You are on page 1of 8

Journal of Materials Processing Technology 169 (2005) 320–327

3-Dimensional CFD modelling of flow round


a threaded friction stir welding tool profile
Paul A. Colegrove a,b,∗ , Hugh R. Shercliff a
a Cambridge University, Department of Engineering, Trumpington Street, Cambridge CB2 1PZ, UK
b TWI Ltd., Granta Park, Great Abington, Cambridge CB1 6AL, UK

Received 29 August 2003; accepted 22 March 2005

Abstract

This paper describes the application of the computational fluid dynamics (CFD) code, FLUENT, to modelling the 3-dimensional metal
flow in friction stir welding (FSW). A standard threaded tool profile is used for the analysis and features such as the tool rake angle, heat
generation and heat flow are included. The primary goal is to gain a better understanding of the material flow around a complex FSW tool and
to demonstrate the effect of the tool rake angle, and weld and rotation speed. The model captured many of the real process characteristics, but
gave poor predictions of the welding forces. The model also generated an excessive amount of heat, which led to a large over-prediction in
the weld temperature. These shortcomings can be overcome by using a viscosity relationship that includes material softening near the solidus
or material slip at the tool interface.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Friction; Stir; Weld; Flow; Modelling

1. Introduction Shercliff and Colegrove [3] have reviewed modelling


activities in FSW. One of the great challenges is the devel-
Friction stir welding (FSW) is a relatively new welding opment and validation of a 3-dimensional flow model. This
process, patented in 1991 by Thomas et al. [1,2]. This pro- paper describes the development of such a model for the 5651
cess has great advantages in welding aluminium alloys that tool profile developed by Dawes et al. [4]. Bendzsak et al. [5]
are difficult to weld. The process gives low distortion, can and Askari et al. [6] have developed flow models for this
weld thick sections in a single pass and produces welds with tool.
excellent mechanical properties. The model published by Bendzsak et al. [5] used the finite
A schematic diagram illustrating the process of FSW is volume method to calculate the flow around the tool. Arrow
shown in Fig. 1. The key components of the FSW tool are: plots and particle tracks were used to visualise the flow field,
which was greater in size than that observed experimen-
• The shoulder: this is the primary means of generating heat tally. The flow indicated by the model was divided into two
during the process, prevents material expulsion and assists regions. Near the shoulder, it was largely rotational while
material movement around the tool. further down it appeared to extrude around both sides of
• The pin: the pin’s primary function is to deform the mate- the tool. Spirals were observed between these two regions
rial around the tool and its secondary function is to generate and it was postulated that these could be the cause of weld
heat. defects.
The work presented by Askari et al. [6] used the CTH
hydro-code and is the most advanced flow modelling pub-
∗ Corresponding author. Tel.: +44 1223 332791.
lished to date. The model assumed that the material stuck to
E-mail addresses: pac44@cam.ac.uk (P.A. Colegrove), the tool surface and used the Johnson and Cook [7] relation-
hrs@eng.cam.ac.uk (H.R. Shercliff). ship to represent the material flow stress. The model used

0924-0136/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2005.03.015
P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327 321

Fig. 1. Schematic diagram of friction stir welding.

