Sie sind auf Seite 1von 10

Two-dimensional CFD modelling of flow round

profiled FSW tooling


P. A. Colegrove and H. R. Shercliff

This paper describes the application of the computational fluid dynamics (CFD) code, FLUENT, to
modelling the two-dimensional metal flow in friction
stir welding (FSW). The primary goal is to assist in
the development of new welding tools, though the
models also improve the understanding of the deformation mechanism in FSW, and enable a first order
visualisation of the flow round the probe. The paper
describes a quantitative method of comparing the flow
round different tool shapes. A novel slip model was
developed, where the interface conditions were governed by the local shear stresses. This revealed
significant differences in behaviour when compared
with the common assumption of material stick. The
two-dimensional model has demonstrated the viability
of the approach for investigating the flow round
practical tool shapes, and gives confidence that the
development of more computer intensive threedimensional models will be justified.
STWJ/413
Keywords: frictions stir welding, modelling, material
deformation
Drs Colegrove (pac44@cam.ac.uk) and Shercliff
(hrs@eng.cam.ac.uk) are in the Dept. of
Engineering, Cambridge University, Trumpington
Street, Cambridge CB2 1PZ, UK. At the time this
work was carried Dr Colegrove was also with TWI
Ltd, Granta Park, Great Abington, Cambridge,
CB1 6AL, UK. Manuscript received 16 January
2003; accepted 12 June 2003.
# 2004 Institute of Materials, Minerals and Mining.
Published by Maney on behalf of the Institute.

INTRODUCTION
Friction stir welding (FSW) is a relatively new welding
process, patented in 1991 by Thomas et al.1,2 This process
has great advantages in welding aluminium alloys that are
difficult to weld. The process gives low distortion, can weld
thick sections in a single pass and produces welds with
excellent mechanical properties.
A schematic diagram illustrating the process of FSW is
shown in Fig. 1. The key components of the FSW tool are:
(i) The Shoulder. This is the primary means of
generating heat during the process, and it prevents
material expulsion and assists material movement
around the tool
(ii) The Pin. The pins primary function is to deform the
material around the tool and its secondary function
is to generate heat.
Macrosections from friction stir welds exhibit four
distinct regions. First, there is the dynamically recrystallised
region (DRZ) where the material is recrystallised from
the high temperature deformation. Outside this is the
DOI 10.1179/136217104225021832

thermo-mechanically affected zone (TMAZ) where the high


temperature and strain rates are not sufficient to recrystallise the material. Finally, there is the heat affected zone
(HAZ) and the parent material which are not deformed.
Weld cross-sections are not symmetrical owing to the tool
rotation. Figure 1 defines the advancing side and
retreating side where the rotation opposes and coincides,
respectively, with the direction of welding.
Modelling of friction stir welding (FSW) has been
reviewed by Shercliff and Colegrove.3 Friction stir models
to date have focused on thermal field prediction, with some
studies leading to prediction of microstructural evolution,
properties and residual stress. A much greater modelling
challenge is the prediction and validation of the threedimensional flow field round the profiled tool. While both
the shoulder and pin are important to the process, a
successful tool is primarily dependent on a well-designed
pin. The three-dimensional flow is complex, but considerable
insight can be obtained by first considering two-dimensional
flow round profiled rotating pins. This captures the
interaction with features, which are predominantly aligned
with the pin axis (such as flutes) and is computationally more
rapid. Therefore, a two-dimensional model of the flow round
a rotating pin has been used to examine and compare the
flow round four different pin profiles.

DESCRIPTION OF MODEL
Basic model
A computational fluid dynamics package, FLUENT,4 was
selected for the modelling work because it enabled
modelling of the high strain rate flow beside the tool and
permitted the application of a slip boundary condition.
Figure 2a shows the standard set-up for the models. The
physical size of the model was kept small to keep solution
times low. The model is divided into two regions: a rotating
region beside the tool where the mesh moves at the rotation
speed of the tool and a surrounding stationary region,
which models the flow in the far field. All walls on the
boundary of the stationary flow region are given a velocity
equal to the welding speed. Note that the material flows
from left to right and the tool rotates in a clockwise
direction in all the models. The four pin profiles that were
examined are shown in Fig. 2b.

