Graphical Abstract Figure
Graphical Abstract Figure
Close modal

Abstract

Electrocatalytically active titanium oxynitride (TiNO) thin films were fabricated on commercially available titanium metal plates using a pulsed laser deposition method for energy storage applications. The elemental composition and nature of bonding were analyzed using X-ray photoelectron spectroscopy (XPS) to reveal the reacting species and active sites responsible for the enhanced electrochemical performance of the TiNO electrodes. Symmetric supercapacitor devices were fabricated using two TiNO working electrodes separated by an ion-transporting layer to analyze their real-time performance. The galvanostatic charge–discharge studies on the symmetric cell have indicated that TiNO films deposited on the polycrystalline titanium plates at lower temperatures are superior to TiNO films deposited at higher temperatures in terms of storage characteristics. For example, TiNO films deposited at 300 °C exhibited the highest specific capacity of 69 mF/cm2 at 0.125 mA/cm2 with an energy density of 7.5 Wh/cm2. The performance of this supercapacitor (300 °C TiNO) device is also found to be ∼22% better compared to that of a 500 °C TiNO supercapacitor with a capacitance retention ability of 90% after 1000 cycles. The difference in the electrochemical storage and capacitance properties is attributed to the reduced leaching away of oxygen from the TiNO films by the Ti plate at lower deposition temperatures, leading to higher oxygen content in the TiNO films and, consequently, a high redox activity at the electrode/electrolyte interface.

1 Introduction

Due to the depletion of fossil resources and the rise in environmental degradation, producing clean energy has become a top concern for the world. Energy has a significant impact on our socioeconomic development and the quality of our lives [1]. Researchers have been motivated by this issue to create effective, sustainable, and clean energy solutions. The two most popular types of energy storage devices are batteries and capacitors. Both batteries and capacitors perform the same function of storing energy; the difference between the two devices comes from the way they perform this task. Battery stores and distributes energy slowly, while capacitors store and distribute energy instantaneously. During the discharge process in batteries, the chemical energy held is converted to electrical energy and discharged. Capacitors, on the other hand, use electrostatic attraction to store electrical energy, making them suitable for devices that require a high power density [2].

Supercapacitors have drawn a lot of attention from electrochemical energy storage devices as one of the alternative energy sources because of their quick charge–discharge speed, high specific capacitance, extended lifespan, safety, and dependability [36]. Based on the charge storage process, supercapacitors are divided into electrical double-layer capacitors (EDLCs) and pseudocapacitors (PCs) [7]. EDLCs store energy through the accumulation of electrostatic charges at the electrolyte/electrode interface due to a non-Faradaic process [8]. However, the energy storage mechanism in PCs incorporates quick redox events brought about by Faradaic reactions [9]. Carbonaceous materials such as activated carbons, graphene, and carbon nanotubes have been exploited as electrode materials for EDLCs due to their high surface area and porosity [4,10,11]. As opposed to that, electrochemical active materials such as nitrides [1214], sulfides [15,16], transition metal oxides (TMO) [1719], and oxynitrides [20,21] are employed as electrode materials for pseudocapacitor due to their energy densities and high capacitance than carbon materials. However, poor electrical conductivity and structure degradation of TMO as a consequence of Faradaic redox reactions eventually lead to low cyclic stability. A means to improve supercapacitor technology is made possible by novel electrode materials with distinctive physicochemical features. As an alternative to TMO, metal nitrides such as titanium nitride (TiN) are used as electrode materials due to their chemical stability and high thermal and electrical conductivity [2224]. Despite these advantages, metal nitrides exhibit poor cycling rate capabilities, low charge storage, and limited long-term durability [25].

Metal-based oxynitrides have recently emerged as promising intermediate compounds between insulating metal oxides and conducting metallic nitrides. These oxynitrides possess a varied range of attractive properties: high electronic conductivity, corrosion resistance, and biocompatibility, which enable them to circumvent the difficulties related to metal oxides and nitrides [26,27]. Wang et al. synthesized titanium oxynitride nano grid film, which exhibits excellent capacitance retention and cyclability [28]. Yu et al. have also reported outstanding cycling stability (∼105 cycles) exhibited by tungsten oxynitride fibers [29].

In this work, we have explored the electrochemical and supercapacitor properties of titanium-based oxynitride (TiNO) thin films synthesized using a pulsed laser deposition (PLD) method. The general formula of the TiNO homolog series is expressed as Ti(1x3)3+Ti(vac)x33+N(1x)3Ox2. The series exists in the whole range of 0x1 [30,31]. In this formula, Ti3+ (vac) is the number of Ti3+ vacancies in each formula unit, which is created to maintain the charge neutrality in the resulting compounds. This molecular formula takes into account the substitution of trivalent N anions by bivalent O anions in an ionic TiN lattice and the maintenance of charge neutrality in the lattice. The formula is also built on the assumption that the valences of Ti, N, and O are maintained at +3, −3, and −2, respectively [31]. The terminal compounds (TiN with x = 0 and TiO with x = 1), as well as all the intermediate compounds with x between 0 and 1, all have rock salt structures, but they possess wide-ranging physicochemical properties [32,33]. One of the terminal compounds of the family, TiN (x = 0), has been known for its well-established properties for decades [19,3438]. The TiNO material system has several favorable properties and lends itself to a variety of applications, making it a suitable electrode material candidate [3942].

To investigate the electrochemical performance of the TiNO film for supercapacitor applications, we have carried out three and two electrode measurements. A symmetric cell is fabricated to evaluate the cyclic stability, energy, and power density for its real-time application.

2 Experimental Method and Materials

TiNO thin films were grown on 10 mm × 15 mm × 1 mm titanium plates using a PLD method. The schematic of the PLD method is shown in Fig. 1(a). In our PLD experiments, a high-purity (99.99%) TiN target was used to deposit the TiNO films by controlling the substrate temperature, oxygen pressure, laser energy density, and laser repetition rate. A KrF laser (Coherent Complex Pro, wavelength 248 nm, pulse duration 30 ns, repetition rate 10 Hz) was used. A fixed number of laser pulses (shots) of 10,000 (deposition time = 1000 s) was used. We set the oxygen pressure at 50 mTorr, varied the substrate temperature from 300 °C to 500 °C, and the laser energy density at 2 J/cm2. The surface morphology of the TiNO electrode before and after the electrocatalysis measurements was studied using a Hitachi SU8000 scanning electron microscopy (SEM). The electrical resistivity was determined using an Ossila T2001A standard four-probe measurement. The unit lattice models were simulated using the visualization for electronic and structural analysis. The film-substrate orientation was analyzed using an X-ray diffractometer (Bruker D8) with a CuKα source (λave = 0.154 nm). XPS measurements were carried out using a Thermo Escalab Xi+ (XPS). Raman spectroscopy measurements were carried out using a WiTec alpha300R Confocal Raman microscope, specifically at the laser excitation wavelength of 532 nm.

