International Journal of Thermal Sciences: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

International Journal of Thermal Sciences 124 (2018) 447–458

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


journal homepage: www.elsevier.com/locate/ijts

Numerical analysis of weld pool behaviors in plasma arc welding with the MARK
lattice Boltzmann method
Junjie Caia,e, Yanhui Fenga,b,∗, Junjie Zhoua, Yan Lic, Xinxin Zhanga,b, Chuansong Wud
a
School of Energy and Environmental Engineering, University of Science and Technology Beijing, Beijing 100083, China
b
Beijing Key Laboratory of Energy Saving and Emission Reduction for Metallurgical Industry, University of Science and Technology Beijing, Beijing 100083, China
c
State Key Laboratory of Heavy Oil Processing, China University of Petroleum-Beijing, Beijing 102249, China
d
MOE Key Lab for Liquid-Solid Structure Evolution and Materials Processing, Institute of Materials Joining, Shandong University, Jinan 250061, China
e
China Nuclear Power Technology Research Institute, China General Nuclear Power Corporation, Shenzhen 518031, China

A R T I C L E I N F O A B S T R A C T

Keywords: The lattice Boltzmann method (LBM) is first applied to both simulate and analyze the heat transfer and fluid flow
Plasma arc welding (PAW) with phase change in a workpiece of stationary plasma arc welding (PAW). The heating and key-holing effects
Weld pool are considered with a modified heat source model. The total enthalpy lattice Boltzmann model is employed to
Modified heat source investigate the phase change process in the weld pool. The predicted weld pool geometry agrees well with the
Lattice Boltzmann method (LBM)
experimental results and average relative errors remain below 12%. It seems that LBM is more intuitive to
Solid-liquid phase change
display the fluid flow in the weld pool, e.g. circulations, than the finite volume method. However, the effect of
fluid flow on the PAW process is indicated to be non-negligible, and particularly responsible for the upper shape
of the weld pool. Mechanisms of the formation of the hump in the fusion line are further analyzed from the
perspective of fluid motion. Finally, the sulfur activity of workpiece and welding current is discussed, showing
that sulfur activity only affects pool geometry. High welding current accelerates the penetration with low time
consumption and high energy efficiency; however, the optimum current is a better choice to ensure high weld
quality.

1. Introduction characterizing the keyhole effect [4–7], (2) coupling the plasma arc and
weld pool [8–10], and (3) developing effective heat source models
With high energy density and a most powerful arc force, plasma arc [11–15]. In the first category, researchers employed the volume of fluid
welding (PAW) is one of the most productive welding technologies. method (VOF) [6] or the Level Set (LS) method [7] to track the dy-
During the PAW process, the plasma arc with extremely high tem- namic-variation keyhole boundary in analyzing the fluid flow and heat
perature instantly provokes local melting, resulting in a molten pool transfer in the weld pool [16]. Although the interface tracking methods
during a short time. Afterwards, the liquid metal evaporates and if the develop rapidly, the features of keyhole and plasma still remain difficult
keyhole emerges, the welding process will be greatly accelerated. to describe. The second category investigates the interaction between
Finally, the workpiece will be cooled by the ambient air and the weld the plasma arc and the weld pool and the complete welding mechanism
bead forms [1]. is better understood. In these models, the keyhole was dynamically
Recently, PAW has been widely used in industrial applications, e.g. formed [8,9] or preset at a certain shape [10]. However, neither of
many delicate components of aircraft and the shell of missiles are PAW them could accurately predict the weld bead shape and the described
welded. Consequently, PAW weld quality is extremely important. The simulations were all cumbersome and time consuming. In the third
weld quality is strongly characterized by the weld pool properties, category, effective heat source models were developed for the energy
which determine the mechanical properties of the weld [2,3]. To equation, considering the keyhole effect, instead of directly describing
achieve a full understanding of the weld pool properties and related the features of the keyhole. Apparently, the requirement of keyhole
processing mechanisms, many numerical studies have been conducted. interface tracking was removed and the computational domain was
It is crucial to model the key-holing process because both the keyhole only in the workpiece. Liu et al. [11] reported that a conical heat source
formation and weld pool development are inseparable. Generally, ex- was superior to the widely used Gaussian surface heat source [17,18].
isting studies can be grouped into three major categories: (1) Wu et al. [12] applied a combined heat source, consisting of a double-


Corresponding author. School of Energy and Environmental Engineering, University of Science and Technology Beijing, Beijing 100083, China.
E-mail address: yhfeng@me.ustb.edu.cn (Y. Feng).

http://dx.doi.org/10.1016/j.ijthermalsci.2017.10.026
Received 3 December 2016; Received in revised form 2 September 2017; Accepted 23 October 2017
1290-0729/ © 2017 Elsevier Masson SAS. All rights reserved.
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Nomenclature Qm Discrete source term (W/m3)


s 0 , s e , sε ,
s j, sq , s p Relaxation factors
a1, a2, b Shape coefficients of heat source
S Relaxation matrix in the moment space
B0 ∼B4 Coefficients of fitting curves
t Time (s)
c Lattice speed (m/s)
T Temperature (K)
Cp Specific heat at a constant pressure (J/(kg·K))
Tl, Ts Liquidus and solidus temperature (K)
ei Discrete lattice velocity in direction i (m/s)
u Velocity (m/s)
f Density distribution function (kg/m3)
ux , uy Horizontal, vertical components of u (m/s)
fi Density distribution in direction i (kg/m3)
U Welding voltage (V)
fl Liquid fraction
x, y Coordinates (m)
Fi Sum of force terms in direction i (N/kg)
F Sum of force terms (N/kg)
Greek symbols
Fm Discrete force term in moment space (N/kg)
Fx, Fy Horizontal, vertical components of F(N/kg)
α Thermal diffusivity (m2/s)
g Temperature distribution function (J/kg)
β The included angle (degree)
gi Temperature distribution in direction i (J/kg)
δt Time step
h Shape coefficient of heat source
δx Lattice space
H Total enthalpy (J/kg)
Λ Relaxation matrix in the velocity space
Hl, H s Enthalpy of liquidus, solidus point (J/kg)
η Welding efficiency coefficient
I Current (A)
μ Dynamic viscosity (kg/(m·s))
I Unit matrix
ρ Density (kg/m3)
kT Thermal conductivity (W/(m·K))
ν Kinetic viscosity (m2/s)
L Thickness of the workpiece (mm)
χ1, χ2 Energy distribution coefficient
Lα Latent heat of melting (J/kg)
m Distribution functions in the moment space
Superscripts and subscripts
M Orthogonal transformation matrix
p Pressure (Pa)
eq Equilibrium distribution function
q Internal heat source (W/m3)
i, j Directions in a lattice
Q0 Thermal energy absorbed from plasma arc (W)
T Transposition of matrix
Qi Heat source in direction i (W/kg)
Q Heat source (W/kg)

