1. Introduction
The evolution of bubbles in turbulence has multiple applications in environmental and industrial contexts, from exchange at the ocean–atmosphere interface (Deike, Melville & Popinet Reference Deike, Melville and Popinet2016; Deike & Melville Reference Deike and Melville2018), as bubble-mediated gas exchange accounts for a significant part of carbon dioxide ($\textrm {CO}_{2}$) uptake by the ocean (Reichl & Deike Reference Reichl and Deike2020), to bubble column chemical reactors (Risso Reference Risso2018), while providing an effective pathway to bring oxygen into lakes (Karn et al. Reference Karn, Monson, Ellis, Hong, Arndt and Gulliver2015). In an engineering context, bubble-mediated gas exchange is controlled by the level of agitation in the flow, which is itself influenced by the presence of bubbles, and in turn affects the bubble velocity statistics controlling the gas transfer from individual bubbles (Risso Reference Risso2018; Mathai, Lohse & Sun Reference Mathai, Lohse and Sun2020).
While empirical formulae (e.g. Colombet et al. Reference Colombet, Legendre, Risso, Cockx and Guiraud2015; Karn et al. Reference Karn, Monson, Ellis, Hong, Arndt and Gulliver2015) have been proposed for describing the gas transfer in bubble swarms, their general applicability and theoretical foundation remain an active research topic. Bubble-mediated gas transfer models in the context of ocean–atmosphere interactions often use formulae based on gas transfer theory (Levich Reference Levich1962) based on bubbles rising in quiescent flow (Woolf & Thorpe Reference Woolf and Thorpe1991; Keeling Reference Keeling1993; Liang et al. Reference Liang, McWilliams, Sullivan and Baschek2011; Deike & Melville Reference Deike and Melville2018). The diffusive gas transfer by a single bubble rising in a quiescent flow has been described theoretically by Boussinesq (Reference Boussinesq1905) and Levich (Reference Levich1962), within the assumption of negligible variation in bubble volume. The non-dimensional transfer rate from a single bubble is the Sherwood number ${Sh} = {k_L d_0}/{\mathscr {D}_l}$, where
$k_L$ is the transfer rate,
$d_0$ the bubble size and
$\mathscr {D}_l$ the gas diffusivity in the liquid, and reads
${Sh} = ({2}/{\sqrt {{\rm \pi} }})\sqrt {{Pe}}$, where
${Pe}=U d_0/\mathscr {D}_l$ is the bubble Péclet number and
$U$ the bubble rise velocity. Several studies dealing with mass transfer from bubbles rising in a quiescent liquid have developed numerical techniques to resolve interphase mass transfer and validated their results against diffusive mass transfer theory (Haroun, Legendre & Raynal Reference Haroun, Legendre and Raynal2010; Marschall et al. Reference Marschall, Hinterberger, Schüler, Habla and Hinrichsen2012; Deising, Marschall & Bothe Reference Deising, Marschall and Bothe2016; Fleckenstein & Bothe Reference Fleckenstein and Bothe2015; Deising, Bothe & Marschall Reference Deising, Bothe and Marschall2018). Direct numerical simulation (DNS) of mass transfer of dilute gas from rising bubble swarms by Roghair (Reference Roghair2012) concluded that the transfer rate increases as the gas hold-up (ratio of gas volume to total volume) increases.
The gas transfer at the interface between two fluids involving a turbulent flow has been described as presenting two regimes depending on the turbulent Reynolds number (Theofanous, Houze & Brumfield Reference Theofanous, Houze and Brumfield1976). At low Reynolds number, the transfer rate can be written as $k_L/u_0 \propto {Sc}^{-1/2}{Re}^{-1/2}$, where
${Sc}=\nu /\mathscr {D}_l$ is the Schmidt number (ratio of kinematic viscosity over mass diffusivity),
$u_0$ a characteristic turbulent velocity and
${Re}$ the turbulence Reynolds number. The high-Reynolds-number regime can be described through the action of the smallest eddies at the Kolmogorov scale enhancing the transfer, leading to
$k_L\sim {Sc}^{-1/2}(\nu \epsilon )^{1/4}$, where
$\epsilon$ is the turbulence dissipation rate. This can be expressed in terms of the turbulence Reynolds number and reads
$k_L/u_0 \propto {Sc}^{-1/2}\,{Re}^{-1/4}$ (Theofanous et al. Reference Theofanous, Houze and Brumfield1976), further discussed by Magnaudet & Calmet (Reference Magnaudet and Calmet2006) and Katul & Liu (Reference Katul and Liu2017). These two regimes and their crossover have been observed experimentally and numerically by Herlina & Wissink (Reference Herlina and Wissink2016, Reference Herlina and Wissink2019), following earlier experimental work by Fortescue & Pearson (Reference Fortescue and Pearson1967) (who used the root mean square of the fluctuating velocity as characteristic velocity).
The same reasoning can be applied in the context of bubbles in a turbulent flow. Levich (Reference Levich1962) provides a brief discussion on gas dissolution from a bubble of diameter $d_0$ in a turbulent stream with a characteristic velocity
$u_0$. The gas transfer rate can then be estimated as
${Sh} \propto {Re}^{3/4} {Sc}^{1/2}$, where
${Sc}=\nu /\mathscr {D}_l$ is the Schmidt number (ratio of kinematic viscosity over mass diffusivity) and
${Re}=d_0 u_0/\nu$ a turbulence Reynolds number, based on the bubble size
$d_0$ and a turbulence velocity
$u_0$ (the velocity scale considered by Levich (Reference Levich1962) is the maximum velocity of the eddies in the liquid that flows past the bubble). This regime is equivalent in terms of scalings to the high-Reynolds-number regime described by Theofanous et al. (Reference Theofanous, Houze and Brumfield1976) and Magnaudet & Calmet (Reference Magnaudet and Calmet2006), and its applicability in the context of mass exchange by a bubble swarm is discussed by Colombet et al. (Reference Colombet, Legendre, Risso, Cockx and Guiraud2015).
