Next Article in Journal
Mechanical and Electrochemical Properties of Titanium Aluminum Nitride Coatings with Different Nitrogen Flow Rates on CrMnSi Steel by Filter Cathode Vacuum Arc Technology
Previous Article in Journal
The Thickness Dependence of the Superconducting Properties of MgB2 Spherical Shells Deposited on 1 mm Diameter Si3N4 Spheres
Previous Article in Special Issue
High-Stability WSe2 Homojunction Photodetectors via Asymmetric Schottky and PIN Architectures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Voltage Tunable Spoof Surface Plasmon Polariton Waveguide Loaded with Ferroelectric Resonators

1
School of Physics and Mechatronics, Guizhou Minzu University, Guiyang 550025, China
2
State Key Lab of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Coatings 2025, 15(4), 378; https://doi.org/10.3390/coatings15040378
Submission received: 22 February 2025 / Revised: 21 March 2025 / Accepted: 22 March 2025 / Published: 23 March 2025

Abstract

:
A real-time tunable planar plasmonic waveguide based on a voltage-adjustable ferroelectric resonator is designed and investigated. The laminated ferroelectric compound resonator is composed of a ferroelectric Ba0.85Ca0.15Zr0.9Ti0.1O3 (BCZT) layer, a PCB layer, as well as a localized spoof plasmonic metal layer, where the BCZT layer is beneficial for enhancing the voltage tunability in the spoof surface plasmon polariton (SSPP) waveguide. The simulated results show that the tuning range of the notch in the transmission curve, generated by the coupling between the ferroelectric compound resonator and the plasmonic waveguide, can achieve a variation of up to 8.8% thanks to the large tunability value in the BCZT ferroelectric layer. In addition, the notches consist of Fano resonant frequencies, the generation mechanism of which is elaborately discussed in terms of the temporal coupled mode theory.

1. Introduction

As is well-known, the spoof surface plasmon polariton (SSPP) concept proposed by J. B. Pendry is a kind of artificial EM wave propagating in corrugated metal, which mimics the natural inherent electron oscillation on the metal surface, named surface plasmon polaritons (SPPs) [1,2,3]. Since SSPPs can expand the SPP effect to GHz and THz bands through subwavelength engineering in metal, they have become more flexible and have generated increasing enthusiasm since their proposal. In 2012, Ander Pors et al. proposed the concept of a localized spoof surface plasmonic (LSSP), which refers to the 2D or 3D periodical plasmonic structures on the surface of metal particles, resonating in the excited standing wave mode [4]. The EM field of the resonating LSSP is highly confined near the metal interface, with a strong enhancement effect, and is highly sensitive to the geometry and local dielectric environment [5,6]. For example, in 2021, Zhang Xuanru proposed a ring plasma resonator to achieve both high efficiency and high Q capture mode excitation [7]. In particular, when utilizing coupled mode theory, multipolar plasmonic modes can be excited [8,9,10,11,12], which have sharper resonances than the basic dipole mode, and hence provide a higher Q value. Owing to these priorities, LSSPs have been widely used in optical antennas [13,14,15,16], chemical and biological sensors [17,18,19,20,21], and superlenses [22,23]. However, the previously designed LSSP was mostly static, which means the resonant properties had been solidified once the structure was fabricated. This unitary resonance is not conducive to modern complex communication environments. Thus, the real-time adjustability and experimental flexibility of multiple LSSPs become a major challenge.
Certain researchers started utilizing diverse measurement techniques to adjust LSSPs. For example, PIN diodes for the electric tuning of LSSPs have been employed [24,25,26]. Yuzan Xiong et al. utilized Yttrium Iron Garnet (YIG) ferrite for magnetically adjusting the ferromagnetic resonance field of the LSSP resonator [27]. Furthermore, a varactor diode and micro-electro-mechanical system (MEMS) have also been used as active elements for tuning LSSPs [28,29,30]. However, due to the complexity and variability of the communication environment, more regulatory measures need to be continuously expanded.
Ferroelectric materials can realize the exotic regulation of dielectric permittivity with an applied DC electric field, which has potential applications in tunable LSSP resonators [31,32,33,34,35]. The advanced features of ferroelectrics include their capability for continuous tunability, with high permittivity, low loss, and fast responses [36,37,38,39]. It is emphasized that the efficiency of the electrical tunability of ferroelectric resonators is larger than that of conventional liquid crystal devices [40,41,42,43]. The frequency stability, precision, durability, reliability, and vibration resistance of ferroelectric resonators are obviously superior to MEMSs and phase-change materials [44,45]. Also, ferroelectric resonators can be flexibly regulated by employing structure engineering as well as composition fluctuation [46,47,48,49,50,51]. However, until now, few works have comprehensively exploited the modulation mechanism of the ferroelectric resonator.
In this paper, we use the promising environmentally friendly ferroelectric Ba0.85Ca0.15Zr0.9Ti0.1O3 (BCZT) layer for fabricating an LSSP resonator. An exhaustive study on the structure engineering as well as the voltage tunability in the BCZT layer is conducted, which possesses high dielectric permittivity as well as a low loss at microwave frequencies, and will be beneficial for a compact microwave device with a low insertion loss [52,53,54,55]. Based on these characteristics, a compound ferroelectric LSSP resonator is designed and loaded onto a plasmonic waveguide to fulfill the real-time and accurate S-parameter regulation; S parameters, representing scattering parameters, are important parameters for describing the characteristics of microwave transmission and radio frequency components. Also, we demonstrate that multiple Fano modes can be aroused by the feeding of the ultracompact plasmonic structure, which can be tailored independently for filtering or sensing applications. A model based on temporal coupled mode theory is provided to describe the triple Fano behaviors, including dipole, quadrupole, and hexapole modes, which are all sensitive to the disk refractive index of the LSPR. Since the proposed LSSP is planar-textured, it provides significant flexibility in engineering the resonant properties, which possesses potential applications in compact filtering applications at microwave frequencies.

