916
Views
2
CrossRef citations to date
0
Altmetric
Article

Nonuniform particle distribution and interference between removal mechanisms during unsteady aerosol deposition from a rising spherical bubble

ORCID Icon, &
Pages 1037-1046 | Received 18 Sep 2021, Accepted 06 Jan 2022, Published online: 22 Feb 2022

ABSTRACT

Most of the practical computer models available today on aerosol pool scrubbing are built on the basis of fundamental assumptions introduced by Fuchs for his model published in 1955 (in Russian). The primary assumptions are that particle concentration in the bubble can be regarded as uniform and that particle removal mechanisms can be regarded as mutually independent. The present paper shows that these assumptions can cause considerable under- or overprediction of removal rates when applied to particles with radii in the submicrometer range. This size range, known to be where removal mechanisms are generally inefficient, is of particular interest in pool scrubbing study. The present paper, while not being the first to point out the problems with Fuchs’s assumptions, provides a comprehensive description on how the assumptions fail in the submictometer range based on comparisons of Fuchs model predictions against Lagrangian tracking simulations. The simulations, performed for particles of 0.1 to 1.0 μm in radius, advected by circulating gas flow in a spherical bubble rising in a water pool, show interference between inertial migration and Brownian diffusion of particles through the development of nonuniform particle concentration profile.

Graphical Abstract

1. Introduction

The radioactive source term to the environment under nuclear reactor severe accident conditions needs to be limited by all available means. Pool scrubbing is among those counted on for retention of radionuclides vapor and aerosols in, e.g. BWR pressure suppression pool and PWR steam generator secondary side [Citation1,Citation2]. Although the pool scrubbing process will show different appearances depending on the injection geometry and gas velocity [Citation3,Citation4], its last stage, spanning the majority of the total height, will always take the form of small bubbles rising buoyantly in water. Since this stage takes over residues from the preceding processes and stages, the bubbles therein will consist primarily of noncondensable gases laden with aerosol particles with a narrowed size distribution around the so-called Greenfield gap [Citation5], a size range between a few tens of nanometer and ~1 μm in aerodynamic radius, where the basic deposition mechanisms (sedimentation, inertial impaction and Brownian diffusion) are all inefficient. Quantification, experimental and analytical, of pool scrubbing performance for such particle sizes has therefore gathered considerable research interest [Citation6–8]. Works published recently on this issue include those focusing on the influence of bubble and pool parameters [Citation9–13].

Existing models for pool scrubbing analysis, including those used in computer codes including VANESA [Citation14], SPARC [Citation15], BUSCA [Citation16], SUPRA [Citation17], SRT [Citation13], POSCAR [Citation11] and I-COSTA [Citation12], are mostly based on the pioneering work of Fuchs in 1964 [Citation18]. Fuchs’s model takes a single-control-volume lumped parameter approach to modeling particle removal from the bubble. This implies that the particles in the bubble are assumed to have an approximately uniform concentration profile. The model further assumes that the particle deposition mechanisms are mutually independent such that the deposition velocity is given by the sum of velocities due to the individual mechanisms. The deposition velocity due to the Brownian diffusion is calculated by applying Higbie’s penetration theory [Citation19] to the particle transfer inside the bubble. This choice has been criticized repeatedly for being fundamentally inadequate [Citation20,Citation21] since the theory is applicable only to transfer at the surface of a large well-mixed bulk. It is noted that this choice fits (unnecessarily) well in the frame of simplification assumptions on uniformity and independence.

The assumption of independent mechanisms may represent a fair approximation when the particle deposition is dictated by a single dominant mechanism. This, however, will not be the case for particle sizes close to the deposition velocity minimum where the inertial and Brownian contributions are both small and comparable to each other. Under such conditions, possible interference between the two mechanisms may affect the particle deposition flux.

The adequacy of the Fuchs model’s assumptions, of concentration uniformity and mechanisms independence, has been addressed by analyzing particle transport and deposition by using an Eulerian (continuum based) model with one-dimensional approximation [Citation22] and three-dimensional Lagrangian particle tracking [Citation23]. While these studies show non-uniform concentration profile evolving with time and deviations from Fuchs model predictions in particle deposition rate, the underlying mechanisms, including the possible interference between the deposition mechanisms, are not discussed in depth.

The present work is based on three-dimensional Lagrangian simulations of particles with sizes between 0.1 and 1.0 μm in radius where the dominant mechanisms for particle retention are inertial impaction and Brownian diffusion. The simulations show that the inertial transport causes faster-than-average particle depletion around the center of recirculating flow inside the bubble, producing a surface-peaked concentration profile like in cyclone separators, with surface-to-average concentration ratio increasing with elapsed time. As such, the inertial deposition rate becomes increasingly greater than the Fuchs model prediction. The influence of diffusion is observable for the smallest particle radius simulated (0.1 μm) but in a much different way from the Fuchs model’s presumption: a quick growth of bubble-surface diffusion layer causes a transition of dominant removal mechanisms from unsteady Brownian diffusion to inertial migration. Fuchs’s model, based on the penetration theory, predicts a time-invariant deposition velocity corresponding to a very thin diffusion layer, and hence largely overpredicts the removal rates for small particles.

