Tunable Casimir equilibria with phase change materials: from quantum trapping to its release
Abstract
A stable suspension of nanoscale particles due to the Casimir force is of great interest for many applications such as sensing, non-contract nano-machines. However, the suspension properties are difficult to change once the devices are fabricated. Vanadium dioxide (VO2) is a phase change material, which undergoes a transition from a low-temperature insulating phase to a high-temperature metallic phase around a temperature of 340 K. In this work, we study Casimir forces between a nanoplate (gold or Teflon) and a layered structure containing a VO2 film. It is found that stable Casimir suspensions of nanoplates can be realized in a liquid environment, and the equilibrium distances are determined, not only by the layer thicknesses but also by the matter phases of VO2. Under proper designs, a switch from quantum trapping of the gold nanoplate (“on” state) to its release (“off” state) as a result of the metal-to-insulator transition of VO2, is revealed. On the other hand, the quantum trapping and release of a Teflon nanoplate is found under the insulator-to-metal transition of VO2. Our findings offer the possibility of designing switchable devices for applications in micro-and nano-electromechanical systems.
I Introduction
Micro- and nano-electromechanical systems (MEMS and NEMS), which integrate electrical and mechanical functionality on the micro- and nano-scales, have attracted enormous attention Lyshevski (2018); Craighead (2000). Thanks to small sizes, the MEMS and NEMS exhibit low mass, high mechanical resonance frequencies and quantum effects, leading to a broad range of applications such as biological/chemical detections Eom et al. (2011), accelerometers Xu et al. (2011) and micro/nanomachines Wang (2013). One major problem in MEMS and NEMS is the which makes the systems collapse and permanent adhesion caused by the attractive Casimir forces Buks and Roukes (2001); Chan et al. (2001). The Casimir force is a macroscopic quantum effect which arises from quantum fluctuations of the electromagnetic field Casimir (1948). In most cases, two neutral, parallel plates consisted of the same materials are attractive to each other, and the magnitudes of the attraction depend on several parameters such as separations, geometric thicknesses, finite conductivities and temperatures (see, e.g., the review Klimchitskaya et al. (2009) and Refs.Yampol’skii et al. (2008, 2010). Therefore, repulsive Casimir forces are highly required for non-contact and low-friction MEMS and NEMS. The repulsive Casimir forces have been intensively studied in many systems Woods et al. (2016) including liquid-separated environments Munday et al. (2009); van Zwol and Palasantzas (2010); Phan and Viet (2011); Dou et al. (2014), meta-materials Rosa et al. (2008); Zhao et al. (2009, 2011); Song et al. (2018), topological insulators Grushin and Cortijo (2011); Chen and Wan (2012); Nie et al. (2013) and specific geometrics Tang et al. (2017); Levin et al. (2010). In addition, the concept of Casimir equilibria was also investigated, using the enclosed geometries Rodriguez et al. (2008); Rahi and Zaheer (2010) and dispersive materials Rodriguez et al. (2010a). Lately, stable Casimir equilibria of nanoplates above a Teflon-coated gold substrate were reported by Zhao et al Zhao et al. (2019). However, the Casimir equilibria of previous studies were mainly in passive systems. Once the devices are fabricated, the trapping properties are difficult to change. Thus, the tunable trapping or even the switching from the trapping to its release by external stimuli (e.g., heating, electric fields or optical waves) is highly desired in MEMS and NEMS.
