10
Proceedings of the ECCOMAS Thematic International Conference on Simulation and Modeling of Biological Flows (SIMBIO 2011) September 21–23, 2011, VUB, Brussels, Belgium Numerical and Experimental Investigation of Ultrafine Particle Transport and Deposition in a Human Upper Airway Model F. Krause 1 , A. Wenk 2 , W. Möller 2 , S. Verbanck 3 , C. Lacor 4 , and W. G. Kreyling 2 1 Vrije Universiteit Brussel, Department of Mechanical Engineering, Fluid Dynamics and Thermodynamics Re- search Group 2 Helmholtz Zentrum München, German Research Center for Environmental Health, Institute of Lung Biology and Disease Pleinlaan 2, 1050 Brussel, Belgium, [email protected] Ingolstaedter Landstrasse 1, 85764 Neuherberg, Ger- many, {alexander.wenk, moeller, kreyling}@helmholtz- muenchen.de 3 University Hospital UZ Brussel, Respiratory Division 4 Vrije Universiteit Brussel, Department of Mechanical Engineering, Fluid Dynamics and Thermodynamics Re- search Group Laarbeeklaan 101, 1090 Brussel, Belgium, [email protected] Pleinlaan 2, 1050 Brussel, Belgium, [email protected] Abstract The majority of the published numerical studies on the deposition of ultrafine particles in the human upper airways lack the validation of the predicted flow field and particle deposition. Thus, this study aims at validating the numer- ically predicted laminar-turbulent flow field and subsequent deposition of ultrafine particles in a realisitic model of the human upper airways by means of experimental in vitro measurements. The simulated flow field is compared to existing particle image velocimetry measurements and the total deposition results are compared to experimental deposition measurements of ultrafine carbon particles in a cast of the same upper airway model. Monodisperse ultrafine carbon particles of different thermodynamic diameter ranging from d p = 14.5nm to d p = 52nm are generated by spark ignition and their deposition in a silicone cast of the upper airway model is measured for three physiologically relevant steady inhalation flow rates Q = 10, 20 and 30 l min . Computations of the laminar-turbulent flow field are carried out using an open-source steady-state solver for incompressible, turbulent flows in conjunction with the k-ω SST RANS turbulence model. For the simulation of the particle motion and the subsequent particle deposition, the Eulerian mass transport model is employed. Both simulations and experiments show that total particle deposition fraction increases with decreasing flow rate and decreasing particle diameter. Furthermore, simulations as well as experiments are in satisfactory quantitative agree- ment with previous numerical and experimental data. A direct comparison in the range of flow rate (Q> 10 l min ) and particle size (d p > 14.5nm) that are achieved experimentally, reveals that the simulated total deposition is approx- imately half the experimental value. However, this discrepancy needs to be considered in the light of the very low deposition efficiencies in this flow and size range (with all experimental deposition fractions less than 5%). Finally, the simulations also serve to identify local deposition hot-spots by contours of the so-called deposition en- hancement factor. Although the local deposition results cannot be validated quantitatively, they show that the deposi- tion pattern is more discrete as particle size decrease. This observation is in line with previous studies. Keywords: human upper airway model, ultrafine particle deposition, Reynolds-averaged Navier-Stokes equations, Eulerian mass transport model Introduction During the last number of decades, the inhalation ex- posure to airborne nano-sized particles has increased sig- nificantly because of its widespread use and production in diverse sectors in science, technology and medicine. The outstanding properties of nanomaterials set them apart from traditional materials. It is expected that they will reach a key position in the aforementioned areas thus re- quiring assessement and examination of their potential for adverse health effects [1-4]. In fact, the World Health Organization has already re- ported more than 2 million premature deaths associated SIMBIO 2011 1

Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

Proceedings of the ECCOMAS Thematic International Conference onSimulation and Modeling of Biological Flows (SIMBIO 2011)September 21–23, 2011, VUB, Brussels, Belgium

Numerical and Experimental Investigation of Ultrafine ParticleTransport and Deposition in a Human Upper Airway Model

F. Krause1, A. Wenk2, W. Möller 2, S. Verbanck3, C. Lacor4, and W. G. Kreyling2

1 Vrije Universiteit Brussel, Department of MechanicalEngineering, Fluid Dynamics and Thermodynamics Re-search Group

2 Helmholtz Zentrum München, German Research Centerfor Environmental Health, Institute of Lung Biology andDisease

Pleinlaan 2, 1050 Brussel, Belgium, [email protected] Ingolstaedter Landstrasse 1, 85764 Neuherberg, Ger-many, {alexander.wenk, moeller, kreyling}@helmholtz-muenchen.de

3 University Hospital UZ Brussel, Respiratory Division 4 Vrije Universiteit Brussel, Department of MechanicalEngineering, Fluid Dynamics and Thermodynamics Re-search Group

Laarbeeklaan 101, 1090 Brussel, Belgium,[email protected]

Pleinlaan 2, 1050 Brussel, Belgium, [email protected]

