Sunteți pe pagina 1din 15

Aerosol Science 35 (2004) 789 803

www.elsevier.com/locate/jaerosci

Particle deposition measurements and numerical simulation


in a highly idealized mouththroat
Y. Zhang, W.H. Finlay , E.A. Matida
Aerosol Research Laboratory of Alberta, Department of Mechanical Engineering, 4-9 Mechanical Engineering
Building, University of Alberta, Edmonton, Alberta, Canada T6G 2G8

Received 11 July 2003; accepted 30 December 2003

Abstract

A new highly idealized mouththroat that mimics in vivo deposition in the human oropharynx is proposed
in the present study. Deposition of aerosols with mass median diameters of approximately 2.5, 5.0, and
6:0 m is studied experimentally in the proposed highly idealized mouththroat and two standard USP throats
(described in the United States Pharmacopeia, 1995 and 2000, respectively) for steady inhalation 8ow rates of
30, 60 and 90 l=min. Gravimetry is used to measure aerosol deposition in these geometries. Both USP throats
show curves of total deposition e;ciency that are far below the in vivo average curve (J. Aerosol Med. 2(3)
(1989) 285), revealing the inadequacy of the USP throats in replicating realistic mouththroat geometries. On
the contrary, the new proposed highly idealized mouththroat geometry follows the in vivo average curve and
is therefore an excellent candidate for use as a standard in the pharmaceutical >eld, replacing the available
and inadequate USP throats.
In addition to the above measurements, numerical simulation of aerosol deposition using computational 8uid
dynamics (CFD) techniques is also performed. Reynolds averaged NavierStokes (RANS) equations with a
k! turbulence model are used to solve the primary 8ow and a Lagrangian random-walk eddy interaction
model with near-wall correction (J. Aerosol Sci. 35 (2004) 1) is applied in the tracking of individual particles
in the highly idealized mouththroat. The total deposition results obtained from CFD simulation show relatively
good agreement when compared with the measured data.
? 2004 Elsevier Ltd. All rights reserved.

Keywords: Highly idealized mouththroat; USP throat; Monodisperse aerosols; Particle deposition; Gravimetry; CFD
simulations


Corresponding author. Tel.: +1-780-492-4707; fax: +1-780-4922200.
E-mail address: warren.>nlay@ualberta.ca (W.H. Finlay).

0021-8502/$ - see front matter ? 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jaerosci.2003.12.006
790 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

1. Introduction

In the treatment of lung diseases, medication in the form of aerosols (Finlay, 2001) is usually
delivered into the lungs through the extrathoracic region (de>ned as the region from the mouth
opening to the trachea). This region plays an important role in respiratory drug delivery, since it
acts as a >lter proximal to the lung (Finlay, 2001). Most of the previous work on mouth deposition
has been performed in vivo using radiolabelled aerosols (Foord, Black, & Walsh, 1978; Stahlhofen,
Gebhart, & Heyder, 1980, 1981; Emmett, Aitken, & Hannan, 1982; Stahlhofen, Gebhard, Heyder,
& Scheuch, 1983; Stahlhofen, Gebhard, Heyder, Scheuch, & Juraske, 1984; Cass et al., 1999). A
comprehensive evaluation of these early works was presented by Stahlhofen, Rudolf, and James
(1989), as shown in Fig. 1. However, due to morphological variation among diKerent individu-
als and the in8uence of inhalation 8ow rate, it is di;cult to understand the 8ow dynamics and
deposition of inhaled aerosols in the extrathoracic region. Since in vivo measurements are costly,
complex to conduct and show considerable experimental variability, recent researchers have adopted
in vitro measurements as an alternative, which are less costly and easier to perform, allowing sys-
tematic study of each parameter relevant to aerosol deposition (Grgic, Finlay, & Heenan, 2004a).
Experimental work with an idealized biological mouththroat model has been done by DeHaan
and Finlay (2001). In that study, a more physiologically realistic extrathoracic model was built
using simple geometric shapes based on computed tomography (CT) scan, magnetic resonance
imaging (MRI) scans, and direct observation of living subjects during tidal breathing (Stapleton,
Guentsch, Hoskinson, & Finlay, 2000). Further measurement and understanding were provided
by Grgic, Finlay, Burnell, and Heenan (2004), which is complementary to 8uid 8ow measure-
ments (Heenan, Matida, Pollard, & Finlay, 2003) and numerical modelling work (Matida, Finlay,
Lange, & Grgic, 2004). The eKect of inhaler mouthpiece geometry is elucidated by DeHaan and
Finlay (2003). However, the geometrical model of the extrathoracic airways used in these previous
studies is still complex and time consuming to build. Thus, it is common practice in the pharma-
ceutical >eld to use highly idealized representations of the mouththroat region, such as the United
States Pharmacopeia (USP) idealized throats when drug formulations or inhalation devices are tested.
Such highly idealized mouththroats are less expensive, easier to manufacture, and simpler to collect

