, MoO2 electrodes exhibit better electronic and ionic conductivity, and the unique tunnel structure can facilitate fast ionic transport during intercalation/deintercalation reactions, expected to result in a high energy density [144,301]. However, MoO2 suffers from instability; hence, most reports on MoO2 electrodes for MSC applications are based on MoO2 composite materials [324,325]. For instance, Lin et al. [325] prepared a laser-induced graphene/MoO2 (LIG/MoO2) composite using Mo ion homo-dispersed hydrogel ink as a precursor for the laser scribing process to fabricate interdigitated MSCs to deliver an areal capacitance of 81.8 mF cm−2 and an areal energy density of 113.7 μWh cm−2 at a power density of 2.5 mW cm−2 with PVA-H3PO4 gel electrolyte. Most of the reported MoO2- and MoO3-based electrodes for MSC applications are usually synthesized via chemical processes [144,326].

Iridium Oxide (IrO2)

Iridium oxide (IrO2) is an electrically conductive transition metal oxide and is projected as one of the most interesting oxide materials for MSC applications owing to its fast reversible redox reactions between the Ir3+ and Ir4+ oxidation states, good conductivity at room temperature, and excellent stability in acidic and basic solutions [327,328]. IrO2 thin films, especially sputtered iridium oxide films (SIROFs), have been widely explored in other applications such as neural stimulation and recording electrodes in implantable bioelectronics due to their high charge capacity, low impedance, and biocompatibility [245], while the applications in MSCs are yet to be widely known. IrO2 exhibits better pseudocapacitive behavior with its hydrated phase in an acidic solution, like most binary transition metal oxides such as RuO2 and MnO2. IrO2 electrodes for aqueous MSCs application deposited by radio frequency (RF) sputtering were first reported by Liu et al. [327]. The sputtered IrO2 thin films electrode was tested in 1 M LiNO3 electrolyte, demonstrating the capacitive behavior; the sputtered IrO2 thin films electrode was tested in 1 M LiNO3 electrolyte, demonstrating the capacitive behavior within the potential window of 0–1.0 V (vs. SCE) (Figure 8d). The IrO2 thin film electrode exhibits excellent cycling stability over 10,000 cycles in LiNO3 solution, and the pseudocapacitive process within the IrO2 thin film electrode was expressed as:
( IrO 2 ) surface + C + + e ( IrO 2 C + ) surface
where C+ is the electrolyte cation
Inspired by the applications in neural interface research, Geramifard et al. [245] studied the electrochemical performance of the interdigitated MSCs based on 300 nm-thick SIROFs; the device was fabricated via lift-off photolithography and was operated in two different physiological aqueous electrolytes, phosphate-buffered saline (PBS) and an inorganic model of interstitial fluid (model-ISF) in a potential range of −0.6 to +0.8 V, exhibiting capacitance of 12.75 mF cm−2 (425 F cm−3) in PBS and 6.7 mF cm−2 (223 F cm−3) in model-ISF, and an energy density of 59.1 mWh cm−3 in PBS and 30.9 mWh cm−3 in model-ISF. The device demonstrated stability over 100,000 cycles at 200 mV s−1 without loss of capacitance in these electrolytes. In addition, a number of studies have investigated IrO2 electrode composites [329].

Iron Oxides

Iron oxides, especially Fe2O3 and Fe3O4, are promising negative electrode materials in aqueous alkaline [330,331,332] and neutral [333,334,335,336] electrolytes; they exhibit a surface redox pseudocapacitance in a relatively wide operating potential window in negative potential (0 to -0.8 V vs. Ag/AgCl) and behave like a battery material when the window is extended [337,338,339,340]. Fe oxides are structurally diverse with multi-valence states (Fe0, Fe2+, Fe3+, etc.), and the rich oxidation-reduction reactions (Fe2+/Fe3+, Fe0/Fe2+, Fe0/Fe3+, etc.) are beneficial to high specific capacitance performance [341]. Crystal structures play a crucial role in adjusting the electrochemical properties/performance of the iron oxide electrode materials. Among the various crystal structures of Fe2O3, α-Fe2O3 (rhombohedral-centered hexagonal) is the most thermodynamically stable and more attractive candidate for MSC applications, while other phases such as β-Fe2O3 (cubic body-centered), γ-Fe2O3 (cubic structure of inverse spinel type), and ε-Fe2O3 (orthorhombic) are intermediary metastable and can easily evolve to α-Fe2O3 when subjected to heat treatment [342,343].
Despite their potential benefits, significant volume expansion, poor rate stability, low cycle lifetime, and poor electrical conductivity limit their practical applications as MSC electrodes [344]. However, attempts have been made to address these challenges, such as combining them with a highly conductive matrix to mitigate the structural distortion that usually leads to the deterioration of the inherently fragile material structure [344,345,346,347].

Tungsten Oxides

Tungsten oxides (W oxides), particularly WO3, exhibit good electronic conductivity (10−6–10 Scm−1), high intrinsic density (>7 gcm−3), multi-valence states (from W2+ to W6+), and high theoretical specific capacitance (1112 Fg−1) [348,349,350,351,352,353]. The charge storage mechanism of WO3 arises from electrochemical intercalation/deintercalation with protons, as expressed in Equation (16).
WO 3 + xH + + xe H x WO 3 0 < x < 1
The demonstration of pseudocapacitive behaviors is considered an important merit for MSC applications [350,353,354]. Just like Mo oxides, W oxides appear in various stoichiometries, from stoichiometric WO3 to more conductive nonstoichiometric WO3-x (2 < x < 3) [355]. The crystal structure of W oxides plays a critical role in the electrochemical performance; WO3 exists in various polymorphs, including tetragonal (α-WO3), orthorhombic (β-WO3), monoclinic (γ-WO3), triclinic (δ-WO3), monoclinic (ε-WO3), and hexagonal (h-WO3) [355,356]. Monoclinic (γ-WO3) is the most stable phase among the crystal structures, while hexagonal (h-WO3) is more attractive due to larger hexagonal tunnels to benefit the insertion of protons without changing the crystal structure, but unfortunately, the phase is metastable and can easily be transformed to monoclinic, which is the most stable phase [345,355,356].
It is worth noting that the bulk state of WO3 exhibits battery-type energy storage behavior with redox phase transformation, so WO3 is mostly used in composite electrodes, especially within a porous conductive carbon network; they exhibit pseudocapacitive behavior [141,353,354,357,358]. For instance, Shi et al. [359] prepared 3D hierarchical porous carbon/WO3 nanocomposites using carbonization and solvothermal processes. The MSC based on the nanocomposites achieved a specific capacitance of 20 mF cm−2 in polyvinyl alcohol–sulfuric acid gel electrolyte. Again, Hapel et al. [360] employed a multi-step process to prepare a nanocomposite comprising WO3-x and rGO by grafting WO3 nuclei onto GO defect sites by electrochemical processing and nanoparticle growth, followed by the electroreduction in unprotected oxidation sites of GO to form a highly conductive reduced graphene oxide (rGO) support. The fabricated MSCs based on the composite electrode achieved a maximum energy density of 19.8 Wh cm−3 and a maximum power density of 80.4 Wcm−3.
In addition, Shinde et al. [355] have proposed that metallic tungsten can be converted into various tungsten oxides through an electrochemical route, initiating from the surface of metallic tungsten interpenetrated with an aqueous electrolyte. The electrochemical properties of the surface oxide on metallic tungsten can be harnessed and greatly benefit the pseudocapacitance [361,362,363,364]. In this context, pure metallic W can be investigated as MSC electrodes.

Vanadium Oxides

Vanadium oxides are pseudocapacitive electrode materials and are known for their high power density, natural abundance, and high theoretical specific capacitance [221,344,365]. V oxides exhibit multiple oxidation states from +2 to +5. V5+ is the most stable, while V4+ is the most unstable, making V2O5 more attention-grabbing [221,344,357,366]. The crystalline form of V2O5 (orthorhombic) does not exhibit capacitive behavior when in bulk form but instead displays battery electrode characteristics, whereas amorphous V2O5 exhibits pseudocapacitive behavior [357,367]. However, poor electronic conductivity (10−2–10−3 S cm−1) and structural instability are major issues faced by vanadium oxides. Various strategies have been exploited to improve their electrochemical performance, including forming composites with highly conductive materials and structural control to stabilize the material [221]. Liu fabricated a 3D MSC with V2O3 to achieve a volumetric energy density of 110 mWh cm−3. Huang et al. [368] fabricated vanadium oxide/carbon nanowire electrode composites to achieve a high areal capacitance of 1.31 F cm−2 at 1 mA cm−2 in a 5 M LiCl aqueous solution, and the MSC obtained a volumetric energy density of 3.61 mWh cm−3 at 0.01 W cm−3, retaining 91% of its capacitance after 10,000 galvanostatic charge–discharge cycles.

Other Binary Metal Oxides

Other binary metal oxides, such as niobium oxides [369,370,371], titanium oxides [372,373], nickel oxides [374,375], cobalt oxides [376,377,378], and copper oxides [323,379] have been studied by various investigators. These pseudocapacitive electrode materials suffer poor electronic conductivity and structural instability, limiting their practical applications as MSC electrode materials. However, they have been explored as composites with highly conductive materials [370,372,373]. Oxygen deficient metal oxides could also propose nice electrical conductivity, which could be suitable for high power applications [315,380,381].

Mixed-Metal Oxide Electrodes

Many polycationic oxides have been found to exhibit a capacitive-like behavior and are considered one of the leading contenders to replace carbon-based electrode materials, especially in conventional supercapacitors. Examples of mixed-metal oxides that exhibit (pseudo)capacitive behavior include spinel oxides (AB2O4, A = divalent metal ion with +2 oxidation state and B = trivalent metal ion with +3 oxidation states) [382,383] such as MnFe2O4 [384] and FeCo2O4 [385]; perovskite oxides (ABO3) such as; and wolframite transition metal tungstate (AWO4, A = Mn, Fe, Co, Ni, Cu, Zn) such as FeWO4 [386,387,388], CoWO4 [389], NiWO4 [390], ZnWO4 [391], and CuWO4 [391]. Unfortunately, the preparation routes of these compounds [386,392] are largely not compatible with microfabrication protocols employed in the electronics industries, thereby restricting their applications for on-chip MSC applications.
However, recent development has shown that there is a possibility of obtaining some of the multicationic mixed-metal oxide electrodes via magnetron sputtering techniques [128,236]. This is a game changer and one of the most promising candidates for on-chip MSC applications. For instance, Lethien group developed mixed-metal oxide electrodes by reactive DC magnetron sputtering, a deposition method widely used in the semiconductor industry to manufacture micro-devices [128,236]. The first study investigated sputtered wolframite-type oxide and iron-tungstate oxide (FeWO4) films in a neutral aqueous electrolyte (5 M LiNO3) (Figure 9a) [128].
The sputtered films exhibited capacitance behavior in the negative potential (vs. Ag/AgCl), in good agreement with the one reported for the FeWO4 bulk counterpart [387]. The sputtered 900 nm-thick FeWO4 film achieved an areal capacitance of 3.5 mFcm−2 at 10 mVs−1 with excellent capacitance retention. Regarding the low capacitance value obtained for this sputtered FeWO4 film, a small amount of Fe or W is redox active. This group also investigated mixed Mn-Fe spinel oxide films in a neutral aqueous electrolyte (1 M Na2SO4) (Figure 9b) [236]. The electrochemical signature of the sputtered (Fe,Mn)3O4 spinel oxide films is comparable to those reported for their spinel-type Mn-Fe bulk counterparts [393,394]. The areal capacitance at 10 mVs−1 is 15.5 mFcm−2 for 1 μm thick film, exhibiting excellent coulombic efficiency and long-term cycle stability after 10,000 cycles.

3.2.2. Metal Nitrides

Metal nitride-based electrode materials have attracted significant attention as promising pseudocapacitive electrodes due to their metal-like conductivity and excellent electrochemical properties [59]. Metal nitrides differ from metal oxide-based electrode materials, which exhibit high capacitance in their hydrous forms and are typically prepared using solution routes. High-performance metal nitrides can be obtained through vacuum deposition techniques, particularly sputtering deposition [139,145,395,396,397,398]. This benefit can also be leveraged to perform large-scale depositions of porous nitride thin films, which is considered an advantage for mass production of on-chip MSCs via microfabrication processes. Table 4 summarizes various nitride electrodes for on-chip MSC applications.

Vanadium Nitride

Vanadium nitride (VN) is the most widely investigated nitride material for MSC applications due to its outstanding electrochemical performance in aqueous media [399], multiple oxidation states (II–V), and high electronic conductivity (about 1.6 × 106 S m−1) [398], bifunctional properties, serving as both the current collector and the electrochemically active material (working electrode) [400], its negative potential in aqueous electrolytes, which endows its utilization as negative electrodes in an asymmetric MSC configuration [45,145]. Brousse’s group [400] investigated the charge storage mechanism in the VN film electrode prepared by reactive magnetron sputtering in both aqueous (KOH) and organic (NEt4BF4) electrolytes; they observed that the charge storage mechanism depends on the nature of the electrolytes. In the presence of KOH, fast and reversible redox reactions were observed, with characteristics of pseudocapacitance, while in the case of NEt4BF4 in acetonitrile enabled, only double layer capacitance was observed. Further investigation on the charge storage mechanism was conducted by Bondarchuk et al. [401], leaving room for speculation due to an uncovered anomalous non-faradaic capacitance.
The Lethien group [399] employed a combination of numerous advanced characterization techniques to further unravel the charge storage mechanism in the sputtered porous VN films (Figure 10a). The VN films were obtained via the sputtering deposition technique. X-ray photoelectron spectroscopy (XPS) with time-of-flight secondary ion mass spectrometry (ToF-SIMS) measurements were performed on the films to reveal the species of VN, vanadium oxide (+III and +IV), and a small amount of vanadium oxynitride in the films. A further analysis was conducted using times-resolved operando X-ray absorption spectroscopy (XAS) under synchrotron radiation in KOH electrolyte. The result clearly showed a thin vanadium oxide layer on the VN, which plays a key role in the fast charge transfer occurring during the redox process, and the V oxidation state in the VN film is close to +4, agreeing with the XPS results. The VN film primarily acts as a conductive scaffold for electron transport, while the amorphous mixed-valence vanadium oxides (+3/+4) serve as the redox-active material, allowing fast charge transfer.
A 16 μm-thick sputtered VN film was tested in 1 M KOH to achieve a high capacitance of 1.2 F cm−2 (>700 F cm−3). It retained 80% of its initial capacitance after 50,000 cycles. Moreover, the MSC based on the sputtered VN electrode delivered 25 μWh cm−2, which is quite competitive with cutting-edge transition metal oxide/nitride materials and exceeds the performance of standard carbon electrodes [399]. This group made the assumption that a thin layer of vanadium oxides is responsible for the charge storage mechanism while the VN is used for fast electron transport, making this material a pertinent material for MSC applications.
The reaction in Equation (17) was proposed to elucidate the charge storage mechanism of the VN films tested in an alkaline KOH electrolyte [399]:
V 4 + O 2 + H 2 O + 2   V e V 3 + O O H + O H
However, it has been reported that the VN stability in aqueous electrolytes is a major issue, as it can cause a structural change that affects the potential use of energy storage systems [402]. To address this issue, efforts have been made to improve stability and electrochemical performance by adjusting the physicochemical properties during the deposition processes [62,260]. For instance, Jrondi et al. [255] tuned the sputtering deposition process to achieve a high surface capacitance value of 1.4 F cm−2, which demonstrates excellent cycling stability, retaining 90% of the initial capacitance after 150,000 cycles. Additionally, the film exhibits an ultra-high rate capability, with 75% of the initial capacitance maintained at 1.6 V s−1 [255].

Ruthenium Nitride

Ruthenium nitride (RuN) electrodes are unexplored for MSC on-chip applications. Studies have shown that RuN electrode films exhibit good electrochemical properties, similar to those of RuO2 in aqueous electrolytes [53,397]. A recent investigation by the Lethien group [53] sputtered porous RuN films with nanofeather morphology; the RuNelectrode properties were modified through an electrochemical oxidation process (EOP). The RuN nanofeather electrode benefited from the charge transfer at the electrode/electrolyte interface between the hydroxide ions (from the 1 M KOH electrolyte) and the feather-edge amorphous RuO2 pre-EOP and the h-RuO2 post-EOP. The electrode demonstrated remarkable performance (3.2 F cm−2 and 3200 F cm−3), with only a slight decrease in rate capability (τ < 10 s).
Table 4. Metal nitride electrodes for on-chip MSCs.
Table 4. Metal nitride electrodes for on-chip MSCs.
Electrode MaterialSubstrateSynthesis/FabricationThickness (nm)ElectrolyteCell Volt./Pot. Window, VAreal Cap.
mF cm−2
Vol. Cap.
F cm−3
ConfigurationRef.
VNTa foilSputtering/nitridation 4001 M KOH0.6375Single electrode [401]
VNSi/Si3N4Sputtering16,0001 M KOH0.61200700parallel plates [399]
VNSi/Si3N4Sputtering34001 M KOH0.6220650Single electrode [59]
VNSi/Si3N4Sputtering20001 M KOH0.640 Interdigitated[59]
VNSi/Si3N4Sputtering32,2001 M KOH0.61400 Single electrode[255]
Cr-doped VNSiSputteringNA1 M KOH1190NASingle electrode [403]
VN-carbon nanosheetsCarbon nanosheetsNitridation17,000PVA-KOH hydrogel0.92046.121203.6Single electrode [404]
W2NSi/Si3N4Sputtering79001 M KOH0.6550700Single electrode [395]
Mo2NTi foilsSputtering12420.5 M Li2SO40.955722Single electrode [405]
RuNSiSputtering4501 M KOH0.96133Single electrode [397]
RuNSi/Si3N4Sputtering10,0001 M KOH0.8532003200Single electrode[53]
VWNzSi/Si3N4Sputtering1501 M KOH0.611.5700Single electrode[406]
TiNSiSputtering15,6000.5 M H2SO40.827.342.6parallel plates [407]
TiNSiSputtering7700.5 M K2SO40.712146.4Single electrode [408]
TiVNSiSputtering2701 M KOH115500Single electrode [409]
Nb4 N5Nb foilElectrodeposition17,4001 M H2SO40.6225.872.3Single electrode [410]
Nb4N5-N-doped carbonNb foilElectrodeposition17,4021 M H2SO41232.9133.85parallel plates [410]
CrNSiSputtering11000.5 M H2SO40.812.8116.36parallel plates [411]
CrNSiSputtering21000.5 M H2SO40.817.783.26parallel plates [412]
CrNSiSputtering11000.5 M H2SO40.831.3284.55parallel plates [413]
Mn3N2Stainless steel SputteringNA1 M KOH0.9118NASingle electrode [414]
MoNx-TiNTi foilsElectrodeposition + nitridation 40001.0 M LiOH0.6121.5NASingle electrode [415]
Co3NSiSputtering9296 M KOH0.679.1851.4Single electrode [416]
CNTs-TiN (3.4/1.2 µm)Si/amorphous carbon filmCVD + Sputtering4600K2SO40.718.13-Single electrode [417]
CNTs-VN (3.4/1.3 µm)Si/SiO2CVD + Sputtering4700K2SO40.537.5-Single electrode [214]
CNTs-VNCNT fibersSolvothermal 30,0003 M KOH1564-Single electrode [215]
Figure 10. Metal nitride electrode materials for on-chip micro-supercapacitor applications. (a) Vanadium nitride electrode films. SEM cross-section images of the sputtered porous VN films with the charge storage process, CV plot of the VN electrode vs. the thickness, evolution of the areal and volumetric capacitances vs. the thickness, and capacitance retention vs. the number of cycles (16 μm-thick film) at 50 mV s−1, respectively. Reproduced with permission from [399]. Copyright (2020), Royal Society of Chemistry. (b) Sputtered tungsten nitride films as pseudocapacitive electrodes for on chip micro-supercapacitors. SEM cross-section images of the sputtered W2N films, CVs of various W2N film thickness measured at 5 mV s−1 in 1 M KOH, and the corresponding areal capacitance, respectively. Reproduced with permission from [395]. Copyright (2019), Elsevier. (c) Wafer-scale performance mapping of magnetron-sputtered vanadium tungsten nitride. VWNz sputtered on silicon wafers, CV plot of the VWNz films, and film stability in KOH electrolyte, respectively. Reproduced with permission from [406]. Copyright © 2024, American Chemical Society.
Figure 10. Metal nitride electrode materials for on-chip micro-supercapacitor applications. (a) Vanadium nitride electrode films. SEM cross-section images of the sputtered porous VN films with the charge storage process, CV plot of the VN electrode vs. the thickness, evolution of the areal and volumetric capacitances vs. the thickness, and capacitance retention vs. the number of cycles (16 μm-thick film) at 50 mV s−1, respectively. Reproduced with permission from [399]. Copyright (2020), Royal Society of Chemistry. (b) Sputtered tungsten nitride films as pseudocapacitive electrodes for on chip micro-supercapacitors. SEM cross-section images of the sputtered W2N films, CVs of various W2N film thickness measured at 5 mV s−1 in 1 M KOH, and the corresponding areal capacitance, respectively. Reproduced with permission from [395]. Copyright (2019), Elsevier. (c) Wafer-scale performance mapping of magnetron-sputtered vanadium tungsten nitride. VWNz sputtered on silicon wafers, CV plot of the VWNz films, and film stability in KOH electrolyte, respectively. Reproduced with permission from [406]. Copyright © 2024, American Chemical Society.
Batteries 10 00317 g010

Other Binary Nitride Electrodes

Besides the most widely exploited VN-based electrodes, other nitride materials issued from microelectronics can be synthesized/deposited and investigated as electrode materials for on-chip MSC applications. Examples of these metal nitrides are tungsten nitride (WN, W2N) [395,418], chromium nitride (CrN, Cr2N) [139,396,411,412], manganese nitride (MnN, Mn3N2) [414,419], titanium nitride (TiN) [408,420], molybdenum nitride (MoN) [421] ruthenium nitride (RuN) [397], niobium nitride (NbN) [422], hafnium nitride (HfN) [423], and cobalt nitride (Co3N) [423]. Wei et al. [411] reported the preparation of CrN thin films by reactive DC magnetron sputtering. The sputtered CrN thin film electrodes achieved a specific areal capacitance of 12.8 mF cm−2 at 1.0 mA cm−2, retaining about 92% of the initial capacitance after 20,000 cycles in a 0.5 M H2SO4 electrolyte, and fabricated MSCs based on CrN (in a face-to-face configuration) to deliver an energy density of 8.2 mW h cm−3 at a power density of 0.7 W cm−3 along with outstanding cycling stability. Qi et al. [412] fabricated symmetric MSCs based on porous CrN thin film electrodes deposited by sputtering to obtain a capacitance of 17.7 mF cm−2 at a current density of 1.0 mA cm−2 in a 0.5 M H2SO4 electrolyte, delivering maximum energy and power densities of 7.4 mWh cm−3 and 18.2 W cm−3, respectively. Ouendi et al. [395] achieved capacitance values up to 0.55 F cm−2(>700 F cm−3) with sputtered 7.9 μm-thick W2N films in 1 M KOH aqueous electrolyte (Figure 10b). Sputtered NbN thin film electrodes on silicon substrate demonstrated a volumetric capacitance of 707.1 F cm−3 in 0.5 M H2SO4 aqueous electrolyte with good cycling stability, recording 92.2% capacitance retention after 20,000 cycles [423]. HfN thin film electrodes deposited by reactive DC magnetron sputtering demonstrated an achieved areal capacitance of 5.6 mF·cm−2 at 1.0 mA·cm−2 in 0.5 M H2SO4 aqueous electrolyte [422]. Sputtered TiN thin films on silicon substrate achieved a volumetric capacitance of 146.4 F cm−3, with outstanding cycling stability over 20,000 cycles [408]. A fabricated MSC based on tailored surface sputtered TiN thin film electrodes achieved energy and power densities of 23 mWh cm−3 and 7.4 W cm−3, respectively [420]. Sputtered Mn3N2 thin film electrodes were comparatively tested in basic, neutral, and acidic electrolytes, they demonstrated areal capacitance values of 118 mF cm−2 for KOH, 68 mF cm−2 for KCl, and 27 mF cm−2 for Na2SO4 at a scan rate of 10 mVs−1. The Mn3N2 electrodes indicated better stability in alkaline media than the neutral and acidic electrolytes [414]. These materials are becoming more competitive with VN-based electrodes due to recent concerted efforts to fine-tune their properties to achieve better electrochemical performance.
In addition, the formation of nitride composite (not limited to VN) materials is considered a viable strategy to mitigate against the electrochemical oxidation and the dissolution of the pristine metal nitride materials, usually encountered in aqueous electrolytes, thus enhancing the electrochemical performance and improving the cycle life [402,404,424,425]. Numerous MSCs based on nitride composite electrodes have been exploited with impressive electrochemical performance. Unfortunately, there is a restriction on the on-chip implementation due to the preparation techniques, which are largely not compatible with electronic microfabrication protocols. The common processing routes to obtain the majority of the composites include hydrothermal [404,418,426], a solvothermal process employing nitridation [215,427], and chemical methods [428]. An example is the solvothermal process to prepare VN-CNTs composite electrodes for wearable MSCs application, achieving a specific capacitance of 564 mFcm−2 (188 Fcm−3) in a 3 M KOH aqueous electrolyte [215]. Another report of an NV-carbon nanosheet composite (VN-CNS) electrode for MSCs prepared by hydrothermal process demonstrated a high volumetric capacitance of 1203.6 F cm−3 at 1.1 A cm−3 and a high rate capability of 703.1 F cm−3 at 210 A cm−3 in 1 M KOH electrolyte, retaining 90% of the initial capacitance after 10,000 cycles [404].

Mixed-Metal Nitride Electrodes

Like mixed-metal oxide electrodes, mixed-metal nitride electrodes are currently receiving notable attention due to the synergy of the participating multivalent metal cations in modifying the properties of their parent compounds to improve electrochemical performance. Taking advantage of the processing techniques, especially physical vapor deposition like sputtering techniques that are CMOS compatible, is capable of obtaining thin films on a large wafer. For instance, the electrochemical behavior of VN electrodes is principally due to pseudocapacitive charge storage, exhibiting high areal capacitance [399], while, unlike VN electrodes, TiN appears to be mainly ruled by EDL type, exhibiting high cycling ability [409,420]. The synergistic electrochemical effects of coupling the participating transition metals, in this case, mixed titanium and vanadium (Ti,V)N, have aroused research interest for MSC applications. Achour et al. [409] reported an example of this synergy, using a reactive DC co-sputtering technique to deposit titanium vanadium nitride (TiVN) thin film electrodes with different Ti/V ratios. The results showed that the V-rich electrode sample exhibits a Faradic behavior that limits its cycling ability. Despite a high capacitance of 24 mFcm−2 at 2 mVs−1, incorporation of Ti in the film drastically improves the cycling ability, with virtually no fade in capacitance after 10,000 cycles. The Ti-rich electrode exhibits 5 mFcm−2, while the TiVN electrode (with a Ti/V ratio close to 1.1) exhibits an areal capacitance value of 15 mFcm−2 (500 Fcm−3) at 2 mVs−1 in 1 M KOH electrolyte solution. Similarly, a thin film of TiVN deposited on stainless steel substrates by a pulsed DC magnetron sputtering technique was tested with 1 M Na2SO4 electrolyte to achieve a volumetric capacitance of 155.94 Fcm−3 [424].
In addition, Niobium titanium nitride (NbTiN) thin film electrodes were deposited using the reactive magnetron co-sputtering technique. The electrodes exhibit a specific capacitance of 59.3 mF cm−2 at 1.0 mA cm−2 (or 154.4 mF cm−2 at 100 mV s−1) along with excellent long-term stability (at least 20,000 cycles), superior to that of TiN (26.9 mF cm−2 at 1.0 mA cm−2) and NbN (39.6 mF cm−2 at 1.0 mA cm−2) electrodes [429]. Dinh et al. [406] investigated the electrochemical properties of sputtered VWNz on a large silicon wafer in alkali aqueous electrolytes (Figure 10c). The electrode film showed excellent electrochemical performance with a volumetric capacitance of 700 Fcm−3 (11.5 mF cm−2) and no loss in capacitance retention after 5000 cycles.

