Reiterated homogenization of a laminate with imperfect contact: gain-enhancement of effective properties
1 Introduction A typical feature of problems in nanofluids[1], bone mechanics[2-3], soil physics[5], and porous media[6], etc. is that the studied objects are heterogeneous media exhibiting several structural scales. In general, the mathematical modeling of such situations involves differential equations with rapidly-oscillating piecewise-smooth coefficients depending on several local or fast variables corresponding to such structural scales. The application of homogenization techniques to determine the macroscopic properties of such media plays an important role in the knowledge of the interest phenomena. Some basic elements from the homogenization theory may be found in Ref. [7]. A rigorous mathematical tool for this purpose is the reiterated homogenization method (RHM) introduced in Ref. [8]. Mathematically, the RHM is an asymptotic method. It allows to transform problems, which involve differential equations with periodic and rapidly oscillating coefficients dependent on several local variables, into the problems where the coefficients are not rapidly oscillating. The latter problems are known as homogenized problems, and their coefficients are the so-called effective coefficients of the original heterogeneous media. A crucial step is to show the proximity between the solutions of both types of the problems. The existence of several scales directly affects the effective coefficients, and it makes other intermediate coefficients appear, which are related to each structural scale. In addition, the presence of the structural imperfections related to the piecewise-continuous character of the coefficients must be considered by adding the contact conditions relevant to each physical situation under the analysis.
The RHM has been applied to multidimensional elliptic problems by various authors. In the pioneering work of Ref. [8], the RHM was first introduced for scalar elliptic problems depending on three scales by using the asymptotic expansion. These results were generalized in Ref. [9] for several microscopic scales by introducing the multiscale convergence method. In Ref. [3], the Γ-convergence method was applied to an elastic problem with the coefficients depending on four scales. A detailed description of the periodic unfolding method and how it was applied to periodic homogenization multiscale elliptic problems could be found in Ref. [10]. However, in the classical books[11-12], the reiterated homogenization was not studied. An interesting review of the different mathematical techniques for conventional two-scale homogenization was reported in Ref. [13].
The goal of this work is to illustrate, from the beginning to the ending, the main stages of the RHM by applying some basic elements of the theory of differential equations and functional analysis so as to provide an accessible material for a wide audience for the applied mathematicians, scientists, and engineers interested in this field. To this aim, the RHM is applied to a family of one-dimensional (1D) elliptic boundary-value problems with periodic and rapidly oscillating coefficients depending on two microscopic scales and a finite number of discontinuities. A mathematical justification of the asymptotic process is given based on an estimate. Although, the investigated family of problems is just as simple as to admit the analytical solutions, it is suitable to unfold the new aspects of the behavior of the RHM solution with respect to the conventional homogenization solution. This study could provide a relevant extension of previous perfect-contact RHM analyses[6, 8, 13].
The following content of this paper is organized as follows. In Section 2, a formal asymptotic solution for a family of 1D two-point boundary-value problems with imperfect contact conditions is constructed. The expressions for the solutions of the local problems and the effective coefficients are given, and so is the statement of the related homogenized problem. A necessary and sufficient condition for the existence of periodic solutions is stated. In Section 3, some elements of the Sobolev space theory are extended to open non-connected sets. The variational formulation of the problem is considered. Finally, with the Lax-Milgram lemma, the convergence of the solution of the original problem to the solution of the homogenized one is proven. In Section 4, an example with piecewise-differentiable coefficients is developed mainly to illustrate the gain of the effective properties of a three-scale heterogeneous 1D laminate with respect to their two-scale counterparts.
2 Reiterated homogenization for the 1D elliptic equation with piecewise differentiable coefficients and discontinuity conditions In this section, the problem is stated. Following Refs. [8], [11], and [14], a formal asymptotic solution is constructed in order to obtain the local problems, the effective coefficients, and the homogenized problem.
2.1 Problem settings Let
The periodic extension to
of the real function a(y, z) is differentiable, positive, and bounded for all (y, z)∈Ωy*×Ωz*. Then, a(y, z) can be regarded as 1-periodic with respect to y and z, and there exist α-, α+∈
such that 0 < α- ≤ a(y, z) ≤α+ for all (y, z).
