You are on page 1of 12

International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

www.elsevier.com/locate/ijhmt

Modelling of radiation transfer in metallic powders


at laser treatment
a,c,* b
A.V. Gusarov , J.-P. Kruth
a
Russian Academy of Sciences, Baikov Institute of Metallurgy, Leninsky Prospect 49, 119991 Moscow, Russia
b
Department of Mechanical Engineering, University of Leuven (K.U. Leuven), PMA, Celestijnenlaan 300B, 3001 Leuven, Belgium
c
Graduate School of Engineering, Department of Aeronautics Astronautics, Kyoto University, Kyoto 606-8501, Japan

Received 12 December 2003; received in revised form 10 November 2004

Abstract

Absorptances and deposited energy profiles are calculated for thin layers of metallic powder placed on a reflective
substrate at normal incidence of collimated radiation. Radiation transfer equation is analytically solved by the two-flux
method. The effective optical parameters of the powder are estimated in geometrical optics approximation taking into
account dependent scattering. The cases of specularly and diffusely reflecting particles are studied. Due to multiple
reflections in an open pore system, laser radiation can penetrate in powder to considerable depths, much greater than
the characteristic particle diameter. Total laser energy absorbed in a thin powder layer on a reflective substrate increases
with its thickness but the deposited energy density decreases. Generally, the absorptance and the energy density for
specular reflection are slightly greater than these values for diffuse reflection. The results obtained in the limit of deep
powder bed essentially agree with known ray tracing simulations. The present model is in good agreement with exper-
imentally observed correlation between effective absorptance of metallic powders and absorptance of corresponding
dense metals.
Ó 2005 Elsevier Ltd. All rights reserved.

Keywords: Powder bed; Radiation transfer equation; Dependent scattering; Effective absorptance; Deposited energy

1. Introduction cladding [1]. Such rapid prototyping and manufacturing


methods as selective laser sintering [2] and selective laser
Consolidation of loose powder by local laser heating melting [3] are being developed. The essential operation
is becoming a promising manufacturing technique be- is laser beam scanning over the surface of a thin powder
cause of easy control over both powder deposition and layer previously deposited on a substrate. Powder bind-
laser radiation. One of the well-known examples is laser ing mechanisms, such as melting and solid-state and
liquid-phase sintering, strongly depend on temperature
and so estimating local temperature fields is very impor-
*
Corresponding author. Address: Graduate School of Engi-
tant for process control.
neering, Department of Aeronautics Astronautics, Kyoto The first problem arising when one attempts to calcu-
University, Kyoto 606-8501, Japan. Tel.: +81 75 753 4943; late the temperature induced by a scanning laser beam is
fax: +81 75 753 4942. laser energy deposition. This includes not only an
E-mail address: av.gusarov@relcom.ru (A.V. Gusarov). absorbed fraction of laser energy but also a depth

0017-9310/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2005.01.044
3424 A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

Nomenclature

A absorptance u dimensionless deposited energy density


a eigenvalue factor w mass fraction
C1 , C2 constants in general solution x radius vector
c fraction of collimated radiation scattered x, y, z coordinates
into the forward hemisphere
D particle diameter Greek symbols
F angular intensity of diffuse radiation a scattering angle
f dimensionless angular intensity of diffuse b extinction coefficient
radiation c scaling factor
G collimated radiation intensity d theoretical density
g dimensionless collimated radiation intensity  porosity
h fraction of radiation uniformly distributed h polar angle
over a hemisphere that remains in the hemi- j absorption coefficient
sphere after scattering K wavelength
I angular radiation intensity k optical thickness
L layer thickness n dimensionless coordinate
n number of particles per unit volume q reflectance
P scattering phase function r scattering coefficient
Q net radiative energy flux u azimuth angle
q dimensionless net radiative energy flux v angle of incidence
R particle radius w angle of reflection
S specific area of powder per unit pore volume X unit vector of direction
s directional specific area per unit pore vol- x albedo
ume
U specific surface area per unit mass

distribution of the absorbed energy because the incident In absorbing and scattering medium the angular radi-
radiation penetrates into the powder due to multiple ation intensity is governed by the radiation transfer
reflections in an open pore system. The penetration is equation (RTE) [6]:
especially important for highly reflective metallic pow- Z
r
ders as indicated by ray tracing simulation [4]. The ray XrI ¼ ðr þ jÞI þ IðX0 Þ P ðX0 ! XÞ dX0 ; ð2Þ
4p 4p
tracing technique [4,5] requires specifying the detailed
powder structure by particles shape, dimension, and where r and j are the scattering and absorption coeffi-
their coordinates that is a complicated problem itself cients, respectively, and P(X 0 ! X) the scattering phase
and requires additional assumptions. To avoid such a function, which represents the probability that the radi-
problem in this work, powder is considered as a homog- ation propagating in direction X 0 is scattered to direc-
enous absorbing and scattering continuum with effective tion X. The phase function is normalised:
radiation transfer properties equivalent to those of a Z
powder bed. 1
P ðX0 ! XÞ dX0 ¼ 1. ð3Þ
Radiation in the powder is characterized by the dis- 4p 4p
tribution of its angular intensity, I(x, X), so that
XI(x, X) dX is the energy flux density transferred by Instead of the pair r and j one often use the extinction
photons directed about the unit vector X within solid coefficient b = r + j and albedo x = r/b.
angle dX at point x. Effective value of the intensity is This article deals with the general problem of normal
meant when considering a heterogeneous medium such incidence of collimated radiation on a thin powder layer
as loose powder. That is, the local value is averaged over placed on a reflecting substrate. This configuration is
a volume much greater than a particle. The net radiative typical for laser manufacturing [2]. Effective radiation
energy flux density is defined as an integral over all transfer properties of metallic powders are theoretically
directions: estimated in Section 2. The RTE equation (2) is analyt-
Z ically solved by the two-flux method [7] in Sections 3 and
Q¼ XI dX. ð1Þ 4. In Section 5, the obtained results on energy deposition
4p profiles are compared with ray tracing simulation [4,8]
A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434 3425