In order to active modulate the Casimir effect, one straight way is to change the dielectric properties of materials under external means Torricelli et al. (2012); Sedighi et al. (2013); Torricelli et al. (2010). Vanadium dioxide (VO2) Shao et al. (2018); Zylbersztejn and Mott (1975) is a phase change material(PCM), which undergoes a transition from a low-temperature insulating phase to a high-temperature metallic phase at critical temperature 340 K. The phase transition of VO2 is accompanied by a structural transformation from the monoclinic phase to the tetragonal one. Meanwhile, the dielectric function of VO2 changes dramatically during the phase transition, leading to many interesting applications Wu et al. (2017); Liu et al. (2017); Kats et al. (2012); van Zwol et al. (2012). In general, the phase transition of VO2 can be induced by changing the temperature of systems. Alternatively, the phase transition can be driven by optical lasers Cavalleri et al. (2001); Rini et al. (2008) or electrical gratings Qazilbash et al. (2008); Nakano et al. (2012) on a sub-picosecond timescale. Recently, VO2 has been employed to study the tunable Casimir effect in the vacuum Galkina et al. (2009); Pirozhenko and Lambrecht (2008); Castillo-Garza et al. (2007). For a large separation (e.g., 1 m), the contrast of Casimir forces due to the phase-transition is quite large (e.g., over 2 times for two semi-infinite plates of VO2, this value could be even larger for the case of finite thickness Galkina et al. (2009); Pirozhenko and Lambrecht (2008)). As the separation is small (e.g., 100 nm), however, the modulation of Casimir forces owning to the phase transition and finite-thickness decreases greatly Pirozhenko and Lambrecht (2008); Castillo-Garza et al. (2007). Nonetheless, the Casimir forces are always attractive and only magnitude modulations have been reported in a vacuum-separated configuration. The influences of phase transition of VO2 on the sign modulation of Casimir forces (e.g., from attraction to repulsion) are yet less explored. In a liquid environment, the function of sign modulation and the related phenomena such as tunable Casimir equilibria are expected based on the phase transition of VO2.
Here, the Casimir forces between a nanoplate and a layered structure separated by a liquid are investigated. The layered structure consists of two kinds of materials, i.e., Vanadium dioxide (VO2) and Teflon. It is found that stable Casimir equilibria of gold nanoplates can be realized when a VO2 film is buried under a semi-infinite Teflon. The properties of Casimir equilibria are determined, not only by the layer thicknesses but also by the matter phases of VO2. For thick-film VO2, the Casimir equilibria and quantum traps can be achieved for both the metallic and insulating phases. On the other hand, a switch from quantum trapping of the gold nanoplate(“on” state) to its release (“off” state) can be triggered by the metal-to-insulator phase transition when the thickness of VO2 is thin (e.g., 20 nm). Finally, stable suspensions of Teflon nanoplates are also proposed with a complementary design, where the Teflon substrate is coated by a VO2 film. Unlike the case of gold nanoplates, the quantum trapping of Teflon nanoplates and its release correspond to the insulating and metallic phases of VO2. Moreover, the switching phenomena can be realized only with a several-nanometers thickness of VO2.
II Theoretical models
The system in this work is schematically shown in Fig. 1(a), where a gold nanoplate with thickness is suspended in a liquid of bromobenzene. The separation between the nanoplate and the substrate is . The substrate is composed of a VO2 film buried under a semi-infinite plate of Teflon. The thicknesses of the top-layer Teflon and VO2 are denoted as and , respectively. The in-plane dimension of the gold nanoplate is much larger than and , and it is considered as a slab during our calculations. The Casimir force is calculated by , where is the Casimir energy between the gold nanoplate and the substrate, having the form Nie et al. (2013); Zhao et al. (2019)
(1) |
where is the reduced Planck constant, is the in-plane area, is the parallel wavevector, is the vertical wavevector, is the speed of light in vacuum, is the permittivity of the intervening liquid evaluated with imaginary frequency , is the reflection matrix for layered structures, having the form
(2) |
where with =1 and =2 are the reflection coefficients for the upper and lower layered structures, and the superscripts and correspond to the polarizations of transverse electric () and transverse magnetic () modes, respectively. Note that the temperature for Eq. (1) equals 0 K and it is an effective approximation as the separation is smaller than 1 for finite temperatures Milton (2004). For a nanoplate suspended in a liquid, the reflection coefficients can be given analytically as follows Zhao et al. (2011)
(3) |
where and , is the thickness of the nanoplate, is the vertical wavevector, is the permittivity of the nanoplate. The subscripts of represent the light is incident from the medium to (0 means the liquid).
Alternatively, the reflection coefficients for layered structures can be calculated by a transfer matrix method. The general form is given as , where and are the elements of the matrixZhan et al. (2013). The matrix is the multiplications of transmission matrices across different interfaces and propagation matrices in different layers. Considering an arbitrary -layer system, the -matrix is given as :
(4) |
where the transmission matrix is given as:
(5) |
where for p-polarization and for s-polarization. The propagation matric in the -th layer (for both and polarizations) is written as:
(6) |
For example, we have for the multilayered substrate in Fig. 1. The matrix is given by , where the subscripts 0, 1, 2 and 3 represent the media of liquid, Teflon, VO2 and Teflon (from top to down); the thicknesses , .
