You are on page 1of 13

Available online at www.sciencedirect.

com

Acta Materialia 56 (2008) 2843–2855


www.elsevier.com/locate/actamat

Comparative analysis of heat generation in friction welding of steel bars


M. Maalekian a,*, E. Kozeschnik a, H.P. Brantner b, H. Cerjak a
a
Institute for Materials Science, Welding and Forming, Graz University of Technology, Kopernikusgasse 24, 8010 Graz, Austria
b
Voestalpine Schienen GmbH, Technology/Research and Development, Kerpelystrasse 199, 8700 Leoben, Austria

Received 28 September 2007; received in revised form 5 February 2008; accepted 9 February 2008
Available online 14 March 2008

Abstract

For the first time a comparative thermal analysis of the friction welding process using various heat generation models is presented.
The heat-generation rate in orbital friction welding of steel bars is analyzed using four different methods; constant Coulomb friction,
sliding–sticking friction, the experimentally measured power data and an inverse heat conduction approach. A comparison between
the calculated temperature profiles and the experimental data shows that the inverse heat conduction approach predicts the heat-gener-
ation rate accurately, whereas the constant friction coefficient approach leads to the most inaccurate temperature profile. Moreover, a
three-dimensional thermomechanical finite element (FE) analysis based on the calculated heat input data and the experimental axial
shortening rate demonstrates that the process can be analyzed in a one-dimensional domain due to the short frictional heating cycle
and the uniform heat-generation rate across the interface. The FE analysis also indicates that the heat-generation rate due to the plastic
deformation in the workpiece away from the interface is negligible compared to the heat-generation rate by friction.
Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Friction welding; Heat generation; Inverse heat conduction problem; Finite element analysis; Sliding–sticking friction

1. Introduction nents are brought together under application of pressure.


In linear friction welding, which has been used since the
Friction welding is a solid-state joining process wherein 1980s, the parts move under friction pressure relative to
heat is generated by rubbing one component against each other in a reciprocating fashion through a small linear
another under pressure. Once sufficient heat has been gen- displacement (amplitude) in the plane of the joint to be
erated, the rubbing action is terminated and the pressure is made. Orbital friction welding, which is a combination of
usually increased in the forging stage to consolidate the linear and rotary friction welding, was introduced in the
weld. The joint has a narrow heat-affected zone, and shows early 1970s. In this process the centre of one component
plastically deformed material which is subject to severe relative to the other component is moved around a two-
plastic deformation due to the high local temperature and dimensional curve, e.g. a circle, to provide the rubbing
internal pressures. The subject has been reviewed thor- action. The two parts to be joined are rotated around their
oughly [1]. longitudinal axes in the same sense with the same constant
There are three variants of this welding technique: angular speed. The two longitudinal axes are parallel
rotary, linear and orbital friction welding (Fig. 1). In rotary except for a small distance offset. When motion of the com-
friction welding, which has been used commercially since ponents ceases, and before forging pressure is applied, the
the 1940s, one component is rotated around its axis while parts are correctly aligned to form a weld.
the other remains stationary. For welding, the two compo- Rotary friction welding has the inherent limitation that
it cannot be used for non-circular cross-section compo-
*
Corresponding author. Tel.: +43 0 316 873 7184; fax: +43 0 316 873
nents. Another major disadvantage is that the rate of
7187. heat generation is not uniform over the interface (Fig. 1).
E-mail address: mehran.maalekian@tugraz.at (M. Maalekian). This causes the heat-affected zone to have a non-uniform

1359-6454/$34.00 Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2008.02.016
2844 M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855

Fig. 1. Three variants of friction welding. A comparison of heat generation over the interface for three types of friction welding is shown with black
arrows.

thickness. These shortcomings can be avoided by using lin- bars. Since then, many researchers have employed a con-
ear or orbital friction welding. The two methods can weld stant friction coefficient to quantitatively describe the fric-
non-circular parts and the interfacial energy is generated tional heat generation at the interface [4–6]. However, it
almost uniformly. Fig. 1 illustrates that the frictional heat should be noted that the relative speed, the temperature
generation at the interface for orbital compared to linear of the rubbing surfaces, the nature of the material, the pres-
friction welding is even more uniform. This is attributed ence of surface films, the normal pressure and the rigidity
to the uniform unidirectional relative velocity between of the friction surfaces are the five factors that influence
the two components over the whole interfacial area [2]. the coefficient of friction in a quantitative manner.
In friction welding, three parameters control the charac- Vill’ [7], for the thermal analysis of rotary friction weld-
ter of a weld: relative velocity between the workpieces, ing of low-carbon steel, took the coefficient of friction as
duration of the process and axial force. Linear and orbital being inversely proportional to the square of the linear
friction welding require the addition of amplitude as a velocity. In 1962, Cheng [8] made the first numerical anal-
fourth parameter. These parameters determine the amount ysis of friction welding using the finite difference (FD)
of energy input to the weld and the rate of heat generation method. He took variable thermal properties, experimental
at the interface. time-dependent heat input, axial shortening and melting
Since the 1950s, friction welding has been widely studied into consideration for his analysis. However, in contrast
experimentally and theoretically to better understand the to Cheng’s results, it is widely accepted [1] that gross melt-
welding process as well as the welded materials [1]. Peak ing does not occur during friction welding.
joint temperature and the temperature profile next to the The first finite element (FE) approach to friction welding
weld have a significant effect on the joint properties, flash is attributed to Sluzalec [9]. He used variable material prop-
formation and heat-affected zone (HAZ). It is, therefore, erties and a temperature-dependent friction coefficient, to
desirable to estimate the heat-generation rate at the friction predict the temperature distribution and deformation pat-
surface accurately to calculate the joint peak temperature terns of the flash in friction welding of mild steel. Moal
and temperature profile in the HAZ. The major obstacle and Massoni [10] developed a thermomechanical FE model
in theoretical analysis of friction welding is the determina- specifically for the inertia welding of two similar bars. An
tion of an analytical representation for the rate of heat input incompressible viscoplastic material behavior was used in
at the interface. The general expression for the rate of heat their model. Using torsion tests, the temperature-dependent
generation is the shearing work at the interface, which is material properties at different strain rates, and the friction
assumed to be converted to frictional heat and is repre- coefficient variation with speed and pressure were identified.
sented in terms of friction coefficient, pressure distribution The mechanical problem was solved by considering the vir-
across the interface, relative velocity, and radius of the tual work principle including the inertia term. The predicted
cylindrical workpiece for rotary friction welding and ampli- evolution of interface temperature showed the final temper-
tude for linear and orbital friction welding. The analytical ature to be in the melting zone of the alloy. In one of the
description for the friction coefficient is the hardest task. very few available three-dimensional FE models, Zhang
The first published quantitative analysis is attributed to et al. [11] recently performed a coupled thermomechanical
Rykalin et al. [3], who used a constant coefficient of friction FE model of rotary friction welding of a nickel based super-
and thermal properties, and uniform heat flux to calculate alloy using the DEFORM software package [12]. The
TM

