516
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Recovery of a degenerate space-dependent heat capacity

&
Pages 3214-3226 | Received 16 Mar 2021, Accepted 05 Sep 2021, Published online: 27 Sep 2021

Abstract

We are concerned with the reconstruction of the heat capacity coefficient of a one-dimensional heat equation from a sequence of observations of solutions at a single fixed point. This new direct method reconstructs the principal eigenfunction by extracting its Fourier coefficients from the observations of the solution. The algorithm presented here is an alternative to Krein's inverse spectral theory, minimization and iterative procedures. Error estimates and numerical experiments are presented.

2010 Mathematics Subject Classifications:

1. Introduction

We are concerned with the reconstruction of the heat capacity coefficient ρL2(0,1) appearing in the one-dimensional heat equation (1) {ut(x,t)=1ρ(x)uxx(x,t)0<x<1,t>0ux(0,t)hu(0,t)=0t>0ux(1,t)+Hu(1,t)=0t>0u(x,0)=f(x)x(0,1)(1) from observations of a sequence of temperatures taken at an arbitrary single point, i.e. u(b,t) where b is a fixed but arbitrary point in [0,1]. Here the real values h and H in the boundary conditions describe the constant heat transfer occurring at the endpoints, and fC(0,1) is the initial condition.

We recall that in [Citation1] the idea was to extract the complete spectral data {λn,αn}n1 of the operator ρ1(x)xx, from the observation of a single solution on the boundary, i.e. u(0,t) and u(1,t) based on Krein's inverse spectral theory of the string [Citation8,Citation10]. The main difficulty of applying Krein's theory is its requirement of having the full and complete spectral data, which is impossible to extract in practice.

This note implements numerically the new method introduced in [Citation4], which applies to multidimensional partial differential equations, and presents an alternative to methods such as the Gelfand–Levitan or Krein inverse spectral theories, optimal control, minimization methods, Newton's and fixed points methods, see [Citation2,Citation6,Citation7,Citation9,Citation11,Citation12]. It is based on a simple idea of computing the Fourier coefficients of the first eigenfunction which allows for a simple approximating scheme of ρ. However, because ρ is unknown, the Fourier coefficients of the principal eigenfunction are not available as in [Citation4,Citation5], and so the analysis needs to be modified which is the purpose of this note. The main result can be stated as follows:

Theorem 1.1

Let {ψj}jN be a basis of L2(0,1) that we take for the initial condition u(x,0)=ψj(x) for x(0,1),jN and let b(0,1). If h+H+hH0, then we can reconstruct 0ρL2(0,1) from the map ψj{u(b,k)}kN. In the Neumann case, i.e. h = H = 0, then we also need the mean value of 01ρ(x)dx.

We now briefly describe the content of the paper. Notations are in Section 2, and in Section 3, the idea on how to reconstruct ρ by computing the first eigenvalue λ1 and φ1(x)ρ(x) from the observation of the solution at one point is given. By solving an auxiliary boundary problem we then find ϕ1 and deduce the sought coefficient ρ. The Dirichlet and the Neumann cases are examined respectively in Section 4 and Section 5. Convergence is proved in Section 6 while the algorithm is exposed in Section 7. Numerical examples are provided in the last section.

2. Preliminaries

By the Fourier method, the solution of Equation (Equation1) can be written as (2) u(x,t)=n1cn(f)eλntφn(x),(2) where the Fourier coefficients are given by (3) cn(f)=01f(x)φn(x)ρ(x)dx,nN(3) and φn are eigenfunctions of the Sturm–Liouville operator A defined by (4) {Aφn(x)=1ρ(x)φn(x)=λnφn(x),0<x<1,φn(0)hφn(0)=0andφn(1)+Hφn(1)=0,(4) where 0ρ(x)L2(0,1), and h,HR. It is also known that A, defined by (Equation4), is a self-adjoint operator acting in the Hilbert space Lρ2(0,1)={f measurable: fρ2=01|f(x)|2ρ(x)dx<}and its spectrum is discrete and simple. Since φn(x)=λnρ(x)φn(x) and ρL2(0,1), then φnH2(0,1), and {φn}n1 forms an orthonormal basis of Lρ2(0,1) normalized by φnρ=1.

