Anda di halaman 1dari 11

Finite Elements in Analysis and Design 93 (2015) 85–95

Contents lists available at ScienceDirect

Finite Elements in Analysis and Design


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

Review

A 2D finite element with through the thickness parabolic temperature


distribution for heat transfer simulations including welding
Darlesson Alves do Carmo n, Alfredo Rocha de Faria
Department of Mechanical Engineering, Instituto Tecnológico de Aeronáutica (ITA), Praça Marechal Eduardo Gomes, 50, São José dos Campos,
SP 12228-900, Brazil

art ic l e i nf o a b s t r a c t

Article history: The arc welding process involves thermal cycles that cause the appearance of undesirable residual
Received 2 May 2014 stresses. The determination of this thermal cycle is the first step to a thermomechanical analysis that
Received in revised form allows the numerical calculation of residual stresses. This study describes the formulation of a 2D finite
8 September 2014
element with through the thickness parabolic temperature distribution, including an element estabil-
Accepted 15 September 2014
ization procedure. The 2D element described in this paper can be used to perform thermal analysis more
economically than 3D elements, especially in plates, because the number of degrees of freedom through
Keywords: the thickness will always be three. A numerical model of a tungsten arc welding (GTAW) setup was made
Finite element based on published experimental results. Size and distribution of the heat source input, thermal
Welding simulation
properties dependent on temperature, surface heat losses by convection and latent heat during phase
Thermal analysis
change were considered. In parallel the same setup was modeled using ANSYS software with 3D
Numerical simulation
elements (SOLID70) to compare against 2D numerical results. The results obtained by 2D model, 3D
model and experimental data showed good agreement.
& 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
2. Problem formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
2.1. Element formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
2.2. Element stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
2.3. Transient and nonlinear problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
2.4. Phase change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
2.5. Input heat model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3. Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.1. Transient thermal analysis of an aluminum plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.2. Transient thermal analysis of a steel plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.3. Transient thermal analysis of arc welding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4. Results and discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

1. Introduction

Welding is a fabrication process widely used in industry, being


employed in various industries such as petrochemical, aerospace,
n
Corresponding author. Tel.: þ 55 12 39476411.
automotive, marine and others. The fusion welding provides the
E-mail addresses: darlessondac@iae.cta.br (D. Alves do Carmo), union between two parts by melting and subsequent solidification
arfaria@ita.br (A. Rocha de Faria). of the welding areas. During this process a small area of the part is

http://dx.doi.org/10.1016/j.finel.2014.09.005
0168-874X/& 2014 Elsevier B.V. All rights reserved.
86 D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95

