Open Access
18 June 2012 Bruggeman formalism versus "Bruggeman formalism": particulate composite materials comprising oriented ellipsoidal particles
Author Affiliations +
Abstract
Two different formalisms for the homogenization of composite materials containing oriented ellipsoidal particles of isotropic dielectric materials are being named after Bruggeman. Numerical studies reveal clear differences between the two formalisms which may be exacerbated: (i) if the component particles become more aspherical, (ii) at mid-range values of the volume fractions, and (iii) if the homogenized component material is dissipative. Only the correct Bruggeman formalism uses the correct polarizability density dyadics of the component particles in the homogenized composite material (HCM) and is directly derivable from the frequency-domain Maxwell postulates specialized for the HCM.

The Bruggeman formalism provides a well-established technique for estimating the effective constitutive parameters of homogenized composite materials (HCMs).13 The scope of its applicability is not restricted to dilute composite materials and it is easy to implement numerically, both of which contribute to its enduring popularity.

The Bruggeman formalism was originally devised for isotropic dielectric HCMs, comprising two (or more) isotropic dielectric component materials distributed randomly as electrically small spherical particles.4 Generalizations of the Bruggeman formalism which accommodate anisotropic and bianisotropic HCMs have been developed.5 A rigorous basis for the Bruggeman formalism—for isotropic dielectric,6 anisotropic dielectric,7,8 and bianisotropic9,10 HCMs—is provided by the strong-property-fluctuation theory, whose lowest-order formulation is the Bruggeman formalism. (In Ref. 7, the formula of Polder and van Santen, for the effective relative permittivity of an isotropic dielectric HCM, yields the same results as the formula of Bruggeman.4)

Our focus in this letter is on HCMs arising from two isotropic dielectric component materials, labeled a and b. Their relative permittivities are ϵa and ϵb, while their volume fractions are fa and fb1fa. Both component materials are assumed to be randomly distributed as electrically small ellipsoidal particles. For simplicity, all component particles have the same shape and orientation. The surface of each ellipsoid, relative to its centroid, may be represented by the vector

Eq. (1)

r̲e(θ,ϕ)=ηU̲̲r̲^(θ,ϕ),
with r̲^ being the radial unit vector from the ellipsoid’s centroid, specified by the spherical polar coordinates θ and ϕ. The linear dimensions of each ellipsoid, as determined by the parameter η, are assumed to be small relative to the electromagnetic wavelength(s). Let us choose our coordinate system to be such that the Cartesian axes are aligned with the principal axes of the ellipsoids. Then the ellipsoidal shape is captured by the dyadic

Eq. (2)

U̲̲=Uxx̲^x̲^+Uyy̲^y̲^+Uzz̲^z̲^,
wherein the shape parameters Ux,y,z>0 and {x̲^,y̲^,z̲^} are unit vectors aligned with the Cartesian axes.

The ellipsoidal shape of the component particles results in the corresponding HCM being an orthorhombic biaxial dielectric material. That is, the Bruggeman estimate of the HCM relative permittivity dyadic has the form

Eq. (3)

ϵ̲̲Br1=ϵxBr1x̲^x̲^+ϵyBr1y̲^y̲^+ϵzBr1z̲^z̲^.
The relative permittivity parameters ϵx,y,zBr1 are given implicitly by the three coupled equations11

Eq. (4)

ϵaϵBr11+D(ϵaϵBr1)fa+ϵbϵBr11+D(ϵbϵBr1)fb=0,({x,y,z}).
Herein D are components of the depolarization dyadic

Eq. (5)

D̲̲=Dxx̲^x̲^+Dyy̲^y̲^+Dzz̲^z̲^,
specified by the double integrals5

Eq. (6)

Dx=14π02πdϕ0πdθsin3θcos2ϕUx2ρDy=14π02πdϕ0πdθsin3θsin2ϕUy2ρDz=14π02πdϕ0πdθsinθcos2θUz2ρ},
which involve the scalar parameter

Eq. (7)

ρ=sin2θcos2ϕUx2ϵxBr1+sin2θsin2ϕUy2ϵyBr1+cos2θUz2ϵzBr1.
The coupled nature of the three equations in Eq. (4) means that numerical methods are generally needed to extract the relative permittivity parameters ϵx,y,zBr1 from them.

