scispace - formally typeset
Search or ask a question
Journal ArticleDOI

The history effect on bubble growth and dissolution. Part 2. Experiments and simulations of a spherical bubble attached to a horizontal flat plate

TL;DR: In this paper, the authors provide an experimental and numerical study of the history effect regarding a spherical bubble attached to a horizontal flat plate and in the presence of gravity, and demonstrate the importance that boundary-induced advection has to accurately describe bubble growth and shrinkage.
Abstract: The accurate description of the growth or dissolution dynamics of a soluble gas bubble in a super- or undersaturated solution requires taking into account a number of physical effects that contribute to the instantaneous mass transfer rate. One of these effects is the so-called history effect. It refers to the contribution of the local concentration boundary layer around the bubble that has developed from past mass transfer events between the bubble and liquid surroundings. In Part 1 of this work (Penas-Lopez et al., J. Fluid Mech., vol. 800, 2016b, pp. 180–212), a theoretical treatment of this effect was given for a spherical, isolated bubble. Here, Part 2 provides an experimental and numerical study of the history effect regarding a spherical bubble attached to a horizontal flat plate and in the presence of gravity. The simulation technique developed in this paper is based on a streamfunction–vorticity formulation that may be applied to other flows where bubbles or drops exchange mass in the presence of a gravity field. Using this numerical tool, simulations are performed for the same conditions used in the experiments, in which the bubble is exposed to subsequent growth and dissolution stages, using stepwise variations in the ambient pressure. Besides proving the relevance of the history effect, the simulations highlight the importance that boundary-induced advection has to accurately describe bubble growth and shrinkage, i.e. the bubble radius evolution. In addition, natural convection has a significant influence that shows up in the velocity field even at short times, although given the supersaturation conditions studied here, the bubble evolution is expected to be mainly diffusive.

Summary (2 min read)

1. Introduction

  • Mass transfer processes involving bubbles have gained a renewed interest over the last few years due to their relevance in modern microfluidic applications connected to topics such as carbon sequestration (Sun & Cubaud 2011; Volk et al. 2015).
  • It has been shown that any recent history of growth and/or dissolution (triggered by past changes in ambient pressure) experienced by a particular bubble may leave, at least for some time, a non-negligible imprint on the current state of the concentration profile surrounding such a bubble.
  • It is worth mentioning that history effects are common to problems in which diffusion plays a central role, such as the viscous drag around a body or, closer to the present mass transfer problem, the heat transfer around a spheroid (Michaelides 2003).
  • With the purpose of maintaining the standard nomenclature, R0 will be used throughout this paper.
  • Section 3 presents the general mathematical formulation of the problem and sheds light on the importance of the different physical effects involved in the experiments.

2. Experimental characterisation of the history effect

  • The authors have carried out experiments to support their theoretical and numerical analyses by subjecting single bubbles to well-controlled, step-like pressure jumps that super- or undersaturate the liquid alternatively.
  • It becomes apparent through the differences in the responses to successive identical pressure–time histories.
  • Once the measurement tank is filled with the carbonated water, the following experimental procedure is followed: (i) The pressure is lowered below the saturation value until a bubble nucleates at the pit and grows up to the desired size, Ri. (ii) The tank pressure is set again to the saturation value.
  • In all cases the growth rate during the second stage lies above that of the first one, although both curves eventually converge at longer times, when the memory of the previous dissolution stage damps out.
  • It is interesting to compare this behaviour with that found when the bubble is forced to first dissolve and then to grow .

3. Numerical analysis: problem formulation

  • Which involves a non-stationary boundary and that must be coupled with the equations of motion for the liquid assuming axisymmetry around the vertical axis.the authors.
  • The initial concentration of dissolved gas is assumed to be uniform throughout the liquid and equal to C∞, equal to the gas concentration in the far field.
  • This knowledge will now be used in the next section when implementing the equations into a numerical model.
  • The vorticity field is then allowed to independently evolve through the vorticity transport equation, advancing with time step 1τv.