Fig. 1
(a) Schematic of a PLD method for thin film deposition and (b) schematic representation of a symmetric cell used in the present study for the supercapacitor properties measurements [43]. The Ti plates on both sides of the electrolyte-soaked separator are coated with electrically conducting TiNO films.
Fig. 1
(a) Schematic of a PLD method for thin film deposition and (b) schematic representation of a symmetric cell used in the present study for the supercapacitor properties measurements [43]. The Ti plates on both sides of the electrolyte-soaked separator are coated with electrically conducting TiNO films.
Close modal

The electrochemical activities of the TiNO films and supercapacitor cells were performed using three-electrode and two-electrode techniques, respectively. The electrolytic solution used for these experiments was an aqueous 1 M KOH solution. For the three-electrode measurement, an Hg/HgO electrode is used as a reference electrode, platinum wire is used as the counter electrode, and TiNO is used as the working electrode. The supercapacitor device was fabricated using two-working electrodes separated by an ion-transporting layer (chromatography paper) (Fig. 1(b)). Before testing, the symmetric cell was soaked in the 1 M KOH electrolyte for 1 h. A symmetric supercapacitor refers to similar supercapacitive electrodes. In the present study, Ti plates coated with TiNO films are used as electrodes on the two sides of the ion-transporting layer [43]. The charge storage capacity of the electrode was studied using cyclic voltammetry (CV) and galvanostatic charge–discharge methods. Electrochemical impedance spectroscopy (EIS) measurement was carried out by applying an AC voltage in a frequency range from 100 kHz to 100 µHz at open circuit potential. Electrochemical measurements were performed on VSP-3e Biologic EC-lab potentiostat. In the two-electrode cell, one electrode acts as a counter electrode, and the other electrode serves as the working electrode. The active electrode surface for both positive and negative electrodes was 0.75 cm2. A constant charging–discharging current was applied to record the charge/discharge curve of supercapacitors, and then the change of cell voltage with time was recorded. All the electrochemical data were analyzed after iR correction, using the solution resistance obtained from the EIS measurement at 0 V versus Hg/HgO.

3 Results and Discussion

Figure 2 shows the X-ray diffraction patterns of the TiNO thin films grown on a polycrystalline titanium (Ti) plate substrate at three different deposition temperatures. The thin films exhibited the characteristic peaks of the face-centered cubic TiNO phase with the appearance of the (111) plane at 36.7 deg on all three samples. All other peaks, indexed s, belong to the polycrystalline Ti substrate. A careful examination of the X-ray diffraction (XRD) patterns shows a slight shift in the (111) peak position with respect to the (111) plane in pure TiN (111), which is presented as a line pattern using the JCPDS data [44,45]. This shift in the peak position is attributed to the partial oxidation of TiN to TiNO. The lattice constant (a) of the TiNO films was calculated using the XRD d (111) values according to the formula 1/d2hkl = (h2 + k2 + l2)/a2, where (hkl) is the Miller indices of the plane [46]. The lattice constant values of TiNO films deposited at 300 °C, 400 °C, and 500 °C were found to be 4.21 Å, 4.23 Å, 4.24 Å with a ratio of a300: a400: a500: 1.000, 1.004, 1.007. The increase in the TiNO lattice constant with an increase in the deposition temperature is explained on the basis of reduced oxygen content in the films since the ionic radius of O2− (1.42 Å) is smaller than that of N3− (1.71 Å), the TiNO films with more oxygen will have a smaller lattice constant [47]. The higher oxygen content in TiNO films deposited at lower temperatures (i.e., more oxidation of TiN) is attributed to a reduced out-diffusion process from films at lower deposition temperatures [48]. The gettering effect of Ti due to its high oxygen-affinity also leads to the leaching away of oxygen from the TiNO films and the formation of a TiO2 layer on the Ti plate at higher temperatures [49,50]. The formation of TiNO takes place via partial oxidation of TiN in the presence of oxygen as the following reaction: TiN1/2(O2)TiNO1/2(O2)TiO2+12N2. The present method of oxidizing TiN to TiNO system via controlled oxidation during the PLD process has an energy advantage over the conventional method of nitridation of TiO2 to TiNO. The energy advantage comes from the higher activation energy of the nitridation process (672 KJ/mol) than that of the oxidation process (496 kJ/mol) [51]. The favorable energetics of an oxidation process coupled with the existence of isostructural rock salt phase between TiN (no oxygen substitution) and TiO (N fully substituting O) can result in the realization of a much wider doping range of O into TiN than N into TiO2 system. However, due to the absence of the isostructural phase between TiN and TiO2, the substitutional nitrogen doping beyond 5.5 atm% has not been realized when TiO2 is subjected to nitration. The possibility of oxygen substituting the nitrogen instead of occupying an interstitial site in the TiN lattice is ruled out due to positive incorporation energy for the intestinal site (+1.59 eV) and a negative incorporation energy (−3.54 e V) at the N substitutional site [52].

Fig. 2
XRD patterns of TiNO deposited at different temperatures. The peaks indexed s belong to the polycrystalline Ti substrate. The unit cells of the TiN and Ti are shown schematically with the red, green, and blue arrows indicating a-, b-, and c-axis, respectively.
Fig. 2
XRD patterns of TiNO deposited at different temperatures. The peaks indexed s belong to the polycrystalline Ti substrate. The unit cells of the TiN and Ti are shown schematically with the red, green, and blue arrows indicating a-, b-, and c-axis, respectively.
Close modal

The (111) textured growth of TiNO films on a polycrystalline substrate is understood using the concept of domain epitaxy [53] and lattice arrangement along (111) plane of TiNO and (200) planes. Figure 3 schematically highlights the lattice arrangement at the film-substrate interface. The lattice constants of two-dimensional TiN (111) and Ti (200) planes are 5.19 Å and 4.11 Å, respectively, which results in a 20.8% compressive misfit strain in the TiN lattice. However, this mismatch reduces to 1.06% via the matching of four unit cells of TiN and five unit cells of Ti based on the paradigm of domain matching concept for epitaxial thin film [53,54].

Fig. 3
Schematic representation of the polyhedral domain structure of TiNO (TiN-111) deposited on polycrystalline Ti substrate (Ti-200): (a) side view along the a-direction, (b) side view along the b-direction, (c) isometric view, all with reference to a unit cell thick layer of the Ti polycrystalline lattice, (d) interfacial (top) plane of nitrogen-terminated-TiN on Ti, and (e) interfacial (top) plane of titanium-terminated-TiN on Ti. The red, green, and blue arrows indicate the a-, b-, and c-axis, respectively.
Fig. 3
Schematic representation of the polyhedral domain structure of TiNO (TiN-111) deposited on polycrystalline Ti substrate (Ti-200): (a) side view along the a-direction, (b) side view along the b-direction, (c) isometric view, all with reference to a unit cell thick layer of the Ti polycrystalline lattice, (d) interfacial (top) plane of nitrogen-terminated-TiN on Ti, and (e) interfacial (top) plane of titanium-terminated-TiN on Ti. The red, green, and blue arrows indicate the a-, b-, and c-axis, respectively.
Close modal

