Download as pdf or txt
Download as pdf or txt
You are on page 1of 3

Article

Cite This: Langmuir 2019, 35, 8294−8307 pubs.acs.org/Langmuir

Modeling Bubble Collisions at Liquid−Liquid and Compound


Interfaces
Travis S. Emery† and Satish G. Kandlikar*,†,‡

Microsystems Engineering Department and ‡Mechanical Engineering Department, Rochester Institute of Technology, 76 Lomb
Memorial Dr., Rochester, New York 14623, United States
*
S Supporting Information

ABSTRACT: The collision of a bubble at liquid−liquid, solid−liquid−liquid, and


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

gas−liquid−liquid interfaces, the latter two of which are referred to as compound


Downloaded via INDIAN INST OF TECH DELHI- IIT on January 16, 2021 at 03:53:25 (UTC).

interfaces, is modeled to predict the bubble’s velocity profile and the pressure buildup
and drainage rate of the film(s) formed at impact. A force balance approach,
previously outlined for bubble collisions at solid and free surfaces, is employed, which
takes into account four forces acting on the bubble: buoyancy, drag, inertia of the
surrounding liquid through an added mass force, and a film force resulting from the
pressure buildup in the liquid film formed between the bubble and the interface upon
impact. The augmented Young−Laplace equation is applied to define the pressure
buildup in the film(s), while lubrication theory is employed to define the film drainage rate(s) through the use of the Stokes−
Reynolds equation. This is the first time this modeling technique has been implemented for bubble collisions with these
interface types as all previous models have relied only on grid-based simulations. The models were validated through
experiments conducted here with water and silicone oils of various viscosities and from data found in literature. A reasonable
agreement is observed between the theoretical and experimental velocity profiles found for these liquid combinations under
varying conditions of impact velocity and top film thickness. The spatiotemporal film thickness and pressure profile evolution,
features not yet able to be captured through experiment, are also presented and discussed.

■ INTRODUCTION
Bubble interactions at multifluid interfaces find relevance in
However, certain assumptions allow some of these to be
neglected, as will be discussed later in The Force Balance
numerous industrial applications and environmental phenom- Model section of the paper. Previously derived numerical
ena. Bubble collisions with free and solid surfaces have been models from this same group for the impact of bubbles with
studied in relation to processes such as mineral flotation, foam mobile7 and immobile8 surface conditions are also discussed.
formation, and direct contact evaporation.1−3 Bubble passage These models couple the force balance model with the
through a liquid−liquid interface has been explored primarily Stokes−Reynolds−Young−Laplace equations to predict the
due to its importance in metallurgical applications, though bubble motion and the pressure buildup and drainage rate of
other processes such as nuclear reactor accident scenarios and the film formed between the bubble and the interface at
the ascent of plumes through the Earth’s mantle have also been impact. The Young−Laplace equation has wide reaching
noted as motivators. The collisions of bubbles with gas− applications including describing the interfaces of compound
liquid−liquid or solid−liquid−liquid interfaces are associated sessile drops and drops floating at a free surface.9,10 The solid
with a newly developed technique for nanoemulsion formation surface bubble collision models have been shown to be highly
and natural phenomena such as bubble collisions with the sea accurate in predicting not only the motion of the bubble but
surface microlayer or the collision of liquid-encapsulated also the spatiotemporal evolution of the film thickness.11 A
bubbles.4,5 In these applications, the flow dynamics associated similar approach was also recently applied to model a bubble
with the bubble collision play a crucial role in governing the collision with a free surface.12 Although the film thickness
overall process. evolution has not yet been validated for such collisions, the
The collision of bubbles at solid and free surfaces has been success of the previous solid surface collision model and the
studied extensively, and comparisons are often made between excellent agreement of this model in the prediction of the
them due to their similarity. A review of the rise and impact of bubble motion provide a strong basis for the model’s accuracy.
bubbles with solid surfaces is offered by Manica et al.6 They Additionally, high-speed monochromatic light interferometry
considered spherical bubbles in Stokes flow and deformable was recently introduced as a means to quantify the film
bubbles with higher Reynolds numbers. In either case, a drainage during the collision of a 365 μm-diameter bubble and
balance of forces was successfully used to predict the bubble
motion from rest to its terminal velocity and its impact with a Received: April 24, 2019
solid surface. The forces considered are buoyancy, drag, added Revised: May 23, 2019
mass, history, film, van der Waals, and electrical double layer. Published: May 29, 2019