3.2.3. Metal Carbides/Carbonitrides

Transition metal carbides/carbonitrides (MXene) are an emerging class of transition metal-based electrode materials for MSCs applications, having the general formula M n + 1 X n T x   n = 1 ,   2 ,   or   3 , where M is an early transition metal (e.g., Sc, Ti, Zr, Hf, V, Nb, Ta, Cr, and Mo), X denotes carbon or nitrogen, and T represents the surface terminations or functional groups (=O, –OH, and/or –F), and x is the number of functional termination groups [430,431,432,433]. The unique structures exhibited by MXenes with the inner conductive transition metal carbide layer enable efficient electron transportation, and the hydrophilicity rendered by the transitional metal oxide-like surface functional groups can act as active sites for fast redox reactions and also influence the Fermi level density of the states, thereby electronic properties [432,434]. MXenes are particularly promising because the 2D nature of the nanosheets can facilitate the intercalation of electrolyte ions between adjacent layers and, in the meantime, shorten the ion diffusion paths between positive and negative electrodes [433,434]. Several groups have reported successful synthesis of MXene thin films for MSC applications [434,435,436,437].
Titanium carbide MXene (Ti3C2Tx) has leapt to the forefront and become the most widely studied among the successfully realized MXene family for MSCs applications. This is due to its ultrahigh conductivity (2.4 × 105 S cm−1) [438], good electrochemical properties, having demonstrated an outstanding volumetric capacitance of ≈1500 F cm−3 with good rate capability in acidic media [439], and excellent mechanical properties [430]. The Gogotsi group [231] unraveled the charge mechanism exhibited by MXenein aqueous media and observed a change in the interplanar spacing between Ti3C2Tx layers, resulting in electrode deformations, i.e., expansion/contraction, which was attributed to the occurrence of spontaneous intercalation of some cations (Li+, Na+, Mg2+, K+, NH4+, and Al3+) when MXene samples are exposed to aqueous electrolytes during the cycling process. Electrochemical in situ XAS measurements further detected changes in the Ti oxidation state during cycling, confirming that the electrochemical behavior of Ti3C2Tx is predominantly pseudocapacitive [440]. The group asserted that the spontaneous ion intercalation provides access to electrochemically active transition metal oxide surfaces, with the Ti oxidation state in Ti3C2Tx being much closer to +2 than +4, unlike the case in TiO2 and cubic TiC, while the conductive carbide layer ensures rapid charge transfer. A summary of the electrochemical performance of recently exploited MXene-based on-chip MSCs is presented in Table 5.
Ti3C2Tx electrodes are usually prepared via wet etching [434], which is largely compatible with microfabrication protocols and offers a great opportunity for implementation in on-chip MSC applications. MXene-based on-chip MSCs can be achieved through various patterning techniques, either by direct patterning, such as laser scribing or reactive ion etching, of MXene solid on the desired substrate [441,442] or by printing of the MXene ink [430,433]. MXene (Ti3C2Tx) electrode materials have demonstrated good electrochemical properties in aqueous media.
Table 5. MXene electrodes for on-chip MSCs.
Table 5. MXene electrodes for on-chip MSCs.
Electrode MaterialSubstrateSynthesis/FabricationThickness (nm)ElectrolyteAreal Cap.
mF cm−2
Vol. Cap
F cm−3
Energy Density
mWh cm−3
Power Density
W cm−3
ConfigurationRef.
Ti3C2TxFree-standingEtching20,0001 M KOH680340NANASingle electrode[231]
Ti3C2TxFree-standingEtching12,000PVA-KOH hydrogel636530NANASingle electrode[438]
Ti3C2TxPaperScreen printing20,000PVA-H2SO4 hydrogel1108---Intedigitated[431]
Ti3C2Tx3D LIGdirect CO2 laser-scribing/lift-off lithographyNA2 M H2SO41348---Intedigitated[230]
Ti3C2TxPET3D PrintingNAPVA-H2SO4 hydrogel2337---Intedigitated[443]
Ti3C2TxGlassExtrusion 3D PrintingNAPVA-H2SO4 hydrogel1035-10.005Intedigitated[432]
Ti3C2TxGlassSpray coating/laser cutting1300PVA-H2SO4 hydrogel27.3356.81815Intedigitated[441]
Ti3C2TxPETStamping690PVA-H2SO4 hydrogel6188444.290Intedigitated[444]
Ti3C2TxPETCO2 laser machining + spray coating4000PVA-H3PO4 hydrogel2357.52.80.744Intedigitated[3]
Ti3C2Tx-GraphenePETSpray coating/Mask-assisted filtration1000PVA-H3PO4 hydrogel3.26333.40.2Intedigitated[445]
Ti3C2TxPaperVacuum-assisted filtration/laser printing2200PVA-H2SO4 hydrogel27.29124.056.10.85Intedigitated[446]
Ti3C2Tx-RuO2PaperScreen printing270PVA–KOH hydrogel23.3864.213.548.5Intedigitated[433]
Ti3C2TxPET/Au/PDMSVacuum-assisted filtration/Laser marking4000PVA-H2SO4 hydrogel7318312.44.38Intedigitated[447]
Ti3C2TxGlassDip coating/scalpel engraving150PVA-H3PO4 hydrogel0.28318.90.676.67Intedigitated[435]
Ti3C2Tx-grapheneSi-SiO2Scratch method5000PVA-H3PO4 hydrogel18.236.42.30.16Intedigitated[448]
Ti3C2Tx-sodium alginate-Fe2+SiEtching/inkjet-printing730PVA-H3PO4 hydrogel123.81696854.6Intedigitated[449]
Jiang et al. [450] fabricated wafer-scale on-chip MXene MSCs for AC-Line filtering applications using a conventional photolithographic lift-off process with 100 nm-thick interdigitated electrodes (Figure 11a) to deliver a volumetric capacitance of 30 F cm−3 at 120 Hz and a very short relaxation time constant of 0.45 ms, surpassing conventional electrolytic capacitors of 0.8 ms. The fabricated device is capable of filtering 120 Hz ripples produced by AC line power at a frequency of 60 Hz.
In a recent study, 3D MXene symmetric MSCs were fabricated using printing techniques [451]. The study leveraged the synergy of a high-voltage “water-in-LiBr” (WiB) gel electrolyte, which has a high voltage window of 1.8 V, and 3D-printed microelectrodes to achieve an ultrahigh areal energy density of 1772 μWh cm−2. However, like other 2D layered materials, MXene nanosheets are prone to self-restacking and agglomeration during electrode microfabrication due to the build-up of van der Waals forces between neighboring nanosheets, leading to a loss of electrochemically active surface to be accessed by electrolyte ions and thus hampering the full utilization of electrochemical reactions [433,452,453]. One of the measures considered to prevent self-restacking is the development of MXene composites by introducing redox-active (such as MnO2 [454,455] or RuO2 [433]) or carbon-based [445,456] materials as molecular spacers between the layers to keep the MXene nanosheets separated within the electrodes, thus increasing the accessible electroactive sites and greatly improving the accessibility of electrolyte ions, thereby enhancing the electrochemical performance [433,453,453,457]. Another strategy is the modification of the fabrication process to prevent restacking [431].
Li et al. [433] fabricated MSCs based on Ti3C2Tx-RuO2 composite electrode by screen-printing technique to achieve volumetric capacitance of 864.2 F cm−3 at 1 mV s−1 with PVA–KOH gel electrolyte, having good capacitance retention of 90% after 10,000 cycles and a good rate capability of 304 F cm−3 at 2 V s−1. The device achieved a remarkable energy density of 13.5 mWh cm−3 and a power density of 48.5 W cm−3. Qin et al. [458] fabricated interdigitated MSCs by conventional photolithography using interconnected porous PEDOT-MXene composite electrodes prepared by electrochemical polymerization (EP). The MSC demonstrated an areal capacitance of 47.4 mF cm−2 and a high volumetric energy density of 20.05 mWh cm−3. In another study, Kim et al. [453], fabricated interdigitated MSCs by electron beam lithography and photolithography using Ti3C2Tx-CNTs composite electrodes to achieve a high areal capacitance of ~317 mF cm−2 at a scan rate of 50 mV s−1. Ma et al. [430] fabricated interdigitated MXene-polymer (MXene/poly(3,4-ethylenedioxyth iophene):poly(styrenesulfonic acid) (MXene/poly(3,4-ethylenedioxyth iophene):poly(styrenesulfonic acid) (MP) composite-based MSCs via inkjet printing to deliver a high volumetric capacitance of 754 F cm−3 and an energy density of 9.4 mWh cm−3 and are considered one of the best performing reported inkjet-printed MSCs to date [179,459,460,461]. The interdigitated MXene-based MSCs fabricated on a flexible substrate by Zheng et al. [433] via screen printing technique exhibit a remarkably high areal capacitance of 1.1 F cm−2 and an energy density of 13.8 μWh cm−2 with an aqueous H2SO4-PVA gel electrolyte, both of which are considered one of the highest values recorded for MXene-based MSCs.
Furthermore, MXene electrodes have also been combined with other (pseudo)capacitive electrodes in an asymmetric configuration to deliver higher energy density [134,136]. An example was demonstrated by Couly et al. [136] (Figure 11b), combining Ti3C2Tx and rGO in a fabricated asymmetric MSC operated at a 1 V voltage window to deliver an energy density of 8.6 mWh cm−3 at a power density of 0.2 W cm−3, while retaining 97% of the initial capacitance after 10,000 cycles. Another study combined Ti3C2Tx and RuO2 to increase the cell voltage of symmetric MXene MSCs to about twice their operating voltage window (as shown in Figure 11c). The asymmetric MSCs operated at a voltage window of 1.5 V to deliver an energy density of 37 µW h cm−2 at a power density of 40 mW cm−2, with 86% capacitance retention after 20,000 charge–discharge cycles [136].
Zhu et al. [175] took advantage of the synergistic effect of 2D MXene and highly conductive rGO nanosheets to produce an interdigitated laser scribed graphene LSG-MXene (LrGO-MXene) composite electrode (thickness of ~4 μm coated with PVA-H3PO4 gel electrolyte) on polyethylene terephthalate (PET) substrate using the laser scribing technique. The device demonstrates outstanding areal capacitance of 2.58 mF cm−2 with cycling stability, maintaining > 97.7% of its initial value after 10,000 cycles, and at a power density of 6.2 μW cm−2, it attained a maximum energy density of 170.67 μWh cm−2.

3.2.4. Conducting Polymers

Conducting polymers are equally receiving attention, but a major drawback is their solution-based synthesis. However, 3D conducting polymers prepared through electrodeposition techniques have been reported for on-chip implementation. In addition, the pseudocapacitive properties of the conducting polymers have been explored with other (pseudo)capacitive materials in the form of composite electrodes for MSC applications [462]. Commonly explored conducting polymers for MSC applications include polyaniline (PANI) [463], poly(3,4-ethylene dioxythiophene) (PEDOT) [226], and polypyrrole (PPy) polypyrrole (PPy) [227].

4. Electrolytes

The importance of the electrolytes’ role in achieving high-performance MSCs cannot be overemphasized; they are the media that enable the transport of ions in between two electrodes [464]. The conduction of the ions determines the rate at which energy stored in the electrodes can be delivered [464]. Most electrolytes are solutions consisting of salts dissolved in solvents, either water (aqueous) or organic molecules (nonaqueous), and are in a liquid state in the service temperature range [464]. The electrolyte plays a crucial role in transferring charges and balancing charges between the two electrodes [47,465]. It has been reported that developing high-performance active electrode materials is extremely important for increasing the energy and power density of MSCs. Another important factor to consider in improving the specific energy and power, long cycling life, and safety of MSCs is the electrolyte, which determines the final operating voltage of the MSCs [47,465]. The electrode–electrolyte interaction is critically considered in all electrochemical processes, as it influences the electrode–electrolyte interface and the internal structure of active materials. The electrolytes can be classified into various categories, as shown in Figure 12. The details on the electrolyte conductivity, dissociation, and ion transport properties have been reported elsewhere [464,465]. Here, we will briefly mention the most frequently used electrolytes for MSC applications.

4.1. Aqueous Electrolytes (Alkaline, Neutral, and Acidic)

Aqueous electrolytes are known for their ease of handling and exhibit higher conductivity than organic and ionic electrolytes, which is beneficial for lowering the equivalent series resistance (ESR) and improving the power delivery of MSCs [276,465,469]. Aqueous electrolytes fall into three categories: acidic, neutral, and alkaline solutions. However, the narrow potential window, limited to around 1 V due to the water splitting at 1.23 V, is a major issue in achieving higher energy density. The commonly used aqueous electrolytes are [465] KOH, NaOH, LiOH, Na2SO4, H2SO4, (NH4)2SO4, K2SO4, Li2SO4, MgSO4, H3PO4,CaSO4, BaSO4, KCl, NaCl, LiCl, HCl, CsCl, CaCl2, KNO3, LiNO3, Na(CH3COO), Li(CH3-COO), Mg(CH3COO)2, Na2HPO4, NaHCO3, Na2B4O7, Li-SiW, NaSiW, and K-SiW.
The effects of various electrolyte cations (H+, Li+, Na+, K+) on the electrochemical performance of MSCs are extensively reviewed by Pal et al., Zhu et al., and Xu et al. [464,465,470]. It has been observed that the hydrated cationic radius influences conductivity and ionic mobility; the hydrated cationic radius is in the order H+-H2Oδ−< K+-H2Oδ− < Na+-H2Oδ− < Li+-H2Oδ−, as observed by Zhu et al. [470]. The smaller cationic radius gives rise to higher conductivity and ionic mobility, and helps in fast charge transfer, offering more ion adsorption at the electrolyte/electrode interface to further facilitate the fast Faraday reaction [465,470]. The specific capacitance is significantly influenced by cationic mobility, hydrated cationic radius, conductivity, charge/ion exchange, and diffusion [465]. It has also been observed that long-term cycling stability is greatly influenced by the type of cationic species present [470]. In addition, another key player in electrochemical performance is the electrolyte anions (OH, SO 4 2 , Cl, and NO 3 ). Higher conductivity and ionic mobility arise from a smaller ionic radius; an investigation by Wu et al. [471] shows that the hydrated ionic radius is in the order OH < Cl < NO 3 < SO 4 2 ; thus, higher conductivity and ionic mobility anions lead to better capacitive behavior.

4.2. Organic Electrolytes

Organic electrolytes typically consist of organic solvents and conducting salts dissolved in them; they are attractive due to the high voltage window (2.6 to 2.9 V). However, high resistivity is a significant disadvantage due to the large size of molecules, which require a large pore size in the electrodes. Much like aqueous electrolyte-based MSCs, the properties of solvents and salts, such as ion size, ion-solvent interaction, conductivity, and viscosity, significantly impact the performance of MSCs. A comprehensive review of the organic electrolytes, reaction mechanism, and effect of the ionic size of various organic electrolytes used in EDLC and pseudocapacitive devices has been reported elsewhere [465,472].

4.3. Ionic Liquids

Ionic liquids (ILs) are salts of organic cations and organic/inorganic anions with melting points below 100 °C, exhibiting unique structures and properties that make them attractive as alternative electrolytes [473,474]. The physical and chemical properties of the ILs can be tuned through various combinations of cations and anions, and the electrochemical performance of the (pseudo)capacitive electrodes can also be enhanced by tailoring the voltage window and working temperature range [473]. They have potential benefits, including non-flammability, high thermal and electrochemical stability, a high voltage window (>3 V), and low volatility compared to organic electrolytes [465,473]. Commonly used ILs in supercapacitors are based on ammonium, sulfonium, imidazolium, pyrrolidinium, and phosphonium cations and hexafluorophosphate (PF6), tetrafluoroborate (BF4), bis(trifluoromethane sulfonyl)imide (TFSI), bis(fluorosulfonyl)imide (FSI), and dicyanamide (DCA) anions [465,473,475,476,477,478]. More insight on the ILs for MSCs can be found in refs [465,473].

4.4. Solid-State Electrolytes

Solid-state electrolytes are usually based on polymer electrolytes and have attracted considerable attention in recent years due to the growing demand for on-chip energy storage microdevices, such as MSCs, which leverage their advantages over other electrolytes, such as high ionic conductivity, simple packaging, and leakage-free [465,479,480,481,482]. They are typically divided into three types: solid polymer electrolytes (SPEs), gel polymer electrolytes (GPEs) (also called quasi-solid-state electrolytes), and polyelectrolytes. Gel polymer electrolytes have the highest ionic conductivity among them owing to the presence of a liquid phase, making them the preferred choice for MSC applications [465]. On the other hand, the use of solid polymer electrolytes (SPEs) is limited since they have relatively poor ionic conductivity [482]. However, the high ionic conductivity of gel polymer electrolytes comes with a drawback—they may suffer from narrow operative temperatures due to the presence of water and relatively poor mechanical strength [465,483].
The gel polymer electrolyte is a combination of a polymer host (PEO, PEG, PVA, etc.) and aqueous electrolytes (KOH, H2SO4, K2SO4, etc.) or a conducting salt dissolved in a solvent [484]. In addition, various polymer hosts have been explored for preparing gel polymer electrolytes such as poly(acrylic acid) (PAA), poly(vinyl alcohol) (PVA), poly(ethyl oxide) (PEO), potassium polyacrylate (PAAK), poly(ether ether ketone) (PEEK), poly-(methylmethacrylate) (PMMA), poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP), and poly(acrylonitrile)-blockpoly(ethylene glycol)-block-poly(acrylonitrile) (PAN-b-PEG-b-PAN) [465,483]. Hydrogel polymer electrolyte is a gel polymer electrolyte due to the use of water as a plasticizer, involving the combination of aqueous electrolyte and poly host; on the other hand, polymers play host for ionic electrolytes in ionogel electrolytes, aiming to improve the thermal stability and the electrochemically stable potential window [465,484].

4.5. Redox-Active Electrolytes

Redox-active electrolytes are a new class of electrolytes developed to enhance the capacitance of pseudocapacitors through a redox mediator in the electrolyte [146,469,472]. Adding electroactive materials to conventional electrolytes aims to increase their electrochemical performance. More on the redox-active electrolytes can be found in references [146,465,468].

5. Conclusions and Perspectives

In recent years, the development of capacitive materials for on-chip power sources has witnessed significant progress, driven by the demand for miniaturized energy storage solutions with high performance and integration capabilities. This review explores various dielectric materials for electrostatic (micro-/nano-) capacitors, including emerging 2D dielectric and 3D nanostructured materials. It underlines the importance of dielectric materials in influencing the energy density and operational stability of electrostatic (micro-nano-) capacitors. Emerging dielectric materials exhibit unique electrical properties, such as high carrier mobility and tunable bandgaps, which hold promise for next-generation nanocapacitors with enhanced performance and functionality.
Further, various emerging electrode materials for electrochemical micro-supercapacitors (MSCs), including carbon-based materials, transition metal-based materials, and conducting polymers, were examined. Recent advancements in microfabrication processes, doping, and electrode configuration strategies have enhanced electrode materials’ performance, paving the way for improved on-chip MSCs. In addition, the preparation of various capacitive materials with a focus on compatibility with on-chip fabrication processes was also examined.
The future development of on-chip energy storage devices, particularly MSCs is poised to significantly impact various industries, notably in the realms of IoT and AI. As technology advances, they are expected to undergo rapid developments in several key areas to enhance their performance, integration capabilities, and application scope. Significant progress is expected in the areas of:
-
2D materials like graphene, MXenes, and transition metal dichalcogenides (TMDs) have a high surface area, excellent electrical conductivity, and tunable properties, affording them the leverage to provide higher energy density and faster charge/discharge cycles.
-
3D material developments using similar 3D templates to fabricate MBs, MSCs, and µ capacitors could be particularly valuable and cost-effective with a streamlined fabrication process, enhancing space efficiency and performance synergy. However, from a technological point of view, some issues may arise, including material compatibility, constraints due to the device’s configuration/topology, and fabrication challenges due to the specific requirements.
-
Surface area enhancements/topology
-
Solid-state electrolytes to offer enhanced safety, ionic conductivity, stability, a wider voltage window to improve energy density, and compatibility with on-chip integration.
-
Cycle life improvements by enhancing the active material properties.
-
Materials preparation techniques compatible with MEMS technology
There are several potential areas for future research and development in electrode materials for on-chip MSCs. One promising approach is exploring ion implantation strategies, which could lead to improved energy storage performance. These materials have unique properties that could be utilized to overcome current limitations and enable new functionalities in on-chip MSCs. Furthermore, it is essential to develop scalable and cost-effective preparation techniques for electrode materials to facilitate their large-scale production and commercialization. To address compatibility challenges and speed up the translation of research findings into practical applications, innovative approaches such as vacuum deposition techniques should be prioritized.
It is important to tailor the capacitive materials and device architectures to meet the specific requirements to unlock the full potential of the autonomous microelectronics deployed in emerging technologies like the Internet of Things (IoT) and artificial intelligence (AI). This requires continued interdisciplinary research efforts in areas such as materials science, electrochemistry, microfabrication, and device integration.