Fig. 1. Extrathoracic deposition data from archival literature, plotted as a function of the inertial parameter d2p Q.
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 791

drug from in chemical assays when compared to more realistic mouththroats. The United States
Pharmacopeia (1995, 2000) describes two simple geometric model throats to mimic the >ltering
capabilities of the human mouth and throat. Although, Srichana, Martin, and Marriott (2000) studied
the deposition e;ciency in the USP throat B (the glass throat from a twin-stage impinger (TSI))
and a human mouththroat cast using monodisperse solid particles mixed with lactose and found
the mouththroat cast had a higher >lter e;ciency than the USP throat B, experimental results on
monodisperse aerosol deposition in these geometries are limited.
Cheng, Zhou, and Chen (1999) suggested that the deposition in curved pipes could be used to
model oral deposition. Actually, inertial deposition of aerosol particles in curved pipes is of impor-
tance in many practical problems, such as aerosol sampling and transport systems. The deposition
of inert particles in curved pipes has been studied experimentally by Landahl and Herrmann (1949)
at Reynolds numbers of 3144 and 6270 for tube bends of curvature ratios (the ratio of the bend
curvature radius to the tube radius) of 5 and 8; by Johnston, Isles, and Muir (1977) at Reynolds
numbers in the range 3112071 and curvature ratios in the range 5 30; by Pui, Novas, and Liu
(1987) at Reynolds numbers in the range 100 10,000 and curvature ratios in the range 5.6 7; and
by McFarland, Gong, Muyshond, Wente, and Anand (1997) at Reynolds numbers in the range 3200
19,800 and curvature ratios in the range 210. The results of these experiments showed that the
curvature ratio only weakly aKects the deposition e;ciency and that the 8ow Reynolds number
also mildly aKects deposition. On the other hand, several theoretical calculations in bends have
appeared in the literature. Landahl and Herrmann (1949) calculated a parabolic 8ow without sec-
ondary motion. Yeh (1974) derived an empirical formula to calculate the deposition e;ciency using
a laminar plug 8ow. Cheng and Wang (1975) developed a more complex solution by assuming a
potential vortex 8ow, and neglected the existence of secondary 8ows. Cheng and Wang (1981) used
a three-dimensional developed laminar 8ow model to calculate the deposition e;ciency in bends and
derived a new theory. Recently, with the advent of computational 8uid dynamics (CFD), numerical
simulation has the potential to provide predictions for particle deposition in bends (McFarland et al.,
1997).
Although USP throats have been used for decades when drug formulations or inhalation devices
were tested, there is little experimental data as to whether the USP throats mimic the >ltering ca-
pabilities of the human mouth and throat. The main goal of this study is to investigate the >ltering
capabilities in the USP throats and seek a more adequate mouththroat model for in vitro mea-
surements with nonballistic aerosols (where the aerosol particles move at speeds near that of the
inhalation air, as opposed to propellant metered dose inhaler sprays).

2. Methodology

2.1. Experimental methods

2.1.1. Mouththroat geometries


The proposed highly idealized mouththroat consisted of a circular mouth with 50 mm of length
and 30 mm diameter followed by a contraction nozzle with 45 mm of length, which smoothly con-
nected this mouth model with a curved 90 pipe (8:5 mm inner diameter and curvature ratio 5.6),
which were based on the curved pipe used by Pui et al. (1987). A schematic of the proposed
792 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

Fig. 2. Schematic of the highly idealized mouththroat.

mouththroat is shown in Fig. 2. Manufacturing of the models was done using stereo-lithography
(model FDM 8000, Stratasys, Eden Prairie, MN), a three dimensional printing process that produces
solid models in acrylonitrilebutadiene-styrene (ABS) plastic. Each model consisted of two halves,
which could be >rmly connected along the sagittal plane. The resultant models surfaces were coated
with an airtight epoxy layer which was found to have high chemical and physical resistivity (DeHaan
& Finlay, 2003). Finally, the inside wall of the models was coated with 8uorocarbon coating FC-725
(3M, St. Paul, MN) to increase liquid-surface contact angle, thus preventing surface wetting (Grgic
et al., 2004). As shown in Figs. 3 and 4, two USP throats were built using the same technique as that
used to build the idealized mouththroat. The USP throat A was built following the description in
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 793

