Previous Article in Journal
Correction: Huang et al. Electrosprayed Ultra-Thin Coating of Ethyl Cellulose on Drug Nanoparticles for Improved Sustained Release. Nanomaterials 2020, 10, 1758
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Porous Silicon and Silicon Nanowires for On-Chip Supercapacitor Electrodes: A Review

by
Daria M. Sedlovets
Institute of Microelectronics Technology and High-Purity Materials, Russian Academy of Science (IMT RAS), 6 Academician Ossipyan Str., 142432 Chernogolovka, Moscow, Russia
Nanomaterials 2025, 15(23), 1826; https://doi.org/10.3390/nano15231826
Submission received: 2 November 2025 / Revised: 28 November 2025 / Accepted: 29 November 2025 / Published: 2 December 2025
(This article belongs to the Section Nanocomposite Materials)

Abstract

Finding efficient ways to store energy is a current topic both at the macro level and at the microscale. As silicon plates are the main platform for the integration of microelectronic devices, it is reasonable to use the silicon structures as the active materials for on-chip microcapacitors. Porous silicon (pSi) and silicon nanowires (SiNWs) are ideal candidates for planar electrodes because these layers are directly embedded into the silicon wafer. The review contains a brief summary of the formation features of pSi/SiNW and their electrochemical performance. The structural characteristics of the silicon matrix (depth and morphology) that influence capacitive electrode properties are examined comprehensively for the first time. Particular attention is given to additional coatings on the pore/wire surface. Various ways of depositing metal, carbon, and polymer layers are considered in detail. Different approaches to filling the silicon matrix are explored. Focusing on pSi/SiNWs coatings allows us to identify the effect of the structure, crystallinity, and methods of additional layer deposition on capacitance, cycling stability, and charge transport of modified electrodes. Although fabrication processes for planar microcapacitors based on pSi/SiNWs are currently underdeveloped, the specific requirements and possible challenges of on-chip integration are discussed and proposed.

1. Introduction

The investigation of alternative energy sources and the search for efficient ways to store energy are critical challenges. In this regard, electrochemical supercapacitors are of increasing interest to researchers. Here, we focus on micro devices. As silicon wafers are the main platform for integrating microelectronic components, porous silicon (pSi) or silicon nanowires (SiNWs) are promising materials for on-chip microcapacitor electrodes. Recent reviews relate either to various energy devices based on pSi/SiNWs [1,2,3] or to supercapacitor applications using various silicon-based materials [4], so pSi/SiNWs capacitors are considered in less detail than we would like. While the mentioned works certainly cover the main achievements in this field, they do not provide a comprehensive overview. Here, we summarize the data on the electrochemical properties of pSi and SiNWs depending on the structure and morphology of the Si matrix, to identify factors that influence their performance. Particular attention is given to additional coatings on the pore/wire surface because they have a great impact on capacitive properties. Various ways for deposition of metal, carbon, and polymer layers are considered in detail. Key coatings for pSi/SiNW electrodes and their features in the context of microcapacitor performance are summarized. Cyclic stability is especially discussed.
We focus on on-chip applications because in this case, the creation of active material in the body of a silicon wafer is an inspiring idea. Although fabrication methods for planar microcapacitors based on pSi/SiNWs are currently underdeveloped, the specific requirements and possible challenges of on-chip integration are discussed and proposed.

2. pSi/SiNW Formation

pSi and SiNWs are structurally similar materials. The main difference between them is their production methods: anodic etching (AE) in hydrofluoric acid solutions/deep reactive ion etching (DRIE) for pSi and metal-assisted chemical etching (MACE)/vapor–liquid–solid (VLS) for SiNWs. We will not focus on these methods—they are discussed in sufficient detail in the relevant reviews: [5,6,7] and [8,9,10], respectively. Here, we briefly describe each approach.

2.1. AE

Generally, pSi is formed by anodizing a silicon wafer in an HF-containing electrolyte [5]. Theoretical studies associate the generation of etching sites with the instability of the physical solution–crystal interface under small disturbances in electrolysis processes [11]. The author of [12] argues that, during anodic treatment of the silicon crystal, narrow “channels” form under the influence of HF and move mainly into the depth of the sample. Various models describe the formation mechanism of pSi [13]:
The Beale model—electric field lines concentrate at surface irregularities, focus current flow at the pore tips, and locally enhance dissolution there.
Diffusion-limited model—during pore generation, a hole diffuses to the silicon surface and reacts with a Si surface atom. The pore tips are the most likely contact site for particle diffusion.
Quantum model—the increase in the pSi band gap significantly reduces the concentration of mobile charge carriers up to “depletion”. The current is then limited to the pore tips by increasing the electric field, and the porous structure is passivated by the quantum effect.
The dissolution of silicon during the etching process follows two main paths of conversion of the unstable silicon divalent ion [14,15]:
Si2+ + 2H+ → Si4+ + H2,
Si2+ + Si2+ → Si + Si4+.
In the oxidation reaction (1), silicon oxide or a water-soluble H2SiF6 is generally formed. Disproportionate reaction (2) yields secondary silicon atoms that are resistant to dissolution and unable to integrate into the crystal lattice. These atoms form small nanocrystallites, in which no free holes are available due to the increased band gap [16]. This insulating powdery material is called amorphous silicon or α-Si [17,18].
The morphological characteristics of pSi (the predominant direction of pore growth, their size distribution, and branching) are determined by numerous factors, including silicon wafer characteristics, etching current density, and electrolyte concentration [6].

2.2. DRIE

DRIE can also be used to obtain silicon nanorods (SiNRs). DRIE is performed on an inductively coupled plasma system. As described in [19], the plasma is generated through a coil at 13.56 MHz. The platen is powered independently with a second generator at the same frequency.

2.3. MACE

The MACE approach for SiNWs is realized by selectively etching silicon in the presence of a catalytic metal.
The two-step MACE method includes systematically depositing a thin layer of metal (usually silver, gold, or platinum) onto the surface of silicon, followed by immersing the sample in an etchant solution containing HF and an oxidizing agent such as H2O2 [1]. Silicon can be considered as a local anode and the metal as a cathode for the current produced at the Si/metal interface [20]. The anodic reaction proceeds as follows:
H2O2 + 2H+ → 2H2O +2h+.
As suggested by the authors [21], the cathode reaction occurs as the usual reduction of protons to hydrogen:
2H+ → H2↑ + 2h+ .
The reduction of the oxidant species generates holes that are injected into the silicon underneath the metal. Hence, in the Si anode region, the silicon is oxidized and dissolved. The dissolution models proceed as reactions I, II, and III, respectively.
Reaction I:
Si + 4h+ + 4HF → SiF4 + 4H+,
SiF4 + 2HF → H2SiF6.
Reaction II:
Si + 4HF2 → SiF62–+2HF+ H2↑ + 2e.
Reaction III:
Si + 2H2O → SiO2+4H+ + 4e,
SiO2 + 6HF → H2SiF6 + 2H2O.
In Reaction I, Si is dissolved directly in its tetravalent state without forming silicon dioxide. In Reaction II, the direct dissolution of Si is still present, but in the divalent state. In Reaction III, the Si atoms in contact with the metal are oxidized and then dissolved in two different processes [8].
The single-step MACE concept is based on the use of silver salt. After the cleaning procedure, the sample is immersed in an aqueous AgNO3:HF solution. The AgNO3 catalyst precipitates, forming Ag nanoparticles (NPs) that are randomly distributed on the silicon flat surface, catalyzing the HF-driven Si etching in a manner similar to that described for H2O2. Indeed, AgNO3 acts as both an oxidant and a metal source [20].
The morphology and characteristics of the resulting SiNWs are influenced by several parameters: the concentration and stirring intensity of the etchant solution, the thickness of the metal catalyst layers, the etching temperature and time, and the doping level of the silicon substrate. The review [1] summarizes the main patterns. The characteristics of SiNWs achieved under various conditions of the MACE process are as follows.
(1)
Higher aspect ratio: thicker metal, longer time, increased stirring, higher doping level, and larger etchant concentration;
(2)
Larger diameter: thicker metal, higher temperature, increased stirring, and lower doping level;
(3)
Higher length: longer time;
(4)
Higher etching rate: higher temperature.
The MACE method can be modified into the metal-assisted anodic etching (MAAE) approach. An Au film is deposited on a silicon wafer and patterned using interference lithography to generate an ordered array of holes in the metal. Then, anodic contact is made to the Au mesh, and a Pt wire is used as a counter electrode to complete the circuit. When a bias is applied between the two electrodes in an aqueous HF solution with a concentration of 4 mol/L (M), the etching at the Au–Si interface leads to the formation of SiNW arrays [22].

2.4. VLS

VLS has long been the most widely used approach for SiNW fabrication [20]. The growth of nanowires is catalyzed by metal droplets that are realized as a product of the melting of metal NPs contaminated with silicon atoms [9,23]. The most commonly used catalyst remains gold due to the high crystalline quality of the resulting SiNWs and the simple thermodynamics of the Si/Au alloy. The droplets of gold–silicon alloys are liquid at the eutectic phase obtained through specific growth conditions. When the silicon concentration in the alloy exceeds the Si concentration at thermodynamic equilibrium of the eutectic phase, the silicon precipitates under the gold, solidifies, and grows as nanowires [9,23,24].
The standard process for SiNW growth is as follows. A very thin gold film (4 nm [25,26,27,28] or 12 nm [29]) is deposited via evaporation [25,26,27,29] or from gold colloid solution [28,30]. It should be noted that after gold deposition from AuCl4 solution [31,32], the grown SiNWs yield a reduced capacitance of several hundred μF/cm2. Chemical vapor deposition (CVD) is carried out through exposure to silane gas at low pressure and relatively high temperature (mainly, 500 °C [25,26,28]), except for work [33] where evaporated Sn acts as the catalyst at 1700 °C. To dope the SiNWs, the synthesis temperature was raised to 600 °C, and additional gases were passed (PH3 [29,30] and B2H6 [27] for n- and p-doping, respectively, and H2 as the carrier gas in both cases).

2.5. Comparison

Figure 1a shows that the number of works devoted to SiNW-based supercapacitors exceeds those devoted to pSi-based supercapacitors. It also follows from Figure 1a that the anodic etched pSi electrodes were first proposed in 1991 and have been continuously improved over time. Two notable exceptions to this pattern occurred in 2017 and relate to DRIE-etched pSi [34,35]. SiNWs were first used for microcapacitors in 2012 and generally exhibited capacitive characteristics across a wide range of values.
As can be seen in Figure 1b, there is a change in trend between the MACE and VLS approaches to obtain SiNWs over the past few years. This could be attributed to several drawbacks of VLS, which can be overcome by the MACE process [20]. Indeed, the VLS process, carried out through CVD, has significant disadvantages compared to MACE, as it requires complicated conditions: low pressure, relatively high temperatures, and toxic gases. For these two approaches, an impact on the achieved capacitance is also observed. Figure 1c shows that VLS primarily yields capacitances between 0.013 [36] and 36.3 [26] mF/cm2, while MACE primarily enables capacitances of 100 mF/cm2 or higher. Therefore, the MACE approach is not only simpler but also more suitable for capacitor applications than the VLS approach (due to the morphology of the resulting SiNWs, as will be discussed in Section 5). It should be noted that the data presented in Figure 1c relate not only to different methods of obtaining SiNWs but also to different measurement conditions, the electrolyte used, additional coatings, and so on. However, the pattern is obvious; works [25,28] are the only exceptions. In these works, high capacitance values were obtained using the VLS approach and by depositing additional layers: amorphous NiB and crystalline CrN, respectively. The influence of the various coatings on pSi/SiNWs capacitances will be studied in more detail in Section 6.

3. Electrochemical Performance

There are two main mechanisms for charge accumulation in supercapacitors:
(1)
An electrical double layer (EDL) is formed due to electrostatic attraction between the charged electrode surface and the counter ions of the electrolyte. EDL-materials exhibit a close-to-rectangular shape of cyclic voltammograms (CVA) and a linear galvanostatic charge–discharge (GCD) curve;
(2)
Pseudocapacitance arises from surface redox reactions, which can cause deviation of the CVA form from a rectangular shape or even the appearance of peaks on the CVA.
While ideal supercapacitors (which operate via the EDL mechanism) can be clearly differentiated from rechargeable batteries (which rely on faradaic processes), pseudocapacitors fall somewhere in between and are not always easy to distinguish [37,38]. The pseudocapacitive behavior is caused by redox reactions, but only on the surface (surface of the pores or SiNWs), in contrast to battery-like materials, where solution-based faradaic processes occur (with redox reactions taking place in the volume). This difference can be determined analytically using the well-known Randles–Sevcik equation: the linear relationship between the peak currents and the scan rate (for the capacitive type) or the square root of the scan rate (for the battery type) [39,40,41].
The specific capacitance is calculated by integrating the area under the CVA curve to obtain the charge value and then dividing this by the surface area, the scan rate, and the potential window. Notably, when evaluating the capacitance, as well as energy and power densities from GCD data, it is necessary to use integral formulas (Equations (10)–(12)) if the GCD curves deviate from the linear pattern [42,43]:
C = 2 I c V d t A · ( V ) 2 ,
E = I c V d t A · 3600 ,
P = I c V d t A · t ,
where C is the area-normalized capacitance in F/cm2, Ic is the constant current density in A/cm2, ∫Vdt is the integrated area of the discharge curve, ∆V is the scanned potential window in V, A is the area exposed to the electrolyte in cm2, E is the energy density in Wh/cm2, P is the power density in W/cm2, and Δt is the discharge time in s.
Impedance measurements are also used for electrochemical characterization of pSi/SiNW structures. Figure 2 depicts the equivalent electrical circuit of the porous electrode (the scheme is described in the Supplementary Materials, Section S2) [44]. Electrochemical impedance spectroscopy (EIS) consists of plotting the so-called Nyquist plots representing the negative of the imaginary part versus the real part of the complex impedance of individual electrodes or the electrochemical cell (see Figure 3). Interpretations of Nyquist plots are usually based on the use of equivalent circuits, but study [45], based on the modified Poisson−Nernst−Planck model, established that the electrode resistance, electrolyte resistance, and the equilibrium differential capacitance can be identified directly from Nyquist plots without relying on equivalent circuits.
The intersections on the x-axis at high frequency values are typically called solution resistance (Rs). The contribution of Rs can come from the electrolyte part, the electrode part, and the electrode–electrolyte interface. The semicircle’s diameter shows electrode resistance or Rct. In the active compound, Rct takes place in the high-frequency region due to the resistance of charge transfer. Capacitive region or battery-type behavior may be demonstrated by the slope at low frequency [46]. A higher curve slope is usually related to the presence of diffuse hindrances. The Rs and Rct values depend on too many factors and do not always correlate with the capacity value. Therefore, they cannot be used to evaluate charge transfer during the comparisons of samples measured under different conditions (reported in different works). However, comparing different materials (e.g., pSi/SiNWs with various coatings) within one study provides useful information. The frequencies are mainly studied between 1 mHz [47] and 7 MHz [48].