Author Contributions

B.J. and G.B. performed methodology, investigation, formal analysis, and visualization, wrote, reviewed, and edited the original draft, resources, and data curation. P.R. and C.L. performed resources, conceptualization, methodology, and validation, wrote, reviewed, and edited the original draft, supervision, funding acquisition, and project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work received funding from Nigerian Government through the Petroleum Technology Development Fund (PTDF)/Campus France and from the French National Research Agency (ARTEMIS project ref ANR-22-ASEN-0001-01).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yun, J.; Song, C.; Lee, H.; Park, H.; Jeong, Y.R.; Kim, J.W.; Jin, S.W.; Oh, S.Y.; Sun, L.; Zi, G.; et al. Stretchable Array of High-Performance Micro-Supercapacitors Charged with Solar Cells for Wireless Powering of an Integrated Strain Sensor. Nano Energy 2018, 49, 644–654. [Google Scholar] [CrossRef]
  2. Guo, R.; Chen, J.; Yang, B.; Liu, L.; Su, L.; Shen, B.; Yan, X.; Guo, R.S.; Chen, J.T.; Yang, B.J.; et al. In-Plane Micro-Supercapacitors for an Integrated Device on One Piece of Paper. Adv. Funct. Mater. 2017, 27, 1702394. [Google Scholar] [CrossRef]
  3. Jiang, Q.; Wu, C.; Wang, Z.; Wang, A.C.; He, J.H.; Wang, Z.L.; Alshareef, H.N. MXene Electrochemical Microsupercapacitor Integrated with Triboelectric Nanogenerator as a Wearable Self-Charging Power Unit. Nano Energy 2018, 45, 266–272. [Google Scholar] [CrossRef]
  4. Wang, X.; Yin, Y.; Yi, F.; Dai, K.; Niu, S.; Han, Y.; Zhang, Y.; You, Z. Bioinspired Stretchable Triboelectric Nanogenerator as Energy-Harvesting Skin for Self-Powered Electronics. Nano Energy 2017, 39, 429–436. [Google Scholar] [CrossRef]
  5. Sudevalayam, S.; Kulkarni, P. Energy Harvesting Sensor Nodes: Survey and Implications. IEEE Commun. Surv. Tutor. 2011, 13, 443–461. [Google Scholar] [CrossRef]
  6. Lethien, C.; Le Bideau, J.; Brousse, T. Challenges and Prospects of 3D Micro-Supercapacitors for Powering the Internet of Things. Energy Environ. Sci. 2019, 12, 96–115. [Google Scholar] [CrossRef]
  7. Shen, C.; Xu, S.; Xie, Y.; Sanghadasa, M.; Wang, X.; Lin, L. A Review of On-Chip Micro Supercapacitors for Integrated Self-Powering Systems. J. Microelectromechanical Syst. 2017, 26, 949–965. [Google Scholar] [CrossRef]
  8. Zheng, S.; Shi, X.; Das, P.; Wu, Z.S.; Bao, X. The Road Towards Planar Microbatteries and Micro-Supercapacitors: From 2D to 3D Device Geometries. Adv. Mater. 2019, 31, 1900583. [Google Scholar] [CrossRef] [PubMed]
  9. Kyeremateng, N.A.; Brousse, T.; Pech, D. Microsupercapacitors as Miniaturized Energy-Storage Components for on-Chip Electronics. Nat. Nanotechnol. 2017, 12, 7–15. [Google Scholar] [CrossRef]
  10. Spurney, R.G.; Sharma, H.; Pulugurtha, M.R.; Tummala, R.; Lollis, N.; Weaver, M.; Gandhi, S.; Romig, M.; Brumm, H. Ultra-High Density, Thin-Film Tantalum Capacitors with Improved Frequency Characteristics for MHz Switching Power Converters. J. Electron. Mater. 2018, 47, 5632–5639. [Google Scholar] [CrossRef]
  11. Liu, H.; Zhang, G.; Zheng, X.; Chen, F.; Duan, H. Emerging Miniaturized Energy Storage Devices for Microsystem Applications: From Design to Integration. Int. J. Extrem. Manuf. 2020, 2, 042001. [Google Scholar] [CrossRef]
  12. Li, Y.; Xiao, S.; Qiu, T.; Lang, X.; Tan, H.; Wang, Y.; Li, Y. Recent Advances on Energy Storage Microdevices: From Materials to Configurations. Energy Storage Mater. 2022, 45, 741–767. [Google Scholar] [CrossRef]
  13. Liu, L.; Niu, Z.; Chen, J. Design and Integration of Flexible Planar Micro-Supercapacitors. Nano Res. 2017, 10, 1524–1544. [Google Scholar] [CrossRef]
  14. Hu, H.; Pei, Z.; Ye, C. Recent Advances in Designing and Fabrication of Planar Micro-Supercapacitors for on-Chip Energy Storage. Energy Storage Mater. 2015, 1, 82–102. [Google Scholar] [CrossRef]
  15. Park, S.H.; Goodall, G.; Kim, W.S. Perspective on 3D-Designed Micro-Supercapacitors. Mater. Des. 2020, 193, 108797. [Google Scholar] [CrossRef]
  16. Qi, D.; Liu, Y.; Liu, Z.; Zhang, L.; Chen, X.; Qi, D.; Liu, Y.; Liu, Z.; Zhang, L.; Chen, X. Design of Architectures and Materials in In-Plane Micro-Supercapacitors: Current Status and Future Challenges. Adv. Mater. 2017, 29, 1602802. [Google Scholar] [CrossRef]
  17. Zhang, J.; Zhang, G.; Zhou, T.; Sun, S.; Zhang, J.; Zhou, T.; Zhang, G.; Sun, S. Recent Developments of Planar Micro-Supercapacitors: Fabrication, Properties, and Applications. Adv. Funct. Mater. 2020, 30, 1910000. [Google Scholar] [CrossRef]
  18. Wu, Z.; Li, L.; Yan, J.M.; Zhang, X.B. Materials Design and System Construction for Conventional and New-Concept Supercapacitors. Adv. Sci. 2017, 4, 1600382. [Google Scholar] [CrossRef]
  19. Song, W.J.; Yoo, S.; Song, G.; Lee, S.; Kong, M.; Rim, J.; Jeong, U.; Park, S. Recent Progress in Stretchable Batteries for Wearable Electronics. Batter. Supercaps 2019, 2, 181–199. [Google Scholar] [CrossRef]
  20. Liu, W.; Song, M.S.; Kong, B.; Cui, Y. Flexible and Stretchable Energy Storage: Recent Advances and Future Perspectives. Adv. Mater. 2017, 29, 1603436. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Guo, J. Development of Flexible Li-Ion Batteries for Flexible Electronics. InfoMat 2020, 2, 866–878. [Google Scholar] [CrossRef]
  22. Chen, D.; Lou, Z.; Jiang, K.; Shen, G. Device Configurations and Future Prospects of Flexible/Stretchable Lithium-Ion Batteries. Adv. Funct. Mater. 2018, 28, 1805596. [Google Scholar] [CrossRef]
  23. Jetybayeva, A.; Umirzakov, A.; Uzakbaiuly, B.; Bakenov, Z.; Mukanova, A. Towards Li–S Microbatteries: A Perspective Review. J. Power Sources 2023, 573, 233158. [Google Scholar] [CrossRef]
  24. Xia, Q.; Zan, F.; Zhang, Q.; Liu, W.; Li, Q.; He, Y.; Hua, J.; Liu, J.; Xu, J.; Wang, J.; et al. All-Solid-State Thin Film Lithium/Lithium-Ion Microbatteries for Powering the Internet of Things. Adv. Mater. 2023, 35, 2200538. [Google Scholar] [CrossRef] [PubMed]
  25. Hu, B.; Wang, X. Advances in Micro Lithium-Ion Batteries for on-Chip and Wearable Applications. J. Micromechanics Microengineering 2021, 31, 114002. [Google Scholar] [CrossRef]
  26. Selinis, P.; Farmakis, F. Review—A Review on the Anode and Cathode Materials for Lithium-Ion Batteries with Improved Subzero Temperature Performance. J. Electrochem. Soc. 2022, 169, 010526. [Google Scholar] [CrossRef]
  27. Yue, C.; Li, J.; Lin, L. Fabrication of Si-Based Three-Dimensional Microbatteries: A Review. Front. Mech. Eng. 2017, 12, 459–476. [Google Scholar] [CrossRef]
  28. Gao, Y.; Tenhaeff, W.E. Synthesis and Characterization of Thin Film Polyelectrolytes for Solid-State Lithium Microbatteries. J. Vac. Sci. Technol. B 2019, 37, 051401. [Google Scholar] [CrossRef]
  29. Sheil, R.; Chang, J.P. Synthesis and Integration of Thin Film Solid State Electrolytes for 3D Li-Ion Microbatteries. J. Vac. Sci. Technol. 2020, 38, 032411. [Google Scholar] [CrossRef]
  30. Jetybayeva, A.; Uzakbaiuly, B.; Mukanova, A.; Myung, S.T.; Bakenov, Z. Recent Advancements in Solid Electrolytes Integrated into All-Solid-State 2D and 3D Lithium-Ion Microbatteries. J. Mater. Chem. A. 2021, 9, 15140–15178. [Google Scholar] [CrossRef]
  31. Ni, J.; Li, L. Self-Supported 3D Array Electrodes for Sodium Microbatteries. Adv. Funct. Mater. 2018, 28, 1704880. [Google Scholar] [CrossRef]
  32. Fang, Z.; Wang, J.; Wu, H.; Li, Q.; Fan, S.; Wang, J. Progress and Challenges of Flexible Lithium Ion Batteries. J. Power Sources 2020, 454, 227932. [Google Scholar] [CrossRef]
  33. Radzuan, R.; Salleh, M.K.M.; Hamzah, M.K.; Zulkefle, H.; Md Sin, N.D.; Rusop, M. Development of Thin Film Capacitors for Power System Applications by PVD Technique. In Proceedings of the SHUSER 2012—2012 IEEE Symposium on Humanities, Science and Engineering Research, Kuala Lumpur, Malaysia, 24–27 June 2012; IEEE: New York, NY, USA, 2012; pp. 1097–1100. [Google Scholar] [CrossRef]
  34. Nguyen, P.; Ng, H.T.; Yamada, Y.; Smith, M.K.; Li, J.; Han, J.; Meyyappan, M.; Lett, N.; Yamada, T.; Chen, Y.P.; et al. High-Dielectric-Constant Silver–Epoxy Composites as Embedded Dielectrics. Adv. Mater. 2005, 17, 1777–1781. [Google Scholar] [CrossRef]
  35. Özçelik, V.O.; Ciraci, S. High-Performance Planar Nanoscale Dielectric Capacitors. Phys. Rev. B Condens. Matter Mater. Phys. 2015, 91, 195445. [Google Scholar] [CrossRef]
  36. Yao, S.; Yuan, J.; Mehedi, H.-A.; Gheeraert, E.; Sylvestre, A. Carbon Nanotube Forest Based Electrostatic Capacitor with Excellent Dielectric Performances. Carbon 2017, 116, 648–654. [Google Scholar] [CrossRef]
  37. Liang, Z.; Liu, M.; Shen, L.; Lu, L.; Ma, C.; Lu, X.; Lou, X.; Jia, C.L. All-Inorganic Flexible Embedded Thin-Film Capacitors for Dielectric Energy Storage with High Performance. ACS Appl. Mater. Interfaces 2019, 11, 5247–5255. [Google Scholar] [CrossRef]
  38. Tombak, A.; Maria, J.P.; Ayguavives, F.T.; Jin, Z.; Stauf, G.T.; Kingon, A.I.; Mortazawi, A. Voltage-Controlled RF Filters Employing Thin-Film Barium-Strontium-Titanate Tunable Capacitors. IEEE Trans. Microw. Theory Tech. 2003, 51, 462–467. [Google Scholar] [CrossRef]
  39. Vorobiev, A.; Gevorgian, S. Fabrication of Ferroelectric Components and Devices. In Ferroelectrics in Microwave Devices, Circuits and Systems: Physics, Modeling, Fabrication and Measurements; Springer: Berlin/Heidelberg, Germany, 2009; pp. 61–113. [Google Scholar]
  40. Han, K.; Wang, Q. 10.40—Polymers for Thin Film Capacitors: Energy Storage—Li Conducting Polymers. In Polymer Science: A Comprehensive Reference; Matyjaszewski, K., Möller, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; Volume 1–10, pp. 811–830. [Google Scholar]
  41. Butzen, N.; Krishnarnurthy, H.; Ahmed, Z.; Weng, S.; Ravichandran, K.; Zelikson, M.; Tschanz, J.; Douglas, J. 14.4 A Monolithic 26A/Mm2Imax, 88.5% Peak-Efficiency Continuously Scalable Conversion-Ratio Switched-Capacitor DC-DC Converter. In Proceedings of the Digest of Technical Papers—IEEE International Solid-State Circuits Conference, San Francisco, CA, USA, 19–23 February 2023; IEEE: New York, NY, USA, 2023; Volume 2023, pp. 232–234. [Google Scholar]
  42. Sharma, S.; Chand, P. Supercapacitor and Electrochemical Techniques: A Brief Review. Results Chem. 2023, 5, 100885. [Google Scholar] [CrossRef]
  43. Bonaccorso, F.; Colombo, L.; Yu, G.; Stoller, M.; Tozzini, V.; Ferrari, A.C.; Ruoff, R.S.; Pellegrini, V. Graphene, Related Two-Dimensional Crystals, and Hybrid Systems for Energy Conversion and Storage. Science 2015, 347, 1246501. [Google Scholar] [CrossRef]
  44. Chmiola, J.; Celine Largeot, P.L.T.; Simon, P.; Gogotsi, Y. Monolithic Carbide-Derived Carbon Films for Micro-Supercapacitors. Science 2010, 328, 480–483. [Google Scholar] [CrossRef]
  45. Dinh, K.H.; Roussel, P.; Lethien, C. Advances on Microsupercapacitors: Real Fast Miniaturized Devices toward Technological Dreams for Powering Embedded Electronics? ACS Omega 2022, 8, 8990. [Google Scholar] [CrossRef] [PubMed]
  46. Huang, C.; Grant, P.S. One-Step Spray Processing of High Power All-Solid-State Supercapacitors. Sci. Rep. 2013, 3, 2393. [Google Scholar] [CrossRef] [PubMed]
  47. Pandolfo, T.; Ruiz, V.; Sivakkumar, S.; Nerkar, J. General Properties of Electrochemical Capacitors. Supercapacitors Mater. Syst. Appl. 2013, 69–109. [Google Scholar] [CrossRef]
  48. Da Silva, L.M.; Cesar, R.; Moreira, C.M.R.; Santos, J.H.M.; De Souza, L.G.; Pires, B.M.; Vicentini, R.; Nunes, W.; Zanin, H. Reviewing the Fundamentals of Supercapacitors and the Difficulties Involving the Analysis of the Electrochemical Findings Obtained for Porous Electrode Materials. Energy Storage Mater. 2020, 27, 555–590. [Google Scholar] [CrossRef]
  49. Ceraolo, M.; Lutzemberger, G.; Poli, D. State-Of-Charge Evaluation Of Supercapacitors. J. Energy Storage 2017, 11, 211–218. [Google Scholar] [CrossRef]
  50. Saha, P.; Dey, S.; Khanra, M. Accurate Estimation of State-of-Charge of Supercapacitor under Uncertain Leakage and Open Circuit Voltage Map. J. Power Sources 2019, 434, 226696. [Google Scholar] [CrossRef]
  51. Gogotsi, Y.; Simon, P. True Performance Metrics in Electrochemical Energy Storage. Science 2011, 334, 917–918. [Google Scholar] [CrossRef]
  52. Burke, A. Ultracapacitors: Why, How, and Where Is the Technology. J. Power Sources 2000, 91, 37–50. [Google Scholar] [CrossRef]
  53. Dinh Khac, H.; Whang, G.; Iadecola, A.; Makhlouf, H.; Barnabé, A.; Teurtrie, A.; Marinova, M.; Huvé, M.; Roch-Jeune, I.; Douard, C.; et al. Nanofeather Ruthenium Nitride Electrodes for Electrochemical Capacitors. Nat. Mater. 2024, 23, 670–679. [Google Scholar] [CrossRef]
  54. Huang, P.; Lethien, C.; Pinaud, S.; Brousse, K.; Laloo, R.; Turq, V.; Respaud, M.; Demortière, A.; Daffos, B.; Taberna, P.L.; et al. On-Chip and Freestanding Elastic Carbon Films for Micro-Supercapacitors. Science 2016, 351, 691–695. [Google Scholar] [CrossRef]
  55. Kim, M.S.; Hsia, B.; Carraro, C.; Maboudian, R. Flexible Micro-Supercapacitors with High Energy Density from Simple Transfer of Photoresist-Derived Porous Carbon Electrodes. Carbon 2014, 74, 163–169. [Google Scholar] [CrossRef]
  56. Beidaghi, M.; Gogotsi, Y. Capacitive Energy Storage in Micro-Scale Devices: Recent Advances in Design and Fabrication of Micro-Supercapacitors. Energy Environ. Sci. 2014, 7, 867–884. [Google Scholar] [CrossRef]
  57. Chen, F.; Xu, Z.-L. Design and Manufacture of High-Performance Microbatteries: Lithium and Beyond. Microstructures 2022, 2, 2022012. [Google Scholar] [CrossRef]
  58. Zhu, Z.; Kan, R.; Hu, S.; He, L.; Hong, X.; Tang, H.; Luo, W. Recent Advances in High-Performance Microbatteries: Construction, Application, and Perspective. Small 2020, 16, 2003251. [Google Scholar] [CrossRef] [PubMed]
  59. Robert, K.; Douard, C.; Demortière, A.; Blanchard, F.; Roussel, P.; Brousse, T.; Lethien, C. On Chip Interdigitated Micro-Supercapacitors Based on Sputtered Bifunctional Vanadium Nitride Thin Films with Finely Tuned Inter- and Intracolumnar Porosities. Adv. Mater. Technol. 2018, 3, 1800036. [Google Scholar] [CrossRef]
  60. 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]
  61. Manoccio, M.; Esposito, M.; Passaseo, A.; Cuscunà, M.; Tasco, V. Focused Ion Beam Processing for 3d Chiral Photonics Nanostructures. Micromachines 2021, 12, 6. [Google Scholar] [CrossRef]
  62. Zhang, P.; Wang, F.; Yu, M.; Zhuang, X.; Feng, X. Two-Dimensional Materials for Miniaturized Energy Storage Devices: From Individual Devices to Smart Integrated Systems. Chem. Soc. Rev. 2018, 47, 7426–7451. [Google Scholar] [CrossRef]
  63. Bao, Z.; Chen, X. Flexible and Stretchable Devices. Adv. Mater. 2016, 28, 4177–4179. [Google Scholar] [CrossRef]
  64. Strambini, L.; Paghi, A.; Mariani, S.; Sood, A.; Kalliomäki, J.; Järvinen, P.; Toia, F.; Scurati, M.; Morelli, M.; Lamperti, A.; et al. Three-Dimensional Silicon-Integrated Capacitor with Unprecedented Areal Capacitance for on-Chip Energy Storage. Nano Energy 2020, 68, 104281. [Google Scholar] [CrossRef]
  65. Ulaby, F.T.; Ravaioli, U. Fundamentals of Applied Electromagnetics, 8th ed.; Pearson: London, UK, 2023; ISBN 13: 978-1-292-43676-0. [Google Scholar]
  66. Wang, M.; Zhang, J.; Wang, Y.; Lu, Y. Material and Structural Design of Microsupercapacitors. J. Solid. State Electrochem. 2022, 26, 313–334. [Google Scholar] [CrossRef]
  67. Banerjee, P.; Perez, I.; Henn-Lecordier, L.; Lee, S.B.; Rubloff, G.W. Nanotubular Metal-Insulator-Metal Capacitor Arrays for Energy Storage. Nat. Nanotechnol. 2009, 4, 292–296. [Google Scholar] [CrossRef]
  68. Jayakrishnan, A.R.; Anina Anju, B.; P Nair, S.K.; Dutta, S.; Silva, J.P.B. Recent Development of Lead-Free Relaxor Ferroelectric and Antiferroelectric Thin Films as Energy Storage Dielectric Capacitors. J. Eur. Ceram. Soc. 2024, 44, 4332–4349. [Google Scholar] [CrossRef]
  69. Jayakrishnan, A.R.; Silva, J.P.B.; Kamakshi, K.; Dastan, D.; Annapureddy, V.; Pereira, M.; Sekhar, K.C. Are Lead-Free Relaxor Ferroelectric Materials the Most Promising Candidates for Energy Storage Capacitors? Prog. Mater. Sci. 2023, 132, 101046. [Google Scholar] [CrossRef]
  70. Shim, J.H.; Choi, H.J.; Kim, Y.; Torgersen, J.; An, J.; Lee, M.H.; Prinz, F.B. Process-Property Relationship in High-: K ALD SrTiO3 and BaTiO3: A Review. J. Mater. Chem. C Mater. 2017, 5, 8000–8013. [Google Scholar] [CrossRef]
  71. Go, D.; Shin, J.W.; Lee, S.; Lee, J.; Yang, B.C.; Won, Y.; Motoyama, M.; An, J. Atomic Layer Deposition for Thin Film Solid-State Battery and Capacitor. Int. J. Precis. Eng. Manuf. -Green. Technol. 2023, 10, 851–873. [Google Scholar] [CrossRef]
  72. Klootwijk, J.; Kemmeren, A.; Wolters, R.; Roozeboom, F.; Verhoeven, J.; Heuvel, E. van Den Extremely High-Density Capacitors with ALD High-K Dielectric Layers. In Defects in High-k Gate Dielectric Stacks; Springer: Heidelberg, Germany, 2006; pp. 17–28. [Google Scholar]
  73. Pan, H.; Li, F.; Liu, Y.; Zhang, Q.; Wang, M.; Lan, S.; Zheng, Y.; Ma, J.; Gu, L.; Shen, Y.; et al. Ultrahigh—Energy Density Lead-Free Dielectric Films via Polymorphic Nanodomain Design. Science 2019, 365, 578–582. [Google Scholar] [CrossRef]
  74. Liu, Y.; Han, H.; Pan, H.; Lan, S.; Lin, Y.; Ma, J. Enhancement of Breakdown Strength in Relaxor Ferroelectric Ba(Zr0.3Ti0.7)O3 Thin Film via Manipulating Growth Oxygen Pressure. J. Alloys Compd. 2023, 937, 168452. [Google Scholar] [CrossRef]
  75. Lv, P.; Yang, C.; Qian, J.; Wu, H.; Huang, S.; Cheng, X.; Cheng, Z. Flexible Lead-Free Perovskite Oxide Multilayer Film Capacitor Based on (Na0.8K0.2)0.5Bi0.5TiO3/Ba0.5Sr0.5(Ti0.97Mn0.03)O3 for High-Performance Dielectric Energy Storage. Adv. Energy Mater. 2020, 10, 1904229. [Google Scholar] [CrossRef]
  76. Lee, S.W.; Han, J.H.; Kwon, O.S.; Hwang, C.S. Influences of a Crystalline Seed Layer during Atomic Layer Deposition of SrTiO3 Thin Films Using Ti(O-IPr)2(Thd)2, Sr(Thd)2, and H2O. J. Electrochem. Soc. 2008, 155, G253. [Google Scholar] [CrossRef]
  77. Robertson, J.; Wallace, R.M. High-K Materials and Metal Gates for CMOS Applications. Mater. Sci. Eng. R Rep. 2015, 88, 1–41. [Google Scholar] [CrossRef]
  78. Jeon, W. Recent Advances in the Understanding of High-k Dielectric Materials Deposited by Atomic Layer Deposition for Dynamic Random-Access Memory Capacitor Applications. J. Mater. Res. 2020, 35, 775–794. [Google Scholar] [CrossRef]
  79. Smitha, P.S.; Babu, V.S.; Shiny, G. Critical Parameters of High Performance Metal-Insulator-Metal Nanocapacitors: A Review. Mater. Res. Express 2019, 6, 122003. [Google Scholar] [CrossRef]
  80. Mondal, S.; Pan, T.M. High-Performance Ni/Lu2O3/TaN Metal-Insulator-Metal Capacitors. IEEE Electron. Device Lett. 2011, 32, 1576–1578. [Google Scholar] [CrossRef]
  81. Austin, D.Z.; Allman, D.; Price, D.; Hose, S.; Conley, J.F. Plasma Enhanced Atomic Layer Deposition of Al2O3/SiO2 MIM Capacitors. IEEE Electron. Device Lett. 2015, 36, 496–498. [Google Scholar] [CrossRef]
  82. Ishikawa, T.; Kodama, D.; Matsui, Y.; Hiratani, M.; Furusawa, T.; Hisamoto, D. High-Capacitance Cu/Ta2O5/Cu MIM Structure for SoC Applications Featuring a Single-Mask Add-on Process. In Proceedings of the Technical Digest—International Electron Devices Meeting, San Francisco, CA, USA, 8–11 December 2002; IEEE: New York, NY, USA, 2002; pp. 940–942. [Google Scholar]
  83. Kim, K.D.; Park, M.H.; Kim, H.J.; Kim, Y.J.; Moon, T.; Lee, Y.H.; Hyun, S.D.; Gwon, T.; Hwang, C.S. Ferroelectricity in Undoped-HfO2 Thin Films Induced by Deposition Temperature Control during Atomic Layer Deposition. J. Mater. Chem. C Mater. 2016, 4, 6864–6872. [Google Scholar] [CrossRef]
  84. Kambale, K.R.; Mahajan, A.; Butee, S.P.; Kulkarni, A.R.; Venkataramani, N. Effect of Sm2O3 Addition on the Dielectric Behaviour of BaTiO3. Mater. Today Proc. 2022, 67, 404–409. [Google Scholar] [CrossRef]
  85. Khomenkova, L.; Merabet, H.; Chauvat, M.P.; Frilay, C.; Portier, X.; Labbe, C.; Marie, P.; Cardin, J.; Boudin, S.; Rueff, J.M.; et al. Comprehensive Investigation of Er2O3 Thin Films Grown with Different ALD Approaches. Surf. Interfaces 2022, 34, 102377. [Google Scholar] [CrossRef]
  86. Singh, E.R.; Alam, M.W.; Singh, N.K. Capacitive and RRAM Forming-Free Memory Behavior of Electron-Beam Deposited Ta2O5 Thin Film for Nonvolatile Memory Application. ACS Appl. Electron. Mater. 2023, 5, 3462–3469. [Google Scholar] [CrossRef]
  87. Nam, M.; Lee, S.; Jeong, H.; Yoon, A.; Park, J.S.; Jeon, W. Optimizing the Crystallinity of a ZrO2 Thin Film Insulator for InGaZnO-Based Metal–Insulator–Semiconductor Capacitors. Adv. Mater. Interfaces 2024, 11, 2300883. [Google Scholar] [CrossRef]
  88. Lee, D.K.; Kim, H.B.; Kwon, S.H.; Ahn, J.H. Enhanced Electrical Properties of ZrO2-TiN Based Capacitors by Introducing Ultrathin Metal Oxides. Mater. Lett. 2020, 279, 128490. [Google Scholar] [CrossRef]
  89. Jeong, J.; Han, Y.; Sohn, H. Effect of La Doping on Dielectric Constant and Tetragonality of ZrO2 Thin Films Deposited by Atomic Layer Deposition. J. Alloys Compd. 2022, 927, 166961. [Google Scholar] [CrossRef]
  90. Mehmood, F.; Alcala, R.; Vishnumurthy, P.; Xu, B.; Sachdeva, R.; Mikolajick, T.; Schroeder, U. Reliability Improvement from La2O3 Interfaces in Hf0.5Zr0.5O2-Based Ferroelectric Capacitors. Adv. Mater. Interfaces 2023, 10, 2202151. [Google Scholar] [CrossRef]
  91. Clark, R.D. Emerging Applications for High K Materials in VLSI Technology. Materials 2014, 7, 2913–2944. [Google Scholar] [CrossRef]
  92. Ho, C.S.; Chang, S.J.; Chen, S.C.; Liou, J.J.; Li, H. A Reliable Si3N4/Al2O3-HfO2 Stack MIM Capacitor for High-Voltage Analog Applications. IEEE Trans. Electron. Devices 2014, 61, 2944–2949. [Google Scholar] [CrossRef]
