You are on page 1of 13

AIAA SciTech Forum 10.2514/6.

2020-1578
6-10 January 2020, Orlando, FL
AIAA Scitech 2020 Forum

3D Simulations of Droplets Impacting Liquid Films: Crown


Parameters Measurements

Daniel Vasconcelos∗ ,Daniela Ribeiro† , André Silva‡ and Jorge Barata§


Universidade da Beira Interior, Covilhã, 6200-001, Portugal

The 3D incompressible Navier-Stokes equations are coupled with the CLSVOF method and
employed to numerically simulate the phenomena of single droplet impact onto liquid films. A
solution-adaptive mesh refinement tool, based on the gradient of the volume fraction scalar, is
adopted in order to reduce computational cost. Three different fluids are taken into account:
100% jet fuel and 75%/25% and 50%/50% of jet fuel and biofuel, respectively. Quantitative
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

analysis of the crown height and outer diameter is performed for different impact conditions,
such as the influence of the impact velocity and dimensionless thickness, between experimental
and numerical results, and the qualitative analysis includes the occurrence of splashing and
overall crown evolution. Numerical results show that the crown outer diameter measurements
are in good agreement with the experimental cases, presenting a slight discrepancy for the
lower liquid film thickness of h∗ = 0.2. The crown height measurements are under-predicted
for the current model, maintaining a similar trend for dimensionless thicknesses of h∗ = 0.5
and h∗ = 1 while, for the lower thickness, the crown disintegrates at earlier stages. The crown
curvature and rim instabilities exhibit significant differences, and the splashing phenomenon
occurs for both the experimental and numerical outcomes.

I. Nomenclature

D0 = Droplet Diameter, mm
Dout = Crown Outer Diameter, mm

Dout = Dimensionless Crown Outer Diameter
h = Liquid Film Thickness, mm
h∗ = Dimensionless Liquid Film Thickness
H = Crown Height, mm
H∗ = Dimensionless Crown Height
t = Time subsequent to Impact, ms
U0 = Droplet Impact Velocity, m/s
α = Volume Fraction
µ = Viscosity, kg/(m2 · s)
ϕ = Level-set function
ρ = Density, kg/m3
σ = Surface Tension, mN/m
τ = Dimensionless Time
θw = Static contact angle, ◦

II. Introduction
Multiphase flows are often encountered in nature and industrial processes in different structures, namely slug flows,
fluidized beds, droplet and bubble flows, among others. More specifically, research on the phenomena of droplet
impact onto liquid films has been motivated due to its relevance in several applications, including fuel spray injection
∗ Ph.D. Student, Aerospace Sciences Department, daniel.vasconcelos.rodrigues@ubi.pt.
† Ph.D. Student, Aerospace Sciences Department, daniela.santo.ribeiro@ubi.pt.
‡ Assistant Professor, Aerospace Sciences Department, and Member of AIAA, andre@ubi.pt.
§ Full Professor, Aerospace Sciences Department, and Associate Fellow of AIAA, jmmbarata@gmail.com.

Copyright © 2020 by Daniel Vasconcelos, Daniela Ribeiro, André Silva and Jorge Barata.
Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
in internal combustion engines, spray cooling and coating, ink-jet printing, pesticide spraying of crops, etc. [1, 2].
In recent decades, there has been substantial development in terms of theoretical, experimental and numerical
studies for this kind of phenomenon. Coghe et al. [3] experimentally studied single water droplets impacting onto
liquids films. The authors measured several crown characteristics, such as its diameter, height, and nominal thickness,
as well as the number and growth of the jets from the crown rim. It was reported that the impact Weber number
influences the non-dimensional crown height maximum value and correspondent non-dimensional time. Also, the
non-dimensional crown thickness increases from 3 to 5 times the initial value prior to impact, regardless of the film
thickness and impact velocity. Cossali et al. [4] performed similar work, studying the morphology of impact and
the influence of time in the development of the crown diameter, height, and secondary drop diameters. Conclusions
were in agreement with Coghe et al. [3], stating that the crown thickness evolution and the growing velocity of the
crown height are almost independent of the Weber number. It was also concluded that the mean secondary droplet size
increases with time for a higher impact velocity. However, it is almost independent of time for the lowest Weber number
analyzed. Davidson [5] numerically simulated a droplet impact onto a liquid film for inviscid flow and axisymmetric
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

