Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Voltage triggered near-infrared light modulation using VO2 thin film

Open Access Open Access

Abstract

Development of compact and fast modulators of infrared light has garnered strong research interests in recent years due to their potential applications in communication, imaging, and sensing. In this study, electric field induced fast modulation near-infrared light caused by phase change in VO2 thin films grown on GaN suspended membranes has been reported. It was observed that metal insulator transition caused by temperature change or application of electric field, using an interdigitated finger geometry, resulted in 7% and 14% reduction in transmitted light intensity at near-infrared wavelengths of 790 and 1550 nm, respectively. Near-infrared light modulation has been demonstrated with voltage pulse widths down to 300 µs at 25 V magnitude. Finite element simulations performed on the suspended membrane modulator indicate a combination of the Joule heating and electric field is responsible for the phase transition.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Modulation of near-infrared (IR) light has broad applications ranging from photoacoustic (PA) spectroscopic sensing, optical communication and to biomedical imaging [13]. Since visible light does not penetrate deep into tissue due to scattering and absorption, near-IR light is used for PA imaging such improved penetration depths and high-quality image resolutions [4,5]. Similarly, due to lower absorption of near IR light in fiber optic material and specific wavelength of absorption of mid-IR light by molecules, they are used extensively in optical communication and molecular spectroscopy. In order to reduce measurement noise, applications in communication, imaging, and sensing technologies typically utilize modulated near-IR light and a lock-in based detection [68]. At present, intensity modulation techniques include optical-chopper systems using a slotted wheel, or electronically pulsed lasers, requiring a separate pulsing circuit [7,8]. However, modulation rates of mechanical choppers are limited to ∼10 kHz [9], while infrared lasers with built-in pulsing circuits are expensive and often bulky [7]. Therefore, these IR modulation platforms are not suitable for applications in portable and miniaturized systems, especially if a broad-wavelength source of IR light is utilized (i.e. mercury lamp). In recent years, graphene based metasurfaces, plasmonic nano particles, electrochromic and thermochromic materials have been extensively studied to develop reliable, compact, low-power near-IR modulators [10,11]. Among these technologies, vanadium dioxide (VO2), a well-known thermo- and electrochromic material, has emerged as one of the most promising candidates.

VO2, an interesting member of oxovanadates family, exhibits unique insulator metal transition (IMT) properties as monoclinic insulating phase of VO2 transforms into a rutile metallic phase near temperatures of 67-68 °C [1215]. In addition to temperature triggered IMT, the phase transition of VO2 can also be achieved with externally applied electric field and pressure pulses [12,16,17]. These IMT induced drastic resistance changes of VO2 thin films, which can reach up to 3-4 orders of magnitude, and take place in nanoseconds [18,19], have opened up potential applications in field effect transistors [20,21], memristors [22], and radiofrequency switching devices [23,24]. In addition to the electrical property change of VO2 films during IMT, optical transmittance and reflectance of the thin film, especially in near-IR and the IR regime, also changes drastically [25,26]. For the near and mid-IR part of the spectrum, the insulator phase of VO2 is quite transparent, while it is much more reflective in the metallic phase [2628]. Taking advantage of this behavior, optoelectronic devices utilizing IMT based modulation of light (caused by reversible phase change of VO2 thin films), such as near and mid-IR modulators for imaging and sensing applications, can be developed. While Si, SiO2, TiO2, quartz, and sapphire are commonly used as substrates to realize VO2 thin film based photonic devices [19,2932], investigations on VO2 thin film grown on wide-bandgap semiconducting materials is rather limited [3337]. Compared to traditional substrates, GaN film for VO2 growth offers several advantages stemming from the unique properties of III-Nitrides, including transparency at visible and infrared wavelengths, chemical inertness, and strong piezoelectric properties [3436]. Especially, the possibility of integration of a VO2 based solid-state optical modulator at infrared wavelengths with III-Nitride based surface acoustic wave (SAW) devices operating in GHz frequencies is very attractive, along with integration of III-Nitride microwave devices in the same substrate, offers interesting high frequency electro-optic modulation possibilities. Additionally, synthesis of high quality GaN films on Si (111) substrate is a mature technology (III-Nitride epilayers on 8-inch Si (111) substrate are routinely grown in high volume for light emitting diode applications), which also offers the possibility of easy integration with Si based integrated circuit in the same chip.

In this work, we have investigated for the first time electrically pulsed optical transmittance characteristics of a VO2 thin film, grown by oxidizing 70 nm vanadium deposited on c-plane GaN (thickness of 1.3 µm) membrane. Using interdigitated electrodes patterned on the suspended membrane, electrical and optical characteristics of the VO2 thin film with a radius of 500 µm were recorded at two near-IR wavelengths of 790 and 1550 nm. The IMT of the VO2 thin film was triggered with temperature as well as, dc and pulsed voltages applied to the interdigitated fingers. For the latter, modulation IMT transitions and hence optical modulation could be achieved using voltage pulses down to 300 µs. COMSOL based finite element simulations were also performed with varying voltages applied to the interdigitated fingers to investigate the IMT in the membrane.