Let F(x) be a piecewise continuous real function in [0, 1]. Define the contrast operator as follows:
Let ε = 1/n (n∈
). The problem is to find uε(x) such that
|
(1) |
|
(2) |
|
(3) |
|
(4) |
where
The numbers xijk represent the discontinuity points of aε(x) in the interval (0, 1), i.e.,
That is, there are l1n+l2n2 discontinuity points, and βkiε=ε-kβki, where βki∈
(i=1, 2, …, lk, and k = 1, 2) indicate the microscales where the related discontinuities take place.
Since the condition in Eq. (2) becomes
as βki→∞,
uε(x) is continuous. In the context of composite media, this continuity condition together with the condition in Eq. (3) are known as perfect contact conditions. The expressions related to the homogenization of the problems with the perfect contact conditions are obtained by taking the limit as βki→∞ in the more general expressions to be obtained for Eqs. (1)-(4).
In the context of heat conduction, uε(x) is the temperature distribution in the heterogeneous medium,
aε(x) is its thermal conductivity, and f is the heat source. In addition, Eq. (3) is the continuity condition on the heat flux, and Eq. (2) is a discontinuity condition for the temperature to state that the size of the heat flux across each discontinuity point is proportional to the temperature change at the point. The proportionality constants are the thermal contact conductances βkiε, at the discontinuity points xijk, related to the Biot number defined in Ref. [15].
2.2 Homogenization, local problems, and effective coefficients The first step in this homogenization process is to construct a formal asymptotic solution for Eqs. (1)-(4) as follows:
|
(5) |
where ui(x, y, z) (i=0, 1, …, m) are piecewise differentiable and 1-periodic with respect to the local or fast variables y=x/ε and z=x/ε2.
By substituting the expansion (5) into Eqs. (1)-(4), using the chain rule
and equating the terms corresponding to equal powers of ε, we can obtain the following recurrent family of partial differential equations:
|
(6) |
|
(7) |
|
(8) |
|
(9) |
|
(10) |
and
|
(11) |
for 1≤ i≤ m-4, where, with p, q∈{x, y, z}, the differential operators Lpq are defined as follows:
Analogously, the contact conditions can be expressed as follows:
(ⅰ) For ε-2,
|
(12) |
|
(13) |
|
(14) |
|
(15) |
(ⅱ) For ε-1,
|
(16) |
|
(17) |
|
(18) |
|
(19) |
(ⅲ) For ε0,
|
(20) |
|
(21) |
|
(22) |
|
(23) |
(ⅳ) For εi (1≤ i≤ m-4),
|
(24) |
|
(25) |
|
(26) |
|
(27) |
The boundary conditions (4) for the unknown functions ui(x, y, z) are
|
(28) |
|
(29) |
|
(30) |
The following lemma guarantees the existence of the 1-periodic solutions of the recurrent chain of Eqs. (6)-(30), and, consequently, allows the construction of a formal asymptotic solution for Eqs. (1)-(4).
Lemma 1 Let a(ξ), F0(ξ),
and F1(ξ) be 1- periodic piecewise-differentiable functions with finite jump discontinuities in ξ∈{ξ1, ξ2, …, ξn} (0 < ξj < 1) and positive a(ξ) bounded over [0, 1]. Then, a necessary and sufficient condition for the existence of a 1-periodic solution N(ξ) of the problem
|
(31) |
|
(32) |
|
(33) |
is
|
(34) |
Remark 1 The lemma is stated for functions depending on a single variable. Functions depending on several variables can be considered as functions of a single variable by assuming that the other variables are parameters with fixed values.
Remark 2 If N(ξ) is a 1-periodic solution of Eqs. (31)-(33), there exists a family of 1-periodic solutions given by
=N(ξ)+c, where c∈
is an arbitrary constant, i.e., the solution is unique up to an additive constant. Thus, without loss of generality, the condition N(0)=0 can be added to Eqs. (31)-(33) in order to guarantee the uniqueness of the 1-periodic solution.
Now, Lemma 1 is applied to each of the problems formed by Eqs. (6)-(10) and the corresponding conditions (12)-(23). First, we consider the functions ui of the form
|
(35) |
where Ni(x, y, z), Mi(x, y), and vi(x) are infinitely piecewise-differentiable functions. The main results obtained after the application of Lemma 1 to Eqs. (6)-(10) are the first and second local problems along with the homogenized problem.