subjected to very high temperatures, while the temperature in the This work presents a new 2D heat transfer element that has
rest of the parts remains low. This strong thermal gradient is parabolic through-the-thickness temperature distribution. The
responsible for the appearance of undesirable residual stresses and quadratic nature of the temperature distribution is motivated by
consequent deformations. findings of previous works that point to parabolic through-the-
The determination of the residual stresses of welding has been thickness stress distributions [11–13]. Assuming a linear relation
studied by several papers [1]. Rosenthal [2] was one of the pioneers in between stresses and temperature increment the parabolic tem-
the study of welded joints with the proposition of an analytical solution perature distribution is thereby justified.
for plates with small thicknesses and concentrated heat source. Ueda [3] It will be shown that the 2D element proposed may be prone to
provided the basis for analyses that consider temperature-dependent numerical illness. This is due to the different time scales of the
properties. The finite element method has proved to be a powerful tool heat transfer process. For small thicknesses heat conduction from
in numerical analysis to predict the temperature field, residual stresses the hot top plate surface to the cool bottom plate surface is almost
and deformations. Lindgren reviews the use of this tool in simulations instantaneous. On the other hand, in-plane heat conduction takes
of welding, including aspects to be considered in the model [4]. In the place at a different time scale since, usually, distances traveled by
majority of analyses a twofold procedure is adopted: a pure thermal the heat flow are at least one order of magnitude larger than heat
analysis followed by a mechanical analysis with the calculated tem- flowing along the transverse direction. A stabilization procedure is
peratures [5]. The finite element technique was used by Long [6] to therefore necessary to avoid the numerical illnesses mentioned.
investigate distortions and residual stresses induced in a MIG welding
process of thin plates. The FE results obtained by Long were compared
with experimental data and empirical predictions, showing reasonable 2. Problem formulation
agreement. Furthermore, the plate thickness and the welding speed
showed to have great influence on residual stresses and welding dist- The transient heat conduction equation in Cartesian coordi-
ortions. Deng and Murakawa [7] used different formulations of thermo- nates is given by
elastic–plastic FE methods to compare welding distortions in a low ∂T ∂q ∂qy ∂qz
ρc ¼  x  þQ ð1Þ
carbon steel thin plate butt-welded. Zhu and Chao [8] investigated the ∂t ∂x ∂y ∂z
effect of temperature-dependent material properties on residual stres-
where T is temperature, ρ is density, c is specific heat, qx, qy, qz are
ses, distortions and temperature field in the numerical simulations of
heat fluxes in the x, y and z directions, and Q is internal heat
welding. They conclude that the thermal conductivity has certain effect
generation per unit volume. The heat fluxes relate to the tempera-
on the distribution of transient temperature fields during welding. Bar-
ture gradient according to the Fourier law of heat conduction
roso et al. [9] studied several simplification hypotheses in numerical 8 9 2 38 9
kxx kxy kxz > ∂T=∂x >
welding models and compared the results with experimental data. They < qx >
> = < =
6 7
reported that, in most cases, the simplifying hypotheses lead to quite q ¼ qy ¼  4 kxy kyy kyz 5 ∂T=∂y ¼ k∇T ð2Þ
acceptable results. :q >
> ; : ∂T=∂z >
> ;
z k kxz k yz zz
Bergman and Oldenburg [10] proposed a thermal shell element
with in-plane linear variation of temperature and quadratic variation where kij are the Cartesian components of the conductivity tensor
along the thickness. This element is similar to the one formulated in k and the temperature gradient vector is ∇T¼ {∂T/∂x ∂T/∂y ∂T/∂z}T.
this work but the explicit formulation is used to numerically integrate The variational principle tied to Eq. (1) is derived multiplying it
the governing equations in time. Another major difference is that the by a virtual temperature variation δT. It should be clear that δT is
present element has a stabilization scheme which permits the use of arbitrary but it must satisfy all conditions of prescribed tempera-
relatively coarse meshes to successfully obtain accurate results in the ture, i.e., if T is known for a given point (x, y, z), then δT must be
presence of thermal shocks or in regions where high temperature zero at that particular point. Multiplication of Eq. (1) by δT and
gradients are known to exist. integration over the plate domain Ω leads to
Heat transfer finite element analyses of 3D media are necessary Z    
∂T ∂qx ∂qy ∂qz
if one intends to determine the through-the-thickness tempera- ρc δT þ þ þ δT Q δT dΩ ¼ 0 ð3Þ
Ω ∂t ∂x ∂y ∂z
ture distribution in welded structures. The downside of 3D models
is the inherent huge number of degrees of freedom and badly Straight application of the divergence theorem yields
2  
conditioned elements if the plate is moderately thin. Turning to 2D Z   Z   3
∂qx ∂qy ∂qz ∂ q δ T ∂ qy δT ∂ qz δT 5
elements almost never provides realistic results since the assump- þ þ δT dΩ ¼ 4 x
þ þ dΩ
tion of constant through-the-thickness temperature distribution Ω ∂x ∂y ∂z Ω ∂x ∂y ∂z
holds only in regions away from the weldment fillet or groove. Z  
∂δT ∂δ T ∂δT
However, the most critical regions in the structure are precisely  qx þ qy þ qz dΩ
∂x ∂y ∂z
those close to the weldment fillet or groove commonly denomi- Z Ω Z
 T
nated heat affected zone. In this zone the heat source is located on ¼ q T nδ T dΓ þ ∇δT kð∇T ÞdΩ ð4Þ
the top surface of the welded plate whereas the bottom surface is Γ Ω
often modeled to be in natural convection. where n is the unit normal vector pointing outwards the boundary
surface Γ of Ω. Notice that Eq. (2) has been used in the derivation
of Eq. (4). Substitution of Eq. (4) into (3) gives
Z Z Z Z
z ∂T  T
y ρc δ T dΩ þ ∇δT kð∇T ÞdΩ ¼ Q δT dΩ  qT nδT dΓ
Ω ∂t Ω Ω Γ
4
ð5Þ
3
Eq. (5) is recognized as the variational principle associated with
heat transfer problems. In the right-hand side two loading terms
x
appear: internal heat generation and boundary heat exchange.
1
There are basically three common forms of boundary heat exc-
2
hange: (i) prescribed heat flux where qTn is a known function, (ii)
Fig. 1. Quadrilateral element arbitrarily oriented in 3D space. heat convection where qTn ¼hc(T  T1), and (iii) radiative flux
D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95 87

where qTn ¼ hr(T  T1). The convective heat transfer coefficient hc Substitution of Eq. (11) into the second term in the left-hand
may depend on T and the radiative heat transfer coefficient hr is side of Eq. (5) yields
definitely a function of the temperature T. T1 is the environment 8 9T 2 3
temperature. Z Z > < δqa >=
∂N T =∂x ∂N T =∂y 0T
 T 6 7
∇δT kð∇T ÞdΩ ¼ δ qb 6 T
4 z∂N =∂x z∂N T =∂y NT 7 5
Ωe Ωe >
: δq > ; T T T
c z2 ∂N =∂x z2 ∂N =∂y 2zN
2.1. Element formulation
2 38 9
∂N=∂x z2 ∂N=∂x > q
Fig. 1 shows the element topology chosen. In order to be econ-
z∂N=∂x < a> =
6 7
k4 ∂N=∂y z∂N=∂y z2 ∂N=∂y 5 qb dΩ ð12Þ
omical in terms of degrees of freedom the element is 4-noded and has >
:q > ;
bilinear shape functions over its local plane xy. 0 N 2zN c

In the local transverse direction z the temperature distribution


Based on Eqs. (10) and (12), the element heat capacitance matrix
is assumed to be parabolic such that, for a given point within the
Me and element conductivity matrices Ke are given by
element, the temperature T is
2 T 3
Z N N zN T N z2 N T N
T ðx; y; z; t Þ ¼ T a ðx; y; t Þ þ zT b ðx; y; t Þ þ z2 T c ðx; y; t Þ ð6Þ
Me ¼ ρc 6 7
4 zN T N z2 N T N z3 N T N 5dΩ ð13Þ
where z is the local transverse coordinate measured along the Ωe
z2 N T N z3 N T N z4 N T N
thickness. Ta is the zero order term, zTb is the linear term and z2Tc
is the quadratic term. These three terms are in turn interpolated 2 3 2 3
Z ∂N T =∂x ∂N T =∂y 0T ∂N=∂x z∂N=∂x z2 ∂N=∂x
using the bilinear shape functions with mapping coordinates 6 7 6 7
Ke ¼ 6 z∂N T =∂x z∂N =∂y T
NT 7 k 4 ∂N=∂y z ∂N=∂y 5dΩ
2
ξ A [  1, þ 1] and η A [  1, þ 1] Ωe
4 5 z∂N=∂y
          z2 ∂N T =∂x z ∂N =∂y
2 T
2zN T 0 N 2zN
N 1 ξ; η ¼ 14 1  ξ 1  η ; N 2 ξ; η ¼ 14 1 þ ξ 1  η
ð14Þ
         