  93. Patil, S.R.; Barhate, V.N.; Patil, V.S.; Agrawal, K.S.; Mahajan, A.M. The Effect of Post-Deposition Annealing on the Chemical, Structural and Electrical Properties of Al/ZrO2/La2O3/ZrO2/Al High-k Nanolaminated MIM Capacitors. J. Mater. Sci. Mater. Electron. 2022, 33, 11227–11235. [Google Scholar] [CrossRef]
  94. Patil, S.R.; Borokar, V.Y.; Rasadujjaman, M.; Zhang, J.; Ding, S.J.; Mahajan, A.M. Investigation of PEALD ZrO2/La2O3-Based High-k Nanolaminates Sandwiched between Al and Ti Electrodes for MIM Capacitors. J. Mater. Sci. Mater. Electron. 2023, 34, 1284. [Google Scholar] [CrossRef]
  95. Li, S.; Lin, Y.; Li, G.; Yu, H.; Tang, S.; Wu, Y.; Li, X.; Tian, W. Improved Dielectric Properties of La2O3–ZrO2 Bilayer Films for Novel Gate Dielectrics. Vacuum 2020, 178, 109448. [Google Scholar] [CrossRef]
  96. Ding, S.J.; Hu, H.; Lim, H.F.; Kim, S.J.; Yu, X.F.; Zhu, C.; Li, M.F.; Cho, B.J.; Chan, D.S.H.; Rustagi, S.C.; et al. High-Performance MIM Capacitor Using ALD High-κ HfO 2-Al2O3 Laminate Dielectrics. IEEE Electron. Device Lett. 2003, 24, 730–732. [Google Scholar] [CrossRef]
  97. Jeannot, S.; Bajolet, A.; Manceau, J.P.; Crémer, S.; Deloffre, E.; Oddou, J.P.; Perrot, C.; Benoit, D.; Richard, C.; Bouillon, P.; et al. Toward next High Performances MIM Generation: Up to 30fF/Μm2 with 3D Architecture and High-k Materials. In Proceedings of the 2007 IEEE Technical Digest—International Electron Devices Meeting International Electron Devices Meeting, Washington, DC, USA, 10–12 December 2007; IEEE: New York, NY, USA, 2007; pp. 997–1000. [Google Scholar] [CrossRef]
  98. Chang, P.; Xie, Y. Evaluation of HfO-Based Ferroelectric Resonant Tunnel Junction by Band Engineering. IEEE Electron. Device Lett. 2023, 44, 168–171. [Google Scholar] [CrossRef]
  99. Xiong, L.; Hu, J.; Yang, Z.; Li, X.; Zhang, H.; Zhang, G. Dielectric Properties Investigation of Metal–Insulator–Metal (MIM) Capacitors. Molecules 2022, 27, 3951. [Google Scholar] [CrossRef] [PubMed]
  100. Lee, S.K.; Kim, K.S.; Kim, S.W.; Lee, D.J.; Park, S.J.; Kim, S. Characterizing Voltage Linearity and Leakage Current of High Density Al2O3/HfO2/Al2/O3 MIM Capacitors. IEEE Electron. Device Lett. 2011, 32, 384–386. [Google Scholar] [CrossRef]
  101. Park, M.H.; Kim, H.J.; Kim, Y.J.; Moon, T.; Kim, K.D.; Hwang, C.S. Thin HfxZr1-XO2 Films: A New Lead-Free System for Electrostatic Supercapacitors with Large Energy Storage Density and Robust Thermal Stability. Adv. Energy Mater. 2014, 4, 1400610. [Google Scholar] [CrossRef]
  102. Kozodaev, M.G.; Chernikova, A.G.; Khakimov, R.R.; Park, M.H.; Markeev, A.M.; Hwang, C.S. La-Doped Hf0.5Zr0.5O2 Thin Films for High-Efficiency Electrostatic Supercapacitors. Appl. Phys. Lett. 2018, 113, 123902. [Google Scholar] [CrossRef]
  103. Pan, H.; Lan, S.; Xu, S.; Zhang, Q.; Yao, H.; Liu, Y.; Meng, F.; Guo, E.J.; Gu, L.; Yi, D.; et al. Ultrahigh Energy Storage in Superparaelectric Relaxor Ferroelectrics. Science 2021, 374, 100–104. [Google Scholar] [CrossRef]
  104. Kursumovic, A.; Li, W.W.; Cho, S.; Curran, P.J.; Tjhe, D.H.L.; MacManus-Driscoll, J.L. Lead-Free Relaxor Thin Films with Huge Energy Density and Low Loss for High Temperature Applications. Nano Energy 2020, 71, 104536. [Google Scholar] [CrossRef]
  105. Fan, Y.; Zhou, Z.; Chen, Y.; Huang, W.; Dong, X. A Novel Lead-Free and High-Performance Barium Strontium Titanate-Based Thin Film Capacitor with Ultrahigh Energy Storage Density and Giant Power Density. J. Mater. Chem. C Mater. 2019, 8, 50–57. [Google Scholar] [CrossRef]
  106. Liang, Z.; Ma, C.; Shen, L.; Lu, L.; Lu, X.; Lou, X.; Liu, M.; Jia, C.L. Flexible Lead-Free Oxide Film Capacitors with Ultrahigh Energy Storage Performances in Extremely Wide Operating Temperature. Nano Energy 2019, 57, 519–527. [Google Scholar] [CrossRef]
  107. Hoffmann, M.; Schroeder, U.; Künneth, C.; Kersch, A.; Starschich, S.; Böttger, U.; Mikolajick, T. Ferroelectric Phase Transitions in Nanoscale HfO2 Films Enable Giant Pyroelectric Energy Conversion and Highly Efficient Supercapacitors. Nano Energy 2015, 18, 154–164. [Google Scholar] [CrossRef]
  108. Lomenzo, P.D.; Chung, C.C.; Zhou, C.; Jones, J.L.; Nishida, T. Doped Hf0.5Zr0.5O2 for High Efficiency Integrated Supercapacitors. Appl. Phys. Lett. 2017, 110, 232904. [Google Scholar] [CrossRef]
  109. He, Y.; Zheng, G.; Wu, X.; Liu, W.J.; Zhang, D.W.; Ding, S.J. Superhigh Energy Storage Density On-Chip Capacitors with Ferroelectric Hf0.5Zr0.5O2/Antiferroelectric Hf0.25Zr0.75O2 Bilayer Nanofilms Fabricated by Plasma-Enhanced Atomic Layer Deposition. Nanoscale Adv. 2022, 4, 4648–4657. [Google Scholar] [CrossRef] [PubMed]
  110. Yi, S.H.; Chan, Y.C.; Mo, C.L.; Lin, H.C.; Chen, M.J. Enhancement of Energy Storage for Electrostatic Supercapacitors through Built-in Electric Field Engineering. Nano Energy 2022, 99, 107342. [Google Scholar] [CrossRef]
  111. Yang, K.; Lee, E.B.; Lee, D.H.; Park, J.Y.; Kim, S.H.; Park, G.H.; Yu, G.T.; Lee, J.I.; Kim, G.H.; Park, M.H. Energy Conversion and Storage Using Artificially Induced Antiferroelectricity in HfO2/ZrO2 Nanolaminates. Compos. B Eng. 2022, 236, 109824. [Google Scholar] [CrossRef]
  112. Haspert, L.C.; Lee, S.B.; Rubloff, G.W. Nanoengineering Strategies for Metal-Insulator-Metal Electrostatic Nanocapacitors. ACS Nano 2012, 6, 3528–3536. [Google Scholar] [CrossRef]
  113. Hourdakis, E.; Botzakaki, M.A.; Xanthopoulos, N.J. Metal-Insulator-Metal Micro-Capacitors for Integrated Energy Storage up to 105 Hz. J. Phys. D Appl. Phys. 2022, 55, 455502. [Google Scholar] [CrossRef]
  114. Zhang, Y.; Li, X.; Song, J.; Zhang, S.; Wang, J.; Dai, X.; Liu, B.; Dong, G.; Zhao, L. AgNbO3 Antiferroelectric Film with High Energy Storage Performance. J. Mater. 2021, 7, 1294–1300. [Google Scholar] [CrossRef]
  115. Ko, E.C.; Kang, W.; Han, J.H. Improved Dielectric Constant and Leakage Current Characteristics of BaTiO3 Thin Film on SrRuO3 Seed Layer. J. Alloys Compd. 2022, 895, 162579. [Google Scholar] [CrossRef]
  116. Chien, C.S.; Tsai, M.H.; Huang, C.L. Electrical Properties and Current Conduction Mechanisms of LaGdO3 Thin Film by RF Sputtering for RRAM Applications. J. Asian Ceram. Soc. 2020, 8, 948–956. [Google Scholar] [CrossRef]
  117. Zhang, L.; Pu, Y.; Chen, M.; Peng, X.; Wang, B.; Shang, J. Design Strategies of Perovskite Energy-Storage Dielectrics for next-Generation Capacitors. J. Eur. Ceram. Soc. 2023, 43, 5713–5747. [Google Scholar] [CrossRef]
  118. Pan, M.J.; Randall, C. A Brief Introduction to Ceramic Capacitors. IEEE Electr. Insul. Mag. 2010, 26, 44–50. [Google Scholar] [CrossRef]
  119. Ulrich, R.; Schaper, L.; Nelms, D.; Leftwich, M. Comparison of Paraelectric and Ferroelectric Materials for Applications as Dielectrics in Thin Film Integrated Capacitors. Int. J. Microcircuits Electron. Packag. 2000, 23, 172–181. [Google Scholar]
  120. Dugu, S.; Pavunny, S.P.; Scott, J.F.; Katiyar, R.S. Si:SrTiO3-Al2O3-Si:SrTiO3 Multi-Dielectric Architecture for Metal-Insulator-Metal Capacitor Applications. Appl. Phys. Lett. 2016, 109, 212901. [Google Scholar] [CrossRef]
  121. Nguyen, M.D. Ultrahigh Energy-Storage Performance in Lead-Free BZT Thin-Films by Tuning Relaxor Behavior. Mater. Res. Bull. 2021, 133, 111072. [Google Scholar] [CrossRef]
  122. Moon, S.; Moon Lee, S.; Lim, H.-K.; Jin, H.-J.; Soo Yun, Y.; Moon, S.; Jin, H.; Lee, S.M.; Lim, H.; Yun, Y.S. Relationship between Multivalent Cation Charge Carriers and Organic Solvents on Nanoporous Carbons in 4 V-Window Magnesium Ion Supercapacitors. Adv. Energy Mater. 2021, 11, 2101054. [Google Scholar] [CrossRef]
  123. Cabana, J.; Monconduit, L.; Larcher, D.; Palacín, M.R. Beyond Intercalation-Based Li-Ion Batteries: The State of the Art and Challenges of Electrode Materials Reacting through Conversion Reactions. Adv. Mater. 2010, 22, E170–E192. [Google Scholar] [CrossRef]
  124. Ni, J.; Dai, A.; Yuan, Y.; Li, L.; Lu, J. Three-Dimensional Microbatteries beyond Lithium Ion. Matter 2020, 2, 1366–1376. [Google Scholar] [CrossRef]
  125. Nasreldin, M.; de Mulatier, S.; Delattre, R.; Ramuz, M.; Djenizian, T. Flexible and Stretchable Microbatteries for Wearable Technologies. Adv. Mater. Technol. 2020, 5, 2000412. [Google Scholar] [CrossRef]
  126. Tan, H.; Feng, Y.; Rui, X.; Yu, Y.; Huang, S.; Tan, H.T.; Rui, X.H.; Huang, S.M.; Feng, Y.Z.; Yu, Y. Metal Chalcogenides: Paving the Way for High-Performance Sodium/Potassium-Ion Batteries. Small Methods 2020, 4, 1900563. [Google Scholar] [CrossRef]
  127. Pan, Q.; Tong, Z.; Su, Y.; Qin, S.; Tang, Y.; Pan, Q.G.; Tong, Z.P.; Su, Y.Q.; Qin, S.; Tang, Y.B. Energy Storage Mechanism, Challenge and Design Strategies of Metal Sulfides for Rechargeable Sodium/Potassium-Ion Batteries. Adv. Funct. Mater. 2021, 31, 2103912. [Google Scholar] [CrossRef]
  128. Buvat, G.; Iadecola, A.; Blanchard, F.; Brousse, T.; Roussel, P.; Lethien, C. A First Outlook of Sputtered FeWO 4 Thin Films for Micro-Supercapacitor Electrodes. J. Electrochem. Soc. 2021, 168, 030524. [Google Scholar] [CrossRef]
  129. Nunes, W.G.; Freitas, B.G.A.; Beraldo, R.M.; Filho, R.M.; Da Silva, L.M.; Zanin, H. A Rational Experimental Approach to Identify Correctly the Working Voltage Window of Aqueous-Based Supercapacitors. Sci. Rep. 2020, 10, 19195. [Google Scholar] [CrossRef]
  130. Huang, M.; Li, F.; Dong, F.; Zhang, Y.X.; Zhang, L.L. MnO2-Based Nanostructures for High-Performance Supercapacitors. J. Mater. Chem. A Mater. 2015, 3, 21380–21423. [Google Scholar] [CrossRef]
  131. Kumar, A.; Sanger, A.; Kumar, A.; Mishra, Y.K.; Chandra, R. Performance of High Energy Density Symmetric Supercapacitor Based on Sputtered MnO2 Nanorods. ChemistrySelect 2016, 1, 3885–3891. [Google Scholar] [CrossRef]
  132. Liu, L.; Zhao, H.; Lei, Y. Advances on Three-Dimensional Electrodes for Micro-Supercapacitors: A Mini-Review. InfoMat 2019, 1, 74–84. [Google Scholar] [CrossRef]
  133. Yu, M.; Qiu, W.; Wang, F.; Zhai, T.; Fang, P.; Lu, X.; Tong, Y. Three Dimensional Architectures: Design, Assembly and Application in Electrochemical Capacitors. J. Mater. Chem. A Mater. 2015, 3, 15792–15823. [Google Scholar] [CrossRef]
  134. Xie, Y.; Zhang, H.; Huang, H.; Wang, Z.; Xu, Z.; Zhao, H.; Wang, Y.; Chen, N.; Yang, W. High-Voltage Asymmetric MXene-Based on-Chip Micro-Supercapacitors. Nano Energy 2020, 74, 104928. [Google Scholar] [CrossRef]
  135. Jiang, Q.; Kurra, N.; Alhabeb, M.; Gogotsi, Y.; Alshareef, H.N. All Pseudocapacitive MXene-RuO2 Asymmetric Supercapacitors. Adv. Energy Mater. 2018, 8, 1703043. [Google Scholar] [CrossRef]
  136. Couly, C.; Alhabeb, M.; Van Aken, K.L.; Kurra, N.; Gomes, L.; Navarro-Suárez, A.M.; Anasori, B.; Alshareef, H.N.; Gogotsi, Y. Asymmetric Flexible MXene-Reduced Graphene Oxide Micro-Supercapacitor. Adv. Electron. Mater. 2018, 4, 1700339. [Google Scholar] [CrossRef]
  137. Xia, C.; Jiang, Q.; Zhao, C.; Beaujuge, P.M.; Alshareef, H.N. Asymmetric Supercapacitors with Metal-like Ternary Selenides and Porous Graphene Electrodes. Nano Energy 2016, 24, 78–86. [Google Scholar] [CrossRef]
  138. Zhang, Y.; Cao, N.; Szunerits, S.; Addad, A.; Roussel, P.; Boukherroub, R. Fabrication of ZnCoS Nanomaterial for High Energy Flexible Asymmetric Supercapacitors. Chem. Eng. J. 2019, 374, 347–358. [Google Scholar] [CrossRef]
  139. Adalati, R.; Kumar, A.; Sharma, M.; Chandra, R. Pt Enhanced Capacitive Performance of Cr2N Electrode toward Flexible Asymmetric Supercapacitor. Appl. Phys. Lett. 2021, 118, 183901. [Google Scholar] [CrossRef]
  140. Priyadarsini, S.S.; Saxena, S.; Ladole, A.H.; Nehru, D.; Dasgupta, S. Inkjet-Printed Transparent Asymmetric Microsupercapacitors with Mesoporous NiCo2O4 and Mn2O3 Electrodes. ACS Appl. Energy Mater. 2024, 7, 715–725. [Google Scholar] [CrossRef]
  141. Peng, H.; Ma, G.; Sun, K.; Mu, J.; Luo, M.; Lei, Z. High-Performance Aqueous Asymmetric Supercapacitor Based on Carbon Nanofibers Network and Tungsten Trioxide Nanorod Bundles Electrodes. Electrochim. Acta 2014, 147, 54–61. [Google Scholar] [CrossRef]
  142. Yang, W.; Zhu, Y.; Jia, Z.; He, L.; Xu, L.; Meng, J.; Tahir, M.; Zhou, Z.; Wang, X.; Mai, L.; et al. Interwoven Nanowire Based On-Chip Asymmetric Microsupercapacitor with High Integrability, Areal Energy, and Power Density. Adv. Energy Mater. 2020, 10, 2001873. [Google Scholar] [CrossRef]
  143. Shen, K.; Ding, J.; Yang, S.; Shen, K.; Ding, J.W.; Yang, S.B. 3D Printing Quasi-Solid-State Asymmetric Micro-Supercapacitors with Ultrahigh Areal Energy Density. Adv. Energy Mater. 2018, 8, 1800408. [Google Scholar] [CrossRef]
  144. Ma, B.; Hao, W.; Ruan, W.; Yuan, C.; Wang, Q.; Teng, F. Unveiling Capacitive Behaviors of MoO2 in Different Electrolytes and Flexible MoO2-Based Asymmetric Micro-Supercapacitor. J. Energy Storage 2022, 52, 104833. [Google Scholar] [CrossRef]
  145. Asbani, B.; Robert, K.; Roussel, P.; Brousse, T.; Lethien, C. Asymmetric Micro-Supercapacitors Based on Electrodeposited Ruo2 and Sputtered VN Films. Energy Storage Mater. 2021, 37, 207–214. [Google Scholar] [CrossRef]
  146. 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] [PubMed]
  147. Zheng, S.; Ma, J.; Wu, Z.S.; Zhou, F.; He, Y.B.; Kang, F.; Cheng, H.M.; Bao, X. All-Solid-State Flexible Planar Lithium Ion Micro-Capacitors. Energy Environ. Sci. 2018, 11, 2001–2009. [Google Scholar] [CrossRef]
  148. Bai, C.; Zhang, J.; Chen, R.; Wu, W.; Li, X.; Wang, J.; Lu, Y.; Zhao, Y. A 4 V Planar Li-Ion Micro-Supercapacitor with Ultrahigh Energy Density. ACS Energy Lett. 2024, 9, 410–418. [Google Scholar] [CrossRef]
  149. Zhou, W.; Liu, Z.; Chen, W.; Sun, X.; Luo, M.; Zhang, X.; Li, C.; An, Y.; Song, S.; Wang, K.; et al. A Review on Thermal Behaviors and Thermal Management Systems for Supercapacitors. Batteries 2023, 9, 128. [Google Scholar] [CrossRef]
  150. Liu, W.; Sun, X.; Yan, X.; Gao, Y.; Zhang, X.; Wang, K.; Ma, Y.; Liu, W.; Sun, X.; Yan, X.; et al. Review of Energy Storage Capacitor Technology. Batteries 2024, 10, 271. [Google Scholar] [CrossRef]
  151. Li, H.C.; Shen, H.R.; Shi, Y.; Wen, L.; Li, F. Progress and Prospects of Graphene for In-Plane Micro-Supercapacitors. New Carbon Mater. 2022, 37, 781–801. [Google Scholar] [CrossRef]
  152. Yu, W.; Zhou, H.; Li, B.Q.; Ding, S. 3D Printing of Carbon Nanotubes-Based Microsupercapacitors. ACS Appl. Mater. Interfaces 2017, 9, 4597–4604. [Google Scholar] [CrossRef]
  153. Kim, S.-K.; Koo, H.-J.; Lee, A.; Braun, P.V.; Kim, S.; Koo, H.; Lee, A.; Braun, P. V Selective Wetting-Induced Micro-Electrode Patterning for Flexible Micro-Supercapacitors. Adv. Mater. 2014, 26, 5108–5112. [Google Scholar] [CrossRef]
  154. Dousti, B.; Choi, Y., II; Cogan, S.F.; Lee, G.S. A High Energy Density 2D Microsupercapacitor Based on an Interconnected Network of a Horizontally Aligned Carbon Nanotube Sheet. ACS Appl. Mater. Interfaces 2020, 12, 50011–50023. [Google Scholar] [CrossRef]
  155. Pech, D.; Brunet, M.; Taberna, P.L.; Simon, P.; Fabre, N.; Mesnilgrente, F.; Conédéra, V.; Durou, H. Elaboration of a Microstructured Inkjet-Printed Carbon Electrochemical Capacitor. J. Power Sources 2010, 195, 1266–1269. [Google Scholar] [CrossRef]
  156. Brousse, K.; Martin, C.; Brisse, A.L.; Lethien, C.; Simon, P.; Taberna, P.L.; Brousse, T. Anthraquinone Modification of Microporous Carbide Derived Carbon Films for On-Chip Micro-Supercapacitors Applications. Electrochim. Acta 2017, 246, 391–398. [Google Scholar] [CrossRef]
  157. Heon, M.; Lofland, S.; Applegate, J.; Nolte, R.; Cortes, E.; Hettinger, J.D.; Taberna, P.L.; Simon, P.; Huang, P.; Brunet, M.; et al. Continuous Carbide-Derived Carbon Films with High Volumetric Capacitance. Energy Environ. Sci. 2010, 4, 135–138. [Google Scholar] [CrossRef]
  158. Pech, D.; Brunet, M.; Durou, H.; Huang, P.; Mochalin, V.; Gogotsi, Y.; Taberna, P.L.; Simon, P. Ultrahigh-Power Micrometre-Sized Supercapacitors Based on Onion-like Carbon. Nat. Nanotechnol. 2010, 5, 651–654. [Google Scholar] [CrossRef]
  159. Kumar, S.; Mukherjee, A.; Telpande, S.; Mahapatra, A.D.; Kumar, P.; Misra, A. Role of the Electrode-Edge in Optically Sensitive Three-Dimensional Carbon Foam-MoS2 Based High-Performance Micro-Supercapacitors. J. Mater. Chem. A Mater. 2023, 11, 4963–4976. [Google Scholar] [CrossRef]
  160. El-Kady, M.F.; Kaner, R.B. Scalable Fabrication of High-Power Graphene Micro-Supercapacitors for Flexible and on-Chip Energy Storage. Nat. Commun. 2013, 4, 1475. [Google Scholar] [CrossRef]
  161. Létiche, M.; Brousse, K.; Demortière, A.; Huang, P.; Daffos, B.; Pinaud, S.; Respaud, M.; Chaudret, B.; Roussel, P.; Buchaillot, L.; et al. Sputtered Titanium Carbide Thick Film for High Areal Energy on Chip Carbon-Based Micro-Supercapacitors. Adv. Funct. Mater. 2017, 27, 1606813. [Google Scholar] [CrossRef]
  162. Eustache, E.; Douard, C.; Retoux, R.; Lethien, C.; Brousse, T.; Eustache, E.; Douard, C.; Brousse, T.; Lethien, C. MnO2 Thin Films on 3D Scaffold: Microsupercapacitor Electrodes Competing with “Bulk” Carbon Electrodes. Adv. Energy Mater. 2015, 5, 1500680. [Google Scholar] [CrossRef]
  163. Mayorov, A.S.; Gorbachev, R.V.; Morozov, S.V.; Britnell, L.; Jalil, R.; Ponomarenko, L.A.; Blake, P.; Novoselov, K.S.; Watanabe, K.; Taniguchi, T.; et al. Micrometer-Scale Ballistic Transport in Encapsulated Graphene at Room Temperature. Nano Lett. 2011, 11, 2396–2399. [Google Scholar] [CrossRef]
  164. Beidaghi, M.; Wang, C. Micro-Supercapacitors Based on Interdigital Electrodes of Reduced Graphene Oxide and Carbon Nanotube Composites with Ultrahigh Power Handling Performance. Adv. Funct. Mater. 2012, 22, 4501–4510. [Google Scholar] [CrossRef]
  165. Xia, J.; Chen, F.; Li, J.; Tao, N. Measurement of the Quantum Capacitance of Graphene. Nat. Nanotechnol. 2009, 4, 505–509. [Google Scholar] [CrossRef] [PubMed]
  166. El-Kady, M.F.; Strong, V.; Dubin, S.; Kaner, R.B. Laser Scribing of High-Performance and Flexible Graphene-Based Electrochemical Capacitors. Science 2012, 335, 1326–1330. [Google Scholar] [CrossRef]
  167. Liang, J.; Mondal, A.K.; Wang, D.W.; Iacopi, F. Graphene-Based Planar Microsupercapacitors: Recent Advances and Future Challenges. Adv. Mater. Technol. 2019, 4, 1800200. [Google Scholar] [CrossRef]
  168. Wu, Z.S.; Parvez, K.; Feng, X.; Müllen, K. Graphene-Based in-Plane Micro-Supercapacitors with High Power and Energy Densities. Nat. Commun. 2013, 4, 2487. [Google Scholar] [CrossRef]
  169. Niu, Z.; Zhang, L.; Liu, L.; Zhu, B.; Dong, H.; Chen, X. All-Solid-State Flexible Ultrathin Micro-Supercapacitors Based on Graphene. Adv. Mater. 2013, 25, 4035–4042. [Google Scholar] [CrossRef] [PubMed]
  170. Ye, J.; Tan, H.; Wu, S.; Ni, K.; Pan, F.; Liu, J.; Tao, Z.; Qu, Y.; Ji, H.; Simon, P.; et al. Direct Laser Writing of Graphene Made from Chemical Vapor Deposition for Flexible, Integratable Micro-Supercapacitors with Ultrahigh Power Output. Adv. Mater. 2018, 30, 1801384. [Google Scholar] [CrossRef]
  171. Velasco, A.; Ryu, Y.K.; Boscá, A.; Ladrón-De-Guevara, A.; Hunt, E.; Zuo, J.; Pedrós, J.; Calle, F.; Martinez, J. Recent Trends in Graphene Supercapacitors: From Large Area to Microsupercapacitors. Sustain. Energy Fuels 2021, 5, 1235–1254. [Google Scholar] [CrossRef]
  172. Pei, S.; Cheng, H.M. The Reduction of Graphene Oxide. Carbon 2012, 50, 3210–3228. [Google Scholar] [CrossRef]
  173. Zhu, Y.; Ni, Z.; Gao, J.; Zhang, D.; Wang, S.; Zhao, J. Fabrication of Flexible Planar Micro-Supercapacitors Using Laser Scribed Graphene Electrodes Incorporated with MXene. J. Phys. Chem. Solids 2023, 183, 111619. [Google Scholar] [CrossRef]
  174. Tarcan, R.; Todor-Boer, O.; Petrovai, I.; Leordean, C.; Astilean, S.; Botiz, I. Reduced Graphene Oxide Today. J. Mater. Chem. C Mater. 2020, 8, 1198–1224. [Google Scholar] [CrossRef]
  175. Shi, X.; Pei, S.; Zhou, F.; Ren, W.; Cheng, H.M.; Wu, Z.S.; Bao, X. Ultrahigh-Voltage Integrated Micro-Supercapacitors with Designable Shapes and Superior Flexibility. Energy Environ. Sci. 2019, 12, 1534–1541. [Google Scholar] [CrossRef]
  176. Wang, Y.; Zhao, Y.; Han, Y.; Li, X.; Dai, C.; Zhang, X.; Jin, X.; Shao, C.; Lu, B.; Wang, C.; et al. Fixture-Free Omnidirectional Prestretching Fabrication and Integration of Crumpled in-Plane Micro-Supercapacitors. Sci. Adv. 2022, 8, eabn8338. [Google Scholar] [CrossRef]
  177. Xiao, H.; Wu, Z.S.; Chen, L.; Zhou, F.; Zheng, S.; Ren, W.; Cheng, H.M.; Bao, X. One-Step Device Fabrication of Phosphorene and Graphene Interdigital Micro-Supercapacitors with High Energy Density. ACS Nano 2017, 11, 7284–7292. [Google Scholar] [CrossRef]
  178. Chen, M.; Shi, X.; Wang, X.; Liu, H.; Wang, S.; Meng, C.; Liu, Y.; Zhang, L.; Zhu, Y.; Wu, Z.S. Low-Temperature and High-Voltage Planar Micro-Supercapacitors Based on Anti-Freezing Hybrid Gel Electrolyte. J. Energy Chem. 2022, 72, 195–202. [Google Scholar] [CrossRef]
  179. Li, J.; Sollami Delekta, S.; Zhang, P.; Yang, S.; Lohe, M.R.; Zhuang, X.; Feng, X.; Östling, M. Scalable Fabrication and Integration of Graphene Microsupercapacitors through Full Inkjet Printing. ACS Nano 2017, 11, 8249–8256. [Google Scholar] [CrossRef] [PubMed]
  180. Zhang, H.; Zhang, X.; Zhang, D.; Sun, X.; Lin, H.; Wang, C.; Ma, Y. One-Step Electrophoretic Deposition of Reduced Graphene Oxide and Ni(OH)2 Composite Films for Controlled Syntheses Supercapacitor Electrodes. J. Phys. Chem. B 2013, 117, 1616–1627. [Google Scholar] [CrossRef]
  181. Shi, X.; Wu, Z.S.; Qin, J.; Zheng, S.; Wang, S.; Zhou, F.; Sun, C.; Bao, X. Graphene-Based Linear Tandem Micro-Supercapacitors with Metal-Free Current Collectors and High-Voltage Output. Adv. Mater. 2017, 29, 1703034. [Google Scholar] [CrossRef] [PubMed]