4. SiNW Length and PSi Depth

To determine the influence of SiNW length on their electrochemical performance, we cannot combine the data obtained by different authors because too many factors, in addition to depth, affect the capacitance. Therefore, we consider the capacitance dependence on pSi depth/SiNW length reported in each work. Figure 4 shows some results on area-normalized values of the SiNWs versus the volumetric values, which were calculated by dividing the same magnitude by the SiNW length.
As can be seen in Figure 4a,b, the volumetric capacitance increases together with the area-normalized capacitance. This indicates a qualitative improvement in the material: longer nanowires probably have a larger surface area due to an increased aspect ratio. Otherwise, as shown in Figure 4c, the increase in area-normalized capacitance with increasing SiNW length may be accompanied by a decrease in volumetric capacitance. This pattern means that no qualitative improvement is observed: the area-normalized capacitance growth arises from an increase in the volume of active material, while the area occupied by the electrode on the plate remains the same (i.e., the surface area in contact with the electrolyte). Nevertheless, longer SiNWs are preferable because they exhibit improved capacitance, even if this is only a quantitative effect.
However, the optimal length of SiNWs is often shorter than the maximum (see Figure 4d). Namely, when a critical value is exceeded, the area-normalized capacitance, which has been growing up to that point, begins to decrease. As reported in work [48], increasing the SiNW length above a critical value leads to wire bending, which probably causes a loss of vertical alignment, making it more difficult to form an additional layer on the SiNWs. The longest SiNWs can also be associated with a lengthened pathway for electron diffusion [52]. The longer wires allow for a larger surface area for loading more active materials, but the pathways of electrons are also increased, increasing the electron transport resistance [53]. The inability to fully cover long SiNWs may degrade the electrochemical performance of the composite with polyaniline (PANI) [54]. Similar results are obtained by Maboudian’s group: the maximum capacitance (325 mF/cm2) is achieved by coating 120 μm SiNWs with carbon in work [49], but subsequent deposition of an additional layer of MnOx is carried out on the shorter (only 80 μm) C/SiNW structure and yielded 381 mF/cm2 [55].
The authors of [56] report increasing capacitance with increasing depth of the porous layer for both pSi and pSi with a graphene-like coating (GLC). The maximum depth of pSi is limited by the geometry and structure of pSi samples when they are exposed to high temperatures and pressure drops during GLC synthesis. Samples with a 150 μm thick porous layer do not change noticeably in geometry or structure; layers with a thickness of 180 μm are characterized by deformation; and cracking of the porous layer occurs when its thickness is increased to 200 μm. Work [56] also shows that the capacitance of intrinsic pSi increases up to great depths (150–180 microns), but the charge accumulation mechanism switches from a capacitive to a battery type, which is also inappropriate. Most likely, this is due to the difficulty of completely impregnating the structure with the electrolyte. Notably, the samples have been exposed to the electrolyte at room temperature for seven or more hours before measurements to ensure satisfactory wettability of the electrolyte to the electrode and, therefore, to obtain accurate results. Boiling in 3 M H2SO4 solution was also used to enhance impregnation, but the results were practically the same as those obtained through simple exposure at room temperature. As discussed in [57], the porosity of electrode materials leads to surface nonequipotentiality. It can complicate the formation of the EDL and cause nonlinearity in charge characteristics, which is not necessarily associated with Faradaic processes.

5. Morphology

As discussed in Section 2.5, MACE provides a higher capacitance for the obtained SiNWs than VLS. Indeed, VLS yields chaotically oriented SiNWs that are either straight (as in Figure 5a) or curved and tangled (as in Figure 5b). In contrast, MACE provides vertically aligned SiNWs in a dense, ordered array (see Figure 5c) or SiNWs stuck together at the top (see Figure 5d).
The superior capacitances (more than 300 mF/cm2) achieved by Maboudian’s group [49,55] were also based on ordered oriented thin and dense SiNWs. In contrast, Shen’s group [59,60] used a fundamentally different approach. They obtained short and relatively sparse wires instead of an array. An additional tetramethylammonium hydroxide (TMAH) process was used to reduce the density of the SiNWs. The subsequent goal was not to coat each nanowire; the silicon matrix acted as a scaffold to be filled with active material. This interesting technique yielded a capacitance of more than 300 mF/cm2. Section 6.4 will describe Shen’s group’s path in detail.
A similar approach was demonstrated in work [61], where the strategy was also to fill voids inside the nanostructured silicon mesh with an active material. A combination of VLS-grown SiNWs, covered with a protective alumina layer, embedded in a gelatinous-like polymer matrix was successfully achieved. The flexibility of the SiNWs ensures the proper electrical and mechanical connection between the nanowires and the electrochemically active polymer material while allowing a high mass loading of the nanostructure without breaking it. While not very high capacitance values were achieved, the gelatinous-like polymer matrix provided excellent cyclic stability of the composite.
The first papers on pSi microcapacitors reported capacitances of less than 1 mF/cm2 for micron-sized pores [62,63]. In recent works, improved capacitances of about 100 mF/cm2 and higher have been achieved on DRIE-etched SiNRs (with diameters of several thousand nanometers) [34,35] and then on anodic etched pSi with nm-sized pores [56,64].
A distinctive feature of pSi/SiNW nanostructures was found: a herringbone-like morphology is observed in both the most capacitive pores [64] and the most capacitive individual wires (see Figure 6a and Figure 6b, respectively). This demonstrates that the herringbone structure can work successfully at the nanoscale despite the difficulty of electrolyte impregnation if other conditions are properly selected.
The formation of nanocrystallites on the pore surface was observed in work [65]. The disproportionate reaction (2) yielded secondary silicon atoms that are resistant to dissolution and unable to integrate into the crystal lattice. These atoms formed small nanocrystallites (α-Si), which could significantly affect electrochemical performance and were destroyed during pulsed photonic annealing. A similar result was obtained by Wu et al. [66], but they reported pore fusing with increasing annealing temperature. A more detailed study of this phenomenon needs to be conducted, since pSi is a promising material for microcapacitors, although less represented in the literature.

6. Coatings

The deposition of various coatings in a complex structure of pSi/SiNWs structures is an interesting challenge. Additional layers have a significant impact on pSi/SiNW capacitive properties. Below is an overview of the coatings, grouped by material type. The following aspects are considered for each coating: deposition method, composition analysis, and influence on pSi/SiNWs’ electrochemical characteristics (maximum capacitance achieved, which is just over 1 mF/cm2, and cyclic stability).

6.1. Metallic Coatings

Ru:
The ruthenium-coated SiNWs were obtained in work [26] through a simple electroless layer deposition (ELD) process using ruthenium salt. The SiNW substrate was immersed in an aqueous solution of RuCl3·H2O (1 mM) and HF (0.15 M) for different times at room temperature. The sample was then rinsed with water and dried under a gentle stream of nitrogen. The result was the deposition of the ruthenium NPs consisting of small nanocrystallites on the silicon surface. Elemental mapping and X-ray photoelectron spectroscopy (XPS) ruled out the surface oxidation of the Ru NPs. High-resolution transmission electron microscopy (TEM) and X-ray diffraction (XRD) analyses exhibited clear fringes for Ru NPs, with a lattice spacing of 2.3 Å, which corresponds to the (100) interplanar distance of hexagonal ruthenium. The specific capacitance values increased from 12.5 to 36.25 mF/cm2 (at 1 mA/cm2) upon increasing the Ru NPs deposition time from 30 to 75 min. An increase in capacitance was explained by the increase in Ru NPs density observed in SEM images. Excellent capacitance (112 mF/cm2 at 0.1 mA/cm2) and cyclic stability (80% capacitance retention after 50,000 cycles) were achieved.

6.1.1. Transition Metal Oxides

RuOx:
Ruthenium oxide was deposited by Zheng et al. [22] on the SiNWs via atomic layer deposition (ALD). In a commercial ALD chamber, diethylruthenocene was heated to 110 °C, and oxygen at room temperature served as the precursor. Different numbers of ALD cycles were used to create layers of varying thickness and continuity. The growth temperature was kept at 290 °C. XPS analysis revealed that the samples contained ~40% oxide phase and ~60% metallic ruthenium. EIS measurements showed that the contribution of Si to the overall electrode performance was negligible for electrodes with 400 ALD cycles. Capacitance was increased by over two orders of magnitude (to 19 mF/cm2 with 92% capacitance retention after 10,000 cycles), compared to bare Si electrodes. This increase in capacitance was due to the relatively continuous RuOx coating on the nanowire surfaces, which was uniform from top to bottom and prevented the Si surface from contacting the aqueous electrolyte.
MnOx:
The MnO2 dispersion was prepared using the electrophoretic deposition (EPD) method in work [33]. Manganese sulfate (MnSO4), ammonium persulfate ((NH4)2S2O8), and ammonium hydroxide (NH4OH) were used as the manganese source, oxidizing agent, and precipitating agent, respectively. Then, a manganese oxide shell layer was deposited on SiNWs from the obtained solution by an EPD process at room temperature. XRD confirmed the presence of the MnO2 NP coating on the SiNWs. Excess mass loading (30 s of EPD) decreased the specific capacitance of the electrode. The components in the core–shell structure greatly affected performance: the MnO2 NPs (deposited for 15 s) stored charge, and the SiNWs offered a shorter pathway for charge transport to the electrode. However, the overall capacitance (2.1 mF/cm2) of the MnO2 NPs-modified electrodes was not particularly high.
Dubal et al. [30] grew ultrathin MnO2 nanoflakes on SiNWs using a simple ELD method. They mixed a 50 mL solution of 2 mM potassium permanganate (KMnO4) with 2 mL of 98 wt% hydrochloric acid (HCl). A silicon wafer with pre-deposited SiNWs was then immersed in the bath at 323 K for 5, 10, 15, or 20 min. Finally, the MnO2/SiNW substrates were removed, rinsed, and dried in a vacuum at 373 K for two hours. The proposed growth mechanism for MnO2/SiNWs was as follows: initially, MnO4 nuclei were produced, adsorbed on the surfaces of SiNWs, and formed MnO2 nuclei upon reduction, which grew, aggregated, and transformed to nanoflakes. Ultimately, the MnO2 nanoflakes became compact and completely covered the surface of the SiNWs. XPS investigations indicated that the oxidation state of Mn in the heterostructure was Mn (IV). Selected area electron diffraction (SAED) and XRD data revealed the formation of polycrystalline birnessite-type MnO2. The unique 3D mesoporous structure of MnO2 on SiNWs provided large-area contact for the electrode and electrolyte; this structure enabled the accommodation of large volume changes and the release of associated strain generated during rapid charge and discharge cycling; the MnO2/SiNWs core–shell heterostructure was strongly supported on a silicon wafer, eliminating the need for a polymer binder or conductive additives. The design of the hierarchical MnO2/SiNWs core–shell heterostructure was improved by coupling it with a novel, Li-ion-doped ionic liquid electrolyte based on LiClO4 and 1-methyl-1-propylpyrrolidinium bis(trifluoromethylsulfonyl)imide (PMPyrrBTA). LiClO4, the primary ionic working species, was reversibly inserted into and out of the lattice tunnels between the MnO6 octahedral subunits. This caused a large amount of the manganese oxide to participate in surface redox reactions. The capacitance of 13 mF/cm2 was achieved with 91% retention after 9000 cycles.
In work [67], MnO2 was deposited on SiNWs by ELD from KMnO4. Prior to deposition, the SiNWs substrate was immersed in HF (5%) solution to remove the native silicon oxide layer and generate a hydrogen termination (H-SiNWs). The latter plays a key role in the deposition mechanism according to the following reactions [68]:
SiNWs/Si + HF → H-SiNWs/Si
H-SiNWs/Si + KMnO4 → MnO2/SiNWs/Si
Energy-dispersive X-ray (EDX) analysis confirmed the presence of MnO2, while the secondary-ion mass spectrometry depth profiles of Mn, O, and Si atoms revealed that SiNWs were partially oxidized and that MnO2 deposition on SiNWs could reach a depth of about 3 μm. The deposition of MnO2 on H-SiNWs increased the capacitance to 21.3 mF/cm2 [67] due to its pseudo-capacitive effect, wherein the quasi-reversible redox reaction involves the insertion/deinsertion of alkali ions (Na+) or protons (H+) in the empty sites of the MnO2 [69,70,71]:
MnO2 + H+/Na+ + e ↔ MnOOH/MnOONa
The MnO2 deposition process may also involve carbon as a reducing agent [72]:
4KMnO4 + 3C + H2O → 4MnO2 + K2CO3 + 2KHCO3
The redox reaction (16) was carried out by the authors of [55] for MnOx layer deposition on the C/SiNWs by means of an ELD method from KMnO4. KMnO4 concentrations were varied between 20 to 100 mM, and ELD time intervals were varied between 5 and 60 min. The electron transfer from nanocarbon to MnO4 initiated the heterogeneous nucleation of MnOx on the active surface. The formation of MnOx was thus closely related to the carbon structure. XPS studies revealed a predominance of MnO over the Mn2O3 phase in the MnOx coating layer. The electrochemical analyses indicated that the best performance was obtained on SiNWs with a length of 75 μm and in an 80 mM KMnO4 solution for 45 min. The capacitance of 381 mF/cm2 was achieved with 84% retention after 5000 cycles. The decrease in capacitance after extended cycling might be due to electrochemical dissolution of the MnOx layer, as described in previous studies [73].
Electrochemical deposition (ECD) and ELD methods were compared in work [34] as routes for manganese oxide deposition onto DRIE-etched silicon taper nanorods (SiTNRs) coated with TiN. The ECD process was realized by a galvanostatic process at a current density of 0.4 mA/cm2 for 16 min in an aqueous solution of 0.01 M Mn(CH3COO)2 and 0.02 M Na2SO4 at room temperature. ELD of MnO2 was performed for 8 h in a 200 mL aqueous solution containing 2 mM KMnO4 and 5 mM KOH. Volatile N2H4 was introduced into the KMnO4 solution as the reductant, thus supporting the growth of MnO2. EDX analysis confirmed that MnO2 was deposited on the SiTNR/TiN structures by using both methods. Both electrodes retained approximately 95.7% of their initial capacitance after up to 5000 cycles, while the main capacitance fading occurred during the first 500 cycles. However, the capacitance value increased noticeably from ~40 mF/cm2 (for ECD-deposited MnO2) to ~80 mF/cm2 (for ELD-deposited MnO2) at 10 mV/s.
Section 6.4 will discuss another study in which manganese oxide was combined with other coatings and used as a filler for a sparse SiNW scaffold [59,60,74].
NiO:
N-doped carbon (N-carbon) underlayer was used to prepare SiNWs/N-carbon/NiO electrode [75] through the pyrolysis of polydopamine (PDOP). The SiNWs/PDOP composites were prepared by a modified oxidative polymerization (OxP) method in a quartz glass reactor equipped with a water jacket connected to cooling circulating water pump equipment, which ensured that the polymerization occurred at 0 °C. 100 mg of dopamine hydrochloride, 100 mg of ammonium persulfate, and 60 mL of the phosphate buffer solution (pH = 6.5) were added to the reactor. The reaction lasted for 24 h and was repeated twice. Raman analysis revealed that the carbon coating was predominantly composed of sp2-type carbon, and XPS results demonstrated that the calcination treatment had transformed PDOP to the N-carbon. Subsequently, NiO deposition was carried out by ELD method for 10 min in the mixed solution of 20 mL of 1 M nickel sulfate, 16 mL of potassium persulfate, and 4 mL of aqueous ammonia, followed by a one-step calcination treatment in an argon atmosphere at 500 °C for 4 h. Field emission SEM confirmed that the NiO coating consisted of nanoflakes (with a thickness of about 100 nm as measured using TEM). The as-prepared SiNWs/N-carbon/NiO composite electrode showed a specific capacitance of up to 110 mF/cm2 with 80.8% retention after 4000 cycles. The N-carbon and nickel oxide played a synergistic role in improving the electrochemical performance in aqueous electrolytes. Enhanced charge transfer arose from the N-carbon coating, while additional pseudocapacitance arose from the highly reversible redox reaction (17): oxidation (forward reaction) and reduction (back reaction) [76].
NiO + OH ↔ NiOOH + e.

