Materials: Direct Observation of Filling Process and Porosity Prediction in High Pressure Die Casting

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

materials

Article
Direct Observation of Filling Process and Porosity
Prediction in High Pressure Die Casting
Hanxue Cao 1,2, *, Chao Shen 1 , Chengcheng Wang 1 , Hui Xu 3 and Juanjuan Zhu 3
1 Materials Forming and Control Department, College of Materials Science and Engineering,
Chongqing University, Chongqing 400044, China; [email protected] (C.S.);
[email protected] (C.W.)
2 National Engineering Research Center for Magnesium Alloys, Chongqing University,
Chongqing 400044, China
3 Chongqing Changan Automobile Co., Ltd., Chongqing 400023, China; [email protected] (H.X.);
[email protected] (J.Z.)
* Correspondence: [email protected]; Tel.: +86-23-65112626

Received: 4 March 2019; Accepted: 31 March 2019; Published: 2 April 2019 

Abstract: Although numerical simulation accuracy makes progress rapidly, it is in an insufficient


phase because of complicated phenomena of the filling process and difficulty of experimental
verification in high pressure die casting (HPDC), especially in thin-wall complex die-castings.
Therefore, in this paper, a flow visualization experiment is conducted, and the porosity at different
locations is predicted under three different fast shot velocities. The differences in flow pattern
between the actual filling process and the numerical simulation are compared. It shows that the flow
visualization experiment can directly observe the actual and real-time filling process and could be
an effective experimental verification method for the accuracy of the flow simulation model in HPDC.
Moreover, significant differences start to appear in the flow pattern between the actual experiment
and the Anycasting solution after the fragment or atomization formation. Finally, the fast shot
velocity would determine the position at which the back flow meets the incoming flow. The junction
of two streams of fluid would create more porosity than the other location. There is a transition
in flow patterns due to drag crisis under high fast shot velocity around two staggered cylinders,
which resulted in the porosity relationship also changing from R1 < R3 < R2 (0.88 m/s) to R1 < R2 <
R3 (1.59 and 2.34 m/s).

Keywords: direct observation; filling process; porosity prediction; high pressure die casting

1. Introduction
The high pressure die casting (HPDC) process has the characteristics of high speed filling and
high pressure solidification, which promotes it to become one of the main manufacturing processes of
complicated thin-wall components [1]. Mold has enormous effect on the quality of aluminum alloy
die-castings, such as surface finish, productivity and microstructure refinement. The melt flow is
controlled by the mold design during the filling process, and the parameters associated with the mold
design need to be carefully considered: sprue, gates, mold positioning, mold lubrication, thickness to
be filled and cooling system [2]. In addition, unreasonable gate system design will reduce the mold
life due to aluminum soldering mechanisms. The treatments applied to the mold surface are extremely
important to prolonging the lifespan of mold, such as TiAlN coatings [3]. A mold surface with excellent
performance can facilitate the part extraction in the ejection phase and acquire the integrated parts
with smooth surfaces. It is well known that porosity is the main defect in die-castings, and porosity
can seriously damage mechanical properties of die-castings. Air entrapment in the liquid metal during