Fig. 3. Schematic of the United States Pharmacopeia (2000) throat A.

the United States Pharmacopeia (2000). The USP throat B was made according to the speci>cation
in the United States Pharmacopeia (1995). All inner walls of the two USP throats were painted with
the same coating as that used in the proposed mouththroat.
794 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

Fig. 4. Schematic of the United States Pharmacopeia (1995) throat B.

2.1.2. Experimental setup and procedure


A schematic of the experimental setup is shown in Fig. 5. Monodisperse aerosol was generated
from di-2-ethylhexyl sebacate oil (DEHS) using a controlled heterogeneous condensation Aerosol
Generator (CMAG, Model 3457, Topas, Germany). The aerosol generation principle is based on
condensation of vaporized aerosol material onto small saline particles, which serve as condensation
nuclei. By changing the number concentration of the condensation nuclei or the amount of saturated
vapor, particle sizes can be varied in the range from 1 to 8 m (for sebacate oil) with geometric
standard deviation less than 1.1. The aerosol stream from the generator was diluted and mixed with
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 795

Fig. 5. Schematic of experimental setup.

ambient air in the main mixing chamber to reduce the concentration of the aerosol. Aerosol particle
sizes and monodispersity were monitored using an Aerosizer Mach II (TSI, St Paul, MN). Three
typical particle sizes of inhalable aerosol, with mass median diameters of 2.5, 5.0, and 6:0 m, were
selected for this study. The monodisperse aerosol 8owed into the test mouththroats through the
mixing cylinder ( 90 145 mm with inlet of  7 mm). Two Marquest Respirgard >lters (Marquest
Medical Products Inc.) were used in series at the outlet of the model. The steady inhalation 8ow
rates of 30, 60 and 90 l=min were used in the experiments. The required 8ow rates were regulated
using a valve downstream of the >lter holder. Excess air was exhausted to the atmosphere by an
exhausting hose plugged into the main mixing chamber. The aerosol was passed through the testing
model for periods ranging from 20 s to 10 min, the ideal time being determined by the deposition
rate and the amount of oil that the model surface could hold before 8ooding the inner wall.
A gravimetry technique (Grgic et al., 2004) was used to measure particle deposition in the models.
The total deposition is given by

Mm
T:D:(%) = 100; (1)
Mm + M f

where Mm and Mf are the masses of particles deposited in the mouththroat model and the down-
stream parts (including >lters and adapter), respectively.
796 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

2.2. Numerical methods

2.2.1. Primary :ow solution


In this study, the governing equations of 8uid motion (NavierStokes equations) were solved in
the highly idealized mouththroat using a commercial >nite-volume-based CFX-TASC8ow software
(Version 2.12, AEA Technology Engineering Software Ltd., now ANSYS Inc.). In order to simulate
the inlet condition of the model mouththroat, the calculation domain included the mixing cylinder
( 90 145 mm with inlet of  7 mm) and mouththroat. The computational mesh was generated
with CFX Build (AEA Technology Engineering Software Ltd., now ANSYS Inc.). The comput-
ing domain had 48 blocks. In the present calculation, three diKerent grid sizes (222,919, 565,504,
1,030,000 elements) were tested for grid independence. Total deposition of the particles, pressure
drop through the mouththroat geometry and the mean velocities in outlet sections were chosen
as dependent parameters for checking grid convergence. The calculation results showed that grid
convergence was achieved since grid dependence of less than a few percent diKerences occurred in
the above parameters for the diKerent grid sizes. The >nal mesh included approximately 1,030,000
hexahedron elements, with biased accumulation of nodes toward the wall. Due to the strong swirl of
turbulence in curved pipes, the traditional engineering k turbulent closure model is not adequate
(McFarland et al., 1997). Instead, a standard version of the k! model (Wilcox, 1988) with Kato
& Launder (1993) modi>cation was used (Matida, DeHaan, Finlay, & Lange, 2003).
For boundary conditions, a constant mass 8ow rate of either 30, 60 or 90 l=min was used as
inlet condition to the mixing cylinder with a turbulence intensity of 10% of the mean velocity and
a turbulence length scale of 10% of the inlet diameter (doubling or halving these turbulence inlet
parameters caused deposition results to vary by less than a few percent). A Dirichlet zero gauge
pressure condition at the outlet was applied.