The XPS analysis was performed after 60 cycles of sputtering to remove absorbed carbon, oxygen, and water vapor at the surface using a 500 eV ion gun. The general survey scan is shown in Fig. 4(a). The core levels of titanium (2s, 2p, 3s, 3p) and oxygen (1s) are shown in Figs. 4(b) and 4(c), respectively. As shown in Fig. 4(b), Ti 2p core-level peaks were observed between 454.2 eV and 467.2 eV as a result of spin-orbital coupling, the Ti 2p doublet of Ti 2p1/2 and Ti 2p3/2 are also observed. Previous publications [22,31,47] have shown that the instability of TiN in oxygen ambient (and even vacuum conditions that contain residual oxygen) leads to the formation of TiNO and TiO2. This was further confirmed by the presence of three peaks in Fig. 4(b) (A1, A2, and A3), which correspond to the Ti 2p3/2 peak. For example, in the sample deposited at 500 °C, the peak A1, observed at 454.88 eV, is connected with the formation of titanium-nitrogen bonding in TiN [47]; peak A2 at 456.83 eV is related to the titanium-oxygen-nitrogen bonding. The peak A3 at 458.70 eV is associated with the titanium-oxygen bonding. Thus, the Ti 2p3/2 peak was a deconvolution of these species. Similarly, the Ti 2p1/2 peaks for the same sample (500 °C), observed at 464.47 eV, 462.63 eV, and 460.89 eV, are characteristic features of TiO2 (anatase), TiNO, and TiN, respectively [55]. The molecular fractions of TiO2, TiNO, and TiN, calculated using the area of these peaks, are plotted in Fig. 4(d) as a function of film deposition temperature. As shown in this figure, the molecular fraction of TiNO is observed to decrease by ∼26.45% while the molecular fractions of TiO2 and the TiN increase by ∼15.47% and ∼30.1% with a standard deviation less than ±3% determined on two different samples. The difference in binding energies between Ti 2p3/2 and Ti 2p1/2 of TiN, TiNO, and TiO2 remained fairly similar, with an average of 5.93, 5.67, and 5.74 eV for samples synthesized under the three different temperatures. The chemical shifts of the deconvolved peaks (to higher binding energies) are evident with the change in deposition temperatures used for making the TiNO films. For example, the TiN 2p3/2, TiNO 2p3/2, and TiO2 2p3/2 peaks demonstrated approximately 0.05%, 0.10%, and 0.07% increase in their binding energy, respectively, signaling higher oxidation state of the TiNO samples deposited at lower temperatures. To avoid errors in XPS data and their analysis, the following care was taken: XPS spectra were recorded with low noise and high peak-to-background ratio, selecting proper model function, reducing the number of free parameters by means of existing information on well-known doublet splitting and intensity ratios, running the fit procedure with Shirley background subtraction and checking the values of the residual standard deviation fittings [56]. For example, the residual standard deviation for our XPS fittings is ∼1.5 units (CPS), which is quite good.

Fig. 4
The XPS spectra recorded for the TiNO films deposited at different temperatures: (a) general survey, (b) deconvoluted high-resolution Ti 2p core level, (c) O 1s core level, and (d) molecular fractions of TiN, TiNO, and TiO2 in the various samples as a function of the deposition temperature with a standard deviation less than ±3% determined on two samples done deposited under identical conditions
Fig. 4
The XPS spectra recorded for the TiNO films deposited at different temperatures: (a) general survey, (b) deconvoluted high-resolution Ti 2p core level, (c) O 1s core level, and (d) molecular fractions of TiN, TiNO, and TiO2 in the various samples as a function of the deposition temperature with a standard deviation less than ±3% determined on two samples done deposited under identical conditions
Close modal

As shown in Fig. 4(c), the core-level O 1s spectrum has been decomposed into three peaks. For example, peak 1 (529.32 eV), peak 2 (529.90 eV), and peak 3 (531.20 eV) in the 500 °C-deposited sample, corresponding to titanium-oxygen-nitrogen bonding in TiNO, titanium-oxygen bonding in TiO2, and adsorbed oxygen, respectively. The O 1s core-level peaks also undergo a minor shift in peak position as a result of the change in deposition temperature. The evidence of more oxygen content in the TiNO deposited at 300 °C with respect to 400 °C and 500 °C-deposited TiNO films was obtained using X-ray photoelectron spectroscopy (XPS). The XPS fractional oxygen content in TiNO films deposited at 300 °C, 400 °C, and 500 °C were 0.72, 0.71, and 0.69, respectively (O300: O400: O500: 1.00, 0.98, 0.96). The fractional XPS oxygen content ratios in the three films are found to be inversely related to lattice constant ratios, which confirms the correlation between oxygen content in the TiNO films and their lattice parameters.

To further confirm the XPS results about the structural conversion of TiN to TiNO and TiO2, Raman spectroscopy analysis was carried out using a laser excitation of 532 nm wavelength. The results obtained from the TiNO films deposited at different temperatures are shown in Fig. 5. The presence of the signature peaks for the anatase TiO2 phase (Eg, B1g, A1g, and Eg) in the spectra at 144, 394, 514, 634 cm−1 indicates the formation of anatase TiO2 [39,40,43,5759]; the intensity of these peaks increases with an increase in the film deposition temperature. The Eg peak for the sample deposited at 300 °C is clearly present but is very broad; the other peaks are naturally non-existent due to their relatively low intensity. The Raman spectrum (RS) of 300 °C sample is reminiscent of an amorphous TiO2 phase with low resistivity [60,61]. Thus, it appears that while TiNO films are structurally qualitatively similar with respect to XRD patterns and crystallinity, the TiO2 phase in the TiN-TiNO-TiO2 composite mixture is mostly amorphous at 300 °C. Since the amorphous TiO2 possesses a low resistivity [61], the samples deposited at 300 °C have the lowest sheet resistance in the present study, too (inset of Fig. 5). As the temperature of deposition increases, the resistivity of the composite structure increases due to the fact that TiNO films lose oxygen, making it more resistive, plus the fraction of crystalline and insulating TiO2 increases. A comparison of these peak positions to the reported Raman peak positions for pure anatase TiO2 indicates a shift in the peak position [59]. This shift in the peak position with respect to peaks in the pure TiO2 is attributed to the doping/modification in the band structure of the TiO2 films due to the hybridization of N (2p) and O (2p) energy levels [47]. As shown in this inset figure, the 300 °C sample has almost five orders of magnitudes lower sheet resistance (3 Ω/square) in comparison to the 400 °C (97,119 Ω/square) and 500 °C (97,371 Ω/square) samples.

Fig. 5
RS intensity signal versus Raman shift recorded from the TiNO deposited at different temperatures. The inset figure shows the variation of sheet resistance as a function of film deposition temperature.
Fig. 5
RS intensity signal versus Raman shift recorded from the TiNO deposited at different temperatures. The inset figure shows the variation of sheet resistance as a function of film deposition temperature.
Close modal

The surface morphology of the TiNO electrode was investigated by recording SEM images prior to and after the electrochemical test. The recording of SEM images of the TiNO film surface before and after the electrochemical measurements is used to examine the surface morphology and surface reconstruction during the applied potential window. The images in Fig. 6 on the left column (a), (c), and (e) indicate the surface morphology before electrochemical measurements, and the set of images on the right of Figs. 6(b), 6(d), and 6(f) in the same row are the surface morphologies of the samples after the electrochemical measurements. The creation of dark discolorations (indicated by the red circle) in the aftermath images is observed. This suggests very moderate degradation of the samples, which indicates the stability of the electrodes under the cycling conditions [20,27].

Fig. 6
SEM images of TiNO films deposited at (a) 300 °C, (c) 400 °C, and (e) 500 °C. The corresponding SEM images of these films after the electrochemical measurements are shown in (b), (d), and (f).
Fig. 6
SEM images of TiNO films deposited at (a) 300 °C, (c) 400 °C, and (e) 500 °C. The corresponding SEM images of these films after the electrochemical measurements are shown in (b), (d), and (f).
Close modal

The CVs of the TiNO thin film, deposited at three different temperatures, are presented in Fig. 7(a). All three CVs, recorded at the same scale rate of 5 mV/s, displayed explicit oxidation/reduction peaks. Figure 7(b) shows the variation of the specific capacitance of these films as a function of TiNO film deposition temperatures; the specific capacitance of TiNO films deposited at 300 °C is nearly 20% lower than that of TiNO films deposited at 500. Shown in Fig. 8(a) are the CV curves at various scan rates for the TiNO sample deposited at 300 °C, which has the highest capacitance values. It is clear from Fig. 8(a) that the oxidation peaks shift to more oxidative potentials (moves to the right), and the reduction peaks shift to more reductive potentials (moves to the left). The shift in the oxidation and reduction potential is attributed to irreversible electron transfer reactions, where the electron transfer kinetics are slow [62]. In the case of reversible electron transfer reactions, which have rapid electron transfer kinetics, the oxidation and reduction peak positions are independent of the CV scan rate [62]. Also, the TiNO sample exhibits a non-linear relationship between current density and scan rate. This non-linear CV behavior suggests that the TiNO material system is pseudocapacitive in nature as opposed to the rectangular shape of the electrical double-layer capacitor without the presence of any redox waves [63].

