You are on page 1of 15

Applied Mathematical Modelling 31 (2007) 1186–1200

www.elsevier.com/locate/apm

A Green’s function model for the analysis of laser


heating of materials
Pradip Majumdar *, Hong Xia
Department of Mechanical Engineering, Northern Illinois University, DeKalb, IL 605115, United States

Received 1 February 2004; received in revised form 1 November 2005; accepted 12 April 2006
Available online 22 June 2006

Abstract

An analytical model based on Green’s function method is developed to analyze the temperature distribution and heated
regions in a material irradiated by a high-energy laser beam. The model is multi-dimensional, transient and incorporates
different types of beam characteristics and boundary conditions. The multi-dimensional integration formulas in the
Green’s function solution equation are evaluated using an adaptive numerical integration algorithm. A parametric study
is conducted to show the effect of various laser beam parameters and material properties on the laser heating process.
 2006 Elsevier Inc. All rights reserved.

1. Introduction

High-energy lasers are increasing being used in the industry for many different material processes due to their
many advantages including high power density; high directionality and smaller heat-affected zones. These indus-
trial applications include heat-treating to modify surface structures through intense heating followed by cooling;
melting the surface layer to modify the properties such as wear and corrosion; surface alloying through surface
melting and addition of alloy elements; drilling and cutting that involve melting and ablation of materials.
In order to enhance the potential for such laser machining process, it is essential to get better understanding
of the physical process associated with the laser material interaction and the effect of various operating param-
eters on the performance. In metallurgical processes such as heat treatment and surface hardening there is a
correlation between the materials properties such as hardness and the heat-affected zone (HAZ). The heat-
affected zone depends on the laser and optical properties such as beam power, beam spot size and material
optical properties such as absorptivity. A mathematical model is a powerful tool for simulating such process
operation and it reduces the need for expensive and time-consuming trial and error approaches normally used
on most of the production floor.
Brugger [1] presented an exact solution for temperature rise in infinitely extended laser-heated slab of finite
thickness. His analytical model is based on an instantaneous point source and a combination of source and

*
Corresponding author. Tel.: +1 815 753 9963; fax: +1 815 753 0416.
E-mail address: majumdar@ceet.niu.edu (P. Majumdar).

0307-904X/$ - see front matter  2006 Elsevier Inc. All rights reserved.
doi:10.1016/j.apm.2006.04.007
P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1187

sinks pairs following the principles of method of images. Warren and Spark [2] developed an exact solution for
the temperature in a laser-heated slab with temperature dependent surface absorptance. The solution was
developed for one-dimensional heat diffusion equation using Laplace transformation techniques. Cline and
Anthony [3] presented an analytical model based on a Green’s function for a unit source and assuming the
media to be semi-infinite. Dabby and Peak [4] presented analytical model for material removal by a high-inten-
sity laser beam. The model assumed volumetric heat absorption in a one-dimensional semi-infinite media. The
analytical solution method involves use of perturbation techniques and use of Laplace transformation for
obtaining a close form solution to the perturbation problems. Rogersan and Chayt [5] presented the exact
solution for the melting time in an ablating slab using a one-dimensional heat conduction model. Minardi
and Bishop [6] studied temperature distribution within metals subjected to laser irradiation of different spatial
intensity profiles. Lemczyk and Yovanovich [7,8] used Hankel integral transform to present analytical solution
in their study of thermal constriction resistance with convection boundary condition in half space contacts and
assuming semi-infinite contact layer attached to half-space. It can be noted here that all these analytical solu-
tions are primarily based on one-dimensional analysis and/or semi-infinite media, and with or with out a uni-
form heat generation.
Barrilot [9] validated his numerical solution of the transient heating of an anisotropic two-object with
Gaussian laser beam using a limiting case analytical solution. The governing heat equation was modified using
co-ordinate transformation technique to immobilize the moving boundary involving ablation. The derived
analytical solution is limited to two-dimensional semi-infinite medium and requires one-dimensional numer-
ical integration involving error and Bessel functions.
Kaylon and Yilbus [10] used Laplace transformation method to derive close form analytical solution for
temperature rise for laser evaporative heating process time exponentially decaying pulse laser. The solution
was limited to one-dimensional and semi-infinite region. Yilbas and Kalyon [11] and Kalyon and Yilbus
[12] also used using Laplace transformation method to obtained close from analytical solution for pulsed laser
heating of metallic substrate with convective boundary condition at the surface.
Yanez et al. [13] presented an analytical solution for a time dependent temperature field in a semi-infinite
medium in their study of laser hardening process. The solution was obtained in terms of Green function, which
was derived using Fourier transformation. Gustavo and Guillermo [14] presented analytical solution for tran-
sient temperature distribution in a three-dimensional finite region using separation of variable. Pantelis [15]
used an analytical model for studying heat-affected zone during laser surface treatment. The model is based
on Duhamel’s principle leading to a convolution integral, which is integrated analytically.
The major objective of this study is to develop an analytical model based on Green’s function method to
analyze the temperature distribution, heat flux distribution and heat-effected zone in a two-dimensional finite
block irradiated by a continuous wave laser beam. The effects of various laser beam operation parameters as
well as effect of material parameters are discussed.