An alternative formalism for the homogenization of the same composite material as in the foregoing paragraph is also referred to as the Bruggeman formalism.1215 Let us write the estimate of the HCM’s relative permittivity dyadic provided by this alternative formalism as

Eq. (8)

ϵ̲̲Br2=ϵxBr2x̲^x̲^+ϵyBr2y̲^y̲^+ϵzBr2z̲^z̲^.
The relative permittivity parameters ϵx,y,zBr2 are given by the three equations

Eq. (9)

ϵaϵBr2ϵBr2+L(ϵaϵBr2)fa+ϵbϵBr2ϵBr2+L(ϵbϵBr2)fb=0,({x,y,z}),
wherein the depolarization factors16

Eq. (10)

L=UxUyUz20ds1(s+U2)(s+Ux2)(s+Uy2)(s+Uz2),({x,y,z})
are components of the depolarization dyadic

Eq. (11)

L̲̲=Lxx̲^x̲^+Lyy̲^y̲^+Lzz̲^z̲^.
Each of the three equations in Eq. (9) is a quadratic equation in ϵBr2 whose solution may be explicitly expressed as

Eq. (12)

ϵBr2=β±β24αγ2α,({x,y,z}),
with α=L1, β=ϵa(faL)+ϵb(fbL), and γ=Lϵaϵb. The sign of the square root term in the solution seen in Eq. (12) may be determined by appealing to the anisotropic dielectric generalization of the Hashin–Shtrikman bounds,17 for example.

Let us illustrate the differences between the estimates ϵ̲̲Br1 and ϵ̲̲Br2 by means of some representative numerical results. The two estimates are identical for the limiting case represented by Ux=Uy=Uz=1 (i.e., for isotropic dielectric HCMs), but differences emerge as the asphericity of the component particles intensifies. Suppose that the shape parameters describing the component ellipsoids have the form Ux=1, Uy=1+(Δ/3), and Uz=1+2Δ. Thus, the asphericity of the ellipsoids is governed by the scalar parameter Δ. We begin with the nondissipative scenario wherein ϵa{0.5,1.5} and ϵb=12. Also, we fix fa=0.5. Plots of the relative permittivity parameters ϵx,y,zBr1,Br2 versus the asphericity parameter Δ are presented in Fig. 1. The difference between ϵxBr1 and ϵxBr2 grows steadily as Δ increases, reaches a maximum for 1<Δ<2, and then slowly shrinks as Δ increases beyond 2. The difference between ϵzBr1 and ϵzBr2 follows a similar pattern. However, in the case of ϵyBr1 and ϵyBr2, the difference increases uniformly as Δ increases. The differences between ϵx,y,zBr1 and ϵx,y,zBr2 are generally greater for ϵa=0.5 than for ϵa=1.5. In the former case the maximum difference is approximately 15%, whereas in the latter case it is approximately 5%.

Fig. 1

The estimates ϵxBr1,2 (blue, dashed curves), ϵyBr1,2 (green, solid curves), and ϵzBr1,2 (red, broken dashed curves) plotted versus the asphericity parameter Δ(0,4.5). The ϵx,y,zBr1 estimates are represented by thick curves and the ϵx,y,zBr2 estimates are represented by thin curves. The ellipsoidal shapes of the component material particles are described by shape parameters Ux=1, Uy=1+(Δ/3), and Uz=1+2Δ. The relative permittivities of the component materials are ϵa{0.5,1.5} and ϵb=12; and the volume fraction fa=0.5.