Schematic diagram of friction stir welding

Science and Technology of Welding and Joining

2004

Vol. 9

No. 6

483

484

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

a Description of model subregions and b diagram of


four tool profiles

7075-T6 aluminium alloy was selected for the modelling


work. To simplify the problem, a temperature of 527uC was
used throughout the model. Since the flow stress of 7075 at
high temperature is not greatly affected by the temperature5
and the temperature variation across the intense deformation zone is relatively small, the assumption is reasonable
for a first analysis and enables faster solution times. The
material property of most interest is the viscosity, which
was interpolated from experimental stress versus strain-rate
data produced by Jin et al.5 and implemented in FLUENT
through user-defined code. The mesh size and number of
iterations were checked to ensure that the solution had
converged.

by material slip against the tool, i.e. different tool


materials exert different shear stresses on the
welded material.
(iv) In the section Traversing force and welding
torque comparison below, it is shown that
different tool profiles cannot be discriminated in
terms of forces and torques when it is assumed that
the material sticks to the tool. Obviously, the tool
shape has a significant effect on the success or
otherwise of the process and Colegrove et al.10
showed that it affected the traverse and down
forces.
In the slip models, the surface shear stress was limited to
some arbitrary value. Therefore two different boundary
conditions can exist simultaneously on different parts of the
pin:
(i) Stick condition. This occurred where the shear stress
was below the limiting shear stress. On a profiled
tool, this usually occurred in the tool flutes.
(ii) Slip condition. Where the shear stress necessary for
a stick condition exceeded the limiting value, the
applied shear stress was truncated to the limiting
value, which resulted in the material slipping across
the tool surface.
Both of the above conditions were implemented in
FLUENT by specifying a shear stress at the boundary.
This was particularly complex for the stick condition where
the shear stress had to be adjusted so that the velocity of the
material equalled the tool velocity. The Appendix describes
how this was implemented in FLUENT.
An alternative to the slip boundary condition used by
Long et al.11 assumed that the viscosity dropped exponentially to zero at temperatures near the solidus. In this case, a
stick boundary condition is used, and the slipping region is
modelled as a very thin boundary layer. The slip model
approach, although cruder, enables the use of a coarser
mesh enabling faster solution times.

Model types
Slip model
Most of the published flow models have assumed that
material is stuck to the tool surface. Although Xu et al.6
demonstrated a slip boundary condition with a cylindrical
two-dimensional tool, such a boundary condition has not
been demonstrated on profiled tools.
There are several reasons that justify the development of
the slip model
(i) Based on the material properties for 7075-T6
provided by Jin et al.,5 a three-dimensional stick
model of the process (without significant softening
near the melt temperature) requires a power input
of 1518 kW for the 6.35 mm welds described in
Colegrove and Shercliff.7 This compares with the
measured heat input of 2.63.5 kW. Therefore the
model significantly overpredicts the power requirement. It is believed that one of the causes of this
discrepancy is the assumption that all the material
sticks to the tool surface.
(ii) The temperature measurements in Colegrove and
Shercliff7 and Song and Kovacevic8 show that the
peak temperature near the tool approached the
solidus. Therefore it is likely that a thin partially
liquid layer exists at the tool/material interface
which reduces the shear stress that can be applied
to the material. This phenomenon has been
discussed extensively by North et al.9
(iii) Colegrove and Shercliff7 demonstrated that the
tool material type affected the weld temperature
and quality. Since heat loss through the tool is
quite minimal (thermal models suggest 2% in
steady state), this observation is best explained
Science and Technology of Welding and Joining

2004

Vol. 9 No. 6

All the models used a rotation speed of 250 rev min1 and a
welding speed of 300 mm min21, which are typical conditions for welding thick (16 mm) 7075 aluminium alloy.
Each of the four tool types was examined with a stick
boundary condition, and a slip boundary condition where
the shear stress was limited to 40 or 80 MPa. These values
are used for demonstration purposes only and compare
with a typical shear strength of 90 MPa (approx) for the
unsoftened material at this temperature.
Two model types were analysed. The first was a steadystate model where the solution is a snap-shot of the flow at
a particular instant in time. Provided the time-dependent
terms in the NavierStokes equation remain relatively
small, this approximation will be reasonably good. The
second type of model was a full transient analysis and was
used only for the mixing models and the force analyses.