Fig. 7
(a) Cyclic voltammogram curves for the TiNO films as a function of potential and (b) specific capacitance as a function of TiNO film deposition temperatures at 5 mV/s scan rate
Fig. 7
(a) Cyclic voltammogram curves for the TiNO films as a function of potential and (b) specific capacitance as a function of TiNO film deposition temperatures at 5 mV/s scan rate
Close modal
Fig. 8
(a) Cyclic voltammogram curves for the 300 °C-deposited TiNO film at various scan rates and (b) variation of specific capacitance as a function of scan rate for TiNO thin films deposited at different temperatures
Fig. 8
(a) Cyclic voltammogram curves for the 300 °C-deposited TiNO film at various scan rates and (b) variation of specific capacitance as a function of scan rate for TiNO thin films deposited at different temperatures
Close modal
The specific capacitance of the TiNO films from the CV curve was calculated using the following equation [64]:
(1)

Here Q is the area under the CV curve, v/t is the scan rate, ΔV is the potential window, and the area is the geometric area of the electrode submerged. The variation of specific capacitance with scan rates at different thin film electrode fabrication temperatures is shown in Fig. 8(b). As shown in this figure, the specific capacitance of the TiNO thin films decreases rapidly with an increase in the scan rates. The lower specific capacitance at the higher scan rate is attributed to insufficient time for the electrolyte to get adsorbed and desorbed on the electrode surface [65]. The decrease in the specific capacitance, however, reaches saturation, which might be attributed to a reduction in the concertation gradient of electroactive species at the electrode surface above a certain scan rate [66], which needs further investigation.

The Nyquist plots obtained from the EIS are shown in Fig. 9. Using a simplified equivalent electrical circuit (EEC) (insert in Fig. 9), the EIS data can be explained [47]. In the EEC, EIS spectra are represented as combinations of circuit elements. R1 is the electrolyte resistance, R2 is the charge transfer resistance at the interface between the electrolyte and the TiNO electrode, and R3 is the resistance of the TiNO film [67,68]. C is the capacitance of the interfacial layer. The circuit shows that R1 and R3 are connected in series with R2 and C. It should be noted that the starting material, TiN, is well known to electrically be very conducting with room temperature resistivity of the order of ∼50 µΩ cm [69]. The resistivity of TiN thin films increases with the incorporation of oxygen in the TiN lattice. The XRD results and XPS do show that the fraction of the TiNO phase (resistivity ∼150 µΩ cm) in the TiN-TiNO-TiO2 in the composite decreases with an increase in the deposition temperature [70,71]. In other words, the fraction of the TiO2 phase (resistivity ∼1018 Ω cm) increases with an increase in the film deposition temperature, and therefore, the net resistivity of TiN-TiNO-TiO2 composite thin film structure increases with an increase in the film deposition temperature. This is manifested in the larger diameter of the semicircles (larger charge transfer resistance) for samples made at higher temperatures. However, it is not possible to quantify iR drop alone and provide specific capacitance values at this time. The equivalent impedance of the circuit is given by Eq. (2)
(2)
Fig. 9
Nyquist plots for the TiNO thin film deposited at different temperatures. The equivalent circuit of the electrochemical cell is presented in the inset diagram.
Fig. 9
Nyquist plots for the TiNO thin film deposited at different temperatures. The equivalent circuit of the electrochemical cell is presented in the inset diagram.
Close modal

In this equation, Zc is the capacitance of the impedance, j=1, and ω is the angular frequency. The Nyquist plot stems from the electrical circuit contained in the semicircle of the EE, where the diameter represents the charge transfer resistance, R2. The charge transfer resistance of the interface of the double-layer structure decreases as the diameter decreases, resulting in faster reaction kinetics [19,72,73]. Galvanostatic charge–discharge tests were used to further examine the electrochemical characteristics and charge storage capacity of TiNO thin films in a two-electrode setup. Figure 10(a) shows the galvanostatic charge–discharge characteristics of the TiNO thin film deposited at 300 °C in 1 M KOH electrolyte.

Fig. 10
(a) Potential versus time plots recorded during charge–discharge measurements at different current densities for 300 °C-deposited TiNO thin film symmetric cell and (b) variation of specific capacitance as a function of applied current density for TiNO thin films deposited at different temperatures
Fig. 10
(a) Potential versus time plots recorded during charge–discharge measurements at different current densities for 300 °C-deposited TiNO thin film symmetric cell and (b) variation of specific capacitance as a function of applied current density for TiNO thin films deposited at different temperatures
Close modal
With increasing current density, the galvanostatic charge–discharge time reduces. The curves show three distinct stages of discharge potential variation. Due to internal resistance in the beginning, the potential suddenly decreases. During the second stage, the potential then decreases linearly as the amount of time increases. The charge separation at the interface between the electrode and electrolyte correlates to a double-layer capacitance characteristic, as shown by this. The correlation between the potential and time becomes non-linear in the final stage, which is related to a typical pseudocapacitance behavior arising from redox reactions, electrochemical adsorption/desorption, or ion intercalation [74,75]. The specific capacitance (Csp) of the TiNO films from the galvanostatic charge–discharge measurement was calculated using the equation given as follows:
(3)

In this equation, I is the discharge current (A), Δt is the discharge time (s), ΔV is the potential window (V), and A is the geometric area of the TiNO film submerged. From the plot (Fig. 10(b)), we can see that the TiNO 300 °C electrode recorded the highest Csp of 66 mF/cm2. This follows our earlier observation from the CV results. The improvement in energy storage could be a result of improved surface conditions as less oxygen dissolves at low deposition temperatures, leading to a high redox activity at the electrode/electrolyte interface.

In addition to specific capacitance, the power and energy densities of the TiNO thin film were calculated using the following equations [65]:
(4)
(5)

The Ragone plot of the TiNO supercapacitor cells is shown in Fig. 11, where the power density is presented as a function of energy density. The obtained energy and power densities were comparable among all three samples performed in this study. The highest energy and power density for the device were recorded to be 7.55 Wh/cm2 and 952.155 W/cm2 for the TiNO300°C and TiNO500C cells, respectively. The TiNO400C had an in-between energy density of 7.37 Wh/cm2 and power of 679.29 W/cm2. The stability of the supercapacitor has been investigated by monitoring the potential of 300 °C-deposited TiNO symmetric cell versus charge–discharge time. Figure 12 shows the 1st and 1000th galvanostatic cycles of the cell. The supercapacitor made with the TiNO300°C electrode material showed a negligible loss of capacitance after 1000 galvanostatic cycles with a capacitance retention ability of 90%.

Fig. 11
Ragone plot for the TiNO thin films deposited at different deposition temperatures
Fig. 11
Ragone plot for the TiNO thin films deposited at different deposition temperatures
Close modal
Fig. 12
Potential versus time plots recorded during charge–discharge measurements of the 300 °C-deposited TiNO thin film cell after the 1st and 1000th cycles
Fig. 12
Potential versus time plots recorded during charge–discharge measurements of the 300 °C-deposited TiNO thin film cell after the 1st and 1000th cycles
Close modal

4 Conclusions

High-performance TiNO thin films were synthesized by a PLD method on polycrystalline titanium metal plates. The highest specific capacitance of 58.7 mF/cm2 was observed for the TiNO thin films deposited at 300 °C, which is ∼22% more than those of TiNO films deposited at higher temperatures. The capacitance retention ability of 300 °C-deposited films was ∼90% after 1000 cycles. These films also showed high power and energy densities of 952.155 W/cm2 and 7.55 Wh/cm2, respectively. The present study has established the thin film deposition temperature as an effective tuning strategy to rationally modulate the morphology and the structure of TiNO films film to achieve a high charge storage performance.