Fuchs-based computer models are used widely for predicting pool scrubbing performance even in the latest journal papers obviously without adequate recognition of the above problems. The present paper intends to contribute to closing such knowledge gap.

2. Mathematical models

2.1. Fluid phase

We consider a single bubble with a radius Rb rising steadily in stationary water at a terminal velocity Ut. At room temperature, bubbles with a radius of less than about 0.5 mm will be spherical [Citation24]. The bubble internal flow is assumed to be steady in the frame moving with the rising bubble, and modeled by Hill’s spherical vortex [Citation25] as has been done in many studies in literature including those cited in Introduction. The contribution of steam condensation (i.e. steam movement to the bubble interface) is ignored since the bubbles to be modeled will consist primarily of noncondensable gases. shows schematically the Hill’s spherical vortex streamlines.

Figure 1. Schematic of Hill’s spherical vortex.

Figure 1. Schematic of Hill’s spherical vortex.

The Stokes stream function ψ for Hill’s spherical vortex, satisfying velocity continuity with an external flow, approximated by a potential flow, can be written for spherical coordinate system r =r,θ,φ as:

(1) ψr,θ=3Ut4Rb2Rb2r2r2sin2θ,    rRb.(1)

The bubble rise velocity Ut is obtained using Peebles and Garber correlation [Citation26,Citation27]

(2) Ut=0.33g0.76νl0.52Rb1.28 \breakfor \break2<Reb<4.02M0.214,(2)

where M=gμl4/ρlσ3 is the Morton number with νl, ρl, g, σ, and μl being the kinematic viscosity of the continuous phase (water), the density of water, the gravitational acceleration, surface tension, and the dynamic viscosity of water, respectively. This correlation has been used in pool scrubbing studies, e.g. [Citation22,Citation23,Citation28], and covers small bubble radii including the value assumed in the present study (0.5 mm).

The bubble internal velocity U=Ur,Uθ,0 is given as follows:

(3) U=rotψrsinθeφ,(3)

where eφ=0,0,1 is the unit vector in the φ direction.

2.2. Particle motion

We consider monodispersed solid spherical particles of radius Rp and density ρp small enough for the Stokes law to hold. The particles are assumed to move independently of each other and do not affect the flow field. The assumption is valid for dilute aerosols: Brownian coagulation can be ignored for particle number concentrations less than 1014/m3 [Citation29] and the fluid-particle interactions can be taken to be one-way when the particle volume fraction is less than 104 [Citation30]. The particle number concentrations and volume fractions considered in the present study are less than 1013/m3 and less than 4×105, respectively. To simulate the motion of a particle of mass mp=4/3πρpRp3, centered instantaneously at xpt and moving with velocity vpt in a steady flow U of gas with density ρg=1.172kg/m3 and dynamic viscosity μg=1.863×105Pas, we use the equation of motion in the form proposed by Maxey and Riley [Citation31] with a Brownian force term:

(4a) dxpdt=vp,(4a)
(4b) mpdvpdt=Fd+FB+Fa+Fs+Fg,(4b)

where Fd, FB, Fa, Fs, and Fg are Stokes drag, Brownian, added mass, fluid stress gradient, and gravitational forces, respectively. The history and lift forces are ignored since they are known to be small in gas-solid flows [Citation32]. The drag force is given by the Stokes law

(5) Fd=fUvp,(5)

where f is the drag coefficient

(6a) f=6πμgRpCc,(6a)
(6b) Cc=1+λRp1.257+0.4exp1.1Rpλ,(6b)

where Cc is the Cunningham correction factor that deviates from unity when the particle radius is not much larger than the mean free path λ of gas molecules. The values of coefficients in EquationEquation (6b) are those obtained by Davies [Citation33] for air. Brownian force is modeled as a Gaussian white noise random process [Citation34]:

(7) FB=ζ2kb2T2DBΔt,(7)

where ζ, kb, DB, T, and Δt are Gaussian random numbers with zero means and unit variances, Boltzmann constant, Brownian diffusion constant, gas temperature, and time step, respectively. The Brownian constant DB is obtained using the Stokes-Einstein relation:

(8) DB=kbTf,(8)

The added mass, fluid stress gradient, and gravitational forces are

(9a) Fa=23πρgRp3DUDtDvpDt,(9a)
(9b) Fs=43πρgRp3DUDt,(9b)

and

(9c) Fg=43πRp3ρpρgg,(9c)

where g=0,0,g is the gravitational acceleration vector. By using these relations, Equation (4) can be rewritten as follows:

(10a) dxpdt=vp,(10a)
(10b) mp+madvpdt=fUvp+ζ2kb2T2DBΔt+2πρgRp3UU+43πRp3ρpρgg,(10b)

where ma=2/3πρgRp3 is the added mass.

