Open Access
25 June 2014 Evolution of surface-plasmon-polariton and Dyakonov–Tamm waves with the ambichirality of a partnering dielectric material
Author Affiliations +
Abstract
The planar interface of an isotropic homogeneous metal and an ambichiral dielectric material can guide surface-plasmon-polariton (SPP) waves. The planar interface of an isotropic, homogeneous dielectric material and an ambichiral dielectric material can guide Dyakonov–Tamm waves. In either instance, we found that, as the ambichiral partnering material evolves into a finely chiral material, the solutions of the dispersion equation for surface-wave propagation evince convergence. The convergence is faster for Dyakonov–Tamm waves than for SPP waves.

1.

Introduction

Anticipating the discovery of cholesteric liquid crystals by about two decades,1,2 Reusch3 proposed in 1869 that a periodically nonhomogeneous multilayered material reflects normally incident circularly polarized light of one handedness, but not of the opposite handedness, provided that all layers are made of the same homogeneous, uniaxial dielectric material such that the optic axis in each layer is rotated about the thickness direction with respect to the optic axis in the adjacent layer by a fixed angle. Such a periodically nonhomogeneous dielectric material is nowadays called a Reusch pile.

Extensive theoretical and experimental work by Joly and colleagues47 showed that the circular-polarization-selective reflection of normally incident light by a Reusch pile may occur in several spectral regimes. This selective reflection of circularly polarized light of one handedness, but very little of the other, in a given spectral regime is commonly called the circular Bragg phenomenon.8,9

A classification scheme based on the number N of layers in each period of a Reusch pile was developed by Hodgkinson et al.10 If N=2, the Reusch pile is classified as an equichiral material; if N>2, but not very large, it can be called an ambichiral material; and if N, it is a finely chiral material. Equichiral materials do not exhibit the circular Bragg phenomenon. Ambichiral materials may exhibit the circular Bragg phenomenon in several spectral regimes, depending on the variations of their constitutive parameters with frequency. A cholesteric liquid crystal11 can be considered as a finely chiral Reusch pile made of uniaxial dielectric layers.

Reusch piles can also be made of biaxial dielectric material such as columnar thin films (CTFs).12 A chiral sculptured thin film (STF)8 can be considered to be a finely chiral Reusch pile comprising biaxial CTFs. Chiral STFs were first fabricated by Young and Kowal13 in 1959 and were rediscovered in the 1990s.14 They have been extensively studied since then for optical applications exploiting the circular Bragg phenomenon.8,9

The effect of the number N of layers in a period on the circular Bragg phenomenon has been studied.15 Both N and the total number of periods have to be substantially large for the circular Bragg phenomenon to fully develop.15

Now, the planar interface of an isotropic homogeneous metal and an ambichiral dielectric material can guide surface-plasmon-polariton (SPP) waves. The planar interface of an isotropic, homogeneous dielectric material and an ambichiral dielectric material can guide Dyakonov–Tamm waves. What is the effect of N on both types of surface waves? The results reported in this communication elucidate the evolution of the solution(s) of the dispersion equation for surface-wave propagation with N.

The plan of this communication is as follows. Section 2 succinctly presents the common formulation of the canonical boundary-value problem for both types of surface waves. Numerical results are presented and discussed in Sec. 3. An exp(iωt) dependence on time t is implicit, with ω denoting the angular frequency and i=1. Furthermore, k0=ωμ0ε0 and λ0=2π/k0, respectively, represent the free-space wavenumber and free-space wavelength, where μ0 is the permeability and ε0 is the permittivity of free space. Vectors are in boldface, dyadics are double-underlined, and the three Cartesian unit vectors are denoted by u^x, u^y, and u^z.

2.