the temperature profile in friction welding of carbon steel analysis was carried out using a rigid viscoplastic material
M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855 2845

model. The torsional friction was introduced into the model


to describe the heat generation during friction welding. The
friction stress was characterized in two stages: Coulomb
friction law was used for the low-temperature range, and,
for the high-temperature range, the shear yield stress of
the material was considered as the friction stress. However,
a general expression for the friction coefficient as a function
of pressure, temperature and linear velocity was proposed
without identifying the constants of the equation. More-
over, in the paper, no criterion that distinguishes low- and
high-temperature ranges was described.
It should be noted that neither the postulated expres-
sions for pressure and coefficient of friction, nor the result-
ing formulae for heat generation have been experimentally
verified to any degree of generality and, in many cases,
experiments have shown them to be grossly oversimplified
and invalid [1]. Therefore, several researchers have pro-
posed that a more fruitful approach for application in tem-
perature studies is to choose the power input as an
approximation of experimental power traces for each set
of friction welding conditions [8,13–16].
To overcome the major difficulty in thermal analysis of
friction welding, Maalekian et al. [17] have very recently
also proposed an inverse heat transfer model for estimation
of heat generation. Measured interior temperature data
were used in their inverse heat transfer model. Using the
calculated heat generation, the predicted temperatures
matched closely to the experimental data. Their simple
inverse model of heat generation can be used to calculate
the apparent coefficient of friction.
Although several analytical and numerical thermal mod- Fig. 2. Variation of (a) yield strength, (b) specific heat (CP) and thermal
els of friction welding have been reported in the literature, conductivity (k) of the steel with temperature.
no previous study has provided a comparative thermal
analysis using different heat generation methods. In this ing. The temperature changes during the friction welding
paper, we compare four different methods of frictional heat process were measured by several thermocouples attached
generation calculation for the orbital friction welding of to the steel bar at distances of 2.5, 5.0 and 7.5 mm from
rectangular steel bars. The resulting temperature fields the rubbing surface.
are compared with the experimental data and advantages
or shortcomings of the methods are discussed. 3. Mathematical modeling

2. Materials and methods 3.1. General assumptions

The chemical composition of the steel studied is 0.75C, For the analysis of friction welding, the heat generation
1.02Mn, 0.28Si, 0.11Cr, 0.05Ni, 0.015S, 0.009P, 0.08Cu and the pressure across the interface are assumed to be uni-
(wt.%). The thermophysical properties of the steel are form. The uniformity of interfacial heat generation in the
shown in Fig. 2. The yield strength of the steel at different orbital friction welding (Fig. 1) was experimentally verified
temperatures was determined experimentally using stan- by the observation of a uniform width of the HAZ reported
dard tensile testing at low temperatures and Gleeble 1500 in Refs. [2,19]. Moreover, this assumption was also mathe-
thermomechanical simulation at high temperatures matically approved for small values of the amplitude [20].
(Fig. 2). The specific heat (CP) is taken from the database Uniform pressure is also a general assumption, which has
of the software DEFORM [12] and the thermal conductiv- been frequently used by many researchers for modeling
ity (k) is taken from Ref. [18]. friction welding (e.g. [4,9,21]).
An orbital friction welding machine was used to produce
a weld between two rectangular steel bars with cross-sec- 3.2. Thermal analysis
tional dimension of 88  20 mm. A PC-based data acquisi-
tion system was designed to monitor and record axial In friction welding, the temperature in the weld region
displacement, rotation speed and axial force during weld- rises sharply due to friction and plastic work. To calculate
2846 M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855