2.3. Fuchs model

The extent of aerosol removal is commonly expressed using the decontamination factor DF or the depletion efficiency ε. The DF and ε, defined based on the total number of particles Np in the system of interest, are

(11a) DFt=Np0Npt1,(11a)
(11b) εt=11DFt.(11b)

Fuchs’s model [Citation18] assumes that the decay of Np with time is given by

(12) dNpdt=Npjαj,(12)

where αj is the removal coefficient for mechanism j. The coefficients for inertia αi and gravity αg are:

(13a) αi=9τpUt22Rb2,(13a)
(13b) αg=3gτp4Rb,(13b)

where τp=mp+ma/f is the particle relaxation time. EquationEquations (12) and (Equation13a) calculate the inertial removal rate as the particle concentration taken as np=Np/Vb, where Vb is the volume of the bubble, times the inertial (centrifugal) terminal velocity of the particle approximated by ||URb,θ||2/Rbτp, integrated over the bubble surface. Similarly, EquationEquations (12) and (Equation13b) calculate the gravitational removal rate based on the terminal velocity for gravity gτp.

The removal coefficient for Brownian diffusion αB is derived based on the penetration theory [Citation19] and is given by:

(13c) αB=1.8UtDBRb3,(13c)

EquationEquation (12) means that Np decays exponentially at a rate proportional to the sum of coefficients. Each coefficient should then be inversely proportional to the characteristic time of respective particle removal mechanisms, even when the linear superposition approach taken by EquationEquation (12) could be justified. In EquationEquations (13a) and (Equation13b), the inertial and gravitational characteristic times are reasonably taken to be the bubble characteristic length Rb divided by the respective particle terminal velocity. In EquationEquation (13c), on the other hand, the diffusional characteristic time is taken, disputably, to be the square root of the product of diffusive and advective characteristic times.

The assumption in EquationEquations (12) and (13) that the particle removal rate is proportional to np=Np/Vb means that a lumped parameter (ideally mixed bubble) approach is taken. The temporal variation of concentration distribution is not taken into account. Then, the relative magnitudes of coefficients, time independent, can be expressed using the nondimensional numbers for transport phenomena.

(14a) αiαg=6Ut2gRb=6Fr2,(14a)
(14b) αiαB=2.5Pe0.5St,(14b)

where Fr=Ut/gRb, Pe=UtRb/DB, and St=τpUt/Rb are Froude, Péclet, and Stokes numbers, respectively. EquationEquation (14a) shows that the importance of gravity relative to inertia depends only on bubble velocity and radius. It is found that the gravity effect on particle removal is much smaller than the inertia effect regardless of the bubble radii.

3. Numerical conditions and validation

The particles are initially distributed randomly in the bubble. Each particle has a velocity equal to the local fluid velocity, vpt=0 = U, but starts being accelerated at t=0, following the equation of motion, Equation (4). Each particle is tracked until the following removal condition is satisfied.

(15) ||xp|| >RbRp,(15)

Parameters and their values for simulations are listed in , where Np0 and P are the initial number of particles and pressure, respectively. Three particle radii are chosen from the range where all the removal mechanisms are known to be inefficient [Citation8]. The Cunningham correction factor used in the Stokes law (EquationEquation (6a)) is obtained using the equation of Davies (EquationEquation (6b)) as Cc=1.865, 1.163, and 1.082 for Rp=0.1 μm,0.5 μm, and 1.0 μm, respectively, showing that the flow around the particle starts deviating from the continuum regime for the smallest radius.

Table 1. Parameters of numerical simulation

The simulation program was tested against a benchmark problem where 1000 particles were initially placed on a single point in a stationary fluid and then allowed to follow the equation of motion, Equation (4), with the gravity term nullified. ) compares the root mean square (RMS) displacement of particles obtained from simulations (dots) with the theoretical values RMSxp=6DBt for Brownian motion (solid lines). The agreement is satisfactory for all the conditions tested.

Figure 2. (a) Root-mean-square displacement for Brownian particles with Rp=0.1 μm, Rp=0.5 μm, and Rp=1.0 μm as a function of time. The lines are theoretical predictions and the dots are the results of the numerical simulation. (b) Fraction of remaining particles 1ε as a function of time calculated with the fourth-order Runge-Kutta (RK) and the first-order Euler schemes. Time steps are Δt=0.1τp, Δt=0.5τp, and Δt=2.0τp. The particle radius Rp=0.5 μm.

Figure 2. (a) Root-mean-square displacement for Brownian particles with Rp=0.1 μm, Rp=0.5 μm, and Rp=1.0 μm as a function of time. The lines are theoretical predictions and the dots are the results of the numerical simulation. (b) Fraction of remaining particles 1−ε as a function of time calculated with the fourth-order Runge-Kutta (RK) and the first-order Euler schemes. Time steps are Δt=0.1τp, Δt=0.5τp, and Δt=2.0τp. The particle radius Rp=0.5 μm.