III Results and discussions
Figure 1(b) shows the permittivity for different materials, where the used models and parameters are given in the Appendixes. The dielectric function of VO2 changes dramatically under different temperatures. For temperature , VO2 is in the metallic phase and it acts as a poor metal. For , it is in the insulating phase (or called semiconducting phase), and the corresponding dielectric function nearly matches that of intrinsic silicon at low frequency Pirozhenko and Lambrecht (2008). To create repulsive Casimir forces between two dissimilar plates separated by a liquid, the permittivity should satisfy for a vast range of frequency Munday et al. (2009). Clearly, the dielectric functions of gold and VO2 (either metallic or insulating phase) are larger than that of bromobenzene over a wide range of frequency. Therefore, the Casimir force is always attractive for the layered structure of gold/bromobenzene/VO2. While the Casimir force for the structure of gold/bromobenzene/Teflon is repulsive instead. Nonetheless, the Casimir equilibria can not be found for above two layered structures.
III.1 Tunable Casimir equilibria for gold nanoplates
Now we consider the Casimir forces as the substrate is composed of a VO2 film and Teflon (see Fig. 1(a)). The Casimir pressure () for the thick film of VO2 is given in Fig. 2(a). The results show that the curves are almost identical for =200, 500 and 1000 nm, indicating the weak impact of the thickness for thick-film configurations. The pressure is repulsive at small separation (e.g., nm), making the nanoplate stay away from the substrate. As the separation increases further, the Casimir equilibria (zero pressure) occur and quantum traps can be realized for both metallic (solid lines) and insulating phases (dashed lines). In addition, the equilibrium distance is shifted under the phase transition of VO2. On the other hand, the thin-film thickness and the phase transition of VO2 can play an important role in Casimir pressure as shown in Fig. 2(b). For the thickness =10 and 20 nm, quantum traps can be realized for the metallic phase, whereas no trap is found for the insulating phase. Under such configurations, a switch from quantum trapping of the nanoplate(“on” state) to its release (“off” state) can be triggered by the metal-insulator transition of VO2. However, the quantum trapping occurs for both metallic and insulating phases as the thickness increases to 30 nm, and the “off” state disappears. Compared with the vacuum-separated configuration Castillo-Garza et al. (2007), not only the magnitude of Casimir forces can be modified in a liquid environment, but also the sign could be switched (e.g., from attraction to repulsion for =100 nm, =30 nm), due to the phase-transition of VO2.
To understand the switch transition from the “on” to the “off” state, the contour plots of Casimir pressure are shown in Fig. 3 under different separations. The sign of pressure is determined by the competition of VO2 film (attraction) and low-refractive-index Teflon (repulsion). For small separation =30 nm, the pressure is dominant by the repulsive component as shown in Figs. 3(a) and 3(b). For the metallic phase, the attractive component increases and it compensates the repulsive one as the separation becomes 85 nm (), resulting in Casimir equilibrium (see Fig. 3(c)). While the repulsion is still dominant for the insulating phase as shown in Fig. 3(d). As increases further to 150 nm, the Casimir pressure turns out to be dominantly attractive in Fig. 3(e) for the metallic phase, resulting in a restoring force for stable trapping. By contrast, the pressure is still dominant by repulsion for the insulating phase as shown in Fig. 3(f). The pressure maps between the metallic and insulating phases are almost identical for large energy (e.g., 2 eV), whereas the discrepancy manifests at low energy. The results indicate that the attractive component appears only at low frequency and small vector for metallic VO2, where the field cannot penetrate the metalZhao et al. (2019). Conversely, the field can penetrate the thin-film of insulating VO2 easily, leading to repulsive Casimir forces.