elasto-viscoplastic material properties and very good agree- Fig. 2. Description of the mesh used for the models.
ment was obtained between the model output and marker
experiments with the flow being largely confined to the region
tion can be neglected. Neither region includes the combined
adjacent to the pin. The temperatures from the model com-
elasto-viscoplastic response of the material.
pared favourably with those obtained experimentally. Com-
The FSW flow model considers a rotating tool with the
parisons between the weld power input and forces were not
material translating past it at the welding speed. The work-
reported.
piece material is subdivided into the five regions shown in
Other authors have analysed the 3-dimensional flow
Fig. 2. Firstly, a rotating mesh that moves at the rotation speed
around unthreaded pins using analytical [8], finite element
of the tool represents a fluid region adjacent to the tool. This is
[9–11], finite difference [12] and Arbitrary Lagrangian Eule-
embedded in a stationary mesh slightly wider than the rotating
rian (ALE) [13–15] techniques. The ALE approach used by
region, extending to the length of the plate, which models the
Schmidt and Hattel [14,15] shows particular promise and can
flow and thermal response of the material that passes through
predict void formation in the weld which is impossible with
the deformation zone. To either side a solid aluminium region
the CFD approach used in this work.
models the thermal response in the non-deforming region,
The aim of the present work is to demonstrate the benefits
which is known to be travelling at the welding speed. Finally,
and pitfalls of using CFD to model FSW. To do this a flow
the tool and backing plate are solid regions that are used to
model was developed that included:
calculate the heat loss to these components.
(a) coupled thermal/flow analysis;
(b) heat generation by viscous dissipation; 2.2. Solver type and boundary conditions
(c) full thermal boundary conditions, such as heat loss to the
backing plate and tool, and the convective heat loss on The model was solved as a steady-state problem. While
the top surface of the plate; this solution technique is valid for an axi-symmetric or cylin-
(d) the tool rake angle. drical tool, the solution will only be an approximation for
a threaded tool. In this case, the solution is a ‘snap-shot’
The CFD package FLUENT [16] was selected for the of the flow at a particular instant in time. The approxima-
modelling work because of its ability to handle the high strain- tion is reasonable provided the time-dependent terms in the
rates that occur near the tool surface. Navier–Stokes equation remain small. The benefit of this
approach over a full transient analysis is much faster solu-
tion times.
2. Description of the model The model assumed that the material sticks to the tool
surface, so the material velocity equalled the rotation speed
2.1. Model regions multiplied by the radius. The translation is prescribed by
allowing the material to flow past the tool. Therefore, the
The models were 3-dimensional and assumed that the material enters the model at a velocity equal to the welding
material stuck to the tool surface. Two different material defi- speed, and at ambient temperature.
nitions are available within the FLUENT solver. Firstly, there Both isothermal and non-isothermal models were solved.
is the fluid region where both the momentum and heat equa- The isothermal models used a representative temperature of
tions are solved. Secondly, there is the solid region where only 800 K, while the non-isothermal models included the heat
the heat equation is solved. This is used where the material equation and calculated the heat generation from viscous
being modelled is either a solid, or where material movement dissipation. In these models, the heat loss through the tool
is sufficiently small that solution of the momentum equa- was calculated by defining a 1-dimensional conduction zone
322 P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327

at the shoulder. This calculation assumed that the far surface


was at ambient, and the material was 100 mm thick tool steel.
To model the imperfect contact between the aluminium and
backing plate a contact resistance of 1000 mm2 K/W from
Colegrove et al. [17] was used. The temperature on the bot-
tom of the backing plate far wall was set to ambient. Finally,
a convective heat transfer coefficient of 10 W/m2 K [17] was
used for the convective heat loss from the top surface of the
plate. Some of the models use a raked region around the tool
shoulder so that the tool could be inclined away from the
direction of travel. This is described in greater detail in Sec-
tion 2.4.

2.3. Material properties

The aluminium alloys 5083 and 7075-T6 aluminium


alloys were used for the analysis. The material property of
most interest was the viscosity, which was found from the
normal stress and strain-rate via the relationship described
in Appendix A. The flow stress of 5083 was found from the Fig. 3. Dimensions of the 5651 tool from Dawes et al. [4] used for the
Sellars and Tegart [18] constitutive law: models.
 
Q
Z = ε̇ exp = A(sinh ασ)n (1) because it avoids large changes in viscosity. However, the
RT strain-rate sensitivity at high temperature diminishes to zero
where ε̇ is the strain-rate (/s), Q the activation energy (J/mol), with this relationship. Experiments on 7075 aluminium alloy
R the universal gas constant (J/kg K), T the temperature (K) have shown that this is not the case. Nevertheless, good pre-
and A, α, n are material constants. dictions can be made with models that use this constitutive
Values of Q = 161 kJ/mol, A = 1.62/s, α = 0.019 /MPa and behaviour.
n = 3.2 from Sheppard and Tutcher [19] were used for the The thermal properties used for the models are shown in
analysis. Table 1. The values are independent of temperature in all
The flow stress for 7075-T6 was found by interpolating cases except for the specific heat and thermal conductivity
experimental stress versus strain-rate data from Jin et al. of the aluminium alloys where temperature-dependent values
[20]. The values were interpolated on a logarithmic scale from Mills [22] were used. Note that the values from Mills for
with strain-rate and on a linear scale with temperature. Both 5083 were actually for 5182, which has a similar composition.
of these relationships were implemented in FLUENT with
user-defined code. Importantly, unlike Seidel and Reynolds 2.4. Model types
[21], neither viscosity relationship included material soften-
ing near the solidus, which led to the over-prediction of weld The models generated were based on the design of the
power and temperature reported in subsequent sections. 5651 tool developed by Dawes et al. [4]. Fig. 3 shows the
Including material softening in FLUENT is difficult because dimensions of this tool. Several steps were required to gen-
the large change in viscosity near the solidus can lead to erate a mesh of the material around the tool. Firstly, a solid
an unstable solution. Further development in FLUENT model of the tool was created in Mechanical Desktop [25].
used a slip boundary condition to capture the low viscosity This was exported into the FLUENT mesh generation pro-
behaviour. gram GAMBIT [26]. In GAMBIT, the geometry was mod-
Other authors [6,14,15] have used the Johnson and Cook ified around the bottom and top of the tool threads to avoid
[7] constitutive model, which is more numerically robust the generation of highly distorted elements. Finally, regions

