Next Article in Journal
Boundary Effect and Critical Temperature of Two-Band Superconducting FeSe Films
Previous Article in Journal
Structural Determination of the Hexacoordinated [Zn(L)2]2+ Complex Isomer Type Using Solution-State NMR, DFT Calculations and X-ray Crystallography
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Metamaterial-Based Cross-Polarization Converter Characterized by Wideband and High Efficiency

1
Guangxi Key Laboratory of Wireless Wideband Communication & Signal Processing, Guilin 541004, China
2
School of Information & Communication, Guilin University of Electronic Technology, Guilin 541004, China
3
National Key Laboratory of Electromagnetic Environment, Qingdao 266107, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(1), 17; https://doi.org/10.3390/cryst13010017
Submission received: 24 November 2022 / Revised: 19 December 2022 / Accepted: 19 December 2022 / Published: 22 December 2022

Abstract

:
Metamaterial-based polarization converters, which have all kinds of polarizations realizable via adjusting metamaterial parameters, have been springing up at an increasing rate. However, the reported metamaterial-based polarization converters suffer from either limited bandwidth or low polarization conversion ratios. In this study, a metamaterial-based polarization converter consisting of multilayer copper split-ring resonators and a copper ground separated by dielectrics was demonstrated and was characterized by the cross-polarization with wideband and high-efficiency. For normal incidence, the simulated results illustrated that the expanded bandwidth benefited from the superposition of cross-polarization electromagnetic resonances around 2.78, 3.09, 3.68, and 4.54 GHz, and the polarization conversion ratio was higher than 99% in the frequency range of 2.73 and 4.63 GHz. For oblique incidence, the design can provide larger angle tolerance in the investigated band, except for a very narrow stopband. Moreover, the experimental results agreed well with the simulations, which verified the reliability of the performance.

1. Introduction

Polarization is an important concept in electromagnetic (EM) theory [1,2,3]. The polarization converter, which is an important class of EM device, manipulates the polarization state of EM waves and achieves significant progress [4,5,6,7,8,9,10,11]. In the past, polarization converters were achieved by using birefringent crystals based on Faraday effects, which were frequently bulky in size and resulted in much difficulty for miniaturization [12,13,14]. Recently, artificial metamaterials designed by subwavelength meta-atom structures provide an alternative to realize many special phenomena and functions, such as hyper lenses [15,16], negative refraction [17,18], invisibility cloaks [19,20], perfect absorption [21,22,23], and electromagnetic wave polarization manipulation [1,24].
In recent years, metamaterial-based reflective or transmissive polarization converters, with all kinds of polarizations realized via adjusting metamaterial parameters, have sprung up at an increasing rate. Zhang et al. reported a simple and broadband reflective polarization converter at gigahertz frequencies by breaking the symmetry of the cross-shaped resonator [25], and the experimental results demonstrated that the polarization conversion ratio (PCR) was over 80%, from 8.3 GHz to 14.3 GHz, with a relative bandwidth 53%. Regrettably, the dual band polarization conversion appeared as the PCR grew to 90%, which implied that the broadband property of polarization conversion disappeared. Peng et al. demonstrated a linear polarization converter based on 3D split-loop resonators arranged orthogonally [26]. Although the crossed polarization reflection coefficient was close to 1 and resulted in an approximately 100% PCR at 2.65 GHz, the polarization bandwidth was very narrow. Loncar et al. designed and fabricated a polarization converting metasurface composed of three metal pattern layers in the X-band [27]. The polarization conversion with a 99.4% PCR was accomplished by guiding and coupling electromagnetic energy between pairs of orthogonal slots in a ground plane, which were placed at the edges of each unit cell (UC). Unfortunately, the relative bandwidth of cross polarization was only 22%. It follows that the previously mentioned metamaterial-based polarization converters suffer from either limited bandwidth or low PCRs.
To address the above problems, this paper designs a reflection-type metamaterial polarization converter (MPC) based on multilayer split-ring resonators (SRRs), which is capable of converting the line polarization into its orthogonal one. Additionally, four electromagnetic resonances generated by the SRRs can significantly extend the reflection bandwidth of the cross-polarization. In addition, numerical simulations showed that the PCR was greater than 99% in the frequency range of 2.73–4.63 GHz, and the relative bandwidth achieved 52%, which basically coincided with the experiments.