Acknowledgment

The authors would like to acknowledge the financial support from the NSF Partnership of the Research and Education in Materials (PREM) program via Grant No. DMR-2122067. M.L., S.S., and S.A. acknowledge receiving partial support in the summer from the Energy Frontier Research Center (EFRC) program via Grant No. DE-SC0023415. This work also made use of the Cornell Center for Materials Research Shared Facilities, which are supported through the NSF MRSEC program (DMR-1719875) and of the Joint School of Nanoscience and Nanoengineering (JSNN), a member of the National Nanotechnology Coordinated Infrastructure (NNCI), which is supported by the National Science Foundation (Grant No. ECCS-2025462).

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

References

1.
Fera
,
M.
,
Macchiaroli
,
R.
,
Iannone
,
R.
,
Miranda
,
S.
, and
Riemma
,
S.
,
2016
, “
Economic Evaluation Model for the Energy Demand Response
,”
Energy
,
112
, pp.
457
468
.
2.
Xie
,
Q.
,
Bao
,
R.
,
Zheng
,
A.
,
Zhang
,
Y.
,
Wu
,
S.
,
Xie
,
C.
, and
Zhao
,
P.
,
2016
, “
Sustainable Low-Cost Green Electrodes With High Volumetric Capacitance for Aqueous Symmetric Supercapacitors With High Energy Density
,”
ACS Sustainable Chem. Eng.
,
4
(
3
), pp.
1422
1430
.
3.
Kyeremateng
,
N. A.
,
Brousse
,
T.
, and
Pech
,
D.
,
2017
, “
Microsupercapacitors as Miniaturized Energy-Storage Components for On-Chip Electronics
,”
Nat. Nanotechnol.
,
12
(
1
), pp.
7
15
.
4.
Wang
,
G.
,
Zhang
,
L.
, and
Zhang
,
J.
,
2012
, “
A Review of Electrode Materials for Electrochemical Supercapacitors
,”
Chem. Soc. Rev.
,
41
(
2
), pp.
797
828
.
5.
Zhao
,
X.
,
Sánchez
,
B. M.
,
Dobson
,
P. J.
, and
Grant
,
P. S.
,
2011
, “
The Role of Nanomaterials in Redox-Based Supercapacitors for Next-Generation Energy Storage Devices
,”
Nanoscale
,
3
(
3
), pp.
839
855
.
6.
Simon
,
P.
,
Gogotsi
,
Y.
, and
Dunn
,
B.
,
2014
, “
Where Do Batteries End and Supercapacitors Begin?
,”
Science
,
343
(
6176
), pp.
1210
1211
.
7.
Simon
,
P.
, and
Gogotsi
,
Y.
,
2010
, “
“Materials for Electrochemical Capacitors,” Nanoscience and Technology: A Collection of Reviews From Nature Journals
,”
World Sci.
,
7
(
11
), pp.
320
329
.
8.
Nugroho
,
A.
,
Erviansyah
,
F.
,
Floresyona
,
D.
,
Mahalingam
,
S.
,
Manap
,
A.
,
Afandi
,
N.
,
Lau
,
K.
, and
Chia
,
C.
,
2022
, “
Synthesis and Characterization NS-Reduced Graphene Oxide Hydrogel and Its Electrochemical Properties
,”
Lett. Mater.
,
12
(
2
), pp.
169
174
.
9.
Ji
,
H.
,
Zhao
,
X.
,
Qiao
,
Z.
,
Jung
,
J.
,
Zhu
,
Y.
,
Lu
,
Y.
,
Zhang
,
L. L.
,
MacDonald
,
A. H.
, and
Ruoff
,
R. S.
,
2014
, “
Capacitance of Carbon-Based Electrical Double-Layer Capacitors
,”
Nat. Commun.
,
5
(
1
), p.
3317
.
10.
Jäckel
,
N.
,
Simon
,
P.
,
Gogotsi
,
Y.
, and
Presser
,
V.
,
2016
, “
Increase in Capacitance by Subnanometer Pores in Carbon
,”
ACS Energy Lett.
,
1
(
6
), pp.
1262
1265
.
11.
Ramavath
,
J. N.
,
Raja
,
M.
,
Kumar
,
S.
, and
Kothandaraman
,
R.
,
2019
, “
Mild Acidic Mixed Electrolyte for High-Performance Electrical Double Layer Capacitor
,”
Appl. Surf. Sci.
,
489
, pp.
867
874
.
12.
Choi
,
D.
,
Blomgren
,
G. E.
, and
Kumta
,
P. N.
,
2006
, “
Fast and Reversible Surface Redox Reaction in Nanocrystalline Vanadium Nitride Supercapacitors
,”
Adv. Mater.
,
18
(
9
), pp.
1178
1182
.
13.
Shi
,
J.
,
Jiang
,
B.
,
Li
,
C.
,
Yan
,
F.
,
Wang
,
D.
,
Yang
,
C.
, and
Wan
,
J.
,
2020
, “
Review of Transition Metal Nitrides and Transition Metal Nitrides/Carbon Nanocomposites for Supercapacitor Electrodes
,”
Mater. Chem. Phys.
,
245
, p.
122533
.
14.
Djire
,
A.
,
Pande
,
P.
,
Deb
,
A.
,
Siegel
,
J. B.
,
Ajenifujah
,
O. T.
,
He
,
L.
,
Sleightholme
,
A. E.
,
Rasmussen
,
P. G.
, and
Thompson
,
L. T.
,
2019
, “
Unveiling the Pseudocapacitive Charge Storage Mechanisms of Nanostructured Vanadium Nitrides Using In-Situ Analyses
,”
Nano Energy
,
60
, pp.
72
81
.
15.
Sahoo
,
M. K.
, and
Rao
,
G. R.
,
2018
, “
Fabrication of NiCo2S4 Nanoball Embedded Nitrogen-Doped Mesoporous Carbon on Nickel Foam as an Advanced Charge Storage Material
,”
Electrochim. Acta
,
268
, pp.
139
149
.
16.
Grace
,
A. N.
,
Ramachandran
,
R.
,
Vinoba
,
M.
,
Choi
,
S. Y.
,
Chu
,
D. H.
,
Yoon
,
Y.
,
Nam
,
S. C.
, and
Jeong
,
S. K.
,
2014
, “
Facile Synthesis and Electrochemical Properties of Co3S4-Nitrogen-Doped Graphene Nanocomposites for Supercapacitor Applications
,”
Electroanalysis
,
26
(
1
), pp.
199
208
.
17.
Zheng
,
J.
,
Cygan
,
P.
, and
Jow
,
T.
,
1995
, “
Hydrous Ruthenium Oxide as an Electrode Material for Electrochemical Capacitors
,”
J. Electrochem. Soc.
,
142
(
8
), pp.
2699
2703
.
18.
Augustyn
,
V.
,
Simon
,
P.
, and
Dunn
,
B.
,
2014
, “
Pseudocapacitive Oxide Materials for High-Rate Electrochemical Energy Storage
,”
Energy Environ. Sci.