Theoretical Preliminaries

The canonical boundary-value problem of surface-wave propagation is depicted schematically in Fig. 1. The half space z<0 is occupied by an isotropic and homogeneous material with relative permittivity εs. The half space z>0 is occupied by an ambichiral dielectric material comprising homogeneous layers each of thickness D, the ’th layer occupying the region (1)D<z<D, [1,). The relative permittivity dyadic is given as

Eq. (1)

ϵ̲̲(z,ω)=S̲̲z(hξ+hγ)·S̲̲y(χ)·ϵ̲̲ref°(ω)·S̲̲y1(χ)·S̲̲z1(hξ+hγ),(1)D<z<D,[1,),
where the reference permittivity dyadic

Eq. (2)

ϵ̲̲ref°(ω)=u^zu^zεa(ω)+u^xu^xεb(ω)+u^yu^yεc(ω)
contains the eigenvalues εa,b,c(ω) of ϵ̲̲(z,ω). The dyadic

Eq. (3)

S̲̲y(χ)=(u^xu^x+u^zu^z)cosχ+(u^zu^xu^xu^z)sinχ+u^yu^y
depends on the tilt angle χ[0deg,90deg] with respect to the xy plane, the dyadic

Eq. (4)

S̲̲z(ξ)=(u^xu^x+u^yu^y)cosξ+(u^yu^xu^xu^y)sinξ+u^zu^z
represents a rotation about the z axis by an angle ξ, ξ=(1)π/N with N1 being the number of layers in each period 2Ω=ND, Ω is the half period, right-handed rotation is represented by h=1 and left-handed rotation by h=1, and γ is an angular offset with respect to the x axis.

Fig. 1

Schematic of the canonical boundary-value problem.

JNP_8_1_083082_f001.png

In the region z<0, the electric field phasor may be written as1618

Eq. (5)

E(x,z)=[a1u^y+a2(αsu^x+qu^zk0ns)]exp[i(qxαsz)],z<0,
where q2+αs2=k02εs, ns=εs, q is the complex-valued wavenumber of the surface wave, Im(αs)>0 for attenuation as z, and a1 and a2 are unknown scalars with the same units as the electric field.

For field representation in the region z>0, let us write18

Eq. (6)

E(x,z)=e(z)exp(iqx),H(x,z)=h(z)exp(iqx).
The Cartesian components of the electric and magnetic field phasors tangential to the xy plane are used to form the column vector

Eq. (7)

[f(z)]=[ex(z)ey(z)hx(z)hy(z)]
which satisfies the matrix differential equation

Eq. (8)

ddz[f(z)]=i[P̲̲(z)]·[f(z)],z>0,
where the 4×4 matrix [P̲̲(z)] depends not only on ε̲̲(z,ω) but also on q.

The piecewise-uniform approximation technique8 can be used to determine the matrix [Q˜̲̲] that appears in the relation

Eq. (9)

[f(2Ω)]=exp{i2Ω[Q˜̲̲]}·[f(0+)]
for specific values of q. Let [t](n), n[1,4], be the eigenvector corresponding to the n’th eigenvalue αn of [Q˜̲̲]. After ensuring that Im(α1,2)>0, we set

Eq. (10)

[f(0+)]=[[t](1)[t](2)]·[b1b2]
for surface-wave propagation, where b1 and b2 are unknown dimensionless scalars; the other two eigenvalues of [Q˜̲̲] pertain to waves that amplify as z and cannot therefore contribute to the surface wave. At the same time, [f(0)] can be obtained from Eq. (5) and the corresponding magnetic field phasor H=i×E/ωμ0.

Enforcement of the usual boundary conditions across the plane z=0 requires that [f(0)]=[f(0+)], which may be rearranged as the matrix equation

Eq. (11)

[Y̲̲]·[a1a2b1b2]=[0000],
leading to the dispersion equation

Eq. (12)

det[Y̲̲(q)]=0.

3.

Numerical Results and Discussion

The dispersion equation (12) was solved using the Newton–Raphson method,19 with λ0 fixed at 633 nm. For all numerical results presented here, the ambichiral dielectric material was taken to comprise CTFs made by directing a collimated evaporant flux of patinal titanium oxide in a low-pressure chamber at a fixed angle χv(0deg,90deg] with respect to a planar substrate.20 For the chosen CTF,

Eq. (13)

{εa=[1.0443+2.7394(2χv/π)1.3697(2χv/π)2]2εb=[1.6765+1.5649(2χv/π)0.7825(2χv/π)2]2εc=[1.3586+2.1109(2χv/π)1.0554(2χv/π)2]2χ=tan1(2.8818tanχv)
according to Hazel and co-workers.20 We fixed Ω=200nm, while varying N[1,15] (so that D=2Ω/N was simultaneously varied) and γ{0deg,45deg,90deg}. Furthermore, χv=20deg was fixed so that εa=2.5135, εb=3.9426, and εc=3.1528.