AbstractThe majority of the published numerical studies on the deposition of ultrafine particles in the human upper airwayslack the validation of the predicted flow field and particle deposition. Thus, this study aims at validating the numer-ically predicted laminar-turbulent flow field and subsequent deposition of ultrafine particles in a realisitic model ofthe human upper airways by means of experimental in vitro measurements. The simulated flow field is comparedto existing particle image velocimetry measurements and the total deposition results are compared to experimentaldeposition measurements of ultrafine carbon particles in a cast of the same upper airway model.Monodisperse ultrafine carbon particles of different thermodynamic diameter ranging fromdp = 14.5nm todp = 52nm are generated by spark ignition and their deposition in a silicone cast of the upper airway model ismeasured for three physiologically relevant steady inhalation flow ratesQ = 10, 20 and30 l

min. Computations of

the laminar-turbulent flow field are carried out using an open-source steady-state solver for incompressible, turbulentflows in conjunction with thek-ω SST RANS turbulence model. For the simulation of the particle motion and thesubsequent particle deposition, the Eulerian mass transport model is employed.Both simulations and experiments show that total particle deposition fraction increases with decreasing flow rate anddecreasing particle diameter. Furthermore, simulations as well as experiments are in satisfactory quantitative agree-ment with previous numerical and experimental data. A direct comparison in the range of flow rate (Q > 10 l

min) and

particle size (dp > 14.5nm) that are achieved experimentally, reveals that the simulated total deposition is approx-imately half the experimental value. However, this discrepancy needs to be considered in the light of the very lowdeposition efficiencies in this flow and size range (with all experimental deposition fractions less than5%).Finally, the simulations also serve to identify local deposition hot-spots by contours of the so-called deposition en-hancement factor. Although the local deposition results cannot be validated quantitatively, they show that the deposi-tion pattern is more discrete as particle size decrease. This observation is in line with previous studies.Keywords: human upper airway model, ultrafine particle deposition, Reynolds-averaged Navier-Stokes equations,Eulerian mass transport model

IntroductionDuring the last number of decades, the inhalation ex-

posure to airborne nano-sized particles has increased sig-nificantly because of its widespread use and productionin diverse sectors in science, technology and medicine.The outstanding properties of nanomaterials set them apart

from traditional materials. It is expected that they willreach a key position in the aforementioned areas thus re-quiring assessement and examination of their potential foradverse health effects [1-4].

In fact, the World Health Organization has already re-ported more than 2 million premature deaths associated

SIMBIO 2011 1

Page 2: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

with particulate matter air pollution each year [5]. Conse-quently, airborne nanoparticles, which are commonly cat-egorized as fine particles with a thermodynamic diameter0.1µm < dp < 1µm and ultrafine particles with a thermo-dynamic diameterdp < 0.1µm, have been the subject ofseveral recent toxicological and epidemiological studies.These have reported an association between the increasedexposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions for treatment of respiratory diseases to increased riskof lung cancer [6, 7].

Apart from several effects such as the toxicity of theparticle and the particle processing and translocation tosecondary organs, the particle deposition and clearancehave significant health implications [8]. The deposition istherefore the main subject of this study. The key factorsinfluencing the total nano-sized particle deposition and thelocation of deposition hot-spots are the particle size, theinhalation and exhalation flow rate and the morphology ofthe human lung.

Reports of in vitro and in vivo ultrafine particle deposi-tion measurements in the human lung are still rare. This ismainly due to the difficulties in the generation of ultrafineparticles, the detection of their number concentration andthe visualization of their distribution within the lung.

On the other hand, Computational Fluid Dynamics(CFD) has already shown its potential to complement ex-perimental studies by predicting the air flow as well as themotion and deposition of nanoparticles in the human air-ways [9]. Applications range from simulations in simpli-fied and realistic upper airway models [10-12] and sim-plified models of the tracheo-bronchial tree with multi-ple bifurcations [13, 14] to complex nasal-oral-tracheo-bronchial models [15]. However, the major weakness ofthe published numerical studies is the lack of a completevalidation of the results by comparing them to experimen-tal flow field and deposition measurements in the samemodel.

For this reason, the goal of the present study is to sim-ulate and subsequently validate numerical results of thelaminar-turbulent flow field and the total and local deposi-tion of ultrafine particles in a human upper airway model,which shows all relevant morphological features of a real-istic computer-tomography based model.

To numerically study the macroscopic motion of ultra-fine particles, it is common practice to use the so-called Eu-lerian mass transport model. As the nanoparticle volumefraction of ambient particulate matter is very small, thismodel assumes that the particle state only depends on thestate of the flow field and that their momentum exchangehas an negligible effect on the flow. Consequently, theparticles are treated as a one-way-coupled passive scalar,which also agrees with the widely accepted classificationmap for particle-laden turbulent flows proposed in [16].The major advantage of this method over discrete methods,where a momentum equation has to be solved for every sin-gle particle, is its’ applicability to a very large number of

Figure 1 Isometric view of the upper airway model (UAM)geometry.

particles at very low computational cost.

Experimental MethodThe conducted experiments aim to generate monodis-

perse ultrafine particles of different thermodynamic diam-eter and to measure their deposition in a cast of the humanupper airway model under three different physiologicallyrelevant inspiratory flow rates. The 3D geometry of theupper airway model, which is depicted in Figure 1, andthe corresponding silicone cast are retrieved from previouswork [17, 18].

Nano-sized Carbon particles are generated by sparkignition using a Carbon electrode in an Argon gas stream(GFG 100, Palas, Karlsruhe, Germany). During igni-tion, small amounts of carbon evaporate on the surface ofthe electrode and subsequently condense to form primarychain-like nanoparticles. A transmission elelectron mis-croscopy (TEM) image of sample nanoparticles after theirgeneration is given in Figure 2. Directly behind the outletof the spark generation chamber, the highly charged parti-cles are quasi-neutralized to Boltzmann equilibrium by us-ing a Kr-85 source. In order to avoid further coagulation,the aerosol is diluted with air. A similar procedure hadpreviously been carried out to generate Titanium Dioxidenanoparticles [19].