The first local problem, for each fixed y, is enunciated as to find the 1-periodic solution N(y, z) for
|
(36) |
|
(37) |
|
(38) |
|
(39) |
This problem also satisfies the hypotheses of Lemma 1 with ξ = z, F0=0, and F1=-a(y, z) for every y. Therefore, it has a solution N(y, z) which is 1-periodic with respect to z. Besides, according to Remark 2, Eq. (39) could ensure the uniqueness of the solution.
Since Eq. (36) states that a(y, z)
+a(y, z) is independent of z, we can introduce the intermediate effective coefficient
(the effective coefficient related to the first microscale of the original problem) as follows:
|
(40) |
Some manipulations of this relation yield
|
(41) |
Note that
is a positive and 1-periodic function of y. It is important to emphasize that the calculation of the effective coefficients depends on the solution of the respective local problems. Typically, these problems will demand numerical solutions for the physical situations of practical interest. The analytical solution of the first local problem (36)-(39) is
|
(42) |
where z∈(zj, zj+1), and j=0, 1, …, l2.
The second local problem consists in finding the 1-periodic solution M(y) for the problem
|
(43) |
|
(44) |
|
(45) |
|
(46) |
This problem also satisfies the hypotheses of Lemma 1 with ξ = y, F0=0, and F1=-
. Therefore, it has a 1-periodic solution M(y). Since Eq. (43) states that
is independent of y, we can introduce the global effective coefficient
related to the macroscale of the original problem as follows:
|
(47) |
Some manipulations in this relation give
|
(48) |
Note that
is a positive constant. Therefore, it is 1-periodic. In the case of a single microscale, in which a(y, z) = a(y) is a piecewise-constant function that takes only two different values,
given by Eq. (48) becomes the 1D realization of the elementary lower bound on the effective conductivity reported in Eq. (3.1) of Ref. [16]. Moreover, it is interesting to note that, in higher-dimension single-microscale problems, the remaining sum in the corresponding version of Eq. (48) becomes an integral over the discontinuity surfaces. In the present case, the analytical solution of the second local problem (43)-(46) is
|
(49) |
where y∈(yj, yj+1), and j=0, 1, …, l1.
The so-called homogenized problem is
|
(50) |
The solutions from those three problems play an important role in the expressions for functions u0, u1, u2, and u3. With Eq. (35), we have that N0(x, y, z)=M0(x, y)=0, and v0(x) is the solution of Eq. (50). For u1(x, y, z), we have N1(x, y, z)=0, and
|
(51) |
where M(y) is the solution of the second local problem. The function v1 will be determined later. For u2(x, y, z), we have
|
(52) |
where N(y, z) is the solution of the first local problem. The function M2(x, y) will be considered to have the form M2(x, y)=M20(x, y)+
, where M20(x, y) is the solution of the problem
|
(53) |
|
(54) |
|
(55) |
|
(56) |
The function
will be introduced later. Note that this problem satisfies the hypotheses of Lemma 1 with
for every x. Then, the solution M20(x, y) exists, and is 1-periodic in y. In addition, the equation of the problem for M20(x, y) states that
does not depend on y. Therefore, it is possible to define
|
(57) |
The function v2(x) will be determined later. For u3(x, y, z), we have
|
(58) |
where N30(x, y, z) is the 1-periodic solution of the following problem:
|
(59) |
|
(60) |
|
(61) |
|
(62) |
The problem (59)-(62) satisfies the hypotheses of Lemma 1 with
for every x and y. Therefore, the solution N30(x, y, z) exists, is 1-periodic in z, and, according to Remark 2, is unique. The solution is
|
(63) |
for z∈(zj, zj+1) and j=0, 1, …, l2, where
|
(64) |
The functions M3(x, y) and v3(x) will be determined later.
Remark 3 Until now, the main results of the homogenization process have been stated with a fourth-order asymptotic expansion. Constructing higher-order formal asymptotic solutions is an induction process that requires the consideration of the asymptotic expansions of orders higher than the fourth in order to provide the information which completes Table 1. In Table 1, all the previous results are summarized, constituting the basis for an induction process developed next.
Table 1 Functions involved in the asymptotic expansion
For the sake of mathematical completeness, from now on, for m≥ 5, the recurrent family of problems formed by Eq. (11) with Eqs. (24)-(27) will be studied.
