Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

MSC 80M40; 80A20, 35B99; 35K20; 35Q41; 35R30;74A30; 74G75

PACS 65.80.-g

8.1 Introduction and Results

In this paper the problem of heat transfer in a complex medium consisting of many small impedance particles of an arbitrary shape is solved. Equation for the effective limiting temperature is derived when the characteristic size a of the particles tends to zero while their number tends to infinity at a suitable rate while the distance d between closest neighboring particles is much larger than a, d > > a.

These results are used for developing a method for creating materials in which heat is transmitted along a line. Thus, the information can be transmitted by a heat signals.

The contents of this paper is based on the earlier papers of the author cited in the bibliography, especially [13, 16, 17].

Let many small bodies (particles) D m , 1 ≤ m ≤ M, of an arbitrary shape be distributed in a bounded domain \(D \subset \mathbb{R}^{3}\), diamD m  = 2a, and the boundary of D m is denoted by \(\mathcal{S}_{m}\) and is assumed twice continuously differentiable. The small bodies are distributed according to the law

$$\displaystyle{ \mathcal{N}(\Delta ) = \frac{1} {a^{2-\kappa }}\int _{\Delta }N(x)dx[1 + o(1)],\quad a \rightarrow 0. }$$
(8.1)

Here \(\Delta \subset D\) is an arbitrary open subdomain of D, κ ∈ [0, 1) is a constant, N(x) ≥ 0 is a continuous function, and \(\mathcal{N}(\Delta )\) is the number of the small bodies D m in \(\Delta \). The heat equation can be stated as follows:

$$\displaystyle\begin{array}{rcl} & u_{t} = \nabla ^{2}u + f(x)\,\mbox{ in }\,\mathbb{R}^{3}\setminus \bigcup _{m=1}^{M}D_{ m},:= \Omega,\quad u\vert _{t=0} = 0,&{}\end{array}$$
(8.2)
$$\displaystyle\begin{array}{rcl} & u_{N} =\zeta _{m}u\mbox{ on }\mathcal{S}_{m},\quad 1 \leq m \leq M,\qquad Re\zeta _{m} \geq 0.&{}\end{array}$$
(8.3)

Here N is the outer unit normal to \(\mathcal{S}\),

$$\displaystyle{\mathcal{S}:=\bigcup _{ m=1}^{M}\mathcal{S}_{ m},\quad \zeta _{m} = \frac{h(x_{m})} {a^{\kappa }},\quad x_{m} \in D_{m},\quad 1 \leq m \leq M,}$$

and h(x) is a continuous function in D, Reh ≥ 0.

Denote

$$\displaystyle{\mathcal{U}:= \mathcal{U}(x,\lambda ) =\int _{ 0}^{\infty }e^{-\lambda t}u(x,t)dt.}$$

Then, taking the Laplace transform of Eqs. (8.2)–(8.3) one gets:

$$\displaystyle\begin{array}{rcl} & -\nabla ^{2}\mathcal{U} +\lambda \mathcal{U} =\lambda ^{-1}f(x)\mbox{ in }\Omega,&{}\end{array}$$
(8.4)
$$\displaystyle\begin{array}{rcl} & \mathcal{U}_{N} =\zeta _{m}\mathcal{U}\mbox{ on }\mathcal{S}_{m},1 \leq m \leq M.&{}\end{array}$$
(8.5)

Let

$$\displaystyle\begin{array}{rcl} & g(x,y):= g(x,y,\lambda ):= \frac{e^{-\sqrt{\lambda }\vert x-y\vert }} {4\pi \vert x-y\vert },&{}\end{array}$$
(8.6)
$$\displaystyle\begin{array}{rcl} & F(x,\lambda ):= \frac{1} {\lambda } \int _{\mathbb{R}^{3}}g(x,y)f(y)dy.&{}\end{array}$$
(8.7)

Look for the solution to (8.4)–(8.5) of the form

$$\displaystyle{ \mathcal{U}(x,\lambda ) = F(x,\lambda ) +\sum _{ m=1}^{M}\int _{ \mathcal{S}_{m}}g(x,s)\sigma _{m}(s)ds, }$$
(8.8)

where

$$\displaystyle{ \mathcal{U}(x,\lambda ):= \mathcal{U}(x):= \mathcal{U}, }$$
(8.9)

and \(\mathcal{U}(x)\) depends on \(\lambda\).

The functions \(\sigma _{m}\) are unknown and should be found from the boundary conditions (8.5). Equation (8.4) is satisfied by \(\mathcal{U}\) of the form (8.8) with arbitrary continuous \(\sigma _{m}\). To satisfy the boundary condition (8.5) one has to solve the following equation obtained from the boundary condition (8.5):

$$\displaystyle{ \frac{\partial \mathcal{U}_{e}(x)} {\partial N} + \frac{A_{m}\sigma _{m} -\sigma _{m}} {2} -\zeta _{m}\mathcal{U}_{e} -\zeta _{m}T_{m}\sigma _{m} = 0\mbox{ on }\mathcal{S}_{m},\quad 1 \leq m \leq M, }$$
(8.10)

where the effective field \(\mathcal{U}_{e}(x)\) is defined by the formula:

$$\displaystyle{ \mathcal{U}_{e}(x):= \mathcal{U}_{e,m}(x):= \mathcal{U}(x) -\int _{\mathcal{S}_{m}}g(x,s)\sigma _{m}(s)ds, }$$
(8.11)

the operator T m is defined by the formula:

$$\displaystyle{ T_{m}\sigma _{m} =\int _{\mathcal{S}_{m}}g(s,s')\sigma _{m}(s')ds', }$$
(8.12)

and A m is:

$$\displaystyle{ A_{m}\sigma _{m} = 2\int _{\mathcal{S}_{m}}\frac{\partial g(s,s')} {\partial N_{s}} \sigma _{m}(s')ds'. }$$
(8.13)

In deriving Eq. (8.10) we have used the known formula for the outer limiting value on \(\mathcal{S}_{m}\) of the normal derivative of a simple layer potential.