2.3. Particle tracking

An eddy interaction model (EIM) (for details see Gosman & Ioannides, 1981), was used to
perform the particle tracking. In EIMs, each particle is allowed to interact successively with various
eddies. The end of interaction between the particle and one eddy occurs when the lifetime of the
eddy is over or when the particle crosses the eddy. Then, a new interaction of the particle with a new
eddy is started. The trajectory of this particular particle is obtained by solving the particle equation
of motion. In order to obtain velocity or deposition statistics, hundreds of particles must be released
into the 8ow. The particles used in the present calculation have relatively small inertia, meaning that
particles will not traverse an eddy before it decays, so crossing trajectory eKects are not relevant and
the length scales can be neglected. Matida, Nishino, and Torii (2000) have shown that the isotropy
assumption used in the EIM caused an over-prediction of particle deposition for small particles in
a turbulent pipe 8ow and also that once anisotropy was considered, particle deposition prediction
was remarkably improved. Similar anisotropy can be found in the 8ows of this study. Therefore,
near-wall corrections (see Matida et al., 2004, for details) were included in this simulation, which
were based on the fact that the 8uctuating velocity component normal to the walls plays the major
role for particle deposition. In the present simulation, the extent of this correction was optimized
and applied up to y+ between 20.0 and 100.0 (depending on the 8ow rate).
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 797

The particle tracking algorithm CODE developed by Matida et al. (2004) was used here and
compared with Lagrangian tracking code available in the CFX-TASC8ow (Version 2.12, AEA Tech-
nology Engineering Software Ltd., now ANSYS Inc.). The basic assumptions in the Lagrangian
tracking model are: particleparticle interactions are not included in the model; there are no particle
source terms to the turbulence equations and therefore, turbulence is not modulated by the discrete
phase; and spherical particles are considered. In our experiments, the volume fractions of the par-
ticles were usually in the order of 106 or less. The mean distance between the particles is an
order of magnitude larger than the diameter of the particles. Thus, particle motion can be treated as
having no eKect on the 8uid motion (Finlay, 2001), and particleparticle interactions can be ignored
as well. Furthermore, the particle material is DEHS oil, which is inert and has low vapor pressure,
so the eKect of evaporation on the particle size is negligible. Therefore, the above assumptions are
reasonable here. The deposition of aerosol through the geometry was calculated by tracking a large
number of simultaneously released particles, recording particle deposition positions and determining
how many particles were deposited on the walls and how many penetrated through the bend. Using
the converged 8ow >eld solution, particles were released at the entrance of the mixing cylinder
with a uniform distribution in the plane perpendicular to the axis of the inlet. Mono-sized particles
of 2, 3, 4, 6 and 8 m and density  = 0:912 g=cm3 were used for simulation of deposition e;-
ciency. The initial particle velocities were set equal to those of the 8uid and one-way coupling was
assumed between the air and particle phase. When using the Lagrangian tracking code available in
the CFX-TASC8ow, n = 10; 000 particles were released from the inlet of the mixing cylinder, and
particle deposition positions were recorded. In tracking with our near-wall correction algorithm, the
simulations in an idealized mouth geometry with 10,000 and 1000 particles showed negligible diKer-
ences (within 0.35%) in total deposition (Matida et al., 2003). Therefore, to reduce the calculation
times, n = 1000 particles were released randomly from the inlet of the mixing cylinder with an even
spatial concentration distribution.

3. Results and discussion

3.1. Experimental results

3.1.1. Deposition e<ciency in proposed mouththroat and USP throats


Fig. 6 summaries the total deposition e;ciency as a function of the inertial parameter d2p Q, where
(g=cm3 ) is the particle density, dp (m) is the particle diameter and Q (cm3 =s) is the inhalation
8ow rate. The average curve (Stahlhofen et al., 1989), based on in vivo experimental work is
shown on the same >gure. Each experimental point represents the average of three repeats for
the same particle size and inhalation rate. The experimental results show that particle deposition
increases with increasing values of the inertial parameter, suggesting that inertial impaction is the
dominant deposition mechanism. This result is consistent with the results of many studies (Yu, Diu,
& Soong, 1981, Stahlhofen et al., 1989) where the deposition of particles ( 1 m) in the oral
region is dominated by inertial mechanisms. Comparing the total deposition e;ciency in the three
model throats to the in vivo average curve (Stahlhofen et al., 1989), the total deposition e;ciency in
the proposed mouththroat agrees well with the in vivo average curve. In contrast, both USP throats
achieve lower deposition over the range of inertial parameter d2p Q tested here and the diKerence
798 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