For particle tracking simulations, it is commonly recommended to keep the time step Δt small enough as compared to the relaxation time τp [Citation35]. Since this requirement becomes demanding when sub-micron particles with small τp are to be simulated, we performed test calculations to find how much the requirement could be relaxed in our particular cases. The results are shown in ) for the fraction of remaining particles 1ε vs. time calculated for Rp=0.5 μm with different values of Δt and different integration schemes. The fourth-order Runge-Kutta, used as the standard scheme in our simulations, shows no systematic deviation with an increase in Δt, even when Δt is increased to 2.0τp. This suggests that a higher order integration scheme can accommodate larger time steps than τp at least for certain cases. In contrast, the first-order Euler scheme requires much smaller time step than τp as shown in the same figure.

4. Results and discussion

Simulation results are presented and compared with Fuchs model predictions for three cases where the particle removal mechanisms were considered fully or partially as summarized in . The mechanisms considered in each case are represented by the forces activated in simulations, and by the removal coefficients deployed in Fuchs model calculations, respectively. In , those in the parentheses are known to have relatively small influences on the results. Namely, the major contributors here are inertia and diffusion.

Table 2. The forces considered in the numerical simulation and the removal coefficients used in the Fuchs model calculations

can be read that Case 1 considers both inertia and diffusion (and others), Case 2 inertia alone, and Case 3 diffusion alone. Case 3 calculation was conducted only with the Fuchs model. To perform simulation for Case 3, i.e. particle motion due to Brownian diffusion (advection-diffusion) without inertial displacement, some artificial treatment would become necessary. Instead of doing that, we use a numerical solution for that case, in an approximation form shown in Mills and Hoseyni [Citation20], for comparison with other cases. Note that inertial displacement of a particle is calculated in simulations as the consequence of the lack of enough fluid forces for retaining the particle on a curved streamline, while it is calculated by Fuchs’s model as caused by an inertial (centrifugal) force.

) compares simulation results for Cases 1 and 2. First of all, it is clear that the particle removal rate decreases quickly as the particle size decreases. For larger sizes, Rp=0.5 and 1.0 μm, no meaningful differences are seen between the two cases, showing that the inertial removal prevails for both sizes, i.e. the inertial effect is much greater than the gravitational and diffusional effects (that are considered in Case 1 but not in Case 2). For a smaller particle size Rp=0.1 μm, however, slightly faster removal is seen for Case 1 than for Case 2. The difference between the two cases occurs immediately after the initiations of removal, and remains almost constant after about 1 s.

Figure 3. Fraction of remaining particles 1ε as a function of time for (a) the Case 1 and Case 2 simulations and (b) the Case 1 simulation and the Fuchs model predictions for Case 1, for particle radii Rp=0.1 μm,0.5 μm, and 1.0 μm.

Figure 3. Fraction of remaining particles 1−ε as a function of time for (a) the Case 1 and Case 2 simulations and (b) the Case 1 simulation and the Fuchs model predictions for Case 1, for particle radii Rp=0.1 μm,0.5 μm, and 1.0 μm.

The Rp dependence of the difference between Cases 1 and 2 should be related to the Rp dependence of forces considered or unconsidered in the two cases. We can express the Rp dependence of the individual forces as FdRp, Fa+FsRp3, FgRp3, and FBRp, if we ignore the weak dependence of the Cunningham correction factor Cc on Rp. This scaling consideration shows that the importance of Brownian diffusion increases with the decrease in particle size, and suggests that the difference between Cases 1 and 2 results for Rp=0.1 μm stems from the diffusion effects in the early phase of the transient. This problem will be further discussed later in this paper.

) shows comparisons between the present simulation and Fuchs model prediction for Case 1 with the three particle radii. Overall, the Fuchs model predictions show considerably smaller dependence on Rp than the simulations. For larger particle sizes Rp=0.5 and 1.0 μm, subject to inertia-dominated removal as found in ), Fuchs model initially shows a fairly good agreement with the simulations, as long as 1ε>0.7, but then starts predicting slower removal than the simulations. In contrast to this, for a smaller particle size Rp=0.1 μm, the Fuchs model predicts much faster removal than the simulation.

shows further comparisons for Rp=0.1 μm, seeking for the cause of the large difference seen in ) between the Fuchs model and simulations, with the aid of Case 2 calculations. The Case 2 results, with consideration of the inertia effect alone, show no significant difference between the Fuchs and simulation results. The difference becomes significant only when the Brownian diffusion is considered as in Case 1, while this would not immediately mean that the diffusion effect alone is responsible for the difference.