After the aerosol generation, the high initial particleconcentration is significantly reduced by inserting a filterdownstream of the spark generation chamber.

Monodisperse particle fractions with a geometric stan-dard deviationgsd < 1.15 are then extracted using aDifferential Mobility Analyser (DMA, Model 3071A, TSIInc., Aachen Germany) and their charge level is againreduced using a Am-241 source. The concentration ofthe extracted monodisperse aerosol is always smaller than104 p

cm3 .In order to change the size distribution and the ther-

modynamic particle diameter of the monodisperse particle

SIMBIO 2011 2

Page 3: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

Figure 2 TEM image of sample chain-like nanoparticle agger-agtes after generation [19]

fraction respectively, the ignition frequency and the elec-trical deflection field (voltage) is varied. A higher ignitionfrequency results in a higher particle concentration andsubsequent coagulation, so that larger particles are formed.Accordingly, higher deflection voltages of the DMA al-low larger particles to pass the device. Overall, particlesof four different thermodynamic diametersdp = 14.5, 20,32 and52nm are generated. The particle size distributionof the monodisperse aerosol is measured with a scanningmobility particle sizing system (SMPS, DMA model 3071and CPC model 3022, both TSI Inc., Aachen, Germany) asshown in Figure 3.

Three different physiologically relevant steady volumeflow rates ofQ1 = 10 l

min,Q2 = 20 l

minandQ3 = 30 l

min

are generated using a pump which is connected to the out-let downstream of the cast of the upper airway model.

Particle number concentrations are calculated at twodifferent measurement points, directly in front of the inletof the upper airway cast (point #1) and directly behind theoutlet of the cast (point #2) using a Concentration ParticleCounter (CPC, Model 3010 TSI Inc., Aachen, Germany).The deposition fraction is calculated based on the detectedparticle concentrationC as

DF =C#1 − C#2

C#1, (1)

where the subscripts denote the corresponding measure-ment point, either in front of the cast inlet or behind thecast outlet. A sketch of the experimental setup is shown inFigure 3.

Numerical MethodThe governing equations describing incompressible

fluid flow are the equation of continuity (1) and the Navier-Stokes equations (2). Using Einstein’s summation conven-tion for the indexing of the spatial direction(x, y, z) andthe velocity vector components(u, v, w), they can be writ-ten as

∂ui

∂xi

= 0, (2)

uj

∂ui

∂xj

= −1

ρ

∂p

∂xi

+∂τi,j∂xj

, (3)

Figure 3 Schematic of the nano-sized carbon particle genera-tion and particle size and concentration measurement.

with τi,j denoting the stress tensor. Performing Reynoldsaveraging on the governing equations and followingReynolds conditions yields the Reynolds-averaged Navier-Stokes (RANS) equations

∂ui

∂xi

= 0 , (4)

uj

∂ui

∂xj

= −1

ρ

∂p

∂xi

+∂

∂xj

(

τ i,j − u′

i u′

j

)

, (5)

with −u′

i u′

j denoting the specific Reynolds stress ten-sor. Following the Boussinesq approximation, the specificReynolds stress tensor is modeled as

−u′

i u′

j = νt

(

∂ui

∂xj

+∂uj

∂xi

)

− 2

3kδi,j , (6)

with νt denoting the turbulent kinematic viscosity.The turbulent kinematic viscosityνt is calculated us-

ing the k-ω SST RANS model [21]. It is based on thesolution of a transport equations for the turbulent kineticenergyk and the specific dissipation rateω. The key ideawhich sets the model apart from the traditionalk-ω RANSmodel is a blend to switch between thek-ω RANS modelin the near wall region and thek-ε RANS model in in thecore region of the flow. More details on the definition ofthe blending function as well as of model constants can befound in [21]. It should be noted, that the model is used inits standard high-Reynolds version, thus leaving out low-Reynolds correction terms.

Since the particles are modeled as a one-way-coupledpassive scalar, their motion is described by solving anadvection-diffusion equation (7) for the particle concen-

SIMBIO 2011 3

Page 4: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

trationC. It can be written as

uj

∂C

∂xj

=∂

∂xj

(

D∂C

∂xj

)

, (7)

with D denoting the diffusion coefficient of the particles.It is calculated as proposed in [22] by

D =kbTCs

3πµdp, (8)

with kb denoting the Boltzmann constant,T the tempera-ture,µ the dynamic viscosity,dp the thermodynamic parti-cle diameter andCs the Cunningham slip correction factorcalculated as proposed in [23] by

Cs = 1 +Kn(

1.142 + 0.558 e−0.999

Kn

)

. (9)

Here, Kn denotes the Knudsen number. PerformingReynolds averaging and following Reynolds conditions onequation (7) yields

uj

∂C

∂xj

= D∂C

∂xj

− ∂u′

iC′

∂xj

. (10)

The closure of the scalar-flux term−u′

iC′ is accomplished

by using a gradient-diffusion hypothesis, so that

−u′

iC′ = Dt

∂C

∂xj

. (11)

Here,Dt denotes the turbulent mass diffusion coefficient,which is modelled as

Dt =νtSct

. (12)

The turbulent Schmidt numberSct is treated as a globalparameter and set toSct = 0.9 [11, 24].