2. Design of the Metamaterial Polarization Converter

2.1. Fundamental Theory

For simplicity, the normal incidence was considered. By starting from anticlockwise and rotating the xy coordinate system for an angle of 45°, the uv coordinate system was reached. Furthermore, the incident wave was supposed to be y polarization and vertically strike the surface of the metamaterial along the –z direction. Therefore, the y polarization wave can be decomposed into u and v orthogonal components, as shown in Figure 1a. Mathematically, the electric field of the incident wave is thus expressed as
E i = e y E i e j k z = e u E i 2 e j k z + e v E i 2 e j k z = E u i + E v i
where e y , e u , and e v are the unit vectors in different coordinate systems, and E i is the amplitude of the electric field. For reflection-type anisotropic metamaterial, even if the incident wave possesses one (u or v) polarization, the reflection beam generally involves both u and v polarizations. The reflected wave of the electric field is then described as
E r = r u u E i e j ( - k z + ϕ u u ) + r u v E i e j ( - k z + ϕ u v ) 2 e u + r v u E i e j ( - k z + ϕ v u ) + r v v E i e j ( - k z + ϕ v v ) 2 e v
where ruu and rvv are the co-polarization reflection coefficient amplitudes; ruv (v-polarized wave is incident and u-polarized wave is reflected) and rvu represent the cross-reflection coefficient amplitudes; ϕuu, ϕvv, ϕuv, and ϕvu are the phases corresponding to the reflection coefficients. For given ruv = rvu = 0, rvv = ruu = r, and ∆ϕ = ϕvvϕuu = 2nπ ± π or ruv = rvu = r, rvv = ruu = 0, and ∆ϕ = ϕvuϕuv = 2 ± π (n is an integer), Equation (2) can be simply stated as
E r = r E i e - j k z e j ϕ u u 2 e u + r E i e - j k z e j ( ϕ u u + Δ ϕ ) 2 e v = r E i e - j k z e j ϕ u u 2 e u - r E i e - j k z e j ϕ u u 2 e v = e x r E i e j ϕ u u e - j k z
or
E r = r E i e - j k z e j ϕ u v 2 e u + r E i e - j k z e j ( ϕ u v + Δ ϕ ) 2 e v = r E i e - j k z e j ϕ u v 2 e u - r E i e - j k z e j ϕ u v 2 e v = e x r E i e j ϕ u v e - j k z
Equation (3) or (4) implies that the reflected wave is linearly polarized, and the polarization direction is perpendicular to that of the original wave. A complete cross-polarization conversion (CPC) is thus realized. And it draws a conclusion that the CPC strongly depends on the reflection coefficient, which relies on the metamaterial parameters. In other words, the limitation on the CPC performance is in the metamaterial parameters.

2.2. Polarization Converter Model

After optimizing the metamaterial parameters, the polarization converter model was proposed. Figure 1b–d depict a UC of the MPC, which consists of three copper split-ring resonators (SRRs) (labeled as SRR1, SRR2, and SRR3 as shown in Figure 1c) and a metal ground separated by four dielectric spacers. The upmost dielectric layer is the PTFE (εr = 2.2) with the thickness h1, and the others FR4 (εr = 4.4) with the thicknesses h2, h3, and h4, respectively. And in Figure 1d, the widths of SRR1, SRR2, and SRR3 are respectively w 1 = r 1 - r 1 , w 2 = r 2 - r 2 , and w 3 = r 3 - r 3 . The parameters relating to the UC are listed in Table 1.

