488
Views
3
CrossRef citations to date
0
Altmetric
Original Articles

Application of the cut-off projection to solve a backward heat conduction problem in a two-slab composite system

, ORCID Icon, , &
Pages 460-483 | Received 25 May 2017, Accepted 14 Apr 2018, Published online: 07 May 2018

ABSTRACT

The main goal of this paper is applying the cut-off projection for solving one-dimensional backward heat conduction problem in a two-slab system with a perfect contact. In a constructive manner, we commence by demonstrating the Fourier-based solution that contains the drastic growth due to the high-frequency nature of the Fourier series. Such instability leads to the need of studying the projection method where the cut-off approach is derived consistently. In the theoretical framework, the first two objectives are to construct the regularized problem and prove its stability for each noise level. Our second interest is estimating the error in L2-norm. Another supplementary objective is computing the eigen-elements. All in all, this paper can be considered as a preliminary attempt to solve the heating/cooling of a two-slab composite system backward in time. Several numerical tests are provided to corroborate the qualitative analysis.

AMS SUBJECT CLASSIFICATIONS:

1. Introduction

The backward-in-time heat conduction problem has been scrutinized for a long time by many approaches with fast-adapting and effectively distinct strategies. This problem plays a critical role in different fields of study, for example, enabling us to determine the initial temperature of a volcano when the eruption has already started, or recover noisy, blurry digital images acquired by camera sensors.

The need for finding numerical solutions of various kinds of such equations is originated from its natural instability due to the drastic growth of the solutions. Once inaccurate final data are applied to the partial differential equation (PDE), one could obtain a totally wrong solution even if the final data are measured as accurately as possible. In fact, the measured data in real-life applications are always inaccurate. Thus, studies on regularization methods have attracted a considerable amount of attention, where stable approximate solutions were developed year by year. Several comprehensive materials are given in the textbook of Lattès and Lions [Citation1], and papers such as [Citation2Citation11].

In this paper, we consider the heat transfer problem in a two slabs composite (-b<x<0) and 0<x<a. We also follow the notations in [Citation12]. In particular, the problem, in which the time interval is defined by J:=t0,tf for t00,tf>0, can be expressed as:(1) ρbcbTbt=Kb2Tbx2,-b<x<0,tJ,(1) (2) ρacaTat=Ka2Tax2,0<x<a,tJ.(2)

The boundary conditions are:(3) Tbx=0atx=-b,Tax=0atx=a,tJ(3) (4) Tb=Ta,KbTbx=KaTaxatx=0,tJ.(4)

The model (Equation1)–(Equation4) involves a number of dimensionless quantities, namely the temperature Tα (C), mass density ρα (g/cm3), specific heat cα (cal/(g·C)), and thermal conductivity Kα (cal/(cm·sec·C)). Here the index α indicates either material a or b, i.e. α{a,b}. The perfect contact in (Equation4) indicates that the temperature and heat flux are continuous at the interface.

The model (Equation1)–(Equation4) is supplemented with the terminal measured data Tαεx,tf which satisfy:(5) Ta·,tf-Taε·,tfL20,a+Tb·,tf-Tbε·,tfL2-b,0ε,(5)

where 0<ε1 represents the maximum bound of errors arising in measurement. Henceforward, our task is to identify the initial value Tαεx,t0,-bxa.

Our main objective is to prove the ill-posedness of the aforementioned model and then deal with it by using the cut-off projection method. Historically, there have been several works that treat ill-posed problems using this method. For instances, Daripa and his co-workers in [Citation13,Citation14] contributed significantly. In particular, the cut-off method [Citation13] was used to control the spurious effects on the solution due to short wave components of the round-off and truncation errors. Recently, the cut-off method was utilized by Regińska and Regiński [Citation15] and Tuan et al. [Citation16] to stabilize the coshx- and sinhx/x-like growth in a Cauchy problem for the 3D Helmholtz equation.

This paper is fivefold. Section 2 is devoted to the eigen-elements and ill-posedness of the model. We compute the approximate eigenvalues and eigenfunctions, along with providing the temperature solution by Fourier-based modes. In Section 3, the cut-off method is described by providing the regularized problem corresponding to the terminal measured data Tαεx,tf. We also show the error estimates which clearly imply the stability of the regularized solution at each noise level. Lastly, some numerical tests are given in Section 4, whilst Section 5 is dedicated to a few conclusions and pointing out problems for possible further development.

2. Eigen-elements

In the context of Fourier representations, the eigen-elements including the eigenvalues and eigenfunctions can be obtained by solving the Sturm–Liouville problem. Such problem would be of the form:(6) ϕα+λα2ϕα=0forαa,b,(6)

associated with the boundary conditions (Equation3) and (Equation4). After some straightforward calculations and using (Equation3) one has(7) ϕb(x)=θbcosλbx+b,ϕa(x)=θacosλax-a,(7)

where θa,θb are constants.

Applying the transmission conditions (Equation4) at the interface x=0, we arrive at(8) θbcosλbb-θacosλaa=0,(8) (9) θbKbλbsinλbb+θaKaλasinλaa=0,(9)

which lead to the transcendental equation for eigenvalues:(10) Kbλbsinλbbcosλaa+Kaλasinλaacosλbb=0.(10)

For brevity, we denote the thermal diffusivities (cm2/sec) byκα=Kαραcαforαa,b.

Since we need the same time dependence in both slabs, the condition(11) κbλb2=κaλa2(11)

is required. Hereby, the above equation enables us to derive the new eigenvalue condition from (Equation10):(12) Kbκbsinλbbcosλaa+Kaκasinλaacosλbb=0.(12)

To sum up, the two transcendental Equations (Equation11) and (Equation12) allow us to compute the eigenvalues λan,λbn for nN.