,
7
(
5
), pp.
1597
1614
.
19.
Som
,
J.
,
Choi
,
J.
,
Zhang
,
H.
,
Reddy Mucha
,
N.
,
Fialkova
,
S.
,
Mensah-Darkwa
,
K.
,
Suntivich
,
J.
,
Gupta
,
R. K.
, and
Kumar
,
D.
,
2022
, “
Effect of Substrate-Induced Lattice Strain on the Electrochemical Properties of Pulsed Laser Deposited Nickel Oxide Thin Film
,”
Mater. Sci. Eng. B
,
280
, p.
115711
.
20.
Kartachova
,
O.
,
Glushenkov
,
A. M.
,
Chen
,
Y.
,
Zhang
,
H.
,
Dai
,
X. J.
, and
Chen
,
Y.
,
2012
, “
Electrochemical Capacitance of Mesoporous Tungsten Oxynitride in Aqueous Electrolytes
,”
J. Power Sources
,
220
, pp.
298
305
.
21.
Kumar
,
U. N.
,
Ghosh
,
S.
, and
Thomas
,
T.
,
2019
, “
Metal Oxynitrides as Promising Electrode Materials for Supercapacitor Applications
,”
ChemElectroChem
,
6
(
5
), pp.
1255
1272
.
22.
Roy
,
M.
,
Mucha
,
N. R.
,
Fialkova
,
S.
, and
Kumar
,
D.
,
2021
, “
Effect of Thickness on Metal-to-Semiconductor Transition in 2-Dimensional TiN Thin Films
,”
AIP Adv.
,
11
(
4
), p.
045204
.
23.
Sarkar
,
K.
,
Jaipan
,
P.
,
Choi
,
J.
,
Haywood
,
T.
,
Tran
,
D.
,
Mucha
,
N. R.
,
Yarmolenko
,
S.
,
Scott-Emuakpor
,
O.
,
Sundaresan
,
M.
, and
Gupta
,
R. K.
,
2020
, “
Enhancement in Corrosion Resistance and Vibration Damping Performance in Titanium by Titanium Nitride Coating
,”
SN Appl. Sci.
,
2
(
5
), pp.
1
14
.
24.
Jaipan
,
P.
,
Nannuri
,
C.
,
Mucha
,
N. R.
,
Singh
,
M. P.
,
Xu
,
Z.
,
Moatti
,
A.
,
Narayan
,
J.
, et al
,
2018
, “
Influence of Gold Catalyst on the Growth of Titanium Nitride Nanowires
,”
Mater. Focus
,
7
(
5
), pp.
720
725
.
25.
Ghosh
,
S.
,
Jeong
,
S. M.
, and
Polaki
,
S. R.
,
2018
, “
A Review on Metal Nitrides/Oxynitrides as an Emerging Supercapacitor Electrode Beyond Oxide
,”
Korean J. Chem. Eng.
,
35
(
7
), pp.
1389
1408
.
26.
Chen
,
T.-T.
,
Liu
,
H.-P.
,
Wei
,
Y.-J.
,
Chang
,
I.-C.
,
Yang
,
M.-H.
,
Lin
,
Y.-S.
,
Chan
,
K.-L.
,
Chiu
,
H.-T.
, and
Lee
,
C.-Y.
,
2014
, “
Porous Titanium Oxynitride Sheets as Electrochemical Electrodes for Energy Storage
,”
Nanoscale
,
6
(
10
), pp.
5106
5109
.
27.
Lee
,
E. J.
,
Lee
,
L.
,
Abbas
,
M. A.
, and
Bang
,
J. H.
,
2017
, “
The Influence of Surface Area, Porous Structure, and Surface State on the Supercapacitor Performance of Titanium Oxynitride: Implications for a Nanostructuring Strategy
,”
Phys. Chem. Chem. Phys.
,
19
(
31
), pp.
21140
21151
.
28.
Wang
,
Z.
,
Li
,
Z.
, and
Zou
,
Z.
,
2015
, “
Application of Binder-Free TiOxN1− x Nanogrid Film as a High-Power Supercapacitor Electrode
,”
J. Power Sources
,
296
, pp.
53
63
.
29.
Yu
,
M.
,
Han
,
Y.
,
Cheng
,
X.
,
Hu
,
L.
,
Zeng
,
Y.
,
Chen
,
M.
,
Cheng
,
F.
,
Lu
,
X.
, and
Tong
,
Y.
,
2015
, “
Holey Tungsten Oxynitride Nanowires: Novel Anodes Efficiently Integrate Microbial Chemical Energy Conversion and Electrochemical Energy Storage
,”
Adv. Mater.
,
27
(
19
), pp.
3085
3091
.
30.
Ghailane
,
A.
,
Oluwatosin
,
A. O.
,
Larhlimi
,
H.
,
Hejjaj
,
C.
,
Makha
,
M.
,
Busch
,
H.
,
Fischer
,
C. B.
, and
Alami
,
J.
,
2022
, “
Titanium Nitride, TiXN(1−X), Coatings Deposited by HiPIMS for Corrosion Resistance and Wear Protection Properties
,”
Appl. Surf. Sci.
,
574
, p.
151635
.
31.
Roy
,
M.
,
Sarkar
,
K.
,
Som
,
J.
,
Pfeifer
,
M. A.
,
Craciun
,
V.
,
Schall
,
J. D.
,
Aravamudhan
,
S.
,
Wise
,
F. W.
, and
Kumar
,
D.
,
2023
, “
Modulation of Structural, Electronic, and Optical Properties of Titanium Nitride Thin Films by Regulated In Situ Oxidation
,”
ACS Appl. Mater. Interfaces
,
15
(
3
), pp.
4733
4742
.
32.
Di
,
J.
,
Zhu
,
H.
,
Xia
,
J.
,
Bao
,
J.
,
Zhang
,
P.
,
Yang
,
S.-Z.
,
Li
,
H.
, and
Dai
,
S.
,
2019
, “
High-Performance Electrolytic Oxygen Evolution With a Seamless Armor Core–Shell FeCoNi Oxynitride
,”
Nanoscale
,
11
(
15
), pp.
7239
7246
.
33.
Haydous
,
F.
,
Dobeli
,
M.
,
Si
,
W.
,
Waag
,
F.
,
Li
,
F.
,
Pomjakushina
,
E.
,
Wokaun
,
A.
,
Gökce
,
B.
,
Pergolesi
,
D.
, and
Lippert
,
T.
,
2019
, “
Oxynitride Thin Films Versus Particle-Based Photoanodes: A Comparative Study for Photoelectrochemical Solar Water Splitting
,”
ACS Appl. Energy Mater.
,
2
(
1
), pp.
754
763
.
34.
Pichler
,
M.
,
Pergolesi
,
D.
,
Landsmann
,
S.
,
Chawla
,
V.
,
Michler
,
J.
,
Döbeli
,
M.
,
Wokaun
,
A.
, and
Lippert
,
T.
,
2016
, “
TiN-Buffered Substrates for Photoelectrochemical Measurements of Oxynitride Thin Films
,”
Appl. Surf. Sci.
,
369
(
1
), pp.
67
75
.
35.
Roy
,
M.
, and
Kumar
,
D.
, “
Blue Shift in Ultraviolet Absorption Spectra of Oxygen Doped Titanium Nitride Thin Films
,”
Proceedings of ASME 2020 International Mechanical Engineering Congress and Exposition
, p.
V003T03A020
.
36.
Roy
,
M.
,
2018
, “
Growth, Structural, and Electrical Properties of TiN Thin Films
,”
M.S. thesis
,
North Carolina Agricultural and Technical State University
,
Ann Arbor, MI
.
37.
Sherman
,
A.
,
1990
, “