  182. Wang, Y.; Niu, J. Facile Self-Assembly of Exfoliated Graphene/PANI Film for High-Energy Zn-Ion Micro-Supercapacitors. Molecules 2023, 28, 4470. [Google Scholar] [CrossRef] [PubMed]
  183. Mao, X.; Zhu, L.; Liu, H.; Chen, H.; Ju, P.; Li, W. Synthesis of Graphene via Electrochemical Exfoliation in Different Electrolytes for Direct Electrodeposition of a Cu/Graphene Composite Coating. RSC Adv. 2019, 9, 35524–35531. [Google Scholar] [CrossRef]
  184. Xiang, L.; Shen, Q.; Zhang, Y.; Bai, W.; Nie, C. One-Step Electrodeposited Ni-Graphene Composite Coating with Excellent Tribological Properties. Surf. Coat. Technol. 2019, 373, 38–46. [Google Scholar] [CrossRef]
  185. Wuttke, M.; Liu, Z.; Lu, H.; Narita, A.; Müllen, K. Direct Metal-Free Chemical Vapor Deposition of Graphene Films on Insulating Substrates for Micro-Supercapacitors with High Volumetric Capacitance. Batter. Supercaps 2019, 2, 929–933. [Google Scholar] [CrossRef]
  186. Gupta, N.; Mogera, U.; Kulkarni, G.U. Ultrafast Planar Microsupercapacitor Based on Defect-Free Twisted Multilayer Graphene. Mater. Res. Bull. 2022, 152, 111841. [Google Scholar] [CrossRef]
  187. Liu, Z.; Chen, Z.; Wang, C.; Wang, H.I.; Wuttke, M.; Wang, X.Y.; Bonn, M.; Chi, L.; Narita, A.; Müllen, K. Bottom-Up, On-Surface-Synthesized Armchair Graphene Nanoribbons for Ultra-High-Power Micro-Supercapacitors. J. Am. Chem. Soc. 2020, 142, 17881–17886. [Google Scholar] [CrossRef]
  188. Das, P.; Zhang, L.; Zheng, S.; Shi, X.; Li, Y.; Wu, Z.S. Rapid Fabrication of High-Quality Few-Layer Graphene through Gel-Phase Electrochemical Exfoliation of Graphite for High-Energy-Density Ionogel-Based Micro-Supercapacitors. Carbon 2022, 196, 203–212. [Google Scholar] [CrossRef]
  189. Lin, J.; Peng, Z.; Liu, Y.; Ruiz-Zepeda, F.; Ye, R.; Samuel, E.L.G.; Yacaman, M.J.; Yakobson, B.I.; Tour, J.M. Laser-Induced Porous Graphene Films from Commercial Polymers. Nat. Commun. 2014, 5, 5714. [Google Scholar] [CrossRef]
  190. Wu, Y.; Chen, J.; Yuan, W.; Zhang, X.; Bai, S.; Chen, Y.; Zhao, B.; Wu, X.; Wang, C.; Huang, H.; et al. Direct Mask-Free Fabrication of Patterned Hierarchical Graphene Electrode for on-Chip Micro-Supercapacitors. J. Mater. Sci. Technol. 2023, 143, 12–19. [Google Scholar] [CrossRef]
  191. Bellani, S.; Petroni, E.; Del Rio Castillo, A.E.; Curreli, N.; Martín-García, B.; Oropesa-Nuñez, R.; Prato, M.; Bonaccorso, F. Scalable Production of Graphene Inks via Wet-Jet Milling Exfoliation for Screen-Printed Micro-Supercapacitors. Adv. Funct. Mater. 2019, 29, 1807659. [Google Scholar] [CrossRef]
  192. Cao, C.; Zhou, Y.; Ubnoske, S.; Zang, J.; Cao, Y.; Henry, P.; Parker, C.B.; Glass, J.T. Highly Stretchable Supercapacitors via Crumpled Vertically Aligned Carbon Nanotube Forests. Adv. Energy Mater. 2019, 9, 1900618. [Google Scholar] [CrossRef]
  193. He, P.; Ding, Z.; Zhao, X.; Liu, J.; Huang, Q.; Peng, J.; Fan, L.Z. Growth of Carbon Nanosheets on Carbon Nanotube Arrays for the Fabrication of Three-Dimensional Micro-Patterned Supercapacitors. Carbon 2019, 155, 453–461. [Google Scholar] [CrossRef]
  194. Dinh, T.M.; Achour, A.; Vizireanu, S.; Dinescu, G.; Nistor, L.; Armstrong, K.; Guay, D.; Pech, D. Hydrous RuO2/Carbon Nanowalls Hierarchical Structures for All-Solid-State Ultrahigh-Energy-Density Micro-Supercapacitors. Nano Energy 2014, 10, 288–294. [Google Scholar] [CrossRef]
  195. Ye, R.; James, D.K.; Tour, J.M. Laser-Induced Graphene: From Discovery to Translation. Adv. Mater. 2019, 31, 1803621. [Google Scholar] [CrossRef]
  196. Wang, S.; Wu, Z.S.; Zheng, S.; Zhou, F.; Sun, C.; Cheng, H.M.; Bao, X. Scalable Fabrication of Photochemically Reduced Graphene-Based Monolithic Micro-Supercapacitors with Superior Energy and Power Densities. ACS Nano 2017, 11, 4283–4291. [Google Scholar] [CrossRef]
  197. Zhou, F.; Huang, H.; Xiao, C.; Zheng, S.; Shi, X.; Qin, J.; Fu, Q.; Bao, X.; Feng, X.; Müllen, K.; et al. Electrochemically Scalable Production of Fluorine-Modified Graphene for Flexible and High-Energy Ionogel-Based Microsupercapacitors. J. Am. Chem. Soc. 2018, 140, 8198–8205. [Google Scholar] [CrossRef]
  198. Rao, Y.; Yuan, M.; Gao, B.; Li, H.; Yu, J.; Chen, X. Laser-Scribed Phosphorus-Doped Graphene Derived from Kevlar Textile for Enhanced Wearable Micro-Supercapacitor. J. Colloid. Interface Sci. 2023, 630, 586–594. [Google Scholar] [CrossRef]
  199. Wu, Z.S.; Tan, Y.Z.; Zheng, S.; Wang, S.; Parvez, K.; Qin, J.; Shi, X.; Sun, C.; Bao, X.; Feng, X.; et al. Bottom-Up Fabrication of Sulfur-Doped Graphene Films Derived from Sulfur-Annulated Nanographene for Ultrahigh Volumetric Capacitance Micro-Supercapacitors. J. Am. Chem. Soc. 2017, 139, 4506–4512. [Google Scholar] [CrossRef] [PubMed]
  200. Yuan, M.; Wang, Z.; Rao, Y.; Wang, Y.; Gao, B.; Yu, J.; Li, H.; Chen, X. Laser Direct Writing O/N/S Co-Doped Hierarchically Porous Graphene on Carboxymethyl Chitosan/Lignin-Reinforced Wood for Boosted Microsupercapacitor. Carbon 2023, 202, 296–304. [Google Scholar] [CrossRef]
  201. El-Kady, M.F.; Ihns, M.; Li, M.; Hwang, J.Y.; Mousavi, M.F.; Chaney, L.; Lech, A.T.; Kaner, R.B. Engineering Three-Dimensional Hybrid Supercapacitors and Microsupercapacitors for High-Performance Integrated Energy Storage. Proc. Natl. Acad. Sci. USA 2015, 112, 4233–4238. [Google Scholar] [CrossRef]
  202. Zhai, S.; Wang, C.; Karahan, H.E.; Wang, Y.; Chen, X.; Sui, X.; Huang, Q.; Liao, X.; Wang, X.; Chen, Y. Nano-RuO2-Decorated Holey Graphene Composite Fibers for Micro-Supercapacitors with Ultrahigh Energy Density. Small 2018, 14, 1800582. [Google Scholar] [CrossRef]
  203. Gao, M.; Dong, X.; Wang, K.; Duan, W.; Sun, X.; Zhu, C.; Wang, W. Laser Direct Preparation and Processing of Graphene/MnO Nanocomposite Electrodes for Microsupercapacitors. J. Energy Storage 2021, 33, 102162. [Google Scholar] [CrossRef]
  204. Lee, H.U.; Kim, S.W. Pen Lithography for Flexible Microsupercapacitors with Layer-by-Layer Assembled Graphene Flake/PEDOT Nanocomposite Electrodes. J. Mater. Chem. A Mater. 2017, 5, 13581–13590. [Google Scholar] [CrossRef]
  205. Chen, Y.; Guo, M.; Xu, L.; Cai, Y.; Tian, X.; Liao, X.; Wang, Z.; Meng, J.; Hong, X.; Mai, L. In-Situ Selective Surface Engineering of Graphene Micro-Supercapacitor Chips. Nano Res. 2022, 15, 1492–1499. [Google Scholar] [CrossRef]
  206. Yao, B.; Chandrasekaran, S.; Zhang, H.; Ma, A.; Kang, J.; Zhang, L.; Lu, X.; Qian, F.; Zhu, C.; Duoss, E.B.; et al. 3D-Printed Structure Boosts the Kinetics and Intrinsic Capacitance of Pseudocapacitive Graphene Aerogels. Adv. Mater. 2020, 32, 1906652. [Google Scholar] [CrossRef]
  207. Wu, C.; Zhang, Z.; Chen, Z.; Jiang, Z.; Li, H.; Cao, H.; Liu, Y.; Zhu, Y.; Fang, Z.; Yu, X. Rational Design of Novel Ultra-Small Amorphous Fe2O3 Nanodots/Graphene Heterostructures for All-Solid-State Asymmetric Supercapacitors. Nano Res. 2021, 14, 953–960. [Google Scholar] [CrossRef]
  208. Zhang, H.; Zhang, Z.; Qi, X.; Yu, J.; Cai, J.; Yang, Z. Manganese Monoxide/Biomass-Inherited Porous Carbon Nanostructure Composite Based on the High Water-Absorbent Agaric for Asymmetric Supercapacitor. ACS Sustain. Chem. Eng. 2019, 7, 4284–4294. [Google Scholar] [CrossRef]
  209. Lin, J.; Zhang, C.; Yan, Z.; Zhu, Y.; Peng, Z.; Hauge, R.H.; Natelson, D.; Tour, J.M. 3-Dimensional Graphene Carbon Nanotube Carpet-Based Microsupercapacitors with High Electrochemical Performance. Nano Lett. 2013, 13, 72–78. [Google Scholar] [CrossRef] [PubMed]
  210. Ghosh, A.; Le, V.T.; Bae, J.J.; Lee, Y.H. TLM-PSD Model for Optimization of Energy and Power Density of Vertically Aligned Carbon Nanotube Supercapacitor. Sci. Rep. 2013, 3, 2939. [Google Scholar] [CrossRef]
  211. Hsia, B.; Marschewski, J.; Wang, S.; In, J.B.; Carraro, C.; Poulikakos, D.; Grigoropoulos, C.P.; Maboudian, R. Highly Flexible, All Solid-State Micro-Supercapacitors from Vertically Aligned Carbon Nanotubes. Nanotechnology 2014, 25, 055401. [Google Scholar] [CrossRef]
  212. Baughman, R.H.; Zakhidov, A.A.; De Heer, W.A. Carbon Nanotubes—The Route toward Applications. Science 2002, 297, 787–792. [Google Scholar] [CrossRef]
  213. Zhu, S.; Ni, J.; Li, Y. Carbon Nanotube-Based Electrodes for Flexible Supercapacitors. Nano Res. 2020, 13, 1825–1841. [Google Scholar] [CrossRef]
  214. Ouldhamadouche, N.; Achour, A.; Lucio-Porto, R.; Islam, M.; Solaymani, S.; Arman, A.; Ahmadpourian, A.; Achour, H.; Le Brizoual, L.; Djouadi, M.A.; et al. Electrodes Based on Nano-Tree-like Vanadium Nitride and Carbon Nanotubes for Micro-Supercapacitors. J. Mater. Sci. Technol. 2018, 34, 976–982. [Google Scholar] [CrossRef]
  215. Guo, J.; Zhang, Q.; Sun, J.; Li, C.; Zhao, J.; Zhou, Z.; He, B.; Wang, X.; Man, P.; Li, Q.; et al. Direct Growth of Vanadium Nitride Nanosheets on Carbon Nanotube Fibers as Novel Negative Electrodes for High-Energy-Density Wearable Fiber-Shaped Asymmetric Supercapacitors. J. Power Sources 2018, 382, 122–127. [Google Scholar] [CrossRef]
  216. Pitkänen, O.; Järvinen, T.; Cheng, H.; Lorite, G.S.; Dombovari, A.; Rieppo, L.; Talapatra, S.; Duong, H.M.; Tóth, G.; Juhász, K.L.; et al. On-Chip Integrated Vertically Aligned Carbon Nanotube Based Super- and Pseudocapacitors. Sci. Rep. 2017, 7, 16594. [Google Scholar] [CrossRef] [PubMed]
  217. Zhang, H.; Wang, B.; Brown, B. Area-Selective Atomic Layer Deposition of Titanium Oxide and Nitride on Vertically Aligned Carbon Nanotubes Patterned by Aerosol Jet Printing for 3D Microsupercapacitors. J. Power Sources 2022, 551, 232154. [Google Scholar] [CrossRef]
  218. Presser, V.; Heon, M.; Gogotsi, Y. Carbide-Derived Carbons—from Porous Networks to Nanotubes and Graphene. Adv. Funct. Mater. 2011, 21, 810–833. [Google Scholar] [CrossRef]
  219. Huang, P.; Heon, M.; Pech, D.; Brunet, M.; Taberna, P.L.; Gogotsi, Y.; Lofland, S.; Hettinger, J.D.; Simon, P. Micro-Supercapacitors from Carbide Derived Carbon (CDC) Films on Silicon Chips. J. Power Sources 2013, 225, 240–244. [Google Scholar] [CrossRef]
  220. Cheng, K.Y.; Bijukumar, D.; Runa, M.; McNallan, M.; Mathew, M. Tribocorrosion Aspects of Implant Coatings: Hip Replacements. In Tribocorrosion: Fundamentals, Methods, and Materials; Academic Press: Cambridge, MA, USA, 2021; pp. 93–126. [Google Scholar]
  221. Forouzandeh, P.; Kumaravel, V.; Pillai, S.C. Electrode Materials for Supercapacitors: A Review of Recent Advances. Catalysts 2020, 10, 969. [Google Scholar] [CrossRef]
  222. Jiang, Y.; Liu, J. Definitions of Pseudocapacitive Materials: A Brief Review. Energy Environ. Mater. 2019, 2, 30–37. [Google Scholar] [CrossRef]
  223. Liu, J.; Wang, J.; Xu, C.; Jiang, H.; Li, C.; Zhang, L.; Lin, J.; Shen, Z.X. Advanced Energy Storage Devices: Basic Principles, Analytical Methods, and Rational Materials Design. Adv. Sci. 2018, 5, 1700322. [Google Scholar] [CrossRef]
  224. Conway, B.E. Two-Dimensional and Quasi-Two-Dimensional Isotherms for Li Intercalation and Upd Processes at Surfaces. Electrochim. Acta 1993, 38, 1249–1258. [Google Scholar] [CrossRef]
  225. Huang, C.; Zhang, W.; Zheng, W. The Debut and Spreading the Landscape for Excellent Vacancies-Promoted Electrochemical Energy Storage of Nano-Architected Molybdenum Oxides. Mater. Today Energy 2022, 30, 101154. [Google Scholar] [CrossRef]
  226. Diao, Y.; Lu, Y.; Yang, H.; Wang, H.; Chen, H.; D’Arcy, J.M. Direct Conversion of Fe2O3 to 3D Nanofibrillar PEDOT Microsupercapacitors. Adv. Funct. Mater. 2020, 30, 2003394. [Google Scholar] [CrossRef]
  227. Sung, J.H.; Kim, S.J.; Lee, K.H. Fabrication of Microcapacitors Using Conducting Polymer Microelectrodes. J. Power Sources 2003, 124, 343–350. [Google Scholar] [CrossRef]
  228. Wang, X.; Myers, B.D.; Yan, J.; Shekhawat, G.; Dravid, V.; Lee, P.S. Manganese Oxide Micro-Supercapacitors with Ultra-High Areal Capacitance. Nanoscale 2013, 5, 4119–4122. [Google Scholar] [CrossRef]
  229. Choi, C.; Robert, K.; Whang, G.; Roussel, P.; Lethien, C.; Dunn, B. Photopatternable Hydroxide Ion Electrolyte for Solid-State Micro-Supercapacitors. Joule 2021, 5, 2466–2478. [Google Scholar] [CrossRef]
  230. Lei, Y.; Zhao, W.; Zhu, Y.; Buttner, U.; Dong, X.; Alshareef, H.N. Three-Dimensional Ti3C2Tx MXene-Prussian Blue Hybrid Microsupercapacitors by Water Lift-Off Lithography. ACS Nano 2022, 16, 1974–1985. [Google Scholar] [CrossRef] [PubMed]
  231. Lukatskaya, M.R.; Mashtalir, O.; Ren, C.E.; Dall’Agnese, Y.; Rozier, P.; Taberna, P.L.; Naguib, M.; Simon, P.; Barsoum, M.W.; Gogotsi, Y. Cation Intercalation and High Volumetric Capacitance of Two-Dimensional Titanium Carbide. Science 2013, 341, 1502–1505. [Google Scholar] [CrossRef] [PubMed]
  232. Feng, J.; Sun, X.; Wu, C.; Peng, L.; Lin, C.; Hu, S.; Yang, J.; Xie, Y. Metallic Few-Layered VS2 Ultrathin Nanosheets: High Two-Dimensional Conductivity for in-Plane Supercapacitors. J. Am. Chem. Soc. 2011, 133, 17832–17838. [Google Scholar] [CrossRef] [PubMed]
  233. Wang, X.; Zhang, Y.; Zheng, J.; Liu, X.; Meng, C. Hydrothermal Synthesis of VS4/CNTs Composite with Petal-Shape Structures Performing a High Specific Capacity in a Large Potential Range for High-Performance Symmetric Supercapacitors. J. Colloid. Interface Sci. 2019, 554, 191–201. [Google Scholar] [CrossRef]
  234. Guo, Z.; Yang, L.; Wang, W.; Cao, L.; Dong, B. Ultrathin VS2 Nanoplate with In-Plane and out-of-Plane Defects for an Electrochemical Supercapacitor with Ultrahigh Specific Capacitance. J. Mater. Chem. A Mater. 2018, 6, 14681–14688. [Google Scholar] [CrossRef]
  235. Karade, S.S.; Banerjee, K.; Majumder, S.; Sankapal, B.R. Novel Application of Non-Aqueous Chemical Bath Deposited Sb2S3 Thin Films as Supercapacitive Electrode. Int. J. Hydrogen Energy 2016, 41, 21278–21285. [Google Scholar] [CrossRef]
  236. Jolayemi, B.; Buvat, G.; Brousse, T.; Roussel, P.; Lethien, C. Sputtered (Fe,Mn)3O4 Spinel Oxide Thin Films for Micro-Supercapacitor. J. Electrochem. Soc. 2022, 169, 110524. [Google Scholar] [CrossRef]
  237. Gogotsi, Y.; Penner, R.M. Energy Storage in Nanomaterials—Capacitive, Pseudocapacitive, or Battery-Like? ACS Nano 2018, 12, 2081–2083. [Google Scholar] [CrossRef]
  238. Nguyen, T.; De Fátima Montemor, M.; Nguyen, T.; Montemor, M.F. Metal Oxide and Hydroxide–Based Aqueous Supercapacitors: From Charge Storage Mechanisms and Functional Electrode Engineering to Need-Tailored Devices. Adv. Sci. 2019, 6, 1801797. [Google Scholar] [CrossRef]
  239. Huang, T.; Jiang, K.; Chen, D.; Shen, G. Recent Progress and Perspectives of Metal Oxides Based On-Chip Microsupercapacitors. Chin. Chem. Lett. 2018, 29, 553–563. [Google Scholar] [CrossRef]
  240. Nithya, V.D.; Arul, N.S. Review on α-Fe2O3 Based Negative Electrode for High Performance Supercapacitors. J. Power Sources 2016, 327, 297–318. [Google Scholar] [CrossRef]
  241. Brousse, K.; Pinaud, S.; Nguyen, S.; Fazzini, P.F.; Makarem, R.; Josse, C.; Thimont, Y.; Chaudret, B.; Taberna, P.L.; Respaud, M.; et al. Facile and Scalable Preparation of Ruthenium Oxide-Based Flexible Micro-Supercapacitors. Adv. Energy Mater. 2020, 10, 1903136. [Google Scholar] [CrossRef]
  242. Guan, M.; Wang, Q.; Zhang, X.; Bao, J.; Gong, X.; Liu, Y. Two-Dimensional Transition Metal Oxide and Hydroxide-Based Hierarchical Architectures for Advanced Supercapacitor Materials. Front. Chem. 2020, 8, 390. [Google Scholar] [CrossRef]
  243. Liu, C.; Li, Z.; Zhang, Z. MoOx Thin Films Deposited by Magnetron Sputtering as an Anode for Aqueous Micro-Supercapacitors. Sci. Technol. Adv. Mater. 2013, 14, 065005. [Google Scholar] [CrossRef]
  244. Kim, J.Y.; Lee, S.H.; Yan, Y.; Oh, J.; Zhu, K. Controlled Synthesis of Aligned Ni-NiO Core-Shell Nanowire Arrays on Glass Substrates as a New Supercapacitor Electrode. RSC Adv. 2012, 2, 8281–8285. [Google Scholar] [CrossRef]
  245. Geramifard, N.; Chakraborty, B.; Dousti, B.; Lee, G.S.; Maeng, J. High-Energy-Density Sputtered Iridium Oxide Micro-Supercapacitors Operating in Physiological Electrolytes. J. Electrochem. Soc. 2022, 169, 050508. [Google Scholar] [CrossRef]
  246. Deng, W.; Ji, X.; Chen, Q.; Banks, C.E. Electrochemical Capacitors Utilising Transition Metal Oxides: An Update of Recent Developments. RSC Adv. 2011, 1, 1171–1178. [Google Scholar] [CrossRef]
  247. Hota, M.K.; Jiang, Q.; Mashraei, Y.; Salama, K.N.; Alshareef, H.N. Fractal Electrochemical Microsupercapacitors. Adv. Electron. Mater. 2017, 3, 1700185. [Google Scholar] [CrossRef]
  248. Wang, Y.T.; Lu, A.H.; Zhang, H.L.; Li, W.C. Synthesis of Nanostructured Mesoporous Manganese Oxides with Three-Dimensional Frameworks and Their Application in Supercapacitors. J. Phys. Chem. C 2011, 115, 5413–5421. [Google Scholar] [CrossRef]
  249. Han, Z.J.; Pineda, S.; Murdock, A.T.; Seo, D.H.; Ostrikov, K.K.; Bendavid, A. RuO2-Coated Vertical Graphene Hybrid Electrodes for High-Performance Solid-State Supercapacitors. J. Mater. Chem. A Mater. 2017, 5, 17293–17301. [Google Scholar] [CrossRef]
  250. Bounor, B.; Asbani, B.; Douard, C.; Favier, F.; Brousse, T.; Lethien, C. On Chip MnO2-Based 3D Micro-Supercapacitors with Ultra-High Areal Energy Density. Energy Storage Mater. 2021, 38, 520–527. [Google Scholar] [CrossRef]
  251. Asbani, B.; Buvat, G.; Freixas, J.; Huvé, M.; Troadec, D.; Roussel, P.; Brousse, T.; Lethien, C. Ultra-High Areal Capacitance and High Rate Capability RuO2 Thin Film Electrodes for 3D Micro-Supercapacitors. Energy Storage Mater. 2021, 42, 259–267. [Google Scholar] [CrossRef]
  252. Bounor, B.; Asbani, B.; Douard, C.; Deresmes, D.; Stiévenard, D.; Roussel, P.; Favier, F.; Lethien, C.; Brousse, T. Nanostructured MnO2 Films for 3D Micro-Supercapacitors: From New Insights of the Growth Mechanism to the Fine Tuning of Areal Capacitance Values. J. Electrochem. Soc. 2023, 170, 030530. [Google Scholar] [CrossRef]
  253. Wei, W.; Cui, X.; Chen, W.; Ivey, D.G. Manganese Oxide-Based Materials as Electrochemical Supercapacitor Electrodes. Chem. Soc. Rev. 2011, 40, 1697–1721. [Google Scholar] [CrossRef] [PubMed]
  254. Tiwari, J.N.; Tiwari, R.N.; Kim, K.S. Zero-Dimensional, One-Dimensional, Two-Dimensional and Three-Dimensional Nanostructured Materials for Advanced Electrochemical Energy Devices. Prog. Mater. Sci. 2012, 57, 724–803. [Google Scholar] [CrossRef]
  255. Jrondi, A.; Buvat, G.; Pena, F.D.L.; Marinova, M.; Huvé, M.; Brousse, T.; Roussel, P.; Lethien, C. Major Improvement in the Cycling Ability of Pseudocapacitive Vanadium Nitride Films for Micro-Supercapacitor. Adv. Energy Mater. 2023, 13, 2203462. [Google Scholar] [CrossRef]
  256. Lim, J.H.; Choi, D.J.; Kim, H.-K.; Cho, W., II; Yoon, Y.S. Thin Film Supercapacitors Using a Sputtered RuO2 Electrode. J. Electrochem. Soc. 2001, 148, A275. [Google Scholar] [CrossRef]
  257. Patnaik, S.G.; Shamsudeen Seenath, J.; Bourrier, D.; Prabhudev, S.; Guay, D.; Pech, D. Porous RuOxNySzElectrodes for Microsupercapacitors and Microbatteries with Enhanced Areal Performance. ACS Energy Lett. 2021, 6, 131–139. [Google Scholar] [CrossRef]
  258. Sugimoto, W.; Iwata, H.; Yokoshima, K.; Murakami, Y.; Takasu, Y. Proton and Electron Conductivity in Hydrous Ruthenium Oxides Evaluated by Electrochemical Impedance Spectroscopy: The Origin of Large Capacitance. J. Phys. Chem. B 2005, 109, 7330–7338. [Google Scholar] [CrossRef]
  259. Yoon, Y.S.; Cho, W.I.; Lim, J.H.; Choi, D.J. Solid-State Thin-Film Supercapacitor with Ruthenium Oxide and Solid Electrolyte Thin Films. J. Power Sources 2001, 101, 126–129. [Google Scholar] [CrossRef]
  260. Ferris, A.; Garbarino, S.; Guay, D.; Pech, D. 3D RuO2 Microsupercapacitors with Remarkable Areal Energy. Adv. Mater. 2015, 27, 6625–6629. [Google Scholar] [CrossRef] [PubMed]
  261. Ponrouch, A.; Garbarino, S.; Bertin, E.; Guay, D. Ultra High Capacitance Values of Pt@RuO2 Core-Shell Nanotubular Electrodes for Microsupercapacitor Applications. J. Power Sources 2013, 221, 228–231. [Google Scholar] [CrossRef]
  262. Dinh, T.M.; Armstrong, K.; Guay, D.; Pech, D. High-Resolution on-Chip Supercapacitors with Ultra-High Scan Rate Ability. J. Mater. Chem. A Mater. 2014, 2, 7170–7174. [Google Scholar] [CrossRef]
  263. Eustache, E.; Douard, C.; Demortière, A.; De Andrade, V.; Brachet, M.; Le Bideau, J.; Brousse, T.; Lethien, C. High Areal Energy 3D-Interdigitated Micro-Supercapacitors in Aqueous and Ionic Liquid Electrolytes. Adv. Mater. Technol. 2017, 2, 1700126. [Google Scholar] [CrossRef]
  264. Li, F.; Hu, A.; Zhao, X.; Wu, T.; Chen, W.; Lei, T.; Hu, Y.; Huang, M.; Wang, X. On-Chip High-Energy Interdigital Micro-Supercapacitors with 3D Nanotubular Array Electrodes. J. Mater. Chem. A Mater. 2022, 10, 14051–14059. [Google Scholar] [CrossRef]
  265. Si, W.; Yan, C.; Chen, Y.; Oswald, S.; Han, L.; Schmidt, O.G. On Chip, All Solid-State and Flexible Micro-Supercapacitors with High Performance Based on MnOx/Au Multilayers. Energy Environ. Sci. 2013, 6, 3218–3223. [Google Scholar] [CrossRef]
  266. Wu, H.; Jiang, K.; Gu, S.; Yang, H.; Lou, Z.; Chen, D.; Shen, G. Two-Dimensional Ni(OH)2 Nanoplates for Flexible on-Chip Microsupercapacitors. Nano Res. 2015, 8, 3544–3552. [Google Scholar] [CrossRef]
  267. Li, L.; Lou, Z.; Han, W.; Shen, G. Flexible In-Plane Microsupercapacitors with Electrospun NiFe2O4 Nanofibers for Portable Sensing Applications. Nanoscale 2016, 8, 14986–14991. [Google Scholar] [CrossRef]
  268. Liu, B.; Zhang, Q.; Zhang, L.; Xu, C.; Pan, Z.; Zhou, Q.; Zhou, W.; Wang, J.; Gu, L.; Liu, H.; et al. Electrochemically Exfoliated Chlorine-Doped Graphene for Flexible All-Solid-State Micro-Supercapacitors with High Volumetric Energy Density. Adv. Mater. 2022, 34, 2106309. [Google Scholar] [CrossRef]
