Open access peer-reviewed chapter

A New Approach to Solve Non-Fourier Heat Equation via Empirical Methods Combined with the Integral Transform Technique in Finite Domains

Written By

Cristian N. Mihăilescu, Mihai Oane, Natalia Mihăilescu, Carmen Ristoscu, Muhammad Arif Mahmood and Ion N. Mihăilescu

Submitted: 18 January 2022 Reviewed: 14 March 2022 Published: 21 April 2022

DOI: 10.5772/intechopen.104499

From the Edited Volume

Matrix Theory - Classics and Advances

Edited by Mykhaylo Andriychuk

Chapter metrics overview

94 Chapter Downloads

View Full Metrics

Abstract

This chapter deals with the validity/limits of the integral transform technique on finite domains. The integral transform technique based upon eigenvalues and eigenfunctions can serve as an appropriate tool for solving the Fourier heat equation, in the case of both laser and electron beam processing. The crux of the method consists in the fact that the solutions by mentioned technique demonstrate strong convergence after the 10 eigenvalues iterations, only. Nevertheless, the method meets with difficulties to extend to the case of non-Fourier equations. A solution is however possible, but it is bulky with a weak convergence and requires the use of extra-boundary conditions. To surpass this difficulty, a new mix approach is proposed with this chapter resorting to experimental data, in order to support a more appropriate solution. The proposed method opens in our opinion a beneficial prospective for either laser or electron beam processing.

Keywords

  • non-Fourier equation
  • integral transforms technique
  • eigenfunctions and values
  • experimental data

1. Introduction

1.1 Mathematical background

The heat equation can be solved in a simpler mode via the Fourier heat equation, which involves the propagation of heat waves with infinite speed. This hypothesis is in particular valid for many applications, such as laser-metal interaction in the frame of two-temperature model [1, 2].

The solution of Fourier equations can be inferred using different mathematical techniques via Green function, integral, Laplace transform, or complex analysis. The predictions of the solutions given by the mentioned methods are of analytical or semianalytical nature and confirm the experimental data for certain situations such as laser–metal interaction.

One basically assumes that the heat waves propagation speed is inversely proportional to the square root of the relaxation time. A smaller relaxation time leads to higher heat speed waves, resulting in a good Fourier approximation. If one requires however a more accurate description of experimental data, one should introduce a more exact method to solve the non-Fourier equation involving a finite heat wave speed.

A mixed solution of the non-Fourier equation combines the theoretical method of finite integral transforms with information from experimental data. Thus, two additional boundary conditions can be imposed, which will lead to a semianalytical solution of the non-Fourier equation. The finite domains of the integral transform method for Fourier equations are eigenfunctions and values, which reach after 10 iterations a quite conform solution for the Fourier equation [3, 4, 5, 6]. This method is applied to the non-Fourier equation, and the final form is obtained, with the support of experimental results.

A new heat transfer model was adopted in order to unify the thermal field distribution in both laser and electron beam processing. An analytical solution using non-Fourier heat equation has been developed corresponding to boundary conditions in the case of material processing. The model has been compared with the experimental data obtained using an in-house developed facility. A simplified and easy-to-use model via MATHEMATICA software stands for the novelty of the current work.

Advertisement

2. Non-Fourier equation

The non-Fourier equation is hyperbolic and can be written as:

2Tr2+1rTr+2Tz2+1r22Tφ21γTtτ0γ2Tt2=PrφztK.E1

Here, T is target temperature, r is the radial coordinate, z is the spatial coordinate on the direction of laser beam propagation, t is time, τ0 is relaxation time, and γ is thermal diffusivity. P stands for the source term,φ is the angular coordinate, while K is the target thermal conductivity. For a simple general solution, one assumes a cylindrical target with angular symmetry, under irradiation with a Gaussian laser beam with the center at r = z = 0. In this case, the temperature does not depend on φ:

Tφ=0.=>T=Trzt.E2

The corresponding boundary conditions are:

KTrztrr=b+hTbzt=0.E3

Here r = b is the cylinder radius, while h is the heat transfer coefficient. The boundary conditions for the z coordinate are:

KTrztzz=0hTr0t=0,E4

and

KTrztzz=a+hTrat=0,E5

where a is the cylinder length. We will pass on the effective solution of the equation. The operator Dr was defined as:

Dr=2r2+1rr.E6

This applies to Eq. (1) with the boundary conditions:

2Krr2+1rKrr+μ2Kr=0,E7

and

KKrrztrr=b+hKrbzt=0.E8

Eqs. (7) and (8) corroborate to Eq. (9):

hKJ0μibμiJ1μib=0.E9

One can deduce, based upon the theory of finite integral transforms, the eigenfunction Krrμi corresponding to the eigenvalue μi:

Krrμi=JOμirr1Ci,E10

where the normalization constant is given by:

Cir=0brKr2rμidr=b22μi2h2k2+μi2J02μib.E11

One defines:

Tμi,zt=1Ci0bTrztrJ0μirdr,E12

and

Pμizt=1Ci0bPrztrJ0μirdr.E13

Eq. (1) in this case converts to:

μi2·T+2Tz21γTtτ0γ2Tt2=PK.E14

One obtains for z coordinate via similar mathematical calculation:

2Kzz2+λ2Kz=0,E15

and

KKZzhKzz=0=0,E16

as well as:

KKZz+hKzz=a=0.E17

The − and + signs for h in Eqs. (16) and (17) denote the target heat absorption and emission, respectively. One has:

Kzzλ=cosλz+hλKsinλz,E18

and

2cotλja=λjkhhλjk.E19

Here, λj denotes eigenvalues along the z-axis. From theory [7], it follows:

Kzzλj=1CjKzzλj,E20

and

Cj=0aKz2zλjdz.E21

Note that Eqs. (12) and (13) discuss the eigenvalues along the r-axis, only. After introducing the eigenvalues along the z-axis, one can step ahead to generalize Eqs. (12) and (13) as:

T¯μiλj,t=1CiCj0a0bTrztrKrμirKzλjzdrdz,E22

and

P¯μiλjt=1CiCj0a0bPrztrKrμirKzλjzdrdz.E23

Eq. (1) becomes now:

μi2+λj2T¯+1γT¯t+τ0γ2T¯t2=P¯K.E24

We next applied the direct and inverse Laplace integral transform to solve Eq. (1) in relation to time. C[1] and C[2] stand for the normalizing coefficients with respect to the experimental data. The results are as follows:

Trztτ0=i=110j=110Pμiλjμi2+λj2+C11γ1γ24μi2τ0γ4λj2τ0γt2τ0γ+C21γ+1γ24μi2τ0γ4λj2τ0γt2τ0γKrμirKzλj)z.E25

and

Tμiλj=1CiCj0a0bPrztrKrμirKzλjzdrdz.E26

We finally mention that for an intermediate point in the experimental curve, one has:

Trzt=Tτ0C1C2rzt.E27

With the boundary conditions:

Trz0C1C2τ0=22°C.E28
TrzC1C2τ0=22°C.E29
Advertisement

3. Two-temperature model in the non-Fourier version

The two-temperature model (TTM) is based upon two coupled equations:

ATeTet+Kτ0γ2Tet2=K2Tex2+2Tey2+2Tez2GTeTi+Part,E30
CiTit=GTeTi.E31

Here Te and Ti stand for the electron and phonon temperatures, respectively. G is the coupling factor between electrons and phonons. Part is the heat source, which is induced via laser-metal interaction. The interaction could be considered of either classical or steady-state quantum mechanical type. A is the electron heat capacity, and K is the thermal conductivity of the metal. According to Ref. [8], G can be determined from:

G=π2mNv26τTiTeTi4×0TeTdx4/ex1dx,E32

where m is the electron mass, N is the conduction electron density, v is the velocity of sound in the metal, τ is the electron–phonon collision time, and TD is the Debye temperature. The exact data for each metal (Cu, Ag, Al, or Fe) are available from text books and current literature (e.g., [8, 9, 10]). As for all metals, Ci>>A, one may assume, in a first approximation, that:

K2Tex2+2Tey2+2Tez2CiTitKτ0γ2Tet2=Part.E33

According to the Nolte model [2], one has:

Ti=κTe,E34

where

κ=τLτL+τi.E35

Here, τi is the lattice cooling time while τL is the pulse duration time. Consequently, one has:

Tit=κTet.E36

Eq. (3) can be rewritten as:

K2Tex2+2Tey2+2Tez2CiκTetKτ0γ2Tet2=Part.E37

It follows that:

2Tex2+2Tey2+2Tez21γTetτ0γ2Tet2=PartKE38

with

γ=KCiκ.E39

Under the most general form, the heat source reads as:

Pa=m,nImnyzαmneαmnx1rSmn+rSmnδx+qcHtHtt0.E40

Here, Imnyz stands for the laser transverse mode {m,n} while αmn is the linear absorption coefficient, and rSmn is the surface absorption coefficient. The quantum corrections (qc) are steady state for the respective mode. H stands for the step function, and t0 for the exposure time. One model explains the continuous laser beam irradiation, while the other one, more realistic in our opinion, illustrates the laser beam in pulse form. The equivalence between the two models requires therefore that the intensity versus time plot should cover the same area.

