Partha Sarathi
Padhi
*ab,
Sanjay K.
Rai
bc,
R. S.
Ajimsha
a and
Pankaj
Misra
*ab
aOxide nano Electronics Lab, Laser Materials Processing Division, Raja Ramanna Centre for Advanced Technology, Indore 452013, India. E-mail: partha.rrcat@gmail.com; pmisra@rrcat.gov.in
bHomi Bhabha National Institute, Training School Complex, Anushakti Nagar, Mumbai 400094, India
cAccelerator Physics and Synchrotrons Utilization Division, Raja Ramanna Centre for Advanced Technology, Indore 452013, India
First published on 6th August 2024
The Al2O3/TiO2 nanolaminates (ATA NLs), with the dominant Maxwell–Wagner interfacial polarization, have been extensively explored in last decade due to their potential for new-generation energy storage applications. Here, we report the fabrication of device-grade sub-nanometric (<1 nm) ATA NLs using an optimized pulsed laser deposition technique, where the interface-confined carrier relaxation and sublayer conductivity contrast-induced Maxwell–Wagner interfacial polarization mechanism was engineered by precisely tailoring the individual Al2O3 and TiO2 sublayer thickness along with the top-bottom capping layer thickness. The formation of oxygen vacancy-generated carriers in reduced titania sublayers across Al2O3/TiO2 heterointerfaces and their relative response towards the applied field were responsible for both charge storage and leakage. An NL with a TiO2 and Al2O3 sublayer thickness of ∼1 and 0.6 nm, respectively, sandwiched between ∼3 nm Al2O3 barrier layers, has demonstrated an improved capacitance density of ∼33.1 fF μm−2 and a high cut-off frequency up to ∼0.5 MHz, along with a low dielectric loss of ∼0.032 and a reduced leakage current density of ∼3.08 × 10−7 A cm−2 at 1 V. The calculated energy density value of ∼4.6 J cm−3 achieved with this optimized subnanometric Al2O3/TiO2 laminate is comparable to those of state-of-the-art capacitive devices. These superior electrical properties and controllable dielectric relaxation make this laminate a promising high-k and low-loss dielectric material for next-generation nano-electronics and high-density energy storage capacitors.
In our previous report, we demonstrated the scope of engineering the interfacial polarization effect in pulsed laser deposited (PLD) ATA NLs by tailoring the sublayer conductivity contrast.9 The best dielectric performance was obtained from the NL with a total stack thickness (d) of ∼60 nm and a bilayer thickness (tbl) of ∼1.6 nm (i.e. ∼0.8 nm sublayer thickness (ts) for both Al2O3 and TiO2 sublayers). However, further scaling down to the subnanometric level, the NLs suffered from degradation of dielectric constant, cut-off frequency (fc) and Eb values along with a high dielectric loss (tanδ) and high leakage current density (Jleak) issues.9 This scaling down-assisted dielectric and electric property degradation in the subnanometric regime is also observed in laminates of different material combinations, which hinders the processing of these NLs in commercial device applications.9,15,16 Recently, a few approaches have been made to reduce the carrier loss in NLs by inserting an insulating interfacial layer across the NL/electrode interface; however, most of them lack an explicit investigation regarding their influence on carrier dynamics and dielectric properties.16–19 Furthermore, the role of interface defect states and their interdiffusion in the carrier relaxation and an analytical testing or evaluation of the energy storage properties have not yet been addressed. Since most of the storage applications demand a simultaneous requirement of high-k, low tanδ and reduced Ileak values,8,9 a detailed investigation towards further improvement in areal capacitance density (capacitance (C)/area (A)) and reduction in charge carrier loss in these ATA NL superlattice structures will be highly beneficial.
