Abstract
Temporal and spatial tuning of the refractive index of optical thin films is desired for flat optics applications. The redistribution of mobile ions in mixed ionic-electronic conductors (MIEC) has been demonstrated to serve as a viable means for achieving optical tuning down to the nanoscale. Here we studied the dynamic range of the optical tuning achievable in the refractive index, in the MIEC oxide – Pr x Ce1−x O2−δ (PCO), for x = 0.1, 0.2 and 0.4, at 500 °C, by in-situ spectrophotometry. Significant increases in the modulation of both the imaginary and real optical constants in the visible and the adjacent spectra were obtained for increased doping levels. Device employing an electrochemical titration method was implemented to modulate the oxygen concentration, and thereby the optical transmission of PCO. Incorporation of a patterned top electrode allowed for the demonstration of spatial control of PCO thin film properties by in-situ video imaging of the optical switching process. The electrochemically induced optical state is shown to remain non-volatile upon quenching the device to room temperature under applied bias.
1 Introduction
Actively tunable optical devices require materials with adjustable optical properties e.g. phase change materials (PCMs) [1–4], liquid crystals [5–8], metal hydrides [9, 10], metal-insulator phase transition materials (e.g. VO2) [11–13], electrochromic polymers [14] and other examples [15–17]. Another option for achieving tunability in optical materials is via the use of mixed ionic-electronic conductors (MIECs) that exhibit electrochromic properties upon electric field driven cation (e.g. H+, Li+) intercalation. A prime example is the use of WO3 [18] and other oxides, for the implementation of light modulators [19, 20], dynamic color displays [21–23] and reprogrammable metasurfaces [24]. There is ongoing search for new tunable optical materials that rely on other physical principles and potentially provide complementary functionality to existing options. The authors previously proposed rare earth and transition MIEC metal oxides, with oxygen ions as the mobile species, as promising tunable optical materials, amenable to changes in their optical properties within the NIR-VIS-UV spectrum [24]. For example, initially transparent, wide band gap CeO2 (E g > 3 eV) becomes optically absorbing and tunable when the Ce cations are partially replaced by Pr [25, 26], which has a mixed 4+/3+ valency in the Pr x Ce1−x O2−δ (PCO) solid solution system. Only Pr in its most oxidized state (Pr4+) is optically absorbing. Successively reducing Pr4+ to Pr3+, via generation of positively charged oxygen vacancies and compensating electrons that localize on the Pr ions within the band gap, enables one to tune the optical absorption of PCO. The Pr dopant readily undergoes a redox reaction, under modest oxygen activity and temperature conditions, that enables modulation of the optically absorbing concentration over the whole doping range, that can reach as high as 40% in PCO. That change is comparable to the cation insertion levels achieved in intercalation materials [21]. However, electronic charges associated with intercalated cations tend to delocalize at high doping concentrations [27], causing significant unwanted optical losses, e.g. for implementation of phase optical actuators [28], especially in the NIR [19, 24]. Thermally tunable materials, e.g. PCM and VO2, that change to a metallic phase at a given transition temperature suffer from the excessive optical losses as well. While for PCO, strong electron localization, at energy levels associated with the observed optical transitions, persists even at very high concentrations, keeping the unwanted optical losses in the NIR region due to the free charge carriers very low. Further, PCM and VO2 materials are largely used for switching optical devices [12, 29] as they lack the ability to achieve continuous optical tuning due to abrupt changes in optical properties associated with thermally induced phase transitions. On the other hand, PCO enables gradual fine tuning of the refractive index by modifying the concentration of optically absorbing ions, providing a versatile material system for implementation of tunable optical components. Recent studies on dynamic color displays report relatively slow switching times, from minutes to hours, for metal hydrides [9, 30] and cation intercalation metal oxides [19, 21, 31]. Faster tuning of optical properties can be achieved in PCO based devices by utilizing thermally enhanced oxygen vacancy transport (activation energy of ∼0.8 eV) [32] at elevated temperatures, that can be also exploited to achieve non-volatility, i.e. long retention times, of the tuned state near room temperature. For fast optical tuning, the temperature of the whole device can be raised [29] as done in this study; for applications that require spatial control of the refractive index, e.g. pixels in a dynamic color display, they can be controlled in a local manner, e.g. by joule heating [2]. In this study switching times on the seconds scale were shown at 500 °C, with further improvements possible by scaling down device dimensions or by applying higher electrical fields [25].