Practically, the influences of gravitation and buoyancy on the force balances should be taken into account. The condition for the force equilibrium is written as =0, where is the unit vector normal to the surface, is the sum of gravity and buoyancy, is the gravitational acceleration, 19.3 g/cm3 and 1.50 g/cm3 is the density of gold and liquid bromobenzene, respectively. The magnitude of is about 7.0 mPa as the thickness =40 nm. Three types of configurations are depicted in the inset of Fig. 4(a) for the cross-section views. The type I configuration corresponds to a zero-projection (or weightlessness in aerospace), where the switching from quantum trapping (metallic state) to its release (insulating state) can be obtained as in a proper range, from about 2 to 22 nm. For type II configuration, the attractive can compensate the long-range repulsive Casimir force at large , leading to stable suspensions for both and . However, the equilibrium distances are different, and it can be inferred that the stiffness of trapping for metallic phase is stronger than that of the insulating phase. For type III configuration (a flipped down system), the switching between trapping and its release can also be realized. Interestingly, there are two equilibrium distances for this configuration. It is not difficult to know that the smaller equilibrium distance (solid lines) is stable, whereas the other one (dashed lines) with larger distance is unstable to small perturbations in position. For both type II and III configurations, the deviations from Type I become strong as is large.
In addition to the thickness of VO2 film, the top-layer Teflon can also play a significant role in the Casimir effect. The plots of Casimir pressure via the thicknesses of the coating Teflon are shown in Figs. 4(b) and 4(c), where =20 nm is fixed. The results show that the switching between quantum trapping and it release occurs only when is larger than about 42 nm (no gravity). The larger the , the larger of the position for the Casimir equilibrium. As is smaller than 42 nm, the equilibrium distance is also small, and quantum trappings can be realized for both metallic and insulating phases. For comparison, the gravitation and buoyancy are taken into account. Again, strong discrepancies among three configurations occur as the equilibrium positions larger than about 150 nm, resulting from the comparable magnitude of and the Casimir force. The impact of can be further reduced by decreasing the thickness near the skin depth (about 22 nm) Lisanti et al. (2005).
III.2 Tunable Casimir equilibria for Teflon nanoplates
The active control of the low-refractive-index nanoplates can also be significant in many applications. Inspiring by the work Zhao et al. (2019), a complementary design is schematically shown in the inset of Fig. 5(a). A Teflon nanoplate is suspended in a liquid of bromobenzene, and the substrate is a semi-infinite plate of Teflon coated by a VO2 film (high refractive index). Under such design, the Casimir force is repulsive at very short separation, due to the dominant interaction between Teflon/bromobenzene/VO2. As the separation increases, the attractive interaction from Teflon/bromobenzene/Teflon can be dominant instead, resulting in a stable Casimir trapping. To verify the design, the Casimir pressure is given quantitatively in Figs. 5(a) and 5(b) as a function of separation. Interestingly, the Casimir pressure shows a long-range repulsive behavior for the metallic VO2, which corresponds to the “off” state. The repulsion pressure becomes stronger as the thickness enlarges from 2 to 6 nm. For = 2 nm, a Casimir equilibria and strong restoring forces can be found when VO2 is in the insulating phase. Therefore, the quantum trapping and release of a Teflon nanoplate can be achieved under the insulator-to-metal transition of VO2. As the thickness is 4 nm, the restoring force decreases and the trapping stiffness drops considerably. The calculation results indicate that the Casimir pressure is quite sensitive to the thickness of VO2. Due to the low density of Teflon (2.1 g/cm3), the pressure for the Teflon nanoplate is about 0.6 mPa, which is reduced significantly compared with those of gold nanoplates.
III.3 Finite temperatures effect
To achieve the phase transition of VO2, the temperatures of the devices need to be changed. We assume that the dielectric functions of the gold and Teflon are temperature-independent. For organic liquids, the change of refractive index due to the temperature Li et al. (1994) is an order of K, and the permittivity of bromobenzene is also treated as temperature-independent. Nonetheless, it is interesting to check the finite temperature effect on Casimir forces. The integral over frequency in Eq. (1) now is replaced by a discrete summation Rahi et al. (2009):
(7) |
where is replaced by discrete Matsubara frequencies is the Boltzmann’s constant and the prime denotes a prefactor 1/2 for the term =0. The Casimir pressures under different temperatures are shown in Figs. 6(a) and 6(b), where two different designs are demonstrated. It is found that the curves for temperature 320 K (insulating phase) overlap with those calculated from Eq. (1). For the temperature of 360 K, there is only a small deviation between 0 K and 360 K. Overall, the calculation results from 320 and 360 K confirm the accuracy of the 0 K approximation. Recently, the switching between repulsive and attractive Casimir forces based on PCM has also been reported Boström et al. (2018), where the equilibrium distances for switching occur only at several nanometers. The equilibrium distances in our work are more accessible to experiments, and it can be tuned by designing the geometric thickness of VO2 and the Teflon.