© 2019 American Chemical Society 8294 DOI: 10.1021/acs.langmuir.9b01209


Langmuir 2019, 35, 8294−8307
Langmuir Article

a free surface.13 This technique could be used in the future to PP11−air interface collisions. This is due to the two mobile
provide further validation of the free surface collision model interfaces of the free surface collision, which offer little
with respect to its film drainage predictions. A number of other resistance to film thinning. The previously mentioned work by
experimental studies further examine various aspects of bubble Liu et al.19 extended this work to further understand the role of
collisions with free and solid surfaces such as the influence of surfactants on film drainage.
surfactants,2,14,15 oscillating interfaces,1,16,17 and bubble kinetic Just recently, the collision of a bubble at a gas−liquid−liquid
energy.3,18 Surfactants in particular introduce an added level of interface, therein labeled a compound interface, has begun to
complexity to the bubble collision process by influencing the be explored.4,5,37,38 Specifically, bubble collisions with an air−
mobility of the liquid−gas interface(s), which in turn affects oil−water surfactant interface were studied as a means of
the liquid film drainage rate. In a recent work from Liu et al.,19 dispersing submicrometer oil droplets in water to form
it was shown that the interface velocity is determined by a nanoemulsions.4,37 In this, a thin layer of oil, 0.1−2.0 mm in
balance of shear and Marangoni stresses at the interface. As the thickness, coats the top of a water bath. Bubbles released in
stresses evolve continuously during the collision, this implies water collide with the multiliquid interface and, upon rupture
that the drainage rate could change continuously during the of the oil and water films ahead of the bubble, generate
collision depending on the surfactant concentration and initial microdroplets of oil in water. The addition of a lipophilic
impact velocity. Other works have focused on bubble−solid material to the oil has also been studied as a scalable platform
surface interactions in relation to their application in flotation for creating functional nanoemulsions for use in areas such as
separation.20−23 Niecikowska et al.21 examined the effect of drug delivery, material science, biology, functional foods, and
surface charge and hydrophobicity in three-phase contact nutraceuticals.4 The stability of the bilayer film formed at
(TPC) formation and found that when surfaces are highly impact was also studied through the derivation of a numerical
hydrophobic, the effects of surface charge are negligible and model considering van der Waals intermolecular attractions to
surface roughness defines the rupture and TPC formation. provide parametric regimes where rupture behavior is
Hydrophilic surfaces were shown to enable the stability of the expected.38 A finite-difference/front-tracking method was
wetting films even in the presence of attractive electrostatic applied to model bubble collisions with this type of compound
forces. Krasowska and Malysa23 found that TPC formation can interface by solving the Navier−Stokes equations on a fixed
vary by over an order of magnitude depending on surface uniform Eulerian back grid and tracking the interface motion
roughness, with surfaces having a roughness below 1 μm on a Lagrangian front grid.5 A close agreement is found in
demonstrating four to five bubble collision cycles prior to comparison to experiments conducted with water as the
attachment and surfaces having a roughness of 50 μm or larger bottom liquid and various viscosity silicone oils as the top
inducing rupture on the first collision. liquid.
The collision of a bubble with a liquid−liquid interface has In the present work, the force balance approach previously
received much less research focus compared to a solid−liquid employed for modeling bubble collisions with solid and free
or gas−liquid interface. Most experimental studies focused on surfaces is extended to model bubble collisions at liquid−liquid
bubbles that pass through the liquid−liquid interface with a and compound interfaces. In addition to the gas−liquid−liquid
volume of the lower liquid entrained around or behind the compound interface, we also consider a bubble collision at a
bubble.24−34 These works have focused on the influence of solid−liquid−liquid compound interface. Previous numerical
upper and lower liquid properties,26 the formation of models of these phenomena have relied on grid-based
encapsulated bubbles, 24 stages of film rupture, 31 and numerical simulations. As noted for solid interface collisions,6
identifying and classifying associated flow regimes.34 Numer- a full solution of the Navier−Stokes equations to model such
ical models of this process have relied on grid-based simulation systems is complex to implement and computationally
techniques, either using commercial software such as ANSYS expensive. Furthermore, it is noted that very fine mesh sizes
Fluent or other customized schemes to solve the flow are needed to capture the film drainage, which is orders of
equations.27,28,35 Confining previous studies to bubbles that magnitude smaller than the bubble size. The force balance
do not pass through the interface but instead experience a approach offers a simplified method that overcomes these
bouncing behavior similar to that seen with solid and free challenges while further providing valuable insight into the
surface collisions significantly reduces the breadth of relevant flow dynamics. In comparison to the previous solid and free
previous works. In the grid-based simulations of such collisions surface collision models employing this approach, changes
from Bonhomme et al.,27 this was the only situation that the must be made in their derivation to account for the variation in
computational predictions differed dramatically from exper- interfacial versus bubble surface tension and the presence of a
imental observations. While the experiments showed the viscous and non-massless fluid above the interface. The present
bubble becoming trapped at the interface, the simulation work outlines this derivation and also provides the
experimental data to be used for validation.