6.1.2. Other Metal Compounds

TiN:
ALD-deposited TiN coating greatly improved the electrode performance by increasing the conductivity and passivating the silicon surface, as has been found by several groups [34,77,78,79]. TiN with near-ideal conformality was deposited via ALD at 450 °C using TiCl4 and NH3 as the precursors and N2 as the carrier gas.
In work [77,78], approximately 6 μm thick pSi was prepared during 30 min anodization at a current density of 5.3 mA/cm2 in 50% HF and ethanol solution (1:4). Using two pPSi/TiN electrodes with a polydimethylsiloxane frame as a separator and as a reservoir for electrolyte, the volume-normalized capacitance of 7.3 F/cm3 was measured in work [77], and during the 5500 cycles, the capacitance changed very little. We obtained the area-normalized capacitance value by dividing the volume-normalized one by the porous layer thickness (6 μm):
7.3 F/cm3 × (6 × 10−4 cm) = 4.38 × 10−3 F/cm2 = 4.38 mF/cm2,
The in-chip electrochemical cell, consisting of two pPSi/TiN electrodes inside silicon plates with an electrolyte reservoir between them, was implemented in work [78]. This device configuration (for more details, see Section 7) allowed the achievement of higher volumetric capacitance (with 70–75% retention after 13,000 cycles) as well as higher energy and power densities.
In the same ALD process, TiN was deposited onto SiNRs fabricated through a cyclic DRIE route (75 etching cycles were performed to obtain SiNRs with a depth as high as ~20 μm) [79]. SEM and EDX studies illustrated the successful growth of TiN on SiNRs. As a result, the capacitance of 1 mF/cm2 was achieved with 95.2% retention after 2000 cycles.
The SiTNR/TiN was coated with MnO2 through ELD, as described in Section 6.1.1 [34], yielding a capacitance of 81.6 mF/cm2 with 95.7% retention after 5000 cycles.
NiB:
The deposition of nickel boride (NiB) onto SiNWs was carried out through ELD [25]. Nickel chloride (NiCl2) aqueous solution was mixed with ethylene diamine C2H4(NH2)2 and then with NaOH solution containing NaBH4 as a reducing agent. The final mixture was heated up to 90 °C, and SiNW/Si substrates were immersed in it for 10, 20, and 30 s. Energy dispersive spectroscopy and XPS data confirmed the NiB formation over the surfaces of SiNWs. SEM results showed uniform coverage of the underlying template along the entire SiNWs length. XRD indicated that the as-deposited NiB is amorphous (the capacitance of crystalline NiB needs to be explored in the future). The reported preparation method led to the formation of an amorphous phase and allowed the achievement of 165.7 mF/cm2 capacitance with 93% retention after 15,000 cycles.
CrN:
Chromium nitride (CrN) layers were deposited onto the SiNWs array by bipolar magnetron sputtering [28]. The usage of bipolar sputtering with two chromium targets, instead of conventional sputtering, was beneficial for well-crystallized nanostructures. SEM, electron diffraction, and XRD analyses revealed that the SiNWs were covered uniformly with small cubic CrN crystallites. EIS measurements indicated that the CrN layers reduced the charge transfer resistance and the electrodes became more conductive as the thickness of the CrN layer increased from 290 to 900 nm. However, the areal capacitance increased with the CrN thickness up to 550 nm, then stabilized at 180 mF/cm2 as the thickness reached 550 nm (the electrode could sustain 92% of its initial capacitance value after 15,000 cycles). A further increase in CrN thickness (from 550 to 900 nm) led to a capacitance decrease, because part of the surface became inaccessible. TEM images showed a herringbone structure due to the deposition of 550 nm CrN on SiNWs. This peculiar nanostructure, which contained highly nanoporous channels, was expected to contribute to the electrode capacitance enhancement. We should recall that the herringbone-like morphology was observed in both the most capacitive AE-etched pSi [64] and the most capacitive MACE-etched SiNWs [55].

6.2. Carbon Coatings

Due to its biocompatibility, chemical stability, high specific surface area, and electrical conductivity, carbon is a promising material for improving the properties of pSi/SiNWs for potential applications in optoelectronics [80], sensors [81,82,83,84,85], and photodetectors [86]. Below, various carbon coatings proposed for improving the electrodes of pSi/SiNW microcapacitors are listed and described.

6.2.1. Diamond

Boron-doped diamond coating [87] was synthesized onto SiNWs for 10 min in a microwave (MW) CVD system from a H2/CH4 gas mixture containing trimethylborane at an MW power of 2750 W, a gas pressure of 40 mbar, and a substrate temperature of 675 ± 10 °C. SEM images showed that a uniform, adherent, and homogeneous coating was deposited along all SiNWs with a thickness of ~100 nm. The nanocrystalline grain sizes varied from tens of nanometers to approximately 150 nm. The boron-based doping and structure of crystalline diamond were confirmed by Raman spectroscopy. A capacitance of 1.5 mF/cm2 was achieved with 65% retention after ultra-long cycling measurements (1,000,000 cycles).
In work [88], SiNWs were successfully coated with diamond-like carbon (DLC) films using EPD in a two-electrode electrolytic cell (graphite sheet was used as the anode, SiNWs as the cathode, and dimethylsulfoxide solution as the electrolyte) at 70 °C for 60 min at an applied voltage of 150 V. SEM analysis showed that the DLC films were uniformly deposited on the whole SiNW surface and composed of compact and small spherical grains with uniform distribution. Raman spectroscopy, EDX, and XRD studies indicated that the coating films were DLC. The electrochemical behavior of the SiNWs was clearly enhanced by the DLC coating, showing a specific capacitance of 2 mF/cm2 (at 0.01 mA/cm2) with capacitance retention of 90% after 16,000 cycles. Notably, a specific capacitance of 400 mF/cm2 was achieved at a scan rate of 5 mV/s, but an unusually large difference between CVA and galvanostatic charge–discharge GCD measurements was not explained in the manuscript, so we present the results obtained from GCD here.

6.2.2. SiC

In work [51], a thin (several tens of nm) SiC layer was chosen for chemically passivating the SiNWs using low-pressure CVD from 1,3-disilabutane (5 sccm) at 800 °C and 5 × 10−6 Torr for 3 min. A capacitance of 1.7 mF/cm2 was achieved for the longest (32 μm) SiNWs with SiC coating, and remained stable (95% retention) over 1000 charge/discharge cycles.

6.2.3. Nanocarbon

A highly activated carbon layer was developed on vertically aligned SiNWs through pyrolysis of glucose followed by annealing at 600 °C [48]. SEM and TEM results revealed that the nanocarbon layer (thickness of ~20 nm) was uniformly coated over the SiNW length (up to 19 μm). The carbon-coated 17 μm thick SiNWs displayed a specific capacitance as high as 25.6 mF/cm2 with 75% retention after 25,000 cycles.
Depositing an ultrathin (sub-nm) carbon sheath over the SiNWs enabled full electrolyte access to the porous surface area while still mitigating Si degradation [49]. Carbonization was performed in an atmospheric pressure CVD chamber at 900 °C in a CH4/Ar flow for 30 min. Raman spectroscopy data indicated a predominantly sp2 hybridized carbon film with a correspondingly high electronic conductivity. The C/SiNW composite yielded 325 mF/cm2 with 83% retention after 5000 cycles.

6.2.4. 1D Nanocarbon

In work [35], fullerene-like carbon (FC) decorated carbon nanotubes (CNTs) were used to fill a DRIE-etched SiTNR scaffold. Ni catalysts for CNT growth were deposited by electron-beam deposition at a high vacuum (5 × 10−6 Torr). The growth of FC-CNTs was carried out via an atmospheric pressure CVD process in Ar/H2/C2H2 (100:20:1 v:v:v) atmosphere at 750 °C for 12 min. SEM images revealed that the interconnection assembly of CNTs readily created abundant interspaced macropores, while the FC NPs contributed to forming mesopores (2 to 10 nm), as shown by TEM. Altogether, 1D nanocarbons incorporated in a tapered silicon nano-scaffold improved the capacitance (123 mF/cm2 with excellent retention of 102% after 5000 cycles).
The CNTs were also used for pSi decoration: new functional nanocomposite materials based on layers of pSi, CNTs, and tin oxide were proposed for supercapacitor electrodes [89]. The SnOx/CNT/pSi hybrid electrode exhibited a specific capacitance of 347 mF/cm2, but its cyclic stability was insufficient (50 charge–discharge cycles were carried out to assess the cyclic stability of the measured electrodes).

6.2.5. Graphene-Based Films

The graphenic layer was grown directly on the pSi structure through Ni-assisted CVD at heating temperatures of 1000–1100 °C under an increased pressure of 60 Torr in a CH4/Ar flow [66]. The coating consisted of few-layer graphene (FLG), which was confirmed by Raman spectra. High-magnified SEM images showed that after CVD at 1000 °C, the average pore diameter was ~11 nm (in the range of mesopores). An increase in the annealing temperature to 1100 °C caused the pore coalescence, with all mesopores merging to form larger macropores. At the annealing temperature of 1050 °C, some mesopores (but not all) fused together, forming a porous structure composed of both mesopores and macropores. This hybrid-porous FLG/SiNWs enabled achieving a high areal capacitance of 6.21 mF/cm2 and an unusual progressive cyclic stability of 131% at 10,000 cycles.
Electropolymerization (ElP) of 2,6-dihydroxynaphthalene (DHN) using CVA or potentiometry with subsequent thermal pyrolysis (800 °C, N2, 4 h) resulted in a GLC strongly adhering to the nanoporous silicon matrix [90]. ElP was carried out using a 2 mM solution of DHN dissolved in the electrolyte solution (10 mM phosphate buffer saline solution, pH 7.40, containing 0.1 M KCl):
-
by potentiometry—at 1 mA for 120 s (“J” sample, active mass = 90.8 μg);
-
by CVA—at 5 mV/s for 6 cycles (“CV” sample, active mass = 74.2 μg).
After annealing, the capacitances were measured: 162 F/g and 100 F/g for the “J” and “CV” electrodes, respectively. Since the active area was equal to 0.49 cm2, the area-normalized capacitances were calculated as follows:
162 F/g = 162 × (90.8 × 10−6 g) × (1/0.49 cm2) = 30 mF/cm2,
100 F/g = 100 × (74.2 × 10−6 g) × (1/0.49 cm2) = 15 mF/cm2.
The capacitance retentions after 1000 cycles were equal to 75% and 80% for the “J” and “CV” electrodes, respectively. XRD data confirmed a low graphitization degree of GLC, which may be the reason for poor cyclic stability.
Continuous GLC, consisting of sub-ten-nanometer graphene flakes, was deposited in the pSi structure in a CVD reactor through high-temperature (950 °C), low-pressure (∼100 Pa) pyrolysis of C2H5OH vapor [56]. For GLC deposition inside the porous structure, a sharp pressure drop (SPD) was required during the synthesis. The SPD mode of CVD allowed the synthesis of GLC throughout the full depth (up to 200 μm) of a complex, herringbone structure of nanopores, which was confirmed by Raman spectroscopy. As a result, the composite electrode showed a capacitance of 87 mF/cm2 and retained 100% of its value after 15,000 cycles.
In work [64], the above CVD procedure was carried out (but acetonitrile was used as the precursor) for N-doped GLC (N-GLC) deposition in pSi structures. XPS analysis revealed that the main configuration of the nitrogen atoms embedded in GLC was graphitic. The capacitance increased (up to 145 mF/cm2) when pSi was coated with N-GLC due to the enhanced conductivity (namely, improved electron transfer) induced by the graphitic configuration of the N-dopant. Good cyclic stability was also achieved (over 20,000 cycles even at low scan rates).
Carbon undercoating was also used for NiO [75] and MnOx [55] deposition as described in Section 6.1.2, and for PEDOT [58] and PANI [91] deposition, as will be discussed in Section 6.3.

6.3. Polymer Coatings

6.3.1. PPy

Polypyrrole (PPy) films were electrochemically deposited onto SiNTrs from a PYR13TFSI (N-propyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)imide) solution containing 0.1 M Py as the monomer [27]. The ElP was carried out in a three-electrode electrochemical cell using a potentiostat–galvanostat at a constant potential of 0.4 V (vs. Ag/Ag+) under a polymerization charge of 750 mC/cm2 controlled by the chronocoulometry technique in an argon-filled glove box with oxygen and water levels below 1 ppm. A uniform and homogeneous PPy coating had a globular morphology as illustrated by SEM. The thickness of the conducting polymer coating was determined to be around 20 nm using TEM. From CVA, a pair of redox peaks were detected, which are associated with the oxidation (forward reaction (21)) and reduction (back reaction (21)) of Ppy (intercalation and deintercalation of electrolyte anions, respectively):
PPy + n TFSI ↔ (PPy)n + (TFSI)n + ne.
The PPy coating increased the capacitance (to 14 mF/cm2). However, 30% capacitance loss was observed after 10,000 cycles, whereas bare SiNTr microcapacitors exhibited no loss of electrochemical stability after 10,000 galvanostatic charge–discharge cycles.

6.3.2. PEDOT

The surface functionalization of SiNWs by electroactive conducting polymers was proposed in work [29]. Poly(3,4-ethylenedioxythiophene) (PEDOT) coatings were electrochemically deposited from a propylene carbonate solution containing 10 mM EDOT as the monomer and 0.1 M tetrabutylammonium tetrafluoroborate (TBABF4) as the electrolyte at a constant potential of 0.9 V (vs Ag/Ag+) for 1 h. SEM study showed that a nanometric PEDOT coating (thickness of 100 nm) with a globular morphology was deposited along all SiNWs, increasing the capacitance to 17 mF/cm2.
In work [58], SiNWs were introduced into an MWCVD system and exposed to a H2/CH4 gas mixture at an MW power of 2750 W and a temperature of 675 °C to coat the electrode with diamond. Subsequent electrochemical deposition of nanometric PEDOT films was carried out from a PYR13TFSI solution containing 0.1 M EDOT as the monomer (as described in [92]). The deposition of diamond and PEDOT was confirmed by Raman spectroscopy and XPS, respectively. Scanning TEM and EDX analyses showed that the PEDOT and diamond conformal coatings were deposited on SiNWs with a 3D multi-hierarchical structure (see Figure 7), which yielded a capacitance of 8.5 mF/cm2 (at 1 mA/cm2) with 80% retention after 15,000 cycles in butyltrimethylammonium bis(trifluoromethylsulfonyl)imide (N1114TFSI) electrolyte.

