gas absorption into a spherical liquid droplet: numerical ... · gas absorption into a spherical...

38
Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock *,a , P. Colinet a , B. Haut a a Transfers, Interfaces and Processes (TIPs),Universit´ e Libre de Bruxelles, 50 av. F.D. Roosevelt CP 165/67, 1050 Brussels, Belgium Abstract This work deals with the modeling and the direct numerical simulation of the absorption of a gas component into a spherical liquid droplet in free fall, where the absorbed component takes part to a chemical reaction. This study is realized by computing the flow fields and the concentration fields simul- taneously in the gas and liquid phases. The time evolution of the droplet global mass absorption rate is studied for various regimes characterized by the Reynolds and the Hatta numbers. The monitoring of the time evolution of the concentration fields in the droplet enables the understanding of the interactions between the diffusive and convective mass transports and the chemical reaction. The phenomena controlling the mass transfer rate and how they evolve during the mass absorption process can then be identified. In a first stage, the case of the physical absorption is studied. Moreover, analytical expressions are developed to correlate the time evolution of the Sherwood number with the Reynolds number and the absorption time. In a second stage, the influence of the coupling with a chemical reaction is stud- ied. It is observed that several rate-limiting mechanisms successively control the global absorption rate, enlightening the complex interactions between convective and diffusive mass transport and the chemical reaction. Key words: droplet, mass transfer, chemical reaction, computational fluid dynamics * Corresponding author Email address: [email protected] (C. Wylock) Preprint submitted to Chemical Engineering Journal July 9, 2012 This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Upload: others

Post on 08-Aug-2020

18 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Gas absorption into a spherical liquid droplet:

numerical and theoretical study

C. Wylock∗,a, P. Colineta, B. Hauta

aTransfers, Interfaces and Processes (TIPs),Universite Libre de Bruxelles, 50 av. F.D.Roosevelt CP 165/67, 1050 Brussels, Belgium

Abstract

This work deals with the modeling and the direct numerical simulation ofthe absorption of a gas component into a spherical liquid droplet in free fall,where the absorbed component takes part to a chemical reaction. This studyis realized by computing the flow fields and the concentration fields simul-taneously in the gas and liquid phases. The time evolution of the dropletglobal mass absorption rate is studied for various regimes characterized bythe Reynolds and the Hatta numbers. The monitoring of the time evolutionof the concentration fields in the droplet enables the understanding of theinteractions between the diffusive and convective mass transports and thechemical reaction. The phenomena controlling the mass transfer rate andhow they evolve during the mass absorption process can then be identified.In a first stage, the case of the physical absorption is studied. Moreover,analytical expressions are developed to correlate the time evolution of theSherwood number with the Reynolds number and the absorption time. In asecond stage, the influence of the coupling with a chemical reaction is stud-ied. It is observed that several rate-limiting mechanisms successively controlthe global absorption rate, enlightening the complex interactions betweenconvective and diffusive mass transport and the chemical reaction.

Key words: droplet, mass transfer, chemical reaction, computational fluiddynamics

∗Corresponding authorEmail address: [email protected] (C. Wylock)

Preprint submitted to Chemical Engineering Journal July 9, 2012

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 2: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

1. Introduction

The absorption of gas materials from flue gas by reactive liquid dropletsis an important mass transfer process for several industrial applications, suchas air pollution control. The presence of a chemical component in the liq-uid phase which reacts with the absorbed component enables a considerableenhancement of the mass transfer rate. One of the most studied case is thescrubbing of sulfur dioxide (SO2) by a limestone slurry in a counter currentspray scrubber, used in order to reduce the emissions [1, 2, 3, 4]. A deepunderstanding of the phenomena taking place during the gas-droplet masstransfer coupled with chemical reaction is of fundamental interest for thedesign and the optimization of such industrial applications.

A detailed modeling, based on a sound mathematical statement, is re-quired to predict correctly the gas-droplet mass transfer rate and its evolutionover time. This is a complex problem, as shown by the large number of math-ematical models and experimental investigations that have been devoted toit. Indeed, the mass transfer rate is influenced by several phenomena andby their intricate couplings: the diffusive transport, the convective transportimposed by the flow fields in both phases, and the chemical reaction in theliquid phase.

This complexity is probably one of the main reason explaining why manyempirical and idealized models can be found in the literature. For instance,the expression for the Sherwood number derived by Frossling from a studyof the evaporation of small water droplet (see [5]) was often used for gasabsorption into a droplet. More recently, Brogren and Karlsson [1] haveproposed a model for the flue gas desulfurization taking into account all thechemical reactions taking place in the liquid but using the Higbie penetrationtheory [6] for the mass transport.

Nowadays, it is well known that for falling droplets having a Reynoldsnumber (defined in section 2.2) higher than 10, the shear stress exerted by thegas phase induces a vortex in the droplet [7], which can have a significantinfluence on the mass transfer rate [8, 9, 10]. To compute the flow fieldin both phases and their influence on the mass transfer is one of the mainmodeling difficulty, which has been studied for decades.

It is now commonly admitted that, for Reynolds numbers smaller than1, the gas flow is creeping and the flow field in the droplet can be describedby the analytical stream function of Hadamard and Rybczinski [7, 11]. Forhigher Reynolds numbers, between 1 and 80, the stream function determined

2

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 3: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

by the approximate expressions of Hameliec and Johnson [12] can be usedinstead. Leclair et al. [7] presented a general theoretical framework to com-pute the flow fields inside and outside water droplets. One of the first worksto compute the mass mass transfer rate with internal circulation was realizedby Kronig and Brink [13] in the framework of extraction. The first incorpo-ration of the flow fields computed using the model of Leclair et al. [7] into amodel of gas absorption was realized by Baboolal et al. [14], as mentioned byElperin and Fominykh [15]. The work of Uribe-Ramırez and Korchinsky [16]presented a theory yielding an analytical solution for the mass transfer rateat intermediate Reynolds number. In the most recent works, an increasingnumbers of authors used Computational Fluid Dynamics to compute the flowfields in both phases, such as in works of Waheed et al. [17], or Paschedaget al. [18].

This work proceeds along these lines, dealing with the modeling and thedirect numerical simulation of the absorption of a component A from a gasphase into a spherical liquid droplet in free fall in this gas phase. A com-ponent B is dissolved in the liquid droplet and reacts with A following thisirreversible reaction:

A+B → C (1)

where C is the product of the reaction.This study is realized by computing the flow fields and the concentration

fields simultaneously in both the gas and liquid phases. This enables to rig-orously account for the coupling between convection, diffusion and reaction.This work is focused on water droplet falling in air but the model equationsare presented in a general dimensionless form, making the model suitable forstudying mass transfer in other systems.

This work has two main goals. The first one is the detailed study ofthe time evolution of the droplet global mass absorption rate for variousflow regimes and chemical regimes, which are characterized by the Reynoldsnumber (defined in section 2.2) and the Hatta number (defined in section 2.3),respectively. The monitoring of the time evolution of the concentration fieldsin the droplet yields a better understanding of the influence of the interactionsbetween the diffusive/convective mass transport and chemical reaction onthe time evolution of the mass transfer rate. This procedure enables theidentification, for each regime, of the phenomena controlling the mass transferrate and how they possibly evolve during the mass absorption. The second

3

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 4: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

goal is to propose analytical expressions to estimate the time evolution ofthe droplet mass transfer rate in the case of physical absorption (no chemicalreaction). This is realized by comparing the simulation results with simplifiedmodelling approaches based on the analysis of the mass transfer rate-limitingphenomena.