assumption with the purpose of examining jetting behavior for early stages of the impact, and predicting the crown
evolution and shape for later stages. The obtained results suggest that, for h∗ < 0.5, an increase in the film thickness
leads to an increase in height and thickness of the crown. A few years later, Mukherjee and Abraham [6] also studied
the behavior of the crown that emerges subsequent to impact of a single droplet onto a liquid film. However, their
conclusions show that both the crown height and diameter display a non-monotonic behavior with the variation of the
film thickness. For h∗ < 0.25, an increase in film thickness causes an increase in the rate of increase of the crown
height and diameter, while, for h∗ > 0.25, these geometrical parameters tend to decrease with the increase of film
thickness.
Yarin [7] reviewed, in detail, droplet impact dynamics onto thin liquid films and dry surfaces, focusing on how the
different physical properties and surface characteristics affect the splashing phenomenon. Besides including several
experimental studies regarding this matter, theoretical and numerical modeling are also presented, in particular the
existence of a kinematic discontinuity in the velocity distribution, predictions of crown position, and crown rim shape.
Conclusions of this review relate to the fact that impacting droplets onto very thin liquid films require more attention, as
well as an understanding of the nature of the splashing transition, crown formation, and propagation. Recently, Ribeiro
et al. [8] focus on studying the implementation of jet fuel (JF) and biofuel (HVO) mixtures in the aviation sector due
to the negative influence of fossil fuels on climate change and greenhouse gas emissions. Their work consisted of
impact characterization and phenomena visualization of single droplets impinging upon liquid films. Several impact
conditions, such as the impact velocity, the droplet diameter and the dimensionless thickness of the liquid film, as well
as four different fluids: water, jet fuel, and mixtures of jet fuel and biofuel (75%/25% and 50%/50%, respectively),
were varied. It was concluded that the physical properties of fluids and the dimensionless thickness of the liquid
film influences the outcome, and the size and number of splashed droplets changes with the impact energy and the
dimensionless thickness.
The main purpose of this study is numerically simulating the three-dimensional phenomena of droplet impact onto
liquid films using the CLSVOF method and a solution-adaptive mesh refinement tool. Regarding qualitative results,
an analysis of the overall droplet impact phenomena in terms of the initial impact, crown formation, and secondary
atomization is performed. Concerning quantitative results, several geometrical parameters of the crown formation,
namely its diameter and height, are measured numerically and experimentally.

III. Numerical Setup/Model


The numerical model solves the 3D incompressible Navier-Stokes equations coupled with the CLSVOF method.
This method couples two different interface tracking mechanisms: the VOF [9] and the level-set [10] methods. The
VOF method is a free-surface technique capable of modeling several immiscible fluids by solving a single set of
momentum equations and tracking the volume fraction, α, of each phase on the computational domain. Therefore, in
each control volume, the volume fraction ranges from α = 0, for the gas-phase, to α = 1, for the liquid-phase, where a
value of α between 0 and 1 consists of the interface between phases. Despite conserving mass quite well, this method,
due to discontinuities across the interface, is weak in calculating spatial derivatives, more specifically the interface
normal vector and curvature. In order to overcome these weaknesses, the level-set method is coupled to the VOF
method. Since the level-set function, ϕ, is smooth and continuous throughout the domain, it provides accurate means in
calculating spatial gradients. Although this method has no conservation properties, the CLSVOF method combines the
advantages of the VOF and the level-set function, guaranteeing mass conservation and higher accuracy calculations in

2
terms of interface and curvature. Both these functions are advected similarly, as displayed by the following equations:
∂α ( )
® =0
+ ∇ · Uα (1)
∂t
∂ϕ ( )
® =0
+ ∇ · Uϕ (2)
∂t
where U® is velocity. The governing equations for mass and momentum conservation are displayed by equations 3
and 4, respectively,

∇ · U® = 0 (3)

∂ ( ) ( ) [ ( )]
ρ (ϕ) U® + ∇ · ρ (ϕ) U® U® = −∇p + ∇ · µ (ϕ) ∇U® + ∇U® T + ρ (ϕ) g® − F®s f
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