Materials 2019, 12, 1099; doi:10.3390/ma12071099 www.mdpi.com/journal/materials


Materials 2019, 12, 1099 2 of 19

the filling process is the major source of porosity [4]. Therefore, to obtain high-performance castings,
it is extremely important to observe the die casting filling process and predict the air entrapment.
Thus, the filling process can be improved, and the defects rate can be reduced by changing the gate
system and venting system at the mold level according to melt flow observed in mold [5].
At present, computational fluid mechanics and experimental fluid mechanics are two main
methods to study the filling process and predict gas entrapment defects. Computational fluid
mechanics could predict the filling process through calculation and provide a powerful and cost
effective tool to control the filling process of die casting, and many new models and methods are
presented by many scholars in the HPDC field. Cao et al. [6] used the gas-liquid multiphase flow
model to predict the air entrainment phenomenon during the die casting filling process of zinc alloy,
and the volume of fluid (VOF) was used to track free surface. A water-filling experiment in an S-shaped
channel was simulated, and the simulated results were basically consistent with the experimental
results. Bi et al. [7] used the coupling of the level-set method and the VOF method to capture the free
interface and obtain its normal vector, and used the continuous surface force (CSF) model to consider
the surface tension, which was verified by experiment and could improve the simulation accuracy on
interface geometries, liquid flows and gas entrapments.
In addition to the above grid method, the grid-free method has received more attention in recent
years, especially smooth particle hydrodynamics (SPH). By solving the governing equation described
by the Lagrangian method, the SPH method can better deal with fluid flow with significant free surface
fragmentation and splashing during the die casting filling. Cleary et al. [8] used the SPH model
to simulate the filling process of laptop chassis, and found that the results were highly consistent
with the experiments results, indicating that the SPH model is very suitable for simulating complex
thin-wall die casting. Nevertheless, liquid metal is injected into the die cavity at a high velocity (around
30~100 m/s) in HPDC. The resulting flow is transient and complex. Although the simulation accuracy
makes progress rapidly, it is in an insufficient phase because of complicated phenomena of the filling
process and difficulty of experimental verification in HPDC [9,10], especially in thin-wall complex
die-castings. Therefore, it is hoped that the actual filling process could be observed directly so as to
provide the reference to improve the accuracy of numerical simulation.
The simulation result adopted a new model and a new method should be validated experimentally.
At the moment, the water analogue experiment is the most common method of experimental fluid
mechanics to observe the filling process and to validate the numerical predictions [6–8,11,12]. However,
although these experiments are designed according to dynamic similarity [13,14], water is not the
actual melt. The physicochemical properties and surface tension of water is much different from
the actual melt. Especially, surface tension plays an extremely important role in various surface and
interface phenomena [7,15]. Moreover, the water analogue experiment is generally conducted at room
temperature and does not take into account the temperature change, so it cannot really accurately
capture the actual filling process. In addition, Ohnaka et al. [10] presented that there was another
approach to directly observe the filling process during HPDC by X-ray diffraction. Although the
filling process could be observed intuitively, an especially designed mold (graphite mold rather than
thick steel mold) should be adopted due to the characteristics of the X-ray absorption image, and the
equipment required is expensive.
In this paper, a new method is introduced to directly observe the actual die casting filling process
with high speed, high temperature and high pressure. Three flow visualization experiments were
conducted. The actual filling process is captured directly by high speed camera. Comparing the
actual filling process observed and the numerical simulation solution, the differences of flow pattern
between the two are obtained. Meanwhile, the porosity of die-castings is predicted at different locations
according to the actual filling process.
Materials 2019, 12, 1099 3 of 19

2. Materials and Methods


A horizontal cold chamber die casting machine (L.K. Technology Holdings Limited, Hong Kong,
China) was used in this experiment. The casting material was 99.7% pure aluminum. Three flow
visualization experiments were conducted. The first stage speed was 0.13 m/s, the second stage
speed was 0.27 m/s and the fast shot set point was 270 mm. The fast shot speed was 0.88, 1.59 and
2.34 m/s, respectively. At the beginning of each test, the liquid metal was taken from the furnace
at the temperature of 750 ± 10 ◦ C, and the liquid metal was manually poured into the shot sleeve
with an inner diameter of 70 mm and a length of 380 mm. The pouring time was about 4 sec. At that
time, the liquid metal (melt volume: ~530,000 mm3 ) accounted for about 36% of the whole shot sleeve.
The die (material: P20 steel) was heated to 200 ◦ C before pouring the molten metal. Then, the plunger
tip was rapidly accelerated to reach the desired injection velocity. The nominal pressure of the HPDC
machine was 13.5 MPa. There was no pressure intensification stage in these experiments.

Table 1. Properties of borosilicate glass.

Properties Value
Density 2.23 g/cm3
Bending Strength 120–160 MPa
Thermal Expansion Coefficient (20-350 ◦ C) 32-35 × 10-6 cm/cm·◦ C
Thermal Conductivity (20 ◦ C) 0.82 W/m·◦ C
Specific Heat 820 J/kg·◦ C
SiO2 81.0%
B2 O3 12.5%
Chemical Composition (wt.%)
Al2 O3 2.32%
Na2 O + K2 O 6.0%

Figure 1. Geometry of die castings for the flow visualization experiment.

Figure 1 shows the geometry of die castings for the flow visualization experiment. Figure 2 shows
a schematic of the flow visualization experimental setup used to record the flow patterns. The transparent
borosilicate glasses were placed on the movable die to visualize the real time flow patterns during filling
the die cavity. Table 1 shows the properties of borosilicate glass used. The position and size of the two
Materials 2019, 12, 1099 4 of 19

transparent windows (the dimension of the bigger is 150 × 101 mm and the dimension of the smaller is
104 × 79 mm) is shown in Figure 2b, and the windows were parallel to the castings. The bottom of the
left window is located in the inner gate, and the left window is a vertical plane, so the flow of aluminum
liquid into the mold cavity can be observed conveniently. There are two staggered cylinders in the right
window, and the flow of liquid aluminum around them could be studied under die casting conditions.
Flow patterns of the aluminum melt were captured using a high speed camera with a sampling rate of
1000 frames per second and a shutter speed of 1/1000 seconds. To determine the start time of plunger tip
moving in the video, a signal light was placed in the shooting area as shown in Figure 2b. As long as the
plunger tip started to move, the signal light went on.