Numerical methods for interfacial mass transfer started with Sato, Jung & Abe (Reference Sato, Jung and Abe2000) and Davidson & Rudman (Reference Davidson and Rudman2002) where the dilute gas concentration is continuous across the interface. Bothe et al. (Reference Bothe, Koebe, Wielage, Prüss and Warnecke2004) introduced a method to simulate the discontinuous concentration due to solubility. A three-dimensional front tracking model with mass transfer was presented by Darmana, Deen & Kuipers (Reference Darmana, Deen and Kuipers2006). A one-fluid formulation for the algebraic volume-of-fluid method was presented independently by Haroun et al. (Reference Haroun, Legendre and Raynal2010) and Marschall et al. (Reference Marschall, Hinterberger, Schüler, Habla and Hinrichsen2012). Bothe & Fleckenstein (Reference Bothe and Fleckenstein2013) introduced a two-field approach using a geometrical volume-of-fluid method for multicomponent conjugate mass transfer. Subgrid-scale models to simulate high-Schmidt-number, bubble-mediated mass exchange have been developed by Bothe & Fleckenstein (Reference Bothe and Fleckenstein2013), Weiner & Bothe (Reference Weiner and Bothe2017) and Claassen et al. (Reference Claassen, Islam, Peters, Deen, Kuipers and Baltussen2020). Recent advances in numerical methods by Tanguy et al. (Reference Tanguy, Sagan, Lalanne, Couderc and Colin2014), Fleckenstein & Bothe (Reference Fleckenstein and Bothe2015), Maes & Soulaine (Reference Maes and Soulaine2020) and Scapin, Costa & Brandt (Reference Scapin, Costa and Brandt2020) allow the simulation of problems of mass transfer with local volume changes.
In the present work, we adapt the classic penetration theory (Higbie Reference Higbie1935) to describe the mass transfer from a bubble in a turbulent flow, by considering a turbulent time scale and predict the transfer rate of low-solubility gases, presented in § 2. We next present a framework in § 3 which combines recent advances in numerical algorithms for interfacial mass transfer, together with progress in turbulent multiphase flow modelling to characterize the diffusive mass transfer of dilute gas from a bubble in turbulence. Dilute gas diffusion from a bubble is similar to conjugate mass transfer from a spherical droplet (Rachih et al. Reference Rachih, Legendre, Climent and Charton2020). However, in the present study, the fluid inside the spherical cavity is considerably less dense than the surrounding fluid and hence is called a bubble. We use the Basilisk flow solver (Popinet & collaborators Reference Popinet2013–2020) which uses adaptive mesh refinement, a momentum-conserving scheme for velocity and a geometric volume-of-fluid method to capture the interface. We implement a module for the concentration advection and diffusion using one-fluid formulation without phase change (Haroun et al. Reference Haroun, Legendre and Raynal2010; Bothe & Fleckenstein Reference Bothe and Fleckenstein2013; Taqieddin Reference Taqieddin2018; Yang et al. Reference Yang, Peters, Fries, Harshe, Kuipers and Baltussen2020). The mass transfer module is validated by comparing the numerical results of diffusion from a static bubble with a solution using inverse Laplace transform, as well as for the classic diffusion from a rising bubble in a quiescent liquid. Finally, in § 4, we perform DNS of bubble-mediated mass transfer in a homogeneous and isotropic turbulent flow, solving the three-dimensional, incompressible, two-phase Navier–Stokes equations coupled with an advection–diffusion equation for the gas concentration. We consider bubbles that can deform but at Weber number below the breakup threshold. The theoretical model is confirmed by the DNS results for a wide range of turbulent Péclet numbers.
The present configuration of mass exchange of a dilute component from a bubble to the surrounding turbulent water is especially relevant for bubble-mediated $\textrm {CO}_2$ gas transfer at the ocean–atmosphere interface, as
$\textrm {CO}_2$ is present in a small concentration in the atmosphere so that its exchange with the surrounding water does not change the overall bubble volume, while gases such as
$\textrm {N}_2$ and
$\textrm {O}_2$ which contribute most of the volume of the bubbles have a much lower solubility and exchange over longer times (see detailed discussions in Woolf & Thorpe (Reference Woolf and Thorpe1991), Keeling (Reference Keeling1993), Liang et al. (Reference Liang, McWilliams, Sullivan and Baschek2011) and Deike & Melville (Reference Deike and Melville2018)).
2. Theory of bubble mass transfer in turbulence
We consider the mass transfer of dilute gas from a bubble of diameter $d_0$. The diffusivity of gas inside and outside the bubble is given by
$\mathscr {D}_g$ and
$\mathscr {D}_l$, respectively. The ratio of momentum to mass diffusivity defines the Schmidt number
${Sc}=\nu _l/\mathscr {D}_l$. The mass transfer predicted by the two-film theory (Whitman Reference Whitman1923) assumes a steady-state diffusion through the film, while in turbulent flows the fluctuations in the velocity field keep the diffusion transient (Treybal Reference Treybal1980). Higbie (Reference Higbie1935) penetration theory states that when eddies of liquid in a turbulent flow are exposed to bubbles for a time scale
$\theta$, then the mass transfer rate is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn1.png?pub-status=live)
In a turbulent flow, the eddies of various sizes will interact with the bubble and advect the gas present in the surrounding thin boundary layers. These interactions will be characterized by a turbulent velocity $\tilde {u}$. Several choices could be considered, in particular the large-scale fluctuation in velocity, the velocity fluctuations at the bubble scale or the small-scale fluctuations. We consider here that the large-scale fluctuations of the flow will drive the transfer process, and can be characterized by the magnitude of velocity fluctuations
$|\tilde {u}|=\sqrt {3}\,u_{{rms}}$. We propose that the time scale of exposure to eddies is given by
$\theta = {d_0}/(\sqrt {3}\,u_{{rms}})$. The transfer rate (2.1) is then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn2.png?pub-status=live)
For a bubble-mediated mass transfer in turbulent flow, we define the non-dimensional transfer rate or Sherwood number, ${Sh}={k_L d_0}/{\mathscr {D}_l}$, and the turbulent Péclet number,
$\widetilde {{Pe}}={u_{{rms}}d_0}/{\mathscr {D}_l}$. Finally, (2.2) can be written
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn3.png?pub-status=live)
This equation is analogous to that for the transfer rate for a bubble rising in a quiescent fluid, where the Péclet number would be defined based on the bubble rise velocity (Levich Reference Levich1962) and mass transfer for bubble swarms where the mean rise velocity of the bubbles in the swarm is considered (Colombet et al. Reference Colombet, Legendre, Risso, Cockx and Guiraud2015).