2.1. A computational approach

The Equations (Equation11) and (Equation12) spontaneously generate a system of nonlinear algebraic equations with 2N variables λbn and λan for n0,...,N. From the numerical point of view, the common way is to use the Newton method which is known as the one-step iterative scheme with quadratic convergence [Citation17] for solving equations of the form Fx=0. Besides, so far there have been several effective approaches such as the Newton-Simpson’s method of [Citation18] and the improved Newton’s method by the Closed-Open quadrature formula of [Citation19]. However, there would be some basic impediments if we exploit such approaches. Noticeably, the transcendental system (Equation11) and (Equation12) has infinitely many roots and furthermore, we cannot ensure that the obtained solutions correspond one-to-one to the indexes n, which means that the result may be missed out or exceeded. Thus, by varying initial guesses x0, it would produce a few confusing results. For instance, taking Kα=κα=1, we easily obtain the family of solutions(13) λαk=kπa+bforkN,(13)

then with a=3,b=5 and the two initial pointsx0:=01N-1N01N-1NTandx1:=0.5x0,

the eigenvalues governed by the standard Newton method (for N=50) are showed in Figure for comparison. It is apparent that there are many differences between the two figures. Basically, the actual final eigenvalue with N=50 is approximately 19.63495 from (Equation13), while the others are, respectively, 49.87278 and 26.70353 in Figure . The results also indicate that with the guess x0 there are 48 admissible values obtained by the elimination of equal values whilst there are only 42 values for the guess x1.

Figure 1. Numerical eigenvalues for the explicit case (Equation13) using the standard Newton method with different initial guesses x0 and x1 at the 10th iteration and with the stopping constant δ=10-10.

Figure 1. Numerical eigenvalues for the explicit case (Equation13(13) λαk=kπa+bfork∈N,(13) ) using the standard Newton method with different initial guesses x0 and x1 at the 10th iteration and with the stopping constant δ=10-10.

In order to overcome this issue, we apply the built-in function fzero of MATLAB to several different initial points over some pre-defined range. Thanks to the structure of the transcendental Equations (Equation11) and (Equation12), we are capable of reducing the system to the following equation:(14) Kbκbsinλbbcosκbκaλba+Kaκasinκbκaλbacosλbb=0,(14)

where we have derived from (Equation11) that(15) λa=κbκaλb(15)

and substituted this to (Equation12). Once λb is found, it is straightforward to compute the corresponding λa from (Equation15). The algorithm to compute the first N non-negative eigenvalues λbn from (Equation14) is described as follows:

Figure 2. Numerical eigenvalues using the MATLAB-built-in function fzero.

Figure 2. Numerical eigenvalues using the MATLAB-built-in function fzero.

Remark 1:

A complicated case is provided in Figure (b) where the eigenvalues cannot be solved explicitly. The materials we illustrate are the copper versus the molybdeum, which was experienced by Josell et al. [Citation20], providing κα0.838,0.339 and Kα3.42,1.05, respectively. Here, this case is governed by the same parameters of the previous explicit case (Equation13), i.e. N=50,a=3,b=5, with d=10 and δ=10-3. By comparing Figures and , it is easy to see the differences and how well the simple algorithm works. Especially, the computed final numerical eigenvalue, i.e. λb50=19.63495 in Figure (a), is very closed to the exact value. The only disadvantage of this approach is the long execution time that may rise in higher dimensions.

Now, let us consider the eigenfunctions ϕα. By choosing appropriate constants θa and θb in (Equation7), we obtain the following eigenfunctions:(16) ϕ~n(x)=ϕbn(x)=cosλbnx+bcosλbnbfor the left slab-b<x<0,(16)

and(17) ϕ~n(x)=ϕan(x)=cosλanx-acosλanafor the right slab0<x<a.(17)

Remark 2:

For different indexes mn, the eigenfunctions ϕ~m and ϕ~n are orthogonal in the sense that(18) Kbκbϕ~m,ϕ~nL2-b,0+Kaκaϕ~m,ϕ~nL20,a=0.(18)

On the other hand, we also obtain the same equality for the derivatives of such eigenfunctions. In fact, these functions are orthogonal in the sense thatKbϕ~m,ϕ~nL2-b,0+Kaϕ~m,ϕ~nL20,a=0.

Let us denoteNn:=Kbκbϕ~nL2-b,02+Kaκaϕ~nL20,a2=bKb2κb1cos2λbnb+aKa2κa1cos2λanafornN, Mn:=Kbϕ~nL2-b,02+Kaϕ~nL20,a2=bKb2λbn2cos2λbnb+aKa2λan2cos2λanafornN,

withN0:=Kbκbϕ~0L2-b,02+Kaκaϕ~0L20,a2=bKbκb+aKaκa,

andM0:=Kbϕ~0L2-b,02+Kaϕ~0L20,a2=0.

3. Main results

3.1. Solution by Fourier-mode and Instability

From the transcendental Equations (Equation11)–(Equation12) and the formulation of the eigenfunctions in (Equation16)–(Equation17), we are competent to construct the temperature solution throughout the two-slab system, denoted by Tx,t. Basically, the Fourier-mode for the solution of this system cannot be obtained in the same manner with the classical backward heat equation due to the non-trivial orthogonality (showed in Remark 2) in which we cannot make equivalent computations. Nevertheless, we are still able to compute it by following the structural property of the solution in the classical case.