2. Mathematical formulation

Fig. 1 shows a general arrangement of the laser installation and the work piece. The laser beam is focused
by a lens through which the laser beam passes and behind which it converges to a minimum beam waist
around the focal point of the lens.
The absorbed heat contributes to the generation of heat in the workpiece. High performance is achieved
through better ability to control this heat input precisely and/or minimize the heat-affected zone (HAZ).
The physical mechanism that occurs at the processing surface is a function of laser beam intensity and char-
acteristics as well as the thermal characteristics of the material. Such interaction dictates the temperature rises,
heat-affected zone (HAZ), penetrative depth and process time.

2.1. Laser beam characteristics and parameters

The basic role of the laser in material processing is to generate heat in the workpiece. The efficiency with
which a material absorbs an incoming laser beam and converts it to heat depends on the properties of mate-
rials and operating conditions as well as on the types and characteristics of the beam, such as wavelength,
1188 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

Laser Beam

Focusing lens

Workpiece

X-Y Stage

Fig. 1. Schematic of laser installation and workpiece.

divergence of the incoming beam, and spatial intensity profile. Various beam characteristics such as beam
diameter, divergence, focusing characteristics, etc. are accommodated in the present model to enable simula-
tion of practical situations. A laser beam can be characterized based on the following parameters: shape and
intensity distribution, beam radius, divergence and defocusing distance.

2.2. Laser beam intensity distribution

The beam shape is assumed to be Gaussian with the intensity distribution in TEM00 mode is expressed as
R2 ð0Þ ðxR22 ð0Þ
þy 2 Þ
I ¼ I0 e ; ð1Þ
R2 ðzÞ
where I0 is the beam power density at the beam center.

2.3. Laser beam radius

Beam radius, R defined as the distance from the beam center at which the beam intensity drops to (1/e2) its
value at the beam center. The beam expands as it propagates through the space but the intensity distribution
remains unaltered with the width of the beam changing.

2.4. Beam divergence and defocusing

The beam divergence describes the expansion of the beam in the space. The effective radius of a diverging
Gaussian beam is given by
"  2 #1=2
W þz
RðzÞ ¼ Rð0Þ 1 þ ; ð2Þ
pR20 =k
where R(0) is the effective radius at the focal plane; W is the defocused distance between the focal plane of the
lens and the material surface.

W < 0 if the focal plane is below the material surface.


W > 0 if the focal plane is above the material surface.

The beam diameter changes as defocusing distance is varied. Both beam radius and defocusing distance are
important laser beam parameters and directly affect the beam power intensity.

