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

Transformation optics for perfect two-dimensional non-magnetic all-mode waveguide couplers

Open Access Open Access

Abstract

Here, we demonstrate that transformation optics can be used to produce 2-D non-magnetic waveguide couplers with no reflections. Our approach consists of using a scaling function for reflection suppression and introducing an auxiliary function in the transformation optics formulation to achieve a non-magnetic medium for coupling the TM polarization. To demonstrate the potential of this method, two non-magnetic waveguide couplers are designed. The first one satisfies the Brewster angle condition for any arbitrary incidence angle (TMn modes), extending the performance of couplers previously reported in the literature that only operate for TEM (TM0 mode), i.e. waves with normal incidence. Our method can be applied to match any given dielectric constant. Our results demonstrate that for a given mode (angle), we achieve a perfect match to a defined dielectric constant. The second design removes the dependence of the reflectionless condition to the incident angle at the boundary. Hence, this coupler works for all incident angles (TMn modes). It is used to compress all the modes into a region with a higher predefined refractive index.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Transformation optics is a powerful tool that enables a systematic design of electromagnetic devices. The concept of transformation optics was introduced by Prof. Pendry [1] and Leonhardt [2,3]. Although the original idea was applied for cloaking [4–11], transformation optics was promptly adopted for a huge number of applications, including polarization splitters [12–17], phase transformers [18–20], flat lenses [21–31], field rotators [32,33], field concentrators [34–36] waveguide bends [37–39] and waveguide couplers [40–46]. Transformation optics is a technique that can be employed to tailor the propagation of electromagnetic fields in a given physical space. Such tailoring can also be realized using properly engineered metasurfaces [47–50] which may alleviate the need for a bulky medium [49].

One straightforward use of transformation optics is to compress/expand the space [37, 51]. This device can be considered as a waveguide coupler that transfers the energy between two waveguides with different widths. In order to compress the space, a coordinate transformation must be applied. For example, if a compression wants to be applied along the y-axis of Cartesian coordinates, the coordinate transformation must be: x′ = x, y′ = m(x)y, z′ = z with m(x) as the transition function. With this transformation, a rectangular virtual space of dimensions L and d is transformed into a semi-trapezoid shape in the physical space as illustrated in Fig. 1, where d′ indicates the length of the physical coupler, and the shape of transition between waveguides is controlled by the function m(x) that satisfies m(0) = 1 and m(d) = M. M × L is the width of the second waveguide where M indicates the amount of compression or expansion. In the particular case illustrated in Fig. 1, the transition function is a cosine with power of 2. Cosine, exponential and polynomial series can be employed as transition functions. These specific functions can be expressed as:

m1(x)=(1M)n=1Nαn(1x/d)n+M.
m2(x)=(1M)n=1Nαncosn(πx/2d)+M.
m3(x)=n=1NαnM(x/d)n.

In the m(x) functions defined in Eqs. (1)(3), n=1Nαn=1 must be satisfied, so m(0) = 1. Additionally, in order to have a smooth variation in the transition, it is advantageous to choose a function in which m′(0) = m′(d) = 0. To have this property for all the proposed transition functions included in Eqs. (1)(3), α1 must be zero and n=2Nnαn=0 for Eq. (1) and (3). Note that as long as the transition function satisfies the conditions at the transition with the waveguides, the device will properly operate. On the other hand, a smooth transition function will lead to smooth changes of material inside the coupler, producing a better coupling of the waves and a device that is easier to manufacture and more robust to fabrication errors.

 figure: Fig. 1

Fig. 1 Scheme of the physical space for a waveguide coupler with a cosine transition function, m(x), with power of 2.

Download Full Size | PDF

When considering a TM mode, the problem can be simplified as two plane waves bouncing off the top and bottom sides with perfect electric conductor (PEC) boundaries [52]. The PEC is used here to realize the waveguide boundaries that lead to the existence of TMn modes and to prevent wave leakage. For an optical incident beam, one may remove PEC and the device will maintain its functionality. The propagation vectors and the angle of their constitutive plane waves, namely β and θ, are not necessarily equal for both sides. This propagation is illustrated in Fig. 1 including the propagation vectors.