Figure 2. The flow visualization setup. (a) A schematic showing the flow visualization experiment setup;
(b) the two shooting windows of high speed camera (the dimension of the bigger is 150 × 101 mm and the
dimension of the smaller is 104 × 79 mm) on the movable die.
Materials 2019, 12, 1099 5 of 19

A popular commercially available software package (Anycasting software) (AnyCasting Co.Ltd.,


Seoul, Korea) for casting simulations and analysis was adopted. The CAD date of mold was imported
into the software. The filling process simulation of the same process was conducted. The boundary
conditions and mesh properties are listed in Table 2. The finite difference meshes were adopted in the
Anycasting simulation. The SOLA-VOF (Solution Algorithm-Volume of Fluid) approach was used
to solve the coupling between the momentum and mass conservation equations and to trace the free
surface in Anycasting software. The thermophysical properties of melt used in the simulation are
listed in Table 3. The properties of borosilicate glass in Table 1 were imported into the Anycasting
simulation. Heat transfer coefficient between casting and die was set to 2093.5 W/m2 ·K. Since the heat
transfer coefficient between glass and casting was hard to measure, it was set to be the same as that
between the mold and casting. Standard continuum surface force (CSF) model was enabled, and RNG
(Renormalization Group) k-ε turbulence model was adopted. Nowadays, RNG k-e turbulence model
is an optional module in the Anycasting software and is widely used to deal with turbulence flow in
die casting. The RNG k-ε turbulence model can be applied to the near wall region directly, and the
turbulent vortex is taken into account in this model. Therefore, the RNG k-e turbulence model is suited
to simulate the jetting fluid and the flow around circular cylinders under the die casting condition.

Table 2. Boundary conditions and mesh properties.

Properties Value
Material P20 Steel
Special Heat (285.5 + 1.076 T) J/kg·K
Density (7920 – 0.4739 T) kg/m3
Thermal Conductivity (46.08 + 0.0022 T) W/m·K
Total Mesh Cells 28,545,644
Mesh Cell Size 0.001702 m

Table 3. Thermophysical properties of melt in the simulation.

Property Value
Material Pure Aluminum
Initial Temperature 750 ◦ C
Liquidus 675 ◦ C
Special Heat 900 J/kg·K
Density 2705 kg/m3
Thermal Conductivity 234 W/m·K
Latent Heat 389,000 J/kg
Surface Tension (−0.0028 T + 2.762) N/m
Temperature Value
660 ◦ C 0.003017 kg/m·s
Dynamic Viscosity 680 ◦ C 0.002762 kg/m·s
700 ◦ C 0.002541 kg/m·s
720 ◦ C 0.002349 kg/m·s

3. Results and Discussion

3.1. The Differences in the Flow Patterns between the Actual Filling Process and the Numerical Simulation
In this paper, the filling process observed under the injection velocity of 2.34 m/s is taken as
an example to study the difference in flow pattern between the actual filling process and the numerical
simulation results. Figure 3 shows the comparison between numerical simulation (right) results and
the flow visualization experiment (left) within the transparent windows. The first frame is determined
when the fluid has just started entering the right window in two cases. Then, the difference of flow
pattern between the two in the same position is observed.
Materials 2019, 12, 1099 6 of 19