We now apply the ideas and methods for solving many-body scattering problems developed in [1215].

Let us call \(\mathcal{U}_{e,m}\) the effective (self-consistent) value of \(\mathcal{U}\), acting on the m-th body. As \(a \rightarrow 0\), the dependence on m disappears, since

$$\displaystyle{\int _{\mathcal{S}_{m}}g(x,s)\sigma _{m}(s)ds \rightarrow 0\mbox{ as }a \rightarrow 0.}$$

One has

$$\displaystyle{ \mathcal{U}(x,\lambda ) = F(x,\lambda ) +\sum _{ m=1}^{M}g(x,x_{ m})Q_{m} + \mathcal{J}_{2},\quad x_{m} \in D_{m}, }$$
(8.14)

where

$$\displaystyle\begin{array}{rcl} & Q_{m}:=\int _{\mathcal{S}_{m}}\sigma _{m}(s)ds, & \\ & \mathcal{J}_{2}:=\sum _{ m=1}^{M}\int _{\mathcal{S}_{m}}[g(x,s') - g(x,x_{m})]\sigma _{m}(s')ds'.&{}\end{array}$$
(8.15)

Define

$$\displaystyle{ \mathcal{J}_{1}:=\sum _{ m=1}^{M}g(x,x_{ m})Q_{m}. }$$
(8.16)

We prove in Lemma 3, Sect. 8.4 (see also [13, 16]) that

$$\displaystyle{ \vert \mathcal{J}_{2}\vert << \vert \mathcal{J}_{1}\vert \mbox{ as }a \rightarrow 0 }$$
(8.17)

provided that

$$\displaystyle{ \lim _{a\rightarrow 0} \frac{a} {d(a)} = 0, }$$
(8.18)

where d(a) = d is the minimal distance between neighboring particles.

If (8.17) holds, then problem (8.4)–(8.5) is solved asymptotically by the formula

$$\displaystyle{ \mathcal{U}(x,\lambda ) = F(x,\lambda ) +\sum _{ m=1}^{M}g(x,x_{ m})Q_{m},\quad a \rightarrow 0, }$$
(8.19)

provided that asymptotic formulas for Q m , as \(a \rightarrow 0\), are found.

To find formulas for Q m , let us integrate (8.10) over \(\mathcal{S}_{m}\), estimate the order of the terms in the resulting equation as \(a \rightarrow 0\), and keep the main terms, that is, neglect the terms of higher order of smallness as a → 0.

We get

$$\displaystyle{ \int _{\mathcal{S}_{m}}\frac{\partial \mathcal{U}_{e}} {\partial N} ds =\int _{D_{m}}\nabla ^{2}\mathcal{U}_{ e}dx = O(a^{3}). }$$
(8.20)

Here we assumed that \(\vert \nabla ^{2}\mathcal{U}_{e}\vert = O(1),a \rightarrow 0\). This assumption is valid since \(\mathcal{U} =\lim _{a\rightarrow 0}\mathcal{U}_{e}\) is smooth as a solution to an elliptic equation. One has

$$\displaystyle{ \int _{\mathcal{S}_{m}}\frac{A_{m}\sigma _{m} -\sigma _{m}} {2} ds = -Q_{m}[1 + o(1)],\,a \rightarrow 0. }$$
(8.21)

This relation is proved in Lemma 2, Sect. 8.4, see also [13]. Furthermore,

$$\displaystyle{ -\zeta _{m}\int _{\mathcal{S}_{m}}\mathcal{U}_{e}ds = -\zeta _{m}\vert \mathcal{S}_{m}\vert \mathcal{U}_{e}(x_{m}) = O(a^{2-\kappa }),\quad a \rightarrow 0, }$$
(8.22)

where \(\vert \mathcal{S}_{m}\vert = O(a^{2})\) is the surface area of \(\mathcal{S}_{m}\). Finally,

$$\displaystyle\begin{array}{rcl} & -\zeta _{m}\int _{\mathcal{S}_{m}}ds\int _{\mathcal{S}_{m}}g(s,s')\sigma _{m}(s')ds' = -\zeta _{m}\int _{\mathcal{S}_{m}}ds'\sigma _{m}(s')\int _{\mathcal{S}_{m}}dsg(s,s')& \\ & = Q_{m}O(a^{1-\kappa }),\qquad a \rightarrow 0. &{}\end{array}$$
(8.23)

Thus, the main term of the asymptotic of Q m , as a → 0, is

$$\displaystyle{ Q_{m} = -\zeta _{m}\vert \mathcal{S}_{m}\vert \mathcal{U}_{e}(x_{m}). }$$
(8.24)

Formulas (8.24) and (8.19) yield

$$\displaystyle{ \mathcal{U}(x,\lambda ) = F(x,\lambda ) -\sum _{m=1}^{M}g(x,x_{ m})\zeta _{m}\vert \mathcal{S}_{m}\vert \mathcal{U}_{e}(x_{m},\lambda ), }$$
(8.25)

and

$$\displaystyle{ \mathcal{U}_{e}(x_{m},\lambda ) = F(x_{m},\lambda ) -\sum _{m'\neq m,m'=1}^{M}g(x_{ m},x_{m'})\zeta _{m'}\vert \mathcal{S}_{m'}\vert \mathcal{U}_{e}(x_{m'},\lambda ). }$$
(8.26)