3. Numerical framework
3.1. The Basilisk solver
We solve the three-dimensional, incompressible, two-phase Navier–Stokes equations using the open-source solver Basilisk (Popinet Reference Popinet2009, Reference Popinet2015; van Hooft et al. Reference van Hooft, Popinet, van Heerwaarden, van der Linden, de Roode and van de Wiel2018):
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn4.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn5.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn6.png?pub-status=live)
where $\boldsymbol {u}$,
$p$,
$\gamma$,
$\mu$,
$\rho$,
$\kappa$,
$n$ and
$\mathcal {T}$ are the velocity, pressure, surface tension coefficient, viscosity, density, curvature, interface normal and volume fraction fields, respectively. The solver has been extensively validated for complex interfacial flows (Popinet Reference Popinet2015; Farsoiya, Mayya & Dasgupta Reference Farsoiya, Mayya and Dasgupta2017; Ruth et al. Reference Ruth, Mostert, Perrard and Deike2019; Gumulya et al. Reference Gumulya, Utikar, Pareek, Evans and Joshi2020; Berny et al. Reference Berny, Deike, Séon and Popinet2020; Mostert & Deike Reference Mostert and Deike2020). It uses the projection method to compute the velocity and pressure and the geometric volume-of-fluid method for the evolution of the interface between two immiscible fluids (Tryggvason, Scardovelli & Zaleski Reference Tryggvason, Scardovelli and Zaleski2011). The Piecewise Linear Interface Calculation geometric interface and flux reconstruction ensures a sharp representation of the interface (Scardovelli & Zaleski Reference Scardovelli and Zaleski1999; Marić, Kothe & Bothe Reference Marić, Kothe and Bothe2020) and is combined with an accurate height-function curvature calculation and a well-balanced, continuum surface tension model (Brackbill, Kothe & Zemach Reference Brackbill, Kothe and Zemach1992; Popinet Reference Popinet2018).
3.2. One-fluid formulation for mass transfer of dilute gas
The continuous formulation for mass transfer that we use in this study has been independently developed by Haroun et al. (Reference Haroun, Legendre and Raynal2010) and Marschall et al. (Reference Marschall, Hinterberger, Schüler, Habla and Hinrichsen2012). We have implemented the concentration diffusion of dilute gas (Haroun et al. Reference Haroun, Legendre and Raynal2010) using the harmonic mean diffusion coefficient as verified by Deising et al. (Reference Deising, Marschall and Bothe2016). Note that the present study investigates the mass transfer of dilute gas present in the bubble where the effects of loss of volume and phase change are ignored. Numerical methods for phase change require changes in the volume-of-fluid advection equation and continuity equation (Tanguy et al. Reference Tanguy, Sagan, Lalanne, Couderc and Colin2014; Fleckenstein & Bothe Reference Fleckenstein and Bothe2015; Maes & Soulaine Reference Maes and Soulaine2020; Scapin et al. Reference Scapin, Costa and Brandt2020). The time evolution of the $j$th gas concentration
$c_{l/g,j}$ for the liquid phase
$l$ or the gas phase
$g$ is given by (Standart Reference Standart1964; Haroun et al. Reference Haroun, Legendre and Raynal2010)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn7.png?pub-status=live)
The continuity of normal fluxes across the interface $\varSigma$ (Standart Reference Standart1964; Bothe & Fleckenstein Reference Bothe and Fleckenstein2013) is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn8.png?pub-status=live)
where $\boldsymbol {u}_\varSigma$ is the interface velocity. As we assume that the transfer of a dilute component does not cause volume change, (3.5) reduces to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn9.png?pub-status=live)
The standard assumption of continuous chemical potentials which is good for most applications at interface $\varSigma$ results in Henry's law (we refer interested readers to Bothe & Fleckenstein (Reference Bothe and Fleckenstein2013) for a discussion on local chemical equilibrium and the generalized Henry's law):
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn10.png?pub-status=live)
where the dimensionless ratio of the liquid-phase concentration to the gas-phase concentration of the component transferred $\alpha _j$ is called the Henry's law solubility constant (Sander Reference Sander2015) (solubility hereafter). The problem investigated is isothermal and there is no bubble breakup or large deformation of the interface which can change the pressure inside the bubble significantly. Hence, solubility which at least depends on temperature and pressure (Bothe & Fleckenstein Reference Bothe and Fleckenstein2013) is assumed to be constant. Introducing variables for the one-fluid formulation which is valid for both phases
$l$ and
$g$ and for gas
$j$, we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn11.png?pub-status=live)
and the flux is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn12.png?pub-status=live)
The diffusivity for the interfacial cells is calculated using the harmonic mean of the two diffusivities of the gas inside and outside the bubble (Haroun et al. Reference Haroun, Legendre and Raynal2010),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn13.png?pub-status=live)
to get a single equation for both phases:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn14.png?pub-status=live)
The coefficient of $c_j$ in the second term on the right-hand side of (3.11) can be written as
$\beta _j$:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn15.png?pub-status=live)
Using a time-implicit Euler discretization,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn16.png?pub-status=live)
and rearranging the implicit terms gives
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn17.png?pub-status=live)
Equation (3.14) is a set of linear equations which is solved efficiently using the multigrid method (Popinet Reference Popinet2015).
The solubility boundary condition (3.7) presents a discontinuity for the concentration field across the interface similar to the volume fraction field $\mathcal {T}$. As discussed by Bothe & Fleckenstein (Reference Bothe and Fleckenstein2013), a non-consistent advection leads to artificial diffusion for concentration. For a consistent advection two tracer fields
$\phi _{g}$ and
$\phi _{l}$ associated with the volume of fluid
$\mathcal {T}$ are defined:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn18.png?pub-status=live)
Using (3.7) and (3.8a,b), we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn19.png?pub-status=live)
The advection equation for $\phi _{g/l}$ reads
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn20.png?pub-status=live)
and is solved using the volume-of-fluid associated fields (López-Herrera et al. Reference López-Herrera, Ganan-Calvo, Popinet and Herrada2015) which guarantees strictly non-diffusive transport close to the interface. The concentration tracer is updated after advection using
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn21.png?pub-status=live)
3.3. Validation
We validate the numerical methods and their implementation for static and rising bubble cases.