In the first frame (at 1 ms), the fluid has just started entering the right windows in two cases.
The volume of fluid that has entered the window is slightly small for Anycasting. By 11 ms (Figure 3b),
this thin (in the third dimension) jet strikes the surface of the raised cylindrical section and starts to
flow both up and around its sides in the right window. Due to the low speed at the moment, the flow
behavior presents a typical detour flow. This flow pattern shows almost no variance between the
two flows. At 27 ms (Figure 3c), two streams of the melt from two ingates both encounter. It can
be speculated that the region that two streams of fluid encounter is almost identical in two cases.
However, a wave from the right ingate enters the left window (marked by the red circle) in the actual
experiment while this region is totally outside the left window in the numerical simulation. It is
extremely easy to produce some defects such as cold shut and air pockets in this region. By 31 ms
(Figure 3d), the liquid metal fills smoothly and is continuous in both cases. Nevertheless, the flow
front is irregular in the actual experiment while the flow front is smooth in the Anycasting simulation.
In Figure 3e–g, the significant differences start to appear in the flow pattern between the actual
experiment and the Anycasting solution. In the actual experiment, many fragments emerge in
the left window. Disintegration leads to an increase of the surface area for the liquid aluminum,
which inevitably results in an increase of oxidation. Oxide film can impact the surface tension of
melt, heat transfer and so on. In addition, these fragments cool rapidly, and their surfaces partially
or completely solidify. When two disintegrated parts collide, it is extremely easy to produce some
defects such as cold shut. Moreover, it is liable to trap small amounts of air around their asperities
when they pack together. However, almost no fragment is observed in the Anycasting solution.
At 104 ms (Figure 4a), the injection process is in the fast shot phase. A jet of high speed (indicated by
the red circle) begins to appear in the left window. At 118 ms (Figure 4b), high gate velocities create an
atomization phenomenon. When a liquid jet is subjected to the high shear forces between the jet and
surrounding gas, the jet disintegrates into some fragments or atomizes. The shear forces mainly depend
on the relative velocity between the jet and the gas at the gate discharge [16]. When the injection
velocity is high enough, the atomization phenomenon would be observed. However, the atomization
phenomenon is not observed in the Anycasting solution.
In Figure 3h, it shows that the last area to be filled is in the left side of the left window. When the
fluid jet reaches the top wall of die, it immediately fills back the lower empty region and joins the
incoming jet. Thus, a recirculation zones is created, and a clockwise vortex is developed there as
marked by the red circle in Figure 3h. This intense vortex would create a near vacuum condition inside
it, and any dissolved or free gases around this region will be sucked into this vortex, which induces
void and leaves a porous region. By particle tracking in the Anycasting simulation, a clockwise
recirculation is also observed at the same position in the Anycasting solution, but the volume of empty
region is slightly larger and its interface is clear and smooth. At 131 ms (Figure 3i), the center of
vortex moves down in the actual experiment, and the dynamic evolution of air entrainment is clearly
observed. Although the vortex also moves downward, its size is different in the Anycasting simulation.
In short, the Anycasting solution is able to make quite accurate predictions for the motion of fluid
fronts in the very early stages. As the injection process is in the second injection phase in the very early
stages, the velocity of aluminum melts is low and the flow is laminar flow. Moreover, the last area to
be filled could also be precisely forecasted. Nevertheless, the significant differences start to appear in
the flow pattern between the actual experiment and the Anycasting solution after the break-up and
disintegration process takes place. The fragment or atomization formations are the main reasons for
the difficulties of modelling the filling process using numerical simulation.
Materials 2019, 12, 1099 7 of 19

Figure 3. Cont.
Materials 2019, 12, 1099 8 of 19

Figure 3. Cont.
Materials 2019, 12, 1099 9 of 19

Figure 3. The comparison between the numerical simulation (right) results and the flow visualization
experiment (left) within the transparent windows (the time shown in (a–i) is the actual experimental
time from when the fluid has just started entering the right window).

Figure 4. The actual filling process within the transparent windows (the time shown in (a,b) is the
same as Figure 3).

3.2. Porosity Prediction by Direct Observation of Filling Process in the Left Window
The simulation accuracy still needs to improve because of complicated phenomena of the filling
process, such as fragment and atomization. The flow visualization experiment can directly observe
the actual and real-time filling process by transparent windows installed on the die in this paper,
which could provide the reference to predict the porosity at different positions in HPDC.
To validate the prediction results, the sampling locations of die-castings under different fast
shot speeds are shown in Figure 5. The porosity was measured by hydrostatic weighing method.
Materials 2019, 12, 1099 10 of 19

All specimens are weighted in air and in water respectively, and then their densities are determined
according to the following formula:
m1
$p = ·$ (1)
m1 − m2 w
where: $p and $w is, respectively, the density of the specimen and water (unit: g/cm3 ); m1 and m2 is
the mass of the specimen in air and in water respectively (unit: g). Next, the porosity of the examined
specimens is calculated from the following relationship:
 $p 
P = 1− ·100% (2)
$wz

where: $wz is true density (unit: g/cm3 ), equal to 2.7 g/cm3 .

Figure 5. The sampling location (L1, L2, and L3 is located in the left window, and R1, R2, and R3 is
located in the right window).
Materials 2019, 12, 1099 11 of 19

Figure 6. The porosity at different locations in the left window under different fast shot velocities.

Figure 6 shows the porosity of die-castings at different locations in the left window under different
fast shot speeds. From Figure 6, the porosity of the L2 sample is obviously higher than that of other
samples, especially under 0.88 m/s. Furthermore, there are small differences between the porosity of
the L1 and L3 samples. Figure 7 shows the flow pattern under the fast shot velocity of 0.88 m/s in
the left window. Liquid aluminum passes through the ingate at high speed, and it is hindered by the
cylindrical hole (the area surrounded by green lines in Figure 7), which results in the flux and velocity
of liquid aluminum being reduced above the cylindrical hole. Consequently, there is a vacant land that
is not occupied by aluminum melt at L1 and L2 regions, as shown in Figure 7a.
Liquid metal flows forward from both sides of the cylindrical hole with less resistance, and a back
flow towards the L2 region is observed when liquid metal collides with the top wall of die and builds up,
as shown in Figure 7b. Sequentially, the back flow from the top wall meets the incoming flow from the
ingate at the symmetrical lines of the L2 region, and the front of incoming flow is jagged. The returning
fluid collided with the filling stream, which would create the turbulence and increase gas entrapment [5].
The fast shot velocity of 0.88 m/s is relatively low, so the temperature and velocity of aluminum melt is
very low where two streams of liquid join. Moreover, the collisions between two streams cause turbulence,
and turbulence can promote the formation of droplets, which tend to freeze rapidly [17]. Therefore,
the back flow and incoming flow are not closely combined, and the L2 sample is filled with a fairly sparse
liquid metal (shown in Figure 7c), which creates a large number of pores. Furthermore, liquid metal
collides with the top wall of die and heaps up at the L1 region, which would also cause the potential risk
of gas entrapment to increase at the L1 region. In addition, air entrainment is not found in the L3 sample.
Hence, the porosity of the L1 sample (7.333%) is higher than that of the L3 sample (6.593%) and is lower
than that of the L2 sample (21.519%).
Materials 2019, 12, 1099 12 of 19