In the fifties, Krein used the inverse spectral theory of the string operator which is generated by the differential operator ddM(x)d+dx+,where dMis a Stieljtes measure defined by a nondecreasing function M, which represents the mass of a string, and d+dx+ is the right derivative at x, see [Citation8,Citation10] for the spectral theory of the string. When M(x) is an absolutely continuous function, the string operator reduces to (Equation4) and Af=gLρ2(0,1) is understood as f(x)=ax+b+0x(xη)g(η)ρ(η)dη,and thus, when ρ vanishes on a subinterval, the solution f is a linear function.

We assume that fC(0,1) so the Fourier coefficients cn(f)in (Equation3) are well defined although the weight ρL2(0,1) is yet unknown. We still have {cn(f)}n12 and f(x)=n=1cn(f)φn(x)holdsinLρ2(0,1).

Remark 2.1

The sum in Equation (Equation2) provides a representation of the solution only, although all its terms are unknown. The objective is to use the observation u(b,t) to extract the Fourier coefficients cn and λn, whenever they are present, as the φn(b) can vanish. Obviously, if a coefficient is not present in the sum then it cannot be extracted.

3. Reconstructing ρ

In principle, to reconstruct ρ it is enough to reconstruct the first eigenfunction φ1, since ρ is a solution of the algebraic equation (5) φ1(x)=λ1ρ(x)φ1(x),x(0,1).(5) To do so we only need the first eigenvalue λ10, as we already have φ1(x)0 for all x(0,1). To reconstruct φ1, we could use a basis of  L2(0,1), say {ψk}k1, to write (6) φ1(x)=k1(01φ1(ξ)ψk(ξ)dξ)ψk(x)(6) with the hope that we can read the coefficients 01φ1(x)ψk(x)dx from the solution u(b,t), as done in [Citation4]. Unfortunately, the observation u(b,t), in (Equation2) contains a different coefficient of the exponential, namely, (7) (01φ1(x)ψk(x)ρ(x)dx)φ1(b)(7) with initial condition f=ψk. In other words, even if we could extract the whole sequence {(01φ1(ξ)ψk(ξ)ρ(ξ)dξ)φ1(b)}k1,from (Equation2), we still cannot reconstruct φ1 by (Equation6), but (8) φ1(x)ρ(x)φ1(b)=k1(01φ1(ξ)ψk(ξ)ρ(ξ)dξ)φ1(b)ψk(x)=k1c1(ψk)φ1(b)ψk(x)inL2(0,1)(8) where we used the notation (Equation3). To proceed further we now show how to extract the coefficients (Equation7) from (Equation2).

3.1. Extracting c1(ψk)φ1(b) from observations

To capture c1(ψk)φ1(b) from (Equation2), it is enough to use the method of limits, [Citation3], on the sequence of observations or data given by {u(b,j)}jN for a fixed b(0,1). Assume that ψ1 does not change sign, say ψ1(x)>0 and since λ1<λ2<<λn<, the sequence {u(b,j)}jN generated by f=ψ1 is nontrivial and has the asymptotic for large j, (9) u(b,j)=eλ1j{c1(ψ1)φ1(b)+n2cn(ψ1)e(λnλ1)jφn(b)}=c1(ψ1)φ1(b)eλ1j+O(e(λ2λ1)j),(9) which yields (10) limju(b,j+1)u(b,j)=eλ1.(10) Although the limit is easy to set up, its existence hinges on the facts that φ1(b)0, as the first eigenfunction does not vanish inside, and also on c1(ψ1)=01φ1(x)ψ1(x)ρ(x)dx0, as the functions do not change sign. The limit can then be used to find the sought Fourier coefficients c1(ψk)φ1(b) (11) limjeλ1ju(b,j)=c1(ψk)φ1(b)=01φ1(x)φ1(b)ψk(x)ρ(x)dx.(11) From (Equation11), and (Equation8) we have (12) φ1(x)φ1(b)ρ(x)=k1c1(ψk)φ1(b)ψk(x)inL2(0,1).(12) Since we have reconstructed φ1(x)φ1(b)ρ(x), we only need φ1(x)φ1(b) to deduce ρ. To this end, since we already have λ1 and φ1(x)φ1(b)ρ(x), it means we have the right-hand side of a rescaled (Equation5), i.e. (13) φ1(x)φ1(b)=λ1φ1(x)φ1(b)ρ(x)L2(0,1),(13) where φ1(x)φ1(b)H2(0,1), which is easily obtained by the Green's function of the boundary value problem (14) {y(x)=f(x)0<x<1,y(0)hy(0)=0andy(1)+Hy(1)=0.(14) The existence and uniqueness is given by the following lemma.