We previously studied optical tuning in PCO by in-situ transmission spectrophotometry and were able to demonstrate that both the real (n r) and imaginary (k) parts of the PCO refractive index could be modified by varying its oxygen stoichiometry, δ, either by chemical means or locally by application of an electric field [25]. In these approaches, gradual changes in n r and k of the PCO refractive index were achieved, reaching as high as 0.1 (∼4% change) and 0.1 (nearly three orders of magnitude change), respectively. Here, we extend such studies by examining the effect of increasing Pr level (for x = 0.1, 0.2 and 0.4) on the dynamic range of the refractive index modulation. We also introduce a tunable optical transmission device (shown in Figure 1A), that employs electrochemical oxygen titration [33] (“oxygen pumping” [34]), to vary the oxygen stoichiometry of the PCO thin film, and thus its optical properties. The voltage controlled transmission device, as we demonstrate, enables access to wide dynamic changes in optical transmittance, exhibiting modulations up to 75% in a 150 nm PCO thin film. Effects of PCO composition on device optical characteristics were analyzed both under steady state and switching (non-steady state) voltage conditions. Finally, in-situ video imaging of the transmission modulator with patterned top electrode, and the subsequent image analysis, show that the electrochemical titration approach enables spatially localized control of PCO’s optical properties and that the electrically programmed optical state can be made non-volatile by quenching to room temperature.

Electrochemical optical modulation device structures.
(A) Cross section of four-layered stack of ITO/PCO/YSZ/Pt used to study the active optical tuning of PCO thin films with different compositions. Optical tuning, centered in the visible spectrum, is achieved via electrochemical titration of oxygen through the YSZ electrolyte (oxygen moves in YSZ and PCO via oxygen vacancies–V O) into and out of the actively tunable PCO layer (see text for details). (B) Conceptual design (cross section) of proposed electrochemically controlled all-solid-state one color device (pixel) for a non-volatile reflective display. The actively tunable PCO layer enables modulation of the device reflectance via the same mechanism as explored in this study in the transmission device.
2 Experimental
PCO powders with x = 0.1, 0.2 and 0.4, denoted here by PCO10, PCO20 and PCO40, respectively, were synthesized by a solution combustion method [35], starting from Ce(NO3)3·6H2O and Pr(NO3)3·6H2O precursors (99.99%, Alfa Aesar). Stoichiometric amounts of nitrate-based cation precursors were dissolved in deionized water and citric acid was added as a fuel for combustion. The combustion, initiated by heating, lasted for 2 h, eventually yielding a reddish PCO powder that was subsequently calcined in a box furnace at 750 °C for 6 h in air. The obtained powder was used to prepare dense ceramic targets for pulsed laser deposition (PLD), following the procedure reported elsewhere [36].
For optical transmission measurements PCO20 and PCO40 thin films (190 ± 5 nm thick) were grown by PLD on single crystal sapphire substrates (c-cut, double side polished, 10 × 10 × 0.43 mm3), following the procedure described elsewhere [25]. The complex refractive indices of the PCO compositions were obtained, first by measuring the optical transmission in a custom-built spectrophotometer, capable of in-situ measurements at 500 °C in a controlled gas atmosphere. This was followed by fitting the total transmittance of the thin films, with the aid of the “RefFIT” program developed by Kuzmenko [37]. Further details regarding the experimental setup and the fitting process can be found elsewhere [25].
The electrochemical titration device, illustrated in Figure 1A, was prepared for thin film optical transmission modulation and switching measurements by providing means to modulate the oxygen vacancy concentration, or equivalently δ, and thereby the oxidation state of Pr in the film. It is composed of a four-layer planar structure, ITO/PCO/YSZ/Pt, where the indium tin oxide–ITO (∼70 nm thick) and PCO (∼150 nm thick) double layer was grown by PLD (same conditions as above) on yttria stabilized zirconia (YSZ) substrates (surface orientation is 100; double side polished; dimensions – 10 × 10 × 0.5 mm3). The top ITO layer is optically transparent and serves, concurrently, as an ionically blocking and electronically ohmic electrode. On the back side of the YSZ substrates, a Pt electrode (a grid of 5 µm wide lines spaced by 40 µm) was prepared by a standard photolithography process, prior to the oxide depositions. The Pt grid electrode passes ∼90% of the incident light and supplies electrons for the electrochemically driven oxygen exchange at the Pt-YSZ-gas triple phase boundary that sustains the oxygen ion’s current flow through the device.
In-situ video imaging of the electrochemical titration device, in a controlled temperature and gas environment, was performed inside a closed mini-chamber with parallel sapphire windows (Linkam stage, Scientific Instruments) with the aid of a long working distance optical microscope (Mitutoyo) in transmission mode. The device was back lit with a white light source, that for improved contrast, was cut-off below 550 nm by a short-pass filter (Thorlabs). Image analysis was performed on the unprocessed video frames (Adobe Photoshop software).
3 Results and discussion
3.1 Effect of PCO composition on the tunable range of optical properties
An initial optical characterization of the PCO thin films, deposited onto sapphire substrates, was performed to investigate the dependence of the dynamic range of actively tunable optical properties (controlled by varying the oxygen activity in the PCO) on the Pr doping level, x. Figure 2A shows the optical transmittance of PCO20 and PCO40 thin films as a function of the equilibration pO2, measured in-situ at 500 °C. The covered pO2 range, from 1 to ∼10−23 bar (from pure O2 to 2000 ppm CO in CO2 mixture, respectively), spans nearly the whole redox range of the optically absorbing Pr4+ ion concentration,
![Figure 2:
Dependence of PCO optical properties on the equilibrium pO2.