6.3.3. Combined Polymers

One-step deposition of PEDOT with poly(styrenesulfonate) (PSS) onto Al2O3-protected SiNWs was reported in work [61]. A 3 nm thick Al2O3 layer was deposited using ALD from trimethylaluminum and H2O at 250 °C under 10−2 Torr. By replacing the uncontrolled, native, and inherent Si oxide layer with alumina, the silicon nanostructure was protected from any oxidative stress from contact with the aqueous electrolyte. The conducting polymer layer was deposited by a simple drop of ethanol, followed by three droplets of the homogenized PEDOT-PSS solution. SEM images showed a mesh of highly doped 50 μm long Al2O3-coated SiNWs (Figure 8a) embedded inside a 5 μm thick gelatinous-like polymer matrix (Figure 8b), ensuring an excellent electrical connection between these two materials. The coating of the polymer flattened the nanostructure, and the flexible nanowires interlocked with each other. This new nanocomposite material exhibited a capacitance of 8.5 mF/cm2. Stability was confirmed by a performance decline of only 5% after at least 500,000 cycles. Throughout the cycling, the electrolyte was percolating inside the composite, reorganizing the polymeric matrix by the successive ionic intercalations and deinsertions. The areal capacitance, therefore, increased until the whole nanocomposite was soaked with the electrolyte, where it reached a stable plateau.
SiNWs/PEDOT/PPy composites were obtained consecutively by a photo- and chemical-OxP method [52]. An in situ photo polymerization (PhP) onto hydrogen-terminated SiNWs (H-SiNWs) was carried out via the reaction with a monomer solution (0.005 mol of EDOT, 25 mL of deionized water, and 25 mL of acetonitrile) under UV irradiation at 365 nm for 3 h in a quartz glass reactor equipped with a blue light filter.
After that, the SiNWs/PEDOT electrodes were suspended in 40 mL of monomer solution (1 M HCl containing 0.01% (v/v) pyrrole), and 10 mL of 1 M FeCl3 aqueous solution was added dropwise. The polymerization reaction was carried out at about 0 °C for 12 h. The Raman spectra proved the successful synthesis of PEDOT and PPy in the composites. As indicated by SEM, the SiNWs/PEDOT/PPy had a core–shell structure, and the thickness of the PPy nanolayers was about 40 nm. XPS data confirmed the interaction between PEDOT NPs and SiNWs in the composite (the photoexcited SiNWs directly initiated the PhP of EDOT at the surface of SiNWs, which built up this interaction). The PEDOT NPs with excellent electrical conductivity not only acted as efficient conducting highways for facilitating electron transfer, but also made the surface of SiNWs more compatible with the PPy nanolayers. The composite exhibited a specific capacitance of 106 mF/cm2 with a retention of 80.2% after 5000 cycles, and its core–shell structure remained stable after the cycling test.

6.3.4. PANI

The in situ polymerization of aniline over SiNWs was carried out via the OxP process [54]. The SiNW substrate was soaked in 0.5 M H2SO4 containing aniline as the monomer for 30 min. Consequently, the mixture of 0.03 M ammonium persulfate ((NH4)2S2O8) and 0.5 M H2SO4 was gradually added to the initial solution under an ice-water bath for 4 h. PANI layers of multifarious morphologies covered over SiNW arrays were obtained by varying the aniline concentrations from 0.4 mM to 6.4 mM. Field emission SEM revealed that the coral-shaped PANI/SiNW hybrid material was obtained at the aniline concentration of 1.6 mM. XPS and Raman peaks registered from the composite surface were assigned to the PANI macromolecular chain. Water contact angle measurements demonstrated that modification of SiNWs with PANI led to greater hydrophilicity and therefore accelerated the exchange rate of H+ ions during the charge–discharge process. As a result, a capacitance of 81 mF/cm2 was achieved with 71.8% retention after 2000 cycles.
The influence of PANI and graphene nanowalls (GNWs) (both separately and together) on the SiNWs’ electrochemical performance was investigated in work [91]. PANI was electrochemically deposited using a monomer solution (a mixture of 0.1 M aniline, 30 mL of ethanol, and 70 mL of 1 M H2SO4) at a constant potential of 0.8 V (vs Ag/Ag+) for 15 min. GNWs were grown on the plasma-etched SiNWs’ surface by an MW plasma-enhanced CVD method from H2/CH4 for 20 min at an RF power of 700 W and at a chamber pressure of 4 kPa. The capacitance achieved for SiNWs/GNW and SiNWs/PANI was 4.6 mF/cm2 and 20.9 mF/cm2, respectively. The sequential deposition of both coatings yielded an excellent improvement: the SiNW/GNW/PANI capacitance was 130 mF/cm2 with 80% retention after 2000 cycles.

6.4. Combined Coatings

The above-discussed combined coatings are mainly carbon layers under transition metal oxides and conductive polymers. The latter are often used in combination with each other.
Here, we describe the original Shen’s group concept that opens up an avenue to prepare 3D silicon nanostructures based on metal oxide and conductive polymers, which not only reduce the contact resistance between the active materials and the current collector but also shorten the ion and electron diffusion paths. Unlike traditional attempts to obtain a dense array of SiNWs and further coat each nanowire, the reduction of SiNW density is achieved in a mixed TMAH solution (TMAH:isopropyl alcohol:distilled water = 0.3:5:25 in volume ratio). It is known that TMAH is an attractive etchant solution for the silicon substrate due to its nontoxic, easy disposal, and high-quality anisotropic features without any metal ion contaminants. The reaction equation between silicon and TMAH is as follows [93]:
Si + 2(CH3)4NOH + 2H2O → 2(CH3)4N+ + SiO2(OH)22− + 2H2↑.
In work [53], the silicon matrix with sparse SiNWs (see Figure 9a) was filled with an active material (see Figure 9b,c): PEDOT + MnO2 were electrodeposited on the SiNWs in an electrolyte containing EDOT, Na2SO4, sodium dodecyl sulfate (SDS), and MnSO4 at a constant voltage of 2 V in a two-electrode cell. The current collector layer was prepared as follows:
-
Reduced graphene oxide (rGO), synthesized by hydrazine reduction, was deposited by ELD from an rGO solution for 10 min (see Figure 9d);
-
Subsequently, the PEDOT-PSS layer incorporated with Ag nanowires (AgNWs:PEDOT-PSS = 1:2) was spin-coated (see Figure 9e,f).
The rGO not only improves the capacitance of the electrode by providing EDL capacitance but also greatly promotes the transfer of electrons and ions. XRD analysis revealed that the birnessite-type MnO2 NPs had poor crystallinity. Nevertheless, the composite electrode material with low crystallinity exhibited higher ionic conductivity and shorter diffusion paths. Both of these properties are beneficial for improving capacitance, reaching 101 mF/cm2 with 81% retention after 2000 cycles.
In further works, Ni particles were electrodeposited on SiNWs in a two-electrode system at a constant current of 3 mA, from an aqueous electrolyte containing NiSO4, NH4Cl, and SDS. The Ni particles connected with each other to generate a highly conductive layer that prevented silicon from oxidation/corrosion, largely improving the stability of SiNWs and the electrode.
In work [74], a PEDOT + MnO2 layer was deposited on SiNWs/Ni structure by the one-step co-electrodeposition method as ascribed in [53]. Finally, Pt NPs decorated the electrode by immersing it in the ethanol solution containing H2PtCl6 for 1 min, followed by annealing at 180 °C for 30 min in vacuum. After decoration with Pt NPs, the conductivity of the composite was improved, and the surface became more hydrophilic, favoring ion/charge transport. EIS measurements also confirmed that Pt NPs decreased ion diffusion resistance and improved electron transfer. This electrode design (see Figure 10) provided a capacitance of 207 mF/cm2, losing 5% of its value after 2000 cycles.
In [59], the capacitance value was increased to 328.6 mF/cm2 (with 79% retention after 7000 cycles) after MnO2 electrodeposition on SiNWs/Ni under a constant current of 0.003 A in a two-electrode system, with a platinum sheet utilized as the counter electrode. The electrolyte contained 0.4 M MnSO4, 0.025 M SDS, and 0.1 M Na2SO4. Excessive mass-loading of MnOx not only hindered electron/ion transfer due to increased distance but also diminished contact areas between the active material and electrolyte, consequently impeding ion/electron transfer within the electrode. Notably, some cracks were observed within the dense MnOx layer. These cracks in the MnOx layer provided more active sites and facilitated electrolyte penetration, thereby enhancing the electrode’s capacitance.
PEDOT was electropolymerized on SiNWs/Ni, followed by dipping in KMnO4 solution [60]. Platinum particles were introduced between PEDOT and MnOx, which not only enhanced the mass loading of MnOx but also improved the conductivity of the composite. After spin-coating a PEDOT-PSS layer doped with silver nanowires on the composite, a hierarchical core–shell electrode comprising active materials and a current collector was formed. As a result, the capacitance of 352 mF/cm2 was achieved with 85% retention after 2000 cycles.
Ppy and nickel–cobalt sulfides were co-electrodeposited on nickel-particle-passivated SiNW arrays to achieve a gravimetric capacitance of 1602 F/g [94].
SiNWs/Ni/NiCoSe-rGO composite provided a gravimetric capacitance of 2268 F/g, which is equivalent to an areal capacitance of 1973 mF/cm2 [95]. The ECD of the nickel–cobalt selenide nanosheets and rGO was carried out at voltages ranging from 1.6 to 0 V with a scan rate of 15 mV/s in an electrolyte solution containing nickel and cobalt nitrates, selenium dioxide, lithium chloride, and rGO. The prepared NiCoSe layer was composed of aggregated nanoparticles, generating a porous layer. The abundant pores within the active layer favored full contact with the electrolyte, facilitating ionic diffusion. The hierarchical core–shell structures not only afforded abundant electrochemical reaction sites but also provided short distances for electron/ion transport, largely improving the charge transfer kinetics during the charging/discharging process.

6.5. Comparison

All the data from the considered works (as well as from works in which the achieved capacitance is less than 1 mF/cm2) are summarized in Table 1 in order of increasing capacitance value.
Clearly, VLS is superior to MACE in terms of using the obtained structures as electrodes of microcapacitors. Interestingly, two studies deviate from this pattern: VLS-synthesized SiNWs, when coated with crystalline chromium nitride [28] or amorphous nickel boride [25], exhibit capacitances greater than 100 mF/cm2. Both materials are pseudocapacitive. CrN layers reduce charge transfer resistance (this is atypical for pseudo-capacitive materials). More attention should be given to such metal-containing coatings, as well as to the relationship between the capacitive properties of SiNW-based composites and the crystalline structure of coatings.
MnOx is the most investigated transition metal oxide used as a coating for pSi/SiNWs electrodes. EPD-deposited MnO2 provides a slight increase in overall capacitance [33] compared to the precipitation method (ELD), which yields higher results [30,67]. A direct comparison of the two methods was also performed in the work [34], which confirmed the advantages of ELD. Moreover, the electroless process occurs at room temperature, making it compatible with microelectronic processing technology.
The other main types of coatings used are carbon-based, primarily graphene, and conductive polymers. Their advantage is high conductivity, which improves charge transfer. Researchers also note that a hierarchical core–shell structure provides protection from oxidative electrolyte contact and a shorter ion path, but a high degree of adhesion is necessary in such structures.
Table 2 summarizes the key coatings for pSi/SiNWs, as well as their main benefits and drawbacks regarding electrochemical performance.
Cyclic stability is an important characteristic that requires specific consideration. The Shen group’s concept, when sparse SiNWs act as a scaffold to be filled with active material, allows for superior capacitive performance. The authors attribute the cycle durability to the following:
-
the smart design without consideration of the passivation/corrosion of the SiNWs in aqueous solution and the synergistic effects of the core–shell configuration and the combination of PsAg and rGO [53];
-
cracks on MnOx can effectively alleviate volume variations in MnOx during electrochemical cycling [59];
-
rGO underlayer used for NiCoSe coating [95].
However, noticeable losses are observed at average cycling times (less than 10,000 cycles) and need to be confirmed.
MnOx loses capacitance as a result of consecutive reduction to soluble Mn2+ ion [104]. Very stable cycling performance has been achieved with MnO2 hierarchical sphere-based electrodes [73]. The results reveal that tiny nanostructures are generated and disappear during the cycling test, causing the reorganization of MnOx and resulting in a transformation of its morphology. The loss of active materials in the form of nanostructures clearly causes capacitance degradation. Currently, it is still unclear what triggers and controls such morphological evolution.
Nevertheless, MnOx possesses low cyclic stability. For instance, the MnOx/C/SiNW structure retains 84% of its initial capacitance after 5000 cycles [55]. However, C/SiNWs demonstrate significant capacitance degradation [49]. These losses probably arise not from MnOx dissolution, but from SiNW instability. Indeed, as seen in Table 1, for SiNW capacitance values greater than 100 mF/cm2, cyclic measurements do not exceed 15,000 cycles; for values greater than 200 mF/cm2, they do not exceed 7000 cycles with about 20% capacitance loss.
Conversely, 100% capacitance retention (or even higher) is observed for pSi (DRIE [35] or anodic [64,66] etched) with carbon coatings. These materials have been little investigated in the literature, while studies on SiNWs clearly outnumber those on pSi. This is likely due to the rapid development of MACE, a simple and accessible method. Certainly, the progress achieved allows us to consider that pSi is a promising material for on-chip supercapacitor applications, especially for long-term stable performance, and needs to be further studied.