Since the transformation is continuous at the input boundary of the coupler, no reflection is produced at this boundary for any incident angle. It has been previously reported that this type of device suffers from reflections at the output boundary because of a metric mismatch between the squeezer and the surrounding space [37]. It has been also demonstrated that under special conditions, 3-D squeezers can be reflectionless [40]. In [40], it was demonstrated that if the compression/expansion is applied with similar rates to both transverse directions, y and z-directions in Fig. 1, the device will exhibit no reflections for normal incidence at both sides when surrounded by vacuum. In [41], the reflection and transmission coefficients of a 3-D squeezer/expander were calculated. The device proposed in [41] had no reflections for a longitudinal compression. Additionally, for the TM polarization, if the compression ratio in the z-direction is square of the compression ratio in the y-direction, the device is omnidirectional reflectionless at the output with a dielectric constant equal to the compression ratio in the z-direction. Finally, if the compression ratio is similar in two transverse directions, the device will be matched to the vacuum at both sides for normal incidence, which is the same result obtained in [40].

Reflection suppression in 2-D and 3-D couplers requires distinct approaches to be followed since the definition of the modes and the Brewster angle are different, being also the transformation different. Strategies for suppressing reflections in a 2-D squeezer/expander were investigated in [44]. This solution is based on an optical transformation that alters the impedance in the virtual space for a specified incidence angle. This coupler can be matched to similar dielectric constants at the input and output. In addition, the material for the normal incidence case was non-magnetic. In [45], an auxiliary function was proposed to produce a coupler that is matched to any given dielectric materials at both boundaries for normal incidence with a medium that is non-magnetic. Normal incidence means that the TEM (TM0) mode can travel from a vacuum waveguide to a waveguide with the desired dielectric constant. Here, we extend the work in [45] for any TMn mode by presenting a Brewster angle coupler that uses only non-magnetic materials. This means that the incidence angle of the wave is no longer normal to the boundary of the coupler. Also, differently to [45], our approach couples all modes to a dielectric constant predefined by the amount of compression.

It is worth to mention that we focus here on TM modes, however, the best strategy to couple TE modes with transformation optics reported in the literature is the use of quasi-conformal mappings that leads to a non-homogeneous solution that can be implemented with fully dielectric and isotropic materials. This implementation reduces the complexity of the manufacturing, although unwanted reflections are seen at the output of the device [30,53].

In this paper, we propose the use of transformation optics to eliminate the reflections from the output boundary of the 2-D waveguide coupler for an arbitrary incident angles, i.e. all the TMn modes. Our generalized formulation of transformation optics provides extra degrees of freedom for a designer to achieve more suitable transition shapes. Using the Brewster angle definition, we present a coupler suitable for a single-mode transmission. This coupler is matched to an arbitrary dielectric at the output. Also, we illustrate that for a given level of compression, an all-mode coupler can be produced for a specified dielectric constant at the output. In order to impose non-magnetic materials, an auxiliary function is employed. The validity of this technique is proven with simulated results calculated with COMSOL Multiphysics.

2. Theoretical basis

2.1. Transformation optics

Transformation optics theory is based on the form-invariance of Maxwell equations under a coordinate transformation. The coordinate transformation between two spaces, named virtual and physical spaces, is defined by the Jacobian matrix J = (x′, y′, z′) / (x, y, z). This theory implies that the material in the virtual space (taken here as vacuum) is transformed to the following relative permittivity ε and permeability μ tensors in the physical space [1]:

ε=μ=JJTdet(J).

It must be remarked that Maxwell equations imply that constitutive parameters that play a role in the propagation of TE (Ez, Hx, Hy) and TM (Hz, Ex, Ey) polarizations are εzz, μxx, μyy, μxy, μyx and μzz, εxx, εyy, εxy, εyx. Furthermore, if the dispersion equation is kept intact, that is, the multiplication of εzz(μzz) by μxx, μyy, μxy, μyx(εxx, εyy, εxy, εyx) is kept unchanged, the wave propagation direction is not altered. Hence, a smooth engineered spatial variation can be applied to permittivity and permeability parameters without changing the functionality of the device. However, reflections may occur at the boundaries due to the impedance mismatch with respect to the surrounding media. This property has been employed for several purposes in the literature including simplification of the transformation material [4–6, 8, 35, 36, 42, 46] and reflection suppression [45].

The dispersion equation for TM modes and the Brewster angle condition for normal incidence TM0 case (TEM mode) were derived in [45]. This formulation can be generalized to oblique incidence, TMn modes, resulting in the following equation for the Brewster angle:

cos2θB=εi(μzzεxx|iεi)εuεv|iεi2,
where θB is the specified Brewster angle (mode angle for our case) and subscript i is 1 or 2, corresponding to the left and right boundaries. εu and εv are the eigen values of the permittivity tensor in the xy plane. Dielectrics in the left and right regions are denoted as εi.