3.1.

Surface-Plasmon-Polariton Waves

Let the isotropic homogeneous partnering material be thin-film aluminum with εs=14.65+i5.85 (Ref. 21) at λ0=633nm. Then, all solutions q of the dispersion equation (12) represent SPP waves.16,18 These solutions are complex valued because εs is complex valued.

Only one solution of the dispersion equation exists for any γ[0deg,90deg] when the anisotropic partnering material is homogeneous (i.e., N=1).22 However, when that partnering material is periodically nonhomogeneous (i.e., N>1), the solutions can be organized into two branches for γ=0deg and 45 deg but only one for γ=90deg, as shown in Fig. 2. Convergence on either branch is monotonic for N>2.

Fig. 2

Real and imaginary parts of the solutions q of Eq. (12) as functions of N[1,15] and γ{0deg,45deg,90deg} for SPP waves guided by the planar interface of aluminum (εs=14.65+i5.85) and an ambichiral dielectric material characterized by Eq. (13) with χv=20deg, when λ0=633nm and Ω=200nm. Top panels: First branch of solutions. Bottom panels: Second branch of solutions.

JNP_8_1_083082_f002.png

The most notable feature of the solutions presented in Fig. 2 is their evolution and convergence as N increases. Although not shown in this figure, both the real and imaginary parts of q on the first branch in Fig. 2 lie within 0.1% of the corresponding solutions for the metal/chiral-STF interface16 when N18, and on the second branch in that figure lie within 0.1% when N55. Also, Re(q) converges faster with respect to N than Im(q) does.

In order to delineate the effect of N on the localization of SPP waves to the interface z=0, the penetration depth δs=1/Im(αs) into aluminum is presented in Fig. 3 in relation to N for all solutions of Eq. (12). The penetration depth converges to 23nm for the first branch and to 24nm for the second branch as N increases.

Fig. 3

The penetration depth δs of SPP waves into aluminum for the (top) first and (bottom) second branches of solutions provided in Fig. 2.

JNP_8_1_083082_f003.png

Since the ambichiral material is periodically nonhomogeneous and anisotropic, two decay constants β1,2=exp[2ΩIm(α1,2)](0,1) are defined18 to quantify localization in the anisotropic partnering material. The smaller that min{β1,β2} is, the higher is that localization of the SPP wave to the interface. The decay constants for all SPP waves found are presented in Fig. 4.

Fig. 4

The decay constants β1,2 of SPP waves for the (top) first and (bottom) second branches of solutions provided in Fig. 2.

JNP_8_1_083082_f004.png

The SPP waves on the first branch are highly localized to the interface within the ambichiral material as both decay constants converge to values <0.05 with increasing N. The SPP waves on the second branch are loosely bound to the interface, and the degree of localization strongly depends on the direction of propagation. Both decay constants converge to 0.6 for γ=0deg, but both converge to 0.9 for γ=45deg. Furthermore, data for N>15 (not shown) indicated that δs, β1, and β2 converge to within 0.1% of their respective values when N29 on the first branch and N30 on the second branch.

3.2.

Dyakonov–Tamm Waves

Next, let the isotropic homogeneous partnering material be magnesium fluoride—a dielectric material with εs=1.896 at λ0=633nm—instead of a metal. Every solution of the dispersion equation (12) represents a Dyakonov–Tamm wave17—named thus because this wave has the attributes of both the Dyakonov wave23,24 and the Tamm wave18,25—since both partnering materials are dielectric materials and one of the two is anisotropic and periodically nonhomogeneous normal to the waveguiding interface for N2.17,18

For every γ{0deg,45deg,90deg}, only one solution of Eq. (12) was found. Dissipation being absent in both partnering materials, q is real valued. The dependence of q on N is shown in Fig. 5. Also provided in the same figure are δs and β1,2 in relation to N.

Fig. 5

Solution q, penetration depth δs into magnesium fluoride (εs=1.896), and the decay constants β1,2 in the ambichiral dielectric material characterized by Eqs. (13) with χv=20deg, as functions of N[1,15] and γ{0deg,45deg,90deg}, for Dyakonov–Tamm waves guided by the interface of magnesium fluoride and the ambichiral dielectric material, when λ0=633nm and Ω=200nm.