At the wall (subscriptw), no-slip boundary conditionsare applied for the velocity as well as for the turbulent ki-netic energy, so thatuw,i = 0 andkw = 0. Furthermore,the wall is assumed to be perfectly absorbing the ultrafineparticle, so thatCw = 0. The specific dissipation rate inthe viscous sublayer is calculated according to [21] as

ωw,j =6 ν

0.075 y20,j, (13)

wherey0,j denotes the distance between the center of awall facefj and the center of the closest cellcj normal tofj. For the pressurepw, zero-gradient boundary conditionare applied at the wall.

The inlet profiles of all flow quantities are assumed tobe block profiles. The inlet velocityuin is set according tothe considered inlet mass flow rate, while the initial inletvalue of the turbulent kinetic energyk is calculated as

kin =3

2(Iin uin)

2 , (14)

whereIin denotes the core turbulence intensity of a fullydeveloped turbulent pipe flow. The intensity at the inlet iscalculated as

Iin = 0.16Re(−0.125)in , (15)

with Rein denoting the Reynolds number based on the in-let diameter and the inlet velocity. The initial value of thespecific dissipation rate is calculated as

ωin = C(−0.25)µ

√k

lt, (16)

whereCµ is a constant set toCµ = 0.9 andlt the turbu-lence length scale in a fully developed turbulent pipe flow.It is calculated as

lt = 0.07 din . (17)

The amount of particles depositing on the wall is quan-tified by the deposition fraction (DF ). In the frameworkof the Eulerian mass transport model, it is calculated as

DF =mw

min

, (18)

where the mass flow rate at the inletmin and the mass flowrate to the wallmw are computed as

min = ρmuin2πrin , (19)

mw =

Nw∑

j=1

(

−ρmAj

(

D +νtSct

)

∂C

∂n

)

. (20)

Here,ρm denotes the air-nanoparticle mixture density de-fined as

ρm = αpρp + (1− αf ) ρf , (21)

where the subscriptsp andf stand for the particle phaseand the fluid phase respectively andα for the phase volumefraction. Further,Nw denotes the number of wall faces andAj the surface area offj . The local particle deposition isquantified by the deposition enhancement factor (DEF ) asproposed in [25] by

DEFj =mw,j/Aj

mw/A, (22)

whereA is the total wall surface area. The depositionenhancement factor (22) is commonly employed for theidentification of local deposition hot-spots [9-15, 26].

The governing equations are discretized andsolved using the open-source CFD software packageOpenFOAM

®[27, 28]. It uses the finite-volume method on

three-dimensional arbitrarily unstructured meshes. Detailson the disctretisation procedure of the gradient, divergenceand Laplace operator can be found in [29].

The laminar-turbulent flow field is computed usingthe already existing solversimpleFoam, a steady-statesolver for incompressible, turbulent flows, which uses

SIMBIO 2011 4

Page 5: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

the SIMPLE (Semi-Implicit Method for Pressure-LinkedEquations) algorithm [30] for the pressure-velocity cou-pling.

The particle motion and equation (10), (11) and (12) inparticular are computed using theturbMassDiffusionFoamsolver. It is an extension of the already existing solverscalarTransportFoam. It is extended by equation (12), sothat the turbMassDiffusionFoamsolver accounts for theeffect of turbulence on the particle motion. Similar tothe simpleFoamsolver, it is a steady-state solver, whichuses under-relaxation of the flow variables to iterate thesolution until convergence.

The procedure to obtain the flow field and the resultingparticle motion and deposition for a specific inlet volumeflow rate and particle diameter is as follows:

1. UsesimpleFoamto solve the flow equations until con-vergence and write out the velocity field and the tur-bulent kinematic viscosity field.

2. Use these fields as the initial field condition in theturbMassDiffusionFoamsolver. If different compu-tational grids are used, the fields are mapped and in-terpolated onto the new grid.

3. Solve the particle equation until convergence usingthe turbMassDiffusionFoamsolver and write out theparticle concentration field.

4. Calculate the total deposition fractionDF and thedeposition enhancementDEF factor using the post-processing utilitiesdepositionFractionanddepositio-nEnhancementFactor, which are developed for thispurpose

According to the conclusions of recently performed meshstyle studies [31, 32], the geometry is subdivided intomultiple blocks and a purely hexahedral grid is generatedusing the commercial meshing software ANSYS

®ICEM

CFDTM

[33]. As can be seen in Figure 4a and 5b, blockswith a circle-like cross-section are meshed with an O-grid.

In order to accurately capture the near-wall flow fieldand the expected flow features, such as the forced flowseperation at the Epiglottis, the first grid point normal tothe wall is systematically placed at a dimensionless walldistancey+ ≤ 1. In fact, the specific dissipation rateboundary conditions as proposed in [21] is only valid inthe viscous sub-layer. By using edge grading, the near-wall region is resolved while keeping the overall cell countas low as possible.

According to the experiments, simulations are per-formed for three different inlet volume flow ratesQ1 =10 l

min, Q2 = 20 l

minandQ3 = 30 l

min. They correspond

to the inlet Reynolds numbersRe1 = 615, Re2 = 1230andRe3 = 2468 based on the inlet velocityuin and the in-let diameterdin. As the flow with the highest inlet flow rateand Reynolds number respectively requires the finest near-wall grid discretization, the grid yielding grid-convergedresults for the case withQ3 = 30 l

minandRe3 = 2468

(a)

(b)

Figure 4 Illustration of the O-grid structure at circle-lik ecross-section of the inlet (a) as well as the Larynx and Tra-chea (b).

will be used for all fluid flow simulations.Additionally to the four generated particle diameters,

the deposition of even smaller particles with thermody-namic diameters ofdp = 1, 2, 5 and10nm are studiedin order to investigate an effect of the particle diameter onthe deposition fraction.