2. Experiments

The BCZT ferroelectrics were synthesized by the plasma-activated sintering (PAS) method. The calcined BCZT powders were plasma-activated for 30 s with the supplied pulsed current before sintering using a PAS device (ED-PAS-111, ELENIX, Japan). The axial pressure during PAS sintering is 50 MPa and the heating rate is 100 °C/min, with a holding time of 3 min at the sintering temperature of 1250 °C. The crystal structure of the BCZT ferroelectric resonator was examined by using a PANalytical Empyrean four-circle diffraction system. In order to test the dielectric properties of the BCZT ferroelectric resonator, a dielectric measurement equipment (ZY6173) loaded with a Keysight fixture 16197A, which can work in frequencies up to 3GHz, was utilized to test the permittivity and loss tangent of the ferroelectric resonator. Then, a compound resonator was designed employing a BCZT and PCB laminated structure, whose dielectric permittivity can be regulated flexibly in a large range by its thickness ratio (see Equations (1) and (2)). Finally, the compound resonator was loaded on the SSPP waveguide for fabricating a real-time tunable notched filter. The dispersion and Extinction Cross Section (ECS) properties of the BCZT compound resonator as well as the tunability of the notching properties were investigated using CST Studio Suite 2021 software simulation. The horizontal plane waves were used to excite the side of the resonator for the calculation of the ECS spectra. The boundary condition was set as an open boundary, and the time domain solver was selected with a grid resolution of 1000 nm in every direction of the mesh cell. A plane EM wave in the frequency range of 0 to 7 GHz, propagating in the x direction and with an electric field amplitude of 1 V/m, was employed as an incident source in the ECS investigation. A far-field/RCS monitor was selected to record the field information with the sampling of 700 frequency points.

3. Results and Discussion

3.1. Crystal Structure and Dielectric Properties of BCZT Ferroelectric

Figure 1 presents the X-ray diffraction (XRD) patterns of the BCZT ferroelectric resonator, sintered at 1250 °C. The indexed diffraction peaks indicate that the ferroelectric resonator displays a single phase without any impurity. The inset of Figure 1 depicts the cross section grain morphology of the as-sintered BCZT ferroelectric resonator. The well-densified grain structure and the even distributed grain size illustrate an optimized sintering condition at 1250 °C. Also, the BCZT ferroelectric resonator became crystallized in the trigonal–tetragonal morphotropic phase boundary (MPB) region at room temperature, with the small full width at half-maximum (FWHM), which suggests that the BCZT ferroelectric resonator has excellent crystallinity and will be beneficial for the large permittivity as well as tunability of the subsequent ferroelectric resonator.
In order to construct a tunable microwave device, the dielectric properties as well as the turnability of the BCZT resonator should be declared firstly. Therefore, we begin to investigate the dielectric permittivity as well as the unloaded quality factor of the BCZT ferroelectric ceramic at the 0.1GHz~1 GHz frequency range; the results are exhibited in Figure 2.
As can be viewed, the BCZT ceramic has a stable permittivity in a wide frequency range from 0.1 GHz to 1 GHz, and the permittivity at the 1GHz frequency band is around 190. The other important issue for ferroelectric devices is the loss tangent of the BCZT dielectric resonator in GHz band ranges from 4.0 × 10−4 to 6.3 × 10−4. The voltage tunability of the BCZT ferroelectric can be altered according to the data reported by D. Mercier et al. [56,57], where a 75% capacitance tuning is obtained in the Pt/BCZT/Pt heterostructure system under the biased voltage of 20 V. Our previous work on the dielectric tunability of the BCZT heterostructure system also shows a larger than 60% capacitance tunability [55]. The large permittivity as well as the high tunability of the BCZT ferroelectric resonator will be beneficial for the following design of the large tunability in the ferroelectric compound resonator.