Lemma 3.1

If H+h+hH0 then λ10and the solution to the boundary value problem  (Equation13) exists and is given explicitly by y(x)=G(f)(x):=01G(x,t)f(t)dt with G(x,t) defined by  (Equation15). Furthermore, the map fy is a bounded operator from L2(0,1)C(0,1).

Proof.

If λ1=0, then φ1(x)=0from (Equation5) and so φ1(x)=ax+b. Using the boundary conditions, at x=0,1, yield the system {ahb=0(1+H)a+Hb=0which gives a nontrivial solution, i.e. a2+b20, if and only if its determinant, which is the Wronskian ω=H+h+Hh=0.The same condition also implies that (Equation14) has a unique solution, which we can obtain using the Green's function (15) G(x,t)={1ω(1+H(1x))(1+ht)for0<t<x,1ω(1+H(1t))(1+hx)forx<t<1,(15) where ω0. From |G(f)(x)|G(x,.)f and the fact that the Green function is continuous, we deduce that (16) sup0x1|G(f)(x)|Mf(x)(16) with M=sup0x1G(x,.).

Once we have reconstructed λ1φ1(x)φ1(b)ρ(x), we can solve (Equation13) to obtain (17) φ1(x)φ1(b)=G(λ1φ1(x)φ1(b)ρ(x))=λ1(1+HxH)H+h+hH0x(1+ht)φ1(t)φ1(b)ρ(t)dtλ1(1+hx)H+h+hHx1(1+HtH)φ1(t)φ1(b)ρ(t)dt.(17) Combining (Equation12) and (Equation17) yield (18) ρ(x)=φ1(x)φ1(b)ρ(x)φ1(x)φ1(b)=S(x)G(S)(x),(18) where (19) S(x)=φ1(x)φ1(b)ρ(x)=k1c1(ψk)φ1(b)ψk(x).(19) Recall that the denominator is non zero since φ1(x)0 for all x(0,1). Thus, we have proved the main result of the paper stated previously in Theorem 1.1.

Remark 3.2

The condition H+h+hH0 excludes two cases, namely the Dirichlet and the Neumann cases which will be treated separately below. Note that their Green's function is easier to compute.