By assigning λ¯n:=κbλbn2=κaλan2, the temperature solution (in the finite series truncated computational sense) would be of the following form:(19) TNx,t=n=0NCneλ¯ntf-tϕ~nx,x,t-b,a×t0,tf,(19)

for some NN and the data associated with the final state of temperature is provided by(20) TNx,tf=n=0NCnϕ~nx.(20)

Before investigating the instability of the solution as well as the cut-off projection, our first concern is to determine the Fourier-type coefficients Cn in (Equation19) for a fixed NN. It is worth noticing that due to the two-slab system being considered, such coefficients vary on each material, hence we denote by Cαn for αa,b to distinguish those different values.

According to the representation (Equation20), the algorithm to compute the coefficient Cαn is provided as follows: Let Jα be the number of discrete nodes in the material α. Suppose these are measured at the time tf, producing a set Tαεxj,tfj=0,Jα¯. Then the choice of a suitable Jα will be discussed later. We are led to the matrix-formed system AX=B where ARJα+1×RN+1, XRN+1, BRJα+1 are expressed by(21) A=ϕα0x0ϕα1x0ϕαNx0ϕα0x1ϕα1x1ϕαNx1ϕα0xJαϕα1xJαϕαNxJα,X=Cα0Cα1CαN,B=Tαεx0,tfTαεx1,tfTαεxJα,tf.(21)

Remark 3:

Multiplying the expression (Equation20) by the nth eigenfunction ϕ~nx using the inner product in L2-b,0 with the weight Kb/κb, then do the same using the inner product in L20,a, we obtain(22) ϑbCbn+ϑaCan=Nn-1ϑbKbκbTb·,tf,ϕbnL2-b,0+ϑaKaκaTa·,tf,ϕanL20,aforϑαR,(22)

which also guarantees that Cαn< for all nN.

The second objective concerns the instability of the temperature in this model manifested by the expression (Equation19). Due to the intrinsic property of the eigenvalues λ¯n which reads0λ¯1<λ¯2<...<λ¯n<...<limnλ¯n=,

the rapid escalation is indeed catastrophic and it would be the same issue that commonly occurs in previous works (e.g. [Citation15,Citation21Citation24]). Note that even if the coefficients Cn may decrease quickly, a large error between the Fourier-based solution (Equation19) and the exact solution always happens. Thus, any computational method are impossible to be applied. To be precise, we give the proof of the following theorem that the temperature through the system is exponentially unstable.

Theorem 3.1:

The truncated approximate solution TN is exponentially unstable over the L2-norm by the following estimate:(23) TN·,tL2-b,a2maxKaκa,Kbκb,1-1n=0NminCbn2,Can2e2λ¯ntf-tNn.(23)

Proof Consider the L2-norm of (Equation19) over -b,a:(24) TN·,tL2-b,a2=n=0NCneλ¯ntf-tϕ~n,n=0NCneλ¯ntf-tϕ~nL2-b,a=n=0Nm=0NCbnCbmeλ¯ntf-teλ¯mtf-tϕbn,ϕbmL2-b,0+n=0Nm=0NCanCameλ¯ntf-teλ¯mtf-tϕan,ϕanL20,amaxKaκa,Kbκb,1-1n=0Nm=0Neλ¯ntf-teλ¯mtf-t×minCbnCbm,CanCam×Kbκbϕbn,ϕbmL2-b,0+Kaκaϕan,ϕanL20,a.(24)

From the orthogonality relation (Equation18), we continue to estimate (Equation24) and obtain (Equation23).

Remark 4:

One needs Jα=N to ensure the unique solvability of the linear system (Equation21). However, the question concerning how big the constant N should be taken severely hinders the choice of Jα. Therefore, the cut-off projection that we consider in the next subsection is important not only in regularization, but also in providing a precise value of Jα needed at each noise level.

3.2. The cut–off projection

This part of work aims at applying the cut-off projection to deal with the aforementioned instability as well as proposes an effective technique to compute the temperature solution Tx,t. The idea is to construct a projection in which a suitable high frequency N is appropriately selected depending on the measurement error ε0,1. In other words, for each ε>0 we now introduce the cut-off mapping Pε such that(25) PεT(x,t)=nΘ(ε)Cneλ¯ntf-tϕ~nx,(25)

where Θ(ε):=nN:λ¯nNε is the set of admissible frequencies and Nε:=Nε>0 satisfying limε0Nε= plays the role of the regularization parameter. Henceforward, the algorithm to compute the Fourier coefficients Cαn in Section 3.1 is now active with Jα=Nε.

As the main objectives of this section, we prove the following theorems to analyse the stability and the convergence of the (approximate) regularized solution governed by the representation (Equation25).

Theorem 3.2:

For each ε>0, the regularized solution depends continuously on the final temperature in the L2-norm by the following estimate:(26) PεT·,tL2-b,a2minKbκb,Kaκa,1-1Nεe2Nεtf-t×KbκbTb·,tfL2-b,02+KaκaTa·,tfL20,a2.(26)

Proof Once again, we use the definition of the norm induced by inner product in L2-b,a and then the proof is straightforward with the non-trivial orthogonality (Equation18). Indeed, expressing the projection onto the solution gives usPεT·,tL2-b,a2:=PεT·,t,PεT·,tL2-b,a=nΘεmΘεC¯bntC¯bmtϕbn,ϕbmL2-b,0+nΘεmΘεC¯antC¯amtϕan,ϕamL20,a,

where we have denoted by C¯αnt=Cαneλ¯ntf-t for αa,b. This leads to(27) PεT·,tL2-b,a2minKbκb,Kaκa,1-1×nΘεKbC¯bn2tκbϕbnL2-b,02+KaC¯an2tκaϕanL20,a2.(27)

We then combine the elementary inequality a1a32+a2a42a1+a2a32+a42 for a1,a20 and a3,a4R with the properties in Remark 1 and (Equation22), to yield from (Equation27) thatPεT·,tL2-b,a2minKbκb,Kaκa,1-1nΘεNn-1e2λ¯ntf-t×KbκbTb·,tf,ϕbnL2-b,02+KaκaTa·,tf,ϕanL20,a2.