2. Materials and methods

The GaN epitaxial wafers used in this study were purchased from NTT advanced technology corporation, Japan. The wafer had 1.3 µm i-GaN buffer layer on 675-750 µm thick Si (111) substrate. The modulator devices utilizing VO2 thin films on GaN membrane were fabricated using a traditional photolithographic fabrication process. First, GaN circular mesa structures with a diameter of 500 µm were defined on a substrate of 1.3 µm GaN epitaxial layer on Si(111) wafer, using an inductively coupled plasma (ICP) dry etching process. Then, 70 nm vanadium was deposited on top of the circular mesa using e-beam evaporation. A home-built low-pressure chemical vapor deposition system was used to oxidize the vanadium to VO2 following an optimized synthesis process reported elsewhere [33,34,38]. After the VO2 growth process, a set of 18 interdigitated metal fingers of Ti (20 nm) / Au (250 nm) stack were deposited on top of the membrane. The gap between the interdigitated fingers is 15 µm while the width of the metal electrodes is 10 µm. Finally, the GaN membrane was released from the Si substrate by through wafer etching of Si using an optimized Bosch process [34,38]. Schematic diagrams (top and side views) of the device structure are shown in Fig. 1(a), while a scanning electron micrograph (SEM) of the final fabricated device is shown in Figs. 1 (b). The x-ray diffraction (XRD) θ-2θ plot for the membrane is shown in Fig. 1(c), with inset showing a semi-log plot of the same. The peaks located at 2θ ≈ 28°, 52°, and 56° indicates the presence of VO2 [39].

 figure: Fig. 1.

Fig. 1. (a) Schematic of the VO2 thin film structure grown on epitaxial GaN membrane on Si (111) substrate. The top schematic illustrates a 3D view of the membrane with interdigitated metal fingers. The side view of the membrane is shown in the bottom schematic along with dimensions. (b) SEM image of VO2 on GaN circular membrane and interdigitated fingers. The diameter of the membrane is 500 µm (equal to the scale bar shown in the image). The metal fingers were 10 µm wide, with a gap of 15 µm between them. (c) XRD characteristics of the VO2 thin film on GaN membrane showing characteristics VO2 peaks. The inset shows the XRD intensity values in log scale to

Download Full Size | PDF

To investigate electrical and optical characteristics of the fabricated VO2 thin film membrane, an experimental setup as shown schematically in Fig. 2(a) was used. The chip containing VO2 on GaN membrane was attached to a ceramic chip carrier that was carefully placed on top of a ceramic heater. The chip carrier (with a circular opening at the bottom of 10 mm diameter) was positioned on the heater, which also had a hole of 6 mm diameter, and they were aligned carefully with the GaN membrane to enable transmission of the laser beam. Two near-IR lasers with wavelengths of 1550 nm (Alphanov) and 790 nm (Worldstartech Inc.) were used in the experiments. The collimated laser beams, with spot sizes of 20 µm (for 790 nm) and 100 µm (for 1550 nm), were focused vertically on the VO2 thin film using an x-y-z micropositioner. The transmitted laser power was measured using a photodiode (Newport 918D-IR) positioned underneath the heater, with its aperture properly aligned to capture the laser beam. Temperature measurements were carried using a data accusation system (Keysight 34972A) and a thermocouple (k type with a typical range of -200 to 1000 °C) attached to the chip carrier near the modulator chip. A computer-controlled source measurement unit (Keysight B2902A) was used to automatically apply the bias voltages to the interdigitated fingers through a 1 kΩ resistor to limit the current in the metallic state of VO2.

 figure: Fig. 2.

Fig. 2. (a) Schematic of the experimental setup used to measure optical modulation by the VO2 film using external voltage pulse. A series resistance of 1 kΩ was used in series with the voltage source to avoid high current related failures in the metallic state. (b) Variation in resistance and transmitted optical power near IMT in the VO2 thin film. The left axis shows resistance changes as the temperature slowly rises and drops (solid and dotted red lines, respectively), while the transmitted optical power changes at 1550 nm wavelength, with temperature rise and drop (blue solid and dotted lines, respectively) are shown on the right axis. Inset presents temperature characteristics of large area 70 nm VO2 thin film grown on GaN epitaxial film on Si substrate. (c) Effects of applied electric field on optical and electrical characteristics of the VO2 thin film. A reduction by 7% and 14% in transmitted optical power is observed at wavelengths of 790 nm (red) and 1550 nm (blue) occurs around applied voltage of ∼15.7 V, respectively, which results in the IMT. The inset shows resistance variations of VO2 thin film due to the applied field induced IMT.

Download Full Size | PDF

3. Results and discussion

At first, the temperature dependence of the electrical resistance and optical transmittance of the VO2 film under 1550 nm near-IR laser exposure was determined. As the temperature was increased from room temperature to 75 °C and then cooled down to room temperature, the resistance of VO2 membrane changed from 5 kΩ and 150 Ω (6 kΩ to 1.15 kΩ measured with the 1 kΩ series resistor RS) and vice-versa. The temperature-induced resistance changes of VO2 thin film are shown in Fig. 2(b), over the range of 30 to 75 °C are shown by red solid (for heating cycle) and dotted (for cooling cycle) lines on the left y-axis. A gradual resistance transition from insulating state to metallic state is observed around 67 °C for the heating process and around 62 °C for the cooling process. For comparison, we have included the resistance characterization results of a large area (1 cm × 1 cm) VO2 thin film grown (from 70 nm V deposition) on GaN substrate in inset of Fig. 2(b). As shown in the inset, the large area VO2 thin film resistance changes from 2 MΩ to 3 kΩ with much sharper transition rates, between 50-55 °C. The contrasting resistance characteristics between large and small area (current device) VO2 thin films might be due to several factors, including structural defects in the VO2 film grown on a small area, which can suffer from larger fluctuation of analyte concentrations over its surface [15,40,41].