3.2. Design of LSSP Resonator with BCZT Ferroelectric Compound Substrate

The following design of the LSSP resonator starts with the design of a compound tunable ferroelectric substrate. As shown in Figure 3, a bilayer compound substrate was constructed using a laminated BCZT and PCB substrate. Then, the LSSP metal structure was printed on the surface of the compound substrate. Therefore, the substrate permittivity of the LSSP resonator can be flexibly adjusted by the applied voltage. According to the superposition principle of EM theory, the final effective permittivity of the compound substrate can be designed by controlling the relative thickness ratio of each layer, as depicted in Equations (1) and (2):
ε i = i ( ε i h i P e , i )
where εi, hi, and Pe,i (i = 1, 2, 3) refer to the permittivity, laminated thickness, and filling coefficient of the ith layer in the LSSP resonator, respectively. As can be seen from Figure 3b, since the value of the dielectric permittivity of the whole substrate employed in the simulation is actually an equivalent average value, the equivalent dielectric permittivity can be adjusted by the Pe,i value, which ranges between 0 and 1 [58].
P e , i = ( 1 / 2 ) V ε r , i E × E * d V i ( ( 1 / 2 ) V ε r , i E × E * d V ) = S ε r , i E θ 2 × r d S i ( S ε r , i E θ 2 × r d S )
where εr,i (i = 1, 2, 3) denotes the permittivity of the ith layer in the compound resonator, E* represents the complex conjugation of the electric field E, while Eθ depicts the component of the electric field oriented along the θ direction. From the above equation, it can be revealed that the precise adjustment of the equivalent dielectric permittivity of the resonator can be achieved by judiciously designing the thickness and filling coefficient of each layer in the composite LSSP. For example, when the compound substrate is composed of two layers, where the lower layer is a PCB dielectric layer with a permittivity of 2.2 and thickness of h1 = 0.5 mm, and the upper layer is a BCZT ferroelectric layer with a permittivity of 190 in the working frequency range and with a thickness of h2 = 5.4 μm, then the permittivity of the compound substrate can be calculated from Equations (1) and (2), and the result exhibits an equivalent permittivity value of 3.5. Then, as the applied voltage is loaded on BCZT layer, the permittivity will experience a 60% reduction, and the equivalent permittivity of the compound substrate will become 2.9. We numerically compared the ECS properties of the LSSP resonator with the compound substrate and with a single substrate of the equivalent permittivity, as illustrated in Figure 3b. The result reveals that the two LSSP resonators with the above two different substrates actually possess similar ECS properties, which indicates the equivalence of the two substrates and verifies the validity of the design strategy of the compound tunable ferroelectric substrate. Also, it is worth pointing out that although the regulation range employed in the LSSP ferroelectric resonator is from 3.5 to 2.9, which is similar to that of the nematic liquid crystals [38], the former actually possesses a larger regulation range than that of the latter, since the permittivity variation range of the former can be adjusted by the thickness ratio in the laminated composition.
Bearing the above substrate equivalence in mind, the LSSP can be constructed on a single equivalent substrate instead of the compound substrate for convenience in the following calculation model. As depicted in panels a and b of Figure 4, the configuration of the LSSP resonator consists of three parts; the upper layer is composed of two disjoint electrodes, the middle layer is the equivalent substrate with a permittivity of ε1 = 3.5 and thickness of t = 0.65 mm, and the bottom layer is composed of a metal disk with radius r1 and six evenly distributed metal spokes. Figure 4c shows the unit structure diagram of the LSSP resonator, where the geometric parameters are listed in Table 1. The top electrode is composed of two parallel and nonintersecting rectangular metals, in which each metal electrode has a width of 0.06 mm and a gap of 0.7 mm. The permittivity of ε1 can be regulated by applying voltage at both ends of the electrode. The metal electrode is annealed copper with a thickness of 0.018 mm.
The resonant mode of the LSSP can be essentially understood as a standing wave mode propagating in a waveguide with circular arranged periodic units. When the disk circumference L of the LSSP resonator is equal to an integer n multiple the effective wavelength of the propagating mode, the standing wave resonance will be aroused [53]. Thus, the dipole mode, quadrupole mode, and hexapole mode can be excited, where the three modes represent n = 1, 2, and 3, respectively, and the wave vector kLSSP of the LSSP resonator can be described as follows:
k L S S P = 2 n π / L
Then, the dispersion characteristics of the LSSP resonator as a function of the permittivity is investigated. The simulated results illustrate that when ε1 = 3.5, the asymptotic frequency is slightly lower than that of ε1 = 2.9, deviating further from the light line, as shown in Figure 4d. The distance between the dispersion curve and the light line indicates the degree of constraint on the field; this means that as ε1 increases, the constrained field gradually increases, resulting in less radiation loss of the LSSP resonator. Moreover, according to the dispersion relationship of the LSSP unit structure and its circumference, it can be seen that when the permittivity is 3.5, the dipole frequency of the resonator should appear at 3.6 GHz, while the quadrupole and the hexapole resonance appear at 5.68 GHz and 6.35 GHz, respectively, in terms of Equation (3). It can be seen that the calculated resonances according to the dispersion curve correspond quite well with the resonant peaks in the ECS spectra, as exhibited in Figure 3a.

3.3. Tunability of SSPP Waveguide Loaded with BCZT LSSP Resonator

According to the above, this investigation is conducive to the higher tunability of εr in the BCZT compound resonator, which, in turn, will be beneficial for the following tunability of the plasmonic waveguide. Figure 5 presents the plasmonic waveguide with the BCZT compound resonator loaded symmetrically along the center of SSPP waveguide. The transmitted power distribution was numerically monitored in the x-y plane 0.5 mm above the coupled LSSP resonator at the resonant frequencies. The plasmonic waveguide is settled on a Rogers RT5880 substrate, with a substrate thickness of 0.5 mm, a permittivity of 2.2, and a loss tangent of 0.0009.
After loading the LSSP resonator, we conducted a numerical simulation on the S parameter of the filter, and the results are shown in Figure 5. The power flow distribution forms counterclockwise closed loops in the LSSP and the first three excited resonant modes are dipole, quadrupole, and hexapole resonant modes, as depicted in Figure 5c. Also, when the dielectric permittivity of the resonator varies from 3.5 to 2.9, the entire resonance peak position undergoes a blue shift, where the resonant frequency of the dipole mode shifts from 3.44 GHz to 3.65 GHz, with a blue shift of 210 MHz, or a 6.1% controllable range. Meanwhile, the resonant frequency of the quadrupole mode shifts from 5.39 GHz to 5.71 GHz, with a blue shift of 320 MHz, or a regulation range of 5.9%, and the hexapole mode shifts from 6.07 GHz to 6.44 GHz, with a tuning range of 6.1%, which is consistent with the variation pattern in the ECS spectra of the LSSP resonator illustrated in Figure 3a. Also, the transmission properties show that the excited dipole, the quadrupoles, as well as the hexapole of the loaded LSSP resonator are all Fano resonance frequencies because of the asymmetrically distributed resonant line shape. Thus, in the spectral domain, the interference between these plasmonic modes can be expressed with an analytical Fano interference model to fit the spectral lines as follows:
T ( ω ) = T 0 + A 0   [ q + 2 ( ω ω 0 ) / Γ ] 2 1 +   [ 2 ( ω ω 0 ) / Γ ] 2
where ω0 represents the Fano resonant frequencies, ω is the working angular frequency, Γ is the resonance line width, T0 characterizes the background scattering parameter, A0 is the coupling parameter between the background and the Fano resonances, and q depicts the Breit–Wigner–Fano parameter, which determines the asymmetry of the resonance spectral line.
The fitting spectral line utilizing the analytical Fano format as well as the simulated spectra are shown in the red and black line, respectively, in Figure 6. As can be seen, a well-matched spectra within the three oscillations (n = 1, 2, 3) can be obtained for the LSSP structure. The resonant frequencies of the dipole, quadrupole, and the hexapole modes center at 3.43 GHz, 5.39 GHz, and 6.06 GHz, respectively, and the corresponding line widths of the modes are 0.0125 GHz, 0.078 GHz, and 0.045 GHz, respectively. Thus, the Q value of the resonator can be calculated using the following formula:
Q = ω 0 Γ
The results show that the Q values of the dipole, quadrupole, and hexapole modes of the LSSP resonator can be calculated as 275, 80, and 135, respectively.

