1. Introduction
Two-phase flows with variable cross section still remain a challenging subject with great interest from the scientific community and industry related to the efficient cooling of computer parts and biomedical devices. Moreover, channels with different corrugation patterns seem to considerably improve the effect of heat dissipation by modifying the drop’s dynamics and thus the liquid film thickness of the two-phase system. Corrugated channel is a class of variable cross section channels that present a periodic pattern along its length and has strong impact in the fluid flow, changing its behavior by increasing recirculation zones and eventually leading to onset of boundary layer detachment. Note that the presence of a deformable interface due to the presence of more fluid phases brings an extra degree of complexity in the fluid flow system. Additionally, the periodic pattern of the corrugated channels requires very large domains to be analyzed, thus dramatically increasing the costs associated to the experiment assembling or the numerical simulations.
Single-phase flows with heat transfer were experimentally analyzed in [
1,
2], where the effects of several key parameters were studied such as channel wavy amplitude and flow velocity in periodic wavy passages. Numerically, several authors have investigated single-phase flows in corrugated channels. For instance, in [
3] a numerical study was done for transition regimes from laminar to turbulent when convective heat transfer plays an important role in those flows. They concluded that the Nusselt number for the wavy wall channel was directly affected when compared to those for a straight channels, while in [
4], a 2-dimensional laminar steady and time-dependent fluid flow solver was developed for heat transfer in periodic wavy channels with sinusoidal and arc shapes. It was concluded that the arc-shaped periodic channel enhances heat transfer due to a higher friction factor when compared to the sinusoidal channel, specially beyond the critical value of the Reynolds number for the unsteady regime. Another interesting numerical approach was suggested by the authors of [
5], where the coordinate transformation method and the spline alternating-direction implicit method was used to study the rates of heat transfer for flow through a sinusoidally curved converging–diverging channel. The steady streamfunction vorticity formulation was used to model the fluid flow with the energy equation using the finite difference method. Several channel amplitude wavelengths were tested at different Reynolds number, and it was found that at a sufficiently larger value of amplitude wavelength ratio the corrugated channel increases heat transfer for large Reynolds numbers. Turbulent forced convection in a wavy channel using a two-phase model was studied in [
6] for water-Al
O
nanofluids using a 2-dimensional numerical solver. It has been seen that if the channel amplitude increases, as well as the Reynolds number and volume fraction of nanoparticles, the Nusselt number is directly affected, leading to a higher value. Additionally, it was observed that the more nanoparticles added to the solution, the higher pressure drop in corrugated channels is seen.
It is already known that two-phase flows bring another level of challenges numerically and experimentally due to the presence of an interface between fluids and different phases in the system. In [
7], a comparative analysis of two-phase flow in sinusoidally corrugated channels was carried out with application to polymer electrolyte membrane fuel cells. Different patterns of corrugated channels were used to numerically analyze single and multiples bubbles dynamics in air–water systems through a 3-dimensional simulation using the Volume-of-Fluid (VOF) method. The authors concluded that the two-phase flow performance in the sinusoidal channels can be substantially improved, decreasing the radius of curvature at the bends and making the confining walls extremely hydrophobic. On the other hand, in [
8], experimental and numerical approaches have been used to tackle CO
bubble behavior flowing in a methanol solution in sinusoidally corrugated channels for fuel cell purposes. A detailed investigation of key two-phase flow parameters was undertaken, and a thermal analysis was made regarding when a single bubble travels along sinusoidal channels under pressure differences between the inlet and outlet regions of the channel. Additionally, a comparison of the same flow condition was carried out for a straight channel. The results reveals that the effect of flow disturbance structure is negligible with small values of channel amplitude and angular frequency of corrugation for the studied two-phase system. Sinusoidal channels with one and more wavy patterns were numerically simulated using the front-tracking and the finite volume methods in [
9]. Two-phase 2-dimensional flow simulations were carried out to investigate single buoyant drops in channel constrictions and expansions and the interaction between two small drops in different simulation conditions with the newly implemented numerical code. Results have been successfully reported for various constricted channels.
