Sie sind auf Seite 1von 8

E.

FilEDiAN
Westinghouse Electric Corp.,
Bettis Atomic Power Laboratory,
West Mifflin, Pa.

Thermomechanical Analysis i f tie Welding


Process Using tie Finite Element Method
Analytical models are developed for calculating temperatures, stresses and distortions
resulting from the welding process. The models are implemented in finite element
formulations
and applied to a longitudinal butt weld. Nonuniform
temperature
transients are shown to result in the characteristic transverse bending
distortions.
Residual stresses are greatest in the weld metal and heat-affected zones, while the accumulated plastic strain is maximum at the interface of tliese two zones on tlie underside of the weldment.

Introduction
The welding of metal structures is aimed at providing a means
of joining together a number of components in such a way as to
minimize the impairment of the properties of those components.
Considerable effort has been expended on developing effective
welding techniques for a large number of metals and alloys. Concurrent with these developments have been efforts aimed at
identifying the various problems that result from welding processes, determining the strength of welded joints and structures
subject to specified loading conditions, and establishing guidelines and criteria for most effective joint design. The information
accumulated in these areas over the years has been overwhelmingly experimental. Though attempts have been made to establish
empirical approaches toward understanding the complex behavior of materials due to welding, comparatively little effort
has been expended in developing and applying analytical models
to explain and predict the transient thermal response and the
transient and residual mechanical response. Survey papers t h a t
deal with many of the analysis methods that do exist have been
published by the Welding Research Council on heat flow due
to welding [l] 1 and welding-induced stresses and distortions [2].
Analytical treatments of temperature transients during welding
have most often been considered by assuming the effective
thermal energy supplied by the heat source to be deposited in
such a narrow band of material that it may be idealized as a
point, or a line source, depending on the geometry of the weld.
The early work done in this area was concerned with the quasistationary transient temperature solution (steady-state with
respect to a moving coordinate system) resulting from a line
source of heat traveling on a straight path at a constant speed

through an infinite plate. The classical line source solution,


which was determined independently by Boulton and LanceMartjn [3] and Rosenthal and Schmerber [4], has been used by
Tallp] in the development of a numerical model to calculate the
longitudinal stresses (stresses in the direction of the weld line)
at a section of the plate normal to the weld line. Plastic flow
near the weld line was included in the calculation, and mechanical
propertiesthough not thermal propertieswere taken to be
temperature-dependent. Masubuchi, Simmons, and Monroe [6]
used Tail's method of analysis to write a digital computer program to calculate temperatures and stresses in welded plates.
Because of computational difficulties encountered when attempting to include inelastic strains in an analysis in which rapid
temperature changes take place, little analytical work has been
carried out which goes significantly beyond Tail's one-dimensional
analysis of longitudinal stress. The complexity of performing
analytical studies of welding stresses and distortions for various
weldment configurations, when considering factors such as transient multidimensional temperature distributions, effects of
weldment fixturing and surrounding structure on heat sink and
constraint characteristics, and realistic models of inelastic material behavior in the vicinity of the weld line, among others,
suggests the use of numerical methods of analysisin particular,
the widely adaptable finite element methodfor predicting transient and residual stresses and distortions.

Though a determination of the complete thermomechanical


response due to welding would require a full three-dimensional,
inelastic analysis of mechanical behavior, accompanied by computation of the three-dimensional transient temperature distribution, computer run times would be very large and the
resulting costs perhaps prohibitive. I t is therefore reasonable at
this time to first develop, for weldments under quasi-stationary
thermal conditions, numerical techniques for planar analysis in
sections normal to the weld direction. Applications of these
techniques yield valuable data that have heretofore been unNumbers in brackets designate References at end of paper.
obtainable, both in the fusion and heat-affected zones, and in
Contributed by the Pressure Vessels and Piping Division and presented
the weldment itself.
at the Second National Congress on Pressure Vessels and Piping, San Francisco,
Calif., June 23-27, 1975, of THE AMERICAN SOCIETY OF MECHANICAL ENThe recent paper b y Hibbitt and Marcal [7] represents an
GINEERS. Manuscript Teeeived at ASME Headquarters, March 28, 1975.
initial step in the development of numerical analysis techniques
Paper No. 75-PVP 27.
206 / A U G U S T