III.4 The effect of Brownian motion
In a real configuration, the position of a nanoplate has a fluctuation around the equilibrium distances due to the Brownian motion. To evaluate the effect of Brownian motion, the total energy of the suspended nanoplate should be known, which are written as , where is the Casimir energy given by Eq. [1], and are respectively the energies caused by the gravity and buoyancy Phan et al. (2012), and represent the density and thickness of the suspended nanoplate. The coefficient is the parameter depending on the gravity projection. For type I configuration (see the inset of Fig. 4), =0. While we have =1 and -1 for type II and type III configurations. The total energy of a gold and Teflon nanoplate are shown in Figs. 7(a) and 7(b), respectively. The minimum of corresponds to the equilibrium distance . Clearly, stable quantum trapping can be realized for a gold (Teflon) nanoplate when VO2 is in the metallic (insulating) phase. Due to the balance of repulsive Casimir force and gravity, stable trapping can also be realized for type II configuration. Theoretically, the transition rate from the equilibrium distance to another position due to the Brownian motion is proportional to Phan et al. (2012); Rodriguez et al. (2010b), where represents the energy barrier between these two positions. The calculated results indicate that the transition rates from Casimir equilibria to stiction are negligible since the energy barriers are quite large (e.g., over ) for the gold and Teflon nanoplates. For a flipped-down system (type III), quantum trapping can be realized for gold (Teflon) nanoplate when VO2 is in the metallic (insulating) phase. However, there is a nonzero possibility that the nanoplates can escape from the equilibrium distances to the free-liquid regime (). Fortunately, the energy barrier for such a transition is the order of as shown in Figs. 7(a) and 7(b), and the transition rate of the escape is also negligible.
IV Conclusions
In summary, the Casimir forces between a nanoplate and a layered structure containing VO2 films are investigated. In a liquid-separated environment, not only the magnitude of Casimir forces can be modified, but also the sign could be switched (e.g., from attraction to repulsion), due to the phase-transition of VO2. Moreover, a stable Casimir suspension of nanoplates and its tunability are revealed. For a gold nanoplate, a switch from the quantum trapping to its release is obtained under the metal-to-insulator transition of VO2. In addition, the quantum trapping and release of a Teflon nanoplate are demonstrated with a complementary design. The switching performances due to the layer thicknesses, gravitation and temperatures are discussed as well. Theoretically, the bromobenzene can be substituted by other high-refractive-index liquids (e.g., glycerol and styrene van Zwol and Palasantzas (2010)) as long as the boiling points are larger than . The Teflon can also be replaced by other low-refractive-index materials (e.g., mesoporous silica Dou et al. (2014)). This work offers the possibility of designing switchable devices in MEMS/NEMS, resulting from the quantum fluctuations of the electromagnetic field.
Acknowledgements.
This work is supported by the National Natural Science Foundation of China (Grant No. 11804288, No. 11704254, No. 61571386 and No. 61974127), and the Innovation Scientists and Technicians Troop Construction Projects of Henan Province. The research of L.X. Ge is further supported by Nanhu Scholars Program for Young Scholars of XYNU.Appendix A The permittivity of gold
Here, a generalized Drude-Lorentz model is applied for the permittivity of gold Sehmi et al. (2017):
(8) |
where the Drude term is given by:
(9) |
where 0.83409, 3134.5 eV, and 0.02334 eV. The Lorentz term is described by four pairs of poles:
(10) |
where and are the generalized conductivity and resonant frequency of the -th Lorentz pole. The star superscripts represent the operation of complex conjugation. The generalized Drude-Lorentz model respects causality, and it can represent the exact physical resonances in the material. The parameters for the model are listed in the Table I.