Figure 4. Fraction of remaining particles 1ε as a function of time in the range of 0.0<t<1.0 for the Case 1 and 2 simulations, Fuchs model predictions for Cases 1 and 2, and Mills and Hoseyni [Citation20] proposed curve for Case 3, all for particle radius Rp=0.1 μm.

Figure 4. Fraction of remaining particles 1−ε as a function of time in the range of 0.0<t<1.0 for the Case 1 and 2 simulations, Fuchs model predictions for Cases 1 and 2, and Mills and Hoseyni [Citation20] proposed curve for Case 3, all for particle radius Rp=0.1 μm.

Inertial displacement occurs necessarily when solid particles are advected by gas flow along curved streamlines, as in the present problem. Therefore, evaluation of Brownian diffusion effect separately from the inertial displacement could be done only through thought experiments. Shown in for comparison with simulations and Fuchs model prediction is a curve proposed by Mill and Hoseyni [Citation20] based on an approximate expression for analytical solution obtained by Calderbank and Korchinski [Citation36] for the unsteady advection-diffusion problem formulated by Kronig and Brink [Citation37] for Hill’s vortex flow. The Kronig and Brink problem concerns the advection-diffusion of scalar quantity (heat or dissolved mass) for the limiting case of Re0,Pe. Mills and Hoseyni proposed application of this solution to Brownian particle diffusion in the same flow. As depicted in , the Mills and Hoseyni curve shows a faster particle removal than the inertia-alone (Case 2) solutions until 0.5 s, but a slower removal thereafter, suggesting, again, that diffusion effect could be important at the beginning of the transient.

The temporal change of the particle concentration profile was derived from the simulation results by counting the particles remaining in each of ten equal-volume, concentric toroidal regions defined to be bounded by stream surfaces (i.e. the isosurface of ψ). shows normalized particle number density npt/np0: np0=Np0/Vb plotted against normalized stream function ξ=ψ/maxψ, where ξ=0 corresponds to the bubble surface and ξ=1 to the vortex center. The results are obtained from simulations conducted with an enhanced initial particle number Np0=10,000 and repeated five times to reduce the random error.

Figure 5. The distribution of particle concentration npt normalized by the initial value np0 and 95% confidence intervals for (a) Rp=1.0 μm and (b) Rp=0.1 μm.

Figure 5. The distribution of particle concentration npt normalized by the initial value np0 and 95% confidence intervals for (a) Rp=1.0 μm and (b) Rp=0.1 μm.

For Rp=1.0 μm ()), nonuniformity in particle concentration starts forming over the entire bubble cross-section upon the initiation of transient, and develops with time, with a maximum concentration near the bubble surface and a minimum at the vortex center. This nonuniformity, resulting from inertial migration of particles, apparently has an effect to raise the particle flow rate across the bubble surface, i.e. the particle removal rate. The Fuchs model does not consider such nonuniformity but simply assumes that the near-surface concentration to be the same as the volume average. This explains why the Fuchs model starts predicting slower particle removal than the simulation soon after the initiation of transient ()).

For Rp=0.1 μm ()), the particle concentration starts indicating a steep drop toward the bubble surface (to be discussed in detail later in the paragraph below the Equation (17). Except for this near-surface change, nonuniformity in concentration forms over the entire bubble cross-section in a similar fashion to that for Rp=1.0 μm. The nonuniformity however develops more slowly because of a slower particle migration. The concentration distribution shown in ) looks similar to the results of an Eulerian one-dimensional calculation by Ghiaasiaan and Yao [Citation22].

For discussing the particle removal rate, dNp/dt, or d1ε/dt, quantitatively in relation to removal mechanisms, it is convenient to use the Sherwood number, or nondimensionalized mass transfer coefficient, Sh=2Rbk/DB, where k is the mass transfer coefficient. shows Sh for Rp=0.1 μm as a function of the non-dimensional time, or Fourier number, tD=tDB/Rb2. The Sherwood number was calculated from the simulation results by using the relation [Citation24]

(16) Sh=231εdεdtD.(16)

Figure 6. Sherwood number (Sh) as a function of non-dimensional time tD=tDB/Rb2 for the Case 1 and Case 2 simulations, Fuchs model predictions for Case 2 and Case 3, and Mills and Hoseyni [Citation20] curve, for particle radius Rp=0.1 μm.

Figure 6. Sherwood number (Sh) as a function of non-dimensional time tD=tDB/Rb2 for the Case 1 and Case 2 simulations, Fuchs model predictions for Case 2 and Case 3, and Mills and Hoseyni [Citation20] curve, for particle radius Rp=0.1 μm.

