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

Maxwell Garnett approximation in random media: tutorial

Open Access Open Access

Abstract

In the third part of the tutorial, we develop a Maxwell Garnett approximation, which is applicable, in particular, to disordered random media with non-spherical inclusions and to multicomponent mixtures. We are especially interested in the case when the inclusions are non-spherical and randomly distributed and oriented in space, but the composite is isotropic on average. The effective medium formula applicable to such media cannot be obtained by direct generalization of the results that were derived in the first two parts of the tutorial. We show that the Maxwell Garnett effective medium approximation can be stated in a very general form. The results derived by us previously as well as a result applicable to the medium described above can be obtained as special cases of this general result.

© 2022 Optica Publishing Group

1. INTRODUCTION

The two-part tutorial on Maxwell Garnett approximation was published in JOSA A in 2016 [1,2]. Given the vast literature on the subject and the variety of applications in which Maxwell Garnett approximation continues to play an important role, the exposition in [1,2] was limited to a subset of topics that, in the author’s opinion, were conceptually important. However, one omission was particularly unfortunate: we did not cover in any detail disordered media. Although the elementary considerations presented in Section 2 of [1] apply equally to periodic and disordered composites, there is a class of problems that requires a more nuanced approach.

Namely, consider a medium consisting of a homogeneous host and many electrically small, non-spherical, randomly oriented inclusions. The dipole polarizabilities of the inclusions depend on their shapes and are, in general, tensorial. Therefore, we expect the shape of inclusions to remain important even in the low-concentration limit. Yet, since the composite is isotropic on average, its effective permittivity must be isotropic (scalar). In the preceding parts of the tutorial, we did not provide any formulas that are applicable to this case, and we will fill this gap below.

This third part of the tutorial is organized as follows. In Section 2, we discuss the homogenization formulas that were derived by us previously and explain why these results are inadequate to the problem formulated above. In Section 3, we briefly review the relevant literature. In Section 4, we summarize all assumptions that must be made to derive the general Maxwell Garnett-type approximation. Derivation of the main result is presented in Section 5. Here we obtain a formula, which allows for many special cases and encompasses the results derived in the first to parts of the tutorial as well as the case we are specifically interested in here. In Section 6, we derive various special cases of the above general result. Here we discuss in detail two-component mixtures (media consisting of only two materials) and multicomponent mixtures in which inclusions are of the same shape and randomly oriented in space. Finally, Section 7 contains a discussion.

As in the preceding two parts of the tutorial, the Gaussian system of units is used. Tensors are denoted by an overhead hat as in $\hat \alpha$. Principal values and axes of a tensor will be labeled by the variable $p$, where $p = 1,2,3$. The unit tensor is denoted by $\hat I$. Some quantities are tensorial in general but can be scalar under some conditions. For example, the effective permittivity ${\hat \epsilon _{\texttt{MG}}}$ is in general a tensor. However, if the effective medium is isotropic, we can write ${\hat \epsilon _{\texttt{MG}}} = \hat I{\epsilon _{\texttt{MG}}}$. In such cases, we say that the effective permittivity is reduced to a scalar and write ${\epsilon _{\texttt{MG}}}$ without a hat in place of ${\hat \epsilon _{\texttt{MG}}}$.

2. SURVEY OF FORMULAS

It is instructive to briefly recount the effective medium formulas that were already derived in the previous parts of this tutorial [1,2], discuss whether any of these formulas are good candidates or can be generalized easily to capture the physical properties of random composites of non-spherical inclusions, and then preview the more general result that will be derived below.

We can start the survey with the anisotropic mixing formula for ellipsoidal inclusions {Eq. (34) in [1]}. In a slightly rearranged form, this equation reads

$${({\epsilon _{{\texttt{MG}}}})_p} = {\epsilon _{\texttt h}}\left[{1 + \frac{f}{{s + (1 - f){\nu _p}}}} \right],\quad p = 1,2,3,$$
where
$$s = \frac{{ {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}$$
is the Bergman–Milton spectral parameter (we will use this notation frequently below). Here $ {\epsilon _{\texttt h}}$ and $ {\epsilon _{\texttt i}}$ are the permittivities of the host and inclusions, and ${\nu _p}$ and $f$ are the depolarization factors and the volume fraction of the inclusions. Derivation of Eq. (1) requires that all ellipsoids are similarly oriented in space. Naturally, the effective permittivity given by Eq. (1) in such a composite is tensorial. Different principal values of the permittivity tensor ${\epsilon _{{\texttt{MG}}}}$ (labeled by $p$) correspond to different directions of the electric field. This is not what we expect of a random mixture, which is isotropic on average. Therefore, Eq. (1) is not what we are looking for.

Alternatively, we can use the isotropic mixing formula {Eq. (18) in [1]}, viz.,

$${\epsilon _{{\texttt{MG}}}} = {\epsilon _{\texttt h}}\left[{1 + \frac{f}{{s + (1 - f)/3}}} \right].$$
This expression [a special case of Eq. (1) obtained by taking ${\nu _p} = 1/3$] is scalar and, thus, independent of the electric field direction. However, it is also independent of the shape or the aspect ratios of inclusions. This is not satisfactory either.

Yet another option is to start from the expression containing the dipole polarizability of the inclusions $\alpha$ explicitly. In what follows, we will frequently use the dimensionless specific polarizability $\beta$ defined as

$$\beta = \frac{{4\pi}}{{ {\epsilon _{\texttt h}}v}}\alpha ,$$
where $v$ is the inclusion volume. The relevant mixing formula {Eq. (24) of [1]} written in terms of $\beta$ reads
$${\epsilon _{{\texttt{MG}}}} = {\epsilon _{\texttt h}}\left[{1 + \frac{{f\beta}}{{1 - f\beta /3}}} \right].$$
Importantly, this equation was derived for spherical inclusions. If we take
$$\beta = \frac{1}{{s + 1/3}}\quad {\rm{(for}}\;{\rm{spherical}}\;{\rm{inclusions)}}$$
and substitute this expression into Eq. (5), the result would be identical to Eq. (3). As we have already discussed, this is inadequate for our purposes. However, Eq. (5) is, apparently, generalizable. It seems that all one needs to do is use an expression for $\beta$ that correctly describes the non-spherical inclusions. Let the shape of inclusions be fixed and known but the orientations be random. In principle, we can compute the tensorial polarizability of one inclusion, $\hat \alpha$. To describe a random mixture, we can average this result over orientations according to $\langle \hat \alpha \rangle = \frac{1}{3}\hat I {\rm{Tr}}(\hat \alpha)$ and then compute $\langle \hat \beta \rangle$ according to Eq. (4), where $\hat I$ is the unit tensor. As we expect the result to be reduced to a scalar, we can write $\langle \hat \beta \rangle = b\hat I$ and use the scalar $b$ in place of $\beta$ in Eq. (5). The resulting expression seems to satisfy our requirements: it describes an isotropic effective medium, and the quantity $b$ depends on the shape of inclusions.

However, direct generalization of Eq. (5) as described above is not the right approach. In what follows, we will show that the correct mixing formula applicable to the problem at hand is

$${\epsilon _{{\texttt{MG}}}} = {\epsilon _{\texttt h}}\left[{1 + \frac{{fb}}{{1 - f + fsb}}} \right].$$
This formula is stated as Eq. (27) below. By comparing Eq. (5) to Eq. (7), we see that $\beta$ is indeed replaced with $b$ in the numerator of the second term in the square brackets. However, the denominators differ more substantially. Moreover, Eq. (7) cannot be written in terms of $b$ alone, as the form of Eq. (5) implies; the correct expression contains the spectral parameter $s$ explicitly. Physically, one can say that direct generalization of Eq. (5) does not capture correctly the local field correction, or, more precisely, it captures the local field correction correctly only in the case of spherical inclusions. Equation (7) is free from this deficiency and applies to inclusions of arbitrary shape.

In what follows, we will show that Eq. (7) is a special case of a more general formula, which is stated in Eq. (23) below. Equations (1), (3), and (5) are also special cases of this general result. Moreover, Eq. (23) encompasses many other special cases including isotropic and anisotropic media and multicomponent mixtures as is discussed in more detail in Section 6.

3. HISTORICAL OVERVIEW

Accounting for non-sphericity of inclusions in Maxwell Garnett-type effective medium theories goes back to the work of Wiener (1912) [3]. In particular, Wiener has derived a formula equivalent to Eq. (1) for two-component mixtures of similarly oriented ellipsoids, and considered two extreme cases: $\nu = 0$ and $\nu = 1$. As can be easily verified, Eq. (1) yields the arithmetic average of $ {\epsilon _{\texttt i}}$ and $ {\epsilon _{\texttt h}}$ for $\nu = 0$ and the harmonic average for $\nu = 1$. These two extreme cases are known as Wiener bounds and play an important role in the theory of electromagnetic homogenization of composites even beyond the low concentration limit. We have illustrated the Wiener bounds in the first two parts of this tutorial [1,2] with numerical examples involving Maxwell Garnett, Bruggeman, and more rigorous homogenization theories.

However, Wiener’s treatise is lengthy, considers many different subjects, and to the best of our knowledge has not been published in an English translation. In 1953, Bragg and Pippard re-derived Wiener’s result (with a reference to the original 1912 work) to explain birefringence in hemoglobin crystals [4]. Wiener’s theory was again re-derived in 1989 and extended to the case when inclusions can contain interstitial voids filled with liquid (essentially, structured or multi-layer inclusions) by Oldenbour and Ruiz [5]. Below, we will follow the derivation these references conceptually but in a substantially more general setting.

In a further development, Stroud [6] and Stroud and Pan [7] applied the mean-field approximation to derive effective parameters of composite materials. The mean-field approach is to start from the rigorous integral equations for the electric field or the potential and derive much simpler algebraic equations by factoring certain correlators. In other words, a function inside an integral transform is replaced by an a priori unknown constant (the mean field) so that the integral can be evaluated; then the constant is found from the ensuing algebraic equations. Typically, the mean-field approximation results in Bruggeman-type symmetric mixing formulas. Some of the physical effects considered in [6,7] include coupling to a static magnetic field and radiative corrections (extending the theory beyond quasistatics; see also [810]).

References [6,7] were limited to spherical isotropic inclusions. However, the formalism of these references is quite general, and it was extended to describe an effect that is of interest to us (wherein a medium contains anisotropic inclusions but is isotropic on average) in 1997 by Levy and Stroud [11]. In this reference, the inclusions were still assumed to be spherical, but the material of inclusions could be anisotropic with arbitrary orientations of the principal axes. The theory was further generalized to include anisotropic host and ellipsoidal inclusions by Levy and Cherkaev in 2013 [12]. The results derived below largely overlap with those of Levy and Cherkaev, although there are some differences. In particular, we do not consider intrinsic anisotropy of the materials while Levy and Cherkaev did not consider multicomponent mixtures or non-ellipsoidal shapes.

Some recent applications of Maxwell Garnett theory to randomly distributed non-spherical particles include calculations of conductivity of mineral aggregates [13] and investigation of optical spectra of composite cholesteric elastomers doped with metallic ellipsoids [14].

4. ASSUMPTIONS

A. Linearity and Frequency Domain

If the electromagnetic response of the medium is linear, as we assume here, it is possible to consider different electromagnetic frequencies independently. We can start by focusing on a single working frequency $\omega$. Any time- and position-dependent, strictly monochromatic field can be written in the form ${\textbf{F}}({\textbf{r}},t) = {\rm{Re}}[{{\textbf{F}}_\omega}({\textbf{r}}){e^{- {\rm{i}}\omega t}}]$, where the factor ${e^{- {\rm{i}}\omega t}}$ is known as the phasor. Once we substitute this ansatz into Maxwell’s equations, the phasor cancels out, and we are left with equations containing only spatial derivatives and the (generally, complex) time-independent fields ${{\textbf{F}}_\omega}({\textbf{r}})$. Non-monochromatic fields can be introduced by using the superposition principle, that is, by simply adding monochromatic solutions at different frequencies together to satisfy some initial conditions for the field. Below we assume that all fields oscillate at a working frequency $\omega$ and say that we are working in the frequency domain as opposed to the time domain wherein time dependence is considered explicitly.

In what follows, we will omit the argument $\omega$ in the notations. Thus, we write $\epsilon$ instead of ${\epsilon _\omega}$ for the dielectric permittivity, ${\textbf{E}}({\textbf{r}})$ instead of ${{\textbf{E}}_\omega}({\textbf{r}})$ for the electric field, etc. We will keep in mind, however, that the permittivities of the host and inclusions can depend on $\omega$ and the same is true for the effective permittivity of the composite. Moreover, the effective permittivity can exhibit optical resonances—the phenomenon wherein a physical quantity such as absorption changes rapidly or has sharp spectral features near one or more special (resonant) frequencies [15].

We should also keep in mind that, for any given geometry of the composite, the frequency that we can consider is not arbitrary but bounded from above by the geometrical conditions, which are discussed next.

B. Geometry

The material sample of a composite medium can be of arbitrary shape and does not need to be small compared to the free-space wavelength. However, we require that the statistical properties of the composite be spatially uniform. That is, each physically small volume of the composite must have the same statistical properties as the composite as a whole.

The physically small volume is defined as follows. Consider a region inside the sample, ${\mathbb{V}}$, of sufficiently “simple” shape [16] and volume $V$ and containing $N \gg 1$ randomly distributed and oriented, relatively small, non-overlapping inclusions. We assume that ${\mathbb{V}}$ is entirely contained inside a large sample whose size and shape are not important. We denote the region occupied by $n$th inclusion and its volume by ${\Omega _n}$ and ${v_n}$, respectively, where $n = 1,2, \ldots ,N$. The permittivity of the host medium at the working frequency is denoted by $ {\epsilon _{\texttt h}}$ and the permittivity of $n$th inclusion by ${\epsilon _n}$. In the case when all inclusions have the same permittivity (i.e., are made of the same material), we use the notation $ {\epsilon _{\texttt i}}$ for the inclusion permittivity.

The volume fraction of inclusions $f$ is given by

$$f: = \frac{1}{V}\sum\limits_{n = 1}^N {v_n}.$$
We also define the average volume of one inclusion as
$$\langle v\rangle : = \frac{{fV}}{N} = \frac{1}{N}\sum\limits_{n = 1}^N {v_n}.$$
Of course, in any particular realization of a random medium, the averages defined in Eq. (8) will depend on the choice of ${\mathbb{V}}$. However, we assume that the size of ${\mathbb{V}}$ and the number $N$ are large enough for the sample means in Eq. (8) to represent accurately the averages computed over the entire sample. Moreover, we assume that other statistical properties of the distribution of inclusions in ${\mathbb{V}}$ are independent of the choice of ${\mathbb{V}}$. In other words, we assume that ${\mathbb{V}}$ is statistically representative of the entire sample.

As any standard effective medium theory, Maxwell Garnett approximation is based on the quasistatic approximation. A necessary condition for this approximation to be applicable is that the characteristic size of each inclusion $a$ be sufficiently small so that $ka \ll 1$, where $k = \omega /c$ is the free-space wavenumber. However, in the case of disordered media, this condition is not sufficient; we will also require that $kL \ll 1$, where $L \gg a$ is the characteristic size of the physically small volume ${\mathbb{V}}$. That is, we assume that the quasistatic approximation holds in ${\mathbb{V}}$.

The assumptions that ${\mathbb{V}}$ is large enough to be statistically representative of the entire sample but small enough for the quasistatic approximation to be accurate are somewhat contradictory. However, these conditions can hold simultaneously at sufficiently low frequencies. If the working frequency is too high, not only Maxwell Garnett approximation can break, but the medium may become not electromagnetically homogeneous in principle. Such media cannot be characterized by spatially uniform effective parameters.

The physically small volume and various physical scales discussed above are illustrated in Fig. 1.

 figure: Fig. 1.

Fig. 1. Schematic illustration of the geometry for which Maxwell Garnett approximation is applicable. The physically small volume occupies region ${\mathbb{V}}$ of volume $V$. Inclusions occupy regions ${\Omega _n} \in {\mathbb{V}}$, $n = 1,2, \ldots ,N$, and the volume of $n$th inclusion is ${v_n}$. The characteristic size of ${\mathbb{V}}$, $L$ is much larger than the characteristic size of one inclusion, $a$, and much smaller than the free-space wavelength, $\lambda$. Inclusions are distributed sparsely so that the volume fraction of inclusions, $f$, is much less than unity. Dielectric permittivity at the working frequency is $ {\epsilon _{\texttt h}}$ in the host and ${\epsilon _n}$ in $n$th inclusion. When all inclusions have the same permittivity, we use the notation $ {\epsilon _{\texttt i}}$ for the permittivity of inclusions.

Download Full Size | PDF

C. Fundamental Assumption of Maxwell Garnett Theory

Even if the above conditions hold with arbitrary precision, we cannot solve the electromagnetic problem in a complex medium analytically; additional assumptions are needed to make progress. We now state the fundamental assumption of the Maxwell Garnett theory, which will allow us to derive the effective permittivity of the medium analytically. Unlike the conditions discussed above, this assumption is not about the medium or the electromagnetic frequency but rather about the electric field that is induced in the medium by external sources.

Namely, we assume that inclusions, when polarized by an externally applied field, do not produce any secondary fields (in addition to the external field that is already present in the host material) outside of their interior. This approximation was already stated in the first part of this tutorial [1], but now we formulate it more generally. In [1], we considered geometries for which the field generated by a polarized inclusion integrates to zero in certain regions outside of the inclusion. For example, the field of a polarized sphere integrates to zero in any co-centric spherical shell, which excludes the region occupied by the inclusion itself. For ellipsoidal inclusions, the integration region is an ellipsoidal volume. The rationale used in [1] was that, since the field in the host material integrates to zero, and since it is produced by many inclusions that are distributed uniformly over the entire composite, the secondary field outside of the inclusions cancels out at any point in the host. However, inside the inclusions, the depolarizing field does not integrate to zero and, therefore, must be taken into account.

There is, however, a problem with the above argument. For example, consider the simplest case of electric field produced by a uniformly polarized sphere. The difficulty is that, even in the quasistatic regime, this field is not integrable. The zero integral can be obtained by selecting a special integration region (a finite spherical shell). Choosing a different region would yield a different integration result. For this reason, we have encountered an ambiguity when deriving the anisotropic Maxwell Garnett mixing formula in [1]: the result seemed to depend on the choice of integration region. Of course, once we include the far-zone field into consideration (which makes sense at non-zero frequencies), the zero integration result becomes even more questionable.

One way to deal with the above difficulty is to acknowledge that the fundamental assumption of Maxwell Garnett theory is heuristic and cannot be rigorously justified mathematically. On one hand, this point of view is not very satisfying since it provides no proof that Maxwell Garnett mixing formulas are accurate to second order in $f$. We can be sure that these formulas are accurate to first order in $f$, but the main thrust of Maxwell Garnett-type theories is to obtain corrections to the $O(f)$ result. On the other hand, once we realize what the nature of approximation is, we can apply it easily to a very general class of problems, which is indeed what we will do in the remainder of this tutorial.

Thus, the main approximation of Maxwell Garnett theory is to disregard the electric field produced by polarized inclusions inside the host material. Simply stated, we assume that the field in the host is the externally applied field. The above point is not fully appreciated in the literature. A commonly encountered point of view is that the local field corrections that arise in Maxwell Garnett theory are due to electromagnetic interaction of inclusions. However, this is not so; the local field corrections simply account for the difference between the external and the average fields in a composite [17].

We recognize that the assumption stated above can hold to various degrees of accuracy. The “best” shapes of inclusions that respect this assumption are spheres or ellipsoids or, more generally, convex particles with a high degree of symmetry such as cubes or cylinders. Non-convex inclusions or inclusions without a center of symmetry can pose a problem. For example, a non-convex inclusion with the shape of two touching or slightly overlapping spheres can concentrate the electric field in the gap, which is occupied by the host medium. This concentration of the field outside of the region of inclusions contradicts the basic assumption. For this reason, Maxwell Garnett approximation can break down if inclusions cluster or otherwise come into close contact with each other: two touching inclusions are equivalent to one strongly non-convex inclusion, and this is problematic for the theory.

Ultimately, the problem of homogenization of random mixtures with an arbitrary geometry is very complicated, and verification of the Maxwell Garnett corrections beyond the first order in $f$ is still an open question. We will, however, proceed with calculations under the assumption that the field produced by polarized inclusions in the host material is zero.

5. DERIVATION

The derivation presented below is conceptually similar to the work of Wiener [3], Bragg and Pippard [4], and Oldenbourg and Ruiz [5], but substantially more general. Most importantly, we will allow for random orientations of inclusions and, thus, capture the somewhat counterintuitive case when the medium is isotropic on average but its effective permittivity depends on the shape of non-spherical inclusions. We will also allow for general or varying shapes of inclusions and for multicomponent mixtures.

Perhaps, the simplest and most general way to derive Maxwell Garnett approximation (the “first avenue” according to Sihvola [18]) is based on the following equation:

$$\left\langle {{\textbf{D}}({\textbf{r}})} \right\rangle \equiv \left\langle {\epsilon ({\textbf{r}}){\textbf{E}}({\textbf{r}})} \right\rangle = {\hat \epsilon _{\texttt{MG}}}\left\langle {{\textbf{E}}({\textbf{r}})} \right\rangle .$$
Here $\langle ...\rangle$ denotes averaging over the physically small volume ${\mathbb{V}}$ whose properties were discussed in Section 4. Equation (10) expresses the idea that the average electric displacement $\langle {\textbf{D}}({\textbf{r}})\rangle$ is related to the average electric field $\langle {\textbf{E}}({\textbf{r}})\rangle$ by an effective permittivity tensor ${\hat \epsilon _{\texttt{MG}}}$, which is independent of the external illumination, the shape of the medium, or the choice of ${\mathbb{V}}$. A tensor with such properties can be defined only if the conditions stated in Section 4 are met, and we proceed under the assumption that this is the case.

Let us start with computing $\langle {\textbf{E}}({\textbf{r}})\rangle$. As discussed above, the fundamental assumption of Maxwell Garnett approximation is that the electric field produced by polarized inclusions is averaged out to zero in the host medium and, therefore, can be completely disregarded. It follows from this assumption that

$$\langle {\textbf{E}}({\textbf{r}})\rangle = (1 - f\,){{\textbf{E}}_{\texttt{ext}}} + \frac{1}{V}\sum\limits_{n = 1}^N \int_{{\Omega _n}} {\textbf{E}}({\textbf{r}}){{\rm{d}}^3}r,$$
where ${{\textbf{E}}_{\texttt{ext}}}$ is a spatially uniform, externally applied electric field. Equation (10) expresses straightforward averaging: the point of observation is in the host, where the field is equal to ${{\textbf{E}}_{\texttt{ext}}}$, with the probability $1 - f$ and in $n$th inclusion, where the field is different from ${{\textbf{E}}_{\texttt{ext}}}$, with the probability ${v_n}/V$. In view of Eq. (8a), the sum of all probabilities is 1. The field inside $n$th inclusion is, in general, not spatially uniform. However, only the integral quantities
$${{\textbf{e}}_n}: = \int_{{\Omega _n}} {\textbf{E}}({\textbf{r}}){{\rm{d}}^3}r$$
enter the expression for $\langle {\textbf{E}}({\textbf{r}})\rangle$. Using the definition given in Eq. (11), we can rewrite Eq. (10) as
$$\langle {\textbf{E}}({\textbf{r}})\rangle = (1 - f\,){{\textbf{E}}_{\texttt{ext}}} + \frac{1}{V}\sum\limits_{n = 1}^N {{\textbf{e}}_n}.$$
In a similar manner, we find that
$$\langle {\textbf{D}}({\textbf{r}})\rangle = (1 - f\,) {\epsilon _{\texttt h}}{{\textbf{E}}_{\texttt{ext}}} + \frac{1}{V}\sum\limits_{n = 1}^N {\epsilon _n}{{\textbf{e}}_n}.$$
To close the equations and compute ${\hat \epsilon _{\texttt{MG}}}$, we need one more ingredient: a relation between ${{\textbf{e}}_n}$ and ${{\textbf{E}}_{\texttt{ext}}}$. From the linearity of Maxwell’s equations (Laplace equation in the case of quasistatics), it follows that, for every inclusion, there is a tensor ${\hat \kappa _n}$ such that ${{\textbf{e}}_n} = {\hat \kappa _n}{{\textbf{E}}_{\texttt{ext}}}$. Naturally, ${\hat \kappa _n}$ can depend on the shape and permittivity of the inclusion but not on ${{\textbf{E}}_{\texttt{ext}}}$.

To establish a connection with more familiar quantities, we will relate ${\hat \kappa _n}$ to the dipole polarizability of $n$th inclusion ${\hat \alpha _n}$ and, further, to the dimensionless specific dipole polarizability ${\hat \beta _n}$. The polarizability ${\hat \alpha _n}$ relates the excess dipole moment [19] of $n$th inclusion, ${{\textbf{d}}_n}$, to the external field ${{\textbf{E}}_{\texttt{ext}}}$. The excess dipole moments are defined as follows. Following [1] (Section 2.D), we split the total polarization field ${\textbf{P}}({\textbf{r}})$ into two contributions as

$${\textbf{P}}({\textbf{r}}): = \frac{{\epsilon ({\textbf{r}}) - 1}}{{4\pi}}{\textbf{E}}({\textbf{r}}) \equiv {{\textbf{P}}_{\texttt h}}({\textbf{r}}) + {{\textbf{P}}_{\texttt i}}({\textbf{r}}),$$
where
$${{\textbf{P}}_{\texttt h}}({\textbf{r}}): = \frac{{ {\epsilon _{\texttt h}} - 1}}{{4\pi}}{\textbf{E}}({\textbf{r}})\;,\;\;\;{{\textbf{P}}_{\texttt i}}({\textbf{r}}): = \frac{{\epsilon ({\textbf{r}}) - {\epsilon _{\texttt h}}}}{{4\pi}}{\textbf{E}}({\textbf{r}}).$$
Clearly, ${{\textbf{P}}_i}({\textbf{r}})$ is zero in the host. The polarization field ${{\textbf{P}}_i}({\textbf{r}})$ is important in the electromagnetic theory of composites because it serves as the secondary source for the electric field. Now, by definition,
$${{\textbf{d}}_n}: = \int_{{\Omega _n}} {{\textbf{P}}_{\texttt i}}({\textbf{r}}){{\rm{d}}^3}r.$$
We say that ${{\textbf{d}}_n}$ is the excess dipole moment because Eq. (14) does not include the term ${{\textbf{P}}_h}({\textbf{r}})$ in the right-hand side. Indeed, the total dipole moment of ${\Omega _n}$ is computed using an integral similar to Eq. (14) but with the total polarization field ${\textbf{P}}({\textbf{r}})$. Conversely, if we take ${\epsilon _n} = {\epsilon _{\texttt h}}$, the corresponding excess moment ${{\textbf{d}}_n}$ is zero even though the region ${\Omega _n}$ can still be polarized by the electric field.

The conventional dipole polarizability of an inclusion in a homogeneous host is introduced as the linear coefficient between the excess dipole moment and the external field, viz.,

$${{\textbf{d}}_n} = {\hat \alpha _n}{{\textbf{E}}_{\texttt{ext}}}.$$
Since
$${{\textbf{d}}_n} = \frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{4\pi}}{{\textbf{e}}_n},$$
we also have
$${\hat \kappa _n} = \frac{{4\pi}}{{{\epsilon _n} - {\epsilon _{\texttt h}}}}{\hat \alpha _n}.$$
We can now use Eq. (17) and the relation ${{\textbf{e}}_n} = {\hat \kappa _n}{{\textbf{E}}_{\texttt{ext}}}$ to rewrite the averages [Eq. (12)] identically as follows:
$$\langle {\textbf{E}}({\textbf{r}})\rangle = {\epsilon _{\texttt h}}\! \left[{(1 - f)\hat I + \frac{{4\pi}}{{ {\epsilon _{\texttt h}}V}}\sum\limits_{n = 1}^N {s_n}{{\hat \alpha}_n}} \right]{{\textbf{E}}_{\texttt{ext}}},$$
$$\langle {\textbf{D}}({\textbf{r}})\rangle = {\epsilon _{\texttt h}} \!\left[{(1 - f)\hat I + \frac{{4\pi}}{{ {\epsilon _{\texttt h}}V}}\sum\limits_{n = 1}^N (1 + {s_n}){{\hat \alpha}_n}} \right]{{\textbf{E}}_{\texttt{ext}}},$$
where $\hat I$ is the unit tensor and
$${s_n}: = \frac{{ {\epsilon _{\texttt h}}}}{{{\epsilon _n} - {\epsilon _{\texttt h}}}}$$
is the Bergman–Milton spectral parameter for $n$th inclusion [compare to Eq. (2)]. We next introduce the dimensionless specific polarizabilities,
$${\hat \beta _n}: = \frac{{4\pi}}{{ {\epsilon _{\texttt h}}{v_n}}}{\hat \alpha _n},$$
and the averages,
$$\langle \hat \beta \rangle : = \frac{1}{N}\sum\limits_{n = 1}^N \frac{{{v_n}}}{{\langle v\rangle}}{\hat \beta _n}\;,\;\;\langle s\hat \beta \rangle : = \frac{1}{N}\sum\limits_{n = 1}^N \frac{{{v_n}}}{{\langle v\rangle}}{s_n}{\hat \beta _n}.$$
Note that the individual values of ${\hat \beta _n}$ in the above averaging are weighted by the ratios ${v_n}/\langle v\rangle$ so that larger inclusions contribute more to the average. If all inclusions are of the same shape, the weight factors are all equal to 1. As discussed above, the number $N$ of inclusions in ${\mathbb{V}}$ is assumed to be large enough for the above averages to approach the theoretical limits. With these notations, Eq. (18) takes the form
$$\left\langle {{\textbf{E}}({\textbf{r}})} \right\rangle \def\LDeqtab{}= {\epsilon _{\texttt h}} \big[{(1 - f)\hat I + f\langle s\hat \beta \rangle} \big]{{\textbf{E}}_{\texttt{ext}}},$$
$$\left\langle {{\textbf{D}}({\textbf{r}})} \right\rangle \def\LDeqtab{}= {\epsilon _{\texttt h}} \big[{(1 - f)\hat I + f\langle s\hat \beta \rangle + f\langle \hat \beta \rangle} \big]{{\textbf{E}}_{\texttt{ext}}}.$$
Since the direction of ${{\textbf{E}}_{\texttt{ext}}}$ is arbitrary, the basic definition of ${\hat \epsilon _{\texttt{MG}}}$ given in Eq. (9) and the two equations in Eq. (22) imply that
$${\hat \epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}}\! \left\{{\hat I + f\langle \hat \beta \rangle {{\left[{(1 - f)\hat I + f\langle s\hat \beta \rangle} \right]}^{- 1}}} \right\}.$$
This simple formula is the most general form of the Maxwell Garnett approximation, and it encompasses many special cases. Note that the right-hand side in Eq. (23) involves a tensor inverse and the tensors $\langle \hat \beta \rangle$ and $\langle s\hat \beta \rangle$ may not commute.

6. SPECIAL CASES

A. Two Component Media

In the case of a two-component media, all inclusions have the same permittivity $ {\epsilon _{\texttt i}}$, and all spectral parameters ${s_n}$ are equal to $s$, where $s$ is given by Eq. (2). Under the circumstances, Eq. (23) simplifies to

$${\hat \epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}} \big\{{\hat I + f\langle \hat \beta \rangle {{\big[{(1 - f)\hat I + fs\langle \hat \beta \rangle} \big]}^{- 1}}} \big\}.$$
In general, $\langle \hat \beta \rangle$ is a symmetric tensor since any sum of symmetric tensors is also symmetric. If the host and inclusions are transparent at the working frequency, then $\langle \hat \beta \rangle$ is real, symmetric, and, therefore, has three mutually orthogonal, real-valued principal axes. This is true for arbitrarily distributed and oriented inclusions. If $\langle \hat \beta \rangle$ is complex due non-negligible absorption, then it may not be diagonalizable in a set of such principal axes. However, there exists a class of geometries for which complexity of $\langle \hat \beta \rangle$ does not pose a problem. Consider absorbing inclusions with a center of symmetry. Each inclusion with such property has three orthogonal principal axes even though its polarizability tensor is complex. If the inclusions are oriented in space randomly, then $\langle \hat \beta \rangle$ is scalar (proportional to $\hat I$) and, therefore, trivially diagonalizable. Otherwise, if all inclusions are oriented in space so that their principal axes coincide, then the principal axes of $\langle \hat \beta \rangle$ coincide with those of any inclusion. This works even if the inclusions have varying shapes. The case when inclusions are oriented not similarly and not completely at random but according to some distribution is more complicated, and we do not discuss it here.

Let us assume that $\langle \hat \beta \rangle$ is diagonalizable in the above sense either because all ${\hat \beta _n}$ are real or due to some symmetry. Then we can find a set of mutually orthogonal, real-valued vectors of unit length ${{\textbf{u}}_p}$ such that

$$\langle \hat \beta \rangle {{\textbf{u}}_p} = {b_p}{{\textbf{u}}_p},\;{{\textbf{u}}_p} \cdot {{\textbf{u}}_q} = {\delta _{\textit{pq}}};\quad p,q = 1,2,3.$$
Here the principal values ${b_p}$ can still be complex. We should also keep in mind that, in general, the eigenvectors and eigenvalues of the average tensor $\langle \hat \beta \rangle$ cannot be obtained by averaging the eigenvectors and eigenvalues of the individual tensors ${\hat \beta _n}$. The most that can be said is that $\sum\nolimits_p {b_p} = \langle {\rm{Tr}}({\hat \beta _n})\rangle$. In other words, the operations of trace and averaging commute.

By using Eq. (25), we can simplify Eq. (24) as

$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}}\! \left[{\hat I + \sum\limits_{p = 1}^3 \frac{{f{b_p} {{\textbf{u}}_p} \otimes {{\textbf{u}}_p}}}{{1 - f + fs{b_p}}}} \right].$$
This is a general Maxwell Garnett approximation for a two-component medium that can be approximated as a biaxial crystal [20]. The expression can be further simplified if the medium is isotropic on average. In the latter case, we have ${b_1} = {b_2} = {b_3} = b$, and using $\sum\nolimits_{p = 1}^3 {{\textbf{u}}_p} \otimes {{\textbf{u}}_p} = \hat I$, we arrive at the mixing formula,
$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}}\! \left[{1 + \frac{{fb}}{{1 - f + fsb}}} \right].$$
Here $b$ is the average of specific polarizabilities of inclusions; that is, we can write (for isotropic composites) $\langle \hat \beta \rangle = \hat Ib$. Equation (27), also stated in the Survey of Formulas [Section 2, Eq. (7) above], solves the problem posed in Section 1. Namely, it gives the effective permittivity of a mixture of randomly placed and oriented non-spherical particles. The particles shape enters the expression via the variable $b$. Characterizing the shape of a three-dimensional particle by a single albeit a complex parameter is clearly an approximation; for one, there surely exist different shapes that correspond to the same value of $b$ (the map from the set of shapes to the complex plane is not uniquely invertible). However, this approximation is better than assuming that the shape is spherical or neglecting it altogether.

Next, we illustrate how these results work in the case of ellipsoids.

1. Similarly Oriented Ellipsoids

If all ellipsoids are similarly shaped and oriented in space, we have

$$\langle \hat \beta \rangle = \sum\limits_{p = 1}^3 \frac{{{{\textbf{u}}_p} \otimes {{\textbf{u}}_p}}}{{s + {\nu _p}}}\;,\;\;\;\;{b_p} = \frac{1}{{s + {\nu _p}}},$$
where ${\nu _p}$ represents the depolarization factors (${\nu _1} + {\nu _2} \,+ {\nu _3} = 1$). Substituting this into Eq. (26), we obtain the mixing formula,
$${\hat \epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}} \!\left[{\hat I + \sum\limits_{p = 1}^3 \frac{{f{{\textbf{u}}_p} \otimes {{\textbf{u}}_p}}}{{s + (1 - f){\nu _p}}}} \right].$$
This result is equivalent to Eq. (1).

2. Randomly Oriented Ellipsoids

If the ellipsoids are of the same shape but randomly oriented, then

$$\langle \hat \beta \rangle = \frac{1}{3}{\rm{Tr}}\sum\limits_{p = 1}^3 \frac{{{{\textbf{u}}_p} \otimes {{\textbf{u}}_p}}}{{s + {\nu _p}}}\;,\;\;b = \frac{1}{3}\sum\limits_{p = 1}^3 \frac{1}{{s + {\nu _p}}}.$$
Using this result in Eq. (27), we obtain
$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}} \!\left[{1 + \frac{{\frac{f}{3}\sum_{p = 1}^3 \frac{1}{{s + {\nu _p}}}}}{{1 - f + \frac{f}{3}\sum_{p = 1}^3 \frac{s}{{s + {\nu _p}}}}}} \right].$$
Equation (31) cannot be further simplified without making assumptions about the depolarization factors. We can, however, consider the physically important cases of spheres, thin circular disks and thin needles. Below, we use Eq. (2) for $s$ to write the mixing formulas in terms of the permittivities $ {\epsilon _{\texttt i}}$ and $ {\epsilon _{\texttt h}}$. The results are as follows:

(a) For spheres (${\nu _1} = {\nu _2} = {\nu _3} = 1/3$):

$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}}\frac{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}} + 2f( {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}})}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}} - f( {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}})}}.$$

(b) For randomly oriented disks (${\nu _1} = {\nu _2} = 0$, ${\nu _3} = 1$):

$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt i}}\frac{{3 {\epsilon _{\texttt h}} + 2f( {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}})}}{{3 {\epsilon _{\texttt i}} - f( {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}})}}.$$

(c) For randomly oriented needles (${\nu _1} = 0$, ${\nu _2} = {\nu _3} = 1/2$):

$${\epsilon _{\texttt{MG}}} = \frac{{3 {\epsilon _{\texttt h}}( {\epsilon _{\texttt h}} + {\epsilon _{\texttt i}}) + f( {\epsilon _{\texttt i}} + 3 {\epsilon _{\texttt h}})( {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}})}}{{3( {\epsilon _{\texttt h}} + {\epsilon _{\texttt i}}) - 2f( {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}})}}.$$

Equation (32a) is equivalent to Eq. (3). It is widely known and has been stated by us in various forms in [1]. However, Equations (32b) and (32c) are less known and were not derived in the first two parts of this tutorial. These results answer the question posed in Section 1 (the form of Maxwell Garnett mixing formula for randomly distributed and oriented non-spherical particles) for the extreme shapes of needles and disks. It can be seen that the result is not the same as the conventional isotropic formula [Eq. (32a)]; the three expressions in Eq. (32) differ even to first order in $f$. This conclusion is not surprising but not widely appreciated.

B. Isotropic Multicomponent Media with Inclusions of Fixed Shape

In general multicomponent media, each inclusion is characterized by shape, orientation, and permittivity. The averages in Eq. (23) must be computed with respect to the joint probability of these three variables, which can be a rather tedious procedure. However, if some of the variables are statistically independent of the others or fixed, the averaging is simplified.

Consider inclusions of a fixed shape whose permittivities can take discrete values ${\epsilon _m}$ ($m = 1,2,...,M$) with the probabilities ${P_m}$. We further assume that the inclusions are randomly distributed in space and randomly rotated with respect to a laboratory frame. Moreover, the spatial orientation and the permittivity of an inclusion are not correlated. Under these assumptions, the weight factors ${v_n}/\langle v\rangle$ in Eq. (21) are all equal to unity, and we can write

$$\langle s\hat \beta \rangle = \sum\limits_{m = 1}^M {P_m}{s_m}{\langle \hat \beta ({s_m})\rangle _{\texttt r}},$$
where the subscript “${\texttt r}$.” indicates averaging over rotations and we have indicated explicitly that $\hat \beta$ depends on $s$. For this reason, the outer average in Eq. (33) cannot be factored. However, we still have a simplification. Namely, we know that rotational averaging must yield a scalar so that
$${\langle \hat \beta ({s_m})\rangle _{\texttt r}} = b({s_m})\hat I.$$
Here $b({s_m})$ is analogous to the averaged specific polarizability $b$, which was defined around Eq. (27) above. However, there are a couple of differences. Previously, the averaging was done over shapes and rotations whereas now we assume that the shape is fixed. Also, previously we assumed that the permittivity of inclusions is fixed whereas now it can vary. The latter point is made explicit in the notation $b({s_m})$, where ${s_m}$ is a proxy for ${\epsilon _m}$; the relationship between the two variables is established by Eq. (19).

With account of Eq. (34), the general Maxwell Garnett formula [Eq. (23)] takes the form

$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}} \left[{1 + \frac{{f{{\langle b(s)\rangle}_\epsilon}}}{{1 - f + f{{\langle sb(s)\rangle}_\epsilon}}}} \right],$$
where
$${\langle b(s)\rangle _\epsilon} = \sum\limits_{m = 1}^M {P_m}b({s_m})\;,\;\;{\langle sb(s)\rangle _\epsilon} = \sum\limits_{m = 1}^M {P_m}{s_m}b({s_m}).$$
Equations (35) and (36) can be used in a numerical computation if $b({s_m})$ and ${P_m}$ are known. To make some progress analytically, we recall that any quasistatic polarizability can be written as a spectral expansion [21]. Let us focus on one particular inclusion, and let its spectral parameter be $s$. The specific polarizability of this inclusion can be written in the form
$$\hat \beta (s) = \sum\limits_k \frac{{{{\textbf{u}}_k} \otimes {{\textbf{u}}_{{k}}}}}{{s + {\nu _k}}},$$
where ${\nu _k}$ represents the generalized depolarization factors (satisfying $0 \lt {\nu _k} \lt 1$) and ${{\textbf{u}}_k}$ represents vectors with the property
$$\sum\limits_k {{\textbf{u}}_k} \otimes {{\textbf{u}}_k} = \hat I\;,\;\;\;\;\sum\limits_k {{\textbf{u}}_k} \cdot {{\textbf{u}}_k} = 3.$$
A few remarks about Eq. (37) should be made. First, summation over $k$ is infinite in the general case and can include integration over intervals of continuous spectrum (which occurs in particles with non-differentiable surface). Second, the vectors ${{\textbf{u}}_k}$ in Eq. (37a) are not required to be of unit length or mutually orthogonal. The latter two properties hold only in the special case of ellipsoids when Eq. (37a) contains three non-zero terms. Third, Eq. (37a) gives the specific polarizability for a particle of fixed shape and fixed orientation. The quantities ${{\textbf{u}}_k}$ and ${\nu _k}$ can both depend on shape, but in this subsection we assume that the shapes of all inclusions are the same. In this case, the set of ${\nu _k}$ is fixed. The only quantities in Eq. (37) that can depend on orientation are the vectors ${{\textbf{u}}_k}$, which are transformed under rotations according to the usual rules. This is, however, not important for us as we will be interested only in rotationally averaged quantities. The point here is that, although the dependence of $\hat \beta$ on orientation is not indicated explicitly in the notations of Eq. (37), we still keep it in mind.

Rotational averaging can be performed according to the general formula $b(s) = \frac{1}{3}{\rm{Tr}}\hat \beta (s)$, from which we find

$$b(s) = \sum\limits_k \frac{{{w_k}}}{{s + {\nu _k}}},\;\;\;{\rm{where}}\;\;\;{w_k} = \frac{1}{3}{{\textbf{u}}_k} \cdot {{\textbf{u}}_k}.$$
The coefficients ${w_k}$ can be interpreted as probabilities since $\sum\nolimits_k {w_k} = 1$ and ${w_k} \ge 0$. The useful property of the spectral expansion given in Eq. (38) is that ${w_k}$ and ${\nu _k}$ depend only on the inclusion shape but not on its orientation or permittivity. Therefore, these numbers can be pre-computed for any given shape and then used in the expression for $b(s)$ regardless of the value of $s$. By using Eq. (38) in Eq. (36), we obtain the following results for the averages with respect to permittivities:
$${\langle b(s)\rangle _\epsilon} = \sum\limits_{\textit{mk}} \frac{{{P_m}{w_k}}}{{{s_m} + {\nu _k}}}\;,\;{\langle sb(s)\rangle _\epsilon} = \sum\limits_{\textit{mk}} \frac{{{P_m}{s_m}{w_k}}}{{{s_m} + {\nu _k}}}.$$
Substituting the result of Eq. (39) into Eq. (35) yields the general mixing formula for a multicomponent mixture of similar-shaped randomly oriented inclusions. Further progress can be achieved either numerically for general shapes of inclusions or analytically in the special case of ellipsoids, which is considered next.

1. Randomly Oriented Ellipsoids

All ellipsoids have three non-zero terms in the summation over $k$ with ${w_1} = {w_2} = {w_3} = 1/3$, but the depolarization factors ${\nu _1},{\nu _2}$ and ${\nu _3}$ depend on the aspect ratios. As in the previous subsection, we limit consideration to the three special cases of spheres, thin circular disks and thin needles.

(a) Spheres (${\nu _1} = {\nu _2} = {\nu _3} = 1/3$):

$${\langle b(s)\rangle _\epsilon} = 3\sum\limits_m {P_m}\frac{{{\epsilon _m} - {\epsilon _{\texttt h}}}}{{{\epsilon _m} + 2 {\epsilon _{\texttt h}}}} = 3\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}}}}} \right\rangle ,$$
$${\langle sb(s)\rangle _\epsilon} = 1 - \sum\limits_m {P_m}\frac{{{\epsilon _m} - {\epsilon _{\texttt h}}}}{{{\epsilon _m} + 2 {\epsilon _{\texttt h}}}} = 1 - \left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}}}}} \right\rangle ,$$
where we have used Eq. (2) for ${s_m}$ and
$$\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}}}}} \right\rangle = \sum\limits_m {P_m}\frac{{{\epsilon _m} - {\epsilon _{\texttt h}}}}{{{\epsilon _m} + 2 {\epsilon _{\texttt h}}}}.$$
Substitution of Eq. (40a) into Eq. (35) yields the mixing formula
$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}} \left[{1 + \frac{{3f\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}}}}} \right\rangle}}{{1 - f\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}}}}} \right\rangle}}} \right].$$
It is easy to show that the above equation is equivalent to
$$\frac{{{\epsilon _{\texttt{MG}}} - {\epsilon _{\texttt h}}}}{{{\epsilon _{\texttt{MG}}} + 2 {\epsilon _{\texttt h}}}} = f\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + 2 {\epsilon _{\texttt h}}}}} \right\rangle = \sum\limits_m {f_m}\frac{{{\epsilon _m} - {\epsilon _{\texttt h}}}}{{{\epsilon _m} + 2 {\epsilon _{\texttt h}}}},$$
where ${f_m} = f{P_m}$ is the volume fraction of $m$th species of inclusions. This formula was stated in [1] as a simple generalization of the two-component mixing formula [Eq. (32a)]. Here we have derived it from more fundamental principles.

(b) Randomly oriented disks (${\nu _1} = {\nu _2} = 0$, ${\nu _3} = 1$):

$${\langle b(s)\rangle _\epsilon} \def\LDeqtab{}= \frac{1}{3}\left[{2\left\langle {\frac{{ {\epsilon _{\texttt i}}}}{{ {\epsilon _{\texttt h}}}}} \right\rangle - \left\langle {\frac{{ {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}}}}} \right\rangle - 1} \right],$$
$${\langle sb(s)\rangle _\epsilon} \def\LDeqtab{}= \frac{1}{3}\left[{2 + \left\langle {\frac{{ {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}}}}} \right\rangle} \right],$$
where
$$\left\langle {\frac{{ {\epsilon _{\texttt i}}}}{{ {\epsilon _{\texttt h}}}}} \right\rangle = \sum\limits_m {P_m}\frac{{{\epsilon _m}}}{{ {\epsilon _{\texttt h}}}}\;,\;\;\left\langle {\frac{{ {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}}}}} \right\rangle = \sum\limits_m {P_m}\frac{{ {\epsilon _{\texttt h}}}}{{{\epsilon _m}}}.$$
Upon substitution of Eq. (43) into Eq. (35), we find the mixing formula
$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}}\frac{{1 + \frac{{2f}}{3}\left[{\left\langle {\frac{{ {\epsilon _{\texttt i}}}}{{ {\epsilon _{\texttt h}}}}} \right\rangle - 1} \right]}}{{1 + \frac{f}{3}\left[{\left\langle {\frac{{ {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}}}}} \right\rangle - 1} \right]}}.$$
In the case when there is only one species of inclusions and the symbols of averaging can be omitted, Eq. (45) becomes equivalent to Eq. (32b).

(c) Randomly oriented needles (${\nu _1} = 0$, ${\nu _2} = {\nu _3} = 1/2$):

$${\langle b(s)\rangle _\epsilon} \def\LDeqtab{}= \frac{1}{3}\left[{\left\langle {\frac{{ {\epsilon _{\texttt i}}}}{{ {\epsilon _{\texttt h}}}}} \right\rangle + 4\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + {\epsilon _{\texttt h}}}}} \right\rangle - 1} \right],$$
$${\langle sb(s)\rangle _\epsilon} \def\LDeqtab{}= 1 - \frac{2}{3}\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + {\epsilon _{\texttt h}}}}} \right\rangle ,$$
where the averages are defined similarly to Eqs. (41) and (44). Upon substitution of Eq. (46) into Eq. (35), we find the mixing formula
$${\epsilon _{\texttt{MG}}} = {\epsilon _{\texttt h}}\frac{{1 + \frac{f}{3}\left[{\left\langle {\frac{{ {\epsilon _{\texttt i}}}}{{ {\epsilon _{\texttt h}}}}} \right\rangle + 2\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + {\epsilon _{\texttt h}}}}} \right\rangle - 1} \right]}}{{1 - \frac{{2f}}{3}\left\langle {\frac{{ {\epsilon _{\texttt i}} - {\epsilon _{\texttt h}}}}{{ {\epsilon _{\texttt i}} + {\epsilon _{\texttt h}}}}} \right\rangle}}.$$
As in the case of randomly oriented disks, Eq. (47) is reduced to Eq. (32c) if there is only one species of inclusions and averaging over permittivities can be omitted.

7. DISCUSSION

Random complex media are very complicated from the theoretical standpoint, and detailed analytical or numerical description of such objects beyond various mean-field approximations is a formidable task. In the three-part tutorial, we discussed the physical setting in which electromagnetic properties of a composite can be approximated by a homogeneous effective medium. The Maxwell Garnett approximation is perhaps the most widely used approach to deriving the effective permittivity of such media. The approximation is, however, applicable only in the limit of low volume fraction of inclusions, $f$. Whereas we can almost always be sure that the Maxwell Garnett theory is accurate to first order in $f$ (assuming the physical conditions described in Section 4 were met), the accuracy of higher-order corrections cannot be easily ascertained. Still, we can hope that, in many cases, especially with simple-shape and non-metallic inclusions, the Maxwell Garnett approximation is accurate at least to second order in $f$.

If the Maxwell Garnett approximation fails, the effective medium description can still be applicable, but one needs to use other, more general approaches to deriving the effective parameters. Generically, the problem is known as electromagnetic homogenization. In the second part of this tutorial [2], we have shown how effective medium parameters can be computed in periodic two-component media with arbitrary volume fractions. In principle, one can extend this approach to random media by generating a quasi-random cell (a computer model of the physically small volume) and replicating it periodically in all directions. The resultant medium would still be periodic with a quasi-random cell, and the methods of [2] can be applied to such cases. However, in any realistic setting, this formulation is intractable due to high computational complexity.

Of course, one should be conscious that complex media exhibit many physical phenomena that cannot be captured, even in principle, by an effective medium theory. Such phenomena include strong multiple scattering, speckles, and localization of light. In the physical regimes wherein these phenomena can be observed, Maxwell Garnett theory is definitely inapplicable, and drastically different theoretical approaches must be used. Therefore, before using the Maxwell Garnett theory, one should decide whether it is applicable. Material in Section 4 was designed to help with this determination, but it is by no means exhaustive. Physically, the question is always whether multiple scattering plays a significant role in a composite at the working frequency. If it does, Maxwell Garnett theory is inapplicable.

The mechanisms by which multiple scattering leads to breakdown of Maxwell Garnett approximation have been extensively studied in the literature. One observation is that multiple scattering in random composites results in strong fluctuations of the electric field on the scale of one inter-inclusion distance [15]. Maxwell Garnett, as well as other mean-field approximations that rely on factorization of field correlators (such as Bruggeman’s approximation or the theories of Stroud and Pan [6,7]), cannot capture such phenomena. As soon as the fast-fluctuating component of the field becomes comparable to the mean field, the respective approximation loses accuracy. Another way to look at the role of multiple scattering is the following. Maxwell Garnett mixing formulas contain the volume fraction of inclusions but not their specific locations. On the other hand, multiple scattering clearly depends on higher-order density correlation functions in a composite. One can view $f = \langle \theta ({\textbf{r}})\rangle$ as the first-order correlation function, where $\theta ({\textbf{r}})$ is unity inside an inclusion and zero in the host. But second-order scattering depends on the two-point correlation function ${\rho _2}({\textbf{R}}) = \langle \theta ({\textbf{r}})\theta ({\textbf{r}} + {\textbf{R}})\rangle$. There can be many composites with the same $f$ but different ${\rho _2}({\textbf{R}})$. According to Maxwell Garnett approximation, all such media have the same effective properties. But in practice the multiply scattered light will be reflected and transmitted differently by samples with same $f$ but different ${\rho _2}({\textbf{R}})$. Therefore, Maxwell Garnet approximation cannot be accurate if second-order scattering is non-negligible. As the strength of scattering increases, higher-order correlators also come to the fore, and complex resonant phenomena occur. The connection between the higher-order density correlation functions, multiple scattering, and breakdown of Maxwell Garnett approximation was pointed out and investigated in [22].

We, thus, conclude that second-order scattering between inclusions is not captured correctly by Maxwell Garnett approximation. In this respect, it would be useful to clear one common misconception. Namely, it is often stated that the local field correction [essentially, the denominator or the inverse in the main result Eq. (23) above] appears due to the electromagnetic interaction of inclusions. For example, this point of view was expressed in [12]. Although this misconception does not affect any results and can, therefore, be seen as a matter of interpretation, we think that it is useful to recognize that there is no interaction of inclusions in Maxwell Garnett theory at all [17]. The local field to which the inclusions are subjected is the same as the external (applied) field, and the latter would not be changed if we remove the inclusions. The only regions of space that are affected by the inclusions (according to the approximation) are those that the inclusions occupy. It is the contribution of these regions to the average electric field and displacement that is important in Maxwell Garnett theory.

Disclosures

The author declares no conflicts of interest.

Data availability

No data were generated or analyzed in the presented tutorial.

REFERENCES AND NOTES

1. V. A. Markel, “Introduction to the Maxwell Garnett approximation: tutorial,” J. Opt. Soc. Am. A 33, 1244–1256 (2016). [CrossRef]  

2. V. A. Markel, “Maxwell Garnett approximation (advanced topics): tutorial,” J. Opt. Soc. Am. A 33, 2237–2255 (2016). [CrossRef]  

3. O. Wiener, Die Theorie des Mischkörpers für das Feld der Stationären Strömung. Erste Abhandlung: Die Mittelwertsätze für Kraft, Polarisation und Energie, Vol. 32 of Abhandlungen der Mathematisch-Physischen Klasse der Königl. Sächsischen Gesellschaft der Wissenschaften (2012), pp. 507–604.

4. W. L. Bragg and A. B. Pippard, “The form birefringence of macromolecules,” Acta Crystallogr. 6, 865–867 (1953). [CrossRef]  

5. R. Oldenbourg and T. Ruiz, “Birefringence of macromolecules. Wiener’s theory revisited, with applications to DNA and tobacco mosaic virus,” Biophys. J. 56, 195–205 (1989). [CrossRef]  

6. D. Stroud, “Generalized effective-medium approach to the conductivity of an inhomogeneous material,” Phys. Rev. B 12, 3368–3373 (1975). [CrossRef]  

7. D. Stroud and F. P. Pan, “Self-consistent approach to electromagnetic wave propagation in composite media: application to model granular metals,” Phys. Rev. B 17, 1602–1610 (1978). [CrossRef]  

8. W. T. Doyle, “Optical properties of a suspension of metal spheres,” Phys. Rev. B 39, 9852–9858 (1989). [CrossRef]  

9. R. Ruppin, “Evaluation of extended Maxwell-Garnett theories,” Opt. Commun. 182, 273–279 (2000). [CrossRef]  

10. C. F. Bohren, “Do extended effective-medium formulas scale properly?” J. Nanophoton. 3, 039501 (2009). [CrossRef]  

11. O. Levy and D. Stroud, “Maxwell Garnett theory for mixtures of anisotropic inclusions: application to conducting polymers,” Phys. Rev. B 56, 8035–8046 (1997). [CrossRef]  

12. O. Levy and E. Cherkaev, “Effective medium approximations for anisotropic composites with arbitrary component orientation,” J. Appl. Phys. 114, 164102 (2013). [CrossRef]  

13. K. Han and S. M. Clark, “Review of calculating the electrical conductivity of mineral aggregates from constituent conductivities,” Solid Earth Sci. 6, 111–128 (2021). [CrossRef]  

14. G. Reyes and J. A. Reyes, “Optical spectra of composite cholesteric elastomers doped with metallic nano-ellipsoids,” Liq. Cryst. 49, 1–16 (2021). [CrossRef]  

15. M. I. Stockman, K. B. Kurlaev, and T. F. George, “Linear and nonlinear optical susceptibilities of Maxwell Garnett composites: Dipolar spectral theory,” Phys. Rev. B 60, 17071–17083 (1999). [CrossRef]  

16. Without going into excessive mathematical formalism, we expect ${\mathbb{V}}$ to be simply connected, convex, and approximately of unit aspect ratio like a cube or a sphere.

17. The statement that Maxwell Garnett approximation ignores interaction of inclusions might be confusing as this seems to imply that the averaged field in any finite sample of the composite medium inserted in a constant external electric field remains constant. There are obvious counter-examples to this conclusion. We note, however, that the interaction is neglected for the purpose of computing the Maxwell Garnett permittivity, not for solving any boundary value problems in the effective medium (these tasks are separate).

18. A. Sihvola, “Two main avenues leading to the Maxwell Garnett mixing rule,” J. Electromagn. Waves Appl. 15, 715–725 (2001). [CrossRef]  

19. We introduce the excess dipole moments ${{\textbf{d}}_n}$ and the dipole polarizabilities ${\hat \alpha _n}$ only for convenience. The derivation does not rely on literal interpretation of the quantities ${{\textbf{d}}_n}$ as dipole moments.

20. Anisotropic media with three different principal values of the dielectric tensor are conventionally referred to as biaxial. Sometimes the definition of biaxial media includes the assumption that the principal values are real and can be ordered. We do not make this assumption here.

21. V. A. Markel, V. N. Pustovit, S. V. Karpov, A. V. Obuschenko, V. S. Gerasimov, and I. L. Isaev, “Electromagnetic density of states and absorption of radiation by aggregates of nanospheres with multipole interactions,” Phys. Rev. B 70, 054202 (2004). [CrossRef]  

22. P. Mallet, C. Guerin, and A. Sentenac, “Maxwell-Garnett mixing rule in the presence of multiple scattering: derivation and accuracy,” Phys. Rev. B 72, 014205 (2005). [CrossRef]  

Data availability

No data were generated or analyzed in the presented tutorial.

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 (1)

Fig. 1.
Fig. 1. Schematic illustration of the geometry for which Maxwell Garnett approximation is applicable. The physically small volume occupies region ${\mathbb{V}}$ of volume $V$ . Inclusions occupy regions ${\Omega _n} \in {\mathbb{V}}$ , $n = 1,2, \ldots ,N$ , and the volume of $n$ th inclusion is ${v_n}$ . The characteristic size of ${\mathbb{V}}$ , $L$ is much larger than the characteristic size of one inclusion, $a$ , and much smaller than the free-space wavelength, $\lambda$ . Inclusions are distributed sparsely so that the volume fraction of inclusions, $f$ , is much less than unity. Dielectric permittivity at the working frequency is $ {\epsilon _{\texttt h}}$ in the host and ${\epsilon _n}$ in $n$ th inclusion. When all inclusions have the same permittivity, we use the notation $ {\epsilon _{\texttt i}}$ for the permittivity of inclusions.

Equations (59)

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

( ϵ MG ) p = ϵ h [ 1 + f s + ( 1 f ) ν p ] , p = 1 , 2 , 3 ,
s = ϵ h ϵ i ϵ h
ϵ MG = ϵ h [ 1 + f s + ( 1 f ) / 3 ] .
β = 4 π ϵ h v α ,
ϵ MG = ϵ h [ 1 + f β 1 f β / 3 ] .
β = 1 s + 1 / 3 ( f o r s p h e r i c a l i n c l u s i o n s )
ϵ MG = ϵ h [ 1 + f b 1 f + f s b ] .
f := 1 V n = 1 N v n .
v := f V N = 1 N n = 1 N v n .
D ( r ) ϵ ( r ) E ( r ) = ϵ ^ MG E ( r ) .
E ( r ) = ( 1 f ) E ext + 1 V n = 1 N Ω n E ( r ) d 3 r ,
e n := Ω n E ( r ) d 3 r
E ( r ) = ( 1 f ) E ext + 1 V n = 1 N e n .
D ( r ) = ( 1 f ) ϵ h E ext + 1 V n = 1 N ϵ n e n .
P ( r ) := ϵ ( r ) 1 4 π E ( r ) P h ( r ) + P i ( r ) ,
P h ( r ) := ϵ h 1 4 π E ( r ) , P i ( r ) := ϵ ( r ) ϵ h 4 π E ( r ) .
d n := Ω n P i ( r ) d 3 r .
d n = α ^ n E ext .
d n = ϵ i ϵ h 4 π e n ,
κ ^ n = 4 π ϵ n ϵ h α ^ n .
E ( r ) = ϵ h [ ( 1 f ) I ^ + 4 π ϵ h V n = 1 N s n α ^ n ] E ext ,
D ( r ) = ϵ h [ ( 1 f ) I ^ + 4 π ϵ h V n = 1 N ( 1 + s n ) α ^ n ] E ext ,
s n := ϵ h ϵ n ϵ h
β ^ n := 4 π ϵ h v n α ^ n ,
β ^ := 1 N n = 1 N v n v β ^ n , s β ^ := 1 N n = 1 N v n v s n β ^ n .
E ( r ) = ϵ h [ ( 1 f ) I ^ + f s β ^ ] E ext ,
D ( r ) = ϵ h [ ( 1 f ) I ^ + f s β ^ + f β ^ ] E ext .
ϵ ^ MG = ϵ h { I ^ + f β ^ [ ( 1 f ) I ^ + f s β ^ ] 1 } .
ϵ ^ MG = ϵ h { I ^ + f β ^ [ ( 1 f ) I ^ + f s β ^ ] 1 } .
β ^ u p = b p u p , u p u q = δ pq ; p , q = 1 , 2 , 3.
ϵ MG = ϵ h [ I ^ + p = 1 3 f b p u p u p 1 f + f s b p ] .
ϵ MG = ϵ h [ 1 + f b 1 f + f s b ] .
β ^ = p = 1 3 u p u p s + ν p , b p = 1 s + ν p ,
ϵ ^ MG = ϵ h [ I ^ + p = 1 3 f u p u p s + ( 1 f ) ν p ] .
β ^ = 1 3 T r p = 1 3 u p u p s + ν p , b = 1 3 p = 1 3 1 s + ν p .
ϵ MG = ϵ h [ 1 + f 3 p = 1 3 1 s + ν p 1 f + f 3 p = 1 3 s s + ν p ] .
ϵ MG = ϵ h ϵ i + 2 ϵ h + 2 f ( ϵ i ϵ h ) ϵ i + 2 ϵ h f ( ϵ i ϵ h ) .
ϵ MG = ϵ i 3 ϵ h + 2 f ( ϵ i ϵ h ) 3 ϵ i f ( ϵ i ϵ h ) .
ϵ MG = 3 ϵ h ( ϵ h + ϵ i ) + f ( ϵ i + 3 ϵ h ) ( ϵ i ϵ h ) 3 ( ϵ h + ϵ i ) 2 f ( ϵ i ϵ h ) .
s β ^ = m = 1 M P m s m β ^ ( s m ) r ,
β ^ ( s m ) r = b ( s m ) I ^ .
ϵ MG = ϵ h [ 1 + f b ( s ) ϵ 1 f + f s b ( s ) ϵ ] ,
b ( s ) ϵ = m = 1 M P m b ( s m ) , s b ( s ) ϵ = m = 1 M P m s m b ( s m ) .
β ^ ( s ) = k u k u k s + ν k ,
k u k u k = I ^ , k u k u k = 3.
b ( s ) = k w k s + ν k , w h e r e w k = 1 3 u k u k .
b ( s ) ϵ = mk P m w k s m + ν k , s b ( s ) ϵ = mk P m s m w k s m + ν k .
b ( s ) ϵ = 3 m P m ϵ m ϵ h ϵ m + 2 ϵ h = 3 ϵ i ϵ h ϵ i + 2 ϵ h ,
s b ( s ) ϵ = 1 m P m ϵ m ϵ h ϵ m + 2 ϵ h = 1 ϵ i ϵ h ϵ i + 2 ϵ h ,
ϵ i ϵ h ϵ i + 2 ϵ h = m P m ϵ m ϵ h ϵ m + 2 ϵ h .
ϵ MG = ϵ h [ 1 + 3 f ϵ i ϵ h ϵ i + 2 ϵ h 1 f ϵ i ϵ h ϵ i + 2 ϵ h ] .
ϵ MG ϵ h ϵ MG + 2 ϵ h = f ϵ i ϵ h ϵ i + 2 ϵ h = m f m ϵ m ϵ h ϵ m + 2 ϵ h ,
b ( s ) ϵ = 1 3 [ 2 ϵ i ϵ h ϵ h ϵ i 1 ] ,
s b ( s ) ϵ = 1 3 [ 2 + ϵ h ϵ i ] ,
ϵ i ϵ h = m P m ϵ m ϵ h , ϵ h ϵ i = m P m ϵ h ϵ m .
ϵ MG = ϵ h 1 + 2 f 3 [ ϵ i ϵ h 1 ] 1 + f 3 [ ϵ h ϵ i 1 ] .
b ( s ) ϵ = 1 3 [ ϵ i ϵ h + 4 ϵ i ϵ h ϵ i + ϵ h 1 ] ,
s b ( s ) ϵ = 1 2 3 ϵ i ϵ h ϵ i + ϵ h ,
ϵ MG = ϵ h 1 + f 3 [ ϵ i ϵ h + 2 ϵ i ϵ h ϵ i + ϵ h 1 ] 1 2 f 3 ϵ i ϵ h ϵ i + ϵ h .
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.