(4)
∂t
where ρ is density, µ is viscosity, p is pressure, g® is the gravitational acceleration constant and F®s f are the surface
tension forces.
The level-set function is used to represent the interface as a signed distance. Correspondingly, this function assumes
the interface as a zero level-set and can be expressed, in a two-phase flow, as :

 +|d|



for the gas-phase
ϕ(x, t) = 0 for the interface (5)


 −|d| for the liquid-phase

where d is the distance from the interface.
The surface tension forces are computed as an explicit term in the momentum equation by the CSF model [11] and
smoothed by a Heaviside function, Hϕ :

Fs f = 2Hϕ σκδ (ϕ) n® (6)


 0 |ϕ| > a for the gas-phase



Hϕ = 1 [ ( )] |ϕ| > a for the liquid-phase (7)

 ϕ πϕ
1 1 + + π1 sin |ϕ| ≤ a for the interface
2 a a

where κ is the curvature of the interface, δ (ϕ) is a delta function, n® is the unit normal vector to the interface,
a = 1.5l is the thickness of the interface and l is the minimum size of the cell. The unit normal vector, n®, and the
curvature, κ, can be approximated as:
∇ϕ
n® = (8)
|∇ϕ|
∇ϕ
κ =∇· (9)
|∇ϕ|
and the delta function, δ (ϕ), is defined as:
{
1+cos(πϕ/a)
|ϕ| < a
δ (ϕ) = 2a (10)
0 |ϕ| ≥ a
The physical properties of density and viscosity can be defined using the level-set function:
( )
ρ (ϕ) = ρg + ρl − ρg H (ϕ) (11)

( )
µ (ϕ) = µg + µl − µg H (ϕ) (12)
where the subscripts g and l refer to the gas-phase and the liquid-phase, respectively.

3
Figure 1 displays the schematic of the numerical model. A liquid droplet, of diameter D0 and impact velocity U0 ,
impinges onto a liquid film of thickness h. The droplet and the liquid film, of density, ρl , viscosity, µl and surface
tension, σ, are surrounded by air, of density, ρg , and viscosity, µg . Due to the 3D nature of the model, and in order to
efficiently consume computational resources, only 1/4 of the droplet impact phenomenon is numerically simulated. In
terms of boundary conditions, six boundaries define the outer limits of the model. The lower boundary is the impact
surface, on which the liquid film is at rest, and it is defined as a stationary wall with a no-slip condition applied. The
respective contact angles of the different fluids are specified on this boundary. Since only 1/4 of the impact domain is
simulated, as priory mentioned, the two planes in contact with the droplet are defined as symmetry planes. Opposed
to these planes are two stationary walls with no-slip condition, which we denominate as radial walls. Unlike the
impact surface, the contact angle for the radial walls is not considered, since the effects of the surface tension forces
are ignored between these walls and the interface. The upper boundary is defined as a pressure-outlet with no static
pressure for an operating pressure of P = 101.33 kPa. Gravity is enabled and acts on the negative direction of the
y-axis, gy = −9.81 m/s2 . Three different fluids are taken into account: 100% jet fuel and 75%/25% and 50%/50% of
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

jet fuel and biofuel, respectively. The physical properties of the fluids and contact angles are presented in table 1. Jet
A-1 and NExBTL (an hydroprocessed vegetable oil) are the jet fuel and biofuel adopted, respectively.

Table 1 Physical properties of density, dynamic viscosity, surface tension and contact angles for the different
fluids [8, 12].

Fluid ρ [kg/m3 ] µ [mPa.s] σ [mN/m] θ w [◦ ]


100% JF 798 1.12 25.4 ≈0
75% JF / 25% HVO 795 1.44 25.5 ≈0
50% JF / 50% HVO 792 1.79 24.6 ≈0
Air 1.23 1.79 × 10−2 - -

Fig. 1 Schematic of the numerical model.