Changes in the transmitted optical power, as the VO2 film underwent MIT, was measured using the formula $Optical\; Power\; Change(\%)= \; {\raise0.7ex\hbox{${({T - {T_{RT}}} )}$} \!\mathord{\left/ {\vphantom {{({T - {T_{RT}}} )} {{T_{RT}}}}} \right.}\!\lower0.7ex\hbox{${{T_{RT}}}$}} \times 100\; $ where T and TRT are the transmitted laser power at a given temperature, and at the room temperature, respectively. The results are shown as blue solid (heating cycle) and dotted (cooling cycle) lines on the right y-axis of Fig. 2(b). We find that the transmitted laser power (1550 nm) reduced approximately 25% when the temperature reached 75 °C. Unlike the gradual decreases in the resistance observed before and around the transition temperature of 67 °C, the optical transmittance remained relatively flat before the transition and changed sharply at the transition temperature. This is because the laser power heated the area on which it was focused, enabling quicker phase transition, unlike the electrical characteristics, which exhibited an aggregate of the transition of various areas (which could be non-uniform) over the entire membrane.

Following the initial characterization, we investigated electrical field induced IMT in the VO2 thin film. A variable voltage from 0.5 V to 20 V was applied to the interdigitated fingers on the VO2 thin film, while the resistance and relative optical transmittance changes were simultaneously measured. Figure 2(c) shows the percentage change in transmitted optical power of lasers with wavelengths of 790 nm (shown as red solid (forward sweep) and red dashed (reverse sweep) lines) and 1550 nm (displayed as blue solid (forward sweep) and blue dashed (reverse sweep) lines under applied electrical field. The change in transmittance was approximately 14% for 1550 nm laser and 7% for 790 nm laser. The 790 nm plot is somewhat noisier compared to the 1550 nm data as different responsivity values (chosen automatically based on input wavelength) was used by the photodetector to calculate incident optical power. Inset of Fig. 2(c) shows the resistance characteristics of VO2 membrane under applied electric field during the 1550 nm laser transmittance measurements. IMT of the VO2 thin film membrane was observed approximately at ∼15.7 V with upward bias sweeping configurations, whereas the film switched back from metal to insulator state at 12 V, under downward bias sweep. We note that the electrical and optical changes observed in Fig. 2(c) are much sharper compared to those in Fig. 2(b) near the IMT of VO2. This can be attributed to faster and more uniform transition process, over the entire VO2 film, enabled by application of electric field using the finger structure that was uniformly distributed, as compared to the heater pad, which raised the temperature of the various regions at a much slower rate causing various degrees of IMT over the various sections of the film.

After investigating IMT characterization experiments of the VO2 thin film membrane, we performed COMSOL simulations to understand the surface temperature characteristics of the VO2 thin film under application of electrical field. Figure 3(a) shows the temperature variation over the membrane when an electrical bias of 15 V is applied to one side of the electrodes while the other side was grounded. Figure 3(b) shows the temperature variation across a metal finger and a VO2 gap as various voltages (5, 10, 15, and 20 V) are applied to the interdigitated fingers. For the simulations, the conductivity of the VO2 film was taken as 13 S/m at the onset of phase transition, and the metal conductivity was 45.6×106 S/m. The conductivity was determined from the film resistance (Rf) of ∼1 K Ω at the onset of transition (2 kΩ as read from Fig. 2(c)–1 kΩ series resistance), the thickness (t) of the VO2 film of ∼70 nm (arising from a V deposition of 70 nm) [42], the finger gap (l) of 15 µm, and assuming an average width (w) of ∼400 µm. The conductivity σ is then calculated using the formula: σ $= l/({tw{R_{gap}}} )$, where Rgap is the resistance between one finger gap given as Rf × 17 (number of gaps). Putting in the numbers, we find σ = 13 S/m. From Fig. 3(b), we find the maximum temperature on the VO2 thin film membrane at 15 V bias is ∼45 °C, which is slightly below the transition temperature of ∼67 °C [43]. The temperature reaches its maximum value of ∼45 °C at the middle of the VO2 thin film section and reduces toward the metal contacts to ∼30 °C. The increase in temperature is not significant on the metal lines compared to the VO2 thin film stripes due to much lower resistance (causing less heat dissipation) and much higher thermal conductivity (helping to carry away the heat faster).

 figure: Fig. 3.

Fig. 3. (a) COMSOL simulations showing surface temperature variations over the circular VO2 on GaN thin film membrane with 500 µm diameter. The voltage applied across the metal fingers was 15 V. (b) Temperature variations across a VO2 thin film stripe and a metal finger simulated for various applied voltages. (c) COMSOL simulations showing electric field variations over the circular VO2 on GaN thin film membrane at 15 V applied bias. (d) Electric field variations across a VO2 thin film stripe and a metal contact for various applied voltages. The results shown in (b) and (d) are displayed along the same cut line.