-th | ||
---|---|---|
1 | -0.01743+0.3059*I | 2.6905-0.16645*I |
2 | 1.0349+1.2919*I | 2.8772-0.44473*I |
3 | 1.2274+2.5605*I | 3.7911-0.81981*I |
4 | 9.85+37.614*I | 4.8532-13.891*I |
Appendix B The permittivity of VO2
For temperature , VO2 is in the metallic phase, and the permittivity is given by Pirozhenko and Lambrecht (2008); Castillo-Garza et al. (2007)
(11) | |||||
where eV, and eV. The parameters and represent respectively the strength and linewidth of the -th oscillator (resonant frequency ).
For temperature , VO2 is in the insulating phase, and the permittivity is described as
(12) |
where and eV. The above equations for metallic and insulating VO2 are valid for a wide range of frequency (up to about 10 eV)Castillo-Garza et al. (2007), which are modified versions of Ref. Verleur et al. (1968). The parameters are listed in Table II.
-th () | |||
1 | 1.816 | 0.86 | 0.95 |
2 | 0.972 | 2.8 | 0.23 |
3 | 1.04 | 3.48 | 0.28 |
4 | 1.05 | 4.6 | 0.34 |
-th () | |||
1 | 0.79 | 1.02 | 0.55 |
2 | 0.474 | 1.30 | 0.55 |
3 | 0.483 | 1.50 | 0.50 |
4 | 0.536 | 2.75 | 0.22 |
5 | 1.316 | 3.49 | 0.47 |
6 | 1.060 | 3.76 | 0.38 |
7 | 0.99 | 5.1 | 0.385 |
-th | ||||
---|---|---|---|---|
1 | 0.0093 | 0.0003 | 0.0544 | 0.00502 |
2 | 0.0183 | 0.0076 | 0.0184 | 0.0309 |
3 | 0.139 | 0.0557 | 0.0475 | 0.111 |
4 | 0.112 | 0.126 | 0.532 | 6.75 |
5 | 0.195 | 6.71 | 0.645 | 13.3 |
6 | 0.438 | 18.6 | 0.240 | 24.0 |
7 | 0.106 | 42.1 | 0.00927 | 99.9 |
8 | 0.0386 | 77.6 |
Appendix C The permittivity of Teflon and bromobenzene
The permittivity for the Teflon and bromobenzene are given by the oscillator model van Zwol and Palasantzas (2010):
(13) |
where corresponds to the oscillator strength for the -th resonance, and is the corresponding resonant frequency. The values of and listed in Table III are fitted from the experimental data in a wide range of frequency.
References
- Lyshevski (2018) S. E. Lyshevski, MEMS and NEMS: systems, devices, and structures (CRC press, 2018).
- Craighead (2000) H. G. Craighead, Science 290, 1532 (2000).
- Eom et al. (2011) K. Eom, H. S. Park, D. S. Yoon, and T. Kwon, Phys. Rep. 503, 115 (2011).
- Xu et al. (2011) R. Xu, S. Zhou, and W. J. Li, IEEE Sens. J. 12, 1166 (2011).
- Wang (2013) J. Wang, Nanomachines: fundamentals and applications (John Wiley & Sons, 2013).
- Buks and Roukes (2001) E. Buks and M. L. Roukes, Phys. Rev. B 63, 033402 (2001).
- Chan et al. (2001) H. Chan, V. Aksyuk, R. Kleiman, D. Bishop, and F. Capasso, Science 291, 1941 (2001).
- Casimir (1948) H. B. Casimir, Proc. Kon. Ned. Akad. Wet. 51, 793 (1948).
- Klimchitskaya et al. (2009) G. L. Klimchitskaya, U. Mohideen, and V. M. Mostepanenko, Rev. Mod. Phys. 81, 1827 (2009).
- Yampol’skii et al. (2008) V. A. Yampol’skii, S. Savel’ev, Z. A. Mayselis, S. S. Apostolov, and F. Nori, Phys. Rev. Lett. 101, 096803 (2008).
- Yampol’skii et al. (2010) V. A. Yampol’skii, S. Savel’ev, Z. A. Maizelis, S. S. Apostolov, and F. Nori, Phys. Rev. A 82, 032511 (2010).