predicted the bubble to pass through the interface. The noted
reason for this is that the flow in the film between the bubble
and the interface is poorly described due to its small thickness EXPERIMENTAL SETUP
in comparison to the grid size. Vakarelski et al. 36 The experimental setup used to capture bubble collisions with liquid−
experimentally compared the impact of varying size bubbles liquid and compound interfaces is shown in Figure 1. The primary
or water droplets with a PP11−water interface and a PP11−air housing consisted of a rectangular polycarbonate column with an
interface to study the influence of interface mobility on inner cross section of 5 cm × 5 cm and height of 15 cm. A glass
capillary tube with an inner diameter of 0.05 mm was positioned at
coalescence dynamics. The liquid−gas interfaces of the system the bottom of the container to generate single bubbles of radius 0.65
were fully mobile, while the liquid−liquid interfaces were ± 0.03 mm in water. This is in perfect agreement with the theoretical
considered immobile. Concerning bubbles, they found that the bubble radius of 0.65 mm predicted by Tate’s law for a capillary tube
coalescence induction time increased by 2 to 3 orders of of this size.39 Wall effects on rising bubbles have been previously
magnitude for PP11−water interface collisions as compared to shown to influence the bubble motion when the container radius is

8295 DOI: 10.1021/acs.langmuir.9b01209


Langmuir 2019, 35, 8294−8307
Langmuir Article

water and was then placed in a hot water bath for several hours. The
container was rinsed several more times with hot water to remove any
remaining contaminants. The system cleanliness can significantly alter
the behavior of the bubble during its rise and collision. Specifically,
the bubble surface mobility is very sensitive to system contamination
levels. For a “clean” bubble, the surface is fully mobile, while a
“contaminated” bubble has an immobile surface. These surface
conditions are reflected in the bubble’s terminal velocity in that clean
bubbles have significantly higher terminal velocities compared to
contaminated bubbles. A theoretical terminal velocity, VT, can be
determined by equating the buoyancy and drag forces to yield VT2 =
8Rg/3CD, where g is the gravitational constant, and CD is the drag
coefficient. The drag coefficient may be found for a clean or
contaminated bubble using the theory compiled by Loth,41 which is
Figure 1. Experimental setup used to capture bubble collisions with given in the Supporting Information. Based on this compiled theory, a
liquid−liquid and compound interfaces. 0.65 mm radius bubble in water has a terminal velocity of 34.0 cm/s if
it is clean and just 14.2 cm/s if it is contaminated. The terminal
less than three times the bubble diameter.40 Since the hydraulic radius velocity reduces very quickly once the contaminant concentration
of the column used here was over 19 times the bubble diameter, these reaches a critical threshold, and intermediate terminal velocities can
effects were deemed negligible. A peristaltic pump was used to control only be attained with very precise control of the surface-active
air flow to the capillary to form single bubbles. In this testing, substance concentration.42 The bubble terminal velocity found in the
deionized water was always used as the bottom liquid and 0.82, 4.59, present experiments coincided with that theoretically predicted for a
or 9.35 mPa·s silicone oil was used as the top liquid. In order to attain clean bubble, as demonstrated in the Experimental Validation section.
various film thicknesses, a prescribed volume of the top liquid, as As such, it was concluded that the bubbles rose with mobile surface
determined by the cross-sectional area of the container multiplied by conditions and the cleaning process was sufficient. Additionally, it was
the desired film thickness, was carefully added to the container. The noted that the bubbles rose in a rectilinear rise path as opposed to a
influence of the meniscus formed at the container edge was neglected zigzag pattern, which occurs with bubbles rising in water with radii