In relation to this scientific goal, here we report the deposition and characterization of device grade ATA NLs using optimized PLD growth conditions, wherein the interface confined carrier relaxation and M–W interfacial polarization process in subnanometric layers have been tailored by precisely controlling the sublayer and capping layer thicknesses. The experimental protocol reported here is designed to tailor the M–W relaxation assisted dielectric properties and subsequently to improve the storage performance of ATA NL-based nanocapacitors, by tailoring the interface confined carrier relaxation time (τ) and sublayer conductivity contrast. As an important part of the research, we have investigated the effect of Al2O3 top-bottom barrier layer thickness on carrier loss and leakage properties. In this work, Al2O3 layers of different thicknesses varying between 1 and 5 nm were inserted across both top and bottom electrode/ATA-NL interfaces, to minimize carrier losses and leakage while maintaining a high-k value up to a higher fc limit. The optimized ATA NL-based capacitors with a 3 nm Al2O3 capping layer have displayed a superior C/A of ∼33.2 fF μm−2, low loss and reduced Jleak of ∼3.08 × 10−7 A cm−2 at 1 V along with a higher breakdown voltage of ∼0.65 MV cm−1. The investigation protocols, in this work, adopted for engineering the carrier relaxation and transport properties across Al2O3/TiO2 interfaces can be implemented for other binary metal oxide combinations in NL structures and the obtained results may open up new avenues for next-generation nanoelectronics and energy storage capacitors.
For convenience, the NL with an Al2O3 sublayer thickness (tA) and TiO2 sublayer thickness (tT) of ∼0.6 nm and 1 nm, respectively, is labelled 0.6A-1T-0.6A NL and the NLs capped with Al2O3 top and bottom barrier layers of 1, 3, and 5 nm thicknesses are denoted as 1A/(0.6A-1T-0.6A)/1A, 3A/(0.6A-1T-0.6A)/3A and 5A/(0.6A-1T-0.6A)/5A NL, respectively. The dielectric and electrical characterization of the as-grown NLs was carried out in a metal–insulator–metal (MIM)-based capacitor (MIMCAP) configuration, where the RF-magnetron sputtered top-bottom conducting TiN (∼300 μΩ cm) circular electrodes of ∼50 nm thickness and ∼200 μm diameter were defined using a stainless-steel shadow mask. For ∼60 nm thick NL-based MIMCAPs, the empty cell (air) capacitance (C0 = ε0A/d) becomes ∼5.6 pF.
To find out the type of defect states and their concentration in these NLs, X-ray photoemission spectroscopy (XPS) measurements were performed using an Al Kα source with ∼1486 eV photon energy and 0.5 eV energy resolution. In order to verify the results from XPS measurements, the valence band resonant photoemission spectroscopy (VB-RPES) measurements of a representative 0.8A-0.8T NL were performed at the Beamline-10 (BL-10) of Indus-2, RRCAT, India. The beamline (BL-10) details can be found elsewhere.20 Prior to measurements, the samples were electrically grounded using silver contacts and Ar+ ion sputtering at 1.5 kV and 6.0 mA for ∼1 min was performed to clean the sample surface. A hemispherical electron energy analyser SPECS Phoibos 150, with a typical energy resolution of 40 meV at 90 eV photon energy, was used to record the high-resolution photoemission spectra. The VB spectra were collected in the angle integrated mode and normalized to the mirror current recorded during the measurement. The binding energy calibration was performed with the Ag 3d lines and Fermi edge using the standard procedure.20,21 The VB-RPES spectra were recorded across the Ti 2p-3d and Al 2p-3s transition i.e. at ∼458.5 eV and 74.3 eV photon excitation energies, respectively. The typical energy resolution around the Ti 2p-3d resonance is 200 meV and around the Al 2p-3s resonance is 35 meV. All the measurements were performed at room temperature and in an ultrahigh vacuum of 7 × 10−11 mbar. The measured XPS spectra were deconvoluted using XPS PeakFit 4.11 software, whereas the inelastic background of the core levels was subtracted using the Shirley method.22
The frequency-dependent dielectric measurements of NL-based MIMCAPs were performed in the range from 10 to 106 Hz using an impedance analyzer (Solartron make: SI-1206) with 100 mV applied AC bias. To understand the influence of ts on the M–W interfacial polarization mechanism and the carrier relaxation time, temperature dependent dielectric measurements were carried out in the range of 303–363 K, at an interval of 10 K and with an accuracy of ±1 K. The temperature was controlled using an Angstrom Sun PID controlled heater (Ang-A1205) combined with a chiller. The current–voltage measurements of these MIMCAPs were carried out using a source measurement unit (Keithley 2636B).