MODELLING RESULTS AND DISCUSSION


Deformation region size and flow pattern
The size of the deformation zone was found by plotting the
surface where the strain rate equalled 2 s21. The choice of
2 s21 was arbitrary, nevertheless a comparison of this
surface with a vector plot of the flow field in Fig. 3a and b
for Tool_1 shows that it approximates the boundary
between the rotating material beside the tool and the
largely undeformed material in the far field.
The deformation region for the slip model is much
smaller on the advancing than retreating side. A close
inspection shows that the deformation region on the
retreating side is greater for the slip model than for
the stick model (6.2 versus 5.7 mm). For the slip model,

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

485

Results for Tool_1: velocity vectors and strain rate52 s21 contour for a stick model and b slip model (limiting shear
stress540 MPa); streamline plots for c stick model and d slip model (limiting shear stress540 MPa)

the distance between the material at the tool surface and the
undeformed region is particularly small on the advancing
side. Nevertheless, the same amount of material still needs
to get past the pin so the deformation region enlarges on the
retreating side. In the stick model, the tool is able to stir
the material more effectively.
A streamline plot of the stick model is shown in Fig. 3c
and for the slip model in Fig. 3d. Note that streamlines are
instantaneously parallel to the flow and are shown for a
particular tool orientation. Particle tracks are much more
difficult to calculate and require a full transient analysis,
however they give a better indication of the flow around the
tool. An attempt to implement particle tracks in FLUENT
was unsuccessful. Both streamline plots are similar to those
reported in Siedel and Reynolds.12 The streamlines are
much nearer the tool for the slip model. In the stick model,
there is an essentially stagnant region, which (although
shearing) does not permit the entry or exit of flow material
because it is shearing parallel to the tool surface. (Note that
flow material refers to the material which flows past the
tool.) This region of parallel shear does not occur with the
slip model and hence the flow material can come into
contact with the tool.
Figure 4a and b show an arrow and deformation plot of
the flow for Tool_4 at two orientations. The flow around
Tool_4 is less circular than that observed for Tool_1,
however it is broadly the same size. There is however
greater perturbation around the features, with the flow
being more dependent on the tool orientation. The
perturbation around the flats leads to the bulging of the
streamlines seen in Fig. 4c and d. The velocity field is
investigated later, but in essence, the material beside the
flats is stuck to the tool surface and therefore has a high
rotational velocity. In comparison the velocity of the
material at the outer circumference is near zero. This
causes the enlargement of the flow field around the flats and
the bulging of the streamlines in this area.

Comparison of model flow pattern with experimental


results
The plots of the streamlines show that the material in the
path of the pin is swept round the retreating side of the tool.
This characteristic of the model is supported by flow
visualisation experiments by London et al.13 and Guerra
et al.14 An image from London et al. is reproduced in
Fig. 5a. This indicates that the deformation region is
greater on the retreating side of the pin. A particle that is
in line with the advancing edge of the pin will flow past the
advancing side, rather than being swept around the
retreating side. This observation is more consistent with
the flow pattern indicated by the slip model.
The streamlines in Fig. 3c and d and Fig. 4c and d are
shaded according to the particle residence time, i.e. the
particles at the transition in colour were initially in a
straight line transverse to the flow direction. This enables
the final position of this plane to be viewed after it has been
swept round the tool. These results can be compared with
the transverse marker experiment in Fig. 5b from
Dickerson et al.15 which is similar to marker experiment
work by Reynolds.16 In this experiment, a transverse
copper foil was placed in the 2024 aluminium before
welding. The image shown is an X-ray of the mid-section of
the weld and demonstrates two key features
(i) There is a bulge in the material behind the pin
(ii) Some material that started on the advancing side
was dragged ahead of the pin.
Note that ahead is in the direction of the tool and not the
material flow. The plots in Fig. 3c and d and Fig. 4c and d
show a similar bulge behind the tool, and the dragging of
material ahead of the pin on the advancing side. None of
the results indicated the abrupt boundary between
deformed and undeformed material that is indicated in
Fig. 5b. This is because the model has ignored several of the
real process characteristics, e.g. the elastic properties of the
material, the temperature gradient and microstructural
Science and Technology of Welding and Joining