Since ε obtained from simulations indicate large fluctuations with time, the time differentiation was applied to ε fitted by a smooth function. also shows the Fuchs models translated into Sherwood numbers:

(17a) Sh=23Rb2DBαi=3PeSt  for inertia,(17a)
(17b) Sh=23Rb2DBαB=1.2Pe0.5  for Brownan diffusion.(17b)

At the beginning of the transient in , two curves reflecting Brownian diffusion, i.e. Case 1 simulation and Mills and Hoseyni [Citation20] curve, show as high Sh as 103, followed by steep decreases with time. This is characteristic to unsteady advection-diffusion (Colombet et al. [Citation38] and Juncu [Citation39]). The initial removal rate is predicted to be high since the transient starts with a uniform particle concentration in the bubble where the bubble surface concentration is forced to be zero at t=0. (Despite high Péclet number Pe=UtRb/DB105, the Sherwood numbers obtained from Case 1 simulation are only about twice as large as the analytical solution for pure diffusion (Re=0 and Pe=0) [Citation38]). The removal rate decreases quickly as the near-surface concentration drops, as has been seen in ), to form a diffusion layer (a vacancy layer with low particle concentration) whose thickness increases with time. While the diffusion-alone model (Mills and Hoseyni) predicts a continuous, exponential decay of Sh, Case 1 simulation starts deviating from that at tD104 and finally converges with Case 2 (inertia-alone) simulation result. Both simulations then indicate a gradual increase in Sh, reflecting the near-surface peaking of concentration shown in ). Fuchs’s inertial model fails to capture this increase in Sh, since the model does not consider nonuniformity in concentration.

Fuchs’s diffusion model predicts time-independent Sh since it does not take into account of the development of the diffusion layer with time. The model is based on the penetration theory [Citation19] which considers a steady diffusion into (or from) a surface flow on a large fluid volume (bulk) unaffected by the surface diffusion. In other words, the theory assumes the surface diffusion layer to be much thinner than the fluid bulk dimension. This basic assumption provides a time independent Sh. It is, therefore, inadequate to apply this theory to the bubble internal flow, a recalculating flow in a small fluid volume (without a bulk), where the near-surface concentration gradient can vary significantly with time, directly affecting the surface transfer coefficient. The inadequacy of Fuchs’s diffusion model has an actual influence as an overprediction of the overall particle removal rate for smaller particle sizes, as shown in ) for Rp=0.1 μm, for which Fuchs’s model (mis)calculates the Sherwood number for diffusion to be about five times larger than that for inertial migration as shown in . The ratio of the two Sherwood numbers, given by the inverse of EquationEquation (14b), decreases with Rp and hence the influence of the diffusion model disappears for larger particle sizes for which 2.5Pe0.5St is of the order of unity or more.

We have observed for Rp=0.1 μm a transition from diffusion- to inertia-driven particle removal. Although the two mechanisms coexist through the end of transient, their relative importance changes with time. During the initial transient before tD104 for Rp=0.1 μm, the particles near the bubble surface are removed by advection-diffusion (i.e. Brownian motions of particles advected by fluid flow) so quickly that the inertial migration is unable to make up the loss in the near-surface region. The resultant self-decay of diffusional removal leads to transition from this diffusion-driven phase, hereafter called Phase I, to the next phase, Phase II, where the removal rate is determined by inertia alone. The advection-diffusion plays only a passive, secondary role. Although the duration of Phase I is very short for Rp=0.1 μm, it will increase with decreasing Rp. The diffusion layer is clearly visible in the concentration map for Rp=0.1 μm ()) but indiscernible for Rp=1.0 μm ()). This can be explained by one-dimensional consideration of particle transport that predicts the layer thickness to be DB/vt where vt is the inertial terminal velocity perpendicular to the stream surfaces that varies approximately as Rp2. Although the diffusion layer continues to exist, it does not add any resistance or driving force for particle removal. Meanwhile, as the near-surface-peaked concentration profile develops with time, as shown in ) for Rp=1.0 μm, the concentration gradient in the bubble internal region increases. Therefore, there is a possibility that an inward back diffusion takes place to such an extent that it starts limiting the outward, inertial migration, as suggested by Laker and Ghiaasiaan [Citation23]. This possible situation, Phase III, has not been observed in our simulations.

shows a normalized histogram of angular location θ/π at which particles reach the bubble surface. Regardless of particle size, particles tend to be removed near the bubble equator θ/π=0.5, where inertial removal is prompted by a high surface velocity and Brownian removal is also accelerated by the increase in concentration gradient caused by dense streamlines. For Rp=0.1 μm, the histogram is approximately symmetrical with respect to the bubble equator, reflecting the sinusoidal variation of surface velocity U with θ given by EquationEquations (1) and (Equation3). The histogram for Rp=1.0 μm, on the other hand, is skewed to show that more particles are removed in the northern hemisphere of the bubble than in the southern hemisphere, suggesting the influence of streamwise (θ-wise) decay of near-surface particle concentration. Kronig and Brink [Citation37] proposed, for advection-diffusion extraction of a substance from a falling droplet, that the concentration of dissolved substance could be taken to be uniform on a given streamline if the characteristic time of diffusion is much larger than the time of circulation along the streamline. This criterion, written as Pe1, is well satisfied in the present calculations where Pe=UtRb/DB105. A more restrictive criterion appropriate for the conditions of present calculations will be that the characteristic time of inertial removal Rb2/τpUt2 must be much larger than the characteristic time of circulation. This is written as St=τpUt/Rb1. Since the Stokes number measures the tendency of a particle to inertially deviate from a curved streamline of viscous fluid, in the present problem it measures the progression tendency of inertia-driven particle migration and removal. In fact, the particle removal velocity normal to the bubble inner surface is approximately given by StUU/Ut, as assumed by Fuchs in deriving his inertial removal coefficient αi (EquationEquation (13a)). The comparison in between two cases, correspond to St=8.2×103 for Rp=1.0 μm and St=1.4×104 for Rp=0.1 μm, suggests that, as the particle removal velocity increases with St, the streamwise decline of near-surface concentration becomes significant enough to cause an appreciable shift in particle flux distribution.