1975

Copyright 1975 by ASME

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

that will simultaneously account for many welding parameters


iind factors that previously had been considered piecemeal or not
I " this work, the thermal input for the finite element
a t all.
stress analysis is developed by use of a finite element transient
thermal analysis. The thermal model formulated by Hibbitt and
Marcal is designed to simulate the gas-metal arc welding process,
in which consumable electrodes are employed. The numerical
thermal analysis is required to account for such phenomena as
temperature-dependent material properties, phase change,
through-the-thickness temperature variations, irregular weldment
geometries, nonuniform distribution of energy from the heat
source, and deposition of filler metal, among others. The line
source solution [3-6] clearly cannot be used to model any of these
phenomena.
The weld thermomechanical model described in the present
work has been developed generally along the lines of Hibbitt
and Marcal [7]; i.e., the finite element method is utilized to
model both the thermal and the mechanical behavior. Emphasis,
however, is placed on applying the analysis technique to gastimgsten arc welding, although suitable modification of the
thermal model will enable the method to be applied to other
welding processes, as well. Though the purpose here is to develop
a capability for predicting the weld-induced transient and residual
stresses and distortions, the nature and characteristics of the
thermal model by itself are obviously of great importance, since
the model can be used, for example, to predict cooling rates,
peak temperature distributions, changes in metallurgical structure, and effects of chill block placement and other fixturing, as
well as to provide input for the mechanical model. Since strain
rates induced by welding are expected to be of the same order
of magnitude as the free thermal expansion rates and since the
(hermoelastic coupling coefficient [8] is small, thermomechanical
coupling is neglected. On this basis, descriptions of first the
thermal model and then the procedure for calculating stresses
and distortions are presented. An application of the thermomechanical welding analysis method is then given for a simple
butt weld.

Thermal Model
Calculation of the transient temperature distribution is based
on the attainment of quasi-stationary conditions, which are developed when the welding heat source is moving at constant
speed on a regular path (i.e., a straight line in a planar weld,
or a circle in an axisymmetric weld), and end effects resulting
from either initiation or termination of the heat source are
neglected. The temperature distribution is then stationary with
respect to a moving coordinate system whose origin coincides
with the point of application of the heat source. Consider the
planar weld illustrated in Fig. 1. The temperature at any point
in the weldment is expressed functionally as:

/ _ - - WELD LINE

f /
/
/
/

-AAJSA-YAAAA^
-SECTION
ANALYZED

//

DIRECTION OF
ELECTRODE TRAVEL

//

//
I

mrf

150 mm

Flf. 1 Weldment configuration

the heat conduction equation. Two-dimensional thermal analysis


at the section x3 = 0, normal to the direction of welding, is
thus treated in the present work.
Perhaps the most critical input data required for welding
thermal analysis are the parameters necessary to describe the
heat input to the weldment from the arc. Not only the magnitude, but the distribution of heat input will influence the dimensions of the weld metal and heat-affected zones, the cooling rates,
and the peak temperature distribution, as well as the temperature
gradients necessary to calculate stresses and distortions. The
welding parameters required to formulate the heat flux boundary
conditions are the magnitude of heat input from the arc Q, the
distribution of the heat input, characterized by a length parameter f, and the weld speed v. Introducing the arc efficiency !j,
Q is found simply from the formula
Q = r]EI

(3)

where E and / are the arc voltage and current, respectively. The
heat deposited per linear inch of weld is merely Q/v. It must be
recognized that an a priori determination of the magnitude and
distribution of heat input cannot, in general, be made due to
T(xs, x%, Xz, t) = T(xi, xi, x3 vt),
(1) the lack of knowledge regarding energy transfer from the arc
to the workpiece. Investigations of the physics of the welding
where v is the welding speed. Thus, given the transient tempera- arc are required to shed light on this area.
ture distribution at any one section of the weldment defined,
Following the approach of Pavelic, et al. [9], the heat from the
say, by x3 = 0, the temperature at any other section is determined
welding arc is, at any given time, assumed to be deposited on the
by an appropriate shift of the time scale as follows:
surface of the weldment as a radially symmetric normal distribution function. Letting r be the distance from the center of the
T(xu xt, x3, t) = T(xu xi, 0, t - x3/i>).
(2)
heat source, which is coincident with the axis of the electrode,
The problem is, therefore, reduced to finding the two-dimenthe heat flux q is given by:
sional, unsteady temperature field at a section normal to the
weld line. A planar analysis may be used for this purpose when
(4)
q(r) = qoC-Cr2
the weld speed, relative to a characteristic diffusion rate for the
where q0 and C are constants determined by the magnitude and
material, is sufficiently high so that the amount of heat condistribution of the heat input. The heat input parameters Q
ducted ahead of the weld torch is very small relative to the total
and f are defined by:
heat input. In this case, the net heat flow across any infmitesimally thin slice of the weldment normal to the weld line is as(5)
sumed to be negligible relative to the heat being diffused within
Q '2w I
q(r) r dr
*/ o
is neglected in
the slice itself; t h a t is, the term
k(T)
dx3
(6)
dxz
q(f) = 0.05 g0.

Journal of Pressure Vessel Technology

A U G U S T 19 7 5 / 207

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

Equation (4) then becomes


9(r)

<*p[-3(r/f)].