Numerical Results and DiscussionNumerical results of velocity and turbulent kinetic en-

ergy for the flow with an inlet volume flow rate ofQ3 =30 l

minare compared to available particle image velocime-

try (PIV) and large-eddy simulation (LES) results [34] infour different section of the same upper airway model:5mm above the Epiglottis (section E) and 1, 2 and 3 tra-cheal diameter downstream of the Larynx (section T1, T2and T3). The approximate location of the four sections aredepicted in Figure 7a. The profiles of the velocity magni-tude and the turbulent kinetic energy measured in the foursections are displayed in Figure 5 and Figure 6.

Comparing all velocity profiles in section E againsteach other, it is apparent that neither the reference LES re-sults nor the presentk-ω SST RANS results are able toaccurately capture the complex velocity profile predictedby the reference PIV measurements. Nevertheless, themain flow features are captured by both the reference LESand the present RANS results. Furthermore, the simula-tions correctly predict the typical velocity profiles down-stream of the Larynx with the distinct velocity peak at0.5 ≤ x/D ≤ 1 refering to the high-velocity laryngeal jet.

SIMBIO 2011 5

Page 6: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

0 0.5 10

0.5

1

1.5

2

x / D

u m /

u in

(a)

0 0.5 10

0.5

1

1.5

2

x / Du m

/ u in

(b)

0 0.5 10

0.5

1

1.5

2

x / D

u m /

u in

(c)

0 0.5 10

0.5

1

1.5

2

x / D

u m /

u in

(d)

Figure 5 Profiles of two-component (ux anduz) velocity mag-nitude um normalized by the inlet velocityuin in sections E(a), T1 (b), T2 (c) and T3 (d). ◦ : experimental results [34],−− : LES results [34], − : present RANS results using thek-ω SST RANS model.

0 0.5 10

0.05

0.1

0.15

0.2

0.25

0.3

x / D

k / u

in2

(a)

0 0.5 10

0.05

0.1

0.15

0.2

0.25

0.3

x / D

k / u

in2

(b)

0 0.5 10

0.05

0.1

0.15

0.2

0.25

0.3

x / D

k / u

in2

(c)

0 0.5 10

0.05

0.1

0.15

0.2

0.25

0.3

x / D

k / u

in2

(d)

Figure 6 Profiles of turbulent kinetic energyk normalized bythe square of the inlet velocityuin in sections E (a), T1 (b),T2 (c) and T3 (d). ◦ : experimental results [34],−− : LESresults [34], − : present RANS results using thek-ω SSTRANS model.

(a) (b)

Figure 7 Two component (ux anduz) velocity magnitudeum

normalized by the inlet velocity (a) and turbulent kinetic en-ergyk normalized by the square of the inlet velocity (b) in thecentral saggital plane.

However, it is observed in the corresponding section T1,T2 and T3 that the RANS simulation systematically over-estimate the peak velocity, while the reference LES resultsshow a good agreement with the experimental measure-ments in these particular regions. The overprediction ofthe laryngeal jet’s peak velocity might be attributed to themissing low-Reynolds correction terms of the employedRANS turbulence model. These correction terms accountfor the viscous damping in the near-wall region. Consid-ering the overall performance of the used RANS model,it can be stated that the predicted results are in accept-able agreement with the reference LES and experimentalresults.

The profiles of turbulent kinetic energy presented inFigure 6 support the oberservations made for the veloc-ity profiles. Although the location of the peak turbulentkinetic energy is predicted correctly in the three sectionsdownstream of the Larynx, its value is significantly over-estimated. This again might be attributed to the missinglow-Reynolds correction terms. As for the the velocity pro-files, the LES results are in quite good agreement with thecorresponding PIV measurements.

Contours of the two component velocity magnitude andthe turbulent kinetic energy in a central saggital plane areshown in Figure 7a and 7b. Both contour plots show therelevant flow structures, i.e. the complex flow structure up-stream of the Epiglottis, the laryngeal jet and the recircu-lation zone downstream of the Larynx constriction. Com-parison with contour plots produced by PIV measurements[17] shows a quite good agreement of the flow field and theturbulent kinetic energy.

The suitability of the computational grid is investi-gated by calculating the dimensionless wall distancey+jfor each wall face. According to [21], the proposed bound-

SIMBIO 2011 6

Page 7: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

(a) (b)

Figure 8 Contour plot of dimensionless wall distancey+ in the range 1 ≤ y+ ≤ 2 for an inlet flow rate ofQ3 = 30 l

min.

ary conditions forω (13) is valid, ify+ < 2. Further, for awall-resolving grid, the value ofy+ of the first grid-pointnormal to the wall should bey+ ≈ 1. Figure 8 shows thecontours ofy+ in the range1 < y+ < 2. It is apparent,that the dimensionless wall distance is sufficiently small atthe entire wall surface.

The total particle deposition fractionDF is comparedto the present experimental results for all flow rates andparticles with a thermodynamic diameter ofdp = 14.5, 20,32 and52nm. Furthermore, comparison is made with nu-merical deposition results in two different simplified mod-els of the human upper airways [10,11] and experimentalin Vitro results [35, 36]. The percentage deposition frac-tion as a function of the particle diameter forQ1 = 10 l

min,

Q2 = 20 lmin

andQ3 = 30 lmin

is presented in Figure 9.Additionaly the numerical as well as the experimental totaldeposition fraction results of the present study are summa-rized in Table 2.