2. Mathematical statement

2.1. Assumptions

This work is focused on the case of a spherical water droplet of constantvolume in free fall in air. It is considered that the droplet falls at its terminalvelocity and that its motion is rectilinear with a non-deformable interface.The two phases are considered incompressible and the flows in both phasesare assumed to be stationary, laminar and axisymmetric. It is consideredthat the interface is completely free from surface-active contaminants andthe continuity of the velocity and of the tangential shear stress is consideredat the interface (in particular, possible Marangoni effects due to surface ten-sion variations are neglected). In addition, it is considered that the massabsorption and the resulting chemical reactions in the liquid phase do notsignificantly affect the density and the viscosity of the two phases. Withthese assumptions, the continuity and momentum transport equations canbe solved separately from the mass transport equations.

The assumption of stationary flow is justified by the fact that the charac-teristic time of the transient droplet motion is much smaller than the charac-teristic time of diffusive mass transport, since the Schmidt number (definedin section 2.3) is much larger than one.

According to Elperin and Fominykh [15], a spherical shape of a dropletfalling in a gas phase is observed until a Reynolds number of 300, which cor-responds approximately to a droplet diameter of 1.1 mm for a water dropletfalling in air. This limit is also mentioned for limestone droplets by Akbar etal. [19]: the droplets remain spherical until 1 mm and their interface is non-oscillating until a Reynolds number between 300 and 400. Beyond this value,the droplet starts to be deformed and their interface oscillates. This valueis also presented as an upper limit for the spherical shape in the numericalstudy performed by Amokrane and Caussade [20]. Clift et al. [21] reportedthat water droplets smaller than 1 mm falling in air can be considered as rigidspheres. Other authors, such as Bandyopadhyay and Biswas [2], claim thatwater droplets falling is gas are spherical and their interface is non-oscillating

4

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 5: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

until a diameter of 0.5 mm but that this has no significant influence on themass transfer rate at this stage since the effects of the oscillation of the in-terface and its deformation on the mass transfer mutually compensate eachother. Therefore, it appears that the assumptions of a spherical shape and anon-deformable interface are valid, for water droplets in air, until a diameterof 1 mm, corresponding to a Reynolds number of 250.

Concerning the laminar flow assumption, Amokrane and Caussade [20]evaluated that for droplets of water falling in air, the flow becomes turbulentin both gas and liquid phases for a diameter close to 1.6 mm. It correspondsto a Reynolds number close to 690 in the assumption of a rigid sphere. More-over, according to Paschedag et al. [18], the flow of the gas phase around adroplet is a creeping flow until a Reynolds number of 0.1. For higher values,wakes are formed behind the droplet but the flow remains laminar until aReynolds number of 500.

The validity limit for the axisymmetric assumption appears to be a di-ameter of 1 mm for water droplets falling in air at their terminal velocity,according to Walcek et al. [22] (and cited by Altwicker and Lindhjem [23]),which claim that the internal circulation pattern is steady and that the ab-sorption characteristics follow the model of Kroning and Brink [13].

Therefore, taking into account this literature review, the assumptionsof the developed model are valid for Re ≤ 250. In addition, in order toprevent too large numerical errors of the simulation, the minimum consideredReynolds number is 0.01. It corresponds, for water droplet falling in air, toa droplet diameter of 17 µm.

The model is based on the balance equations written in dimensionlessform in an inertial reference frame attached to the center of mass of thedroplet. A two-dimensional axisymmetric computational domain is used. Itis presented schematically in Fig. 1. This computational domain is dividedin two distinct subdomains: one for the gas phase and one for the liquidphase. The droplet diameter dd is used as the reference length. Therefore,the dimensionless radius of the droplet is 0.5. A domain diameter 10 timeslarger than the droplet is used in this work. In the reference frame, the gasis flowing upward. r and z are the dimensionless radial and axial componentof the cylindrical coordinate system, respectively.

Fig. 1

5

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 6: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

2.2. Flow fields

The flow fields are computed by solving the stationary incompressibleNavier-Stokes and continuity equations in both gas and liquid phases.

The droplet diameter dd is used as the reference length and the terminalfalling velocity U is used as the reference velocity. The reference pressure isdefined as ρgU

2, where ρg is the gas density. The Reynolds number is definedas Re = ρgddU/ηg, where ηg is the gas dynamic viscosity.

In both phases, let u = (u, v) be the dimensionless velocity and p thedimensionless pressure. The subscript l and g are used to refer to the liquidand the gas phases, respectively.

The Navier-Stokes and continuity equations in the gas phase read:

ug∂ug∂r

+ vg∂ug∂z

= −∂pg∂r

+1

Re

(∂2ug∂r2

+1

r

∂ug∂r

− ugr2

+∂2ug∂z2

)(2)

ug∂vg∂r

+ vg∂vg∂z

= −∂pg∂z

+1

Re

(∂2vg∂r2

+1

r

∂vg∂r

+∂2vg∂z2

)(3)

∂ug∂r

+ugr

+∂vg∂z

= 0 (4)

while the Navier-Stokes and continuity equations in the liquid phase are:

ρ∗(ul∂ul∂r

+ vl∂ul∂z

)= −∂pl

∂r+η∗

Re

(∂2ul∂r2

+1

r

∂ul∂r

− ulr2

+∂2ul∂z2

)(5)

ρ∗(ul∂vl∂r

+ vl∂vl∂z

)= −∂pl

∂z+η∗

Re

(∂2vl∂r2

+1

r

∂vl∂r

+∂2vl∂z2

)(6)

∂ul∂r

+ulr+∂vl∂z

= 0 (7)

where ρ∗ = ρl/ρg and η∗ = ηl/ηg are the density and the dynamic viscosityratios, respectively.

At the droplet interface, which is considered completely free from surface-active contaminants, it is assumed, on the one hand, that the component ofthe shear stress tangent to the interface is continuous across the interface. Onthe other hand, the balance of the component of the shear stress normal to theinterface is unnecessary since the interface is considered as non-deformable.

6

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 7: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Therefore, the following equation is written:(2∂ug∂r

nrtr +

(∂ug∂z

+∂vg∂r

)(nrtz + nztr) + 2

∂vg∂z

nztz

)= η∗

(2∂ul∂r

nrtr +

(∂ul∂z

+∂vl∂r

)(nrtz + nztr) + 2

∂vl∂z

nztz

)(8)

where n = (nr, nz) and t = (tr, tz) are unit vectors, normal and tangent tothe interface, respectively.

In addition, it is assumed, at the droplet interface, that the componentof the velocity tangent to the interface is continuous across the interface andthat the component of the velocity normal to the interface equals zero inboth phases:

ugnr + vgnz = ulnr + vlnz = 0 (9)

ugtr + vgtz = ultr + vltz (10)

At the top boundary of the domain, corresponding to the gas flow outlet,a free flow outlet boundary condition is imposed. The resulting equationswrite:

−pg = 0 (11)

1

Re

(2∂ug∂r

nr +

(∂ug∂z

+∂vg∂r

)nz

)= 0 (12)

1

Re

((∂ug∂z

+∂vg∂r

)nr + 2

∂vg∂z

nz

)= 0 (13)

At the bottom boundary of the domain, corresponding to the gas flowinlet, a uniform upward velocity is imposed:

ug = 0 (14)

vg = 1 (15)

2.3. Concentration fields

Since the mass transfer resistance can evolve with time due to the compe-tition between convection, diffusion and reaction phenomena, the concentra-tion fields are computed by solving the transient mass transport equationsin both phases, imposing the gas-liquid equilibrium and the continuity of themass transfer rate of the component A at the interface.

7

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 8: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

The reference time τ is defined as τ = dd/U . Let [A]eq be the con-centration of A in the droplet when the equilibrium with the gas phase isreached. Considering dilute liquid solution and perfect gas, [A]eq is defined as[A]eq = h[Ag]∞. [Ag]∞ is the concentration of A in the gas phase far from thedroplet and h = HRT , where H is the Henry, R is the perfect gas constantand T is the absolute temperature. Let [B]0 be the initial concentration ofB in the droplet.