Using the Cauchy-Schwarz inequality and the basic inequality a12a32+a22a42a12+a22a32+a42 for all ajR,j1,...,4 we getKbκbTb·,tf,ϕbnL2-b,02+KaκaTa·,tf,ϕanL20,a2NnKbκbTb·,tfL2-b,02+KaκaTa·,tfL20,a2,

and by definition of the set Θε, we arrive at (Equation26)

Theorem 3.2 shows us that the cut-off projection solution is bounded at each noise level, and such an interesting estimate proves that the regularized solution is stable even when the final temperature is corrupted by noise. Furthermore, its appearance enables us to derive the L2-type estimate between the solutions PεTx,t and PεTεx,t (i.e. we apply the projection method to design the solution associated with the terminal measured data), which we refer to the following lemma.

Lemma 3.3:

Let PεTεx,t be the regularized solution associated with the square–integrable measured data Tαε·,tf satisfying (Equation5). Then the following estimate holds:(28) PεT·,t-PεTε·,tL2-b,a2minKbκb,Kaκa,1-1Nεe2Nεtf-tKbκb+Kaκaε2.(28)

Proof Similar to the proof of Theorem 3.2, we arrive atPεT·,t-PεTε·,tL2-b,a2minKbκb,Kaκa,1-1Nεe2Nεtf-t×KbκbTb·,tf-Tbϵ·,tfL2-b,02+KaκaTa·,tf-Taϵ·,tfL20,a2.

The proof of the lemma is completed by applying the terminal condition (Equation5) to the above estimate.

At this moment, it is possible for us to prove the convergence result. From the triangle inequality, it followsT·,t-PεTε·,tL2-b,aPεT·,t-PεTε·,tL2-b,a+T·,t-PεT·,tL2-b,a,

which then questions us how to estimate the norm T·,t-PεT·,tL2-b,a.

We remind that in the Fourier-based sense, the quantity N in the representation (Equation19) tends to infinity and that, consequently, requires the regularization parameter to satisfy limε0Nε=. To compare the difference between Tx,t and PεTx,t, we suppose that (Equation19) reaches the most ideal case, i.e. it can be represented by(29) Tx,t=n=0Cneλ¯ntf-tϕ~nx.(29)

Subtract (Equation29) and (Equation25), we thus see thatTx,t-PεTx,t=nΘεCneλ¯ntf-tϕ~nx.

Therefore, very much in line with the computations in the proof of Theorem 3.2, the L2-norm of the above difference can be upper-bounded by(30) T·,t-PεT·,tL2-b,a2minKbκb,Kaκa,1-1×nΘεKbC¯bn2tκbϕbnL2-b,02+KaC¯an2tκaϕanL20,a2.(30)

On the other side, we compute thatxTx,t=n=0Cneλ¯ntf-tϕ~x,

which givesxT·,tL2-b,a2=n=0m=0C¯bntC¯bmtϕbn,ϕbmL2-b,0+n=0m=0C¯antC¯amtϕan,ϕamL20,a,

and using (Equation18) we obtain(31) xT·,tL2-b,a2maxKa,Kb,1-1×n=0KbC¯bn2tϕbnL2-b,02+KaC¯an2tϕanL20,a2.(31)

Furthermore, it reveals from (Equation30) that(32) T·,t-PεT·,tL2-b,a2minKbκb,Kaκa,1-1×nΘελ¯n-1KbC¯bn2tϕbnL2-b,02+KaC¯an2tϕanL20,a2minKbκb,Kaκa,1-1Nε-1×nΘεKbC¯bn2tϕbnL2-b,02+KaC¯an2tϕanL20,a2.(32)

Combining (Equation31) and (Equation32) we observe that(33) T·,t-PεT·,tL2-b,a2minKbκb,Kaκa,1-1maxKa,Kb,1Nε-1xT·,tL2-b,a2.(33)

At this stage, we conclude that the error T·,t-PεT·,tL2-b,a will tend to zero as ε0 if the gradient of the temperature solution is bounded over L2-b,a. By using the triangle inequality and the errors in (Equation28) and (Equation33), we easily obtain the estimate T·,t-Tε·,tL2-b,a. We thus have the following theorem.

Theorem 3.4:

Assume that T·,tH1-b,a for all tJ and let PεTεx,t be the regularized solution associated with the square–integrable measured data Tαε·,tf. Then the following estimate holds:(34) T·,t-PεTε·,tL2-b,aminKbκb,Kaκa,1-12Kbκb+KaκaNε12eNεtf-tε+minKbκb,Kaκa,1-12maxKa,Kb,112Nε-12xT·,tL2-b,a.(34)

Remark 5:

In general, the error estimate (Equation34) can be rewritten asT·,t-PεTε·,tL2-b,aCNε12eNεtf-tε+Nε-12,

in which C is a generic positive constant. Then by choosing Nε=βlnε-γ/tf for β0,1 and γ(0,1] we obtain thatT·,t-PεTε·,tL2-b,aCβtfln12ε-γε1-γ1-ttfβ+tfβln-12ε-γ.