The current study is based upon the experiments in [
10], where the authors have studied experimentally the effects of buoyancy-driven motion of drops in a periodically constricted capillary. Several two-phase systems were used in the experiment such as glycerol–water, diethylene glycol, and diethylene glycol–glycerol as suspending fluid, and silicon oil, Dow Corning fluids, and UCON fluid emollient as drop fluids, and an extensive study has been reported for sinusoidal shape channels with fixed dimensionless amplitude
and wavelength
for single drops. However, no further investigation has been conducted for different channel amplitude and its effect on interface break-up for single and multiple drops. Additionally, it is clear from the literature review presented above that much is still required for the complete understanding of two-phase flows in short and large periodic arrays with different corrugation levels. It is indeed of great significance that single and multiple buoyant drops in sinusoidal channels deliver another level of flow complexity due to the possible significant channel amplitude perturbation in the two-phase system, leading eventually to drop coalescence and/or interface breakup. The aim of this work is to investigate numerically two-phase flow key parameters such as the drop’s rising velocity, film thickness, and interface perimeter change for one and three drops in buoyancy-driven motion for a diethylene glycol/UCON-1145 (DEG3) two-phase system in sinusoidal channels with dimensionless wavelength
and wave amplitude
. Additionally, a parametric study is also carried out to investigate the influence of channel wavelength and drop volume in coalescence during slug flows. A state-of-the-art model is employed to accurately compute the dynamics of the drop’s interface motion using a modern moving frame/moving mesh technique within the arbitrary Lagrangian–Eulerian framework. The presented results show the drop’s dynamics in large periodic channels and coalescence even for large liquid slug distances.
This article is organized with a literature review of single- and two-phase corrugated channels, followed by the mathematical description of the modeling equations in axisymmetric formulation based on the Finite Element Method (FEM). The next section provides details concerning the numerical modeling used to discretize the mathematical equations, followed by the result section where several important test cases are presented to assess and analyze the dynamics of two-phase flow in corrugated channels with single and multiples bubbles. Finally, this text ends with remarks in the conclusion section.
4. Results
We present the results of single and multiples drops rising in channels with two levels of corrugation amplitude (
) obtained with the ALE-FE Two-Phase flow solver for axisymmetric coordinate system. Note that such a code has been extensively validated against several single and two-phase flows problems and have been reported in different well-recognized international journals. All numerical results presented here refers to fluid the two-phase system diethylene glycol/UCON-1145 from [
10] also known as DEG3 system. According to the experiments, the drop fluid (UCON-1145) presented the dynamic viscosity of
and density of
. The suspending fluid (diethylene–glycol) presented dynamic viscosity and density of
and
, respectively. Gravity was set to standard condition:
. The surface tension coefficient was set to
, which is approximately 10 times lower than the original value found in the aforementioned reference (
), thus meaning the dimensionless parameters used at all simulations were the Archimedes number
= 15,417 and the Eötvös number
. These dimensionless numbers were chosen to allow large drop deformation due to low surface tension force while keeping numerical stability with large time steps. The initial mesh to all test cases had approximately 16,500 nodes including centroid and 10,060 triangle elements and finished with 43,000 nodes including centroid and 28,200 triangle elements due to the adaptive remeshing described in the previous section.
A total of 10 test cases using the geometries presented in
Figure 2 for the straight channel (
) and sinusoidal channel (
) with single and three consecutive drops were made, namely,
single drop in straight channel,
single drop in sinusoidal channel,
slug flows in straight channel with initial slug length, and
slug flows in sinusoidal channel, both with initial slug length ranging from
; thus, the same slug lengths
s were used as comparison between different channel shapes. Additionally, 14 test cases were carried out to investigate drop’s coalescence time
in slug flows, varying systematically the drop’s size factor
and keeping constant the wavelength
and the initial slug length
, and varying systematically the channel’s wavelength
and keeping constant the drop’s size factor
and the initial slug length
s.