Table 1
Thermal properties of the materials used in the thermal model
Material name Density (kg/m3 ) Specific heat (J/kg K) Thermal conductivity (W/m K)
5083/7075-T6 Aluminium 2660/2800b Millsa Millsa
Steel (0.5% C steel) at 150 ◦ C 7833c 465c 50c
H13 tool steel (5% chrome) at 300 ◦ C 7833c 460c 36c
a Ref. [22].
b Ref. [23].
c Ref. [24].
P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327 323

Table 2 model’. The other models are described where their results
Summary of the welding conditions used for the models are significantly different.
Rotation speed Welding speed (mm/min)
(rpm) 3.1. Arrow plots and deformation region
60 90 120
750 5651-raked
500 5651-raked 5651-raked 5651-raked Fig. 4 shows vector plots of in-plane velocity on horizontal
5651-normal (+ a model planes at positions z = 0.1, 3.175 and 6 mm from the bottom
that uses 7075 aluminium face, for the standard model. Referring to Fig. 4a, the material
with isothermal under the dome of the tool (root) is influenced by the rotat-
properties)
Unthreaded tool
ing tool above. At mid-thickness (Fig. 4b) there is a large
250 5651-raked rotating region around the tool. Overlayed on this image is a
Both isothermal (800 K) and thermal models were solved using the properties
line where the strain-rate equals 2/s, which approximates the
of 5083 aluminium alloy. boundary between the rotating and stationary flow regions.
Finally, Fig. 4c shows the rotating material under the shoul-
der. Note that in all cases, there is a stagnation point on the
were created around the tool in accordance with Section 2.1, advancing side, and all the material up to this point flows
which were then meshed. round the retreating side of the tool.
Two models of the 5651 tool were generated: in one the Fig. 5 shows a longitudinal vertical section that compares
tool was normal to the workpiece (5651-normal) and in the the size of the deformation regions between an isothermal
other it was raked at 2.5◦ away from the direction of travel and a thermal model. Interestingly very little difference is
(5651-raked). Several alternative methods for raking the observed between the two, which contrasts with the dramatic
tool were attempted, however, the mesh shown in Fig. 2 reduction observed in a model for 7075-T6 that included a
which used a ‘prow’ around the shoulder was the only one thermal profile in Colegrove and Shercliff [27]. In the latter
that was successful. A final model (unthreaded tool) was an work, a large increase in material strength occurs outside
unthreaded version of the ‘5651-normal’ model described the deformation region, but this is far more gradual with the
above. Sellars and Tegart material properties for 5083 used in the
As seen in Fig. 2, the inner part of the tool protruded above current work.
the shoulder. This avoided the creation of highly distorted Streamlines of the flow around the tool are shown in Fig. 6.
elements around the fillet where the pin meets the shoulder. Note that streamlines are instantaneously tangential to the
Tetrahedral elements were used and a spacing of 0.3 mm was flow and are shown for a particular tool orientation. Parti-
used on the surface of the tool. cle tracks are much more difficult to calculate and require
In the ‘5651-raked’ model the plate thickness was a full transient analysis. The planar view (Fig. 6a) demon-
6.35 mm, while the other two models where the tool was strates how the material is swept around the retreating side
normal to the surface used a plate thickness of 6.95 mm. of the tool. The deformation zone is large compared with 2-
This is because the tool protrudes further into the plate when dimensional models [28], experimental results [29,30] and
it is normal to the surface. All the models used a back- later modelling work [27]. The deformation zone is larger
ing plate with a thickness of 40 mm and width of 140 mm. than the 2-dimensional models because of the additional
Finally, all the models extended 70 mm ahead of the weld and stirring induced by the shoulder. The later modelling work
100 mm behind. In this preliminary study, tests on the influ- uses a different constitutive model and includes slip at the
ence of mesh density and the effect of plate size were not tool/material interface. The large tetrahedral mesh used for
undertaken. the model may also expand the flow region.
Table 2 describes the models and welding parameters used. The longitudinal streamlines in Fig. 6b show very little
Note that unless otherwise stated, two versions of each were deviation in height as the material passes the tool. This is
solved: one was an isothermal (800 K) model and the other in contrast to the quite large vertical movement of mate-
was a thermal model. Both used the material properties of rial observed by Colligan [31]. The problem is caused
5083 aluminium alloy. The 7075 properties were only used by the model’s over-prediction of the deformation region
for one isothermal model, since the full thermal model could size.
not be solved due to the high heat generation and excessive
temperatures produced with these properties. 3.2. Pressure plot