Denote

$$\displaystyle{\mathcal{U}_{e}(x_{m},\lambda ):= \mathcal{U}_{m},\quad F(x_{m},\lambda ):= F_{m},\quad g(x_{m},x_{m'}):= g_{mm'},}$$

and write (8.26) as a linear algebraic system for \(\mathcal{U}_{m}\):

$$\displaystyle{ \mathcal{U}_{m} = F_{m} - a^{2-\kappa }\sum _{ m'\neq m}g_{mm'}h_{m'}c_{m'}\mathcal{U}_{m'},\quad 1 \leq m \leq M, }$$
(8.27)

where h m = h(x m), \(\zeta _{m'} = \frac{h_{m'}} {a^{\kappa }}\), \(c_{m'}:= \vert S_{m'}\vert a^{-2}\).

Consider a partition of the bounded domain D, in which the small bodies are distributed, into a union of P < < M small nonintersecting cubes \(\Delta _{p}\), 1 ≤ p ≤ P, of side b,

$$\displaystyle{b >> d,\quad b = b(a) \rightarrow 0\quad \text{as}\,\,a \rightarrow 0\quad \lim _{a\rightarrow 0}\frac{d(a)} {b(a)} = 0.}$$

Let \(x_{p} \in \Delta _{p}\), \(\vert \Delta _{p}\vert =\) volume of \(\Delta _{p}\). One has

$$\displaystyle\begin{array}{rcl} & a^{2-\kappa }\sum _{m'=1,m'\neq m}^{M}g_{mm'}h_{m'}c_{m'}\mathcal{U}_{m'} = a^{2-\kappa }\sum _{p'=1,p'\neq p}^{P}g_{pp'}h_{p'}c_{p'}\mathcal{U}_{p'}\sum _{x_{m'}\in \Delta _{p'}}1 =& \\ & =\sum _{p'\neq p}g_{pp'}h_{p'}c_{p'}\mathcal{U}_{p'}N(x_{p'})\vert \Delta _{p'}\vert [1 + o(1)],\quad a \rightarrow 0. &{}\end{array}$$
(8.28)

Thus, (8.27) yields a linear algebraic system (LAS) of order P < < M for the unknowns \(\mathcal{U}_{p}\):

$$\displaystyle{ \mathcal{U}_{p} = F_{p} -\sum _{p'\neq p,p'=1}^{P}g_{ pp'}h_{p'}c_{p'}N_{p'}\mathcal{U}_{p'}\vert \Delta _{p'}\vert,\quad 1 \leq p \leq P. }$$
(8.29)

Since P < < M, the order of the original LAS (8.27) is drastically reduced. This is crucial when the number of particles tends to infinity and their size a tends to zero. We have assumed that

$$\displaystyle{ h_{m'} = h_{p'}[1 + o(1)],\quad c_{m'} = c_{p'}[1 + o(1)],\quad \mathcal{U}_{m'} = \mathcal{U}_{p'}[1 + o(1)],\,\,a \rightarrow 0,\quad }$$
(8.30)

for \(x_{m'} \in \Delta _{p'}\). This assumption is justified, for example, if the functions h(x), \(\mathcal{U}(x,\lambda )\),

$$\displaystyle{c(x) =\lim _{x_{m'}\in \Delta _{x},a\rightarrow 0}\frac{\vert S_{m'}\vert } {a^{2}},}$$

and N(x) are continuous, but these assumptions can be relaxed.

The continuity of the \(\mathcal{U}(x,\lambda )\) is a consequence of the fact that this function satisfies elliptic equation, and the continuity of c(x) is assumed. If all the small bodies are identical, then \(c(x) = c = const\), so in this case the function c(x) is certainly continuous.

The sum in the right-hand side of (8.29) is the Riemannian sum for the integral

$$\displaystyle\begin{array}{rcl} & lim_{a\rightarrow 0}\sum _{p'=1,p'\neq p}^{P}g_{pp'}h_{p'}c_{p'}N(x_{p'})\mathcal{U}_{p'}\vert \Delta _{p}'\vert =& {}\\ & \int _{D}g(x,y)h(y)c(y)N(y)\mathcal{U}(y,\lambda )dy & {}\\ \end{array}$$

Therefore, linear algebraic system (8.29) is a collocation method for solving integral equation

$$\displaystyle{ \mathcal{U}(x,\lambda ) = F(x,\lambda ) -\int _{D}g(x,y)c(y)h(y)N(y)\mathcal{U}(y,\lambda )dy. }$$
(8.31)

Convergence of this method for solving equations with weakly singular kernels is proved in [10], see also [11, 20].

Applying the operator \(-\nabla ^{2}+\lambda\) to Eq. (8.31) one gets an elliptic differential equation:

$$\displaystyle{ (-\Delta +\lambda )\mathcal{U}(x,\lambda ) = \frac{f(x)} {\lambda } - c(x)h(x)N(x)\mathcal{U}(x,\lambda ). }$$
(8.32)

Taking the inverse Laplace transform of this equation yields

$$\displaystyle{ u_{t} = \Delta u + f(x) - q(x)u,\quad q(x):= c(x)h(x)N(x). }$$
(8.33)

Therefore, the limiting equation for the temperature contains the term q(x)u. Thus, the embedding of many small particles creates a distribution of source and sink terms in the medium, the distribution of which is described by the term q(x)u.