3.3.1. Diffusion from a static bubble
We propose a new test case of diffusion from a constant-size, static, spherical bubble. Test cases are available for concentration profiles in the case of a planar interface in Bird, Stewart & Lightfoot (Reference Bird, Stewart and Lightfoot2002) and Haroun et al. (Reference Haroun, Legendre and Raynal2010). We provide the solution for the transient concentration both inside and outside the spherical bubble in the form of integrals which are then evaluated numerically.
Consider a static axisymmetric bubble of radius $R_0=d_0/2$. The diffusivities of the gas are
$D_g$ and
$D_l$ for the gas and liquid phase, respectively. The one-dimensional transient concentration diffusion for a spherical geometry for both inside and outside the bubble is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn22.png?pub-status=live)
Applying the Laplace transform to the above equations, we get $\tilde {c}(s)=\mathcal {L}[c(t)]$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn23.png?pub-status=live)
Given boundary conditions $r\tilde {c}_g\rightarrow 0$ as
$r\rightarrow 0$,
$\tilde {c}_l\rightarrow 0$ as
$r\rightarrow \infty$,
$[\kern-1pt[ \mathscr {D} \partial \tilde {c}/\partial r ]\kern-1pt] = 0$ and
$\tilde {c}_l/\tilde {c}_g=\alpha$ at
$r=R_0$, the concentrations (Laplace transformed) inside the bubble
$\tilde {c}_g$ and outside
$\tilde {c}_l$ are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn24.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn25.png?pub-status=live)
Equation (3.21a,b) can be inverted using the Cauchy residue theorem (using keyhole contour as discussed in Farsoiya, Roy & Dasgupta (Reference Farsoiya, Roy and Dasgupta2020)):
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn26.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn27.png?pub-status=live)
where $\mathrm {Im}(\cdot )$ is the imaginary part of a complex number. We validate the results of the Basilisk solver against (3.23) and (3.24) (using the numerical integration functions based on double-exponential quadrature in Wolfram Research, Inc. (2020)) in figure 1.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig1.png?pub-status=live)
Figure 1. Diffusion from a static bubble, comparing the numerical results with (3.23) and (3.24). (a) Concentration inside the bubble at $r/d_0=0.25$. (b) Concentration outside the bubble at
$r/d_0=0.75$. (c) Radial profile at time
$t\mathscr {D}_l/d_0^{2}=1.2$. (d) Maximum relative error at different resolutions
$\max |c_{11}-c_n|/c_{11}$, where
$c_{11}$ and
$c_n$ are numerical solutions at resolution
$2^{11}$ and lower, respectively, and for different solubilities
$\alpha$, displaying first-order convergence. The scripts sufficient to reproduce these results are provided in Farsoiya, Popinet & Deike (Reference Farsoiya, Popinet and Deike2020a).
A static bubble of diameter $d_0/L=0.2$, diffusivity ratio
$\mathscr {D}_g/\mathscr {D}_l=10$ and solubility
$\alpha =10^{-3}$ to
$10^{-1}$ is considered for this test. Figures 1(a) and 1(b) show the transient concentrations inside the bubble at
$r/d_0=0.25$ and outside the bubble at
$r/d_0=0.75$, respectively. The accuracy of the numerical solution is very good above 25 cells per diameter. The stringent solubility condition
$\alpha =10^{-3}$ is achieved at the interface
$r/d_0=0.5$ even at a low resolution of six cells per diameter (figure 1c). To quantify the convergence, we compute the norm
$\max |c_{11}-c_n|/c_{11}$ at different resolutions, where
$c_{11}$ and
$c_n$ are numerical solutions at uniform resolutions
$2^{11}\times 2^{11}$ and lower, respectively. Figure 1(d) shows first-order convergence with respect to the grid size
$\Delta x$. The effect of solubility on the error has not been discussed in earlier studies. We show in figure 1(d) that decreasing the solubility at the interface increases the error in the solution for a given resolution, but that the error remains small at high resolution.
3.3.2. Diffusion from a rising bubble
For the validation of the advection–diffusion scheme, we consider a bubble in a quiescent fluid rising due to buoyancy. We set up the test cases with the parameters considered in earlier studies (Darmana et al. Reference Darmana, Deen and Kuipers2006; Roghair Reference Roghair2012; Deising et al. Reference Deising, Marschall and Bothe2016; Jia, Xiao & Kang Reference Jia, Xiao and Kang2019), which encompass significant variations in bubble conditions, through variations of the Bond and Archimedes (or Morton) numbers. The rise velocity of a bubble in a quiescent liquid is indeed determined by the bubble–liquid physical parameters, summarized by the Archimedes number ${Ar}= g d_0^{3} \rho _l(\rho _l-\rho _g)/\mu ^{2}$ or Morton number
${Mo} = {g \mu _l^{4}}/{(\rho _l \gamma ^{3})}$ and Bond number
${Bo} = {\rho _l g d_0^{2}}{\gamma }$ (Moore Reference Moore1965; Maxworthy et al. Reference Maxworthy, Gnann, Kürten and Durst1996; Clift et al. Reference Clift, Grace and Weber2005; Cano-Lozano et al. Reference Cano-Lozano, Martinez-Bazan, Magnaudet and Tchoufag2016). The terminal rise velocity
$U$ can be computed theoretically (Moore Reference Moore1965; Clift et al. Reference Clift, Grace and Weber2005), and expressed as a non-dimensional bubble Reynolds number
${Re} = {\rho _l U d_0 }/{\mu _l}$. We consider four bubble configurations leading to Re from 5 to 100, for Bond numbers ranging from 1 to 40 and Archimedes numbers ranging from 100 to 8000. The mass transfer for these bubble conditions is then computed, considering the ratio of momentum to mass diffusivity given by the Schmidt number
${Sc} = {\nu _l}/{\mathscr {D}_l} =1$ and the gas solubility
$\alpha =1/30$, as in Deising et al. (Reference Deising, Marschall and Bothe2016). The fourth configuration is further tested for high Schmidt numbers of 10 and 100 in
$\textrm {D}_{2}$ and