JNP_8_1_083082_f005.png

For N=1, the anisotropic partner is homogeneous and the solution q for γ=0deg represents a Dyakonov wave;23,24 no solution was found for γ{45deg,90deg}. For N>1, only one solution was found regardless of the value of γ, and that solution represents a Dyakonov–Tamm wave. Figure 5 indicates the typical difference between Dyakonov and Dyakonov–Tamm waves:17 The range of γ is much larger for surface waves of the latter type than for the surface waves of the former type.

Just like the solutions presented in Fig. 2 for SPP waves, those presented in Fig. 5 for Dyakonov–Tamm waves also evolve as N increases. Specifically, the solutions in Fig. 5 converge monotonically to within 0.1% of the corresponding solution for the isotropic-dielectric/chiral-STF interface17 when N9.

Figure 5 shows that the penetration depth of the Dyakonov–Tamm wave into the homogeneous partnering material depends upon γ and, hence, the direction of propagation, but is about four times greater than for SPP waves. The plots of the decay constants show that the degree of localization of the Dyakonov–Tamm wave in the ambichiral material increases as γ increases. The values of β1,2 converge to 0.8 for γ=0deg and to 0.5 for γ=45deg and 90 deg. Furthermore, δs, β1, and β2 converge to within 0.1% when N32.

3.3.

Sufficient Value of N

In order to fabricate a structurally chiral material with continuous variation of ε̲̲(z,ω) with z as an ambichiral material, a sufficient value of N must be chosen based upon the tolerances in the values of the wavenumber q, penetration depth δs, and the smaller decay constant min{β1,β2}. If all tolerances are chosen to be 0.1%, as in the previous subsections, N=55 is sufficient for the SPP wave, and N=32 for the Dyakonov–Tamm wave. However, if all tolerances are chosen to be 1%, N=19 suffices for the SPP wave and N=11 for the Dyakonov–Tamm wave.

Let us note that the thickness D of each CTF in the ambichiral material is 2Ω/N21nm for N=19 and 36nm for N=11. If the thickness of each CTF either equals or is less than one-tenth of the smallest wavelength inside the ambichiral material, then each CTF is electrically thin.26 For the chosen parameters, one-tenth of the smallest wavelength inside the ambichiral material is λ0/(10εb)32nm since εb>εc>εa>1. Therefore, if Dλ0/(10max{εa,εb,εc}), the ambichiral material has a sufficiently smooth variation of ε̲̲(z,ω) with z and all important quantities characterizing a surface wave will converge to within 1%.

4.

Concluding Remarks

The canonical boundary-value problem of surface-wave propagation guided by the planar interface of an isotropic homogeneous material and an ambichiral material was set up and solved. Both SPP and Dyakonov–Tamm waves were investigated. As the number N of layers per period in the ambichiral partnering material was increased, the solutions of the dispersion equation for surface-wave propagation were found to converge to those for the ambichiral material replaced by the corresponding finely chiral material. The convergence is faster when the homogeneous partnering material is dielectric than when it is metallic. The real part of the wavenumber q of an SPP wave converges faster than the imaginary part with respect to N. The real and imaginary parts of the surface wavenumber, the penetration depth into the isotropic partner, and the decay constants in the ambichiral dielectric material converge to within 1% if the layers in the ambichiral material are electrically thin.

Acknowledgments

MF and AL are grateful for partial support from Grant No. DMR-1125591 of the US National Science Foundation. AL is also grateful to the Charles Godfrey Binder Endowment at the Pennsylvania State University for partial support of this work.

References

1. 

F. Reinitzer, “Beiträge zur Kenntiss des Cholesterins,” Monat. Chem. (Wien), 9 (1), 421 –441 (1888). http://dx.doi.org/10.1007/BF01516710 0026-9247 Google Scholar

2. 

T. J. SluckinD. A. DunmurH. Stegemeyer, Crystals That Flow: Classic Papers From the History of Liquid Crystals, Taylor & Francis, London (2004). Google Scholar

3. 