The pressure distribution showed how raking the tool


3. Results increased the pressure behind the tool, but gave little insight
into the effect of tool shape on the process. Further details
The results primarily focus on the thermal model of the are provided in Colegrove [32]. Force comparisons, which
‘5651-raked’ model that used a welding speed of 90 mm/min are discussed below are more easily quantified and provide
and a rotation speed of 500 rpm. This is the so-called ’standard more practical guidance.
324 P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327

Fig. 5. Comparison of deformation zones between isothermal and thermal


versions of the 5651-raked model.

3.3. Temperature, force and power comparison

The temperature contours from the raked model in Fig. 7


demonstrate how the contact resistance inhibits the heat flow
to the backing plate. The peak temperature of all the thermal
models that used a welding speed of 90 mm/min and a rotation
speed of 500 rpm was approximately 860 ◦ C, which is well in
excess of the welding temperature of 550 ◦ C expected for this
material [17]. The model type had little effect on this value.
As discussed in Colegrove and Shercliff [33], one possible
reason for this discrepancy is that in a real weld slip may occur
between the tool and welded material, which reduces the heat
input and avoids material melting. An alternative explanation
suggested by Seidel and Reynolds [21] is that the material
significantly softens at temperatures approaching the solidus
which reduces the heat generation and the weld temperature
predicted by the model. Either effect will significantly reduce
the heat generation.

Fig. 4. Arrow plots across the (a) z = 0.1, (b) z = 3.18 and (c) z = 6 planes for
the 5651-raked thermal model. Note: (b) includes an overlay of the strain-
rate = 2/s boundary; although advancing and retreating side labels are shown
only for (a) they apply equally to (b) and (c). Fig. 6. Streamlines for 5651-raked thermal model (a) planar from a start
height of 3 mm; (b) transverse longitudinal section.
P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327 325

Fig. 7. Temperature contours for 5651-raked model. (Units of temperature


are ◦ C.)

Fig. 8a shows the effect of the welding parameters on the


peak temperature. The welding speed had little effect, while
increasing the rotation speed caused a significant increase in
the peak temperature. Obviously, this is caused by the greater
heat production from the higher rotational speed. Experi-
mentally, it has been shown that when welding 7075-T6
[33] and 6061 [34] aluminium alloys, the effect of chang-
ing the welding or rotation speeds on the peak temperature is
small, for conditions that produce sound welds. Note that the
effect of changing the rotation speed is greater at low rotation
speeds.
The weld power was found by multiplying the torque
by the rotation speed. The power input from translation is
approximately 1% of this value and can be ignored. The
results showed that the weld power was independent of model
type with 3.3 kW required for the thermal models. Fig. 8b
shows the effect of the welding parameters on the weld power.
These results reflect the peak temperature results in Fig. 8a in
that increasing the rotation speed increases the power input
to the weld, while the welding speed had little impact. The
Fig. 8. Comparisons between (a) peak temperature, (b) weld power and (c),
latter result suggests that, because of the relatively low weld- traversing force for the 5651-raked tool.
ing speed being used, the process is largely axi-symmetric
with the heat being primarily lost through the tool, backing
plate and by convection. The heat required to raise the mate-
rial to weld temperature is a relatively small component. This the isothermal model. This was a consequence of the lower
model effect is not reflected in practice [32,35]. The rotation flow stress at the higher temperature. Nevertheless, both are
speed had a much stronger influence on the heat input for a factor of 10 lower than the experimental values from John-
the isothermal models because there was no softening effect, son [37] for this alloy and tool type. It is believed that this
i.e., in the thermal models the flow strength reduces as the result could be improved by using a model that either included
temperature increases which offsets the increase in heat from viscosity softening near the solidus or material slip. Both
the higher rotation speed. would reduce the bulk movement of material around the tool,
Fig. 8c shows the effect of the welding parameters on increasing the traversing force.
the traversing force. It increased with welding speed because Finally, some interesting results were obtained when com-
of the greater amount of material needing to pass the tool. paring the forces between the three models. Fig. 9 shows how
The traversing force reduced with increasing rotation speed, the traversing force with the ‘5651-raked’ model was consid-
which is not observed experimentally [33], with the opposite erably lower than the models where the tool was normal to the
often being observed [36]. This demonstrates the inadequa- surface. The main cause of this effect is the pressure redistri-
cies of the current modelling approach. The traversing forces bution behind the tool, i.e., the higher pressure caused by the
with the thermal models were slightly lower than those with forging of material lowers the force needed to push the tool
326 P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327