[A]eq is used as the reference concentration of A, in both the gas andliquid phases, while [B]0 is used as the reference concentration for B.

Let t be the dimensionless time. Let CAg be the dimensionless concentra-tion of A in the gas phase. CA and CB are the dimensionless concentrationsof A and B in the liquid phase, respectively.

As there is no chemical reaction in the gas phase, the following masstransport equation, involving only diffusive and convective transports, is con-sidered for A in the gas phase:

∂CAg

∂t+ ug

∂CAg

∂r+ vg

∂CAg

∂z=βgPe

(∂2CAg

∂r2+

1

r

∂CAg

∂r+∂2CAg

∂z2

)(16)

where βg = DAg/DAland Pe = Udd/DAl

. Pe is the Peclet number for themass transport of A in the liquid phase. Note that this Peclet number can becalculated from the Reynolds number and the Schmidt number Sc = ηl/ρlDAl

of A in the liquid by Pe = Re Sc ρ∗/η∗

Two mass transport equations, involving diffusion, convection and chem-ical reaction, are considered for A and B in the liquid phase. They write:

∂CA

∂t+ ul

∂CA

∂r+ vl

∂CA

∂z=

1

Pe

(∂2CA

∂r2+

1

r

∂CA

∂r+∂2CA

∂z2

)− ψ (17)

∂CB

∂t+ ul

∂CB

∂r+ vl

∂CB

∂z=

βbPe

(∂2CB

∂r2+

1

r

∂CB

∂r+∂2CB

∂z2

)− χψ (18)

where χ = h[Ag]∞/[B]0 is the concentration ratio between the maximumconcentration of A that can be reached in the droplet and the initial con-centration of the B in the droplet and βb = DB/DAl

. ψ is the dimensionlesschemical reaction rate.

It is assumed that the reaction (see Eq. (1)) is of the first order withrespect to the concentrations of A and B, and that its kinetic constant is k.Therefore, ψ is calculated by:

ψ = Ha2CACB (19)

8

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 9: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

where Ha =√k[B]0τ is the Hatta number.

The Hatta number compares the reference time τ to the characteristictime of the reaction [24, 25]. A Hatta number lower than unity means thatthe reaction is slow compared to the convective time scale τ . When theHatta number is of order one, the chemical reaction rate and τ have thesame order of magnitude. Finally, when the Hatta number is larger thanone, the reaction occurs much faster than the time it takes for the droplet totravel a distance equal to its diameter.

At the droplet interface, the gas-liquid equilibrium for A and the continu-ity across the interface of the flux of A normal to the interface are assumed.Therefore, the following boundary conditions can be written:

CA = hCAg (20)

∂CA

∂rnr +

∂CA

∂znz = βg

(∂CAg

∂rnr +

∂CAg

∂znz

)(21)

It is assumed that B does not cross the droplet interface. Therefore, thefollowing no flux boundary condition can be written:

∂CB

∂rnr +

∂CB

∂znz = 0 (22)

At the top boundary, it is assumed that the transport of A is only con-vective. Therefore, the top boundary condition reads:

∂CAg

∂rnr +

∂CAg

∂znz = 0 (23)

At the bottom boundary, a uniform concentration of A entering the gassubdomain is assumed:

CAg = h−1 (24)

At the initial time t = 0, it is supposed that the concentrations in bothphases are homogeneous and at equilibrium in the liquid phase. The initialconditions for the concentrations write therefore:

CAg(t = 0) = h−1 (25)

CA(t = 0) = α (26)

CB(t = 0) = 1 (27)

α = [A]0/h[Ag]∞ is the ratio between the initial concentration of A in thedroplet, denoted [A]0, and the maximum concentration of A that can bereached in the droplet.

9

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 10: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

3. Simulations

The equations of the model are solved numerically using the COMSOLMultiphysics 3.4 software. This commercial code uses the finite elementmethod to solve the boundary-value problem. The “Incompressible Navier-Stokes mode” coupled with the “Convection and Diffusion mode”, of the“Chemical Engineering Module”, are used in a 2D axisymmetric geometry.

Since finite elements are used, a meshing of the computational domainhas to be defined. The equations are discretized using quadratic Lagrangianelements.

3.1. Meshing

The two subdomains are meshed independently and so-called boundary-layer meshes are used in both gas and liquid phases near the interface. Thismesh type is characterized by the coexistence of two mesh shapes in thedomain. On the domain zones which are characterized by the presence of veryimportant gradients, such as in boundary layers, rectangular shape meshesare used. Triangular shape meshes are used outside these zones, in the bulkof the phases.

Fig. 2

In the liquid subdomain (on the left in Fig. 2), a boundary-layer meshis defined near the interface boundary and near the symmetry axis. Thisboundary-layer zone contains 1080 meshes. It is made of a layer of 15 rect-angles with a stretching factor equal to 1.15 in the direction normal to theinterface and the symmetry axis, towards the droplet interior. The firstlayer is made of 44 rectangles, regularly spaced along the interface and 28rectangles regularly spaced along the symmetry axis. The thickness of thefirst layer equals 7.5 10−4. The remaining subdomain is meshed with 880triangular meshes.

In the gas subdomain (on the right in Fig. 2), a boundary-layer mesh isdefined near the gas-liquid interface. This boundary-layer zone contains 280meshes. It is made of a layer of 20 rectangles with a stretching factor equal to1.2 in the direction normal to the interface, towards the bulk of the gas phase.The first layer is made of 14 rectangles, regularly spaced along the interface.The thickness of the first layer equal 1.75 10−3. The remaining subdomain ismeshed with 3098 triangular meshes, with 40 mesh points regularly spaced

10

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 11: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

on each domain boundary and with 23 mesh points set on each part (up anddown) of the symmetry axis.

It is observed that this choice of meshing configuration leads to the bestcompromise between the convergence quality and the number of elements. Ifonly a triangular meshing was used, the convergence quality would not begood enough even with very small meshes in the interface zone. A rectangularmeshing only would lead to good convergence results but the number ofelements required using this mesh shape would be too high and would leadto meshing errors at the center of the droplet. Moreover, it was verifiedthat further refinement of this meshing configuration does not influence thesimulation results.

3.2. Numerical procedure

According to the assumptions (see section 2.1), the mass transfer has noinfluence on the liquid flow; the momentum transport and the mass transportcan be solved in a segregated way. For each step, the direct solver UMFPACKis used.

The steady Navier-Stokes and continuity equations (Eqs. (2)-(7)) are firstsolved without mass transfer using the interface and boundary conditions(Eqs. (8)-(10) and (11)-(15), respectively). The initial guess is no flow andthe pressure is set to zero in both phases:

ug = 0 , vg = 0 , pg = 0 (28)

ul = 0 , vl = 0 , pl = 0 (29)

The resulting computed flows are then stored.In the case of a simulation of the physical absorption, the diffusion-

convection problem is solved without chemical reaction. Only Eqs. (16) and(17) are solved with ψ = 0, using the flow fields calculated at the previousstep, the interface and boundary conditions (Eqs. (20)-(21) and (23)-(24),respectively) and the initial conditions (Eqs. (25)-(26)). Note that the con-centrations of B does not vary in this case and remains equal to its initialvalue (Eq. (27)).

In the case of a simulation of an absorption coupled with a chemicalreaction, the convection-diffusion-reaction problem (Eqs. (16)-(18) with ψcalculated by Eq. (19)) is solved, using the flow fields from the first step,the interface and boundary conditions (Eqs. (20)-(22) and (23)-(24), respec-tively) and the initial conditions (Eqs. (25)-(27)).

11

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 12: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

3.3. Post-processing and engineering aspects

The dimensionless flow and concentration fields obtained by the numericalresolution enables the calculation of some macroscopic data characterizingthe flow and the dynamics of the mass transfer.