2.5. Mathematical model

The physical phenomena involved in laser material processing are basically thermal in nature. When a laser
beam strikes a material surface, several effects take place: reflection and absorption of the beam, diffusion of
P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1189

∂T ∂T
k = h(T − T∞ ) k = h(T − T∞ )
∂z ∂z

z
-R) R (2L, H)
(0, H)

T = T∞ T = T∞

(0, 0) (2L, 0) x
dT/dz=0

Fig. 2. A general schematic of a laser heating problem.

heat in the material and loss of heat by convection and/or radiation from the material surfaces. The amount of
energy absorbed and utilized in processing the material depends on the optical and thermophysical properties
of the material. Fig. 2 shows a general arrangement of the laser heating problem and associated physical con-
ditions. A mathematical model that describes the process of heating the work piece subjected to high-energy
laser beam is presented here. The following simplifying assumptions are made in defining the model: (i) The
material is homogeneous and isotropic. (ii) Material is opaque with constant optical property. (iii) The laser
beam is perpendicular to the surface of block. (iv) CO2 laser is continuous Gaussian beam and TEM00 mode.
Based on these assumptions, the mathematical statement of the problem is as follows:
Governing equation

oh o2 h o2 h
¼ þ þ gðx; z; tÞ: ð3Þ
ot ox2 oz2
Boundary conditions:

1: hðt; 0; zÞ ¼ 0; ð4Þ
2: hðt; 2L; zÞ ¼ 0; ð5Þ
oh
3: at z ¼ 0; ¼ 0; ð6Þ
oz
4: at z ¼ H ;

Region I: 0 < x < L  R, L + R < x < 2L


oh
¼ B1 h: ð7Þ
oz
Region II: L  R < x < L + R
2
oh  x
¼ B2 I 0 e R2 ð0Þ : ð8Þ
oz

Initial condition
hð0; x; zÞ ¼ h0 ¼ T 0  T 1 : ð9Þ
The variables and parameters are

h ¼ T  T 1; t ¼ at; B1 ¼ h=k; B2 ¼ aa =k: ð10Þ


1190 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

3. The Green’s function solution

The governing equation and boundary condition presented in the proceeding chapter is solved by Green’s
function method. The Green function solution equation (GFSE) is derived using the Eqs. (3)–(9) and an aux-
iliary problem for an instantaneous heat source inside the workpiece. The solution to the auxiliary problem is
the Green’s function G(x, y, tjx 0 , y 0 , s), where the instantaneous source is located at position (x 0 , y 0 ) and at time
s; and (x 0 , y 0 ) is the location at which the temperature is observed at time t. The auxiliary problem has homo-
geneous boundary conditions and a zero initial temperature as described by

0.1000%

0.0900%

0.0800%
Percent relative error E

0.0700%

0.0600%

0.0500%

0.0400%

0.0300%

0.0200%

0.0100%

0.0000%
0

0
0

0
0

0
0

0
10

60

10

60
10

60
31

36
41

46

51
56

61

66
71

76

81

86

91

96
10
10

11

11
12

12
Integration point, N

Fig. 3. Percent relative error with varying number of integration points.

1200

N=1290
1000
N=100
N=100
800
N=500
Temperature θ

N=1140
N=1180
600
N=1200
N=1240
400 N=1260
N=1290

200

0
05

15

25

35

45

55

65

75

85

95
0. 0

0. 1

0. 2

0. 3

0. 4

0. 5

0. 6

0. 7

0. 8

0. 9

01
00

00

00

00

00

00
00

00

00
00

00

00

00

00

00
00

00

00

00

0.
0.

0.

0.

0.

0.

0.
0.

0.

0.

Axial distance, x

Fig. 4. Evaluation of numerical integration for temperature with varying number of integration point.
P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1191

1 1 oG
r20 G þ dð~ r ~ r0 Þdðt  sÞ ¼  ðt > sÞ; ð11Þ
a a os
Gðx0 ; y 0 ; sjx; y; tÞ ¼ 0 ðt < sÞ; ð12Þ
oG
k i 0 þ hi G ¼ 0 ðt > sÞ; ð13Þ
oni
R
where d is the Direc delta function such that R dð~ r  eÞd~ r ¼ 1 and dð~r  eÞ ¼ 0 except at e. A general proce-
dure is applied to obtain the two-dimensional Green’s function solution equation (GFSE). Applying initial
and boundary conditions, and simplifying each term leads to the two-dimensional Green’s function solution
equation (GFSE) as
Z 2L Z H Z Z Z H
a t 2L
hðx; z; tÞ ¼ Gðx; z; tjx0 ; z0 ; sÞh0 dx0 dz0 þ Gðx; z; tjx0 ; z0 ; s0 Þgðx0 ; z0 ; sÞ
0 0
x ¼0 z ¼0 k 0 0 0
x ¼0 z ¼0
Z t Z LþR 02
0 x 2 0
þ I0 Gðx; z; tjx ; H ; sÞe R dx ds: ð14Þ
s¼0 x0 ¼LR