The unsteady behavior of all flows in sinusoidal channel can be easily noted at all plots presented below due to their wavy response with respect to the evolution with time of drop rising velocity and minimum film thickness, which is the minimum distance between any interface node to the closest wall node. On the other hand, the straight channel flow presents negligible shape variation after a short initial transient regime and one can be noted by checking out the full straight lines in the same images colored in red and blue. Furthermore, recall that the initial bubble shape at all simulations were the equal with drop’s size factor and a short transient period for all test cases were expected with time for the straight channel with single and multiples drops and for all sinusoidal channels simulations.
Figure 4 shows the steady shape of a single drop in straight channel with blue color representing the suspending fluid (diethylene–glycol) and the drop fluid (UCON-1145). As can be seen, the large minimum film thickness at the tail of the drop stabilizes the flow and no further variation in its shape is seen. Additionally, due to the size of the drop relative to the channel’s diameter, the channel wall constricts the drop’s shape to an elongated drop with a round nose due to the hydrodynamics and compacts the tail from which the minimum liquid film thickness
is measured. The drop’s shape, drop’s velocity, and the minimum liquid film thickness remain constant for
.
A more challenging test case is presented in
Figure 5 where a single drop shape evolution with time in sinusoidal channel can be seen during its passage in one corrugated length where the drop undergoes severe topological change. The amplitude of the wavy channel wall is
, and a detailed description of the geometry is presented in
Figure 2. At time
, the drop leaves one divergent section and approaches the convergent section; however, the channel’s constriction, where the minimum cross-sectional area is found, stretches out the drop’s shape to its minimum thickness where the interface approaches considerably the symmetry wall; therefore, suggesting a possible topological change due to the interface breakup. However, for the two-phase system numerically investigated here, no breakup has been seen for any of the presented simulations with viscosity ratio
. According to the work in [
10], the breakup of drops was observed in systems with lower viscosity ratios, more specifically in GW5 systems with (
). Later, with time
, the drops shrinks to its minimum shape length and reaches the minimum film thickness to the wall before the convergent section begins. In time
the drop stretches over again and a fixed periodic shape motion is noted with time
.
Slug flow is now analyzed, and the drop’s shape evolution with time in straight channel with amplitude
and initial liquid slug length
is presented in
Figure 6. As can be noted, the center drop (green color) approaches the front drop (red color) at time
, where coalescence takes place. It can also be seen that before drops collision, the center drop elongates and the minimum liquid film thickness
becomes larger if compared to time
. Such a behavior is due to the increase of drop’s velocity as result of lower pressure behind the front drop. The same physical process was also seen when the initial slug length
s is smaller, i.e.,
. However, the due to the smaller initial distance of the initial slug, the coalescence time is shorter; therefore, for
, the following coalescence times were computed:
, respectively. The drop shapes of the rear and front drops remained fairly constant during the simulations.
A slug flow is analyzed in a corrugated channel with amplitude
, and the drop’s shape evolution with time is presented in
Figure 7 for the initial liquid slug length
. The center drop (green color) approaches the front drop (red color) at time
, where coalescence takes place, which is faster than the straight channel with same initial conditions. The same physical process was also seen when the initial slug length
s was smaller, i.e.,
. As seen in the straight channel, due to the smaller initial distance of the initial slug, the coalescence time is shorter, therefore for
the following coalescence time were computed
respectively. The drop shapes of the rear and front drops presented fairly the same dynamical shape of the single drop in sinusoidal channel of
Figure 5.
The evolution of the drop’s rising velocity with time is presented for all test cases simulated in this work in
Figure 8. The subplots are divided according to the initial slug length
s, and the single drop flow in straight and sinusoidal channels is used as a reference at all plots. As can be seen, the smaller is the initial slug length
s, the faster the coalescence takes place, thus resulting in topological change of the two front drops. The drops flow in sinusoidal presented same wavy pattern of velocity variation with time to all test cases. The full lines were used to represent the drop velocity of single drop in straight channel (red color) and sinusoidal (blue color) and the others line types were used for the slug flows in both channels.