JNP_6_1_069501_f001.png

We turn now to the effect of volume fraction. The calculations of Fig. 1 are repeated for Fig. 2 except that here the relative permittivity parameters ϵx,y,zBr1,Br2 are plotted versus the volume fraction fa, while the asphericity parameter is fixed at Δ=4.5. The differences between the estimates of the two formalisms are clearly greatest at mid-range values of fa, and they are generally greater for ϵa=0.5 than for ϵa=1.5.

Fig. 2

As Fig. 1 except that Δ=4.5 and the estimates ϵx,y,zBr1,2 are plotted versus the volume fraction fa(0,1).

JNP_6_1_069501_f002.png

Lastly, the effects of dissipation are considered. We repeated the calculations of Fig. 1 but with Δ=4.5 and ϵa{0.5+iδ,1.5+iδ}. Here δ>0 governs the degree of dissipation exhibited by component material a. The real and imaginary parts of the relative permittivity parameters ϵx,y,zBr1,Br2 are plotted versus the dissipation parameter δ in Fig. 3. The differences between the real parts of the estimates ϵx,y,zBr1 and ϵx,y,zBr2 are largest when component material a is nondissipative and they decrease uniformly as δ increases. In contrast, the differences between the imaginary parts of the estimates ϵx,y,zBr1 and ϵx,y,zBr2 increase as δ increases. These differences in the imaginary parts generally reach a maximum for mid-range values of δ and thereafter decrease as δ increases. For both the real and imaginary parts of the estimates ϵx,y,zBr1 and ϵx,y,zBr2, generally larger differences arise for ϵa=0.5+iδ than for ϵa=1.5+iδ.

Fig. 3

As Fig. 1 except that Δ=4.5, ϵa{0.5+iδ,1.5+iδ}, and the real and imaginary parts of the estimates ϵx,y,zBr1,2 are plotted versus the dissipation parameter δ(0,2).

JNP_6_1_069501_f003.png

Thus, there are significant differences between the estimates ϵ̲̲Br1 and ϵ̲̲Br2 when ellipsoidal component particles are considered. These differences may be exacerbated: (i) if the component particles become more aspherical, (ii) at mid-range values of the volume fractions of the component materials, and (iii) if the HCM is dissipative. The differences between the two estimates may be further exacerbated if one of the component materials has a positive-valued relative permittivity which is less than unity (or a relative permittivity whose real part is positive-valued and less than unity). Relative permittivities in this range are associated with novel materials possessing engineered nanostructures; these artificial materials have been the subject of intense research lately.1820 (The parameter regime wherein one of the component materials has a positive-valued relative permittivity while the other has a negative-valued relative permittivity (or likewise for the real parts of the relative permittivities in the case of dissipative HCMs) is avoided here because the Bruggeman formalism can deliver estimates in this regime which are not physically plausible.21)

The differences between the two formalisms stem from the differences between the depolarization dyadics D̲̲ and L̲̲. The Bruggeman formalism conceptually employs an average-polarizability-density approach:22 Suppose the composite material has been homogenized into an HCM. Into this HCM, let additional aligned ellipsoidal particles of the two component materials be dispersed in such a way as to maintain the overall volume fractions of a and b. This dispersal cannot change the effective properties of the HCM—because the volume-fraction-average of the polarization density dyadics of two particles, one of each component material, embedded in the HCM with relative permittivity ϵ̲̲Br1 is exactly zero. In computing the polarizability density dyadic of each particle, it must therefore be assumed that the particle is surrounded by the HCM. This fact legitimizes the use of D̲̲, which contains the anisotropic HCM’s effective constitutive properties via the scalar ρ of Eq. (7). Indeed, Eq. (4) is directly derivable from the frequency-domain Maxwell postulates specialized for the actual HCM.610

On the other hand, use of L̲̲ to compute the polarizability density dyadic of a particle implies that it is surrounded by an isotropic HCM, which is clearly incorrect. The alternative formalism that delivers ϵ̲̲Br2 is a heuristic extrapolation of the Bruggeman formalism for isotropic dielectric HCMs,12,13 and it lacks the rigorous basis that underpins the estimate of ϵ̲̲Br1. The dispersal of aligned ellipsoidal particles of the two component materials without disturbing the overall volume fractions into the HCM with relative permittivity ϵ̲̲Br2 would change the effective properties of the HCM—because the volume-fraction-average of the polarization density dyadics of two particles, one of each component material, embedded in the HCM with relative permittivity ϵ̲̲Br2 would not be zero.