the temperature profile, the heat transfer analysis is under- oT 
k ¼ qðtÞ ð3Þ
taken by considering the frictional heat generation at the ox x¼0
interface, heat generation by plastic deformation and heat
where q(t) is the frictional heat generation at the interface.
loss to the environment. For the homogeneous and isotro-
As was mentioned above, the aim of this paper is to com-
pic continuum with temperature-dependent material char-
pare different approaches for estimating the heat-genera-
acteristics, the following equation of heat conduction
tion rate q(t) as described below.
applies, which is based on the principle of conservation
of energy:
  3.3. Heat generation
oT o oT oT
qC P ¼ k þ qC P u þ S_ ð1Þ
ot oxi oxi ox1 Four methods for the calculation of q(t) are presented.
Two approaches are based on the coefficient of friction, the
where q is the material density, CP is the specific heat, k is
third method is based on the experimental power dissipa-
the thermal conductivity, T is the temperature, t is the time,
tion and, finally, an inverse heat conduction model for esti-
and xi with i = 1, 2 and 3 represents x, y and z directions,
mation of the heat generation is described.
respectively. The convective term on the right-hand side of
the equation accounts for the shortening of the specimen
3.3.1. Constant friction coefficient (Coulomb friction)
during friction welding, which is given in the usual way
The difficulties in determining the friction value lie in the
by the product of the temperature gradient and the short-
complexity of the phenomena and in the inability to accu-
ening velocity u. S_ is the heat-generation rate due to plastic
rately measure shear stresses [25]. Therefore, far-reaching
deformation in the workpiece away from the interface,
approximations are used to explain friction behavior dur-
which can be defined as
ing friction welding. These approximations deal with
S_ ¼ are
_ ð2Þ apparent friction rather than with the fundamental phe-
nomenon. One of the consequences of this approach is that
where r  is the effective stress, e_ is the effective strain rate
the friction value must be adapted separately for each fric-
and a is the thermal efficiency of plastic deformation,
tion welding process and for each material.
which is assumed to be 90–95% [22–24]. The rest of this en-
In the study of friction welding, similar to metal-forming
ergy is principally stored as dislocations and vacancies. For
processes, it is usually assumed that the resistance to sliding
the benefit of simplicity, in the calculation, the term S_ is ig-
along the interface between the workpieces is uniform over
nored due to its low value compared to the frictional heat
the entire contact surface. The most common simplifying
generation, as will be discussed later in the paper. Fig. 3
assumption made with regard to friction stress (sfric)
shows the coordinate system and boundary conditions used
between the workpieces involves Coulomb friction. For
for the model.
Coulomb friction, it is assumed that the contact shear stress
The major boundary condition is the frictional heat gen-
(sfric) is proportional to the contact (normal) pressure (P) as
eration at the interface, taken as follows:
sfric ¼ l  P ð4Þ
where the proportionality factor (l) is called the Coulomb
coefficient of friction.
The heat flux (q) generated by friction between the two
loaded specimens rubbing together with relative sliding
velocity v can be expressed as
q ¼ sfric  v ð5Þ
The relative velocity is related to the rotational speed (n)
as
v ¼ 2p  n  e ð6Þ
where e is the amplitude (offset distance) in the orbital fric-
tion welding.
Therefore, if we combine the above formulations (4)–(6),
the frictional heat flux for orbital friction welding is
obtained as a function of process parameters
q ¼ 2p  n  e  l  P ð7Þ
The major unknown value for estimation of heat flux is
the friction coefficient (l). Many factors affect frictional
behavior and the forces that resist sliding between the
Fig. 3. Geometry, coordinate system and boundary conditions used for two surfaces. For example, according to Blau [26], a num-
the model. ber of possible variables influencing frictional behavior,
M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855 2847

and thus l, are contact geometry (e.g. surface roughness),


temperature, relative motion (e.g. unidirectional or recipro-
cating motion, magnitude of relative surface velocity),
applied forces, material properties (e.g. stiffness and flow
properties), etc. In the literature, many different friction
coefficient values even for the same sliding materials are
reported, which is attributed to the large number of poten-
tial friction-affecting factors. Consequently, the friction
coefficient is an established, but somewhat misunderstood
quantity in the field of science and engineering [26]. There-
fore, it is virtually impossible to find the exact coefficient of
friction, particularly for friction welding processes, by
looking up values in published tables of friction coefficients
because the conditions used to obtain those values are usu-
ally unlike the conditions encountered during friction
welding. Fig. 4. Characterization of frictional shear stress as a function of normal
As a result, it is necessary to select a value for the fric- pressure. When the contact shear stress is smaller than the shear yield
tion coefficient that gives reasonable results for our prob- stress (i.e. sfric < sy), the sliding condition is met and the friction force is
proportional to the applied force (Amontons’ law). If sfric is equal to sy,
lem. Here, a value of 0.2 is taken for the friction the sticking state is fulfilled and the real area of contact is equal to the
coefficient. This value provided the best possible agreement apparent one. In this schematic diagram, the ideal condition is shown with
between computed and experimental temperatures. the solid line, whereas in the actual condition it is difficult to define a
transition point between the two regions; therefore, the real behavior is
3.3.2. Sliding–sticking friction shown by the dotted curve.
The normal interpretation of Amontons’ (or Cou-
lomb’s) friction law (Eq. (4)) is based on rigid contact sur- are large compared to the molecular dimension, are called
faces (also called dry friction) and does not take into asperities. In sliding conditions, due to the inevitable
account internal stresses and plastic deformation. How- roughness over the contact surfaces, contacts are made
ever, when heavy plastic deformation is involved, as in between the asperities of the two contact bodies and the
metal forming and friction welding, the Coulomb friction real contact area is a small fraction of the apparent contact
is not sufficiently representative for the model. Therefore, area. In this case the frictional shear stress is directly pro-
it is more reasonable to take the friction stress along the portional to the normal load (see Fig. 4). On the other
interface of the two bodies as a function of the normal hand, when the real and apparent areas of contact are
pressure in the friction welding process (Fig. 4). approximately equal, the fully plastic flow at the interface
At a microscopic level, even well-machined metal sur- occurs (as shown by the dotted curve in Fig. 4b). In this
faces are not smooth homogeneous planes, but rough con- case, as illustrated in Fig. 4, the friction stress and material
tours with numerous irregularities are seen (as shown by shear stress are equal and the sticking condition is fulfilled
the dotted curve in Fig. 4a). These irregularities, which (see Fig. 5).