To verify the artificial periodic structure in the as-grown ATA NLs, XRR measurements of a representative ATA NL with 0.8 nm sublayer thickness grown on TiN coated Si, sapphire and Si substrates are carried out and the measured XRR profiles are depicted in Fig. 1(b) with Y-offsets. The main purpose of growing NLs on Si and sapphire was to investigate their interfacial and optical properties, whereas the growth of NLs on the TiN coated Si substrate was intended to construct the MIM capacitor structure for dielectric and electrical characterization. The presence of Bragg peaks in the XRR profiles is clearly indicating the maintained distinct layer feature in [0.8A-0.8T-0.8A] NL grown on all three substrates. The relative reduction in Bragg peak intensity observed for the NL grown on TiN coated Si, as compared to the bare Si and sapphire substrate, indicates slightly higher interface intermixing, which is likely due to the propagation of TiN layer roughness into the subsequent sublayers of the NL.8 Furthermore, the Bragg peak intensity drastically reduced for lower ts values, which indicates higher interface roughness and enhanced interface intermixing. The observed Bragg peak broadness indicates a slight aperiodicity in the bottom and topmost bilayer thickness of the NL, likely due to slight non uniformities across these substrates arising from the narrow angular distribution profile of the PLD plasma plume. To calculate the measured XRR profiles, a small aperiodicity in ts values and a Si-SiO2-(Al2O3-TiO2)30-Al2O3 based multilayer model, mimicking the NL structure, are implemented. An excellent match between the measured and calculated XRR profiles for [0.8A-0.8T-0.8A] NL can be observed in Fig. 1(c). The alternating alumina and titania sublayer growth in the NL structure can be confirmed from the electron density profile (EDP), illustrated in the inset of Fig. 1(c). The best fit results including the calculated aperiodicity, layer thickness, densities and interface width for 0.8A-0.8T NL grown on Si and TiN coated Si obtained from XRR curve fitting are summarized in Table S1 of the ESI.†
Cross-sectional transmission electron microscopy (TEM) measurement of a representative 0.8A-0.8T NL, grown on Si substrates, is carried out to demonstrate the controllability of optimized PLD parameters in growing subnanometric layers and to validate the multilayer periodicity in the as-grown NLs. The distinctive layer structure, shown in Fig. 1(d), confirms the long-range sublayer thickness uniformity and distinct interfaces even in the subnanometric regime. From the intensity distribution profiles, as depicted in Fig. S1(b) of the ESI,† the measured ts values for Al2O3 and TiO2 sublayers are measured to be ∼0.79 and 0.81 nm, respectively, which indicate the control in the deposition of subnanometric layers using the optimized PLD technique. The absence of any discernable crystalline phases in the selected area electron diffraction (SAED) pattern of the 0.8A-0.8T NL, shown in the inset of Fig. 1(d), confirms the amorphous nature of the deposited subnanometric layers. The finite interface interdiffusion observed from the TEM images also corroborates with the XRR results.