If one solves Eq. (8.31) for \(\mathcal{U}(x,\lambda )\), or linear algebraic system (8.29) for \(\mathcal{U}_{p}(\lambda )\), then one can Laplace-invert \(\mathcal{U}(x,\lambda )\) for \(\mathcal{U}(x,t)\). Numerical methods for Laplace inversion from the real axis are discussed in [4, 19].

If one is interested only in the average temperature, one can use the relation

$$\displaystyle{ \lim _{T\rightarrow \infty }\frac{1} {T}\int _{0}^{T}u(x,t)dt =\lim _{\lambda \rightarrow 0}\lambda \mathcal{U}(x,\lambda ). }$$
(8.34)

Relation (8.34) is proved in Lemma 1, Sect. 8.4. It holds if the limit on one of its sides exists. The limit on the right-hand side of (8.34) let us denote by ψ(x). From Eqs. (8.7) and (8.31) it follows that ψ satisfies the equation

$$\displaystyle{\psi =\varphi -B\varphi,}$$

where

$$\displaystyle\begin{array}{rcl} & \varphi:=\int _{\Omega }g_{0}(x,y)f(y)dy, & {}\\ & g_{0}(x,y):= \frac{1} {4\pi \vert x-y\vert }, & {}\\ & B\psi:=\int _{\Omega }g_{0}(x,y)q(y)\psi (y)dy,& {}\\ \end{array}$$

and

$$\displaystyle{q(x):= c(x)h(x)N(x).}$$

The function ψ can be calculated by the formula

$$\displaystyle{ \psi (x) = (I + B)^{-1}\varphi. }$$
(8.35)

From the physical point of view the function h(x) is non-negative because the flux −∇u of the heat flow is proportional to the temperature u and is directed along the outer normal N: \(-u_{N} = h_{1}u\), where \(h_{1} = -h < 0\). Thus, q ≥ 0.

It is proved in [5, 6] that zero is not an eigenvalue of the operator \(-\nabla ^{2} + q(x)\) provided that q(x) ≥ 0 and

$$\displaystyle{q = O\big( \frac{1} {\vert x\vert ^{2+\epsilon }}\big),\quad \vert x\vert \rightarrow \infty,}$$

and ε > 0.

In our case, q(x) = 0 outside of the bounded region D, so the operator \((I + B)^{-1}\) exists and is bounded in C(D).

Let us formulate our basic result.

Theorem 1

Assume (8.1)(8.18), and h ≥ 0. Then, there exists the limit \(\mathcal{U}(x,\lambda )\) of \(\mathcal{U}_{e}(x,\lambda )\) as \(a \rightarrow 0\), \(\mathcal{U}(x,\lambda )\) solves Eq. (8.31), and there exists the limit (8.34), where ψ(x) is given by formula (8.35).

Methods of our proof of Theorem 1 are quite different from the proof of homogenization theory results in [1, 3].

The author’s plenary talk at Chaos-2015 Conference was published in [18].

8.2 Creating Materials Which Allows One to Transmit Heat Signals Along a Line

In applications it is of interest to have materials in which heat propagates along a line and decays fast in all the directions orthogonal to this line.

In this section a construction of such material is given. We follow [17] with some simplifications.

The idea is to create first the medium in which the heat transfer is governed by the equation

$$\displaystyle{ u_{t} = \Delta u - q(x)u\quad \text{in }\ D,\quad u\vert _{S} = 0,\quad u\vert _{t=0} = f(x), }$$
(8.36)

where D is a bounded domain with a piece-wise smooth boundary S, D = D 0 × [0, L], \(D_{0} \subset \mathbb{R}^{2}\) is a smooth domain orthogonal to the axis x 1, x = (x 1, x 2, x 3), x 2, x 3 ∈ D 0, 0 ≤ x 1 ≤ L.

Such a medium is created by embedding many small impedance particles into a given domain D filled with a homogeneous material. A detailed argument, given in Sect. 8.1 (see also [13, 16]), yields the following result.

Assume that in every open subset \(\Delta \) of D the number of small particles is defined by the formula:

$$\displaystyle{ \mathcal{N}(\Delta ) = \frac{1} {a^{2-\kappa }}\int _{\Delta }N(x)dx[1 + o(1)],\quad a \rightarrow 0, }$$
(8.37)

where a > 0 is the characteristic size of a small particle, κ ∈ [0, 1) is a given number and N(x) ≥ 0 is a continuous in D function.

Assume also that on the surface S m of the m-th particle D m the impedance boundary condition holds. Here

$$\displaystyle{1 \leq m \leq M = \mathcal{N}(D) = O\left ( \frac{1} {a^{2-\kappa }}\right ),\quad a \rightarrow 0,}$$

and the impedance boundary conditions are:

$$\displaystyle{ u_{N} =\zeta _{m}u\quad \text{on }\ S_{m},\quad \text{Re}\zeta _{m} \geq 0, }$$
(8.38)

where

$$\displaystyle{\zeta _{m}:= \frac{h(x_{m})} {a^{\kappa }} }$$

is the boundary impedance, x m  ∈ D m is an arbitrary point (since D m is small the position of x m in D m is not important), κ is the same parameter as in (8.37) and h(x) is a continuous in D function, Reh ≥ 0, N is the unit normal to S m pointing out of D m . The functions h(x), N(x) and the number κ can be chosen as the experimenter wishes.