0
0
0
0
0
0
0
0
0
0
0
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
-10000000
-20000000
-30000000 N=100
N=100
Heat flux qz

N=1290 N=500
-40000000 N=1140
N=1180
-50000000
N=1200
-60000000 N=1240
N=1260
-70000000 N=1290

-80000000
-90000000
Axial distance, x

Fig. 5. Evaluation of numerical integration for heat flux with varying number of integration point.

1200

1000 0(GF)
Surface Temperature

0(FEA)
800

600

400

200

0
0.001 0.002 0.003 0.004 0.005 0.005 0.005 0.004 0.003 0.002 0.001
Axial distance, x

Fig. 6. Evaluation of surface temperature by Green’s function and finite element code.
1192 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

3.1. Green’s function

In order to evaluate Eq. (14) for analyzing two-dimensional laser heating problems, the Green’s function G
has to be obtained. This function G can be expressed as a function of two other Green’s functions G1 and G2
due to the property of multiplication as follows:
G ¼ G1  G2 ; ð15Þ

where G1 and G2 are Green’s function along x- and z-direction, respectively. Because there are two kinds of
boundary conditions along the z-direction, the corresponding Green’s function can be written as
G2 ¼ G02 þ G002 ; ð16Þ

0
0.001 0.002 0.003 0.004 0.0045 0.005 0.0045 0.004 0.003 0.002 0.001
-5000000

-10000000
Heat flux, qz

-15000000

-20000000

-25000000 qz(GF)
qz(FEM)
-30000000

-35000000

-40000000
Axial distance, x

Fig. 7. Evaluation of heat flux qz by Green’s function and finite element code.

1200

1000 R=0.0003
R=0.0004
Surface temperature

R=0.0005
800
R=0.0035

600

400

200

0
0. 5

0. 5

0. 5

0. 5

0. 5

0. 5

0. 5

0. 5

0. 5

95
0. 0

0. 1

0. 2

0. 3

0. 4

0. 5

0. 6

0. 7

0. 8

0. 9

01
0

8
00

00
00

00

00

00

00

00

00
00

00

00

00

00

00
00

00

00

00

0.

Fig. 8. Effect of laser beam radius on surface temperature distribution.


P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1193

where G02 is the Green’s function associated with boundary condition oh


oz
þ B1 h ¼ 0 in Region I (0 6 x 6 L  R,
2
x2
L + R 6 x 6 2L). G002 is the Green’s function associated with boundary condition oh
oz
¼ I 0e R in Region II
(L  R 6 x 6 L + R).
So the Green’s function for Eq. (14) becomes

G ¼ G1 ðG02 þ G002 Þ: ð17Þ

In general, two-dimensional Green’s functions can be found by simple multiplication of one-dimensional


Green’s functions for the rectangular co-ordinate system for most of the boundary conditions provided that
the body is homogeneous. The Green’s function needed for the solution is represented by the solution of the
homogeneous linear problem derived for plane slab geometry. Several one-dimensional Green’s functions used
in the solution the GFSE are given below The Green ’s function G1 is derived from the solution of an initial
value problem with zero temperature boundary conditions and arbitrary initial condition as

0.005 0.005

21.491 85.964 257.892


236.40185.964 21.491 14.594 72.971 204.319379.45218.913 72.971 14.594
0.004 0.004 29.188 87.565 145.942 116.754 43.783
42.982107.455214.91 171.928 64.473
64.473128.946150.437 43.783 102.159
58.377
0.003 42.982 0.003 29.188