To eliminate the reflections for all modes (i.e. all incident angles), the constitutive parameters of the transformation medium must satisfy a specific condition at the boundaries with the surrounding medium. Using Eq. (5) and aiming to eliminate the dependence of the equation to the Brewster angle θB and hence achieving an all-mode reflectionless coupler, the conditions at the boundaries along the y-direction are:

εuεv|i=εi2.
εxxμzz|i=εi.
From Eqs. (5)(7), we can conclude that the reflections at a boundary along the y-direction in the transformed medium are governed by two values, εuεv and εxx μzz.

2.2. Non-magnetic Brewster-angle waveguide coupler

Let’s consider the case where the goal is to match two waveguides with arbitrary dielectric constants. Similar to [45], this matching can be achieved by scaling to the constitutive parameters and multiplying (dividing) εxx, εxy, εyx, εyy(μzz) by the scaling function s(x). In this case, the dispersion equation (also the εxx μzz quantity) is kept unchanged while εuεv is multiplied by s2(x). Using Eq. (5) and the transformation material of the coupler, the output-matched dielectric constant for the device with scaled parameters will be:

cos2θB|x=d=ε2(1/M2ε2)s2(d)ε22ε2=1±1M4s2(d)sin2(2θB)2M2sin2θB.
Figure 2 illustrates the matched dielectric constants versus the corresponding Brewster angles (TMn modes angles) for different values of M2s(d) and M = 0.5.

 figure: Fig. 2

Fig. 2 Output matched dielectric constant versus the Brewster angle based on Eq. (8) for a device with scaled parameters and M = 0.5. (a) M2s(d) < 1 and “+” sign and (b) M2s(d) < 1 and “−” sign. (c) M2s(d) > 1 and “+” sign and (d) M2s(d) > 1 and “−” sign.

Download Full Size | PDF

From Eq. (8) and as illustrated in Fig. 2(a) and Fig. 2(b), we can conclude that the values of matched dielectric constant vary in the range of [∞, M−2] or [M2s2(d), 0] for the Brewster angle range of [0, 90]. Additionally, from Fig. 2(c) and Fig. 2(d) we observe that for the dielectric constants range of [∞, (M2M4s2(d))1] or [M2s2(d), (M2M4s2(d))1], the valid Brewster angle values cover the range of [0, 0.5sin−1(M−2s−1(d))].

Note that the dielectric constant of the right waveguide ε2 in Eq. (8) and Fig. 2 is a desired design parameter and not a required value. Hence, for a given value of ε2 and the incident angle of the TMn mode, the value of s(d), which is used in the coupler design procedure, is derived from Eq. (8), as represented in Fig. 2. The incident angles of the TMn modes are θi = sin−1(fc/f), where fc is the cut-off frequency of the mode and f is the operating frequency [52]. After calculating the value of s(d), the full function, s(x), can be any continuous function that satisfies the values of s(0) = 1 and s(d). The proposed series functions in Eqs. (1)(3) are good candidates for this purpose since one can also engineer the slope of the function at the beginning and the end of the coupler.

To make the medium non-magnetic, we introduce the auxiliary function f(x) that applies a transformation to x-constant lines. The transformation formula is now x′ = f(x), y′ = m(x)y, z′ = z and the transformation medium parameters can be found in Eq. (7) of [45]. This auxiliary function does not perturb the reflectionless condition as it only compresses/expands the x-constant coordinate line in the longitudinal direction. Note that the magnetic parameter of the coupler is μzz = 1/(s(x)m(x)fx(x)). This indicates that if the function f(x) is selected in such a way to satisfy fx(x) = 1/s(x)m(x), the transformation medium becomes non-magnetic. This auxiliary function can also change the coupler length if needed since d′ = f(d). Hence, in summary, s(x) produces a reflectionless waveguide coupler, while f(x) enforces only non-magnetic materials.

2.3. Non-magnetic all-mode waveguide coupler

In this section, we enforce the use of only non-magnetic materials while maintaining the reflectionless property at the output boundary for all modes. An all-mode reflectionless device implies that different waveguide modes, or waves with different incident angles, passing through the coupler do not have reflections. As in the previous section, we apply here again the scaling function s(x), being Eq. (6) and Eq. (7) changed to:

{εuεv|x=0=s2(0)εuεv|x=d=s2(d)
{εxxμzz|x=0=1εxxμzz|x=d=1/M2

Hence, for an all-mode coupler, s(x) should satisfy s(0) = 1 and s(d) = 1/M2 conditions. Such a device is impedance-matched to ε1 = 1 at the input port and to ε2 = 1/M2 at the output port. As previously stated, the scaling function s(x) can be any continuous function that satisfies the above conditions. One choice is s(x) = 1/m2(x) that is both smooth and differentiable and also meets the above requirements.