(A) Optical transmittance of PCO20 (cross) and PCO40 (filled circle) at 500 °C, equilibrated at various pO2 (in bar units), as indicated in plot. Symbols – experimental data points, curves–modeled transmittance (see text for details). Bare Al2O3 substrate transmittance, ∼0.87, is shown as reference (grey curve). (B) n
r of the PCO compositions in the most reduced (dashed curves) and most oxidized states (solid curves). The PCO10 data was measured at 600 °C (adapted from reference [25]). (C) Change in the real part of the refractive index, Δn
r, between the most oxidized and reduced states in (B). (D) k – the extinction coefficient of PCO compositions corresponding to the n
r as shown in (B) (the same legend applies to the both plots) (E) Change in the extinction coefficient, Δk, between the most oxidized and reduced states in (C).](/document/doi/10.1515/nanoph-2022-0079/asset/graphic/j_nanoph-2022-0079_fig_002.jpg)
Dependence of PCO optical properties on the equilibrium pO2.
(A) Optical transmittance of PCO20 (cross) and PCO40 (filled circle) at 500 °C, equilibrated at various pO2 (in bar units), as indicated in plot. Symbols – experimental data points, curves–modeled transmittance (see text for details). Bare Al2O3 substrate transmittance, ∼0.87, is shown as reference (grey curve). (B) n r of the PCO compositions in the most reduced (dashed curves) and most oxidized states (solid curves). The PCO10 data was measured at 600 °C (adapted from reference [25]). (C) Change in the real part of the refractive index, Δn r, between the most oxidized and reduced states in (B). (D) k – the extinction coefficient of PCO compositions corresponding to the n r as shown in (B) (the same legend applies to the both plots) (E) Change in the extinction coefficient, Δk, between the most oxidized and reduced states in (C).
3.2 Steady state optical tuning of the PCO–electrochemical titration device
Active tuning of the optical constants in the PCO compositions was studied here by applying electrochemical oxygen titration [33] (“oxygen pumping” [34]) that augments the previously reported chemical and electro-tuning methods [25]. We developed a transmission modulation device, see Figure 1A and experimental section for details, that enables precise electrochemical control over the oxygen vacancy concentration, likewise the oxygen activity, inside a PCO thin film supported on the YSZ solid oxygen ion electrolyte. Figure 3A–C shows the optical transmittance modulation in the electrochemical devices for different PCO compositions. The devices were operated and characterized at 500 °C to enable sufficiently rapid surface oxygen exchange at the Pt/YSZ interface and subsequent transport through the YSZ bulk. Most of the transmittance change is due to amplitude attenuation as the k changes significantly with δ (and thus

Actively tuned transmittance of the ITO/PCO/YSZ/Pt device, shown in Figure 1A, measured in a controlled environment in a pO2 of 1 bar, at 500 °C. The actively tunable PCO layer reduced and oxidized by electrochemical titration with voltage biases, indicated in legends or plotted, were applied on the top ITO electrode. Equilibrium transmittances, under voltage bias, are shown for (A) PCO10, (B) PCO20, (C) PCO40 compositions and for (D) PCO20 under extremely reducing conditions. (E) Time relaxation of the optical transmittance (blue circles, lines are a guide to the eye), normalized by a most transparent state, at photon energy 2.5 eV (495 nm) for the PCO20 device. Switching is controlled by voltage bias indicated on the same plot (red line). Full range switching for PCO20 is shown, where transmittance changes by more than 50%. Switching time constants are τ on ∼ 3 s, τ off ∼ 25 s. (F) Transmittance switching in PCO20 device, lower modulation depth, ∼15% instead of 50% shown in (E) (τ on ∼ 1 s, τ off ∼ 25 s) (G) Five times faster transmittance off switching, than in (G), achieved by, at first, voltage overshoot and then the set to the new value, c.f. with the applied voltage in (G) (τ on ∼ 1 s, τ off ∼ 5 s) (H) A full range transmittance switching for the PCO40 device exhibits a modulation of more than 75% percent (τ on ∼ 5 s, τ off ∼ 10 s).
An extremely reduced PCO20 device (biased below −1.5 V), on the contrary, shows a significant reverse in transmittance change with increasing voltage, though with a somewhat different wavelength dependence, as shown in Figure 3D. The latter observation can be explained by noting that while nearly all of the Pr ions are fixed in a Pr3+ state at these high magnitudes of negative applied bias, and thus cannot contribute to further decreases in transmittance, the majority cation Ce that remains stable as Ce4+ to much lower pO2 that Pr4+, does begin to reduce to Ce3+ [42, 43], while Zr4+ ions in YSZ begin to reduce to Zr3+, forming “black zirconia” [44]. The latter, reported previously to exhibit a strong absorption in the VIS, may be the reason behind the reverse in the transmittance change upon strong electrochemical reduction.