$\textrm {D}_{3}$, respectively. Our results are shown in figure 2 and validated against previous work as shown in table 1. Note that to achieve the correct mass transfer, the correct rise velocity must be obtained, since the mass transfer will directly depend on the bubble velocity. Figure 2(a–d) shows examples of the shape of the bubble and the surrounding adaptive grid once the terminal velocity is reached. The average gas concentrations inside and outside the bubble are computed as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn28.png?pub-status=live)
where $V_g$ and
$V_l$ are the volumes of bubble and liquid, respectively. The mass transfer rate
$k_L$ is calculated as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn29.png?pub-status=live)
where $A_g$ is the instantaneous surface area of the bubble. The non-dimensionalized transfer rates are computed for both the axisymmetric and three-dimensional simulations with adaptive mesh refinement and a maximum refinement corresponding to a resolution on the bubble of
$d_0/\Delta x = 100$. Levich (Reference Levich1962) derived the mass transfer rates from a rising bubble with constant size and surface concentration, which reads
${Sh} = {2}/{\sqrt {{\rm \pi} }}\sqrt {{Pe}}$, where
${Pe}={Re} \, {Sc}$ is the bubble Péclet number. Upon reaching terminal velocity the transfer rates are steady after
$t\;U/d_0 \geqslant 2$ as shown in figure 2(e) and are accurately predicted by Levich (Reference Levich1962) as shown in figures 2(e) and 2(f). The values of the non-dimensional transfer rate, scaled by the diffusion velocity scale, or Sherwood number
${Sh} = {k d_0}/{\mathscr {D}_l}$ are also provided in table 1. Overall, the agreement between theoretical and numerical transfer rates is very good. The deviation from the spherical shape in case C, visible in figure 2(c), causes a difference with the predicted transfer rate which assumes a spherical bubble shape.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig2.png?pub-status=live)
Figure 2. Mass transfer of dilute gas from a bubble rising in a quiescent flow. (a–d) Concentration $c$ and bubble interface (three dimensional) at
$t\,U/d_0\approx 4$ for the four cases considered (increasing bubble Re number). (e) Evolution with time of transfer rates for axisymmetric and three-dimensional simulations and from Levich (Reference Levich1962) using the computed terminal velocity. (f) Steady-state transfer rate (for
$t\,U/d_0\geqslant 2$) as a function of Péclet number compared against Levich (Reference Levich1962). Very good agreement between the theoretical and numerical mass transfer is observed except for case C where the terminal shape is far from spherical. The scripts sufficient to reproduce these results are provided in Farsoiya, Popinet & Deike (Reference Farsoiya, Popinet and Deike2020b).
Table 1. Reynolds and Sherwood numbers (for $t\,U/d_0 \geqslant 2$) for axisymmetric and three-dimensional rising bubbles obtained from our numerical study, indicated as present work (PW Axi and PW 3D), compared with existing theoretical (Levich Reference Levich1962; Clift, Grace & Weber Reference Clift, Grace and Weber2005) and numerical (Roghair Reference Roghair2012; Deising et al. Reference Deising, Marschall and Bothe2016) results.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_tab1.png?pub-status=live)
4. Mass transfer in homogeneous and isotropic turbulence
We now present DNS of diffusion of dilute gas from a bubble inside a surrounding homogeneous and isotropic turbulent flow. We follow recent work on bubble deformation in turbulence (Perrard et al. Reference Perrard, Rivière, Mostert and Deike2021; Rivière et al. Reference Rivière, Mostert, Perrard and Deike2021) and first prepare the turbulent flow by solving the momentum equation with a forcing term, and then insert a bubble at the centre of the homogeneous and isotropic turbulent flow, once a statistically stationary state has been reached. A similar approach has been used to study the interaction of droplets with isotropic turbulence (Dodd & Ferrante Reference Dodd and Ferrante2016; Elghobashi Reference Elghobashi2019a). As the bubble is inserted, the volume-of-fluid advection (3.3) and mass advection–diffusion (3.11) are solved coupled with the momentum equations (3.1). We then calculate the mass transfer rate from the bubble and compare it with the theory presented in § 2.
4.1. Precursor simulation for isotropic turbulence
The turbulent flow is generated by adding a linear volumetric forcing term $f=A \boldsymbol {u}(\boldsymbol {x},t)$ in (3.1). This approach has been introduced by Rosales & Meneveau (Reference Rosales and Meneveau2005) and yields turbulence properties similar to those using a forcing in spectral space and leads to a well-characterized homogeneous and isotropic turbulent flow. Such an approach has been applied to study bubble rising in turbulence by Loisy, Naso & Spelt (Reference Loisy, Naso and Spelt2017), and more recently we used this approach within the Basilisk solver to study bubble deformation in turbulence (Perrard et al. Reference Perrard, Rivière, Mostert and Deike2021; Rivière et al. Reference Rivière, Mostert, Perrard and Deike2021). We consider a three-dimensional periodic box of size
$L$ for a precursor simulation to achieve isotropic turbulence. We use adaptive mesh refinement on the velocity field, and the maximum level of refinement can be used to compare the resolution with that of a fixed grid. The turbulent flow is generated for increasing resolutions with the maximum level of refinement going from 6 to 8, corresponding to an equivalent of
$64^{3}$ to
$256^{3}$ grid points. The resolution will be increased once we insert the bubble. The turbulence state is characterized by the kinetic energy density
$K$, turbulence dissipation rate
$\epsilon$ and Taylor microscale Reynolds number
${Re}_\lambda$, which are given by (Pope Reference Pope2001)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn30.png?pub-status=live)
and are computed over time to characterize the turbulent flow. The root mean square of the velocity is $u_{{rms}} = \sqrt {2K/3\rho _l}$, and the eddy turnover time at the scale of the bubble of diameter
$d_0$ is given by
$t_c=d_0^{2/3}\epsilon ^{-1/3}$ (Pope Reference Pope2001; Perrard et al. Reference Perrard, Rivière, Mostert and Deike2021).