The impedance spectroscopy measurements of the five as-grown TiN/ATA-NL/TiN-based MIMCAPs are carried out to examine the effect of variation in tT and tA on the interfacial and dielectric properties. The schematic of one of the MIMCAPs is depicted in the inset of Fig. 2(a), and the top view of an actual device structure is shown in the inset of Fig. 2(b). It is relevant to discuss a few key observations from the frequency dispersed εr and tanδ spectra, as depicted in Fig. 2(a) and (b), respectively. From a cursory overview of frequency dispersed εr and tanδ spectra, as depicted in Fig. 2(a) and (b), respectively, the εr plateau, tanδ valley, and the fc values are found to be directly proportional to tT and inversely proportional to tA values up to a certain sublayer thickness limit. The εr plateau and tanδ values demonstrate a monotonous increment from ∼260 to 720 and ∼0.31 to 0.7 with an increase in tT from 0.4 to 1 nm and a simultaneous decrease in tA from 1 to 0.4 nm, respectively. Since the total number of interfaces of these ATA NLs is kept constant at ∼76, the variation in dielectric contributions is independent of interface density and is solely decided by the change in sublayer conductivity contrast.
Furthermore, with the increase in TiO2 content, the M–W interfacial carrier relaxation frequency peak positions in frequency dispersed tanδ plots are found to shift towards the high frequency side, as shown by arrow marks in Fig. 2(b). This clearly indicates an improvement in the semiconducting nature of titania sublayers with the increase in TiO2 content up to ∼63% in a given bilayer, which is possibly due to the increase in the Ti3+ and Ov defect generated carrier concentration.8,13 This observation concludes that the concentration of charge carriers in the ATA NLs is increasing with tT, which is increasing the sublayer conductivity contrast assisted M–W interfacial polarization and resulting monotonous increment in εr values. On the other hand, due to a fixed tbl of ∼1.6 nm, the Al2O3 barrier layers are simultaneously getting thinner with an increase in TiO2 content, hence resulting in an increase in the tanδ values. We can also observe that with an increase in tT up to 1 nm, the cut-off frequency (fc) for high k values of 0.6A-1T-0.6A NL extends up to ∼3 × 105 Hz. With an increase in TiO2/Al2O3 content up to ∼66.3%, the increase and decrease in fc and τ values, respectively, can be attributed to the increase in carrier concentration of TiO2 layers. Meanwhile, further reduction in tT down to ∼0.4 nm leads to trend reversal in fc values, possibly due to the enhanced interface intermixing of sublayers. With an increase in tT from ∼0.4 to 1 nm, there is an improvement in the working frequency domain from 16 Hz to 198 Hz in 1.2A-0.4 T NL and 1.58 KHz to 126 KHz in 0.6A-1T NL, as depicted in Fig. 2(c) and (d). The above observations strongly support that the εr and fc values of these ATA NLs are mainly dependent on the amount of charge carriers in the oxygen deficient TiO2 (TiO2−δ) sublayers. Meanwhile, the tanδ and Jleak values rely on the insulating nature of the Al2O3 barrier layers. Furthermore, at room temperature, the M–W relaxation process and hence the dielectric properties of these PLD-grown ATA NLs can be tailored by only varying individual sublayer thicknesses. Since the dielectric measurements for all the NL samples were carried out at room temperature, the change in τ and fc is definitely not a thermally induced effect.
