International Journal of Thermal Sciences 167 (2021) 107011
Contents lists available at ScienceDirect
International Journal of Thermal Sciences
journal homepage: www.elsevier.com/locate/ijts
Balling phenomenon in metallic laser based 3D printing process
M’hamed Boutaous a, Xin Liu a, Dennis A. Siginer b, *, Shihe Xin a
a
b
Université de Lyon, CNRS, INSA-Lyon, CETHIL, UMR5008, F-69621, Villeurbanne, France
Centro de Investigación en Creatividad y Educación Superior & Departamento de Ingeniería Mecánica, Universidad de Santiago de Chile, Santiago, Chile
A R T I C L E I N F O
A B S T R A C T
Keywords:
Additive manufacturing
Selective laser melting
3D discrete element method
Balling phenomenon
A comprehensive model is developed coupling major physical phenomena inherent to the Selective Laser Melting
process together with a 3D numerical model based on the discrete element method to study the effect of the
process parameters on the generation of balling droplets in the laser melting process. Effects of several process
parameters as well as material properties on the balling phenomenon are investigated. Simulation results
compare well to the experimental findings in the literature.
1. Introduction
Selective laser melting (SLM) is the most innovative and promising
non-contact additive manufacturing process developed and in wide
spread use today with many applications and advantages easy control
and flexibility among them, Loh et al. [1], Hu et al. [2], Ho et al. [3] and
Schwab et al. [4]. Several components of the physical phenomenon
underlying the SLM process are still not well understood including the
balling phenomenon an extremely undesirable occurrence in the
manufacturing process. Particles in the metallic powder bed in the SLM
may either not be sintering at all or may be coalescing into rather large
droplets triggering the balling phenomenon if the process and laser
parameters are not correctly chosen resulting in a powder bed which is
neither sintered at all or completely melted resulting ultimately into
particles joined into rather large drops, Bourel et al. [5], Tolochko et al.
[6] and Li et al. [7]. Usually the drops quickly spread out and their sizes
may exceed the diameter of the laser spot, which triggers instability and
distortion of the molten track, Fig. 1. As the physics underlying the SLM
is not well understood the process cannot be thoroughly modeled. Thus,
it is not possible at this time to devise accurate numerical tools useful in
improving the quality of the parts made through the SLM.
Increasing applications of the laser melting/sintering technologies
over the years brought the balling phenomenon more in focus and
studies were undertaken to clarify the underlying physics. To our
knowledge the balling phenomenon has only been studied experimentally. Niu and Chang [8,9] are among the pioneers to study the balling
phenomenon in the laser melting process. They investigated laser sintering of steel powders by scanning a single track on the powder bed at
various laser powers and scanning speeds to establish an understanding
of the underlying principles of the laser sintering of these powders and to
determine a window of optimum processing conditions to avoid the
balling phenomenon. Tolochko et al. [10,11] investigated the occurrence of the balling phenomenon under different processing conditions,
especially analyzing the relationship between laser parameters and
formation of “melted cakes”. Das [12] identified some of the important
physical mechanisms in the direct SLM process of metals and provided
an insight into the phenomena observed during the direct SLM of a variety of metallic materials. In his work, he concluded that balling occurs
when the laser melted powder layer does not wet the underlying substrate. Two kinds of balling phenomena during direct SLM of 316L
stainless steel powders were investigated by Gu and Shen [13] and Gu
[14]. They found that using a low laser power gives rise to the first kind
of balling characterized by highly coarsened balls possessing an interrupted dendritic structure in the surface layer. On the other hand, a high
scan speed caused the second kind of balling featured by a large amount
of micrometer-scale (10μm) balls on the laser sintered surface. Recently,
Zhou et al. [15] studied the balling phenomenon in the SLM of pure
tungsten powder bed. The surface morphology of the SLM manufactured
part was investigated by SEM images. They concluded that the dominant
solidification yields balling of large melt droplets causing surface
roughness and instability due to the high surface tension and viscosity of
tungsten. Few works focused on the numerical simulation of the balling
phenomenon compared to experimental studies. This is because the SLM
is a very complex process coupling many physical phenomena, which is
hard to be modeled numerically, Yadroitsev et al. [16]. However, a
numerical model of the SLM is desired and necessary for industrial
manufacturers to optimize the process, reduce the cost and increase the
product quality. This is the original motivation of this work.
The aim of this work is to develop an accurate model to simulate the
* Corresponding author.
E-mail address: dennis.siginer@usach.cl (D.A. Siginer).
https://doi.org/10.1016/j.ijthermalsci.2021.107011
Received 21 August 2016; Received in revised form 28 March 2021; Accepted 8 April 2021
1290-0729/© 2021 Elsevier Masson SAS. All rights reserved.