5. Simulation results and discussion

  • The simulation predictions for the bubble size history are compared with the experiments in figure 11.
  • The history effect on the growth rate is evident: the concentration contours in (b) are noticeably closer together than those in (a).
  • The results of their simulations can also be used to validate the hypothesis made by Enríquez et al. (2014) about the existence of a ‘dead zone’ near the contact point where mass transfer is almost zero.
  • In figure 14 the authors plot the local mass flux distribution along the bubble surface for different time instants during the first growth cycle of experiment 2.
  • In dissolution, a low velocity recirculation region surrounding the bubble is observed.

6. Conclusions

  • The authors have experimentally and numerically explored the influence of the past history of the ambient pressure experienced by a bubble on its instantaneous rate of mass transfer – the so-called history effect.
  • This effect is caused by a history-induced preexisting concentration boundary layer of dissolved gas that surrounds the bubble at the beginning of a given growth or dissolution stage.
  • The authors would naively expect that such a situation would lead to a monotonic dissolution, since the liquid is undersaturated during the whole process.
  • By performing order of magnitude analyses, the authors show that their experiments belong to a regime dominated by mass and viscous diffusion.
  • Thus, the momentum equation can be decoupled from the mass transfer problem.

Did you find this useful? Give us your feedback

Figures (17)

Content maybe subject to copyright    Report

This is a postprint version of the following published document:
Peñas-López, P., Moreno Soto, A., Parrales, M.A., van der Meer,
D. (2017). The history effect on bubble growth and dissolution.
Part 2. Experiments and simulations of a spherical bubble
attached to a horizontal flat plate. Journal of Fluid Mechanics,
vol 820, pp 479-510.
DOI: 10.1017/jfm.2017.221
©Cambridge University Press 2017

The history effect on bubble growth and
dissolution. Part 2. Exper iments and
simulations of a spherical bubble attached
to a hori zont al flat plate
Pablo Peñas-López
1,
, Álvaro Moreno Soto
2,
, Miguel A. Parrales
3
,
Devaraj van der Meer
2
, Detlef Lohse
2
and Javier Rodríguez-Rodríguez
1
1
Fluid Mechanics Group, Universidad Carlos III de Madrid, Avda. de la Universidad 30,
28911 Leganés (Madrid), Spain
2
Physics of Fluids Group, Faculty of Science and Technology, University of Twente, P.O. Box 217,
7500 AE Enschede, The Netherlands
3
Departamento de Ingeniería Energética y Fluidomecánica, Escuela Técnica Superior de Ingenieros
Industriales, Universidad Politécnica de Madrid, C. José Gutiérrez Abascal 2, 28006 Madrid, Spain
The accurate description of the growth or dissolution dynamics of a soluble gas
bubble in a super- or undersaturated solution requires taking into account a number
of physical effects that contribute to the instantaneous mass transfer rate. One of
these effects is the so-called history effect. It refers to the contribution of the local
concentration boundary layer around the bubble that has developed from past mass
transfer events between the bubble and liquid surroundings. In Part 1 of this work
(Peñas-López et al., J. Fluid Mech., vol. 800, 2016b, pp. 180–212), a theoretical
treatment of this effect was given for a spherical, isolated bubble. Here, Part 2
provides an experimental and numerical study of the history effect regarding a
spherical bubble attached to a horizontal flat plate and in the presence of gravity. The
simulation technique developed in this paper is based on a streamfunction–vorticity
formulation that may be applied to other flows where bubbles or drops exchange
mass in the presence of a gravity field. Using this numerical tool, simulations are
performed for the same conditions used in the experiments, in which the bubble is
exposed to subsequent growth and dissolution stages, using stepwise variations in the
ambient pressure. Besides proving the relevance of the history effect, the simulations
highlight the importance that boundary-induced advection has to accurately describe
bubble growth and shrinkage, i.e. the bubble radius evolution. In addition, natural
convection has a significant influence that shows up in the velocity field even
at short times, although given the supersaturation conditions studied here, the
bubble evolution is expected to be mainly diffusive.
Key words: bubble dynamics, convection, multiphase and particle-laden flows
Email addresses for correspondence: papenasl@ing.uc3m.es, a.morenosoto@utwente.nl
1