3.4. Coupling Mechanism Between BCZT LSSP Resonator and SSPP Waveguide

To reveal the formation of the Fano response of the LSSP resonator, the coupling mechanism of the near-field EM between the three Fano modes and the SSPP modes in plasmonic waveguide is investigated in terms of temporal coupled mode theory (TCMT). The excited Fano modes should stem from the interference of the two transmission paths, as illustrated in Figure 7. An equivalent network is therefore employed to describe the EM field coupling characteristics between the two plasmonic structures, as illustrated in the inset of Figure 7.
A dielectric loss as well as weak radiative loss of the LSSP resonator are considered azimuthally as a resistance R, where the amplitude damping factor in the LSSP is denoted as γ. Regarding the distinction between γ and Γ, Γ represents the resonance linewidth, which is a measure of the energy dissipation and directly determines the quality factor (Q) of the resonance. In contrast, γ describes the system’s intrinsic damping characteristics, which contribute to the total loss, but are treated as a separate parameter in our model. The normalized EM power due to the interference of the two transmission paths can thus be depicted as follows:
E 2 = t E 1 + j r E 4
E 3 = t E 4 + j r E 2
E 4 = ( 1 γ ) e x p ( j θ ) E 3
Then, the transmission coefficient E2/E1 can be deduced as follows:
E 2 E 1 = t + ( 1 γ ) e x p ( j θ ) 1 + t ( 1 γ ) e x p ( j θ )
where t represents the coupling through nonradiative damping channels, r describes the coupling process from the LSSP resonator back to the plasmonic waveguide, j stands for imaginary unit, and E stands for electric field strength. For the lossless coupling process between the SSPP waveguide and LSSP resonator, we assume t2 + r2 = 1. The rotated phase delay θ distributed along the azimuthal direction of the LSSP resonator is in the range of 0 to 2π for the dipole Fano mode. In terms of the low loss of the BCZT substrate in the LSSP resonator, we consider the weak value of R in the equivalent circuit. This parameter represents both the dielectric and radiative losses of the LSSP resonator, thereby influencing the system’s damping characteristics. We set γ = 0.05 ( ω ω 0 ) / ω [54], where ω 0 represents the dipole resonant frequency, and t = 0.996 to fit the normalized transmission coefficient.
This evidently indicates that the coupling effect between the SSPP waveguide and the LSSP resonator can deeply trap the power flow, appearing with a strong band notch in the transmission spectrum, as exhibited in Figure 7. Also, the theoretical calculated real part ratio |E2/E1| in Equation (9), standing for transmission intensity ratio, matches quite well with the numerically simulated one. It can thus be validated that the asymmetric field fed by the SSPP waveguide could lead to a unidirectional EM coupling into the LSSP resonator, and therefore create an asymmetric resonant line shape, forming the Fano resonant frequencies.

4. Conclusions

In summary, a LSSP resonator with a BCZT compound ferroelectric substrate is designed and employed for the tunability of the SSPP waveguide. The results show that high dielectric tunability can be achieved in the BCZT laminated structure, which is beneficial for the deep tunability of the notch in the plasmonic waveguide. When a pair of mirror-symmetrical LSSP resonators are coupled to the SSPP WG, the central frequency of the notch can be dynamically adjusted in real-time by about 8.8% due to the high dielectric tunability of the BCZT substrate. In addition, the notch generated by the plasmonic coupling enables the excitation of triple Fano resonances, which stem from a unidirectional EM coupling between the LSSP resonator and the SSPP waveguide. The proposal of this study achieves the real-time dynamic adjustment of waveguide performance, which is a significant innovation in the field of waveguide design. Meanwhile, the BCZT layer, as a ferroelectric material, exhibits excellent voltage response characteristics. By applying a voltage to it, the dielectric constant can be significantly changed, thereby enabling the real-time control of the SSPP waveguide performance. This enhancement of voltage tunability is a major highlight of this study.
The method provides a new idea for a compact physical size, minimal insertion loss, large tunability, and real-time flexible regulation in notch filter scenarios. The proposed novel waveguide structure provides new ideas and methods for the development of waveguide technology. By adjusting the waveguide performance in real-time, it can meet the needs of different application scenarios and promote the application of waveguide technology in fields such as communication, sensing, and imaging. Also, since the proposed LSSP is a planar-textured disk, it is more flexible in engineering a Fano line shape, which can be of practical interest for a variety of applications including enhanced biosensing, on-chip communication, and optical information processing.