Fig. 6. The total deposition e;ciency in model throats for diKerent particle size and 8ow rate as a function of inertial
parameter d2p Q and the empirical curve >t proposed by Stahlhofen et al. (1989). Each point represents the average of
three repeats and error bars refer to standard deviation.

increases with increasing d2p Q. For relatively larger particle sizes and higher inhalation 8ow rate, in
vitro deposition results using the USP throats will give unrealistically low particle deposition when
compared to the average curve (Stahlhofen et al., 1989) for in vivo experiments in humans.
The inertial impaction parameter d2p Q does not take into account cross-sectional area inter-subject
variability of the mouththroat region. Grgic et al. (2004) suggested instead adopting a Stokes
number based on the mean equivalent dimension, since Stokes number includes the relevant length
and velocity scales. For this purpose, the Stokes number (Stk) is de>ned as
Cp d2p U0
Stk = ; (2)
18dt
where C is the Cunningham slip correction, p is the particle density, dp is the particle diameter,
and  is the dynamic viscosity of air. The velocity U0 is obtained by dividing 8ow rate by the area,
obtained using a mean diameter (dt ) that is calculated by dividing the total volume by the length of
the central sagittal axis of each mouththroat model. Assuming a circular equivalent cross-sectional
area, the equivalent diameters are derived for the three mouththroat models. Fig. 7 reveals the
deposition e;ciency as a function of the Stokes number derived in the above manner. Also shown
in Fig. 7 is data obtained by Grgic et al. (2004a) for deposition in seven diKerent realistic mouth
throat replicas created from MRI scans of actual human subjects during inhalation. The experimental
results show that the particle deposition increases with increasing Stokes number, and that the highly
idealized geometry provides a reasonable representation of deposition seen in the realistic mouth
throats of Grgic et al. (2004a).

3.2. The CFD results

3.2.1. Particle tracking with di>erent tracking models


Total particle deposition results for monodisperse particles of 2.0, 3.0, 4.0, 6.0 and 8:0 m and
the steady inhalation 8ow rates of Q = 30, 60 and 90 l=min with the standard k! model (with
Kato & Launder, 1993, modi>cation) are shown in Figs. 810. In comparison with the experimental
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 799

Fig. 7. The total deposition e;ciency in the model throats for diKerent particle size and 8ow rate as a function of Stokes
number. Each point represents the average of three repeats and error bars refer to standard deviation. Also shown for
comparison is data from Grgic et al. (2004a) for seven realistic mouththroat replicas created from MRI scans of subjects
during inhalation.

Fig. 8. Comparison of experimental data and numerical calculation results: inhalation 8ow rate, Q = 30 l=min.

results (four experiments with diKerent particle diameters, dp = 2:5, 5.0, 6.0 and 7:0 m), the dotted
line represents turbulent tracking without the near-wall correction. The inertial parameter, d2p Q
is used as abscissa. It can be seen clearly in the >gures that turbulent tracking without near-wall
correction overestimates total deposition over the entire range of inertial parameter (d2p Q), when
compared to the experimental data. When the near-wall correction, up to a dimensionless distance
from the wall of y+ = 100 (for 8ow rate 60, 90 l=min) and y+ = 20 (for 8ow rate 30 l=min), was
applied, better agreement with the experimental results is seen as shown by the solid line. The
simulated results show the near-wall corrections for EIM are necessary to circumvent overestimation
800 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

Fig. 9. Comparison of experimental data and numerical calculation results: inhalation 8ow rate, Q = 60 l=min.

Fig. 10. Comparison of experimental data and numerical calculation results: inhalation 8ow rate, Q = 90 l=min.

of particle deposition. Comparison of the near-wall correction results vs. mean 8ow tracking reveals
that turbulence mixing plays only a limited role in the particle deposition in this geometry.