1. Introduction
Mass transfer processes involving bubbles have gained a renewed interest over the
last few years due to their relevance in modern microfluidic applications connected
to topics such as carbon sequestration (Sun & Cubaud 2011; Volk et al. 2015). Due
to the small size of these bubbles they are spherical once they become smaller
than the channel’s size and are detached from the channel’s wall. Thus, in general
terms, the theory of Epstein & Plesset (1950) describing the diffusion-driven growth
or dissolution of an isolated, spherical particle should be applicable. However, as
discussed in Part 1 of this work (Peñas-López et al. 2016b), a number of effects not
included in the Epstein–Plesset theory, e.g. flow around the bubble, must be taken
into account to properly describe various experimental observations. Bubbles may also
interact with nearby surfaces or they may contain more than one chemical species
(Shim et al. 2014; Peñas-López, Parrales & Rodríguez-Rodríguez 2015). Another
effect that contributes to the diffusion-driven dynamics of a bubble is the so-called
history effect, discussed in Part 1 and more recently in Chu & Prosperetti (2016b).
It has been shown that any recent history of growth and/or dissolution (triggered by
past changes in ambient pressure) experienced by a particular bubble may leave, at
least for some time, a non-negligible imprint on the current state of the concentration
profile surrounding such a bubble. Consequently, the mass transfer rate is affected as
well. In Part 1, we proposed a modification to the theory of Epstein & Plesset to
take into account the history effect through a memory integral term for the case of
spherical, isolated bubbles. Moreover, we applied this modified equation to calculate
the bubble radius evolution when the bubble is subjected to some simple, yet relevant,
pressure–time histories. It is worth mentioning that history effects are common to
problems in which diffusion plays a central role, such as the viscous drag around
a body or, closer to the present mass transfer problem, the heat transfer around a
spheroid (Michaelides 2003).
The primary goal of the present paper is to quantify the relative importance of the
history effect in a canonical, yet experimentally relevant, configuration that does not
exhibit spherical symmetry, namely, that of a single spherical bubble tangent to a
horizontal flat plate that grows and dissolves in response to changes in the ambient
pressure and in the presence of gravity. In this configuration, the existence of the
history effect may become noticeable with a simple experiment: let us consider such
a spherical CO
2
bubble that dissolves when the pressure is above saturation (see
figure 1). At a given time t 60 s, the pressure is lowered to a new value still above
saturation (figure 1b). Despite the pressure being at all times above saturation, after
changing the pressure, the bubble is observed to grow for some time (figure 1a).
Naturally, part of this growth is due to the expansion of the gas. Thus, to observe
the effect purely due to diffusion, it is convenient to plot the ambient radius, R
0
. It
is defined as the radius one would observe if the liquid surroundings were at the
reference ambient pressure, P
0
, instead of the actual ambient pressure P
(t):
R
0
(t) =R(t)
P
(t)
P
0
1/3
. (1.1)
Here, R(t) is the measured bubble radius. Still, the ambient radius can be seen
to grow until approximately t 100 s, an effect purely driven by diffusion. Note
that R
0
was referred to in Part 1 of this article (Peñas-López et al. 2016b) as the
pressure-corrected radius R
corr
. However, with the purpose of maintaining the standard
nomenclature, R
0
will be used throughout this paper.
2