Figure 3(a) shows the evolution of the turbulent kinetic energy with time for increasing Reynolds numbers. It shows that at $t/t_c\approx 25$ the flow has reached a statistically stationary state. We show that the state is grid-independent when using an adaptive mesh refinement of
$\textrm {L}7\equiv 2^{7}$ and
$\textrm {L}8\equiv 2^{8}$ for the three cases. The turbulence Reynolds number and the turbulence dissipation rate have similar time evolutions. The range of turbulent Taylor Reynolds numbers is
${Re}_{\lambda }\approx 38$ to 77 which is a typical value for current two-phase simulations of turbulent flow (Loisy et al. Reference Loisy, Naso and Spelt2017; Elghobashi Reference Elghobashi2019b). We characterize the turbulent stationary state using the second-order structure functions in the longitudinal direction
$D_{LL}(r)$ and the transverse direction
$D_{NN}(r)$, given by
$D_{LL}(r)=({1}/{3})\sum _i\langle (u_i(\boldsymbol {r},t)-u_i(\boldsymbol {r}+\textrm {d}\boldsymbol {\hat {r}}_i,t))^{2}\rangle$ and
$D_{NN}(r)=({1}/{6})\sum _{i\neq j}\langle (u_i(\boldsymbol {r},t)-u_i(\boldsymbol {r}+\textrm {d} \boldsymbol {\hat {r}}_j,t))^{2}\rangle$, where
$\boldsymbol {\hat {r}}_i$ is the unit vector along the
$i$th direction. Figure 3(b) shows that the scaled structure functions plateau at
$C=2$ (Pope Reference Pope2001) in the inertial range. The relation
$D_{LL} = 3/4 D_{NN}$ is verified and the inertial range is relatively limited due to the relatively coarse resolution and limited turbulence Reynolds number. The bubble is inserted once the turbulent stationary state is reached, and is of a size within the inertial range, the turbulence at this scale being reasonable, as described in Perrard et al. (Reference Perrard, Rivière, Mostert and Deike2021) and Rivière et al. (Reference Rivière, Mostert, Perrard and Deike2021).
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig3.png?pub-status=live)
Figure 3. Properties of the homogeneous and isotropic turbulent flow. (a) Turbulent kinetic energy as a function of time. After a short transient, a statistically stationary state is reached. The bubble is inserted once the statistically stationary state is reached. (b) Second-order structure functions $D_{LL}$ and
$D_{NN}$ in the longitudinal and transverse directions, respectively, compensated by the homogeneous and isotropic turbulence scaling
$(r \epsilon )^{-2/3}$ and
$D_{LL} = 3/4 D_{NN}$. Turbulence theory
$4/3 D_{LL}(r) (r \epsilon )^{2/3}$ is superimposed as a red dashed line. Parameter
$\eta$ is the Kolmogorov length scale. The bubble has a size comparable to the Taylor turbulence scale, within the inertial range.
4.2. Bubble insertion
The bubble is inserted at the centre of the box after reaching isotropic turbulence, i.e. for $t_0 > 25t_c$ (see figure 3a). The bubble is of diameter
$d_0$, viscosity
$\mu _b$ and density
$\rho _b$ surrounded by a liquid of viscosity
$\mu _l$ and density
$\rho _l$. The solubility of dilute gas is
$\alpha _j = 0.3$ which is transferred across the interface. The Weber number
${We} \equiv \rho _l u_{{rms}}^{2} d_0/\gamma = 1.3$ is below the critical number for bubble breakup (Perrard et al. Reference Perrard, Rivière, Mostert and Deike2021; Rivière et al. Reference Rivière, Mostert, Perrard and Deike2021), so that all results are for bubbles that can deform but do not break. Theoretical discussion and experimental data in Theofanous et al. (Reference Theofanous, Houze and Brumfield1976) provide the approximation that the transfer rate is a weak function of surface tension within three orders of magnitude
$(0.001 \lesssim We \lesssim 1)$ and depends mainly on the bulk turbulence properties. Diffusion rates for six different gases are calculated corresponding to Schmidt numbers ranging from 1 to 100. The bubble size with respect to the Taylor microscale length (
$\lambda = \sqrt {15 \nu u_{{rms}}^{2}/\epsilon }$) is in the range 1.72–2.82, and the box size is
$L = 7.5 d_0$. The turbulence properties and simulation parameters are given in table 2. Higher resolutions are used to properly resolve the diffusion of mass and the bubble deformation dynamics, with effective resolution using an adaptive mesh refinement of
$\textrm {L}10\equiv 2^{10}$,
$\textrm {L}11\equiv 2^{11}$ and
$\textrm {L}12 \equiv 2^{12}$. As is discussed in detail later, higher Schmidt numbers lead to thinner diffusive boundary layers which require smaller grid sizes. The resolution of smallest momentum and mass length scales are discussed in the Appendix.
Table 2. Simulation parameters (with adaptive mesh refinement) of the turbulence simulation of mass transfer. Three Reynolds numbers are used, with two effective resolutions, and a range of Schmidt numbers. The Weber number, density and viscosity ratio are kept constant.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_tab2.png?pub-status=live)
4.3. Mass transfer from the bubble to the surrounding turbulent flow
As the bubble moves with the flow, vortices of outer liquid come in contact with the interface, as seen in figure 4(a,b). Molecular diffusion of gas around the bubble interacts with the eddies of the flow and unsteady boundary layers are formed. The corresponding concentration fields for ${Sc}=1$ and
${Sc}=10$ are shown in figures 4(c,d) and 4(e,f). It can be observed that the concentration fields have followed the flow field due to advection. The gas with high diffusivity (
${Sc}=1$) has thicker boundary layers (figure 4(c,d) compared to the low-diffusivity gas (
${Sc}=10$; figure 4e,f) subjected to the exact same flow. Figure 5 shows a three-dimensional rendering of the same times of the simulations, displaying the three-dimensional concentration field around the bubble, as the bubble moves in the flow, with lower
${Sc}=1$ in figure 5(a,b) and higher
${Sc}=10$ in figure 5(c,d). Again, higher Schmidt numbers lead to thinner boundary layers around the bubble for the same flow field.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig4.png?pub-status=live)
Figure 4. Mass diffusion from the bubble in turbulence at two different times, $(t-t_0)/t_c=1.5$ (a,c,e) and 2 (b,d,f), showing a two-dimensional planar cut of the magnitude of the vorticity field (a,b) and concentration field for
${Re}_{\lambda }=77$,
${Sc}=1$ (
$Pe^{(t)}=205$) (c,d) and
${Re}_{\lambda }=77$,
${Sc}=10$ (
$Pe^{(t)}=2050$) (e,f). The wake in the vorticity field presents similarities to the structure of the concentration field. Higher Schmidt numbers lead to a thinner boundary layer around the bubble and a thinner wake structure.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig5.png?pub-status=live)
Figure 5. three-dimensional rendering of the concentration field for $(t-t_0)/t_c=1.5$ (a,c) and 2 (b,d). (a,b) Concentration field for
${Re}_{\lambda }=77$,
${Sc}=1$ (
$Pe^{(t)}=205$) and (c,d) concentration field for
${Re}_{\lambda }=77$,
${Sc}=10$ (
${Pe}^{(t)}=2050$). Thinner boundary layers are observed at higher Schmidt numbers.