Two more interesting results are presented in
Figure 9 for initial slug length
, where the variation of the ratio of the drop’s perimeter
relative to its initial perimeter
is investigated in two corrugated lengths in
Figure 9a, and the minimum film thickness
evolution is presented for two corrugated lengths in
Figure 9b. As reference, the single drop flow in the straight channel (red color) and the sinusoidal channel (blue color) is plotted against the slug flow in sinusoidal channel. As can be noted, the drop’s perimeter ratio
of the slug flow differs to that of the single drop flow (blue color) when the liquid film thickness hits its maximum length; however, its evolution returns to the same behavior of the single drop flow. On the other hand, the liquid film thickness in the slug flow do not present significant variations if compared to the single drop flow in sinusoidal channel; however, it can be noted that the sinusoidal channel geometry reduces considerably the mean liquid film thickness when compared to the single drop in straight channel (red color).
The axisymmetric streamlines
are plotted along with the velocity components
and
and pressure
p in
Figure 10a to highlight the wake structure of center and front drop’s before coalescence in slug flow with drop’s initial slug lengths
. In
Figure 10b,c the velocity components,
and
, and pressure are interpolated axially between
at radius
, respectively, showing the variation of drop’s velocity and the pressure drop between the center and front drops.
Figure 11 depicts the trend of the coalescence time
relative to the initial slug length
s for the straight and sinusoidal channels. As can be noted, for initial slug length
, the coalescence time increases linearly as
s increases. Additionally, as noted previously, the sinusoidal channel anticipates the coalescence time for the same flow parameters of the straight channel.
In the first parametric study, the wavelength has been systematically changed for values while keeping constant the drop’s size factor and all dimensionless numbers used at all simulations for two initial slug lengths . The standard test case for this parametric study is the straight channels with in which the wavelength approaches infinity, where the coalescence time was , respectively, for the initial slug lengths . In the sinusoidal channels, the wavelength parameters presented coalescence time of , respectively, indicating that the coalescence time is not linearly related to the channel’s wavelength as can be seen in Figure 13.
In
Figure 12, the final solution before coalescence is shown for all wavelength tested in this parametric study for initial slug length
and the respective coalescence time. For the wavelength
, no coalescence between the drops took place for time
; therefore, such a wavelength seems to be critical for the simulation parameters used in this work. All the others wavelengths have presented lower coalescence time relative to the straight channel with same parameters.
Figure 13 reveals the coalescence time
for the straight
and sinusoidal
channels relative to the wavelengths. Is is clear that the wavelength
is critical for the presented simulation parameters where the coalescence time
is much larger than the straight channels with same simulation parameters. On the other hand, all the others wavelengths anticipates the coalescence process when compared to the straight channels.
In the last parametric study, the drop size factor is systematically changed for values
while keeping constant the channel’s wavelength
, initial slug length
, and amplitude
for the corrugated channel as well as for the straight channel
. In
Figure 14, the center drop’s perimeter ratio
within two corrugated lengths is presented. As expected, the larger is the drop’s size factor
, the larger is the perimeter ratio variation.
Figure 15 presents a parametric study of coalescence time
between drops in slug flow for the sinusoidal channel with amplitude
, initial slug length
, and several drop’s size factor
. For
the rear and center drops coalesce, where for the others, the coalescence takes place between the center and front drops. Before coalescence, the drop geometries are similar for
and it takes place in the same channel region. The drop’s body increases and the drop’s size factor increases.
Figure 16 shows the trend of the coalescence time relative to the initial slug length
s for the sinusoidal channel with amplitude
. The drop’s size factor
presented larger coalescence time
for corrugated channels relative to the straight channels with same parameters. The others drop’s size factors presented smaller coalescence time for the corrugated channel with amplitude
relative to the straight channel
. Therefore, it is expected to have smaller coalescence time in corrugated channels.