0.24
0.26
0.28
0.30
0.32
(a)
Simulation
t (s)
0 50 100 150
6.0
6.5
7.0
7.5
01234
0
10
20
30
40
(b)
(c)
FIGURE 1. Dissolution of a CO
2
spherical-cap bubble tangent to a flat chip immersed
in a CO
2
-water solution under pressurised conditions (see later figure 9). The bubble is
subjected to (b) a pressure jump P
(t), from P
(0) = P
0
= 7.4 bar to 6.5 bar. Both
pressures are above the saturation pressure, P
sat
=6.1 bar (according to simulation). Panel
(a) shows the evolution in time of the measured bubble radius R(t) (white markers)
and ambient radius R
0
(t) (dark markers). The former is compared to simulation, which
in addition was employed to depict (c) the concentration profile along the z-axis above
the bubble at three different instants in time. The employed experimental and numerical
techniques are detailed in the main text.
This phenomenon may be explained by examining the concentration of dissolved
CO
2
near the bubble (figure 1c). Indeed, although the concentration at the bubble
surface, given by Henry’s law, responds instantaneously to pressure changes, there
exists a boundary layer around the bubble where the concentration of CO
2
is higher
than the instantaneous saturation one, as a result of the dissolution stage that took
place before the pressure drop. In the example depicted in figure 1(c), it can be seen
how the concentration gradient at the bubble’s top is actually positive at t = 65 s,
which explains the growth of the ambient radius. In this figure, numerical simulations
such as the ones described in §§ 35 have been used to compute the concentration
field along the z axis. These simulations are validated by comparing the predicted
bubble radius with the experimental one (see figure 1a).
This simple example illustrates that, to properly describe the time evolution of the
bubble radius observed in experiments, the history effect must be taken into account.
However, a question that was left open in our previous work (Peñas-López et al.
2016b) was the relative importance of this effect in a realistic experimental condition
where other effects such as the interference with a wall and natural convection may
greatly influence the diffusion-driven bubble dynamics, as was shown by Enríquez
et al. (2014). With this idea in mind, another objective of the present work is
to propose a numerical approach able to accurately describe the evolution of a
bubble attached to a horizontal flat plate and growing/dissolving in the presence of a
gravitational field.
While this work only deals with bubbles composed of a single soluble gas, it is
important to realise that the history effect is omnipresent in multicomponent bubbles.
In Part 1, the history effect was described as ‘the acknowledgement that at any
given time the mass flux across the bubble is conditioned by the preceding time
3

history of the concentration at the bubble interface’. Thus, in dissolving/growing
multicomponent bubbles, the flow rate of a particular species across the bubble
interface will likely be different from the rest. The species composition inside the
bubble will thus change over time, which amounts to time-dependent partial pressures
and hence time-dependent interfacial concentrations. It is possible to artificially
discern the contribution of the history effect numerically, as was done by Chu &
Prosperetti (2016b) for the case of a dissolving two-gas bubble. Isolating the history
effect experimentally, on the other hand, is anticipated to be much harder.
Finally, it is worth mentioning that the history effect is naturally present in the
evaporation of multicomponent drops. Chu & Prosperetti (2016a) have recently
developed a formulation that includes a memory integral to describe the diffusion-
driven dynamics of multicomponent drops in the presence of a solvent, a phenomenon
of relevance in modern techniques of chemical analysis (Lohse 2016). In this problem,
the faster or slower dissolution of one of the components yields a time-varying
composition at the drop’s interface, which makes the inclusion of the history integral
in Fick’s law essential, even when the ambient pressure remains constant.
The paper is structured as follows: § 2 presents the experimental results that
illustrate the effect of history in the growth–dissolution of CO
2
bubbles tangent to
a flat plate. Section 3 presents the general mathematical formulation of the problem
and sheds light on the importance of the different physical effects involved in the
experiments. In § 4, a formulation based on the streamfunction–vorticity method
is described to simulate the mass transfer and flow field around the bubbles. The
simulation results are then presented and discussed in § 5. Finally, § 6 summarises
the main conclusions.
2. Experimental characterisation of the history effect
We have carried out experiments to support our theoretical and numerical analyses
by subjecting single bubbles to well-controlled, step-like pressure jumps that super- or
undersaturate the liquid alternatively. This way, we can make bubbles grow and shrink
under repeatable conditions to expose the history effect. It becomes apparent through
the differences in the responses to successive identical pressure–time histories.
2.1. Experimental set-up and procedure
Although the experimental set-up has been described in a previous work (Enríquez
et al. 2013), a brief description is included here for convenience. The facility is fed
with water that is demineralised in a purifier (MilliQ A10) and degassed by making
it flow through a filter (MiniModule, Liquicel, Membrana). This water enters into the
mixing chamber (see figure 2), that has been previously flushed with CO
2
to purge
the air from the system. There the water is stirred in the presence of CO
2
, kept at
the desired saturation pressure for approximately 45 min. Finally, the experimental
tank is pressurised with CO
2
at this same pressure and then slowly flooded with
the carbonated water, so bubbles do not appear during the filling. This preparation
procedure ensures that in the experimental tank there are no other gas species present
within the liquid or gas phases apart from CO
2
(at least in quantifiable amounts).
Placed at the centre of the experimental tank there is a silicon chip, treated to
become hydrophilic, with a black-silicon hydrophobic pit (50 µm in radius) at its
centre. The role of this pit is to force a single bubble to nucleate at a fixed location
in a repeatable way. Furthermore, in order to avoid slight temperature variations
4