Author Contributions

Conceptualization, M.H. and C.W.; methodology, J.S. and S.L. (Shun Lei); validation, J.S. and C.Z.; formal analysis, S.L. (Shengyun Luo) and C.Z.; investigation, C.Z. and C.W.; data curation, S.L. (Shun Lei) and S.L. (Shengyun Luo); writing—original draft preparation, J.S.; writing—review and editing, M.H. and C.W.; supervision, M.H.; project administration, C.W.; funding acquisition, M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No.62261008), the High Level Innovative Talent Project of Science and Technology Department of Guizhou Province (GCC2023086), and the Fundamental Research Funds for the Central Universities (WUT: 2019III029).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data generated by the process can be found in the text, and no additional data need to be added.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Pendry, J.B.; Holden, A.J.; Stewart, W.J.; Youngs, I. Extremely low frequency plasmons in metallic mesostructures. Phys. Rev. Lett. 1996, 76, 4773. [Google Scholar] [PubMed]
  2. Zayats, A.V.; Smolyaninov, I.I.; Maradudin, A.A. Nano-optics of surface plasmon polaritons. Phys. Rep. 2005, 408, 131–314. [Google Scholar]
  3. Garcia-Vidal, F.J.; Martin-Moreno, L.; Pendry, J.B. Surfaces with holes in them: New plasmonic metamaterials. J. Opt. A Pure Appl. Opt. 2005, 7, S97. [Google Scholar]
  4. Pendry, J.B.; Martin-Moreno, L.; Garcia-Vidal, F.J. Mimicking surface plasmons with structured surfaces. Science 2004, 305, 847–848. [Google Scholar]
  5. Ritchie, R.H. Plasma losses by fast electrons in thin films. Phys. Rev. 1957, 106, 874. [Google Scholar]
  6. Liebsch, A. Electronic Excitations at Metal Surfaces; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  7. Zhang, X.; Cui, T.J. Deep-subwavelength and high-Q trapped mode induced by symmetry-broken in toroidal plasmonic resonator. IEEE Trans. Antennas Propag. 2020, 69, 2122–2129. [Google Scholar]
  8. Maier, S.A. Plasmonics: Fundamentals and Applications; Springer: New York, NY, USA, 2007. [Google Scholar]
  9. Baimuratov, A.S.; Tepliakov, N.V.; Gun’ko, Y.K.; Baranov, A.V.; Fedorov, A.V. Mixing of quantum states: A new route to creating optical activity. Sci. Rep. 2016, 6, 5. [Google Scholar]
  10. Homola, J. Present and future of surface plasmon resonance biosensors. Anal. Bioanal. Chem. 2003, 377, 528–539. [Google Scholar]
  11. Pan, L.; Wu, Y.; Wang, W.; Wei, Y.; Yang, Y. A flexible high-selectivity single-layer coplanar waveguide bandpass filter using interdigital spoof surface plasmon polaritons of bow-tie cells. IEEE Trans. Plasma Sci. 2020, 48, 3582–3588. [Google Scholar]
  12. Jaiswal, R.K.; Pandit, N.; Pathak, N.P. Spoof plasmonic-based band-pass filter with high selectivity and wide rejection bandwidth. IEEE Photonics Technol. Lett. 2019, 31, 1293–1296. [Google Scholar]
  13. Wang, J.; Zhao, L.; Hao, Z.C. A band-pass filter based on the spoof surface plasmon polaritons and CPW-based coupling structure. IEEE Access 2019, 7, 35089–35096. [Google Scholar] [CrossRef]
  14. Guo, Y.J.; Xu, K.D.; Deng, X.; Cheng, X.; Chen, Q. Millimeter-wave on-chip bandpass filter based on spoof surface plasmon polaritons. IEEE Electron Device Lett. 2020, 41, 1165–1168. [Google Scholar] [CrossRef]
  15. Zhang, D.; Sun, Y.; Zhang, K.; Wu, Q.; Jiang, T. Short-circuited stub-loaded spoof surface plasmon polariton transmission lines with flexibly controllable lower out-of-band rejections. Opt. Lett. 2021, 46, 4354–4357. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, Z.; Zhang, B.; Chen, W.; Yang, T. Rejection of spoof SPPs using the second resonant mode of vertical split-ring resonator. IEEE Microw. Wirel. Compon. Lett. 2018, 29, 23–25. [Google Scholar] [CrossRef]
  17. Li, W.; Qin, Z.; Wang, Y.; Ye, L.; Liu, Y. Spoof surface plasmonic waveguide and its band-rejection filter based on H-shaped slot units. J. Phys. D Appl. Phys. 2019, 52, 365303. [Google Scholar] [CrossRef]
  18. Yan, D.; Li, X.; Ma, C. Terahertz refractive index sensing based on gradient metasurface coupled confined spoof surface plasmon polaritons mode. IEEE Sens. J. 2021, 22, 324–329. [Google Scholar] [CrossRef]
  19. Kumari, A.; Singh, S.P.; Tiwari, N.K. Design of a differential spoof surface plasmon sensor for dielectric sensing and defect detection. IEEE Sens. J. 2022, 22, 3188–3195. [Google Scholar] [CrossRef]
  20. Fu, J.H.; Wu, W.J.; Wang, D.W. High-sensitivity microfluidic sensor based on quarter-mode interdigitated spoof plasmons. IEEE Sens. J. 2022, 22, 23888–23895. [Google Scholar] [CrossRef]
  21. Dai, L.H.; Zhao, H.Z.; Zhao, X. Flexible and printed microwave plasmonic sensor for noninvasive measurement. IEEE Access. 2020, 8, 163238–163243. [Google Scholar]
  22. Xu, K.D.; Lu, S.; Guo, Y.J.; Chen, Q. High-order mode of spoof surface plasmon polaritons and its application in bandpass filters. IEEE Trans. Plasma Sci. 2020, 49, 269–275. [Google Scholar] [CrossRef]
  23. Mira, F.; Mateu, J.; Collado, C. Mechanical tuning of substrate integrated waveguide filters. IEEE Trans. Microw. Theory Tech. 2015, 63, 3939–3946. [Google Scholar] [CrossRef]