2004 Vol. 9

No. 6

486

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

Results for Tool_4: a velocity vectors and strain rate52 s21 contour for Tool_4 slip model (limiting shear
stress540 MPa) and b rotated 60u; c streamlines for Tool_4 slip model (limiting shear stress540 MPa) and d rotated
60u

softening. Xu et al.6 and Siedel and Reynolds12 have shown


similar results in their 2D modelling work.
Finally, Fig. 5c and d shows horizontal planar sections
through a 2014-T6 weld and a dissimilar weld between
2024-T3 and cast AlSiMg, respectively. The latter shows
the microstructure after a stop-action test, i.e. the weld was
stopped and the pin unscrewed from the material to reveal
the flow while welding. Both figures show a very distinct
boundary between the recrystallised nugget microstructure
and the material in the Thermo-Mechanically Affected
Zone on the advancing side. The boundary between the two
is far less distinct on the retreating side, particularly for
Fig. 5c. This can be explained by the two-dimensional
model. The material that flowed round the tool and was
deposited on the advancing side underwent significant
deformation and was therefore recrystallised. The adjacent
material that flowed past the tool underwent relatively little
deformation and was therefore not recrystallised, i.e. the
boundary represents the surface where the divided flows
met and welded together. On the retreating side, the change
from the highly deformed to undeformed material is
continuous, so an abrupt boundary between the two is
not observed.

Onion ring investigations


FSW microstructures in the Dynamically Recrystallised
Zone (DRZ) are far from continuous and show features
which are periodic with the tool rotation. This has
traditionally been labelled the onion ring and is clearly
seen in the planar macrosection in Fig. 5c. While the grain
size within the DRZ is uniform in some alloys,17 others
reveal layers of coarse and fine grains.1820 It is proposed
that this feature is caused by cyclic or layered flow around
the pin, which causes high and low rates of deformation.
The higher the deformation, the finer the resulting grain
size from dynamic recrystallisation though this may be
masked in some alloys by static evolution of the microstructure in the wake of the weld. The models were analysed
Science and Technology of Welding and Joining

2004

Vol. 9 No. 6

to determine whether there were any flow characteristics


that suggested this type of behaviour.
For a preliminary qualitative analysis, consider the
streamline plots for the Tool_4 profile in Fig. 4c and d.
These show a significant difference in the final position of
the material for the two tool orientations. The analysis is
slightly unrealistic in that the final position of the material
is being determined from a single tool orientation in a
steady-state model. Nevertheless, the results show that the
final position of the material is highly dependent on the tool
orientation with a much flatter bulge being observed in
Fig. 4d. This result suggests a cyclic fluctuation in the
material throughput past the pin that may cause the onion
ring.
In the second analysis, the velocity in the x direction
along the weld centreline behind the tool was plotted for six
tool orientations and is shown in Fig. 6. The corresponding
vector plots for two of these tool orientations (1 and 4)
are shown in Fig. 4a and b. The x position is the distance
from the maximum periphery of the tool (6.91 mm), and
each profile is rotated 20u clockwise from the previous one.
This plot clearly shows the fluctuating flow in the x
direction. In orientation 1, very little material is deposited
due to the large negative velocity in the x direction
(material is being drawn in to the region over the tool
flat). The strain rate experienced by this material is also
high. In orientation 4 a large amount will be deposited as
material is being expelled from the region over the tool flat
by the advancing prow of the tool profile. The strain rate is
lower, potentially giving a contrast in grain size.
The final two methods involve the use of the transient
mixing model. This model numerically separated the flow
into two different zones with identical properties, and
analysed how these mixed as they flowed round the tool.
Figure 7 examines material that was separated along the
centreline and is similar to the dissimilar weld in Fig. 5d.
Figure 8 shows material that was separated in a transverse
direction and is similar to the transverse marker experiment

Colegrove and Shercliff

487

a Pathline markers reproduced from London et al.,13 b transverse copper marker in 2024 aluminium alloy mid-section
(courtesy T. Dickerson); c horizontal planar section through a FSW in 2014-T6 (scale on left shows 1 mm intervals);
d dissimilar weld between 2024-T3 and cast AlSiMg (courtesy T. Dickerson)

in Fig. 5b. Both of these models used the slip model with a
limiting shear stress of 40 MPa and were incremented 5u
per iteration. An extra fine mesh was used near the tool so
that the boundary between the two materials could be
tracked.
In models where the two materials were separated along
the joint line, the light material in the flutes was trapped
during start-up, i.e. the model started with the dark
material below the y axis and the light material above.
The discontinuous flow of material round the tool can be
clearly seen in the stick (Fig. 7a) and slip (Fig. 7b) models