3. Simulations and Discussions

In this paper, CST Microwave Studio was used to simulate the performance parameters of the design. Again, the normal incidence electromagnetic wave possessing y polarization propagated along the -z direction.
The simulation results in the uv coordinate system are presented in Figure 2a, in which rvu and ruv were equal to 0, ruu was consistent with rvv, and the phase difference was approximately −180° or 180° in the 2.73–4.63 GHz band. It showed that the reflected electric field agreed with that in Equation (3), and the proposed polarization converter effectively converted y-polarized incident waves into cross-polarized reflected waves in the mentioned broadband range. To estimate the polarization conversion performance, the cross-polarization reflection coefficient amplitude r x y = E x r / E y i , the co-polarization reflection coefficient amplitude r y y = E y r / E y i , and the cross-polarization conversion rate P C R = r x y 2 / ( r x y 2 + r y y 2 ) were obtained. Obviously, it can be seen from Figure 2b,c that rxy (green) is far greater than ryy (blue), and the PCR was above 99% in the range of 2.73–4.63 GHz, which was superior to the ones found in [28,29,30,31,32] (in which, the reported optimal converters could only operate with a PCR ≥ 99% at certain frequencies, and operated with a PCR < 99% and even at 90% in most of frequency bands) and indicated that the proposed polarization converter was characterized by high efficiency and a wide band, with a relative bandwidth of 52%. Moreover, Figure 2b shows that four resonances, such as 2.78, 3.09, 3.68, and 4.54 GHz, contributed to the perfecter cross-polarization conversion. In other words, four resonant frequencies were brought closer to result in the broadband cross-polarization conversion [33].
To reveal the physical mechanism of the proposed MPC, the co-polarization reflection coefficient magnitudes and the surface current distributions for normal incidence at resonant frequencies of 2.78, 3.09, 3.68, and 4.54 GHz were plotted, with the results shown in Figure 3. In Figure 3a, the black stars and solid lines (red, green, blue, pink, and brown) express the yy polarized reflection coefficient magnitudes for models with SRR1, SRR2, SRR3, then with SRR1 and SRR2, next with SRR1and SRR3, and finally with SRR2 and SRR3, respectively. The comparison of results (green, pink, and brown solid lines) implied that the SRR1 made no impression on the resonance at 2.78 GHz when depending on SRR2 & SRR3, which is clear in Figure 3b. Furthermore, it is observed from Figure 3b,f that the strong surface current induced on the SRR2 metal was antiparallel to that on the ground sheet at 2.78 GHz, which was equivalent to loop currents and resulted in the magnetic dipole moments m1 exciting the magnetic field H1. In addition, H1 decomposed into H1y and H1x components, which were parallel and perpendicular to Ey, respectively. Obviously, H1y resulted in a x-direction reflected electric field, which indicated that the structure was capable of achieving y to x cross-polarized conversion. In contrast to other designs of polarization manipulating devices in which the coupling between different resonators is exploited [34], in the proposed design, the coupling between corresponding SRRs and the ground plane was dominant.
Furthermore, Figure 3a,c–e demonstrate that SRR2 and SRR3, SRR2 and SRR1, and SRR2 determined the resonances at 3.09, 3.68, and 4.54 GHz respectively, at which a similar physical mechanism occurred as depicted in Figure 3g–i. In brief, the x components (H1x, H2x, H3x and H4x) of the induced magnetic field were perpendicular to Ey, which signified that the induced electric field was parallel to Ey and could not excite the cross-polarization. Furthermore, the y components (H1y, H2y, H3y and H4y) of the induced magnetic field were parallel to Ey and induced an electronic field (Ex) perpendicular to Ey, which resulted in the cross-polarization conversion. It can be seen from the simulations that the wideband operation resulted from the superposition of four cross-polarization peaks around 2.78, 3.09, 3.68, and 4.54 GHz.
Next, the polarization converter performance for oblique incidence was considered. For incidence angles θ = 0°, 15°, 30°, 40°, 55°, 70°, and 80°, the cross-polarization reflection coefficient amplitudes are shown in Figure 4. Obviously, the cross-polarization performance could be sustained for θ ≤ 15° when varying from 2.73 to 4.63 GHz. For 30° ≤ θ ≤ 55°, the design presented a narrow stopband near 3.45 GHz, as shown in the inset, which resulted in a dual-band cross polarization conversion. For θ > 55°, the converter performance sharply declined. Clearly, the incidence angle had a great influence on the cross-polarization conversion. The main reason can be seen in Figure 5. For the normal incidence, as shown in Figure 5a, the phase difference ( φ = 2 β z ) of directly reflected wave and emergent wave satisfied the ideal destructive interference condition. Nevertheless, for the oblique incidence, as shown in Figure 5b, φ = 2 β z = 2 β z / cos θ increased with the increase in incidence angle, which broke the ideal destructive interference condition [28]. Finally, the polarization conversion performance degradation occurred, which resulted in some dips. In spite of this, the results show that the proposed design can provide a larger tolerance of oblique incidence in the investigated band, except for a very narrow stopband.