Fig. 5. Schematic of the two different interfacial conditions that may arise in friction welding. (a) The real contact area is the asperities with small fraction
of the apparent contact area and deformation occurs only in the asperities (i.e. the sliding condition). (b) Full plasticity at the interface arises when real and
apparent areas of contact are equal (i.e. the sticking condition).
2848 M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855

Therefore, while Coulombic friction controls the inter-


face forces at low loads according to Amontons’ law, as
the load increases to the point where the real area of contact
is equal to the apparent area of contact, friction becomes
independent of pressure and takes on the value sy, which
is the shear yield stress of the workpiece material (Fig. 4).
By considering sliding and sticking conditions, the inter-
facial heat generation rate at the interface of the two work-
pieces may be defined as
q ¼ ½ð1  dÞlPv þ ðdgsy vÞ ð8Þ
where g is the mechanical efficiency, i.e. the amount of
mechanical energy converted to heat energy, and d is a
state variable, which denotes the extent of sticking. When
d is 1, full plasticity at the interface (full sticking) is present
and the entire heat is generated by plastic deformation,
whereas heat is generated only by Coulombic friction if
d = 0. Recently, similar approach has been used to model
the heat generation in friction stir welding [27,28].
During friction welding, the interface temperature rises
rapidly and, consequently, the yield strength of the mate-
rial decreases (Fig. 2a). To calculate q, three stages accord-
ing to the variation in shear yield stress are defined. When
the temperature is low, i.e. at the beginning of friction
welding, and the friction stress (sfric) is smaller than the
shear yield stress (sy), pure sliding interface conditions
(i.e. d = 0) described by Coulomb friction are assumed.
At high temperatures, when the shear yield stress is lower
than the friction stress, a pure sticking condition (i.e.
Fig. 6. (a) Variation of shear yield strength with temperature obtained
d = 1) is adopted. For the third situation, which is the tran- from the von Mises yield criterion and tensile yield strength (Fig. 2a).
sition from Coulomb friction to full plasticity at the inter- When the interface temperature is lower than 720 °C, the state variable d is
face (see Fig. 4), a partial sliding–sticking condition (i.e. 0, and when the temperature reaches 860 °C, d is assumed to be 1. (b) It is
0 < d < 1) is assumed. The transition temperature range is assumed that the transition from Coulomb friction (d = 0) to full plasticity
defined according to the variation of shear yield stress as (d = 1) takes place within 0.1 s.
a function of temperature (Fig. 6).
The shear yield stress is related to yielding in the tension the heat flux for the analysis and modeling of friction
test using the von Mises yield criterion as welding processes [8,13–16]. Therefore, the experimentally
1 measured power data is used to define the heat flux bound-
sy ¼ pffiffiffi ry ð9Þ ary condition at the weld interface, as the third variant of
3 the heat-generation estimate. The variation of recorded
where ry is the material tensile yield stress. The shear yield power with time during friction welding is shown schemat-
stress, obtained from Eq. (9) and Fig. 2a, is shown in ically in Fig. 7, which shows that the dissipated power rises
Fig. 6a. Onset of partial sticking is assumed at 720 °C, to an initial peak value. This is the required power in order
where the shear yield stress is equal to the friction stress to reach a predetermined rotational speed. In this stage
(i.e. sfric = sy shown in Fig. 6a). The transition from slip there is no contact (no friction force) between the speci-
to stick condition (i.e. 0 < d < 1) is assumed to take place mens, and therefore it is included in the calculation of
within 0.1 s. Then, after this transition period, the full plas- the heat-generation rate. Then, the power increases to its
ticity condition (i.e. d = 1) is assumed (Fig. 6b). The tem- peak value as the frictional force is exerted. Since the
perature for the full plasticity condition is about 860 °C. recorded power (Fig. 7) is not entirely converted to heat
At this temperature, which is about 0.6 Tm (where Tm is at the friction interface, an efficiency factor of 35% is
the melting temperature of the material in Kelvin), the yield defined for the machine. This value gives the best possible
stress is low, and dynamic recrystallization usually occurs. correspondence between the calculated and measured tem-
perature values.
3.3.3. Experimental power dissipation
During the experiment, the power dissipation by the 3.3.4. Inverse heat conduction method
machine is monitored. The power dissipation characteristic In the direct heat conduction problem (DHCP), the tem-
as a function of time has been employed frequently as perature distribution in the interior of a body is determined
M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855 2849