It is proved in Sect. 8.1 (see also [13, 16]) that, as a → 0, the solution of the problem

$$\displaystyle\begin{array}{rcl} & u_{t} = \Delta u\quad \text{in }D\setminus \bigcup _{m=1}^{M}D_{ m},\,u_{N} =\zeta _{m}u\quad \text{on }\ S_{m},\,1 \leq m \leq M,\,&{}\end{array}$$
(8.39)
$$\displaystyle\begin{array}{rcl} & u\vert _{S} = 0,&{}\end{array}$$
(8.40)

and

$$\displaystyle{ u\vert _{t=0} = f(x), }$$
(8.41)

has a limit u(x, t). This limit solves problem (8.36) with

$$\displaystyle{ q(x) = c_{S}N(x)h(x), }$$
(8.42)

where

$$\displaystyle{ \quad c_{S}:= \frac{\vert S_{m}\vert } {a^{2}} = const, }$$
(8.43)

and | S m  | is the surface area of S m . By assuming that c S is a constant, we assume, for simplicity only, that the small particles are identical in shape, see [13].

Since N(x) ≥ 0 is an arbitrary continuous function and h(x), Reh ≥ 0, is an arbitrary continuous function, and both functions can be chosen by experimenter as he/she wishes, it is clear that an arbitrary real-valued potential q can be obtained by formula (8.42).

Suppose that

$$\displaystyle{ (-\Delta + q(x))\phi (x) =\lambda _{n}\phi _{n},\quad \phi _{n}\vert _{S} = 0,\quad \vert \vert \phi _{n}\vert \vert _{L^{2}(D)} = \vert \vert \phi _{n}\vert \vert = 1, }$$
(8.44)

where {ϕ n } is an orthonormal basis of L 2(D): = H. Then the unique solution to (8.36) is

$$\displaystyle{ u(x,t) =\sum _{ n=1}^{\infty }e^{-\lambda _{n}t}(f,\phi _{ n})\phi _{n}(x). }$$
(8.45)

If q(x) is such that \(\lambda _{1} = 0\), \(\lambda _{2} \gg 1\), and \(\lambda _{2} \leq \lambda _{3} \leq \ldots\), then, as \(t \rightarrow \infty \), the series (8.45) is well approximated by its first term

$$\displaystyle{ u(x,t) = (f,\phi _{1})\phi _{1} + O(e^{-10t}),\quad t \rightarrow \infty. }$$
(8.46)

If \(\lambda _{1} > 0\) is very small, then the main term of the solution is

$$\displaystyle{u(x,t) = (f,\phi _{1})\phi _{1}e^{-\lambda _{1}t} + O(e^{-10t})}$$

as \(t \rightarrow \infty \). The term \(e^{-\lambda _{1}t} \sim 1\) if \(t << \frac{1} {\lambda _{1}}\).

Thus, our problem is solved if q(x) has the following property:

$$\displaystyle{ \vert \phi _{1}(x)\vert \mbox{decays as $\rho$ grows},\quad \rho = (x_{2}^{2} + x_{ 3}^{2})^{1/2}. }$$
(8.47)

Since the eigenfunction is normalized, | | ϕ 1 | |  = 1, this function will not tend to zero in a neighborhood of the line ρ = 0, so information can be transformed by the heat signals along the line ρ = 0, that is, along s−axis. Here we use the cylindrical coordinates:

$$\displaystyle{x = (x_{1},x_{2},x_{3}) = (s,\rho,\theta ),\quad s = x_{1},\quad \rho = (x_{2}^{2} + x_{ 3}^{2})^{1/2}.}$$

In Sect. 8.3 the domain D 0 is a disc and the potential q(x) does not depend on \(\theta\).

The technical part of solving our problem consists of the construction of q(x) = c S N(x)h(x) such that

$$\displaystyle{ \lambda _{1} = 0,\quad \lambda _{2} \gg 1;\quad \vert \phi _{1}(x)\vert \ \mbox{ decays as $\rho $ grows}. }$$
(8.48)

Since the function N(x) ≥ 0 and h(x), Reh ≥ 0, are at our disposal, any desirable q, Re q ≥ 0, can be obtained by embedding many small impedance particles in a given domain D. In Sect. 8.3, a potential q with the desired properties is constructed. This construction allows one to transform information along a straight line using heat signals.

8.3 Construction of q(x)

Let

$$\displaystyle{q(x) = p(\rho ) + Q(s),}$$

where s: = x 1, \(\rho:= (x_{2}^{2} + x_{3}^{2})^{1/2}\). Then the solution to problem (8.44) is u = v(ρ)w(s), where

$$\displaystyle\begin{array}{rcl} - v_{m}'' -\rho ^{-1}v_{ m}' + p(\rho )v_{m} =\mu _{m}v_{m},\quad 0& \leq & \rho \leq R, \\ & & \vert v_{m}(0)\vert < \infty,\quad v_{m}(R) = 0,\quad {}\end{array}$$
(8.49)

and

$$\displaystyle\begin{array}{rcl} & & -w_{l}'' + Q(s)w_{l} =\nu _{l}w_{l},\quad 0 \leq s \leq L, \\ & & \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad w_{l}(0) = 0,\quad w_{l}(L) = 0.\quad {}\end{array}$$
(8.50)

One has

$$\displaystyle{ \lambda _{n} =\mu _{m} +\nu _{l},\quad n = n(m,l). }$$
(8.51)

Our task is to find a potential Q(s) such that ν 1 = 0, ν 2 ≫ 1 and a potential p(ρ) such that μ 1 = 0, μ 2 ≫ 1 and | v m (ρ) | decays as ρ grows.