and the results on absorptance are compared with exper- only restriction that the curvature radius of the surface
iments [9]. is greater than the wavelength. According to Eq. (7), this
allows applying geometrical optics. The structure of PB
is characterized by such parameters as porosity e (vol-
2. Effective radiation-transfer properties of metallic ume fraction of pores) and specific surface area (total
powders surface area of particles) per unit mass U. In the studied
powder beds of opaque particles, radiation can not pen-
Light scattering by spherical particles is generally de- etrate into particles and is concentrated in pores. There-
scribed by the Mie theory [10]. Currently, powders most fore, specific surface area per unit pore volume
widely used in laser manufacturing consist of particles S = d(1  e)Ue is introduced with d being the density
with diameter D of several tens microns [2]. Scattering of the solid phase (theoretical density). Distribution of
of laser radiation by such large particles often reduces specific surface S over directions is specified by direc-
to refraction and reflection in the geometrical optics tional volumetric specific surface area of the powder
approximation. Thus, in a cloud of opaque spheres of s(X), so that s(X) dX is the area of the particlesÕ surface
radius R considered as independent scattering centres, with the external normal about unit vector X within
extinction coefficient bi and albedo xi are [7]: solid angle dX per unit pore volume of PB.
bi ¼ pR2 n; xi ¼ q; ð4Þ Consideration of radiation transfer through a thin
powder layer gives its effective parameters. Thus, the
where n is the number of particles per unit volume and q extinction and scattering coefficients are obtained by
the hemispherical reflectance of the particle surface. It is averaging over directions of the surface normal within
noteworthy that the extinction coefficient is determined a hemisphere where incidence angle v is less than p/2
by geometrical parameters only and is independent of (see Fig. 1):
the material and wavelength. In contrast, the albedo, Z
which is equal to reflectance, is independent of the b¼ cos vsðXÞ dX; ð8Þ
geometry but includes the mentioned physical depen- 2p

dences. The specular and diffuse reflection laws give Z


phase functions Ps and Pd, respectively [7]: r¼ q0 ðvÞ cos v sðXÞ dX; ð9Þ
p  a. 2p

P s ¼ q0 q; ð5Þ where dX = du d cos v and u is the azimuth angle.


2
8
Pd ¼ ðsin a  a cos aÞ; ð6Þ
3p
where a is the scattering angle and q 0 (v) the directional
specular reflectance depending generally on incidence
angle v = (p  a)/2. Diffuse reflection is independent of
incidence angle. Eq. (5) gives isotropic scattering in case
of directional reflectance independent of incidence angle.
It is known [10] that taking into account small-angle
diffraction effects at particle diameter D much greater
than wavelength K, leads to doubling extinction coeffi-
cient (4) while a strong forward-directed diffraction term
arises in phase function (5) and (6). However, the signif-
icant difference is only apparent. The issue is whether to
consider the small-angle diffraction as extinction or to
neglect it. In practice, the geometrical optics formulas
(4)–(6) can be applied [7] when
pD=K P 5. ð7Þ
The simple theory described above is valid only if the
particles are spherical and each particle scatters radia-
tion independently of its neighbours.
Fig. 1. Schemes of scattering: (a), specular reflection, fixed
surface normal orientation for a given scattering direction; (b)
2.1. Statistical scattering model diffuse reflection, the radiation scattered into a specified
direction arises from surface elements oriented within the
Consider a powder bed (PB) composed of opaque shadowed region of the unit sphere. Angles: v, incidence; w,
particles of arbitrary shapes and dimensions with the reflection; a, scattering; u, azimuth.
3426 A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