Growth and Properties of LPCVD Titanium Nitride as a Diffusion Barrier for Silicon Device Technology
,”
J. Electrochem. Soc.
,
137
(
6
), pp.
1892
1897
.
38.
Naik
,
G. V.
,
Schroeder
,
J. L.
,
Ni
,
X.
,
Kildishev
,
A. V.
,
Sands
,
T. D.
, and
Boltasseva
,
A.
,
2012
, “
Titanium Nitride as a Plasmonic Material for Visible and Near-Infrared Wavelengths
,”
Opt. Mater. Expr.
,
2
(
4
), pp.
478
489
.
39.
Jia
,
L.
,
Lu
,
H.
,
Ran
,
Y.
,
Zhao
,
S.
,
Liu
,
H.
,
Li
,
Y.
,
Jiang
,
Z.
, and
Wang
,
Z.
,
2019
, “
Structural and Dielectric Properties of Ion Beam Deposited Titanium Oxynitride Thin Films
,”
J. Mater. Sci.
,
54
(
2
), pp.
1452
1461
.
40.
Maeda
,
K.
, and
Domen
,
K.
,
2011
, “
Oxynitride Materials for Solar Water Splitting
,”
MRS Bull.
,
36
(
1
), pp.
25
31
.
41.
Rawal
,
S. K.
,
Chawla
,
A. K.
,
Chawla
,
V.
,
Jayaganthan
,
R.
, and
Chandra
,
R.
,
2010
, “
Effect of Ambient Gas on Structural and Optical Properties of Titanium Oxynitride Films
,”
Appl. Surf. Sci.
,
256
(
13
), pp.
4129
4135
.
42.
Yoo
,
J. B.
,
Yoo
,
H. J.
,
Jung
,
H. J.
,
Kim
,
H. S.
,
Bang
,
S.
,
Choi
,
J.
,
Suh
,
H.
,
Lee
,
J.-H.
,
Kim
,
J.-G.
, and
Hur
,
N. H.
,
2016
, “
Titanium Oxynitride Microspheres With the Rock-Salt Structure for Use as Visible-Light Photocatalysts
,”
J. Mater. Chem. A
,
4
(
3
), pp.
869
876
.
43.
Miao
,
Z.
,
Huang
,
Y.
,
Xin
,
J.
,
Su
,
X.
,
Sang
,
Y.
,
Liu
,
H.
, and
Wang
,
J.-J.
,
2019
, “
High-Performance Symmetric Supercapacitor Constructed Using Carbon Cloth Boosted by Engineering Oxygen-Containing Functional Groups
,”
ACS Appl. Mater. Interfaces
,
11
(
19
), pp.
18044
18050
.
44.
Tian
,
X.
,
Luo
,
J.
,
Nan
,
H.
,
Zou
,
H.
,
Chen
,
R.
,
Shu
,
T.
,
Li
,
X.
, et al
,
2016
, “
Transition Metal Nitride Coated with Atomic Layers of Pt as a Low-Cost, Highly Stable Electrocatalyst for the Oxygen Reduction Reaction
,”
J. Am. Chem. Soc.
,
138
(
5
), pp.
1575
1583
.
45.
Pichler
,
M.
,
Si
,
W.
,
Haydous
,
F.
,
Téllez
,
H.
,
Druce
,
J.
,
Fabbri
,
E.
,
Kazzi
,
M. E.
, et al
,
2017
, “
LaTiOxNy Thin Film Model Systems for Photocatalytic Water Splitting: Physicochemical Evolution of the Solid–Liquid Interface and the Role of the Crystallographic Orientation
,”
Adv. Funct. Mater.
,
27
(
1
), p.
1605690
.
46.
Callister
,
W. D.
, and
Rethwisch
,
D. G.
,
2007
,
The Structure of Crystalline Solids Materials Science and Engineering an Introduction
, 7th ed.,
John Wiley and Sons, Inc.
,
Hoboken, NJ
.
47.
Mucha
,
N. R.
,
Som
,
J.
,
Choi
,
J.
,
Shaji
,
S.
,
Gupta
,
R. K.
,
Meyer
,
H. M.
,
Cramer
,
C. L.
,
Elliott
,
A. M.
, and
Kumar
,
D.
,
2020
, “
High-Performance Titanium Oxynitride Thin Films for Electrocatalytic Water Oxidation
,”
ACS Appl. Energy Mater.
,
3
(
9
), pp.
8366
8374
.
48.
Kumar
,
D.
,
Oktyabrsky
,
S.
,
Kalyanaraman
,
R.
,
Narayan
,
J.
,
Apte
,
P. R.
,
Pinto
,
R.
,
Manoharan
,
S. S.
,
Hegde
,
M. S.
,
Ogale
,
S. B.
, and
Adhi
,
K. P.
,
1997
, “
Role of Silver Doping in Oxygen Incorporation of Oxide Thin Film
,”
Mater. Sci. Eng. B
,
45
(
1
), pp.
55
58
.
49.
Yen
,
S. S.
,
Chiu
,
Y. C.
,
Cheng
,
C. H.
,
Chen
,
P. C.
,
Yeh
,
Y. C.
,
Tung
,
C. H.
,
Hsu
,
H. H.
, and
Chang
,
C. Y.
,
2016
, “
Gettering Effect Induced by Oxygen-Deficient Titanium Oxide in InZnO and InGaZnO Channel Systems for Low-Power Display Applications
,”
J. Disp. Technol.
,
12
(
3
), pp.
219
223
.
50.
Müller
,
J.
,
Singh
,
B.
, and
Surplice
,
N. A.
,
1972
, “
The Gettering Action of Evaporated Films of Titanium and Erbium
,”
J. Phys. D: Appl. Phys.
,
5
(
6
), pp.
1177
1184
.
51.
Avasarala
,
B.
, and
Haldar
,
P.
,
2010
, “
Electrochemical Oxidation Behavior of Titanium Nitride Based Electrocatalysts Under PEM Fuel Cell Conditions
,”
Electrochim. Acta
,
55
(
28
), pp.
9024
9034
.
52.
McKenna
,
K. P.
,
2018
, “
Structure, Electronic Properties, and Oxygen Incorporation/Diffusion Characteristics of the Σ 5 TiN (310)[001] Tilt Grain Boundary
,”
J. Appl. Phys.
,
123
(
7
), p.
075301
.
53.
Narayan
,
J.
, and
Larson
,
B.
,
2003
, “
Domain Epitaxy: A Unified Paradigm for Thin Film Growth
,”
J. Appl. Phys.
,
93
(
1
), pp.
278
285
.
54.
Moatti
,
A.
,
Bayati
,
R.
, and
Narayan
,
J.
,
2016
, “
Epitaxial Growth of Rutile TiO2 Thin Films by Oxidation of TiN/Si{100} Heterostructure
,”
Acta Mater.
,
103
, pp.
502
511
.
55.
Nawaz
,
R.
,
Kait
,
C. F.
,
Chia
,
H. Y.
,
Isa
,
M. H.
, and
Huei
,
L. W.
,
2019
, “
Glycerol-Mediated Facile Synthesis of Colored Titania Nanoparticles for Visible Light Photodegradation of Phenolic Compounds
,”
Nanomaterials
,
9
(
11
), p.
1586
.
56.
Végh
,
J.
,
2006
, “
The Shirley Background Revised
,”
J. Electron Spectrosc. Relat. Phenom.
,
151
(
3
), pp.
159
164
.
57.
Iwashita
,
S.
,
Aoyama
,
S.
,
Nasu
,
M.
,
Shimomura
,
K.
,
Noro
,
N.
,
Hasegawa
,
T.
,
Akasaka
,
Y.
, and
Miyashita
,