4. Experimental Validations

Based on the proposed design, a sample with an array of 20 × 20 units, was fabricated and measured. As shown in Figure 6a,b, the measurement for normal incidence exploited the free space method, which was performed in an anechoic chamber using a vector network analyzer. Concretely speaking, the experimental configurations were as follows:
  • Two ridge horns operating at 1–6 GHz were connected to a vector network analyzer, and were respectively utilized to generate the y-polarized plane wave at nearly normal incidence (the oblique angle was less than 10°) and received both the co-polarization and cross-polarization reflective waves;
  • The fabricated sample was placed in front of the horn antennas, and they were surrounded by absorbing materials. Then, original co-polarized reflection wave amplitude ryy and cross-polarized reflection wave amplitude rxy were recorded by the network analyzer respectively, as detailed in Figure 6c.
  • As plotted in Figure 6c, the reflective wave amplitude r of a same-sized metal plate was also measured for the calibration. Consequently, ryy_Exp = ryyr and rxy_Exp = rxyr were the available reflection coefficient amplitudes in dB, as shown in Figure 2b.
Furthermore, the experimental PCR could be obtained, as plotted in Figure 2c. When compared with the simulation results, the measured ones suffered from a slight tolerance, which was probably due to the unavoidable manufacturing error in the sample and the difference in incident waves (i.e., the incident waves used in the simulation and the measurement were the ideal plane wave and the quasi-plane wave respectively). Basically, the measured results were consistent with the simulation results, which verifies the polarization conversion performance of the design.

5. Conclusions

In this paper, we proposed a multilayer broadband and high-performance reflective polarization converter based on metamaterial. The significant bandwidth expansion was attributed to the four electromagnetic resonances generated in multilayer SRRs. The nature of the polarization conversion was elucidated by the reflection amplitude decompositions and surface current distributions at four resonant frequencies. Furthermore, the measurement results were consistent with the numerical simulation, which validates the excellent performance of the proposed polarization converter. The design can be used in antenna radiation, target stealth, and electromagnetic measurement, etc.