7. Devices

Micropower sources have been in development since the 1980s. In 2003, the fabrication of microcapacitors was proposed in a study [105], causing a wave of interest in the field. After the report of an ultrahigh-power, micrometer-sized supercapacitor based on onion-like carbon [106], graphene-based planar micro-supercapacitors advanced significantly [107]. Other carbon materials, such as CNTs and porous carbon [108], as well as transition metal oxides, conductive polymers, and MXenes, have also been successfully integrated into microdevices or chip-based systems for energy storage [109]. Recently, the last class of 2D coatings opened up an exciting new field of 2D inorganic functional materials due to their intrinsic electronic conductivity, superior hydrophilicity, rich surface chemistry, and layered structure [110].
The above review focuses on nanostructured silicon electrodes and summarizes data on surface capacitance (which is important for on-chip applications as opposed to gravimetric ones) and the factors influencing it. PSi/SiNW-based microelectrochemical capacitors offer scalability and integration of on-chip energy storage components, but as discussed below, fabrication methods for such devices are currently underdeveloped.
Figure 11a [78] illustrates the fabrication of the in-chip supercapacitor device (a detailed description is provided in the Supplementary Materials, Section S3). To the best of our knowledge, the solid-state supercapacitor structure shown in Figure 11a has only been proposed in one study [78]. Most of the reported microcapacitors are fabricated by sandwiching the gel electrolyte between two pSi/SiNW-based electrodes, as shown in Figure 11b.
Table 3 contains data for some of the reported devices. The capacitance values for a single electrode (measured using a three-electrode electrochemical cell) are provided for comparison (certainly, these values are significantly higher than the device’s capacitance). As can be seen, energy and power densities can be adjusted by the properties of the electrodes and the measurement conditions.
Interestingly, the last two results are obtained in organic and aqueous electrolytes (works [55] and [95], respectively). Generally, organic electrolytes are preferable to aqueous ones because ionic liquids are stable over a much wider voltage window. This improves the theoretical maximum energy and power densities [49], which scale with the square of the potential window. For comparison, the authors of work [55] report a higher energy density (20–146 μWh/cm2) in an organic electrolyte with a 3.5 V voltage window than in an aqueous electrolyte with a 1.6 V voltage window (34–109 μWh/cm2 [95]), although the capacitance in the former work is almost six times lower than in the latter. Furthermore, hydrogen ions formed by water dissociation can diffuse through the silicon crystal lattice and passivate its dopants, thereby reducing the silicon matrix conductivity [111].
A sandwich-type electrode connection is not suitable for on-chip applications. Further development of on-chip microcapacitors also requires studying the features of devices with planar topology (e.g., as shown in Figure 12).
The electrochemical window can be extended when two different active materials are used in the negative and positive electrodes to create an asymmetric configuration (as shown in Figure 12), combining the potential windows of both electrodes [112]. Using conventional lithography, the gap between the electrodes should be reduced to micrometers, greatly increasing the resultant performance due to shorter diffusion paths [109]. Furthermore, it is essential to develop scalable and cost-effective preparation techniques for electrode materials to facilitate their large-scale production and commercialization [113]. Photo- or electron resists are unstable in HF solutions. Chemically resistant dielectric pattern layers must be applied to serve as both an etch mask and an insulator between the silicon substrate and the electrical contact. Although conventional microfabrication can reach resolutions of up to micrometers, high resolution can be hindered by multi-stage deposition processes. This requires a compromise between achieving the best electrochemical performance and using the simplest technological approaches to keep costs down when scaling up. To address compatibility challenges and accelerate the translation of research findings into practical applications, innovative approaches such as printable conductive inks should be prioritized [114].
The performance of the microcapacitors is influenced not only by topology, the width of the electrochemical window, and electrical contact formation, but also by electrolyte properties. Liquid electrolyte systems are poorly suited for on-chip integration or the satisfactory performance of embedded microelectronic components. In the reported pSi/SiNWs microcapacitor devices, gel polymers are mainly used, except in work [115], in which a solid polymeric separator is proposed. Since silicon-based anodes in advanced lithium-ion batteries have been studied and developed more extensively, we can apply their experience. General conclusions from reviews [116,117] that also apply to capacitors are as follows:
The drawback of all-solid-state electrolytes stems from inadequate interfacial contact between the solid electrolyte and the electrode surface. This leads to substantially elevated interfacial resistance and generally lower ionic conductivity.
Quasi-solid-state electrolytes employ gelation strategies to strike a balance between liquid- and solid-state characteristics. This achieves synergistic optimization of ionic conductivity and mechanical robustness. Additives such as ceramic and propylene carbonate and artificial interlayers have been shown to improve the structural integrity of the electrode/electrolyte interface.

8. Conclusions

On-chip electrodes for microcapacitors are a task that is of growing interest to researchers, as both alternative energy and portable devices are very relevant issues today. A summary of the literature data, including their compilation and analysis, leads us to conclusions about the most effective way to obtain the best electrochemical characteristics of pSi/SiNWs:
Aligned MACE-etched SiNWs are better than twisted ones synthesized by VLS.
Herringbone-like architecture of nanostructured silicon is preferred.
Nanopores in pSi are more successful compared to micropores.
The use of organic ionic liquids is preferable to aqueous electrolytes (quasi-solid-state electrolytes are promising).
The capacitance increases together with the growth of pSi depth/SiNW length, but the maximum depth/length can be limited by the following:
The difficulty of completely impregnating the structure with the electrolyte;
The difficulty of completely covering long SiNWs/deep pSi with additional coatings;
The lengthened pathway for electron diffusion.
SiNWs provide the achievement of high capacitance values, while pSi structures are very promising for long-term stable performance. Therefore,
The cyclic stability, especially for Shen’s and Maboudian’s group approaches, should be additionally investigated;
More coatings (carbon, primarily graphene-based, conductive polymers, and metal compounds, especially oxides) should be tested to coat pSi;
The influence of DRIE depth on pSi capacitance should be studied.
Most of the reported results are devoted to the materials of the electrodes. For further development of on-chip microcapacitors, it is necessary to study the features of performing devices. On the way to on-chip microcapacitors, many problems still need to be solved, both in their creation (micropatterning, insulation, and selection of a suitable electrolyte) and in their operation (cyclic stability and particle diffusion).
It is also necessary to use integral formulas to calculate the capacitance, as well as energy and power densities from GCD data, when the GCD curves deviate from the linear pattern.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15231826/s1, Section S1: Data table for Figures 1a and 1c; Section S2: The equivalent electrical circuit of the porous electrode; Section S3: Fabrication steps of the in-chip supercapacitor; Section S4: Areal calculations from reported gravimetric data.

Funding

This work was carried out within the framework of the State Task No. 075-00295-25-00.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
pSiPorous silicon
SiNWSilicon nanowire
AEAnodic etching
DRIEDeep reactive ion etching
MACEMetal-assisted chemical etching
VLSVapor–liquid–solid
SiNRSilicon nanorod
NPNanoparticles 
MAAEMetal-assisted anodic etching
Mmol/L 
CVDChemical vapor deposition
SMSupplementary materials
EDLElectrical double layer
CVACyclic voltammetry
GCDGalvanostatic charge–discharge
EISElectrochemical impedance spectroscopy 
PANIPolyaniline
GLCGraphene-like coating
SEMScanning electron microscopy
TMAHTetramethylammonium hydroxide
ELDElectroless layer deposition
XPSX-ray photoelectron spectroscopy 
TEMTransmission electron microscopy
XRDX-ray diffraction
ALDAtomic layer deposition
EPDElectrophoretic deposition 
SAEDSelected area electron diffraction 
PMPyrrBTA1-methyl-1-propylpyrrolidinium bis(trifluoromethylsulfonyl)imide
EDXEnergy-dispersive X-ray
ECDElectrochemical deposition
SiTNRSilicon taper nanorod
N-carbonN-doped carbon
PDOPPolydopamine
OxPOxidative polymerization
MWMicrowave
DLCDiamond-like carbon
FCFullerene-like carbon 
CNTCarbon nanotube
FLGFew-layer graphene
ElPElectropolymerization
DHN2,6-dihydroxynaphthalene
SPDSharp pressure drop
PPyPolypyrrole
PYR13TFSIN-propyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)imide
PEDOTPoly(3,4-ethylenedioxythiophene)
TBABF4Tetrabutylammonium tetrafluoroborate
N1114TFSIButyltrimethylammonium bis(trifluoromethylsulfonyl)imide
PSSPoly(styrenesulfonate)
PhPPhoto polymerization
GNWGraphene nanowall
SDSSodium dodecyl sulfate
rGOReduced graphene oxide
EMIM-TFSI1-Ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide
TEABF4Tetraethylammonium tetrafluoroborate
Et3NH TFSITriethylammonium bis(trifluoromethylsulfonyl)imide