To investigate the effect of the relative variation in tA and tT on the carrier relaxation and transport mechanism in the aforesaid NLs, the temperature dependent dielectric measurements are carried out in a temperature range from ∼303 to 373 K. The frequency dispersed εr and tanδ profiles of a few representative NLs, labelled 1A-0.6T NL, 0.8A-0.8T NL, 0.6A-1T NL, are depicted in Fig. 3(a)–(c) and in their corresponding insets, respectively. With an increase in temperature, the shift in the εr profile and the relaxation peaks towards the high frequency side indicates the thermally activated nature of this M–W relaxation process. The two sets of relaxation peaks observed in the tanδ vs. f plots support the M–W relaxation in these NLs and the decrease in τ for interface confined carriers at elevated temperature is owing to the increase in charge carrier mobility. With an increase in temperature, the tanδ values of 1A-0.6T NL initially increased and saturated at ∼353 K, whereas the trend reversed with further increase in temperature to 373 K. A higher Al interdiffusion into the TiO2 sublayer is expected, with an increase in temperature, owing to the smaller Al cation size as compared to Ti.3,4,9 Additionally, an abrupt increase in the measured εr and tanδ values for 0.6A-1T NL is observed around 363 K, as shown in Fig. 3(c), and is likely owing to the thermal soft breakdown of thinner Al2O3 barrier layers (∼0.6 nm).9 This observation further indicates that decreasing the tA below 0.6 nm severely degrades the charge holding capability, which subsequently deteriorates the M–W relaxation assisted dielectric properties. The reciprocal temperature variation of the M–W relaxation peak positions is linearly fitted using the Arrhenius equation, f(T) = f0exp(−Ea/KβT), as shown in Fig. 3(d), and the activation energies (Ea) for interfacial carrier relaxation is calculated to be in the range from ∼0.395 to 0.489 eV. This Ea values are closely matching with the literature reported activation energy values of oxygen vacancy (OV) defect states in TiO2, and it can be assigned responsible for M–W relaxation mechanism in ATA NLs.8
The X-ray photoelectron spectroscopy (XPS) measurement of a representative 0.8A-0.8T NL was carried out using Al kα (∼1486 eV) photon energy, to verify the role of OV related defect states in the origin of M–W relaxation induced enhanced dielectric properties in these ATA NLs. O 1s core level spectra are deconvoluted into four peaks, as depicted in Fig. 4(a), where the 1st and 3rd peaks around 529.7 ± 0.1 eV and 531.15 ± 0.1 are assigned to the lattice oxygens in Ti–O and Al–O bonds, respectively. The shoulder peaks observed around 530.45 ± 0.1 and 532.2 ± 0.1 eV are assigned to OV related defect states in TiO2 sublayers and surface adsorbed oxygen in the form of C–O & O–H bonds on the top Al2O3 sublayer (labelled as Al–O–OH), respectively.8 The recorded Ti 2p and Al 2p core level scans and their deconvoluted spectra are depicted in Fig. 4(b) and the inset, respectively. The Ti core level peak is deconvoluted into two sets of spin orbit splitting Ti 2p3/2 and Ti 2p1/2 peaks, corresponding to Ti4+ and Ti3+ states, with a peak separation of ∼5.9 and ∼5.78 eV, respectively. The presence of both Ti4+ and Ti3+ states in the Ti 2p core level deconvoluted spectra confirms the OV induced reduced titania (Ti2O3) phase, in the TiO2 sublayers of 0.8A-0.8T NL.13 These defect states, as a source of high concentration unpaired electrons, are known to improve the conductivity of thin TiO2 sublayers making them n-type semiconducting in nature. The deconvoluted spectra of high-resolution Al 2p core levels and the corresponding binding energy (BE) positions, shown in the inset of Fig. 4(b), clearly indicate the absence of any Al related defect states except an Al–O bonding and Al–OH related surface contamination contribution.8
Furthermore, VB-RPES measurements of a representative 0.8A-0.8T NL are performed to identify the source of defect states through a detailed understanding of the sublayer electronic structure and the recorded spectra are shown in Fig. 4(c). In the Ti 2p to 3d VB-RPES spectra, ∼1 eV below the Fermi level, the appearance of the in-gap-state (IGS) feature ‘A’ clearly signifies the presence of OV defect states as trap centers for available charge carriers.23–25 The appearance of IGS A′ in the Al 2p-3s RPES spectra and its close matching of the BE positions with those of feature A indicate the Al 3s and Ti 3d states' hybridization effect due to finite sublayer interface intermixing in the 0.8A-0.8T NL, as verified from XRR and cross-sectional TEM measurements.26
Additionally, one more important aspect can be observed from Fig. 3(d), i.e. the increment in Ea values from ∼0.395 to 0.489 eV with an increase in Al2O3 barrier layer thickness from 0.6 to 1 nm and decrease in TiO2 quantum well (QW) thickness from 1 to 0.6 nm. This increasing trend in Ea with an increase in tA can be explained by considering NLs as multiple QW structures and a strong quantum confinement effect in this 1A-0.6T NL.27 To verify this effect, UV-VIS transmission spectroscopy measurements of the aforesaid three representative NLs, grown on sapphire substrates, were carried out and are depicted in Fig. 4(b). With an increase in tA from 0.6 to 1 nm and a simultaneous decrement in the tT value from 1 to 0.6 nm, a blue shift in the measured T% vs. wavelength spectra can be observed. This confirms an improvement in the carrier confinement effect with an increase in tA and decrease in tT, which further supports the decrement in carrier loss observed from frequency dispersed tanδ profiles, shown in Fig. 2(b).