Figure 7. Flow pattern under the fast shot velocity of 0.88 m/s in the left window. (a–c) are arranged
according to the order of filling.

Figure 8 shows the flow pattern under the fast shot velocity of 1.59 m/s in the left window. It is
observed that this flow pattern is similar to that under the fast shot velocity of 0.88 m/s. The liquid
metal is also diverted by the cylindrical hole when it passes through the ingate, which generates an
unoccupied space by aluminum melt at L1 and L2 regions, as shown in Figure 8a. Liquid melt on
both sides of the L2 region collides with the top wall of die and meets in the L1 region, which would
involve a large number of pores, as shown in Figure 8b. However, compared with the fast shot velocity
of 0.88 m/s, there is a larger unoccupied space by aluminum melt before the back flow joining with
the incoming flow. This area includes both the L2 region and part of the L3 region, as opposed to
including only the L2 region under the fast shot speed of 0.88 m/s, as shown in Figure 8b. It is the main
reason that the higher the gate speed, the more focused the aluminum liquid is that reaches the left
window plane. Therefore, under the fast shot velocity of 1.59 m/s, only a very narrow stream of liquid
aluminum could flow upward from the left side of the cylindrical hole, which results in that part of L3
region not being filled. Moreover, the position where the incoming flow meets the back flow is lower
than the symmetrical lines of the L2 region, as shown in Figure 8c, so the area of the junction is reduced
in the L2 sample relative to under the fast shot velocity of 0.88 m/s. This is because the back flow
has more momentum under the fast shot velocity of 1.59 m/s. Besides, due to more momentum and
less heat dissipation, two streams of liquid bond more closely when they join. Therefore, the porosity
of L2 sample is lower than that under the fast shot velocity of 0.88m/s. In addition, part of the L3
region is located at the junction of two streams of fluid. Hence, under the fast shot velocity of 1.59 m/s,
the porosity of the L3 sample (9.222%) is higher than that of the L1 sample (6.704%) and is lower than
that of the L2 sample (11.148%).
Materials 2019, 12, 1099 13 of 19

Figure 8. Flow pattern under the fast shot velocity of 1.59 m/s in the left window. (a–c) are arranged
according to the order of filling.

Figure 9 shows the flow pattern under the fast shot velocity of 2.34 m/s in the left window.
It exhibits that this flow pattern is nearly identical to that under the fast shot velocity of 1.59 m/s.
There is an unoccupied space by aluminum melt at L1 and L2 regions, as shown in Figure 9a. Liquid
metal collides with the top wall of die and accumulates at the L1 region. Both the L2 region and a small
fraction of the L3 region is not filled with liquid aluminum before the back flow joining with the
incoming flow, as shown in Figure 9b. Nevertheless, compared with the fast shot velocity of 1.59 m/s,
the position where the incoming flow meets the back flow is higher. This is because although the
velocity of back flow increases, the rate at which the cylindrical hole is filled also increases under the
fast shot velocity of 2.34 m/s. Consequently, the blocking effect of the cylinder hole fails in advance,
and the back flow and the incoming flow meet at the symmetrical lines of the L2 region. The area of
junction becomes bigger in the L2 sample, which results in that the porosity of L2 sample is higher than
that under the fast shot velocity of 1.59 m/s. In addition, the L3 region is not located at the junction of
two streams of fluid, as shown in Figure 9c. Therefore, the porosity of the L1 sample (7.370%) is higher
than that of the L3 sample (7.037%) and is lower than that of the L2 sample (12.185%).