  269. Wu, Z.-S.; Zheng, Y.; Zheng, S.; Wang, S.; Sun, C.; Parvez, K.; Ikeda, T.; Bao, X.; Müllen, K.; Feng, X.; et al. Stacked-Layer Heterostructure Films of 2D Thiophene Nanosheets and Graphene for High-Rate All-Solid-State Pseudocapacitors with Enhanced Volumetric Capacitance. Adv. Mater. 2017, 29, 1602960. [Google Scholar] [CrossRef]
  270. Liu, L.; Lu, J.Y.; Long, X.L.; Zhou, R.; Liu, Y.Q.; Wu, Y.T.; Yan, K.W. 3D Printing of High-Performance Micro-Supercapacitors with Patterned Exfoliated Graphene/Carbon Nanotube/Silver Nanowire Electrodes. Sci. China Technol. Sci. 2021, 64, 1065–1073. [Google Scholar] [CrossRef]
  271. Yuan, Y.; Jiang, L.; Li, X.; Zuo, P.; Xu, C.; Tian, M.; Zhang, X.; Wang, S.; Lu, B.; Shao, C.; et al. Laser Photonic-Reduction Stamping for Graphene-Based Micro-Supercapacitors Ultrafast Fabrication. Nat. Commun. 2020, 11, 6185. [Google Scholar] [CrossRef]
  272. Jung, J.; Jeong, J.R.; Lee, J.; Lee, S.H.; Kim, S.Y.; Kim, M.J.; Nah, J.; Lee, M.H. In Situ Formation of Graphene/Metal Oxide Composites for High-Energy Microsupercapacitors. NPG Asia Mater. 2020, 12, 50. [Google Scholar] [CrossRef]
  273. Zhang, P.; Zhu, F.; Wang, F.; Wang, J.; Dong, R.; Zhuang, X.; Schmidt, O.G.; Feng, X.; Zhang, P.; Wang, F.; et al. Stimulus-Responsive Micro-Supercapacitors with Ultrahigh Energy Density and Reversible Electrochromic Window. Adv. Mater. 2017, 29, 1604491. [Google Scholar] [CrossRef] [PubMed]
  274. Li, H.; Tang, P.; Shen, H.; Hu, T.; Chen, J.; Chen, K.; Qi, F.; Yang, H.; Wen, L.; Li, F. Scalable Fabrication of Vanadium Carbide/Graphene Electrodes for High-Energy and Flexible Microsupercapacitors. Carbon 2021, 183, 840–849. [Google Scholar] [CrossRef]
  275. Tian, H.; Qin, J.; Hou, D.; Li, Q.; Li, C.; Wu, Z.-S.; Mai, Y.; Tian, H.; Hou, D.; Li, Q.; et al. General Interfacial Self-Assembly Engineering for Patterning Two-Dimensional Polymers with Cylindrical Mesopores on Graphene. Angew. Chem. 2019, 131, 10279–10284. [Google Scholar] [CrossRef]
  276. Lee, H.Y.; Goodenough, J.B. Supercapacitor Behavior with KCl Electrolyte. J. Solid. State Chem. 1999, 144, 220–223. [Google Scholar] [CrossRef]
  277. Toupin, M.; Brousse, T.; Bélanger, D. Influence of Microstucture on the Charge Storage Properties of Chemically Synthesized Manganese Dioxide. Chem. Mater. 2002, 14, 3946–3952. [Google Scholar] [CrossRef]
  278. Li, S.H.; Liu, Q.H.; Qi, L.; Lu, L.H.; Wang, H.Y. Progress in Research on Manganese Dioxide Electrode Aterials for Electrochemical Capacitors. Fenxi Huaxue/Chin. J. Anal. Chem. 2012, 40, 339–346. [Google Scholar] [CrossRef]
  279. Nam, K.W.; Lee, C.W.; Yang, X.Q.; Cho, B.W.; Yoon, W.S.; Kim, K.B. Electrodeposited Manganese Oxides on Three-Dimensional Carbon Nanotube Substrate: Supercapacitive Behaviour in Aqueous and Organic Electrolytes. J. Power Sources 2009, 188, 323–331. [Google Scholar] [CrossRef]
  280. Yugang, S.; Yuzi, L.; Tu, T.T.; Yang, R. Thermal Transformation of 7sigma;-MnO2 Nanoflowers Studied by in-Situ TEM. Sci. China Chem. 2012, 55, 2346–2352. [Google Scholar] [CrossRef]
  281. Fischer, A.E.; Pettigrew, K.A.; Rolison, D.R.; Stroud, R.M.; Long, J.W. Incorporation of Homogeneous, Nanoscale MnO2 within Ultraporous Carbon Structures via Self-Limiting Electroless Deposition: Implications for Electrochemical Capacitors. Nano Lett. 2007, 7, 281–286. [Google Scholar] [CrossRef] [PubMed]
  282. Qin, J.; Wang, S.; Zhou, F.; Das, P.; Zheng, S.; Sun, C.; Bao, X.; Wu, Z.S. 2D Mesoporous MnO2 Nanosheets for High-Energy Asymmetric Micro-Supercapacitors in Water-in-Salt Gel Electrolyte. Energy Storage Mater. 2019, 18, 397–404. [Google Scholar] [CrossRef]
  283. Aricò, A.S.; Bruce, P.; Scrosati, B.; Tarascon, J.M.; Van Schalkwijk, W. Nanostructured Materials for Advanced Energy Conversion and Storage Devices. Nat. Mater. 2005, 4, 366–377. [Google Scholar] [CrossRef]
  284. Liu, W.; Lu, C.; Wang, X.; Tay, R.Y.; Tay, B.K. High-Performance Microsupercapacitors Based on Two-Dimensional Graphene/Manganese Dioxide/Silver Nanowire Ternary Hybrid Film. ACS Nano 2015, 9, 1528–1542. [Google Scholar] [CrossRef]
  285. 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]
  286. Pang, S.-C.; Anderson, M.A.; Chapman, T.W. Novel Electrode Materials for Thin-Film Ultracapacitors: Comparison of Electrochemical Properties of Sol-Gel-Derived and Electrodeposited Manganese Dioxide. J. Electrochem. Soc. 2000, 147, 444. [Google Scholar] [CrossRef]
  287. Hung, C.J.; Hung, J.H.; Lin, P.; Tseng, T.Y. Electrophoretic Fabrication and Characterizations of Manganese Oxide/Carbon Nanotube Nanocomposite Pseudocapacitors. J. Electrochem. Soc. 2011, 158, A942. [Google Scholar] [CrossRef]
  288. Wang, Y.; Zhitomirsky, I. Electrophoretic Deposition of Manganese Dioxide—Multiwalled Carbon Nanotube Composites for Electrochemical Supercapacitors. Langmuir 2009, 25, 9684–9689. [Google Scholar] [CrossRef]
  289. Broughton, J.N.; Brett, M.J. Investigation of Thin Sputtered Mn Films for Electrochemical Capacitors. Electrochim. Acta 2004, 49, 4439–4446. [Google Scholar] [CrossRef]
  290. Agrawal, R.; Wang, C. On-Chip Asymmetric Microsupercapacitors Combining Reduced Graphene Oxide and Manganese Oxide for High Energy-Power Tradeoff. Micromachines 2018, 9, 399. [Google Scholar] [CrossRef]
  291. Byun, S.; Shim, Y.; Min Yuk, J.; Lee, C.W.; Yoo, J. Reduced Graphene Oxide as a Charge Reservoir of Manganese Oxide: Interfacial Interaction Promotes Charge Storage Property of MnOx-Based Micro-Supercapacitors. Chem. Eng. J. 2022, 439, 135569. [Google Scholar] [CrossRef]
  292. Dubal, D.P.; Dhawale, D.S.; Salunkhe, R.R.; Pawar, S.M.; Lokhande, C.D. A Novel Chemical Synthesis and Characterization of Mn3O4 Thin Films for Supercapacitor Application. Appl. Surf. Sci. 2010, 256, 4411–4416. [Google Scholar] [CrossRef]
  293. Yadav, A.A.; Jadhav, S.N.; Chougule, D.M.; Patil, P.D.; Chavan, U.J.; Kolekar, Y.D. Spray Deposited Hausmannite Mn3O4 Thin Films Using Aqueous/Organic Solvent Mixture for Supercapacitor Applications. Electrochim. Acta 2016, 206, 134–142. [Google Scholar] [CrossRef]
  294. Wu, Y.; Liu, S.; Wang, H.; Wang, X.; Zhang, X.; Jin, G. A Novel Solvothermal Synthesis of Mn3O4/Graphene Composites for Supercapacitors. Electrochim. Acta 2013, 90, 210–218. [Google Scholar] [CrossRef]
  295. Shi, X.; Zeng, Z.; Liao, C.; Tao, S.; Guo, E.; Long, X.; Wang, X.; Deng, D.; Dai, Y. Flexible, Planar Integratable and All-Solid-State Micro-Supercapacitors Based on Nanoporous Gold/ Manganese Oxide Hybrid Electrodes via Template Plasma Etching Method. J. Alloys Compd. 2018, 739, 979–986. [Google Scholar] [CrossRef]
  296. Li, X.J.; Song, Z.-W.; Zhao, Y.; Wang, Y.; Zhao, X.C.; Liang, M.; Chu, W.G.; Jiang, P.; Liu, Y. Vertically Porous Nickel Thin Film Supported Mn3O4 for Enhanced Energy Storage Performance. J. Colloid. Interface Sci. 2016, 483, 17–25. [Google Scholar] [CrossRef] [PubMed]
  297. Gao, M.; Dong, X.; Mei, X.; Wang, K.; Wang, W.; Zhu, C.; Duan, W.; Sun, X. Laser Direct Writing of Graphene/MnO-Mn3O4 Doped with Sulfur for High-Performance Microsupercapacitors. J. Energy Storage 2022, 49, 104118. [Google Scholar] [CrossRef]
  298. Li, Y.Q.; Shi, X.M.; Lang, X.Y.; Wen, Z.; Li, J.C.; Jiang, Q. Remarkable Improvements in Volumetric Energy and Power of 3D MnO2 Microsupercapacitors by Tuning Crystallographic Structures. Adv. Funct. Mater. 2016, 26, 1830–1839. [Google Scholar] [CrossRef]
  299. Bounor, B.; Seenath, J.S.; Patnaik, S.G.; Bourrier, D.; Tran, C.C.H.; Esvan, J.; Weingarten, L.; Descamps-Mandine, A.; Rochefort, D.; Guay, D.; et al. Low-Cost Micro-Supercapacitors Using Porous Ni/MnO2 Entangled Pillars and Na-Based Ionic Liquids. Energy Storage Mater. 2023, 63, 102986. [Google Scholar] [CrossRef]
  300. Liu, C.; Li, Z.; Zhang, Z. Molybdenum Oxide Film with Stable Pseudocapacitive Property for Aqueous Micro-Scale Electrochemical Capacitor. Electrochim. Acta 2014, 134, 84–91. [Google Scholar] [CrossRef]
  301. Zhu, Y.; Ji, X.; Cheng, S.; Chern, Z.Y.; Jia, J.; Yang, L.; Luo, H.; Yu, J.; Peng, X.; Wang, J.; et al. Fast Energy Storage in Two-Dimensional MoO2 Enabled by Uniform Oriented Tunnels. ACS Nano 2019, 13, 9091–9099. [Google Scholar] [CrossRef] [PubMed]
  302. Wang, S.; Yong, H.; Yao, J.; Ma, J.; Liu, B.; Hu, J.; Zhang, Y. Influence of the Phase Evolution and Hydrogen Storage Behaviors of Mg-RE Alloy by a Multi-Valence Mo-Based Catalyst. J. Energy Storage 2023, 58, 106397. [Google Scholar] [CrossRef]
  303. Vincent, R.C.; Luo, Y.; Andrews, J.L.; Zohar, A.; Zhou, Y.; Yan, Q.; Mozur, E.M.; Preefer, M.B.; Weker, J.N.; Cheetham, A.K.; et al. High-Rate Lithium Cycling and Structure Evolution in Mo4O11. Chem. Mater. 2022, 34, 4122–4133. [Google Scholar] [CrossRef]
  304. Ge, H.; Kuwahara, Y.; Yamashita, H. Development of Defective Molybdenum Oxides for Photocatalysis, Thermal Catalysis, and Photothermal Catalysis. Chem. Commun. 2022, 58, 8466–8479. [Google Scholar] [CrossRef] [PubMed]
  305. Wang, Y.; Wang, H.; Qu, G. Molybdenum-Based Electrode Materials Applied in High-Performance Supercapacitors. Batteries 2023, 9, 479. [Google Scholar] [CrossRef]
  306. Maheswari, N.; Muralidharan, G. Controlled Synthesis of Nanostructured Molybdenum Oxide Electrodes for High Performance Supercapacitor Devices. Appl. Surf. Sci. 2017, 416, 461–469. [Google Scholar] [CrossRef]
  307. Tang, W.; Liu, L.; Tian, S.; Li, L.; Yue, Y.; Wu, Y.; Zhu, K. Aqueous Supercapacitors of High Energy Density Based on MoO3 Nanoplates as Anode Material. Chem. Commun. 2011, 47, 10058–10060. [Google Scholar] [CrossRef] [PubMed]
  308. Liang, R.; Cao, H.; Qian, D. MoO3 Nanowires as Electrochemical Pseudocapacitor Materials. Chem. Commun. 2011, 47, 10305–10307. [Google Scholar] [CrossRef]
  309. Zhao, X.Y.; Gong, L.G.; Wang, C.X.; Wang, C.M.; Yu, K.; Zhou, B.B. Silver-Doped Molybdenum Oxide Ternary Nanoparticles with Excellent Supercapacitor and Hydrogen Peroxide-Sensing Performance. Mater. Today Energy 2021, 20, 100774. [Google Scholar] [CrossRef]
  310. Xu, Y.; Deng, P.; Chen, R.; Xie, W.; Xu, Z.; Yang, Y.; Liu, D.; Huang, F.; Zhuang, Z.; Zhitomirsky, I.; et al. Electric Discharge Direct Writing of 3D Mo-MoOx Pseudocapacitive Micro-Supercapacitors with Designable Patterns. Ceram. Int. 2023, 49, 22586–22594. [Google Scholar] [CrossRef]
  311. Liu, C.; Li, Z.; Zhang, Z. Growth of [010] Oriented α-MoO3 Nanorods by Pulsed Electron Beam Deposition. Appl. Phys. Lett. 2011, 99, 223104. [Google Scholar] [CrossRef]
  312. Prakash, N.G.; Dhananjaya, M.; Narayana, A.L.; Shaik, D.P.; Rosaiah, P.; Hussain, O.M. High Performance One Dimensional α-MoO3 Nanorods for Supercapacitor Applications. Ceram. Int. 2018, 44, 9967–9975. [Google Scholar] [CrossRef]
  313. Shaheen, I.; Ahmad, K.S.; Zequine, C.; Gupta, R.K.; Thomas, A.G.; Malik, M.A. Electrochemical Energy Storage by Nanosized MoO3/PdO Material: Investigation of Its Structural, Optical and Electrochemical Properties for Supercapacitor. J. Energy Storage 2021, 36, 102447. [Google Scholar] [CrossRef]
  314. Brezesinski, T.; Wang, J.; Tolbert, S.H.; Dunn, B. Ordered Mesoporous α-MoO3 with Iso-Oriented Nanocrystalline Walls for Thin-Film Pseudocapacitors. Nat. Mater. 2010, 9, 146–151. [Google Scholar] [CrossRef]
  315. Kim, H.S.; Cook, J.B.; Lin, H.; Ko, J.S.; Tolbert, S.H.; Ozolins, V.; Dunn, B. Oxygen Vacancies Enhance Pseudocapacitive Charge Storage Properties of MoO3-x. Nat. Mater. 2017, 16, 454–460. [Google Scholar] [CrossRef]
  316. Gurusamy, L.; Karuppasamy, L.; Anandan, S.; Liu, N.; Lee, G.J.; Liu, C.H.; Wu, J.J. Enhanced Performance of Charge Storage Supercapattery by Dominant Oxygen Deficiency in Crystal Defects of 2-D MoO3-x Nanoplates. Appl. Surf. Sci. 2021, 541, 148676. [Google Scholar] [CrossRef]
  317. Xiao, X.; Ding, T.; Yuan, L.; Shen, Y.; Zhong, Q.; Zhang, X.; Cao, Y.; Hu, B.; Zhai, T.; Gong, L.; et al. WO3-x/MoO3-x Core/Shell Nanowires on Carbon Fabric as an Anode for All-Solid-State Asymmetric Supercapacitors. Adv. Energy Mater. 2012, 2, 1328–1332. [Google Scholar] [CrossRef]
  318. Huang, L.; Yao, B.; Sun, J.; Gao, X.; Wu, J.; Wan, J.; Li, T.; Hu, Z.; Zhou, J. Highly Conductive and Flexible Molybdenum Oxide Nanopaper for High Volumetric Supercapacitor Electrode. J. Mater. Chem. A Mater. 2017, 5, 2897–2903. [Google Scholar] [CrossRef]
  319. Guru Prakash, N.; Dhananjaya, M.; Lakshmi Narayana, A.; Hussain, O.M. One-Dimensional MoO3/Pd Nanocomposite Electrodes for High Performance Supercapacitors. Mater. Res. Express 2019, 6, 085543. [Google Scholar] [CrossRef]
  320. Cheng, A.; Shen, Y.; Hong, T.; Zhan, R.; Chen, E.; Chen, Z.; Chen, G.; Liang, M.; Sun, X.; Wang, D.; et al. Self-Assembly Vertical Graphene-Based MoO3 Nanosheets for High Performance Supercapacitors. Nanomaterials 2022, 12, 2057. [Google Scholar] [CrossRef]
  321. Zhang, X.; Zhao, W.; Wei, L.; Jin, Y.; Hou, J.; Wang, X.; Guo, X. In-Plane Flexible Solid-State Microsupercapacitors for on-Chip Electronics. Energy 2019, 170, 338–348. [Google Scholar] [CrossRef]
  322. Chen, J.; Han, S.; Zhao, H.; Bai, J.; Wang, L.; Sun, G.; Zhang, Z.; Pan, X.; Zhou, J.; Xie, E. Robust Wire-Based Supercapacitors Based on Hierarchical A-MoO3 Nanosheet Arrays with Well-Aligned Laminated Structure. Chem. Eng. J. 2017, 320, 34–42. [Google Scholar] [CrossRef]
  323. Singh, B.K.; Shaikh, A.; Badrayyana, S.; Mohapatra, D.; Dusane, R.O.; Parida, S. Nanoporous Gold-Copper Oxide Based All-Solid-State Micro-Supercapacitors. RSC Adv. 2016, 6, 100467–100475. [Google Scholar] [CrossRef]
  324. Zhang, L.; Chen, Z.; Zheng, S.; Qin, S.; Wang, J.; Chen, C.; Liu, D.; Wang, L.; Yang, G.; Su, Y.; et al. Shape-Tailorable High-Energy Asymmetric Micro-Supercapacitors Based on Plasma Reduced and Nitrogen-Doped Graphene Oxide and MoO2 Nanoparticles. J. Mater. Chem. A Mater. 2019, 7, 14328–14336. [Google Scholar] [CrossRef]
  325. Lin, N.; Chen, H.; Wang, W.; Lu, L. Laser-Induced Graphene/MoO2 Core-Shell Electrodes on Carbon Cloth for Integrated, High-Voltage, and In-Planar Microsupercapacitors. Adv. Mater. Technol. 2021, 6, 2000991. [Google Scholar] [CrossRef]
  326. Zhang, L.; Yang, G.; Chen, Z.; Liu, D.; Wang, J.; Qian, Y.; Chen, C.; Liu, Y.; Wang, L.; Razal, J.; et al. MXene Coupled with Molybdenum Dioxide Nanoparticles as 2D-0D Pseudocapacitive Electrode for High Performance Flexible Asymmetric Micro-Supercapacitors. J. Mater. 2020, 6, 138–144. [Google Scholar] [CrossRef]
  327. Liu, D.-Q.; Yu, S.-H.; Son, S.-W.; Joo, S.-K. Electrochemical Performance of Iridium Oxide Thin Film for Supercapacitor Prepared by Radio Frequency Magnetron Sputtering Method. ECS Trans. 2008, 16, 103. [Google Scholar] [CrossRef]
  328. Liao, P.C.; Ho, W.S.; Huang, Y.S.; Tiong, K.K. Characterization of Sputtered Iridium Dioxide Thin Films. J. Mater. Res. 1998, 13, 1318–1326. [Google Scholar] [CrossRef]
  329. Beknalkar, S.A.; Teli, A.M.; Harale, N.S.; Shin, J.C.; Patil, P.S. Construction of IrO2@Mn3O4 Core-Shell Heterostructured Nanocomposites for High Performance Symmetric Supercapacitor Device. J. Alloys Compd. 2021, 887, 161328. [Google Scholar] [CrossRef]
  330. Sethuraman, B.; Purushothaman, K.K.; Muralidharan, G. Synthesis of Mesh-like Fe2O3/C Nanocomposite via Greener Route for High Performance Supercapacitors. RSC Adv. 2014, 4, 4631–4637. [Google Scholar] [CrossRef]
  331. Tang, Q.; Wang, W.; Wang, G. The Perfect Matching between the Low-Cost Fe2O3 Nanowire Anode and the NiO Nanoflake Cathode Significantly Enhances the Energy Density of Asymmetric Supercapacitors. J. Mater. Chem. A Mater. 2015, 3, 6662–6670. [Google Scholar] [CrossRef]
  332. Kulal, P.M.; Dubal, D.P.; Lokhande, C.D.; Fulari, V.J. Chemical Synthesis of Fe2O3 Thin Films for Supercapacitor Application. J. Alloys Compd. 2011, 509, 2567–2571. [Google Scholar] [CrossRef]
  333. Xie, K.; Li, J.; Lai, Y.; Lu, W.; Zhang, Z.; Liu, Y.; Zhou, L.; Huang, H. Highly Ordered Iron Oxide Nanotube Arrays as Electrodes for Electrochemical Energy Storage. Electrochem. Commun. 2011, 13, 657–660. [Google Scholar] [CrossRef]
  334. Gund, G.S.; Dubal, D.P.; Chodankar, N.R.; Cho, J.Y.; Gomez-Romero, P.; Park, C.; Lokhande, C.D. Low-Cost Flexible Supercapacitors with High-Energy Density Based on Nanostructured MnO2 and Fe2O3 Thin Films Directly Fabricated onto Stainless Steel. Sci. Rep. 2015, 5, 12454. [Google Scholar] [CrossRef] [PubMed]
  335. Shivakumara, S.; Penki, T.R.; Munichandraiah, N. High Specific Surface Area α-Fe2O3 Nanostructures as High Performance Electrode Material for Supercapacitors. Mater. Lett. 2014, 131, 100–103. [Google Scholar] [CrossRef]
  336. Wang, D.; Wang, Q.; Wang, T. Controlled Synthesis of Mesoporous Hematite Nanostructures and Their Application as Electrochemical Capacitor Electrodes. Nanotechnology 2011, 22, 135604. [Google Scholar] [CrossRef]
  337. Zeng, Y.; Yu, M.; Meng, Y.; Fang, P.; Lu, X.; Tong, Y. Iron-Based Supercapacitor Electrodes: Advances and Challenges. Adv. Energy Mater. 2016, 6, 1601053. [Google Scholar] [CrossRef]
  338. Xia, Q.; Xu, M.; Xia, H.; Xie, J. Nanostructured Iron Oxide/Hydroxide-Based Electrode Materials for Supercapacitors. ChemNanoMat 2016, 2, 588–600. [Google Scholar] [CrossRef]
  339. Ma, J.; Guo, X.; Yan, Y.; Xue, H.; Pang, H. FeOx-Based Materials for Electrochemical Energy Storage. Adv. Sci. 2018, 5, 1700986. [Google Scholar] [CrossRef]
  340. Yu, S.; Hong Ng, V.M.; Wang, F.; Xiao, Z.; Li, C.; Kong, L.B.; Que, W.; Zhou, K. Synthesis and Application of Iron-Based Nanomaterials as Anodes of Lithium-Ion Batteries and Supercapacitors. J. Mater. Chem. A Mater. 2018, 6, 9332–9367. [Google Scholar] [CrossRef]
  341. El Issmaeli, Y.; Lahrichi, A.; Kalanur, S.S.; Natarajan, S.K.; Pollet, B.G. Recent Advances and Prospects of FeOOH-Based Electrode Materials for Supercapacitors. Batteries 2023, 9, 259. [Google Scholar] [CrossRef]
  342. Sakurai, S.; Namai, A.; Hashimoto, K.; Ohkoshi, S.I. First Observation of Phase Transformation of All Four Fe2O3 Phases (γ → ε → β → α-Phase). J. Am. Chem. Soc. 2009, 131, 18299–18303. [Google Scholar] [CrossRef]
  343. Danno, T.; Nakatsuka, D.; Kusano, Y.; Asaoka, H.; Nakanishi, M.; Fujii, T.; Ikeda, Y.; Takada, J. Crystal Structure of β-Fe2O3 and Topotactic Phase Transformation to α-Fe2O3. Cryst. Growth Des. 2013, 13, 770–774. [Google Scholar] [CrossRef]
  344. An, C.; Zhang, Y.; Guo, H.; Wang, Y. Metal Oxide-Based Supercapacitors: Progress and Prospectives. Nanoscale Adv. 2019, 1, 4644–4658. [Google Scholar] [CrossRef]
  345. Wang, T.; Chen, H.C.; Yu, F.; Zhao, X.S.; Wang, H. Boosting the Cycling Stability of Transition Metal Compounds-Based Supercapacitors. Energy Storage Mater. 2019, 16, 545–573. [Google Scholar] [CrossRef]
  346. Zhang, H.; Cao, Y.; Chee, M.O.L.; Dong, P.; Ye, M.; Shen, J. Recent Advances in Micro-Supercapacitors. Nanoscale 2019, 11, 5807–5821. [Google Scholar] [CrossRef]
  347. Sheng, S.; Liu, W.; Zhu, K.; Cheng, K.; Ye, K.; Wang, G.; Cao, D.; Yan, J. Fe3O4 Nanospheres in Situ Decorated Graphene as High-Performance Anode for Asymmetric Supercapacitor with Impressive Energy Density. J. Colloid. Interface Sci. 2019, 536, 235–244. [Google Scholar] [CrossRef]
  348. Zhu, M.; Meng, W.; Huang, Y.; Huang, Y.; Zhi, C. Proton-Insertion-Enhanced Pseudocapacitance Based on the Assembly Structure of Tungsten Oxide. ACS Appl. Mater. Interfaces 2014, 6, 18901–18910. [Google Scholar] [CrossRef]
  349. Zheng, F.; Gong, H.; Li, Z.; Yang, W.; Xu, J.; Hu, P.; Li, Y.; Gong, Y.; Zhen, Q. Tertiary Structure of Cactus-like WO3 Spheres Self-Assembled on Cu Foil for Supercapacitive Electrode Materials. J. Alloys Compd. 2017, 712, 345–354. [Google Scholar] [CrossRef]
  350. Yoon, S.; Kang, E.; Kon Kim, J.; Wee Lee, C.; Lee, J. Development of High-Performance Supercapacitor Electrodes Using Novel Ordered Mesoporous Tungsten Oxide Materials with High Electrical Conductivity. Chem. Commun. 2011, 47, 1021–1023. [Google Scholar] [CrossRef]
  351. Yang, P.; Sun, P.; Du, L.; Liang, Z.; Xie, W.; Cai, X.; Huang, L.; Tan, S.; Mai, W. Quantitative Analysis of Charge Storage Process of Tungsten Oxide That Combines Pseudocapacitive and Electrochromic Properties. J. Phys. Chem. C 2015, 119, 16483–16489. [Google Scholar] [CrossRef]
  352. Liu, X.; Sheng, G.; Zhong, M.; Zhou, X. Dispersed and Size-Selected WO3 Nanoparticles in Carbon Aerogel for Supercapacitor Applications. Mater. Des. 2018, 141, 220–229. [Google Scholar] [CrossRef]
  353. Jia, J.; Liu, X.; Mi, R.; Liu, N.; Xiong, Z.; Yuan, L.; Wang, C.; Sheng, G.; Cao, L.; Zhou, X.; et al. Self-Assembled Pancake-like Hexagonal Tungsten Oxide with Ordered Mesopores for Supercapacitors. J. Mater. Chem. A Mater. 2018, 6, 15330–15339. [Google Scholar] [CrossRef]
  354. Huang, X.; Liu, H.; Zhang, X.; Jiang, H. High Performance All-Solid-State Flexible Micro-Pseudocapacitor Based on Hierarchically Nanostructured Tungsten Trioxide Composite. ACS Appl. Mater. Interfaces 2015, 7, 27845–27852. [Google Scholar] [CrossRef]
  355. Shinde, P.A.; Jun, S.C. Review on Recent Progress in the Development of Tungsten Oxide Based Electrodes for Electrochemical Energy Storage. ChemSusChem 2020, 13, 11–38. [Google Scholar] [CrossRef]
  356. Cong, S.; Geng, F.; Zhao, Z. Tungsten Oxide Materials for Optoelectronic Applications. Adv. Mater. 2016, 28, 10518–10528. [Google Scholar] [CrossRef] [PubMed]
  357. Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive Oxide Materials for High-Rate Electrochemical Energy Storage. Energy Environ. Sci. 2014, 7, 1597–1614. [Google Scholar] [CrossRef]