K.
,
2016
, “
Periodic Oxidation for Fabricating Titanium Oxynitride Thin Films Via Atomic Layer Deposition
,”
J. Vac. Sci. Technol., A
,
34
(
1
), p.
01A145
.
58.
Fakhouri
,
H.
,
Pulpytel
,
J.
,
Smith
,
W.
,
Zolfaghari
,
A.
,
Mortaheb
,
H. R.
,
Meshkini
,
F.
,
Jafari
,
R.
,
Sutter
,
E.
, and
Arefi-Khonsari
,
F.
,
2014
, “
Control of the Visible and UV Light Water Splitting and Photocatalysis of Nitrogen Doped TiO2 Thin Films Deposited by Reactive Magnetron Sputtering
,”
Appl. Catal., B
,
144
(
1
), pp.
12
21
.
59.
El-Deen
,
S. S.
,
Hashem
,
A. M.
,
Abdel Ghany
,
A. E.
,
Indris
,
S.
,
Ehrenberg
,
H.
,
Mauger
,
A.
, and
Julien
,
C. M.
,
2018
, “
Anatase TiO2 Nanoparticles for Lithium-Ion Batteries
,”
Ionics
,
24
(
10
), pp.
2925
2934
.
60.
Bradley
,
J. D.
,
Evans
,
C. C.
,
Choy
,
J. T.
,
Reshef
,
O.
,
Deotare
,
P. B.
,
Parsy
,
F.
,
Phillips
,
K. C.
,
Lončar
,
M.
, and
Mazur
,
E.
,
2012
, “
Submicrometer-Wide Amorphous and Polycrystalline Anatase TiO2 Waveguides for Microphotonic Devices
,”
Opt Expr.
,
20
(
21
), pp.
23821
23831
.
61.
Park
,
G. S.
,
Lee
,
S.
,
Kim
,
D.-S.
,
Park
,
S. Y.
,
Koh
,
J. H.
,
Won
,
D. H.
,
Lee
,
P.
,
Do
,
Y. R.
, and
Min
,
B. K.
,
2023
, “
Amorphous TiO2 Passivating Contacts for Cu(In,Ga)(S,Se)2 Ultrathin Solar Cells: Defect-State-Mediated Hole Conduction
,”
Adv. Energy Mater.
,
13
(
8
), p.
2203183
.
62.
Elgrishi
,
N.
,
Rountree
,
K. J.
,
McCarthy
,
B. D.
,
Rountree
,
E. S.
,
Eisenhart
,
T. T.
, and
Dempsey
,
J. L.
,
2018
, “
A Practical Beginner’s Guide to Cyclic Voltammetry
,”
J. Chem. Educ.
,
95
(
2
), pp.
197
206
.
63.
Fleischmann
,
S.
,
Mitchell
,
J. B.
,
Wang
,
R.
,
Zhan
,
C.
,
Jiang
,
D.-E.
,
Presser
,
V.
, and
Augustyn
,
V.
,
2020
, “
Pseudocapacitance: From Fundamental Understanding to High Power Energy Storage Materials
,”
Chem. Rev.
,
120
(
14
), pp.
6738
6782
.
64.
Gomez
,
J.
, and
Kalu
,
E. E.
,
2013
, “
High-Performance Binder-Free Co–Mn Composite Oxide Supercapacitor Electrode
,”
J. Power Sources
,
230
, pp.
218
224
.
65.
Zequine
,
C.
,
Ranaweera
,
C. K.
,
Wang
,
Z.
,
Singh
,
S.
,
Tripathi
,
P.
,
Srivastava
,
O. N.
,
Gupta
,
B. K.
, et al
,
2016
, “
High Performance and Flexible Supercapacitors Based on Carbonized Bamboo Fibers for Wide Temperature Applications
,”
Sci. Rep.
,
6
(
1
), p.
31704
.
66.
Khalafi
,
L.
,
Cunningham
,
A. M.
,
Hoober-Burkhardt
,
L. E.
, and
Rafiee
,
M.
,
2021
, “
Why Is Voltammetric Current Scan Rate Dependent? Representation of a Mathematically Dense Concept Using Conceptual Thinking
,”
J. Chem. Educ.
,
98
(
12
), pp.
3957
3961
.
67.
Mavrokefalos
,
C. K.
, and
Patzke
,
G. R.
,
2019
, “
Water Oxidation Catalysts: The Quest for New Oxide-Based Materials
,”
Inorganics
,
7
(
3
), p.
29
.
68.
Wilson
,
J. R.
,
Schwartz
,
D. T.
, and
Adler
,
S. B.
,
2006
, “
Nonlinear Electrochemical Impedance Spectroscopy for Solid Oxide Fuel Cell Cathode Materials
,”
Electrochim. Acta
,
51
(
8
), pp.
1389
1402
.
69.
Mucha
,
N.
,
Som
,
J.
,
Shaji
,
S.
,
Fialkova
,
S.
,
Apte
,
P.
,
Balasubramanian
,
B.
,
Shield
,
J.
,
Anderson
,
M.
, and
Kumar
,
D.
,
2020
, “
Electrical and Optical Properties of Titanium Oxynitride Thin Films
,”
J. Mater. Sci.
,
55
(
12
), pp.
5123
5134
.
70.
Yan
,
L.
,
Chen
,
G.
,
Tan
,
S.
,
Zhou
,
M.
,
Zou
,
G.
,
Deng
,
S.
,
Smirnov
,
S.
, and
Luo
,
H.
,
2015
, “
Titanium Oxynitride Nanoparticles Anchored on Carbon Nanotubes as Energy Storage Materials
,”
ACS Appl. Mater. Interfaces
,
7
(
43
), pp.
24212
24217
.
71.
Braic
,
M.
,
Balaceanu
,
M.
,
Vladescu
,
A.
,
Kiss
,
A.
,
Braic
,
V.
,
Epurescu
,
G.
,
Dinescu
,
G.
,
Moldovan
,
A.
,
Birjega
,
R.
, and
Dinescu
,
M.
,
2007
, “
Preparation and Characterization of Titanium Oxy-Nitride Thin Films
,”
Appl. Surf. Sci.
,
253
(
19
), pp.
8210
8214
.
72.
Zequine
,
C.
,
Bhoyate
,
S.
,
Wang
,
F.
,
Li
,
X.
,
Siam
,
K.
,
Kahol
,
P.
, and
Gupta
,
R.
,
2019
, “
Effect of Solvent for Tailoring the Nanomorphology of Multinary CuCo2S4 for Overall Water Splitting and Energy Storage
,”
J. Alloys Compd.
,
784
, pp.
1
7
.
73.
Mitchell
,
E.
,
Gupta
,
R. K.
,
Mensah-Darkwa
,
K.
,
Kumar
,
D.
,
Ramasamy
,
K.
,
Gupta
,
B. K.
, and
Kahol
,
P.
,
2014
, “
Facile Synthesis and Morphogenesis of Superparamagnetic Iron Oxide Nanoparticles for High-Performance Supercapacitor Applications
,”
New J. Chem.
,
38
(
9
), pp.
4344
4350
.
74.
Dubal
,
D. P.
,
Fulari
,
V. J.
, and
Lokhande
,
C. D.
,
2012
, “
Effect of Morphology on Supercapacitive Properties of Chemically Grown β-Ni(OH)2 Thin Films
,”
Microporous Mesoporous Mater.
,
151
, pp.
511
516
.
75.
Pimsawat
,
A.
,
Tangtrakarn
,
A.
,
Pimsawat
,
N.
, and
Daengsakul
,
S.
,
2019
, “
Effect of Substrate Surface Roughening on the Capacitance and Cycling Stability of Ni(OH)2 Nanoarray Films
,”
Sci. Rep.
,
9
(
1
), p.
16877
.