mate the surface heat flux q(t) on the boundary generated


by friction provided that certain transient temperature
measurements inside the specimen are known.
In the orbital friction welding process, the interfacial
heat generation is approximately uniform, and, since the
duration of the friction welding cycle is very short, i.e. typ-
ically a few seconds, the radiation and convection heat
losses are insignificant [8,16,17]. Therefore, in the model,
the lateral surfaces of the specimen are assumed to have
an adiabatic boundary condition and the temperature dis-
tribution existing across the plane parallel to rubbing sur-
face is uniform (Fig. 8). Consequently, the temperature of
the cross-section (plane perpendicular to the x-axis in
Fig. 8) is approximated by a single value at the node in
Fig. 7. Schematic trace of dissipated power during the friction welding
the one-dimensional (1-D) FD analysis.
process.
Because the solution of the inverse heat conduction
problem (IHCP) is always based on the corresponding
from data given on its surface. However, it is sometimes direct problem, the formulation of the DHCP is out-
necessary to estimate surface temperature and surface heat lined first. Then, the inverse heat conduction model is
flux from the temperature data measured at some location presented.
inside a body rather than use the surface data to calculate The partial differential equation describing 1-D
bulk values. This is an inverse problem because it is con- unsteady heat conduction with variable thermal properties
cerned with the estimation of the unknown surface condi- and consideration of the translational movement due to
tion. In the direct problem, the causes are given, the shortening (burn-off) is given by
effect is determined, whereas, in the inverse problem, the  
effect is given, the cause (or causes) is estimated [29]. oT o oT oT
qC P ¼ k þ qC P u ð10Þ
The aim here is to develop an inverse heat conduction ot ox ox ox
model to determine the heat generation at the welding
interface during the friction welding process. The experi- The boundary and initial conditions when solving Eq. (10)
mental temperature data measured by thermocouples are
located near the friction heated region are used to estimate 
@T 
the heat flux generated on the rubbing surface. Based on k  ¼ qðtÞ ¼ ? for t > 0 ð11Þ
this approach, an explicit FD model is developed to resolve @x x¼0
the inverse heat conduction problem and to find the heat T ðx; tÞ ¼ T 0 ; for t ¼ 0 ð12Þ
flux based on experimental measurement.
where T0 is the initial temperature of the specimen, which is
3.3.4.1. Inverse heat conduction problem concept. If the tem- taken as 25 °C.
perature gradient inside the solid at the heat transfer sur- The objective is to calculate the surface heat flux evolu-
face is determined from experimental measurements, the tion, q(t), at the friction surface (x = 0), based on the tem-
heat flux can be calculated as a product of the thermal con- perature measurements during the welding experiment.
ductivity of the solid and the temperature gradient at the
surface. Due to the practical difficulty involved in the exact
measurement of surface temperature or surface tempera-
ture gradient, it has been suggested that the surface condi-
tions be estimated from temperature measurements made
at some convenient point inside the solid. Such a problem
is termed the inverse problem of heat conduction, wherein
the temperature history at an interior point is prescribed
and the surface conditions are calculated from the solution
of the transient heat conduction equation satisfying the
prescribed interior conditions. This is in contrast to the
direct problem of heat conduction, where the boundary
conditions are prescribed at the surfaces and the interior
conditions are calculated [30].

3.3.4.2. Mathematical modeling. The inverse heat conduc- Fig. 8. Boundary conditions used for the inverse modeling of friction
tion problem investigated here uses the FD method to esti- welding of a bar.
2850 M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855

Fig. 9. Space–time grids for interior, inward and outward extrapolation regions are shown.

3.3.4.3. Explicit finite difference method. The transient heat The increment values are chosen such that the stability
conduction problem can be solved numerically by transfer- criterion for the explicit numerical solution is obeyed,
ring the partial differential equation of heat conduction to which is given by
FD equations in both space and time domains. Based on a  Dt 1
the explicit FD method and using the notations of Fig. 9, 2
6 ð16Þ
ðDxÞ 2
the transient differential equation (10), which applies to
all interior nodes, can be written as
3.3.4.4. Inverse heat conduction solution. The only unknown
Dt aiþ1;j þ ai;j þ ai1;j
T i;jþ1 ¼ T i;j þ 2
 in the FD model presented above is the heat flux qj gener-
ðDxÞ 3 ated by friction. The inverse heat conduction method is
Dt applied to obtain qj. The inverse problem is solved by min-
 ðT i1;j  2T i;j þ T iþ1;j Þ þ u ðT iþ1;j  T i;j Þ ð13Þ
Dx imizing the objective function, R, which is the least-squares
norm defined as
The boundary nodes at the heating surface (x = 0) with
the index i = 1 can be expressed as X
J X
I
R¼ ðY ij  T ij Þ2 ð17Þ
Dt a1;j þ a2;j þ a3;j þ a4;j þ a5;j j¼1 i¼1
T 1;jþ1 ¼ T 1;j þ 2

ðDxÞ 5 where Yij and Tij are the vectors containing the measured
ð25T 1;j þ 48T 2;j  36T 3;j þ 16T 4;j  3T 5;j Þ and predicted temperatures, and the superscripts i and j
 indicate distance and time increments, respectively.
12
qj Dt Suppose that a matrix of temperature measurements,
þ  Yij, is given for locations xi and times tj and a reasonable
qC P Dx
initial guess q0j is available for the unknown vector qj .
ð14Þ
Then, the temperatures Tij are computed based on the
And for the boundary nodes at x = L (right end of the direct FD formulation and compared with the measured
bar), the mirror points shown in Fig. 9 are used: values Yij. If the objective function R, which represents
the squared mean deviation between Yij and Tij, is smaller
Dt than a given threshold tolerance, the present values of qj
T I;jþ1 ¼ T I;j þ 2
 ðaI;j þ aI1;j Þ  ðT I1;j  T I;j Þ
ðDxÞ are accepted. Otherwise, by calculation of the error gradi-
Dt ent and performing a correction step into qj and repeating
þ u ðT I;j  T I1;j Þ ð15Þ the above steps, the objective function is further minimized.
Dx
Finally, a suitable heat flux delivered by friction is esti-
where i = 1, 2, 3, . . . , I is the number of distance increments mated. Fig. 10 shows the flowchart of the algorithm, which
of Dx, j = 1, 2, 3, . . ., J is the number of time increments of has been written in MATLAB software [31].
Dt , a (m2 s-1) is the thermal diffusivity of the steel and qj For the above method, it is necessary to describe the
(W m-2) is the heat flux generated by friction. qj is the vec- variation of q(t). Linear interpolation is used for the
tor form of q(t) in the formulation. As pointed out previ- approximation of q(t) as follows:
ously, in friction welding, the heating stage is very short (of
t  tj
the order of a few seconds), and therefore for the sake of qðtÞ ¼ qj þ ðqjþ1  qj Þ  for tj < t < tjþ1 ð18Þ
Dt
accuracy of the model at the heat front boundary nodes,
the fourth-order error approximation is used (Eq. (14)) to where tj =j Dt, j = 1, 2, 3, . . . , J.
achieve a good representation of the corresponding steep The governing equations and the boundary conditions
gradients. are implemented in a 1-D FD computer code. The accuracy
M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855 2851