ellipsoidal heat source and a cylindrical heat source. The predicted regarded as a more suitable approach. Jiaung et al. [33] first proposed
cross-section of the weld geometry matched well with the measured an enthalpy-based model for heat conduction. In conjunction with an
cross-section. Li et al. [13] modified the previous model and established enthalpy formulation for the treatment of solid–liquid phase change,
a volumetric model, combining a double-ellipsoidal heat source with a one-dimensional half-space melting and solidification as well as two-
conical heat source. The resulting shape of the modified heat source dimensional solidification were simulated in a corner. A large amount
model is more approximate to the actual configuration of welds. of iterations were required at each time step. To simplify the iteration
All these studies are based on macroscopic continuum equations, procedure, Chatterjee and Chakraborty [34] employed a thermo-
and it is difficult to mathematically model the underlying details in dynamically-consistent enthalpy-updating scheme that was previously
microscopic physics [19]. In recent years, the lattice Boltzmann method proposed by Brent et al. [35]. In their follow-up research, Chatterjee
(LBM) has become an alternative and promising numerical scheme for and Chakraborty modeled the phase change in the presence of fluid
modeling fluid flows. Due to its mesoscopic nature, the LBM has ad- flow, viscous heat dissipation, and crystal growth [32,36–38]. To avoid
vantages over conventional methods including finite element, finite iteration steps and to improve both the formulation stability and effi-
volume, and finite difference methods. The LBM succeeds in simula- ciency, Eshraghi and Felicelli [39] treated the latent heat source term
tions involving interfacial dynamics and complex boundaries [20]. It is by solving linear equations. Huang et al. [40] proposed a new model to
easy to perform parallel computation with a simple but efficient cal- avoid the iteration steps or by solving a group of linear equations. The
culation procedure [21]. More detailed introductions of the origin and phase interface was traced by updating the total enthalpy. Huang's
development of the LBM can be found in previous publications [19,21]. model simplified the calculation procedure, which saved much time.
Developed over decades, the LBM has been widely used to simulate
phase change problems such as crystal growth in foundry [22], foam
production [23], melting of fatty acids [24], and natural convection in
porous media [25,26]. Generally, two types of methods are applied
including the phase-field method and the enthalpy-based method. The
phase-field method was first employed by De Fabritiis et al. [22]. The
phase transition, which was redefined as a chemical reaction by ana-
logy, was traced by introducing an auxiliary parameter and a series of
extended work was conducted subsequently [27–31]. The solidification
problems with anisotropic crystal growth and dendritic growth in a
binary-alloy system were solved via the phase-field model. Although
the phase-field method possesses a strong fundamental basis, it is not
effective in practice due to certain stringent computational constraints
[32]. To resolve the interfacial region, extremely finer grid spacings are
required. Therefore, the fixed-grid enthalpy-based method was Fig. 1. Physical model of PAW process.

448
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Table 1
Welding conditions in PAW experiments [1].

Parameters Values

Distance from torch to workpiece 6.0 mm


Orifice diameter 3.8 mm
Plasma gas (pure Argon) flow rate 3 L/min
Shielding gas (pure Argon) flow rate 20 L/min
Welding voltage (Case A) 21.4 V
Welding voltage (Case B) 24.2 V
Welding current (Case A) 140 A
Welding current (Case B) 169 A

Table 2
Chemical composition of 0Cr18Ni9 [1].

Fig. 2. Combined heat source model.


Alloy element C Cr Ni Mn Si S P Fe

Content (wt %) 0.08 17–19 8–11 2.0 1.0 0.2 0.035 Balanced

Table 3
Thermophysical properties of 0Cr18Ni9 [1].

Symbol Nomenclatures Values

Ts Solidus temperature 1663 K


Tl Liquidus temperature 1723 K
Lα Latent heat of material 2.6 × 105 J/kg
ρ Density 7200 kg/m3
Cp Specific heat 630 J/(kg·K)

change in welding process has been accomplished via the LBM.


To date, most numerical simulations of the PAW process are based
on the continuum equations and the calculation is time-consuming with
cumbersome procedures. Furthermore, few analyses have been per-
formed on the formation mechanism of weld bead, due to the defi-
ciencies of conventional methods. In this paper, to raise the numerical
efficiency and to better understand the inherent mechanism, we de-
veloped a two-dimensional lattice Boltzmann (LB) model to describe
the heat transfer and fluid flow during the stationary PAW process,
which was based on the total enthalpy model. An effective heat source
Fig. 3. Schematic sketch of discrete velocities for D2Q9 model. was applied to take the heat effect of the plasma arc into account. The
predicted results were compared to the experimental data and the re-
Afterwards, Huang et al. [41] modified their model, thus considering sults of the finite volume method (FVM). To better understand the weld
the variations of properties. In engineering applications, the LBM has pool properties, the weld bead, the velocity distribution, and the tem-
been applied in casting [42], injection moulding [43], and jet flows perature field were further analyzed in detail. The effect of varied sulfur
[44]. However, to the best of our knowledge, no research on phase concentrations in material and welding currents were studied. Our

Fig. 4. Experimental system setup. (a) Schematic sketch and (b) platform photograph.

449
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Fig. 5. Comparison of the fusion zones between experimental images and simulated results. (a) Case A and (b) Case B.