Download Full Size | PDF

The electric field mapping on the VO2 thin film membrane at an applied bias of 15 V is shown in Fig. 3(c). The maximum value of the electric field on the top of the VO2 thin film membrane is found to be ∼106 V/m, which is very close to the average electric field expected from simply dividing the 15 V applied bias with the width of the VO2 strip of 15 µm. The simulated electric fields at applied voltages of 5, 10, 15, and 20 V are shown in Fig. 3(d) across a VO2 gap and a metal line. Noting that the critical electric field triggering phase transition process in VO2 thin films is ∼107 V/m [18,44], the experimentally observed phase transition of VO2 thin film on GaN membrane at an applied bias of 15.7 V is not likely to be solely driven by the applied electric field. Also, the ∼45 °C temperature rise along the VO2 thin film caused by Joule heating is not sufficient to trigger the phase transition, which occurs at ∼67 °C. Therefore, we conclude that a combination of electric field and Joule heating is responsible for phase transition in the VO2 thin film, with more significant role played by the electric field. This is in agreement with earlier studies reporting that electric field assisted phase transition is generally a combination of electric field effect and temperature rise due to Joule heating [14,43,45].

Taking advantage of the abrupt drop and jumps in the optical transmittance data, the VO2 membranes can be utilized to modulate near-IR light using a pulsed electric field. To experimentally demonstrate such optical modulation using applied electric field, pulsed voltage signals with varying pulse widths and magnitudes, and a constant duty cycle of 20%, were applied to the membrane, while the relative optical transmittance of the membrane under continuous near-IR laser exposure (790 or 1550 nm) was measured. Figures 4(a) – (c) show the optical transmittances of the VO2 film under various applied voltage pulses with pulse widths of 100 ms, 1 ms, and 300 µs, respectively. Figure 4(d) displays the magnitude of the applied voltage pulses with respect to pulse width, while the relative change in transmission of the laser power at wavelengths of 1550 nm (blue) and 790 nm (green) was kept approximately constant, assuring the complete phase transition as shown in inset of Fig. 4(d). As demonstrated in Fig. 4(d), while ∼16 V bias was utilized in the pulse range of 1 s – 10 ms for IMT induced transmitted optical power changes, which resulted in 14% and 7% reduction for 1550 nm and 790 nm, respectively, the required voltages to trigger phase transition were increased to 17.5 V and 25 V for the pulse widths of 1 ms and 300 µs, respectively. Due to Joule heating effects during application of the externally applied electric field, both electronic and electrothermal mechanisms contribute to IMT, in agreement with our COMSOL simulations and previous reports [14,46,47]. As the duration of the applied voltage was decreased, the contribution of Joule heating on facilitating IMT in VO2 thin film reduced. Consequently, the applied bias required to enable IMT in VO2 thin film increased with a reduction in pulse width. Such an increase led to application of higher electric field and also led to higher Joule heating.

 figure: Fig. 4.

Fig. 4. Electric field induced modulation of near-IR (1550 nm shown as blue, 790 nm shown as green, with powers of 0.38 and 0.22 mW, respectively) transmitted light using voltage pulses applied across the metal fingers on the VO2 on GaN thin film membrane, with pulse widths of (a) 100 ms, (b) 1 ms, (c) 300 µs (all with 20% duty cycle). (d) Voltage required for triggering IMT increases with reduction in voltage pulse width. Inset shows optical power changes ΔPopt corresponding to laser wavelengths of 1550 nm (blue) and 790 nm (green) maintained approximately constant at various voltage pulses.

Download Full Size | PDF

In order to verify the contribution of temperature on the electric field necessary to facilitate IMT, we measured the applied voltage required to cause a phase change in VO2 thin film at various temperatures ranging from 23 °C to 50 °C. As shown in Fig. 5, the voltage needed to trigger phase transition reduces as the membrane temperature increases. While the external voltage required for transition was 15.7 V at 23 °C displayed in Fig. 5(a), it decreased significantly to 12 V and 9.5 V when the film temperatures were increased to 40 °C, and then 50 °C as shown in Figs. 5 (b) and (c), respectively.

 figure: Fig. 5.

Fig. 5. Variation in optical power of 1550 nm laser transmitted through the VO2 membrane while increasing (solid lines) and decreasing (dotted lines) applied voltage bias, applied across the interdigitated metal fingers on the VO2 thin film on GaN membrane, at various temperatures of (a) 23 °C, (b) 40 °C, and (c) 50 °C. Increasing the VO2 film temperature resulted in lowering of the critical voltage required for its IMT transition. While the transition voltage is ∼15.7 V at 23 °C, it reduces to 12 V, and further to 9.5 V, as the temperature is increased to 40 °C and then to 50 °C.