  24. Zhang, C.; Zhou, Y.J. Electronically controlled spoof localized surface plasmons on the corrugated ring with a shorting pin. J. Mod. Opt. 2018, 65, 1535–1541. [Google Scholar]
  25. Ding, J.; Zhao, P.; Chen, H.; Fu, H. Ultraviolet photodetectors based on wide bandgap semiconductor: A review. J. Appl. Phys. A 2024, 130, 350. [Google Scholar]
  26. Pande, S.; Patil, D.; Kumar, A.; Tyavlovsky, A.K.; Muthanna, A. Design of a metasurface loaded with RF varactor and PIN diode integration dual-band reconfigurable antenna for bluetooth, Wi-Fi, WLAN and wireless 5G applications. Phys. Scr. 2025, 100, 025537. [Google Scholar]
  27. Xiong, Y.Z.; Christy, A.; Dong, Y. Combinatorial split-ring and spiral metaresonator for efficient magnon-photon coupling. Phys. Rev. Appl. 2024, 21, 034034. [Google Scholar] [CrossRef]
  28. Ma, J.; Hu, J.M.; Li, Z.; Nan, C.W. Recent progress in multiferroic magnetoelectric composites: From bulk to thin films. Adv. Mater. 2011, 42, 1062–1087. [Google Scholar]
  29. Li, S.Q.; Du, C.H. Efficient magnetic-coupling excitation of LSSPs on high-Q multilayer planar-circular-grating resonators. Opt. Express 2021, 29, 25189–25201. [Google Scholar] [CrossRef] [PubMed]
  30. Zhu, B.; Hu, M.Z. Plasmonic dual-band waveguide with independently controllable band-notched characteristics. Appl. Phys. Express 2023, 16, 086001. [Google Scholar]
  31. Ramesh, R.; Spaldin, N.A. Multiferroics: Progress and prospects in thin films. Nat. Mater. 2007, 6, 21–29. [Google Scholar]
  32. Zhang, G.; Dong, S.; Yan, Z.; Guo, Y.; Zhang, Q.; Yunoki, S.; Dagotto, E.; Liu, J.M. Multiferroic properties of CaMn7O12. Phys. Rev. B 2011, 84, 174413. [Google Scholar] [CrossRef]
  33. Lou, J.; Wang, J.; Ma, H. Tunable spoof surface plasmon polariton transmission line based on ferroelectric thick film. Appl. Phys. A 2019, 125, 737. [Google Scholar] [CrossRef]
  34. Miranda, F.A.; Subramanyam, G.; Vankeuls, F.W.; Romanofsky, R.R. Design and development of ferroelectric tunable microwave components for Ku- and K-band satellite communication systems. ZEEE Trans. Microw. Theory Tech. 2000, 48, 1181–1189. [Google Scholar]
  35. Schileo, G.; Pascual-Gonzalez, C.; Alguero, M.; Reaney, I.M.; Postolache, P.; Mitoseriu, L.; Reichmann, K.; Venet, M.; Feteira, A. Multiferroic and magnetoelectric properties of Pb0.99[Zr0.45Ti0.47(Ni1/3Sb2/3)0.08] O3-CoFe2O4 multilayer composites fabricated by tape casting. J. Eur. Ceram. Soc. 2018, 38, 1473–1478. [Google Scholar] [CrossRef]
  36. Liu, W.; Ren, X. Large piezoelectric effect in Pb-free ceramics. Phys. Rev. Lett. 2009, 103, 257602. [Google Scholar]
  37. Nan, C.W.; Bichurin, M.I.; Dong, S.; Viehland, D.; Srinivasan, G. Multiferroic magnetoelectric composites: Historical perspective, status, and future directions. J. Appl. Phys. 2008, 103, 03110. [Google Scholar]
  38. Li, T.; Zhang, F.; Fang, H.; Li, K.; Yu, F. The magnetoelectric properties of La0.7Sr0.3MnO3/BaTiO3 bilayers with various orientations. J. Alloy. Compd. 2013, 560, 167–170. [Google Scholar] [CrossRef]
  39. Li, T.; Wang, H.; Ma, D.; Li, K.; Hu, Z. Influence of clamping effect in BaTiO3 film on the magnetoelectric behavior of layered multiferroic heterostructures. Mater. Res. Bull. 2019, 115, 116–120. [Google Scholar] [CrossRef]
  40. Martin, L.W.; Ramesh, R. Multiferroic and magnetoelectric heterostructures. Acta Mater. 2012, 60, 2449–2470. [Google Scholar] [CrossRef]
  41. Fetisov, Y.K.; Srinivasana, G. Electric field tuning characteristics of a ferrite-piezoelectric microwave resonator. Appl. Phys. Lett. 2006, 88, 143503. [Google Scholar] [CrossRef]
  42. Ustinov, A.; Srinivasan, G.; Kalinikos, B.A. Ferrite-ferroelectric hybrid wave phase shifters. Appl. Phys. Lett. 2007, 90, 031913. [Google Scholar] [CrossRef]
  43. Pettiford, C.; Dasgupta, S.; Lou, J.; Yoon, S.D.; Sun, N.X. Bias field effects on microwave frequency behavior of PZT/YIG magnetoelectric bilayer. IEEE Trans. Magn. 2007, 43, 3343. [Google Scholar]
  44. Zine-El-Abidine, I.; Okoniewski, M. A tunable radio frequency MEMS inductor using MetalMUMPs. J. Micromech. Microeng. 2007, 17, 2280. [Google Scholar]
  45. Hummel, G.; Hui, Y.M.; Rinaldi, Y. Reconfigurable piezoelectric MEMS resonator using phase change material programmable vias. J. Microelectromech. Syst. 2015, 24, 2145–2151. [Google Scholar] [CrossRef]
  46. Wang, J.; Li, Z.; Wang, J.; He, H.; Nan, C. Effect of thickness on the stress and magnetoelectric coupling in bilayered Pb (Zr0.52Ti0.48) O3-CoFe2O4 films. J. Appl. Phys. 2015, 117, 759. [Google Scholar]
  47. Fan, K.B.; Averitt, R.D.; Padilla, W.J. Active and tunable nanophotonic metamaterials. Nanophotonics 2022, 11, 3769–3803. [Google Scholar]
  48. Dai, Y.Q.; Dai, J.M.; Tang, X.W.; Zhang, K.J.; Zhu, X.B.; Yang, J.; Sun, Y.P. Thickness effect on the properties of BaTiO3-CoFe2O4 multilayer thin films prepared by chemical solution deposition. J. Alloy. Compd. 2014, 587, 681–687. [Google Scholar] [CrossRef]