In order to make a comparison with experiments, one needs besides analytical description, concrete numerical values. The next step is therefore to estimate the eigenvalues, numerically. For this purpose, Eq. (7) can be solved using the integral transform technique, and eigenfunctions and eigenvalues could be calculated. One has three differential equations as follows (K represents the eigenfunctions, while λ,μ,andξ are the eigenvalues) [5]:

2Kxx2+λi2Kx=0,E41
2Kyy2+μj2Ky=0,E42
2Kzz2+ξk2Kz=0.E43

The final solutions could be achieved on the basis of Eqs. (41)(43):

Kx=cosλix+hKλisinλix,E44
Ky=cosμjy+hKμjsinμjy,E45
Kz=cosξkz+hKξksinξkz.E46

The boundary conditions are:

KxxhKxKx=0=0;Kxx+hKxKx=a=0,E47
Kyy+hKyKy=0=0;Kyy+hKyKy=b=0,E48
Kzz+hKzKz=0=0;Kzz+hKzKz=c=0.E49

Here, a, b, and c are the metal sample dimensions. As for the boundary conditions, the eigenvalues (h is the heat transfer coefficient) can be inferred from Eqs. (47) to (49), as:

2cotλia=λiKhhKλi,E50
2cotμjb=μjKhhKμj,E51
2cotξkc=ξkKhhKξk.E52

The solution is obtained via integral transform technique as:

TxyztC1C2τ0=i=110j=110k=110Pμiλj,ξkμi2+λj2+ξk2+C11γ1γ24μi2τ0γ4λj2τ0γ4ξk2τ0γt2τ0γ+C21γ+1γ24μi2τ0γ4λj2τ0γ4ξk2τ0γt2τ0γ×KxμixKyλjyKzξkzE53

The advantage of Eq. (53) in our model is related to a quick converging series. Thus, after 10 iterations, the solution’s accuracy reaches already 10−2 K in the case of thermal distribution [11].

Advertisement

4. Experimental details

The experimental setup is operated by a Nd:YAG pulsed laser source (λ = 355 nm) (Surelite II from Continuum), generating pulses of 6 ns duration with (130 ± 0.6) mJ energy at a frequency repetition rate of 10 Hz. The laser beam had a spatial top-hat distribution. The laser beam was focused onto the metallic target surface by a lens with 240 mm focal lengths. An Al bulk target of (10 × 10 × 5) mm3 was used in experiments. The laser fluence was set at ∼7.5 J/cm2 to surpass the ablation threshold but also to avoid the excessive plasma formation. A crater of 18 μm depth was dig into the sample after the application of 1000 subsequent laser pulses, as checked up by a Vernier Caliper instrument. During the multipulse laser irradiation, the thermal distribution was monitored on the sample back side via a FLUKA thermocouple connected to a computer having Lab view software, while the sample was irradiated at the top. All experiments were performed on an in-house developed equipment at Laser Department, National Institute for Laser, Plasma and Radiation Physics (INFLPR), Magurele, Romania.

Advertisement

5. Results and discussion

Experiments and simulations were carried out during the heating of a metallic target. The boundary conditions were described by Eq. (28). In all figures, the experimental data are plotted with dots while the simulations are represented by a continuous line. Relaxation time,τ0, was assumed 0.5 ps (Figure 1), 1 ns (Figure 2), and 1 μs (Figure 3), respectively. For simulation, a heat transfer coefficient = 3 × 10−7 W mm−2 K−1 was selected. As known [12, 13, 14], for a very low heat transfer coefficient, the eigenvalues are positive very small numbers, resulting in a linear thermal distribution curve, as visible in Figures 13.

Figure 1.

Time evolution of temperature for a relaxation time of 0.5 ps: experiments (dotted line) vs. simulation (continuous line).

Figure 2.

Time evolution of the temperature for a relaxation time of 1 ns: experiments (dotted line) vs. simulation (continuous line).

Figure 3.

Time evolution of the temperature for a relaxation time of 1 μs: experiments (dotted line) vs. simulation (continuous line).

The best agreement between theory and experiment was achieved for a relaxation time,τ0,of 0.5 ps, as visible from Figure 1. We note that this is in accordance with available literature on subject [15].

Advertisement

6. Conclusions and outlook

The two-temperature model was generalized to the case of the non-Fourier approach via the electron-phonon relaxation time. Boundary conditions, Eq. (28) for heating and Eq. (29) for cooling, were considered to this purpose. The obtained solutions prove useful for experimental data analysis. The mathematical method belongs to the eigenvalues and functions family, while details on software are available from Ref. [6].