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
Nomenclature
I(s)
Iλ0
Re
P
RL
x, y
z
w
Qc;i,j
Qe;i,j
Qi,j
Hc
kc
ki
Ac
a
d
Ri
r0
He
H′e
Cp
Csp,i
Clp,i
Hp,i
T
Tm
Ta
Hm
h
ke
Pl
Ps
ΔPp
Radiation energy intensity
Initial radiation energy intensity
Reflective coefficient on the surface
Power of the laser beam
Radius of the laser beam
Coordinates of the center of the laser beam
Depth
Correction coefficient in the Beer-Lambert law
Conduction through the contact area between two particles
i and j
Conduction through air between two particles i and j
Total heat flow between two particles i and j
Heat transfer coefficient
Harmonic mean of the conductivity of the particles i and j
Conductivity of the ith particle
Contact area between particles
/ radius of the contact area between two particles i and j/
radius of the grains
Thickness of the contact region between two particles i and
j
Radius of the ith particle
Initial radius of the sphere
Thermal conductance for two particles in contact
Heat flow from particle i to particle j when particle i is in
contact with N particles
Specific heat [J/(kg⋅K)]
a(ΔPp )
ȧ
Specific heat of the solid phase
Specific heat of the liquid phase
Volumetric enthalpy
Temperature
Melting temperature
Ambient temperature
Latent heat of melting
Convective heat transfer coefficient [w/m2K]
Effective thermal conductivity
Liquid pressure
Sintering atmospheric pressure
Difference in gas pressure between the pore and the
atmosphere
Critical grain size under ΔPp ∕
=0
Rate of change of the grain radius in time
Greek Symbols
λ
Wave length
α
Extinction coefficient
Particle density [kg.m−3 ]
ρp
σ
Stefan-Boltzmann constant
εR
Surface emissivity of the material
Γ
Surface tension
η
Dynamic viscosity
θ
Coalescence angle
Radius of the liquid meniscus at the pore
ρs
Radius of the liquid meniscus at the compact surface
ρp
Strack [17], to study the effect of the process and material parameters on
the generation of balling droplets in the laser melting process. The heat
source with Gauss distribution is assumed to be volumetric and attenuates following the Beer-Lambert law. A heat transfer sub-model capable
of describing thermal conduction in a discrete system is developed. The
model makes it possible to estimate the effect of air between grains on
the laser melting process. The effect of scanning speed on the generation
of balling droplets is studied at a constant input heat flux. To highlight
the importance of introducing the discrete model to describe the thermal
behavior of granular materials and to validate our model, simulation
results are compared with experimental works of other researchers
taken from the literature. Finally, the effect of several laser parameters
on the balling phenomenon is also analyzed.
2. Theoretical model of the SLM
Physical phenomena that occur during the SLM process are multiple
and complex. They can be divided into three groups: radiation transfer,
heat conduction and melting behavior, Dong et al. [18], Defauchy [19],
Jodhpur [20] and Peyré et al. [21]. Fig. 2 shows the operational process
of the integrated model: the radiation transfer sub-model uses laser
beam parameters, such as the laser power, scanning speed and radius of
the laser beam. It also makes use of the optical properties of the material
and the geometry of the grains to calculate the energy absorbed by the
powder bed. The heat flux from radiation transfer sub-model is regarded
as the heat source in the heat conduction sub-model. The temperature
field of the powder bed is calculated by the discrete heat conduction
sub-model. The model has the capability of describing the thermal history of each grain in the powder bed. Sintering sub-model uses temperature data and calculates values of the rheological properties
corresponding to the temperature, and then estimates the change in the
geometry of the powder bed caused by the coalescence and densification. The single layer simulation consists of these three sub-models as
described and shown in Fig. 2.
Fig. 1. Instability of the laser sintered tracks of 316L powder on steel substrate [12].
balling phenomenon in metallic SLM process, to better understand the
multiple physical phenomena occurring in the material and to study the
influence of each parameter on the quality of the manufactured parts.
This will extend the areas of application of this technology, which
promises to be very innovative, especially if we succeed in making
structurally improved parts of higher mechanical and thermal characteristics. A comprehensive model is developed coupling multiple phenomena of the full SLM process including radiation transfer, thermal
conduction, phase change, coalescence and air exhaust together with a
3D numerical model based on the discrete element method, Cundall and
2
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
Fig. 2. Operational process of the integrated model for the SLM process.
Fig. 4. Conduction between two elastic spheres in contact.
Fig. 3. Illustration of the luminance within the homogeneous absorbing scattering medium.
3. Radiation and heat transfer in metallic powder bed
I(s) = Iλ0 e−αs
(2)
Idepth (z) = Isurface e−αz
(4)
Therefore, the intensity of the laser beam conforms nearly to Gauss
distribution on the surface of the powder bed and attenuates following
the Beer-Lambert law in depth, Defauchy [19],
]
[
2(x2 +y2 )
−
2P
RL
w2
Isurface = (1 − Re )Iλ0 e
, Iλ0 = 2 w =
(3)
2.146
πw
3.1. Radiation transfer in metallic powder bed
A powerful fiber laser beam either of the CO2 or the Fiber laser type is
used as the input energy source in the SLM process. The CO2 type is
generally used with polymeric materials, and the Fiber laser and
sometimes Chrysler laser (Nd:YAG) are more suitable in metallic applications, Yu et al. [22], Martinez et al. [23] and Sing et al. [24].