Author Contributions

Conceptualization, Y.J. and J.W.; Formal analysis, M.L. and X.H.; Methodology, Y.J. and J.W.; Validation, M.L., X.H. and S.Z.; Writing—original draft, Y.J. and J.W.; Writing—review and editing, Y.J. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 62261015), the Natural Science Foundation of Guangxi (Grant No. 2019GXNSFFA245002), the Key Laboratories for National Defense Science and Technology (Grant No. 202003007), and the Dean Project of Guangxi Key Laboratory of Wireless Wideband Communication and Signal Processing (Grant No. GXKL06190118).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank the editors for technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Menzel, C.; Helgert, C.; Rockstuhl, C.; Kley, E.; Tünnermann, A.; Pertsch, T.; Lederer, F. Asymmetric transmission of linearly polarized light at optical metamaterials. Phys. Rev. Lett. 2010, 104, 253902. [Google Scholar] [CrossRef] [PubMed]
  2. Fedotov, V.; Mladyonov, P.; Prosvirnin, S.; Rogacheva, A.; Chen, Y.; Zheludev, N. Asymmetric Propagation of Electromagnetic Waves through a Planar Chiral Structure. Phys. Rev. Lett. 2006, 97, 167401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Mutlu, M.; Akosman, A.; Serebryannikov, A.; Ozbay, E. Diodelike asymmetric transmission of linearly polarized waves using magnetoelectric coupling and electromagnetic wave tunneling. Phys. Rev. Lett. 2012, 108, 213905. [Google Scholar] [CrossRef] [PubMed]
  4. Pors, A.; Nielsen, M.G.; Bozhevolnyi, S.I. Broadband plasmonic half-wave plates in reflection. Opt. Lett. 2013, 38, 513–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Chen, C. Scattering by a two-dimensional periodic array of conducting plates. IEEE Trans. Antennas Propag. 2003, 18, 660–665. [Google Scholar] [CrossRef]
  6. Pfeiffer, C.; Grbic, A. Cascaded metasurfaces for complete phase and polarization control. Appl. Phys. Lett. 2013, 102, 1232009. [Google Scholar] [CrossRef]
  7. Tierney, B.B. Advances in Emerging Electromagnetics Topics: Metamaterials and Wireless Power Transfer. Ph.D. Thesis, University of Michigan, Ann Arbor, MI, USA, 2016. [Google Scholar]
  8. Boriskin, A.; Sauleau, R. (Eds.) Aperture Antennas for Millimeter and Sub-Millimeter Wave Applications; Springer: Berlin, Germany, 2017. [Google Scholar]
  9. Tharp, J.; López-Alonso, J.; Ginn, J.; Lail, B.; Middleton, C.; Munk, B.; Boreman, G. Demonstration of a single layer meanderline phase retarder at IR. Opt. Lett. 2006, 31, 2687–2689. [Google Scholar] [CrossRef]
  10. Tharp, J.; Alda, J.; Boreman, G. Off-axis behavior of an infrared meanderline waveplate. Opt. Lett. 2007, 32, 2852–2854. [Google Scholar] [CrossRef]
  11. Tharp, J.; Lail, B.; Munk, B.; Boreman, G. Design and demonstration of an infrared meanderline phase retarder. IEEE Tantenn. Propag. 2007, 55, 2983–2988. [Google Scholar] [CrossRef]
  12. Meissner, T.; Wentz, F.J. Polarization rotation and the third Stokes parameter: The effects of spacecraft attitude and Faraday rotation. IEEE Trans. Geosci. Remote Sens. 2006, 44, 506–515. [Google Scholar] [CrossRef]
  13. Mutlu, M.; Ozbay, E. A transparent 90° polarization rotator by combining chirality and electromagnetic wave tunneling. Appl. Phys. Lett. 2012, 100, 051909. [Google Scholar] [CrossRef] [Green Version]
  14. Zhao, Y.; Alu, A. Tailoring the dispersion of plasmonic ’nanorods to realize broadband optical meta-wave plates. Nano Lett. 2013, 13, 1086–1091. [Google Scholar] [CrossRef]
  15. Lee, H.; Liu, Z.; Xiong, Y.; Sun, C.; Zhang, X. Development of optical hyperlens for imaging below the diffraction limit. Opt. Express 2007, 15, 15886–15891. [Google Scholar] [CrossRef] [Green Version]
  16. Jacob, Z.; Alekseyev, L.V.; Narimanov, E. Optical hyperlens: Far-field imaging beyond the diffraction limit. Opt. Express 2006, 14, 8247–8256. [Google Scholar] [CrossRef] [Green Version]