The results displayed in Figure 9 clearly show twotrends, which are already reported [9]. First, the depositionfraction is decreasing with an increasing particle diame-ter and second, the deposition fraction is decreasing withan increasing inhalation volume fow rate. In fact, againstthe background that the diffusion coefficient is decreasingwith an increasing particle diameter and that the particleresidence time is decreasing with an increasing flow rate,both trends are expected. The influence of the morphologyof the upper airway model on the deposition fraction be-comes visible when comparing the deposition fraction fordp = 1nm up todp = 10nm reported in [10, 11]. An ab-solute difference of up to20% is observed fordp = 1nm.Furthermore, it can already be seen, that the amount of de-posited ultrafine particles in the upper airways is very smallfor dp > 10nm. Considering, that the deposition data aregenerated using not only different geometries but also dif-ferent numerical as well as experimental methods, they are

Table 1 Percentage deposition fractionDF [%] obtained byexperiments and numerical simulation in the same humanupper airway model as a function of the particle diame-ter dp and percantage deviation between the present exper-imental and numerical total deposition fraction results forQ1 = 10 l

min(a),Q2 = 20 l

min(b) andQ3 = 30 l

min(c).

(a)

dp[nm] 14.5 20 32 52DFexp.[%] 4.59 3.00 3.17 2.12DFnum.[%] 2.83 1.90 1.07 0.65ε[%] 38.34 36.67 66.25 69.34

(b)

dp[nm] 14.5 20 32 52DFexp.[%] 4.11 2.07 1.56 1.31DFnum.[%] 2.07 1.38 0.78 0.48ε[%] 49.64 33.33 50.00 63.36

(c)

dp[nm] 14.5 20 32 52DFexp.[%] 3.13 1.30 1.05 0.77DFnum.[%] 1.70 1.12 0.64 0.40ε[%] 45.67 13.85 39.05 48.05

in good quantitative agreement.Comparing the achieved total deposition fraction

data, it is apparent that there is still a significant dis-crepancy between numerical simulations and experiments.The percentage deviation ranges fromε[%] = 13.85 fordp = 20nm andQ3 = 30 l

minup to ε[%] = 69.34 for

dp = 52nm andQ1 = 10 lmin

. However, the importanceof a higher accuracy is questionable, as the overall amountof deposited ultrafine particles is very small.

Contour plots of the deposition enhancement factor(DEF) are presented in Figure 10 and 11 for the depo-sition of dp = 14.5nm and dp = 52nm particles atQ1 = 10 l

min.

Although the overall deposition pattern is quite sim-ilar for both particle diameters, the DEF contours fordp = 10nm shows more discrete deposition hot-spots withDEF > 4. This can be clearly seen at the undersite of thetongue and the pharyngeal wall, Figure 10, as well as at theepiglottal stop, Figure 11.

Conclusions and OutlookThe obtained flow field results for the volume flow rate

Q = 30 lmin

are in acceptable agreement to the referenceLES and experimental results. Nevertheless, it is observedthat the peak velocity of the laryngeal jet is overestimatedin sections T1, T2 and T3 and that the reference LES re-sults match the experimental results condsiderably better.This behaviour is attributed to the missing low-Reynoldscorrection terms which account for viscous damping ef-fects in the viscous sublayer. This effect is also observedfor the profiles of turbulent kinetic energy. Although thepeak location is correctly predicted, the peak value is over-predicted systematically downstream of the Larynx.

The achieved local and total deposition fraction results

SIMBIO 2011 7

Page 8: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

100

1010

10

20

30

40

50

60

70

dp [nm]

DF

[%]

(a)

100

1010

10

20

30

40

50

60

70

dp [nm]

DF

[%]

(b)

100

1010

10

20

30

40

50

60

70

dp [nm]

DF

[%]

(c)

Figure 9 Total percentage deposition fractionDF [%] asa function of the particle diameter for Q1 = 10 l

min(a),

Q2 = 20 l

min(b) and Q3 = 30 l

min(c). x : Eulerian mass

transport model [11], + : Euler-Lagrange method [11], ∗ :Eulerian mass transport model [10],⊳ : experimental results[35], ⊲ : experimental results [36], • : present experimen-tal results, see Table 2 ,−�− : present numerical results, seeTable 2.

(a) (b)

Figure 10 Isometric back view on contours of deposition en-hancement factor DEF predicted at the Epiglottis and Larynxcorresponding to the deposition of nanoparticles with a ther-modynamic diameter dp = 14.5nm (a) and dp = 52nm(b) for a volume flow rate ofQ1 = 10 l

min.

(a) (b)

Figure 11 Contour of deposition enhancement factor (DEF)predicted at the Epiglottis and Larynx corresponding to thedeposition of nanoparticles with a thermodynamic diameterdp = 14.5nm (a) and dp = 52nm (b) for a volume flowrate of Q1 = 10 l

min.

SIMBIO 2011 8

Page 9: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

show the expected effects of the inhalation flow rate andthe particle diameter on the amount of the deposited par-ticles and the deposition pattern. Nonetheless, the com-parison of experimental and numerical deposition fractionvalues shows a significant difference, which might be re-lated to both the inaccurate prediction of the turbulent ki-netic energy and the turbulent kinematic viscosity respec-tively and the near-wall concentration gradient. However,for small deposition fractions ofDF < 2% as monitoredfor dp = 20nm it is questionable, if a more accurate pre-diction is necessary or if a statement of the approximateamount of deposited particles, i.e.DF < 5%, is sufficient.