Two-dimensional CFD modelling of flow round profiled FSW tooling

Material velocity in x direction along weld centreline


behind tool for slip model (limiting shear stress
540 MPa)

and is similar to the layers between the two materials


observed in the dissimilar weld in Fig. 5d. Even though the
layering is more distinct in the slip model than the stick, it
still does not match that observed experimentally. One of
the causes of this discrepancy is the size of the mesh used in
the model. With a finite mesh size, it is difficult to capture
the boundary between the two materials, especially where
the distance between the layers becomes small. Eventually
the layers collapse on the reverse side of the tool. A finer
mesh may resolve this problem. The second difference
between the two results is the progressive banding observed
around the tool. Although this is partially observable in
Fig. 5d, it is far more progressive in the model. This may be
another characteristic of using an isothermal model.
The horizontal distance in the x direction was measured
between the bands for the slip model and found to be
initially quite large but reduced to 0.4 mm near the end.
This corresponds to three bands for every revolution of the
tool, which matches the period observed experimentally
with Triflute tools.
In Fig. 8 where the material was separated in a transverse
direction, the layered material flow is seen even more
clearly. The final shape of the deposited material in Fig. 8c
is similar to that shown in Fig. 5b, however the model
deformation region size is clearly larger in part because of
the isothermal assumption in the model. Finally, the mesh
Science and Technology of Welding and Joining

2004 Vol. 9

No. 6

488

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

Understanding the slip boundary condition

Material mixing model using Tool_2; a stick model; b


slip model (limiting shear stress540 MPa). Plots show
material position after 18 revolutions

in the far field is much larger which blurs the banded


features (Fig. 8c).
This work has shown the importance of tool profiling to
promote intense shearing and unsteady flow. The flow is
cyclic with the number of features on the tool. Therefore
with three sided tools, the flow is cyclic with every third
revolution and a conventional threaded tool is cyclic with
every revolution. In the latter case, the threads also induce
vertical movement, however the flow is still largely
rotational with the material flowing around an eccentric
profile.

Figure 9 shows polar plots of the absolute velocities and


shear stresses for the slip and stick models that used the
Tool_4 profile. The absolute velocity plots in Fig. 9a and c
are overlayed with the tool velocity, i.e. the stick velocity.
Similarly, the shear stress plots are overlayed with the shear
stress for a stick boundary condition. To provide further
explanation of these plots, Fig. 10 shows the velocity and
shear stresses plotted along the tool surface for the 40 MPa
slip model, i.e. the plots are equivalent to Fig. 9a and b.
Note that the thin dashed line in the surface velocity plot in
Fig. 10a represents the surface velocity if stick were to
occur. This reduces in size toward the centre of the flat
owing to the reduction in radius, i.e. n5rv.
Three points on the profile were selected to help explain
the mechanics behind the model. Referring to Fig. 10,
point A is on the outer circumference of the tool where slip
occurs at the maximum shear stress of 40 MPa, so the
velocity of the material is relatively low, cf. Fig. 9a. At
Point B, the surface normal does not act through the tool
axis, and quite a different condition holds. The material
sticks to the surface and the shear stress is below the
limiting value of 40 MPa. The tool and material velocity
are then equal. Finally, at Point C slip occurs, but the
velocity of the material is much higher than at Point A
because the surrounding material is stuck to the tool. The
shear stress equals the limiting value of 40 MPa. Note that
slip occurs at Point C because the surface is aligned with the
flow.
Figure 9a shows that the slipping material has a slightly
higher velocity on the retreating than advancing side. The
shear stress is the same on both sides, but the material flow
on the retreating side is assisted by the bulk material flow
causing the higher velocity. The same plot for a model with
a limiting shear stress of 80 MPa is shown in Fig. 9c. In this
model, the material sticks along the entire length of the flat
while slipping against the outer circumference. The slipping
material has a much higher velocity than the previous
model as a result of the higher limiting shear stress. Once
again the material velocity is slightly higher on the
retreating side.
The corresponding plots of the resultant shear stress on
the tool surface are shown in Fig. 9b and d. Where the shear
stress is limited to 40 MPa (Fig. 9b), the shear stress equals
this value along most of the tool surface. The six small dips
correspond to where the material sticks to the flats. The
plot for the limiting shear stress of the 80 MPa model is
very different. Three large spikes are observed in the centre
of the flats. This shear stress is required to prevent material
slippage, as occurred in the 40 MPa model in this region.