N 3 ξ; η ¼ 14 1 þ ξ 1 þ η ; N 4 ξ; η ¼ 14 1  ξ 1 þ η ð7Þ
Both Me and Ke are symmetric and 12  12. Moreover, notice
such that that the local coordinate z is explicitly present in their computa-
tions. One might be tempted to integrate analytically through the
T a ¼ T a1 N 1 þ T a2 N 2 þT a3 N 3 þ T a4 N4 thickness. However, the thermal properties c and k are likely to
T b ¼ T b1 N1 þ T b2 N 2 þ T b3 N 3 þ T b4 N 4 vary with temperature, particularly in welding problems where
elevated temperatures are common. Therefore the most indicated
T c ¼ T c1 N1 þ T c2 N2 þ T c3 N3 þ T c4 N4 ð8Þ method to compute Me and Ke is numerical integration along x, y
and z. The Gauss quadrature recipe applicable to the bilinear
where Ta1, …, Tc4 are nodal degrees of freedom. The units of Ta, Tb
isoparametric element requires 2 integration points along x and
and Tc are not the same since they are multiplied by 1, z and z2,
2 along y. In the transverse direction z the terms containing z4
respectively.
require at least 3 integration points.
Based on Eq. (8) element local vectors of degrees of freedom are
defined as: qa ¼ {Ta1 Ta2 Ta3 Ta4}T, qb ¼ {Tb1 Tb2 Tb3 Tb4}T and qc ¼{Tc1
Tc2 Tc3 Tc4}T. The variation counterparts are then δqa ¼{δTa1 δTa2 2.2. Element stabilization
δTa3 δTa4}T, δqb ¼{δTb1 δTb2 δTb3 δTb4}T and δqc ¼{δTc1 δTc2 δTc3
δTc4}T. Also, the matrix of in-plane shape functions is defined as In order to illustrate a basic flaw of element formulated using
N¼[N1 N2 N3 N4]. Using these definitions Eq. (6) and its variational parabolic temperature distribution through the thickness consider
counterpart can be written as the simple 1D steady state problem of a beam shown in Fig. 2. It
T ðx; y; z; t Þ ¼ Nðqa þ zqb þ z2 qc Þ has total length 2L, thickness h and both ends have Ta, Tb and Tc
prescribed.
δT ðx; y; z; t Þ ¼ Nðδqa þ zδqb þ z2 δqc Þ ð9Þ Considering the practical case of a diagonal conductivity tensor
k¼diag(kxx kyy kzz) and constant thermal properties the differen-
Two terms are in the left-hand side of Eq. (5). These can be tial equations that govern the problem are:
calculated for each element e and its domain Ωe. Using Eq. (9) the
2
first term is ∂2 T a h ∂2 T c
kxx þkxx ¼ 0; ð15aÞ
8 9T 2 T 38 9 ∂x 2 12 ∂x2
Z
∂T < δqa >
>Z = N N zN T N z2 N T N <> q_ a >
=
ρc δT dΩ ¼ ρc δqb 6 4 zN T N z2 N T N 3 T
z N N5
7 q_
b dΩ 2
h ∂2 T b
Ωe ∂t Ωe >: δq >; >
: q_ > ; kzz T b  kxx ¼ 0; ð15bÞ
c z2 N T N z3 N T N z4 N T N c
12 ∂x2
ð10Þ
2
1 ∂2 T a h ∂2 T c
The second term in the left-hand side of Eq. (5) is more kzz T c  kxx  kxx ¼0 ð15cÞ
4 ∂x 2 16 ∂x2
elaborate since it requires computation of ∇T and ∇δT. This is
achieved, again, using Eq. (9). Eqs. (15a) and (15c) couple Ta and Tc. However, decoupling is
8 9 2 38 9 simple and, in view of the homogeneous boundary conditions for
∂T=∂x >
>
< = ∂N=∂x z∂N=∂x z2 ∂N=∂x > < qa >
= Ta and Tc, leads to the solution Ta(x)¼ Tc(x) ¼0. The analytical
6 7
∇T ¼ ∂T=∂y ¼ 4 ∂N=∂y z∂N=∂y z2 ∂N=∂y 5 qb
>
: ∂T=∂z ;> >
:q > ;
0 N 2zN c z
8 9 2 38 9 Ta = 0 Ta = 0
< ∂δT=∂x >
> = ∂N=∂x z∂N=∂x z2 ∂N=∂x > < δ qa >
= Tb = T 0
x
Tb = T 0
6 7
∇δT ¼ ∂δT=∂y ¼ 4 ∂N=∂y z∂N=∂y z2 ∂N=∂y 5 δqb ð11Þ Tc = 0 Tc = 0
>
: ∂δT=∂z >
; >
: δq > ; 2L
0 N 2zN c
Fig. 2. Steady state 1D beam problem.
88 D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95