Since these data are calculated by the shell integration of the computedflow and concentration fields, a polar coordinate system (R, θ) is introducedfor the sake of clarity. R is the radial coordinate and θ is the polar coordinate(from the top pole to the bottom pole of the droplet). The presented post-processed data are therefore calculated using this coordinate system.

The drag coefficient CD is calculated from the total drag force FD ex-erted by the gas phase on the droplet using the classical definition CD =8FD/πd

2dρgU

2. This drag force is calculated by the shell integration of thez-component of the stress on the droplet interface. It can be demonstratedthat CD is given by:

CD = 4

∫ π

0

τz(θ) sin θdθ (30)

where τz(θ) is the z-component of the dimensionless stress exerted by the gason the droplet interface, which is calculated by:

τz(θ) =

(−pg|θ +

2

Re

∂vg∂z

∣∣∣∣θ

)nz(θ) +

1

Re

(∂vg∂r

∣∣∣∣θ

+∂ug∂z

∣∣∣∣θ

)nr(θ) (31)

where (nr(θ), nz(θ)) is the unit vector normal to the interface (directed to-wards the gas) at the considered point.

In order to monitor the mass absorption process over time, a saturationlevel of the droplet ϕ(t) is followed. It is defined as the ratio between theabsorbed amount of A at the time t over the maximum amount of A thatthe droplet can absorb. ϕ = 0 at t = 0 and ϕ = 1 when the droplet cannotabsorb more of the component A.

ϕ(t) can be evaluated using spatially-averaged dimensionless concentra-tions in the droplet at the time t, written < Cξ >|t, where ξ = A, B.< Cξ >|t is given by:

< Cξ >|t = 12

∫ Rd

0

∫ π

0

Cξ|R, θ, tR2 sin θ dR dθ (32)

where Rd = 0.5 is the dimensionless radius of the droplet.

12

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 13: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

In the case of a physical absorption, ϕ(t) is calculated by:

ϕ(t) =< CA >|t − α

1− α(33)

In the case of an absorption coupled with a chemical reaction, the ab-sorbed A can be accumulated in the droplet or can be consumed by thereaction with B. The presence of B increases the droplet absorption capac-ity. The droplet can become saturated of A only if all the B is consumed.Therefore, ϕ(t) is calculated by:

ϕ(t) =(< CA >|t − α) + χ−1 (1− < CB >|t)

1− α+ χ−1(34)

Concerning the evaluation of the mass transfer rate, the local dimension-less flux density of A at the time t on a point of the droplet interface locatedby θ is computed by:

jA(t, θ) =∂CA

∂r

∣∣∣∣t, θ

nr(θ) +∂CA

∂z

∣∣∣∣t, θ

nz(θ) (35)

The dimensionless total flux of A across the overall droplet interface at thetime t is then calculated by:

JA(t) =π

2

∫ π

0

jA(t, θ) sin θ dθ (36)

The Sherwood number at time t, Sh(t), which is a dimensionless masstransfer coefficient, is evaluated by dividing JA(t) by the product of thedimensionless droplet surface area (equal to π) and the dimensionless masstransfer driving force at time t. This driving force is the difference betweenthe maximum dimensionless concentration of A in the droplet (equal to 1)and < CA >|t. Therefore, the instantaneous Sherwood number at time treads:

Sh(t) =JA(t)

π (1− < CA >|t)(37)

It is more common in engineering to describe the mass transfer efficiencyby Sherwood number averaged over the absorption time. The time-averagedSherwood number Sh over a dimensionless fall time tf is calculated by:

Sh(tf ) =1

tf

∫ tf

0

Sh(t)dt (38)

13

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 14: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

In the case of a physical absorption, the concentration variation of A inthe droplet is directly related to the mass transfer rate. By expressing a massbalance, the following equation can be written:

JA(t) =π

6Pe

d < CA >

dt

∣∣∣∣t

(39)

The time-averaged Sherwood number Sh (defined by Eq. (38)) is calcu-lated using Eq. (39) in Eq. (37), leading to:

Sh(tf ) =Pe

6 tflog

(1− α

1− < CA >|tf

)(40)

According to the expression of ϕ in the physical absorption case (see Eq.(33)), the following relation between Sh(tf ) and ϕ(tf ) is immediately de-duced:

Sh(tf ) =Pe

6 tflog

(1

1− ϕ(tf )

)(41)

In the results section, Sh(tf ) is presented as a function of the Fouriernumber Fo = 4 t/Pe (such as in [9, 17, 18, 10]), for various Re and vari-ous Ha. Using the Fourier number enables the comparison of the differentregimes considered with a “normalized” dimensionless time scale, since thecharacteristic time scales of the different cases can vary in a large range(the investigated Reynolds number variation in this study lies on 4 order ofmagnitude).

Note that the Fourier number of a droplet can be related to the falldistance covered by this droplet. By the definitions of the reference time andthe Fourier number, the value of Fo reached after a dimensional fall distanceLf is given by the following expression:

Fo =4Lf η

ddReSc ρ∗(42)

Knowing CD for a given Re, a force balance enables to calculate the dimen-sional droplet diameter dd, which leads to the following equation:

dd =3

√3

4

η2g CD Re2

ρg (ρl − ρg) g(43)

14

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 15: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

where g is the gravity acceleration. Therefore, the value of Fo reached by adroplet after a fall of Lf can be evaluated if its Reynolds number and its dragcoefficient are known. For instance, in the experimental device of Amokraneand Caussade [20] where the fall height was 1 m, a droplet with a diameterof 1 mm (Re close to 250) reached the bottom at Fo = 2 10−3. For dropletswith a diameter smaller than 0.15 mm (Re < 4), Fo was beyond 1 at thedevice bottom.

4. Results and discussion

4.1. Validation of the flow field

The numerical simulation results are first compared with literature cor-relations for the drag coefficient. Since droplets having a diameter below 1mm can be considered as rigid spheres as far as the gas flow is concerned, thecomputed drag coefficient is compared with correlations for a rigid spherecited in [21]: Lapple (Eq. L3 of Table 5.1), Schiller and Naumann (Eq.S1 of Table 5.1), and the experimental values presented in their Tables 5.2and 5.6. The results are presented in Fig. 3 for Reynolds number varyingfrom Re = 0.02 to Re = 250. The result corresponding to the Stokes’law(CD = 24/Re) is also presented.

Fig. 3

It is observed that the calculated drag coefficients are in good agree-ment with the literature correlations for the range of Reynolds number used,though they tend to be slightly larger than the correlations for Re → 0.According to these results, the numerical flow fields are considered to bevalidated.

Note that this good agreement with correlations developed for rigid sphereshows that the gas flow is well estimated by solving only the momentum equa-tions in the gas phase (Eqs. (2)-(4)), imposing a velocity equal to zero atthe droplet interface. However, it is essential to compute the flows simul-taneously in the gas and liquid phases (Eqs. (2)-(7)), with the appropriateconditions at the droplet interface (Eqs. (8)-(10)), to determine the flowinduced in the droplet by its fall in the gas.

15

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 16: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

4.2. Physical absorption

Droplets having a Reynolds number up to 250 are considered. Typicalvalues for an air-water system are used. The used density and viscosity ratiosare ρ∗ = 830 and η∗ = 55, respectively. It is assumed that the gas componentis highly soluble in the liquid, i.e. h = 1 is selected. The diffusivity ratiobetween the gas the and liquid phases is βg = 103 and the Schmidt numberis Sc = 500. It is considered that, at the initial state, the concentration of Ais zero in the droplet: α = 0.

During the simulation of the transient mass absorption, the level of sat-uration ϕ is monitored. When ϕ reaches 0.9999, the simulation is stopped inorder to prevent the computation of an undetermined Sh. Indeed, close tothe saturation, the mass transfer driving force tends to zero as well as themass transfer rate.

In order to follow the absorption process evolution over time for the var-ious cases, the time evolution of ϕ and Sh are presented as a function of Foin Fig. 4 for several Reynolds number values. Note that it is verified for Shthat Eq. (38) leads to the same results than Eq. (41).