Citations
More filters
Journal ArticleDOI
05 Apr 2018
TL;DR: In this paper, asymptotic analysis is used to characterize collective shielding effects in bubble lattices, the resulting bubble lifetime and dissolution dynamics, which is driven by the diffusion of gas within the surrounding liquid.
Abstract: Microscopic bubble dissolution is driven by the diffusion of gas within the surrounding liquid. Asymptotic analysis is used to characterize collective shielding effects in bubble lattices, the resulting bubble lifetime and dissolution dynamics.

37 citations

Journal ArticleDOI
31 Oct 2017-Langmuir
TL;DR: It is observed that H2 bubble successions at large gas-evolving substrates first experience a stagnation regime, followed by a fast increase in the growth coefficient before a steady state is reached, which clearly contradicts the common assumption that constant current densities must yield time-invariant growth rates.
Abstract: Control over the bubble growth rates forming on the electrodes of water-splitting cells or chemical reactors is critical with respect to the attainment of higher energy efficiencies within these devices. This study focuses on the diffusion-driven growth dynamics of a succession of H2 bubbles generated at a flat silicon electrode substrate. Controlled nucleation is achieved by means of a single nucleation site consisting of a hydrophobic micropit etched within a micrometer-sized pillar. In our experimental configuration of constant-current electrolysis, we identify gas depletion from (i) previous bubbles in the succession, (ii) unwanted bubbles forming on the sidewalls, and (iii) the mere presence of the circular cavity where the electrode is being held. The impact of these effects on bubble growth is discussed with support from numerical simulations. The time evolution of the dimensionless bubble growth coefficient, which is a measure of the overall growth rate of a particular bubble, of electrolysis-gene...

35 citations

Journal ArticleDOI
TL;DR: The current results show that the dynamical evolution of bubbles is influenced by comprehensive effects combining chemical catalysis and physical mass transfer, and the size of the bubbles at the moment of detachment is determined by the balance between buoyancy and surface tension and by the detailed geometry at the bubble’s contact line.
Abstract: Whereas bubble growth out of gas-oversatured solutions has been quite well understood, including the formation and stability of surface nanobubbles, this is not the case for bubbles forming on catalytic surfaces due to catalytic reactions, though it has important implications for gas evolution reactions and self-propulsion of micro/nanomotors fueled by bubble release. In this work we have filled this gap by experimentally and theoretically examining the growth and detachment dynamics of oxygen bubbles from hydrogen peroxide decomposition catalyzed by gold. We measured the bubble radius R(t) as a function of time by confocal microscopy and find R(t) ∝ t1/2. This diffusive growth behavior demonstrates that the bubbles grow from an oxygen-oversaturated environment. For several consecutive bubbles detaching from the same position in a short period of time, a well-repeated growing behavior is obtained from which we conclude the absence of noticeable depletion effect of oxygen from previous bubbles or increasing oversaturation from the gas production. In contrast, for two bubbles far apart either in space or in time, substantial discrepancies in their growth rates are observed, which we attribute to the variation in the local gas oversaturation. The current results show that the dynamical evolution of bubbles is influenced by comprehensive effects combining chemical catalysis and physical mass transfer. Finally, we find that the size of the bubbles at the moment of detachment is determined by the balance between buoyancy and surface tension and by the detailed geometry at the bubble's contact line.

34 citations

Journal ArticleDOI
TL;DR: In this paper, the authors study the successive quasi-static growth of many bubbles from the same nucleation site described in this paper and show that the radius-versus-time curves of subsequent bubbles differ from each other due to this phenomenon.
Abstract: When a gas bubble grows by diffusion in a gas-liquid solution, it affects the distribution of gas in its surroundings. If the density of the solution is sensitive to the local amount of dissolved gas, there is the potential for the onset of natural convection, which will affect the bubble growth rate. The experimental study of the successive quasi-static growth of many bubbles from the same nucleation site described in this paper illustrates some consequences of this effect. The enhanced growth due to convection causes a local depletion of dissolved gas in the neighbourhood of each bubble beyond that due to pure diffusion. The quantitative data of sequential bubble growth provided in the paper show that the radius-versus-time curves of subsequent bubbles differ from each other due to this phenomenon. A simplified model accounting for the local depletion is able to collapse the experimental curves and to predict the progressively increasing bubble detachment times.