1.0 cosh ½sðxi þ le Þ


T b;i þ 1 ¼ T 0 ð20Þ
0.9 coshðsLÞ
Exact solution
0.8 The stencil presented in Eq. (19) is of the form
Numerical solution (no stabilization)
0.7 α1 T b;i  1 þ α2 T b;i þ α3 T b;i þ 1 ¼ 0 ð21Þ
0.6
where α1, α2, α3 are coefficients to be determined. Substitution of
0.5 Eq. (20) into (21) and grouping terms the following conditions
0.4 must be met:
Tb/T0

0.3 α1 ¼ α3 ; α2 ¼  2α1 coshðsle Þ ð22Þ


0.2 Hence, substitution of Eq. (22) into (21) yields
0.1
 T b;i  1 þ 2 coshðsle ÞT b;i  T b;i þ 1 ¼ 0 ð23Þ
0.0
which can be rewritten as
-0.1      
1 1 1 1 1 1
-0.2 T b;i  1  2 þ 2 coshðsle ÞT b;i 2  þ T b;i þ 1  2 ¼0 ð24Þ
6 r r 6 6 r
-0.3
-1.0 -0.5 0.0 0.5 1.0 Comparison between Eqs. (24) and (19) reveals the error
x/L incurred when the numerical FE solution is adopted: the coeffi-
Fig. 3. Deficiencies of the proposed element. cient in Tb,i is wrong. The question is now how much must be
subtracted from the FE stencil to exactly reproduce Eq. (19). This is
T b,i−1 T b,i+1 given by
T b,i    
1 1 1 1 1 1 1 1
þ 2  α ¼ coshðr Þ 2  ) α ¼ þ 2 þ coshðr Þ  2 ð25Þ
le le 3 r r 6 3 r 6 r

Fig. 4. Uniform 1D FE mesh. The value of r given in Eq. (18) is usually large since kxx and kzz
are equal and the relation le/h is very large. Thus one approxima-
solution for Tb is tion is α Ecosh(r)/6. Additionally, since r is large, it is acceptable
that cosh(r)/6 may be replaced by a large number.
coshðsxÞ 12kzz
T b ðxÞ ¼ T 0 ; s2 ¼ 2
ð16Þ In practice the procedure adopted for correction of the FE
coshðsLÞ h kxx equations is to subtract α ¼hlekzz cosh(r)/6 from the main diagonal
Fig. 3 shows a comparison between the exact solution given in related to the Tb degrees of freedom of the element conductivity
Eq. (16) and the numerical solution obtained with no stabilization. matrix Ke, cosh(r)/6 replaced by a large number (1000). In the case
It is obvious from Fig. 3 that wide oscillations occur. Increasing the of 2D problems the correction depends on the element area Ae and
number of elements in the FE mesh certainly alleviates the is given by α ¼hAekzz cosh(r)/6. Similar reasoning shows that, in
numerical problem but it cannot be completely eliminated by that general, the main diagonal related to the Tc degrees of freedom
subterfuge alone. Moreover, one of the most attractive features of must also be corrected using the same procedure just presented.
the FE method is its ability to approximate solutions with a
relatively small number of simple elements. Within this context, 2.3. Transient and nonlinear problems
mesh refinement enhances the results but tends to overburden the
numerical procedures. The heat transfer problem posed in Eq. (5) can be discretized in
The question to be answered now is why the proposed element space using the heat capacitance matrix Me and element con-
does not deliver accurate results. The answer is obtained by comparing ductivity matrices Ke given in Eqs. (13) and (14). Additionally,
the exact solution of the problem, Eq. (16), against the finite element element thermal load vectors fe may be derived using Eqs. (6)–(8).
solution. Considering a uniform FE mesh and two adjacent elements of When the global matrix equations are assembled they result in
length le as shown in Fig. 4 it is possible to write the stencil of the
M T_ þKT ¼ f ð26Þ
finite element scheme associated with line i:
where M, K and f are the global counterparts of Me, Ke and fe, and T
kxx h   kzz le
2
 T b;i  1 þ 2T b;i  T b;i þ 1 þ ðT b;i  1 þ 4T b;i þ T b;i þ 1 Þ ¼ 0 is the global vector of nodal temperatures.
12le 6 In steady state problems the temperature is assumed constant
ð17Þ leading to a problem in the form
After defining KT ¼ f ð27Þ
2
12kzz le Eq. (27) is nonlinear when K and f depend on temperature.
r2 ¼ 2
ð18Þ
kxx h Nonlinearity may also arise depending on the thermal loading and
boundary conditions. A simple recurring scheme to solve the non-
Eq. (17) can be recast as
linear Eq. (27) is the Picard method[14] given by
     
1 1 1 1 1 1
KðT n ÞT n þ 1 ¼ f
n
T b;i  1  2 þ 2T b;i þ 2 þ T b;i þ 1  2 ¼0 ð19Þ ð28Þ
6 r 3 r 6 r
where the superscript n represents the iteration number. The proce-
In order to compare the exact and the numerical solutions it is
dure progresses until the tolerance ε is achieved, computed as follows:
important to determine what should have been the FE scheme
stencil to precisely reproduce the exact solution given by Eq. (16). ðT n þ 1  T n ÞT ðT n þ 1  T n Þ
o ε2 ð29Þ
Assuming that node i has coordinate xi and using Eq. (16), ðT n þ 1 ÞT T n þ 1
cosh ½sðxi  le Þ coshðsxi Þ The Picard method is usually good enough to achieve conver-
T b;i  1 ¼ T 0 ; T b;i ¼ T 0 ;
coshðsLÞ coshðsLÞ gence in a few iterations. However, in highly nonlinear problems,
D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95 89