Download Full Size | PDF

In the literature, investigations on changes in transmittance properties of VO2 thin films in near-IR and mid-IR regimes during IMT have been reported quite extensively [2628,36,48,49]. Hiltunen et al [28] demonstrated temperature triggered IMT in 70 nm VO2 thin film deposited on Al2O3 substrate causing light modulation from 49% to 14% at 1550 nm. Using VO2 nanoparticle film on ITO glass substrate, Li et al. also reported comparable modulation of 38.8% at 1550 nm [48]. In the mid IR range at 2.5 µm, much larger transmittance change of ∼50% was demonstrated by Nakano et al. for 50 nm VO2 thin film on TiO2 substrate, utilizing electrically and thermally induced IMT [27]. After an extensive search, we found only a handful of studies reporting VO2 thin film characterization on GaN film [3537], and only one reported measurement of optical modulation (∼50% observed at 1.8 µm wavelength), through thermal effect using conductive GaN as the heating layer [36]. Although larger modulation on VO2 films has been reported with superior electrical and optical modulation characteristics, most of the films are deposited on standard substrates like quartz, Si, SiO2, TiO2, glass or sapphire, or GaN/sapphire; using magnetron sputtering technique for film deposition [14,19,26,35]. However, for integration with mainstream Si devices and to reduce substrate costs, Si substrates are attractive, while growth on GaN entails its own favorable properties, including transparency to near-IR light, chemical inertness, and simple fabrication procedure of micro-size suspended membranes (by etching away the Si substrate under). In addition, direct oxidation of vanadium film in a CVD furnace is simpler and lower cost alternative to the magnetron sputtering technique. Indeed, this study is the first of its kind on demonstrating near-IR light modulation, with sub-ms response time, using VO2 film grown on GaN epitaxial layer, through direct application of a pulsed voltage across the VO2 layer to trigger IMT. The main reason behind for observing lower optical transmittance change in the VO2 films used in this study can be attributed to poorer material quality of the VO2 thin film grown over a microscale area. As pointed out before (inset of Fig. 2(b)), large area VO2 thin film on GaN epitaxial layer, grown by the same CVD technique, provides 3-4 orders of magnitude change in resistance at IMT, which is comparable to the best results reported in the literature [33,35]. Microscale dimension films of comparable quality can be expected to exhibit much larger percentage change in transmitted optical light intensity upon phase transition.

We note here that although nanosecond switching rates of electrically driven IMT phase of VO2 on various substrates has been demonstrated [18,19] using electrical bias, in our case, the IMT phase of the VO2 thin film on GaN membrane was prompted utilizing both electric field and Joule heating, which likely limited it’s operational speed, since Joule heating triggered IMT tends to be slower than non-thermal switching methods [14,50]. We also note that the only report of optical modulation of VO2 film on GaN/sapphire resulted in switching time on the order of seconds, which is much larger than the response time for our devices. The operating speed can be improved significantly by further improving the quality VO2 film (leading to quicker and abrupt transition), reducing the spatial distance between fingers (which will enhance electric field for same applied voltage), doping the VO2 film so that near room temperature operation, so that role of temperature change in the phase transition is minimized.

4. Conclusions

In conclusion, we have investigated electrical and near-IR optical transmittance characteristics near the insulator to metallic phase transition of VO2 thin film deposited on suspended GaN membrane on a Si substrate. IMT in the VO2 thin film, triggered by voltage pulses applied across interdigitated metal fingers, resulted in modulation of optical transmittances at multiple near-IR wavelengths, with sub ms time constant. COMSOL based finite element simulation indicated that the phase transition was triggered by a combination of Joule heating and applied electric field, which was corroborated by experimental observations.

Funding

National Science Foundation (IIP-1602006, ECCS-1809891).

Acknowledgments

This research was financially supported by the National Science Foundation, grants number ECCS-1809891 and IIP-1602006.

Disclosures

The authors declare no conflicts of interest.

Data availability

All relevant data is available from the corresponding author upon reasonable request.

References

1. J. Laufer, C. Elwell, D. Delpy, and P. Beard, “In vitro measurements of absolute blood oxygen saturation using pulsed near-infrared photoacoustic spectroscopy: accuracy and resolution,” Phys. Med. Biol. 50(18), 4409–4428 (2005). [CrossRef]  

2. P. K. Upputuri and M. Pramanik, “Photoacoustic imaging in the second near-infrared window: a review,” J. Biomed. Opt. 24(04), 1 (2019). [CrossRef]  

3. G. Li, S. Zhang, C. Guo, and S. Liu, “Absorption and electrochromic modulation of near-infrared light: realized by tungsten suboxide,” Nanoscale 8(18), 9861–9868 (2016). [CrossRef]  

4. D. Fried, R. E. Glena, J. D. Featherstone, and W. Seka, “Nature of light scattering in dental enamel and dentin at visible and near-infrared wavelengths,” Appl. Opt. 34(7), 1278–1285 (1995). [CrossRef]  

5. D. Jung, S. Park, C. Lee, and H. Kim, “Recent progress on near-infrared photoacoustic imaging: imaging modality and organic semiconducting agents,” Polymers 11(10), 1693 (2019). [CrossRef]  

6. P. Beard, “Biomedical photoacoustic imaging,” Interface Focus. 1(4), 602–631 (2011). [CrossRef]  

7. K. I. Maslov and L. V. Wang, “Photoacoustic imaging of biological tissue with intensity-modulated continuous-wave laser,” J. Biomed. Opt. 13(2), 024006 (2008). [CrossRef]  

8. P. K. Upputuri and M. Pramanik, “Performance characterization of low-cost, high-speed, portable pulsed laser diode photoacoustic tomography (PLD-PAT) system,” Biomed. Opt. Express 6(10), 4118–4129 (2015). [CrossRef]  