  358. Lin, J.; Du, X. High Performance Asymmetric Supercapacitor Based on Hierarchical Carbon Cloth in Situ Deposited with H-Wo3 Nanobelts as Negative Electrode and Carbon Nanotubes as Positive Electrode. Micromachines 2021, 12, 1195. [Google Scholar] [CrossRef]
  359. Shi, F.; Li, J.; Xiao, J.; Zhao, X.; Li, H.; An, Q.; Zhai, S.; Wang, K.; Wei, L.; Tong, Y. Three-Dimensional Hierarchical Porous Lignin-Derived Carbon/WO3 for High-Performance Solid-State Planar Micro-Supercapacitor. Int. J. Biol. Macromol. 2021, 190, 11–18. [Google Scholar] [CrossRef]
  360. Hepel, M.; Petrukhina, M.A.; Samuilov, V. High Power-Density WO3-x–Grafted Corannulene-Modified Graphene Nanostructures for Micro-Supercapacitors. J. Electroanal. Chem. 2023, 928, 116990. [Google Scholar] [CrossRef]
  361. Lillard, R.S.; Kanner, G.S.; Butt, D.P. The Nature of Oxide Films on Tungsten in Acidic and Alkaline Solutions. J. Electrochem. Soc. 1998, 145, 2718. [Google Scholar] [CrossRef]
  362. Anik, M.; Osseo-Asare, K. Effect of PH on the Anodic Behavior of Tungsten. J. Electrochem. Soc. 2002, 149, B224. [Google Scholar] [CrossRef]
  363. Di Paola, A.; Di Quarto, F.; Sunseri, C. Electrochromism in Anodically Formed Tungsten Oxide Films. J. Electrochem. Soc. 1978, 125, 1344. [Google Scholar] [CrossRef]
  364. El-Basiouny, M.S.; Hassan, S.A.; Hefny, M.M. On the Electrochemical Behaviour of Tungsten: The Formation and Dissolution of Tungsten Oxide in Sulphuric Acid Solutions. Corros. Sci. 1980, 20, 909–917. [Google Scholar] [CrossRef]
  365. 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]
  366. Yan, Y.; Li, B.; Guo, W.; Pang, H.; Xue, H. Vanadium Based Materials as Electrode Materials for High Performance Supercapacitors. J. Power Sources 2016, 329, 148–169. [Google Scholar] [CrossRef]
  367. Lai, W.-Y.; Wei Huang, P.; Kang, J.; Zhang, Q.; Wang, Y.; Wen-Yong Lai, S.; Zhang, Y.-Z.; Wang, Y.; Cheng, T.; Yao, L.-Q.; et al. Printed Supercapacitors: Materials, Printing and Applications. Chem. Soc. Rev. 2019, 48, 3229–3264. [Google Scholar] [CrossRef]
  368. Huang, Z.H.; Song, Y.; Liu, X.X. Boosting Operating Voltage of Vanadium Oxide-Based Symmetric Aqueous Supercapacitor to 2 V. Chem. Eng. J. 2019, 358, 1529–1538. [Google Scholar] [CrossRef]
  369. Arico, C.; Ouendi, S.; Taberna, P.L.; Roussel, P.; Simon, P.; Lethien, C. Fast Electrochemical Storage Process in Sputtered Nb2O5 Porous Thin Films. ACS Nano 2019, 13, 5826–5832. [Google Scholar] [CrossRef]
  370. Ahmad, M.W.; Dey, B.; Syed, A.; Bahkali, A.H.; Verma, M.; Yang, D.J.; Choudhury, A. MOFs-Derived Niobium Oxide Nanoparticles/Carbon Nanofiber Hybrid Paper as Flexible Binder-Free Electrode for Solid-State Asymmetric Supercapacitors. J. Alloys Compd. 2023, 957, 170269. [Google Scholar] [CrossRef]
  371. Brousse, K.; Taberna, P.L.; Simon, P. Current Collector-Free Interdigitated Nb2O5//LiFePO4 Micro-Batteries Prepared by a Simple Laser-Writing Process. J. Electrochem. Soc. 2022, 169, 070534. [Google Scholar] [CrossRef]
  372. Yuan, Y.; Li, X.; Jiang, L.; Liang, M.; Zhang, X.; Wu, S.; Wu, J.; Tian, M.; Zhao, Y.; Qu, L. Laser Maskless Fast Patterning for Multitype Microsupercapacitors. Nat. Commun. 2023, 14, 3967. [Google Scholar] [CrossRef] [PubMed]
  373. Zhang, H.; Wang, B.; Brown, B. Atomic Layer Deposition of Titanium Oxide and Nitride on Vertically Aligned Carbon Nanotubes for Energy Dense 3D Microsupercapacitors. Appl. Surf. Sci. 2020, 521, 146349. [Google Scholar] [CrossRef]
  374. Siddiqui, S.A.; Das, S.; Rani, S.; Afshan, M.; Pahuja, M.; Jain, A.; Rani, D.; Chaudhary, N.; Jyoti; Ghosh, R.; et al. Phosphorus-Doped Nickel Oxide Micro-Supercapacitor: Unleashing the Power of Energy Storage for Miniaturized Electronic Devices. Small 2023, 20, 2306756. [Google Scholar] [CrossRef]
  375. Zhang, K.; Ying, G.; Liu, L.; Ma, F.; Su, L.; Zhang, C.; Wu, D.; Wang, X.; Zhou, Y. Three-Dimensional Porous Ti3C2 Tx-NiO Composite Electrodes with Enhanced Electrochemical Performance for Supercapacitors. Materials 2019, 12, 188. [Google Scholar] [CrossRef]
  376. Rao, Y.; Yuan, M.; Luo, F.; Li, H.; Yu, J.; Chen, X. Laser In-Situ Synthesis of Metallic Cobalt Decorated Porous Graphene for Flexible In-Plane Microsupercapacitors. J. Colloid. Interface Sci. 2022, 610, 775–784. [Google Scholar] [CrossRef]
  377. Zhu, Y.G.; Wang, Y.; Shi, Y.; Wong, J.I.; Yang, H.Y. CoO Nanoflowers Woven by CNT Network for High Energy Density Flexible Micro-Supercapacitor. Nano Energy 2014, 3, 46–54. [Google Scholar] [CrossRef]
  378. Li, M.; Jia, C.; Zhang, D.; Luo, Y.; Ma, Y.; Luo, G.; Zhao, L.; Wang, L.; Li, Z.; Lin, Q.; et al. An Efficient Cobalt-Nickel Phosphate Positive Electrode for High-Performance Hybrid Microsupercapacitors. J. Energy Storage 2023, 64, 107144. [Google Scholar] [CrossRef]
  379. Mahendra, G.; Malathi, R.; Kedhareswara, S.P.; Lakshmi-Narayana, A.; Dhananjaya, M.; Guruprakash, N.; Hussain, O.M.; Mauger, A.; Julien, C.M. RF Sputter-Deposited Nanostructured CuO Films for Micro-Supercapacitors. Appl. Nano 2021, 2, 46–66. [Google Scholar] [CrossRef]
  380. Sayed, D.M.; Salem, K.E.; Allam, N.K. Optimized Lithography-Free Fabrication of Sub-100 Nm Nb2O5 Nanotube Films as Negative Supercapacitor Electrodes: Tuned Oxygen Vacancies and Cationic Intercalation. ACS Appl. Mater. Interfaces 2022, 14, 25545–25555. [Google Scholar] [CrossRef] [PubMed]
  381. Zhai, X.; Liu, J.; Zhao, Y.; Chen, C.; Zhao, X.; Li, J.; Jin, H. Oxygen Vacancy Boosted the Electrochemistry Performance of Ti4+ Doped Nb2O5 toward Lithium Ion Battery. Appl. Surf. Sci. 2020, 499, 143905. [Google Scholar] [CrossRef]
  382. Taffa, D.H.; Dillert, R.; Ulpe, A.C.; Bauerfeind, K.C.L.; Bredow, T.; Bahnemann, D.W.; Wark, M. Photoelectrochemical and Theoretical Investigations of Spinel Type Ferrites (MxFe3xO4) for Water Splitting: A Mini-Review. J. Photonics Energy 2016, 7, 012009. [Google Scholar] [CrossRef]
  383. Szotek, Z.; Temmerman, W.M.; Ködderitzsch, D.; Svane, A.; Petit, L.; Winter, H. Electronic Structures of Normal and Inverse Spinel Ferrites from First Principles. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 74, 174431. [Google Scholar] [CrossRef]
  384. Lin, Y.P.; Wu, N.L. Characterization of MnFe2O4/LiMn2O4 Aqueous Asymmetric Supercapacitor. J. Power Sources 2011, 196, 851–854. [Google Scholar] [CrossRef]
  385. Zhu, B.; Tang, S.; Vongehr, S.; Xie, H.; Zhu, J.; Meng, X. FeCo2O4 Submicron-Tube Arrays Grown on Ni Foam as High Rate-Capability and Cycling-Stability Electrodes Allowing Superior Energy and Power Densities with Symmetric Supercapacitors. Chem. Commun. 2016, 52, 2624–2627. [Google Scholar] [CrossRef]
  386. Lannelongue, P.; Le Vot, S.; Fontaine, O.; Brousse, T.; Favier, F. Electrochemical Study of Asymmetric Aqueous Supercapacitors Based on High Density Oxides: C/Ba0.5Sr0.5Co0.8Fe0.2O3-δ and FeWO4/Ba0.5Sr0.5Co0.8Fe0.2O3-δ. Electrochim. Acta 2019, 326, 134886. [Google Scholar] [CrossRef]
  387. Goubard-Bretesché, N.; Crosnier, O.; Payen, C.; Favier, F.; Brousse, T. Nanocrystalline FeWO4 as a Pseudocapacitive Electrode Material for High Volumetric Energy Density Supercapacitors Operated in an Aqueous Electrolyte. Electrochem. Commun. 2015, 57, 61–64. [Google Scholar] [CrossRef]
  388. Goubard-Bretesché, N.; Crosnier, O.; Douard, C.; Iadecola, A.; Retoux, R.; Payen, C.; Doublet, M.L.; Kisu, K.; Iwama, E.; Naoi, K.; et al. Unveiling Pseudocapacitive Charge Storage Behavior in FeWO4 Electrode Material by Operando X-Ray Absorption Spectroscopy. Small 2020, 16, 2002855. [Google Scholar] [CrossRef]
  389. Xing, X.; Gui, Y.; Zhang, G.; Song, C. CoWO4 Nanoparticles Prepared by Two Methods Displaying Different Structures and Supercapacitive Performances. Electrochim. Acta 2015, 157, 15–22. [Google Scholar] [CrossRef]
  390. Xu, X.; Gao, J.; Huang, G.; Qiu, H.; Wang, Z.; Wu, J.; Pan, Z.; Xing, F. Fabrication of CoWO4@NiWO4 Nanocomposites with Good Supercapacitve Performances. Electrochim. Acta 2015, 174, 837–845. [Google Scholar] [CrossRef]
  391. Yang, Y.; Zhu, J.; Shi, W.; Zhou, J.; Gong, D.; Gu, S.; Wang, L.; Xu, Z.; Lu, B. 3D Nanoporous ZnWO4 Nanoparticles with Excellent Electrochemical Performances for Supercapacitors. Mater. Lett. 2016, 177, 34–38. [Google Scholar] [CrossRef]
  392. Jadhav, S.; Donolikar, P.D.; Chodankar, N.R.; Dongale, T.D.; Dubal, D.P.; Patil, D.R. Nano-Dimensional Iron Tungstate for Super High Energy Density Symmetric Supercapacitor with Redox Electrolyte. J. Solid. State Electrochem. 2019, 23, 3459–3465. [Google Scholar] [CrossRef]
  393. Kuo, S.-L.; Lee, J.-F.; Wu, N.-L. Study on Pseudocapacitance Mechanism of Aqueous MnFe2O4 Supercapacitor. J. Electrochem. Soc. 2007, 154, A34. [Google Scholar] [CrossRef]
  394. Kuo, S.L.; Wu, N.L. Electrochemical Capacitor of MnFe2O4 with NaCl Electrolyte. Electrochem. Solid-State Lett. 2005, 8, A495. [Google Scholar] [CrossRef]
  395. Ouendi, S.; Robert, K.; Stievenard, D.; Brousse, T.; Roussel, P.; Lethien, C. Sputtered Tungsten Nitride Films as Pseudocapacitive Electrode for on Chip Micro-Supercapacitors. Energy Storage Mater. 2019, 20, 243–252. [Google Scholar] [CrossRef]
  396. Arif, M.; Sanger, A.; Singh, A. Sputter Deposited Chromium Nitride Thin Electrodes for Supercapacitor Applications. Mater. Lett. 2018, 220, 213–217. [Google Scholar] [CrossRef]
  397. Bouhtiyya, S.; Lucio Porto, R.; Laïk, B.; Boulet, P.; Capon, F.; Pereira-Ramos, J.P.; Brousse, T.; Pierson, J.F. Application of Sputtered Ruthenium Nitride Thin Films as Electrode Material for Energy-Storage Devices. Scr. Mater. 2013, 68, 659–662. [Google Scholar] [CrossRef]
  398. Adalati, R.; Sharma, M.; Sharma, S.; Kumar, A.; Malik, G.; Boukherroub, R.; Chandra, R. Metal Nitrides as Efficient Electrode Material for Supercapacitors: A Review. J. Energy Storage 2022, 56, 105912. [Google Scholar] [CrossRef]
  399. Robert, K.; Stiévenard, D.; Deresmes, D.; Douard, C.; Iadecola, A.; Troadec, D.; Simon, P.; Nuns, N.; Marinova, M.; Huvé, M.; et al. Novel Insights into the Charge Storage Mechanism in Pseudocapacitive Vanadium Nitride Thick Films for High-Performance on-Chip Micro-Supercapacitors. Energy Environ. Sci. 2020, 13, 949–957. [Google Scholar] [CrossRef]
  400. Lucio-Porto, R.; Bouhtiyya, S.; Pierson, J.F.; Morel, A.; Capon, F.; Boulet, P.; Brousse, T. VN Thin Films as Electrode Materials for Electrochemical Capacitors. Electrochim. Acta 2014, 141, 203–211. [Google Scholar] [CrossRef]
  401. Bondarchuk, O.; Morel, A.; Bélanger, D.; Goikolea, E.; Brousse, T.; Mysyk, R. Thin Films of Pure Vanadium Nitride: Evidence for Anomalous Non-Faradaic Capacitance. J. Power Sources 2016, 324, 439–446. [Google Scholar] [CrossRef]
  402. Zhou, Y.; Guo, W.; Li, T. A Review on Transition Metal Nitrides as Electrode Materials for Supercapacitors. Ceram. Int. 2019, 45, 21062–21076. [Google Scholar] [CrossRef]
  403. Durai, G.; Kuppusami, P.; Maiyalagan, T.; Theerthagiri, J.; Vinoth Kumar, P.; Kim, H.S. Influence of Chromium Content on Microstructural and Electrochemical Supercapacitive Properties of Vanadium Nitride Thin Films Developed by Reactive Magnetron Co-Sputtering Process. Ceram. Int. 2019, 45, 12643–12653. [Google Scholar] [CrossRef]
  404. Li, Q.; Chen, Y.; Zhang, J.; Tian, W.; Wang, L.; Ren, Z.; Ren, X.; Li, X.; Gao, B.; Peng, X.; et al. Spatially Confined Synthesis of Vanadium Nitride Nanodots Intercalated Carbon Nanosheets with Ultrahigh Volumetric Capacitance and Long Life for Flexible Supercapacitors. Nano Energy 2018, 51, 128–136. [Google Scholar] [CrossRef]
  405. Chen, L.; Liu, C.; Zhang, Z. Novel [111] Oriented γ-Mo2N Thin Films Deposited by Magnetron Sputtering as an Anode for Aqueous Micro-Supercapacitors. Electrochim. Acta 2017, 245, 237–248. [Google Scholar] [CrossRef]
  406. Dinh, K.H.; Robert, K.; Thuriot-Roukos, J.; Huvé, M.; Simon, P.; Troadec, D.; Lethien, C.; Roussel, P. Wafer-Scale Performance Mapping of Magnetron-Sputtered Ternary Vanadium Tungsten Nitride for Microsupercapacitors. Chem. Mater. 2023, 35, 8654–8663. [Google Scholar] [CrossRef]
  407. Wei, B.; Liang, H.; Zhang, D.; Qi, Z.; Shen, H.; Wang, Z. Magnetron Sputtered TiN Thin Films toward Enhanced Performance Supercapacitor Electrodes. Mater. Renew. Sustain. Energy 2018, 7, 11. [Google Scholar] [CrossRef]
  408. Achour, A.; Porto, R.L.; Soussou, M.A.; Islam, M.; Boujtita, M.; Aissa, K.A.; Le Brizoual, L.; Djouadi, A.; Brousse, T. Titanium Nitride Films for Micro-Supercapacitors: Effect of Surface Chemistry and Film Morphology on the Capacitance. J. Power Sources 2015, 300, 525–532. [Google Scholar] [CrossRef]
  409. Achour, A.; Lucio-Porto, R.; Chaker, M.; Arman, A.; Ahmadpourian, A.; Soussou, M.A.; Boujtita, M.; Le Brizoual, L.; Djouadi, M.A.; Brousse, T. Titanium Vanadium Nitride Electrode for Micro-Supercapacitors. Electrochem. Commun. 2017, 77, 40–43. [Google Scholar] [CrossRef]
  410. Cui, H.; Zhu, G.; Liu, X.; Liu, F.; Xie, Y.; Yang, C.; Lin, T.; Gu, H.; Huang, F. Niobium Nitride Nb4N5 as a New High-Performance Electrode Material for Supercapacitors. Adv. Sci. 2015, 2, 1500126. [Google Scholar] [CrossRef] [PubMed]
  411. Wei, B.; Liang, H.; Zhang, D.; Wu, Z.; Qi, Z.; Wang, Z. CrN Thin Films Prepared by Reactive DC Magnetron Sputtering for Symmetric Supercapacitors. J. Mater. Chem. A Mater. 2017, 5, 2844–2851. [Google Scholar] [CrossRef]
  412. Qi, Z.; Wei, B.; Wang, J.; Yang, Y.; Wang, Z. Nanostructured Porous CrN Thin Films by Oblique Angle Magnetron Sputtering for Symmetric Supercapacitors. J. Alloys Compd. 2019, 806, 953–959. [Google Scholar] [CrossRef]
  413. Wei, B.; Mei, G.; Liang, H.; Qi, Z.; Zhang, D.; Shen, H.; Wang, Z. Porous CrN Thin Films by Selectively Etching CrCuN for Symmetric Supercapacitors. J. Power Sources 2018, 385, 39–44. [Google Scholar] [CrossRef]
  414. Durai, G.; Kuppusami, P.; Maiyalagan, T.; Ahila, M.; Vinoth kumar, P. Supercapacitive Properties of Manganese Nitride Thin Film Electrodes Prepared by Reactive Magnetron Sputtering: Effect of Different Electrolytes. Ceram. Int. 2019, 45, 17120–17127. [Google Scholar] [CrossRef]
  415. Xie, Y.; Tian, F. Capacitive Performance of Molybdenum Nitride/Titanium Nitride Nanotube Array for Supercapacitor. Mater. Sci. Eng. B 2017, 215, 64–70. [Google Scholar] [CrossRef]
  416. Zhang, H.; Hu, W.; Wei, B.; Zheng, J.; Qi, Z.; Wang, Z. Freestanding Co3N Thin Film for High Performance Supercapacitors. Ceram. Int. 2021, 47, 3267–3271. [Google Scholar] [CrossRef]
  417. Achour, A.; Ducros, J.B.; Porto, R.L.; Boujtita, M.; Gautron, E.; Le Brizoual, L.; Djouadi, M.A.; Brousse, T. Hierarchical Nanocomposite Electrodes Based on Titanium Nitride and Carbon Nanotubes for Micro-Supercapacitors. Nano Energy 2014, 7, 104–113. [Google Scholar] [CrossRef]
  418. Salman, A.; Padmajan Sasikala, S.; Kim, I.H.; Kim, J.T.; Lee, G.S.; Kim, J.G.; Kim, S.O. Tungsten Nitride-Coated Graphene Fibers for High-Performance Wearable Supercapacitors. Nanoscale 2020, 12, 20239–20249. [Google Scholar] [CrossRef]
  419. Zhang, W.-B.; Ma, X.-J.; Kong, L.-B.; Luo, Y.-C.; Kang, L. Capacitive Intermetallic Manganese Nitride with High Volumetric Energy Densities. J. Electrochem. Soc. 2016, 163, A2830. [Google Scholar] [CrossRef]
  420. Sun, N.; Zhou, D.; Liu, W.; Shi, S.; Tian, Z.; Liu, F.; Li, S.; Wang, J.; Ali, F. Tailoring Surface Chemistry and Morphology of Titanium Nitride Electrode for On-Chip Supercapacitors. ACS Sustain. Chem. Eng. 2020, 8, 7869–7878. [Google Scholar] [CrossRef]
  421. Lv, F.; Ma, H.; Shen, L.; Jiang, Y.; Sun, T.; Ma, J.; Geng, X.; Kiran, A.; Zhu, N. Wearable Helical Molybdenum Nitride Supercapacitors for Self-Powered Healthcare Smartsensors. ACS Appl. Mater. Interfaces 2021, 13, 29780–29787. [Google Scholar] [CrossRef]
  422. Shen, H.; Wei, B.; Zhang, D.; Qi, Z.; Wang, Z. Magnetron Sputtered NbN Thin Film Electrodes for Supercapacitors. Mater. Lett. 2018, 229, 17–20. [Google Scholar] [CrossRef]
  423. Gao, Z.; Wu, Z.; Zhao, S.; Zhang, T.; Wang, Q. Enhanced Capacitive Property of HfN Film Electrode by Plasma Etching for Supercapacitors. Mater. Lett. 2019, 235, 148–152. [Google Scholar] [CrossRef]
  424. Anusha Thampi, V.V.; Nithiyanantham, U.; Nanda Kumar, A.K.; Martin, P.; Bendavid, A.; Subramanian, B. Fabrication of Sputtered Titanium Vanadium Nitride (TiVN) Thin Films for Micro-Supercapacitors. J. Mater. Sci. Mater. Electron. 2018, 29, 12457–12465. [Google Scholar] [CrossRef]
  425. Yi, C.-Q.; Zou, J.-P.; Yang, H.-Z.; Leng, X. Recent Advances in Pseudocapacitor Electrode Materials: Transition Metal Oxides and Nitrides. Trans. Nonferrous Met. Soc. China (Engl. Ed.) 2018, 28, 1980–2001. [Google Scholar] [CrossRef]
  426. Zhao, J.; Li, C.; Zhang, Q.; Zhang, J.; Wang, X.; Lin, Z.; Wang, J.; Lv, W.; Lu, C.; Wong, C.P.; et al. An All-Solid-State, Lightweight, and Flexible Asymmetric Supercapacitor Based on Cabbage-like ZnCo2O4 and Porous VN Nanowires Electrode Materials. J. Mater. Chem. A Mater. 2017, 5, 6928–6936. [Google Scholar] [CrossRef]
  427. Wei, B.; Shang, C.; Shui, L.; Wang, X.; Zhou, G. TiVN Composite Hollow Mesospheres for High-Performance Supercapacitors. Mater. Res. Express 2019, 6, 025801. [Google Scholar] [CrossRef]
  428. Su, Y.; Zhitomirsky, I. Hybrid MnO2/Carbon Nanotube-VN/Carbon Nanotube Supercapacitors. J. Power Sources 2014, 267, 235–242. [Google Scholar] [CrossRef]
  429. Wei, B.; Ming, F.; Liang, H.; Qi, Z.; Hu, W.; Wang, Z. All Nitride Asymmetric Supercapacitors of Niobium Titanium Nitride-Vanadium Nitride. J. Power Sources 2021, 481, 228842. [Google Scholar] [CrossRef]
  430. Ma, J.; Zheng, S.; Cao, Y.; Zhu, Y.; Das, P.; Wang, H.; Liu, Y.; Wang, J.; Chi, L.; Liu, S.; et al. Aqueous MXene/PH1000 Hybrid Inks for Inkjet-Printing Micro-Supercapacitors with Unprecedented Volumetric Capacitance and Modular Self-Powered Microelectronics. Adv. Energy Mater. 2021, 11, 2100746. [Google Scholar] [CrossRef]
  431. Zheng, S.; Wang, H.; Das, P.; Zhang, Y.; Cao, Y.; Ma, J.; Liu, S.; Wu, Z.S. Multitasking MXene Inks Enable High-Performance Printable Microelectrochemical Energy Storage Devices for All-Flexible Self-Powered Integrated Systems. Adv. Mater. 2021, 33, 2005449. [Google Scholar] [CrossRef] [PubMed]
  432. Orangi, J.; Hamade, F.; Davis, V.A.; Beidaghi, M. 3D Printing of Additive-Free 2D Ti3C2Tx (MXene) Ink for Fabrication of Micro-Supercapacitors with Ultra-High Energy Densities. ACS Nano 2020, 14, 640–650. [Google Scholar] [CrossRef] [PubMed]
  433. Li, H.; Li, X.; Liang, J.; Chen, Y. Hydrous RuO2-Decorated MXene Coordinating with Silver Nanowire Inks Enabling Fully Printed Micro-Supercapacitors with Extraordinary Volumetric Performance. Adv. Energy Mater. 2019, 9, 1803987. [Google Scholar] [CrossRef]
  434. Jiang, Q.; Lei, Y.; Liang, H.; Xi, K.; Xia, C.; Alshareef, H.N. Review of MXene Electrochemical Microsupercapacitors. Energy Storage Mater. 2020, 27, 78–95. [Google Scholar] [CrossRef]
  435. Salles, P.; Quain, E.; Kurra, N.; Sarycheva, A.; Gogotsi, Y. Automated Scalpel Patterning of Solution Processed Thin Films for Fabrication of Transparent MXene Microsupercapacitors. Small 2018, 14, e1802864. [Google Scholar] [CrossRef]
  436. Hu, H.; Bai, Z.; Niu, B.; Wu, M.; Hua, T. Binder-Free Bonding of Modularized MXene Thin Films into Thick Film Electrodes for on-Chip Micro-Supercapacitors with Enhanced Areal Performance Metrics. J. Mater. Chem. A Mater. 2018, 6, 14876–14884. [Google Scholar] [CrossRef]
  437. Biswas, A.; Natu, V.; Puthirath, A.B. Thin-Film Growth of MAX Phases as Functional Materials. Oxf. Open Mater. Sci. 2021, 1, itab020. [Google Scholar] [CrossRef]
  438. Ling, Z.; Ren, C.E.; Zhao, M.Q.; Yang, J.; Giammarco, J.M.; Qiu, J.; Barsoum, M.W.; Gogotsi, Y. Flexible and Conductive MXene Films and Nanocomposites with High Capacitance. Proc. Natl. Acad. Sci. USA 2014, 111, 16676–16681. [Google Scholar] [CrossRef]
  439. Lukatskaya, M.R.; Kota, S.; Lin, Z.; Zhao, M.Q.; Shpigel, N.; Levi, M.D.; Halim, J.; Taberna, P.L.; Barsoum, M.W.; Simon, P.; et al. Ultra-High-Rate Pseudocapacitive Energy Storage in Two-Dimensional Transition Metal Carbides. Nat. Energy 2017, 6, 17105. [Google Scholar] [CrossRef]
  440. Lukatskaya, M.R.; Bak, S.M.; Yu, X.; Yang, X.Q.; Barsoum, M.W.; Gogotsi, Y. Probing the Mechanism of High Capacitance in 2D Titanium Carbide Using in Situ X-Ray Absorption Spectroscopy. Adv. Energy Mater. 2015, 5, 1500589. [Google Scholar] [CrossRef]
  441. Peng, Y.Y.; Akuzum, B.; Kurra, N.; Zhao, M.Q.; Alhabeb, M.; Anasori, B.; Kumbur, E.C.; Alshareef, H.N.; Ger, M.-D.; Gogotsi, Y. All-MXene (2D Titanium Carbide) Solid-State Microsupercapacitors for on-Chip Energy Storage. Energy Environ. Sci. 2016, 9, 2847–2854. [Google Scholar] [CrossRef]
  442. Jiao, S.; Zhou, A.; Wu, M.; Hu, H. Kirigami Patterning of MXene/Bacterial Cellulose Composite Paper for All-Solid-State Stretchable Micro-Supercapacitor Arrays. Adv. Sci. 2019, 6, 1900529. [Google Scholar] [CrossRef] [PubMed]
  443. Yuan, M.; Wang, L.; Liu, X.; Du, X.; Zhang, G.; Chang, Y.; Xia, Q.; Hu, Q.; Zhou, A. 3D Printing Quasi-Solid-State Micro-Supercapacitors with Ultrahigh Areal Energy Density Based on High Concentration MXene Sediment. Chem. Eng. J. 2023, 451, 138686. [Google Scholar] [CrossRef]
  444. Zhang, C.J.; Kremer, M.P.; Seral-Ascaso, A.; Park, S.H.; McEvoy, N.; Anasori, B.; Gogotsi, Y.; Nicolosi, V. Stamping of Flexible, Coplanar Micro-Supercapacitors Using MXene Inks. Adv. Funct. Mater. 2018, 28, 1705506. [Google Scholar] [CrossRef]