3.3 Non-steady state optical switching–electrochemical titration device
Figure 3E–H show repeated optical switching of the transmission tunable devices by a rectangular waveform voltage (red line shown on the plot, below the transmittance), under the same atmosphere and temperature conditions as before. At 495 nm (2.5 eV), the transmittance of the PCO20 device is modulated by more than a factor of 2, shown in Figure 3E, by switching voltage repeatedly between 0.2 to −1.5 V. The ON (from low to high transmittance) and the OFF (from high to low) switching time constants are, ∼3 and 25 s, respectively. The switching times are asymmetric, with ON leading to faster and OFF to slower transients (values shown in caption of Figure 3). The device’s switching dynamics, in general, can be attributed to different rate determining processes controlling the oxidation and reduction of the PCO film e.g. the transport of oxygen through the YSZ thick conductor and exchange of oxygen with the gas phase at the Pt/YSZ interface [27]. Redistribution of oxygen through thin film PCO deposited on a thick electrolyte has been demonstrated, previously, not to be a rate limiting process [45]. For the PCO20 device the ON switching time decreases for lower transmittance modulation, c.f. Figure 3E and F, as less oxygen needs to be shifted in and out of the film. The OFF switching time can be further decreased, by temporarily applying an “overshoot” voltage followed by a new set value, as shown in Figure 3G. The device containing the PCO40 thin film switches at a similar speed, c.f. Figure 3F and Figure 3H, but provides an enhanced contrast, factor of 3, between the ON and OFF states. In all studied cases, the optical transmittance change is reversible and shows high repeatability in the voltage programmed state. The proposed conceptual design of a reflective device illustrated in Figure 1B, employing electrochemical oxygen titration as well, should enable faster switching speed, as the ionic conductor there is a thin film with a correspondingly faster bulk transport time constant. Additionally, oxygen ions are stored in a solid oxygen “reservoir” layer that would replace the need for the likely slow gas-solid interface reaction by a much faster solid-solid interface reaction, where oxygen ions are transferred between the two layers directly [46].
3.4 In-situ video imaging of electrochemical switching device with patterned ITO surface
For the experiments on spatially selective control of PCO optical properties, the ITO top layer of the ITO/PCO40/YSZ/Pt device used previously for electrochemical titration experiments, was patterned into two separate regions (ITO covered with patterned positive photoresist was etched with HCl acid). The middle region, with “K” letter shape, is separated from the rest of the ITO film (background area) by an etched groove, 50 µm wide. During switching experiments, shown in Figure 4, the background region was biased with respect to the Pt electrode on the back, while the K region was left floating (assuming that a current through an intact PCO40 layer is at least an order of magnitude less than through the ITO). Figure 4A shows a sequence of video frames taken in-situ during transmission switching of the device (from 2 to 12 s) and two additional frames (after 760 s and finally 16 h) taken in the thermally quenched-in state. At the initial equilibrium state, 0 V bias in pO2 of 1 bar at 500 °C, the whole device area has the same transmission (see Figure 4A). After a constant bias of −2 V is applied on the background region ITO electrode, its transmission gradually increases until a saturation level is reached after 12 s. In parallel, the transmission of the floating “K” region is also changing, but at slower rate, the change results due to non-perfect electrical insulation, for the present prototype device, e.g. current leaks from the PCO layer. Uneven change in the transmission of the biased versus floating regions, demonstrates that a spatial control of PCO optical constants can be achieved, in principle, by locally applied electrical fields. Eventually both regions reach the same steady state transmission after ∼12 s. Once the switching process is completed, the transmission of the devices increased by ∼3–7 times in the 2.3–3.2 eV range, as can be deduced from for the measurement taken on the same PCO40 device before the ITO layer was patterned, shown in Figure 3C. The thermally activated oxygen vacancies (or oxygen ions) transport in PCO [32, 38, 43] requires, on the one hand, elevated temperature for fast optical modulation, while on other hand, it enables implementation of non-volatility of the written state by thermal quenching. Here the non-volatility of a tuned transmission state is demonstrated by cooling the voltage biased device after switching, c.f. video frames in Figure 4A at 12 s, 760 s and 16 h, to room temperature. The transmission remains nearly the same following thermal quenching after 16 h, even though the device was kept in an atmosphere of pure O2, extremely far from the equilibrium state. Figure 4B visualizes the PCO40 thin film in the quenched initial state (here PCO40 on the Al2O3 substrate is shown), and the PCO40 device in the quenched final, electrochemically induced out-of-equilibrium state.

Patterned electrochemical switching device.