(7)

It is easily shown that f defines the region in which 95 percent


of the heat flux is deposited. Let xi be the distance from the weld
line in the section x3 = 0, and t = 0 the time at which the center
of the heat source (electrode) passes over this section (Fig. 1).
The heat flux distribution on the surface of the weldment,
directed parallel to the electrode, is then given by;
q{xltt)

30
!!
= ,exp[-3(:ti/?)
]exp[-3(t></F) a ].
2

(8)

ST

The three welding parameters, Q, f, and v, are all embodied in


this formula.
Liquid-solid phase change and the associated latent heat are
included in the finite element weld thermal model by means of a
recently developed direct iteration method [10]. The inaccuracies
that may arise when the latent heat is treated numerically as
an apparent increase in the specific heat, as was done, for example, by Hibbitt and Marcal [7], are eliminated here by directly
accounting for the latent heat in each time step of the thermal
analysis. The energy balance then depends on the temperature
at the end of the step, thus necessitating an iterative procedure.
The method is applicable to alloy systems, in which melting and
freezing occur over a range of temperatures bounded by the
liquidus and solidus.
Strictly speaking, the finite element method used here cannot
be applied to materials in which the solid-to-liquid phase change
takes place at a unique temperature level, as is the case with
pure metals. This arises because of the discontinuity of heat
flow across the liquid/solid interface. Since, for the presently
employed finite element method, the temperature is the dependent variable, continuity of temperature gradient within, but not
on the boundary of, each finite element is guaranteed. Thus,
the melting isotherm, which is a discontinuity surface, cannot
penetrate the interior of an element. In the direct iteration
method, this difficulty manifests itself in a noncoixvergent temperature solution. However, in any engineering application,
impurities and alloying additions render this purely an academic
question. Thus, for the cases of interest, the heat flow remains
continuous through the phase transformation, although its
derivatives are, in fact, discontinuous across the liquidus and
solidus lines. The temperatures within and in the vicinity of the
phase change range can then be better approximated by employing a finer finite element mesh, which can better simulate the
changes in heat flow. In any event, based on the size of the
material's phase change range, sufficiently small time steps
assure convergence of the solution. The rapidity with which the
solution converges, furthermore, depends on the phase change
range, size of time steps, and fineness of mesh. A continual
narrowing of the phase transformation range leads to a situation
approaching that of a pure metal and, hence, of a discontinuity
surface.
The iterative phase change calculations are confined to that
portion of the weldment t h a t undergoes phase transformation.
This significantly reduces computer run times.
The finite element used for the thermal analysis is the twodimensional, biquadratic, isoparametric element [11]. Based on
the discussion of phase change, and as pointed out by Hibbitt
and Marcal [7], it may, in retrospect, have been more appropriate
to employ a finer mesh with lower-order elements to better
model the progression of the melting and solidification fronts.
In any event, these fronts can be traced quite easily since both
the solidus and liquidus surfaces are defined by isotherms, the
locations of which are found by interpolation of nodal point
temperatures.
The finite element formulation for transient thermal problems
may be developed by variational methods [12], Galerkin's
208 / A U G U S T 19 7 5

principle [7, 11], or suitable application of the principle of virtual


work for heat conduction problems (see, for example, [10]).
Regardless of the method used, a set of simultaneous equations,
most conveniently expressed in the concise matrix form, is obtained. For the phase change technique used in the present
work, these equations are suitably modified to account for the
latent heat effects [10]. The other nonlinearities inherent in
welding thermal analysis, namely those resulting from temperature-dependent material properties and surface heat losses by
radiation, are handled by specifying the temperature-dependent
material and boundary coefficients for a given time increment
to be those evaluated at the known temperatures at the beginning
of the increment. The coefficients are updated for every time
step and thus are made to be piecewise-constant functions of
temperature. Therefore, except for those regions undergoing
phase change, the analysis procedure is linear and noniterative
in each time step.
The development of the finite element matrix equations will
not be repeated here, except to note that the finite element discretization, in effect, reduces the transient heat conduction equation, which is a partial differential equation, to a set of simultaneous ordinary differential equations which contain time derivatives
of the nodal temperatures. The numerical scheme employed to
integrate these equations is based on the use of the CrankNicholson operator, as modified by Wilson and Nickell [12].
Use of this method implies that temperature at a node point
varies linearly between the beginning and the end of a time step.
Though the Crank-Nicholson integration scheme is both highly
accurate and stable, the resultant transient temperature solution
is often oscillatory to some degree. To substantially reduce these
oscillations, averages of the temperatures at the beginning and
end of each time step (i.e., the temperatures at the middle of the
time/step) are computed.

Formulation for Stress and Distortion


Material subject to the welding thermal cycle is postulated
to behave' mechanically as an initially isotropic, elastoplastic,
strain hardening continuum, such that a component of total
strain e,y, at an instant of time t, is given by [13]:
l 7 (n = e^Ht)
where

+ i-ifHt)

+ fijTa(t),

(9)

e.v^ft) = Aijkl(T)<ni(t)

(10)

*nPL(t) = f

(11)

i>jPL(t)dt

,-/*() = M ( r ) 5 i . .

(12)

are components of elastic, plastic and thermal strains, respectively. Ami and oti represent components of the isotropic, temperature-dependent tensor of elastic moduli and of stress, respectively, while inPL is the rate of plastic deformation, eTH is the
free linear thermal expansion, and Si,- is the Kronecker delta.
Though creep effects could be included in equation (11), ratedependency is neglected in the present treatment. Applying
incremental plasticity theory, the plastic strain rates are replaced
by plastic strain increments, such that the total plastic strain
is accumulated in a finite number of increments; that is,
N(t)

n-l

where the end of the Arth increment is associated with time t.