Consequently, PεTε is a good approximation of T as the small noise ε tends to zero. On top of that, this rate of convergence is uniform in time.

It is worth pointing out that due to the structural property of the eigen-elements, one may deduce that for pN and mn the functions ϕ~m(2p) and ϕ~n(2p) are orthogonal with weights Kb/κb and Ka/κa whilst the functions ϕ~m(2p+1) and ϕ~n(2p+1) are orthogonal with weights Kb and Ka. On the other hand, we see the following relationsKbκbϕbn(2p)L2-b,02=λ¯n-1Kbϕbn(2p+1)L2-b,02,Kaκaϕan(2p)L20,a2=λ¯n-1Kaϕan(2p+1)L20,a2.

Combining these two relations with the fact that ϕαn(2p)=-1pλαn2pϕαn leads us to the L2-norm equivalence between ϕαn and their high-order derivatives. Therefore, if we wish to obtain the error estimate over the space Hp-b,a, the requirement will be T·,tHp+1-b,a. Note that to tackle the error estimate in this context, only the difference between the high-order derivatives of the exact solution and the regularized is needed. Such result can be found explicitly in [Citation21].

4. Numerical tests

Following the example from Remark 1, in this section we consider three examples including the smooth, piecewise smooth and discontinuous cases in a short-time observation J=[0,0.1]. Note that herein we do not know the exact solution, but at some point we can generate artificial data to see how the approximation works. All computations were done using MATLAB, and only using 5 digit floating point arithmetic. Recall that with b=5cm and a=3cm, the materials we illustrate herein are the copper versus the molybdeum providing κα0.838,0.339 and Kα3.42,1.05, respectively. We also select β=0.05, γ=0.5 and fix 20 discrete spatial points for each slab throughout this illustration. Accordingly, we are able to compute the regularization parameter Nε which has been chosen in Remark 5.

4.1. Example 1

In this part, the stability of the approximate solution is investigated to prove the approximation as ε0. Practically, we merely have the measured final data given byTbεxj,tf=cosλb1xj+ba+bcosλb1be-tf2+randxjε,xj(-b,0),Taεxj,tf=cosλa1xj-aa+bcosλa1ae-tf2+randxjε,xj(0,a),

with randxj2maxa,b being a random number generator.

4.2. Example 2

In this example, we assume that the exact initial data are known, albeit of the unknown exact temperature solution. They are given by piecewise smooth functions, as follows:Tbx,0=x,x-b2,0-b2,x-b,b2andTax,0=-a2,xa2,a,-x,x0,a2.

Figure 3. The instability illustration with ε=10-2 at a cut-off high-frequency (selected randomly) N=50. (Example 4.1).

Figure 3. The instability illustration with ε=10-2 at a cut-off high-frequency (selected randomly) N=50. (Example 4.1).

Figure 4. The stability illustration of the proposed cut-off projection, showing the approximate temperature throughout the two-slab system at t0=0 with ε10-2,10-4,10-6. (Example 4.1).

Figure 4. The stability illustration of the proposed cut-off projection, showing the approximate temperature throughout the two-slab system at t0=0 with ε∈10-2,10-4,10-6. (Example 4.1).

Figure 5. The approximate temperature throughout the two-slab system with ε=10-2 in a short-time observation. (Example 4.1).

Figure 5. The approximate temperature throughout the two-slab system with ε=10-2 in a short-time observation. (Example 4.1).

In order to observe the approximation PεTεx,t to Tx,t, the following steps are carried out:

(1)

Find the final values of Tαx,tf from the forward problem which admits the solution expressed as: for MN (for comparison, we take M=Nε) Tαx,t=n=0MC~αne-λ¯ntϕαnx, and to find C~αn, we arrive at Tαx,t0=n=0MC~αnϕαnx.

(2)

Obtain the values Tαεx,tf from those values of Tαx,tf by perturbing them, as follows: Tαε(xj,tf)=Tα(xj,tf)+randxjε=n=0MC~αne-λ¯ntfϕαn(xj)+randxjε, with randxj(2b)-1/4 in order to fulfill (Equation5).

(3)

Pass the resulting values to the approximate solution PεTεx,t in accordance with (Equation25) to find back the initial data Tαεx,t0.

Similar steps are also used in Example 4.3 with the same type of noise. For our way of choosing M=Nε in Step 1, this make things convenient since we can both obtain numerical results for the discrete exact solution and approximate solution with the same dimensions.

Figure 6. The comparison between the (artificial) exact initial data and the computed approximations at t0=0 with ε10-2,10-4,10-6. (Example 4.2).

Figure 6. The comparison between the (artificial) exact initial data and the computed approximations at t0=0 with ε∈10-2,10-4,10-6. (Example 4.2).

4.3. Example 3

We finalize this section by testing the example of Parker et al. [Citation25] in which the prescribed initial data are provided byTbx,0=Qρbcbσ,x-b,0,Tax,0=0,x0,a,

where Q is the total heat energy in the initial pulse (cal·cm-2) and the depth σ at the front surface x=-b is performed as a very small constant compared to the length a+b. These two dimensionless parameters will be selected as Q=5 and σ=10-3.

Figure 7. The comparison (using log-scale for the y-axis) between the (artificial) exact initial data and the computed approximations at t0=0 with ε10-2,10-4,10-6. (Example 4.3).

Figure 7. The comparison (using log-scale for the y-axis) between the (artificial) exact initial data and the computed approximations at t0=0 with ε∈10-2,10-4,10-6. (Example 4.3).