In case of specular reflection, scattered radiation 2.2. Powder mixtures


intensity at scattering angle a and azimuth angle u is pro-
portional to surface area at the direction satisfying the The above model can be extended to powder mix-
specular reflection condition with incidence angle v = tures. Let the mixture of porosity e be composed of K
(p  a)/2 and azimuth angle u as shown in Fig. 1(a): components with mass fractions wk, theoretical densities
dI(a, u)  q 0 (v)cosvs(v, u)dudcosv  q0 (v)s(v, u)dudcosa. dk, and specific surface areas per unit mass Uk,
Taking into account normalising condition (3), the k = 1, . . . , K. Total specific area is
scattering phase function at specular reflection is X
U¼ U k wk . ð16Þ
expressed as
pq0 ðvÞsðv; uÞ The scattering properties of each component are
P s ða; uÞ ¼ R 1 R 2p . ð10Þ characterised by albedo xk and phase function Pk.
0
q0 ðvÞ cos v d cos v 0 sðv; uÞ du
Extinction coefficient b, albedo x, and phase function
Diffuse reflection gives the scattered intensity propor- P of the mixture are then calculated as
tional to the integral over surface elements with normals 1e U
oriented within the shadowed region of the unit sphere b¼ P ; ð17Þ
4e wk =dk
as shown in Fig. 1(b):
Z p=2þa Z 1 P
xk U k wk
P d ða; uÞ  da0 q0 ðvÞ x¼ ; ð18Þ
p=2 1
U
P
 cos v cos w sðu0 ; a0 Þ d cos u0 . ð11Þ xk P k U k wk
P¼ . ð19Þ
where v is the incidence and w the reflection angles. Inte- xU
gration is performed in Eq. (11) in a spherical coordinate Eqs. (17)–(19) estimate radiation transfer properties
system with the polar angle u 0 and the azimuth angle a 0 of a powder mixture with known properties of its
and the axis u 0 = 0 corresponding to the scattering polar components.
angle a = p/2 and the scattering azimuth angle u  p/2.
The incidence and reflection angles are obtained from 2.3. Dependent scattering
the following equations:
In dense PBs where particles touch each other, radi-
cos v ¼ sin u0 cos a0 ; cos w ¼ sin u0 cosða0  aÞ. ð12Þ
ation transfer can not be treated as scattering by sepa-
If there is no preferential orientation of particles, sur- rate particles [11]. This effect is referred to as
face area distribution is isotropic: dependent scattering and has been studied in a number
sðXÞ dX ¼ S dX=ð4pÞ; ð13Þ of experimental [12–14] and theoretical [5,11,15,16]
works.
where S is the specific surface area of the powder per For a PB of equal spherical particles with radius R,
unit pore volume. Then Eqs. (8) and (9) give the follow- specific area per unit pore volume is
ing expressions for extinction b and scattering r coeffi-
cients and albedo x = r/b: 4pR2 n
S¼ ; ð20Þ
e
b ¼ S=4; r¼q S=4; x ¼ q; ð14Þ
where n is the number of particles per unit volume of PB
where the definition of the hemispherical reflectance [7], and e = 1  4pR3n/3 is the porosity (total volume of
Z pores per unit volume of PB). This gives the following
1
q¼ q0 ðvÞ cos v dX; ð15Þ relation between independent scattering by opaque
p 2p spheres (4) and the statistical model (14):
is taken into account. One can see from Eq. (14) that b ¼ cbi ; r ¼ cri ; x ¼ xi ; ð21Þ
only a quarter of the powder surface is effective in scat-
tering. A factor of 1/2 arises because, in average, only a where a scaling factor is introduced,
half of particle surface is exposed to the radiation of a
c ¼ 1=e. ð22Þ
specified direction. Another factor of 1/2 originates from
integrating of cos v over a hemisphere. Physically, this Eqs. (21) and (22) reduce to independent scattering
means that not the surface but its cross-section visible (4) at high porosities, e ! 1, when scaling factor c tends
in the direction of irradiation is important. In the isotro- to unity. A considerable difference arises in dense PBs
pic case, integration in Eqs. (10) and (11) gives the same due to dependent scattering, which is actually a mutual
result as for spherical particles. The phase function is influence of scattering particles. Dependent scattering
given by Eq. (5) for specular reflection and by Eq. (6) has no effect on the phase function according to the pres-
for diffuse reflection. ent model. A simple interpretation of Eq. (22) is that
A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434 3427

radiation does not penetrate into opaque particles and factor of 1.5 at the porosity e = 0.4, which is typical for
so the apparent concentration of scattering centres powders employed in rapid manufacturing [2].
should be calculated not per unit volume of PB but Thus, the DOTS model is an example where the sim-
per unit volume of pores. ple statistical scattering model described in Section 2.1
Singh and Kaviany [5] obtained similar scaling rela- and Section 2.2, is not accurate. This discrepancy may
tions (21) describing their ray tracing simulation results originate from the general limitation of description of
in a wide range of porosity 0.3 < e < 1. They also con- heterogeneous medium by effective properties. In partic-
cluded that dependent scattering does not influence the ular, the extinction length is comparable with particlesÕ
phase function and that the scaling factor is a function size in dense PBs. On the other hand, the DOTS struc-
of porosity. The analytical approximation of the func- ture is rather complicate that can correspond to powders
tion c(e) derived in [5] differs from Eq. (22) but actually with highly irregular particles only. Powders with about
gives very close values in the considered porosity range spherical particles are better described by DOOS struc-
as shown in Fig. 2. ture where Eqs. (14), (21) and (22) are rigorous [16].
Kamiuto [15] attempted to estimate the dependent The same rigorous result can be obtained for non-over-
scattering effects theoretically and obtained different lapping opaque spheres in a transparent matrix consid-
scaling relations: ering the chord-length distribution estimated in [17].
b ¼ cbi ; ð1  xÞ ¼ ð1  xi Þ=c. ð23Þ Monte Carlo identification [16] indicates that, in gen-
eral, the phase function of a dense PB differs from that
However, they are equivalent to Eq. (21) in case of given by the independent theory, Eqs. (5) and (6). How-
highly reflective metallic powders with albedo x ! 1. ever, in case of DOOS model with the specular reflection
Function c(e) reported in [15] is plotted in Fig. 2. It law, the phase function becomes about isotropic at high
agrees with the present model at high porosities e > 0.7 reflectances [16]. This justifies applying Eq. (5) for metal-
and considerably differs for more dense powders. lic powders.
Tancrez and Taine [16] studied two models for por-
ous structures: Dispersed radius Overlapping Opaque
Spheres (DOOS) in a transparent matrix and Dispersed 3. Radiation transfer model
radius Overlapping Transparent Spheres (DOTS) in an
opaque matrix. They showed that Eq. (14) are rigorous Fig. 3(a) shows the general configuration of PB used
for optically thin layers of the both structures and for in modelling. A powder layer of thickness L is placed on
DOOS of arbitrary thickness. On the other hand, Monte a specularly reflecting substrate. Thus, one can simulate
Carlo investigation of the DOTS model [16] gives stron- thin powder layers on metallic substrates as well as deep
ger extinction in thick layers than predicted by Eq. (14). powder beds at L ! 1 where the influence of the sub-
This can be described by Eq. (21) where the scaling fac- strate is negligible. The Cartesian frame is introduced,
tor is higher than specified by Eq. (22). The difference in so that axis (OZ) is directed along the inner normal to
scaling factor increases with the density and reaches the the powder bed surface and axes (OX) and (OY) lay
on the surface. Thus, z-coordinate is in fact the depth.
Normally incident laser radiation is assumed to be dis-
tributed over a spot at the powder bed surface with a
size much greater than the radiation penetration depth.
Therefore, radiation intensity distribution in lateral