more elaborate schemes may be required that adopt the concept of Eq. (1) can be rewritten as
relaxation. In the problems tackled in this work the Picard method
dH ∂T ∂q ∂qy ∂qz
has proven to be good enough. ¼  x  þQ ð37Þ
dT ∂t ∂x ∂y ∂z
If T is time dependent some type of numerical time marching
scheme must be employed. The most basic time integration proce- And, Me is consequently given by
dure available is the forward Euler method. However, stability issues 2 T 3
Z N N zN T N z2 N T N
disavow the forward Euler method because it would require extre- dH 6 7
mely small time steps. A more general approach, called generalized Me ¼ 4 zN N z2 N N z3 N N 5dΩ
T T T
ð38Þ
Ωe dT T T T
midpoint method, consists in evaluating Eq. (26) in an intermediate z2 N N z3 N N z4 N N
time step, in between n and nþ1 such that
Latent heat effect causes a large increase in the heat capacity in
1 nþ1   the mushy zone, between the liquidus (Tl) and solidus (Ts) tem-
T_ ¼
a
ðT  T n Þ; T a ¼ 1  θ T n þ θT n þ 1 ð30Þ
Δt peratures of material, as shown in the scheme in Fig. 5.
The direct use of ceff on time integration process should be done
where θ is an arbitrary parameter with 0r θ r1. When Eq. (26) is
with caution due to its step-like behavior. A node that is changing
evaluated at Ta it results in
phase must be forced to attain a temperature value in the interval
1   ΔT. If a node does not pass through this interval the latent heat
MðT a ÞðT n þ 1  T n Þ þ θKðT a ÞT n þ 1 ¼ f ðT a Þ  1  θ KðT a ÞT n ð31Þ
Δt effect is lost for that node, resulting in an incorrect energy balance
where Δt¼tn þ 1  tn is the time step. Different values of θ lead to and consequently incorrect temperatures. This may cause numer-
different time marching schemes. When θ ¼0 the forward Euler ical oscillations, making difficult the solution convergence [14].
method is produced. When θ ¼1 the backward Euler method is There are several averaging techniques that overcome difficulties of
produced which is an implicit method. In general θ 40 produces the direct use of ceff. These averaging techniques are generally referred
implicit schemes but the only second order accurate is the Crank– to as enthalpy method. A simple backward difference approximation is
Nicolson obtained with θ ¼1/2. Substitution of θ ¼ 12 into (30) and an averaging scheme that gives good results and is computationally
(31) gives: fast [4,16,17]:

Tn þTnþ1 dH Hn  Hn  1
Ta ¼ ð32Þ ¼ cef f ¼ n ð39Þ
2 dT T Tn1
This was the method used in this work.
1 1 1
MðT a ÞðT n þ 1  T n Þ þ KðT a ÞT n þ 1 þ KðT a ÞT n ¼ f ðT a Þ ð33Þ
Δt 2 2
2.5. Input heat model
Considering that T n þ 1 ¼ 2T a  T n , the Crank–Nicolson scheme
applied to Eq. (31) yields
The welding arc can generally be modeled in two different ways:
 
2 2 by prescribing the temperature in the weld or by specifying the heat
MðT a Þ þ KðT a Þ T a ¼ f ðT a Þ þ MðT a ÞT n ð34Þ
Δt Δt input. In prescribed temperature techniques the temperature is given
in the nodes inside an area or volume that simulates the weld pool.
which is an implicit equation whose solution delivers Ta. The Picard The prescribed temperature in the specific nodes is increased linearly
method presented in Eq. (28) can be effectively used to solve Eq. up to the temperature considered necessary for melting the material,
(34). Once Ta is computed the algorithm advances in time through and is maintained at this value for a period of time [5]. This procedure
Tn þ 1 ¼ 2Ta  Tn. One important aspect of the Crank–Nicolson met- is repeated in volumes along the weld bead for simulating the torch
hod is that it is theoretically unconditionally stable, i.e., any Δt is movement.
guaranteed not to diverge, although temporal oscillations may The prescribed heat input method, which is the most commonly
occur if Δt is too big. used method, can be applied in two most known ways: by a surface
heat flux with gaussian distribution or a volumetric heat flux called
2.4. Phase change Goldak's double ellipsoid. Pavelic was the first to suggest a gaussian
distribution of heat flux applied on the surface of workpiece [1].
The phase change process in the weld pool needs to absorb or Friedman [18] and Krutz [19] extended this model to a moving heat
release thermal energy described by latent heat. There are several source model by adapting Pavelic's disc for use in finite element
methods to take into account the latent heat. These methods are applications (Fig. 6). It is computed as
usually divided into fixed and moving mesh methods. In the moving 3Q
e  3½y þ vðτ  tÞ
2
e  3x
2 =c2 =c2
mesh method the phase change interface is determined as a moving qðx; y; t Þ ¼ ð40Þ
π c2
boundary and solid and liquid region are analyzed separately. The
fixed mesh methods account for the phase change condition implici
tly, involving a solution of a continuous system, without attempting to
establish the position of the interface. The moving mesh method req-
uires a sophisticated front tracking algorithm, while with a simpler
approach is possible to use a fixed grid [15,16].
The enthalpy is defined by
Z T
H¼ ρðT ÞcðT ÞdT ð35Þ
Tr