  17. Grigorenko, A.N. Negative refractive index in artificial metamaterials. Opt. Lett. 2006, 31, 2483–2485. [Google Scholar] [CrossRef] [Green Version]
  18. Valentine, J.; Zhang, S.; Zentgraf, T.; Ulin-Avila, E.; Genov, D.A.; Bartal, G.; Zhang, X. Three-dimensional optical metamaterial with a negative refractive index. Nature 2008, 455, 376–379. [Google Scholar] [CrossRef]
  19. Alù, A.; Engheta, N. Plasmonic and metamaterial cloaking: Physical mechanisms and potentials. J. Opt. A Pure Appl. Opt. 2008, 10, 93002. [Google Scholar] [CrossRef] [Green Version]
  20. Boyvat, M.; Hafner, C.V. Molding the Flow of Magnetic Field with Metamaterials: Magnetic Field Shielding. Prog. Electromagn. Res. 2012, 126, 303–316. [Google Scholar] [CrossRef] [Green Version]
  21. Landy, N.I.; Sajuyigbe, S.; Mock, J.J.; Smith, D.R.; Padilla, W.J. Perfect metamaterial absorber. Phys. Rev. Lett. 2008, 100, 207402. [Google Scholar] [CrossRef]
  22. Cui, Y.; Fung, K.H.; Xu, J.; Ma, H.; Jin, Y.; He, S.; Fang, N.X. Ultra-broadband Light Absorption by a Sawtooth Anisotropic Metamaterial Slab. Nano Lett. 2011, 12, 1443–1447. [Google Scholar] [CrossRef]
  23. Dhillon, A.S.; Mittal, D.; Bargota, R. Triple band ultrathin polarization insensitive metamaterial absorber for defense, explosive detection and airborne radar applications. Microw. Opt. Techn. Let. 2019, 61, 89–95. [Google Scholar] [CrossRef]
  24. Wang, S.Y.; Liu, W.; Geyi, W. Dual-band transmission polarization converter based on planar-dipole pair frequency selective surface. Sci. Rep. 2018, 8, 3791. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, L.B.; Zhou, P.H.; Lu, H.; Zhang, L.; Xie, J.; Deng, L. Realization of broadband reflective polarization converter using asymmetric cross-shaped resonator. Opt. Mater. Express 2016, 6, 1393–1404. [Google Scholar] [CrossRef]
  26. Xu, P.; Wang, S.Y.; Wen, G.Y. A linear polarization converter with near unity efficiency in microwave regime. J. Appl. Phys. 2017, 121, 144502. [Google Scholar] [CrossRef]
  27. Loncar, J.; Grbic, A.; Hrabar, S. A Reflective Polarization Converting Metasurface at X-Band Frequencies. IEEE Trans. Antennas Propag. 2018, 66, 3213–3218. [Google Scholar] [CrossRef]
  28. Yu, H.; Wang, X.; Su, J.; Qu, M.; Guo, Q.; Li, Z.; Song, J. Ultrawideband and high-efficient polarization conversion metasurface based on multi-resonant element and interference theory. Opt. Express 2021, 29, 35938–35950. [Google Scholar] [CrossRef]
  29. Kamal, B.; Chen, J.; Zeng, Y.; Ren, J.; Ullah, S.; Khan, W. High efficiency and ultra-wideband polarization converter based on an L-shaped metasurface. Opt. Mater. Express 2021, 11, 1343. [Google Scholar] [CrossRef]
  30. Chen, H.; Wang, J.; Ma, H.; Qu, S.; Xu, Z.; Zhang, A.; Yan, M.; Li, Y. Ultra-wideband polarization conversion metasurfaces based on multiple plasmon resonances. J. Appl. Phys. 2014, 115, 154504. [Google Scholar] [CrossRef]
  31. Sun, H.; Gu, C.; Chen, X.; Li, Z.; Liu, L.; Martin, F. Ultra-wideband and broad-angle linear polarization conversion metasurface. J. Appl. Phys. 2017, 121, 174902. [Google Scholar] [CrossRef] [Green Version]
  32. Khan, M.; Fraz, Q.; Tahir, F. Ultra-wideband cross polarization conversion metasurface insensitive to incidence angle. J. Appl. Phys. 2017, 121, 045103. [Google Scholar] [CrossRef]
  33. Wang, J.; Gao, C.; Jiang, Y.; Akwuruoha, C. Ultra-broadband and polarization-independent planar absorber based on multilayered graphene. Chin. Phys. B 2017, 26, 114102. [Google Scholar] [CrossRef]
  34. Cho, S.H.; Lee, H.; Kwon, H.; Shin, D.W.; Joh, H.K.; Han, K.; Park, J.H.; Cho, B. Association of underweight status with the risk of tuberculosis: A nationwide population-based cohort study. Sci. Rep. 2022, 12, 3518. [Google Scholar] [CrossRef]