Fig. 3. Scheme of the considered configuration of powder bed


Fig. 2. Scaling factor to take into account dependent scattering: (a) and typical angular diagram of radiation intensity in the
full line, present model, Eq. (22); dashed line, theoretical two-flux approximation (b). Radiation components: g+, for-
estimations by Kamiuto [15]; dotted line, fit to ray tracing ward collimated; g, backward collimated; f+, forward diffuse;
simulation by Singh and Kaviany [5]. f, backward diffuse.
3428 A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

directions follows the distribution of incident laser en- The solution of this problem is:
ergy flux density over the surface G0(x, y) and the consid- Gþ ¼ G0 expðnÞ; G ¼ G0 expðn  2kÞ; ð31Þ
ered problem is one-dimensional. Lateral coordinates x
and y are omitted below to simplify notation, although where n = bz is the dimensionless coordinate and k = bL
they are actually taken into account through function the optical thickness. One can obtain from Eqs. (24),
G0(x, y). (26) and (27) the following boundary problem for diffuse
In one-dimensional case, the general RTE (2) reduces radiation:
to: of x x
 Z 1  cos h ¼ P r ð0; hÞgþ þ P r ðp; hÞg
oI x on 2 2
cos h ¼ b Iðh0 ÞP r ðh0 ; hÞ d cos h0  IðhÞ ; ð24Þ Z
oz 2 1 x 1
þ f ðh0 ÞP r ðh0 ; hÞ d cos h0  f ; ð32Þ
2 1
where angular radiation intensity I(z, h) and reduced
phase function Pr(h 0 , h) are independent of azimuth f ð0; hÞ ¼ 0; at h < p=2; ð33Þ
angle. In isotropic media, phase function P(a) depends
on scattering angle a only, like in Eqs. (5) and (6), and f ðk; hÞ ¼ f ðk; p  hÞ; ð34Þ
the scattering angle, in turn, depends on polar angles where the following dimensionless parameters are used:
of incident radiation, h 0 , and scattered radiation,h, and f = F/G0 is the diffuse radiation intensity and g+ = G+/
on the difference of azimuth angles u 0 : cos a = (2pG0) and g = G/(2pG0) the forward and backward
cos h cos h 0 + sin h sin h 0 cos u 0 . This allows defining the collimated radiation intensities, respectively. The first
reduced phase function as term in the right-hand side of Eq. (32) is the diffuse radi-
Z 2p ation source due to scattering of direct collimated radi-
1
P r ðh0 ; hÞ ¼ P ðaÞ du0 ; ð25Þ ation. The second term is due to scattering of back-
2p 0 reflected collimated radiation. The third term is the
Eq. (24) is solved in segment z = 0, . . . , L. Incident laser source due to scattering of the diffuse radiation from
radiation is assumed to be directed along the normal to other directions, and the forth term is the sink caused
the powder bed surface that gives the boundary condi- by scattering and absorption.
tion of normal incidence at z = 0: Eqs. (32)–(34) are solved by the two-flux method [7]
as follows. The diffuse radiation intensity is approxi-
G0 p mated as
Ið0; hÞ ¼ dðcos h  1Þ; at h < ; ð26Þ
2p 2
f ðn; hÞ ¼ fþ ðnÞUþ ðhÞ þ f ðnÞU ðhÞ; ð35Þ
where d(x) denotes the Dirac delta-function. At plane
z = L the boundary condition of complete specular where function U+(h) is equal to unity in the for-
reflection is used: ward hemisphere, h < p/2, and zero otherwise. Function
U(h), on the contrary, is equal to unity in the backward
IðL; hÞ ¼ IðL; p  hÞ. ð27Þ hemisphere, h > p/2, and zero in the forward one. As
shown in Fig. 3(b), total radiation intensity is approxi-
Angular radiation intensity can be expressed as the
mated by a superposition of two collimated and two dif-
sum
fuse components.
Gþ ðzÞ G ðzÞ Integrating equation (32) multiplied by U+(h) or
Iðz; hÞ ¼ dðcos h  1Þ þ dðcos h þ 1Þ þ F ðz; hÞ; U(h) over cos h from 1 to 1 gives two moment
2p 2p
ð28Þ equations:
1 dfþ
of collimated direct laser radiation and laser radiation ¼ ð1  hxÞfþ þ ð1  hÞxf þ x½cgþ þ ð1  cÞg ;
2 dn
reflected from the substrate with energy flux densities
ð36Þ
of G+ and G, respectively, and diffuse radiation F.
The diffuse term arises when the direct and back re- 1 df
flected laser radiation are scattered. Integrating Eqs.  ¼ ð1  hxÞf þ ð1  hÞxfþ þ x½ð1  cÞgþ þ cg ;
2 dn
(24), (26) and (27) over cosh in small vicinities of for-
ð37Þ
ward, h = 0, and backward, h = p, directions gives the
following boundary problem to obtain the direct and where
back-reflected collimated radiation: Z Z 1
1 1
h¼ d cos h P r ðh0 ; hÞ d cos h0 ; ð38Þ
dGþ dG 2 0 0
¼ bGþ ; ¼ bG ; ð29Þ
dz dz is the fraction of radiation uniformly distributed over a
hemisphere that remains in the hemisphere after scatter-
Gþ ð0Þ ¼ G0 ; Gþ ðLÞ ¼ G ðLÞ. ð30Þ ing and
A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434 3429