It is expected, that extending thek-ω SST RANS modelwith low-Reynolds corrections terms as proposed in [21]will significantly improve the near-wall as well as the over-all flow field predictions. An even more advanced model isthe recently proposedγ-Reθ RANS model [37]. It extendsthe usedk-ω SST model by a transport equation for theturbulence intermittencyγ and for the momentum thick-ness Reynolds numberReθ as an indication for the onsetof transition. However, the computational cost will also in-crease with extending the turbulence model. Furthermore,previous studies have already shown that LES is superiorto RANS for the accurate prediction of the complex meanflow field human upper airways [10, 33]. It allows not onlythe analysis of the time-averaged flow fields, but also of thelarge scale turbulent structures.

However, even if the flow field is computed using LESand accurate dynamic subgrid-scale models, the scalar-fluxterm is commonly modeled using the gradient diffusion hy-pothesis [10]. The major drawback of this hypothesis is theassumption of isotropic turbulence which is invalid for thelaminar-turbulent flow in the human upper airways. Thus,a major accuracy improvement is expected by solving thefluid and particle equations in a coupled manner using LESand a localised dynamic subgrid-scale closure not only forthe residual stress tensor but also for the scalar-flux term[38]. Although this method will be computationally ex-pensive, it gives accurate instantaneous and time-averagedvelocity and particle concentration profiles as well as in-sight in the dynamic interaction between the turbulent flowand the particle motion and local deposition.

There are several other ways to increase the realism ofthe model to mimic the real world inhalation and exhala-tion of airborne fine and ultrafine particles. They rangefrom applying a physiologically realistic breathing pat-tern and consideration of a polydisperse aerosol to extend-ing the upper airway model with a realistic model of thetracheo-bronchial tree with multiple generations.

Despite the huge potential of CFD for the simulation ofthe very complex flow in the human airways, their valida-tion by means of experimental results will remain one ofthe major issues.

AcknowledgementsThe authors gratefully acknowledge partial financial

support by the concerted research (GOA) grant from theVrije Universiteit Brussel.

References[1 ] M. Geiser, W. G. Kreyling, Deposition and biokinet-

ics of inhaled nanoparticles,Particle and Fibre Toxi-cology7, pp. 1-17 (2010)

[2 ] W. G. Kreyling, M. Semmler-Behnke, W. Möller,Health implications of nanoparticles,Journal ofNanoparticle Research8, pp. 543-56 (2006)

[3 ] A. Ibald-Mulli, H.-E. Wichmann W. G. Kreyling, A.Peters, Epidemiological Evidence on Health Effectsof Ultrafine Particles,Journal of Aerosol Medicine15,pp. 189-201 (2002)

[4 ] H. Schulz, V. Harder, A. Ibald-Mulli, A. Khandoga,W. König, F. Krombach, R. Radykewicz, A. Stampfl,B. Thorand, A. Peters, Cardiovascular Effects of Fineand Ultrafine Particles,Journal of Aerosol Medicine18, pp. 1-22 (2005)

[5 ] World Health Organization (WHO), Air qualityguidelines. Global update 2005. Particulate matter,ozone, nitrogen dioxide and sulfur dioxide (2006)

[6 ] G. Oberdörster, Pulmonary effects of inhaled ultra-fine particles,International Archives of Occupationaland Environmental Health74, pp. 1-8 (2001)

[7 ] U. S. Akhtar, J. A. Scott, A. Chu, G. J. Evans, In vivoand In vitro Assessment of Particulate Matter Toxi-cology, Environmental Science and Engineering74,pp. 427-449 (2011)

[8 ] O. Schmid, W. Möller, M. Semmler-Behnke, G. A.Ferron, E. Karg, J. Lipka, H. Schulz, W. G. Kreylingand T. Stöger, Dosimetry and toxicology of inhaledultrafine particles,Biomarkers14, pp. 67-73 (2009)

[9 ] C. Kleinstreuer and Z. Zhang, Airflow and ParticleTransport in the Human Respiratory System,Annu.Rev. Fluid Mech.42, pp. 301-334 (2010)

[10 ] Z. Zhang and C. Kleinstreuer, Laminar-to-turbulentfluid-nanoparticle dynamics simulations: Modelcomparisons and nanoparticle-deposition applica-tions,International Journal of Numerical Methods inBiomedical Engineering(2011)

[11 ] P. W. Longest and J. Xi, Effectiveness of Direct La-grangian Tracking Models for Simulating Nanoparti-cle Deposition in the Upper Airways,Aerosol Scinceand Technology41, pp. 380-397 (2007)

[12 ] Z. Zhang, C. Kleinstreuer, J. F. Donohue and C. S.Kim, Comparison of micro- and nano-size particle de-positions in a human upper airway model,Journal ofAerosol Science36, pp. 211-233 (2005)

[13 ] A. Comerford, G. Bauer and W. A. Wall, Nanopar-ticle transport in a realistic model of the tracheo-bronchial region,International Journal for NumericalMethods in Biomedical Engineering26, pp. 904-914(2010)

SIMBIO 2011 9

Page 10: Numerical and Experimental Investigation of Ultrafine Part ... · exposure to airborne fine and ultrafine particles and ad-verse health effects, ranging from increased hospitaliza-tions

[14 ] Z. Zhang, C. Kleinstreuer and C. S. Kim, Airflowand Nanoparticle Deposition in a 16-Generation Tra-cheobronchial Airway Model,Annals of BiomedicalEngineering36, pp. 2095-2110 (2008)