Fig. 4

It is observed in Fig. 4-b that for Fo → 0, all the curves converge to asingle curve (called hereafter LC0), meaning that the mass transfer rate isindependent of the Reynolds number. The mass transfer rate is controlledonly by the diffusion in the vicinity of the interface. For Fo → ∞, all thecurves converge to another single curve (called hereafter LC∞), distinct fromLC0. For intermediate Fo, a sudden change in the time evolution of Sh (aswell in the time evolution of ϕ in Fig. 4-a) is observed, after a time decreasingas the Re increases. This change reflects a transition in the limiting step ofthe mass transport mechanism.

The time evolution of the concentration field of A in the droplet is ex-amined in detail for three Reynolds number values: Re = 0.1, Re = 5 andRe = 200.

At Re = 0.1, the dimensionless A concentration field of A is presented fortwo Fo in Fig. 5.

Fig. 5

16

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 17: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

In this case, it seems that the mass absorption is dominated by the diffu-sive transport of A in the droplet, as the time evolution of the concentrationfield is very close to what would be observed in a purely diffusive process.

The dimensionless concentration field of A is presented for four Fo valuesat Re = 5 in Fig. 6.

Fig. 6

It is observed that the absorbed component A is simultaneously trans-ported by convection in the periphery of the droplet from the lower poleto the upper pole and penetrates in the droplet by diffusion. In this case,it seems that the convective and diffusive transport have the same order ofmagnitude.

The dimensionless concentration field of A is presented for four Fo valuesat Re = 200 in Fig. 7.

Fig. 7

It can be observed that the absorbed component A is, in a first stage,mainly transported by convection in the periphery of the droplet from thelower to the upper pole. A starts to penetrate in the vortex interior onlyafter the vortex periphery almost reaches saturation in A. A penetrates thengradually from the vortex periphery towards the interior by a mainly diffusiveprocess.

Based on this result analysis, it can be concluded that three differentsteps can be observed during the overall absorption process. The first step ispurely diffusive, as demonstrated by the fact that for any Re, the behaviorat sufficiently small time is given by a single curve (LC0), independent of Re.The reason why convection is not directly effective is probably that when theboundary layer is very thin, it is uniform along the interface, such that con-vection affects the concentration distribution only at the poles. Anyway, thisfirst step is short in practice, especially for large Reynolds number. Duringthe second step, the convection begins to deplete this boundary layer, first atthe upper pole, the boundary layer becomes less saturated and this enhancesthe transfer. That moment corresponds to the sharper increase of ϕ(t) andthe less pronounced decrease of Sh. In the third step, this enhancement thenstops after a certain time, when the liquid richer in component A starts toreach the lower pole. At this moment, the “loop is closed”, and the concentra-tion of A in the periphery of the vortex re-increases, progressively saturating

17

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 18: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

the interfacial region and decreasing the transfer. Consequently, the transferrate is again limited by the diffusion of the component A accumulated in theperiphery, towards the vortex interior.

The identification of these three successive limiting steps for the masstransfer rate enables us to propose simplified mechanisms to describe themass transfer rate evolution for any Re.

For Re → 0, only the first step is observed. In this case, the convectivecontribution to the mass transport becomes negligible and the mass absorp-tion is dominated by the diffusive transport in the droplet. The mass transferin the droplet can therefore be modeled assuming the purely diffusive masstransport in a stagnant liquid droplet, which writes:

∂CA

∂t=

1

Pe

1

r2∂

∂r

(r2∂CA

∂r

)(44)

Using Eq. (25) as initial condition and using CA|t, r=1 = 1 and ∂CA

∂r

∣∣t, r=0

= 0

as boundary conditions, Eq. (44) can be solved analytically and the solutionreads:

CA(t, r) = 1 + (1− α)∞∑i=1

(−1)i2Rd

iπrsin

(iπr

Rd

)exp

(− 1

Pe

(iπ

Rd

)2

t

)(45)

By spatially averaging Eq. (45) and expressing it as a function of Fo, thefollowing expression can be written:

< CA >|Fo = 1− (1− α)6

π2

∞∑i=1

1

i2exp

(− (iπ)2 Fo

)(46)

Therefore, using Eq. (41) and by injecting Eq. (46) in Eq. (33), Sh can beestimated as a function of Fo by the following analytical expression:

Shstagnant(Fo) = −2

3

1

Folog

(6

π2

∞∑i=1

1

i2exp

(− (iπ)2 Fo

))(47)

Note that this expression tends to Shstagnant = 2π2/3 for Fo → ∞.For Re → ∞, the third step is directly observed. The vortex periphery

can be considered as instantaneously saturated and the mass absorption iscontrolled by the diffusive transport within the vortex. The vortex zone, in

18

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 19: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

the three-dimensional droplet, has the form of a ring torus. If this torus istransversally cut and unfold along the longitudinal direction, the generatedvolume is a cylinder. The absorbed mass which has to penetrate by a mainlydiffusive process through this toroidal vortex can then be modeled by thediffusive mass transport through a cylinder along the cylinder radius coordi-nate rcyl with an imposed concentration at its surface. This model equationwrites:

∂CA

∂t=

1

Pe

1

rcyl

∂rcyl

(rcyl

∂CA

∂rcyl

)(48)

Using Eq. (25) as initial condition and using CA|t, rcyl=1 = 1 and

∂CA

∂rcyl

∣∣∣t, rcyl=0

= 0 as boundary conditions, Eq. (48) can be solved analytically

and the solution reads:

CA(t, rcyl) = 1− (1− α)∞∑i=1

γiζiJ0

(λircylRcyl

)exp

(− λ2iR2

cyl

t

Pe

)(49)

where Rcyl is the dimensionless radius of the cylinder equivalent to the torusformed by the vortex zone, λi is the i

th zero of the zero-order Bessel functionof the first kind J0, γi =

∫ 1

0xJ0 (λix) dx and ζi =

∫ 1

0x (J0 (λix))

2 dx.The spatially-averaged concentration of A over the cylinder volume, ex-

pressed as a function of the Fourier number, writes then:

< CA >cyl|Fo = 1− (1− α)∞∑i=1

2γ2iζi

exp

(−λ2i

(Rd

Rcyl

)2

Fo

)(50)

where Vcyl is the dimensionless volume of the cylinder equivalent to the torusformed by the vortex zone. Since the vortex periphery is considered as in-stantaneously saturated (with CA = 1 in this zone), the spatially-averagedconcentration of A in the overall droplet is related to the spatially-averagedconcentration in the torus modeled by the cylinder byVd < CA >= Vcyl < CA >cyl +(Vd − Vcyl). Therefore, the spatially-averagedconcentration of A over the droplet volume writes:

< CA >|Fo = 1− (1− α)VcylVd

∞∑i=1

2γ2iζi

exp

(−λ2i

(Rd

Rcyl

)2

Fo

)(51)

19

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 20: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Using Eq. (41) and by injecting Eq. (51) in Eq. (33), Sh can be estimatedas a function of Fo by the following analytical expression:

Shcyl(Fo) = −2

3

1

Folog

(∞∑i=1

2γ2iζi

VcylVd

exp

(−λ2i

(Rd

Rcyl

)2

Fo

))(52)

This equation tends to Shcyl =23λ21

(Rd

Rcyl

)2for Fo → ∞. Rcyl and Vcyl are

adjustable parameters of the model and their values are estimated by thecomparison with the curve LC∞ of Fig. 4-b. Rcyl is estimated to Rd/2.4 andVcyl is estimated to 0.97Vd. Note that these fitted values are geometricallycoherent for a ring torus contained in a spherical droplet.