Therefore, with Aeijt'L(t) representing the change of the plastic
strain component in the increment bounded by ( At and (-,
equation (13) becomes
HjFL{t)

= inPL{t

- At) + AtaPL{t)

(I'D

pi

where e,y (( At) is the plastic strain accumulated in the first

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

jV-1 increments. Consider yielding to be governed by the Mises


criterion and plastic flow by the Prandtl-Reuss law. Then, as
for example derived by Hill [14], the increment of plastic strain
may be expressed in the form
3<3..

(15)

2er

where $,/ = a,/ 1/3 ov* 8,7 is a component of deviatoric stress,

* = (f S S j

is the generalized stress, and &ePL is an in-

crement of the generalized plastic strain. A conventional power


law relation is postulated to characterize the generalized plastic
strain increment, such t h a t
AePL(l)

= maximum {(a(t)/KY

-^PL(t

- At); 0)] ,

(16)

where K and n are material parameters obtained from the uniaxial stress-strain curve and, in general, are temperature-dependent. Defining the yield strength cr, as the stress level at which
the plastic strain in a uniaxial, isothermal test is 0.2 percent,
K is evaluated by
K(T)

= <ry(T) (0.002)-"".

(17)

The requirement that A~iPL be either positive or zero is the means


by which continued plastic loading, accompanied by isotropic
expansion of the yield surface, and elastic unloading from the
yield surface are accommodated in the formulation [13]. The
generalized plastic strain is accumulated by adding the incremental value obtained from equation (16) at each load step,
to the previously accumulated plastic strain.
During the welding process, material in close proximity to the
center line of the weld experiences loading to a plastic state of
stress prior to melting. The plastic deformation accumulated
during this time is completely relieved when melting takes place.
At temperatures below the melting point TM, the plastic strain is
reduced by annealing processes, whose effects are significant
above the so-called recrystallization temperature. It is not the
intent of the present work to dwell on the establishment of
precise formulae to describe the effects of recrystallization-induced relief of plastic deformation.
Instead, the following
empirical expression is used to approximate the reduction of
previously accumulated plastic strain at high temperatures:
tnPL(t

- At) - f(T)iiPL(t

- At) ,

The biquadratic, isoparametric element employed in the


thermal analysis is used in the stress and distortion analysis as
well, though the finite element mesh configurations need not be
the same for the two sets of calculations.

Material Properties
The weldment material properties employed are those of
Inconel Alloy 600, an alloy of nickel, chromium and iron [16],
Its thermomechanical behavior is known from room temperature
to between 1140K and 1370K. The temperature-dependent
properties required for the welding analysis are plotted in Fig. 2.

' ii

^--'Ti-

(18)

where f(T) = e<r-rm) i f T{t) < TM; f(T) = 0 if T(t) > TM.
The extent of the recrystallization range is embodied in a. Using
a value of a 0.02, for example, it is found that plastic strain
relief greater then 5 percent occurs at temperatures above the
level T = TM 150. This value of a has been used in the
present analysis.
The incremental plasticity relations used in the welding analysis are incorporated into a small strain, small displacement finite
element representation by linearizing the stress-strain equation
in each increment of loading and using the tangent stiffness
method to evaluate the material stiffness [15]. The linearized
element stiffness matrices are obtained by using a NewtonRaphson type of iteration procedure in each increment [15].
The nonlinear stress-strain law is linearized about some prescribed value of stress (equivalent to using the first two terms
of a Taylor series expansion about this value). Finite element
calculations are performed, new values of stress calculated, and
the process repeated until suitable convergence criteria are met.
Since the transient and residual stress and distortion patterns
that are generated during welding are based on the build-up of
irreversible deformations, the accuracy of the accumulated
plastic strain is of paramount importance. The convergence
criterion is thus based on the cumulative plastic strain, and is
set equal to 10 - 6 . In order to facilitate convergence, suitable
acceleration parameters are employed in the iterative process.
Based on the discussion of the thermal model, it has been

Journal of Pressure essel Technology

assumed that, for sufficiently high weld speeds, a two-dimensional


transient analysis of temperatures in a section normal to the
direction of welding is, under quasi-stationary conditions, adequate for calculation of the welding cycle. The results of the
temperature calculations are then used as input for an analysis
of stresses and distortions.
A completely thorough determination of the mechanical response due to the welding thermal cycle, as determined by the
quasi-stationary temperature analysis, would require a full threedimensional incremental plasticity analysis with, at the least,
recalculation of the stiffness coefficients at each time step. Computer run times would be very large and the cost, therefore, may
be prohibitive. As an initial step, then, the computational
technique is developed for analysis in sections normal to the weld
direction. This implies that all sections normal to the weld line
remain plane during the entire welding process. Though this
assumption is probably adequate in regions somewhat removed
from the weld puddle, it may not be valid in the neighborhood
of the molten metal. The approach employed at present therefore
prevents calculation of distortions ahead of the welding arc,
and may inhibit accurate computations of deformations in the
immediate vicinity of the weld puddle. It nevertheless enables
the essential features of the mechanical response during cooldown to be modelled, and the resultant residual stresses and
distortions to be calculated.