21.491
14.594
0.002 0.002

0.001 0.001

0 0
0 0.002 0.004 0.006 0.008 0.01 0 0.002 0.004 0.006 0.008 0.01

(a) R = 0.0003 m (b) R = 0.0004 m

0.005 0.005

64.316 4.1466.219 8.292 14.51


10.719 53.59796.475182.23 139.352 10.36412.437 10.364
0.004 21.439 64.316 107.194 53.597 10.719 0.004 2.073 8.292 4.146
32.158 75.036 85.755 32.158 6.219
42.878 4.146 2.073
0.003 0.003
21.439 2.073

10.719
0.002 0.002

0.001 0.001

0 0
0 0.002 0.004 0.006 0.008 0.01 0 0.002 0.004 0.006 0.008 0.01

(d) R = 0.0035 m (c) R = 0.0005 m

Fig. 9a. Isotherms with different beam radius.


1194 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

1X 1
2 2 bx b x0
Gðx; tjx0 ; sÞ ¼ ebn aðtsÞ=4L sin n sin n ; ð18Þ
L n¼1 2L 2L

where the bn values are


bn ¼ np; n ¼ 1; 2; . . . ð19Þ

The Green’s function G02 is derived from the solution of a boundary value problem with adiabatic boundary
condition at Z = 0 and convective boundary condition at Z = H, and an arbitrary initial condition as
 z  
2 X1
2 2 b2 þ B2 z0
G02 ðz; t; z0 ; sÞ ¼ ebm ðtsÞ=H 2 m 2 2 cos bm cos bm ; ð20Þ
H m¼1 bm þ B2 þ B2 H H

where bm is characteristics values and are root of the equation given by


bm tan bm ¼ B2 ð21Þ

0.005 0.005

950
750 850 550600 650
0.00492 650 900 0.00492
800 500
850 600 550
750 650
700 800 700
0.00484 0.00484 550
600 750 450 500
650 500
550 700 450
0.00476 0.00476
650 600
400 450
600
0.00468 550 0.00468
550
500 500 400

0.0046 0.0046
0.0046 0.0048 0.0049 0.00508 0.00524 0.0054 0.0046 0.0048 0.0049 0.00508 0.00524 0.0054

(a) R = 0.0003 m (b) R= 0.0004 m

0.005

450 500
0.00492
400 400
450

0.00484
350 400 350

0.00476
350
300
0.00468
300

0.0046
0.0046 0.0048 0.0049 0.00508 0.00524 0.0054

(c) R = 0.0005 m

Fig. 9b. Enlarged view of isotherm with different beam radius.


P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1195

and B1, B2 are constants given by


B2 ¼ B1 H ; B1 ¼ h=k: ð22Þ
00
The Green’s function G2 is derived from the solution of a boundary value problem with an arbitrary initial
condition and adiabatic boundary conditions as
" #
1 X1
mpz mpz0
00 0 m2 p2 ðtsÞ=H 2
G2 ðz; t; z ; sÞ ¼ 1þ2 e m cos cos ; ð23Þ
H m¼1
H H

where
bm ¼ mp; m ¼ 1; 2; . . . ð24Þ
Substituting these Green’s functions into Eq. (17), we have following solution for temperature distribution in
the workpiece subjected to a CW laser beam:
Z H Z LR Z LþR Z 2L 
0 0 0 00 0 0 0
hðx; z; tÞ ¼ h0 dz G1 G2 dx þ G1 G2 dx þ G1 G2 dx
z0 ¼0 x0 ¼0 x0 ¼LR x0 ¼LþR
Z t Z LþR Z t Z LR Z H
02
x 2 a
þ I0 ds G1 G002 e R dx0 þ G1 G02 gðx0 ; z0 ; sÞdz0 dx0 ds
s¼0 x0 ¼LR k 0 x0 ¼0 z0 ¼0
Z t Z LþR Z H Z t Z 2L Z H
a a
þ G1 G002 gðx0 ; z0 ; sÞdz0 dx0 ds þ G1 G02 gðx0 ; z0 ; sÞdz0 dx0 ds: ð25Þ
k 0 x0 ¼LR z0 ¼0 k 0 x0 ¼LþR z0 ¼0

It can be noted here that the Green’s function solution given by Eq. (25) is also applicable to cases where the
heat generation term varies with space and time. So, it can provide solution for problems with non-uniform
heat generation due to a partially transparent materials. It can also provide solutions for cases with temper-
ature dependent heat generation due to the variation of optical properties such as absorptivity and transmis-
sivity. Also, the solution can be extended to a three-dimensional analysis with use of a three-dimensional
Green’s function.