For intermediate Re and limited Fo, it is observed during the initial stageof the mass absorption that the mass transfer rate is limited by the pen-etration in the vicinity of the interface, i.e. by a diffusive mechanism. Athin layer of liquid in contact with the interface can then be considered asa film and the width of this layer depends on the toroidal vortex angularvelocity, which decreases as the Reynolds number increases. Therefore, thepenetration-film model of Toor and Marchello [26] is proposed to model themass transfer rate before the saturation of the vortex periphery. The masstransport equation writes

∂CA

∂t=

1

Pe

∂2CA

∂x2(53)

where x is the dimensionless distance normal to the interface, oriented tothe droplet interior (x = 0 at the interface). Let δf be the dimensionlessfilm thickness of the model (i.e. the ratio of the dimensional film thicknessover the dimensional droplet diameter). Eq. (53) is solved analytically usingCA(x, t = 0) = α as initial conditions and CA(x = 0, t) = 1 and CA(x =δf , t) = α as boundary conditions. The solution reads:

CA(x, t) = α+ (1− α)∞∑i=0

(erfc

((2iδf + x)

√Pe

2√t

)− erfc

((2(i+ 1)δf − x)

√Pe

2√t

))(54)

The mass flux density is then calculated by:

jA(t) = − ∂CA

∂x

∣∣∣∣x=0, t

=

√Pe

πt

(1 + 2

∞∑i=1

exp

(−i2δ2fPe

t

))(55)

20

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 21: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Using JA(t) = 4πR2djA(t), Sh(t) = JA(t)/ (π(1− α)) and the definition of Sh

(see Eq. (38)), Sh can then be estimated as a function of Fo by the followinganalytical expression:

ShPF(Fo) =4√πFo

1 + 2∞∑i=1

exp

(−i2δ2fFo

)−

√πi2δ2fFo

erfc

√i2δ2fFo

(56)

δf is an adjustable parameter of this model. The values of δf are estimatedfor various Re (between 1 and 250) by fitting the curves computed using Eq.(56) with the initial part (i.e. before the transition to the third step) of theSh curves in Fig. 4-b.

It is observed that the values obtained for δf decrease as Re increases. Itis found that δf can be correlated to Re by the following expression:

δf = 0.1825Re−0.587 (57)

To summarize, thanks to these analysis, Sh for a droplet with a given Recan be estimated at any Fo using analytical expressions. For Re < 1, Shcan be approached by Shstagnant, given by Eq. (47). For Re > 1, Sh can beapproached by ShPF, given by Eq. (56) and using Eq. (57), during the initialstage and then by Shcyl, given by Eq. (52), when the vortex periphery issaturated. The transition is reached when ShPF becomes larger than Shcyl.In other words, Sh can be approached by: min

[Shcyl(Fo), ShPF(Re,Fo)

].

The time evolution of the analytical approximations of Sh are presentedin Fig. 8 and superimposed to the numerical simulation results. Eq. (47)is used for Re → 0, min

[Shcyl(Fo), ShPF(Re,Fo)

]is used for Re between 1

and 250, and Eq. (52) is used for Re → ∞. The analytical expressions arepresented in solid lines and the numerical simulation results are presented indashed lines. It is observed that the analytical expressions agree very wellwith the numerical simulation results.

Fig. 8

To end this section, a last comment is done on the curve presented in Fig.4-b. For Re > 1, it seems that the Fourier number at which the transitionoccurs is proportional to 1/Re. It would mean that the transition occurat a Fourier number proportional only to the number of diameter that the

21

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 22: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

droplet has covered during its fall (when the “loop is closed”). For a givenphysico-chemical system, this number would be constant and independent ofthe droplet size. This point will be analyzed in details in a future work.

4.3. Mass absorption enhanced by chemical reaction

In this section is studied the effect of a chemical reaction on the masstransfer rate, for various Reynolds and Hatta numbers. The initial concen-tration of B is considered in excess compared to the concentration of A atsaturation, as it is the case in the industrial practice. [B]0 is chosen as tentimes larger (χ = 0.1) than [A]eq. The simulations are realized for four val-ues of the Reynolds number, corresponding to the different regimes whichhave been described for the physical absorption: Re = 0.1, Re = 1, Re = 10and Re = 100. For each value of the Reynolds number, three values of theHatta number, corresponding to different chemical regimes, are considered:Ha = 10, Ha = 1 and Ha = 0.1. The same values than those chosen in theprevious section are used for ρ∗, η∗, Sc, βg, h and α. It is considered that thediffusion coefficient of the reactant B is smaller than the diffusion coefficientof A, then βb = 0.5 is selected.

The effect of a chemical reaction on the mass transfer rate and the sat-uration evolutions, for various Reynolds and Hatta numbers are presentedin the following figures. The time evolution of Sh and of ϕ as functions ofFo for the three values of Ha are presented and compared with the physicalabsorption case in Figs. 9, 11, 13 and 15 for Re = 0.1, Re = 1, Re = 10and Re = 100, respectively. The dimensionless concentration fields of A andB, for the three values of Ha, are presented at Fo = 10−1 in Fig. 10, atFo = 10−1 in Fig. 12, at Fo = 10−2 in Fig. 14 and at Fo = 10−3 in Fig. 16,for Re = 0.1, Re = 1, Re = 10 and Re = 100, respectively.

Fig. 9

Fig. 10

Fig. 11

Fig. 12

Fig. 13

Fig. 14

22

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 23: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 15

Fig. 16

For Ha = 0.1, it is observed in Figs. 9-a, 11-a, 13-a and 15-a that thepresence of B has a significant influence on the time evolution of ϕ. Thesaturation of the droplet is delayed and it is observed that this delay increasesas the Reynolds number decreases. By contrast, it is observed that thepresence of the chemical reactant does not have a significant influence onthe mass transfer rate. Indeed, the time evolution of Sh is always close tothe evolution observed for physical absorption, as it can be observed in Figs.9-b, 11-b, 13-b and 15-b. A slight enhancement is observed after Fo = 10−3

only for the case Re = 100 (Fig. 15-b).Note that it might seem surprising that the effect of the chemical reaction

is so pronounced on the evolution of the saturation ϕ (see Fig. 9-a), especiallyfor small values of Ha, while it has only a very small impact on Sh (see Fig. 9-b). Actually, even though the overall rate of absorption is almost unchangedby the chemical reaction when the latter is slow, ϕ is calculated in a differentway than in the physical absorption case, i.e. the full storage capacity of thedroplet is now considered (see the difference in Eqs. (33) and (34), the latterincorporating the additional storage capability due to chemical conversionby reaction with the reactant B, which increases as χ decreases).

On the one hand, the analysis of the concentration fields of B (in Figs. 10,12, 14 and 16) shows that, for any Re, the concentration of B decreases grad-ually in the droplet but its concentration field remains almost homogeneous.On the other hand, it is observed that the evolutions of the concentrationfields of A with Re are similar to those observed in the physical absorptioncase. Therefore, the mass transfer rate remains controlled only by the masstransport of A but the saturation is delayed by the presence of B.

For Ha = 1, it appears that the time evolutions of Sh and ϕ with Fo isdetermined by a combination of the phenomena controlling the mass transferfor Ha = 0.1 and Ha = 10 (analyzed hereunder), especially by the couplingof the transports of A and B. Indeed, for any Re, an intermediate situationbetween the cases Ha = 0.1 and Ha = 10 is observed for the considered Re.

For Ha = 10, it is observed in the time evolution of ϕ for any Re (seeFigs. 9-a, 11-a, 13-a and 15-a) that the saturation is always reached aroundthe same Fourier number than in the physical absorption case. By contrast,significantly different behaviors are observed for the time evolution of Sh,depending on Re.