Table 4 3) The thermal conductivity and dynamic viscosity depend on the


Detailed comparison of experimental data and calculated results. temperature and other thermal properties such as density and spe-
cific heat, which are assumed to be constant [1].
Parameters Case A Case B
4) The model is two dimensional and symmetrical along the y-axis
Measured LBM FVM Measured LBM FVM since the welding torch is stationary [9].
5) The thermal boundary conditions are set to be adiabatic [5,45,62].
Top surface width (mm) 4.89 5.02 4.92 5.13 5.09 5.34
Bottom surface width 1.12 1.03 1.03 1.00 0.90 0.99
(mm)
Based on the above assumptions, the governing equations are ex-
pressed as [35]:
Average relative error – 11.75 15.27 – 11.88 22.04
(%)
∂ρ
+ ∇ ·(ρu) = 0
Maximum relative error – 21.06 30.63 – 28.49 47.26 ∂t (1)
(%)
∂ (ρ u)
+ ∇•(ρuu) = −∇ p+ ∇•(μ∇u) + ρFe
∂t (2)
work offers several indications to further study and improve the PAW
process. The model and the method presented in this paper can be ∂ (ρCp T ) ∂ (ρfl La)
+ ∇•(ρCp T u) = ∇•(kT ∇T ) − +q
modified and adopted for other welding processes with similar me- ∂t ∂t (3)
chanisms such as the tungsten inert gas welding (TIG), gas tungsten arc Where ρ, u, p and T represent the density, velocity, pressure, and
welding (GTAW), and gas metal arc welding (GMAW), provided that temperature, respectively. μ, Cp and kT represent the dynamic viscosity,
some modifications upon the heat source are employed. specific heat at constant pressure, and thermal conductivity, respec-
tively. fl and Lα represent the liquid fraction and latent heat of melting,
2. Mathematical model respectively. Fε represents the sum of external forces including the
Marangoni shear on the free surface, equivalent frictional resistance on
2.1. Physical model the solid-liquid interface (i.e. the Darcy's resistance), buoyancy force,
and electromagnetic force. These forces have previously been defined
During the common PAW process, the plasma arc impinges on the [10,13]. q represents the internal heat source. The heat source model
workpieces, creates a weld pool, and then passes along the joint area. [13] is redeveloped as schematically demonstrated in Fig. 2. It consists
However, in a stationary PAW, the welding torch is fixed so that the of a semi-ellipsoidal and a trapezoidal heat source. The heat source
welding speed is equal to zero and the numerical model can be sim- distributions are expressed as follows:
plified as two-dimensional. As schematically presented in Fig. 1, the Upper part: ellipsoidal heat source:
Cartesian coordinate is established at the cross-section of the workpiece
(the shadowed part) to simplify the calculation. AB (15 mm), BC 1200 3 Q0 χ1 3x 2 3y 2
q (x , y ) = b
exp ⎛− 2 ⎞ exp ⎛− 2 ⎞
⎜ ⎟ ⎜ ⎟

(6 mm), CO (15 mm), and OA (6 mm) are the boundaries of the hπ π ⎝ b ⎠ ⎝ h ⎠ (4)
2
workpiece, i.e. the simulation domain 15 mm × 6 mm, respectively. In
the Results and discussions section, the numerical results from 0 to Lower part: trapezoidal heat source:
6 mm on the x-axis are shown while those from 6 to 15 mm are hidden 900Q0 χ2 3x 2
because the temperature at the rear end (far away from the arc) of the q (x , y ) = exp ⎛⎜3 − 2 ⎞⎟
π (L−h)(exp(3) − 1)(a12 + a22 + a1 a2) ⎝ x0 ⎠ (5)
workpiece is too low to aid further analysis.
L−h−y
2.2. Governing differential equations x 0 (y) = a2 − (a2 − a1)
L−h (6)

The continuum model is developed by introducing the following where Q0 = ηUI represents the thermal energy absorbed from plasma
assumptions. arc, which depends on the welding efficiency coefficient η, voltage U
and current I. η is set to be 0.5 in this paper according to [12,46]. χ1 and
1) The effects of both plasma arc and keyhole are indirectly considered χ2, which are energy distribution coefficients, are set to be 0.75 and
via the effective heat source model. 0.25, respectively and the sum of both remains at 1 [13]. a1, a2, b and h
2) The flow is treated as a laminar and incompressible Newtonian fluid represent shape coefficients of the heat source. L represents the thick-
flow [13]. ness of the workpiece. This heat source model is built by previous ex-
perience [12], experiments [46], and a number of tests. Although it

450
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Fig. 6. Temperature fields and velocity distributions.

451
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Fig. 7. Geometric developments of the weld pool versus time.

Fig. 8. Comparison of solidus lines between the conditions with and without fluid flow.

Fig. 10. (a) Heat flux, (b) velocity, (c) temperature gradient and (d) included angle be-
tween velocity vector and temperature gradient adjacent to the solidus lines at 2.0s.

directs at the 6-mm thickness workpiece, it can also be applied to dif-


ferent thicknesses in the form of Eqs. (4–6) and only the coefficients
should be revised correspondingly. In the above equations, all proper-
ties of welding material (stainless steel 0Cr18Ni9) and other related
parameters, such as μ, Cp, kT, and μ0 are given by previous publications
Fig. 9. Streamlines and solidus line at 2.0s.

452
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Fig. 11. Variation of weld pool properties with sulfur concentration at 2.0s. (a) Surface tension coefficient, maximum flow speed in the weld pool and top surface width and (b) solidus
lines.

[10,13]. The model is suitable for stainless steel and other similar m g (x, t ) = m g (x, t ) − Sg [m g (x, t ) − meq
g (x , t )] + δt (I−Sg /2) Qm (11)
materials with a high melting point (beyond 1000 K) and low thermal
conductivity (below 100 W m−1 K−1). If the object is some type of where mf and m g represent the post-collision distribution functions in
material like an aluminum alloy, which conducts heat and melts swiftly, the moment space. I is the unit matrix. With an orthogonal transfor-
the stability and accuracy of the simulation should be studied and mation matrix M given by Lallemand and Luo [48].
tested.
⎛ 1 1 1 1 1 1 1 1 1 ⎞
−4 −1 −1 −1 −1 2 2 2 2
⎜ ⎟
2.3. Lattice Boltzmann equations 4 −2 −2 −2 −2 1 1 1 1
⎜ 0 1 0 −1 0 1 −1 −1 1 ⎟
M=⎜ 0 −2 0 2 0 1 −1 −1 1 ⎟
By neglecting the viscous dissipation and compression work, the ⎜ 0 0 1 0 − 1 1 1 − 1 − 1⎟
double-distribution-function lattice Boltzmann model first proposed by ⎜ 0 0 −2 0 2 1 1 − 1 − 1⎟
Shan and Chen [47] is redeveloped. The velocity and temperature fields ⎜ 0 1 − 1 1 − 1 0 0 0 0 ⎟⎟

are solved via the distribution functions for density and temperature, ⎝ 0 0 0 0 0 1 − 1 1 − 1⎠ (12)
respectively. To simplify the calculation, the total enthalpy-based
model obtains the temperature field and phase-fraction distribution. The distribution functions, relaxation matrix and other terms can be
The distribution functions with MRT collision scheme can be ex- transformed into new forms in moment space.
pressed as follows: For the velocity field, mf = Mf is the distribution function and
Sf = MɅfM−1 is the relaxation matrix, which is generally given as:
fi (x + ei δt , t + δt ) = fi (x, t ) − Λf − ij ⎡f j (x, t ) − f jeq (x, t ) ⎤ + δt Fi Sf = diag (sf − 0, sf − e, sf − e, sf − j, sf − q, sf − j, sf − q, sf − p, sf − p) (13)
⎢ ⎥ (7)
⎣ ⎦
The equilibrium function can be expressed according to Lallemand
gi (x + ei δt , t + δt ) = gi (x, t ) − Λg − ij ⎡gj (x, t ) − g jeq (x, t ) ⎤ + δt Qi and Luo [48] as:
⎢ ⎥ (8)
⎣ ⎦ T
meq 2 2 2 2
f = (ρ , − 2ρ + 3u , ρ − 3u , u x , − u x , u y , − u y , u x −u y ,u x u y ) (14)
where f and g denote the distributions for density and temperature,
respectively. The subscripts i and j represent the directions in a lattice. where the superscript “T” denotes the transposition of matrix. To con-
As Fig. 3 depicts, for the D2Q9 discrete velocity model, the nine discrete sider the discrete lattice effects and to recover the exact Navier-Stokes
velocities are given by: equations, the discrete force term in moment space is given as [49,50]:

⎧ (0,0) i=0 Fm = (0,6F⋅u , −6F⋅u , Fx , −Fx , Fy , −Fy , 2Fx u x −2Fy u y , Fx u y + Fy u x ) T


ei = c(cos [(i − 1) π /2], sin[(i − 1) π /2]) i = 1,2,3,4 (15)

⎩ 2 c(cos [(2i − 1) π /4], sin[(2i − 1) π /4]) i = 5,6,7,8 (9) where F=(Fx,Fy)T. With the post-collision distribution function mf , the
where c = δx/δt represents the lattice speed. δx and δt are set as distribution function f in the velocity space can be updated through the
5 × 10−5 m and 10−5 s, respectively. In the direction i, fi (or gi), and inverse transformation f = M−1mf . After collision process and inverse
fieq (orgieq ) are distribution function and equilibrium distribution func- transformation, the advection process f (x + eiδt,t+δt) = f (x,t) is
tion for density (or temperature), respectively. Λf − ij (or Λg − ij ) are the executed.
relaxation parameters. Fi and Qi represent discrete force term and heat Similarly, the temperature field evolution can be performed. The
source term, respectively. Accordingly, f (or g), feq (or geq), Λf f (or Λg ),F equilibrium function and discrete source term are given as:
and Q are in the form of matrices.
⎛ u2 u2
In the calculation, the evolution of the model consists in two steps: meq
g = ⎜H , − 4H + 2Cp T + 3Cp T 2 , 4H − 3Cp T − 3Cp T 2 ,
collision and advection. The collision and advection are conducted in ⎝ c c
T
moment space and velocity space, respectively. Generally, the collision u u uy uy u 2
x − u 2
y ux u y ⎞
is executed as [48–50]: Cp T x , −Cp T x , Cp T , −Cp T , Cp T ,Cp T ⎟
c c c c c2 c2 ⎠
mf (x, t ) = mf (x, t ) − Sf [mf (x, t ) − meq
f (x , t )] + δt (I−Sf /2) Fm (10) (16)

453
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Fig. 12. Velocity profiles under different sulfur concentrations at 2.0s. (a) 0.4% (b) 1.2% (c) 2.0% (d) 2.8%.

T
q −2q q ⎧ 0, H ≤ Hs
Qm = ⎜⎛ , , , 0,0,0,0,0,0⎟⎞ ⎪ Hl − H
⎝ρ ρ ρ ⎠ (17) fl = H − H , Hs < H < Hl
⎨ l s
⎪1, H ≥ Hl (20)

where H represents the total enthalpy. The macroscopic variables such
as density, velocity, total enthalpy temperature, and liquid fraction are where Hs (or Hl) represent the total enthalpy at the solidus (or liquidus)
obtained by: temperature Ts (or Tl). The kinetic viscosity ν and the thermal diffusivity
8 8 8 α satisfy ν = δt (1/sf-p-0.5)/3 and α = δt (1/sj-g-0.5)/3, respectively.
δt δt q To simplify the calculation, the boundary conditions for energy are
ρ= ∑ fi , u= ∑ ei fi +
2
F, H = ∑ gi + 2 ρ
i=0 i= 0 i=0 (18) set to be adiabatic [5,45,62]. For momentum, as shown in Fig. 1, the
boundary OA is set to a symmetrical boundary. The bounce-back
H −H
scheme [62] is employed to the boundary BC to ensure the conservation
s
⎧Ts − Cp , H ≤ Hs of mass and momentum, which always remains at solid state. The non-

Hl − H H − Hs equilibrium extrapolation scheme [51] is applied to the top and bottom
T = Ts H − H + Tl H − H , Hs < H < Hl
⎨ l s l s boundaries where the Marangoni force and viscous shearing stress
⎪Tl + H − Hl , H ≥ Hl counterbalance with each other for better stability and applicability.
⎪ Cp (19)

The simulation was self-programmed with the C++ language. The
CPU time for one simulation was about 4 h on an Intel Xeon E5-2687 W,
3.10 GHz computer.

454
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

4. Results and discussions

4.1. Experimental and simulated results

To validate the welding simulations, numerical results of the LBM


are compared with experimental data and calculation results in our
previous publication [13] with FVM, as shown in Fig. 5. The weld beads
simulated via LBM agree well with the experimental images for two
cases. The simulated results of FVM are also coincident with the actual
cross-section of the fusion zone. It should be pointed out that there are
46 × 39 grids in FVM [13], but 300 × 120 grids in LBM, respectively.
However, the computational time for LBM is about half of that for FVM.
This indicates that LBM is more efficient in the calculation, an ad-
vantage that will be far more significant for parallel computing.
Comparisons of the weld bead widths and relative errors between
LBM simulations and experiments are listed in Table 4. Since the geo-
metry of the molten pool is normally discussed depending on different
fusion depths, data points with the same value of “y” are extracted. The
x-distance errors from each measured point to the corresponding si-
mulated point are firstly calculated. Then, the relative errors are ob-
tained with the experimental points divided by the errors, as Eq. (21).
relative error = (x experiment − x calculation )/ x experiment × 100% (21)

The average relative errors of both cases are below 12%. In addi-
tion, either average or maximum relative errors are smaller for LBM
than for FVM.