Although metallic grains are opaque, the metallic powder bed is not
opaque due to porosity. A powder bed with opaque particles can be
approximated by an equivalent homogeneous absorbing scattering medium, Gusarov and Smurov [25]. During a transient radiative heat
transfer in a two-dimensional slab, the distribution of the luminance
within the homogeneous absorbing scattering medium involves three
mechanisms: emission, absorption and diffusion, Fig. 3.
If emission is negligibly small, the radiation transfer equation can be
simplified into the Beer-Lambert Law, Gusarov and Smurov [25],
describing laser metal interaction:
∫
I(s) = Iλ0 e−αλ s dλ
(1)
where Re is the reflective coefficient on the surface, r is the radial distance from the center of the laser beam, P is the power of the laser beam,
RL is the radius of the laser beam, α is the extinction coefficient, z the
depth location and w the correction coefficient in the Beer-Lambert law.
3.2. Thermal conduction in granular systems and the DEM
Unlike heat transfer in homogenous media, thermal conduction in
granular media like powder beds is defined as transfer of energy by
diffusion between objects in physical contact, Watson et al. [26]. The
traditional continuum conduction model is not suitable for the laser
melting process. A discrete model is needed to describe the thermal
behavior of the powder bed. Heat flow by conduction between two
particles i and j in a granular system is illustrated in Fig. 4. It consists of
two components: conduction through the contact area Qc and conduction through air Qe . Total heat conduction between two particles is then
given by:
λ
where αλ is the attenuation coefficient in the wave length λ, I(s) is the
radiation energy intensity at the given depth s and Iλ0 is the initial radiation energy intensity. For the laser heat source, the radiation wave
length is usually unique. Then equation (1) can be written as:
3
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
Table 1
Thermo-physical properties and other parameters used in the simulation, Jodhpur [20].
Qi,j = Qc;i,j + Qe;i,j
Qc could be estimated based on the Fourier equation:
Qc;i,j = − kc Ac
dTi,j
dx
(5)
2k k
i j
where kc = ki +k
is the harmonic mean of the conductivity of the particles
j
[J/(K⋅s⋅m)] and Ac = πa2 is the contact area between particles [m2 ].
Because the size of the particles i and j is very small, usually less than
100μm, the thickness of the contact region d is assumed to be the same as
the thickness of the thermal conduction distance, so the thermal
gradient between two particles could be approximated by:
dT ΔTi,j
≈
dx
d
(6)
kc Ac
ΔTi,j = −Hc ΔTi,j ,
d
Hc =
kc Ac π kc a2
=
d
d
(7)
where ΔTi,j = Ti − Tj [K] and Hc is the heat transfer coefficient [J/(K⋅
√̅̅̅̅̅̅
s)]. The radius of the contact area is defined as a = Rd based on the
Hertz contact theory with d representing the thickness of the contact
region, R = 1 +1 1 is the effective radius and R1 and R2 are the radii of the
R1
R2
Ri Rj
Ri + Rj
(8)
where Ri and Rj are the radii of particles i and j, respectively. Heat
conduction from particle i to j via air Qe;i,j is given by,
⎡ [
⎤
( )2 ]
2π 1 − 12 ar
(r − a) ⎥
⎢
⎥
Qe;i,j = −He ΔTi,j , He = kair ⎢
(9)
⎣
⎦
1 − π4
where kair is the conductivity of air and He is the thermal conductance for
two particles in contact. In a multiple contact model, if the particle i is in
contact with N particles, considering all the contact areas, heat flow
between the particle i and neighboring j is given by:
)⎤
( )2 ](
⎡ [
π 2 − N2 ar
r−a
⎢
⎥
H ′e = kair ⎣
(10)
⎦
1 − π4
Tit+Δt = Tit +
∑(
)
dTp,i
Qc;i,j + Qe;i,j + Qp,i source
=−
dt
Qti
Δt,
mi Cpi
Qti = −
)
∑(
Qtc;i,j + Qte;i,j + Qtp,isource
(15)
3.3. Thermal boundary conditions
Energy lost by radiation and convection at the upper surface of the
powder bed is given by (16)1. The bottom of the powder bed is insulated
as represented by (16)2:
⃒
(
)
)
(
∂T
∂T ⃒
−ke |surf = h Ta − T|surf + εR σ Ta4 − T|4surf , − ke ⃒⃒
=0
(16)
∂z
∂z bottom
(11)
where Ta is the ambient temperature, h is the convective heat transfer
coefficient, assumed as constant h = 10 w/m2K, εR is the surface emissivity of the material, σ is the Stefan-Boltzmann constant and ke is the
effective thermal conductivity.