For each i, Eqs. (11), (25), and (27) constitute a problem satisfying the hypotheses of Lemma 1 with
for every x and y. Therefore, the existence of a solution ui+4(x, y, z) of Eq. (11), which is 1-periodic in z, can be ensured by satisfying the necessary and sufficient condition
|
(65) |
In fact, the condition (65) is satisfied if F0=0 as it does not depend on z. Indeed, proceeding as above, it follows that
|
(66) |
Thus, equating Eq. (66) to zero leads to the equation for ui+4(x, y, z), Lzzui+4=-Lzyui+3-Lzxui+2, and
This equation, together with the corresponding appropriate conditions, satisfies the hypotheses of Lemma 1 with
for every x. The necessary and sufficient condition
is satisfied if F0=0 as F0 does not depend on y, which together with the boundary conditions (30) lead to the problem for vi(x) as follows:
|
(67) |
Then, the solution Mi+2(x, y) exists, is 1-periodic in y, and, according to Remark 2, is unique. Similarly,
|
(68) |
where Mi+20(x, y) is the solution of the problem
|
(69) |
|
(70) |
|
(71) |
|
(72) |
This problem satisfies the hypotheses of Lemma 1 with
for every x. Therefore, the solution Mi+20(x, y) exists, is 1-periodic in y, and, according to Remark 2, is unique.
In addition, the equation of the problem for Mi+20(x, y) states that
does not depend on y. Therefore, it is possible to define
and Eq. (70) becomes
Now, the deduced equation for ui+4(x, y, z) can be written as
which leads to an expression that suggests
|
(73) |
With these considerations, ui+4(x, y, z) takes the following form:
|
(74) |
where Ni+40(x, y, z) is the 1-periodic solution for the problem
This problem satisfies the hypotheses of Lemma 1 with
for every x and y. Therefore, the solution Ni+40(x, y, z) exists, is 1-periodic in z, and, according to Remark 2, is unique. The solution of this problem is
|
(75) |
for z∈(zj, zj+1) and j=0, 1, …, l2, where
|
(76) |
Remark 4 In the previous section, the basis of the induction process developed here is established (see Remark 3), which corresponds to the consideration i = -4, -3, …, 0 in Eq. (11). In particular, the necessary and sufficient condition ensuring that the existence of the 1-periodic solutions ui+4 = Ni+4 + Mi+4 + vi+4 is satisfied if the sum of the last six terms of Eq. (11) is equal to zero, i.e., F0 = 0, leading to problems (or expressions) to obtain some of the functions Ni,
Mi, and vi. Now, note that, for i=1, 2, …, m-4 in Eq. (11), such a condition also leads to Eqs. (73), (68), and (67) for the functions Ni, Mi, and vi and that F0 = 0. This ensures that if F0 = 0 for i, F0=0 for i+1. Thus, the process of constructing higher-order formal asymptotic solutions is justified by mathematical induction, which follows a recurrent cycle as follows:
where negative-index functions are considered to be null.
3 Mathematical justification In this section, some of the results of Chapter 2 of Ref. [17] are extended to the case of piecewise differentiable coefficients in order to provide a mathematical justification of the homogenization process.
3.1 Preliminaries Consider Ω*=(0, 1)\{x1, x2, …, xl}
(l∈
), and denote x0=0 and xl+1=1. Then, the generalized (weak) derivative of a function u∈L2((0, 1)) can be defined as follows:
where
for k∈
. If Du(φ) is bounded for all u∈L2((0, 1)), it follows from the Riesz representation theorem[17] that there exists a unique function v∈L2((0, 1)) such that Du(φ)=〈 v, φ〉L2((0, 1)) for all φ∈L2((0, 1)). Then, it is said that the weak derivative belongs to L2((0, 1)), and
=v. In the particular case of u∈
(Ω*),
is the usual derivative of u in each interval (xj, xj+1).
Now, the space H1(Ω*) can be defined as the space of all functions on L2(Ω*) with weak derivatives belonging to L2(Ω*) as follows:
An equivalent definition of H1(Ω*) is
The space H1(Ω*) is equipped with the usual norm, denoted as ‖·‖1, and is defined as follows:
where ‖·‖L2(Ω*) = ‖·‖L2((0, 1)). Therefore, it is necessary to equip H1(Ω*) with a norm accounting for the behavior of functions in the neighborhood of the discontinuity points x = xj (j = 1, 2, …, l).