4.2. Evolution of the temperature, velocity and weld pool geometry

In this section and in section 4.3, all results and discussions are
given upon Case A. At the bottom right corner of each panel of Fig. 6,
Fig. 13. (a) Heat flux and (b) resultant flow speed adjacent to the solidus lines at 2.0s.
the reference vector 0.1 m/s may help to recognize the magnitude of
the flow speed. From Fig. 6: (a) at 0.4 s, as marked in the contour lines,
3. Experimental description the temperature increases to 1663 K (i.e. the melting point) and solid-
liquid phase change begins at the top of the workpiece. The mushy zone
The plasma arc welding machine Trans Synergic 5000, manu- (i.e. solid-liquid two-phase zone) only appears, where the fluid hardly
factured by the FRONIUS Company, Austria, was utilized in our ex- flows; (b) at 0.6 s, the molten pool with liquid zone and mushy zone
periments. The machine included a digital welding power source, a emerges, and continues to enlarge both in width and depth. At the
plasma module, a PAW torch, and a working table. A control PC was upper part, the temperature keeps increasing, along with apparent fluid
used to adjust the process parameters and tp monitor the process signals flow, while the fluid is nearly stagnant at the lower part; (c) at 0.8 s, the
via input/output interfaces. The vision sensor consisted of an image workpiece has been completely melted through. Rapid penetration is
acquisition card and a CCD camera with filter. The CCD camera was caused by a large amount of heat transferred to the bottom, which
fixed at the backside of the workpiece. The observation angle between stems from the distribution of the heat source. The shape of the fusion
the camera axis and the horizontal plane was 70°. The distance between zone appears like a “reversed bugle”, which coincides nicely with
the camera and the target was 200 mm. This enabled a rapid feedback previous research [4,6,13]; (d) at 1.0 s (after penetration), the contour
and adjustment for different welding conditions Via digital DSP mi- lines mainly develop in the horizontal direction, along with the
croprocessor that can supervise the entire welding procedure. Expected widening molten pool. The fluid flow diffuses to the lower part, but the
arc plasma can be produced in the FPM plasma module by adjusting speed is still low; (e) - (f) between 1.2 and 1.4 s, at the upper part,
both welding current and voltage, as shown in Fig. 4. The shielding gas primarily driven by the Marangoni shear, the flow speed increases
surrounds the workpiece fixed on the platform and the plasma arc can sharply and the maximum emerges with 0.07 m/s. The fluid flows along
straight impinge on the workpiece. More details of the system can be the top surface, follows along the edge of pool (or solidus line), then
obtained from previous publications [52,53]. Two cases of experiments swerves to the center of the pool, and finally turns back to the top
with different welding power were conducted and the related condi- surface. As a result, a clockwise circulation flow emerges. Furthermore,
tions are listed in Table 1. After the experiments, the workpiece was at the lower part, under the effects of electromagnetic force and upper
processed in a metallographical way. For this, the fusion zones at the flow, the fluid speed increases gradually and another circulation is
transverse cross-section of the workpiece (or weld beads) were visua- found, which flows in counterclockwise direction. Two circulations
lized as shown in the following section. The chemical composition and become more and more apparent, as illustrated in Fig. 6(g)–(j) between
thermophysical properties of the material are listed in Tables 2 and 3, 1.6 and 2.2 s. At 2.0 s, the maximum temperature increases to 3200 K
respectively. and the contour lines indicate that the top and bottom widths of the

Table 5
Top surface width at the moment of penetration.

Sulfur concentration (%) 0.2 0.6 1.0 1.4 1.8 2.2 2.6 3.0 Conduction

Size (mm) 6.53 6.51 6.49 6.48 6.47 6.47 6.46 6.45 6.44

455
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

Fig. 14. Variation of weld properties with welding current at penetration. (a) Time consumption, energy consumption and top surface width and (b) average growth rates of depth and
top surface width and the ratio between them.

molten pool enlarge to 5.02 mm and 1.03 mm, respectively. The Due to the limits of macroscopic continuum models, few researchers
maximum flow velocity is about 0.11 m/s (at x = 3 and approximately provided a reasonable explanation for how the fluid flow affected the
y = 5.9), which is consistent with most published research results shape of the weld pool. With a mesoscopic kinetic background, the LBM
[5,54,55]. Circulation flows have been observed in the simulation with makes it easier to understand the hydromechanical principles in the
unified FVM models, which couples the plasma arc and the weld pool weld pool. Fig. 9 displays the streamlines in the weld pool at 2.0 s. Two
[9,10], but the calculations were cumbersome and time consuming. circulations exist at the weld pool, one at the upper part and in clock-
However, no circulations were displayed in the FVM simulations of the wise direction and the other at the lower part in counterclockwise di-
weld pool alone [6,7,13]. We attribute this to the fact that the plasma rection. According to Ref. [1], both are primarily driven by the Mar-
arc effect was taken into consideration by introducing the heat source angoni shear and the electromagnetic force, respectively. At the upper
in the energy equation; however, its effect on the fluid flow in the part, the fluid particles flow along the top surface of the weld pool until
welding pool has been underestimated. Interestingly, in this paper, it is they run into an obstacle (or unmelted metal). Due to Darcy's re-
found that the LB model (with only an effective heat source in the sistance, the fluid particles are bounced back or stream along the so-
energy equation) could reveal the circulation flows in the weld pool, lidus line. In a similar way, at the lower part, the fluid particles flow in
with the benefit of with much less computational load. This is possibly the opposite direction. The two fluid flows collide with each other,
because the LBM is based on a particle picture, which it is physically where a speed drop appears (see insets in Fig. 9). It can be rationally
more intuitive than the FVM. concluded that near the solid surface, where the two flows are con-
Fig. 7 shows variations of fusion depth, top surface width and fluent, the temperature increases by inches and the heat transfer rate
bottom surface width of the weld pool. Prior to 0.35 s, no metal has drops remarkably.
melted yet and all sizes are equal to zero. At 0.4 s, the weld pool forms, To validate this theory and to obtain deep insight into the heat
with a top surface width of 3.1 mm and a depth of 0.3 mm. At 0.4–0.6 s, transfer adjacent to the solid phase, the heat flux adjacent to the solidus
as shown by the local enlargement plot in Fig. 7, the top surface width lines are shown in Fig. 10 (a). Firstly, the heat flux under fluid flow is
rises quickly in a nearly constant rate. In contrast to the top surface close to the conduction case between 1 mm and 1.6 mm on the x-axis
width, the growth pattern of depth is “increasing sharply - turning (i.e. the lower part) and much larger beyond 3.2 mm (i.e. the upper
slowly - increasing again”, which coincides with previous publications part), which is consistent with the conclusions drawn in the above
[7,56]. At 0.71 s, the workpiece is completely penetrated through and sections. Secondly, focusing only on the case of fluid flow, the heat flux
the bottom metal is also fused. The bottom surface width keeps growing drops at the area between 1.8 and 2.8 mm, corresponding to the hump
slowly and at the meanwhile, the fusion depth stops rising. Finally, the shown in Fig. 8. Therefore, the hump appearing at this area is caused by
top and bottom surface widths enlarge to their maximum sizes of the lower heat flux compared to the nearby areas on the solidus line. In
10.04 mm and 2.06 mm, respectively. addition, it is remarkable that the undulation of the heat flux keeps
pace with the variation of the resultant speed in Fig. 10 (b), especially
at the area beyond 2.2 mm on the x-axis. This implies that the fluid flow
4.3. Analysis of the weld bead
actually contributes to the widening of the weld pool. According to the
field synergy principle proposed by Guo et al. [57,58], the convection
Fig. 8 demonstrates how the fluid flow affects the formation of the
heat transfer depends on three fields: velocity vector, temperature
weld bead. In comparison to the solidus lines formed via conduction
gradient, and the included angle β between them. The heat transfer
only (i.e. without consideration of fluid flow), those under the condi-
enhancement requires large values of velocity and temperature gradient
tion with fluid flow are slightly wider. This means that during the PAW
and small included angle. As shown in Fig. 10 (b), (c), (d), these three
process, the conduction is the dominant mechanism of heat transfer,
requirements are simultaneously met at around 3.5–4.0 mm. Moreover,
especially at the lower part of the weld pool. However, above 4.8 mm
at the hump area (i.e. 1.8–2.8 mm), a β smaller than 90° does not en-
on the y-axis (i.e. the upper part), the fluid flow markedly promotes
hance the heat transfer rate because of the speed drop. Among the three
pool enlargement. In addition, at 2.0 s, an apparent hump (marked with
requirements, it is clear that the flow speed is most influential in this
a red circle in Fig. 8) in the solidus line is observed in both experiment
paper. Intrinsically, the pattern of fluid flow (two circulations in our
and simulation with fluid flow, which is different from the smooth line
case) determines the effect of field synergy and the geometrical char-
obtained via conduction. Therefore, it can be shown that the fluid flow
acteristics of the weld bead (e.g. the hump).
is mainly responsible for the shape of the weld bead. A similar con-
clusion was drawn in previous research [13].