23

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 24: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

For Re = 0.1, it is observed in Fig. 9-b that the mass transfer rate ishigher than in the physical absorption case, especially around Fo = 10−3, butit tends to be the same than in the physical absorption case for Fo → 0. Theanalysis of the concentration fields of A and B (see the right part of Fig. 10)shows that A is depleted by its consumption in the reaction. The concentra-tion of A close to the interface cannot increase as long as B is present in thiszone, maintaining a high concentration gradient of A near the interface andenhancing the mass transfer rate. For this value of Re, it has been shownthat the mass transport mechanism is mostly diffusive. Therefore, the masstransfer rate is mostly controlled by the diffusive transport of B in a firststage, and it becomes controlled by the diffusive transport of A in a secondstage, when B is depleted.

For Re = 1, it is observed in Fig. 11-b that the mass transfer rate becomessignificantly higher than in the physical absorption, until the droplet is closeto the saturation (see the corresponding Fo in Fig. 11-a). The analysis ofthe concentration fields of A and B (right part of Fig. 12) reveals that themass transfer rate is controlled by the transport of B in a first stage, andit becomes controlled by the transport of A in a second stage, but this timethe mass transport is determined by the combination of the convection andthe diffusion.

For Re = 10 and Re = 100, it is observed in Figs. 13-b and 15-b thatthe mass transfer rate is strongly enhanced (even more for Re = 100 thanfor Re = 10) until the liquid is close to the saturation. A sudden change inthe curve of Sh is also observed around Fo = 10−2 for Re = 10 and aroundFo = 10−3 for Re = 100, reflecting a transition in the mechanism controllingthe mass transfer rate, such as in the physical absorption case. Note thatthese transition occur at approximately the same Fo than those observed inthe physical absorption. The analysis of the concentration fields of A andB (right part of Figs. 14 and 16) shows that B is completely depleted inthe periphery of the toroidal vortex. The concentration of B is still highwithin the vortex and the concentration of A is just starting to fill the vortexperiphery. After that, the concentration of B decreases gradually within thevortex, while the concentration of A increases and the transport of A and Bwithin the toroidal vortex is mostly diffusive. Therefore, the mass transferrate is controlled in a first stage by the convective transport of B in thevicinity of the interface. It is strongly enhanced in a second stage by theconvective transport of A in the vortex periphery bo be finally controlled bythe coupling of the reaction with the diffusive transport of A and B within

24

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 25: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

the vortex.

5. Conclusion and perspectives

This work has considered direct numerical simulations of the gas-dropletmass transfer of a component A in gas phase into a liquid spherical dropletin free fall. A component B is dissolved in the liquid droplet, and thiscomponent reacts with A in the liquid phase, leading to an increase of thestorage capability and an enhancement of the mass transfer rate.

The mass transfer is studied by computing simultaneously the flow fieldsand the concentration fields in both gas and liquid phases. The analysis of thetime evolution of the concentration fields for various regimes, distinguishedby the Reynolds and the Hatta numbers, has enabled us to highlight thecoupling between the phenomena for each regime and to identify which onescontrol the mass transfer rate.

In section 4.2, the physical absorption (without reaction) is studied. It isconcluded from the results analysis that three different steps can be observedduring the overall absorption process. The first step is purely diffusive, asdemonstrated by the fact that the behavior at sufficiently small time is inde-pendent of Re. The diffusive transport is directed towards the droplet center.During the second step, the convection begins to deplete this boundary layer,first at the upper pole. The boundary layer becomes less saturated and thisenhances the transfer. In the third step, this enhancement then stops af-ter a certain time, when the liquid richer in component A starts to reachthe lower pole. The concentration of A in the periphery of the vortex re-increases, progressively saturating the interfacial region and decreasing thetransfer. Consequently, the transfer rate is again limited by the diffusion ofthe component A accumulated in the periphery, towards the vortex inside.

For Re → 0, only the first step is observed. For Re → ∞, the third stepis directly observed. For intermediate Re, the mass transfer rate is controlledsuccessively by these three limiting steps and the time of transition betweenthe steps decreases as Re increases.

The identification of these three successive limiting steps for the masstransfer rate enables us to propose simplified mechanisms to describe themass transfer rate evolution for any flow regime. Based on these simplifiedmechanisms, analytical expressions are developed to evaluate Sh as a functionof Fo for any Re.

25

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 26: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

In the limit of Re → 0 (first step), Sh is modeled by Shstagnant (Eq. (47)),while in the limit of Re → ∞ (third step), Sh is modeled by Shcyl (Eq. (52))(with the adjustable parameters Rcyl and Vcyl fitted to Rd/2.4 and 0.97Vd,respectively).

For Re < 1, Sh(Re,Fo) can be approached by Shstagnant(Fo). For Re > 1,Sh can be approached by ShPF during the initial stage (given by Eq. (56) andusing the correlation (57) developed by comparison with the numerical sim-ulation results), and later by Shcyl, when the periphery of the toroidal vortexis saturated. The transition is reached when ShPF becomes larger than Shcyl,therefore Sh(Re,Fo) can be approached by min

[Shcyl(Fo), ShPF(Re,Fo)

].

It is worth mentioning that these analytical expressions are compared tothe numerical simulation results (see Fig. (8)) and that an excellent agree-ment is observed.

In section 4.3, the influence of a chemical reaction on the mass absorptionrate is investigated. In this case, the chemical reaction can enhance the masstransfer rate and can delay the saturation of the droplet, depending on theReynolds and the Hatta number.

For Ha < 1, it is observed that the presence of B does not have a signif-icant influence on the mass transfer rate and it is concluded that the masstransfer rate remains mostly controlled by the mass transport of A in thedroplet, as without reaction. However, the saturation is delayed by the pres-ence of B as it increases the droplet absorption capability due to chemicalconversion by reaction and the delay increases as the Reynolds number de-creases.

For Ha = 1, an intermediate situation is observed. The mass transferrate of A also gets influenced by the transport of B in the droplet.

Finally, for Ha > 1, the mass transfer rate is controlled by the transportof B in the droplet in a first stage. Later, it becomes controlled by theinteraction between the reaction and the transport of A and B. The prevalentmass transport mechanism depends on the value of Re. For Re < 1, the masstransport is mostly diffusive towards the droplet center. For Re = 1, themass transport is determined by the combination of the convection and thediffusion. For Re > 1, the mass transfer rate is controlled in a first stage bythe convective transport of B in the vicinity of the interface. It is controlledin a second stage by the convective transport of A in the vortex peripheryand finally it is controlled by the interaction of the reaction and the diffusivetransport of A and B within the vortex.

The results presented in this paper and their detailed analysis show the

26

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 27: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

importance of taking into account all the phenomena simultaneously in gas-droplet mass transfer modeling. However, a further analysis of some issuesfor Re > 1 would be worth of interest, such as the observation that in thephysical absorption, the value of Fo at the transition between the second andthe third step seems to be proportional to 1/Re and the observation thatthe value of Fo at the transition for Ha = 10 seems to be the same thanin the physical absorption. A criterion determining the value of Fo at thetransition could be identified. It would also be full of interest to develop asimplified model incorporating the effect of chemical reaction. In the limitHa → ∞, it would be possible to determine a correlation for the mass transferrate enhancement, based of the mass transport mechanism observed in thephysical absorption for any Re, at least for the initial stage (when the masstransfer rate is controlled by the transport of B).

Another perspectives are to model the effects of the presence of surface-active contaminants and to extend this analysis to larger drops. In thislatter case, it will be necessary to take into account the deformations andthe oscillations of the drop interface and the turbulence in the flow fields.Several methods are available and have been described in the literature forphysical absorption, such as the moving element method [8] or the level setmethod [27] .

Acknowledgements

F.R.S.-FNRS (Belgian National Fund for Scientific Research) Postdoc-toral Researcher Christophe Wylock and Senior Research Associate PierreColinet gratefully acknowledge the financial support from the Fund.

References

[1] C. Brogren, H. T. Karlsson, Modeling of the absorption of SO2 in aspray scrubber using the penetration theory, Chem. Eng. Sci. 52(18)(1997) 3085–3099.