Table 1 Table 2
Coefficients of the two-flux equations for different scattering Dimensionless optical characteristics of powder
models
Conditions Name Definition
Coefficients h c
Thin powder layer Energy flux qk = Qz/G0
Specular reflection 1 1 on a reflective uk ¼ rQ=ðbG0 Þ
2 2 Deposited energy
substrate ¼ dqk =dn
1 1 Absorptance k
Ae ¼ qk ð0Þ
Diffuse reflection
3 6
Deep powder bed Energy flux q = limk!1 qk
Z u ¼ rQ=ðb G0 Þ
1 Deposited energy
1 ¼ dq=dn
c¼ P r ð0; hÞ d cos h; ð39Þ
2 0 Absorptance Ae = q(0)

is the fraction of collimated radiation scattered into the


forward hemisphere. Coefficients h and c corresponding
Dimensionless net energy flux density in z-direction
to the two particular cases of specular and diffuse reflec-
qk is obtained from Eq. (1):
tion specified by Eqs. (5) and (6) are listed in Table 1.
Specular reflection is assumed to be independent of inci- qk ¼ Qz =G0 ¼ pðfþ  f Þ þ 2pðgþ  g Þ
dence angle.
Eqs. (36) and (37) are to be solved with the following ð1  xÞ2  a2
¼
boundary conditions: ð1  hÞð1  4a2 Þ
3c þ 2xð1  h  cÞ þ ½1  c  2xðh  cÞ expð2kÞ
fþ ð0Þ ¼ 0; f þ ðkÞ ¼ f ðkÞ. ð40Þ 
ð1  x þ aÞ expð2akÞ  ð1  x  aÞ expð2akÞ
 fexp½2aðk  nÞ  exp½2aðn  kÞg
4. Results ð1  xÞ½1  4xðh  cÞ þ 2a2
 ½expðnÞ
1  4a2
The general solution of Eqs. (36) and (37) can be
 expðn  2kÞ; ð45Þ
expressed as
When k ! 1, Eq. (45) gives energy flux density in a
fþ ¼ C 1 expð2anÞ þ C 2 expð2anÞ deep powder bed:
1 x
 f½3c þ 2xð1  h  cÞ expðnÞ ð1  x  aÞ½3c þ 2xð1  h  cÞ
p 1  4a2 q ¼ lim qk ¼ 
þ ½1  c  2xðh  cÞ expðn  2kÞg; ð41Þ k!1 ð1  hÞð1  4a2 Þ expð2anÞ
ð1  xÞ½1  4xðh  cÞ þ 2a2
 ð46Þ
1 ð1  4a2 Þ expðnÞ
f ¼ ½C 1 ð1  hx  aÞ expð2anÞ
ð1  hÞx
þ C 2 ð1  hx þ aÞ expð2anÞ The definitions of powder optical characteristics used
1 x below are summarized in Table 2. Two of them are spec-
 f½1  c  2xðh  cÞ expðnÞ ified by Eqs. (45) and (46) and the others can be derived
p 1  4a2
from these equations. A simple expression is obtained
þ ½3c þ 2xð1  h  cÞ expðn  2kÞg; ð42Þ
for absorptance of a deep powder bed:

where a is the eigenvalue factor, a2 = 1  2hx  ð1  h  cÞð1  xÞ þ ½2ð1  hÞ þ ca


Ae ¼ ; ð47Þ
x2(1  2h). Constants C1 and C2 are found from bound- ð1  hÞð1 þ 2aÞ
ary conditions (40):
which reduces to
1 xð1  x þ aÞ
C1 ¼ 3a
p 1  4a2 Ase ¼ ; a ¼ ð1  xÞ1=2 ; ð48Þ
1 þ 2a
3c þ 2xð1  h  cÞ þ ½1  c  2xðh  cÞ expð2kÞ
 ; in the case of specular reflection, and to
1  x þ a  ð1  x  aÞexpð4akÞ
 1=2
ð43Þ 3 1  x þ 3a 2 1
Ade ¼ ; a ¼ 1  x  x2 ; ð49Þ
4 1 þ 2a 3 3
1xa
C 2 ¼ C 1 expð4akÞ. ð44Þ
1xþa in the case of diffuse reflection.
3430 A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