1200

w = 1 mm
w = 5 mm
1000
w = 9 mm

800
Surface temperature θ

600

400

200

0
0
0
0
0
0
0
0
0
0
0
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01

Fig. 10. Effect of defocusing distance.


1196 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

4. Results and discussion

4.1. Numerical integration

Green’s function solution equation (GFSE) for 2D laser heating problem involves a two-dimensional inte-
gration formula. These integrals are performed using Romberg’s integration algorithm. One advantage of this
algorithm is that the method can be efficiently used to obtain results of any desired accuracy. The results are
improved iteratively by combining two successive integration values with number of integration points as N
and 2N until percent relative ea is less than assigned tolerance limit es = 0.01%. Results for maximum temper-
ature h and corresponding percent relative error are calculated with increase in integration points. Fig. 3 shows
variation of percent relative error with increase in number of integration points. Results show that the percent
relative decreases exponentially and reaches a converged value of less than 1% and maximum temperature of
1096 K for N = 1290.

0.005 0.005

104.661
20.93241.864 167.458 293.051 18.86937.738 150.953 264.167
94.345
0.004 104.661 188.39 146.526 62.797 20.932 0.004 94.345 169.822 132.084 56.607 18.869
62.797 125.593 56.607 113.214
83.729 75.476
0.003 41.864 0.003 37.738

20.932 18.869
0.002 0.002

0.001 0.001

0 0
0 0.002 0.004 0.006 0.008 0.01 0 0.002 0.004 0.006 0.008 0.01

(a) w = 1 mm (b) w = 5 mm

0.005

85.427
17.08534.171 136.683 239.195
0.004 85.427 153.768 119.597 51.256 17.085
51.256 102.512
68.341
0.003 34.171

17.085
0.002

0.001

0
0 0.002 0.004 0.006 0.008 0.01

(c) w = 9 mm

Fig. 11a. Isotherm with different defocusing distance.


P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1197

The surface temperature h and surface heat flux are evaluated with increase integration points N, and
results are presented in Figs. 4 and 5, respectively. It can be seen that distributions of surface temperature
and surface heat reaches convergence for number of integration points at around 1290.

4.2. Comparison of Green’s function solution with finite element results

The analytical solution given by the Green’s function method is compared with that given by the commer-
cial finite element code ANSYS 5.3. The finite element model was developed for the numerical solution of
two-dimensional laser heating problem. A mesh refinement study was conducted to establish the accuracy
and convergence of the finite element code. Mesh was refined from 100 · 25 to 200 · 100, and, results for
surface temperature and surface heat flux converge with an increase in number of elements to 200 · 50 with
percent relative error less than 0.07%.
In this section, analytical results given by Green’s function are compared with that given by the commercial
finite element code. Results for surface temperature and heat flux are presented in Figs. 6 and 7. Results show
close agreement between results estimated by Green’s function and FEA with errors less than 1%. Further
improvement can be achieved by using finer mesh size distribution in FEA.

0.005
0.005
800 850 800
750 950 750 600 750 600
900 0.00492
0.00492 800 550 650 750 650
650 850
700 700
600 700 800 550
0.00484
0.00484 750 600
500 600 650 600
550 700 650
650 0.00476 550
0.00476
600 550
600
550 500
0.00468 0.00468 500
550
500 500 450 450

0.0046 0.0046
0.0046 0.0048 0.0049 0.00508 0.00524 0.0054 0.0046 0.0048 0.0049 0.00508 0.00524 0.0054

(a) w = 1 mm (b) w = 5 mm

0.005

550 750 600


650
0.00492 700
500 600 550
650
500
0.00484
550 600
450
550
0.00476 500

500
450
0.00468 450
400 400

0.0046
0.0046 0.0048 0.0049 0.00508 0.00524 0.0054

(c) w = 9 mm

Fig. 11b. Enlarged view of isotherms with different defocusing distance.