Thus, using equations (5) and (11), the conductive heat transfer
equation becomes:
ρp C p
(14)
where the superscript t means at time t
with the assumption that the temperature within each particle is uniform, thermal conduction between particles in contact is calculated as:
Qe;i,j = − H ′e ΔTi,j
1673.0 K
285.0 kJ. kg−1
500.0 J.kg-1.K −1
7800.0 kg.m−3
40.0 W.m−1. K −1
0.2
mono-dispersed
30 μm
140 μm
where the volumetric enthalpy Hp,i is related to temperature T by the
thermal equation of phase state. Csp,i and Clp,i are the specific heats of the
solid and liquid phases, respectively, and ρp is the particle density. Tm is
the melting temperature and Hm is the latent heat of melting, Qp,isource is
given by the Beer-Lambert law (4).
DEM allows following the evolution in time of each particle in a
batch of particles initially randomly distributed. The number of neighboring particles to each particle is known at each time step, and the
associated temperature is calculated assuming it is uniformly distributed
within each particle. Several integration schemes have been developed
for DEM. We use the classic Euler explicit integration scheme. At each
time step, the force, displacement and heat flux are calculated by submodels introduced in previous sections. Detailed description of the
sub-models is given in Xin et al. [27]. The update scheme applied for the
temperature is:
spheres in contact, respectively. The heat transfer coefficient can then be
modified as:
Hc = π kc
Value
Melting point (Tm)
Enthalpy of fusion (Hm)
Specific heat (Cp)
Density (ρ)
Thermal conductivity (kc)
Emissivity (ε)
Size distribution
Grain size
Laser beam size
⎧ H
p,i
⎪
: Hp,i ≤ ρp Csp,i *Tm
⎪
⎪
⎪ ρp Csp,i
⎪
⎪
⎨
Tpi = Tm : ρp Csp,i *Tm < Hp,i < ρp Csp,i *Tm + Hm
⎪
⎪
⎪
⎪
Hp,i − Csp,i *Tm − Hm
⎪
⎪
: Hp,i ≥ ρp Csp,i *Tm + Hm
⎩ Tm +
ρp Clp,i
Then (5) can be modified as:
Qc;i,j = −
Property
(12)
where Cp is the specific thermal capacity [J/(kg⋅K)] and ρp is the particle
3.4. Validation of thermal conduction approach in granular systems
∑(
)
dHp,i
=−
Qc;i,j + Qe;i,j + Qp,i source
dt
In this section, the discrete thermal conduction model developed in
this paper is validated by comparing the results with the published data
from Jodhpur [20], who developed a similar approach, a 3D discrete
thermal conduction model, and numerically simulated the temperature
distribution in the powder bed using a Discrete Element Method. The
simulations for validation are done under the same initial and boundary
conditions used by Jodhpur [20]. The powder bed is made of
stainless-steel grains. The properties of powder material are listed in
Table .1.
density [kg.m−3 ]. To account for the phase change phenomena during
the SLM process, the conductive heat transfer equation (12) is modified
as:
(13)
4
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
4.1. Viscous coalescence model
Metallic melt flow can be regarded as a viscous liquid. Therefore, the
viscous coalescence model is valid in modeling of the SLM process. The
coalescence described as the overlapping of spherical particles and neck
growth between contacting particles is an important aspect of sintering,
Fig. 7. Particles are assumed to be single crystals and they remain
spherical throughout the coalescence process.
The sintering rate equation is derived by Frenkel from the energy
balance equation assuming the surface energy reduction rate equal to
the viscous dissipation rate during the deformation, Smoluchowski [29].
Pokluda et al. [30] derived a first order equation for the sintered angle
through a modification of the Frenkel’s model:
5
dθ
Γ 2−3 cosθ sinθ (2 − cosθ)1/3
=
dt ηr0
(1 − cosθ)(1 + cosθ)1/3
,
θ(0) = 0
(17)
where Γ, η and r0 represent the surface tension, the dynamic viscosity
and the initial radius of the grains (spheres). The evolution of the sintering neck radius in time is given by
Fig. 5. Temperature distribution in the powder bed for increasing scan speed at
P = 50 W: (a) (c) Results from Jodhpur [20]; (b) (d) Results from our
simulations.
The top view of temperature distribution in the powder bed with
different scanning velocities is illustrated in Fig. 5. The temperature
history of a given particle in the laser scan path is shown in Fig. 6. Figs. 5
and 6 show that simulations are in good agreement with those from
published literature, Jodhpur [20]. In Fig. 5, the scan velocity of 0.6 m/s
with a power of 50 W is enough to melt the particles in the laser scan
path to form a continuous melted track. The peak temperature in powder
bed is about 2000 K, which is significantly above the melting point.