24 citations

Journal ArticleDOI
TL;DR: In this article, the formation of gas bubbles at gas cavities located in walls bounding the flow has been studied in many technical applications, but it is usually hard to observe.
Abstract: The formation of gas bubbles at gas cavities located in walls bounding the flow occurs in many technical applications, but is usually hard to observe. Even though, the presence of a fluid flow undoubtedly affects the formation of bubbles, there are very few studies that take this fact into account. In the present paper new experimental results on bubble formation (diffusion-driven nucleation) from surface nuclei in a shear flow are presented. The observed gas-filled cavities are micrometre-sized blind holes etched in silicon substrates. We measure the frequency of bubble generation (nucleation rate), the size of the detaching bubbles and analyse the growth of the surface nuclei. The experimental findings support an extended understanding of bubble formation as a self-excited cyclic process and can serve as validation data for analytical and numerical models.

23 citations

References
More filters
Journal ArticleDOI
TL;DR: An experimental system to study gas bubble growth in slightly supersaturated liquids by working with carbon dioxide dissolved in water, pressurized at a maximum of 1 MPa and applying a small pressure drop from saturation conditions.
Abstract: We have designed and constructed an experimental system to study gas bubble growth in slightly supersaturated liquids. This is achieved by working with carbon dioxide dissolved in water, pressurized at a maximum of 1MPa and applying a small pressure drop from saturation conditions. Bubbles grow from hydrophobic cavities etched on silicon wafers, which allows us to control their number and position. Hence, the experiment can be used to investigate the interaction among bubbles growing in close proximity when the main mass transfer mechanism is diffusion and there is a limited availability of the dissolved species

64 citations

Journal ArticleDOI
TL;DR: In this paper, the Stokes equations are solved in a wedge geometry of arbitrary contact angle and the flow field is described by similarity solutions, with exponents that match the singular boundary condition due to evaporation.
Abstract: The evaporation of sessile drops in quiescent air is usually governed by vapour diffusion. For contact angles below , the evaporative flux from the droplet tends to diverge in the vicinity of the contact line. Therefore, the description of the flow inside an evaporating drop has remained a challenge. Here, we focus on the asymptotic behaviour near the pinned contact line, by analytically solving the Stokes equations in a wedge geometry of arbitrary contact angle. The flow field is described by similarity solutions, with exponents that match the singular boundary condition due to evaporation. We demonstrate that there are three contributions to the flow in a wedge: the evaporative flux, the downward motion of the liquid–air interface and the eigenmode solution which fulfils the homogeneous boundary conditions. Below a critical contact angle of , the evaporative flux solution will dominate, while above this angle the eigenmode solution dominates. We demonstrate that for small contact angles, the velocity field is very accurately described by the lubrication approximation. For larger contact angles, the flow separates into regions where the flow is reversing towards the drop centre.

52 citations

Journal ArticleDOI
TL;DR: In this paper, the dissolution process of small (equivalent) radius R0 < 1 mm) long-chain alcohol sessile droplets in water is studied, disentangling diffusive and convective contributions.
Abstract: The dissolution process of small (initial (equivalent) radius R0 < 1 mm) long-chain alcohol (of various types) sessile droplets in water is studied, disentangling diffusive and convective contributions. The latter can arise for high solubilities of the alcohol, as the density of the alcohol–water mixture is then considerably less than that of pure water, giving rise to buoyancy-driven convection. The convective flow around the droplets is measured, using micro-particle image velocimetry (mPIV) and the schlieren technique. When non-dimensionalizing the system, we find a universal Sh Ra1=4 scaling relation for all alcohols (of different solubilities) and all droplets in the convective regime. Here Sh is the Sherwood number (dimensionless mass flux) and Ra is the Rayleigh number (dimensionless density difference between clean and alcohol-saturated water). This scaling implies the scaling relation c / R5=4 0 of the convective dissolution time c, which is found to agree with experimental data. We show that in the convective regime the plume Reynolds number (the dimensionless velocity) of the detaching alcohol-saturated plume follows Rep Sc