9. H. Chen, W. J. Padilla, M. J. Cich, A. K. Azad, R. D. Averitt, and A. J. Taylor, “A metamaterial solid-state terahertz phase modulator,” Nat. Photonics 3(3), 148–151 (2009). [CrossRef]  

10. N. Dabidian, I. Kholmanov, A. B. Khanikaev, K. Tatar, S. Trendafilov, S. H. Mousavi, C. Magnuson, R. S. Ruoff, and G. Shvets, “Electrical switching of infrared light using graphene integration with plasmonic Fano resonant metasurfaces,” ACS Photonics 2(2), 216–227 (2015). [CrossRef]  

11. Z. Li and N. Yu, “Modulation of mid-infrared light using graphene-metal plasmonic antennas,” Appl. Phys. Lett. 102(13), 131108 (2013). [CrossRef]  

12. G. Stefanovich, A. Pergament, and D. Stefanovich, “Electrical switching and Mott transition in VO2,” J. Phys.: Condens. Matter 12(41), 8837–8845 (2000). [CrossRef]  

13. M. M. Qazilbash, M. Brehm, B. G. Chae, P. C. Ho, G. O. Andreev, B. J. Kim, S. J. Yun, A. V. Balatsky, M. B. Maple, F. Keilmann, H. T. Kim, and D. N. Basov, “Mott transition in VO2 revealed by infrared spectroscopy and nano-imaging,” Science 318(5857), 1750–1753 (2007). [CrossRef]  

14. Y. Kalcheim, A. Camjayi, J. del Valle, P. Salev, M. Rozenberg, and I. K. Schuller, “Non-thermal resistive switching in Mott insulator nanowires,” Nat. Commun. 11(1), 2985 (2020). [CrossRef]  

15. R. R. Kumar, B. Karunagaran, D. Mangalaraj, S. K. Narayandass, P. Manoravi, and M. Joseph, “Characteristics of amorphous VO 2 thin films prepared by pulsed laser deposition,” J. Mater. Sci. 39(8), 2869–2871 (2004). [CrossRef]  

16. L. Fan, S. Chen, Z. Luo, Q. Liu, Y. Wu, L. Song, D. Ji, P. Wang, W. Chu, and C. Gao, “Strain dynamics of ultrathin VO2 film grown on TiO2 (001) and the associated phase transition modulation,” Nano Lett. 14(7), 4036–4043 (2014). [CrossRef]  

17. B. Hu, Y. Ding, W. Chen, D. Kulkarni, Y. Shen, V. V. Tsukruk, and Z. L. Wang, “External-strain induced insulating phase transition in VO2 nanobeam and its application as flexible strain sensor,” Adv. Mater. 22(45), 5134–5139 (2010). [CrossRef]  

18. Y. Zhou, X. Chen, C. Ko, Z. Yang, C. Mouli, and S. Ramanathan, “Voltage-triggered ultrafast phase transition in vanadium dioxide switches,” IEEE Electron Device Lett. 34(2), 220–222 (2013). [CrossRef]  

19. P. Markov, R. E. Marvel, H. J. Conley, K. J. Miller, R. F. Haglund Jr, and S. M. Weiss, “Optically monitored electrical switching in VO2,” ACS Photonics 2(8), 1175–1182 (2015). [CrossRef]  

20. S. Sengupta, K. Wang, K. Liu, A. K. Bhat, S. Dhara, J. Wu, and M. M. Deshmukh, “Field-effect modulation of conductance in VO2 nanobeam transistors with HfO2 as the gate dielectric,” Appl. Phys. Lett. 99(6), 062114 (2011). [CrossRef]  

21. N. Shukla, A. V. Thathachary, A. Agrawal, H. Paik, A. Aziz, D. G. Schlom, S. K. Gupta, R. Engel-Herbert, and S. Datta, “A steep-slope transistor based on abrupt electronic phase transition,” Nat. Commun. 6, 1–6 (2015). [CrossRef]  

22. S. Bae, S. Lee, H. Koo, L. Lin, B. H. Jo, C. Park, and Z. L. Wang, “The memristive properties of a single VO2 nanowire with switching controlled by self-heating,” Adv. Mater. 25(36), 5098–5103 (2013). [CrossRef]  

23. S. D. Ha, Y. Zhou, C. J. Fisher, S. Ramanathan, and J. P. Treadway, “Electrical switching dynamics and broadband microwave characteristics of VO2 radio frequency devices,” J. Appl. Phys. 113(18), 184501 (2013). [CrossRef]  

24. W. Li, M. Vaseem, S. Yang, and A. Shamim, “Flexible and reconfigurable radio frequency electronics realized by high-throughput screen printing of vanadium dioxide switches,” Microsyst. Nanoeng. 6(1), 1–12 (2020). [CrossRef]  

25. H. Wang, X. Yi, S. Chen, and X. Fu, “Fabrication of vanadium oxide micro-optical switches,” Sensors and Actuators A: Physical 122(1), 108–112 (2005). [CrossRef]  

26. Q. Hao, W. Li, H. Xu, J. Wang, Y. Yin, H. Wang, L. Ma, F. Ma, X. Jiang, and O. G. Schmidt, “VO2/TiN plasmonic thermochromic smart coatings for room-temperature applications,” Adv. Mater. 30(10), 1705421 (2018). [CrossRef]  