1-

S**^*"^

" ^-"-""""'"^
/

'

K-

/11

-1

^ / ^

^ ^ ^ ^

i*

-yS
PH&SE C H A N G E R

J
\Y

H*

I I"
250

500

750

IOOO

I250

S500

I750

Fig, Z Material properties

AUGUST

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

1 9 7 5 / 209

Solid lines are representative of known data, while dashed lines


depict assumed high temperature properties. The density of
Alloy 600 is 8430 kg/m 3 and its melting range is 1630 K - 1690 K
[16]. The latent heat of nickel, which is 309 k j / k g , is used for
phase change calculations.
The thermal properties are extrapolated linearly to the melting
temperature. The adequacy of this assumption is of course open
to question, and only the experimental determination of the
required high temperature data can serve to test its validity.
The characterization of heat transfer in the weld puddle itself
is an even more formidable problem. Until more is known about
the mechanisms of heat flow in the puddle, only estimates of
the molten metal thermal characteristics can be made by specification of appropriate values of conductivity and specific heat to
be used in the finite element analysis. T h e heat capacity of the
molten metal is assumed to be equal to its value at the melting
point. The thermal conductivity, on the other hand, has been
taken to be half t h a t of the solid material at the solidus temperature. This is based on the expected behavior of molten metal
that is not in motion. The motion of the weld puddle, however,
results in temperatures that do not greatly exceed the melting
temperature. The relatively low puddle conductivity used appears to have caused excessive temperatures to be computed,
as will be seen in the next section. The effective conductivity
above the melting point should realistically be somewhat greater
than t h a t of the solid material, in order to simulate the heat
transfer mechanisms in the puddle, and thus to lower the molten
weld metal temperatures to more realistic levels.
Since the strength and elastic modulus are negligible at the
melting point, both these properties are assumed to decrease
linearly from their respective values at the highest temperatures
at which data are available, to extremely low values at 1630 K,
The strain hardening characteristics of the material, characterized by the parameter n, are not known at very high temperatures. A value of n 11, however, yields an analytical stressstrain relationship in good agreement with test data at 590 K,
and is used to characterize hardening at all temperatures. Though
the uniaxial stiffness becomes negligible as the melting range is
approached, the material is relatively incompressible in the
molten state. In order to model this behavior and, at the same
time, to avoid possible ill-behavior of the solution process and
thus facilitate convergence, the elastic modulus approaches zero
and the Poisson's ratio one-half, in such a way t h a t the bulk
modulus is maintained at its room temperature value through
the entire high temperature range, including the phase transition
and the liquid state. Since compressibility has no effect on the
generalized plastic strain, no plastic strain is, as expected, accumulated in the weld puddle.

bends [17], consists of relaxing the plane strain requirement by


letting the longitudinal strain be a linear function of these added
degrees of freedom, which represent longitudinal extension, and
rotation about the xt and x2 axes of Fig. 1. For the present,
however, the constraints associated with plane strain are main,
tained.
The welding parameters chosen for this analysis are as follows:
heat input Q = 703 W, characteristic radius of heat flux distribn.
tion f = 5.08 mm, weld speed v = 2.12 mm/s. An accurate
determination of both the magnitude and distribution of heat
input from the welding arc is predicated upon knowledge of the
heat transfer mechanisms from the arc to the workpiece. These
values, however, were selected to illustrate the thermomechanical
response of a full-penetration weld with a weld metal crosssection t h a t varies in width through the weldment thickness.
The transient thermal analysis was performed using the finite
element mesh shown in Fig. 3(a). A total of 47 time steps was
used in the analysis from the time the section being analyzed
(x3 = 0) is initially subject to heating from the welding arc, to
complete cool-down. Temperature transients are plotted in Fig.
4, while the resulting weld metal zone, which is by definition
t h a t portion of the weldment in which the peak temperatures
exceed the liquidus temperature, is depicted in Fig. 5. Tliese
two figures demonstrate the variation through the thickness of
the thermal response, resulting from the thermal energy from
the welding arc being distributed over a finite portion of the

^ ^ H E A T FLUX

^ ^ - ^

X^^,

(a)

y i i

i i

~i

~i

(b)
Fig. 3 Finite element mesh: (a) Temperature analysis; (fa) Stress and
distortion analysis

Analysis of a Butt Weld


Consider a 2.54-mm thick flat weldment whose configuration
is shown in Fig. 1. Since the temperature distribution is symmetric about the weld line xi = 0, only half the weldment need
be modelled. The plane section normal to the direction of travel
of the electrode is assumed to be in a state of plane strain. Thus
motion is completely restrained in the weld, or longitudinal,
direction. To illustrate transverse bending distortions that are
characteristic of thin weldments of this type, the weldment is
completely free to expand and bend in the transverse direction.
Though the plane strain approach prevents characterization
of the longitudinal bow mode of distortion also common to these
types of welds, the method nevertheless enables one to determine
complete residual stress distributions, transverse distortions, and
estimates of damage accumulated as a result of the welding
process. I t should be noted, however, t h a t longitudinal extension
and bowing could be considered without resorting to a threedimensional analysis by introducing up to three additional degrees
of freedom to the system. This approach, called generalized
plane strain, and used by Marcal in his analysis of curved pipe
210 / AOJ G U S T