Figure 7. Normalized histograms of angular location θ/π at which particles reach the bubble surface, for particle radii Rp=1.0 and 0.1 μm.

Figure 7. Normalized histograms of angular location θ/π at which particles reach the bubble surface, for particle radii Rp=1.0 and 0.1 μm.

5. Conclusion

Brownian and inertial removals of aerosol particles from a bubble rising in water have been studied by performing Lagrangian particle tracking simulations. The results have been presented and discussed with a focus on the development of particle concentration profile and its influence on removal rates in conjunction with the interference between the removal mechanisms. Comparison has been made with predictions by Fuchs’s model (1964) used widely although it is based on simplifications including lumped parameter approach for concentration and linear superposition assumption for the influence of different removal mechanisms.

The results from simulations conducted for particles with radii ranging between 0.1 and 1.0 μm, moving in a 0.5 mm radius bubble rising steadily at 0.159 m/s can be summarized as follows. Brownian diffusion plays a definitive role only at the very beginning of transient. The rest of the transient is dominated by inertial migration. Fuchs’s model for diffusional removal is unphysical in that it assumes the process to be quasi-steady and unaffected by inertial migration while the real process is highly transient and is quickly integrated into inertial removal. Such fundamental difference results in a significant overprediction of the total removal rate for small Rp. The inertial migration produces a near-surface-peaked concentration profile in the bubble that results in faster particle removal than predicted by Fuchs’s model based on a lumped parameter (ideally mixed bubble) approach. It appears that the particle concentration tends to be nonuniform also in the streamwise direction with an increase in particle Stokes number, affecting the distribution of the removal rate on the bubble surface. In summary, the present Lagrangian particle tracking simulations have provided useful information for improved modeling of macroscopic phenomena controlling the particle removal from a rising bubble. The main conclusions summarized above are qualitative and concern the inadequacy of fundamental assumptions in the Fuchs model. Naturally, they are not sensitive to parameter values, but are based on a steady-flow assumption that may provide a reasonable approximation to reality for bubble radii smaller than a few millimeters. The dependence of quantitative results (1ε vs. time) on particle and bubble parameters represents mixed influence of different mechanisms. We plan to address this problem through dimensional analyses.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • López-Jiménez J, Herranz LE, Escudero MJ, et al. Pool scrubbing. Spain: Centro de Investigaciones Energeticas; 1996. (CIEMAT–805).
  • Berzal ME, Crespo MJM, Swiderska-Kowalczyk M, et al. State of the art review on fission product aerosol pool scrubbing under severe accident conditions. Poland: European Commission; 1994. (INIS-PL–0001).
  • Berna C, Escrivá A, Munõz-Cobo JL, et al. Enhancement of the SPARC90 code to pool scrubbing events under jet injection regime. Nucl Eng Des. 2016;300:563–577.
  • Zhao YF, Irons GA. The breakup of bubbles into jets during submerged gas injection. Metall Trans B. 1990;21:997–1003.
  • Cherrier G, Belut E, Gerardin F, et al. Aerosol particles scavenging by a droplet: microphysical modeling in the Greenfield gap. Atmos Environ. 2017;166:519–530.
  • Dehbi A, Suckow D, Guentay S. Aerosol retention in low-subcooling pools under realistic accident conditions. Nucl Eng Des. 2001;203:229–241.
  • Kanai T, Furuya M, Arai T, et al. Development of an aerosol decontamination factor evaluation method using an aerosol spectrometer. Nucl Eng Des. 2016;303:58–67.
  • Pich J, Schütz W. On the theory of particle deposition in rising gas bubbles: the absorption minimum. J Aerosol Sci. 1991;22:267–272.
  • Park SH, Park C, Lee JY, et al. A general solution for the removal of polydisperse aerosol particles and application to pool scrubbing. Part Sci Technol. 2018;36:552–559.