27. M. Nakano, K. Shibuya, N. Ogawa, T. Hatano, M. Kawasaki, Y. Iwasa, and Y. Tokura, “Infrared-sensitive electrochromic device based on VO2,” Appl. Phys. Lett. 103(15), 153503 (2013). [CrossRef]  

28. J. Hiltunen, J. Puustinen, A. Sitomaniemi, S. Pearce, M. Charlton, and J. Lappalainen, “Self-modulation of ultra-fast laser pulses with 1550 nm central wavelength in VO2 thin films,” Appl. Phys. Lett. 102(12), 121111 (2013). [CrossRef]  

29. J. D. Ryckman, K. A. Hallman, R. E. Marvel, R. F. Haglund, and S. M. Weiss, “Ultra-compact silicon photonic devices reconfigured by an optically induced semiconductor-to-metal transition,” Opt. Express 21(9), 10753–10763 (2013). [CrossRef]  

30. S. Cueff, J. John, Z. Zhang, J. Parra, J. Sun, R. Orobtchouk, S. Ramanathan, and P. Sanchis, “VO2 nanophotonics,” APL Photonics 5(11), 110901 (2020). [CrossRef]  

31. V. Devthade and S. Lee, “Synthesis of vanadium dioxide thin films and nanostructures,” J. Appl. Phys. 128(23), 231101 (2020). [CrossRef]  

32. Y. Zhang, W. Xiong, W. Chen, and Y. Zheng, “Recent Progress on Vanadium Dioxide Nanostructures and Devices: Fabrication, Properties, Applications and Perspectives,” Nanomaterials 11(2), 338 (2021). [CrossRef]  

33. S. Azad, R. Singh, M. Munna, F. Bayram, D. Khan, H. Li, and G. Koley, “Investigation of VO 2 Thin Film Grown on III-Nitride Epitaxial Layer,” in 2020 IEEE 20th International Conference on Nanotechnology (IEEE-NANO)Anonymous (IEEE, 2020).

34. F. Bayram, D. Gajula, D. Khan, and G. Koley, “Investigation of AlGaN/GaN HFET and VO2 Thin Film Based Deflection Transducers Embedded in GaN Microcantilevers,” Micromachines 11(9), 875 (2020). [CrossRef]  

35. Y. Zhou and S. Ramanathan, “Heteroepitaxial VO2 thin films on GaN: Structure and metal-insulator transition characteristics,” J. Appl. Phys. 112(7), 074114 (2012). [CrossRef]  

36. L. Fan, Y. Chen, Q. Liu, S. Chen, L. Zhu, Q. Meng, B. Wang, Q. Zhang, H. Ren, and C. Zou, “Infrared response and optoelectronic memory device fabrication based on epitaxial VO2 film,” ACS Appl. Mater. Interfaces 8(48), 32971–32977 (2016). [CrossRef]  

37. H. W. Yang, J. I. Sohn, J. H. Yang, J. E. Jang, S. N. Cha, J. Kim, and D. J. Kang, “Unusual M2-mediated metal-insulator transition in epitaxial VO2 thin films on GaN substrates,” EPL 109(2), 27004 (2015). [CrossRef]  

38. D. Gajula, F. Bayram, I. Jahangir, D. Khan, and G. Koley, “Dynamic response of VO 2 mesa based GaN microcantilevers for sensing applications,” Proceedings of the 2019 IEEE Sensors, Montreal, QC, Canada27–30 (2019).

39. D. Brassard, S. Fourmaux, M. Jean-Jacques, J. Kieffer, and M. El Khakani, “Grain size effect on the semiconductor-metal phase transition characteristics of magnetron-sputtered VO 2 thin films,” Appl. Phys. Lett. 87(5), 051910 (2005). [CrossRef]  

40. L. Fan, S. Chen, Y. Wu, F. Chen, W. Chu, X. Chen, C. Zou, and Z. Wu, “Growth and phase transition characteristics of pure M-phase VO2 epitaxial film prepared by oxide molecular beam epitaxy,” Appl. Phys. Lett. 103(13), 131914 (2013). [CrossRef]  

41. H. Kim, N. Charipar, M. Osofsky, S. Qadri, and A. Piqué, “Optimization of the semiconductor-metal transition in VO2 epitaxial thin films as a function of oxygen growth pressure,” Appl. Phys. Lett. 104(8), 081913 (2014). [CrossRef]  

42. C. O. Ba, S. T. Bah, M. D’Auteuil, V. Fortin, P. Ashrit, and R. Vallée, “VO2 thin films based active and passive thermochromic devices for energy management applications,” Current Applied Physics 14(11), 1531–1537 (2014). [CrossRef]  

43. Z. Yang, C. Ko, and S. Ramanathan, “Oxide electronics utilizing ultrafast metal-insulator transitions,” Annu. Rev. Mater. Res. 41(1), 337–367 (2011). [CrossRef]  

44. C. Ko and S. Ramanathan, “Observation of electric field-assisted phase transition in thin film vanadium oxide in a metal-oxide-semiconductor device geometry,” Appl. Phys. Lett. 93(25), 252101 (2008). [CrossRef]  

45. G. Liao, S. Chen, L. Fan, Y. Chen, X. Wang, H. Ren, Z. Zhang, and C. Zou, “Dynamically tracking the joule heating effect on the voltage induced metal-insulator transition in VO2 crystal film,” AIP Adv. 6(4), 045014 (2016). [CrossRef]  