Figure 9. Flow pattern under the fast shot velocity of 2.34 m/s in the left window. (a–c) are arranged
according to the order of filling.
Materials 2019, 12, 1099 14 of 19

3.3. Porosity Prediction by Direct Observation of Filling Process in the Right Window
According to the flow pattern observed in the flow visualization video, the flow in the right
window could be simplified down to the planar flow around two staggered cylinders. Figure 10 shows
the arrangement of the two cylinders in the right window. Figure 11 shows the porosity of die-castings
at different locations in the right window under different fast shot speeds. When the fast shot speed is
0.88 m/s, the relationship of porosity could be descript as the R2 sample is highest; the R3 sample is
next; while R1 sample is lowest. From Figure 12, it is observed that the inner separated shear layer
of upstream cylinder reattaches to the outer surface of the downstream cylinder, whereas the outer
separated shear layer of the upstream cylinder does not contact the downstream cylinder. In addition,
although there is a gap between the two cylinders, the gap is surrounded by the reattached shear layer
on one side and by the outer shear layer of the upstream cylinder on the other side, which effectively
prevents the incoming flow from penetrating this gap. Consequently, almost no oncoming flow runs
through the gap, and there is only nearly stagnant fluid in the gap.

Figure 10. Schematic diagram showing (a) the arrangement of the two cylinders in the right window;
(b) shear layer designations.

Figure 11. The porosity at different locations in the right window under different fast shot velocities.
Materials 2019, 12, 1099 15 of 19

Sumner et al. [18] summarized the nine low subcritical regime flow patterns for two staggered
circular cylinders of equal diameter in cross-flow. Under the fast shot velocity of 0.88 m/s, the flow
pattern in the right window is similar to the “shear layer reattachment flow pattern (SLR)” above.
Besides, Gu and Sun [19] made also similar discoveries for two staggered circular cylinders in the high
subcritical Reynolds numbers, and denominated the flow pattern as “Pattern IB .” The principal part
of R3 sample is submerged in the wake of the large cylinder, as shown in Figure 12a. A two-eddy
configuration pattern is observed in the near-wake region, and a rapid increase in the frequency
of vortex shedding when the flow is under the Reynolds numbers [20]. A large negative pressure
appears in the wake, and the maximum turbulent intensity is located on the near-wake region [21],
which would result in that large quantities of gases are sucked and involved. The main part of the R2
sample is located in the near-wake region of upstream cylinder, and R2 sample is filled later than R3
sample, as shown in Figure 12b, which causes the porosity of R2 sample to be higher than that of the
R3 sample. The incoming flow, which is close to the outer separated shear layer of upstream cylinder,
reaches the inner surface of the downstream cylinder directly (R1 region), so the R1 region is almost
free from the interference of wake. Therefore, under the fast shot velocity of 0.88 m/s, the porosity of
the R3 sample (14.704%) is higher than that of the R1 sample (11.000%) and is lower than that of the R2
sample (15.815%).

Figure 12. Flow pattern under the fast shot velocity of 0.88 m/s in the right window. (a,b) are arranged
according to the order of filling.

Under the fast shot speed of 1.59 m/s, the relationship of porosity is: the R3 sample is highest;
next is the R2 sample; the R1 sample is lowest between them in the right window. From Figure 13,
the near-wake region becomes narrow behind the upstream cylinder and deviates from the downstream
cylinder and the flow axis. Meanwhile, the inner separated shear layer of the upstream cylinder has no
capacity to reattach to the outer surface of the downstream cylinder, and the oncoming flow is allowed
to penetrate the gap between the cylinders. This flow pattern is similar to “induced separation flow
pattern (IS)” denoted by Sumner et al. [18]. This pattern is also similar to “Pattern IIB ” denominated
by Gu and Sun [19]. The near-wake region is highly compressed by the gap flow, and the R2 sample
keeps away from this region, so the wake of the small cylinder has only a little effect on the porosity
of the R2 sample. Moreover, the R3 region situated in the rear of downstream cylinder is filled last.
The oncoming flow closed to the outer separated shear layer of upstream cylinder reaches the R1 region
directly and runs through the gap between the cylinders up to the R1 region. Hence, the porosity of
the R2 sample (11.148%) is higher than that of the R1 sample (10.630%) and is lower than that of the R3
sample (15.519%).
Materials 2019, 12, 1099 16 of 19

The Reynolds number plays an important role in the flow pattern around a circular cylinder. It is
well known that the boundary layer in the front of the circular cylinder undergoes a transition from
laminar to turbulent when Reynolds numbers approach 2 × 105 [22]. In the critical regime (2 × 105
< Re < 3.5 × 105 ) [23], the turbulence transition in the boundary layer makes the pressure difference
between the rear of the cylinder and front less, and the drag force reduce suddenly. When the fluid
goes around this cylinder, it has more inertia and is able to go around a little bit more, which causes the
delay of the separation point of the shear layer and a little bit more compact wake. In addition, in the
supercritical regime (3.5 × 105 < Re < 1.5 × 106 ) [23], the boundary layer becomes turbulent completely,
and the drag force around the cylinder begins to recover. The Reynolds number (Re = UD1 /ν) is
defined in terms of the incoming flow velocity U, the upstream cylinder diameter D1 and the kinematic
viscosity of fluid ν. Under the three fast shot velocities, Reynolds numbers are 1.9 × 105 (0.88 m/s),
3.3 × 105 (1.59 m/s), and 4.9 × 105 (2.34 m/s), respectively. Therefore, in the same arrangement of the
two cylinders, when the fast shot velocity increases from 0.88 to 1.59 m/s, the turbulence transition
takes place in the boundary layer and the shear layer is delayed to separate from the upstream cylinder.
Thereby, the relationship of porosity of three samples also changes from R1 < R3 < R2 (0.88 m/s) to R1
< R2 < R3 (1.59 m/s).