Table 1. Numerical results at t0=0 with various different points in space.

4.4. Comments on numerical results

When approximating the temperature at t0=0 for various amounts of noise ε10-2,10-4,10-6, we first observe in Figure that the stability of solution ruins drastically for ε=10-2 for Example 4.1. Since in that example we do not have an analytical exact solution, the convergence of the scheme is affirmed by its stability (inspired from the Lax equivalence theorem). Thereby, it can be seen in Figure that the approximation by the cut-off projection almost remains unchanged. For instance, at (x,t)=(0,0) the approximation in the material a for ε=10-2 yields 0.11787, whilst the obtained values ε=10-4,10-6 are 0.11553 and 0.11556, respectively. As a result, the stable approximate solution has been depicted in Figure .

Aside from the first example, the measured final data in the next two examples is established from the artificial initial data by solving the direct problem. Then the approximations can be applied to find back the approximate initial data.

Figure shows the accuracy of reconstructing the (artificial) exact initial data by the cut-off projection. In fact, we can observe a steady approach of the approximations (nodes) towards the exact solution (black line) as ε gets from 10-2 to 10-6. In line with that, Figure also confirms a good approximation in the discontinuous case (Example 4.3).

In Table , numerical results with several different spatial points are provided. It is apparent that the numerical solutions in the nonsmooth and discontinuous cases are less stable (accurate) than that of the smooth example. It is simply because of the well-known Gibbs phenomenon, and that the reconstructed data near the nonsmooth and discontinuous points are not well-approximated. Nevertheless, taking into consideration the approximate values in Table at the boundaries as well as the instability in Example 4.1 (Figure ), the results in Example 4.2 and Example 4.3 are acceptable.

5. Discussion

5.1. Non-homogeneous case

We have successfully applied the cut-off method to solve the heat transfer problem backward in time in a 1D two-slab system with perfect contact. This system can be solved similarly when the heat source or sink is applied to the PDE. Indeed, let Fα:=Fαx,tL2-b,a represent the heat source or sink. The system (Equation1)–(Equation2) now reads(35) ρbcbTbt=Kb2Tbx2+Fb,-b<x<0,tJ,(35) (36) ρacaTat=Ka2Tax2+Fa,0<x<a,tJ.(36)

As in Section 2, the system (Equation35)–(Equation36) together with the boundary conditions (Equation3)–(Equation4) also lead to the transcendental conditions (Equation10)–(Equation11), which allow direct computation of the eigenvalues λan,λbn and the eigenfunctions ϕan,ϕbn of the Sturm-Liouville equation. The temperature solution (in the finite series truncated computational sense) of the system (Equation35),(Equation36), (Equation3), (Equation4) has the following form:(37) T(x,t)=n=0NCneλ¯n(tf-t)-ttfDn(s)eλ¯n(s-t)dsϕ~n(x),(x,t)[-b,a]×[t0,tf],(37)

for some NN. Here, we follow the notation λ¯n=κbλbn2=κaλan2 in Section 3, whilst the formula of ϕ~n has already been introduced in (Equation16)–(Equation17). The coefficients Cn can be derived as in Section 3 from the final temperature condition (Equation20), while the algorithm for the time-dependent coefficient Dαn is rather difficult and can only be computed when Cαn are known. First of all, let us return to Equations (Equation35)–(Equation36) and then after plugging the representation (Equation37) into each quantity, we see thatTαt=n=0NDαnt+λ¯nttfDαnseλ¯ns-tds-Cαneλ¯ntf-tϕαn,κα2Tαx2=-n=0Nλ¯nCαneλ¯ntf-t-ttfDαnseλ¯ns-tdsϕαn,Tαt=κα2Tαx2+1ραcαFα,

where we have used the fundamental differentiation under the integral sign. Therefore, it follows that(38) n=0NDαntϕαnx=1ραcαFαx,tforαa,b.(38)

As the second part of work, fix t in (Equation38) and once again, with the nodes xjj=0,Jα¯ it enables us to formulate the matrix-formed system AX=B then solve it through. Thereby, we have sufficient values for approximating the integral ttfDnseλ¯ns-tds, and that provides the way to numerically compute the temperature Tx,t by discrete nodes on the plane consisting of the time and spatial variables.

It is worthy noticing that Fα should satisfy the following condition so that (Equation38) is solvable:1ρacaFa(x,t),ϕan(x)L2(0,a)=1ρbcbFb(x,t),ϕbn(x)L2(-b,0)for allnN.

The above condition is a necessary condition for Fα. We do not aim at finding the sufficient condition to guarantee the existence of solution of the system (Equation35)–(Equation36) in this manuscript.

5.2. The bilayer system in 2D

In this subsection, we generalize the homogeneous problem to two dimensional space. We consider the backward heat transfer problem between two plates: the left plate is the rectangle [-b,0]×[0,c], while the right plate is [0,a]×[0,c]. The heat distribution in the time interval [t0,tf] is described by:(39) ρbcbTbt=KbΔTb,x,y-b,0×0,c,tJ,(39) (40) ρacaTat=KaΔTa,x,y0,a×0,c,tJ.(40)

Unlike the problem (Equation1)–(Equation2), we have the boundary conditions on each side of the rectangle and also at the intersection of them, i.e. x,y0×[0,c],(41) Tbx=0atx=-b,Tax=0atx=a,tJ,(41) Tby=0atx,y[-b,0]×0,c,Tay=0atx,y[0,a]×0,c,tJ, (42) Tb=Ta,KbTbx=KaTaxatx,y0×[0,c],tJ.(42)

Using the same approach as in Section 2, we can construct the eigen-elements as the solution of(43) Δϕαx,y+λα2ϕαx,y=0forαa,b,(43)

with the assumption that the time dependence of each plates is the same, and that ϕα inherits the boundary conditions (Equation41)–(Equation42).