Increasing the scan velocity up to 1.0 m/s and keeping laser power
constant decreases interaction time and few particles in the laser scan
path seem to undergo melting. In this case, the peak temperature is
barely beyond the melting point. Thus, energy absorbed is not enough to
generate a continuous molten track. Temperature history of a particle in
the scan path with different powers is shown in Fig. 6. Clearly increasing
power raises the peak temperature of particles, which is opposite of the
observed trends in Fig. 5 where scan velocity is increased. This suggests
that the ratio between the power and scan speed can play an important
role in characterizing thermal behavior of heated particles in powder
beds.
Fig. 7. The neck geometry applied to the sintering of two spheres.
4. Sintering model
Fig. 8. Illustration of the liquid menisci at the specimen surface and around a
pore containing a gas of pressure ΔPp during liquid phase sintering, Cho
et al. [31].
The sintering process mainly consists of two phenomena: coalescence and densification, Suk-Joong [28].
Fig. 6. Temperature history of a particle in the laser scan path: (a) Results from Jodhpur [19]; (b) Results from our simulations.
5
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
h
= sinθ
a
(18)
Thus, yielding an equation to trace out the evolution of the grain
equivalent radius, a, in time as the grains are coalescing.
4.2. Entrapped gases and pore filling theory
Solid and liquid particles coexist at the interface between the molten
and solid regions. In this region, molten fluid flows into interstices between solid grains pushed by gravity. Substantial volume of gas remains
entrapped in this region. The filling of the pores roughly corresponds to
the escaping air because of the difference between external and internal
pressure of entrapped gases. When the gas pressure in an isolated pore is
different from that outside the compact area, the pore filling is either
delayed with an excess internal pressure or accelerated with an excess
external pressure. Fig. 8 shows schematically the compact surface and
the internal surface of a pore containing insoluble gases, Cho et al. [31],
Park et al. [32]. Densification is mainly caused by escaping entrapped
gases in the powder bed. Thus, the pore filling theory is suitable in
modeling densification in metallic SLM process.
Since hydrostatic pressure is maintained in the liquid, the bulk liquid
pressure at the surface is the same as that at the pore surface, Suk-Joong
[28],
2Γl
Pl = Ps −
,
ρs
Pl = ΔPp + Ps −
2Γl
Fig. 9. Original powder bed before laser treatment.
(19)
ρp
where Pl is the liquid pressure, Ps is the sintering atmospheric pressure,
ΔPp is the difference in gas pressure between the pore and the atmosphere, ρs and ρp are the radii of the liquid menisci at the pore and at the
compact surface, respectively. ρs is linearly proportional to the grain
radius a
(20)
ρs (t)∝a(t)
= 0; the critical condition for the wetting of a
ρp is equal to rp when ΔPp ∕
pore surface leading to a spherical pore is reached, Park et al. [32].
Based on this hypothesis of spherical pores, let a(ΔPp ) and a(0) be the
critical grain sizes under ΔPp ∕
= 0 and ΔPp = 0, respectively. Then
following Cho et al. [31] and Park et al. [32],
)
)
(
)
(
)
(
(
a ΔPp
ρ ΔPp
ρ ΔPp
a ΔPp
1
= s
=
,
(21)
= s
r
a(0)
a(0)
rp
1 − 2Γp l ΔPp
ρs (0)
Fig. 10. Heat diffusion and sintering process in a powder bed with the laser
power and the scanning speed set at 50W and 0.25 m/s, respectively. (a) Time
= 0.0001s; (b) Time = 0.001s; (c) Time = 0.002s.
each grain, their surface energy Γ and the evolution of the viscosity η
versus temperature via an Arrhenius law. Runge-Kutta method is used to
solve Eq. (17) coupled with Eqs. (19) and (22) to update the grain radius
at each time step. The evolution of the neck radius of the joined grains
during coalescence gives the evolution of the shape of the metal bulls
formed on the surface. The temperature decrease in time after the laser
sweep and the growth of the viscosity as the temperature decreases at
the limits of the melted zone fix the size of the metallic bulls.