To demonstrate the reliability of the measured dielectric properties, the εr and tanδ values of 12 MIM devices from each of the S0, S1, S2, S3 NLs are depicted in the form of box chart representation, in Fig. 5(c) and the inset. A relatively broad distribution of εr and tanδ values for the S0 NL was found with the median value of ∼730; however, a decrease in this distribution profile with an increase in tAl2O3 indicates an improvement in the device stability. Furthermore, the distributions in the εr and tanδ values of ∼68%, 75%, 86%, and 95% of the tested devices of the S0, S1, S2, and S3 NLs, respectively, are found within the upper and lower quartile profiles of the box chart, which clearly suggests a consistent improvement in the stability of the device performance parameters with the increase in tAl2O3. This might be owing to the better flat and amorphous platform provided by thicker alumina layers across top-bottom electrodes, which restricts the propagation of bottom electrode roughness into the subsequent interfaces of the NL.
To investigate the role of tAl2O3 in the conduction mechanism as well as to calculate the bias dependent Jleak values, current–voltage (I–V) measurements were performed on the S0, S1, S2 and S3 NL-based MIMCAPs in the voltage range from 0 to 8 V. From the I–V profiles, as depicted in Fig. 5(d), the Jleak values for all the NLs at a low applied bias (Vapp) decreased with the increase in Vapp, reaching a minimum. Non-zero crossing current features appeared due to the M–W interfacial polarization induced built-in potential or induced back emf in these NL structures, which opposed the applied field and resulted in a reduction in charge carrier leakage across the interfaces.8,29 The electrons can quantum mechanically tunnel through ultrathin alumina barrier layers in these NLs and strongly couple with the neighboring quantum wells to form delocalized minibands, which help in vertical carrier transport and increase the tunneling current through the NLs.30 With the increase in Vapp, the Jleak values increased exponentially passing through current kinks/spikes, which are assigned to negative differential resistance (NDR) features in these ATA NL-based multiple QW structures.8,31 Further increase in Vapp promotes electron tunneling through the triangular potential barriers, resulting in an exponential increase in Jleak, as shown in Fig. 5(d),32 which supports the Fowler–Nordheim tunneling (FNT) transport in these NLs.33 Hence, the dominant leakage current mechanism and the observed NDR feature can be explained by considering the low bias trap-assisted and higher bias triangle barrier quantum mechanical tunneling mechanism through these subnanometric Al2O3 barrier layers.28,30
For the S0 NL, the higher Jleak and higher number of increasing current steps along with the diminished NDR feature were possibly owing to the soft breakdown of ultrathin (∼0.6 nm) barrier layers at low Vapp.27 By introducing 1 to 5 nm thick Al2O3 barrier layers across the NL/electrode interface, the Jleak reduced drastically from 1.18 × 10−5 to 2.11 × 10−9 @ 1 V. Considering the rapid current surge steps in each NL as the soft breakdown of Al2O3 barrier layers, the cumulative probability distributions of the electrical soft breakdown field strengths (Ebreak) of the 12 MIM devices from the S0, S1, S2 and S3 NL stacks are calculated and are displayed in the inset of Fig. 5(d). Although the distribution in Ebreak is relatively broad for the S0 NL, an improvement in the average distribution of the Ebreak values and also in the median values (from ∼0.3 MV cm−1 to 1.01 MV cm−1) can be observed with the increase in tAl2O3 from 1 to 5 nm. This improvement in the magnitude and distribution of Ebreak indicates an increase in the insulating properties and stability of the performance parameters of NLs.13 The overall fluctuation in the measured dielectric and electrical parameters observed from device to device in a given NL can be attributed to the thickness non-uniformity issues in PLD-grown thin films and/or the randomness in manually applied probe tip pressure.