To make the medium non-magnetic, the magnetic parameter is μzz = 1/(s(x)m(x)fx(x)) after the scaling. Considering s(x) = 1/m2(x), then μzz = m(x)/fx(x). Hence, the auxiliary function can lead to the non-magnetic property if fx(x) = m(x) condition holds. In summary, it is possible to have an all-mode reflectionless non-magnetic coupler by choosing s(x) = 1/m2(x) and f(x) = ∫ m(x)dx. Note that if another scaling function s(x) is used, the auxiliary function f(x) changes too.

3. Simulation results

To confirm the functionality and the reflectionless property of our proposed designs, we carried out numerical simulations. In all simulations, the width ratio of the waveguides is 2:1 meaning that M = 0.5. TM modes are excited and the normalized real part and magnitude of Hz component are represented. The simulation frequency for all cases except the last one is f=8fcTM1 where fcTM1=c0/2L is the cut-off frequency of the vacuum left waveguide for the TM1 mode and c0 is the speed of light in the vacuum. The incident wave is excited with a port on the left side boundary.

In the first case, we aim to compare the coupling performance between two vacuum waveguides with and without the non-magnetic coupler. The design parameters are L = 0.7m and d′ = 0.4m. Based on Eq. (1), the following transition function is used.

m(x)=(1M)(3(1x/d)22(1x/d)3)+M,
which is a polynomial series and satisfies m(0) = 1, m(d) = M conditions as well as m′(0) = m′(d) = 0. Note that we assign later a value to the parameter d in order to achieve the desired coupler length d′. The left port is excited with the TM0 (TEM) mode which corresponds to the incidence angle of zero. Eq. (8) shows that s(d) = 2 ensures the reflectionless coupling for this mode. Since the scaling function must satisfy two conditions, s(0) = 1 and s(d) = 2, one smooth choice can be the following equation:
s(x)=(1s(d))(3(1x/d)22(1x/d)3)+s(d).
The auxiliary function should now fulfill f(x) = ∫ 1/s(x)m(x) to make the material non-magnetic. The calculation is done numerically. Under these conditions, d = 0.42m is needed in order to have d′ = f(d) = 0.4m. Figure 3 represents the effect of non-magnetic coupler between the vacuum waveguides.

 figure: Fig. 3

Fig. 3 Simulated magnetic field for the TM0 mode. The medium in both left and right waveguides is a vacuum. (a) Real part and (b) magnitude of Hz without the non-magnetic coupler. (c) Real part and (d) magnitude of Hz for the non-magnetic coupler with the Brewster angle of zero.

Download Full Size | PDF

The reflection coefficient in the absence of the coupler medium is Γ ≃ 0.21 whereas Γ ≃ 0.004 is achieved in the presence of the non-magnetic coupler. Figure 3(a) and Fig. 3(b) illustrated that the wave does not only have reflections, but the phase is also distorted.

Next, we excite the left waveguide boundary with a TM3 mode. This means that the coupler should be designed for the incidence angle of θB = sin−1(fc/f) = sin−1(3/4) ≃ 49°, and based on Eq. (8), s(d) = 2.82. The transition function m(x) and scaling function s(x) follow the same formula as Eq. (11) and Eq. (12), respectively. The auxiliary function need to be recalculated for the new values. In this case, d = 0.52m leads to d′ = f(d) = 0.4m.

Figure 4 illustrates the comparison between a coupler designed for θB ≃ 49° and a coupler designed for the normal incidence case.

 figure: Fig. 4

Fig. 4 Simulated magnetic field for the non-magnetic Brewster-angle coupler for the TM3 mode. The medium in both left and right waveguides is a vacuum. (a) Real part and (b) magnitude of Hz for the design with Brewster angle of 49°. (c) Real part and (d) magnitude of Hz for the design with Brewster angle of 0 (normal incidence).

Download Full Size | PDF

The reflection coefficient of the Brewster-angle design is Γ ≃ 0 whereas the normal incidence design has a Γ ≃ 0.17 reflection coefficient. The reflections are actually easy to notice since the magnitude plot is highly modulated as a sign of the reflected wave presence in Fig. 4(d). Clearly, the performance of the normal incident design degrades when the incident angle deviates highly from zero.