Proposition 1 Consider u∈H1(Ω*),
and define ‖·‖1* as follows:
Then, ‖·‖1* is a norm in H1(Ω*) equivalent to the norm ‖·‖1, and H1(Ω*) equipped with ‖·‖1* is a Banach space.
In fact, one can prove that
where
Such an equivalence of norms implies that, since H1(Ω*) equipped with the norm ‖·‖1 is a Banach space, it is also a Banach space if it is equipped with the norm ‖·‖1*.
Remark 5 Consider the space H01(Ω*), which is the subspace of H1(Ω*) defined as follows:
The seminorm
is a norm in H01(Ω*).
The following three propositions ensure that H01(Ω*) equipped with |·|1* is a Banach space.
Proposition 2 H01(Ω*) is a closed subspace of H1(Ω*).
Proposition 3
(Ω*)={u∈
(Ω*) : u(0)=u(1)=0} is dense in H01(Ω*).
Proposition 4 The norms ‖·‖1* and |·|1* are equivalent in H01(Ω*).
The following is a proof for Proposition 4.
Proof Consider x∈Ω* and u∈
(Ω*). Then,
Taking the absolute values and using the Cauchy-Schwarz inequalities for the integral and the sum on the right-hand side, we have
|
(77) |
Squaring both sides, using the inequality (x+y)2≤ 2(x2+y2), and integrating with respect to x∈(0, 1), we have
Adding
to both sides of the above equation leads to
Then, the equivalence can be obtained from the relation
Since
(Ω*) is dense in H01(Ω*), the equivalence holds in H01(Ω*).
Remark 6 H01(Ω*) equipped with the norm |·|1* is a closed subspace of a Banach space. Hence, it is also a Banach space. Moreover, the norm |·|1* is associated with the scalar product
Therefore, H01(Ω*) is a Hilbert space.
Remark 7 It follows from Eq. (77) that, for any u∈H01(Ω*),
Then,
Remark 8 Both constants in the equivalence relation of the norms ‖·‖1* and |·|1* and in the relation from Remark 7 depend on l, which is the number of the discontinuity points of functions in H1(Ω*).
3.2 A variational problem for equation -Lξξw=f0-f'1 Consider Ω* = (0, 1)\{ξ1, …, ξl}
(l∈
), ξ0 = 0, and ξl+1 = 1. Then, the problem is to find w∈
(Ω*) such that
|
(78) |
|
(79) |
|
(80) |
|
(81) |
where a, f1∈
(Ω*), and f0∈
(Ω*), such that α-≤ a(ξ)≤α+ for every ξ∈Ω*. In the above equations, α-, α+, βj∈
,
cj1, cj2∈
, and j = 1, 2, …, l.
The variational formulation of Eqs. (78)-(81) is to find w∈H01(Ω*) such that
|
(82) |
for all φ∈H01(Ω*), where
|
(83) |
|
(84) |
A solution of Eqs. (82)-(84) is called a generalized or weak solution of Eqs. (78)-(81).
The Lax-Milgram lemma (Theorem A.3 in Ref. [17]) allows proving that Eq. (82) is well-posed if b(w, φ) given by Eq. (83) is a bounded coercive bilinear form and L(φ) given by Eq. (84) is a bounded linear form, both in H01(Ω*).
Proposition 5 The bilinear form b(w, φ) given by Eq. (83) is coercive and bounded in H01(Ω*).
Proof Since a(ξ)≥α-,
|
(85) |
where
Then, b(w, φ) is coercive.
Since a(ξ)≤α+, from the Cauchy-Schwarz inequality, we have
|
(86) |
where
Then, b(w, φ) is bounded.
Proposition 6 The linear form L(φ)
given by Eq. (84) is bounded in H01(Ω*).
Proof Remark 7 ensures the boundedness of the lateral limits and contrast operators over H01(Ω*) through the relations
Furthermore, from the Cauchy-Schwarz inequality and the equivalence of the norms ‖·‖1, ‖·‖1*, and |·|1*, we have
where
Consequently, |L(φ)|≤ C2|φ|1*, where
Remark 9 It follows from Propositions 5 and 6 that Eqs. (82), (83), and (84) satisfy the hypotheses of the Lax-Milgram lemma. Then, there exists a unique function w∈H01(Ω*) such that Eqs. (82), (83), and (84) hold for all φ∈H01(Ω*), i.e., there exists a unique solution for the variational formulation of Eqs. (82), (83), and (84). Additionally, |w|1*≤ C2/C0 holds for such a solution that is an immediate consequence of the chain of inequalities C0|w|1*2≤ b(w, w)=L(w)≤ C2 |w|1* and, as noted above, depends on the number l of the function discontinuity points in H01(Ω*).