In this study, to start the calculation, the initial radius a(ΔPp ) and rp
are taken constant for all the powder bed and evaluated, respectively,
from the initial mean porosity of the powder bed, and the initial radius
of the solid grains, Fig. 8. Considering that rp remains constant, as the
cooling process following the melting and coalescence phenomena is
very fast in the case of the SLM process, and as it is established that
porosity remains during powder sintering, then the rate of change of the
grain radius in time ȧ can be written as:
ȧ = a(0) [
rp
2Γl
ΔṖp
r
1 − 2Γp l ΔPp
]2 = a(0)
1
r
rp
2Γl
ΔṖp
r
1 − 2Γpl ΔPp 1 − 2Γp l ΔPp
)
(
= a ΔPp
rp
2Γl
5. Results and discussion
To validate the numerical model, simulation results are compared to
the experimental data from published literature, Johnson and Rahaman
[33]. 13,909 particles of diameter 130 μm are deposited in a box with
dimensions of 8mm × 4mm × 4mm, Fig. 9. For solid particles, the
density is 8470 kg/m3 , specific heat is 444 J/kg⋅K and conductivity is
14.9 W/m⋅K. For molten particles, the density is 7880 kg/m3 , specific
heat is 611 J/kg⋅K and conductivity is 27. 5 W/m⋅K. The melting point of
Ni-alloy is 1475 K. The latent heat of melting is 2. 516 GJ/m3 . The
surface tension and viscosity of Ni-alloy at the melting point are 1850
mN/m and 5 mPa⋅s, respectively. The power of input heat flux is 50 W
and the radius of laser beam is 0.65 mm. In the SLM process, the laser
beam scans the powder bed in the longitudinal direction with various
scanning speeds. The entire numerical simulation of the laser sintering
process in Ni-alloy powder bed is calculated using self-developed
ΔṖp
r
1 − 2Γp l ΔPp
(22)
Equation (22) gives a relationship between the pressure evolution in
the pore zone and the radius of the grains. Its solution needs the
calculation of the evolution of the grain radius, given by the sintering
theory, equations (17) and (18). Hence, Eq. (22) coupled with the coalescence Eqs. (17) and (18), which determine the evolution of the radius
of the metal bulls, form a complete set to define the kinetics of the
balling phenomenon on the surface of the grain bed. The initial grain
radius is fixed for each test, and the formation of metallic bulls due to
coalescence and pore evolution are governed by the thermal state of
6
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
Fig. 11. Effect of scanning speed on the balling phenomenon: (a) Conclusion from Guo & Shen [13]; (b) Simulation results with the power set at 50W and various
scanning speeds: 0.15 m/s, 0.25 m/s, 0.65 m/s; (c) Experimental results from Tolochko et al. [6] and Yadroitsev and Smurov [16].
Fig. 12. Numerical estimation of the effect of scanning speed on the width of the molten track.
in-house code based on Fortran 90 software. Details on the Discrete
Element method and all the couplings are already presented in our
previous paper, Xin et al. [27].
Fig. 10 illustrates the numerical results of heat diffusion and the
sintering process in a powder bed with laser treatment. The top view of
the molten tracks and the calculated temperature distribution on the top
of the powder surface are presented. There is no molten particle at the
beginning of the sintering process because of insufficient input heat flux,
Fig. 10 (a). Particles begin melting with increasing input laser energy,
Fig. 10 (b), and molten particles combine with each other and grow into
larger grains through coalescence, Fig. 10 (c). The temperature of the
molten region decreases because of convection cooling at the surface
after laser spot has passed.
The effect of the scanning speed on the balling phenomenon is presented in Fig. 11. The width of the molten track is in a highly unstable
state when the scanning speed is low because the amount of molten
particles generated is excessive, Figs. 11 (b-1). The surface energy of
molten particles will keep decreasing to get a final equilibrium state as
explained by Gu and Shen [13], and as represented in Fig. 11 (a). The
absorbed heat flux decreases leading to the shrinkage of the molten
region with increasing scanning speed. If the scanning speed is too high,
the molten region becomes unstable resulting in the break-up of the
molten track as illustrated in Figs. 11 (b-3). Fig. 11 provides the evidence that the simulation methodology presented in this paper is quite
good as confirmed by comparison with experimental findings in the
literature [6–13]. We conclude that the discrete numerical model of the
7
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
Fig. 13. Numerical estimation of the effect of the laser power on the width of the molten track.
metallic laser sintering developed in this paper is useful in predicting the
suitable input power and scanning speed for the laser melting process to
avoid the triggering of strong balling.
The calculated width of molten track is presented in Fig. 12 to
quantitatively analyze the effect of the scanning speed on the balling.
The lines represent the limits of the melted zones. Note that the variations are because in DEM we consider each grain at homogeneous
temperature, and the variation around the diameter of the grains appears in the estimation of the melting temperature location. The average
width of molten track is largest when the scanning speed is too low
(0.15 m/s), and the shape of the molten track is distorted. The average
width of the molten track is smallest, and break-ups of the molten track
may occur (width is zero) when the scanning speed is too high (0.65 m/
s). The laser power is also a very important parameter in the SLM process. We simulated the SLM process with constant scanning speed (0.25
m/s) and different laser powers (20 W, 50W and 80 W) to study the effect
of the laser power on the balling phenomenon.
The coalescence of molten grains is directly controlled by surface
tension and viscosity according to Frenkel (Eq. (17)). To study the influence of these two material properties on the balling phenomenon, we
simulated the SLM process with constant process parameters (laser
power 50 W and scanning speed 0.25 m/s) and different material
properties (surface tension 1350 mN/m, 1850 mN/m, 2350 mN/m and
viscosity 3.75 mPa⋅s, 5 mPa⋅s, 6.25 mPa⋅s ).