(A) Selected video frames (time indicated on images) of patterned ITO/PCO40/YSZ/Pt device taken during electrochemical transmission switching and in the quenched state. The sides of each frame are 1.6 mm long with microscope focused on device’s back side; dark vertical lines are back Pt electrodes. The switching changes the transmission of the background area (around the “K” region) at first, and is then followed by oxygen ion redistribution within the whole non-patterned PCO40 layer, including the “K” region (see text for details). Last two frames show the quenched device in the most transparent (ON) state that remains non-volatile, beyond 16 h. (B) Images of the PCO40 thin film, left – on the Al2O3 substrate, equilibrated in O2 at 500 °C (oxidized) and quenched to room temperature in the highly absorbing state; right – PCO40 thin film inside the ITO/PCO/YSZ/Pt device, electrochemically tuned to the most transparent state (fully reduced), 16 h after quenching. Note easily distinguishable color contrast between the two thin films that remains non-volatile. (C) Quantitative analysis, blue color histogram (BCH), of change in the video frames shown in (A), (only the background region was included in the analysis). (D) Quantitative comparison analysis of the transmission switching of the biased background and the floating “K” regions in the PCO40 device. The dominant color tone (y-axis) was calculated at BCH center that evolves with time (see text for details). An inset shows a zoom-in to the time period when the switching transition occurs, highlighting a difference in the rate of the color change of the background (biased) and “K” (floating) regions.
To quantify the spatial transmission changes of the device and their dynamics during the switching experiment, a color histogram analysis (using the blue component from the RGB color space, coded into the 0–255 range) of the video frames was implemented. Figure 4C shows evolution of the blue color histogram (BCH) during the transmission modulation, calculated for the background region only (see also Figure 4A). Initially, the PCO40 film is oxidized (more light absorbing) and the darker blue tones dominate (see in Figure 4A the video frame at 2 s), thus the corresponding BCH is centered around the 110 value, where it peaks. After the voltage bias is applied, the center value of BCH shifts to the lighter tones, though the change is not entirely homogeneous. Thus, the non-steady state BCHs are more smeared than those at the initial and final steady states (see Figure 4A). After 16 h, the peak position of the BCH of the quenched non-volatile state, stays close, but slightly below, the most transparent state of the device. Most of the observed change in color of the quenched state is due to a slight dependence of the PCO optical properties on temperature [25]. Analyzing a separate BCH of the background and the “K” regions, enables comparison of their switching dynamics, i.e. of the PCO40 under bias and the floating parts of the ITO electrode. The background and the “K” region BCHs were calculated from the video frames, e.g. the ones shown in Figure 4A. Next, Figure 4D shows the time evolution of the calculated BCH center values (approximated by the peak position) during and after the switching process. The analysis shows that following a rapid,
4 Conclusions
Tuning of the optical constant in the PCO compositions, with Pr doping levels as high as 0.4, was studied by varying the oxygen stoichiometry, δ, chemically, under oxidizing and reducing atmospheres, and by an electrochemical titration method operated in pure O2, at 500 °C. Calculating optical constants from the transmittance of the PCO thin films on sapphire substrate in the 1.1–3.5 eV range shows that increasing the Pr doping level extends the dynamic range of the refractive index modulation of both the real and the imaginary parts, while the latter scales nearly proportionally with x. Actively tunable transmission devices, based on electrochemical titration, were studied in-situ, at 500 °C, providing access to the equilibrium and non-steady state parameters related to transmission control of the PCO films, including the optical contrast, switching speed, reversibility and repeatability. Finally, the PCO40 device’s top ITO electrode was patterned to enable spatially selective optical tuning. In-situ video imaging during the electrochemical titration switching process, at elevated temperature, and following the quenching process, enabled quantitative analysis of the spatially selective optical tuning and non-volatility of the programmed, out-of-equilibrium, device state. The results of this study offer guidelines to achieve further enhancements in actively tunable transmission, and similar reflective devices, based on MIECs, e.g. as illustrated in Figure 1, with potential for future full color palette non-volatile display implementation.