Material mixing model using Tool_2 and transverse marker; a 4.9 rotations, b 9.9 rotations and c 19.4 rotations. All
are from slip model with limiting shear stress of 40 MPa

Science and Technology of Welding and Joining

2004

Vol. 9 No. 6

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

489

Surface velocity and shear stress results for Tool_4: a surface velocity and b resultant shear stress for slip model ()
with limiting shear stress of 40 MPa and stick model (- - -); c surface velocity and d resultant shear stress for slip
model () with limiting shear stress of 80 MPa and stick model (- - -)

Traversing force and welding torque comparison


Figure 11 shows the traversing force and welding torque.
The smaller the limiting shear stress, the larger the
traversing force and the lower the welding torque.
Limiting the shear stress reduces the tangential force at
the tool surface and hence the resultant torque. It is
believed that this reduced torque reduces the stirring
action of the tool, increasing the traversing force.
The difference in the torque value between steady state
and transient models is small. The variation over a onethird revolution in the transient model output is also
relatively small (134 N for the traversing force and less than
6 N m for the welding torque). Therefore the steady-state
model gives a good estimate of the transient torque.
The comparison between the welding torques in Fig. 11a
shows some interesting trends. The stick model and the
model that used a limiting shear stress of 80 MPa gave the
greatest torques for Tool_1 and the smallest for Tool_4.
Since all the material is sticking, (or almost) the tool shapes
that maximise the circumferential area maximise the
torque. The effect of pin shape is quite weak. Of the
models that have a limiting shear stress of 40 MPa, Tool_4
has the greatest torque and Tool_1 the least. In the case of
Tool_1, most of the material slips round the outer
circumference, which limits the torque. The other tools,
especially Tool_4 have a smaller cross-sectional area and a
much greater amount of trapped material that is moving
with the tool profile. This enables the application of a much
greater torque to the material.
The comparison between the traversing forces in Fig. 11b
shows that there is little or no difference in the traversing
force for the stick models. There is however a significant
difference for the models that have a limiting shear stress of
40 MPa. Tool_1 is seen to have the highest value, Tool_2
and Tool_3 have nearly identical values and Tool_4 has the
lowest traversing force. Interestingly, reversing the shaped

feature from Tool_2 to Tool_3 had no effect. The difference


in the traversing forces is a reflection of the difference in the
welding torques seen previously. The higher the torque,
the greater the stirring action of the tool and the lower the
traversing force. Note that subsequent work has shown a
deviation from this general rule, when using tools with
convex features, which is explained further in Colegrove
and Shercliff.10

CONCLUSIONS
The aim of the current work has been the development of a
model that enables the comparison of different tool shapes
and a better understanding of the science behind friction
stir welding. The application of the CFD code FLUENT,
and the development of a novel slip model have made
some significant progress toward that aim.
Good visualisation is provided by plots of the flow. In
particular, the streamlines correlated well with embedded
marker experiments. Qualitatively, the model compared
favourably, though the deformation zone size predicted by
the model was greater than that indicated by the experiments. Great quantitative accuracy is not sought given the
assumptions in the present model, i.e. isothermal conditions
and 2D flow. Finally, the mixing model indicated oscillating
flow which was thought to be the cause of the banded
onion-ring microstructure.
To demonstrate the modelling approach, four different
tool profiles were evaluated, with three types of boundary
conditions: the stick model and two slip models with
limiting shear stresses of 40 and 80 MPa, respectively. The
slip model revealed significant differences with different
tool shapes, which was not evident with the conventional
stick model. Comparison of the traversing force and
welding torque for the constant shear stress models
indicated that Tool_4 was the most efficient, followed by
Tool_2 and Tool_3 and finally Tool_1. It was considered
Science and Technology of Welding and Joining

2004 Vol. 9

No. 6

490

10

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

Quarter sections of a surface velocity, b velocity vectors and c shear stresses for slip model shown in Fig. 9

that Tool_4 had the ability to stir the material most


effectively.
Finally, the slip model provided some useful insights into
the process mechanics and suggested regions where the
material was likely to stick to the tool surface. The
technique, which will be extended to three dimensions and
experimentally validated, shows great potential for assisting
the development of new tool shapes.