52 citations

Journal ArticleDOI
TL;DR: It is concluded that the equilibrium regime is obtained by gas exchange between the bubbles and liquid phase, and SDS molecules on the gas-liquid interface form a diffusion barrier, which controls the dissolution behaviour and the eventual equilibrium radius of the bubble.
Abstract: We studied the dissolution dynamics of CO2 gas bubbles in a microfluidic channel, both experimentally and theoretically. In the experiments, spherical CO2 bubbles in a flow of a solution of sodium dodecyl sulfate (SDS) first shrink rapidly before attaining an equilibrium size. In the rapid dissolution regime, the time to obtain a new equilibrium is 30 ms regardless of SDS concentration, and the equilibrium radius achieved varies with the SDS concentration. To explain the lack of complete dissolution, we interpret the results by considering the effects of other gases (O2, N2) that are already dissolved in the aqueous phase, and we develop a multicomponent dissolution model that includes the effect of surface tension and the liquid pressure drop along the channel. Solutions of the model for a stationary gas bubble show good agreement with the experimental results, which lead to our conclusion that the equilibrium regime is obtained by gas exchange between the bubbles and liquid phase. Also, our observations from experiments and model calculations suggest that SDS molecules on the gas–liquid interface form a diffusion barrier, which controls the dissolution behaviour and the eventual equilibrium radius of the bubble.

46 citations

Journal ArticleDOI
TL;DR: The dissolution process of small (equivalent) radius $R_0 Ra_t, where $Ra_t = 12$ is the transition Ra-number as extracted from the data, was studied in this paper.
Abstract: The dissolution process of small (initial (equivalent) radius $R_0 Ra_t$, where $Ra_t = 12$ is the transition Ra-number as extracted from the data. For $Ra < Ra_t$ and smaller, convective transport is progressively overtaken by diffusion and the above scaling relations break down.

44 citations

Frequently Asked Questions (10)
Q1. What are the contributions mentioned in the paper "The history effect on bubble growth and dissolution. part 2. experiments and simulations of a spherical bubble attached to a horizontal flat plate" ?

Peñas-López et al. this paper proposed a modification to the theory of Epstein & Plesset to take into account the history effect through a memory integral term for the case of spherical, isolated bubbles. 

Mass transfer processes involving bubbles have gained a renewed interest over the last few years due to their relevance in modern microfluidic applications connected to topics such as carbon sequestration (Sun & Cubaud 2011; Volk et al. 2015). 

The total gas pressure inside the bubble, Pg, considering liquid–gas surface tension γlg, but neglecting inertial and viscous effects inside the gas phase, is given byPg = P∞ + 2γlg/R. (3.4)The mass transfer problem is closed with Fick’s first law, which sets the molar flow rate of gas across the bubble surface S to beṅ=D ∫S ∇C · n̂ dS, (3.5)where dS is an infinitesimal area element of the bubble surface, and n̂ is the outwardpointing unit normal from the bubble surface. 

It can be concluded that, although the instantaneous rate of mass transfer may only be slightly affected by advection, its effect accumulates over time and becomes important to describe the evolution of the bubble when subjected to successive expansion–compression cycles. 

The authors may bypass this limitation by modelling the effect of stratification essentially through just an effective increase (decrease) of mass transfer towards (from) the bubble. 

the assumption of perfectly spherical bubble at all time yields a relative error of less than 3 % as compared to the actual gas volume of the spherical cap and the pit. 

The vorticity field is then allowed to independently evolve through the vorticity transport equation, advancing with time step 1τv. 

despite the changes that convection induces in the velocity field, its effect on the concentration boundary layer near the bubble is minute, as is revealed by the comparison between figures 15(c) and 16(c). 

Another effect that contributes to the diffusion-driven dynamics of a bubble is the so-called history effect, discussed in Part 1 and more recently in Chu & Prosperetti (2016b). 

Let us define the a priori unknown corresponding (dimensionless) apparent velocity field as urel(η, ξ, τ )= urel,η êη + urel,ξ êξ .