4. The Dirichlet case

With few modifications, we can also treat the Dirichlet case which corresponds to the limiting case h=H=, in (Equation1) i.e. (20) {ut(x,t)=1ρ(x)uxx(x,t),0<x<1,t>0,u(0,t)=0andu(1,t)=0,t>0u(x,0)=f(x),x(0,1),(20) whose solution is given by (21) u(x,t)=n1dn(f)ϕn(x)eμnt,(21) where ϕn are the eigenfunctions, (22) {ϕn(x)=μnρ(x)ϕn(x)x(0,1),ϕn(0)=ϕn(1)=0.(22) It is readily seen that μ10, otherwise ϕ1(x)=0, and clearly all eigenvalues are positive 0<μ1<<μn< . The Fourier coefficients are given by dn(f)=01f(x)ϕn(x)ρ(x)dx.Next, we use the basis {sin(πnx)}n1, to extract μ1 and the coefficients d1(sin(nπx))ϕ1(b)=01sin(nπx)ϕ1(x)ϕ1(b)ρ(x)dx,fornNfrom the observation u(b,t), from (Equation21) by the method of limits. Using the sequence {d1(sin(nπx))}n1 we reconstruct (23) ϕ1(x)ϕ1(b)ρ(x)=n12d1(sin(nπx))ϕ1(b)sin(nπx)inL2(0,1).(23) One of the advantages of the Dirichlet problem is the simplicity of its Green's function. From (Equation22) we have (24) ϕ1(x)ϕ1(b)=μ1ϕ1(x)ϕ1(b)ρ(x).(24) From (Equation24) and (Equation23) it follows that (25) ϕ1(x)ϕ1(b)=2μ1n1d1(sin(nπx))ϕ1(b)sin(nπx)n2π2.(25) In order words, we deduce from (Equation25) (26) ρ(x)=n1d1(sin(nπx))ϕ1(b)sin(nπx)μ1n1d1(sin(nπx))ϕ1(b)sin(nπx)n2π2.(26) Thus after obtaining μ1 and ϕ1(x)ρ(x), we can reconstruct ρL2(0,1) by (Equation26). Thus we have proved the following theorem

Theorem 4.1

In the Dirichlet case we can reconstruct ρL2(0,1) by  (Equation26) from the map sin(nπx){u(b,k)}kN for any fixed b(0,1), where u is a solution of  (Equation20).

Remark 4.2

Although ϕ1(b)0, we still cannot simplify (Equation26) because it is unknown.

We now work out the Neumann boundary condition which is very different because 0 is an eigenvalue.

5. The Neumann case

Neumann boundary conditions means we have h=H=0, in (Equation1) i.e. (27) {ut(x,t)=1ρ(x)uxx(x,t),0<x<1,t>0,ux(0,t)=0andux(1,t)=0,t>0,u(x,0)=f(x)x(0,1),(27) whose solution is given by (28) u(b,t)=h0(f)χ0(b)+n1hn(f)χn(b)eξnt.(28) Here χn are the normalized eigenfunctions in Lρ2(0,1) with χnρ=1, and are solutions of (29) {χn(x)=ξnρ(x)χn(x)0<x<1,n=0,1,2,,χn(0)=χn(1)=0(29) with eigenvalues (30) 0=ξ0<ξ1<ξ2<<ξn<andχ0(x)=(01ρ(x)dx)1/2.(30) The Fourier coefficients in (Equation28) are given by hn(f)=01f(x)χn(x)ρ(x)dxforn=0,1,2,.It follows from (Equation28) that u(b,t)h0(f)χ0(b)ast.Note we are not looking for the first eigenvalue, as it is already known to be zero. Next, we use the basis {cos(nxπ)}n0, as initial conditions to extract the first coefficients in (Equation28), and use (Equation30) (31) h0(cos(nxπ))χ0(b)=01cos(nxπ)ρ(x)χ0(b)dxχ0(b)=(01ρ(x)dx)101cos(nxπ)ρ(x)dxforn=0,1,2,(31) Next observe that for n=0, (Equation31) reduces to h0(1)=1and the Fourier cosine coefficient of ρin terms of the observation (32) 01cos(nxπ)ρ(x)dx=h0(cos(nxπ))χ0(b)01ρ(x)dx.(32) Recall that the Fourier cosine series of ρ in L2(0,1) is ρ(x)=01ρ(x)dx+2n101cos(nπx)ρ(x)dxcos(nπx)=01ρ(x)dx[1+2n1h0(cos(nxπ))χ0(b)cos(nπx)]inL2(0,1).Thus we need the value 01ρ(x)dxin order to reconstruct ρ.

Theorem 5.1

In the Neumann case we can reconstruct ρ provided we know 01ρ(x)dx and the map cos(nπx){u(b,k)}kN for b[0,1].

Remark 5.2

If we could extract the second eigenvalue, ξ10, then we can recover ρ by a similar formula to (Equation26) and we would not need the value 01ρ(x)dx.