of these problems can be addressed in FLUENT by either


using a slip model or changing the viscosity so that its value
reduces dramatically near the solidus temperature.

Acknowledgements

The authors wish to express their thanks to the Uni-


versity of Adelaide, The Welding Institute (TWI), the UK
Department of Trade and Industry (DTI), the University of
Cambridge, the Post-Graduate Training Partnership (PTP),
and the Cambridge Commonwealth Trust for their financial
support. The advice given by Philip Threadgill was greatly
Fig. 9. Force comparison between the different tool types. All models used
appreciated.
a welding speed of 90 mm/min and a rotation speed of 500 rpm.

through the material. It is unclear whether the lower force is Appendix A. Determining the viscosity from the
real or an artefact of the model. normal stress and strain-rate
The unthreaded tool had a lower traversing force than
the threaded version. This result is consistent with the 2- The equation for the effective stress is given by Bathe [38]:
dimensional stick model analysis in Colegrove and Shercliff 
[28], which showed that tools with features had a higher 3
traversing force than their cylindrical counterparts. Slip mod- σ̄ = S·S (A.1)
2
els of the same profiles indicated that the tools with features
assisted the movement of material round the tool, which low- where S is the deviatoric stress tensor. The equation for the
ered the traversing force. The latter result is more consistent effective strain-rate is given by Bathe [38]:
with experimental findings. Finally, the model that used 7075 
2
properties had an even higher traversing force due to the ¯ε̇ = ε̇ · ε̇ (A.2)
3
higher flow stress of this material.
where ε̇¯ is the strain-rate tensor given by:
1 
4. Conclusions ε̇ = ∇v + (∇v)T (A.3)
2
This work has demonstrated how the CFD package, FLU- where v is the velocity vector. The equation relating the devi-
ENT can be used to analyse material flow in FSW. The atoric stress tensor to the strain-rate is given in FLUENT [16].
models primarily used a threaded tool, which was raked 2.5◦ (Note that the bulk viscosity term is ignored.)
away from the direction of travel. This model was compared
against ones that used a normal threaded tool and a normal S = 2µε̇ (A.4)
unthreaded tool. Two model types were solved: an isothermal where µ is the viscosity. The deviatoric stress is defined as:
model and one that included the viscous heat generation and
solved the resulting thermal profile. S = ␴ + pI (A.5)
All the models predicted that the material in line with the
deformation zone was swept around the retreating side of the where ␴ is the stress and p is the pressure (positive in com-
pin. The amount of material swept around the pin increased pression). Note that p is defined as:
at locations closer to the shoulder. It was noted, however, that 1
the size of the deformation zone was much larger than that σm = σii = −p (A.6)
3
observed experimentally.
In addition to the excessive deformation zone size, the Substitute equation (A.4) into (A.1):
models experienced several additional problems. Firstly, 
σ̄ = 6µ2 ε̇ · ε̇
using representative material properties for 5083 and 7075
(without viscosity softening near the solidus), there is a large 
over-prediction of the weld temperature and the weld power. 2
σ̄ = µ 9 × ε̇ · ε̇
Secondly, the rotation speed had a large impact on the peak 3
welding temperature, which is relatively small in practice. Substituting from equation (A.2) gives:
Thirdly, the traversing force was found to be an order of
magnitude lower than that measured experimentally. Many σ̄ = 3µε̇¯
P.A. Colegrove, H.R. Shercliff / Journal of Materials Processing Technology 169 (2005) 320–327 327