1975

I
J

-0.1

^
1

TOP SURFACE OF WELD


__ BOTTOM SURFACE OF WELD

0.1

0.2
TIME-MIN

0.3

0.4

0.5

f i g . 4 Temperature histories at various distances from the we IP


certterlins

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

WELD

Fig, 5 Weld metal zone


O

10

20

30

40

DISTANCE FROM WELD L I N E - m m


Fig. 7

WELD

Transient and residual longitudinal stress distributions

-ORIGINAL GEOMETRY

^-DISTORTED GEOMETRY
Fig. 6

T r a n s v e r s e d i s t o r t i o n of w e i d m e n t

heated surface of the weldment (as opposed to a line source of


heat, which yields no thickness variations whatsoever, employed
in previous investigations [3-6]). This localized temperature nonuniformity yields a variation of shrinkage through the thickness
during cool-down, thus causing the characteristic bending distortions, which are accompanied by a nonuniform accumulation
of plastic strain in the weld metal and heat-affected zones.
In the weld puddle, calculated peak temperatures are considerably higher than the melting temperature (Fig. 4). As
pointed out in the previous section, these apparently excessive
temperatures arise from the relatively low value of molten metal
conductivity used in the analysis. The cooling curves for weld
metal material passing through the liquid-solid phase transition
range are characteristic of solidifying alloys, in that the latent
heat liberated as a result of the phase transformation results in
a momentary cooling rate decrease during the freezing period.
The incremental elastoplastic stress and distortion analysis
was carried out using the finite element mesh shown in Fig. 3(6).
A total of 24 loading increments was used for the entire weld
cycle, from initial heatup to final cooldown. Proper modelling
of the bending distortion mode in the region in close proximity
to the weld centerline requires a relatively fine grid both in the
thickness and transverse directions. The final (residual) distorted shape of the weldment cross-section, evaluated when cooldown is complete and the temperature is uniform at its room
temperature value, is shown in Fig. 6. The maximum deflection
(at the weld centerline) is 0.066 mm.
The unconstrained motion of the weldment in the plane normal
to the weld line produces transient and residual stress patterns
that are dominated by the longitudinal stress. Longitudinal
stress distributions at a number of points in time, including the
residual stress distribution at final cool-down, are plotted in Fig.
7. Stress histories at various locations are plotted in Fig. 8. I t
is of interest to note some of the characteristics of the transient
stress distribution.
Upon initial heatup (prior to the time at which the center of
the electrode passes over the section x3 0, being analyzed),
the localization of severe temperature gradients in the immediate
vicinity of the weld line produces compressive yielding in this

Journal of Pressure essel Technology

region. As more energy is being supplied by the arc and temperatures increase, the yield strength quickly decreases until, at
the melting point, it is negligible. During the time period prior
to solidification of the weld metal, all material outside the puddle
is in compression with t h a t region immediately adjacent to t h e
molten zone in a plastic state of stress. Since temperatures in
this region are extremely nonuniform, the yield stresses vary
from zero at the melting point to about 260 M N / m 1 , which is
only slightly lower than the room temperature yield strength
(291 M N / m 2 ) . Plastic deformation in molten material has been
completely relieved.
Upon solidification, the fusion zone material yields initially
in compression at rather low stress levels. Upon further cooling,
yield strength is increased and unloading from the yield surface
proceeds elastically until the material yields in tension. In the
weld metal and heat-affected zones, unloading and reversed
yielding occur over a very narrow stress range. Hence, as cooldown proceeds, t h a t portion of the weldment that is in tension
longitudinally grows steadily until, at final cool-down, the
residual longitudinal stresses, which are appreciable only in the
region within about 50 mm of the weld centerline, are completely
tensile. This is due to the plane strain restrictions placed on the
analysis. Though relaxation of this constraint would yield
residual stresses in compression outside this region, it is important
to note that the tensile stresses in the highly stressed region of
interest (within about 25 mm of the centerline) are governed
primarily by plastic behavior and should change little.
The residual stress distribution is such that there is little
variation of stress through the thickness. The residual stresses
are essentially uniform at 390 M N / m 2 in the weld metal and heataffected zones and then decrease uniformly to the room temperature yield strength at a location 23 mm from the weld centerline.
In the elastic region beyond this location, the stresses drop off
quite rapidly. The residual stresses in the plastic region exceed
the room temperature yield strength because of material strain
hardening, which is such that the initial short-time compressive
yielding produces an expansion of the yield surface followed,
during cool-down, by yielding in tension at a stress level higher
than the mono tonic yield strength. This type of behavior is
consistent with the residual stresses measured for bead-on-plate
welds by Nagaraja Rao and Tall [18]. I t was noted in this study
that residual stresses in (or in the close vicinity of) the weld metal
zone are usually about 50 percent above the yield strength of
the base metal. The results presented here, therefore, appear to
be in good qualitative agreement with these test data.
Plastic deformations are confined to a localized region as