In Table S1 of the ESI,† the important fundamental and derived dielectric performance parameters of our S0, S1, S2 and S3 NL stack-based MIM capacitors are compared with the literature reports on other metal oxide NL-based devices produced through different fabrication technologies and having different electrodes, device areas, layer thicknesses, etc. The device performance parameters for our optimised NL stacks seem to be comparable and/or superior, among previous reports, even with a relatively larger electrode area of ∼3.9 × 104 μm2. Although the series capacitances of the low dielectric constant Al2O3 barrier layers impose an adverse effect on the overall high dielectric constant, it is highly beneficial for the reduction in carrier loss and leakage current in these NLs. Hence, to maintain a sizeable M–W effect along with a simultaneous reduction of the tanδ and Jleak values in ATA NLs, an optimized tAl2O3 is needed before integrating these NLs into the final electronic circuits. The International Technology Roadmap of Semiconductors (ITRS)-2023 has set required device performance parameters for energy storage applications of MIM capacitors, such as a C/A > 13 fF um−2, Jleak < 1 × 10−8 A cm−2, equivalent oxide thickness (EOT) <2 nm, etc.8,34 Considering these requirements, the observed high C/A (∼33.2 fF μm−2), low tanδ (∼0.032), low EOT (∼1.04 nm), and reduced Jleak (∼3.08 × 10−7 A cm−2 at 1 V bias) values for the S2 NL capacitor make it highly suitable for high-density energy storage and advanced MIM capacitor applications. The volumetric energy density [U = (1/2) (C/(Ad)) (Eb)2] of this optimized NL structure (S2 NL) is computed to compare with the other dielectric thin film capacitors, documented in the literature. Using the values of C/A = ∼33.2 fF μm−2, d = 60 nm, and Eb = 0.65 MV cm−1, the energy density (U) value is calculated to be ∼4.6 J cm−3, which is on par with those reported for commercial dielectric thin film capacitors (1–2 J cm−3),35 lead lanthanum zirconated titanate (PLZT) dielectrics (<3 J cm−3),36 and nanocomposite multilayer capacitors (<2.5 J cm−3).37 The calculated U for our optimized ATA NL is not only superior to pure TiO2 (0.76 J cm−3) and Al2O3/TiO2 nanocomposites, (1.05 J cm−3),38 but also comparable to the volume energy density values reported for atomic layer deposited Al2O3/TiO2 NLs.39 The further increment in the calculated energy density values of our optimized NL structure can be achieved through improving the C/A and the Eb values, by increasing the NL interface density and employing high-work function electrodes like Au or Pt, respectively. Device performance optimization in this direction is ongoing.
Footnote |
† Electronic supplementary information (ESI) available: The interfacial analysis of representative ATA NLs is presented. Comparison of performance parameters of our optimised TiN/ATA NL/TiN capacitors with the previously reported NLs in the literature is also illustrated (PDF). Fig. S1. (a) Measured XRR curves for three representative ATA NLs deposited on a TiN coated Si substrate with Y-offsets. (b) Measured and fitted XRR profiles of [1A-1T-1A] NL grown on a TiN coated Si substrate and the depth variation of the electron density profile is displayed. (c) High-resolution transmission electron microscopy image and (d) the intensity variation profile of a representative [0.8A-0.8T-0.8T] NL. Table S1 Structural parameters of [0.8A-0.8T-0.8T] NL deposited on Si and TiN coated Si substrates extracted from best fit results of XRR data. Table S2: Comparison of the performance parameters of our PLD grown TiN/ATA NL/TiN capacitors with previously reported NLs in the literature. See DOI: https://doi.org/10.1039/d4lf00125g |
This journal is © The Royal Society of Chemistry 2024 |