We apply the separation variable technique for (Equation43) by assuming that ϕαx,y can be expressed as ϕαx,y=XαxYαy. Then, (Equation43) becomes(44) XαXα+λα2=-YαYα=μα2.(44)

Direct computation yields that Yα(y)=ηαcos(μαy), where ηα=2c is a constant and μα=μ=mπc for mN. Since we are solely interested in the case λα2>μα2, we can assume that να2=λα2-μα2. In the same manner with Section 2, we obtainXa=θacosνax-a;Xb=θbcosνbx+b,

where θa,θb are constants. The transcendental condition (Equation10) now reads(45) Kbνbsinνbbcosνaa+Kaνasinνaacosνbb=0.(45)

Similarly, a 2D version of condition (Equation11) is(46) κaνa2+μa2=κbνb2+μb2.(46)

The temperature solution of the system (Equation39)–(Equation40) together with the boundary conditions (Equation41)–(Equation42) isTαx,y,t=m,n0Cαmneκαναn2+μαm2tf-tXαnxYαmy,

where the coefficient Cαmn can be derived from the temperature distribution at the final time:Tα(x,y,tf)=m,n0CαmnXαnxYαmy.

For the time being, the algorithm proposed in Section 2 can be applied again in searching for the eigenvalues ναnnN which are solutions to (Equation45) and (Equation46). Note from (Equation46) that if the thermal diffusivities of two plates are identical (i.e. κa=κb), it turns back to solving the system (Equation14)–(Equation15).

Applying again the cut-off projection, our regularized solution in the 2D two-slab system is given byTε(x,y,t)=m,nΘ(ε)Cmneλ¯mn(tf-t)Xn(x)Ym(y),

associated with the terminal measured data Tαεx,y,tf satisfying(47) Ta·,tf-Taε·,tfL2(0,a)×(0,c)+Tb·,tf-Tbε·,tfL2(-b,0)×(0,c)ε.(47)

Here, the set of admissible frequencies is defined by Θ(ε):=(m,n)N2:λ¯mnNε with λ¯mn:=κaνan2+μam2=κbνbn2+μbm2.

Taking again κα0.838,0.339 and Kα3.42,1.05 as in Section 4, we test the cut-off projection with a simple 2D problem in a normalized square two-slab system, i.e. a=b=c=1, and with tf=0.1, β=0.01, γ=1.

In this test, since the values of thermal diffusivities are known, we are allowed to write from (Equation46) that(48) νa=κbκaνb2+κbκa-1μ2,(48)

and thus get the transcendental condition in (Equation45) that(49) Kbνbsin(νbb)cosaκbκaνb2+κbκa-1μ2+Kaκbκaνb2+κbκa-1μ2sinaκbκaνb2+κbκa-1μ2cos(νbb)=0.(49)

As we can see from (Equation49) , the eigenvalues cannot be computed directly. In order to depict the behavior of those eigenvalues, we demonstrate a simple numerical example by giving an exact smooth form of initial data, viz.Tbx,y,0=cosπx+bcosπy,Tax,y,0=cosπx-acosπy.

Notice that these are the initial values for the forward problem. Following the steps mentioned in Section 4.2, we set the limit of our summation in the expression of Fourier form solution, i.e. (m,n)Θ(ε) and calculate the eigenvalues according to (Equation48) and (Equation49). In Figure , we compute several numerical eigenvalues driven by (Equation48) and (Equation49) by using the built-in function fzero from MATLAB (this has already been introduced in Section 2).

In this example we consider the system at y0=0. The algorithm for this 2D case is similar to that for the 1D case (see in Section 4). Basically, the following steps are carried out:

(1)

Find the final values from the forward problem Tα(x,y0,t)=m,nΘ(ε)C~αmne-λ¯ntXαn(x)Ym(y0), and to find C~αmn, we solve a linear system AX=B where A=Xα0(x0)Y0(y0)Xα0(x0)Y1(y0)Xαn(x0)Ym(y0)Xα0(x1)Y0(y0)Xα0(x0)Y1(y0)Xαn(x1)Ym(y0)Xα0(xJα)Y0(y0)Xα0(xJα)Y1(y0)Xαn(xJα)Ym(y0)R(Jα+1)2, X=C~α00C~α01C~αnmRJα+1,B=Tα(x0,y0,t0)Tα(x1,y0,t0)Tα(xJα,y0,t0)RJα+1. In this case we choose Jα=Θ(ε)-1 where Θ(ε) is the cardinality of Θ(ε).

(2)

Obtain the values Tαεx,y0,tf from those values of Tαx,y0,tf by perturbing them, as follows: Tαε(xj,y0,tf)=Tα(xj,y0,tf)+randxjε=m,nΘ(ε)C~αmne-λ¯ntfXαn(x)Ym(y0)+randxjε, with randxj((a+b)c)-1/2.

(3)

Perform the proposed method on generated data to find back the initial data Tαεx,y0,t0.

Figure 8. Numerical eigenvalues λα2=να2+μα2 for the two-slab system in 2D at ε=10-6.

Figure 8. Numerical eigenvalues λα2=να2+μα2 for the two-slab system in 2D at ε=10-6.