AUGUST

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

1 9 7 5 / 211

> ' ' '^H^^^iiUU


300

ZOO

STRESSES AT MfOSURFACE

SOLIDIFICATION,

100

MELTfNG^
.,-0 25-,
\

/ i f

/Jf

X/jr

//-"

\\ \

100

200

\\

1 f\
\ / /

~r

"i"3 72

V/

*,-U.43v

~~^

\.
/

y.1-31.75

\__^/
,

, , , , ,,i

, ,,|

Fig. 8 Longitudinal stress histories at various distances from the weld center!?!?

curves defining the farthest extent of the liquidus and solidus


isotherms (recall that the former defines the weld metal zone),
indicate that there is considerable variation of plastic strain
through the thickness of the weld in the fusion and heat-affected
zones. The maximum residual plastic strain exists on the underside of the weldment at the weld metal/heat-affected zone interface and reaches a level of 4.5 percent. Maximum damage can
thus be expected to occur in this vicinity. This agrees qualitatively with conclusions based on plastic strain distributions calculated by Hibbitt and Marcal [7],

Conclusions

IO
15
20
DISTANCE FROM WELD L I N E - m m

25

Fig. 8 Transient and residua! generalized plastic strain distributions

illustrated by Fig. 9, in which transient and residual generalized


plastic strain distributions are plotted. Consistent with the
plastic zone associated with the residual stress distribution,
plastic strains greater than 0.2 percent are generated within 23
mm of the weld centerline. The highest plastic strain levels
exist in the weld metal and heat-affected zones throughout the
entire welding cycle, with the exception of t h a t time period in
which the weld metal is molten and the plastic strain nonexistent.
Beyond the heat-affected zone the plastic strain decreases quite
rapidly. The plots of plastic strain against time at various locations in Fig. 10 illustrate the manner in which plastic deformation is accumulated in the weld. In the weld metal zone, plastic
strain is accumulated rapidly as the material yields in compression. At the melting point, the plastic strain, which has reached
a level of 2.3 percent, is completely relieved. Upon solidification,
the plastic strain very rapidly reaches a level in close proximity
to that existing just prior to melting. As cooling continues and
yielding in tension progresses, the plastic strain steadily increases to its final value of 4.3 percent. In the heat-affected zone,
plastic strain is accumulated during heat-up in much the same
way as in the weld metal zone, but not quite as rapidly. At the
very high temperature levels just below the solidus, the accumulated plastic strain is partially relieved. As the material cools,
it yields in tension and further plastic strain is accumulated. In
regions removed from the weld metal and heat-affected zones,
there is no relief of plastic strain and elastic withdrawal from a
plastic state of compressive stress is characterized by no accumulation of plastic strain until yielding in tension commences.
The generalized plastic strain, which characterizes permanent
deformation, may be used as an indicator of cumulative "damage" in the weld metal and heat-affected zones during the welding
process. Contours of the accumulated generalized plastic strain
at final cool-down are plotted in Fig. 11 and, superposed with
212 / A U G U S T

1975

A thermomechanical model of the welding process has been


developed using the finite element method of analysis. The
finite element approach has been shown to be a powerful tool
both for determining the welding thermal cycle and for evaluating
t h e stresses and distortions generated as a result of the temperature transients. The analysis procedures are applicable to planar
or axisymmetric welds under quasistationary conditions.
The method used for determining temperatures is featured by
a direct iteration procedure to accurately account for the latent
heat liberated during solidification of the weld. The finite element calculations enable, in particular, the effects of the heat
input distribution on the heat flow patterns through the thickness of the weldment to be determined. The short-time thermal
response, which yields the dimensions of the fusion and heataffected zones, thus greatly affects the resulting nonuniform
shrinkage in these zones.
The mechanical response is calculated by implementing the
equations of time-independent incremental plasticity with temperature-dependent material properties. The transient longitudinal stress history of compressive loading followed by load
reversal and a tensile residual stress state is produced. Peak
residual stresses 30 percent in excess of the room temperature
yield strength are predicted both in the fusion zone and in the
heat-affected zone. A significant portion of the weldment outside
these zones is in a plastic state of stress as well. Nonuniform
shrinkage in the vicinity of the weld centerline produces the
transverse bending distortions typical of these types of welds.
The plastic strains, or permanent deformations, accumulated
during welding are maximum at the interface of the weld metal
and heat-affected zones on the underside of the weldment. With
plastic strain as an indicator of "damage" generated during
welding, the analytical prediction that maximum damage occurs
in this region is consistent with expected damage in full-penetration welds. Thus, in addition to predicting residual stresses and
distortions due to welding, the analytical technique can potentially be used for damage assessments as well.
Though experimental studies of weld temperatures, residual
stresses and distortions have been made, none appears to describe the weld process and the material properties in sutficient
detail to adequately define the finite element thermal and me-

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

'

1 1 I 1 M|

PLASTIC STRAINS AT MIDSURFACE

SOLIDIFICATION-v

MELTING.