Funding source: Basic Energy Sciences
Award Identifier / Grant number: DE-SC0002633
Funding source: National Science Foundation
Award Identifier / Grant number: DMR-1419807
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: We thank the U.S. Department of Energy, Basic Energy Sciences Program: “Chemomechanics of Far-From-Equilibrium Interfaces” (COFFEI), project DE-SC0002633 for funding. This work made use of the MRL Shared Experimental Facilities at MIT, supported by the National Science Foundation under award number DMR-1419807, and of MIT.nano clean lab facilities.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
[1] Y. Zhang, J. B. Chou, J. Li, et al.., “Broadband transparent optical phase change materials for high-performance nonvolatile photonics,” Nat. Commun., vol. 10, no. 1, pp. 1–9, 2019. https://doi.org/10.1038/s41467-019-12196-4.Search in Google Scholar PubMed PubMed Central
[2] Y. Zhang, C. Fowler, J. Liang, et al.., “Electrically reconfigurable non-volatile metasurface using low-loss optical phase-change material,” Nat. Nanotechnol., vol. 16, no. 6, pp. 661–666, 2021. https://doi.org/10.1038/s41565-021-00881-9.Search in Google Scholar PubMed
[3] Q. Wang, E. T. F. Rogers, B. Gholipour, et al.., “Optically reconfigurable metasurfaces and photonic devices based on phase change materials,” Nat. Photonics, vol. 10, no. 1, pp. 60–65, 2016. https://doi.org/10.1038/nphoton.2015.247.Search in Google Scholar
[4] W. Dong, Y. Qiu, X. Zhou, et al.., “Tunable mid-infrared phase-change metasurface,” Adv. Opt. Mater., vol. 6, no. 14, pp. 1–6, 2018. https://doi.org/10.1002/adom.201701346.Search in Google Scholar
[5] S. Ishii, A. V. Kildishev, V. M. Shalaev, and V. P. Drachev, “Controlling the wave focal structure of metallic nanoslit lenses with liquid crystals,” Laser Phys. Lett., vol. 8, no. 11, pp. 828–832, 2011. https://doi.org/10.1002/lapl.201110077.Search in Google Scholar
[6] B. Maune, R. Lawson, G. Gunn, A. Scherer, and L. Dalton, “Electrically tunable ring resonators incorporating nematic liquid crystals as cladding layers,” Appl. Phys. Lett., vol. 83, no. 23, pp. 4689–4691, 2003. https://doi.org/10.1063/1.1630370.Search in Google Scholar
[7] Y. J. Liu, G. Y. Si, E. S. P. Leong, N. Xiang, A. J. Danner, and J. H. Teng, “Light-driven plasmonic color filters by overlaying photoresponsive liquid crystals on gold annular aperture arrays,” Adv. Mater., vol. 24, no. 23, pp. 131–135, 2012. https://doi.org/10.1002/adma.201104440.Search in Google Scholar PubMed
[8] O. Buchnev, N. Podoliak, M. Kaczmarek, N. I. Zheludev, and V. A. Fedotov, “Electrically controlled nanostructured metasurface loaded with liquid crystal: toward multifunctional photonic switch,” Adv. Opt. Mater., vol. 3, no. 5, pp. 674–679, 2015. https://doi.org/10.1002/adom.201400494.Search in Google Scholar
[9] X. Duan, S. Kamin, and N. Liu, “Dynamic plasmonic colour display,” Nat. Commun., vol. 8, pp. 1–9, 2017. https://doi.org/10.1038/ncomms14606.Search in Google Scholar PubMed PubMed Central
[10] J. Li, S. Kamin, G. Zheng, F. Neubrech, S. Zhang, and N. Liu, “Addressable metasurfaces for dynamic holography and optical information encryption,” Sci. Adv., vol. 4, no. 6, pp. 1–8, 2018. https://doi.org/10.1126/sciadv.aar6768.Search in Google Scholar PubMed PubMed Central
[11] A. Tognazzi, A. Locatelli, M. A. Vincenti, C. Giannetti, and C. De Angelis, “Tunable optical antennas using vanadium dioxide metal-insulator phase transitions,” Plasmonics, vol. 14, no. 5, pp. 1283–1288, 2019. https://doi.org/10.1007/s11468-019-00917-w.Search in Google Scholar
[12] K. Dong, S. Hong, Y. Deng, et al.., “A lithography-free and field-programmable photonic metacanvas,” Adv. Mater., vol. 30, no. 5, pp. 1–7, 2018. https://doi.org/10.1002/adma.201703878.Search in Google Scholar PubMed
[13] M. J. Dicken, K. Aydin, I. M. Pryce, et al.., “Frequency tunable near-infrared metamaterials based on VO2 phase transition,” Opt. Express, vol. 17, no. 20, p. 18330, 2009. https://doi.org/10.1364/oe.17.018330.Search in Google Scholar PubMed