References

  1. Mateen, A.; Khan, A.J.; Zhou, Z.; Mujear, A.; Farid, G.; Yan, W.; Li, H.; Li, J.; Bao, Z. Silicon Nanowires via Metal-Assisted Chemical Etching for Energy Storage Applications. ChemSusChem 2024, 18, e202400777. [Google Scholar] [CrossRef] [PubMed]
  2. Raman, S.; Sindhuja, M. Advances in silicon nanowire applications in energy generation, storage, sensing, and electronics: A review. Nanotechnology 2023, 34, 182001. [Google Scholar] [CrossRef]
  3. Ghosh, R.; Giri, P.K. Silicon nanowire heterostructures for advanced energy and environmental applications: A review. Nanotechnology 2016, 28, 012001. [Google Scholar] [CrossRef]
  4. Wu, Z.; Zhang, X.; Jin, X.; Li, T.; Ge, J.; Li, Z. A Review on Cutting Edge Technologies of Silicon-Based Supercapacitors. J. Nanomater. 2021, 2021, 6650131. [Google Scholar] [CrossRef]
  5. Granitzer, P.; Rumpf, K. Porous silicon—A versatile host material. Materials 2010, 3, 943–998. [Google Scholar] [CrossRef]
  6. Zhang, G.X. Porous silicon: Morphology and formation mechanisms. In Modern Aspects of Electrochemistry; Springer: Berlin/Heidelberg, Germany, 2006; pp. 65–133. [Google Scholar]
  7. Starkov, V. Production, properties, and application of porous silicon. All Mater. Encycl. Ref. Book 2009, 4, 13–21. [Google Scholar]
  8. Huang, Z.; Geyer, N.; Werner, P.; De Boor, J.; Gösele, U. Metal-assisted chemical etching of silicon: A review. Adv. Mater. 2011, 23, 285–308. [Google Scholar] [CrossRef]
  9. Schmidt, V.; Wittemann, J.V.; Senz, S.; Gösele, U. Silicon nanowires: A review on aspects of their growth and their electrical properties. Adv. Mater. 2009, 21, 2681–2702. [Google Scholar] [CrossRef]
  10. Huo, C.; Wang, J.; Fu, H.; Li, X.; Yang, Y.; Wang, H.; Mateen, A.; Farid, G.; Peng, K.Q. Metal-assisted chemical etching of silicon in oxidizing HF solutions: Origin, mechanism, development, and black silicon solar cell application. Adv. Funct. Mater. 2020, 30, 2005744. [Google Scholar] [CrossRef]
  11. Kang, Y.; Jorné, J. Morphological stability analysis of porous silicon formation. J. Electrochem. Soc. 1993, 140, 2258. [Google Scholar] [CrossRef]
  12. Canham, L.T. Silicon quantum wire array fabrication by electrochemical and chemical dissolution of wafers. Appl. Phys. Lett. 1990, 57, 1046–1048. [Google Scholar] [CrossRef]
  13. Smith, R.; Collins, S. Porous silicon formation mechanisms. J. Appl. Phys. 1992, 71, R1–R22. [Google Scholar] [CrossRef]
  14. Goryachev, D.; Belyakov, L.; Sreseli, O. On the mechanism of porous silicon formation. Semiconductors 2000, 34, 1090–1093. [Google Scholar] [CrossRef]
  15. Kolasinski, K.W. The mechanism of Si etching in fluoride solutions. Phys. Chem. Chem. Phys. 2003, 5, 1270–1278. [Google Scholar] [CrossRef]
  16. Ben-Chorin, M.; Möller, F.; Koch, F. Nonlinear electrical transport in porous silicon. Phys. Rev. B 1994, 49, 2981. [Google Scholar] [CrossRef]
  17. Pérez, J.M.; Villalobos, J.; McNeill, P.; Prasad, J.; Cheek, R.; Kelber, J.; Estrera, J.; Stevens, P.; Glosser, R. Direct evidence for the amorphous silicon phase in visible photoluminescent porous silicon. Appl. Phys. Lett. 1992, 61, 563–565. [Google Scholar] [CrossRef]
  18. Zhao, Y.; Yang, D.; Li, D.; Jiang, M. Annealing and amorphous silicon passivation of porous silicon with blue light emission. Appl. Surf. Sci. 2005, 252, 1065–1069. [Google Scholar] [CrossRef]
  19. Leopold, S.; Mueller, L.; Kremin, C.; Hoffmann, M. Online monitoring of the passivation breakthrough during deep reactive ion etching of silicon using optical plasma emission spectroscopy. J. Micromechanics Microengineering 2013, 23, 074001. [Google Scholar] [CrossRef]
  20. Leonardi, A.A.; Faro, M.J.L.; Irrera, A. Silicon nanowires synthesis by metal-assisted chemical etching: A review. Nanomaterials 2021, 11, 383. [Google Scholar] [CrossRef]
  21. Li, X.; Bohn, P. Metal-assisted chemical etching in HF/H2O2 produces porous silicon. Appl. Phys. Lett. 2000, 77, 2572–2574. [Google Scholar] [CrossRef]
  22. Zheng, W.; Cheng, Q.; Wang, D.; Thompson, C.V. High-performance solid-state on-chip supercapacitors based on Si nanowires coated with ruthenium oxide via atomic layer deposition. J. Power Sources 2017, 341, 1–10. [Google Scholar] [CrossRef]
  23. Dubrovskii, V.; Sibirev, N.; Harmand, J.; Glas, F. Growth kinetics and crystal structure of semiconductor nanowires. Phys. Rev. B Condens. Matter Mater. Phys. 2008, 78, 235301. [Google Scholar] [CrossRef]
  24. Artoni, P.; Pecora, E.F.; Irrera, A.; Priolo, F. Kinetics of Si and Ge nanowires growth through electron beam evaporation. Nanoscale Res. Lett. 2011, 6, 162. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, B.; Tang, H.; Zhang, N.; Sun, Q. Amorphous Nickel Boride Deposited on Silicon Nanowires and Carbon Nanowall Templates for High-Performance Micro-supercapacitors. J. Mater. Eng. Perform. 2024, 33, 2268–2278. [Google Scholar] [CrossRef]
  26. Bencheikh, Y.; Harnois, M.; Jijie, R.; Addad, A.; Roussel, P.; Szunerits, S.; Hadjersi, T.; Abaidia, S.E.H.; Boukherroub, R. High performance silicon nanowires/ruthenium nanoparticles micro-supercapacitors. Electrochim. Acta 2019, 311, 150–159. [Google Scholar] [CrossRef]
  27. Aradilla, D.; Gaboriau, D.; Bidan, G.; Gentile, P.; Boniface, M.; Dubal, D.; Gómez-Romero, P.; Wimberg, J.; Schubert, T.J.; Sadki, S. An innovative 3-D nanoforest heterostructure made of polypyrrole coated silicon nanotrees for new high performance hybrid micro-supercapacitors. J. Mater. Chem. A 2015, 3, 13978–13985. [Google Scholar] [CrossRef]
  28. Guerra, A.; Haye, E.; Achour, A.; Harnois, M.; Hadjersi, T.; Colomer, J.-F.; Pireaux, J.-J.; Lucas, S.; Boukherroub, R. High performance of 3D silicon nanowires array@CrN for electrochemical capacitors. Nanotechnology 2019, 31, 035407. [Google Scholar] [CrossRef] [PubMed]
  29. Lé, T.; Bidan, G.; Gentile, P.; Billon, F.; Debiemme-Chouvy, C.; Perrot, H.; Sel, O.; Aradilla, D. Understanding the energy storage mechanisms of poly (3, 4-ethylenedioxythiophene)-coated silicon nanowires by electrochemical quartz crystal microbalance. Mater. Lett. 2019, 240, 59–61. [Google Scholar] [CrossRef]
  30. Dubal, D.P.; Aradilla, D.; Bidan, G.; Gentile, P.; Schubert, T.J.; Wimberg, J.; Sadki, S.; Gomez-Romero, P. 3D hierarchical assembly of ultrathin MnO2 nanoflakes on silicon nanowires for high performance micro-supercapacitors in Li-doped ionic liquid. Sci. Rep. 2015, 5, 9771. [Google Scholar] [CrossRef]
  31. Gaboriau, D.; Aradilla, D.; Brachet, M.; Le Bideau, J.; Brousse, T.; Bidan, G.; Gentile, P.; Sadki, S. Silicon nanowires and nanotrees: Elaboration and optimization of new 3D architectures for high performance on-chip supercapacitors. RSC Adv. 2016, 6, 81017–81027. [Google Scholar] [CrossRef]
  32. Gaboriau, D.; Boniface, M.; Valero, A.; Aldakov, D.; Brousse, T.; Gentile, P.; Sadki, S. Atomic layer deposition alumina-passivated silicon nanowires: Probing the transition from electrochemical double-layer capacitor to electrolytic capacitor. ACS Appl. Mater. Interfaces 2017, 9, 13761–13769. [Google Scholar] [CrossRef] [PubMed]
  33. Soam, A.; Parida, K.; Kumar, R.; Dusane, R.O. Silicon-MnO2 core-shell nanowires as electrodes for micro-supercapacitor application. Ceram. Int. 2019, 45, 18914–18923. [Google Scholar] [CrossRef]
  34. Lu, P.; Halvorsen, E.; Ohlckers, P.; Müller, L.; Leopold, S.; Hoffmann, M.; Grigoras, K.; Ahopelto, J.; Prunnila, M.; Chen, X. Ternary composite Si/TiN/MnO2 taper nanorod array for on-chip supercapacitor. Electrochim. Acta 2017, 248, 397–408. [Google Scholar] [CrossRef]
  35. Lu, P.; Müller, L.; Hoffmann, M.; Chen, X. Taper silicon nano-scaffold regulated compact integration of 1D nanocarbons for improved on-chip supercapacitor. Nano Energy 2017, 41, 618–625. [Google Scholar] [CrossRef]
  36. Soam, A.; Arya, N.; Singh, A.; Dusane, R. Fabrication of silicon nanowires based on-chip micro-supercapacitor. Chem. Phys. Lett. 2017, 678, 46–50. [Google Scholar] [CrossRef]
  37. Gogotsi, Y.; Penner, R.M. Energy storage in nanomaterials–capacitive, pseudocapacitive, or battery-like? ACS Nano 2018, 12, 2081–2083. [Google Scholar] [CrossRef]
  38. Zhao, J.; Burke, A.F. Electrochemical capacitors: Materials, technologies and performance. Energy Storage Mater. 2021, 36, 31–55. [Google Scholar] [CrossRef]
  39. Kissinger, P.T.; Heineman, W.R. Cyclic voltammetry. J. Chem. Educ. 1983, 60, 702. [Google Scholar] [CrossRef]
  40. Pletcher, D.; Greff, R.; Peat, R.; Peter, L.; Robinson, J. Instrumental Methods in Electrochemistry; Elsevier: Amsterdam, The Netherlands, 2001. [Google Scholar]
  41. Zanello, P.; Nervi, C.; De Biani, F.F. Inorganic Electrochemistry: Theory, Practice and Application; Royal Society of Chemistry: London, UK, 2019. [Google Scholar]
  42. Mai, L.-Q.; Minhas-Khan, A.; Tian, X.; Hercule, K.M.; Zhao, Y.-L.; Lin, X.; Xu, X. Synergistic interaction between redox-active electrolyte and binder-free functionalized carbon for ultrahigh supercapacitor performance. Nat. Commun. 2013, 4, 2923. [Google Scholar] [CrossRef]
  43. Chen, G.Z. Understanding supercapacitors based on nano-hybrid materials with interfacial conjugation. Prog. Nat. Sci. Mater. Int. 2013, 23, 245–255. [Google Scholar] [CrossRef]
  44. Huang, J.; Gao, Y.; Luo, J.; Wang, S.; Li, C.; Chen, S.; Zhang, J. Editors’ choice—Review—Impedance response of porous electrodes: Theoretical framework, physical models and applications. J. Electrochem. Soc. 2020, 167, 166503. [Google Scholar] [CrossRef]
  45. Mei, B.-A.; Munteshari, O.; Lau, J.; Dunn, B.; Pilon, L. Physical interpretations of Nyquist plots for EDLC electrodes and devices. J. Phys. Chem. C 2018, 122, 194–206. [Google Scholar] [CrossRef]
  46. Perdana, M.Y.; Johan, B.A.; Abdallah, M.; Hossain, M.E.; Aziz, M.A.; Baroud, T.N.; Drmosh, Q.A. Understanding the behavior of supercapacitor materials via electrochemical impedance spectroscopy: A review. Chem. Rec. 2024, 24, e202400007. [Google Scholar] [CrossRef]
  47. Rowlands, S.; Latham, R.; Schlindwein, W. Supercapacitor devices using porous silicon electrodes. Ionics 1999, 5, 144–149. [Google Scholar] [CrossRef]
  48. Devarapalli, R.R.; Szunerits, S.; Coffinier, Y.; Shelke, M.V.; Boukherroub, R. Glucose-derived porous carbon-coated silicon nanowires as efficient electrodes for aqueous micro-supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 4298–4302. [Google Scholar] [CrossRef]
  49. Alper, J.P.; Wang, S.; Rossi, F.; Salviati, G.; Yiu, N.; Carraro, C.; Maboudian, R. Selective ultrathin carbon sheath on porous silicon nanowires: Materials for extremely high energy density planar micro-supercapacitors. Nano Lett. 2014, 14, 1843–1847. [Google Scholar] [CrossRef] [PubMed]
  50. Thissandier, F.; Gentile, P.; Pauc, N.; Hadji, E.; Le Comte, A.; Crosnier, O.; Bidan, G.; Sadki, S.; Brousse, T. Highly N-doped silicon nanowires as a possible alternative to carbon for on-chip electrochemical capacitors. Electrochemistry 2013, 81, 777–782. [Google Scholar] [CrossRef]
  51. Alper, J.P.; Vincent, M.; Carraro, C.; Maboudian, R. Silicon carbide coated silicon nanowires as robust electrode material for aqueous micro-supercapacitor. Appl. Phys. Lett. 2012, 100, 163901. [Google Scholar] [CrossRef]
  52. Zhou, Q.; Zhou, Y.; Bao, M.; Ni, X. Modified silicon nanowires@ polypyrrole core-shell nanostructures by poly (3, 4-ethylenedioxythiophene) for high performance on-chip micro-supercapacitors. Appl. Surf. Sci. 2019, 487, 236–243. [Google Scholar] [CrossRef]
  53. Shen, X.; Wang, T.; Wei, X.; Li, S. Facile synthesis of metal oxide and conductive polymers around silicon nanowire arrays for a high-performance aqueous supercapacitor. ACS Appl. Energy Mater. 2022, 5, 2596–2605. [Google Scholar] [CrossRef]
  54. Liu, H.; Chen, N.; Umar, A.; Li, H.; Li, S.; Bai, P.; Frans de Rooij, N.; Wang, Y.; Zhou, G. In situ construction of the coral-like polyaniline on the aligned silicon nanowire arrays for silicon substrate on-chip supercapacitors. ACS Appl. Energy Mater. 2020, 3, 11792–11802. [Google Scholar] [CrossRef]
  55. Ortaboy, S.; Alper, J.P.; Rossi, F.; Bertoni, G.; Salviati, G.; Carraro, C.; Maboudian, R. MnOx-decorated carbonized porous silicon nanowire electrodes for high performance supercapacitors. Energy Environ. Sci. 2017, 10, 1505–1516. [Google Scholar] [CrossRef]
  56. Sedlovets, D.M.; Naumov, A.P.; Korotitsky, V.I.; Starkov, V.V. Nanoporous Silicon with Graphene-like Coating for Pseudocapacitor Application. Nanomaterials 2022, 12, 2191. [Google Scholar] [CrossRef]
  57. Esarev, I.V.; Agafonov, D.V.; Surovikin, Y.V.; Nesov, S.N.; Lavrenov, A.V. On the causes of non-linearity of galvanostatic charge curves of electrical double layer capacitors. Electrochim. Acta 2021, 390, 138896. [Google Scholar] [CrossRef]
  58. Aradilla, D.; Gao, F.; Lewes-Malandrakis, G.; Müller-Sebert, W.; Gentile, P.; Boniface, M.; Aldakov, D.; Iliev, B.; Schubert, T.J.; Nebel, C.E. Designing 3D multihierarchical heteronanostructures for high-performance on-chip hybrid supercapacitors: Poly (3, 4-(ethylenedioxy) thiophene)-coated diamond/silicon nanowire electrodes in an aprotic ionic liquid. ACS Appl. Mater. Interfaces 2016, 8, 18069–18077. [Google Scholar] [CrossRef]
  59. Liu, P.; Wei, M.; Shen, X. MnOx Embedded in Silicon Nanowire Arrays for High-Performance Supercapacitors. J. Inorg. Organomet. Polym. Mater. 2024, 35, 2022–2032. [Google Scholar] [CrossRef]
  60. Shen, X.; Wei, X.; Wang, T.; Li, S.; Li, H. Solution-processable hierarchical SiNW/PEDOT/MnOx electrodes for high-performance supercapacitors. Mater. Chem. Front. 2022, 6, 2894–2904. [Google Scholar] [CrossRef]