> , = 2 ? 9 mm

^ - - . , - 2 79
,=0 25-v

/ /

J/y ^<^^7

. , 3 ! 75-,

- ^ ^ U _ _ ^ ^ H

Fig. U Generalized plastic strain histories at various distances from th wId e t n M f Htte

FARTHEST EXTENT OF UOUIDUS (I), SOUDUS 12) CONTOURS


043

.042

.Q4* 04g

Fig. U

043

.041

.038 .036 .034 .032 .030

Contours of residual generaiized plastic strain

chanical models. It is clear that further assessments of both


models can be made only when numerical results are correlated
with experimental d a t a generated from test welds designed
specifically for this purpose. Nevertheless, the potential of
utilizing finite element, methods of analysis for determining the
effects of the welding parameters, weldment geometry, material
properties, heat sink and constraint conditions, etc. on the
thermomechanical response, is evident.
Development of the weld thermomechanical model has also
served to identify other areas of work t h a t would yield a more
thorough understanding of the welding process, in addition to
providing more reliable input data to the finite element analyses.
Among these are determinations of material properties at very
high temperatures, studies of heat transfer mechanisms in the
weld puddle, and investigations of the physics of the welding
arc to determine the means by which thermal energy is transferred from the arc to the workpiece.

Acknowledgment
The author acknowledges the contributions of Dr. C. M.
Friedrich in the development of the incremental plasticity
formulations employed for stress and distortion analysis.

References
1 Myers, P. S., Uyehara, 0 . A., and Borman, G. L., " F u n damentals of Heat Flow in Welding," Welding Research Council
Bulletin No. 123, July 1967.
2 Masubuchi, K., "Control of Distortion and Shrinkage in
Welding," Welding Research Council Bulletin No. 149, Apr. 1970.
3 Boulton, N . S., and Lance Martin, H. E., "Residual
Stresses in Arc Welded Plates," Proc. Inst. Mech. Engrs., Vol.
133, 1936, pp. 295-339.
4 Rosenthal, D., and Schmerber, R., "Thermal Study -of
Arc Welding," Welding Journal, Research Supplement, Vol. 17,
No. 4, Apr. 1938, pp. 2s-8s.

5 Tall, L., "Residual Stresses in Welded PlatesA Theoretical Study," Welding Journal, Research Supplement, Vol. 43,
No. 1, Jan. 1964, pp. 10s-23s.
6 Masubuchi, K., Simmons, F . B., and Monroe, R. E.,
"Analysis of Thermal Stresses and Metal Movement, During
Welding," Battelle Memorial Institute, RSIC-820, Redstone
Scientific Information Center, NASA-TM-X-61300, N68-37857,
July 1968.
7 Hibbitt, H. D., and Marcal, P. V., "A Numerical ThermoMechanical Model for the Welding and Subsequent Loading of a
Fabricated Structure," Computers and Structures, Vol. 3, No. 5,
Sept. 1973, pp. 1145-1174.
8 Boley, B. A., and Weiner, J. H., Theory of Thermal
Stresses, Wiley, New York, 1960, pp. 42-44.
9 Pavelic, V., et al., "Experimental and Computed Temperature Histories in Gas Tungsten-Arc Welding of Thin Plates,"
Welding Journal, Research Supplement, Vol. 48, No. 7, July 1969,
pp. 295s-305s.
10 Friedman, E., "A Direct Iteration Method for the Incorporation of Phase Change in Finite Element Heat Conduction
Programs," WAPD-TM-1133, Bettis Atomic Power Laboratory,
Mar. 1974.
11 Zienkiewicz, O. C , and Parekh, C. J., "Transient Field
Problems: Two-Dimensional and Three-Dimensional Analysis
by Isoparametric Finite Elements," Intl. J. Num.
Methods
Engr., Vol. 2, 1970, pp. 61-71.
12 Wilson, E. L., and Nickell, R. E "Application of the
Finite Element Method to Heat Conduction Analysis," Nuclear
Engr. and Design, Vol. 4, 1966, pp. 276-286.
13 Fung, Y. C., Foundations of Solid Mechanics, Prentice
Hall, Inc., Englewood Cliffs, N . J., 1965, pp. 131-152.
14 Hill, It., The Mathematical Theory of Plasticity, Oxford
University Press, London, 1967, pp. 38-39.
15 Zienkiewicz, O. C , The Finite Element Method in Engineering Science, McGraw-Hill, London, 1971, pp. 374-377.
16 "Engineering Properties of Inconel Alloy 600," Technical
Bulletin T-7, International Nickel Co., Huntington, W. Va,, 1964.
17 Marcal, P. V., "Elastic-Plastic Behavior of Pipe Bends
with In-Plane Bending," J. Strain Analysis, Vol. 2, No. 1, 1967,
pp. 84-90.
18 Nagaraja Rao, N. It., and Tall, L., "Residual Stresses m
Welded Plates," Welding Journal, Research Supplement, Vol. 40,
No. 10, October 1961, pp. 468s-480s.

Joyrnal of Pressure essel Technology


Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 05/10/2014 Terms of Use: http://asme.org/terms

AUGUST

1 9 7 5 / 213

Das könnte Ihnen auch gefallen