It is known how to construct q(s) with the desired properties: the Gel’fand-Levitan method allows one to do this, see [7]. Let us recall this construction. One has ν l0 = l 2, where we set L = π and denote by ν l0 the eigenvalues of the problem (8.50) with Q(s) = 0. Let the eigenvalues of the operator (8.50) with Q ≠ 0 be \(\nu _{1} = 0,\nu _{2} = 11,\nu _{3} = 14,\nu _{l} =\nu _{l0}\) for l ≥ 4.

The kernel L(x, y) in the Gel’fand-Levitan theory is defined as follows:

$$\displaystyle{L(x,y) =\int _{ -\infty }^{\infty }\frac{\sin (\sqrt{\lambda }x)} {\sqrt{\lambda }} \frac{\sin (\sqrt{\lambda }y)} {\sqrt{\lambda }} d(\varrho (\lambda ) -\varrho _{0}(\lambda )),}$$

where \(\varrho (\lambda )\) is the spectral function of the operator (8.50) with the potential Q = Q(s), and \(\varrho _{0}(\lambda )\) is the spectral function of the operator (8.50) with the potential Q = 0 and the same boundary conditions as for the operator with Q ≠ 0.

Due to our choice of ν l and the normalizing constants α j , namely: \(\alpha _{j} = \frac{\pi } {2}\) for j ≥ 2 and \(\alpha _{1} = \frac{\pi ^{3}} {3}\), the kernel L(x, y) is given explicitly by the formula:

$$\displaystyle\begin{array}{rcl} L(x,y) = \frac{3xy} {\pi ^{3}} & +& \frac{2} {\pi } \Big(\frac{\sin (\sqrt{\nu _{2}}x)} {\sqrt{\nu _{2}}} \frac{\sin (\sqrt{\nu _{2}}y)} {\sqrt{\nu _{2}}} + \frac{\sin (\sqrt{\nu _{3}}x)} {\sqrt{\nu _{3}}} \frac{\sin (\sqrt{\nu _{3}}y)} {\sqrt{\nu _{3}}} \Big) - \\ & &-\frac{2} {\pi } \Big(\sin x\sin y +\sin (2x)\sin (2y) +\sin (3x)\sin (3y)\Big),{}\end{array}$$
(8.52)

where ν 1 = 0, ν 2 = 11 and ν 3 = 14. This is a finite rank kernel. The term xy is the value of the function \(\frac{\sin \nu x} {\nu } \frac{\sin \nu y} {\nu }\) at ν = 0, and the corresponding normalizing constant is \(\frac{\pi ^{3}} {3} = \vert \vert x\vert \vert ^{2} =\int _{ 0}^{\pi }x^{2}dx\).

Solve the Gel’fand-Levitan equation:

$$\displaystyle{ K(s,\tau ) +\int _{ 0}^{s}K(s,s')L(s',\tau )ds' = -L(s,\tau ),\quad 0 \leq \tau \leq s, }$$
(8.53)

which is uniquely solvable (see [7]). Since Eq. (8.53) has finite-rank kernel it can be solved analytically being equivalent to a linear algebraic system.

If the function K(s, τ) is found, then the potential Q(s) is computed by the formula [2, 7]:

$$\displaystyle{ Q(s) = 2\frac{dK(s,s)} {ds}, }$$
(8.54)

and this Q(s) has the required properties: \(\nu _{1} = 0,\nu _{2} \gg 1,\nu _{l} \leq \nu _{l+1}\).

Consider now the operator (8.49) for v(ρ). Our problem is to calculate p(ρ) which has the required properties:

$$\displaystyle{\mu _{1} = 0,\quad \mu _{2} \gg 1,\quad \mu _{m} \leq \mu _{m+1},}$$

and | ϕ m (ρ) | decays as ρ grows.

We reduce this problem to the previous one that was solved above. To do this, set \(v = \frac{\psi } {\sqrt{\rho }}\). Then equation

$$\displaystyle{-v'' -\frac{1} {\rho } v' + p(\rho )v =\mu v,}$$

is transformed to the equation

$$\displaystyle{ -\psi ''- \frac{1} {4\rho ^{2}}\psi + p(\rho )\psi =\mu \psi. }$$
(8.55)

Let

$$\displaystyle{ p(\rho ) = \frac{1} {4\rho ^{2}} + Q(\rho ), }$$
(8.56)

where Q(ρ) is constructed above. Then Eq. (8.55) becomes

$$\displaystyle{ -\psi '' + Q(\rho )\psi =\mu \psi, }$$
(8.57)

and the boundary conditions are:

$$\displaystyle{ \psi (R) = 0,\quad \psi (0) = 0. }$$
(8.58)

The problem (8.57)–(8.58) has the desired eigenvalues \(\mu _{1} = 0,\mu _{2} \gg 1,\mu _{m} \leq \mu _{m+1}\).

The eigenfunction

$$\displaystyle{\phi _{1}(x) = v_{1}(\rho )w_{1}(s),}$$

where \(v_{1}(\rho ) = \frac{\psi _{1}(\rho )} {\sqrt{\rho }}\), decays as ρ grows, and the eigenvalues \(\lambda _{n}\) can be calculated by the formula:

$$\displaystyle{\lambda _{n} =\mu _{m} +\nu _{l},\quad m,l \geq 1,\quad n = n(m,l).}$$

Since \(\mu _{1} =\nu _{1} = 0\) one has \(\lambda _{1} = 0\). Since ν 2 = 11 and μ 2 = 11, one has \(\lambda _{2} = 11 \gg 1\).