[14] T. Xu, E. C. Walter, A. Agrawal, et al.., “High-contrast and fast electrochromic switching enabled by plasmonics,” Nat. Commun., vol. 7, pp. 1–6, 2016. https://doi.org/10.1038/ncomms10479.Search in Google Scholar PubMed PubMed Central
[15] C. U. Hail, A. K. U. Michel, D. Poulikakos, and H. Eghlidi, “Optical metasurfaces: evolving from passive to adaptive,” Adv. Opt. Mater., vol. 7, no. 14, pp. 1–29, 2019. https://doi.org/10.1002/adom.201801786.Search in Google Scholar
[16] Y. Che, X. Wang, Q. Song, Y. Zhu, and S. Xiao, “Tunable optical metasurfaces enabled by multiple modulation mechanisms,” Nanophotonics, vol. 9, no. 15, pp. 4407–4431, 2020. https://doi.org/10.1515/nanoph-2020-0311.Search in Google Scholar
[17] L. Kang, R. P. Jenkins, and D. H. Werner, “Recent progress in active optical metasurfaces,” Adv. Opt. Mater., vol. 7, no. 14, pp. 1–26, 2019. https://doi.org/10.1002/adom.201801813.Search in Google Scholar
[18] C. G. Granqvist, “Electrochromics for smart windows: oxide-based thin films and devices,” Thin Solid Films, vol. 564, pp. 1–38, 2014. https://doi.org/10.1016/j.tsf.2014.02.002.Search in Google Scholar
[19] M. Muñoz-Castro, N. Walter, J. K. Prüßing, W. Pernice, and H. Bracht, “Self-holding optical actuator based on a mixed ionic-electronic conductor material,” ACS Photonics, vol. 6, no. 5, pp. 1182–1190, 2019. https://doi.org/10.1021/acsphotonics.8b01708.Search in Google Scholar
[20] E. Hopmann, B. N. Carnio, C. J. Firby, B. Y. Shahriar, and A. Y. Elezzabi, “Nanoscale All-solid-state plasmochromic waveguide nonresonant modulator,” Nano Lett., vol. 21, no. 5, pp. 1955–1961, 2021. https://doi.org/10.1021/acs.nanolett.0c04315.Search in Google Scholar PubMed
[21] Y. Lee, J. Yun, M. Seo, et al.., “Full-color-tunable nanophotonic device using electrochromic tungsten trioxide thin film,” Nano Lett., vol. 20, no. 8, pp. 6084–6090, 2020. https://doi.org/10.1021/acs.nanolett.0c02097.Search in Google Scholar PubMed
[22] Z. Wang, X. Wang, S. Cong, et al.., “Towards full-colour tunability of inorganic electrochromic devices using ultracompact fabry-perot nanocavities,” Nat. Commun., vol. 11, no. 1, pp. 1–9, 2020. https://doi.org/10.1038/s41467-019-14194-y.Search in Google Scholar PubMed PubMed Central
[23] M. Huang, A. J. Tan, F. Büttner, et al.., “Voltage-gated optics and plasmonics enabled by solid-state proton pumping,” Nat. Commun., vol. 10, no. 1, pp. 1–8, 2019. https://doi.org/10.1038/s41467-019-13131-3.Search in Google Scholar PubMed PubMed Central
[24] S. Zanotto, A. Blancato, A. Buchheit, et al.., “Metasurface reconfiguration through lithium-ion intercalation in a transition metal oxide,” Adv. Opt. Mater., vol. 5, no. 2, pp. 1–6, 2017. https://doi.org/10.1002/adom.201600732.Search in Google Scholar
[25] D. Kalaev and H. L. Tuller, “Active tuning of optical constants in the visible – UV : praseodymium-doped ceria — a model mixed ionic – electronic conductor,” Adv. Opt. Mater., vol. 2001934, pp. 1–16, 2021. https://doi.org/10.1002/adom.202001934.Search in Google Scholar
[26] J. J. Kim, S. R. Bishop, N. Thompson, Y. Kuru, and H. L. Tuller, “Optically derived energy band gap states of Pr in ceria,” Solid State Ionics, vol. 225, pp. 198–200, 2012. https://doi.org/10.1016/j.ssi.2012.03.047.Search in Google Scholar
[27] X. Yao, K. Klyukin, W. Lu, et al.., “Protonic solid-state electrochemical synapse for physical neural networks,” Nat. Commun., vol. 11, no. 1, pp. 1–10, 2020. https://doi.org/10.1038/s41467-020-16866-6.Search in Google Scholar PubMed PubMed Central
[28] S. Zanotto, F. Morichetti, and A. Melloni, “Fundamental limits on the losses of phase and amplitude optical actuators,” Laser Photon. Rev., vol. 9, no. 6, pp. 666–673, 2015. https://doi.org/10.1002/lpor.201500101.Search in Google Scholar
[29] M. Y. Shalaginov, S. An, Y. Zhang, et al.., “Reconfigurable all-dielectric metalens with diffraction-limited performance,” Nat. Commun., vol. 12, no. 1, pp. 1–8, 2021. https://doi.org/10.1038/s41467-021-21440-9.Search in Google Scholar PubMed PubMed Central
[30] X. Duan, S. T. White, Y. Cui, et al.., “Reconfigurable multistate optical systems enabled by VO2 phase transitions,” ACS Photonics, vol. 7, no. 11, pp. 2958–2965, 2020. https://doi.org/10.1021/acsphotonics.0c01241.Search in Google Scholar PubMed PubMed Central