5. Discussion density $Q in iron powder calculated according to


Eqs. (45) and (46). Definitions of the corresponding
The statistical scattering model described in Section 2 dimensionless parameters are presented in Table 2.
is applicable to particles of arbitrary shape. According Two phase functions given by isotropic specular (5)
to this model, radiative properties (extinction and scat- and diffuse (6) reflection models are used. Two laser
tering coefficients and scattering phase function) in wavelengths of 1.06 and 10.6 lm are examined that cor-
packed beds of non-spherical particles without preferen- respond to Nd:YAG and CO2 lasers, respectively. Albe-
tial orientation (statistically isotropic) are the same as in dos x are estimated to be equal to reflectances of dense
packed beds of spherical particles with the same porosity material q according to Eq. (14). The reflectance
and specific surface. This result is similar to that well- strongly depends on wavelength: Fe surface is consider-
known from independent scattering theory [10]: ran- ably more reflective at 10.6 lm than at 1.06 lm. The
domly oriented irregular convex particles scatter like numerical values of reflectance used in the calculations
spheres. One can take into account preferential orienta- are taken from Ref. [4] and listed in Table 3.
tion of non-spherical particles through distribution of
specific surface over orientation s(X) in Eqs. (8)–(11).
All calculations are made for statistically isotropic Table 3
powder beds without regard to the shape of particles. Parameters of Fe–Cu and WC–Co powder mixtures accepted in
ray tracing simulation [4,8]
However, this model is validated by comparison with
ray tracing [4,8] only for spherical particles. Validation Parameter Fe Cu WC Co
of the statistical scattering model for non-spherical Porosity of the mixture 0.75 0.75
particles requires further ray tracing simulation or Weight fraction 0.7 0.3 0.91 0.09
experimental measurements with characterization of Theoretical density (g/cm3) 7.87 8.96 15.7 8.90
particlesÕ shape. Particle diameter (lm) 50 30 50 20
Fig. 4 shows typical examples of depth profiles of Reflectance at 1.06 lm 0.7 0.9 0.45 0.69
radiative energy flux density Q and deposited energy Reflectance at 10.6 lm 0.965 0.985 – –

Fig. 4. Penetration of laser radiation into an iron powder bed: (a)–(c), radiative energy flux; (d)–(f), deposited energy density; (a), (b),
(d), (e), thin powder layer on a completely reflecting substrate; (e), (f), deep powder bed. Dimensionless characteristics are defined in
Table 2. The following parameters are marked in the graphs: laser wavelength, albedo x equal to reflectance q of dense material, and
optical thickness k = bL. Full and broken lines represent the specular and diffuse reflection models, respectively.
A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434 3431

Net radiative energy flux in a thin powder layer on a Co) were accepted to consist of spherical monodispersed
completely reflecting substrate qk monotonously de- specularly reflecting particles. The parameters taken for
creases with depth (see Fig. 4(a) and (b)) and vanishes these simulations are listed in Table 3. Applying Eqs.
at the substrate where reflected radiation compensates (17) and (18) one can obtain the optical properties of
the incident one. The flux at a given depth increases with the mixtures listed in Table 4. These parameters are used
the layer thickness because the back-reflected radiation, to obtain radiation flux and deposited energy profiles
which reduces the net energy flux in the forward direc- according to the present RTE model with the isotropic
tion, is more pronounced for thinner layers. The depth phase function of specular reflection (5). Fig. 5 compares
of laser energy penetration into a deep powder bed in- RTE solutions (full lines) with ray tracing simulation
creases with reflectance of dense material as shown in (broken lines). The present calculations based on RTE
Fig. 4(c). In the case of CO2 laser where the reflectance agree in general with the ray tracing [4,8].
of iron is as high as 0.965, laser radiation is still signifi- The extinction length, 1/b, for the both powder mix-
cant at the dimensionless depth of 10. tures is less than 0.1 mm (see values of b in Table 4).
Energy density deposited into a deep powder bed However, radiation can penetrate into the powder beds
either monotonously decreases with depth at low reflec- (PB) substantially deeper, as shown in Fig. 5. For in-
tance (see 1.06 lm curves in Fig. 4(f)) or reaches a max- stance, radiation flux in the Fe–Cu mixture at 10.6 lm
imum in the bulk at high reflectance (see 10.6 lm curves is still considerable at the depth of 0.7 mm (see Fig.
in Fig. 4(f)). Deposited energy distribution in a thin 5(c)). Direct laser radiation attenuates proportionally
layer can be monotonous or have a local maximum to exp(bz) (see Eq. (31)) and is negligible at such
depending on the layer thickness (see Fig. 4(d)–(e)). distance. Only diffuse radiation is responsible for the
The layer thickness impact on the deposited energy is observed deep penetration that is due to multiple reflec-
opposite to the impact on the energy flux: the deposited tions of the incident radiation in the open pore system.
energy density increases with decreasing the thickness. The importance of the diffuse radiation increases with
This behaviour is explained by substrate influence: the albedo x, which can be considered as the average reflec-
substrate reflects radiation and makes it pass through tance of particles in the mixture (see Eqs. (14) and (18)).
the powder once more in the backward direction. The This tendency is clear from Fig. 5.
increase of the deposited energy density in thin powder The only significant difference between the RTE and
layers on reflective substrates is of particular importance the ray tracing results is in a thin layer adjacent to the
for laser treatment of powders. PB surface where RTE technique considerably overesti-
As follows from Fig. 4, the difference in both radia- mates the deposited energy relative to ray tracing (see
tion flux and deposited energy between the specular (full Fig. 5(d)–(f)). Note that the difference occurs within a
lines) and diffuse (broken lines) reflection models is sur- particle diameter from the surface. The origin of the dis-
prisingly low. Radiation flux at diffuse reflection is al- crepancy is that the structure of PB near the surface
ways less than that at specular reflection (see Fig. changes. In particular, the porosity, e, gradually grows
4(a)–(c)). This can be explained by preferential back- to unity within about one surface monolayer. This effect
ward scattering in the diffuse reflection model according is not taken into account by the present RTE model
to Eq. (6). This is why less radiation enters the powder at where a stepwise change of the porosity is implicitly as-
diffuse reflection. The dependence of deposited energy sumed at the surface of PB.
on reflection law is more complicated (see Fig. 4(d)– The effective absorptance of a powder layer on a
(f)) but one can conclude that, in average, the deposited completely reflecting substrate, Ake (defined in Table 2),
energy density is less at diffuse reflection. increases with layer thickness (see Fig. 6) and reaches
Laser radiation transfer in a deep powder bed has a plateau for thicker layers that corresponds to a deep
been already studied by ray tracing Monte Carlo simu- powder bed. The higher is reflectance of dense material,
lation [4,8]. Iron–copper (Fe–Cu) [4] and tungsten car- the slower effective absorptance approaches the plateau.
bide–cobalt (WC–Co) [8] powder mixtures have been The effective absorptance value at the plateau decreases
studied. All the powder components (Fe, Cu, WC, and with the dense material reflectance. Absorptance for the
specular reflection model (full lines in Fig. 6) is slightly
higher than for the diffuse reflection model (broken
Table 4 lines). The difference in absorptance corresponds to the
Estimated optical parameters of the powder mixtures difference in radiation flux shown in Fig. 4(a)–(c). The
Parameter WC–Co Fe–Cu Fe–Cu absorptance at diffuse reflection is lower because of pref-
erential backward scattering according to Eq. (6). Sum-
1.06 lm 1.06 lm 10.6 lm
marising the behaviour of thin powder layers on
Extinction coefficient, reflecting substrates, one can say that the total laser en-
b (mm1) 12.23 11.82 11.82 ergy absorbed in the layer increases with its thickness
Albedo, x 0.52 0.78 0.973
but the energy density decreases.
3432 A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