Thus, the desired potential is constructed:

$$\displaystyle{q(x) = Q(s) + ( \frac{1} {4\rho ^{2}} + Q(\rho )),}$$

where Q(s) is given by formula (8.54).

This concludes the description of our procedure for the construction of q.

Remark 1

It is known (see, for example, [2]) that the normalizing constants

$$\displaystyle{\alpha _{j}:=\int _{ 0}^{\pi }\varphi _{ j}^{2}(s)ds}$$

and the eigenvalues \(\lambda _{j}\), defined by the differential equation

$$\displaystyle{-\frac{d^{2}\varphi _{j}} {ds^{2}} + Q(s)\varphi _{j} =\lambda _{j}\varphi _{j},}$$

the boundary conditions

$$\displaystyle{\varphi _{j}'(0) = 0,\quad \varphi _{j}'(\pi ) = 0,}$$

and the normalizing condition \(\varphi _{j}(0) = 1\), have the following asymptotic:

$$\displaystyle{\alpha _{j} = \frac{\pi } {2} + O( \frac{1} {j^{2}})\quad \text{as }\ j \rightarrow \infty,}$$

and

$$\displaystyle{\sqrt{\lambda _{j}} = j + O(\frac{1} {j})\quad \text{as }\ j \rightarrow \infty.}$$

The differential equation

$$\displaystyle{-\Psi _{j}^{''} + Q(s)\Psi _{ j} =\nu _{j}\Psi _{j},}$$

the boundary condition

$$\displaystyle{\Psi _{j}(0) = 0,\quad \Psi _{j}(\pi ) = 0,}$$

and the normalizing condition \(\Psi _{j}'(0) = 1\) imply

$$\displaystyle\begin{array}{rcl} & \sqrt{\lambda _{j}} = j + O(\frac{1} {j})\quad \text{as }\quad j \rightarrow \infty,& {}\\ & \Psi _{j}(s) \sim \frac{\sin (\,js)} {j} \quad \text{as}\quad j \rightarrow \infty. & {}\\ \end{array}$$

The main term of the normalized eigenfunction is:

$$\displaystyle{ \frac{\Psi _{j}} {\vert \vert \Psi _{j}\vert \vert } \sim \sqrt{2/\pi }\sin (\,js)\quad \text{as}\quad j \rightarrow \infty,}$$

and the main term of the normalizing constant is:

$$\displaystyle{\alpha _{j} \sim \frac{\pi } {2j^{2}}\quad \text{as}\quad j \rightarrow \infty.}$$

8.4 Auxiliary Results

Lemma 1

If one of the limits \(\lim _{t\rightarrow \infty }\frac{1} {t} \int _{0}^{t}u(s)ds\) or \(\lim _{\lambda \rightarrow 0}\lambda \mathcal{U}(\lambda )\) exists, then the other also exists and they are equal to each other:

$$\displaystyle{\lim _{t\rightarrow \infty }\frac{1} {t}\int _{0}^{t}u(s)ds =\lim _{\lambda \rightarrow 0}\lambda \mathcal{U}(\lambda ),}$$

where

$$\displaystyle{\mathcal{U}(\lambda ):=\int _{ 0}^{\infty }e^{-\lambda t}u(t)dt:=\bar{ u}(\lambda ).}$$

Proof

Denote

$$\displaystyle{\frac{1} {t}\int _{0}^{t}u(t)dt:= v(t),\quad \bar{u}(\sigma ):=\int _{ 0}^{\infty }e^{-\sigma t}u(t)dt.}$$

Then

$$\displaystyle{\bar{v}(\lambda ) =\int _{ \lambda }^{\infty }\frac{\bar{u}(\sigma )} {\sigma } d\sigma }$$

by the properties of the Laplace transform.

Assume that the limit \(v(\infty ):= v_{\infty }\) exists:

$$\displaystyle{ \lim _{t\rightarrow \infty }v(t) = v_{\infty }. }$$
(8.59)

Then,

$$\displaystyle{v_{\infty } =\lim _{\lambda \rightarrow 0}\lambda \int _{0}^{\infty }e^{-\lambda t}v(t)dt =\lim _{\lambda \rightarrow 0}\lambda \bar{v}(\lambda ).}$$

Indeed \(\lambda \int _{0}^{\infty }e^{-\lambda t}dt = 1\), so

$$\displaystyle{\lim _{\lambda \rightarrow 0}\lambda \int _{0}^{\infty }e^{-\lambda t}(v(t) - v_{ \infty })dt = 0,}$$

and (8.59) is verified.

One has

$$\displaystyle{ \lim _{\lambda \rightarrow 0}\lambda \bar{v}(\lambda ) =\lim _{\lambda \rightarrow 0}\int _{\lambda }^{\infty }\frac{\lambda } {\sigma }\bar{u}(\sigma )d\sigma =\lim _{\lambda \rightarrow 0}\lambda \bar{u}(\lambda ), }$$
(8.60)

as follows from a simple calculation:

$$\displaystyle{ \lim _{\lambda \rightarrow 0}\int _{\lambda }^{\infty }\frac{\lambda } {\sigma }\bar{u}(\sigma )d\sigma =\lim _{\lambda \rightarrow 0}\int _{\lambda }^{\infty } \frac{\lambda } {\sigma ^{2}}\sigma \bar{u}(\sigma )d\sigma =\lim _{\sigma \rightarrow 0}\sigma \bar{u}(\sigma ), }$$
(8.61)

where we have used the relation \(\int _{\lambda }^{\infty } \frac{\lambda } {\sigma ^{2}}d\sigma = 1\).