- Woods et al. (2016) L. M. Woods, D. A. R. Dalvit, A. Tkatchenko, P. Rodriguez-Lopez, A. W. Rodriguez, and R. Podgornik, Rev. Mod. Phys. 88, 045003 (2016).
- Munday et al. (2009) J. N. Munday, F. Capasso, and V. A. Parsegian, Nature 457, 170 (2009).
- van Zwol and Palasantzas (2010) P. J. van Zwol and G. Palasantzas, Phys. Rev. A 81, 062502 (2010).
- Phan and Viet (2011) A. D. Phan and N. A. Viet, Phys. Rev. A 84, 062503 (2011).
- Dou et al. (2014) M. Dou, F. Lou, M. Boström, I. Brevik, and C. Persson, Phys. Rev. B 89, 201407(R) (2014).
- Rosa et al. (2008) F. S. S. Rosa, D. A. R. Dalvit, and P. W. Milonni, Phys. Rev. Lett. 100, 183602 (2008).
- Zhao et al. (2009) R. Zhao, J. Zhou, T. Koschny, E. N. Economou, and C. M. Soukoulis, Phys. Rev. Lett. 103, 103602 (2009).
- Zhao et al. (2011) R. Zhao, T. Koschny, E. N. Economou, and C. M. Soukoulis, Phys. Rev. B 83, 075108 (2011).
- Song et al. (2018) G. Song, R. Zeng, M. Al-Amri, J. Xu, C. Zhu, P. He, and Y. Yang, Opt. Express 26, 34461 (2018).
- Grushin and Cortijo (2011) A. G. Grushin and A. Cortijo, Phys. Rev. Lett. 106, 020403 (2011).
- Chen and Wan (2012) L. Chen and S. Wan, Phys. Rev. B 85, 115102 (2012).
- Nie et al. (2013) W. Nie, R. Zeng, Y. Lan, and S. Zhu, Phys. Rev. B 88, 085421 (2013).
- Tang et al. (2017) L. Tang, M. Wang, C. Y. Ng, M. Nikolic, C. T. Chan, A. W. Rodriguez, and H. B. Chan, Nat. Photonics 11, 97 (2017).
- Levin et al. (2010) M. Levin, A. P. McCauley, A. W. Rodriguez, M. T. Homer Reid, and S. G. Johnson, Phys. Rev. Lett. 105, 090403 (2010).
- Rodriguez et al. (2008) A. W. Rodriguez, J. N. Munday, J. D. Joannopoulos, F. Capasso, D. A. R. Dalvit, and S. G. Johnson, Phys. Rev. Lett. 101, 190404 (2008).
- Rahi and Zaheer (2010) S. J. Rahi and S. Zaheer, Phys. Rev. Lett. 104, 070405 (2010).
- Rodriguez et al. (2010a) A. W. Rodriguez, A. P. McCauley, D. Woolf, F. Capasso, J. D. Joannopoulos, and S. G. Johnson, Phys. Rev. Lett. 104, 160402 (2010a).
- Zhao et al. (2019) R. Zhao, L. Li, S. Yang, W. Bao, Y. Xia, P. Ashby, Y. Wang, and X. Zhang, Science 364, 984 (2019).
- Torricelli et al. (2012) G. Torricelli, P. J. Van Zwol, O. Shpak, G. Palasantzas, V. B. Svetovoy, C. Binns, B. J. Kooi, P. Jost, and M. Wuttig, Adv. Funct. Mater. 22, 3729 (2012).
- Sedighi et al. (2013) M. Sedighi, W. H. Broer, G. Palasantzas, and B. J. Kooi, Phys. Rev. B 88, 165423 (2013).
- Torricelli et al. (2010) G. Torricelli, P. J. van Zwol, O. Shpak, C. Binns, G. Palasantzas, B. J. Kooi, V. B. Svetovoy, and M. Wuttig, Phys. Rev. A 82, 010101(R) (2010).
- Shao et al. (2018) Z. Shao, X. Cao, H. Luo, and P. Jin, NPG Asia Mater. 10, 581 (2018).
- Zylbersztejn and Mott (1975) A. M. N. F. Zylbersztejn and N. F. Mott, Phys. Rev. B 11, 4383 (1975).