Next, the non-magnetic all-mode design is simulated. As previously derived, for M = 0.5, the device is impedance-matched to ε2 = M−2 = 4. Here, the following transition function is used, corresponding to Eq. (2):

m(x)=(1M)cos2(πx/2d)+M.
This function has a zero slope when connecting with the waveguides and also satisfies m(0) = 1, m(d) = M conditions. As a result, the scaling function equals s(x) = 1/m2(x) and the auxiliary function is derived as follows:
f(x)=m(x)dx=x(M+1)/2d(M1)sin(πx/d)/(2π).
To have the same coupler length as the previous simulation case, d = 0.533m is chosen leading to d′ = f(d) = d(M + 1)/2 = 0.4m. Figure 5 and Fig. 6 illustrate the simulated results for this case. The left waveguide is excited with TM0 and TM1 in Fig. 5 and with TM2 and TM3 modes in Fig. 6. These TM modes correspond to the incident angles of 0, 7, 14 and 22 degrees. The reflection coefficient of these cases is Γ ≃ 0.004. This indicates that the input power is delivered almost perfectly to the output.

 figure: Fig. 5

Fig. 5 Simulated magnetic field for the non-magnetic all-mode coupler impedance-matched to ε2 = M−2 = 4. (a) Real part and (b) magnitude of Hz for TM0 mode. (c) Real part and (d) magnitude of Hz for TM1 mode.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Simulated magnetic field for the non-magnetic all-mode coupler impedance-matched to ε2 = M−2 = 4. (a) Real part and (b) magnitude of Hz for TM2 mode. (c) Real part and (d) magnitude of Hz for TM3 mode.

Download Full Size | PDF

The coupling efficiency of this coupler is tested for different incidence angles for the TM3 mode. To reach high incidence angles, the frequency should get close to the cut-off frequency of the mode. Figure 7 represents the coupling efficiency versus the incidence angle. Note that in case of the incidence angle of 70 degrees, the simulation frequency is f=fcTM3/sin(70)=1.06fcTM3, which is near cut-off and is not commonly used in practical applications.

4. Conclusion

Unwanted reflections from the output boundary of a waveguide coupler are investigated and mitigated in this work and a non-magnetic medium is obtained for TM polarization. Using a suitable parameter scaling function, a Brewster-angle design is proposed that suppresses the reflection for a specified incident angle (mode) and has the refractive index of the output waveguide as a free parameter. Also, an all-mode coupler is introduced that is reflectionless for all modes but its impedance is matched to a predefined refractive index at the output. For both devices, the auxiliary function makes the material non-magnetic for TM polarization and eases the fabrication for optical applications.

 figure: Fig. 7

Fig. 7 Coupling efficiency of the non-magnetic coupler versus the incidence angle for TM3 mode.

Download Full Size | PDF

References

1. J. B. Pendry, “Controlling electromagnetic fields,” Science 312, 1780–1782 (2006). [CrossRef]   [PubMed]  

2. U. Leonhardt, “Optical conformal mapping,” Science 312, 1777–1780 (2006). [CrossRef]   [PubMed]  

3. U. Leonhardt and T. G. Philbin, “General relativity in electrical engineering,” New J. Phys. 8, 247 (2006). [CrossRef]  

4. D. Schurig, J. J. Mock, B. J. Justice, S. A. Cummer, J. B. Pendry, A. F. Starr, and D. R. Smith, “Metamaterial electromagnetic cloak at microwave frequencies,” Science 314, 977–980 (2006). [CrossRef]   [PubMed]  

5. W. Cai, U. K. Chettiar, A. V. Kildishev, V. M. Shalaev, and G. W. Milton, “Nonmagnetic cloak with minimized scattering,” Appl. Phys. Lett. 91, 111105 (2007). [CrossRef]  

6. W. Cai, U. K. Chettiar, A. V. Kildishev, and V. M. Shalaev, “Optical cloaking with metamaterials,” Nat. Photonics 1, 224–227 (2007). [CrossRef]  

7. W. Yan, M. Yan, Z. Ruan, and M. Qiu, “Coordinate transformations make perfect invisibility cloaks with arbitrary shape,” New J. Phys. 10, 043040 (2008). [CrossRef]  

8. W. Yan, M. Yan, and M. Qiu, “Non-magnetic simplified cylindrical cloak with suppressed zeroth order scattering,” Appl. Phys. Lett. 93, 021909 (2008). [CrossRef]  

9. J. Li and J. B. Pendry, “Hiding under the carpet: A new strategy for cloaking,” Phys. Rev. Lett. 101, 203901 (2008). [CrossRef]  

10. R. Liu, C. Ji, J. J. Mock, J. Y. Chin, T. J. Cui, and D. R. Smith, “Broadband ground-plane cloak,” Science 323, 366–369 (2009). [CrossRef]   [PubMed]  

11. H. F. Ma and T. J. Cui, “Three-dimensional broadband ground-plane cloak made of metamaterials,” Nat. Commun. 1, 1–6 (2010).