1198 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

4.3. Parameter study using Green’s function solution equation

A parameter study is carried out to analyze the effect of different laser beam parameters such as beam
radius and defocused distance using Carbon steel workpiece. A comparison study is also carried out using
two different materials: carbon steel and aluminum. Laser beam parameters and material properties are given
in Table 1.

4.4. Effect of laser beam radius

The beam radius along with beam power determines the power intensity distribution. The beam spot size is
varied by using lenses of different focal lengths. In this study, three different beam sizes, 0.3 mm, 0.4 mm,
0.5 mm and 3.5 mm are considered. Results for surface temperature, h and surface heat flux qz with different
laser beam parameters were shown in Fig. 8. Results show that surface temperature distribution closely resem-
bles the Gaussian distribution of the incident laser beam. The maximum surface temperature increases from
590 C to 1079 C as the beam radius decreases from 0.5 mm to 0.3 mm. It can also be noticed that at a beam

Table 1
Properties of materials and laser beam parameters
Material Melting Density Specific Thermal Thermal diffusion h w/(m2 C)
point (C) q (kg/m3) heat C (kJ/kg C) Conductivity a (m2/s) · 105
k (w/m C)
Carbon steel 1537 7,801 0.437 43 1.172 5
Aluminum 983 2,702 0.903 237 9.71 5

Laser beam radius R0 (m) Width of Work piece H (m)


2
0.003 · 10 0.5

1200

1000 steel
aluminum
Surface temperature θ

800

600

400

200

0
05

15

25

35

45

55

65

75

85

95
0. 0

0. 1

0. 2

0. 3

0. 4

0. 5

0. 6

0. 7

0. 8

0. 9

01
00

00

00

00

00

00
00

00

00
00

00

00

00

00

00
00

00

00

00

0.
0.

0.

0.

0.

0.

0.
0.

0.

0.

Fig. 12. Effect of material on temperature distribution.


P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200 1199

0.005 0.005

61.584 102.64 205.28 46.312 169.809 355.055


20.528 41.056 184.752 15.437 61.749 77.186
0.004 82.112123.168
143.696
102.64 0.004 123.497200.683 138.935 15.437
30.874 108.06
82.112 77.186 154.372
41.056 92.623 61.749
0.003 20.528 0.003 46.312
61.584 30.874

0.002 0.002 15.437


41.056

0.001 20.528 0.001


20.528

0 0
0 0.002 0.004 0.006 0.008 0.01 0 0.002 0.004 0.006 0.008 0.01
(a) Aluminum (b) Carbon steel

Fig. 13. Isotherm with different material.

radius of 3.5 mm, the beam power intensity is not high enough to cause any noticeable temperature rise. How-
ever, with the use of a smaller beam radius, higher beam intensity can be achieved that causes in a significantly
high temperature rise in the material. A smaller beam radius not only causes high surface temperature, but
also creates a wider and deeper heated region. This is more evident from the isotherm plots given in Figs.
9a and 9b. The maximum depth and width of the heated zone increases from 0.0025 m · 0.0023 m to
0.003 m · 0.0025 m as the radius decreases from 0.5 mm to 0.3 mm.

4.5. Effect of laser beam defocusing distance

So far, we have considered material surface to be located at the focal plane of the laser beam, i.e. with a
defocused distance of W = 0. As the defocused distance is varied, the beam power intensity is computed based
on the beam spot size given by Eqs. (1) and (2). In the present study, simulation is carried out for a defocused
distance of w = 1 mm, 3 mm and 5 mm with R(0) = .3 mm, and results are presented in Figs. 10, 11a and 11b.
As expected, an increase in defocused distance results in an increase in spot size, but with a decrease in
beam power intensity. This result in a decrease in the temperature rise and decrease in the size of the heated
region. This study shows that the heated region as well as the heat-affected region (HAZ) can be controlled in
a heat treatment process by selecting a suitable combination of the beam focusing lens and defocusing
distance.