And
dH
¼ cef f ¼ ρðT ÞcðT Þ ð36Þ
dT
where ceff is the effective heat capacity. Fig. 5. Variation of ceff and enthalpy with temperature.
90 D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95

where x and y are system axes as shown in Fig. 6, v is the welding The prescribed temperature technique is the simpler way to
speed, c is the characteristic radius of heat flux distribution and τ is model the weld arc, but it can be less accurate than the heat flux
a lag factor that defines the position of the source at time t¼ 0. Q is model. Goldak's double ellipsoid is the most widely used model
the power from the arc and, among the models presented, is what best describes the heat
Q ¼ ηEI ð41Þ source. In this work the heat flux surface model with gaussian
distribution was used, because it provides reasonably accurate
where η is the efficiency of the welding source, E is the arc voltage results and is easier to implement than Goldak's model.
and I is the welding current.
In order to consider the digging action of the arc, and also the
high rate of heat transportation in the weld pool due to convection,
3. Models
Goldak proposed the double ellipsoid model, that distributes, in
gaussian way, a power density (W/m3) in two combined ellipsoidal
3.1. Transient thermal analysis of an aluminum plate
shapes [20]. The heat source is composed of a front half formed with
a quadrant of an ellipsoid, and a rear half composed by a quadrant of
An application example of the developed element is shown in
another ellipsoid (Fig. 7). The power density distribution inside the
Fig. 8. An aluminum plate of 1 mm thick is initially subjected in
front half is given by Eqs. (42) and (43) describes the power density
one of the edges to a linear temperature distribution through the
in the rear half.
pffiffiffi thickness such that Ta ¼1000, Tb ¼ 2  106 and Tc ¼0. The other
6 3f f Q  3x2 =a2  3y2 =b2  3½z þ vðτ  tÞ2 =c2 edges of the plate are kept at 0 1C. This heating condition is maint-
qðx; y; z; t Þ ¼ pffiffiffiffi e e e 1 ð42Þ
abc1 π π ained for 5 s, and then the temperature imposition is removed
from the hot side, causing the cooling of the plate. In this example
pffiffiffi
6 3f r Q  3x2 =a2  3y2 =b2  3½z þ vðτ  tÞ2 =c2 the thermal properties were considered constant, making the pro-
qðx; y; z; t Þ ¼ pffiffiffiffi e e e ð43Þ
blem linear: c ¼900 J/kg 1C, ρ ¼2700 kg/m3 and k ¼237 W/m 1C.
2
abc2 π π
The initial plate temperature is 0 1C and the convection coefficient
where ff and fr are the fractions of heat deposited in the front and
h is zero on all surfaces. The mesh used in the spatial discretization
rear quadrants respectively, and must satisfy ff þfr ¼2. The ellipsoid
was 20  20 elements.
parameters a, b, c1 and c2 correspond to dimensions of the weld pool
In this example the upper and lower surfaces are isolated, so
and can be obtained from experimental data. In absence of these data
there is rapid temperature equalization in thickness. The tempera-
it is possible to approximate the distance in front of heat source, c1,
ture histories at line located at y ¼250 mm on points x¼ 425 mm,
equal to one half weld width a and the distance behind source, c2,
450 mm and 475 mm is shown in Fig. 9. At these points there is no
equal the double of weld width [20].

Fig. 8. Aluminum plate model.

Fig. 6. Surface heat flux model with Gaussian distribution.


580
x=425mm

x=450mm
480
x=475mm

380
temperature (ºC)

280

180

80

-20
0 50 100 150 200
time (s)

Fig. 7. Goldak's double ellipsoidal heat source model. Fig. 9. Temperature history at x ¼475 mm, 450 mm and 425 mm.
D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95 91

difference between top and bottom temperatures because even for top surface (z¼ 0.001 m) and in a point between these two planes
small time steps of the order 0.01 s, Tb decreases instantaneously (z ¼0.0005 m). In this graph it is possible to observe a temperature
to zero. There is a high temperature gradient near the hot edge, difference up to 4 1C. If we consider the symmetry of the problem,
and from the graph it is observed that the temperature evolution the difference between the upper and lower surfaces may reach
in time becomes smoother as the point under analysis is far from 8 1C. A rapid equalization of temperature through thickness is also
the hot side. In the graph it is also possible to observe the natural noted when the heat input is removed.
tendency of stabilization and equalization of temperature until the The graph of Fig. 12 shows the temperature history for the
entire plate is at 0 1C. thickness located at x ¼0.001 m and y¼0.05 m. In this region there
is no significant variation of temperature through thickness.