As shown in Fig. 13, the average width of the molten track is largest
when laser power is too high (80 W), and the shape of the molten track is
highly irregular. The average width of the molten track is smallest, and
the molten track breaks up when the laser power is too low (20 W).
Clearly scanning speed and laser power produce opposite effects. The
interactive relationship between these two process parameters is
embedded in the laser energy density (ED) parameter (ED = P/vϕ)
commonly used in the literature to characterize the SLM process in energetic terms. In (ED) P, v and ϕ represent the laser power, the scanning
speed and the diameter of the laser beam, respectively. Too many grains
melt resulting in a state of high distortion and irregularity if (ED) is too
high. The driving mechanism behind this state of high distortion is the
Fig. 14. Effect of the surface tension on the width of the molten track.
8
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
Fig. 15. Effect of the viscosity on the width of the molten track.
Data availability
decreasing surface energy of the liquid molten track in search of a final
equilibrium. If (ED) is too low, the formation of molten grains is limited
resulting in an unstable state and the occurrence of break-ups.
The width of the molten track with different properties is presented
in Figs. 14 and 15. If we consider the variance of the dimension of the
molten track to the mean value as a criterion, it turns out that the
average widths of the molten track with different surface tension and
viscosity values are nearly the same. In other words, the variation of
surface tension and viscosity has little influence on the balling phenomenon in the SLM process. This is because the coalescence process of
molten metallic grains is too fast. It only takes less than 0.001s for two
molten Ni-alloy grains with a diameter of 130 μm to join as one larger
grain according to the coalescence model of Frenkel developed in
Pokluda et al. [30]. Therefore, process parameters have much larger
influence than material properties in the metallic SLM process.
No data was used for the research described in the article.
References
[1] L.E. Loh, C.K. Chua, W.Y. Yeong, J. Song, M. Mapar, S.L. Sing, Z.H. Liu, D.
Q. Zhang, Numerical investigation and an effective modeling on the selective laser
melting process with aluminum alloy 6061, Int. J. Heat Mass Tran. 80 (2015)
288–300.
[2] H. Hu, X. Ding, L. Wang, Numerical analysis of heat transfer during multi-layer
selective laser melting of AlSi10Mg, Optik 127 (2016) 8883–8891.
[3] J.Y. Ho, K.K. Wong, K.C. Leong, Saturated pool boiling of FC-72 from enhanced
surfaces produced by selective laser melting, Int. J. Heat Mass Tran. 99 (2016)
107–121.
[4] H. Schwab, F. Palm, U. Kühn, J. Eckert, Microstructure and mechanical properties
of the near-beta titanium alloy Ti-5553 processed by selective laser melting, Mater.
Des. 105 (2016) 75–80.
[5] D.I. Bourel, H.L. Marcus, J.W. Barlow, J.J. Beaman, Selective laser sintering of
metal and ceramics, Int. J. Powder Metall. 28 (4) (1992) 369–381.
[6] N.K. Tolochko, N.V. Sobolenko, S.E. Mozzharov, I.A. Yadroitsev, Peculiarities of
Laser Sintering of Loose Powder Layers of Metal/Polymer Type, Fisikai Khimiya
Obrabotki Materialov (Physics and Chemistry of Treatment of Materials), No. 5,
Moscow, 1996, pp. 13–17.
[7] R. Li, J. Liu, Y. Shi, Balling behavior of stainless steel and nickel powder during
selective laser melting process, Int. J. Adv. Manuf. Technol. 59 (2012) 1025–1035.
[8] H.J. Niu, I.T.H. Chang, Instability of scan tracks of selective laser sintering of highspeed steel powder, Scripta Mater. 41 (1999) 1229–1234.
[9] H.J. Niu, I.T.H. Chang, Selective laser sintering of gas and water atomized high
speed steel powders, Scripta Mater. 41 (1999) 25–30.
[10] N.K. Tolochko, I.A. Yadroitsev, S.E. Mozzharov, Selective laser sintering: some
questions of physics and technology, in: Proceedings of PM 98 World Congress, 1822 October 1998, Granada, Spain, vol. 5, 1998, 407-12.
[11] N.K. Tolochko, S.E. Mozzharov, I.A. Yadroitsev, T. Laoui, L. Froyen, V.I. Titov, M.
B. Ignatiev, Balling processes during selective laser treatment of powders, Rapid
Prototyp. J. 10 (2004) 78–87.
[12] S. Das, Physical aspects of process control in selective laser sintering of metals,
Adv. Eng. Mater. 5 (2003) 701–711.
[13] D. Gu, Y. Shen, Balling phenomena in direct laser sintering of stainless-steel
powder: metallurgical mechanisms and control methods, Mater. Des. 30 (2009)
2903–2910.
[14] D. Gu, Densification behavior, microstructure evolution, and wear performance of
selective laser melting processed commercially pure titanium, Acta Mater. 60
(Issue 9) (2012) 3849–3860.