4.6. Effect of work piece materials

Heated region as well as HAZ depends not only on the beam and optical properties but also on the thermo-
physical properties of the workpiece materials. In order to see these effects on the heated region or the HAZ,
simulation is carried out using two different materials: carbon steel and aluminum. Fig. 12 shows a signifi-
cantly higher temperature rise in steel than in aluminum. However, the heated region is much wider and dee-
per for aluminum due to its higher thermal diffusivity value than steel as depicted in Fig. 13.

5. Conclusions

An analytical model based on Green’s function method is developed to analyze the temperature distribu-
tions and heat effected zones in materials irradiated by a high-energy laser. A parametric study shows the
significant effect of parameters such as beam power diameter, defocused distance and material properties
on the heat shape and size of heated region or HAZ. The study demonstrates the effectiveness of the Green’s
1200 P. Majumdar, H. Xia / Applied Mathematical Modelling 31 (2007) 1186–1200

function-based analytical model as an effective tool in heat treatment process for deciding operating condi-
tions for a specific material. The solution could be extended to cases that involve melting by transforming
the temperature-based heat equation into an enthalpy-based heat equation.

References

[1] K. Brugger, Exact solution for the temperature rise in a laser-heated slab, J. Appl. Phys. 43 (1972).
[2] R.E. Warren, M. Spark, Laser heating of a slab having temperature-dependent surface absorption, J. Appl. Phys. 50 (1990).
[3] H.E. Cline, T.R. Anthony, Heat treating and melting with scanning laser or electron beam, J. Appl. Phys. 48 (1977).
[4] F.W. Dabby, U.C. Peak, High-intensity laser-induces vaporization and explosion of solid, IEEE J. Quantum Electron QE-8 (1972).
[5] J.E. Rogerson, G.A. Chayt, Total melting time in the ablating slab problem, J. Appl. Phys. 42 (1971).
[6] A. Minardi, P.J. Bishop, Temperature distribution within a metal subjected irradiation by a laser of spatially varying intensity
Heat Transfer in Manufacturing and Materials Processing, HTD vol. 113, ASME, New York, 1989, pp. 39–44.
[7] T.F. Lemczyk, M.M. Yovanovich, Thermal constriction resistance with convection boundary conditions – 1. Half-space contacts, Int.
J. Heat Mass Transfer 31 (9) (1988) 1861–1872.
[8] T.F. Lemczyk, M.M. Yovanovich, Thermal constriction resistance with convection boundary conditions – 2. Layered half-space
contacts, Int. J. Heat Mass Transfer 31 (9) (1988) 1873–1883.
[9] Ph. Barrilot, Numerical simulation of crater formation heating by laser beam, J. Numer. Heat Transfer, Part B 17 (1990) 245–256.
[10] M. Kaylon, B.S. Yilbus, Analytical solution for laser evaporative heating process: time exponentially decaying pulse, J. Appl. Phys. 34
(2001) 3303–3311.
[11] B.S. Yilbas, M. Kaylon, Analytical solution for pulsed laser heating process: convective boundary condition case, Int. J. Heat Mass
Transfer 45 (2002) 1571–1582.
[12] M. Kalyon, B.S. Yilbus, Closed form solution for exponentially decaying laser pulse heating evaporation at the surface, Jpn. J. Appl.
Phys. 41 (2002) 3737–3746.
[13] A. Yanez, J.C. Alvarez, A.J. Lopez, G. Nicolas, J.A. Perez, A. Ramil, E. Saavedra, Modelling of temperature evolution on metals
during laser hardening process, J. Appl. Surf. Sci. 186 (2002) 611–616.
[14] G. Gustavo, A.J. Guillermo, Analytical solution for transient three-dimensional temperature distribution in laser assisted machining
processes, in: Proceedings of the ASME Heat Transfer/Fluid Engineering Summer Conference, HT/FED 2004, 2004, pp. 1055–1063.
[15] D.I. Pantelis, Development and experimental validation of analytical thermal models for the evaluation of the depth of laser-treated
zones, J. Appl. Phys. A67 (1998) 435–439.

You might also like