We next compute the total transfer rates $k_L$ of gases corresponding to increasing Schmidt numbers, and show their evolution with time in figure 6 for increasing numerical resolution:
$d_0/\Delta x=136$ points per bubble diameter (figure 6a,b) and
$d_0/\Delta x=273$ points per bubble diameter (figure 6c,d); and increasing turbulence Reynolds number:
${Re}_\lambda =38$ (figure 6a,c) and
${Re}_{\lambda }=77$ (figure 6b,d). The time is shifted to the bubble insertion time
$t_0$ and normalized with the eddy turnover time
$t_c$. After a short transient, the transfer rates reach a steady state for
$(t-t_0)/t_c>0.25$. The transfer rates in steady state are compared with the predicted rates given by (2.3), which are shown as dashed lines. At lower resolution and Reynolds number (figure 6a;
$d_0/\Delta x=136$ and
${Re}_\lambda =38$), all cases exhibit good agreement between computed and predicted transfer rates, and results are unchanged when the resolution is increased (see figure 6c;
$d_0/\Delta x=273$ and
${Re}_\lambda =38$). The decrease in rates over time for lower-Schmidt-number cases in figure 6(a,b) is due to a considerable decrease in concentration of the transferred dilute gas. When increasing the Reynolds number, higher resolution is required for the highest Schmidt numbers, as visible in figure 6(b,c). To summarize, the DNS results are in close agreement with the predicted rates for
$50<\widetilde {{Pe}}<10^{4}$, with
$\widetilde {{Pe}}={u_{{rms}}d_0}/{\mathscr {D}_l}$ the turbulent Péclet number.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig6.png?pub-status=live)
Figure 6. Non-dimensional mass transfer rates ${Sh}$ as a function of time, as the bubble is exposed to the turbulent flow. (a,b) Lower resolution
$d_0/\Delta x=136$ (level 10), for
${Re}_{\lambda } = 38$ (a) and
${Re}_{\lambda } = 77$ (b). (c,d) Higher resolution (solid line for level 11
$d_0/\Delta x=273$, dotted line for level 12
$d_0/\Delta x=546$), for
${Re}_{\lambda } = 38$ (c) and
${Re}_{\lambda } = 77$ (d). Dashed lines represent the theoretical prediction (2.3) for the different Schmidt numbers. At steady state, very good agreement between simulations and theory is achieved for
$\widetilde {{Pe}}={u_{{rms}}d_0}/{\mathscr {D}_l}={Sc} (u_{{rms}} d_0/\mathscr {D}_l)\leqslant 10^{4}$, which corresponds to a diffusive boundary layer
$\delta _{k}$ resolved with more than approximately four grid points.
For higher turbulent Péclet numbers ($1.02\times 10^{4},\; 2.05\times 10^{4}$), we observe an overprediction of transfer rates. These results for the cases of
$\widetilde {{Pe}}>10^{4}$ (yellow and black curves in figure 6b,d) can be understood by considering the resolution of boundary layers. The thickness of the hydrodynamic boundary layer
$\delta _\nu$ around a spherical bubble is of
${\textit{O}}({Re}^{-1/2})$ (Moore Reference Moore1963). The concentration boundary layer thickness
$\delta _k \propto \delta _\nu \,{Sc}^{-1/2}$ (Levich Reference Levich1962; Bothe & Fleckenstein Reference Bothe and Fleckenstein2013) scales with the bubble diameter
$d_0$, and is given by
${\delta _k}/{d_0} \approx {Pe}^{-1/2}$. The numerical framework used in the present work uses adaptive mesh refinement with respect to the norm of the second-derivative of velocity and concentration (van Hooft et al. Reference van Hooft, Popinet, van Heerwaarden, van der Linden, de Roode and van de Wiel2018), with an error threshold of
$0.2$ times the average of fields over the entire domain. The average of concentration fields is very small as most of the domain has trace amounts of gas concentration. This leads to a resolution of the boundary layers ranging from 40 grid points (
$\widetilde {{Pe}}\approx 50, \delta _\nu /d_0\approx 0.13,\delta _k/d_0\approx 0.13$) to 3 grid points (
$\widetilde {{Pe}}\approx 2 \times 10^{4},\delta _\nu /d_0\approx 0.07,\delta _k/d_0\approx 0.007$). When the resolution is increased, we are able to resolve thinner boundary layers, hence higher Péclet numbers (
$1.02\times 10^{4}, 2.05\times 10^{4}$), as shown in figures 6(b) and 6(d) where the rates for
$\widetilde {{Pe}}>10^{4}$ are converging to the predicted value as the resolution increases (L11 in solid lines and L12 in dotted lines).
Finally, figure 7 shows the Sherwood numbers obtained from the bubble mass transfer in homogeneous and isotropic turbulence simulations for a wide range of turbulent Péclet numbers ($50 < \widetilde {{Pe}} \leqslant 2\times 10^{4}$) for three different resolutions (simulation conditions are summarized in table 2). Very good agreement is observed between (2.3), up to
$\widetilde {{Pe}} \sim 10^{4}$. As such, figure 7 validates (2.3), which predicts the transfer rates from the bubble in homogeneous and isotropic turbulent flow. We note that for
${Pe}\approx 2 \times 10^{4}$, an effective resolution L12 brings the simulated transfer rate closer to our theoretical prediction, as a result of increased resolution in the turbulent–diffusive boundary layer next to the bubble interface. We recall that figure 7 shows the scaling for the transfer rate
$k_L$, as the non-dimensional Sherwood number, which reads
$Sh \propto {Re}^{0.5}{Sc}^{0.5} = \sqrt {\widetilde {{Pe}}}$, and considering the definition of the Sherwood number, this corresponds to
$k_L \propto {Sc}^{-0.5}\sqrt {{u_{{rms}}\nu _l}/{d_0}}$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig7.png?pub-status=live)
Figure 7. The transfer rates (for $(t-t_0)/t_c\geqslant 0.2$) with respect to Péclet number. Symbols are for different
${Re}_{\lambda }$ and resolution. For
$\widetilde {{Pe}}\leqslant 10^{4}$, the numerical results are grid converged (first order) between L10 and L11, and agree very well with the theoretical prediction (2.3),
${Sh}=2(3)^{1/4}\sqrt {\widetilde {{Pe}}/{\rm \pi} }$. For higher Péclet numbers (
$1.02\times 10^{4}, 2.05\times 10^{4}$), higher resolutions may converge to the prediction as visible when using L12 simulations. The transfer rates from Dodd et al. (Reference Dodd, Mohaddes, Ferrante and Ihme2021) correspond to the transfer rates from simulations of an evaporating drop at early times and agree with our data and model.