3.2. Transient thermal analysis of a steel plate


3.3. Transient thermal analysis of arc welding
In this application example it is possible to observe the thro-
ugh-thickness temperature variation. A steel plate with 0.1  0.1  A computer code was developed with the new 2D element
0.002 m3 dimension, receives in the midline of the left side a thermal formulation for solving a transient nonlinear problem. The model
load of 50 W equally distributed. The plate is heated for 5 s and then is was based on Depradeux [21] in order to compare the results of the
allowed to cool. The initial temperature as well as ambient tempera- new element with experimental results. In addition, a 3D model
ture is 0 1C, with h¼40 W/m2 1C applied to all surfaces, except the left created in ANSYS was used to evaluate the through thickness temp-
surface. The constant thermal properties considered were c¼500 J/ erature in comparison with the new element model. A plate of stai-
kg 1C, ρ ¼7800 kg/m3 and k¼40 W/m 1C. Elements near the hot edge nless steel AISI 316L with 0.250  0.160  0.010 m3 is the part under
have size 0.001  0.001 m2, which gradually increases toward the cold analysis (Fig. 13).
side. The model mesh is illustrated by Fig. 10. Considering the enthalpy definition in Eq. (35) the tempera-
The temperature history for the thickness located at x ¼0 m ture–enthalpy curve (Fig. 14) was obtained with data of ρ(T) and c
and y ¼0.05 m is shown in Fig. 11. This graph contains tempera- (T) used in Depradeux. As in Depradeux, the curve was linearly
tures in three different points in the thickness: midplane (z ¼0 m), extrapolated up to the Tsolidus temperature and it is considered that
above Tliquidus the curve varies linearly with the same slope in the
region before the Tsolidus.
The temperature-dependent thermal conductivity is shown in
Fig. 15. For temperature above the melting point the value of
thermal conductivity is doubled to consider fluid flow in weld pool.
The convection heat transfer coefficient considered is 10 W/m2 1C
and the ambient temperature 29 1C. The radiative heat transfer was
not accounted in this article. The consistent nodal loads resulting
from the welding torch were accounted by gauss quadrature of the
right term of Eq. (5), where the input heat function is described by
Eq. (40) applied on top surfaces of elements located in the weld
region. The welding parameters are E¼10 V, I¼150 A, η ¼0.68,
v¼0.001 m/s. Based in the weld pool obtained by Depradeux the
parameter c of Eq. (40) is equal to 6 mm. The welding torch moves
through time steps of 1 s in both 2D and 3D models, starts at
y¼ 10 mm and stops at y¼ 240 mm.
The mesh used in the 2D model is shown in Fig. 16. In the weld
region the elements have 0.001  0.001 m2 dimension and gradu-
ally increase to 0.01  0.001 m2. Only one half of the plate was
modeled because of symmetry.
The 3D model has SOLID70 elements with 0.001  0.001 
Fig. 10. Model mesh of steel plate. 0.001 m3 dimension in the weld region and 0.01  0.001  0.001 m3

45

50 40
z=0m z=0m
z=0.0005m 35 z=0.0005m
40 z=0.001m
Temperature (ºC)

z=0.001m
Temperature (ºC)

30

30 25

20
20
15

10
10
5

0 0
0 2 4 6 8 10 0 2 4 6 8 10
time (s) time (s)
Fig. 11. Temperature history at x ¼0 m and y ¼0.05 m of steel plate. Fig. 12. Temperature history at x¼ 0.001 m and y¼0.05 m of steel plate.
92 D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95

Fig. 13. Plate and measurement points: (a) plate under analysis (dimensions in mm), (b) measured points in top surface, and (c) measured points in bottom surface.

12 4. Results and discussion


Enthalpy (109 J/m3)

10
The 2D model mesh generated 6777 nodes, totaling 20,331
8 degrees of freedom (dof), while in the 3D model were utilized
6 74,547 nodes or dof. The temperature field calculated by the new
2D element at 85 s and 230 s are shown in Fig. 18.
4 The section located 95 mm from the origin of the y-axis is
2 shown in Fig. 19 at time 85 s. In this time the torch is located on
this section, so that the section reaches its peak temperature.
0 As can be seen in Figs. 18 and 19, there is a high gradient of
0 500 1000 1500 2000
Temperature (°C)
temperature in the weld zone; furthermore, the through-thickness
gradient becomes insignificant when the distance to the weld line is
Fig. 14. Enthalpy–temperature curve. 10 mm or bigger.
Figs. 20 and 21 show the numerical results obtained with 2D code,
3D Ansys model and experimental setup (Depradeux). It shows the
60
Conductivity (W/m/°C)

temperature history on specific points at top and bottom surfaces of


50 the section located at y¼95 mm.
40 It can be seen in Figs. 20 and 21 that there is a good agreement
between the numerical results and the measured data. The results
30
obtained from the 2D code were very close to those obtained with
20 the 3D Ansys model. The greatest variations between the models
10
occurs in the points near weld line, where there is a high gradient of
temperature. For points away from the weld line, where the
0 temperature gradient is smaller, the numerical and experimental
0 500 1000 1500
results have better agreement. This can be explained by the model of
Temperature (°C)
the heat source, which has a strong influence on the gradient of the
Fig. 15. Conductivity–temperature curve. weld zone. The calibration of the source model can approximate the
numerical and experimental results in the region of high gradient.
dimension in the cool edge. Fig. 17 shows the 3D ANSYS mesh. The Fig. 22 shows the through thickness temperature located at
latent heat was considered in the ANSYS software by inputting directly y¼95 mm and x¼0 over time. From this chart, it is possible to
enthalpy–temperature curve points. compare the evolution of the through thickness temperature in
D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95 93

Fig. 17. 3D ANSYS model mesh.

Fig. 16. 2D model mesh.

between the 2D and 3D models. As can be seen in Fig. 22, the major
difference between temperatures occurs at the time 85 s, at which
the torch is located on the line in question, creating the largest
temperature gradient. Before and also after the passing of the torch,
the difference between the curves is small, which is an evidence that
the parabolic approximation of temperature on thickness adopted in
the development of the 2D model is reasonable.

5. Conclusions

A formulation of a 2D heat transfer element with parabolic Fig. 18. Temperature field at 85 s and 230 s: (a) top surface at 85 s, (b) bottom
through-the-thickness temperature distribution was carried out. A surface at 85 s, (c) top surface at 230 s, and (d) bottom surface at 230 s.
94 D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95

Fig. 19. Section 95 mm at 85 s calculated by 2D element model.