Figure 13. Flow pattern under the fast shot velocity of 1.59 m/s in the right window. (a,b) are arranged
according to the order of filling.

When the fast shot speed increases to 2.34 m/s, the flow pattern at the moment is the same as
that under the fast shot velocity of 1.59 m/s, as shown in Figure 14. The near-wake region is highly
confined in the gap between the two cylinders, and the incoming flow is allowed to pass through
the gap. Consequently, the relationship of porosity is identical to that under the fast shot velocity of
1.59 m/s and could be descript as the R3 sample is highest; next is the R2 sample; the R1 sample is
lowest between them.
Materials 2019, 12, 1099 17 of 19

Figure 14. Flow pattern under the fast shot velocity of 2.34 m/s in the right window. (a,b) are arranged
according to the order of filling.

4. Conclusions
In this paper, the actual filling process of HPDC is directly observed by the flow visualization
experiment. The differences in flow pattern between the actual filling process and the Anycasting
software solution are compared. Meanwhile, the porosity of samples at different locations is predicted
under three different fast shot velocities. We draw the following conclusions from our investigations:
(a) The flow visualization experiment can clearly observe the whole die casting filling process,
including complex flow phenomena, such as fragment and atomization. Therefore, it could be
a valuable tool for validating experimentally the outcome of flow simulations.
(b) The significant differences start to appear in the flow pattern between the actual experiment
and the Anycasting solution after the fragment or atomization formation. In addition, the dynamic
evolution of air entrainment is clearly observed, the volume of the empty region is small, and its
interface is rough in the flow visualization experiment while the volume of the empty region is slightly
larger and its interface is clear and smooth in the Anycasting solution.
(c) According to direct observation of the filling process, the porosity distribution of die-castings
can be accurately predicted. The fast shot velocity would determine the position at which the back flow
from the top of die meets the incoming flow from the inner gate. The junction of two streams of fluid
would create more porosity than other locations. There is a transition in flow patterns due to drag crisis
under high fast shot velocity around two staggered cylinders, which result in the porosity relationship
also changing from R1 < R3 < R2 (0.88 m/s) to R1 < R2 < R3 (1.59 and 2.34 m/s). Meanwhile, the region
submerged in the wake of circular cylinder would produce greater porosity than other regions.

Author Contributions: Conceptualization, H.C.; Data curation, C.S. and C.W.; Formal analysis, C.S.; Investigation,
C.S. and C.W.; Methodology, H.C.; Software, H.X. and J.Z.; Supervision, H.X. and J.Z.; Writing—original draft,
C.S.; Writing—review & editing, H.C.
Funding: This research received no external funding.
Acknowledgments: Manuscript is approved by all authors for publication. The authors would like to express
their thanks to Hui Xu and Juanjuan Zhu who provided their technical assistance.
Conflicts of Interest: The authors declare no conflict of interest.
Materials 2019, 12, 1099 18 of 19