We have thus delineated the differences between the two formalisms and identified one of them as the correct Bruggeman formalism. We hope that this exposition will prevent confusion between the two formalisms from perpetuating.

References

1. 

L. Ward, The Optical Constants of Bulk Materials and Films, 2nd ed.Institute of Physics, Bristol, UK (2000). Google Scholar

2. 

Selected Papers on Linear Optical Composite Materials, SPIE Optical Engineering Press, Bellingham, WA (1996). Google Scholar

3. 

T. G. Mackay, “Effective constitutive parameters of linear nanocomposites in the long-wavelength regime,” J. Nanophoton., 5 (1), 051001 (2011). http://dx.doi.org/10.1117/1.3626857 JNOACQ 1934-2608 Google Scholar

4. 

D. A. G. Bruggeman, “Berechnung verschiedener physikalischer Konstanten von heterogenen Substanzen, I. Dielektrizitätskonstanten und Leitfähigkeiten der Mischkörper aus isotropen Substanzen,” Ann. Phys. Lpz., 24 636 –679 (1935). Google Scholar

5. 

W. S. WeiglhoferA. LakhtakiaB. Michel, “Maxwell Garnett and Bruggeman formalisms for a particulate composite with bianisotropic host medium,” Microw. Opt. Technol. Lett., 15 (4), 263 –266 (1997). http://dx.doi.org/10.1002/(ISSN)1098-2760 MOTLEO 0895-2477 Google Scholar Microw. Opt. Technol. Lett., 22 221 (1999). http://dx.doi.org/10.1002/(ISSN)1098-2760 MOTLEO 0895-2477 Google Scholar

6. 

L. TsangJ. A. Kong, “Scattering of electromagnetic waves from random media with strong permittivity fluctuations,” Radio. Sci., 16 (3), 303 –320 (1981). http://dx.doi.org/10.1029/RS016i003p00303 RASCAD 0048-6604 Google Scholar

7. 

Z. D. Genchev, “Anisotropic and gyrotropic version of Polder and van Santen’s mixing formula,” Waves Random Media, 2 99 –110 (1992). http://dx.doi.org/10.1088/0959-7174/2/2/001 WRMEEV 0959-7174 Google Scholar

8. 

N. P. Zhuck, “Strong–fluctuation theory for a mean electromagnetic field in a statistically homogeneous random medium with arbitrary anisotropy of electrical and statistical properties,” Phys. Rev. B, 50 (21), 15636 –15645 (1994). http://dx.doi.org/10.1103/PhysRevB.50.15636 PRBMDO 0163-1829 Google Scholar

9. 

T. G. MackayA. LakhtakiaW. S. Weiglhofer, “Strong–property–fluctuation theory for homogenization of bianisotropic composites: formulation,” Phys. Rev. E, 62 (5), 6052 –6064 (2000). http://dx.doi.org/10.1103/PhysRevE.62.6052 PLEEE8 1063-651X Google Scholar Phys. Rev. E, 63 049901 (2001). http://dx.doi.org/10.1103/PhysRevE.63.049901 PLEEE8 1063-651X Google Scholar

10. 

T. G. MackayA. LakhtakiaW. S. Weiglhofer, “Third-order implementation and convergence of the strong-property-fluctuation theory in electromagnetic homogenisation,” Phys. Rev. E, 64 (6 Pt 2), 066616 (2001). http://dx.doi.org/10.1103/PhysRevE.64.066616 PLEEE8 1063-651X Google Scholar

11. 