The mesh consists of a structured orthogonal grid and, in order to reduce the overall mesh size and computational
cost, a solution-adaptive mesh refinement is performed. This adaption method will refine the liquid-gas interface while
maintaining the remaining mesh domain (in cells where the volume fraction is α = 1 or α = 0) to a minimum number
of points. The properties of this adaptive mesh refinement consist of a gradient adaption approach based on the volume
fraction scalar, where the boundary value for the refinement/coarsening of the grid is equal to 0.1. Therefore, it refines
or coarsens only the regions that display higher or lower values than the boundary value, respectively. This adaption

4
is performed dynamically for each time-step to ensure the execution of the gradient adaption automatically. Figure 2
displays how the refinement/coarsening occurs over the domain for a dimensionless time of τ = tU0 /D0 = 0.

Fig. 2 Adaptive mesh refinement of the liquid-gas interface for τ = 0.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

The numerical analysis is performed using the CFD simulation software ANSYS® 2019 R3. The solution methods
to be used in the calculation are based on a pressure-based solver, meaning that mass conservation of the velocity field is
achieved by solving a pressure-correction equation. The correspondent coupling of the pressure and velocity is achieved
by using the PISO scheme. Several schemes are adopted for the discretization of the numerical model: Least Squares
Cell Based scheme for the gradient discretization, PRESTO! for the pressure discretization, the Third-Order MUSCL
for the momentum and level-set function discretization and, for the reconstruction of the interface (based on the volume
fraction), the Geometrical Reconstruction scheme. The finite-volume method (FVM) is used to discretize the domain
into a finite number of control volumes. The time-step is adjusted according to the CFL (Courant-Friedrichs-Lewy)
condition, and it is defined as CFL=0.2 for the numerical simulations.
For the computational domain, in order to guarantee that the height and radial distance have no influence in the
impact phenomena in terms of crown geometrical parameters and influence of capillary waves, a radial independence
study is required. The impact conditions for this and the following independence studies refer to the case A3 described
in table 2, where the fluid is 100% JF, U0 = 3m/s, D0 = 3.0mm and h∗ = 0.5. Figure 3 presents measurements of the
dimensionless outer diameter of the crown, Dout ∗ = Dout /D0 , as a function of dimensionless time for radial distances
of 5D0 , 6D0 and 7D0 . The maximum relative error between the radial distance of 6D0 and 7D0 is 1.32% and, therefore,
a radial distance of 6D0 is adopted for the numerical model, as displayed by figure 1. The height of the computational
domain is defined as 5D0 and it is sufficient for the visualization of the crown development and secondary atomization.

4
5D0
D*out

3 6D
0
7D0
2

0
1 2 3 4 5 6 7 8 9 10 11

Fig. 3 Dimensionless crown outer diameter measurements as a function of dimensionless time for the radial
independence study.

5
Once the computational domain sizes are established, one has to guarantee that the solution is independent of
the mesh resolution. As previously mentioned, the mesh is structured and orthogonal and, in order to perform the
mesh independence study, a base mesh with a cell size of 1/3 of the droplet diameter is considered. From this point
forward, several layers of adaptive mesh refinement are applied to the gradient of the volume fraction (which refers
to the interface between the liquid and the gas phase) until mesh independence is obtained. Figure 4 displays the
dimensionless crown outer diameter measurements as a function of dimensionless time for three different levels of
adaptive refinement: 1/12, 1/24 and 1/48 of the droplet diameter, which corresponds to a level 2, 3 and 4 adaptive-mesh
refinement, respectively. Similarly to the radial independence study, the maximum relative error between the level 3
and 4 adaptive-mesh refinement is 3.9%. Due to these reasons, a base mesh with 1/3 and a level 3 interface refinement
with 1/24 of the droplet diameter is adopted for the following numerical simulations.

6
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

4
D0/12
D*out

3 D0/24
D0/48
2

0
1 2 3 4 5 6 7 8 9 10 11

Fig. 4 Mesh independence study based on dimensionless crown outer diameter measurements as a function of
dimensionless time.

IV. Results and Discussion


The qualitative and quantitative analysis are performed with the objective of understanding the influence of different
parameters, such as the droplet diameter, impact velocity, dimensionless thickness, and the physical properties of the
fluids on the crown geometrical parameters. Figure 5 displays a developing crown and how its properties, namely the
height and diameter, are determined. The crown height is evaluated from the crown base to the free rim and the crown
outer diameter is measured at the free rim height.