Figure 1. Schematics of the uv coordinate system and the proposed MPC UC. (a) The electric field decomposition of the incident wave in the uv coordinate system. (b) 3D perspective, (c) side view, and (d) top view of the UC.
Figure 1. Schematics of the uv coordinate system and the proposed MPC UC. (a) The electric field decomposition of the incident wave in the uv coordinate system. (b) 3D perspective, (c) side view, and (d) top view of the UC.
Crystals 13 00017 g001
Figure 2. Normal incidence results of the proposed model. (a) Reflection coefficient amplitudes and phase differences in uv coordinate system. (b) Reflection coefficient amplitudes in the x-y coordinate system. (c) PCR.
Figure 2. Normal incidence results of the proposed model. (a) Reflection coefficient amplitudes and phase differences in uv coordinate system. (b) Reflection coefficient amplitudes in the x-y coordinate system. (c) PCR.
Crystals 13 00017 g002
Figure 3. Co-polarization reflection coefficient amplitudes and the surface current distributions on SRRs and ground. (a) Co-polarization reflection coefficient amplitudes. 3D SRRs surface current distributions at (b) 2.78 GHz, (c) 3.09 GHz, (d) 3.68 GHz, and (e) 4.54 GHz. 2D ground surface current distributions at (f) 2.78 GHz, (g) 3.09 GHz, (h) 3.68 GHz, and (i) 4.54 GHz.
Figure 3. Co-polarization reflection coefficient amplitudes and the surface current distributions on SRRs and ground. (a) Co-polarization reflection coefficient amplitudes. 3D SRRs surface current distributions at (b) 2.78 GHz, (c) 3.09 GHz, (d) 3.68 GHz, and (e) 4.54 GHz. 2D ground surface current distributions at (f) 2.78 GHz, (g) 3.09 GHz, (h) 3.68 GHz, and (i) 4.54 GHz.
Crystals 13 00017 g003aCrystals 13 00017 g003b
Figure 4. Cross-polarization reflection coefficient amplitudes for various incidence angles.
Figure 4. Cross-polarization reflection coefficient amplitudes for various incidence angles.
Crystals 13 00017 g004
Figure 5. Schematic of incident waves propagating in the MPC described by the reflection, transmission, and emergent wave resulting from the multiple reflections. (a) Normal incidence; (b) Oblique incidence.
Figure 5. Schematic of incident waves propagating in the MPC described by the reflection, transmission, and emergent wave resulting from the multiple reflections. (a) Normal incidence; (b) Oblique incidence.
Crystals 13 00017 g005
Figure 6. Experimental environment and measurement data. (a) Diagram of measurement. (b) Measurement setup in the microwave anechoic chamber. (c) Measured reflection wave amplitudes.
Figure 6. Experimental environment and measurement data. (a) Diagram of measurement. (b) Measurement setup in the microwave anechoic chamber. (c) Measured reflection wave amplitudes.
Crystals 13 00017 g006
Table 1. Structural Parameters of Converter. (Unit: mm).
Table 1. Structural Parameters of Converter. (Unit: mm).
ParameterValueParameterValue
r14g21.2
r26g32
r37.7h12.4
r14.8h23
r26.8h32.4
r38.8h45
g11.2p22
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

Jiang, Y.; Li, M.; Wang, J.; Huang, X.; Zhang, S. A Metamaterial-Based Cross-Polarization Converter Characterized by Wideband and High Efficiency. Crystals 2023, 13, 17. https://doi.org/10.3390/cryst13010017

AMA Style

Jiang Y, Li M, Wang J, Huang X, Zhang S. A Metamaterial-Based Cross-Polarization Converter Characterized by Wideband and High Efficiency. Crystals. 2023; 13(1):17. https://doi.org/10.3390/cryst13010017

Chicago/Turabian Style

Jiang, Yannan, Mianji Li, Jiao Wang, Xialin Huang, and Shitian Zhang. 2023. "A Metamaterial-Based Cross-Polarization Converter Characterized by Wideband and High Efficiency" Crystals 13, no. 1: 17. https://doi.org/10.3390/cryst13010017

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