E. Reusch, “Untersuchung über Glimmercombinationen,” Ann. Phys. Chem. Lpz., 138 (12), 628 –638 (1869). http://dx.doi.org/10.1002/(ISSN)1521-3889 ARPLAP 0066-426X Google Scholar

4. 

G. JolyJ. Billard, “Quelques champs électromagnétiques dans les piles de Reusch I.—Les vibrations propres d’une pile de deux lames a biréfringence rectiligne ne sont pas orthogonales,” J. Opt. (Paris), 12 (5), 323 –329 (1981). http://dx.doi.org/10.1088/0150-536X/12/5/006 JOOPDB 0150-536X Google Scholar

5. 

G. JolyJ. Billard, “Quelques champs électromagnétiques dans les piles de Reusch II. Piles éclairées sous l’incidence normale par des ondes monochromatiques planes et uniformes,” J. Opt. (Paris), 13 (4), 227 –238 (1982). http://dx.doi.org/10.1088/0150-536X/13/4/008 JOOPDB 0150-536X Google Scholar

6. 

G. JolyN. Isaert, “Quelques champs électromagnétiques dans les piles de Reusch III. Biréfringence elliptique des vibrations itératives; activité optique de piles hélicoïdales d’extension finie,” J. Opt. (Paris), 16 (5), 203 –213 (1985). http://dx.doi.org/10.1088/0150-536X/16/5/001 JOOPDB 0150-536X Google Scholar

7. 

G. JolyN. Isaert, “Quelques champs électromagnétiques dans les piles de Reusch IV—Domaines multiples de réflexion sélective,” J. Opt. (Paris), 17 (5), 211 –221 (1986). http://dx.doi.org/10.1088/0150-536X/17/5/001 JOOPDB 0150-536X Google Scholar

8. 

A. LakhtakiaR. Messier, Sculptured Thin Films: Nanoengineered Morphology and Optics, SPIE Press, Bellingham, Washington (2005). Google Scholar

9. 

M. FaryadA. Lakhtakia, “The circular Bragg phenomenon,” Adv. Opt. Photon., 6 (2), 225 –292 (2014). http://dx.doi.org/10.1364/AOP.6.000225 AOPAC7 1943-8206 Google Scholar

10. 

I. J. Hodgkinsonet al., “Ambichiral, equichiral and finely chiral layered structures,” Opt. Commun., 239 (4–6), 353 –358 (2004). http://dx.doi.org/10.1016/j.optcom.2004.06.005 OPCOB8 0030-4018 Google Scholar

11. 

P. G. de GennesJ. A. Prost, The Physics of Liquid Crystals, 2nd ed.Clarendon, Oxford (1993). Google Scholar

12. 

I. J. HodgkinsonQ.-h. Wu, Birefringent Thin Films and Polarizing Elements, World Scientific, Singapore (1997). Google Scholar

13. 

N. O. YoungJ. Kowal, “Optically active fluorite films,” Nature, 183 (4654), 104 –105 (1959). http://dx.doi.org/10.1038/183104a0 NATUAS 0028-0836 Google Scholar

14. 

K. RobbieM. J. BrettA. Lakhtakia, “First thin film realization of a helicoidal bianisotropic medium,” J. Vac. Sci. Technol. A, 13 (6), 2991 –2993 (1995). http://dx.doi.org/10.1116/1.579626 JVTAD6 0734-2101 Google Scholar

15. 

I. Abdulhalim, “Effect of the number of sublayers on axial optics of anisotropic helical structures,” Appl. Opt., 47 (16), 3002 –3008 (2008). http://dx.doi.org/10.1364/AO.47.003002 APOPAI 0003-6935 Google Scholar

16. 

J. A. Polo Jr.A. Lakhtakia, “On the surface plasmon polariton wave at the planar interface of a metal and a chiral sculptured thin film,” Proc. R. Soc. Lond. A, 465 (2101), 87 –107 (2009). http://dx.doi.org/10.1098/rspa.2008.0211 PRLAAZ 1364-5021 Google Scholar

17. 

A. LakhtakiaJ. A. Polo Jr., “Dyakonov–Tamm wave at the planar interface of a chiral sculptured thin film and an isotropic dielectric material,” J. Eur. Opt. Soc.–Rapid Publ., 2 (1), 07021 (2007). http://dx.doi.org/10.2971/jeos.2007.07021 1990-2573 Google Scholar

18. 