Fig. 5 Height and outer diameter measurements of a developing crown.

6
Table 2 shows the different study cases that were considered for the experimental and numerical analysis regarding
the fluid, droplet diameter, impact velocity, and dimensionless thickness, h∗ = h/D0 . The following subsections
correlate the effect of kinematic and physical properties on the crown outer diameter and height measurements, as well
as a qualitative evaluation of the splashing phenomenon, crown shape and overall growth.

Table 2 Cases of study for quantitative and qualitative measurements.

Case Fluid D0 [mm] U0 [m/s] h∗


A1 100% JF 3.0 1.8 0.2
A2 100% JF 3.0 3.0 0.2
A3 100% JF 3.0 3.0 0.5
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

A4 100% JF 3.0 4.1 0.2


B1 75% JF / 25% HVO 3.0 3.0 0.2
B2 75% JF / 25% HVO 3.0 3.0 0.5
B3 75% JF / 25% HVO 3.0 3.0 1.0
C1 50% JF / 50% HVO 3.1 3.0 0.5

A. Impact Velocity
Figures 6 and 8 exhibit experimental and numerical measurements of the dimensionless crown outer diameter and
height, H ∗ = H/D0 , as a function of dimensionless time for the cases A1, A2 and A4, which consist of the same fluid,
droplet diameter and dimensionless thickness for different values of the impact velocity. According to the experimental
results, an increase in the impact velocity prompts an increase in both the crown outer diameter and height, being the
effect in the latter more pronounced than in the former. The peak values for the crown outer diameter and height, due to
the dependence of the impact velocity on the dimensionless time, occur for later stages with the increase of the impact
velocity. Regarding the numerical simulations, the crown outer diameter measurements, despite being slightly lower
than the experimental results, follow a similar trend. On the contrary, the crown height measurements are lower than
the experimental cases and, for cases A2 and A4, the crown disintegrates at approximately τ = 5, causing a gradual
decrease of its height.

4
D*out

3 A1, Exp.
A2, Exp.
2 A4, Exp.
A1, Num.
1 A2, Num.
A4, Num.
0
2 4 6 8 10 12 14

Fig. 6 Comparison between experimental and numerical results of crown outer diameter measurements as a
function of time, both dimensionless, for impact velocity variations.

7
τ=0 τ=0

τ = 2.70 τ = 2.70
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

τ = 5.40 τ = 5.40

τ = 8.10 τ = 8.10

τ = 10.81 τ = 10.81

a) b)

Fig. 7 Visualization of the phenomenon of droplet impact for 100%JF (D0 =3.0mm, U0 =4.1m/s and h* =0.2): a)
Experimental Results; b) Numerical Analysis.

8
3

2.5

A1, Exp.
2 A2, Exp.
A4, Exp.
H*

1.5 A1, Num.


A2, Num.
1 A4, Num.
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

0.5

0
2 4 6 8 10 12 14

Fig. 8 Comparison between experimental and numerical results of crown height measurements as a function
of time, both dimensionless, for impact velocity variations.

Figure 7 exhibits qualitative comparisons between the experimental and numerical results of the A4 case. The
moment prior to the droplet connecting with the liquid film is defined as τ = 0. Due to the initial impact, the fluid
begins to expand up and outwards of the impact region, and a crown starts forming, as can be visualized for τ = 2.7,
with the occurrence of prompt splashing at earlier stages. This type of splashing is characterized by the ejection of tiny
secondary droplets from the tip of the lamella, usually for t << D0 /U0 . With the development of the phenomenon, for
τ = 5.40, the crown will continue to grow in terms of height and diameter, both experimentally and numerically. The
height obtained numerically is noticeably lower than the experimental case, which can also be correlated with figure 8.
From this point forward, for τ = 8.10 and τ = 10.81, crown splashing occurs for the experimental crown, which relates
to secondary droplets being ejected from the rim at later stages of the phenomenon. This also occurs for the numerical
case, however the crown is not able to maintain its shape and disintegrates, which explains the height measurements
discrepancies in figure 8.