Fig. 10. Flowchart of the inverse method for estimation of heat flux
generated by friction.

of the 1-D FD program will then be examined by a 3-D FE


analysis based on the commercial software DEFORM [12].

4. Results and discussion

4.1. Experimental temperature and axial shortening

The thermal cycles measured at three distances, i.e. 2.5, Fig. 11. (a) Time–temperature profiles at several distances of 2.5, 5 and
7.5 mm from the weld interface. The thermal profile at location
5 and 7.5 mm, from the weld interface and the axial short-
x = 2.5 mm is used for the IHCP. (b) Experimentally recorded axial
ening recorded during the friction welding process are shortening. For the calculations, the two linear shortening segments
shown in Fig. 11. The measured temperature of the nearest (dashed-line) are used.
position to the weld interface (i.e. x = 2.5 mm) is used for
the inverse heat conduction model. In order to simplify
the calculation of heat-generation rate, a linear axial short-
ening rate shown in Fig. 11 is employed in the models.

4.2. Heat-generation rates

The estimated heat-generation rate at the friction inter-


face using the three analytical methods, i.e. constant fric-
tion coefficient, sliding–sticking friction and inverse heat
conduction analysis, as well as the measured power dissipa-
tion method (with 35% efficiency factor) are shown in
Fig. 12.
The first method (constant friction coefficient) shows a
constant heat-generation rate through the whole friction
cycle, whereas the sliding–sticking friction method depicts
a peak value in the heat-generation rate. This maximum
value occurs in the transition zone from Coulomb friction Fig. 12. Comparison between the calculated heat-generation rates at the
to the full plasticity condition at the interface when d = 0.5 friction interface with the four different methods described in the text.
(i.e. when the friction time is 2.1 s), because, in this zone,
the sliding (Coulomb) and sticking (shear yield stress) fric-
tion both contribute in the heat-generation rate (Eq. (8)). contribution of the partial sticking friction to the heat-gen-
Moreover, due to the high shear yield stress value at this eration rate is very significant. At friction time t = 2.2 s,
moment (t = 2.1 s), which is about 70 MPa (Fig. 6a), the where the full sticking condition takes over (d = 1), the
2852 M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855

shear yield stress falls to 42 MPa (Fig. 6a) and, conse- tions based on the constant friction and slip–stick models
quently, the heat-generation rate decreases. Only 0.1 s do not generally match the actual data except for the last
later, at t = 2.3 s, the heat-generation rate drops drastically 0.4 s friction time. The two models based on friction theory
as a result of significant reduction in the shear yield stress predict a very rapid increase in temperature, whereas the
of the steel (i.e. sy = 22 MPa). inverse model and the actual data show a gradual increase
The heat-generation rate obtained from the trace of dis- of temperature. This disagreement between the two models
sipated power during the friction welding process (Fig. 7) (constant and slip–stick friction) and experiment is
shows a gradual increase to a peak value close to the end expected, because both are based on a constant high heat
of the process. Since the welding time is very short, i.e. input rate from the beginning of the process, which is not
3 s, the steady-state power level, which is usually expected realistic. In fact, the heat-generation rate in friction weld-
in the extended-time friction welding condition, does not ing usually does not start from high values, and remains
appear. constant up or close to the end of the friction stage. How-
The heat-generation rate estimated with the inverse heat ever, the two models, particularly the slip–stick model, are
conduction method increases continuously to a peak value fairly successful at representing the data at the end of the
at friction time about t = 2.2 s, which in this regard is friction welding cycle (i.e. in time interval 2.6–3 s).
almost similar to the slip–stick method. The peak values The predicted temperature profile based on the modified
of the frictional heat estimated with the sliding–sticking power dissipation curve better represent the data when
friction and the inverse methods are comparable, which compared to the two models based on friction theory.
supports the presumed transition zone from Coulomb fric- However, the measured power data also does not represent
tion to full plastic interface with a plastic layer (shown by the actual temperature evolution accurately.
the dotted lines in Fig. 4) used in the calculation of heat A measure of how well the four models represent the
generation with the second method (i.e. sliding–sticking actual heat-generation rate has been deduced by comparing
friction). As it will be discussed later in the paper, the the predicted and measured temperature data. It is con-
IHCP is a suitable method for reproducing the temperature cluded that the inverse heat conduction method is the most
profiles in the specimen precisely, thus, based on the accu- accurate model for representing the actual heat input in the
racy of the IHCP, the occurrence of the transition condi- friction welding process. The experimental power dissipa-
tion assumed in the sliding–sticking friction method is tion model data can be used for the analysis of friction
supported by the similar peak values occurring at almost welding, provided that an appropriate efficiency factor is
the same friction time. defined. Although the sliding–sticking friction and the con-
stant friction coefficient methods provide less accurate
4.3. Temperature profiles data, these variants are useful for quick and rough estima-
tion of the heat-generation rate if the friction coefficient
For the four proposed heat inputs, the thermal profiles value and transition zone are chosen properly.
will now be compared. The computed temperature profiles Upon frictional heating, the interface temperature
at the initial 2.5 mm distance from the interface are shown climbs to a peak value, which is very hard to measure
in Fig. 13 and compared to the actual data. The heat gen- experimentally. However, the models can at least provide
eration obtained from the inverse heat conduction method good estimates of the interface temperature, as shown in
predicts the temperature profile accurately, whereas predic- Fig. 14. Except for the sliding–sticking model, numerical