Remark 10 If the solution w of Eqs. (82), (83), and (84), which is a weak solution of Eqs. (78)-(81), belongs to H01(Ω*)∩
(Ω*), and satisfies Eqs. (79) and (80), it is the solution of Eqs. (78)-(81).
Remark 11 Note that, for large values of βj
(j=1, 2, …, l), |w|1*≤ C2/C0 is independent of βj, since C0=min{α0, β-}=α-. Then, to ensure that Eq. (79) holds in the limit as βj→∞, it is necessary that
.
3.3 An auxiliary problem Recall the statement of the original problem in Section 2, and consider the problem to find vε(x) such that
|
(87) |
|
(88) |
|
(89) |
|
(90) |
|
(91) |
where δk1 is Kronecker's delta, cijε∈
are such that |cijε|≤εC, and C∈
does not depend on ε.
From Eqs. (87) and (90), we have
|
(92) |
where cε∈
. Then, Eq. (92) in combination with Eq. (91) yields
|
(93) |
The condition (88) implies that
|
(94) |
The substitution of Eqs. (94) into (93) followed by some manipulations yields
|
(95) |
As a(y, z)≤α+, it follows that
|
(96) |
which is a bound that does not depend on ε. Moreover, since |cijε|≤εC and ε=1/n, it follows that
|
(97) |
which is also a bound that does not depend on ε. Thus, the combination of Eqs. (96) and (97) with Eq. (95) yields
|
(98) |
As a(y, z)≥α-, the combination of Eq. (98) with Eq. (92) leads to
|
(99) |
Moreover, since |cijε|≤εC, the combination of Eq. (98) with Eq. (94) leads to
|
(100) |
which implies that
|
(101) |
|
(102) |
Let x∈(0, 1). Then, the condition (91) implies that
where Dxε is the set of discontinuity points of vε(x) contained between 0 and x. Taking absolute values on both sides yields
Applying the Cauchy-Schwarz inequality on the integral, with Eqs. (99), (101), and (102), we have
|
(103) |
Squaring both sides and integrating from 0 to 1 with respect to x, we have
|
(104) |
The relation (103) implies pointwise convergence of vε to 0 when ε→0+, while Eq. (104) implies strong convergence on L2((0, 1)).
In what follows, the convergence of the solution uε(x) of the original problem (1)-(4) to the solution u0(x) of the homogenized problem (50) as ε→0+ will be proven.
3.4 Proximity between uε(x) and u0(x) Let ε=1/n be fixed. Let uε(x) be the solution of the original problem (1)-(4) and u(m)(x, ε) be its formal asymptotic solution constructed in Section 2. With the results from the homogenization process, we have that the function wε(x)=uε(x)-u(m)(x, ε) is the solution of the following problem:
|
(105) |
|
(106) |
|
(107) |
|
(108) |
|
(109) |
where
which is given by
is bounded in L2(0, 1) as
In addition, the constants cijk1 and cijk2 given by
are bounded as
Moreover, the second term on the right-hand side of Eq. (106) is also bounded. Indeed,
where
Since Kr (r=1, 2, 3, 4) do not depend on ε, recalling Subsection 3.2, we can see that Eqs. (105)-(109) are well posed with the solution w∈H01(Ωε*). Consequently, the auxiliary problem
|
(110) |
|
(111) |
|
(112) |
|
(113) |
|
(114) |
satisfies the hypotheses of Eqs. (87)-(91) in Subsection 3.3. Therefore, ‖vε‖L2((0, 1))→0 as ε→0+.