due to the relatively large cross-sectional area of the container. Film larger than ∼0.91 mm.43,44
thicknesses of 0.25, 0.50, and 1.00 mm were tested as compound The experimental procedure went as follows: after cleaning, the
interfaces. For solid−liquid−liquid compound interfaces, a solid container was filled with water to produce the desired distance
polypropylene surface was additionally placed above the top liquid. between the capillary and liquid−liquid interface, L. The specific
Polypropylene was chosen due to its oleophilic and hydrophobic volume of the top liquid was then carefully added to form the desired
properties, which promoted the stability of the oil layer formed top layer thickness. The polypropylene surface was then lowered into
between the water and the solid surface. For liquid−liquid interface place if we were testing for the solid−liquid−liquid interface
collisions, the top liquid layer thickness was increased to ∼10 mm, a configuration. The peristaltic pump was then actuated to form a
value much larger than the expected interface deformation. single bubble at the capillary orifice. The bubble collision at the
High-speed videos of the collision were captured using a Photron interface was captured using the high-speed camera. Video analysis
FASTCAM camera at 3000 fps. From these videos, the bubble size was then used to determine the bubble size, trajectory, and velocity.
and location were measured. A calibration for converting pixels to There are three primary liquid properties that govern the behavior
millimeters was determined prior to recording using an object of of the systems considered here: density (ρ), viscosity (μ), and surface
known size, such as the capillary tube outer diameter as shown in tension (σ). The subscripts B and T will be used to designate these
Figure 2a. The equivalent undeformed bubble radius, R, was properties in association with either the bottom or top liquid,
respectively. The experiments conducted here employed liquid
combinations made up from four different liquids: water and three
different silicone oils (SO) with viscosities ranging from 0.82−9.35
mPa·s. The pertinent properties for each of these liquids are listed in
Table 1. The properties of the silicone oils were taken as specified by

Table 1. Properties of Experimental Liquids at 25 °C


Figure 2. (a) 4 mm outer diameter capillary used for spatial density, ρ viscosity, μ surface tension, σ
liquid (kg/m3) (mPa·s) (mN/m)
calibration and horizontal and vertical bubble diameter measurement
and (b) bubble displacement used in velocity calculation. water 1000 1.00 72.0
SO-1 818 0.82 17.4
SO-5 918 4.59 19.7
calculated as 2R = (Dh2Dv)1/3,18 where Dh and Dv are the horizontal SO-10 935 9.35 20.1
and vertical bubble diameters, respectively, as shown in Figure 2a. The
bubble trajectory was determined by tracking the location of the
bubble center or side using the Photron FASTCAM Analysis software.
The instantaneous bubble velocity was determined from the trajectory
data using a second-order central finite difference scheme: the manufacturer,45−47 while the properties of water are found in the
y − yi − 1 literature.12 In the present experiments, water was always used as the
Vi = i + 1 bottom liquid, while one of the three silicone oils was used as the top
2Δt (1)
liquid. The interfacial tension of each of these combinations was
where y refers to the vertical position of the bubble as shown in Figure determined using pendant drop analysis. Full details of this testing
2a,b, the subscript i refers to the corresponding frame or time, and Δt and the analysis employed to determine the interfacial tension are
is the time between frames.
Between experiments, the system container was thoroughly washed discussed in the Supporting Information. The measured values for the
using hot water and dish soap to remove any contaminants or leftover interfacial tension between water and 0.82, 4.59, or 9.35 mPa·s
liquid. Once clean, the container was rinsed several times with hot silicone oil were 50, 51, and 48 mN/m, respectively.

8296 DOI: 10.1021/acs.langmuir.9b01209


Langmuir 2019, 35, 8294−8307

You might also like