6. Convergence

In practice, we can only use a finite sum to approximate S(x)=φ1(x)φ1(b)ρ(x)=k1c1(ψk)φ1(b)ψk(x)given in (Equation19). Using the notation of section 3, let us denote the approximation of ρ by (33) ρN(x)=SN(x)G(SN)(x)whereSN(x)=k=1k=Nc1(ψk)φ1(b)ψk(x).(33) If SN(x)=φ1(x)φ1(b)ρ(x)+εN(x)then ρ(x)ρN(x)=G(SN(x))S(x)G(S(x))SN(x)G(SN)(x)G(S)(x)=S(x)G(εN(x))G(S(x))εN(x)G(S)(x)G(S)(x)+G(S)(x)G(εN)(x).Note that as εN0 in L2(0,1), it follows that sup|G(εN)(x)|0as N. Recall that we have G(S)(x)=φ1(x)φ1(b)>0 for 0<x<1 and thus for any interval (a,b) such that 0<a<b<1 we have infa<x<b|G(S)(x)G(S)(x)+G(S)(x)G(εN)(x)|=δ(a,b,N)>0forNlargeenough.Then, it follows ab|ρ(x)ρN(x)|2dx1δ(a,b,N)(Ssup0x1|G(εN(x))|+sup0x1|G(S(x))|εN)1δ(a,b,N)(SMεN+MSεN)2Mδ(a,b,N)SεN,where M is defined in (Equation16). Therefore we have proved the following theorem.

Theorem 6.1

ρN defined by (Equation33) converges to ρ on any subinterval of (0,1), i.e. ρNρ in L2(a,b) for any (a,b)(0,1).

7. Algorithm

The whole reconstruction procedure is based on the infinite limits (Equation10) and (Equation11). Since we cannot take them in practice, we shall use the tolerance, ε say, to find a stopping time numerically. Thus from the limit in (Equation10) we have (34) for any ε>0,there exists T0, such that for T>T0 we have |u(b,T+1)u(b,T)eλ1|<ε,(34) where T0 is determined numerically. We also use the same idea to compute the Fourier coefficients c1(ψk)φ1(b) from the limits in (Equation11). As we shall see in the examples, due to the rapid decay of the exponentials, in general, T03. We now work out few numerical examples to illustrate the method.

8. Numerical examples

For the sake of simplicity, we shall use Dirichlet and Neumann boundary conditions where the initial condition f(x) is chosen from the basis {sin(kxπ)}kN for the former and {cos(kxπ)}kN for the latter. The observations of the solution are taken at either b=1/2 or b=1/4 and generated by solving (Equation1) using the Crank–Nicholson method up to T. We recall that it follows from (Equation10) that the sequence of ratios r(k)=u(1/2,(k+1)/100)u(1/2,k/100)yieldslimkr(k)=eλ1/100.Example 1. Consider {ut(x,t)=12sin(πx)uxx(x,t)x(0,1),t(0,T)u(0,t)=0,u(1,t)=0t(0,T)u(x,0)=sin(kxπ)x(0,1),kN

Figure 1. 2sin(πx) –  and ρ20 +++.

Figure 1. 2sin⁡(πx) –  and ρ20 +++.

Figure 2. (2x1)2 – and ρ20 +++

Figure 2. (2x−1)2 – and ρ20 +++
To find the stopping time T0, we observe the sequence u(1/2,k/100) for k1 and

k,  r(k)

35, 0.9437854846474063

36, 0.9437854846439294

37, 0.9437854846420766

38, 0.9437854846410607

39, 0.9437854846405333

40, 0.9437854846402476

41, 0.9437854846400894

42, 0.9437854846400063

Thus, tolerance ε=1012 is achieved at time 39/100,0.943785484640_5333 which means that T040×1/100=0.4. This yields λ1:=5.7856379507 from exp(λ1/100)=0.943785484640. Next we obtain the sequence of Fourier coefficients by evaluating u(1/2,0.4)exp(λ10.4):

0.1794, 7.333 e−16, −0.0390, −3.667e−17, −0.4194e−2, −2.649e−17, −0.1497e−2, 1.111 e−16, −6.936e−4, −4.778e−16, −3.768e−4, −1.235e−16, −2.272e−4, −5.787e−16, −1.474e−4, 4.971 e−16, −1.011e−4, −4.776e−16, −7.232e−5, 9.872 e−16.