Fig. 5. Comparison of the present calculations based on RTE (full lines) with ray tracing simulation (broken lines) for the WC–Co [8]
((a), (d)) and Fe–Cu [4] ((b)–(c), ((e)–(f)) powder mixtures at the wavelengths of 1.06 lm and 10.6 lm: (a)–(c), energy flux; (d)–(f),
deposited energy. Dimensionless values q and u are defined in Table 2.

Fig. 6. Effective absorptance of a powder layer on a completely


Fig. 7. Effective absorptance Ae of a deep powder bed versus
reflecting substrate versus layer thickness: full lines, isotropic
absorptance A of dense material: full line, isotropic specular
specular reflection model (5); broken lines, diffuse reflection
reflection model (5); broken line, diffuse reflection model (6);
model (6). Powder material, laser wavelength, and reflectance of
points, experimental data [9,18–22]. Powder material and laser
dense material q taken for the calculations are marked near the
wavelength are marked near the experimental points.
corresponding curves.

Fig. 7 shows the effective absorptance of a deep pow- and {A = 1; Ae = 1} and are monotonous in the interval
der bed, Ae versus the absorptance of dense material, 0 < A < 1. Effective absorptance of powder Ae can be
A = 1  q, for the specular (full line) and diffuse (broken considerably higher than the absorptance of the corre-
line) reflection laws calculated according to Eqs. (47)– sponding dense material A in the interval 0 < A < 1. This
(49). There is no much difference between the two reflec- behaviour seems to be reasonable because only a part of
tion laws. Both curves pass the points {A = 0; Ae = 0} radiation is reflected by particles at the surface of PB.
A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434 3433

The rest enters the pores and is absorbed inside PB. The particles. For example, albedo is equal to hemispherical
ray tracing method [4] gives qualitatively the same func- reflectance when there is no preferential orientation of
tions of Ae versus A. particles. Dependent scattering can be neglected when
The calculated curves are compared with experimen- estimating the albedo and the phase function of metallic
tal measurements for metallic powders (points in Fig. 7). powders, even at high densities.
The powder effective absorptances are taken from [9] The typical configuration of laser treatment of pow-
and the corresponding dense material absorptances are ders is normal incidence of collimated radiation on a thin
taken from [18–22]. Note that different methods of mea- powder layer deposited on a reflective substrate. The
surement and sample preparation give rather different radiation transfer equation for this configuration is ana-
values of absorptance of dense material [18–22]. The lytically solved by the two-flux method. The models of
points generally correspond to direct reflectance mea- isotropic specular and diffuse reflection are applied. En-
surements of a smooth surface of a polycrystalline sam- ergy deposition profiles and integral absorptances are
ple [18]. Horizontal bars indicate uncertainty in the calculated for the considered system. The results ob-
absorptance of dense material. The experimental data tained in the limit of deep powder bed essentially agree
demonstrate a strong correlation between the absorp- with ray tracing simulation [4,8]. Due to multiple reflec-
tance of powder and that of dense material that approx- tions in an open pore system, laser radiation can pene-
imately agrees with the calculated curves. The trate into the powder to considerable depths, much
experimental error is about the difference between the greater than the characteristic particle diameter. Scat-
specular (full line) and diffuse (broken line) scattering tered radiation component formed by multiple reflec-
models. Therefore, one can not conclude which of the tions in metallic powders becomes more intensive than
two models describes the experiments better. More care- collimated component originated from the incident laser
ful characterization of powder structure and the state of radiation. Summarising the behaviour of thin powder
the particlesÕ surface is necessary to study the influence layers on reflecting substrates, one can say that the total
of these factors on the optical properties. laser energy absorbed in the layer increases with its thick-
In this work only perfectly reflecting substrates are ness but the deposited energy density decreases. Gener-
studied. Ray tracing simulation [4,5] and two-flux ana- ally, absorptance and energy density for the specular
lytical solutions [5] have been reported for layers with reflection model are slightly greater than these values
transmission boundary conditions at the back surface for the diffuse reflection model. This is because backward
that are equivalent to completely absorbing substrate. scattering is more intensive at diffuse reflection.
One can qualitatively estimate the influence of non-per- The obtained theoretical results are in good agree-
fectly reflecting substrate by comparison of these limit- ment with experimentally observed correlation between
ing cases. Absorption by the substrate should reduce effective absorptance of metallic powders [9] and absorp-
radiation flux propagating in the back direction. This tance of corresponding dense metals.
increases absorptance of the system powder layer-
substrate but decreases energy deposited in the powder.
References