12. D.-H. Kwon and D. H. Werner, “Polarization splitter and polarization rotator designs based on transformation optics,” Opt. Express 16, 18731 (2008). [CrossRef]  

13. S. S. S. Mousavi, M. S. Majedi, and H. Eskandari, “Design and simulation of polarization transformers using transformation electromagnetics,” Optik 130, 1099–1106 (2017). [CrossRef]  

14. H. Eskandari, M. S. Majedi, and A. R. Attari, “Non-reflecting non-magnetic homogeneous polarization splitter and polarization deflector design based on transformation electromagnetics,” Optik 135, 407–416 (2017). [CrossRef]  

15. H. Eskandari, M. S. Majedi, and A. R. Attari, “Design of reflectionless non-magnetic homogeneous polarization splitters with minimum anisotropy based on transformation electromagnetics,” J. Opt. Soc. Am. B 34, 1191 (2017). [CrossRef]  

16. P. A. Teixeira, D. G. Silva, L. H. Gabrielli, D. H. Spadoti, and M. A. F. C. Junqueira, “General multimode polarization splitter design in uniaxial media,” Opt. Eng. 57, 1 (2018). [CrossRef]  

17. H. Eskandari, A. R. Attari, and M. S. Majedi, “Design of polarization splitting devices with ideal transmission and anisotropy considerations,” J. Opt. Soc. Am. B 35, 1585 (2018). [CrossRef]  

18. H. Ma, S. Qu, Z. Xu, and J. Wang, “General method for designing wave shape transformers,” Opt. Express 16, 22072 (2008). [CrossRef]   [PubMed]  

19. L. Lin, W. Wang, J. Cui, C. Du, and X. Luo, “Design of electromagnetic refractor and phase transformer using coordinate transformation theory,” Opt. Express 16, 6815 (2008). [CrossRef]   [PubMed]  

20. W. Shu, S. Yang, W. Yan, Y. Ke, and T. Smith, “Flat designs of impedance-matched nonmagnetic phase transformer and wave-shaping polarization splitter via transformation optics,” Opt. Commun. 338, 307–312 (2015). [CrossRef]  

21. D.-H. Kwon and D. H. Werner, “Transformation optical designs for wave collimators, flat lenses and right-angle bends,” New J. Phys. 10, 115023 (2008). [CrossRef]  

22. D.-H. Kwon and D. H. Werner, “Flat focusing lens designs having minimized reflection based on coordinate transformation techniques,” Opt. Express 17, 7807 (2009). [CrossRef]   [PubMed]  

23. N. Kundtz and D. R. Smith, “Extreme-angle broadband metamaterial lens,” Nat. Mater. 9, 129–132 (2009). [CrossRef]   [PubMed]  

24. D. A. Roberts, N. Kundtz, and D. R. Smith, “Optical lens compression via transformation optics,” Opt. Express 17, 16535 (2009). [CrossRef]   [PubMed]  

25. H. F. Ma and T. J. Cui, “Three-dimensional broadband and broad-angle transformation-optics lens,” Nat. Commun. 1, 124 (2010). [CrossRef]   [PubMed]  

26. W. Tang, C. Argyropoulos, E. Kallos, W. Song, and Y. Hao, “Discrete coordinate transformation for designing all-dielectric flat antennas,” IEEE Transactions on Antennas Propag. 58, 3795–3804 (2010). [CrossRef]  

27. Q. Wu, Z. H. Jiang, O. Quevedo-Teruel, J. P. Turpin, W. Tang, Y. Hao, and D. H. Werner, “Transformation optics inspired multibeam lens antennas for broadband directive radiation,” IEEE Transactions on Antennas Propag. 61, 5910–5922 (2013). [CrossRef]  

28. C. Mateo-Segura, A. Dyke, H. Dyke, S. Haq, and Y. Hao, “Flat luneburg lens via transformation optics for directive antenna applications,” IEEE Transactions on Antennas Propag. 62, 1945–1953 (2014). [CrossRef]  

29. S. Jain, R. Mittra, and S. Pandey, “Flat-base broadband multibeam luneburg lens for wide-angle scan,” J. Electromagn. Waves Appl. 29, 1329–1341 (2015). [CrossRef]  

30. M. Ebrahimpouri and O. Quevedo-Teruel, “Bespoke lenses based on quasi-conformal transformation optics technique,” IEEE Transactions on Antennas Propag. 65, 2256–2264 (2017). [CrossRef]  