Using (Equation26) we get

Example 2: We now treat the Neumann case {ut(x,t)=1(2x1)2uxx(x,t)0<x<1,t>0ux(0,t)=0,ux(1,t)=0t>0u(x,0)=cos(kxπ)x(0,1)kNρ(x)=(2x1)2 and so the extra data is 01ρ(x)dx=13. We also know that the first eigenvalue is ξ0=0 and we shall observe the u(1/4,T0). To find T0 use the ratios k,  r(k)

284, 0.99996174

285, 0.99998622

286, 0.99998188

287, 1.00001679

288, 1.00000472

289, 1.00033535

290, 0.99944850

291, 1.00031139

292, 1.00000228 we conclude that T0 =3 for ε=103. The sequence of Fourier coefficients is seq(c[k],k= 0.20):

1.0000, 3.7746 e−13 , 0.6079, −1.1207e−11, 0.1519, 5.1966 e−11, 0.0674, −1.4377e−10, 0.0379,3.1113 e−10 , 0.0242, −5.8620e−10, 0.0168, 1.0112 e−9, 0.0124, −1.6288e−9, 0.00956, 2.4390 e−9 , 0.0076, −3.2949e−9, 0.0063.

9. Conclusion

We have shown that tools from signal processing can help extract spectral data to reconstruct an unknown weight of a parabolic equation. It also shows that good imaging and non-destructive testing in industry can be achieved with few measurements done with one accurate sensor.

Acknowledgements

The authors sincerely thank the referees for their valuable comments and insight. Both authors also thank KFUPM for its support.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work is supported by KFUPM [grant number SB191035].

References

  • Al-Musallam F, Boumenir A. Recovery of a parabolic equation generated by a Krein string. J Math Anal Appl. 2014;420(2):1408–1415.
  • Boumenir A, Ghattassi M, Laleg-Kirati TM. Monitoring the temperature of a direct contact membrane distillation. Math Methods Appl Sci. 2020;43(3):1399–1408.
  • Boumenir A, Tuan VK. Recovery of the heat coefficient by two measurements. Inverse Probl Imaging. 2011;5(4):775–791.
  • Boumenir A, Tuan VK. Inverse problems for multidimensional heat equations by measurements at a single point on the boundary. Numer Funct Anal Optim. 2009;30(11-12):1215–1230.
  • Boumenir A, Tuan VK, Nguyen N. The recovery of a parabolic equation from measurements at a single point. Evol Equ Control Theory. 2018;7(2):197–216.
  • Cao K, Lesnic D. Reconstruction of the perfusion coefficient from temperature measurements using the conjugate gradient method. Int J Comput Math. 2018;95(4):797–814.
  • Cordaro PD, Kawano A. A uniqueness result for the recovery of a coefficient of the heat conduction equation. Inverse Probl. 2007;23(3):1069–1085.
  • Dym H, McKean HP. Gaussian processes, function theory, and the inverse spectral problem. New York: Academic Press; 1976. (Probability and Mathematical Statistics; vol. 31).
  • Huntul MJ, Lesnic D, Hussein MS. Reconstruction of time-dependent coefficients from heat moments. Appl Math Comput. 2017;301:233–253.
  • Kac IS, Krein MG. On the spectral functions of the string. Amer Math Soc Transl. 1974;103:19–102.
  • Kac A. An introduction to the mathematical theory of inverse problems. New York: Springer; 1996. (Applied Mathematical Sciences; vol. 120).
  • Lowe BD, Rundell W. An inverse problem for a Sturm-Liouville operator. J Math Anal Appl. 1994;181:188–199.

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.