Figure 9. The comparison between the (artificial) exact initial data and the computed approximations at y0=0,t0=0 with ε10-2,10-4,10-6. (Example for a two-slab system in 2D).

Figure 9. The comparison between the (artificial) exact initial data and the computed approximations at y0=0,t0=0 with ε∈10-2,10-4,10-6. (Example for a two-slab system in 2D).

We have drawn in Figure the numerical results of the exact initial data and the computed approximations based on our proposed cut-off projection at y0=0 and at t0=0. We can easily observe that these approximations (colored nodes) are closer to the exact initial data (black line) as ε goes from 10-2 to 10-6. This corroborates that our theoretical analysis works well in this 2D example and thus implies the applicability of the proposed method in higher dimensions.

5.3. Boundary conditions

So far, we merely have considered the two-slab problem under idealistic conditions, such as the system has a perfect contact, there is no heat entering or escaping at the boundary of the two slabs. More realistic conditions should be more interesting, for example, an imperfect contact or Robin-type condition motivated by Newton’s Law of Cooling. However, the more complicated boundary conditions, the more severe difficulties of computation we have, especially for the eigen-elements. We shall leave detailed studies of such realistic models for further research.

Acknowledgements

V.A.K. acknowledges Prof. Errico Presutti (GSSI, Italy) for his invaluable help during the PhD years. The authors wish to express their sincere thanks to the anonymous referees and the associate editors for many constructive comments leading to the improved versions of this paper.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work is supported by the Institute for Computational Science and Technology (Ho Chi Minh city) under project named ‘Some ill-posed problems for partial differential equations’ [312/Q -KHCNTT].

References

  • Lattès R, Lions J. Méthode de Quasi-réversibilité et Applications. Paris: Dunod; 1967.
  • Buzbee B, Carasso A. On the numerical computation of parabolic problems for preceding times. Math Comput. 1973;27(122):237–266.
  • Clark G, Oppenheimer S. Quasireversibility methods for non-well-posed problems. Elect J Differ Equ. 1994;1994(8):1–9.
  • Denche M, Bessila K. A modified quasi-boundary value method for ill-posed problems. J Math Anal Appl. 2005;301:419–426.
  • Hào D, Duc N, Sahli H. A non-local boundary value problem method for parabolic equations backward in time. J Math Anal Appl. 2008;345:805–815.
  • Seidman T. Optimal filtering for the backward heat equation. SIAM J Numer Anal. 1996;33(1):162–170.
  • Trong D, Tuan N. A nonhomogeneous backward heat problem: regularization and error estimates. Elect J Differ Equ. 2008;2008(33):1–14.
  • Lesnic D, Elliott L, Ingham D. An iterative boundary element method for solving the backward heat conduction problem using an elliptic approximation. Inverse Probl Eng. 1998;6(4):255–279.
  • Johansson B, Lesnic D, Reeve T. A method of fundamental solutions for radially symmetric and axisymmetric backward heat conduction problems. Int J Comput Math. 2012;89(11):1555–1568.
  • Hào D, Duc N, Lesnic D. Regularization of parabolic equations backward in time by a non-local boundary value problem method. IMA J Appl Math. 2010;75(2):291–315.
  • Lesnic D. The decomposition method for forward and backward time-dependent problems. J Comput Appl Math. 2002;147(1):27–39.
  • Mei C, Vernescu B. Homogenization methods for multiscale mechanics. Singapore: World Scientific Publishing; 2010.
  • Ternat F, Daripa P, Orellana O. On an inverse problem: recovery of non-smooth solutions to backward heat equation. Appl Math Model. 2012;36:4003–4019.
  • Ternat F, Orellana O, Daripa P. Two stable methods with numerical experiments for solving the backward heat equation. Appl Numer Math. 2011;61:266–284.
  • Regińska T, Regiński K. Approximate solution of a Cauchy problem for the Helmholtz equation. Inverse Probl. 2006;22:975–989.
  • Tuan N, Khoa V, Minh M, et al. Reconstruction of the electric field of the Helmholtz equation in three dimensions. J Comput Appl Math. 2017;309:56–78.
  • Ortega J, Rheinboldt W. Iterative solution of nonlinear equations in several variables. New York: Academic Press; 1970.
  • Cordero A, Torregrosa J. Variants of Newton’s method using fifth-order quadrature formulas. Appl Math Comput. 2007;190:686–698.
  • Noor M, Waseem M. Some iterative methods for solving a system of nonlinear equations. Comput Math Appl. 2009;57(1):101–106.
  • Josell D, Cezairliyan A, van Heerden D, et al. Thermal diffusion through multilayer coatings: theory and experiment. Nanostruct Mater. 1997;9:727–736.
  • Nam P, Trong D, Tuan N. The truncation method for a two-dimensional nonhomogeneous backward heat problem. Appl Math Comput. 2010;216:3423–3432.
  • Tuan N, Au V, Khoa V, et al. Identification of the population density of a species model with nonlocal diffusion and nonlinear reaction. Inverse Probl. 2017;33(5):055019.
  • Mera N, Elliott L, Ingham D, et al. An iterative boundary element method for solving the one-dimensional backward heat conduction problem. Int J Heat Mass Transfer. 2001;44(10):1937–1946.
  • Johansson B, Lesnic D, Reeve T. A comparative study on applying the method of fundamental solutions to the backward heat conduction problem. Math Comput Model. 2011;54(1–2):403–416.
  • Parker W, Jenkins R, Butler C, et al. Flash method of determining thermal diffusivity, heat capacity and thermal conductivity. J Appl Phys. 1961;32(9):1679–1684.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.