  445. Li, H.; Hou, Y.; Wang, F.; Lohe, M.R.; Zhuang, X.; Niu, L.; Feng, X.; Li, H.; Hou, Y.; Wang, F.; et al. Flexible All-Solid-State Supercapacitors with High Volumetric Capacitances Boosted by Solution Processable MXene and Electrochemically Exfoliated Graphene. Adv. Energy Mater. 2017, 7, 1601847. [Google Scholar] [CrossRef]
  446. Hu, H.; Hua, T. An Easily Manipulated Protocol for Patterning of MXenes on Paper for Planar Micro-Supercapacitors. J. Mater. Chem. A Mater. 2017, 5, 19639–19648. [Google Scholar] [CrossRef]
  447. Huang, H.; Su, H.; Zhang, H.; Xu, L.; Chu, X.; Hu, C.; Liu, H.; Chen, N.; Liu, F.; Deng, W.; et al. Extraordinary Areal and Volumetric Performance of Flexible Solid-State Micro-Supercapacitors Based on Highly Conductive Freestanding Ti3C2Tx Films. Adv. Electron. Mater. 2018, 4, 1800179. [Google Scholar] [CrossRef]
  448. Li, P.; Shi, W.; Liu, W.; Chen, Y.; Xu, X.; Ye, S.; Yin, R.; Zhang, L.; Xu, L.; Cao, X. Fabrication of High-Performance MXene-Based All-Solid-State Flexible Microsupercapacitor Based on a Facile Scratch Method. Nanotechnology 2018, 29, 445401. [Google Scholar] [CrossRef]
  449. Wang, G.; Zhang, R.; Zhang, H.; Cheng, K. Aqueous MXene Inks for Inkjet-Printing Microsupercapacitors with Ultrahigh Energy Densities. J. Colloid. Interface Sci. 2023, 645, 359–370. [Google Scholar] [CrossRef]
  450. Jiang, Q.; Kurra, N.; Maleski, K.; Lei, Y.; Liang, H.; Zhang, Y.; Gogotsi, Y.; Alshareef, H.N. On-Chip MXene Microsupercapacitors for AC-Line Filtering Applications. Adv. Energy Mater. 2019, 9, 1901061. [Google Scholar] [CrossRef]
  451. Zhu, Y.; Zhang, Q.; Ma, J.; Das, P.; Zhang, L.; Liu, H.; Wang, S.; Li, H.; Wu, Z.S. Three-Dimensional (3D)-Printed MXene High-Voltage Aqueous Micro-Supercapacitors with Ultrahigh Areal Energy Density and Low-Temperature Tolerance. Carbon Energy 2024, 6, e481. [Google Scholar] [CrossRef]
  452. Fan, Z.; Wang, Y.; Xie, Z.; Wang, D.; Yuan, Y.; Kang, H.; Su, B.; Cheng, Z.; Liu, Y. Modified MXene/Holey Graphene Films for Advanced Supercapacitor Electrodes with Superior Energy Storage. Adv. Sci. 2018, 5, 1800750. [Google Scholar] [CrossRef]
  453. Kim, E.; Lee, B.J.; Maleski, K.; Chae, Y.; Lee, Y.; Gogotsi, Y.; Ahn, C.W. Microsupercapacitor with a 500 Nm Gap between MXene/CNT Electrodes. Nano Energy 2021, 81, 105616. [Google Scholar] [CrossRef]
  454. Liu, W.; Wang, Z.; Su, Y.; Li, Q.; Zhao, Z.; Geng, F. Molecularly Stacking Manganese Dioxide/Titanium Carbide Sheets to Produce Highly Flexible and Conductive Film Electrodes with Improved Pseudocapacitive Performances. Adv. Energy Mater. 2017, 7, 1602834. [Google Scholar] [CrossRef]
  455. Tian, Y.; Yang, C.; Que, W.; Liu, X.; Yin, X.; Kong, L.B. Flexible and Free-Standing 2D Titanium Carbide Film Decorated with Manganese Oxide Nanoparticles as a High Volumetric Capacity Electrode for Supercapacitor. J. Power Sources 2017, 359, 332–339. [Google Scholar] [CrossRef]
  456. Zhao, M.-Q.; Ren, C.E.; Ling, Z.; Lukatskaya, M.R.; Zhang, C.; Van Aken, K.L.; Barsoum, M.W.; Gogotsi, Y.; Zhao, M.; Ren, C.E.; et al. Flexible MXene/Carbon Nanotube Composite Paper with High Volumetric Capacitance. Adv. Mater. 2015, 27, 339–345. [Google Scholar] [CrossRef]
  457. Yan, J.; Ren, C.E.; Maleski, K.; Hatter, C.B.; Anasori, B.; Urbankowski, P.; Sarycheva, A.; Gogotsi, Y. Flexible MXene/Graphene Films for Ultrafast Supercapacitors with Outstanding Volumetric Capacitance. Adv. Funct. Mater. 2017, 27, 1701264. [Google Scholar] [CrossRef]
  458. Qin, L.; Tao, Q.; Liu, X.; Fahlman, M.; Halim, J.; Persson, P.O.Å.; Rosen, J.; Zhang, F. Polymer-MXene Composite Films Formed by MXene-Facilitated Electrochemical Polymerization for Flexible Solid-State Microsupercapacitors. Nano Energy 2019, 60, 734–742. [Google Scholar] [CrossRef]
  459. Zhang, C.J.; McKeon, L.; Kremer, M.P.; Park, S.H.; Ronan, O.; Seral-Ascaso, A.; Barwich, S.; Coileáin, C.; McEvoy, N.; Nerl, H.C.; et al. Additive-Free MXene Inks and Direct Printing of Micro-Supercapacitors. Nat. Commun. 2019, 10, 1795. [Google Scholar] [CrossRef]
  460. Wang, Y.; Zhang, Y.Z.; Dubbink, D.; ten Elshof, J.E. Inkjet Printing of δ-MnO2 Nanosheets for Flexible Solid-State Micro-Supercapacitor. Nano Energy 2018, 49, 481–488. [Google Scholar] [CrossRef]
  461. Li, L.; Secor, E.B.; Chen, K.S.; Zhu, J.; Liu, X.; Gao, T.Z.; Seo, J.W.T.; Zhao, Y.; Hersam, M.C. High-Performance Solid-State Supercapacitors and Microsupercapacitors Derived from Printable Graphene Inks. Adv. Energy Mater. 2016, 6, 1600909. [Google Scholar] [CrossRef]
  462. Van Toan, N.; Kim Tuoi, T.T.; Li, J.; Inomata, N.; Ono, T. Liquid and Solid States On-Chip Micro-Supercapacitors Using Silicon Nanowire-Graphene Nanowall-Pani Electrode Based on Microfabrication Technology. Mater. Res. Bull. 2020, 131, 110977. [Google Scholar] [CrossRef]
  463. Wang, K.; Zou, W.; Quan, B.; Yu, A.; Wu, H.; Jiang, P.; Wei, Z. An All-Solid-State Flexible Micro-Supercapacitor on a Chip. Adv. Energy Mater. 2011, 1, 1068–1072. [Google Scholar] [CrossRef]
  464. Xu, K. Nonaqueous Liquid Electrolytes for Lithium-Based Rechargeable Batteries. Chem. Rev. 2004, 104, 4303–4418. [Google Scholar] [CrossRef]
  465. Pal, B.; Yang, S.; Ramesh, S.; Thangadurai, V.; Jose, R. Electrolyte Selection for Supercapacitive Devices: A Critical Review. Nanoscale Adv. 2019, 1, 3807–3835. [Google Scholar] [CrossRef]
  466. Mendhe, A.; Panda, H.S. Discover Materials A Review on Electrolytes for Supercapacitor Device. Discov Mater 2023, 3, 29. [Google Scholar] [CrossRef]
  467. Yu, L.; Chen, G.Z. Ionic Liquid-Based Electrolytes for Supercapacitor and Supercapattery. Front. Chem. 2019, 7, 272. [Google Scholar] [CrossRef]
  468. Bhat, M.Y.; Hashmi, S.A.; Khan, M.; Choi, D.; Qurashi, A. Frontiers and Recent Developments on Supercapacitor’s Materials, Design, and Applications: Transport and Power System Applications. J. Energy Storage 2023, 58, 106104. [Google Scholar] [CrossRef]
  469. Wang, R.; Li, Q.; Cheng, L.; Li, H.; Wang, B.; Zhao, X.S.; Guo, P. Electrochemical Properties of Manganese Ferrite-Based Supercapacitors in Aqueous Electrolyte: The Effect of Ionic Radius. Colloids Surf. A Physicochem. Eng. Asp. 2014, 457, 94–99. [Google Scholar] [CrossRef]
  470. Zhu, J.; Xu, Y.; Wang, J.; Lin, J.; Sun, X.; Mao, S. The Effect of Various Electrolyte Cations on Electrochemical Performance of Polypyrrole/RGO Based Supercapacitors. Phys. Chem. Chem. Phys. 2015, 17, 28666–28673. [Google Scholar] [CrossRef] [PubMed]
  471. Wu, H.; Wang, X.; Jiang, L.; Wu, C.; Zhao, Q.; Liu, X.; Hu, B.; Yi, L. The Effects of Electrolyte on the Supercapacitive Performance of Activated Calcium Carbide-Derived Carbon. J. Power Sources 2013, 226, 202–209. [Google Scholar] [CrossRef]
  472. Lim, Y.; Yoon, J.; Yun, J.; Kim, D.; Hong, S.Y.; Lee, S.J.; Zi, G.; Ha, J.S. Erratum: Biaxially Stretchable, Integrated Array of High Performance Microsupercapacitors. ACS Nano 2014, 8, 11639–11650, Erratum in ACS Nano 2015, 9, 6634. [Google Scholar] [CrossRef]
  473. Armand, M.; Endres, F.; MacFarlane, D.R.; Ohno, H.; Scrosati, B. Ionic-Liquid Materials for the Electrochemical Challenges of the Future. Nat. Mater. 2009, 8, 621–629. [Google Scholar] [CrossRef]
  474. Liew, C.W.; Arifin, K.H.; Kawamura, J.; Iwai, Y.; Ramesh, S.; Arof, A.K. Effect of Halide Anions in Ionic Liquid Added Poly(Vinyl Alcohol)-Based Ion Conductors for Electrical Double Layer Capacitors. J. Non Cryst. Solids 2017, 458, 97–106. [Google Scholar] [CrossRef]
  475. Rennie, A.J.R.; Sanchez-Ramirez, N.; Torresi, R.M.; Hall, P.J. Ether-Bond-Containing Ionic Liquids as Supercapacitor Electrolytes. J. Phys. Chem. Lett. 2013, 4, 2970–2974. [Google Scholar] [CrossRef] [PubMed]
  476. Pohlmann, S.; Olyschläger, T.; Goodrich, P.; Alvarez Vicente, J.; Jacquemin, J.; Balducci, A. Azepanium-Based Ionic Liquids as Green Electrolytes for High Voltage Supercapacitors. J. Power Sources 2015, 273, 931–936. [Google Scholar] [CrossRef]
  477. Timperman, L.; Skowron, P.; Boisset, A.; Galiano, H.; Lemordant, D.; Frackowiak, E.; Béguin, F.; Anouti, M. Triethylammonium Bis(Tetrafluoromethylsulfonyl)Amide Protic Ionic Liquid as an Electrolyte for Electrical Double-Layer Capacitors. Phys. Chem. Chem. Phys. 2012, 14, 8199–8207. [Google Scholar] [CrossRef]
  478. Kurig, H.; Vestli, M.; Tõnurist, K.; Jänes, A.; Lust, E. Influence of Room Temperature Ionic Liquid Anion Chemical Composition and Electrical Charge Delocalization on the Supercapacitor Properties. J. Electrochem. Soc. 2012, 159, A944. [Google Scholar] [CrossRef]
  479. Łatoszyńska, A.A.; Zukowska, G.Z.; Rutkowska, I.A.; Taberna, P.L.; Simon, P.; Kulesza, P.J.; Wieczorek, W. Non-Aqueous Gel Polymer Electrolyte with Phosphoric Acid Ester and Its Application for Quasi Solid-State Supercapacitors. J. Power Sources 2015, 274, 1147–1154. [Google Scholar] [CrossRef]
  480. Francisco, B.E.; Jones, C.M.; Lee, S.H.; Stoldt, C.R. Nanostructured All-Solid-State Supercapacitor Based on Li2S-P2S5 Glass-Ceramic Electrolyte. Appl. Phys. Lett. 2012, 100, 103902. [Google Scholar] [CrossRef]
  481. Chong, M.Y.; Numan, A.; Liew, C.W.; Ng, H.M.; Ramesh, K.; Ramesh, S. Enhancing the Performance of Green Solid-State Electric Double-Layer Capacitor Incorporated with Fumed Silica Nanoparticles. J. Phys. Chem. Solids 2018, 117, 194–203. [Google Scholar] [CrossRef]
  482. Murugan, R.; Thangadurai, V.; Weppner, W. Fast Lithium Ion Conduction in Garnet-Type Li7La3Zr2O12. Angew. Chem. -Int. Ed. 2007, 46, 7778–7781. [Google Scholar] [CrossRef] [PubMed]
  483. Verma, M.L.; Minakshi, M.; Singh, N.K. Synthesis and Characterization of Solid Polymer Electrolyte Based on Activated Carbon for Solid State Capacitor. Electrochim. Acta 2014, 137, 497–503. [Google Scholar] [CrossRef]
  484. Pal, B.; Yasin, A.; Kunwar, R.; Yang, S.; Yusoff, M.M.; Jose, R. Polymer versus Cation of Gel Polymer Electrolytes in the Charge Storage of Asymmetric Supercapacitors. Ind. Eng. Chem. Res. 2019, 58, 654–664. [Google Scholar] [CrossRef]
Figure 1. Performance of energy storage microdevices. (a) Ragone plot relating energy and power density for the micro-energy storage devices. (b) Energy versus time constant of the various energy storage microdevices.
Figure 1. Performance of energy storage microdevices. (a) Ragone plot relating energy and power density for the micro-energy storage devices. (b) Energy versus time constant of the various energy storage microdevices.
Batteries 10 00317 g001
Figure 2. Microfabrication of functional materials for on-chip energy storage microdevices. (a) Thin film deposition techniques for electrode materials. (b) Microfabrication techniques for on-chip electronics energy storage systems. Laser scribing, photolithography, screen printing, inkjet printing, injection, and transferring printing. Reproduced with permission from [11]. © IOP Publishing Ltd. CC BY 3.0. 3D printing. Reproduced with permission from [60]. Copyright (2016), American Chemical Society. Focused ion beam (FIB) technology. Reproduced with permission from [61]. © MDPI. CC BY 4.0. 3D holographic lithography. Reproduced with permission from [12]. Copyright (2022), Elsevier. Mask-assisted filtering and plasma etching. Reproduced with permission from [17]. Copyright (2020), John Wiley & Sons.
Figure 2. Microfabrication of functional materials for on-chip energy storage microdevices. (a) Thin film deposition techniques for electrode materials. (b) Microfabrication techniques for on-chip electronics energy storage systems. Laser scribing, photolithography, screen printing, inkjet printing, injection, and transferring printing. Reproduced with permission from [11]. © IOP Publishing Ltd. CC BY 3.0. 3D printing. Reproduced with permission from [60]. Copyright (2016), American Chemical Society. Focused ion beam (FIB) technology. Reproduced with permission from [61]. © MDPI. CC BY 4.0. 3D holographic lithography. Reproduced with permission from [12]. Copyright (2022), Elsevier. Mask-assisted filtering and plasma etching. Reproduced with permission from [17]. Copyright (2020), John Wiley & Sons.
Batteries 10 00317 g002aBatteries 10 00317 g002b
Figure 3. Schematic diagram of a typical electrostatic (micro-) capacitor.
Figure 3. Schematic diagram of a typical electrostatic (micro-) capacitor.
Batteries 10 00317 g003
Figure 4. Band gap versus dielectric constant of dielectric materials.
Figure 4. Band gap versus dielectric constant of dielectric materials.
Batteries 10 00317 g004
Figure 5. MIM nanocapacitor. (a,b) SEM plan-view and cross-section of an AAO MIM structure. (c) The bottom of the tube, showing the AAO barrier layer and three layers corresponding to the TiN bottom electrode (BE), Al2O3 and the TiN top electrode (TE). (d) Two-inch wafer with nanocapacitors. Reproduced with permission from [67]. Copyright (2009), Springer Nature.
Figure 5. MIM nanocapacitor. (a,b) SEM plan-view and cross-section of an AAO MIM structure. (c) The bottom of the tube, showing the AAO barrier layer and three layers corresponding to the TiN bottom electrode (BE), Al2O3 and the TiN top electrode (TE). (d) Two-inch wafer with nanocapacitors. Reproduced with permission from [67]. Copyright (2009), Springer Nature.
Batteries 10 00317 g005
Figure 8. Metal oxide electrode materials for on-chip micro-supercapacitor applications. (a) Sputtered anhydrous RuO2 thin-film electrodes. CV measurement, and stability of the interdigitated MSC of 10,000 cycles. Reproduced with permission from [247]. Copyright (2017), John Wiley & Sons-Books. (b) All-solid-state interdigitated microdevice based on porous RuOxNySz. CV plots in doped PVA and doped [EMIM] [TFSI] at different scan rates, and cell capacitance retention with the number of charge/discharge curves at 10 mA cm−2 (inset: 1st and 5000th potential–time curve during GCD). Reproduced with permission from [257]. Copyright © 2024, American Chemical Society. (c) 3D interdigitated MSCs based on electrodeposited MnO2. Top view SEM images of the 3D MSC, CV of a 3D MSC in parallel plate configuration measured in 5 M LiNO3 vs. the sweep rate, evolution of the surface capacitance, capacitance retention, and real energy density of the MSC, respectively. Reproduced with permission from [250]. Copyright (2021), Elsevier. (d) Interdigitated MSCs operating in physiological electrolytes based on sputtered iridium oxide films (SIROFs). SEM images of SIROF in cross-sectional view, CV measurements, and cycle stability of SIORF MSC over 100,000 CV cycles at 200 mV s−1. Reproduced with permission from [245]. Copyright (2022), IOP Publishing.
Figure 8. Metal oxide electrode materials for on-chip micro-supercapacitor applications. (a) Sputtered anhydrous RuO2 thin-film electrodes. CV measurement, and stability of the interdigitated MSC of 10,000 cycles. Reproduced with permission from [247]. Copyright (2017), John Wiley & Sons-Books. (b) All-solid-state interdigitated microdevice based on porous RuOxNySz. CV plots in doped PVA and doped [EMIM] [TFSI] at different scan rates, and cell capacitance retention with the number of charge/discharge curves at 10 mA cm−2 (inset: 1st and 5000th potential–time curve during GCD). Reproduced with permission from [257]. Copyright © 2024, American Chemical Society. (c) 3D interdigitated MSCs based on electrodeposited MnO2. Top view SEM images of the 3D MSC, CV of a 3D MSC in parallel plate configuration measured in 5 M LiNO3 vs. the sweep rate, evolution of the surface capacitance, capacitance retention, and real energy density of the MSC, respectively. Reproduced with permission from [250]. Copyright (2021), Elsevier. (d) Interdigitated MSCs operating in physiological electrolytes based on sputtered iridium oxide films (SIROFs). SEM images of SIROF in cross-sectional view, CV measurements, and cycle stability of SIORF MSC over 100,000 CV cycles at 200 mV s−1. Reproduced with permission from [245]. Copyright (2022), IOP Publishing.
Batteries 10 00317 g008
Figure 9. Multicationic oxide electrode materials for on-chip MSC applications. (a) Sputtered mixed Fe-W oxide thin film. SEM cross-section images, CV plots of the sputtered FeWO4 films at different sweep rates in neutral aqueous electrolyte (5 M LiNO3), and stability of the electrode over 10,000 cycles, respectively. Reproduced with permission from [129]. Copyright (2021), IOP Publishing. (b) Sputtered mixed Mn-Fe oxide thin film. SEM cross-section images of the sputtered films at different post-deposition heat treatments, CV plots of the sputtered (Mn,Fe)3O4 films at 10 mV s−1 in neutral aqueous electrolyte (1 M NaSO4), stability of the electrode over 10,000 cycles, respectively. Reproduced with permission from [239]. Copyright (2022), IOP Publishing.
Figure 9. Multicationic oxide electrode materials for on-chip MSC applications. (a) Sputtered mixed Fe-W oxide thin film. SEM cross-section images, CV plots of the sputtered FeWO4 films at different sweep rates in neutral aqueous electrolyte (5 M LiNO3), and stability of the electrode over 10,000 cycles, respectively. Reproduced with permission from [129]. Copyright (2021), IOP Publishing. (b) Sputtered mixed Mn-Fe oxide thin film. SEM cross-section images of the sputtered films at different post-deposition heat treatments, CV plots of the sputtered (Mn,Fe)3O4 films at 10 mV s−1 in neutral aqueous electrolyte (1 M NaSO4), stability of the electrode over 10,000 cycles, respectively. Reproduced with permission from [239]. Copyright (2022), IOP Publishing.
Batteries 10 00317 g009
Figure 11. MXene electrode for on-chip micro-supercapacitor applications. (a) Wafer scale fabrication of the MXene. Digital photograph showing the wafer scale fabrication of the MXene MSCs, cross-sectional SEM images showing uniform coating of Ti3C2Tx on gold, CV plots of Ti3C2Tx – -0.3 µm MSCs using PVA/H3PO4 gel electrolyte, and C′ and C″ versus frequency, respectively. Reproduced with permission from [450]. Copyright (2019), John Wiley & Sons-Books. (b) MXene-rGO asymmetric MSC. Digital photograph of a top-view of the asymmetric MSC device, indicating Ti3C2Tx and rGO interdigitated electrode configurations; CVs of the MSC and the individual electrodes recorded at 2 mV s−1 in a 3-electrode configuration; comparison between asymmetric interdigitated MSC and all-MXene symmetric interdigitated MSC at 2 mV s−1; Ragone plot of the asymmetric interdigitated MSC, respectively. Reproduced with permission from [136]. Copyright (2017), John Wiley & Sons. (c) MXene-RuO2 asymmetric MSC. Field emission scanning electron microscopy (FSEM) images of Ti3C2Tx on carbon fibers (CF) and RuO2/CF; CVs of RuO2/CF, Ti3C2Tx/CF, and the asymmetric device at a scan rate of 50 mV s−1, normalized real (C′) and imaginary (C″) parts of capacitance versus frequency of the device; and cycling stability and Coulombic efficiency of the asymmetric device over 20,000 cycles in 1 m H2SO4 electrolyte at a current density of 20 A g−1 (inset: the typical charge–discharge profiles for the first and last charge–discharge cycles). Reproduced with permission from [135]. Copyright (2018), John Wiley & Sons.
Figure 11. MXene electrode for on-chip micro-supercapacitor applications. (a) Wafer scale fabrication of the MXene. Digital photograph showing the wafer scale fabrication of the MXene MSCs, cross-sectional SEM images showing uniform coating of Ti3C2Tx on gold, CV plots of Ti3C2Tx – -0.3 µm MSCs using PVA/H3PO4 gel electrolyte, and C′ and C″ versus frequency, respectively. Reproduced with permission from [450]. Copyright (2019), John Wiley & Sons-Books. (b) MXene-rGO asymmetric MSC. Digital photograph of a top-view of the asymmetric MSC device, indicating Ti3C2Tx and rGO interdigitated electrode configurations; CVs of the MSC and the individual electrodes recorded at 2 mV s−1 in a 3-electrode configuration; comparison between asymmetric interdigitated MSC and all-MXene symmetric interdigitated MSC at 2 mV s−1; Ragone plot of the asymmetric interdigitated MSC, respectively. Reproduced with permission from [136]. Copyright (2017), John Wiley & Sons. (c) MXene-RuO2 asymmetric MSC. Field emission scanning electron microscopy (FSEM) images of Ti3C2Tx on carbon fibers (CF) and RuO2/CF; CVs of RuO2/CF, Ti3C2Tx/CF, and the asymmetric device at a scan rate of 50 mV s−1, normalized real (C′) and imaginary (C″) parts of capacitance versus frequency of the device; and cycling stability and Coulombic efficiency of the asymmetric device over 20,000 cycles in 1 m H2SO4 electrolyte at a current density of 20 A g−1 (inset: the typical charge–discharge profiles for the first and last charge–discharge cycles). Reproduced with permission from [135]. Copyright (2018), John Wiley & Sons.
Batteries 10 00317 g011
Figure 12. Electrolyte classification for MSC applications [465,466,467,468].
Figure 12. Electrolyte classification for MSC applications [465,466,467,468].
Batteries 10 00317 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jolayemi, B.; Buvat, G.; Roussel, P.; Lethien, C. Emerging Capacitive Materials for On-Chip Electronics Energy Storage Technologies. Batteries 2024, 10, 317. https://doi.org/10.3390/batteries10090317

AMA Style

Jolayemi B, Buvat G, Roussel P, Lethien C. Emerging Capacitive Materials for On-Chip Electronics Energy Storage Technologies. Batteries. 2024; 10(9):317. https://doi.org/10.3390/batteries10090317

Chicago/Turabian Style

Jolayemi, Bukola, Gaetan Buvat, Pascal Roussel, and Christophe Lethien. 2024. "Emerging Capacitive Materials for On-Chip Electronics Energy Storage Technologies" Batteries 10, no. 9: 317. https://doi.org/10.3390/batteries10090317

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