Alternatively, let \(\sigma ^{-1} =\gamma\). Then,

$$\displaystyle{ \int _{\lambda }^{\infty } \frac{\lambda } {\sigma ^{2}}\sigma \bar{u}(\sigma )d\sigma = \frac{1} {1/\lambda }\int _{0}^{1/\lambda }\frac{1} {\gamma } \bar{u}(\frac{1} {\gamma } )d\gamma = \frac{1} {\omega } \int _{0}^{\omega }\frac{1} {\gamma } \bar{u}(\frac{1} {\gamma } )d\gamma. }$$
(8.62)

If \(\lambda \rightarrow 0\), then \(\omega =\lambda ^{-1} \rightarrow \infty,\) and if

$$\displaystyle{\psi:=\gamma ^{-1}\bar{u}(\gamma ^{-1}),}$$

then

$$\displaystyle{ \lim _{\omega \rightarrow \infty }\frac{1} {\omega } \int _{0}^{\omega }\psi d\gamma =\psi (\infty ) =\lim _{\gamma \rightarrow 0}\gamma ^{-1}\bar{u}(\gamma ^{-1}) =\lim _{\sigma \rightarrow \infty }\sigma \bar{u}(\sigma ). }$$
(8.63)

Lemma 1 is proved.

Lemma 2

Equation (8.21)holds.

Proof

As \(a \rightarrow 0\), one has

$$\displaystyle{ \frac{\partial } {\partial N_{s}} \frac{e^{-\sqrt{\lambda }\vert s-s'\vert }} {4\pi \vert s - s'\vert } = \frac{\partial } {\partial N_{s}} \frac{1} {4\pi \vert s - s'\vert } + \frac{\partial } {\partial N_{s}} \frac{e^{-\sqrt{\lambda }\vert s-s'\vert }- 1} {4\pi \vert s - s'\vert }. }$$
(8.64)

It is known (see [8]) that

$$\displaystyle{ \int _{\mathcal{S}_{m}}ds\int _{\mathcal{S}_{m}} \frac{\partial } {\partial N_{s}} \frac{1} {4\pi \vert s - s'\vert }\sigma _{m}(s')ds' = -\frac{1} {2}\int _{\mathcal{S}_{m}}\sigma _{m}(s')ds' = -\frac{1} {2}Q_{m}. }$$
(8.65)

On the other hand, as \(a \rightarrow 0\), one has

$$\displaystyle{ \bigg\vert \int _{\mathcal{S}_{m}}ds\int _{\mathcal{S}_{m}}\frac{e^{-\sqrt{\lambda }\vert s-s'\vert }- 1} {4\pi \vert s - s'\vert } \sigma _{m}(s')ds'\bigg\vert \leq \vert Q_{m}\vert \int _{\mathcal{S}_{m}}ds\frac{1 - e^{-\sqrt{\lambda }\vert s-s'\vert }} {4\pi \vert s - s'\vert } = o(Q_{m}). }$$
(8.66)

The relations (8.65) and (8.66) justify (8.21).

Lemma 2 is proved.

Lemma 3

If assumption (8.18)holds, then inequality (8.17)holds.

Proof

One has

$$\displaystyle{ \mathcal{J}_{1,m}:= \vert g(x,x_{m})Q\vert = \frac{\vert Q_{m}\vert e^{-\sqrt{\lambda }\vert x-x_{m}\vert }} {4\pi \vert x - x_{m}\vert }, }$$
(8.67)

and

$$\displaystyle{ \mathcal{J}_{2,m} \leq \frac{e^{-\sqrt{\lambda }\vert x-x_{m}\vert }} {4\pi \vert x - x_{m}\vert }\max \bigg(\sqrt{\lambda }a, \frac{a} {\vert x - x_{m}\vert }\bigg)\int _{\mathcal{S}_{m}}\vert \sigma _{m}(s')\vert ds' }$$
(8.68)

where | xx m  | ≥ d, and d > 0 is the smallest distance between two neighboring particles. One may consider only those values of \(\lambda\) for which \(\lambda ^{1/4}a < \frac{a} {d}\), because for the large values of \(\lambda\), such that \(\lambda ^{1/4} \geq \frac{1} {d}\) the value of \(e^{-\sqrt{\lambda }\vert x-x_{m}\vert }\) is negligibly small. The average temperature depends on the behavior of \(\mathcal{U}\) for small \(\lambda\), see Lemma 1.

One has \(\vert Q_{m}\vert =\int _{\mathcal{S}_{m}}\vert \sigma _{m}(s')\vert ds' > 0\) because \(\sigma _{m}\) keeps sign on \(\mathcal{S}_{m}\), as follows from Eq. (8.24) as \(a \rightarrow 0\).

It follows from (8.67)–(8.68) that

$$\displaystyle{ \bigg\vert \frac{\mathcal{J}_{2,m}} {\mathcal{J}_{1,m}}\bigg\vert \leq O\bigg(\bigg\vert \frac{a} {x - x_{m}}\bigg\vert \bigg) \leq O\bigg(\frac{a} {d}\bigg) << 1. }$$
(8.69)

From (8.69) by the arguments similar to the given in [9] one obtains (8.17).

Lemma 3 is proved.