Fig. 13. Comparison between predicted and measured temperature Fig. 14. Predicted temperature profiles at the friction interface using the
profiles at an initial distance of x = 2.5 mm from the weld interface. four different models.
M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855 2853

simulation predicts a continuous increase in the interface


temperature. The reason for the discontinuity appearing
in the second model is the transition from sliding to full
plasticity. As shown in Fig. 12, the heat-generation rate
in the transition zone rises rapidly at a friction time of
2.1 s, and therefore a jump in the interface temperature is
observed. In other words, the local variation in heat-gener-
ation rate leads to a significant variation in local tempera-
ture. The peak temperature values predicted with the
inverse method (975 °C), the measured power method
(986 °C) and the sliding–sticking model (951 °C) are com-
parable, whereas the constant friction model predicts a sig-
nificantly lower peak temperature (908 °C) compared to
the other models.

4.4. Verification with finite element analysis

As indicated previously, for the analysis of the orbital


friction welding process, a 1-D FD model was used. Due
to the assumption of adiabatic boundary conditions, the Fig. 15. FE model of friction welded rectangular bars. The effective strain
profile and the size of the deformation zone are shown in the image.
radiation and convection heat transfer to the environment
are neglected. Moreover, in the deforming zone next to the
friction surface, the heat-generation rate owing to the plas-
tic deformation S_ was also neglected as explained earlier.
Therefore, in order to verify these simplifications, a 3-D
thermomechanical FE analysis using the commercial FE
package DEFORM [12] is carried out.
For the FE analysis, the heat-generation rate estimated
with the inverse heat conduction method is utilized because
it represents the frictional heat input more accurately than
the other methods (see Fig. 13). The input for deformation
in the FE analysis is based on the experimental axial short-
ening data, which is shown in Fig. 11b. The bar is meshed
using cuboid elements with a refined mesh close to the weld
zone. The data used for the calculation are presented in
Table 1.
The deformation pattern and the distribution of effective
strain calculated with the FE model are shown in Fig. 15.
The maximum effective strain in the range of 0.34–0.42 is
located next to the weld interface. The predicted width of
the deformation zone is about 1.8 mm. Variation of maxi-
mum effective strain rate and effective stress with friction
time predicted by the model are depicted in Fig. 16. Since
the analysis is based on the recorded upset data
(Fig. 11b), the stress and strain rate do not show any
change in the early stages of the friction time (i.e. 1 s).
However, the stress and strain rate intensify as the axial
shortening increases.

Table 1
Material data used in calculations
Property Value
3
Density, kg m 7800
Emissivity 0.7
Fig. 16. (a) Variation of maximum effective stress and (b) maximum
Heat transfer coefficient, W m2 K1 20
effective strain rate with time calculated by the FE model.
2854 M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855

ent approaches. The first approach is based on the constant


friction coefficient (Coulomb friction). For the second
approach, sliding, sticking and the transition between slid-
ing and sticking conditions are considered and, in the ana-
lytical heat-generation rate formulation, a state variable (d)
is introduced, which accounts for the operating heat gener-
ation mechanisms for each condition (i.e. sliding, sticking
and partial sliding–sticking). A criterion based on friction
stress, shear yield stress and the interface temperature is
introduced to accommodate different contact conditions.
For the third heat generation estimation method, the
experimentally measured power data with an efficiency fac-
tor is used to define the heat flux boundary condition at the
weld interface. For the last approach, an inverse heat con-
duction model is used. The inverse problem is solved by
Fig. 17. Comparison of temperature profiles between centre (P1) and minimizing the objective function, which is the squared
surface (P2) of a cross-section close (x = 1 mm) and parallel to the friction mean deviation between the calculated and measured tem-
surface calculated from FE analysis. Location of the points P1 and P2 are
perature values. The minimization is performed numeri-
shown in Fig. 15.
cally with a 1-D explicit FE model.
For the calculation, temperature-dependent thermal
The maximum effective stress and strain rate profiles conductivity, specific heat and yield strength are consid-
(Fig. 16) can now be utilized to calculate the maximum ered. A measure of how well the four approaches represent
heat-generation rate due to plastic deformation (Eq. (2)). the actual heat-generation rate is deduced by comparing
Using Fig. 16, the total volumetric energy generation by the predicted and measured temperature data. The results
deformation (i.e. r _ during friction welding is obtained
  e) show that the inverse heat conduction method is the most
with 10 MW m-3. Taking into account the rubbing surface accurate approach, and best represents the actual heat
area (1760 mm2) and the width of the deformation zone input in the friction welding process. The experimental
(1.8 mm), the overall power generation by deformation is power dissipation method can be used to analyze friction
obtained with 30 W. On the other hand, using the area welding, provided that an appropriate efficiency factor is
below the heat-generation rate diagram calculated with defined. Although the sliding–sticking friction and the
the IHCP (Fig. 12) and the interfacial area (1760 mm2), constant friction coefficient methods provide less accu-
the total frictional heat input is obtained with 24 kW. That rate data, those variants are useful for quick and rough
is, the amount of heat generated by deformation next to the estimations of heat-generation rate if suitable value for
friction interface zone is only approximately 0.1% of the the friction coefficient value and transition zone are
overall heat input by friction. It is, therefore, reasonable chosen.
to assume that the heat generation by deformation is neg- Moreover, a 3-D thermomechanical FE analysis, based
ligible. This was assumed in the present paper. on the estimated heat input obtained from the inverse heat
The validity of the thermal boundary conditions, i.e. conduction model and the experimental axial shortening
insignificant heat loss to the environment and uniform tem- data, verifies that the adiabatic boundary conditions
perature profile across the planes parallel to the friction together with a uniform temperature distribution across
interface is confirmed by comparison of the temperature the plane parallel to the rubbing surface are reasonable
profiles between two points at the surface (P1) and the cen- assumptions. The cause is found in the short heating cycle
tre (P2) of a cross-section parallel and close (x = 1 mm) to during the orbital friction welding process. It is also shown
the friction interface. The two points are illustrated in that the heat-generation rate due to plastic deformation in
Fig. 15, and the FE calculated temperature profile is shown the workpiece away from the interface is negligible com-
in Fig. 17. The two temperature profiles are very close pared to the frictional heat-generation rate.
together and the maximum difference is almost negligible.
This implies that the temperature distributions at cross-sec- Acknowledgments
tions parallel to the heat source during the orbital friction
welding cycle reported here are homogeneous and it is thus The authors gratefully acknowledge the financial sup-
reasonable to assume that the lateral surfaces can be port of this work as part of K–net JOIN granted by the
described with adiabatic boundary conditions. Federal Ministry of Economy and Labor, Austria.