  61. Valero, A.; Mery, A.; Gaboriau, D.; Gentile, P.; Sadki, S. One step deposition of PEDOT–PSS on ALD protected silicon nanowires: Toward ultrarobust aqueous microsupercapacitors. ACS Appl. Energy Mater. 2018, 2, 436–447. [Google Scholar] [CrossRef]
  62. Desplobain, S.; Gautier, G.; Semai, J.; Ventura, L.; Roy, M. Investigations on porous silicon as electrode material in electrochemical capacitors. Phys. Status Solidi C 2007, 4, 2180–2184. [Google Scholar] [CrossRef]
  63. Ferreira, N.; Azevedo, A.; Beloto, A.; Amaral, M.; Almeida, F.; Oliveira, F.; Silva, R. Nanodiamond films growth on porous silicon substrates for electrochemical applications. Diam. Relat. Mater. 2005, 14, 441–445. [Google Scholar] [CrossRef]
  64. Sedlovets, D.M.; Starkov, V.V.; Ulianova, V.V. N-doped graphene-like coating for improved microcapacitance of nanoporous silicon. Mater. Sci. Semicond. Process. 2025, 198, 109741. [Google Scholar] [CrossRef]
  65. Sedlovets, D.M.; Ulianova, V.V.; Starkov, V.V.; Knyazev, M.A. Photonic annealing effect on nanoporous silicon structures and their electrochemical capacitance. Ceram. Int. 2025, 51, 27184–27189. [Google Scholar] [CrossRef]
  66. Wu, T.-H.; Chang, C.-T.; Wang, C.-C.; Parwaiz, S.; Lai, C.-C.; Chen, Y.-Z.; Lu, S.-Y.; Chueh, Y.-L. Few-layer graphene sheet-passivated porous silicon toward excellent electrochemical double-layer supercapacitor electrode. Nanoscale Res. Lett. 2018, 13, 242. [Google Scholar] [CrossRef]
  67. Moulai, F.; Hadjersi, T.; Ifires, M.; Khen, A.; Rachedi, N. Enhancement of electrochemical capacitance of silicon nanowires arrays (SiNWs) by modification with manganese dioxide MnO2. Silicon 2019, 11, 2799–2810. [Google Scholar] [CrossRef]
  68. Gao, W.; Shao, M.; Yang, L.; Zhuo, S.; Ye, S.; Lee, S.-T. Manganese dioxide modified silicon nanowires and their excellent catalysis in the decomposition of methylene blue. Appl. Phys. Lett. 2012, 100, 063104. [Google Scholar] [CrossRef]
  69. Moulai, F.; Cherchour, N.; Messaoudi, B.; Zerroual, L. Electrosynthesis and characterization of nanostructured MnO2 deposited on stainless steel electrode: A comparative study with commercial EMD. Ionics 2017, 23, 453–460. [Google Scholar] [CrossRef]
  70. Julien, C.M.; Mauger, A. Nanostructured MnO2 as electrode materials for energy storage. Nanomaterials 2017, 7, 396. [Google Scholar] [CrossRef]
  71. Bélanger, D.; Brousse, T.; Long, J. Manganese oxides: Battery materials make the leap to electrochemical capacitors. Electrochem. Soc. Interface 2008, 17, 49. [Google Scholar] [CrossRef]
  72. Lee, S.-W.; Bak, S.-M.; Lee, C.-W.; Jaye, C.; Fischer, D.A.; Kim, B.-K.; Yang, X.-Q.; Nam, K.-W.; Kim, K.-B. Structural changes in reduced graphene oxide upon MnO2 deposition by the redox reaction between carbon and permanganate ions. J. Phys. Chem. C 2014, 118, 2834–2843. [Google Scholar] [CrossRef]
  73. Wang, Z.; Qin, Q.; Xu, W.; Yan, J.; Wu, Y. Long cyclic life in manganese oxide-based electrodes. ACS Appl. Mater. Interfaces 2016, 8, 18078–18088. [Google Scholar] [CrossRef]
  74. Liu, P.; Sun, S.; Wang, T.; Shen, X.; Zhu, M. Passivation of silicon nanowires with Ni particles and a PEDOT/MnOx composite for high-performance aqueous supercapacitors. Energy Adv. 2024, 3, 1111–1120. [Google Scholar] [CrossRef]
  75. Zhou, Q.; Bao, M.; Ni, X. A novel surface modification of silicon nanowires by polydopamine to prepare SiNWs/NC@NiO electrode for high-performance supercapacitor. Surf. Coat. Technol. 2021, 406, 126660. [Google Scholar] [CrossRef]
  76. Lu, F.; Qiu, M.; Qi, X.; Yang, L.; Yin, J.; Hao, G.; Feng, X.; Li, J.; Zhong, J. Electrochemical properties of high-power supercapacitors using ordered NiO coated Si nanowire array electrodes. Appl. Phys. A 2011, 104, 545–550. [Google Scholar] [CrossRef]
  77. Grigoras, K.; Keskinen, J.; Grönberg, L.; Ahopelto, J.; Prunnila, M. Coated porous Si for high performance on-chip supercapacitors. J. Phys. Conf. Ser. 2014, 557, 012058. [Google Scholar]
  78. Grigoras, K.; Keskinen, J.; Grönberg, L.; Yli-Rantala, E.; Laakso, S.; Välimäki, H.; Kauranen, P.; Ahopelto, J.; Prunnila, M. Conformal titanium nitride in a porous silicon matrix: A nanomaterial for in-chip supercapacitors. Nano Energy 2016, 26, 340–345. [Google Scholar] [CrossRef]
  79. Lu, P.; Ohlckers, P.; Müller, L.; Leopold, S.; Hoffmann, M.; Grigoras, K.; Ahopelto, J.; Prunnila, M.; Chen, X. Nano fabricated silicon nanorod array with titanium nitride coating for on-chip supercapacitors. Electrochem. Commun. 2016, 70, 51–55. [Google Scholar] [CrossRef]
  80. Garzon-Roman, A.; Cuate-Gomez, D. Graphene nanoflakes and carbon nanotubes on porous silicon layers by spin coating, for possible applications in optoelectronics. Sens. Actuators A Phys. 2019, 292, 121–128. [Google Scholar] [CrossRef]
  81. Olenych, I.B.; Aksimentyeva, O.I.; Monastyrskii, L.S.; Horbenko, Y.Y.; Partyka, M.V.; Luchechko, A.P.; Yarytska, L.I. Effect of graphene oxide on the properties of porous silicon. Nanoscale Res. Lett. 2016, 11, 43. [Google Scholar] [CrossRef]
  82. Moretta, R.; De Stefano, L.; Terracciano, M.; Rea, I. Porous silicon optical devices: Recent advances in biosensing applications. Sensors 2021, 21, 1336. [Google Scholar] [CrossRef]
  83. Shiraz, H.G. Efficient room temperature hydrogen gas sensing based on graphene oxide and decorated porous silicon. Int. J. Hydrog. Energy 2017, 42, 15966–15972. [Google Scholar] [CrossRef]
  84. Hussein, H.T.; Nayef, U.M.; Hussien, A.M.A. Synthesis of graphene on porous silicon for vapor organic sensor by using photoluminescence. Optik 2019, 180, 61–70. [Google Scholar] [CrossRef]
  85. Hussien, A.M.A.; Hussein, H.T.; Nayef, U.M.; Mahdi, M.H. Study of changing the intensity of photoluminescence spectra as a humidity sensor for graphene on porous silicon. Optik 2019, 193, 163015. [Google Scholar] [CrossRef]
  86. Ismail, R.A.; Kadhim, R.G.; Wasna’a, M.A. Effect of multiwalled carbon nanotubes incorporation on the performance of porous silicon photodetector. Optik 2016, 127, 8144–8152. [Google Scholar] [CrossRef]
  87. Aradilla, D.; Gao, F.; Lewes-Malandrakis, G.; Müller-Sebert, W.; Gaboriau, D.; Gentile, P.; Iliev, B.; Schubert, T.; Sadki, S.; Bidan, G. A step forward into hierarchically nanostructured materials for high performance micro-supercapacitors: Diamond-coated SiNW electrodes in protic ionic liquid electrolyte. Electrochem. Commun. 2016, 63, 34–38. [Google Scholar] [CrossRef]
  88. Rachedi, N.; Hadjersi, T.; Moulai, F.; Dokhane, N. Diamond-like carbon-coated silicon nanowires as a supercapacitor electrode in an aqueous LiClO4 electrolyte. Silicon 2022, 14, 2533–2543. [Google Scholar] [CrossRef]
  89. Bolotov, V.; Nesov, S.; Ponomareva, I.; Knyazev, E.; Ivlev, K.; Stenkin, Y.A.; Roslikov, V. The formation of nanocomposites carbon nanotubes/porous silicon for supercapacitor electrodes. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2020. [Google Scholar]
  90. Romanitan, C.; Varasteanu, P.; Mihalache, I.; Culita, D.; Somacescu, S.; Pascu, R.; Tanasa, E.; Eremia, S.A.; Boldeiu, A.; Simion, M. High-performance solid state supercapacitors assembling graphene interconnected networks in porous silicon electrode by electrochemical methods using 2, 6-dihydroxynaphthalen. Sci. Rep. 2018, 8, 9654. [Google Scholar] [CrossRef]
  91. Van Toan, N.; Tuoi, T.T.K.; 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]
  92. Aradilla, D.; Bidan, G.; Gentile, P.; Weathers, P.; Thissandier, F.; Ruiz, V.; Gómez-Romero, P.; Schubert, T.J.; Sahin, H.; Sadki, S. Novel hybrid micro-supercapacitor based on conducting polymer coated silicon nanowires for electrochemical energy storage. Rsc Adv. 2014, 4, 26462–26467. [Google Scholar] [CrossRef]
  93. Shen, X.; Ma, B.; Chen, L.; Zhao, J. High efficiency conjugated polymer/Si hybrid solar cells with tetramethylammonium hydroxide treatment. RSC Adv. 2017, 7, 688–693. [Google Scholar] [CrossRef]
  94. Shen, X.; Wei, X.; Wang, T.; Li, S.; Li, H. Polypyrrole embedded in nickel-cobalt sulfide nanosheets grown on nickel particles passivated silicon nanowire arrays for high-performance supercapacitors. Chem. Eng. J. 2023, 461, 141745. [Google Scholar] [CrossRef]
  95. Shen, X.; Sun, S.; Wei, X. Co-electrodeposition hybrid reduced graphene oxide and nickel-cobalt selenide nanosheets on silicon nanowires for high-performance supercapacitor. J. Energy Storage 2023, 72, 108770. [Google Scholar] [CrossRef]
  96. Behera, A.K.; Lakshmanan, C.; Viswanath, R.; Poddar, C.; Mathews, T. Electrochemical studies on wafer-scale synthesized silicon nanowalls for supercapacitor application. Bull. Mater. Sci. 2020, 43, 291. [Google Scholar] [CrossRef]
  97. Berton, N.; Brachet, M.; Thissandier, F.; Le Bideau, J.; Gentile, P.; Bidan, G.; Brousse, T.; Sadki, S. Wide-voltage-window silicon nanowire electrodes for micro-supercapacitors via electrochemical surface oxidation in ionic liquid electrolyte. Electrochem. Commun. 2014, 41, 31–34. [Google Scholar] [CrossRef]
  98. Gao, F.; Lewes-Malandrakis, G.; Wolfer, M.T.; Müller-Sebert, W.; Gentile, P.; Aradilla, D.; Schubert, T.; Nebel, C.E. Diamond-coated silicon wires for supercapacitor applications in ionic liquids. Diam. Relat. Mater. 2015, 51, 1–6. [Google Scholar] [CrossRef]
  99. Lé, T.; Gentile, P.; Bidan, G.; Aradilla, D. New electrolyte mixture of propylene carbonate and butyltrimethylammonium bis (trifluoromethylsulfonyl) imide (n1114 tfsi) for high performance silicon nanowire (sinw)-based supercapacitor applications. Electrochim. Acta 2017, 254, 368–374. [Google Scholar] [CrossRef]
  100. Leonardi, A.A.; Arrigo, A.; Lo Faro, M.J.; Nastasi, F.; Irrera, A. 2D Fractal Arrays of Ultrathin Silicon Nanowires as Cost-Effective and High-Performance Substrate for Supercapacitors. Adv. Energy Sustain. Res. 2024, 5, 2400080. [Google Scholar] [CrossRef]
  101. Han, H.-C.; Chong, C.-W.; Wang, S.-B.; Heh, D.; Tseng, C.-A.; Huang, Y.-F.; Chattopadhyay, S.; Chen, K.-H.; Lin, C.-F.; Lee, J.-H. High K nanophase zinc oxide on biomimetic silicon nanotip array as supercapacitors. Nano Lett. 2013, 13, 1422–1428. [Google Scholar] [CrossRef]
  102. Burchak, O.; Keller, C.; Lapertot, G.; Salaün, M.; Danet, J.; Chen, Y.; Bendiab, N.; Pépin-Donat, B.; Lombard, C.; Faure-Vincent, J. Scalable chemical synthesis of doped silicon nanowires for energy applications. Nanoscale 2019, 11, 22504–22514. [Google Scholar] [CrossRef]
  103. Thamri, S.; Raouadi, M. Improved capacitance of NiO and nanoporous silicon electrodes for micro-supercapacitor application. J. Mater. Sci. Mater. Eng. 2025, 20, 1. [Google Scholar] [CrossRef]
  104. Bakardjieva, S.; Bezdička, P.; Grygar, T.; Vorm, P. Reductive dissolution of microparticulate manganese oxides. J. Solid State Electrochem. 2000, 4, 306–313. [Google Scholar] [CrossRef]
  105. 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]
  106. 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]
  107. Wu, Z.-S.; Feng, X.; Cheng, H.-M. Recent advances in graphene-based planar micro-supercapacitors for on-chip energy storage. Natl. Sci. Rev. 2014, 1, 277–292. [Google Scholar] [CrossRef]
  108. Su, Y.; Fan, X.; Xiong, Q.; Xiong, C. Multi-dimensional optimization and diversified applications of carbon fiber-based flexible electrode materials–synergistic design of cross-type supercapacitors and electrolyte systems. J. Alloys Compd. 2025, 1038, 182768. [Google Scholar] [CrossRef]
  109. Wang, X.; Zhang, Q. On-chip microsupercapacitors: From material to fabrication. Energy Technol. 2019, 7, 1900820. [Google Scholar] [CrossRef]
  110. Hu, M.; Zhang, H.; Hu, T.; Fan, B.; Wang, X.; Li, Z. Emerging 2D MXenes for supercapacitors: Status, challenges and prospects. Chem. Soc. Rev. 2020, 49, 6666–6693. [Google Scholar] [CrossRef]
  111. Rizk, R.; De Mierry, P.; Ballutaud, D.; Aucouturier, M.; Mathiot, D. Hydrogen diffusion and passivation processes in p-and n-type crystalline silicon. Phys. Rev. B 1991, 44, 6141. [Google Scholar] [CrossRef] [PubMed]
  112. Dinh, K.H.; Roussel, P.; Lethien, C. Advances on microsupercapacitors: Real fast miniaturized devices toward technological dreams for powering embedded electronics? ACS Omega 2023, 8, 8977–8990. [Google Scholar] [CrossRef] [PubMed]
  113. Jolayemi, B.; Buvat, G.; Roussel, P.; Lethien, C. Emerging capacitive materials for on-chip electronics energy storage technologies. Batteries 2024, 10, 65. [Google Scholar] [CrossRef]
  114. Dimitriou, E.; Michailidis, N. Printable conductive inks used for the fabrication of electronics: An overview. Nanotechnology 2021, 32, 502009. [Google Scholar] [CrossRef]
  115. León-Silva, U.; Cajero-Sotelo, L.; Nicho, M.; Antunez, E.E.; Escobedo-Alatorre, J.; Sandoval-Espino, J.; Marban-Salgado, J.; Diaz-Guillen, M. Biopolymeric separator for capacitors based on porous silicon. J. Phys. Chem. Solids 2025, 200, 112597. [Google Scholar] [CrossRef]
  116. Nam, J.; Lee, H.; Chae, O.B. Overcoming Challenges in Silicon Anodes: The Role of Electrolyte Additives and Solid-State Electrolytes. Micromachines 2025, 16, 800. [Google Scholar] [CrossRef] [PubMed]
  117. Li, Y.; Ni, Z.; Geng, J.; Wang, Z.; Li, Y.; Zhao, Y.; Shao, H.; Li, Y.; Xiong, S.; Feng, J. Advancements in Electrolytes: From Liquid to Solid for Low-Cost and High-Energy-Density Micro-Sized Silicon-Based Batteries. Adv. Energy Mater. 2025, 15, 2502284. [Google Scholar] [CrossRef]