B. Dimensionless Thickness
Figures 9 and 10 display the dimensionless crown outer diameter and height measurements, respectively, as a
function of dimensionless time, for the cases B1, B2 and B3, both experimentally and numerically, to understand
the influence of the dimensionless thickness on the geometrical parameters of the crown. The tendencies of the
experimental results indicate that an increase of the dimensionless thickness leads to a decrease in the crown diameter
and height. This is more noticeable for later stages, as the differences in the measured parameters is more perceptible.
With regard to the numerical analysis, the crown outer diameter measurements are in good agreement with the
experimental results, with a slight disparity present in case B1, which refers to a dimensionless thickness of h∗ = 0.2.
The crown height measurements present similar discrepancies to figure 8, where the values are considerably lower than
the experimental results. However, B2 and B3, despite presenting lower crown height values, maintain an identical
trend to the experimental evaluation, while the case B1 shows a significant decrease at early stages, concluding that the
current model does not correctly predict the crown height formation and, specifically, does not estimate the trend of the
crown height for lower dimensionless thicknesses (h∗ = 0.2). Figure 11 presents the visualization of the experimental
and numerical results for the B3 case. Once the droplet impacts the liquid film (τ = 0), a small crown starts to develop
and splashing occurs in its rim for both cases, at τ = 2. The crown keeps advancing and, for τ = 4, the experimental
crown displays several instabilities at the crown rim. On the contrary, the numerical crown rim is smooth and free of
disturbances, suggesting that the model does not correctly simulate these instabilities. For later stages of the impact,
τ = 6 and τ = 8, the crown shape does not vary significantly. It is also possible to visualize that the overall curvature
of the crown is not identical between the experimental and numerical results.

9
7
B1, Exp.
6 B2, Exp.
B3, Exp.
5 B1, Num.
B2, Num.
4 B3, Num.
D*out

2
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

0
2 4 6 8 10 12

Fig. 9 Comparison between experimental and numerical results of crown outer diameter measurements as a
function of time, both dimensionless, for liquid film thickness variations.

3
B1, Exp.
2.5 B2, Exp.
B3, Exp.
B1, Num.
2 B2, Num.
B3, Num.
H*

1.5

0.5

0
2 4 6 8 10 12

Fig. 10 Comparison between experimental and numerical results of crown height measurements as a function
of time, both dimensionless, for liquid film thickness variations.

10
τ=0 τ=0

τ=2 τ=2
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

τ=4 τ=4

τ=6 τ=6

τ=8 τ=8

a) b)

Fig. 11 Visualization of the phenomenon of droplet impact for the 75%/25% mixture (D0 =3.0mm, U0 =3.0 m/s
and h* =1): a) Experimental Results; b) Numerical Analysis.

11
C. Fluids
Figures 12 and 13 present experimental and numerical results of crown outer diameter and height measurements, as a
function of time, all nondimensionalized, for the cases A3, B2 and C1, which refer to similar impact conditions in terms
of impact velocity, droplet diameter and liquid film thickness for different fluids. Despite these fluids not presenting
significant differences regarding density, viscosity, surface tension, and contact angles, the numerical simulations and
correspondent crown geometrical measurements will aid in comprehending the limitations of the numerical model.
The experimental results indicate that the small variance of the physical properties of the fluids is not sufficient to
obtain clear conclusions of their influence, as the crown diameter and height measurements are similar for all the cases.
The numerical simulations exhibit analogous behavior in comparison with the previous quantitative analysis, where the
diameters of the crown are in good agreement with the experimental results, while the crown height is under-predicted
for the different stages of the droplet impact phenomenon with similar trends.
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

4
D*out

3 A3, Exp.
B2, Exp.
2 C1, Exp.
A3, Num.
1 B2, Num.
C1, Num.
0
2 4 6 8 10 12

Fig. 12 Comparison between experimental and numerical results of crown outer diameter measurements as a
function of time, both dimensionless, for different fluids.

3
A3, Exp.
2.5 B2, Exp.
C1, Exp.
A3, Num.
2 B2, Num.
C1, Num.
H*

1.5

0.5

0
2 4 6 8 10 12

Fig. 13 Comparison between experimental and numerical results of crown height measurements as a function
of time, both dimensionless, for different fluids.