  • Li Y, Sun Z, Gu H, et al. Deposition characteristic of micro-nano soluble aerosol under bubble scrubbing condition. Ann Nucl Energy [Internet]. 2019;133:881–888.
  • Kim YH, Kam DH, Yoon J, et al. The importance of representative aerosol diameter and bubble size distribution in pool scrubbing. Ann Nucl Energy. 2020;147:107712.
  • Lee Y, Cho YJ, Ryu I. Preliminary analyses on decontamination factors during pool scrubbing with bubble size distributions obtained from EPRI experiments. Nucl Eng Technol. 2021;53:509–521.
  • Becker KF, Anderson MH. Experimental study of SRT scrubbing model in water coolant pool. Nucl Eng Des. 2021;377:111130.
  • Powers DA, Brockmann JE, Shiver AW. VANESA : a mechanistic model of radionuclide release and aerosol generation during core debris. Albuquerque, NM: Sandia National Labs; 1986. (NUREG/CR-4308; SAND-85-1370).
  • Owczarski PC, Burk KW. SPARC-90: a code for calculating fission product capture in suppression pools. United States: Nuclear Regulatory Commission; 1991. (NUREG/CR–5765).
  • Ramsdale SA, Güntay S, Friederichs HG. BUSCA-JUN91 reference manual. Switzerland: Paul Scherrer Institute(PSI); 1995. (PSI–95-05).
  • Wassel AT, Mills AF, Bugby DC. Analysis of radionuclide retention in water pools. Nucl Eng Des. 1985;90:87–104.
  • Fuchs NA. The mechanics of aerosols. revised and enlarged ed. (translated from the Russian by R.E. Daisley and Marina Fuchs, translation edited by C.N. Davies). New York (NY): Pergamon Press; 1964.
  • Higbie R. The rate of absorption of a pure gas into a still liquid during short periods of exposure. Trans AIChE. 1935;31:365–389.
  • Mills AF, Hoseyni MS. Diffusive deposition of aerosols in a rising bubble. Aerosol Sci Technol. 1988;8:103–105.
  • Ghiaasiaan SM, Yao GF. Diffusive and convective deposition of aerosols in rising spherical bubbles with internal circulation. Int J Multiph Flow. 1995;21:907–918.
  • Ghiaasiaan SM, Yao GF. A theoretical model for deposition of aerosols in rising spherical bubbles due to diffusion, convection, and inertia. Aerosol Sci Technol. 1997;26:141–153.
  • Laker TS, Ghiaasiaan SM. Monte-carlo simulation of aerosol transport in rising spherical bubbles with internal circulation. J Aerosol Sci. 2004;35:473–488.
  • Clift R, Grace JR, Weber ME. Bubbles, drops, and particles. NewYork (NY): Academic Press; 1978.
  • Hill MJM. On a spherical vortex. Philos Trans R Soc London. 1894;185:213–245.
  • Peebles FN, Garber HJ. Studies on the motion of gas bubbles in liquids. Chem Eng Prog. 1953;49:88–97.
  • Holland FA, Bragg R. Fluid flow for chemical engineers. Great Britain: Elsevier; 1973.
  • Akbar MK, Ghiaasiaan SM. Monte Carlo simulation of aerosol transport in rising gas bubbles undergoing shape deformation. J Aerosol Sci. 2006;37:735–749.
  • Hinds WC. Aerosol technology: properties, behavior, and measurement of airborne particles. New York (NY): John Wiley & Sons; 1999.
  • Crowe GT. Review-numerical models for dilute gas-particle flows. J Fluids Eng Trans ASME. 1982;104:297–303.
  • Maxey MR, Riley JJ. Equation of motion for a small rigid sphere in a nonuniform flow. Phys Fluids. 1983;26:883.
  • Squires K. Point-particle methods for disperse flows. Prosperetti A, Tryggvason G, editors. Cambridge: Computational Methods for Multiphase Flow; 2007.
  • Davies CN. Definitive equations for the fluid resistance of spheres. Proc Phys Soc. 1945;57:259–270.
  • Li A, Ahmadi G. Dispersion and deposition of spherical particles from point sources in a turbulent channel flow. Aerosol Sci Technol. 1992;16:209–226.
  • Abouali O, Nikbakht A, Ahmadi G, et al. Three-dimensional simulation of Brownian motion of nano-particles in aerodynamic lenses. Aerosol Sci Technol. 2009;43:205–215.
  • Calderbank PH, Korchinski IJO. Circulation in liquid drops (A heat-transfer study). Chem Eng Sci. 1956;6:65–78.
  • Kronig R, Brink JC. On the theory of extraction from falling droplets. Appl Sci Res. 1951;2:142–154.
  • Colombet D, Legendre D, Cockx A, et al. Mass or heat transfer inside a spherical gas bubble at low to moderate Reynolds number. Int J Heat Mass Transf. 2013;67:1096–1105.
  • Juncu G. A numerical study of the unsteady heat/mass transfer inside a circulating sphere. Int J Heat Mass Transf. 2010;53:3006–3012.