The exact nature of the metallic target (in our case aluminum) could be detected from the electron-phonon relaxation time using integral transform technique mix via acquired experimental data. The method can be extended to any experimental sample (metal) with the high accuracy.

Advertisement

Acknowledgments

CNM, MO, NM, and CR acknowledge for financial support by Romanian Ministry of Research, Innovation and Digitalization, under Romanian National NUCLEU Program LAPLAS VI–contract no. 16N/2019. CNM, NM, CR, and INM thank for the financial support from a grant of the Romanian Ministry of Education and Research, CNCS-UEFISCDI, project number ID code RO-NO-2019-0498 and UEFISCDI under the TE_196/2021 and PED_306/2020. M.A.M. received financial support from the European Union’s Horizon 2020 (H2020) research and innovation program under the Marie Skłodowska–Curie grant agreement no. 764935.

References

  1. 1. Schoenlein RW, Lin WZ, Fujimoto JG, Eesley GL. Femtosecond studies of nonequilibrium electronic processes in metals. Physical Review Letters. 1987;58(16):1680-1683. DOI: 10.1103/PhysRevLett.58.1680
  2. 2. Nolte S, Momma C, Jacobs H, Tunnermann A, Chichkov BN, Wellegehaussen B, et al. Ablation of metals by ultrashort laser pulses. Journal Optical Society of America B. 1997;14(10):2716-2722. DOI: 10.1364/JOSAB.14.002716
  3. 3. Oane M, Peled A, Medianu R. Notes on Laser Processing. Germany: Lambert Academic Publishing; 2013. ISBN 978-3-659-487-48739-2
  4. 4. Oane M, Ticoş D, Ticoş CM. Charged particle beams processing versus laser processing. Germany: Scholars’ Press; 2015. ISBN 978-3-639-66753-0
  5. 5. Oane M, Sporea D. Temperature profiles modelling in IR optical components during high power laser irradiation. Infrared Physics & Technology. 2001;42(1):31-40. DOI: 10.1016/S1350-4495(00)00065-7
  6. 6. Oane M, Mahmood MA, Popescu A. A state-of-the-art review on integral transform technique in laser-material interaction: Fourier and non-Fourier heat equations. Materials. 2021;14(16):4733. DOI: 10.3390/ma14164733
  7. 7. Koshlyakov NS, Smirnov MM, Gliner EB. Differential Equation of Mathematical Physics. Amsterdam: North-Holland; 1964
  8. 8. American Institute of Physics Handbook. 3rd ed. New York: McGraw-Hill; 1972
  9. 9. Du G, Chen F, Yang Q, Jinhai S, Hou X. Ultrafast temperature evolution in Au film under femtosecond laser pulses irradiation. Optics Communications. 2010;283:1869-1872. DOI: 10.1016/J.OPTCOM.2009.12.038
  10. 10. Damian V, Oane M, Buca A. The Fourier approach of the two temperature model for laser beam-metal interaction: Experiment versus theory. Lasers in Engineering. 2016;33(1-3):181-186
  11. 11. Visan T, Sporea D, Dumitru G. Computing method for evaluating the absorption coefficient of infrared optical elements. Infrared Physics & Technology. 1998;39(5):335-346
  12. 12. Oane M, Vutova K, Mihailescu IN, Donchev V, Florescu G, Munteanu L, et al. The study of vacuum influence on spatial-temporal dependence of thermal distributions during laser-optical components interaction. Vacuum. 2012;86:1440-1442
  13. 13. Oane M, Toader D, Iacob N, Ticos CM. Thermal phenomena induced in a small tungsten sample during irradiation with a few MeV electron beam: Experiment versus simulations. Nuclear Instruments and Methods in Physics Research B. 2014;337:17-20
  14. 14. Oane M, Toader D, Iacob N, Ticos CM. Thermal phenomena induced in a small graphite sample during irradiation with a few MeV electron beam: Experiment versus theoretical simulations. Nuclear Instruments and Methods in Physics Research B. 2014;318:232-236
  15. 15. Girardeau-Montaut JP, Afif M, Girardeau-Montaut C, Moustaizis SD, Papadogiannis N. Aluminium electron-phonon relaxation-time measurement from subpicosecond nonlinear single-photon photoelectric emission at 248 nm. Applied Physics A. 1996;62:3-6. DOI: 10.1007/BF01568079

Written By

Cristian N. Mihăilescu, Mihai Oane, Natalia Mihăilescu, Carmen Ristoscu, Muhammad Arif Mahmood and Ion N. Mihăilescu

Submitted: 18 January 2022 Reviewed: 14 March 2022 Published: 21 April 2022