31. Y. Su and Z. N. Chen, “A flat dual-polarized transformation-optics beamscanning luneburg lens antenna using PCB-stacked gradient index metamaterials,” IEEE Transactions on Antennas Propag. 66, 5088–5097 (2018). [CrossRef]  

32. H. Chen and C. T. Chan, “Transformation media that rotate electromagnetic fields,” Appl. Phys. Lett. 90, 241105 (2007). [CrossRef]  

33. H. Chen, B. Hou, S. Chen, X. Ao, W. Wen, and C. T. Chan, “Design and experimental realization of a broadband transformation media field rotator at microwave frequencies,” Phys. Rev. Lett. 102, 183903 (2009). [CrossRef]  

34. M. Rahm, D. Schurig, D. A. Roberts, S. A. Cummer, D. R. Smith, and J. B. Pendry, “Design of electromagnetic cloaks and concentrators using form-invariant coordinate transformations of maxwell’s equations,” Photonics Nanostructures - Fundamentals Appl. 6, 87–95 (2008). [CrossRef]  

35. B. Bian, S. Liu, S. Wang, X. Kong, Y. Guo, X. Zhao, B. Ma, and C. Chen, “Cylindrical optimized nonmagnetic concentrator with minimized scattering,” Opt. Express 21, A231 (2013). [CrossRef]   [PubMed]  

36. S.-Y. Wang, B. Yu, S. Liu, and B. Bian, “Optimization for nonmagnetic concentrator with minimized scattering,” J. Opt. Soc. Am. A 30, 1563 (2013). [CrossRef]  

37. M. Rahm, D. A. Roberts, J. B. Pendry, and D. R. Smith, “Transformation-optical design of adaptive beam bends and beam expanders,” Opt. Express 16, 11555 (2008). [CrossRef]   [PubMed]  

38. L. H. Gabrielli, D. Liu, S. G. Johnson, and M. Lipson, “On-chip transformation optics for multimode waveguide bends,” Nat. Commun. 3, 2232 (2012). [CrossRef]   [PubMed]  

39. S. Viaene, V. Ginis, J. Danckaert, and P. Tassin, “Transforming two-dimensional guided light using nonmagnetic metamaterial waveguides,” Phys. Rev. B 93, 085429 (2016). [CrossRef]  

40. C. D. Emiroglu and D.-H. Kwon, “Impedance-matched three-dimensional beam expander and compressor designs via transformation optics,” J. Appl. Phys. 107, 084502 (2010). [CrossRef]  

41. C. García-Meca, M. M. Tung, J. V. Galán, R. Ortuño, F. J. Rodríguez-Fortuño, J. Martí, and A. Martínez, “Squeezing and expanding light without reflections via transformation optics,” Opt. Express 19, 3562 (2011). [CrossRef]   [PubMed]  

42. H. Y. Xu, B. Zhang, G. Barbastathis, and H. D. Sun, “Compact optical waveguide coupler using homogeneous uniaxial medium,” J. Opt. Soc. Am. B 28, 2633 (2011). [CrossRef]  

43. P. Markov, J. G. Valentine, and S. M. Weiss, “Fiber-to-chip coupler designed using an optical transformation,” Opt. Express 20, 14705 (2012). [CrossRef]   [PubMed]  

44. J. Cao, L. Zhang, S. Yan, and X. Sun, “Reflectionless design of optical elements using impedance-tunable transformation optics,” Appl. Phys. Lett. 104, 191102 (2014). [CrossRef]  

45. H. Eskandari, M. S. Majedi, and A. R. Attari, “Reflectionless compact nonmagnetic optical waveguide coupler design based on transformation optics,” Appl. Opt. 56, 5599 (2017). [CrossRef]   [PubMed]  

46. H. Eskandari, A. R. Attari, and M. S. Majedi, “Reflectionless design of a nonmagnetic homogeneous optical waveguide coupler based on transformation optics,” J. Opt. Soc. Am. B 35, 54 (2017). [CrossRef]  

47. F. Aieta, P. Genevet, M. A. Kats, N. Yu, R. Blanchard, Z. Gaburro, and F. Capasso, “Aberration-free ultrathin flat lenses and axicons at telecom wavelengths based on plasmonic metasurfaces,” Nano Lett. 12, 4932–4936 (2012). [CrossRef]   [PubMed]  

48. X. Ni, S. Ishii, A. V. Kildishev, and V. M. Shalaev, “Ultra-thin, planar, babinet-inverted plasmonic metalenses,” Light. Sci. & Appl. 2, e72 (2013). [CrossRef]  