T. G. MackayA. Lakhtakia, “Electromagnetic fields in linear bianisotropic mediums,” Prog. Optics, 51 121 –209 (2008). http://dx.doi.org/10.1016/S0079-6638(07)51003-6 POPTAN 0079-6638 Google Scholar

12. 

G. B. Smith, “Effective medium theory and angular dispersion of optical constants in films with oblique columnar structure,” Opt. Commun., 71 (5), 279 –284 (1989). http://dx.doi.org/10.1016/0030-4018(89)90008-4 OPCOB8 0030-4018 Google Scholar

13. 

C. G. GranqvistD. Le BellacG. A. Niklasson, “Angular selective window coatings: effective medium theory and experimental data on sputter-deposited films,” Renew. Energ., 8 (1–4), 530 –539 (1996). http://dx.doi.org/10.1016/0960-1481(96)88913-0 RNENE3 0960-1481 Google Scholar

14. 

D. SchmidtE. SchubertM. Schubert, “Optical properties of cobalt slanted columnar thin films passivated by atomic layer deposition,” Appl. Phys. Lett., 100 011912 (2012). http://dx.doi.org/10.1063/1.3675549 APPLAB 0003-6951 Google Scholar

15. 

T. Hofmannet al., “THz dielectric anisotropy of metal slanted columnar thin films,” Appl. Phys. Lett., 99 (8), 081903 (2011). http://dx.doi.org/10.1063/1.3626846 APPLAB 0003-6951 Google Scholar

16. 

D. PolderJ. H. van Santen, “The effective permeability of mixtures of solids,” Physica, 12 (5), 257 –271 (1946). http://dx.doi.org/10.1016/S0031-8914(46)80066-1 PHYSAG 0031-8914 Google Scholar

17. 

Z. HashinS. Shtrikman, “A variational approach to the theory of the effective magnetic permeability of multiphase materials,” J. Appl. Phys., 33 3125 –3131 (1962). http://dx.doi.org/10.1063/1.1728579 JAPIAU 0021-8979 Google Scholar

18. 

A. Alùet al., “Epsilon-near-zero metamaterials and electromagnetic sources: tailoring the radiation phase pattern,” Phys. Rev. B, 75 155410 (2007). http://dx.doi.org/10.1103/PhysRevB.75.155410 PRBMDO 0163-1829 Google Scholar

19. 

G. Lovatet al., “Combinations of low/high permittivity and/or permeability substrates for highly directive planar metamaterial antennas,” IET Microw. Antennas Propagat., 1 177 –183 (2007). http://dx.doi.org/10.1049/iet-map:20050353 1751-8725 Google Scholar

20. 

M. N. Navarro-Cíaet al., “Enhanced lens by ϵ and μ near-zero metamaterial boosted by extraordinary optical transmission,” Phys. Rev. B, 83 115112 (2011). http://dx.doi.org/10.1103/PhysRevB.83.115112 PRBMDO 0163-1829 Google Scholar

21. 

T. G. Mackay, “On the effective permittivity of silver–insulator nanocomposites,” J. Nanophoton., 1 019501 (2007). http://dx.doi.org/10.1117/1.2472372 JNOACQ 1934-2608 Google Scholar

22. 

B. M. RossA. Lakhtakia, “Bruggeman approach for isotropic chiral mixtures revisited,” Microw. Opt. Technol. Lett., 44 (6), 524 –527 (2005). http://dx.doi.org/10.1002/(ISSN)1098-2760 MOTLEO 0895-2477 Google Scholar
© 2012 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2012/$25.00 © 2012 SPIE
Tom G. Mackay and Akhlesh Lakhtakia "Bruggeman formalism versus "Bruggeman formalism": particulate composite materials comprising oriented ellipsoidal particles," Journal of Nanophotonics 6(1), 069501 (18 June 2012). https://doi.org/10.1117/1.JNP.6.069501
Published: 18 June 2012
Lens.org Logo
CITATIONS
Cited by 18 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Particles

Composites

Dielectrics

Homogenization

Polarizability

Polarization density

Spherical lenses

Back to Top