456
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

4.4. Effects of sulfur concentration in the specimen and welding current electromagnetic force also increases, which opposes the Marangoni
shear [15]. In this case, the development of the top surface width is
The above sections showed that the weld pool geometry is greatly inhibited to a certain degree. If those factors eliminate the effect of
influenced by the Marangoni shear [59], which in turn depends on the increasing current, the top surface might be narrowed and vice versa.
surface tension coefficient. According to Sahoo [60], the surface tension This uncertainty leads to a distinct difficulty to theoretically predict the
coefficient varies inversely with the sulfur concentration in the spe- geometry of the weld pool. However, our tests provide the preliminary
cimen at high temperature. Using the numerical model in this paper, we instruction to obtain the optimum value within definite currents. As
tested the weld pool behavior under different sulfur concentrations shown in Fig. 14 (b), both growth rates of top surface width and depth
(0.2–3.0%), as displayed in Figs. 11–13. are increasing with welding current. However, the ratio between them
In Fig. 11 (a), the surface tension coefficient decreases with in- (depth to top surface width) is not stable and the maximum is ap-
creasing sulfur concentration. With the weakening Marangoni effect, proximately 94%. This indicates that the peak of depth-to-width ratio of
the maximum flow speed drops from 0.11 m/s to about 0.04 m/s. This the weld pool geometry as well as the best weld quality are most likely
indicates that the heat transfer is also weakened at the upper part of the accomplished below 150 A.
weld pool. Therefore, the top surface width is narrowed. The detailed
geometry is shown in Fig. 11 (b). At the lower part, the fusion zones 5. Conclusions
almost overlap because theoretically, the conductive and electro-
magnetic effects have no correlation with the chemical composition of In this paper, a thermal LB model with an effective heat source was
the material. At the hump area (see insets in figure), higher sulfur built to describe the unsteady transport phenomena that accompany
concentration leads to a flattering solidus line, which is closely related melting in the welding pool of PAW. The model was verified via ex-
to the slower fluid flow and the weaker heat transfer at the upper part periment and further compared to FVM simulation results. The evolu-
of the weld pool. We illustrated the velocity profiles under different tion of temperature field, velocity field, and weld bead geometry have
sulfur concentrations in Fig. 12. This shows that as the sulfur con- been presented, and the related mechanisms were further analyzed. A
centration increases, while the fluid velocity in the upper molten pool discussion has been provided on the effects of sulfur concentration in
decreases significantly, resulting in weaker heat transfer. Fig. 13 shows the specimen and welding current. The main conclusions are summar-
the heat flux adjacent to the solidus lines under different sulfur con- ized as follows.
centrations. Beyond 2.2 mm on the x-axis, the heat flux strongly de-
pends on the fluid flow speed. It is remarkable that the flow speed (1) The predicted weld beads were in good agreement with the mea-
below 0.4% sulfur concentration is much higher than the others, sured data and the average relative errors were below 12%.
combined with a larger heat flux. Especially at the peak around 4 mm, Compared to FVM, the computational time-consumption of LBM
the heat flux is twice that of the case under 2.8% concentration. could be decreased without compromising numerical accuracy.
However, the slump is also distinct at around 2.6 mm, which is why the (2) The outline of the weld bead was shaped as a “reversed bugle” and
hump sharpens as the sulfur concentration decreases in Fig. 11 (b). the maximal velocity of the fluid flow reached a magnitude of
Basically, the weld bead is considered to be high-quality if its depth- 0.1 m/s. The weld pool appeared at 0.35 s and the complete pe-
to-width ratio is large [2]. Accordingly, the Marangoni shear, which netration through the workpiece cost 0.71 s. The width of fusion
immensely widens the geometry of the weld pool, damages the joining increased at a stable rate and the depth grew with an “increasing
of workpieces. However, it is interesting to note that when the work- sharply - turning slow - increasing again” pattern.
piece is just penetrated, the top surface width seems to be invariable, as (3) The fluid flow played an important role in the weld process at the
listed in Table 5. From this, it can be inferred that at the outset of the upper part of the weld pool. Benefiting from the mesoscopic kinetic
welding, the fluid flow is still and then the heat transfer has not been background of the LBM, two circulation flows were found in the
substantially promoted. Therefore, if the welding is stopped at the weld pool and the principle of weld bead formation could be ex-
moment of penetration, the Marangoni effect could be weakened and plained.
high-quality welding could be accomplished. In general, the penetra- (4) Low sulfur concentration in the material led to a wider top surface
tion mainly depends on the distance between torch and workpiece, width of the weld pool and a sharper hump on the solidus line.
plasma gas flow rate, welding current, and welding speed. Among these Within a reasonable range, the high welding current lead to the
parameters, the welding current directly affects the melt of the work- welding process being efficient and timesaving. However, con-
piece and is intuitively described in our heat source model. To improve sidering the weld quality, the optimum welding current among our
the weld performance, tests with 0.2% sulfur concentration under dif- tests was 150 A.
ferent welding currents (135–170 A) were conducted.
The energy consumption typically grows with an increasing welding Acknowledgement
current when the voltage is constant. However, in our tests, the total
energy cost drops from about 2.2 kJ to 1.5 kJ with an increase of This work was supported by the National Natural Science
current from 135 A to 170 A, as shown in Fig. 14 (a). This energy saving Foundation of China (No. 51422601).
is attributed to the shortened time cost from 0.75 s to 0.41 s. It can be
concluded that with high welding current, the welding process is ac- References
celerated and the energy efficiency is improved. However, the optimum
current still requires to be studied because excessive current harms both [1] C.S. Wu, Welding thermal processes and weld pool behaviors, CRC Press/China
workpiece and electrode [61]. In addition, depicted by the red dots in Machine Press, 2011.
[2] R. Sacks, Welding: principles and practices, glencoe division, McGraw-Hill, New
Fig. 14 (a), the top surface width varies irregularly, which is incon- York, 1981.
sistent with common principles. Generally, the geometry of the weld [3] S. Wang, J. Goldak, J. Zhou, S. Tchernov, D. Downey, Simulation on the thermal
pool should be widened when the welding power (i.e. the welding cycle of a welding process by space–time convection–diffusion finite element ana-
lysis, Int J Therm Sci 48 (2009) 936–947.
current in our tests) increases. However, three factors are possible to [4] Y. Hsu, B. Rubinsky, Two-dimensional heat transfer study on the keyhole plasma
counteract the effect of increasing current. First, the penetration is ac- arc welding process, Int J Heat Mass Transf 31 (1988) 1409–1421.
complished rapidly so that there is not enough time for the top surface [5] R.G. Keanini, B. Rubinsky, Three-dimensional simulation of the plasma arc welding
process, Int J Heat Mass Transf 36 (1993) 3283–3298.
to grow. Second, the ratio of transferred heat to the bottom relative to [6] H. Fan, R. Kovacevic, Keyhole formation and collapse in plasma arc welding, J Phys
the top continuously increases, corresponding to the sharper and D Appl Phys 32 (1999) 2902.
sharper keyhole effect of the practical PAW process. Third, the [7] T. Li, C. Wu, Y. Feng, L. Zheng, Modeling of the thermal fluid flow and keyhole