12
V. Conclusions
3D numerical simulations of droplets impacting onto thin liquid films for distinct impact conditions and fluids
were performed and validated with available experimental data. Qualitative and quantitative analysis regarding crown
geometrical parameters, i.e. the diameter and height, and the overall development of the phenomenon were presented,
respectively. The numerical crown outer diameter measurements are in good agreement with the experimental results,
maintaining the overall increasing tendency and showing a slight disparity in cases for lower thicknesses (h∗ = 0.2).
In relation to the crown height, the numerical model strongly under-predicts its progression, with the measurements
being considerably lower than the experimental results for all the cases. For dimensionless thicknesses of h∗ = 0.5 and
h∗ = 1, despite the under-prediction of the height, the trend is similar to the experimental rate of increase. However,
for the lower thickness cases, h∗ = 0.2, the crown collapses at earlier stages, causing a rapid decrease in its height.
In qualitative terms, the crown overall curvature and rim instabilities are not accurately simulated and the splashing
phenomenon occurs for both the numerical and experimental results. Future work includes different adaptive-mesh
refinement approaches, such as the curvature of the interface or the vorticity, base mesh sizes and solution methods for
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on May 30, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-1578

height stabilization.

Acknowledgments
The present work was performed under the scope of Aeronautics and Astronautics Research Center (AEROG) of the
Laboratório Associado em Energia, Transportes e Aeronáutica (LAETA) activities and it was supported by Fundação
para a Ciência e Tecnologia (FCT) through the project UID/EMS/50022/2019 and by the Ph.D. scholarships with the
references SFRH/BD/140009/2018 and SFRH/BD/143307/2019.

References
[1] Panão, M. O., and Moreira, A. L. N., “Flow characteristics of spray impingement in PFI injection systems,” Experiments in
Fluids, Vol. 39, No. 2, 2005, pp. 364–374.

[2] van Dam, D. B., and Le Clerc, C., “Experimental study of the impact of an ink-jet printed droplet on a solid substrate,” Physics
of Fluids, Vol. 16, No. 9, 2004, pp. 3403–3414.

[3] Coghe, A., Brunello, G., Cossali, G., Marengo, M., and TeMPE-CNR, M.-I., “Single drop splash on thin film: measurements
of crown characteristics,” ILASS Europe, Vol. 99, 1999.

[4] Cossali, G., Marengo, M., Coghe, A., and Zhdanov, S., “The role of time in single drop splash on thin film,” Experiments in
Fluids, Vol. 36, No. 6, 2004, pp. 888–900.

[5] Davidson, M. R., “Spreading of an inviscid drop impacting on a liquid film,” Chemical Engineering Science, Vol. 57, No. 17,
2002, pp. 3639–3647.

[6] Mukherjee, S., and Abraham, J., “Crown behavior in drop impact on wet walls,” Physics of fluids, Vol. 19, No. 5, 2007, p.
052103.

[7] Yarin, A. L., “Drop impact dynamics: splashing, spreading, receding, bouncing,” Annu. Rev. Fluid Mech., Vol. 38, 2006, pp.
159–192.

[8] Ribeiro, D., Cunha, N., Barata, J., and Silva, A., “Dynamic Behaviour of Single Droplets Impinging upon Liquid Films with
Variable Thickness: Jet A-1 and HVO Mixtures,” 14th Triennial International Conference on Liquid Atomization and Spray
Systems, Chicago, IL, USA, 2018.

[9] Hirt, C. W., and Nichols, B. D., “Volume of fluid (VOF) method for the dynamics of free boundaries,” Journal of computational
physics, Vol. 39, No. 1, 1981, pp. 201–225.

[10] Osher, S., and Sethian, J. A., “Fronts propagating with curvature-dependent speed: algorithms based on Hamilton-Jacobi
formulations,” Journal of computational physics, Vol. 79, No. 1, 1988, pp. 12–49.

[11] Brackbill, J. U., Kothe, D. B., and Zemach, C., “A continuum method for modeling surface tension,” Journal of computational
physics, Vol. 100, No. 2, 1992, pp. 335–354.

[12] Vasconcelos, D., “Numerical Analysis of a Single Droplet Impinging upon Liquid Films using the VOF Method,” Master’s
thesis, Universidade da Beira Interior, Covilhã, Portugal, 2018.

13

You might also like