[15 ] Z. Zhang and C. Kleinstreuer, Computational anal-ysis of airflow and nanoparticle deposition in acombined nasal-oral-tracheobronchial airway model,Journal of Aerosol Science42, pp. 174-194 (2011)

[16 ] S. Elgobashi, An updated Classification Map ofParticle-Laden Turbulent Flows,Fluid Mechanicsand Its Applications81, pp. 3-10 (2006)

[17 ] Mark Brouns, Numerical and experimental studyof flows and deposition of aerosols in the upper hu-man airways, PhD Thesis, Vrije Universiteit Brussel,Brussels, Belgium (2007)

[18 ] S. Verbanck, H. S. Kalsi, M. F. Biddiscombe, V. Ag-nihotri, B. Belkassem, C. Lacor and O. S. Usmani,Inspiratory and expiratory aerosol deposition in theupper airway,Inhalation Toxicology23, pp. 104-111(2011)

[19 ] C. Roth, G. A. Ferron, E. Karg, B. Lentner, G. Schu-mann, S. Takenaka and J. Heyder, Generation of ultra-fine particles by spark discharging,Aerosol Scienceand Technology38, pp. 228-235 (204)

[20 ] W. G. Kreyling, P. Biswas, M. E. Messing, N. Gib-son, M. Geiser, A. Wenk, M. Sahu, K. Deppert, I.Cydzik, C. Wigge, M. Semmler-Behnke, Generationand characterization of stable, highly concentrated ti-tanium dioxide nanoparticle aerosols for rodent in-halation studies ,Journal of Nanoparticle Research13, pp. 511-524 (2011)

[21 ] F. Menter, J. Carregal Ferreira, T. Esch and B.Konno, The SST Turbulence Model with ImprovedWall Treatment for Heat Transfer Predictions in GasTurbines, In proceedings of theInternational GasTurbine Congress, Tokyo, Japan (2003)

[22 ] W. H. Finlay, The Mechanics of Inhaled Pharma-ceutical Aerosols: An Introduction,Academic Press(2001)

[23 ] M. D. Allen and O. G. Raabe, Slip correction mea-surements for aerosol particles of doublet and trian-gular triplet aggregates of spheres,Journal of AerosolScience16, pp. 57-67 (1985)

[24 ] H. Shi, C. Kleinstreuer, Z. Zhang and C. S. Kim,Nanoparticle transport and deposition in bifurcatingtubes with different inlet conditions,Physics of Fluids16, pp. 2199-2213 (2004)

[25 ] I. Balashazy, W. Hofmann, T. Heistracher, Compu-tation of local enhancement factors for the quantifi-cation of particle deposition patterns in airway bifur-cations,Journal of Aerosol Science30, pp. 185-203(1999)

[26 ] I. Balashazy and W. A. Hofmann and T. Heistracher,Local particle deposition patterns may play a key rolein the development of lung cancer,Journal of AppliedPhysiology94, pp. 1719-1725 (2003)

[27 ] OpenFOAM®, www.openfoam.com

[28 ] H. G. Weller, G. Tabor, H. Jasak, C. Fureby, Atensorial approach to computational continuum me-chanics using object-oriented techniques,Computersin Physics12, pp. 620-631 (1998)

[29 ] H. Jasak, Error Analysis and Estimation for the finiteVolume Method with Applications to Fluid Flows,PhD Thesis, Imperial College of Science, Technologyand Medicine, London, United Kingdom (1996)

[30 ] J. H. Ferziger, M. Peric, Computational Methods forFluid Dynamics, 3rd Edition,Springer(2001)

[31 ] P. W. Longest, S. Vinchurkar, Effects of mesh styleand grid convergence on particle deposition in bir-furcating airway models with comparisons to exper-imental data,Medical Engineering & Physics29, pp.350-366 (2007)

[32 ] S. Vinchurkar, P. W. Longest, Evaluation of hexa-hedral, prismatic and hybrid mesh styles for simulat-ing respiratory aerosol dynamics,Computers & Flu-ids37, pp. 317-331 (2008)

[33 ] ANSYS®ICEM CFD

TM, www.ansys.com

[34 ] S. T. Jayaraju, M. Brouns, C. Lacor, B. Belkassemand S. Verbanck, Large eddy and detached eddy sim-ulations of fluid flow and particle deposition in a hu-man mouth-throat,Journal of Aerosol Science39, pp.862-875 (2008)

[35 ] Y. S. Cheng, Y. F. Su. H. C. Yeh and D. L. Swift,Deposition of Thoron Progeny in Human Head Air-ways,Aersol Science and Technology18, pp. 359-375(1993)

[36 ] K. H. Cheng, Y. S. Cheng, H. C. Yeh and D. L. Swift,An Experimental Method for Measuring Aerosol De-position Efficiency in the Human Oral Airway,Amer-ican Industrial Hygiene Association Journal58, pp.207-213 (1997)

[37 ] R. B. Langtry and F. R. Menter, Correlation-BasedTransition Modeling for Unstructured ParallelizedComputational Fluid Dynamics Codes,AIAA Journal47, pp. 2894-2906 (2009)

[38 ] W. Cabot and P. Moin, Large eddy simulationof sclar Transport with the dynamic subgrid-scalemodel, InLarge Eddy Simulation of Complex Engi-neering and Geophysical flows(ed. B. Galperin andS. A. Orszag), Cambridge University Press, pp. 141-158 (1993)

SIMBIO 2011 10