700
Measured 10mm
Measured 20mm
600 Measured 50mm
Ansys 10mm
Ansys 20mm
500
Ansys 50mm
2D 10mm
Temperature °C

2D 20mm
400
2D 50mm

300

200

100

0
0 50 100 150 200 250 300 350 400
time s
Fig. 20. Experimental, Ansys and 2D code results for y¼ 95 mm section at 10, 20 and 50 mm x points on top surface.

900
Mesured 0mm
Mesured 8mm
800
Mesured 20mm
Mesured 35mm
700 Ansys 0mm
Ansys 8mm
600 Ansys 20mm
Temperature °C

Ansys 35mm
2D 0mm
500
2D 8mm
2D 20mm
400
2D 35mm

300

200

100

0
0 50 100 150 200 250 300 350 400
time s
Fig. 21. Experimental, Ansys and 2D code results for y ¼95 mm section at 0, 8, 20 and 35 mm x points on bottom surface.

3D model of ANSYS thermal analysis was developed in parallel to GTAW process was performed. Nonlinear properties of the steel and
compare the numerical results, focusing on the temperature distribu- heat loss by convection were considered. Furthermore, experimental
tion in the thickness. A numerical transient heat transfer analysis of a results obtained from the literature were used to validate the model.
D. Alves do Carmo, A. Rocha de Faria / Finite Elements in Analysis and Design 93 (2015) 85–95 95

Fig. 22. Through thickness temperature at y¼ 95 mm and x¼ 0 for 2D and 3D models.

The 2D model with parabolic temperature variation in thickness [8] X.K. Zhu, Y.J. Chao, Effects of temperature-dependent material properties on
showed a good correlation of results with the 3D model and the exp- welding simulation, Comput. Struct. 80 (2002) 967–976.
[9] A. Barroso, J. Cañas, R. Picón, F. París, C. Méndez, I. Unanue, Prediction of
erimental data. welding residual stresses and displacements by simplified models, Exp. Valid.
Thermal analysis on plates where the through thickness tempera- Mater. Des. 31 (2010) 1338–1349.
ture is important can be made more economically when using the 2D [10] G. Bergman, M. Oldenburg, A finite element model for thermomechanical
analysis of sheet metal forming, Int. J. Numer. Methods Eng. 59 (2004)
element described in this work. The plate discretization using 2D
1167–1186.
elements (12 degrees of freedom each) needs only one element in the [11] J.A. Free, R.F.D. Porter Goff, Predicting residual stresses in multi-pass weld-
thickness, while in the plate discretization with 3D elements (8 degrees ments with the finite element method, Comput. Struct. 32 (1989) 365–378.
of freedom each) is required more elements in the thickness direction. [12] Y.V. Murthy, G.V. Rao, P.K. Yver, Numerical simulation of welding and
quenching processes using transient thermal and thermo-elastic–plastic
formulations, Comput. Struct. 60 (1996) 131–154.
References [13] J.K. Hong, C.L. Tsai, P. Dong, Assessment of numerical procedures for residual
stress analysis of multipass welds, Weld. J. 77 (1998) 372–382.
[14] J.N. Reddy, D.K. Gartling, The Finite Element Method in Heat Transfer and Fluid
[1] J.A. Goldak, M. Akhlaghi, SpringerLink (Online service), Computational Dynamics, CRC Press, Boca Raton, FL (2001) 92–111.
Welding Mechanics, Springer Science þ Business Media, Inc., Boston, MA, [15] R.W. Lewis, P.M. Roberts, Finite element simulation of solidification problems,
2005, vol. 321, pp. 1–11. Appl. Sci. Res. 44 (1987) 61–92.
[2] D. Rosenthal, The theory of moving sources of heat and its application to metal [16] R.W. Lewis, The Finite Element Method in Heat Transfer Analysis, Wiley,
treatment, Trans. ASME 68 (1946) 849–865. Chichester, England (1996) 123–161.
[3] Y. Ueda, T. Amakawa, Analysis of thermal elastic–plastic stress and strain [17] L.-E. Lindgren, H. Runnemalm, M.O. Nasstrom, Simulation of multipass
during welding by finite element method, Trans. Jpn. Weld. Soc. 2 (1971) welding of a thick plate, Int. J. Numer. Methods Eng. 44 (1999) 1301–1316.
186–196. [18] E. Friedman, Thermomechanical analysis of the welding process using the
[4] L.E. Lindgren, Numerical modelling of welding, Comput. Methods Appl. Mech. finite element method, J. Press. Vessel Technol. 97 (1975) 206.
Eng. 195 (2006) 6710–6736. [19] G.W. Krutz, L.J. Segerlind, Finite element analysis of welded structures, Weld.
[5] L.-E. Lindgren, Finite element modeling and simulation of welding part 1: J. Res. Supplement, 57 (1978) 211–216.
increased complexity, J. Therm. Stress. 24 (2001) 141–192. [20] J. Goldak, A. Chakravarti, M. Bibby, A new finite element model for welding
[6] H. Long, D. Gery, A. Carlier, P.G. Maropoulos, Prediction of welding distortion heat sources, Metall. Trans. B 15 (1984) 299–305.
in butt joint of thin plates, Mater. Des. 30 (2009) 4126–4135. [21] L. Depradeux, Simulation Numerique du Soudage – Acier 316L – Validation sur
[7] D. Deng, H. Murakawa, Prediction of welding distortion and residual stress in a Cas Tests de Complexite Croissante (Doctorate thesis), Institut National des
thin plate butt-welded joint, Comput. Mater. Sci. 43 (2008) 353–365. Sciences Appliquees de Lyon, 2004.

Anda mungkin juga menyukai