457
J. Cai et al. International Journal of Thermal Sciences 124 (2018) 447–458

shape in stationary plasma arc welding, Int J Heat Fluid Flow 34 (2012) 117–125. [35] A. Brent, V. Voller, K.t.J. Reid, Enthalpy-porosity technique for modeling convec-
[8] X. Jian, C. Wu, G. Zhang, J. Chen, A unified 3D model for an interaction mechanism tion-diffusion phase change: application to the melting of a pure metal, Numer Heat
of the plasma arc, weld pool and keyhole in plasma arc welding, J Phys D Appl Phys Transf Part A Appl 13 (1988) 297–318.
48 (2015) 465504. [36] D. Chatterjee, S. Chakraborty, A hybrid lattice Boltzmann model for solid–liquid
[9] X. Jian, C.S. Wu, Numerical analysis of the coupled arc–weld pool–keyhole beha- phase transition in presence of fluid flow, Phys Lett A 351 (2006) 359–367.
viors in stationary plasma arc welding, Int J Heat Mass Transf 84 (2015) 839–847. [37] D. Chatterjee, Lattice Boltzmann simulation of incompressible transport phenomena
[10] Y. Li, Y. Feng, Y. Li, X. Zhang, C. Wu, Plasma arc and weld pool coupled modeling of in macroscopic solidification processes, Numer Heat Transf Part B Fundam 58
transport phenomena in keyhole welding, Int J Heat Mass Transf 92 (2016) (2010) 55–72.
628–638. [38] D. Chatterjee, S. Chakraborty, An enthalpy-source based lattice Boltzmann model
[11] W.-l. Liu, S.-s. Hu, L. Ma, Numerical simulation of fluid flow field in plasma arc for conduction dominated phase change of pure substances, Int J Therm Sci 47
welding with 3-D static conical heat source, Trans China Weld Inst 6 (2006) 008. (2008) 552–559.
[12] C. Wu, Q. Hu, J. Gao, An adaptive heat source model for finite-element analysis of [39] M. Eshraghi, S.D. Felicelli, An implicit lattice Boltzmann model for heat conduction
keyhole plasma arc welding, Comput Mater Sci 46 (2009) 167–172. with phase change, Int J Heat Mass Transf 55 (2012) 2420–2428.
[13] Y. Li, Y.-H. Feng, X.-X. Zhang, C.-S. Wu, An improved simulation of heat transfer [40] R. Huang, H. Wu, P. Cheng, A new lattice Boltzmann model for solid–liquid phase
and fluid flow in plasma arc welding with modified heat source model, Int J Therm change, Int J Heat Mass Transf 59 (2013) 295–301.
Sci 64 (2013) 93–104. [41] R. Huang, H. Wu, Phase interface effects in the total enthalpy-based lattice
[14] N. Yadaiah, S. Bag, Development of egg-configuration heat source model in nu- Boltzmann model for solid–liquid phase change, J Comput Phys 294 (2015)
merical simulation of autogenous fusion welding process, Int J Therm Sci 86 (2014) 346–362.
125–138. [42] I. Ginzburg, K. Steiner, Lattice Boltzmann model for free-surface flow and its ap-
[15] A. Traidia, F. Roger, E. Guyot, Optimal parameters for pulsed gas tungsten arc plication to filling process in casting, J Comput Phys 185 (2003) 61–99.
welding in partially and fully penetrated weld pools, Int J Therm Sci 49 (2010) [43] J. Latt, G. Courbebaisse, B. Chopard, J.L. Falcone, Lattice Boltzmann modeling of
1197–1208. injection moulding process, Cellular automata, Springer, 2004, pp. 345–354.
[16] W. Tan, Y.C. Shin, Multi-scale modeling of solidification and microstructure de- [44] J.G. Zhou, Axisymmetric lattice Boltzmann method, Phys Rev E 78 (2008) 036701.
velopment in laser keyhole welding process for austenitic stainless steel, Comput [45] A.-K. Nehad, Enthalpy technique for solution of Stefan problems: application to the
Mater Sci 98 (2015) 446–458. keyhole plasma arc welding process involving moving heat source, Int Commun
[17] V. Pavelic, R. Tanbakuchi, O. Uyehara, P. Myers, Experimental and computed heat mass Transf 22 (1995) 779–790.
temperature histories in gas tungsten-arc welding of thin plates, Weld J 48 (1969) [46] C.-B. Jia, C.-S. Wu, Y.-M. Zhang, Sensing controlled pulse key-holing condition in
295. plasma arc welding, Trans Nonferrous Metals Soc china 19 (2009) 341–346.
[18] N. Tsai, T. Eagar, Distribution of the heat and current fluxes in gas tungsten arcs, [47] X. Shan, H. Chen, Lattice Boltzmann model for simulating flows with multiple
Metall Trans B 16 (1985) 841–846. phases and components, Phys Rev E 47 (1993) 1815.
[19] S. Chen, G.D. Doolen, Lattice Boltzmann method for fluid flows, Annu Rev fluid [48] P. Lallemand, L.-S. Luo, Theory of the lattice Boltzmann method: dispersion, dis-
Mech 30 (1998) 329–364. sipation, isotropy, Galilean invariance, and stability, Phys Rev E 61 (2000) 6546.
[20] Z. Guo, T.-S. Zhao, Explicit finite-difference lattice Boltzmann method for curvi- [49] Z. Guo, C. Zheng, B. Shi, Discrete lattice effects on the forcing term in the lattice