From Eqs. (105)-(109) and (110)-(114), we have that the function defined as χε=wε-vε is the solution of the problem
Now, it is possible to apply Remark 9 to obtain the estimate
|
(115) |
where K=
α--1max{K1, K2+K3}, which holds for small values of ε. Then, with Remark 7 and Eq. (115), we have
|
(116) |
The right-hand side of Eq. (116) is of order εm-5. Then, choosing an asymptotic expansion u(m)(x, ε) with m≥6 and taking the limit as ε→0+, we have
|
(117) |
Moreover, the triangular inequality leads to
|
(118) |
Then, using Eq. (117), recalling that ‖vε‖L2((0, 1))→0 as ε→0+, and taking the limit as ε→0+ in Eq. (118), we have
|
(119) |
Remark 12 Consider ε to be fixed. Then, Eq. (117) follows from Eq. (116) by taking the limit as m→∞.
Moreover, it follows from Eq. (5) that
Consequently,
Then, taking the limit as ε→0+, we have
|
(120) |
Finally, the combination of Eqs. (117) and (120) with the triangular inequality
as ε→0+ leads to
|
(121) |
i.e., uε→u0 in L2((0, 1)) as ε→0+, which completes the mathematical justification of the reiterated homogenization process in Section 2.
4 Some numerical examples 4.1 Comparison of solutions In order to illustrate the theoretical results presented above, consider the original problem (1)-(4), in which f(x) =0, β1i=β1, β2i=β2, t0=0,
t1=1, and a(y, z) given by
where
and
In the above equations,
The expression (41) for the intermediate effective coefficient leads to
whereas Eq. (48) yields the global effective coefficient
|
(122) |
Moreover, the solution of the first local problem (36)-(39) is
Furthermore, the solution M(y) of the second local problem (43)-(46) is
The solution of the homogenized problem (50) is v0(x)≡u0(x)=x. The functions N(y, z) and M(y) allow to obtain the function N2(x, y, z), Eq. (52), which in the present case is
In this case,
which leads to Ni(x, y, z)=0 for i≥3, Mi(x, y)=0 for i≥2, and vi(x)=0 for i≥1. Therefore,
with y=x/ε and z=x/ε2 for all m≥ 2 is the exact solution of the problem. Figure 1 illustrates the convergence of the exact solution uε(x) to the solution u0(x) of the homogenized problem as ε→0+ for β1=β2=1. From the figure, we can see that, for ε=1/64, there are 2(642+64)=8 320 discontinuity points.
4.2 Gain enhancement In this section, a problem with only one micro-scale will be homogenized through conventional homogenization, and the corresponding effective coefficient will be compared with the one obtained after applying the RHM to an equivalent problem with two micro-scales. Besides, it will be illustrated that there is a gain of macroscopic properties from the effective coefficient of the second problem with respect to the first one (see Ref. [18]). Consider h1(y)=(2+sin(2πy))-1, β1, β2, h2∈
, y0∈(0, 1), Φ∈[0, 1-y0), and
where
and
Apply the conventional homogenization method to the problem
|
(123) |
|
(124) |
|
(125) |
|
(126) |
where
and x1j=ε(y0+j) and x2j=ε(y0+Φ+j) are discontinuity points from aε(x). Then, we have
Consider now y1, z1∈(0, 1), ΦY∈[0, 1-y1),
ΦZ∈[0, 1-z1), and
with
where
Then, we have
and
It is important to remark that if y1=y0, ΦY=Φ, and ΦZ=0, the functions bε(x)=b(x/ε, x/ε2) and aε(x) coincide. Let κi1=βi, and κi2=∞. Then,
|
(127) |
|
(128) |
|
(129) |
|
(130) |
where κkiε=ε-kκik (i, k=1, 2), and x1j1=ε(y1+j),
x2j1=ε(y1+ΦY+j), x1j2=ε(z1+j), and x2j2=ε(z1+ΦZ+j) are discontinuity points from bε(x), which coincide with Eqs. (123)-(126). For ΦZ=0, there are no discontinuity points at z=z1, and the discontinuity conditions from z1 and z1+ΦZ overlap each other and must be fulfilled simultaneously. In that case, the problem will only have a solution if κ12=κ22. Applying the RHM to it gives exactly the same effective coefficient from Eqs. (123)-(126).
The aimed goal is to investigate the effect on that effective coefficient of the parameters ΦY, ΦZ, and κik with some constrains so as to ensure the equivalence between the problems (123)-(126) and (127)-(130), i.e., to determine the effect of the cell geometry on the effective coefficient.