Figure 1. (a) Capacitance values of pSi and SiNWs in chronological order. (b) Bibliometric analysis of MACE compared to VLS. (c) Capacitance values of SiNWs obtained by MACE and VLS in increasing order of magnitude. (a,c) The data with reference numbers are given in the Supplementary Materials (SM), Section S1. (b) Ref. [20] is reproduced with permission, Copyright © 2021 by the authors.
Figure 1. (a) Capacitance values of pSi and SiNWs in chronological order. (b) Bibliometric analysis of MACE compared to VLS. (c) Capacitance values of SiNWs obtained by MACE and VLS in increasing order of magnitude. (a,c) The data with reference numbers are given in the Supplementary Materials (SM), Section S1. (b) Ref. [20] is reproduced with permission, Copyright © 2021 by the authors.
Nanomaterials 15 01826 g001
Figure 2. The equivalent electrical circuit of a cylindrical pore with uniformly distributed electrolyte resistance Re, and the specific impedance of the interface between the electronic phase and the electrolyte phase, Zdl. Ref. [44] is reproduced with permission, Copyright © 2020 by the authors.
Figure 2. The equivalent electrical circuit of a cylindrical pore with uniformly distributed electrolyte resistance Re, and the specific impedance of the interface between the electronic phase and the electrolyte phase, Zdl. Ref. [44] is reproduced with permission, Copyright © 2020 by the authors.
Nanomaterials 15 01826 g002
Figure 3. Schematic of typical Nyquist plots for EDL capacitor electrodes or devices. Ref. [45] is reproduced with permission, Copyright © 2018, American Chemical Society.
Figure 3. Schematic of typical Nyquist plots for EDL capacitor electrodes or devices. Ref. [45] is reproduced with permission, Copyright © 2018, American Chemical Society.
Nanomaterials 15 01826 g003
Figure 4. Area-normalized and volumetric capacitances depending on pSi depth/SiNW length. The data from (a) ref. [49]; (b) ref. [50]; (c) ref. [51]; and (d) ref. [48].
Figure 4. Area-normalized and volumetric capacitances depending on pSi depth/SiNW length. The data from (a) ref. [49]; (b) ref. [50]; (c) ref. [51]; and (d) ref. [48].
Nanomaterials 15 01826 g004
Figure 5. Characteristic scanning electron microscope (SEM) images of (a,b) VLS-grown disorderly interwoven SiNWs and (c,d) vertically aligned SiNWs obtained by MACE. (a) Ref. [58] is reproduced with permission, Copyright © 2016, American Chemical Society; (b) ref. [30] is reproduced with permission, Copyright © 2015, The Author(s); (c) ref. [49] is reproduced with permission, Copyright © 2014, American Chemical Society; (d) ref. [54] is reproduced with permission, Copyright © 2020, American Chemical Society.
Figure 5. Characteristic scanning electron microscope (SEM) images of (a,b) VLS-grown disorderly interwoven SiNWs and (c,d) vertically aligned SiNWs obtained by MACE. (a) Ref. [58] is reproduced with permission, Copyright © 2016, American Chemical Society; (b) ref. [30] is reproduced with permission, Copyright © 2015, The Author(s); (c) ref. [49] is reproduced with permission, Copyright © 2014, American Chemical Society; (d) ref. [54] is reproduced with permission, Copyright © 2020, American Chemical Society.
Nanomaterials 15 01826 g005
Figure 6. SEM images of (a) pSi (145 mF/cm2 capacitance with N-doped carbon coating) and (b) SiNWs (381 mF/cm2 capacitance after coating with carbon and MnOx). (a) ref. [65] is reproduced with permission, Copyright © 2025 Elsevier Ltd. and Techna Group S.r.l.; (b) ref. [49] is reproduced with permission, Copyright © 2014, American Chemical Society.
Figure 6. SEM images of (a) pSi (145 mF/cm2 capacitance with N-doped carbon coating) and (b) SiNWs (381 mF/cm2 capacitance after coating with carbon and MnOx). (a) ref. [65] is reproduced with permission, Copyright © 2025 Elsevier Ltd. and Techna Group S.r.l.; (b) ref. [49] is reproduced with permission, Copyright © 2014, American Chemical Society.
Nanomaterials 15 01826 g006
Figure 7. Cross-sectional view of PEDOT/diamond/SiNWs. Scale bar of inset: 100 nm. Ref. [58] is reproduced with permission, Copyright © 2016, American Chemical Society.
Figure 7. Cross-sectional view of PEDOT/diamond/SiNWs. Scale bar of inset: 100 nm. Ref. [58] is reproduced with permission, Copyright © 2016, American Chemical Society.
Nanomaterials 15 01826 g007
Figure 8. SEM cross-section images of (a) pristine Al2O3/SiNWs and (b) gelatinous-like PEDOT-PSS/Al2O3/SiNW composite. (b) Ref. [61] is reproduced with permission, Copyright © 2019, American Chemical Society.
Figure 8. SEM cross-section images of (a) pristine Al2O3/SiNWs and (b) gelatinous-like PEDOT-PSS/Al2O3/SiNW composite. (b) Ref. [61] is reproduced with permission, Copyright © 2019, American Chemical Society.
Nanomaterials 15 01826 g008
Figure 9. SEM images of (a) SiNWs and (b) SiNWs/PEDOT + MnO2. (c) TEM image of SiNWs/PEDOT + MnO2. SEM images of (d) SiNWs/PEDOT + MnO2/rGO and (e) SiNWs/PEDOT + MnO2/rGO/AgNWs:PEDOT-PSS. (f) Cross-sectional SEM image of SiNWs/PEDOT + MnO2/rGO/AgNWs:PEDOT-PSS. Ref. [53] is reproduced with permission, Copyright © 2022, American Chemical Society.
Figure 9. SEM images of (a) SiNWs and (b) SiNWs/PEDOT + MnO2. (c) TEM image of SiNWs/PEDOT + MnO2. SEM images of (d) SiNWs/PEDOT + MnO2/rGO and (e) SiNWs/PEDOT + MnO2/rGO/AgNWs:PEDOT-PSS. (f) Cross-sectional SEM image of SiNWs/PEDOT + MnO2/rGO/AgNWs:PEDOT-PSS. Ref. [53] is reproduced with permission, Copyright © 2022, American Chemical Society.
Nanomaterials 15 01826 g009
Figure 10. Schematic illustration of the SiNWs/Ni/PEDOT + MnO2/Pt electrode. Ref. [74] is reproduced with permission, Copyright © 2022, American Chemical Society.
Figure 10. Schematic illustration of the SiNWs/Ni/PEDOT + MnO2/Pt electrode. Ref. [74] is reproduced with permission, Copyright © 2022, American Chemical Society.
Nanomaterials 15 01826 g010
Figure 11. (a) Fabrication steps of the in-chip supercapacitor. (b) Schematic of the symmetric solid-state on-chip supercapacitors based on PANI/SiNW electrodes. (a) Ref. [78] is reproduced with permission, Copyright © 2016 The Authors. (b) Ref. [54] is reproduced with permission, Copyright © 2020, American Chemical Society.
Figure 11. (a) Fabrication steps of the in-chip supercapacitor. (b) Schematic of the symmetric solid-state on-chip supercapacitors based on PANI/SiNW electrodes. (a) Ref. [78] is reproduced with permission, Copyright © 2016 The Authors. (b) Ref. [54] is reproduced with permission, Copyright © 2020, American Chemical Society.
Nanomaterials 15 01826 g011
Figure 12. MSC based on interdigitated topologies with a symmetric configuration or an asymmetric configuration. Ref. [112] is reproduced with permission, Copyright © 2023, The Authors.
Figure 12. MSC based on interdigitated topologies with a symmetric configuration or an asymmetric configuration. Ref. [112] is reproduced with permission, Copyright © 2023, The Authors.
Nanomaterials 15 01826 g012
Table 1. Capacitance data, measurement conditions, and electrode materials extracted from various works.
Table 1. Capacitance data, measurement conditions, and electrode materials extracted from various works.
Ref.C, mF/cm2MeasuredMatrix (Method)Coating (Method)ElectrolyteCap. Retention
%Cycles
[36]0.013at 10 mV/sSiNWs (VLS)EMIM-TFSI *981k
[96]0.021at 10 mV/s11.9 μm SiNWs (MACE)1 M TEABF4830.5k
[97]0.031at 0.25 mA/cm220 μm SiNWs (VLS)EMIM-TFSI
[32]0.038at 0.25 mA/cm25 μm SiNWs (VLS)3 nm Al2O3 (ALD)EMIM-TFSI961000k
[50]0.051at 5 μA/cm250 μm SiNWs (VLS)TEABF4 *97200k
[98]0.108at 5 mV/s5 μm SiNWs (VLS)diamond (CVD)PMPyrr-TFSI~9310k
[99]0.18at 2.2 mA/cm2SiNWs (VLS)N1114TFSI703000k
[47]0.2at 10 mV/spSi (AE)0.25 M TEABF4
[100]0.274at 50 mV/s2.7 μm SiNWs (MACE)0.1 M Li2SO4741k
[101]0.31 μm SiNWs (VLS)20 nm ZnO (ALD);
10 nm Al2O3(ALD)
[62]0.3243 μm pSi (AE)Au20% H2SO4
[102]0.75at 0.14 mA/cm2SiNW powder (VLS)0.5 M TBABF4801000k
[63]0.99at 100 mV/spSi (AE)nanodiamond (CVD)0.1 M KCl
[31]1.25at 1 mA/cm250 μm SiNTrs
(2-step VLS)
801000k
[87]1.5at 10 mA/cm250 μm SINWs (VLS)diamond (CVD)Et3NH-TFSI *651000k
[79]1.55at 2 mV/s20 μm SiNRs (DRIE)30 nm TiN (ALD)1 M Na2SO495.22k
[51]1.7at 50 mV/sSINWs (MACE)SiC (CVD)1 M KCl951k
[88]2at 0.01 mA/cm260 μm SINWs (MACE)DLC (EPD)0.5 M LiClO49016k
[33]2.1at 0.04 mA/cm2SINWs (VLS)MnO2 (EPD)1 M Na2SO4905k
[77]4.38 *6 μm pSi (AE)TiN (ALD)TEABF4stable5.5 k
[61]6.4at 20 mV/s50 μm SINWs (VLS)Al2O3 (ALD);
PEDOT-PSS
(drop casting)
0.5 M Na2SO495500 k 
[66]8.16at 1000 mV/s15 μm pSi (AE)FLG (Ni-assist CVD) 0.5 M Na2SO413010k
[58]8.5at 1 mA/cm250 μm SINWs (VLS)diamond (CVD);
PEDOT (ElP)
N1114TFSI8015k
[103]9.64at 1 mA/cm2pSi (AE)NiO (sol-gel)1 M NaOH975k
[30]13at 0.4 mA/cm250 μm SINWs (VLS)MnO2 (ELD)LiClO4-PMPyrrBTA915k
[27]14at 1 mA/cm2SiNTrs (VLS) Ppy (ElP)PYR13TFSI7010k
[29]17at 100 mV/s10 μm SINWs (VLS)PEDOT (ECD)TBABF4
[22]19at 5 mV/s6 μm SINWs (MAAE)RuO2 (ALD)1 M Na2SO49210k
[67]21.3at 1000 mV/s10 μm SINWs (MACE)MnO2 (ELD)1 M Na2SO4
[48]25.6at 0.1 mA/cm217 μm SINWs MACEnanocarbon
(glucose pyrolysis)
1 M Na2SO47525k
[90]30at 0.5 A/g 5 μm pSiGLC
(DHN pyrolysis)
PVA-H2SO4751k
[26]36.25at 1 mA/cm2SiNWs (VLS)Ru NPs (ELD)1 M Na2SO48025k
[34]81.6at 5 mV/s~5 μm SiTNR (DRIE)TiN (ALD);
MnO2 (ELD)
1 M Na2SO495.75k
[56]87at 5 mV/s80 μm pSi (AE)GLC (CVD)3 M H2SO410015k
[54]95.8at 10 mV/s5.5 μm SINWs (MACE)PANI (OxP)1 M H2SO471.82k
[53]100.98at 1.5 mA/cm25.6 μm SINWs (MACE+TMAH)PEDOT + MnO2 (ECD); rGO (ELD); AgNWs + PEDOT-PSS (spin-coating)1 M Na2SO4812k
[52]106.1at 1 mA/cm217 μm SINWs (MACE)PPy (OxP);
PEDOT (PhP)
PYR13TFSI80.25k
[75]110at 1 mA/cm210 μm SINWs (MACE)N-carbon (PDOP pyrolysis);
NiO (ELD)
6 M KOH814k
[35]123
192
at 1000 mV/s
at 1 mV/s
~20 μm SiTNRs (DRIE)FC-CNT (CVD)H2SO41025k
[91]130at 10 mV/s10 μm SINWs (MACE)GNWs (CVD);
PANI (ElP)
PVA-H2SO4802k
[64]145at 5 mV/s80 μm pSi (AE)N-GLC (CVD)3M H2SO49320k
[25]165.7at 0.1 mA/cm2several μm SINWs (VLS)NiB
(ELD)
PVA-Na2SO49310k
[28]180
31.8
at 5 mV/s
at 1.6 mA/cm2
SiNWs (VLS)CrN (magnetron sputtering)0.5 M Na2SO49215k
[74]207.43at 1 mA/cm26 μm SiNWs (MACE+TMAH)Ni + PEDOT + MnO2 (ECD co-deposition); Pt NPs (ELD)1 M Na2SO4955k
[49]325at 1 mA/cm2120 μm SINWs (MACE)nanocarbon (CVD)EMIM-TSFI835k
[59]328.6at 1 mA/cm210 μm SINWs (MACE+TMAH)Ni (ECD);
MnO2 (ECD)
1 M Na2SO4797k
[60]352at 2 mA/cm2several μm SINWs (MACE + TMAH)Ni (ECD), PEDOT (ElP), Pt NPs (ELD); MnO2 (ELD)
AgNWs+PEDOT-PSS (spin-coating)
 852k
[55]381at 4 mA/cm280 μm SINWs (MACE)nanocarbon (CVD);
MnOx (ELD)
0.1 M EMIM-TSFI845k
[95]19730.87 A/cm210 μm SiNWs (MACE + TMAH)Ni (ECD);
NiCoSe-rGO (ECD)
6 M KOH80.52k
* 1-Ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide (EMIM-TFSI); tetraethylammonium tetrafluoroborate (TEABF4); triethylammonium bis(trifluoromethylsulfonyl)imide (Et3NH-TFSI).
Table 2. Key coatings for pSi/SiNW electrodes and their features in the context of microcapacitor applications.
Table 2. Key coatings for pSi/SiNW electrodes and their features in the context of microcapacitor applications.
CoatingCharacteristicAdvantageDisadvantage and/or Expectations
NiB, CrNpseudocapacitivehigh capacitancethe capacitance of crystalline NiB needs to be explored
MnO2pseudocapacitiveELD is compatible with microelectronic technologymoderate capacitance
MnOx with underlayer (TiN or carbon)highly conductive underlayermentioned above + high or ultra-high capacitancelow cyclic stability
Carbonimproved charge transfer, good adhesionhigh capacitance, high cyclic stability (for pSi)CVD methods are poorly compatible with microelectronic technology
PEDOT-PSSgelationous-like structureultra-high long-term cyclicmoderate capacitance
PANI with GNW underlayersynergetic effect of both coatingshigh capacitancelow cyclic stability
In combined coatings:
Pt NPS, rGOimproved charge transferultra-high capacitancelong-term cyclic tests need to be carried out
NiCoSeporous structure facilitates ionic diffusion
MnOxcracks on the MnOx facilitate electrolyte penetration
Table 3. Device characteristics for two-electrode supercapacitors reported in some works.
Table 3. Device characteristics for two-electrode supercapacitors reported in some works.
Ref.Celectrode, mF/cm2DeviceCdevice, mF/cm2Edevice, mWh/cm2Pdevice, mW/cm2
[53]100.98 (at 1.5 mA/cm2)symmetric24.7 (at 1 mA/cm2)0.00342.652
[52]106.1 (at 1 mA/cm2)symmetric46.5 (at 0.5 mA/cm2)0.01460.375
[35]123 (at 1000 mV/s)
192 (at 1 mV/s)
symmetric178 (at 5 mV/s)0.0115
0.0096
2
34.7
[91]130 (at 10 mV/s)liquid state
solid state
84.4 (at 100 mV/s)0.0117
0.0108
0.42
0.782
[74]207.43 (at 1 mA/cm2)symmetric64 (at 1 mA/cm2)0.2503
1.5115
0.0103
0.006
[59]328.6 (at 1 mA/cm2)asymmetric95 (at 1 mA/cm2)0.021
0.00835
0.7998
7.9446
[55]381 (at 4 mA/cm2)symmetric49 (at 2 mA/cm2)0.146 *
0.0204 *
0.128 *
14.4 *
[95]1973 (0.87 A/cm2)asymmetric273 * (at 2.06 mA/cm2)0.109 *
0.034 *
1.6 *
16.4 *
* Calculated from the gravimetric data given in the corresponding papers (See details in the Supplementary Materials, Section S4).
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

Sedlovets, D.M. Porous Silicon and Silicon Nanowires for On-Chip Supercapacitor Electrodes: A Review. Nanomaterials 2025, 15, 1826. https://doi.org/10.3390/nano15231826

AMA Style

Sedlovets DM. Porous Silicon and Silicon Nanowires for On-Chip Supercapacitor Electrodes: A Review. Nanomaterials. 2025; 15(23):1826. https://doi.org/10.3390/nano15231826

Chicago/Turabian Style

Sedlovets, Daria M. 2025. "Porous Silicon and Silicon Nanowires for On-Chip Supercapacitor Electrodes: A Review" Nanomaterials 15, no. 23: 1826. https://doi.org/10.3390/nano15231826

APA Style

Sedlovets, D. M. (2025). Porous Silicon and Silicon Nanowires for On-Chip Supercapacitor Electrodes: A Review. Nanomaterials, 15(23), 1826. https://doi.org/10.3390/nano15231826

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