- Wu et al. (2017) S.-H. Wu, M. Chen, M. T. Barako, V. Jankovic, P. W. C. Hon, L. A. Sweatlock, and M. L. Povinelli, Optica 4, 1390 (2017).
- Liu et al. (2017) H. Liu, J. Lu, and X. R. Wang, Nanotechnology 29, 024002 (2017).
- Kats et al. (2012) M. A. Kats, D. Sharma, J. Lin, P. Genevet, R. Blanchard, Z. Yang, M. M. Qazilbash, D. Basov, S. Ramanathan, and F. Capasso, Appl. Phys. Lett. 101, 221101 (2012).
- van Zwol et al. (2012) P. J. van Zwol, L. Ranno, and J. Chevrier, Phys. Rev. Lett. 108, 234301 (2012).
- Cavalleri et al. (2001) A. Cavalleri, C. Tóth, C. W. Siders, J. A. Squier, F. Ráksi, P. Forget, and J. C. Kieffer, Phys. Rev. Lett. 87, 237401 (2001).
- Rini et al. (2008) M. Rini, Z. Hao, R. W. Schoenlein, C. Giannetti, F. Parmigiani, S. Fourmaux, J. C. Kieffer, A. Fujimori, M. Onoda, S. Wall, et al., Appl. Phys. Lett. 92, 181904 (2008).
- Qazilbash et al. (2008) M. M. Qazilbash, Z. Q. Li, V. Podzorov, M. Brehm, F. Keilmann, B. G. Chae, H.-T. Kim, and D. N. Basov, Appl. Phys. Lett. 92, 241906 (2008).
- Nakano et al. (2012) M. Nakano, K. Shibuya, D. Okuyama, T. Hatano, S. Ono, M. Kawasaki, Y. Iwasa, and Y. Tokura, Nature 487, 459 (2012).
- Galkina et al. (2009) E. G. Galkina, B. A. Ivanov, S. Savel’ev, V. A. Yampol’skii, and F. Nori, Phys. Rev. B 80, 125119 (2009).
- Pirozhenko and Lambrecht (2008) I. Pirozhenko and A. Lambrecht, Phys. Rev. A 77, 013811 (2008).
- Castillo-Garza et al. (2007) R. Castillo-Garza, C.-C. Chang, D. Jimenez, G. L. Klimchitskaya, V. M. Mostepanenko, and U. Mohideen, Phys. Rev. A 75, 062114 (2007).
- Milton (2004) K. A. Milton, J. Phys. A 37, R209 (2004).
- Zhan et al. (2013) T. Zhan, X. Shi, Y. Dai, X. Liu, and J. Zi, J. Phys.: Condens. Matter 25, 215301 (2013).
- Lisanti et al. (2005) M. Lisanti, D. Iannuzzi, and F. Capasso, Proc. Natl. Acad. Sci. U.S.A. 102, 11989 (2005).
- Li et al. (1994) W. Li, P. N. Segre, R. Gammon, J. V. Sengers, and M. Lamvik, J. Chem. Phys. 101, 5058 (1994).
- Rahi et al. (2009) S. J. Rahi, T. Emig, N. Graham, R. L. Jaffe, and M. Kardar, Phys. Rev. D 80, 085021 (2009).
- Boström et al. (2018) M. Boström, M. Dou, O. I. Malyi, P. Parashar, D. F. Parsons, I. Brevik, and C. Persson, Phys. Rev. B 97, 125421 (2018).
- Phan et al. (2012) A. D. Phan, L. M. Woods, D. Drosdoff, I. V. Bondarev, and N. A. Viet, Appl. Phys. Lett. 101, 113118 (2012).
- Rodriguez et al. (2010b) A. W. Rodriguez, D. Woolf, A. P. McCauley, F. Capasso, J. D. Joannopoulos, and S. G. Johnson, Phys. Rev. Lett. 105, 060401 (2010b).
- Sehmi et al. (2017) H. S. Sehmi, W. Langbein, and E. A. Muljarov, Phys. Rev. B 95, 115444 (2017).
- Verleur et al. (1968) H. W. Verleur, A. S. Barker Jr, and C. N. Berglund, Phys. Rev. 172, 788 (1968).