[15] X. Zhou, X. Liu, D. Zhang, Balling phenomena in selective laser melted tungsten,
J. Mater. Process. Technol. 222 (2015) 33–42.
[16] I.A. Yadroitsev, I. Smurov, Selective laser melting technology: from the single laser
melted track stability to 3D parts of complex shape, Physics Procedia 5 (2010)
551–560.
[17] P.A. Cundall, O.D.L. Strack, A discrete numerical model for granular assemblies,
Géotechnique 29 (1) (1979) 47–65.
6. Conclusion
The SLM process involves several physical phenomena essential to
account for in the simulation to accurately model the laser-powder
interaction, to understand the sintering process in the powder and to
optimize the manufacturing process. In this work, a discrete element
method is developed to model the coupled phenomena of radiative and
conductive heat transfer in a granular medium, coalescence and densification and the moving laser heat source. The model is validated by
comparison with findings from the literature. The laser heat source
scanning speed and the melting power needed by the material are well
captured by the model to avoid the balling phenomenon, an undesirable
consequence of the sintering process. The findings in this paper represent a crucial step for the optimization of the process of laser sintering
for additive manufacturing. We determine that the variation of surface
tension and viscosity has little influence on the balling phenomenon,
and that process parameters have much larger influence than material
properties in the metallic SLM process.
Declaration of competing interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
9
M. Boutaous et al.
International Journal of Thermal Sciences 167 (2021) 107011
[18] L. Dong, A. Makradi, S. Ahzi, Y. Remond, Three-dimensional transient finite
element analysis of the selective laser sintering process, J. Mater. Process. Technol.
209 (2009) 700–706.
[19] D. Defauchy, Simulation du Procédé de Fabrication Directe de Pièces
Thermoplastiques Après Fusion Laser de Pouder, Ph.D. Thesis, Arts et Métiers
ParisTech, 2013. Paris, France.
[20] R.D. Jodhpur, Numerical Modelling of Process Governing Selective Laser Sintering,
2014. Ph.D. Thesis, Darmstadt, Germany.
[21] P. Peyré, Y. Rouchausse, D. Defauchy, G. Regnier, Experimental and numerical
analysis of the selective laser sintering (SLS) of PA12 and PEKK semi-crystalline
polymers, J. Mater. Process. Technol. 225 (2015) 326–336.
[22] W. Yu, S.L. Sing, C.K. Chua, C.N. Kuo, X.L. Tian, Particle-reinforced metal matrix
nanocomposites fabricated by selective laser melting: a state-of-the-art review,
Prog. Mater. Sci. 104 (2019) 330–379.
[23] R. Martinez, I. Todd, K. Mumtaz, In situ alloying of elemental Al-Cu12 feedstock
using selective laser melting, Virtual Phys. Prototyp. 14 (3) (2019) 242–252.
[24] S.L. Sing, F.E. Wiria, W.Y. Yeong, Selective laser melting of titanium alloy with 50
Wt% tantalum: effect of laser process parameters on Part Quality, Int. J. Refract.
Metals Hard Mater. 77 (2018) 120–127.
[25] A.V. Gusarov, I. Smurov, Radiation transfer in metallic powder beds used in laser
processing, J. Quant. Spectrosc. Radiat. Transf. 111 (2010) 2517–2527.
[26] L. Watson, Vargas-Escobar, J.J. McCarthy, Conductivity of granular media with
stagnant interstitial fluids via thermal particle dynamics simulation, Int. J. Heat
Mass Tran. 45 (2002) 4847–4856.
[27] L. Xin, M. Boutaous, S. Xin, D.A. Siginer, Multiphysical modeling of the heating
phase in the polymer powder bed fusion process, Additive Manufacturing 18
(2017) 121–135.
[28] L.K. Suk-Joong, Sintering: Densification, Grain Growth & Microstructure, Elsevier,
Oxford, UK, 2005.
[29] M. Smoluchowski, Drei Vorträge über Diffusion, Brownsche Molekularbewegung
und Koagulation von Kolloidteilchen, Phys. Z. 17 (1916) 557–571, 585–599.
[30] O. Pokluda, C.T. Bellehumeur, J. Vlachopoulos, Modification of Frenkel’s model
for sintering, AIChE J. 43 (1997) 3253–3256.
[31] S.J. Cho, S.J. Kang, D.N. Yoon, Effect of entrapped inert gas on pore filling during
liquid phase sintering, Metall. Mater. Trans. 17 (A) (1986) 2175–2182.
[32] H. Park, S.-J. Cho, D.N. Yoon, Pore filling process in liquid phase sintering,
Metallur. Trans. A 15A (1984) 1075.
[33] L. de Johnson, M.N. Rahaman, Sintering stress of homogeneous and heterogeneous
powder compacts, Acta Metall. 36 (223) (1988).
10