  49. Pertsev, N.A.; Zembilgotov, A.G.; Tagantsev, A.K. Effect of mechanical boundary conditions on phase diagrams of epitaxial ferroelectric thin films. Phys. Rev. Lett. 1998, 80, 1988–1991. [Google Scholar]
  50. Ion, V.; Craciun, F.; Scarisoreanu, N.D.; Moldovan, A.; Andrei, A.; Birjega, R.; Ghica, C. Impact of thickness variation on structural, dielectric and piezoelectric properties of (Ba, Ca) (Ti, Zr) O3 epitaxial thin films. Sci. Rep. 2018, 8, 2056. [Google Scholar]
  51. Pei, H.; Zhang, Y.; Guo, S.; Ren, L.; Yan, H.; Luo, B. Orientation-dependent optical magnetoelectric effect in patterned BaTiO3/La0. 67Sr0. 33MnO3 heterostructures. ACS Appl. Mater. Interfaces 2018, 10, 30895–30900. [Google Scholar]
  52. Li, S.B.; Wang, C.B.; Shen, Q.; Zhang, L.M. Enhanced dielectric properties in Ba0.85Ca0.15Zr0.10Ti0.90O3/La0.67Ca0.33MnO3 laminated composite. Scr. Mater. 2018, 144, 40–43. [Google Scholar]
  53. Hu, M.Z.; Li, S.; Wang, C.B. Orientation-dependent multiferroic properties in BCZT/LCMO thin films. Ceram. Int. 2021, 103, 385–991. [Google Scholar]
  54. Mercier, D.; Niembro-Martin, A.; Sibuet, H.; Baret, C.; Chautagnat, J. X Band Distributed Phase Shifter Based on Sol-Gel BCTZ Varactors. In Proceedings of the 2017 European Radar Conference (EURAD), Nuremberg, Germany, 11–13 October 2017; IEEE: New York, NY, USA; pp. 382–385. [Google Scholar]
  55. Mercier, D.; Sibuet, H.; Dieppedale, C.; Chautagnat, J.; Baret, C.; Bonnard, C.; Guillaume, J.; Gardes, P.; Poveda, P.; Le Rhun, G.; et al. 8 to 13 GHz tunable filter based on sol-gel BCTZ varactors. In Proceedings of the 2017 IEEE Asia Pacific Microwave Conference (APMC), Kuala Lumpur, Malaysia, 13–16 November 2017; IEEE: New York, NY, USA; pp. 686–689. [Google Scholar]
  56. Zhang, X.R.; Yan, R.T.; Cui, T.J. High-FoM Resonance in Single Hybrid Plasmonic Resonator via Electro-Magnetic Modal Interference. IEEE Trans. Antennas Propag. 2020, 68, 6447–6451. [Google Scholar]
  57. Hu, M.Z.; Gu, H.S.; Chu, X.C.; Qian, J.; Xia, Z. Crystal structure and dielectric properties of (1−x) Ca0. 61Nd0. 26TiO3+ xNd (Mg1/2Ti1/2) O3 complex perovskite at microwave frequencies. J. Appl. Phys. 2008, 104, 124104. [Google Scholar]
  58. Lei, S.; Hu, M.Z.; Xu, J.; Zhou, C.; Zhao, Q.; Zhang, L.; Zhang, H. Real-time tunable notched waveguide based on voltage controllable ferroelectric resonator. Opt. Express 2024, 32, 10587–10598. [Google Scholar] [PubMed]
Figure 1. XRD pattern and corresponding SEM image (the inset) of the BCZT ferroelectric resonator sintered at the optimized temperature of 1250 °C.
Figure 1. XRD pattern and corresponding SEM image (the inset) of the BCZT ferroelectric resonator sintered at the optimized temperature of 1250 °C.
Coatings 15 00378 g001
Figure 2. The dielectric permittivity and loss tangent (tanδ) of the BCZT ferroelectric resonator in the 0.1GHz~1GHz frequency range.
Figure 2. The dielectric permittivity and loss tangent (tanδ) of the BCZT ferroelectric resonator in the 0.1GHz~1GHz frequency range.
Coatings 15 00378 g002
Figure 3. (a) The variation in the simulated ECS spectra with the permittivity ε1 of the substrate. (b) The comparison of the ECS spectra of the ferroelectric composite dielectric resonator vs. the effective single layer LSSP res-nator. The inset is the schematic diagram of the composite substrate, which consists of two layers. The upper layer is the BCZT ferroelectric layer with a permittivity of 190 and lower layer is the PCB dielectric layer with a permittivity of 2.2. The spoke electrodes of the LSSP resonator is on the surface. The resonator has a radius of 6.25 mm and a thickness of 0.65 mm. (c) The electric field distribution of the first three resonant peaks of the proposed resonator when the dielectric constant ε1 = 3.5. (d) The magnetic field distribution of the first three resonant peaks of proposed resonator when the dielectric constant ε1 = 3.5. (e) The electric field distribution of the first three resonant peaks of proposed resonator when the dielectric constant ε1 = 2.9. (f) The magnetic field distribution of the first three resonant peaks of proposed resonator when the dielectric constant ε1 = 2.9. (g) The excitation direction of the plane wave.
Figure 3. (a) The variation in the simulated ECS spectra with the permittivity ε1 of the substrate. (b) The comparison of the ECS spectra of the ferroelectric composite dielectric resonator vs. the effective single layer LSSP res-nator. The inset is the schematic diagram of the composite substrate, which consists of two layers. The upper layer is the BCZT ferroelectric layer with a permittivity of 190 and lower layer is the PCB dielectric layer with a permittivity of 2.2. The spoke electrodes of the LSSP resonator is on the surface. The resonator has a radius of 6.25 mm and a thickness of 0.65 mm. (c) The electric field distribution of the first three resonant peaks of the proposed resonator when the dielectric constant ε1 = 3.5. (d) The magnetic field distribution of the first three resonant peaks of proposed resonator when the dielectric constant ε1 = 3.5. (e) The electric field distribution of the first three resonant peaks of proposed resonator when the dielectric constant ε1 = 2.9. (f) The magnetic field distribution of the first three resonant peaks of proposed resonator when the dielectric constant ε1 = 2.9. (g) The excitation direction of the plane wave.