[31] Y. Li, J. Van De Groep, A. A. Talin, and M. L. Brongersma, “Dynamic tuning of gap plasmon resonances using a solid-state electrochromic device,” Nano Lett., vol. 19, no. 11, pp. 7988–7995, 2019. https://doi.org/10.1021/acs.nanolett.9b03143.Search in Google Scholar PubMed
[32] D. Kalaev, T. Defferriere, C. Nicollet, T. Kadosh, and H. L. Tuller, “Dynamic current–voltage analysis of oxygen vacancy mobility in praseodymium-doped ceria over wide temperature limits,” Adv. Funct. Mater., vol. 30, no. 11, p. 1907402, 2020. https://doi.org/10.1002/adfm.201907402.Search in Google Scholar
[33] H. Rickert, Electrochemistry of Solids, vol. 7, Berlin, Heidelberg, Germany, Springer-Verlag, 1982.10.1007/978-3-642-68312-1Search in Google Scholar
[34] J. G. Swallow, J. J. Kim, J. M. Maloney, et al.., “Dynamic chemical expansion of thin-film non-stoichiometric oxides at extreme temperatures,” Nat. Mater., vol. 16, no. 7, pp. 749–754, 2017. https://doi.org/10.1038/nmat4898.Search in Google Scholar PubMed
[35] A. Varma, A. S. Mukasyan, A. S. Rogachev, and K. V. Manukyan, “Solution combustion synthesis of nanoscale materials,” Chem. Rev., vol. 116, no. 23, pp. 14493–14586, 2016. https://doi.org/10.1021/acs.chemrev.6b00279.Search in Google Scholar PubMed
[36] H. G. Seo, Y. Choi, and W. C. Jung, “Exceptionally enhanced electrode activity of (Pr,Ce)O2−δ-based cathodes for thin-film solid oxide fuel cells,” Adv. Energy Mater., vol. 8, no. 19, pp. 1–7, 2018. https://doi.org/10.1002/aenm.201703647.Search in Google Scholar
[37] A. B. Kuzmenko, “Kramers-Kronig constrained variational analysis of optical spectra,” Rev. Sci. Instrum., vol. 76, no. 8, pp. 1–9, 2005. https://doi.org/10.1063/1.1979470.Search in Google Scholar
[38] T. S. Stefanik and H. L. Tuller, “Nonstoichiometry and defect chemistry in praseodymium-cerium oxide,” J. Electroceram., vol. 13, nos. 1–3, pp. 799–803, 2004. https://doi.org/10.1007/s10832-004-5195-7.Search in Google Scholar
[39] D. P. Fagg, J. R. Frade, V. V. Kharton, and I. P. Marozau, “The defect chemistry of Ce(Pr, Zr)O2-δ,” J. Solid State Chem., vol. 179, no. 5, pp. 1469–1477, 2006. https://doi.org/10.1016/j.jssc.2006.01.052.Search in Google Scholar
[40] J. Maier, Physical Chemistry of Ionic Materials: Ions and Electrons in Solids, vol. 1, Chichester, England, John Wiley & Sons, 2004.10.1002/0470020229Search in Google Scholar
[41] I. Riess, “Current-voltage relation and charge distribution in mixed ionic electronic solid conductors,” J. Phys. Chem. Solid., vol. 47, no. 2, pp. 129–138, 1986. https://doi.org/10.1016/0022-3697(86)90121-6.Search in Google Scholar
[42] D. Chen, S. R. Bishop, and H. L. Tuller, “Nonstoichiometry in oxide thin films operating under anodic conditions: a chemical capacitance study of the praseodymium-cerium oxide system,” Chem. Mater., vol. 26, no. 22, pp. 6622–6627, 2014. https://doi.org/10.1021/cm503440v.Search in Google Scholar
[43] S. R. Bishop, T. S. Stefanik, and H. L. Tuller, “Electrical conductivity and defect equilibria of Pr0.1Ce 0.9O2-δ,” Phys. Chem. Chem. Phys., vol. 13, no. 21, pp. 10165–10173, 2011. https://doi.org/10.1039/c0cp02920c.Search in Google Scholar PubMed
[44] A. Sinhamahapatra, J. P. Jeon, J. Kang, B. Han, and J. S. Yu, “Oxygen-deficient zirconia (ZrO2-x): a new material for solar light absorption,” Sci. Rep., vol. 6, no. April, pp. 1–8, 2016. https://doi.org/10.1038/srep27218.Search in Google Scholar PubMed PubMed Central
[45] D. Chen and H. L. Tuller, “Voltage-controlled nonstoichiometry in oxide thin films: Pr0.1Ce0.9O2-δ case study,” Adv. Funct. Mater., vol. 24, no. 48, pp. 7638–7644, 2014. https://doi.org/10.1002/adfm.201402050.Search in Google Scholar
[46] I. Riess and J. Schoonman, “Electrochemistry of mixed ionic-electronic conductors,” in CRC Handbook of Solid State Electrochemistry, P. J. Gellings and H. J. M. Bouwmeester, Eds., Boca Raton, FL, CRC Press, Inc, 1997.10.1201/9781420049305.ch7Search in Google Scholar
© 2022 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.