5. Summary and conclusions References


The heat generation term during orbital friction welding [1] Maalekian M. Sci Technol Weld Join 2007;12(8):738–59.
of rectangular steel bars has been studied using four differ- [2] Maalekian M. PhD Thesis, Graz University of Technology; 2007.
M. Maalekian et al. / Acta Materialia 56 (2008) 2843–2855 2855

[3] Rykalin NN, Pugin AI, Vasil’eva VA. Svarochnoe Proizvodstvo weld phenomena 8. Graz: Verlag der Technischen Universität Graz;
(Weld Prod) 1959:42–52. 2007. p. 881–90.
[4] Midling OT, Grong Ø. Acta Metall Mater 1994;42(5):1611–22. [18] Özisik MN. Heat transfer: a basic approach. Singapore: McGraw-
[5] Mitelea I, Radu B. In: Cerjak H, editor. Mathematical modelling of Hill; 1985.
weld phenomena 4. London: IOM Communications Ltd.; 1998. p. [19] Maalekian M, Kozeschnik E, Brantner HP, Cerjak H. Metall Mater
444–54. Trans A [in press].
[6] Ghanimi Y, Faes K. In: Cerjak H, Bhadeshia HKDH, editors. [20] Craine RE, Francis A. Wear 1987;114:355–65.
Mathematical modelling of weld phenomena 5. Cambridge: Wood- [21] Hollander MB. Met Eng Quart 1962;14–24.
head; 2001. p. 897–911. [22] Fu L, Duan LY, Du SG. Weld J 2003:65s–70s.
[7] Vill’ VI. Friction welding of metals. New York: American Welding [23] Zhang QZ, Zhang LW, Liu WW, Zhang XG, Zhu WH, Qu S. Sci
Society; 1962 [translated from Russian]. Technol Weld Join 2006;11(6):737–43.
[8] Cheng CJ. Weld J 1962;41(12):542–50. [24] Hosford WF, Caddel RM. Metal forming: mechanics and metal-
[9] Sluzalec A. Int J Mech Sci 1990;32(6):467–78. lurgy. Englewood Cliffs (NJ): Prentice-Hall; 1983.
[10] Moal A, Massoni E. Eng Comput 1995;12:479–512. [25] Blau PJ, editor. ASM handbook, friction, lubrication, and wear
[11] Zhang QZ, Zhang LW, Liu WW, Zhang XG, Zhu WH, Qu S. Sci technology, vol. 18. USA: ASM International; 1992. p. 27–69.
Technol Weld Join 2006;11(6):737–43. [26] Blau PJ. Tribol Int 2001;34:585–91.
[12] Scientific Forming Technologies Corporation. <http:// [27] Schmidth H, Hattel J, wert J. Model Simul Mater Sci 2004;12:143–57.
www.deform.com>. [28] Nandan R, Roy GG, Lienert TJ, Debroy T. Acta Mater
[13] Rich T, Roberts R. Brit Weld J 1971;93–98(March). 2007;55:883–95.
[14] Fu L, Duan LY. Weld Res Suppl 1998:202s–7s. [29] Özisik MN, Orlande HRB. Inverse heat transfer: fundamentals and
[15] Dave VR, Cola MJ, Hussen GNA. Weld Res Suppl 2001:246s–52s. applications. New York: Taylor & Francis; 2000.
[16] Nguyen TC, Weckman DC. Metall Mater Trans B 2006;37:275–92. [30] D’Sousa N. ASME Winter Ann Meet 1975;98:1–9.
[17] Maalekian M, Kozeschnik E, Brantner HP, Cerjak H. In: Cerjak H, [31] <http://www.mathworks.com/products/matlab>.
Bhadeshia HKDB, Kozeschnik E, editors. Mathematical modelling of

You might also like