To this end, the concept of volumetric fraction for each one of the phases plays an important role. In the periodic cell from the function a(y), there is a set, in which the function is equal to h2, i.e., Ω2. The volume fraction for this phase is defined as the total length of the phase divided by the total length of the periodic cell which is equal to 1. Therefore, the volumetric fraction for this phase is Φ. To ensure certain equivalence between the problems (123)-(126) and (127)-(130), the volumetric fraction of the phase, where b(y, z) and a(y) are equal to h2, must be the same. Consequently,
|Ω2YZ|=|Ω2|, or Φ=ΦY+ΦZ-ΦYΦZ. Moreover, it is also necessary to analyze how the coefficients κik affect the effective coefficient. Those constants must fulfill certain conditions to ensure that Eqs. (123)-(126) can be obtained as a limit case of Eqs. (127)-(130). Beside, Eqs. (127)-(130) have a solution for that limit case. As mentioned before, it is important that, if ΦZ=0 (and therefore ΦY=Φ), κ12=κ22, and κi1=βi. In the same way, if ΦY=0, κ11=κ21. The constants κik depend on the parameters ΦY and ΦZ, and also have effects on the conductivity. It will be ruled here that
|
though this is not the only possible consideration. Note that, if ΦZ=0, κ12=κ22=∞; if ΦZ=Φ,
κi2=βi. Finally, to obtain Eqs. (123)-(126) from Eqs. (127)-(130) when ΦY=Φ and ΦZ=0, y1 will be chosen in such a way that, when ΦY=Φ, y1=y0. This choice is not unique. Here, y1=y0 will be used for any value of ΦY.
With these conditions, the global effective coefficient obtained with the RHM to Eqs. (127)-(130) is
As ΦY depends on ΦZ, the effective coefficient depends on this parameter as well. This makes it possible to analyze the effective coefficient gain kgain=
as a function of ΦZ and to find the optimal values of ΦZ∈[0, Φ].
Figures 2, 3, and 4 show the behaviors of kgain as a function of ΦZ with the parameters Φ=0.5, y0=0.25, and the variable h2 for β1=β2=1, β1=β2=3.5, and β1=β2=40, respectively. From Figs. 2 and 4, we can see that, the absolute value of the effective coefficient gain is different from those for all the variation range of ΦZ.
Figure 3 also shows that there exist only two values of Φ∈(0, 0.5) so that the effective coefficient gain is equal to 1. It is interesting to mention that a similar behavior has also been found in Ref. [16] for the effective gain thermal conductivity of two-dimensional (2D) and three-dimensional (3D) composites with perfect contact conditions at the interfaces.
Figures 5, 6, and 7 show the behaviors of kgain as a function of ΦZ with different values of variables. From the figures, we can see that, in all cases, there exists enhancement in the gain. In particular, for large values of β1, β2,
y, and h2, the gain could be greater than 10%.
In Figs. 2, 3, and 4, the value of ΦZ, where there is no gain (kgain=1), is the same for all values of h2. The reason is that, for those values of y0 and Φ, the equation kgain=1, after some manipulation, becomes
which does not depend on h2. Consequently, its solution is independent of h2. In the same way, the minimum and maximum points of kgain are independent of h2. The stationary points of kgain can be obtained by solving k'gain=0, where the derivative is taken with respect to ΦZ.
After some manipulation, the equation k'gain=0 becomes
|
(131) |
This equation does not depend on h2, neither do its solutions. Since kgain has no critical points for ΦZ∈(0, 0.5), all maximum and minimum points are independent of h2.
5 Concluding remarks The RHM is applied to a family of two-point boundary-value problems with periodic and rapidly oscillating coefficients depending on two microscopic scales. A necessary and sufficient condition for the existence of periodic solutions is proven to guarantee the construction of a formal asymptotic solution of arbitrary order. A variational formulation is derived to obtain an estimate for the solution so as to provide a mathematical justification for the homogenization process. An example is presented to illustrate the possibilities of obtaining a gain in the macroscopic response of heterogeneous materials provided by different microscopic scales as well as under the effects of imperfect contact conditions at the interfaces.
Acknowledgements
The authors acknowledge the Cátedra Extraordinaria IIMAS-UNAM, México and PREI-DGAPA, UNAM, and are grateful to Ana Pérez Arteaga and Ramiro Chávez for the computational support. We would also like to thank to the reviewers for their valuable comments and to the editors for their inestimable assistance and suggestions.