References
1. Hamasaiid, A.; Dour, G.; Dargusch, M.S.; Loulou, T.; Davidson, C.; Savage, G. Heat-Transfer Coefficient and
In-Cavity Pressure at the Casting-Die Interface during High-Pressure Die Casting of the Magnesium Alloy
AZ91D. Metall. Mater. Trans. A 2008, 39, 853–864. [CrossRef]
2. Silva, F.J.G.; Campilho, R.D.S.G.; Ferreira, L.P.; Pereira, M.T. Establishing Guidelines to Improve the High-Pressure
Die Casting Process of Complex Aesthetics Parts. In Transdisciplinary Engineering Methods for Social Innovation
of Industry 4.0, 2nd ed.; Peruzzini, M., Pellicciari, M., Bil, C., Stjepandić, J., Wognum, N., Eds.; IOS Press:
Amsterdam, Holland, 2018; Volume 7, pp. 887–896.
3. Nunes, V.; Silva, F.J.G.; Andrade, M.F.; Alexandre, R.; Baptista, A.P.M. Increasing the lifespan of high-pressure
die cast molds subjected to severe wear. Surf. Coat. Technol. 2017, 332, 319–331. [CrossRef]
4. Hernandez-Ortega, J.J.; Zamora, R.; Palacios, J.; Lopez, J.; Faura, F. An Experimental and Numerical
Study of Flow Patterns and Air Entrapment Phenomena During the Filling of a Vertical Die Cavity.
J. Manuf. Sci. Eng.-Trans. ASME 2010, 132, 9. [CrossRef]
5. Pinto, H.; Silva, F.J.G. Optimisation of Die Casting Process in Zamak Alloys. Procedia Manuf. 2017, 11,
517–525. [CrossRef]
6. Cao, L.; Liao, D.; Sun, F.; Chen, T.; Teng, Z.; Tang, Y. Prediction of gas entrapment defects during zinc alloy
high-pressure die casting based on gas-liquid multiphase flow model. Int. J. Adv. Manuf. Technol. 2018, 94,
807–815. [CrossRef]
7. Bi, C.; Guo, Z.P.; Xiong, S.M. An improved mathematical model to simulate mold filling process in high
pressure die casting using CLSVOF method and CSF model. China Foundry 2015, 12, 180–188.
8. Cleary, P.W.; Savage, G.; Ha, J.; Prakash, M. Flow analysis and validation of numerical modelling for a thin
walled high pressure die casting using SPH. Comput. Part. Mech. 2014, 1, 229–243. [CrossRef]
9. Otsuka, Y. Experimental Verification and Accuracy Improvement of Gas Entrapment and Shrinkage Porosity
Simulation in High Pressure Die Casting Process. Mater. Trans. 2013, 55, 154–160. [CrossRef]
10. Itsuo Ohnaka, A.S.; Ikeda, T.; Yasuda, H. Mold Filling and Prevention of Gas Entrapment in High-pressure
Die-casting. J. Mater. Sci. Technol. 2008, 24, 139–140.
11. Niu, X.F.; Fang, Z.; Liang, W.; Hou, H.; Wang, H.X. New numerical algorithm of gas-liquid two-phase flow
considering characteristics of liquid metal during mold filling. Trans. Nonferrous Metals Soc. China 2014, 24,
790–797. [CrossRef]
12. Bi, C.; Guo, Z.; Xiong, S. Modelling and simulation for die casting mould filling process using Cartesian cut
cell approach. Int. J. Cast Metals Res. 2015, 28, 234–241. [CrossRef]
13. Lee, W.B.; Lu, H.Y.; Lui, Y.B. A computer simulation of the effect of wall thickness on the metal flow in
diecasting dies. J. Mater. Process. Technol. 1995, 52, 248–269. [CrossRef]
14. Cleary, P.; Ha, J.; Alguine, V.; Nguyen, T. Flow modelling in casting processes. Appl. Math. Model. 2002, 26,
171–190. [CrossRef]
15. Sakuragi, T. Mould filling simulation with consideration of surface tension and its application to a practical
casting problem. Int. J. Cast Metals Res. 2005, 18, 202–208. [CrossRef]
16. Homayonifar, P.; Babaei, R.; Attar, E.; Shahinfar, S.; Davami, P. Numerical modeling of splashing and air
entrapment in high-pressure die casting. Int. J. Adv. Manuf. Technol. 2008, 39, 219–228. [CrossRef]
17. Armillotta, A.; Fasoli, S.; Guarinoni, A. Cold flow defects in zinc die casting: Prevention criteria using
simulation and experimental investigations. Int. J. Adv. Manuf. Technol. 2016, 85, 605–622. [CrossRef]
18. Sumner, D.; Price, S.J.; Paidoussis, M.P. Flow-pattern identification for two staggered circular cylinders in
cross-flow. J. Fluid Mech. 2000, 411, 263–303. [CrossRef]
19. Gu, Z.; Sun, T. On interference between two circular cylinders in staggered arrangement at high subcritical
Reynolds numbers. J. Wind Eng. Ind. Aerodyn. 1999, 80, 287–309. [CrossRef]
20. Rodríguez, I.; Lehmkuhl, O.; Chiva, J.; Borrell, R.; Oliva, A. On the flow past a circular cylinder from critical
to super-critical Reynolds numbers: Wake topology and vortex shedding. Int. J. Heat Fluid Flow 2015, 55,
91–103. [CrossRef]
21. Djeridi, H.; Braza, M.; Perrin, R.; Harran, G.; Cid, E.; Cazin, S. Near-Wake Turbulence Properties around
a Circular Cylinder at High Reynolds Number. Flow Turbul. Combust. 2003, 71, 19–34. [CrossRef]
Materials 2019, 12, 1099 19 of 19

22. Rodríguez, I.; Lehmkuhl, O.; Borrell, R.; Paniagua, L.; Pérez-Segarra, C.D. High Performance Computing of
the Flow Past a Circular Cylinder at Critical and Supercritical Reynolds Numbers. Procedia Eng. 2013, 61,
166–172. [CrossRef]
23. Yeon, S.M.; Yang, J.; Stern, F. Large-eddy simulation of the flow past a circular cylinder at sub- to super-critical
Reynolds numbers. Appl. Ocean Res. 2016, 59, 663–675. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like