[1] S. Barnes, N. Timms, B. Bryden, I. Pashby, High power


6. Conclusion diode laser cladding, J. Mater. Process. Technol. 138 (2003)
411–416.
[2] J.-P. Kruth, P. Peeters, T. Smolderen, J. Bonse, T. Laoui,
Laser radiation transfer in metallic powders can be
L. Froyen, Comparison between CO2 and Nd:YAG lasers
calculated on the basis of the radiation transfer equation for use with selective laser sintering of steel-copper
where a powder bed is characterised by the following powders, Int. J. CAD/CAM Comput. Graphics 13 (1998)
effective parameters: extinction coefficient b, albedo x, 95–110.
and scattering phase function P. In the geometrical op- [3] C. Over, W. Meiners, K. Wissenbach, J. Hutfless, M.
tics approximation, the extinction coefficient is deter- Lindemann, Rapid manufacturing of metal parts and tools
mined by the structure of powder, i.e. by size and using laser melting, in: R. Poprawe, A. Otto (Eds.), Lasers
shape of particles and by their arrangement, but is inde- in Manufacturing 2003, AT-Verlag, Stuttgart, 2003, pp.
pendent of optical properties of the material of particles 265–270.
[4] X.C. Wang, J.-P. Kruth, A simulation model for direct
like reflectance. In case of opaque particles, one can esti-
selective laser sintering of metal powders, in: B.H.V.
mate extinction coefficient as a quarter of powder sur-
Topping (Ed.), Computational Techniques for Materials,
face per unit pore volume. This takes into account Composites and Composite Structures, Civil-Comp, Edin-
dependent scattering that increases the extinction coeffi- burg, 2000, pp. 57–71.
cient in inverse proportion to porosity. On the contrary, [5] B.P. Singh, M. Kaviany, Modelling radiative heat transfer
the scattering characteristics as albedo and phase func- in packed beds, Int. J. Heat Mass Transfer 35 (1992) 1397–
tion depend on reflective properties of the material of 1405.
3434 A.V. Gusarov, J.-P. Kruth / International Journal of Heat and Mass Transfer 48 (2005) 3423–3434

[6] D. Baillis, J.F. Sacadura, Thermal radiation properties of [14] P.D. Jones, D.G. McLeod, D.E. Dorai-Raj, Correlation of
dispersed media: theoretical prediction and experimental measured and computed radiation intensity exiting a
characterisation, J. Quant. Spectrosc. Radiat. Transfer 67 packed bed, J. Heat Transfer 118 (1996) 94–102.
(2000) 327–363. [15] K. Kamiuto, Correlated radiative transfer in packed-
[7] R. Siegel, J.R. Howell, Thermal Radiation Heat Transfer, sphere systems, J. Quant. Spectrosc. Radiat. Transfer 43
Hemisphere, New York, 1981. (1990) 39–43.
[8] X.C. Wang, T. Laoui, J. Bonse, J.P. Kruth, B. Lauwers, [16] M. Tancrez, J. Taine, Direct identification of absorption
L. Froyen, Direct selective laser sintering of hard metal and scattering coefficients and phase function of a porous
powders: experimental study and simulation, Int. J. Adv. medium by a Monte Carlo technique, Int. J. Heat Mass
Manuf. Technol. 19 (2002) 351–357. Transfer 47 (2004) 373–383.
[9] N.K. Tolochko, T. Laoui, Yu.V. Khlopkpv, S.E. Moz- [17] S. Torquato, B. Lu, Chord-length distribution function for
zharov, V. Titov, M.B. Ignatiev, Absorptance of powder two-phase random media, Phys. Rev. E 47 (1993) 2950–
materials suitable for laser sintering, Rapid Prototyping J. 2953.
6 (2000) 155–160. [18] D.E. Gray (Ed.), American Institute of Physics Handbook,
[10] H.C. Van de Hulst, Light Scattering by Small Particles, McGraw-Hill, New York, 1972.
Dover, New York, 1981. [19] E.D. Palik (Ed.), Handbook of Optical Constants of
[11] B.L. Drolen, C.L. Tien, Independent and dependent Solids, Academic Press, New York, 1985.
scattering in packed-sphere system, J. Thermophys. 1 [20] E.D. Palik (Ed.), Handbook of Optical Constants of Solids
(1987) 63–68. II, Acad. Press, New York, 1991.
[12] J.C. Chen, S.W. Churchill, Radiant heat transfer in packed [21] M. Bass (Ed.), Handbook of Optics, vol. 2, McGraw-Hill,
beds, AIChE J. 9 (1963) 35–41. New York, 1995.
[13] K. Kamiuto, M. Iwamoto, M. Sato, T. Nishimura, [22] D.R. Lide (Ed.), CRC Handbook of Chemistry and
Radiation-extinction coefficients of packed-sphere systems, Physics, CRC Press, New York, 2000.
J. Quant. Spectrosc. Radiat. Transfer 45 (1991) 93–96.

You might also like