[2] A. Bandyopadhyay, M. N. Biswas, Modelling of SO2 scrubbing in spraytowers, Sci. Total Environ. 383 (2007) 25–40.

[3] S. Sarkar, B. C. Meikap, S. G. Chatterjee, Modeling of removal of sulfurdioxide from flue gases in a horizontal cocurrent gasliquid scrubber,Chem. Eng. J. 131 (2007) 263–271.

27

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 28: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

[4] M. K. Akbar, S. M. Ghiaasiaan, Modeling the gas absorption in a sprayscrubber with dissolving reactive particles, Chem. Eng. Sci. 59 (2004)967–976.

[5] T. K. Sherwood, R. L. Pigford, Absorption and extraction, McGraw-Hill, New York, 1963.

[6] R. Higbie, The rate of absorption of pure gas into a still liquid duringshort periods of exposure, Trans. Am. Inst. Chem. Ing. 31 (1935) 365.

[7] B. P. Leclair, A. E. Hamielec, H. R. Pruppacher, W. D. Hall, A theo-retical and experimental study of the internal circulation in water dropsfalling at terminal velocity in air, J. Atmos. Sci. 29(4) (1972) 728–740.

[8] J. Petera, L. Weatherly, Modelling of mass transfer from falling droplets,Chem. Eng. Sci. 56 (2001) 4929–4947.

[9] B. I. Brounshtein, A. S. Zheleznyak, G. A. Fishbein, Heat and masstransfer in interaction of spherical drops and gas bubbles with a liquidflow, Int. J. Heat & Mass Transfer 13 (1970) 963–973.

[10] B. I. Broushtein, G. A. Fishbein, V. Y. Rivikind, Mass transfer accom-panied by chemical conversion of substance in a drop, Int. J. Heat &Mass Transfer 19 (2) (1976) 193–199.

[11] Lamb, Hydrodynamics, 6th Edition, Dover Publications, New York,1932.

[12] A. E. Hameliec, A. I. Johnson, Viscous flow around fluid spheres atintermediate Reynolds number, Can. J. Chem. Eng. 40 (2) (1962) 41–45.

[13] R. Kronig, J. C. Brink, On the theory of extraction from falling droplets,Appl. Sci. Res. A 2 (1950) 142–154.

[14] L. Baboolal, H. Pruppacher, J. Topalian, A sensitivity study of a the-oretical model of SO2 scavenging by water drops in air, J. Atmos. Sci.38 (1980) 856–870.

[15] T. Elperin, A. Fominykh, Conjugate mass transfer during gas absorptionby falling liquid droplet with internal circulation, Atmos. Env. 39 (2005)4575–4582.

28

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 29: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

[16] A. R. Uribe-Ramırez, W. J. Korchinsky, Fundamental theory for predic-tion of single-component mass transfer in liquid drops at intermediateReynolds numbers (10 ≤ Re ≤ 250), Chem. Eng. Sci. 55(16) (2000)3305–3318.

[17] M. Waheed, M. Henschke, A. Pfennig, Mass transfer by free and forcedconvection from single spherical liquid drops, Int. J. Heat & Mass Trans-fer 45 (22) (2002) 4507–4514.

[18] A. Paschedag, W. H. Piarah, M. Kraume, Sensitivity study for the masstransfer at a single droplet, Int. J. Heat & Mass Transfer 48 (2005)3402–3410.

[19] M. K. Akbar, J. Yan, S. M. Ghiaasiaan, Mechanism of gas absorptionenhancement in a slurry droplet containing reactive, sparingly solublemicroparticles, Int. J. Heat & Mass Transfer 46 (2003) 4561–4571.

[20] H. Amokrane, B. Caussade, Gas absorption into a moving spheroidalwater drop, J. Atmos. Sci. 56(12) (1999) 1808–1829.

[21] R. Clift, J. Grace, M. Weber, Bubbles, drops and particles, Dover Pub-lications, Mineola, 1978.

[22] C. J. Walcek, H. R. Pruppacher, J. H. Topalian, S. K. Mitra, On thescavenging of SO2 by cloud and raindrops II: An experimental study ofSO2 absorption and desorption for water drops in air, J. Atmos. Chem.1 (3) (1984) 291–306.

[23] E. R. Altwicker, C. E. Lindhjem, Absorption of gases into drops,A.I.Ch.E. Journal 34(2) (1988) 329–332.

[24] P. Trambouze, J.-P. Euzen, Chemical Reactors: from design to opera-tion, Institut Francais du petrole publications, Technip, 2004.

[25] R. B. Bird, W. E. Stewart, E. N. Lightfoot, Transport Phenomena,seconde Edition, John Wiley&Sons, 2002.

[26] H. L. Toor, J. M. Marchello, Film-penetration model for mass and heattransfer, A.I.Ch.E. Journal 4 (1) (1958) 97–101.

29

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 30: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

[27] K. B. Deshpande, W. B. Zimmerman, Simulation of interfacial masstransfer by droplet dynamics using the level set method, Chem. Eng.Sci. 61 (2006) 6486–6498.

30

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 31: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

List of Figures

1 Sketch of the computational domain including the referenceframe. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2 Pictures of the mesh in each domain. . . . . . . . . . . . . . . 323 CD versus Re: comparison of the model simulation results

with literature correlation. . . . . . . . . . . . . . . . . . . . . 334 ϕ (a) and Sh (b) versus Fo for various Re in the case of physical

absorption. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335 Fields of CA at Re = 0.1 for various Fo. . . . . . . . . . . . . . 336 Fields of CA at Re = 5 for various Fo. . . . . . . . . . . . . . . 347 Fields of CA at Re = 200 for various Fo. . . . . . . . . . . . . 348 Sh versus Fo evaluated with the analytical expressions (solid

lines) and numerically simulated (dashed lines). . . . . . . . . 349 ϕ (a) Sh (b) versus Fo for Re = 0.1 and for various Ha. . . . . 3510 Fields of CA (top) and CB (bottom) at Fo = 10−1 for Re = 0.1

and for Ha = 0.1 (left), Ha = 1 (middle) and Ha = 10 (right). 3511 ϕ (a) and Sh (b) versus Fo for Re = 1 and for various Ha. . . . 3512 Fields of CA (top) and CB(bottom) at Fo = 10−1 for Re = 1

and for Ha = 0.1 (left), Ha = 1 (middle) and Ha = 10 (right). 3613 ϕ (a) and Sh (b) versus Fo for Re = 10 and for various Ha. . . 3614 Fields of CA (top) and CB(bottom) at Fo = 10−2 for Re = 10

and for Ha = 0.1 (left), Ha = 1 (middle) and Ha = 10 (right). 3715 ϕ (a) and Sh (b) versus Fo for Re = 100 and for various Ha. . 3716 Fields of CA (top) and CB (bottom) at Fo = 10−3 for Re = 100

and for Ha = 0.1 (left), Ha = 1 (middle) and Ha = 10 (right). 38

31

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 32: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 1

Fig. 2

32

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 33: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 3

(a) (b)

Fig. 4

Fig. 5

33

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 34: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 6

Fig. 7

Fig. 8

34

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 35: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

(a) (b)

Fig. 9

Fig. 10

(a) (b)

Fig. 11

35

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 36: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 12

(a) (b)

Fig. 13

36

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 37: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 14

(a) (b)

Fig. 15

37

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085

Page 38: Gas absorption into a spherical liquid droplet: numerical ... · Gas absorption into a spherical liquid droplet: numerical and theoretical study C. Wylock,a, P. Colineta, B. Hauta

Fig. 16

38

This is the preprint version of a manuscript published in Chemical Engineering Journal 207-208 (2012), p. 851-864 DOI: 10.1016/j.cej.2012.07.085 Permalink: http://dx.doi.org/10.1016/j.cej.2012.07.085