We note that Dodd et al. (Reference Dodd, Mohaddes, Ferrante and Ihme2021) present simulations of a droplet evaporating in turbulence and their evaporation rates at early times before significant volume change has occurred are in agreement with our model, as shown in figure 7.
The computed Sherwood numbers show convergence to the predicted value as the grid resolution increases. For the thinner boundary layers, corresponding to ${Re}_\lambda =77$ and high Schmidt number at
$d_0/\Delta x=273$ resolution, the total number of cells in the domain reached 27.5 million and used
$2.16\times 10^{4}$ CPU hours for
$0.5$ eddy turnover time when the transfer rates approach a steady state. A corresponding constant-resolution DNS would have required
$(2^{11})^{3} \approx 8.6$ billion cells, which illustrates the gain in efficiency brought by adaptive mesh refinement. It highlights the potential of adaptive mesh refinement methods for complex two-phase flows at high Péclet number, complementary to subgrid-scale modelling approaches.
We note that in § 2, we proposed the large-scale velocity fluctuations as the controlling velocity scale. As already mentioned, another choice could have been the velocity fluctuations at the scale of the bubble. Such a choice would not have changed the results obtained in figure 7, as these velocities are similar at the Reynolds number at the Taylor scale we consider.
Finally let us comment that at high Reynolds numbers, in the case of a flat interface, the scaling of the gas transfer is shifted from $Re^{1/2}$ to
$Re^{3/4}$ due to the role of small eddies in enhancing the gas exchange (Theofanous et al. Reference Theofanous, Houze and Brumfield1976; Herlina & Wissink Reference Herlina and Wissink2016). Whether such a regime could be observed in the case of bubble mass exchange at high Péclet or Reynolds number remains to be investigated.
5. Concluding remarks
We propose a simple theoretical formula for mass transfer rate of dilute gas from bubbles for a dilute component within a turbulent flow, extending the classic formulation of Higbie (Reference Higbie1935) to a homogeneous and isotropic turbulent flow, by considering that the renewal time scale is controlled by the turbulence fluctuations at large scale and the bubble size:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_eqn31.png?pub-status=live)
We developed and presented a versatile numerical framework to perform DNS of multiphase flows, and tested the theoretical prediction against numerical simulations. The agreement between theory and numerical results is very good, as long as the diffusive boundary layer thickness is correctly resolved. The numerical framework can now be used for more complex physical configurations such as bubbles under a breaking wave and bubble swarms.
The proposed theoretical scaling for dilute mass transfer is remarkable as it could be leveraged for multiscale systems, such as bubbles entrained by breaking waves in the upper ocean, or bubbles evolving in turbulent swarms, and avoid solving for the very high Schmidt number necessary in practical problems (for example, the Schmidt number of $\textrm {CO}_2$ at
$20^{\circ }$C is
$\sim$660). Finally, our formulation could also be used to improve recent bubble-mediated mass transfer models which have, up to now, used the bubble transfer rate in a quiescent flow (Liang et al. Reference Liang, McWilliams, Sullivan and Baschek2011; Deike & Melville Reference Deike and Melville2018).
As part of this work, a video has been published in the gallery of fluid motion (Farsoiya, Popinet & Deike Reference Farsoiya, Popinet and Deike2020c).
Acknowledgements
This work was supported by the NSF CAREER award 1844932, the Catalysis Initiative at Princeton and the Cooperative Institute for Modeling the Earth System (CIMES) between Princeton and NOAA-GFDL. We would like to acknowledge high-performance computing support from Tiger, Princeton research computing and Cheyenne (doi:10.5065/D6RX99HX) provided by NCAR's Computational and Information Systems Laboratory, sponsored by the National Science Foundation.
Declaration of interests
The authors report no conflict of interest.
Appendix
The accepted resolution criteria for DNS in the literature (Overholt & Pope Reference Overholt and Pope1996; Pope Reference Pope2001; Schumacher, Sreenivasan & Yeung Reference Schumacher, Sreenivasan and Yeung2005; Dodd et al. Reference Dodd, Mohaddes, Ferrante and Ihme2021) are typically $k_{{max}}\eta > 1.5$ and
$k_{{max}}\eta _B > 1.5$, where
$k_{{max}}$,
$\eta$ and
$\eta _B$ are the maximum resolved wavenumber
$k_{{max}}={\rm \pi} N/L$, the Kolmogorov scale and the Batchelor scale, respectively. The Kolmogorov length scale
$\eta =(\nu _l^{3}/\epsilon )^{1/4}$ defines the length scale at which viscous dissipation becomes dominant while the Batchelor scale is defined as
$\eta _B=\eta /\sqrt {Sc}$. Figure 8 shows these characteristic length scales as functions of the turbulent Péclet number for all our simulations. In all of the cases the Kolmogorov length scale is well resolved, with
$k_{max}\eta > 8$, in agreement with the fact that convergence is already achieved in lower resolutions as shown in figure 3(a). For the highest Péclet number, the Batchelor length scale is resolved up to
$k_{{max}}\eta _B \approx 3.2$ (refinement
$d_0/\Delta x = 546$). The boundary layer thickness
$\delta _{\nu } > 2.5 \eta$ and
$\delta _{k} > 2.5 \eta _B$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20210611182439062-0482:S002211202100447X:S002211202100447X_fig8.png?pub-status=live)
Figure 8. Numerical resolution of Kolmogorov ($\eta$) and Batchelor (
$\eta _B$) length scales as functions of Péclet number, normalized by the maximum resolved wavenumber
$k_{{max}}={\rm \pi} N/L$.