46. E. Janod, J. Tranchant, B. Corraze, M. Querré, P. Stoliar, M. Rozenberg, T. Cren, D. Roditchev, V. T. Phuoc, and M. Besland, “Resistive switching in Mott insulators and correlated systems,” Adv. Funct. Mater. 25(40), 6287–6305 (2015). [CrossRef]  

47. I. Valmianski, P. Wang, S. Wang, J. G. Ramirez, S. Guénon, and I. K. Schuller, “Origin of the current-driven breakdown in vanadium oxides: Thermal versus electronic,” Phys. Rev. B 98(19), 195144 (2018). [CrossRef]  

48. M. Li, H. Wu, L. Zhong, H. Wang, Y. Luo, and G. Li, “Active and dynamic infrared switching of VO 2 (M) nanoparticle film on ITO glass,” Journal of Materials Chemistry C 4(8), 1579–1583 (2016). [CrossRef]  

49. S. Zhao, Y. Tao, Y. Chen, Y. Zhou, R. Li, L. Xie, A. Huang, P. Jin, and S. Ji, “Room-Temperature Synthesis of Inorganic–Organic Hybrid Coated VO2 Nanoparticles for Enhanced Durability and Flexible Temperature-Responsive Near-Infrared Modulator Application,” ACS Appl. Mater. Interfaces 11(10), 10254–10261 (2019). [CrossRef]  

50. J. Del Valle, P. Salev, F. Tesler, N. M. Vargas, Y. Kalcheim, P. Wang, J. Trastoy, M. Lee, G. Kassabian, and J. G. Ramírez, “Subthreshold firing in Mott nanodevices,” Nature 569(7756), 388–392 (2019). [CrossRef]  

Data availability

All relevant data is available from the corresponding author upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) Schematic of the VO2 thin film structure grown on epitaxial GaN membrane on Si (111) substrate. The top schematic illustrates a 3D view of the membrane with interdigitated metal fingers. The side view of the membrane is shown in the bottom schematic along with dimensions. (b) SEM image of VO2 on GaN circular membrane and interdigitated fingers. The diameter of the membrane is 500 µm (equal to the scale bar shown in the image). The metal fingers were 10 µm wide, with a gap of 15 µm between them. (c) XRD characteristics of the VO2 thin film on GaN membrane showing characteristics VO2 peaks. The inset shows the XRD intensity values in log scale to
Fig. 2.
Fig. 2. (a) Schematic of the experimental setup used to measure optical modulation by the VO2 film using external voltage pulse. A series resistance of 1 kΩ was used in series with the voltage source to avoid high current related failures in the metallic state. (b) Variation in resistance and transmitted optical power near IMT in the VO2 thin film. The left axis shows resistance changes as the temperature slowly rises and drops (solid and dotted red lines, respectively), while the transmitted optical power changes at 1550 nm wavelength, with temperature rise and drop (blue solid and dotted lines, respectively) are shown on the right axis. Inset presents temperature characteristics of large area 70 nm VO2 thin film grown on GaN epitaxial film on Si substrate. (c) Effects of applied electric field on optical and electrical characteristics of the VO2 thin film. A reduction by 7% and 14% in transmitted optical power is observed at wavelengths of 790 nm (red) and 1550 nm (blue) occurs around applied voltage of ∼15.7 V, respectively, which results in the IMT. The inset shows resistance variations of VO2 thin film due to the applied field induced IMT.
Fig. 3.
Fig. 3. (a) COMSOL simulations showing surface temperature variations over the circular VO2 on GaN thin film membrane with 500 µm diameter. The voltage applied across the metal fingers was 15 V. (b) Temperature variations across a VO2 thin film stripe and a metal finger simulated for various applied voltages. (c) COMSOL simulations showing electric field variations over the circular VO2 on GaN thin film membrane at 15 V applied bias. (d) Electric field variations across a VO2 thin film stripe and a metal contact for various applied voltages. The results shown in (b) and (d) are displayed along the same cut line.
Fig. 4.
Fig. 4. Electric field induced modulation of near-IR (1550 nm shown as blue, 790 nm shown as green, with powers of 0.38 and 0.22 mW, respectively) transmitted light using voltage pulses applied across the metal fingers on the VO2 on GaN thin film membrane, with pulse widths of (a) 100 ms, (b) 1 ms, (c) 300 µs (all with 20% duty cycle). (d) Voltage required for triggering IMT increases with reduction in voltage pulse width. Inset shows optical power changes ΔPopt corresponding to laser wavelengths of 1550 nm (blue) and 790 nm (green) maintained approximately constant at various voltage pulses.
Fig. 5.
Fig. 5. Variation in optical power of 1550 nm laser transmitted through the VO2 membrane while increasing (solid lines) and decreasing (dotted lines) applied voltage bias, applied across the interdigitated metal fingers on the VO2 thin film on GaN membrane, at various temperatures of (a) 23 °C, (b) 40 °C, and (c) 50 °C. Increasing the VO2 film temperature resulted in lowering of the critical voltage required for its IMT transition. While the transition voltage is ∼15.7 V at 23 °C, it reduces to 12 V, and further to 9.5 V, as the temperature is increased to 40 °C and then to 50 °C.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.