49. X. Ni, Z. J. Wong, M. Mrejen, Y. Wang, and X. Zhang, “An ultrathin invisibility skin cloak for visible light,” Science 349, 1310–1314 (2015). [CrossRef]   [PubMed]  

50. Z.-L. Deng, J. Deng, X. Zhuang, S. Wang, K. Li, Y. Wang, Y. Chi, X. Ye, J. Xu, G. P. Wang, R. Zhao, X. Wang, Y. Cao, X. Cheng, G. Li, and X. Li, “Diatomic metasurface for vectorial holography,” Nano Lett. 18, 2885–2892 (2018). [CrossRef]   [PubMed]  

51. R. C. Mitchell-Thomas and O. Quevedo-Teruel, Transformation Optics Applied to Antennas and Focusing Systems (Springer International Publishing, 2018), pp. 387–406.

52. D. M. Pozar, Microwave Engineering (Wiley, 2011), 4th ed.

53. O. Quevedo-Teruel, W. Tang, R. C. Mitchell-Thomas, A. Dyke, H. Dyke, L. Zhang, S. Haq, and Y. Hao, “Transformation optics for antennas: why limit the bandwidth with metamaterials?” Sci. reports 3, 1903 (2013). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Scheme of the physical space for a waveguide coupler with a cosine transition function, m(x), with power of 2.
Fig. 2
Fig. 2 Output matched dielectric constant versus the Brewster angle based on Eq. (8) for a device with scaled parameters and M = 0.5. (a) M2s(d) < 1 and “+” sign and (b) M2s(d) < 1 and “−” sign. (c) M2s(d) > 1 and “+” sign and (d) M2s(d) > 1 and “−” sign.
Fig. 3
Fig. 3 Simulated magnetic field for the TM0 mode. The medium in both left and right waveguides is a vacuum. (a) Real part and (b) magnitude of Hz without the non-magnetic coupler. (c) Real part and (d) magnitude of Hz for the non-magnetic coupler with the Brewster angle of zero.
Fig. 4
Fig. 4 Simulated magnetic field for the non-magnetic Brewster-angle coupler for the TM3 mode. The medium in both left and right waveguides is a vacuum. (a) Real part and (b) magnitude of Hz for the design with Brewster angle of 49°. (c) Real part and (d) magnitude of Hz for the design with Brewster angle of 0 (normal incidence).
Fig. 5
Fig. 5 Simulated magnetic field for the non-magnetic all-mode coupler impedance-matched to ε2 = M−2 = 4. (a) Real part and (b) magnitude of Hz for TM0 mode. (c) Real part and (d) magnitude of Hz for TM1 mode.
Fig. 6
Fig. 6 Simulated magnetic field for the non-magnetic all-mode coupler impedance-matched to ε2 = M−2 = 4. (a) Real part and (b) magnitude of Hz for TM2 mode. (c) Real part and (d) magnitude of Hz for TM3 mode.
Fig. 7
Fig. 7 Coupling efficiency of the non-magnetic coupler versus the incidence angle for TM3 mode.

Equations (14)

Equations on this page are rendered with MathJax. Learn more.

m 1 ( x ) = ( 1 M ) n = 1 N α n ( 1 x / d ) n + M .
m 2 ( x ) = ( 1 M ) n = 1 N α n cos n ( π x / 2 d ) + M .
m 3 ( x ) = n = 1 N α n M ( x / d ) n .
ε = μ = J J T det ( J ) .
cos 2 θ B = ε i ( μ z z ε x x | i ε i ) ε u ε v | i ε i 2 ,
ε u ε v | i = ε i 2 .
ε x x μ z z | i = ε i .
cos 2 θ B | x = d = ε 2 ( 1 / M 2 ε 2 ) s 2 ( d ) ε 2 2 ε 2 = 1 ± 1 M 4 s 2 ( d ) sin 2 ( 2 θ B ) 2 M 2 sin 2 θ B .
{ ε u ε v | x = 0 = s 2 ( 0 ) ε u ε v | x = d = s 2 ( d )
{ ε x x μ z z | x = 0 = 1 ε x x μ z z | x = d = 1 / M 2
m ( x ) = ( 1 M ) ( 3 ( 1 x / d ) 2 2 ( 1 x / d ) 3 ) + M ,
s ( x ) = ( 1 s ( d ) ) ( 3 ( 1 x / d ) 2 2 ( 1 x / d ) 3 ) + s ( d ) .
m ( x ) = ( 1 M ) cos 2 ( π x / 2 d ) + M .
f ( x ) = m ( x ) d x = x ( M + 1 ) / 2 d ( M 1 ) sin ( π x / d ) / ( 2 π ) .
Select as filters


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