Appendix
Applying the slip boundary condition in two dimensions
Defining shear stress and relative velocity in two
dimensions
Define three vectors
n~(xn ,yn ,0)
v
~(0,0,{1)
jvj
dtan ~n|

11

a Torque and b traversing force comparison for different tool profiles

Science and Technology of Welding and Joining

2004

Vol. 9 No. 6

v
~({yn ,xn ,0)
jvj

The first is the normal unit vector, which points toward the
tool surface. The second is the unit direction vector for the
rotation. Its value is negative for clockwise rotation.
Finally, dtan gives the direction of the shear stress, which
will always be parallel to the surface in the direction of
rotation. These vectors are shown in Fig. 12a.

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

491

The following steps were used when implementing the slip


boundary condition:
(i) Step 1. Solve the system using a stick boundary
condition and determine the shear stress at the
wall.
(ii) Step 2. If
jtj > tmax
then the shear stress at the wall is given by
equation (1) and Shear_Max is given a value of 1.
If
jtjftmax
then
t0 ~t
and Shear_Max will be given a value of 0.
(iii) Step 3. Solve for several iterations until the
solution is reasonably well converged (typically
500 after which the residuals which measure the
stability of the solution flatten).
(iv) Step 4. Continue iterating the solution while
updating the wall shear stress according to:
a. Case 1. If Shear_Max51 and
12

vrelt f0

a Description of normal vector and shear stress at


wall; b description of velocity vectors at tool surface

The shear stress at the wall is limited by the


maximum value and slip is occurring. There is
no need to change from this condition because
the material is slipping.
b. Case 2. If Shear_Max51 and

It will be assumed that the wall shear stress is limited to


some maximum value tmax, i.e.
jtwall jftmax

vrelt > 0

If twall equals tmax it will given by


twall ~tmax dtan :

: :

The shear stress at the wall is limited by the


maximum value from the previous step.
However, the shear stress is such that the
material at this position is actually moving
faster than the tool which is clearly impossible. Therefore it is necessary to change the
boundary condition so that the maximum
shear stress is no longer defined, i.e. set
Shear_Max50 and go to Case 3.
c. Case 3. If Shear_Max50, then the material is stuck to the tool and the shear stress
needs to be updated according to

: (1)

The relative velocity between the tool and adjacent material


is given by
vrel ~vm {vtool

: :

: (2)

where vtool5tool velocity (52r6w); vm5material velocity;


vrel5velocity of the material relative to that of the tool.
Note that this velocity must always be parallel to the
surface of the tool. This is shown diagrammatically in
Fig. 12b.

t0 ~tzCvrel

Slip boundary condition logic in two dimensions


A shear stress boundary condition is not only used where
the material is slipping but where it is stuck to the tool. This
is done because FLUENT does not allow the user to have a
boundary where two different boundary conditions are set,
i.e. stick where the material velocity is prescribed, and slip
where the shear stress is prescribed. Therefore where the
material is stuck to the tool, the shear stress must be set so
as to give zero relative velocity between the tool and
material.
This section uses the following variables in addition to
those defined in the first section of this Appendix:
(i) Shear_Max. Boolean variable, which has a value of
1 if the maximum shear stress is defined on the
surface and slip occurs, and 0 if a no-slip boundary
condition is defined.
(ii) vrelt. Variable that defines whether the relative
velocity is in the same or opposite direction to the
tool velocity. It is found by
vrelt ~vrel {vtool
If the material velocity is greater than the tool
velocity then vrel is in the same direction as the tool
and vrelt.0. Similarly, if it is less than or equal to the
tool velocity, then vrelt(0.

where C is a constant. If C is too large, the


system will go unstable, and if too small, the
system will take a long time to solve. vrel is
given by equation (2). Suitable values were
found by trial-and-error.
This equation forces the value of vrel to converge on zero,
by adjusting the wall shear stress based on the value of this
slip velocity. Where the new value of the shear stress
exceeds the maximum allowable, i.e.
jt0 j > tmax
then a slip boundary condition is applied by allowing
Shear_Max51. The shear stress is set as for Case 1.