linear coordinates, Phys Rev E 67 (2003) 066709. Boltzmann method, Phys Rev E 65 (2002) 046308.
[21] S. Succi, The lattice Boltzmann equation: for fluid dynamics and beyond, Oxford [50] Z. Guo, C. Zheng, Analysis of lattice Boltzmann equation for microscale gas flows:
university press, 2001. relaxation times, boundary conditions and the Knudsen layer, Int J Comput Fluid
[22] G. De Fabritiis, A. Mancini, D. Mansutti, S. Succi, Mesoscopic models of liquid/solid Dyn 22 (2008) 465–473.
phase transitions, Int J Mod Phys C 9 (1998) 1405–1415. [51] Z.-L. Guo, C.-G. Zheng, B.-C. Shi, Non-equilibrium extrapolation method for velocity
[23] E. Attar, C. Körner, Lattice Boltzmann model for thermal free surface flows with and pressure boundary conditions in the lattice Boltzmann method, Chin Phys 11
liquid–solid phase transition, Int J Heat Fluid Flow 32 (2011) 156–163. (2002) 366.
[24] J.M. Fuentes, K. Johannes, F. Kuznik, M. Cosnier, J. Virgone, Melting with con- [52] G. Zhang, C.S. Wu, X. Liu, Single vision system for simultaneous observation of
vection and radiation in a participating phase change material, Appl Energy 109 keyhole and weld pool in plasma arc welding, J Mater Process Technol 215 (2015)
(2013) 454–461. 71–78.
[25] D. Gao, Z. Chen, L. Chen, A thermal lattice Boltzmann model for natural convection [53] Z. Liu, C.S. Wu, J. Gao, Vision-based observation of keyhole geometry in plasma arc
in porous media under local thermal non-equilibrium conditions, Int J Heat Mass welding, Int J Therm Sci 63 (2013) 38–45.
Transf 70 (2014) 979–989. [54] X. Wang, C. Wu, M. Chen, Numerical simulation of weld pool keyholing process in
[26] D. Gao, Z. Chen, Lattice Boltzmann simulation of natural convection dominated stationary plasma arc welding, Acta Metall Sin 46 (2010) 984–990.
melting in a rectangular cavity filled with porous media, Int J Therm Sci 50 (2011) [55] M. Tanaka, M. Ushio, J.J. Lowke, Time dependent numerical analysis of stationary
493–501. GTA welding process (physics, processes, instruments & measurements), Trans
[27] W. Miller, The lattice Boltzmann method: a new tool for numerical simulation of the JWRI 32 (2003) 259–263.
interaction of growth kinetics and melt flow, J Cryst growth 230 (2001) 263–269. [56] Y. Li, Y. Feng, X. Zhang, C. Wu, Energy propagation in plasma arc welding with
[28] W. Miller, S. Succi, D. Mansutti, Lattice Boltzmann model for anisotropic liquid- keyhole tracking, Energy 64 (2014) 1044–1056.
solid phase transition, Phys Rev Lett 86 (2001) 3578. [57] Z. Guo, D. Li, B. Wang, A novel concept for convective heat transfer enhancement,
[29] W. Miller, S. Succi, A lattice Boltzmann model for anisotropic crystal growth from Int J Heat Mass Transf 41 (1998) 2221–2225.
melt, J Stat Phys 107 (2002) 173–186. [58] Z.-Y. Guo, W.-Q. Tao, R. Shah, The field synergy (coordination) principle and its
[30] W. Miller, I. Rasin, F. Pimentel, Growth kinetics and melt convection, J Cryst applications in enhancing single phase convective heat transfer, Int J Heat Mass
growth 266 (2004) 283–288. Transf 48 (2005) 1797–1807.
[31] W. Miller, I. Rasin, S. Succi, Lattice Boltzmann phase-field modelling of binary-alloy [59] A. Kidess, S. Kenjereš, B.W. Righolt, C.R. Kleijn, Marangoni driven turbulence in
solidification, Phys A Stat Mech Appl 362 (2006) 78–83. high energy surface melting processes, Int J Therm Sci 104 (2016) 412–422.
[32] S. Chakraborty, D. Chatterjee, An enthalpy-based hybrid lattice-Boltzmann method [60] P. Sahoo, T. DebRoy, M. McNallan, Surface tension of binary metal—surface active
for modelling solid–liquid phase transition in the presence of convective transport, J solute systems under conditions relevant to welding metallurgy, Metall Trans B 19
Fluid Mech 592 (2007) 155–175. (1988) 483–491.
[33] W.-S. Jiaung, J.-R. Ho, C.-P. Kuo, Lattice Boltzmann method for the heat conduction [61] V.A. Kosovich, V.A. Polupan, V.S. Sedykh, V.V. Nikonov, A.V. Panin,
problem with phase change, Numer Heat Transf Part B Fundam 39 (2001) 167–187. Y.L. Yaroviniskii, Effect of design and material of non-consumable electrodes for arc
[34] D. Chatterjee, S. Chakraborty, An enthalpy-based lattice Boltzmann model for dif- welding on their durability, Weld Int 5 (1991) 885–887.
fusion dominated solid–liquid phase transformation, Phys Lett A 341 (2005) [62] A.A. Mohamad, Lattice Boltzmann method: fundamentals and engineering appli-
320–330. cations with computer codes, Springer Science & Business Media, 2011.

458

You might also like