J. A. Polo Jr.T. G. MackayA. Lakhtakia, Electromagnetic Surface Waves: A Modern Perspective, Elsevier, Waltham, MA (2013). Google Scholar

19. 

Y. Jaluria, Computer Methods for Engineering, Taylor & Francis, Washington, DC (1996). Google Scholar

20. 

I. J. HodgkinsonQ. h. WuJ. Hazel, “Empirical equations for the principal refractive indices and column angle of obliquely-deposited films of tantalum oxide, titanium oxide and zirconium oxide,” Appl. Opt., 37 (13), 2653 –2659 (1998). http://dx.doi.org/10.1364/AO.37.002653 APOPAI 0003-6935 Google Scholar

21. 

A. LakhtakiaY.-J. JenC.-F. Lin, “Multiple trains of same-color surface plasmon-polaritons guided by the planar interface of a metal and a sculptured nematic thin film. Part III: Experimental evidence,” J. Nanophoton., 3 (1), 033506 (2009). http://dx.doi.org/10.1117/1.3249629 JNOACQ 1934-2608 Google Scholar

22. 

J. A. Polo Jr.A. Lakhtakia, “Morphological effects on surface-plasmon-polariton waves at the planar interface of a metal and a columnar thin film,” Opt. Commun., 281 (21), 5453 –5457 (2008). http://dx.doi.org/10.1016/j.optcom.2008.07.051 OPCOB8 0030-4018 Google Scholar

23. 

M. I. Dyakonov, “New type of electromagnetic wave propagating at an interface,” Sov. Phys. JETP, 67 (4), 714 –716 (1988). SPHJAR 0038-5646 Google Scholar

24. 

O. Takayamaet al., “Dyakonov surface waves: a review,” Electromagnetics, 28 (3), 126 –174 (2008). http://dx.doi.org/10.1080/02726340801921403 ETRMDV 0272-6343 Google Scholar

25. 

P. YehA. YarivC.-S. Hong, “Electromagnetic propagation in periodic stratified media. I. General theory,” J. Opt. Soc. Am., 67 (4), 423 –438 (1977). http://dx.doi.org/10.1364/JOSA.67.000423 JOSAAH 0030-3941 Google Scholar

26. 

P. S. ReeseA. Lakhtakia, “Low-frequency electromagnetic properties of an alternating stack of thin uniaxial dielectric laminae and uniaxial magnetic laminae,” Z. Naturforsch. A, 46 (5), 384 –388 (1991). ZTAKDZ 0340-4811 Google Scholar

Biography

Muhammad Faryad received his MSc and MPhil degrees in electronics from Quaid-i-Azam University in 2006 and 2008, respectively, and his PhD (2012) degree in engineering science and mechanics from the Pennsylvania State University, where he is presently a postdoctoral research scholar. His research interests include modeling of thin-film solar cells, electromagnetic surface waves, chemical sensing, circular Bragg phenomenon, and sculptured thin films.

Akhlesh Lakhtakia is the Charles Godfrey Binder (Endowed) professor of Engineering Science and Mechanics at the Pennsylvania State University. He is a fellow of SPIE, Optical Society of America, American Association for the Advancement of Science, American Physical Society, and Institute of Physics (United Kingdom). He was the sole recipient of the 2010 SPIE Technical Achievement Award. His current research interests include nanophotonics, surface multiplasmonics, complex materials including mimumes and sculptured thin films, bone refacing, bioreplication, and forensic science.

© 2014 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2014/$25.00 © 2014 SPIE
Muhammad Faryad and Akhlesh Lakhtakia "Evolution of surface-plasmon-polariton and Dyakonov–Tamm waves with the ambichirality of a partnering dielectric material," Journal of Nanophotonics 8(1), 083082 (25 June 2014). https://doi.org/10.1117/1.JNP.8.083082
Published: 25 June 2014
Lens.org Logo
CITATIONS
Cited by 1 scholarly publication.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Dielectrics

Interfaces

Wave propagation

Waveguides

Contrast transfer function

Aluminum

Metals


CHORUS Article. This article was made freely available starting 25 June 2015

Back to Top