ACKNOWLEDGEMENTS
The authors wish to express their thanks to the University of
Adelaide, The Welding Institute (TWI), the UK Department
of Trade and Industry (DTI), the Post-Graduate Training
Partnership (PTP), the University of Cambridge, and the
Cambridge Commonwealth Trust for their financial support.
The images provided by Terry Dickerson and Blair London,
and the advice given by Philip Threadgill, Chris Dawes,
Nikos Nikiforakis, Wayne Thomas, Mike Russell, and
Richard Johnson were greatly appreciated.
Science and Technology of Welding and Joining

2004 Vol. 9

No. 6

492

Colegrove and Shercliff

Two-dimensional CFD modelling of flow round profiled FSW tooling

10.

REFERENCES
1.

2.

3.

4.
5.

6.
7.
8.
9.

W. M. THOMAS, E. D. NICHOLAS, J. C. NEEDHAM, M. G. MURCH,


P. TEMPLE-SMITH and C. J. DAWES: Friction stir butt welding,
Int. Patent Application No. PCT/GB92/02203; GB Patent
Application No. 9125978.8, 1991; US Patent No. 5460317,
1995.
W. M. THOMAS, E. D. NICHOLAS, J. C. NEEDHAM, M. G. MURCH,
P. TEMPLE-SMITH and C. J. DAWES: Improvements relating to
friction welding, International Patent Classifications B23K 20/
12, B29C 65/06, 1993.
H. R. SHERCLIFF and P. A. COLEGROVE: in Mathematical
modelling of weld phenomena 6, (ed. H. Cerjak and H. K.
H. D. Bhadeshia), 927974; 2002, London, Maney Publishing.
FLUENT, Release 6.0.20, Fluent Incorporated, 2002.
Z. JIN, W. A. CASSADA, C. M. CADY and G. T. GRAY: Material
science forum Vols. 331337, 527532; 2000, Switzerland,
Trans Tech Publications.
S. XU, X. DENG, A. P. REYNOLDS and T. U. SIEDEL: Sci. Technol.
Weld. Joining, 2001, 6, 191193.
P. A. COLEGROVE and H. R. SHERCLIFF: Sci. Technol. Weld.
Joining, 2004, 9, (5), 360368.
M. SONG and R. KOVACEVIC: Trans. NAMRI/SME, 2002, XXX,
565572.
T. H. NORTH, G. J. BENDZSAK, C. B. SMITH and G. H. LUAN: Proc.
7th JWS International Symposium (7WS), Kobe, Japan,
November 2001, 621632.

Science and Technology of Welding and Joining

2004

Vol. 9 No. 6

11.
12.
13.
14.
15.
16.
17.
18.
19.
20.

P. A. COLEGROVE, H. R. SHERCLIFF and P. L. THREADGILL: Sci.


Technol. Weld. Joining, 2004, 9, (4), 345351.
T. LONG, T. U. SEIDEL, W. TANG and A. P. REYNOLDS: in Hot
deformation of aluminum alloys III, (ed. A. Beaudoin et al.);
2003, TMS.
T. U. SIEDEL and A. P. REYNOLDS: Sci. Technol. Weld. Joining,
2003, 8, (3), 175183.
B. LONDON, M. MAHONEY, W. BINGEL, M. CALABRESE and D.
WALDRON: Proc. 3rd Int. Symp. on Friction stir welding,
Kobe, Japan, 2001, Session 4, Paper 4.
M. GUERRA, J. C. MCCLURE, L. E. MURR, A. C. NUNES: in Friction
stir welding and processing, (ed. K. V. Jata et al.), 2534; 2001,
TMS.
T. L. DICKERSON, H. R. SHERCLIFF and H. SCHMIDT: Proc. 4th Int.
Symp. on Friction stir welding, Park City, UT, May, 2003,
Poster 5.
A. P. REYNOLDS: Sci. Technol. Weld. Joining, 2000, 5, 120124.
H. LARSSON, L. KARLSSON and L. E. SVENSSON: Proc. 6th Int.
Conf. on Aluminium alloys, Japan, Toyohashi, July 1998,
14711476.
A. J. LEONARD: Proc. 2nd Int. Symp. on Friction stir welding,
Gothenburg, Sweden, 2000, Session 7, Paper 1.
A. F. NORMAN, I. BROUGH and P. B. PRANGNELL: Material science
forum Vols. 331337, 17131718; 2000, Switzerland, Trans
Tech Publications.
M. W. MAHONEY, C. G. RHODES, J. G. FLINTOFF, R. A. SPURLING and
W. H. BINGEL: Metall. Mater. Trans. A, 1998, 29A, 19551964.

Das könnte Ihnen auch gefallen