[16] FLUENT, Release 6.0.20, Fluent Incorporated, Lebanon, NH,


Hence, the viscosity is given by:
2002.
σ̄ [17] P. Colegrove, M. Painter, D. Graham, T. Miller, 3-Dimensional flow
µ= . (A.7) and thermal modeling of the friction stir welding process, in: Pro-
3ε̇¯
ceedings of the Second International Symposium on ‘Friction Stir
Welding’, TWI Ltd., Gothenburg, Sweden, 2000.
[18] C.M. Sellars, W.J.Mc.G. Tegart, Int. Met. Rev. 17 (1972) 1–24.
References [19] T. Sheppard, M.G. Tutcher, Effect of process parameters on structure
and properties of Al–5Mg–0.8Mn alloy (AA5456), Met. Tech. 8
[1] W.M. Thomas, E.D. Nicholas, J.C. Needham, M.G. Murch, P. (1981) 319–327.
Temple-Smith, C.J. Dawes, Friction stir butt welding, Int. Patent [20] Z. Jin, W.A. Cassada, C.M. Cady, G.T. Gray, Mechanical Response
Application no. PCT/GB92/02203; GB Patent Application no. of AA7075 Aluminium Alloy over a Wide Range of Temperatures
9,125,978.8, 1991; US Patent no. 5,460,317, 1995. and Strain Rates, vol. 331–337, Material Science Forum, Switzer-
[2] W.M. Thomas, E.D. Nicholas, J.C. Needham, M.G. Murch, P. land, 2000, pp. 527–532.
Temple-Smith, C.J. Dawes, Improvements relating to friction weld- [21] T.U. Seidel, A.P. Reynolds, A 2-D friction stir welding process model
ing, International Patent Classifications B23K 20/12, B29C 65/06, based on fluid mechanics, Sci. Technol. Weld. Joining 8 (3) (2003)
1993. 175–183.
[3] H.R. Shercliff, P.A. Colegrove, Modelling of friction stir welding, in: [22] K.C. Mills, Recommended Values of Thermophysical Properties for
H. Cerjak, H.K.H.D. Bhadeshia (Eds.), Mathematical Modelling of Commercial Alloys, Woodhead Publishing Ltd., Cambridge, UK,
Weld Phenomena 6, Maney Publishing, London, 2002, pp. 927–974. 2002.
[4] C.J. Dawes, P.L. Threadgill, E.J.R. Spurgin, D.G. Staines, Develop- [23] J.R. Davis, ASM Speciality Handbook: Aluminium and Aluminium
ment of the new friction stir technique for welding aluminium, TWI Alloys, ASM International, 1993.
Group Sponsored Report no.: 5651/35/95, Cambridge, UK, 1995. [24] J.P. Holman, Heat transfer, in: SI Units, seventh ed., McGraw-Hill
[5] G.J. Bendzsak, T.H. North, C.B. Smith, An experimentally vali- Book Company, London, 1990.
dated 3D model for friction stir welding, in: Proceedings of the [25] Mechanical Desktop 5, Autodesk® , San Rafael, CA, 2000.
Second International Symposium on ‘Friction Stir Welding’, TWI [26] GAMBIT, Release 2.0.4, Fluent Incorporated, Lebanon, NH,
Ltd., Gothenburg, Sweden, 2000. 2002.
[6] A. Askari, S. Silling, B. London, M. Mahoney, Modeling and anal- [27] P.A. Colegrove, H.R. Shercliff, Development of the TrivexTM fric-
ysis of friction stir welding processing, in: K.V. Jata, et al. (Eds.), tion stir welding tool. Part II. 3-Dimensional flow modelling, Sci.
Friction Stir Welding and Processing, TMS, Warrendale, PA, 2001, Technol. Weld. Joining 9 (4) (2004) 352–361.
pp. 43–54. [28] P.A. Colegrove, H.R. Shercliff, 2-Dimensional CFD modelling of
[7] G.R. Johnson, W.H. Cook, A constitutive model and data for metals flow round profiled FSW tooling, Sci. Technol. Weld. Joining 9 (6)
subjected to large strains, high strain rates and high temperatures, in: (2004) 483–492.
Proceedings of the Seventh International Symposium on ‘Ballistics’, [29] A.P. Reynolds, Visualisation of material flow in an autogenous