Coatings 15 00378 g003
Figure 4. (a) Schematic diagram of the top of the resonator, with the parameters listed in Table 1. (b) Schematic diagram of the bottom of the resonator with the LSSP structure of periodic straight spokes. (c) Local unit structure diagram of the LSSP resonator with the parameters listed in Table 1. (d) Dispersion curve of the LSSP resonator with a different substrate permittivity ε1. The substrate thickness is 0.5 mm. (e,f) The internal electric and magnetic field distribution diagram at three resonant points, A, B, and C, of the LSSP resonator, respectively, where ε1 = 3.5.
Figure 4. (a) Schematic diagram of the top of the resonator, with the parameters listed in Table 1. (b) Schematic diagram of the bottom of the resonator with the LSSP structure of periodic straight spokes. (c) Local unit structure diagram of the LSSP resonator with the parameters listed in Table 1. (d) Dispersion curve of the LSSP resonator with a different substrate permittivity ε1. The substrate thickness is 0.5 mm. (e,f) The internal electric and magnetic field distribution diagram at three resonant points, A, B, and C, of the LSSP resonator, respectively, where ε1 = 3.5.
Coatings 15 00378 g004
Figure 5. (a,b) The schematic structure of the plasmonic waveguide. (c) The simulated transmission spectra loaded with the BCZT compound resonator. (d,e) The internal electric and magnetic field distribution diagram at three notched points, 1, 2, and 3, respectively, where ε1 = 3.5. (f,g) The internal electric and magnetic field distribution diagram at three notched points 1*, 2*, and 3*, respectively, where ε1 = 2.9.
Figure 5. (a,b) The schematic structure of the plasmonic waveguide. (c) The simulated transmission spectra loaded with the BCZT compound resonator. (d,e) The internal electric and magnetic field distribution diagram at three notched points, 1, 2, and 3, respectively, where ε1 = 3.5. (f,g) The internal electric and magnetic field distribution diagram at three notched points 1*, 2*, and 3*, respectively, where ε1 = 2.9.
Coatings 15 00378 g005
Figure 6. Simulation and Fano fitting of (a) dipole peak, (b) quadrupole peak, and (c) hexapole peak of the LSSP resonator, where ε1 = 3.5.
Figure 6. Simulation and Fano fitting of (a) dipole peak, (b) quadrupole peak, and (c) hexapole peak of the LSSP resonator, where ε1 = 3.5.
Coatings 15 00378 g006
Figure 7. When ε1 is equal to 3.5, the calculated (blue line) and simulation results (red line) of the real part |E2/E1| are shown in terms of TCMT. The inset shows the equivalent network model revealing the near-field EM power transferring between the plasmonic waveguide and the LSSP resonator.
Figure 7. When ε1 is equal to 3.5, the calculated (blue line) and simulation results (red line) of the real part |E2/E1| are shown in terms of TCMT. The inset shows the equivalent network model revealing the near-field EM power transferring between the plasmonic waveguide and the LSSP resonator.
Coatings 15 00378 g007
Table 1. The denotation and physical dimensions of each part of the designed SSPP waveguide loaded with the LSSP resonator.
Table 1. The denotation and physical dimensions of each part of the designed SSPP waveguide loaded with the LSSP resonator.
StructureSymbolSize (mm)
the lower surface metal strip lengthl15.25
the upper metal strip lengthl24.71
the gap between upper center circle and metal stripw10.1
the upper surface metal strip widthw20.08
the upper and lower center metal circle radiusr11
the upper metal ring widthw0.44
the lower surface metal strip widtha0.7
the periodic of the circle slotsp6.545
the thickness of the dielectric substratet0.65
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shen, J.; Lei, S.; Hu, M.; Zhou, C.; Luo, S.; Wang, C. Voltage Tunable Spoof Surface Plasmon Polariton Waveguide Loaded with Ferroelectric Resonators. Coatings 2025, 15, 378. https://doi.org/10.3390/coatings15040378

AMA Style

Shen J, Lei S, Hu M, Zhou C, Luo S, Wang C. Voltage Tunable Spoof Surface Plasmon Polariton Waveguide Loaded with Ferroelectric Resonators. Coatings. 2025; 15(4):378. https://doi.org/10.3390/coatings15040378

Chicago/Turabian Style

Shen, Jiaxiong, Shun Lei, Mingzhe Hu, Chaobiao Zhou, Shengyun Luo, and Chuanbin Wang. 2025. "Voltage Tunable Spoof Surface Plasmon Polariton Waveguide Loaded with Ferroelectric Resonators" Coatings 15, no. 4: 378. https://doi.org/10.3390/coatings15040378

APA Style

Shen, J., Lei, S., Hu, M., Zhou, C., Luo, S., & Wang, C. (2025). Voltage Tunable Spoof Surface Plasmon Polariton Waveguide Loaded with Ferroelectric Resonators. Coatings, 15(4), 378. https://doi.org/10.3390/coatings15040378

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop