This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An extremal problem arising in the dynamics of two-phase materials that directly reveals information about the internal geometry

Ornella Mattei1, Graeme W. Milton2, and Mihai Putinar3
(1 Department of Mathematics, San Francisco State University, CA 94132, USA,
2Department of Mathematics, University of Utah, Salt Lake City, UT 84112, USA,
3Department of Mathematics, University of California at Santa Barbara, CA 93106, USA, and School of Mathematics, Statistics and Physics, Newcastle University, NE1 7RU Newcastle upon Tyne, UK.
Emails: [email protected], [email protected], [email protected], [email protected]
)
Abstract

In two phase materials, each phase having a non-local response in time, it has been found that for some driving fields the response somehow untangles at specific times, and allows one to directly infer useful information about the geometry of the material, such as the volume fractions of the phases. Motivated by this, and to obtain an algorithm for designing appropriate driving fields, we find approximate, measure independent, linear relations between the values that Markov functions take at a given set of possibly complex points, not belonging to the interval [-1,1] where the measure is supported. The problem is reduced to simply one of polynomial approximation of a given function on the interval [-1,1] and to simplify the analysis Chebyshev approximation is used. This allows one to obtain explicit estimates of the error of the approximation, in terms of the number of points and the minimum distance of the points to the interval [-1,1]. Assuming this minimum distance is bounded below by a number greater than 1/2, the error converges exponentially to zero as the number of points is increased. Approximate linear relations are also obtained that incorporate a set of moments of the measure. In the context of the motivating problem, the analysis also yields bounds on the response at any particular time for any driving field, and allows one to estimate the response at a given frequency using an appropriately designed driving field that effectively is turned on only for a fixed interval of time. The approximation extends directly to Markov-type functions with a positive semidefinite operator valued measure, and this has applications to determining the shape of an inclusion in a body from boundary flux measurements at a specific time, when the time-dependent boundary potentials are suitably tailored.

Keywords: Composites, best rational approximation, volume fraction estimation, bounds on transient response, Calderon problem, Markov functions

1 Introduction

Many systems have responses that are nonlocal in time as this is naturally a consequence of the fact that it takes time for subelements of the system to respond. Usually, this leads one to either: (1) examine the response at each, or one or more, frequencies as the convolution in time characterizing the response of the system becomes a simple multiplication in the frequency domain; or (2) examine the response to a delta function or Heaviside function as this directly reveals the integral kernel characterizing the response. In this paper we show that desired information about the system can be directly obtained from selectively designed input signals that are neither at constant frequency, nor delta or Heaviside functions.

Our initial motivation comes from the work [31, 33] where we derived microstructure-independent bounds on the viscoelastic response at a given time tt of two-phase periodic composites (in antiplane shear) with prescribed volume fractions f1f_{1} and f2=1f1f_{2}=1-f_{1} of the phases and with an applied average stress or strain prescribed as a function of time. We found that the bounds were sometimes extremely tight at particular times t=t0t=t_{0}: see Figure 1. This was quite a surprise because the response of each phase is nonlocal in time, yet somehow this response is untangled at these particular times. Thus, the bounds could be used in an inverse fashion to determine the volume fractions from measurements at time t0t_{0}. Determining volume fractions of phases is important in the oil industry, where one wants to know the proportions occupied by oil and water in the rock, to detecting breast cancer, to assessing the porosity of sea-ice and other materials, and even to determining the volume of holes in swiss cheese. While our bounds were very tight at specific times in some examples, they were far from tight at all times in other examples: see Figure 2.

At that time it was totally unclear as to whether appropriate input signals could produce the desired tight bounds at a specific time and, if so, what algorithm should be followed to design these input signals. The primary goal of this paper is to address this problem in the case where the input function has a finite number of frequencies. In particular, for the example of Figure 2, our algorithm produces the much tighter bounds of Figure 3. It is an open question as to whether one can find smooth input signals, containing a continuum of frequencies, such that the response Re[v(t0)]\mathop{\rm Re}\nolimits[v(t_{0})] of the material at a specific moment of time t0t_{0} is totally measure independent, while Re[v(t)]\mathop{\rm Re}\nolimits[v(t)] has a smooth dependence on tt, with Re[v(t)]0\mathop{\rm Re}\nolimits[v(t)]\to 0 when tt\to-\infty. The example of Figure 1 suggests that it may be possible.

We emphasize that our results are applicable not just to determining the volume fractions of the phases in a two-phase composite but also determining the volume and shape of an inclusion in a body from exterior boundary measurements. This is shown in Sections 10 and 11. It is a classical and important inverse problem with a long history and many contributions: see [10, 26, 27, 47, 46, 22, 1, 19, 9, 11, 4, 3, 25, 44, 28] and references therein.

A secondary goal of this paper is to solve an accompanying mathematical approximation problem which we now outline, and which is essential to achieve the primary goal.

Refer to caption
Figure 1: Comparison between the lower and upper bounds on the output average stress with an input applied average strain of H(t)H(t), where H(t)H(t) is the Heaviside function, 0 for t<0t<0 and 11 for t0t\geq 0. This is called a stress relaxation test. One phase is purely elastic (G=6000G=6000), while the other phase is viscoelastic and modeled by the Maxwell model (G=12000G=12000 and η=20000\eta=20000) (the results are normalized by the response of the elastic phase). The following three cases are graphed: no information about the composite is given; the volume fraction of the components is known (f1=0.4f_{1}=0.4); and the composite is isotropic with given volume fractions. The bounds become tighter and tighter as more information on the composite structure is included, so that if color is missing from the figure the outermost pair of bounds are those with no information, the middle pair include just the volume fraction, and the innermost pair include both volume fraction and isotropy. Reproduced from Figure 6.2 in [33].
Refer to caption
Figure 2: Comparison between the lower and upper bounds on the output stress relaxation in the “badly-ordered case”, when the responses on the pure phases as a function of time do not cross with an input applied average strain of H(t)H(t), where H(t)H(t) is the Heaviside function. Here the purely elastic phase has shear modulus G=12000G=12000, while the Maxwell parameters for the viscoelastic phase are G=6000G=6000 and η=20000\eta=20000 (again, the results are normalized by the response of the elastic phase). The three subcases are the same as for the previous figure. However the bounds remain quite wide except near t=0t=0. Reproduced from Figure 6.5 in [33]. The approach developed in this paper can yield tight bounds with a suitably designed input function as shown in Figure 3.
Refer to caption
Figure 3: Comparison between the lower and upper bounds on the output stress relaxation in the “badly-ordered case” (G=12000G=12000 for the elastic phase, and G=6000G=6000 and η=20,000\eta=20,000 for the viscoelastic Maxwell phase), when the input function is chosen accordingly to equation (4.3), which represents the main result of this paper. Specifically, equation (4.3) provides the amplitude of the applied field that gives extremely tight bounds at a chosen moment of time (here t=0t=0) when the volume fraction is known. Indeed, the bounds incorporating the volume fraction (the innermost bounds, in red) take the value 0.40.4 at t=0t=0, which coincides exactly with the volume fraction of the viscoelastic phase. Here, the applied loading is the sum of three time-harmonic fields with frequencies ω=0.1,0.5,1.5\omega=0.1,0.5,1.5.

2 A fundamental mathematical question of approximation applicable to Markov functions

Here we formulate a problem, not just relevant to systems with a nonlocal response, but potentially to other application where Markov functions arise. Such functions are Cauchy transforms of positive measures with compact support on the real line. Some authors may suitably call them Herglotz functions, or Nevanlinna functions, or Stieltjes transforms.

Suppose Fμ(z)F_{\mu}(z) is a Markov function having the integral representation

Fμ(z)=11dμ(λ)λz,F_{\mu}(z)=\int_{-1}^{1}\frac{d\mu(\lambda)}{\lambda-z}, (2.1)

where the Borel measure μ\mu is positive with unit mass:

11𝑑μ(λ)=1.\int_{-1}^{1}d\mu(\lambda)=1. (2.2)

Given mm (possibly complex) points z1,z2,,zmz_{1},z_{2},\ldots,z_{m} not belonging to the interval [1,1][-1,1], we are interested in finding complex constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} such that

k=1mαkFμ(zk)1\sum_{k=1}^{m}\alpha_{k}F_{\mu}(z_{k})\approx 1 (2.3)

for all probability measures μ\mu. Optimal bounds correlating the possible values of the mm-tuple (Fμ(z1),Fμ(z2),,Fμ(zm))(F_{\mu}(z_{1}),F_{\mu}(z_{2}),\ldots,F_{\mu}(z_{m})) as μ\mu varies over all probability measures are well known, as derived from the well charted analysis of the Nevanlinna-Pick interpolation problem [29]. Indeed, the nonlinear constraints among the values Fμ(z1),Fμ(z2),,Fμ(zm)F_{\mu}(z_{1}),F_{\mu}(z_{2}),\ldots,F_{\mu}(z_{m}), and standard convexity theory provide optimal bounds on the range of the left hand side of (2.3) for given constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m}, see [29] for details. But this is not our main concern.

We would rather like to choose mm points z1,z2,,zmz_{1},z_{2},\ldots,z_{m}, and find associated constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m}, for every prescribed integer mm, having the property

supμ|k=1mαkFμ(zk)1|ϵm\sup_{\mu}\left|\sum_{k=1}^{m}\alpha_{k}F_{\mu}(z_{k})-1\right|\leq\epsilon_{m} (2.4)

for some bound ϵm\epsilon_{m} subject to ϵm0\epsilon_{m}\to 0 as mm\to\infty. The geometry of the locus of these points is obviously essential and it will be detailed in the sequel. The faster the convergence, the better.

Since we deal with probability measures, condition (2.4) is equivalent to

supλ[1,1]|k=1mαkλzk1|ϵm.\sup_{\lambda\in[-1,1]}\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-1\right|\leq\epsilon_{m}. (2.5)

And this is good news. We seek the minimal deviation from zero on the interval [1,1][-1,1] of a rational function R(z)R(z) satisfying R()=1R(\infty)=1 and possessing simple poles at the points z1,,zmz_{1},\ldots,z_{m}. Or equivalently, denoting q(z)=(zz1)(zz2)(zzm)q(z)=(z-z_{1})(z-z_{2})\ldots(z-z_{m}) and w(λ)=|q(λ)|1w(\lambda)=|q(\lambda)|^{-1} we aim at finding the minimal deviation from zero of a monic polynomial pp of degree mm, with respect to the weighted norm pw=supλ[1,1]|p(λ)w(λ)|.\|p\ w\|_{\infty}=\sup_{\lambda\in[-1,1]}|p(\lambda)w(\lambda)|.

Both perspectives align to well-known classical studies in approximation theory. The first one is an extremal problem in rational approximation with prescribed poles, a subject going back at least to Walsh [51]. A great deal of information in this respect was systematized in Walsh’s book [52]. The second approach is a genuine weighted Chebyshev approximation problem, and here we are on solid ground. First, note that the functions

w(λ),λw(λ),λ2w(λ),w(\lambda),\lambda w(\lambda),\lambda^{2}w(\lambda),\ldots (2.6)

form a Chebyshev system on the interval [1,1][-1,1], that is, they are linearly independent and any linear combination of w(λ),λw(λ),λ2w(λ),,λmw(λ)w(\lambda),\lambda w(\lambda),\lambda^{2}w(\lambda),\ldots,\lambda^{m}w(\lambda) has at most mm zeros in [1,1][-1,1]. Even more, a stronger so-called Markov property of this system of functions holds. The classical Chebyshev approximation in the uniform norm theorem has an analog for such non-orthogonal bases [23, 29]. To be more precise, there exists a unique monic polynomial pp of degree mm minimizing the norm pw\|p\ w\|_{\infty}: this polynomial is characterized by the fact that |p(λ)||p(\lambda)| attains its maximal value at m+1m+1 points, and the sign of p(λ)p(\lambda) alternates there, see also [34]. In case w(λ)=1w(\lambda)=1, the optimal polynomial is of course the normalized Chebyshev polynomial of the first kind: p(λ)=Tm(λ)2m,Tm(cosx)=cos(mx),m0.p(\lambda)=\frac{T_{m}(\lambda)}{2^{m}},\ T_{m}(\cos x)=\cos(mx),\,m\geq 0. The constructive aspects of weighted Chebyshev approximation are rather involved, see for instance the early works of Werner [54, 55, 56]. In the same vein, the asymptotics of the optimal bound of our minimization problem inherently involves potential theory or operator theory concepts. We cite for a comparison basis a few remarkable results of the same flavor [16, 45, 6].

Without seeking sharp bounds and guided by the specific applications we aim at, we propose a compromise and relaxation of our extremal problem:

infppwwinfpp.\inf_{p}\|p\ w\|_{\infty}\leq\|w\|_{\infty}\inf_{p}\|p\|_{\infty}. (2.7)

At this point we can invoke Chebyshev original theorem and his polynomial TmT_{m}, obtaining in this way the benefit, very useful for applications, of computing in closed form the residues αk\alpha_{k}. Details and some ramifications will be given in the Section 4.

3 Relevance of the approximation problem to systems with a non-local time response and the viscoelasticity problem in particular.

Without going into the specific details, as these will be provided later in Section 7, 8 and 9, in many linear systems with an input function u(t)u(t) varying with time tt, of the form

u(t)=k=1mβkeiωk(tt0),u(t)=\sum_{k=1}^{m}\beta_{k}e^{-i\omega_{k}(t-t_{0})}, (3.1)

where the ωk\omega_{k} are a set of (possibly complex) frequencies, and t0t_{0} is a given time, the output function v(t)v(t) takes the form

v(t)=k=1mαka0Fμ(z(ωk))eiωk(tt0),v(t)=\sum_{k=1}^{m}\alpha_{k}a_{0}F_{\mu}(z(\omega_{k}))e^{-i\omega_{k}(t-t_{0})}, (3.2)

in which the function Fμ(z)F_{\mu}(z) is given by (2.1),

αk=βkc(ωk),\alpha_{k}=\beta_{k}c(\omega_{k}), (3.3)

and the functions z(ω)z(\omega) and c(ω)c(\omega) depend on ω\omega in some known way: z=z(ω)z=z(\omega) and c=c(ω)c=c(\omega). The real constant a0>0a_{0}>0 and the unknown measure dμd\mu depend on the system. In our viscoelasticity study [33] the connection with Markov functions comes from the fact that the effective shear modulus G(ω)G_{*}(\omega), that relates the average stress to the average strain at frequency ω\omega, as a function of the shear moduli G1(ω)G_{1}(\omega) and G2(ω)G_{2}(\omega) of the two phases, has the property that [(G/G1)1]/(2f1)[(G_{*}/G_{1})-1]/(2f_{1}), in which f1f_{1} is the volume fraction of phase 1, is a Markov function of z=(G1+G2)/(G2G1)z=(G_{1}+G_{2})/(G_{2}-G_{1}) taking the form (2.2) [7, 35, 15].

Henceforth, we adopt the notational simplification

f(z)=Fμ(z).f(z)=F_{\mu}(z).

Thus, at time t=t0t=t_{0}, the output function is

v(t0)=a0k=1mαkf(zk)withzk=z(ωk),v(t_{0})=a_{0}\sum_{k=1}^{m}\alpha_{k}f(z_{k})\quad\text{with}\quad z_{k}=z(\omega_{k}), (3.4)

and we seek an input signal so that the output v(t0)v(t_{0}) is almost system independent with v(t0)a0v(t_{0})\approx a_{0}. So, by measuring v(t0)v(t_{0}) we can determine the system parameter a0a_{0}. In the viscoelastic problem that we studied [31, 33], a0a_{0} is the volume fraction f1f_{1} (see also [7]) and it is useful to be able to determine this from indirect measurements. Typically, one may assume the frequencies ωk\omega_{k} have a positive imaginary part so that the input signal u(t)u(t) is essentially zero in the distant past. In (3.1) one could just take a signal with m1m-1 frequencies ωk\omega_{k}, k=1,2,,m1k=1,2,\ldots,m-1. Then, we have

v(t0)+αma0f(z(ωm))=a0k=1mαkf(zk)a0withzk=z(ωk).v(t_{0})+\alpha_{m}a_{0}f(z(\omega_{m}))=a_{0}\sum_{k=1}^{m}\alpha_{k}f(z_{k})\approx a_{0}\quad\text{with}\quad z_{k}=z(\omega_{k}). (3.5)

Then, if a0a_{0} is known, a measurement of v(t0)v(t_{0}) will allow us to estimate the output a0f(z(ωm))eiωm(tt0)a_{0}f(z(\omega_{m}))e^{-i\omega_{m}(t-t_{0})} at a desired (possibly real) frequency ωm\omega_{m} given the time harmonic input eiωm(tt0)e^{-i\omega_{m}(t-t_{0})}.

It is often the situation, such as in the viscoelastic problem, that only the real part of v(t)v(t) has a direct physical significance and, hence, one might want to find constants αk\alpha_{k} such that, say,

2Re[v(t0)]\displaystyle 2\mathop{\rm Re}\nolimits[v(t_{0})] =a0(k=1mαkf(zk)+k=1mαk¯f(zk)¯)\displaystyle=a_{0}\left(\sum_{k=1}^{m}\alpha_{k}f(z_{k})+\sum_{k=1}^{m}\overline{\alpha_{k}}\overline{f(z_{k})}\right)
=a0(k=1mαkf(zk)+k=1mαk¯f(zk¯))a0,\displaystyle=a_{0}\left(\sum_{k=1}^{m}\alpha_{k}f(z_{k})+\sum_{k=1}^{m}\overline{\alpha_{k}}f(\overline{z_{k}})\right)\approx a_{0}, (3.6)

where the overline denotes complex conjugation. This, again, reduces to a problem of the form (2.3) where, after renumbering, the complex values of zkz_{k} come in pairs, zkz_{k} and zk+1=zk¯z_{k+1}=\overline{z_{k}} and we may take αk+1=αk¯\alpha_{k+1}=\overline{\alpha_{k}} so that the left hand side of (2.3) is real.

It may be the case that the first nn moments of the probability measure dμd\mu are known,

M=11λ𝑑μ(λ),=1,2,,n,M_{\ell}=\int_{-1}^{1}\lambda^{\ell}d\mu(\lambda),\quad\ell=1,2,\ldots,n, (3.7)

in addition to the zeroth moment M0=1M_{0}=1 and that mm (possibly complex) points z1,z2,,zmz_{1},z_{2},\ldots,z_{m} not on the interval [1,1][-1,1] are given. We then may seek complex constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} and γ0,γ1,γ2,γn\gamma_{0},\gamma_{1},\gamma_{2},\ldots\gamma_{n}, with say γn=1\gamma_{n}=1, such that

k=1mαkf(zk)=0nγM\sum_{k=1}^{m}\alpha_{k}f(z_{k})\approx\sum_{\ell=0}^{n}\gamma_{\ell}M_{\ell} (3.8)

for all probability measures μ\mu with the prescribed moments. We will treat this problem in Section 3. We can use these results to determine an approximate linear relation among the nn moments if v(t0)v(t_{0}) is measured. This may be used to estimate one moment if the rest are known. This can be useful when the moments have an important physical significance: in the viscoelastic problem, for instance, M1M_{1} depends only on the volume fraction f1f_{1} if one assumes that the composite has sufficient symmetry to ensure that its response remains invariant as the material is rotated [7]. So, incorporating the moment M1M_{1} and measuring the response at time t0t_{0}, then, allows us to obtain tighter bounds on f1f_{1} as shown in [31, 33].

Mutatis mutandis, we may seek complex constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} and γ1,γ2,γn\gamma_{1},\gamma_{2},\ldots\gamma_{n} (each constant depending both on mm and nn) such that

inf𝐀k=1mαk[𝐀zk𝐈]1=0nγ𝐀ϵm(n),\inf_{\bf A}\left\|\sum_{k=1}^{m}\alpha_{k}[{\bf A}-z_{k}{\bf I}]^{-1}-\sum_{\ell=0}^{n}\gamma_{\ell}{\bf A}^{\ell}\right\|\leq\epsilon_{m}^{(n)}, (3.9)

where the infimum is over all self adjoint operators 𝐀{\bf A} with spectrum in [1,1][-1,1] and for fixed nn, ϵm(n)0\epsilon_{m}^{(n)}\to 0 as mm\to\infty. Our analysis extends easily to treat this problem too. The relevance is that in many linear systems with an input field 𝐮(t){\bf u}(t) varying with time tt, of the form

𝐮(t)=k=1mβkeiωk(tt0)𝐮0,{\bf u}(t)=\sum_{k=1}^{m}\beta_{k}e^{-i\omega_{k}(t-t_{0})}{\bf u}_{0}, (3.10)

the output field 𝐯(t){\bf v}(t) takes the form

𝐯(t)=k=1mαkeiωk(tt0)a0[𝐀z(ωk)𝐈]1𝐮0withαk=βkc(ωk),{\bf v}(t)=\sum_{k=1}^{m}\alpha_{k}e^{-i\omega_{k}(t-t_{0})}a_{0}[{\bf A}-z(\omega_{k}){\bf I}]^{-1}{\bf u}_{0}\quad\text{with}\quad\alpha_{k}=\beta_{k}c(\omega_{k}), (3.11)

where the real constant a0a_{0} and the self-adjoint operator 𝐀{\bf A} characterize the response of the system, and the system parameters z(ω)z(\omega) and c(ω)c(\omega) depend on the frequency ω\omega in some known way. Then, the bound (3.9) implies

|𝐯(t0)a0=0nγ𝐀𝐮0|a0ϵm(n)|𝐮0|.\left|{\bf v}(t_{0})-a_{0}\sum_{\ell=0}^{n}\gamma_{\ell}{\bf A}^{\ell}{\bf u}_{0}\right|\leq a_{0}\epsilon_{m}^{(n)}\left|{\bf u}_{0}\right|. (3.12)

4 The solution to the main approximation problem

The present section contains the main result which provides the theoretical foundation of our explorations. As explained in the introduction, we try to balance the computational accessibility and simplicity with the loss of sharp bounds. A few comments about the versatility of the following theorem are elaborated after its proof.

Theorem 1
Letting

d(zk)=minλ[1,1]|λzk|d(z_{k})=\min_{\lambda\in[-1,1]}|\lambda-z_{k}| (4.1)

denote the distance from zkz_{k} to the line segment [1,1][-1,1], and assuming

dmin=minkd(zk)>1/2,d_{\rm min}=\min_{k}d(z_{k})>1/2, (4.2)

one can find complex constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} each depending on mm, such that (2.4) holds with ϵm0\epsilon_{m}\to 0 as mm\to\infty. In particular, with

αk=Tm(zk)2m1jk(zkzj),\alpha_{k}=-\frac{T_{m}(z_{k})}{2^{m-1}\prod_{j\neq k}(z_{k}-z_{j})}, (4.3)

where Tm(z)T_{m}(z) is the Chebyshev polynomial of the first kind, of degree mm, (2.4) holds with ϵm=2/(2dmin)m\epsilon_{m}=2/(2d_{\rm min})^{m} which tends to zero as mm\to\infty.

Proof

We have

|k=1mαkf(zk)1|11𝑑μ(λ)|k=1mαkλzk1|.\left|\sum_{k=1}^{m}\alpha_{k}f(z_{k})-1\right|\leq\int_{-1}^{1}d\mu(\lambda)\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-1\right|. (4.4)

Since the right hand side of (4.4) is linear in dμd\mu its maximum over all probability measures is achieved when μ(λ)\mu(\lambda) is an extremal measure, namely the point mass

μ(λ)=δ(λλ0),\mu(\lambda)=\delta(\lambda-\lambda_{0}), (4.5)

where λ0\lambda_{0} is varied in [1,1][-1,1] so as to get the maximum value of the right hand side of (4.4). In fact, for this extreme measure one has equality in (4.4). Equivalently, we have

infαsupμ|k=1mαkf(zk)1|\displaystyle\inf_{\alpha}\sup_{\mu}\left|\sum_{k=1}^{m}\alpha_{k}f(z_{k})-1\right| =\displaystyle= infαsupμ11𝑑μ(λ)|k=1mαkλzk1|\displaystyle\inf_{\alpha}\sup_{\mu}\int_{-1}^{1}d\mu(\lambda)\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-1\right| (4.6)
=\displaystyle= infαsupλ0[1,1]|k=1mαkλ0zk1|.\displaystyle\inf_{\alpha}\sup_{\lambda_{0}\in[-1,1]}\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda_{0}-z_{k}}-1\right|.

Thus, we seek a set of constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} (each dependent on mm) and sequence ϵm\epsilon_{m} such that ϵm0\epsilon_{m}\to 0 as mm\to\infty and

|k=1mαkλzk1|ϵm for all λ[1,1].\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-1\right|\leq\epsilon_{m}\text{ for all }\lambda\in[-1,1]. (4.7)

More clearly, direct substitution of (4.7) into (4.4) shows that (2.4) holds.

Now we may write

k=1mαkλzk=p(λ)q(λ)=R(λ),\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}=\frac{p(\lambda)}{q(\lambda)}=R(\lambda), (4.8)

where q(λ)q(\lambda) is the known monic polynomial

q(λ)=j=1m(λzj)q(\lambda)=\prod_{j=1}^{m}(\lambda-z_{j}) (4.9)

of degree mm, and p(λ)p(\lambda) is a polynomial of degree at most m1m-1 that remains to be determined. The constants αk\alpha_{k} can then be identified with the residues at the poles λ=zk\lambda=z_{k} of R(λ)R(\lambda):

αk=p(zk)jk(zkzj).\alpha_{k}=\frac{p(z_{k})}{\prod_{j\neq k}(z_{k}-z_{j})}. (4.10)

Then, the problem becomes one of choosing p(λ)p(\lambda) such that

supλ[1,1]|p(λ)q(λ)1|=supλ[1,1]|p(λ)q(λ)||q(λ)|\sup_{\lambda\in[-1,1]}\left|\frac{p(\lambda)}{q(\lambda)}-1\right|=\sup_{\lambda\in[-1,1]}\frac{\left|p(\lambda)-q(\lambda)\right|}{|q(\lambda)|} (4.11)

is close to zero. Clearly, the problem is now one of polynomial approximation of the monic polynomial q(λ)q(\lambda) of degree mm by the polynomial p(λ)p(\lambda). A natural choice is to take

p(λ)=q(λ)Tm(λ)/2m1,p(\lambda)=q(\lambda)-T_{m}(\lambda)/2^{m-1}, (4.12)

where Tm(λ)/2m1T_{m}(\lambda)/2^{m-1} is the Chebyshev polynomial Tm(λ)T_{m}(\lambda) of degree mm, normalized to be monic. This choice minimizes the sup-norm of |p(λ)q(λ)||p(\lambda)-q(\lambda)| over the interval λ[1,1]\lambda\in[-1,1] and

|p(λ)q(λ)|=|Tm(λ)/2m1|1/2m1|p(\lambda)-q(\lambda)|=|T_{m}(\lambda)/2^{m-1}|\leq 1/2^{m-1} (4.13)

provides a bound on the numerator in (4.11). To bound the denominator, we have

|q(λ)|=k=1m|λzk|k=1md(zk),|q(\lambda)|=\prod_{k=1}^{m}|\lambda-z_{k}|\geq\prod_{k=1}^{m}d(z_{k}), (4.14)

where d(zk)d(z_{k}) is given by (4.1). Using (4.2) and the bounds (4.13) and (4.14) we see that (2.4) is satisfied with ϵm=2/(2dmin)m\epsilon^{m}=2/(2d_{\rm min})^{m}. Finally, with p(λ)p(\lambda) given by (4.12) we see that the residues αk\alpha_{k} at the poles λ=zk\lambda=z_{k} of g(λ)g(\lambda), given by (4.10) correspond to those given by (4.3).

Remark 1

The use of Chebyshev polynomials is convenient as bounds on their sup-norm over the interval [1,1][-1,1] are readily available. An alternative approach, also accessible from the numerical/computational point of view, is to work with the L2L^{2} norm and find the polynomial p(λ)p(\lambda) of degree m1m-1 that approximates the given monic polynomial q(λ)q(\lambda) of degree mm in the precise sense that

11|(p(λ)q(λ)|2dν(λ),withdν(λ)=dλ/|q(λ)|2\int_{-1}^{1}|(p(\lambda)-q(\lambda)|^{2}d\nu(\lambda),\quad\text{with}\quad d\nu(\lambda)=d\lambda/|q(\lambda)|^{2} (4.15)

is minimized. Subsequently, one has to invoke Bernstein-Markov’s inequality which bounds an L2L^{2} norm by uniform norm. This first step is a standard problem in the theory of orthogonal polynomials: one chooses p(λ)q(λ)p(\lambda)-q(\lambda) to be the monic polynomial of degree mm that is orthogonal to all polynomials of degree at most m1m-1 with respect to the measure dν(λ)d\nu(\lambda). Separating the contribution of the denominator, by selecting ν\nu to be the measure dλ/1λ2d\lambda/\sqrt{1-\lambda^{2}} we recover the Chebyshev polynomials we have advocated in the proof of the main result.

Remark 2

Assumption (4.2) is more than we really need. To underline the dependence on nn of all data we set

qn(z)=(zz1(n))(zz2(n))(zzn(n)),n1,q_{n}(z)=(z-z_{1}(n))(z-z_{2}(n))\ldots(z-z_{n}(n)),\ \ n\geq 1,

and

wn(z)=1|qn(z)|.w_{n}(z)=\frac{1}{|q_{n}(z)|}.

For the proof of Theorem 1 above we only need

lim supnwn1/n<2.\limsup_{n}\|w_{n}\|_{\infty}^{1/n}<2.

That is, there exists r<2r<2, so that for large nn, the inequality

wn(λ)rn,λ[1,1],w_{n}(\lambda)\leq r^{n},\ \ \lambda\in[-1,1],

holds true.

By taking the natural logarithm, we are led to enforce the condition

lim supnsupλ[1,1]1nj=1nln1|λzj(n)|<ln2.\limsup_{n}\sup_{\lambda\in[-1,1]}\frac{1}{n}\sum_{j=1}^{n}\ln\frac{1}{|\lambda-z_{j}(n)|}<\ln 2.

In other terms, an evenly distributed probability mass on the points z1(n),,zn(n)z_{1}(n),\ldots,z_{n}(n) should have its logarithmic potential asymptotically bounded from above by a prescribed constant, on the interval [1,1][-1,1]. Again, this turns out to be a rather typical problem of approximation theory, at least when restricting the poles of qnq_{n} to belong to some Jordan curve surrounding [1,1][-1,1]. A natural choice being an ellipse with foci at ±1\pm 1, see also [45, 6].

Remark 3

We can gain more flexibility in the choice of the input signal if we replace Tm(zk)T_{m}(z_{k}) in the formula (4.3) for the residues αk\alpha_{k} with (zkz0)Tm1(zk)(z_{k}-z_{0})T_{m-1}(z_{k}), where z0z_{0} is a prescribed (possibly complex) zero of p(λ)q(λ)=(λz0)Tm1(λ)p(\lambda)-q(\lambda)=(\lambda-z_{0})T_{m-1}(\lambda). In particular, we may choose z0z_{0} to, say, minimize

maxtt0|v(t)|/|v(t0)|maxtt0|[k=1mαkf(zk)eiωk(tt0)]|\max_{t\leq t_{0}}|v(t)|/|v(t_{0})|\approx\max_{t\leq t_{0}}\left|\left[\sum_{k=1}^{m}\alpha_{k}f(z_{k})e^{-i\omega_{k}(t-t_{0})}\right]\right| (4.16)

to help ensure that the output signal is not too wild. If we are only interested in Re[v(t)]\mathop{\rm Re}\nolimits[v(t)] so that the zkz_{k} come in complex conjugate pairs, then we may replace Tm(zk)T_{m}(z_{k}) in (4.3) with (zkz0)(zkz0¯)Tm2(zk)(z_{k}-z_{0})(z_{k}-\overline{z_{0}})T_{m-2}(z_{k}), and choose z0z_{0} to, say, minimize

maxtt0|Re[v(t)]|/|Re[v(t0)]|maxtt0|Re[k=1mαkf(zk)eiωk(tt0)]|.\max_{t\leq t_{0}}|\mathop{\rm Re}\nolimits[v(t)]|/|\mathop{\rm Re}\nolimits[v(t_{0})]|\approx\max_{t\leq t_{0}}\left|\mathop{\rm Re}\nolimits\left[\sum_{k=1}^{m}\alpha_{k}f(z_{k})e^{-i\omega_{k}(t-t_{0})}\right]\right|. (4.17)

In the first case, note that the signal u(t)u(t) (3.1) is linear in z0z_{0} while in the second case it is linear in the real coefficients of the quadratic (λz0)(λz0¯)(\lambda-z_{0})(\lambda-\overline{z_{0}}). So in either case we have a linear space of possible signals (though |z0||z_{0}| should not be too large for the approximation to hold at time t0t_{0}). Also αk0\alpha_{k}\to 0 as z0zkz_{0}\to z_{k} so in this limit the frequency ωk\omega_{k} is absent from the input and output signals. More generally, to help minimize (4.16) or (4.17) one might replace Tm(zk)T_{m}(z_{k}) with sM(zk)TmM(zk)s_{M}(z_{k})T_{m-M}(z_{k}) where sM(λ)s_{M}(\lambda) is a polynomial of fixed degree M<mM<m.

5 Incorporating moments of the measure

Here we assume that the first nn moments M1,M2,,MnM_{1},M_{2},\ldots,M_{n} of the probability measure dμd\mu, given by (2.1), are known, in addition to M0=1M_{0}=1 and that mm (possibly complex) points z1,z2,,zmz_{1},z_{2},\ldots,z_{m} not on the interval [1,1][-1,1] are given. We seek complex constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} and γ1,γ2,γn\gamma_{1},\gamma_{2},\ldots\gamma_{n}, with say γn=1\gamma_{n}=1 such that

|k=1mαkf(zk)=0nγM|\left|\sum_{k=1}^{m}\alpha_{k}f(z_{k})-\sum_{\ell=0}^{n}\gamma_{\ell}M_{\ell}\right| (5.1)

is small for all probability measures μ\mu with the prescribed nn moments. The analysis proceeds as before, only now we introduce the polynomial

r(λ)==0nγλ,r(\lambda)=\sum_{\ell=0}^{n}\gamma_{\ell}\lambda^{\ell}, (5.2)

and set p(λ)p(\lambda) and q(λ)q(\lambda) to be the polynomials defined by (4.8) and (4.9). The goal is now to choose polynomials p(λ)p(\lambda) and r(λ)r(\lambda) of degrees m1m-1 and nn, respectively, such that r(λ)r(\lambda) is monic and

supλ[1,1]|p(λ)q(λ)r(λ)|=supλ[1,1]|p(λ)q(λ)r(λ)||q(λ)|\sup_{\lambda\in[-1,1]}\left|\frac{p(\lambda)}{q(\lambda)}-r(\lambda)\right|=\sup_{\lambda\in[-1,1]}\frac{\left|p(\lambda)-q(\lambda)r(\lambda)\right|}{|q(\lambda)|} (5.3)

is close to zero. We choose p(λ)p(\lambda) and r(λ)r(\lambda) such that

Tm+n(λ)/2m+n1=q(λ)r(λ)p(λ).T_{m+n}(\lambda)/2^{m+n-1}=q(\lambda)r(\lambda)-p(\lambda). (5.4)

This is simply the Euclidean division of the normalized Chebyshev polynomial Tm+n(λ)/2m+n1T_{m+n}(\lambda)/2^{m+n-1} by q(λ)q(\lambda) with r(λ)r(\lambda) being identified as the quotient polynomial and p(λ)-p(\lambda) being identified as the remainder polynomial. Then, assuming (4.2) and using (4.14), we have

supλ[1,1]|p(λ)q(λ)r(λ)|ϵm(n),withϵm(n)=22n(2dmin)m\sup_{\lambda\in[-1,1]}\left|\frac{p(\lambda)}{q(\lambda)}-r(\lambda)\right|\leq\epsilon_{m}^{(n)},\quad\text{with}\quad\epsilon_{m}^{(n)}=\frac{2}{2^{n}(2d_{\rm min})^{m}} (5.5)

satisfying ϵm(n)0\epsilon_{m}^{(n)}\to 0 as mm\to\infty, with nn being fixed. With constants αk\alpha_{k} given by (4.10) and constants γ\gamma_{\ell} being the coefficients of the polynomial r(λ)r(\lambda), as in (5.2), it follows that

supμ|k=1mαkf(zk)=0nγM|\displaystyle\sup_{\mu}\left|\sum_{k=1}^{m}\alpha_{k}f(z_{k})-\sum_{\ell=0}^{n}\gamma_{\ell}M_{\ell}\right| (5.6)
=supμ11𝑑μ(λ)|k=1mαkλzk=0nγλ|ϵm(n).\displaystyle\quad=\sup_{\mu}\int_{-1}^{1}d\mu(\lambda)\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-\sum_{\ell=0}^{n}\gamma_{\ell}\lambda^{\ell}\right|\leq\epsilon_{m}^{(n)}.

Remark

The results exposed in the previous sections extend immediately to the resolvent case using the spectral representation

𝐀=σ(𝐀)λ𝑑𝐏λ,{\bf A}=\int_{\sigma({\bf A})}\lambda d{\bf P}_{\lambda}, (5.7)

where σ(𝐀)\sigma({\bf A}) is the spectrum of 𝐀{\bf A}, assumed to be contained in the interval [1,1][-1,1] and d𝐏λd{\bf P}_{\lambda} is an orthogonal projection valued measure satisfying

𝐈=σ(𝐀)𝑑𝐏λ.{\bf I}=\int_{\sigma({\bf A})}d{\bf P}_{\lambda}. (5.8)

Then, we have

inf𝐀k=1mαk[𝐀zk𝐈]1=0nγ𝐀\displaystyle\inf_{\bf A}\left\|\sum_{k=1}^{m}\alpha_{k}[{\bf A}-z_{k}{\bf I}]^{-1}-\sum_{\ell=0}^{n}\gamma_{\ell}{\bf A}^{\ell}\right\| (5.9)
=supd𝐏λσ(𝐀)[k=1mαkλzk=0nγλ]𝑑𝐏λ\displaystyle\quad\quad=\sup_{d{\bf P}_{\lambda}}\left\|\int_{\sigma({\bf A})}\left[\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-\sum_{\ell=0}^{n}\gamma_{\ell}\lambda^{\ell}\right]d{\bf P}_{\lambda}\right\|
supd𝐏λσ(𝐀)|k=1mαkλzk=0nγλ|d𝐏λ.\displaystyle\quad\quad\leq\sup_{d{\bf P}_{\lambda}}\left\|\int_{\sigma({\bf A})}\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-\sum_{\ell=0}^{n}\gamma_{\ell}\lambda^{\ell}\right|d{\bf P}_{\lambda}\right\|.

Choosing constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} and γ1,γ2,γn\gamma_{1},\gamma_{2},\ldots\gamma_{n}, with γn=1\gamma_{n}=1, as in the previous section, the bound (5.5) substituted in (5.9) implies that the desired bound (3.9) holds with ϵm(n)=2/[2n(2dmin)m]\epsilon_{m}^{(n)}=2/[2^{n}(2d_{\rm min})^{m}].

6 Bounds on the output function v(t)v(t) at any time tt

Supposing any constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} are given, it is easy to get bounds on v(t)v(t) given by (3.2) at any time tt that incorporate the nn known moments M1,M2,,MnM_{1},M_{2},\ldots,M_{n}. One introduces an angle θ\theta and Lagrange multipliers γ1,γ2,,γn\gamma_{1},\gamma_{2},\ldots,\gamma_{n} and takes the minimum value of

11𝑑μ(λ)Re[eiθk=1mαkeiωk(tt0)λz(ωk)+=1nγλ],\int_{-1}^{1}d\mu(\lambda)\mathop{\rm Re}\nolimits\left[e^{i\theta}\sum_{k=1}^{m}\frac{\alpha_{k}e^{-i\omega_{k}(t-t_{0})}}{\lambda-z(\omega_{k})}+\sum_{\ell=1}^{n}\gamma_{\ell}\lambda^{\ell}\right], (6.1)

as μ\mu varies over all probability measures supported on [1,1][-1,1] with unconstrained moments. The minimum will be achieved by the point masses μ=δ(λλ0)\mu=\delta(\lambda-\lambda_{0}), where λ0\lambda_{0} may take one or more values. Typically we will need to choose the Lagrange multipliers γ1,γ2,,γn\gamma_{1},\gamma_{2},\ldots,\gamma_{n} (that depend on θ\theta) so that the minimum is achieved at nn values λ0=λ0()\lambda_{0}=\lambda_{0}^{(\ell)}, =1,2,,n\ell=1,2,\ldots,n, and then adjust the measure to be distributed at these points

dμ(λ)==1nwδ(λλ0()),d\mu(\lambda)=\sum_{\ell=1}^{n}\mathrm{w}_{\ell}\delta(\lambda-\lambda_{0}^{(\ell)}), (6.2)

with the non-negative weights w\mathrm{w}_{\ell}, that sum to 11, chosen so that the moments take their desired values. Then with this measure we obtain the bound

Re[eiθv(t)]a0=1nwRe[eiθk=1mαkeiωk(tt0)λ0()z(ωk)].\mathop{\rm Re}\nolimits[e^{i\theta}v(t)]\geq a_{0}\sum_{\ell=1}^{n}\mathrm{w}_{\ell}\mathop{\rm Re}\nolimits\left[e^{i\theta}\sum_{k=1}^{m}\frac{\alpha_{k}e^{-i\omega_{k}(t-t_{0})}}{\lambda_{0}^{(\ell)}-z(\omega_{k})}\right]. (6.3)

By varying θ\theta from 0 to 2π2\pi we obtain bounds that confine v(t)v(t) to a convex region in the complex plane. Of course, if we are only interested in bounding Re[v(t)]\mathop{\rm Re}\nolimits[v(t)], then it suffices to take θ=0\theta=0 or π\pi.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: (a) Bounds on the real part of the response of the system, Re[v(t)]\mathrm{Re}[v(t)] (6.3), when the system is such that z(ω)=2+i/ωz(\omega)=2+i/\omega and the input signal Re[u(t)]\mathrm{Re}[u(t)] is the one depicted in (b) (with c(ω)c(\omega)=1). We choose the frequencies ωk\omega_{k} to be [1+1i;0.5+0.3i;2+0.5i][1+1i;0.5+0.3i;2+0.5i], and we select the coefficients αk\alpha_{k} according to (4.3) so that the bounds are extremely tight at t0=0t_{0}=0, whereas the point masses λ0()\lambda_{0}^{(\ell)} and the weights w\mathrm{w}_{\ell} are chosen for each moment of time tt such that the minimum value of (6.1) is attained while the moments of the measure take their desired values. Specifically, the bounds on Re[v(t)]\mathrm{Re}[v(t)] are plotted for three different scenarios, as shown by the legend.
Refer to caption
(a)
Refer to caption
(b)
Figure 5: (a) Bounds on the real part of the response of the system, Re[v(t)]\mathrm{Re}[v(t)] (6.3), when the system is such that z(ω)=22/ω2z(\omega)=2-2/\omega^{2} and the input signal Re[u(t)]\mathrm{Re}[u(t)] is the one depicted in (b) (with c(ω)c(\omega)=1). We choose the frequencies ωk\omega_{k} to be [1+1i;0.5+0.3i;2+0.5i][1+1i;0.5+0.3i;2+0.5i], like in the case depicted in Figure 4.

Figure 4 and Figure 5 depict the lower and upper bounds on Re[v(t)]\mathrm{Re}[v(t)] for two systems (z(ω)=2+i/ωz(\omega)=2+i/\omega in Figure 4, thus mimicking the low frequency dielectric response of a lossy dielectric material, and z(ω)=22/ω2z(\omega)=2-2/\omega^{2} in Figure 5, thus mimicking the dielectric response of a plasma), when the coefficients αk\alpha_{k} in (6.3) are chosen such that the bounds are extremely tight at t0=0t_{0}=0, according to (4.3). For both systems, the bounds on Re[v(t)]\mathrm{Re}[v(t)] are tighter the higher the amount of pieces of information on the system is incorporated. Notice that the bounds colored in black (the largest ones) correspond to the case where only the zeroth order moment M0M_{0} of the measure is known but not the value of a0a_{0}: in such a case, as shown by the zoomed graph in the blue box, at t=0t=0, the upper bound takes value 1 and the lower bound takes value 0, that are the smallest and the highest values a0a_{0} can take. On the other hand, when a0a_{0} is assigned, the value that the corresponding bounds take at t=0t=0 is exactly a0=0.6a_{0}=0.6, as shown by the zoomed graph in the blue box. The graphs show clearly that, in order to estimate the system parameter a0a_{0}, one has just to measure the response of the system at a specific moment of time t0t_{0} (if the applied field is carefully chosen).

These are the type of bounds used in [33] to bound the temporal response of two-phase composites in antiplane elasticity. It is not yet clear whether those bounds can be derived from variational principles. In general, in the theory of composites, variational methods have proven to be more powerful than analytic approaches. Variational methods produce tighter bounds that often easily extend to multiphase composites: see the books [13, 49, 37, 2, 48] and references therein. For example, the variational approach gives tighter bounds on the complex permittivity at constant frequency of two-phase lossy composites [24], than the bounds obtained by the analytic approach [35, 8]. It also produces bounds on the complex effective bulk and shear moduli of viscoelastic composites [14, 42]. An exception is bounds that correlate the complex effective dielectric constant at more than two frequencies [36] that have yet to be obtained by a systematic variational approach. Variational bounds in the time domain are available [12, 32], but these are nonlocal in time.

7 Using an appropriate input signal to predict the response at a given frequency

Naturally, if one is interested in the response v0(t)v_{0}(t) at a given (possibly complex) frequency ω0\omega_{0}, the easiest solution is to take an input signal u0(t)u_{0}(t) at that frequency. However, it might not be easy to experimentally generate a signal at that frequency or it might not be easy to measure the response at that frequency. The problem becomes: find complex constants α1,α2,,αm\alpha_{1},\alpha_{2},\ldots,\alpha_{m} such that

supλ[1,1]|k=1mαkλzk1λz0|ϵm,\sup_{\lambda\in[-1,1]}\left|\sum_{k=1}^{m}\frac{\alpha_{k}}{\lambda-z_{k}}-\frac{1}{\lambda-z_{0}}\right|\leq\epsilon_{m}, (7.1)

with zk=z(ωk)z_{k}=z(\omega_{k}), k=0,1,,mk=0,1,\dots,m. Defining the polynomials p(λ)p(\lambda) and q(λ)q(\lambda) as in (4.8) and (4.9) one needs to find p(λ)p(\lambda) of degree m1m-1 such

supλ[1,1]|(λz0)p(λ)q(λ)(λz0)q(λ)|ϵm.\sup_{\lambda\in[-1,1]}\left|\frac{(\lambda-z_{0})p(\lambda)-q(\lambda)}{(\lambda-z_{0})q(\lambda)}\right|\leq\epsilon_{m}. (7.2)

Proceeding as before we choose

(λz0)p(λ)=q(λ)bmTm1(λ)withbm=q(z0)/Tm1(z0),(\lambda-z_{0})p(\lambda)=q(\lambda)-b_{m}T_{m-1}(\lambda)\quad\text{with}\,\,b_{m}=q(z_{0})/T_{m-1}(z_{0}), (7.3)

where bmb_{m} has been chosen so that the polynomial q(λ)bmTm1(λ)q(\lambda)-b_{m}T_{m-1}(\lambda) has a factor of (λz0)(\lambda-z_{0}). Then the residues of R(λ)=p(λ)/q(λ)R(\lambda)=p(\lambda)/q(\lambda) are given by

αk=bmTm1(zk)(zkz0)jk(zkzj)=Tm1(zk)j0(z0zj)Tm1(z0)(zkz0)j0,k(zkzj),\alpha_{k}=-b_{m}\frac{T_{m-1}(z_{k})}{(z_{k}-z_{0})\prod_{j\neq k}(z_{k}-z_{j})}=-\frac{T_{m-1}(z_{k})\prod_{j\neq 0}(z_{0}-z_{j})}{T_{m-1}(z_{0})(z_{k}-z_{0})\prod_{j\neq 0,k}(z_{k}-z_{j})}, (7.4)

and

supλ[1,1]|(λz0)p(λ)q(λ)|=supλ[1,1]|bmTm1(λ)|=|bm|,\sup_{\lambda\in[-1,1]}|(\lambda-z_{0})p(\lambda)-q(\lambda)|=\sup_{\lambda\in[-1,1]}|b_{m}T_{m-1}(\lambda)|=|b_{m}|, (7.5)

so that (7.1) holds with

ϵm=|bm|d0infλ[1,1]|q(λ)|,\epsilon_{m}=\frac{|b_{m}|}{d_{0}\inf_{\lambda\in[-1,1]}|q(\lambda)|}, (7.6)

where d0d_{0} denotes the distance from z0z_{0} to the interval [-1,1]. Joukowski’s map yields:

z0=12(ζ0+1ζ0),withR=|ζ0|>1,z_{0}=\frac{1}{2}\left(\zeta_{0}+\frac{1}{\zeta_{0}}\right),\,\,\text{with}\,\,R=|\zeta_{0}|>1, (7.7)

whence

Tm1(z0)=12(ζ0m1+1ζ0m1).T_{m-1}(z_{0})=\frac{1}{2}\left(\zeta_{0}^{m-1}+\frac{1}{\zeta_{0}^{m-1}}\right). (7.8)

Moreover, since ζ0\zeta_{0} runs over a circle of radius RR, we have

d0=infλ[1,1]|ζ022ζ0λ+1|2(R1)22R,d_{0}=\inf_{\lambda\in[-1,1]}\frac{|\zeta_{0}^{2}-2\zeta_{0}\lambda+1|}{2}\geq\frac{(R-1)^{2}}{2R}, (7.9)

and

|Tm1(z0)|12(Rm1R1m),m2,|T_{m-1}(z_{0})|\geq\frac{1}{2}(R^{m-1}-R^{1-m}),\ \ m\geq 2, (7.10)

implying

|bm|2|q(z0)|Rm1R1m.|b_{m}|\leq\frac{2|q(z_{0})|}{R^{m-1}-R^{1-m}}. (7.11)

All in all, the relevant bound ϵm\epsilon_{m} satisfies

|ϵm|4R(R1)21Rm1R1msupλ[1,1]|q(z0)||q(λ)|.|\epsilon_{m}|\leq\frac{4R}{(R-1)^{2}}\frac{1}{R^{m-1}-R^{1-m}}\sup_{\lambda\in[-1,1]}\frac{|q(z_{0})|}{|q(\lambda)|}. (7.12)

We obtain an exponential decay ϵm0\epsilon_{m}\rightarrow 0 as mm\rightarrow\infty provided the geometry of the loci z1,z2,,zmz_{1},z_{2},\ldots,z_{m} is subject to the following condition: for a positive constant r<Rr<R, each zjH(r)=H1(r)H2(r)H3(r)z_{j}\in H(r)=H_{1}(r)\cup H_{2}(r)\cup H_{3}(r) where

H1(r)\displaystyle H_{1}(r) =\displaystyle= {z:|zz0z+1|r,Rez1},\displaystyle\left\{z:\left|\frac{z-z_{0}}{z+1}\right|\leq r,\ \ \mathop{\rm Re}\nolimits z\leq-1\right\},
H2(r)\displaystyle H_{2}(r) =\displaystyle= {z:|zz0Imz|r,Rez[1,1]},\displaystyle\left\{z:\left|\frac{z-z_{0}}{\mathop{\rm Im}\nolimits z}\right|\leq r,\ \ \mathop{\rm Re}\nolimits z\in[-1,1]\right\},
H3(r)\displaystyle H_{3}(r) =\displaystyle= {z:|zz0z1|r,Rez1}.\displaystyle\left\{z:\left|\frac{z-z_{0}}{z-1}\right|\leq r,\ \ \mathop{\rm Re}\nolimits z\geq 1\right\}. (7.13)

In other words, all of the zjz_{j} must be close to z0z_{0} in the precise sense that zjH(r)z_{j}\in H(r). Note that, as shown in Figure 6a, in case r<1r<1, H1H_{1} and H3H_{3} are sectors of disks, while H2H_{2} is a portion of an ellipse. For r(1,R)r\in(1,R) these regions are complements of disks/ellipse, containing the point z0z_{0}, as shown in Figure 6c. Some of these regions can be empty, depending on the position of z0z_{0}.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 6: Representation of the loci zkz_{k} for a system for which z0=0.3088240.764706iz_{0}=0.308824-0.764706\,i, and R=2.061R=2.061.

A conservative choice would be r=1r=1 (see Figure 6b), in which situation H1H_{1} and H3H_{3} are bounded by straight lines, while H2H_{2} is a parabola. To fix ideas, let us assume z0=x0+iy0z_{0}=x_{0}+iy_{0} with x00x_{0}\geq 0 and y00y_{0}\geq 0. All other cases being symmetrical. Then the euclidean region H(1)H(1) where z1,z2,,zmz_{1},z_{2},\ldots,z_{m} are allowed consists of points z=x+iyz=x+iy subject to the constraints:

x1anddist(z,z0)dist(z,1),x\geq 1\ \ {\rm and}\ \ {\rm dist}(z,z_{0})\leq\ {\rm dist}(z,1), (7.14)

union with

x[1,1]and(xx0)2+y022y0y.x\in[-1,1]\ \ {\rm and}\ \ (x-x_{0})^{2}+y_{0}^{2}\leq 2y_{0}y. (7.15)

If y0=0y_{0}=0 then necessarily x0>1x_{0}>1 and HH is simply the right-half plane x>1+x02x>\frac{1+x_{0}}{2}, while in the case y0>0y_{0}>0, H(1)H(1) is the interior of a parabola with vertex at (x0,y02)(x_{0},\frac{y_{0}}{2}), within the band |x|1|x|\leq 1, union with the polygonal region defined by the first distance inequality (in x1x\geq 1).

Now with an input signal of the form (3.1), with βk=αk/c(ωk)\beta_{k}=\alpha_{k}/c(\omega_{k}), generating the output function v(t)v(t) given by (3.2), (7.1) implies the bound

|v(t0)v0(t0)|a0ϵm,|v(t_{0})-v_{0}(t_{0})|\leq a_{0}\epsilon_{m}, (7.16)

where

v0(t0)=a0Fμ(z(ω0))v_{0}(t_{0})=a_{0}F_{\mu}(z(\omega_{0})) (7.17)

is the response at time t0t_{0} to the single frequency input signal

u0(t)=eiω0(tt0)/c(ω0).u_{0}(t)=e^{-i\omega_{0}(t-t_{0})}/c(\omega_{0}). (7.18)

Of course, because this response v0(t)v_{0}(t) is for a single frequency, v0(t0)v_{0}(t_{0}) determines v0(t)v_{0}(t) for all tt.

In Figure 7 we depict the response v0(t)v_{0}(t) of a given system subject to an input signal at the frequency ω0\omega_{0} and we compare the value it takes at t0=0t_{0}=0 with the value taken by the bounds on the response v(t)v(t) of a system having the same values of the moments of the measure but subject to a multiple-frequency signal with amplitudes αk\alpha_{k} chosen such that the bounds are extremely tight at t0=0t_{0}=0: v0(t0)v_{0}(t_{0}) lies, as expected, between the bounds on v(t)v(t) at t=t0t=t_{0},

Refer to caption
(a) z=2iω\displaystyle z=2-\frac{i}{\omega}
Refer to caption
(b) z=22ω2\displaystyle z=2-\frac{2}{\omega^{2}}
Figure 7: Comparison between the response v0(t)v_{0}(t) of a given system with point masses at -0.5 and 0.5, due to an input at the frequency ω0=0.7i\omega_{0}=0.7{i}, and the upper and lower bounds on the response v(t)v(t) of a system having the same value of the moments of the measure MiM_{i} (M0=1M_{0}=1 and M1=0.4M_{1}=0.4) and subject to an input signal of the type (3.1), with ωk\omega_{k} given by [1+1i;0.5+0.3i;2+0.5i][1+1i;0.5+0.3i;2+0.5i] and coefficients βk\beta_{k} chosen accordingly to (3.3) and (7.4). Notice that in both cases the value of v0(t)v_{0}(t) at t0=0t_{0}=0 lies between the bounds on v(t)v(t) at t0=0t_{0}=0.

Remark

The analysis is easily extended to the case where the response v0(t)v_{0}(t) is known for a given ω0\omega_{0} but one wants to predict the derivative

v0(t0)dω0=a0dFμ(z)dz|z=z(ω0)dz(ω0)dω0.\frac{v_{0}(t_{0})}{d\omega_{0}}=a_{0}\frac{dF_{\mu}(z)}{dz}\bigg{|}_{z=z(\omega_{0})}\frac{dz(\omega_{0})}{d\omega_{0}}. (7.19)

As

dFμ(z)dz=11dμ(λ)(λz)2,\frac{dF_{\mu}(z)}{dz}=\int_{-1}^{1}\frac{d\mu(\lambda)}{(\lambda-z)^{2}}, (7.20)

the problem becomes: find complex constants α0,α1,α2,,αn\alpha_{0},\alpha_{1},\alpha_{2},\ldots,\alpha_{n} such that

supλ[1,1]|j=1mαjλzj+α0λz01(λz0)2|ϵm.\sup_{\lambda\in[-1,1]}\left|\sum_{j=1}^{m}\frac{\alpha_{j}}{\lambda-z_{j}}+\frac{\alpha_{0}}{\lambda-z_{0}}-\frac{1}{(\lambda-z_{0})^{2}}\right|\leq\epsilon_{m}. (7.21)

Defining the polynomials p(λ)p(\lambda) and q(λ)q(\lambda) as in (4.8) and (4.9) one needs to find p(λ)p(\lambda) of degree m1m-1 such that

supλ[1,1]|(λz0)2p(λ)q(λ)[1α0(λz0)](λz0)2q(λ)|ϵm.\sup_{\lambda\in[-1,1]}\left|\frac{(\lambda-z_{0})^{2}p(\lambda)-q(\lambda)[1-\alpha_{0}(\lambda-z_{0})]}{(\lambda-z_{0})^{2}q(\lambda)}\right|\leq\epsilon_{m}. (7.22)

We now choose

(λz0)2p(λ)=q(λ)[1α0(λz0)]bmTm1(λ),(\lambda-z_{0})^{2}p(\lambda)=q(\lambda)[1-\alpha_{0}(\lambda-z_{0})]-b_{m}T_{m-1}(\lambda), (7.23)

with

bm=q(z0)/Tm1(z0),α0=q(λ)bmTm1(λ)q(z0)b_{m}=q(z_{0})/T_{m-1}(z_{0}),\quad\alpha_{0}=\frac{q^{\prime}(\lambda)-b_{m}T_{m-1}^{\prime}(\lambda)}{q(z_{0})} (7.24)

selected so that the polynomial on the right hand side of (7.23) has a factor of (λz0)2(\lambda-z_{0})^{2}, in which q(λ)=dq(λ)/dλq^{\prime}(\lambda)=dq(\lambda)/d\lambda and Tm1(λ)=dTm1(λ)/dλT_{m-1}^{\prime}(\lambda)=dT_{m-1}(\lambda)/d\lambda. So the residues αk\alpha_{k}, for k0k\neq 0, are now given by

αk=bmTm(zk)(zkz0)2jk(zkzj)=Tm(zk)j0(z0zj)Tm(z0)(zkz0)2j0,k(zkzj),\alpha_{k}=-b_{m}\frac{T_{m}(z_{k})}{(z_{k}-z_{0})^{2}\prod_{j\neq k}(z_{k}-z_{j})}=-\frac{T_{m}(z_{k})\prod_{j\neq 0}(z_{0}-z_{j})}{T_{m}(z_{0})(z_{k}-z_{0})^{2}\prod_{j\neq 0,k}(z_{k}-z_{j})}, (7.25)

where bmb_{m} is still given by (7.3) and

supλ[1,1]|(λz0)2p(λ)q(λ)[1α0(λz0)]|=supλ[1,1]|bmTm1(λ)|=|bm|\sup_{\lambda\in[-1,1]}|(\lambda-z_{0})^{2}p(\lambda)-q(\lambda)[1-\alpha_{0}(\lambda-z_{0})]|=\sup_{\lambda\in[-1,1]}|b_{m}T_{m-1}(\lambda)|=|b_{m}| (7.26)

so that (7.21) holds with

ϵm=|bm|d02infλ[1,1]|q(λ)|.\epsilon_{m}=\frac{|b_{m}|}{d_{0}^{2}\inf_{\lambda\in[-1,1]}|q(\lambda)|}. (7.27)

Apart from an extra factor of d0d_{0} this is exactly the same as the formula (7.6), and so the convergence ϵm0\epsilon_{m}\to 0 as mm\to\infty is assured provided for a positive constant r<Rr<R, with each zjH(r)=H1(r)H2(r)H3(r)z_{j}\in H(r)=H_{1}(r)\cup H_{2}(r)\cup H_{3}(r).

8 Framework for a wide variety of time-dependent problems

Suppose that, in some Hilbert (or vector space) {\cal H}, one is interested in solving for 𝐉{\bf J} the equations

𝐉=𝐋𝐄,𝐐𝐉=𝐉,𝐐𝐄=𝐄0,{\bf J}={\bf L}{\bf E},\quad{\bf Q}{\bf J}={\bf J},\quad{\bf Q}{\bf E}={\bf E}_{0}, (8.1)

for a prescribed field 𝐄0{\bf E}_{0}, where 𝐋:{\bf L}:{\cal H}\to{\cal H} is an operator satisfying appropriate boundedness and coercivity conditions, and 𝐐{\bf Q} is a selfadjoint projection onto a subspace 𝒮{\cal S} of {\cal H}, so that both 𝐄0{\bf E}_{0} and 𝐉{\bf J} lie in 𝒮{\cal S}. Note that we can rewrite (8.1) as

𝐉=𝐋𝐄𝐬,𝚪1𝐄=𝐄,𝚪1𝐉=0,{\bf J}={\bf L}{\bf E}^{\prime}-{\bf s},\quad\mbox{\boldmath${\Gamma}$}_{1}{\bf E}^{\prime}={\bf E}^{\prime},\quad\mbox{\boldmath${\Gamma}$}_{1}{\bf J}=0, (8.2)

with 𝐄=𝐄𝐄0{\bf E}^{\prime}={\bf E}-{\bf E}_{0}, 𝐬=𝐋𝐄0{\bf s}=-{\bf L}{\bf E}_{0} being the source term, and 𝚪1=𝐈𝐐\mbox{\boldmath${\Gamma}$}_{1}={\bf I}-{\bf Q} being the projection onto the orthogonal complement of 𝒮{\cal S} in {\cal H}. These equations arise in the extended abstract theory of composites and apply to an enormous plethora of linear continuum equations in physics: see, for example, the books [37, 43] and the articles [38, 39, 40, 41].

The simplest example is for electrical conductivity (and equivalent equations), where one has

𝐣(𝐱)=𝝈(𝐱)𝐞(𝐱)𝐬(𝐱),𝚪1𝐞=𝐞,𝚪1𝐣=0,with𝚪1=(2)1,{\bf j}^{\prime}({\bf x})=\mbox{\boldmath${\sigma}$}({\bf x}){\bf e}({\bf x})-{\bf s}({\bf x}),\quad\mbox{\boldmath${\Gamma}$}_{1}{\bf e}={\bf e},\quad\mbox{\boldmath${\Gamma}$}_{1}{\bf j}^{\prime}=0,\quad\text{with}\quad\mbox{\boldmath${\Gamma}$}_{1}=\nabla(\nabla^{2})^{-1}\nabla\cdot, (8.3)

where 𝝈(𝐱)\mbox{\boldmath${\sigma}$}({\bf x}) is the conductivity tensor, while 𝐬\nabla\cdot{\bf s}, 𝐣=𝐣+𝐬{\bf j}={\bf j}^{\prime}+{\bf s}, and 𝐞{\bf e} are the current source, current, and electric field, and (2)1(\nabla^{2})^{-1} is the inverse Laplacian (there is obviously considerable flexibility in the choice of 𝐬(𝐱){\bf s}({\bf x}), the only constraints being square integrability and that 𝐬\nabla\cdot{\bf s} equals the current source). As current is conserved, 𝐣=𝐬\nabla\cdot{\bf j}=\nabla\cdot{\bf s}, implying 𝐣=0\nabla\cdot{\bf j}^{\prime}=0, which is clearly equivalent to 𝚪1𝐣=0\mbox{\boldmath${\Gamma}$}_{1}{\bf j}^{\prime}=0. In Fourier space 𝚪1(𝐤)=𝐤𝐤/k2\mbox{\boldmath${\Gamma}$}_{1}({\bf k})={\bf k}\otimes{\bf k}/k^{2}, and 𝚪1𝐞=𝐞\mbox{\boldmath${\Gamma}$}_{1}{\bf e}={\bf e} implies the Fourier components 𝐞^(𝐤)\widehat{{\bf e}}({\bf k}) of 𝐞{\bf e} satisfy 𝐞^=i𝐤(i𝐤𝐞^)/k2\widehat{{\bf e}}=-i{\bf k}(i{\bf k}\cdot\widehat{\bf e})/k^{2}. So 𝐞{\bf e} is the gradient of a potential with Fourier components i𝐤𝐞^/k2-i{\bf k}\cdot\widehat{\bf e}/k^{2}. In antiplane elasticity one takes a material with a cross-section in the (x1,x2)(x_{1},x_{2})-plane that is independent of x3x_{3}, applies shearing in the x3x_{3} direction and observes warping of the cross section. The displacement u3(𝐱)u_{3}({\bf x}) in the x3x_{3}-direction that is associated with this warping satisfies a conductivity type equation Gu3=s\nabla\cdot G\nabla u_{3}=\nabla s, where s\nabla s is a shearing source term (dependent on (x1,x2)(x_{1},x_{2})), G(x1,x2)G(x_{1},x_{2}) is the shear modulus, and correspondingly 𝐞=u3{\bf e}=-\nabla u_{3} and 𝐣=Gu3{\bf j}=G\nabla u_{3}. The antiplane response also governs the warping of rods under torsion for rods that have a non-circular cylindrical shape and are composed of long fibers aligned with the cylinder axis and embedded in a matrix such that the fiber separation is much less than the cylinder circumference.

One approach to solving (8.1) is to apply 𝐐{\bf Q} to both sides of the relation 𝐄=𝐋1𝐉{\bf E}={\bf L}^{-1}{\bf J} to obtain 𝐄0=𝐐𝐋1𝐐𝐉{\bf E}_{0}={\bf Q}{\bf L}^{-1}{\bf Q}{\bf J}, giving

𝐉=[𝐐𝐋1𝐐]1𝐄0,{\bf J}=[{\bf Q}{\bf L}^{-1}{\bf Q}]^{-1}{\bf E}_{0}, (8.4)

where the inverse is on the subspace 𝒮{\cal S}. In general the operator 𝐋{\bf L} depends on the frequency ω\omega and 𝐄0{\bf E}_{0} could depend on ω\omega too. Then the response at this frequency is

𝐉^(ω)=[𝐐(𝐋(ω))1𝐐]1𝐄0^(ω).\widehat{{\bf J}}(\omega)=[{\bf Q}({\bf L}(\omega))^{-1}{\bf Q}]^{-1}\widehat{{\bf E}_{0}}(\omega). (8.5)

We are interested in the response in the time domain when 𝐄0^(ω)=β(ω)𝐄¯0\widehat{{\bf E}_{0}}(\omega)=\beta(\omega)\underline{{\bf E}}_{0} for some complex amplitude β(ω)\beta(\omega) and 𝐄¯0𝒮\underline{{\bf E}}_{0}\in{\cal S} does not depend on ω\omega. In particular, for a sum of a finite number of (possibly complex) frequencies in the time domain the input signal is

𝐄0(t)=k=1meiωk(tt0)𝐄0^(ωk)=k=1mβkeiωk(tt0)𝐄¯0withβk=β(ωk).{\bf E}_{0}(t)=\sum_{k=1}^{m}e^{-i\omega_{k}(t-t_{0})}\widehat{{\bf E}_{0}}(\omega_{k})=\sum_{k=1}^{m}\beta_{k}e^{-i\omega_{k}(t-t_{0})}\underline{{\bf E}}_{0}\quad\text{with}\quad\beta_{k}=\beta(\omega_{k}). (8.6)

The resulting field 𝐉(t){\bf J}(t) is then

𝐉(t)=k=1mβkeiωk(tt0)[𝐐(𝐋(ω))1𝐐]1𝐄¯0,{\bf J}(t)=\sum_{k=1}^{m}\beta_{k}e^{-i\omega_{k}(t-t_{0})}[{\bf Q}({\bf L}(\omega))^{-1}{\bf Q}]^{-1}\underline{{\bf E}}_{0}, (8.7)

and we want this to have a simple approximate formula at time t0t_{0}.

To make progress we use another approach to solving (8.1). We introduce a “reference medium” 𝐋0=c0𝐈{\bf L}_{0}=c_{0}{\bf I} where the real constant c0c_{0} is chosen so that 𝐋𝐋0{\bf L}-{\bf L}_{0} is coercive and introduce the so-called “polarization field”

𝐆=(𝐋𝐋0)𝐄=(𝐋c0𝐈)𝐄=𝐉c0𝐄.{\bf G}=({\bf L}-{\bf L}_{0}){\bf E}=({\bf L}-c_{0}{\bf I}){\bf E}={\bf J}-c_{0}{\bf E}. (8.8)

Applying the projection 𝐈𝐐{\bf I}-{\bf Q} to this equation gives

(𝐈𝐐)𝐆=c0(𝐄𝐄0)=c0𝐄0c0(𝐋c0𝐈)1𝐆,({\bf I}-{\bf Q}){\bf G}=-c_{0}({\bf E}-{\bf E}_{0})=c_{0}{\bf E}_{0}-c_{0}({\bf L}-c_{0}{\bf I})^{-1}{\bf G}, (8.9)

and solving this for 𝐆{\bf G} yields

𝐆=c0[(𝐈𝐐)+c0(𝐋c0𝐈)1]1𝐄0.{\bf G}=c_{0}[({\bf I}-{\bf Q})+c_{0}({\bf L}-c_{0}{\bf I})^{-1}]^{-1}{\bf E}_{0}. (8.10)

Finally, applying 𝐐{\bf Q} to both sides gives

𝐉=c0{𝐐+𝐐[(𝐈𝐐)+c0(𝐋c0𝐈)1]1𝐐}𝐄0.{\bf J}=c_{0}\left\{{\bf Q}+{\bf Q}[({\bf I}-{\bf Q})+c_{0}({\bf L}-c_{0}{\bf I})^{-1}]^{-1}{\bf Q}\right\}{\bf E}_{0}. (8.11)

By comparing (8.4) and (8.10) we have

[𝐐𝐋1𝐐]1\displaystyle[{\bf Q}{\bf L}^{-1}{\bf Q}]^{-1} =\displaystyle= c0𝐐+c0𝐐[(𝐈𝐐)+c0(𝐋c0𝐈)1]1𝐐\displaystyle c_{0}{\bf Q}+c_{0}{\bf Q}[({\bf I}-{\bf Q})+c_{0}({\bf L}-c_{0}{\bf I})^{-1}]^{-1}{\bf Q} (8.12)
=\displaystyle= c0{𝐐2𝐐[𝚿(𝐋+c0𝐈)(𝐋c0𝐈)1]1𝐐},\displaystyle c_{0}\left\{{\bf Q}-2{\bf Q}[\mbox{\boldmath${\Psi}$}-({\bf L}+c_{0}{\bf I})({\bf L}-c_{0}{\bf I})^{-1}]^{-1}{\bf Q}\right\},

where 𝚿=2𝐐𝐈\mbox{\boldmath${\Psi}$}=2{\bf Q}-{\bf I} has eigenvalues ±1\pm 1. It is not obvious at all that the right hand side of (8.12) is independent of c0c_{0} but the preceding derivation shows this. This type of solution using a reference medium 𝐋0{\bf L}_{0} (that need not be proportional to 𝐈{\bf I}) is well known in the theory of composites: see, for example Chapter 14 of [37], [57], and references therein.

Now assume 𝐋{\bf L} takes the form

𝐋=c1𝐏+c2(𝐈𝐏),{\bf L}=c_{1}{\bf P}+c_{2}({\bf I}-{\bf P}), (8.13)

where 𝐏{\bf P} is a projection operator onto a subspace 𝒫{\cal P} of {\cal H}. In the theory of composites for two phase composites one frequently has

𝐋=c1𝐈χ(𝐱)+c2𝐈(1χ(𝐱)),{\bf L}=c_{1}{\bf I}\chi({\bf x})+c_{2}{\bf I}(1-\chi({\bf x})), (8.14)

where the characteristic function χ(𝐱)\chi({\bf x}) is 11 in phase 1 and 0 in phase 2, and c1c_{1} and c2c_{2} could be the material moduli. For the antiplane elasticity problem one has c1=G1c_{1}=G_{1} and c2=G2c_{2}=G_{2}, where G1G_{1} and G2G_{2} are the shear moduli of the phases. We take the limit c0c2c_{0}\to c_{2} and then (8.12) becomes

[𝐐𝐋1𝐐]1=c2𝐐+2c2𝐐𝐏[𝐏𝚿𝐏z𝐏]1𝐏𝐐,[{\bf Q}{\bf L}^{-1}{\bf Q}]^{-1}=c_{2}{\bf Q}+2c_{2}{\bf Q}{\bf P}[{\bf P}\mbox{\boldmath${\Psi}$}{\bf P}-z{\bf P}]^{-1}{\bf P}{\bf Q}, (8.15)

where the operator inverse is to be taken on the subspace 𝒫{\cal P} and

z=c1+c2c1c2.z=\frac{c_{1}+c_{2}}{c_{1}-c_{2}}. (8.16)

Note that 𝐏𝚿𝐏{\bf P}\mbox{\boldmath${\Psi}$}{\bf P}, like 𝚿{\Psi}, has norm at most 11. In general, the two moduli c1c_{1} and c2c_{2} depend on the frequency ω\omega and hence zz defined by (8.16) will also, i.e. z=z(ω)z=z(\omega). Given an input field of the form (8.6) and letting

𝐉2(t)=𝐐k=1mβkc2(ωk)eiωk(tt0)𝐄¯0{\bf J}_{2}(t)={\bf Q}\sum_{k=1}^{m}\beta_{k}c_{2}(\omega_{k})e^{-i\omega_{k}(t-t_{0})}\underline{{\bf E}}_{0} (8.17)

denote the response when 𝐏=0{\bf P}=0, i.e. when 𝐋(ω)=c2(ω)𝐈{\bf L}(\omega)=c_{2}(\omega){\bf I}, the corresponding output field can be taken to be

𝐯(t)=𝐉(t)𝐉2(t)=𝐐k=1mαkeiωk(tt0)2𝐏[𝐏𝚿𝐏zk𝐏]1𝐏𝐄¯0,{\bf v}(t)={\bf J}(t)-{\bf J}_{2}(t)={\bf Q}\sum_{k=1}^{m}\alpha_{k}e^{-i\omega_{k}(t-t_{0})}2{\bf P}[{\bf P}\mbox{\boldmath${\Psi}$}{\bf P}-z_{k}{\bf P}]^{-1}{\bf P}\underline{{\bf E}}_{0}, (8.18)

with

zk=z(ωk)=c1(ωk)+c2(ωk)c1(ωk)c2(ωk),αk=βkc2(ωk),z_{k}=z(\omega_{k})=\frac{c_{1}(\omega_{k})+c_{2}(\omega_{k})}{c_{1}(\omega_{k})-c_{2}(\omega_{k})},\quad\quad\alpha_{k}=\beta_{k}c_{2}(\omega_{k}), (8.19)

and we arrive back at the problem we have been studying. In particular, with constants αk\alpha_{k} given by (4.3) the inequality (3.9)(\ref{0.10}) with n=0n=0 implies

|𝐉(t0)𝐉2(t0)2𝐐𝐏𝐄¯0|4|𝐏𝐄¯0|/(2dmin)m.|{\bf J}(t_{0})-{\bf J}_{2}(t_{0})-2{\bf Q}{\bf P}\underline{{\bf E}}_{0}|\leq 4|{\bf P}\underline{{\bf E}}_{0}|/(2d_{\rm min})^{m}. (8.20)

Alternatively, we could have chosen c0=c1c_{0}=c_{1} and let

𝐉1(t)=𝐐k=1mβkc1(ωk)eiωk(tt0)𝐄¯0{\bf J}_{1}(t)={\bf Q}\sum_{k=1}^{m}\beta_{k}c_{1}(\omega_{k})e^{-i\omega_{k}(t-t_{0})}\underline{{\bf E}}_{0} (8.21)

denote the response when 𝐏=𝐈{\bf P}={\bf I}, i.e. when 𝐋(ω)=c1(ω)𝐈{\bf L}(\omega)=c_{1}(\omega){\bf I}. Then, similarly to (8.18), we would have

𝐉(t)𝐉1(t)=𝐐k=1mαkeiωk(tt0)2𝐏[(𝐏𝚿𝐏+zk𝐏]1𝐏𝐄¯0,{\bf J}(t)-{\bf J}_{1}(t)={\bf Q}\sum_{k=1}^{m}\alpha_{k}e^{-i\omega_{k}(t-t_{0})}2{\bf P}_{\perp}[({\bf P}_{\perp}\mbox{\boldmath${\Psi}$}{\bf P}_{\perp}+z_{k}{\bf P}_{\perp}]^{-1}{\bf P}_{\perp}\underline{{\bf E}}_{0}, (8.22)

where zkz_{k} is still given by (8.19), but now with αk=βkc1(ωk)\alpha_{k}=\beta_{k}c_{1}(\omega_{k}), where 𝐏=𝐈𝐏{\bf P}_{\perp}={\bf I}-{\bf P} is the projection onto the subspace perpendicular to 𝒫{\cal P}. The problem, with n=0n=0 and with the same choice of coefficients αk\alpha_{k}, requires a different input signal, i.e. a different choice of the βk\beta_{k} given by βk=βk/c1(ωk)\beta_{k}=\beta_{k}/c_{1}(\omega_{k}), to ensure that

|𝐉(t0)𝐉1(t0)+2𝐐𝐏𝐄¯0|4|𝐏𝐄¯0|/(2dmin)m.|{\bf J}(t_{0})-{\bf J}_{1}(t_{0})+2{\bf Q}{\bf P}_{\perp}\underline{{\bf E}}_{0}|\leq 4|{\bf P}\underline{{\bf E}}_{0}|/(2d_{\rm min})^{m}. (8.23)

9 Framework in the context of the theory of composites and its generalizations

In the theory of composites and its generalizations, one can identify a subspace of 𝒮{\cal S} that we call 𝒰{\cal U} of “source free” fields, and we may wish to confine 𝐄0{\bf E}_{0} to this subspace. Then (8.1) can be rewritten as

𝐉=𝐋𝐄,𝚪2𝐄=0,𝚪1𝐉=0,𝚪0𝐄=𝐄0,{\bf J}={\bf L}{\bf E},\quad\mbox{\boldmath${\Gamma}$}_{2}{\bf E}=0,\quad\mbox{\boldmath${\Gamma}$}_{1}{\bf J}=0,\quad\mbox{\boldmath${\Gamma}$}_{0}{\bf E}={\bf E}_{0}, (9.1)

where 𝚪0\mbox{\boldmath${\Gamma}$}_{0} is the projection onto 𝒰{\cal U}, 𝚪1\mbox{\boldmath${\Gamma}$}_{1} is the projection onto {\cal E}, defined as the orthogonal complement of 𝒮{\cal S}, and 𝚪2\mbox{\boldmath${\Gamma}$}_{2} is the projection onto 𝒥{\cal J}, defined as the orthogonal complement of 𝒰{\cal U} in the subspace 𝒮{\cal S}. Then 𝐐=𝚪0+𝚪2{\bf Q}=\mbox{\boldmath${\Gamma}$}_{0}+\mbox{\boldmath${\Gamma}$}_{2} and the Hilbert space {\cal H} has the decomposition

=𝒰𝒥,{\cal H}={\cal U}\oplus{\cal E}\oplus{\cal J}, (9.2)

and the projections onto these three subspaces are respectively 𝚪0\mbox{\boldmath${\Gamma}$}_{0}, 𝚪1\mbox{\boldmath${\Gamma}$}_{1}, and 𝚪2\mbox{\boldmath${\Gamma}$}_{2}.

In particular, as observed independently in Sections 2.4 and 2.5 of [17], and in Chapter 3 of [43], the Dirichlet-Neumann problem can be reformulated as a problem in the theory of composites. In the simplest case of electrical conductivity, where one has an inclusion DD (not necessarily simply connected) of (isotropic) conductivity c1c_{1} in a simply connected body Ω\Omega having smooth boundary, with c2c_{2} being the (isotropic) conductivity of ΩD\Omega\setminus D, we may take {\cal H} as the space of vector fields that are square integrable with the usual normalized L2L^{2} inner product,

(𝐀1,𝐀2)=1|Ω|Ω𝐀1(𝐱)𝐀2(𝐱)¯𝑑𝐱,({\bf A}_{1},{\bf A}_{2})=\frac{1}{|\Omega|}\int_{\Omega}{\bf A}_{1}({\bf x})\cdot\overline{{\bf A}_{2}({\bf x})}\,d{\bf x}, (9.3)

where |Ω||\Omega| is the volume of Ω\Omega, and take

  • 𝒰{\cal U} to consist of gradients of harmonic fields 𝐮0=V{\bf u}_{0}=-\nabla V with 2V=0\nabla^{2}V=0 in Ω\Omega,

  • {\cal E} to consist of gradients 𝐞=V{\bf e}=-\nabla V with V=0V=0 on the boundary Ω\partial\Omega of Ω\Omega,

  • 𝒥{\cal J} to consist of divergent free vector fields 𝐣{\bf j} with 𝐣=0\nabla\cdot{\bf j}=0 and 𝐣𝐧=0{\bf j}\cdot{\bf n}=0 on Ω\partial\Omega, where 𝐧{\bf n} is the outwards normal to Ω\partial\Omega.

The conductivity of the body may be identified with 𝐋{\bf L} given by (8.13) where 𝐏{\bf P} is the projection onto those fields that are zero outside DD. As we are considering time-dependent problems in the quasistatic limit, where the body is small compared to the wavelength and attenuation lengths of electromagnetic waves at the frequencies ωk\omega_{k}, the moduli c1c_{1} and c2c_{2} and the fields are typically complex and frequency-dependent. The fields in 𝒰{\cal U} can be identified either by the values that VV takes on the boundary Ω\partial\Omega or by the values that the flux 𝐧V{\bf n}\cdot\nabla V takes on the boundary Ω\partial\Omega. Thus the equations (9.1) are nothing other than the Dirichlet problem in the body Ω\Omega,

𝐣=𝐋𝐞,𝐞=V,𝐣=0,𝐞0=V0,2V0=0,V=V0,onΩ,{\bf j}={\bf L}{\bf e},\quad{\bf e}=-\nabla V,\quad\nabla\cdot{\bf j}=0,\quad{\bf e}_{0}=-\nabla V_{0},\quad\nabla^{2}V_{0}=0,\quad V=V_{0},\quad\text{on}\quad\partial\Omega, (9.4)

and the mapping from 𝚪0𝐞\mbox{\boldmath${\Gamma}$}_{0}{\bf e} to 𝚪0𝐣\mbox{\boldmath${\Gamma}$}_{0}{\bf j} is nothing other than the Dirichlet to Neumann map giving 𝐧𝐣{\bf n}\cdot{\bf j} in terms of VV on Ω\partial\Omega.

For periodic two-phase conducting composites, with unit cell Ω\Omega, the framework is similar. We take {\cal H} as the space of vector fields that are Ω\Omega-periodic with the usual normalized L2L^{2} inner product, given by (9.3), and take

  • 𝒰{\cal U} to consist of gradients of constant fields 𝐮0{\bf u}_{0} (that do not depend on 𝐱{\bf x}),

  • {\cal E} to consist of gradients 𝐞=V{\bf e}=-\nabla V with VV being an Ω\Omega-periodic potential,

  • 𝒥{\cal J} to consist of Ω\Omega-periodic divergent free vector fields 𝐣{\bf j} with 𝐣=0\nabla\cdot{\bf j}=0, having zero average over Ω\Omega.

The conductivity of the body may be identified with 𝐋{\bf L} given by (8.13) where 𝐏{\bf P} is the projection onto those fields in {\cal H} that are zero outside phase 1, and c1c_{1} is the (isotropic) conductivity of phase 1 while c2c_{2} is the (isotropic) conductivity of phase 2.

Remark

More generally, the conductivity in the periodic composite could be anisotropic, with the conductivity tensor having the special form

𝐋(ω)=c1(ω)𝐋0𝐏+c2(ω)𝐋0(𝐈𝐏),{\bf L}(\omega)=c_{1}(\omega){\bf L}_{0}{\bf P}+c_{2}(\omega){\bf L}_{0}({\bf I}-{\bf P}), (9.5)

where 𝐋0{\bf L}_{0} is a constant positive definite tensor. As 𝐋0{\bf L}_{0} commutes with 𝚪0\mbox{\boldmath${\Gamma}$}_{0} and 𝐏{\bf P}, we can define new orthogonal spaces

=𝐋01/2,𝒥=𝐋01/2𝒥,𝒰=𝐋01/2𝒰=𝐋01/2𝒰=𝒰,{\cal E}^{\prime}={\bf L}_{0}^{1/2}{\cal E},\quad{\cal J}^{\prime}={\bf L}_{0}^{-1/2}{\cal J},\quad{\cal U}^{\prime}={\bf L}_{0}^{1/2}{\cal U}={\bf L}_{0}^{-1/2}{\cal U}={\cal U}, (9.6)

and rewrite (9.1) in the form

𝐉=𝐋𝐄,𝚪2𝐄=0,𝚪1𝐉=0,𝚪0𝐄=𝐄0,{\bf J}^{\prime}={\bf L}^{\prime}{\bf E}^{\prime},\quad\mbox{\boldmath${\Gamma}$}_{2}^{\prime}{\bf E}^{\prime}=0,\quad\mbox{\boldmath${\Gamma}$}_{1}^{\prime}{\bf J}^{\prime}=0,\quad\mbox{\boldmath${\Gamma}$}_{0}^{\prime}{\bf E}^{\prime}={\bf E}_{0}^{\prime}, (9.7)

where

𝐉\displaystyle{\bf J}^{\prime} =\displaystyle= 𝐋01/2𝐉,𝐄=𝐋01/2𝐄,𝐄0=𝐋01/2𝐄0,\displaystyle{\bf L}_{0}^{-1/2}{\bf J},\quad{\bf E}^{\prime}={\bf L}_{0}^{1/2}{\bf E},\quad{\bf E}^{\prime}_{0}={\bf L}_{0}^{1/2}{\bf E}_{0},
𝐋\displaystyle{\bf L}^{\prime} =\displaystyle= 𝐋01/2𝐋𝐋01/2=c1(ω)𝐏+c2(ω)(𝐈𝐏),\displaystyle{\bf L}_{0}^{-1/2}{\bf L}{\bf L}_{0}^{-1/2}=c_{1}(\omega){\bf P}+c_{2}(\omega)({\bf I}-{\bf P}), (9.8)

and

𝚪0=𝚪0,𝚪1=𝐋01/2𝚪1(𝚪1𝐋0𝚪1)1,𝚪2=𝐈𝚪1𝚪2\mbox{\boldmath${\Gamma}$}_{0}^{\prime}=\mbox{\boldmath${\Gamma}$}_{0},\quad\mbox{\boldmath${\Gamma}$}_{1}^{\prime}={\bf L}_{0}^{-1/2}\mbox{\boldmath${\Gamma}$}_{1}(\mbox{\boldmath${\Gamma}$}_{1}{\bf L}_{0}\mbox{\boldmath${\Gamma}$}_{1})^{-1},\quad\mbox{\boldmath${\Gamma}$}_{2}^{\prime}={\bf I}-\mbox{\boldmath${\Gamma}$}_{1}^{\prime}-\mbox{\boldmath${\Gamma}$}_{2}^{\prime} (9.9)

are the projections onto 𝒰=𝒰{\cal U}^{\prime}={\cal U}, {\cal E}^{\prime}, and 𝒥{\cal J}^{\prime}, in which the inverse in the formula for 𝚪1\mbox{\boldmath${\Gamma}$}_{1}^{\prime} is to be taken on the subspace {\cal E}. As 𝐋{\bf L}^{\prime} now takes the same form as (8.13) we are back to the same problem.

Similarly, in a body where the conductivity tensor has the special form (9.5) we may take

  • 𝒰{\cal U}^{\prime} to consist of gradients of fields 𝐮0=𝐋01/2V{\bf u}_{0}=-{\bf L}_{0}^{1/2}\nabla V with 𝐋0V=0\nabla\cdot{\bf L}_{0}\nabla V=0 in Ω\Omega,

  • {\cal E}^{\prime} to consist of fields 𝐞=𝐋01/2V{\bf e}^{\prime}=-{\bf L}_{0}^{1/2}\nabla V with V=0V=0 on the boundary Ω\partial\Omega of Ω\Omega,

  • 𝒥{\cal J}^{\prime} to consist of fields 𝐣{\bf j}^{\prime} with 𝐋01/2𝐣=0\nabla\cdot{\bf L}_{0}^{1/2}{\bf j}^{\prime}=0 and (𝐋01/2𝐣)𝐧=0({\bf L}_{0}^{1/2}{\bf j}^{\prime})\cdot{\bf n}=0 on Ω\partial\Omega, where 𝐧{\bf n} is the outwards normal to Ω\partial\Omega

as our three orthogonal subspaces. Letting 𝚪0\mbox{\boldmath${\Gamma}$}_{0}, 𝚪1\mbox{\boldmath${\Gamma}$}_{1}, and 𝚪2\mbox{\boldmath${\Gamma}$}_{2} denote the projections onto these three subspaces, respectively, and setting 𝐋=c1(ω)𝐏+c2(ω)(𝐈𝐏){\bf L}^{\prime}=c_{1}(\omega){\bf P}+c_{2}(\omega)({\bf I}-{\bf P}), the equations (9.6) hold and we may proceed as before.

10 Application to solving the Calderon problem with time varying fields

Let us now use ideas from the Calderon problem to solve the inverse problem of finding the inclusion DD from boundary measurements on Ω\partial\Omega. With Ω\Omega being a three-dimensional body, we can take

V0=ei𝐤𝐱withk1,k2real andk3=ik12+k22,V_{0}=e^{i{\bf k}\cdot{\bf x}}\quad\text{with}\quad k_{1},k_{2}\quad\text{real and}\quad k_{3}=i\sqrt{k_{1}^{2}+k_{2}^{2}}, (10.1)

where the last condition implies 𝐤𝐤=0{\bf k}\cdot{\bf k}=0 which ensures that V0V_{0} is harmonic. Then (8.20) implies

(𝐉(t0)𝐉1(t0)+2𝐐𝐏𝐄¯0,ei𝐤𝐱)4|𝐤||𝐤|/(2dmin)m({\bf J}(t_{0})-{\bf J}_{1}(t_{0})+2{\bf Q}{\bf P}\underline{{\bf E}}_{0},\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}})\leq 4|{\bf k}||{\bf k}^{\prime}|/(2d_{\rm min})^{m} (10.2)

for all real or complex 𝐤{\bf k}^{\prime}. We now choose 𝐤{\bf k}^{\prime} with

k3=k3,k1,k2real and with(k1)2+(k1)2=k12+k22k_{3}^{\prime}=-k_{3},\quad k_{1}^{\prime},k_{2}^{\prime}\,\,\text{real and with}\quad(k_{1}^{\prime})^{2}+(k_{1}^{\prime})^{2}=k_{1}^{2}+k_{2}^{2} (10.3)

to ensure that ei𝐤𝐱e^{i{\bf k}^{\prime}\cdot{\bf x}} is harmonic and so that

(2𝐐𝐏𝐄¯0,ei𝐤𝐱)\displaystyle(2{\bf Q}{\bf P}\underline{{\bf E}}_{0},\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}}) =\displaystyle= 2(𝐏𝐄¯0,𝐐ei𝐤𝐱)=2(𝐏𝐄¯0,ei𝐤𝐱)\displaystyle 2({\bf P}\underline{{\bf E}}_{0},{\bf Q}\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}})=2({\bf P}\underline{{\bf E}}_{0},\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}}) (10.4)
=\displaystyle= 2(k1k1+k2k2k12k22)1|Ω|Dei(k1k1)x1+i(k2k2)x2𝑑𝐱\displaystyle 2(k_{1}k^{\prime}_{1}+k_{2}k_{2}^{\prime}-k_{1}^{2}-k_{2}^{2})\frac{1}{|\Omega|}\int_{D}e^{i(k_{1}-k^{\prime}_{1})x_{1}+i(k_{2}-k^{\prime}_{2})x_{2}}\,d{\bf x}

only depends on the Fourier coefficients of the characteristic function associated with DD. Then, using integration by parts,

(𝐉(t0)𝐉1(t0),ei𝐤𝐱)=1|Ω|Ω[𝐉(t0)𝐉1(t0)]𝐧ei𝐤𝐱𝑑S,({\bf J}(t_{0})-{\bf J}_{1}(t_{0}),\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}})=\frac{1}{|\Omega|}\int_{\partial\Omega}[{\bf J}(t_{0})-{\bf J}_{1}(t_{0})]\cdot{\bf n}\,e^{i{\bf k}^{\prime}\cdot{\bf x}}\,dS, (10.5)

where 𝐉(t0)𝐧{\bf J}(t_{0})\cdot{\bf n} can be measured, while 𝐉1(t0)𝐧{\bf J}_{1}(t_{0})\cdot{\bf n} can be computed. As there is nothing special about the x3x_{3} axis we may rotate the cartesian coordinates to get estimates of other Fourier coefficients of the characteristic function associated with DD. We may also take 𝐄¯0\underline{{\bf E}}_{0} as constant and replace ei𝐤𝐱\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}} by 𝐄¯0\underline{{\bf E}}_{0} to get

(𝐉(t0)𝐉1(t0)+2𝐐𝐏𝐄¯0,𝐄¯0)=(𝐉(t0)𝐉1(t0),𝐄¯0)+|𝐄¯0|2|D|/|Ω|4|𝐄¯0|2/(2dmin)m,({\bf J}(t_{0})-{\bf J}_{1}(t_{0})+2{\bf Q}{\bf P}\underline{{\bf E}}_{0},\underline{{\bf E}}_{0})=({\bf J}(t_{0})-{\bf J}_{1}(t_{0}),\underline{{\bf E}}_{0})+|\underline{{\bf E}}_{0}|^{2}|D|/|\Omega|\leq 4|\underline{{\bf E}}_{0}|^{2}/(2d_{\rm min})^{m}, (10.6)

thus giving an estimate of the volume fraction |D|/|Ω||D|/|\Omega| that DD occupies in the body (i.e., the Fourier coefficient at 𝐤=0{\bf k}=0).

With Ω\Omega being a two-dimensional body, the situation is similar. We take

k1real andk2=ik1,k1=k1,k2=ik1,k_{1}\,\,\text{real and}\,\,k_{2}=ik_{1},\quad k_{1}^{\prime}=-k_{1},\quad k_{2}^{\prime}=-ik_{1}, (10.7)

and (10.4) is replaced by

(2𝐐𝐏𝐄¯0,ei𝐤𝐱)=4k12De2ik1x1𝑑𝐱,(2{\bf Q}{\bf P}\underline{{\bf E}}_{0},\nabla e^{i{\bf k}^{\prime}\cdot{\bf x}})=-4k_{1}^{2}\int_{D}e^{2ik_{1}x_{1}}\,d{\bf x}, (10.8)

while (10.5) and (10.6) still hold. Again we approximately recover the Fourier coefficients of the characteristic function associated with DD from measurements of 𝐉(t0)𝐧{\bf J}(t_{0})\cdot{\bf n} and computations of 𝐉1(t0)𝐧{\bf J}_{1}(t_{0})\cdot{\bf n}.

In the usual Calderon problem one solves the inverse problem by taking |𝐤||{\bf k}| to be very large, according to the so-called complex geometric optics approach [47]. Here we see that there is no need to take |𝐤||{\bf k}| to be very large if we allow time dependent applied fields. For electromagnetism in non-magnetic media the measurements are difficult as the time response is typically extremely rapid (From Table 7.7.1 in [18] we see that electromagnetic relaxation times in seconds for copper, distilled water, corn-oil, and mica are 1.5×10191.5\times 10^{-19}, 3.6×1063.6\times 10^{-6}, 0.550.55, and 5.1×1045.1\times 10^{4} respectively, and measurements would need to be taken on these time scales). On the other hand, for the equivalent magnetic permeability, fluid permeability, or antiplane elasticity problems the relaxation times are much more reasonable [5, 20, 30] and measurements in the time domain become feasible. Even in electrical systems one can get long relaxation times, such as the time to charge a capacitor.

Remark

Instead of taking 𝐄0(t)=𝚪0𝐄(t){\bf E}_{0}(t)=\mbox{\boldmath${\Gamma}$}_{0}{\bf E}(t) and 𝐉(t){\bf J}(t) as our input and output fields, one could take 𝐉0(t)=𝚪0𝐉(t){\bf J}_{0}(t)=\mbox{\boldmath${\Gamma}$}_{0}{\bf J}(t) and 𝐄(t){\bf E}(t) as our input and output fields. Then one has

𝐄=𝐋1𝐉,𝚪1𝐉=0,𝚪2𝐄=0,𝚪0𝐉=𝐉0,{\bf E}={\bf L}^{-1}{\bf J},\quad\mbox{\boldmath${\Gamma}$}_{1}{\bf J}=0,\quad\mbox{\boldmath${\Gamma}$}_{2}{\bf E}=0,\quad\mbox{\boldmath${\Gamma}$}_{0}{\bf J}={\bf J}_{0}, (10.9)

which is exactly of the same form as (9.1), but with 𝐋{\bf L} replaced by 𝐋1{\bf L}^{-1} and the roles of 𝚪1\mbox{\boldmath${\Gamma}$}_{1}, 𝚪2\mbox{\boldmath${\Gamma}$}_{2}, and 𝐄{\bf E} and 𝐉{\bf J}, and 𝐄0{\bf E}_{0} and 𝐉0{\bf J}_{0} interchanged. So all the preceding analysis immediately applies to this dual problem too.

11 Generalizations

In many problems of interest the fields in {\cal H} take values in a, say, ss-dimensional tensor space 𝒯{\cal T} and the operator 𝐋:{\bf L}:{\cal H}\to{\cal H} in (8.1), appropriately defined, is frequency dependent with the properties that

  • 𝐋(ω){\bf L}(\omega) is an analytic function of ω\omega in the upper half plane Im(ω)>0\mathop{\rm Im}\nolimits(\omega)>0,

  • Im[ω𝐋(ω)]0\mathop{\rm Im}\nolimits[\omega{\bf L}(\omega)]\geq 0 when Im(ω)>0\mathop{\rm Im}\nolimits(\omega)>0,

  • 𝐋(ω)¯=𝐋(ω¯)\overline{{\bf L}(\omega)}={\bf L}(-\overline{\omega}) when Im(ω)>0\mathop{\rm Im}\nolimits(\omega)>0,

where the overline denotes complex conjugation. By appropriately defined we mean that 𝐋(ω){\bf L}(\omega) (and accordingly 𝐉{\bf J}) may need to be multiplied by a function of ω\omega, for example ii, ω\omega, or iωi\omega, to achieve these properties. In the case of materials where 𝐋{\bf L} acts locally in real space, i.e. if 𝐐=𝐋𝐏{\bf Q}={\bf L}{\bf P}, then 𝐐(𝐱)=𝐋(𝐱)𝐏(𝐱){\bf Q}({\bf x})={\bf L}({\bf x}){\bf P}({\bf x}) for some 𝐋(𝐱){\bf L}({\bf x}), the first property is a consequence of causality, the second a consequence of passivity (that the material does not generate energy - see, for example, [53]), and the third a consequence of 𝐋(ω){\bf L}(\omega) being the Fourier transform of a real kernel. It follows that 𝐋{\bf L} is an analytic function of ω2-\omega^{2} with spectrum on the negative real ω2-\omega^{2} axis (corresponding to real values of ω\omega) having the implied properties that

  • Im(𝐋)0\mathop{\rm Im}\nolimits({\bf L})\geq 0 when Im(ω2)0\mathop{\rm Im}\nolimits(-\omega^{2})\leq 0,

  • 𝐋{\bf L} is real and 𝐋0{\bf L}\geq 0 when ω2\omega^{2} is real and ω20-\omega^{2}\geq 0.

In other words, 𝐋(ω){\bf L}(\omega) is an operator-valued Stieltjes function of ω2-\omega^{2}. The operator 𝐁=[𝐐𝐋1𝐐]1{\bf B}=[{\bf Q}{\bf L}^{-1}{\bf Q}]^{-1} entering (8.4) has the property that it is an analytic function of 𝐋{\bf L} with

Im(𝐁)\displaystyle\mathop{\rm Im}\nolimits({\bf B}) \displaystyle\geq 0whenIm(𝐋)0,\displaystyle 0\,\,\text{when}\,\,\mathop{\rm Im}\nolimits({\bf L})\geq 0,
𝐁 is real and 𝐁\displaystyle{\bf B}\text{ is real and }{\bf B} \displaystyle\geq 0when𝐋 is real and 𝐋0.\displaystyle 0\,\,\text{when}\,\,{\bf L}\text{ is real and }{\bf L}\leq 0. (11.1)

Hence, the Stieltjes properties of 𝐋{\bf L} as a function of ω2-\omega^{2} pass to those of 𝐁{\bf B} as a function of ω2-\omega^{2}:

Im(𝐁)\displaystyle\mathop{\rm Im}\nolimits({\bf B}) \displaystyle\geq 0whenIm(ω2)0,\displaystyle 0\,\,\text{when}\,\,\mathop{\rm Im}\nolimits(-\omega^{2})\leq 0,
𝐁 is real and 𝐁\displaystyle{\bf B}\text{ is real and }{\bf B} \displaystyle\geq 0whenω2 is real and ω20.\displaystyle 0\,\,\text{when}\,\,\omega^{2}\text{ is real and }-\omega^{2}\geq 0. (11.2)

Introducing

z=ω2cω2+c=12cω2+c,z=\frac{\omega^{2}-c}{\omega^{2}+c}=1-\frac{2c}{\omega^{2}+c}, (11.3)

for some real c>0c>0, ensures that the spectrum of 𝐁(z){\bf B}(z) is on the interval [1,1][-1,1] and

Im(𝐁(z))\displaystyle\mathop{\rm Im}\nolimits({\bf B}(z)) \displaystyle\geq 0whenIm(z)0,\displaystyle 0\,\,\text{when}\,\,\mathop{\rm Im}\nolimits(z)\geq 0,
𝐁 is real and 𝐁\displaystyle{\bf B}\text{ is real and }{\bf B} \displaystyle\geq 0whenz is real and z>1 or z<1.\displaystyle 0\,\,\text{when}\,\,z\text{ is real and }z>1\text{ or }z<-1. (11.4)

Note that this choice of zz is quite different to that in (8.16), and not restricted to two-phase composites. Thus, 𝐁(z){\bf B}(z) has the integral representation

𝐁(z)=𝐁0+11d𝐌(λ)λz,{\bf B}(z)={\bf B}_{0}+\int_{-1}^{1}\frac{d{\bf M}(\lambda)}{\lambda-z}, (11.5)

where 𝐁0{\bf B}_{0} is a positive definite operator and d𝐌(λ)d{\bf M}(\lambda) is a positive definite real operator-valued measure, satisfying the constraint

11d𝐌(λ)1λ𝐁0.\int_{-1}^{1}\frac{d{\bf M}(\lambda)}{1-\lambda}\leq{\bf B}_{0}. (11.6)

To begin, suppose we are only interested in the quadratic form (𝐁𝐄0,𝐄0)({\bf B}{\bf E}_{0},{\bf E}_{0}) associated with 𝐁{\bf B}. Then,

(𝐁(z)𝐄0,𝐄0)=k0[1+11(1λ)dη(λ)λz]=k0{1+11[1+1zλz]𝑑η(λ)},({\bf B}(z){\bf E}_{0},{\bf E}_{0})=k_{0}\left[1+\int_{-1}^{1}\frac{(1-\lambda)d\eta(\lambda)}{\lambda-z}\right]=k_{0}\left\{1+\int_{-1}^{1}\left[-1+\frac{1-z}{\lambda-z}\right]d\eta(\lambda)\right\}, (11.7)

where k0=(𝐁0𝐄0,𝐄0)k_{0}=({\bf B}_{0}{\bf E}_{0},{\bf E}_{0}) is real and positive and

dη(λ)=(d𝐌(λ)𝐄0,𝐄0)/[k0(1λ)]d\eta(\lambda)=(d{\bf M}(\lambda){\bf E}_{0},{\bf E}_{0})/[k_{0}(1-\lambda)] (11.8)

is a positive real valued measure, satisfying the constraint

11𝑑η(λ)1.\int_{-1}^{1}d\eta(\lambda)\leq 1. (11.9)

Note that k0k_{0} can be identified with (𝐁(z)𝐄0,𝐄0)({\bf B}(z){\bf E}_{0},{\bf E}_{0}) in the limit zz\to\infty, i.e. as ωic\omega\to i\sqrt{c}.

If we are interested in finding complex coefficients ξk\xi_{k}, k=1,2,mk=1,2\ldots,m, such that

(𝐁(z0)𝐄0,𝐄0)k=1mξk(𝐁(zk)𝐄0,𝐄0)\displaystyle({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0})-\sum_{k=1}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0}) (11.10)
=k0{(1k=1mξk)[111𝑑η(λ)]+11[1z0λz0k=1mξk(1zk)λzk]𝑑η(λ)}\displaystyle\quad=k_{0}\left\{(1-\sum_{k=1}^{m}\xi_{k})\left[1-\int_{-1}^{1}d\eta(\lambda)\right]+\int_{-1}^{1}\left[\frac{1-z_{0}}{\lambda-z_{0}}-\sum_{k=1}^{m}\frac{\xi_{k}(1-z_{k})}{\lambda-z_{k}}\right]d\eta(\lambda)\right\}

is small, we require that

supλ[1,1]|1λz0k=1mξk(1zk)/(1z0)λzk|ϵm.\sup_{\lambda\in[-1,1]}\left|\frac{1}{\lambda-z_{0}}-\sum_{k=1}^{m}\frac{\xi_{k}(1-z_{k})/(1-z_{0})}{\lambda-z_{k}}\right|\leq\epsilon_{m}. (11.11)

In particular, with λ=1\lambda=1, this implies

|1k=1mξk||1z0|ϵm,|1-\sum_{k=1}^{m}\xi_{k}|\leq|1-z_{0}|\epsilon_{m}, (11.12)

and so we obtain

(𝐁(z0)𝐄0,𝐄0)k=1mξk(𝐁(zk)𝐄0,𝐄0)2k0|1z0|ϵm.({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0})-\sum_{k=1}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0})\leq 2k_{0}|1-z_{0}|\epsilon_{m}. (11.13)

By setting αk=ξk(1zk)/(1z0)\alpha_{k}=\xi_{k}(1-z_{k})/(1-z_{0}) we see this is exactly the problem encountered in Section 6, and we may take the coefficients αk\alpha_{k} to be given by (7.4). The motivation for studying this problem is that the response at special frequencies can sometimes directly reveal information about the geometry. This is the case for elastodynamics in the quasistatic limit when only two materials are present. The material parameters are the bulk moduli κ1(ω)\kappa_{1}(\omega), κ2(ω)\kappa_{2}(\omega) and shear moduli μ1(ω)\mu_{1}(\omega), μ2(ω)\mu_{2}(\omega) of the two phases. It may happen that μ1(ω0)=μ2(ω0)\mu_{1}(\omega_{0})=\mu_{2}(\omega_{0}) for certain complex frequencies ω0\omega_{0} and if κ1(ω0)κ2(ω0)\kappa_{1}(\omega_{0})\neq\kappa_{2}(\omega_{0}) the response at frequency ω0\omega_{0} can reveal the volume fraction of phase 1 in a composite, or more generally in a two-phase body.

Remark 1

It is not much more difficult to treat bilinear forms. Then we have

4(𝐁(z)𝐄0,𝐄0)\displaystyle 4({\bf B}(z){\bf E}_{0},{\bf E}_{0}^{\prime}) =\displaystyle= (𝐁(z)(𝐄0+𝐄0),𝐄0+𝐄0)(𝐁(z)(𝐄0𝐄0),𝐄0𝐄0)\displaystyle({\bf B}(z)({\bf E}_{0}+{\bf E}_{0}^{\prime}),{\bf E}_{0}+{\bf E}_{0}^{\prime})-({\bf B}(z)({\bf E}_{0}-{\bf E}_{0}^{\prime}),{\bf E}_{0}-{\bf E}_{0}^{\prime}) (11.14)
=\displaystyle= k0(1){1+11[1+1zλz]𝑑η1(λ)}\displaystyle k_{0}^{(1)}\left\{1+\int_{-1}^{1}\left[-1+\frac{1-z}{\lambda-z}\right]d\eta_{1}(\lambda)\right\}
k0(2){1+11[1+1zλz]𝑑η2(λ)},\displaystyle-k_{0}^{(2)}\left\{1+\int_{-1}^{1}\left[-1+\frac{1-z}{\lambda-z}\right]d\eta_{2}(\lambda)\right\},

where

k0(1)=(𝐁0(𝐄0+𝐄0),𝐄0+𝐄0),k0(2)=(𝐁0(𝐄0𝐄0),𝐄0𝐄0)k_{0}^{(1)}=({\bf B}_{0}({\bf E}_{0}+{\bf E}_{0}^{\prime}),{\bf E}_{0}+{\bf E}_{0}^{\prime}),\quad k_{0}^{(2)}=({\bf B}_{0}({\bf E}_{0}-{\bf E}_{0}^{\prime}),{\bf E}_{0}-{\bf E}_{0}^{\prime}) (11.15)

are both real and positive, while

dη1(λ)\displaystyle d\eta_{1}(\lambda) =\displaystyle= (d𝐌(λ)(𝐄0+𝐄0),𝐄0+𝐄0)/[k0(1)(1λ)],\displaystyle(d{\bf M}(\lambda)({\bf E}_{0}+{\bf E}_{0}^{\prime}),{\bf E}_{0}+{\bf E}_{0}^{\prime})/[k_{0}^{(1)}(1-\lambda)],
dη2(λ)\displaystyle d\eta_{2}(\lambda) =\displaystyle= (d𝐌(λ)(𝐄0𝐄0),𝐄0𝐄0)/[k0(2)(1λ)]\displaystyle(d{\bf M}(\lambda)({\bf E}_{0}-{\bf E}_{0}^{\prime}),{\bf E}_{0}-{\bf E}_{0}^{\prime})/[k_{0}^{(2)}(1-\lambda)] (11.16)

are positive real valued measures, satisfying the constraints that

11𝑑η1(λ)1,11𝑑η2(λ)1.\int_{-1}^{1}d\eta_{1}(\lambda)\leq 1,\quad\int_{-1}^{1}d\eta_{2}(\lambda)\leq 1. (11.17)

We seek complex coefficients ξk\xi_{k}, k=1,2,mk=1,2\ldots,m, such that

(𝐁(z0)𝐄0,𝐄0)k=1mξk(𝐁(zk)𝐄0,𝐄0)\displaystyle({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0}^{\prime})-\sum_{k=1}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0}^{\prime}) (11.18)
=\displaystyle= k0(1){(1k=1mξk)[111𝑑η1(λ)]+11[1z0λz0k=1mξk(1zk)λzk]𝑑η1(λ)}/4\displaystyle\quad k_{0}^{(1)}\left\{(1-\sum_{k=1}^{m}\xi_{k})\left[1-\int_{-1}^{1}d\eta_{1}(\lambda)\right]+\int_{-1}^{1}\left[\frac{1-z_{0}}{\lambda-z_{0}}-\sum_{k=1}^{m}\frac{\xi_{k}(1-z_{k})}{\lambda-z_{k}}\right]d\eta_{1}(\lambda)\right\}/4
k0(2){(1k=1mξk)[111𝑑η2(λ)]+11[1z0λz0k=1mξk(1zk)λzk]𝑑η2(λ)}/4\displaystyle-k_{0}^{(2)}\left\{(1-\sum_{k=1}^{m}\xi_{k})\left[1-\int_{-1}^{1}d\eta_{2}(\lambda)\right]+\int_{-1}^{1}\left[\frac{1-z_{0}}{\lambda-z_{0}}-\sum_{k=1}^{m}\frac{\xi_{k}(1-z_{k})}{\lambda-z_{k}}\right]d\eta_{2}(\lambda)\right\}/4

is small. Using the bounds (11.11) we obtain

|(𝐁(z0)𝐄0,𝐄0)k=1mξk(𝐁(zk)𝐄0,𝐄0)|(k0(1)+k0(2))|1z0|ϵm/2.\left|({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0}^{\prime})-\sum_{k=1}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0}^{\prime})\right|\leq(k_{0}^{(1)}+k_{0}^{(2)})|1-z_{0}|\epsilon_{m}/2. (11.19)

Remark 2

Noting that

ddz(𝐁(z)𝐄0,𝐄0)=k011(1λ)dη(λ)(λz)2=k0111λz[1+1zλz]𝑑η(λ),\frac{d}{dz}({\bf B}(z){\bf E}_{0},{\bf E}_{0})=k_{0}\int_{-1}^{1}\frac{(1-\lambda)d\eta(\lambda)}{(\lambda-z)^{2}}=k_{0}\int_{-1}^{1}\frac{1}{\lambda-z}\left[-1+\frac{1-z}{\lambda-z}\right]d\eta(\lambda), (11.20)

we can easily obtain bounds that correlate this derivative at z0z_{0} with the values of (𝐁(zk)𝐄0,𝐄0)({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0}), k=0,1,2,mk=0,1,2\ldots,m. We seek complex constants γk\gamma_{k}, k=0,1,2,mk=0,1,2\ldots,m, such that

[ddz0(𝐁(z0)𝐄0,𝐄0)]k=0mξk(𝐁(zk)𝐄0,𝐄0)\displaystyle\left[\frac{d}{dz_{0}}({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0})\right]-\sum_{k=0}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0}) (11.21)
=k0{(k=0mξk)[111𝑑η(λ)]+11[1λz0+1z0(λz0)2k=0mξk(1zk)λzk]𝑑η(λ)}\displaystyle\quad=k_{0}\left\{\left(\sum_{k=0}^{m}\xi_{k}\right)\left[1-\int_{-1}^{1}d\eta(\lambda)\right]+\int_{-1}^{1}\left[-\frac{1}{\lambda-z_{0}}+\frac{1-z_{0}}{(\lambda-z_{0})^{2}}-\sum_{k=0}^{m}\frac{\xi_{k}(1-z_{k})}{\lambda-z_{k}}\right]d\eta(\lambda)\right\}

is small, and this is ensured if

supλ[1,1]|k=0mξk(1zk)/(1z0)λzk+ξ0+[1/(1z0)]λz01(λz0)2|ϵm,\sup_{\lambda\in[-1,1]}\left|\sum_{k=0}^{m}\frac{\xi_{k}(1-z_{k})/(1-z_{0})}{\lambda-z_{k}}+\frac{\xi_{0}+[1/(1-z_{0})]}{\lambda-z_{0}}-\frac{1}{(\lambda-z_{0})^{2}}\right|\leq\epsilon_{m}, (11.22)

and ϵm0\epsilon_{m}\to 0 as mm\to\infty. Observe that (11.22) with λ=1\lambda=1 implies

|k=0mξk||1z0|ϵm.\left|\sum_{k=0}^{m}\xi_{k}\right|\leq|1-z_{0}|\epsilon_{m}. (11.23)

Comparing (11.22) with (7.21) we see that we should choose

ξ0=α0[1/(1z0)],ξk=αk(1z0)/(1zk),\xi_{0}=\alpha_{0}-[1/(1-z_{0})],\quad\xi_{k}=\alpha_{k}(1-z_{0})/(1-z_{k}), (11.24)

and then, with bmb_{m} and coefficients αk\alpha_{k} given by (7.24) and (7.25), (11.22) holds with ϵm\epsilon_{m} given by (7.27). Then

|[ddz0(𝐁(z0)𝐄0,𝐄0)]k=0mξk(𝐁(zk)𝐄0,𝐄0)|2|k0||1z0|ϵm,\left|\left[\frac{d}{dz_{0}}({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0})\right]-\sum_{k=0}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0})\right|\leq 2|k_{0}||1-z_{0}|\epsilon_{m}, (11.25)

holds, and similarly one has

|[ddz0(𝐁(z0)𝐄0,𝐄0)]k=0mξk(𝐁(zk)𝐄0,𝐄0)|(k0(1)+k0(2))|1z0|ϵm.\left|\left[\frac{d}{dz_{0}}({\bf B}(z_{0}){\bf E}_{0},{\bf E}_{0}^{\prime})\right]-\sum_{k=0}^{m}\xi_{k}({\bf B}(z_{k}){\bf E}_{0},{\bf E}_{0}^{\prime})\right|\leq(k_{0}^{(1)}+k_{0}^{(2)})|1-z_{0}|\epsilon_{m}. (11.26)

The convergence of ϵm\epsilon_{m} to zero as mm\to\infty is again ensured provided for a positive constant r<Rr<R, where RR is defined by (7.7), each zkH(r)=H1(r)H2(r)H3(r)z_{k}\in H(r)=H_{1}(r)\cup H_{2}(r)\cup H_{3}(r), where the regions HiH_{i}, i=1,2,3i=1,2,3, are given by (7.13). The motivation for studying this problem is that the response may be trivial at certain frequencies ω0\omega_{0} while the derivative of the response with respect to ω\omega at ω=ω0\omega=\omega_{0} directly reveals some information about the body. This is the case for electromagnetism when only two non-magnetic materials are present (with magnetic permeabilities μ1=μ2=μ0\mu_{1}=\mu_{2}=\mu_{0} where μ0\mu_{0} is the permeability of the vacuum). It may happen that the electric permittivities of the two phases satisfy ε1(ω0)=ε2(ω0)\varepsilon_{1}(\omega_{0})=\varepsilon_{2}(\omega_{0}) for certain complex frequencies ω0\omega_{0}. At this frequency ω0\omega_{0} the body is homogeneous and its response can be easily calculated. Using perturbation theory and assuming dε1(ω0)/dω0dε2(ω0)/dω0d\varepsilon_{1}(\omega_{0})/d\omega_{0}\neq d\varepsilon_{2}(\omega_{0})/d\omega_{0} the derivative of the response with respect to ω\omega at ω=ω0\omega=\omega_{0} reveals information about the distribution of the two phases in the body.

Acknowledgements

GWM and OM are grateful to the National Science Foundation for support through the Research Grants DMS-1211359 and DMS-1814854, and DMS-2008105, respectively. MP was partially supported by a Simons Foundation collaboration grant for mathematicians.

References

  • [1] Giovanni Alessandrini and Edi Rosset. The inverse conductivity problem with one measurement: Bounds on the size of the unknown object. SIAM Journal on Applied Mathematics, 58(4):1060–1071, August 1998.
  • [2] Grégoire Allaire. Shape Optimization by the Homogenization Method, volume 146 of Applied Mathematical Sciences. Springer-Verlag, Berlin, Germany / Heidelberg, Germany / London, UK / etc., 2002.
  • [3] H. Ammari and H. Kang. Polarization and moment tensors: with applications to inverse problems and effective medium theory, volume 162. Springer Science & Business Media, New York, 2007.
  • [4] Habib Ammari and Hyeonbae Kang. Reconstruction of Small Inhomogeneities from Boundary Measurements, volume 1846 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, Germany / Heidelberg, Germany / London, UK / etc., 2004.
  • [5] M. Bałlanda. AC susceptibility studies of phase transitions and magnetic relaxation: Conventional, molecular and low-dimensional magnets. Acta Physica Polonica A, 124(6):964–976, December 2013.
  • [6] L. Baratchart, V. A. Prokhorov, and E. B. Saff. Best meromorphic approximation of Markov functions on the unit circle. Found. Comput. Math., 1(4):385–416, 2001.
  • [7] David J. Bergman. The dielectric constant of a composite material — A problem in classical physics. Physics Reports, 43(9):377–407, July 1978.
  • [8] David J. Bergman. Rigorous bounds for the complex dielectric constant of a two-component composite. Annals of Physics, 138(1):78–114, 1982.
  • [9] Martin Brühl, Martin Hanke, and Michael S. Vogelius. A direct impedance tomography algorithm for locating small inhomogeneities. Numerische Mathematik, 93(4):635–654, February 2003.
  • [10] Alberto-P. Calderón. On an inverse boundary value problem. In Seminar on Numerical Analysis and its Applications to Continuum Physics: 24 a 28 de Março 1980, volume 12 of Coleção Atas, pages 65–73. Sociedade Brasiliera de Mathemática, Rio de Janeiro, Brazil, 1980.
  • [11] Yves Capdeboscq and Michael S. Vogelius. Optimal asymptotic estimates for the volume of internal inhomogeneities in terms of multiple boundary measurements. Mathematical Modelling and Numerical Analysis = Modelisation mathématique et analyse numérique: M2ANM^{2}AN, 37(2):227–240, March/April 2003.
  • [12] A. Carini and Ornella Mattei. Variational formulations for the linear viscoelastic problem in the time domain. European Journal of Mechanics, A, Solids, 54:146–159, November/December 2015.
  • [13] Andrej V. Cherkaev. Variational Methods for Structural Optimization, volume 140 of Applied Mathematical Sciences. Springer-Verlag, Berlin, Germany / Heidelberg, Germany / London, UK / etc., 2000.
  • [14] Leonid V. Gibiansky and Graeme W. Milton. On the effective viscoelastic moduli of two-phase media. I. Rigorous bounds on the complex bulk modulus. Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 440(1908):163–188, January 1993.
  • [15] Kenneth M. Golden and George C. Papanicolaou. Bounds for effective parameters of heterogeneous media by analytic continuation. Communications in Mathematical Physics, 90(4):473–491, 1983.
  • [16] A. A. Gončar and Giermo Lopes L. Markov’s theorem for multipoint Padé approximants. Mat. Sb. (N.S.), 105(147)(4):512–524, 639, 1978.
  • [17] Yury Grabovsky. Composite Materials: Mathematical Theory and Exact Relations. IOP Publishing, Bristol, UK, 2016.
  • [18] Hermann A. Haus and James R. Melcher. Electromagnetic Fields and Energy. Prentice-Hall, Upper Saddle River, New Jersey, 1989.
  • [19] M. Ikehata. Size estimation of inclusion. Journal of Inverse and Ill-Posed Problems, 6(2):127–140, 1998.
  • [20] David Linton Johnson, Joel Koplik, and Roger Dashen. Theory of dynamic permeability and tortuosity in fluid-saturated porous media. Journal of Fluid Mechanics, 176(??):379–402, March 1987.
  • [21] Stephen M. Kajiura, Anthony D. Cornett, and Kara E. Yopak. Sensory adaptations to the environment: Electroreceptors as a case study. In Jeffrey C. Carrier and John A. Musick, editors, Sharks and Their Relatives II: Biodiversity, Adaptive Physiology, and Conservation, CRC Marine Biology Series, pages 393–433, Boca Raton, Florida, February 2010. CRC Press.
  • [22] Hyeonbae Kang, Jin Keun Seo, and Dongwoo Sheen. The inverse conductivity problem with one measurement: Stability and estimation of size. SIAM Journal on Mathematical Analysis, 28(6):1389–1405, November 1997.
  • [23] Samuel Karlin and William J. Studden. Tchebycheff systems: With applications in analysis and statistics. Pure and Applied Mathematics, Vol. XV. Interscience Publishers John Wiley & Sons, New York-London-Sydney, 1966.
  • [24] Christian Kern, Owen Miller, and Graeme W. Milton. On the range of complex effective permittivities of isotropic two-phase composites and related problems. Submitted., 2020.
  • [25] Andreas Kirsch. An Introduction to the Mathematical Theory of Inverse Problems, volume 120 of Applied Mathematical Sciences. Springer-Verlag, Berlin, Germany / Heidelberg, Germany / London, UK / etc., second edition, 2011.
  • [26] Robert V. Kohn and Michael S. Vogelius. Determining conductivity by boundary measurements. Communications on Pure and Applied Mathematics (New York), 37(3):289–298, May 1984.
  • [27] Robert V. Kohn and Michael S. Vogelius. Relaxation of a variational method for impedance computed tomography. Communications on Pure and Applied Mathematics (New York), 40(6):745–777, November 1987.
  • [28] T. Kolokolnikov and A. E. Lindsay. Recovering multiple small inclusions in a three-dimensional domain using a single measurement. Inverse Problems in Science and Engineering, 23(3):377–388, 2015.
  • [29] M. G. Kreĭn and A. A. Nudel\cprime man. The Markov moment problem and extremal problems. American Mathematical Society, Providence, R.I., 1977. Ideas and problems of P. L. Čebyšev and A. A. Markov and their further development, Translated from the Russian by D. Louvish, Translations of Mathematical Monographs, Vol. 50.
  • [30] Roderic S. Lakes and John Quackenbush. Viscoelastic behaviour in indium tin alloys over a wide range of frequency and time. Philosophical Magazine Letters, 74(4):227–232, 1996.
  • [31] Ornella Mattei. On bounding the response of linear viscoelastic composites in the time domain: The variational approach and the analytic method. Ph.D. thesis, University of Brescia, Brescia, Italy, 2016.
  • [32] Ornella Mattei and Angelo Carini. Bounds for the overall properties of composites with time-dependent constitutive law. European Journal of Mechanics, A, Solids, 61:408–419, January/February 2017.
  • [33] Ornella Mattei and G. W. Milton. Bounds for the response of viscoelastic composites under antiplane loadings in the time domain. In Milton [43], pages 149–178. See also arXiv:1602.03383 [math-ph].
  • [34] C. A. Micchelli. Characterization of Chebyshev approximation by weak Markoff systems. Computing (Arch. Elektron. Rechnen), 12(1):1–8, 1974.
  • [35] Graeme W. Milton. Bounds on the complex permittivity of a two-component composite material. Journal of Applied Physics, 52(8):5286–5293, August 1981.
  • [36] Graeme W. Milton. Bounds on the transport and optical properties of a two-component composite material. Journal of Applied Physics, 52(8):5294–5304, August 1981.
  • [37] Graeme W. Milton. The Theory of Composites, volume 6 of Cambridge Monographs on Applied and Computational Mathematics. Cambridge University Press, Cambridge, UK, 2002. Series editors: P. G. Ciarlet, A. Iserles, Robert V. Kohn, and M. H. Wright.
  • [38] Graeme W. Milton. A unifying perspective on linear continuum equations prevalent in physics. Part I: Canonical forms for static and quasistatic equations. Available as arXiv:2006.02215 [math.AP]., 2020.
  • [39] Graeme W. Milton. A unifying perspective on linear continuum equations prevalent in physics. Part II: Canonical forms for time-harmonic equations. Available as arXiv:2006.02433 [math-ph]., 2020.
  • [40] Graeme W. Milton. A unifying perspective on linear continuum equations prevalent in physics. Part III: Canonical forms for dynamic equations with moduli that may, or may not, vary with time. Available as arXiv:2006.02432 [math-ph], 2020.
  • [41] Graeme W. Milton. A unifying perspective on linear continuum equations prevalent in physics. Part IV: Canonical forms for equations involving higher order gradients. Available as arXiv:2006.03161 [math-ph]., 2020.
  • [42] Graeme W. Milton and James G. Berryman. On the effective viscoelastic moduli of two-phase media. II. Rigorous bounds on the complex shear modulus in three dimensions. Proceedings of the Royal Society A: Mathematical, Physical, & Engineering Sciences, 453(1964):1849–1880, September 1997.
  • [43] Graeme W. Milton (editor). Extending the Theory of Composites to Other Areas of Science. Milton–Patton Publishers, P.O. Box 581077, Salt Lake City, UT 85148, USA, 2016.
  • [44] Jennifer L. Mueller and Samuli Siltanen. Linear and Nonlinear Inverse Problems with Practical Applications. Computational Science & Engineering. SIAM Press, Philadelphia, 2012.
  • [45] Vasiliy A. Prokhorov. On rational approximation of Markov functions on finite sets. J. Approx. Theory, 191:94–117, 2015.
  • [46] John Sylvester. Linearizations of anisotropic inverse problems. In Lassi Päivärinta and Erkki Somersalo, editors, Inverse Problems in Mathematical Physics: Proceedings of The Lapland Conference on Inverse Problems Held at Saariselkä, Finland, 14–20 June 1992, volume 422 of Lecture Notes in Physics, pages 231–241, Berlin, Germany / Heidelberg, Germany / London, UK / etc., 1993. Springer-Verlag.
  • [47] John Sylvester and Gunther Uhlmann. A global uniqueness theorem for an inverse boundary value problem. Ann. of Math. (2), 125(1):153–169, 1987.
  • [48] Luc Tartar. The General Theory of Homogenization: a Personalized Introduction, volume 7 of Lecture Notes of the Unione Matematica Italiana. Springer-Verlag, Berlin, Germany / Heidelberg, Germany / London, UK / etc., 2009.
  • [49] Salvatore Torquato. Random Heterogeneous Materials: Microstructure and Macroscopic Properties, volume 16 of Interdisciplinary Applied Mathematics. Springer-Verlag, Berlin, Germany / Heidelberg, Germany / London, UK / etc., 2002.
  • [50] G. von der Emde. Non-visual environmental imaging and object detection through active electrolocation in weakly electric fish. Journal of Comparative Physiology A, 192:601–612, 2006.
  • [51] J. L. Walsh. On interpolation and approximation by rational functions with preassigned poles. Trans. Amer. Math. Soc., 34(1):22–74, 1932.
  • [52] J. L. Walsh. Interpolation and approximation by rational functions in the complex domain. Fourth edition. American Mathematical Society Colloquium Publications, Vol. XX. American Mathematical Society, Providence, R.I., 1965.
  • [53] Aaron T. Welters, Yehuda Avniel, and Steven G. Johnson. Speed-of-light limitations in passive linear media. Physical Review A (Atomic, Molecular, and Optical Physics), 90(2):023847, August 2014.
  • [54] Helmut Werner. Die konstruktive Ermittlung der Tschebyscheff-Approximierenden im Bereich der rationalen Funktionen. Arch. Rational Mech. Anal., 11:368–384, 1962.
  • [55] Helmut Werner. Ein Satz über diskrete Tschebyscheff-Approximation bei gebrochen linearen Funktionen. Numer. Math., 4:154–157, 1962.
  • [56] Helmut Werner. Tschebyscheff-Approximation im Bereich der rationalen Funktionen bei Vorliegen einer guten Ausgangsnäherung. Arch. Rational Mech. Anal., 10:205–219, 1962.
  • [57] John R. Willis. Variational and related methods for the overall properties of composites. Advances in Applied Mechanics, 21:1–78, 1981.