The Hague, Netherlands, 1983, pp. 541–548. friction stir weld, Sci. Technol. Weld. Joining 5 (2) (2000) 120–
[8] P. Heurtier, C. Desrayaud, F. Montheillet, A thermomechanical anal- 124.
ysis of the friction stir welding process Materials Science Forum, [30] B. London, M. Mahoney, W. Bingel, M. Calabrese, D. Waldron,
vol. 396–402, Trans Tech Publications, Switzerland, 2002, pp. Experimental methods for determining material flow in friction stir
1537–1542. welds, in: Proceedings of the Third International Symposium on
[9] S. Xu, X. Deng, Two and three-dimensional finite element models ‘Friction Stir Welding’, TWI Ltd., Kobe, Japan, 2001.
for the friction stir welding process, in: Proceedings of the Fourth [31] K.J. Colligan, Material flow behaviour during friction stir welding
International Symposium on ‘Friction Stir Welding’, TWI Ltd., Park of aluminium, Weld. J. 78 (7) (1999) 229–237.
City, Utah, May 2003. [32] P. A. Colegrove, Modelling of Friction Stir Welding, Ph.D. Thesis,
[10] P. Dong, F. Lu, J.K. Hong, Z. Cao, Coupled thermomechanical anal- The University of Cambridge, Cambridge, UK, 2003.
ysis of friction stir welding process using simplified models, Sci. [33] P.A. Colegrove, H.R. Shercliff, Experimental and numerical analysis
Technol. Weld. Joining. 6 (5) (2001) 281–287. of aluminium alloy 7075-T7351 friction stir welds, Sci. Technol.
[11] P. Ulysse, Three-dimensional modeling of the friction stir-welding Weld. Joining 8 (5) (2003) 360–368.
process, Int. J. Mach. Tools Manufacture 42 (2002) 1549–1557. [34] M. Song, R. Kovacevic, Numerical and Experimental Study of the
[12] F. Palm, U. Henneböhle, V. Erofeev, E. Earpuchin, O. Zaitzev, Heat Transfer Process in Friction Stir Welding, Proc. Inst. Mech.
Improved verification of FSW-process modelling relating to the Eng. Part B: J. Eng. Manufacture 1 (2003) 73–85.
origin of material plasticity, in: Proceedings Fifth International Sym- [35] A.P. Reynolds, W. Tang, Alloy, tool geometry and process parameter
posium on ‘Friction Stir Welding’, TWI Ltd., Metz, France, 2004. effects on friction stir weld energies and resultant FSW joint proper-
[13] L. Fourment, S. Geurdoux, M. Miles, T. Nelson, Numerical sim- ties, in: K.V. Jata, et al. (Eds.), Friction Stir Welding and Processing,
ulation of the friction stir welding process using both Lagrangian TMS, Warrendale, PA, 2001, pp. 15–23.
and Arbitrary Lagragian Eulerian formulations, in: Proceedings of [36] J. Yan, M.A. Sutton, A.P. Reynolds, Process-structure-property rela-
the Fifth International Symposium on ‘Friction Stir Welding’, TWI tionships for nugget and HAZ regions of AA2524-T351 FSW Joints,
Ltd., Metz, France, 2004. in: Proceedings of the Fifth International Symposium on ‘Friction
[14] H. Schmidt, J. Hattel, Modelling thermomechanical conditions at the Stir Welding’, TWI Ltd., Metz, France, 2004.
tool/matrix interface in friction stir welding, in: Proceedings of the [37] R. Johnson, Forces in friction stir welding of aluminium alloys—
Fifth International Symposium on ‘Friction Stir Welding’, TWI Ltd., further studies, TWI Core Research Report no. 716/2000, Cambridge,
Metz, France, 2004. UK, 2000.
[15] H. Schmidt, J. Hattel, A local model for the thermomechanical con- [38] K.J. Bathe, Finite Element Procedures, Prentice Hall, Englewood
ditions in friction stir welding, Modell. Simul. Mater. Sci. Eng. 13 Cliffs, NJ, 1996.
(2005) 77–93.

You might also like