3.3. Discussion

Our results show that substantial deposition occurs in the highly idealized mouththroat similar to
in vivo data in human volunteers, which is in8uenced by particle size and inhalation rate. However,
deposition e;ciency in the USP throats is much diKerent from the in vivo measurements. This can
be explained by the lack of constricted high speed 8ow regions in these geometries. In particular,
due to morphological features of the human mouththroat region, both the cross-sectional area and
perimeter varies from section to section, which cause the air 8ow to accelerate, decelerate and rotate,
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 801

especially in the pharynx and larynx regions, where a typical equivalent diameter is 0:817 cm (Cheng
et al., 1999), so that the local average axial velocity can reach 9 m=s at a 8ow rate of 30 l=min. The
small dimension and higher local velocity in the narrowest region of the larynx give a high local
Stokes number. It is well known that higher Stokes number means higher probability of deposition
by inertial impaction. However, within the USP throats, there is no region that can mimic the high
local Stokes number regions where most of deposition of inhalable size particles occurs by inertial
impaction. Pui et al. (1987) pointed out that the deposition e;ciency approaches zero for >nite local
Stokes number (i.e., Stk = 0:1 and 0.2 for Re = 1000 and 100, respectively). Although this critical
Stokes number varies with the geometry and 8ow rate, we can deduce reasonably that when the
Stokes number is less than some particular value, little particle deposition occurs in the geometry.
Local Stokes numbers in the USP throats are an order of magnitude lower than in our proposed
geometry (e.g. local Stk based on minimal diameter is 0.072 for the highly idealized vs. 0.0064 or
0.0032 for the USP A or B throats, respectively, at Q = 30 l=min, Dp = 5 m) as well as being much
lower than expected in vivo values in the constricted regions. This may explain the low deposition
in the USP throats.
As for the CFD simulation, it is well known that turbulence models perform adequately in simple
high Reynolds number and boundary layer 8ows. However, the 8ow in the proposed mouththroat
geometry has many complex characteristics, such as curved streamlines and secondary 8ow regions.
In addition, the Reynolds number is on the order of a few thousand, so that transitional 8ows
that are neither fully laminar nor fully turbulent may occur. Current turbulence models are known
to perform poorly in such 8ows (Stapleton, Guentsch, Hoskinson, & Finlay, 2000). Improvement
in aerosol deposition modelling with near-wall correction shows an increase in the accuracy of the
particle deposition prediction. In contrast, commercial CFD software with a turbulent particle tracking
model can result in signi>cant errors in prediction of particle deposition in the present geometry.

4. Conclusions

Both USP throats show curves of total deposition e;ciency that are discrepant with the in vivo
average curve (Stahlhofen et al., 1989), revealing the inadequacy of the USP throats in replicating
realistic mouththroat geometries. The diKerence between the in vivo data and the data from the USP
throats limits the ability of cascade impactor tests to predict in vivo deposition performance, when
drug formulations or inhalation devices are tested. In contrast, a newly proposed highly idealized
mouththroat geometry follows the in vivo average curve as well as can be expected for such a
simpli>ed geometry and is therefore an excellent candidate for use as a standard in the pharmaceutical
>eld, replacing the available and inadequate USP throats for dry powder inhaler and nebulizer testing.
Future work is needed to examine the ability of the proposed highly idealized geometry to mimic
in vivo deposition of high speed propellant metered dose inhaler (pMDI) sprays.
Our present simulation shows that once adequate near-wall correction (Matida et al., 2004) is ap-
plied, a relatively simple Reynolds averaged NavierStokes (RANS) model for the primary 8ow with
Schuens eddy interaction model (EIM) (Schuen, Chen, & Faeth, 1983) for the particulate phase
can be used in the prediction of the deposition of particles in our highly idealized mouththroat
geometry. When near-wall corrections of the turbulence kinetic velocity are used in the EIM, sim-
ulated results of total deposition of particles in the highly idealized mouththroat geometry show
802 Y. Zhang et al. / Aerosol Science 35 (2004) 789 803

relatively good agreement with experimental data obtained in the experiments, improving signi>-
cantly over results obtained without near-wall corrections, namely, the turbulent tracking which is
commonly used in commercial codes.

Acknowledgements

The authors gratefully acknowledge Biljana Grgic for her help with the aerosol deposition tests.

References

Cass, M. R. L., Brown, J., Ickford, P., Fayinka, S., Newman, P. S., Johansson, J. C., & Bye, A. (1999).
Pharmacoscintigraphic evaluation of lung deposition of inhaled zanamivir in healthy volunteers. Clinical
Pharmacokinetics, 36(Suppl.1), 2131.
Cheng, Y. S., & Wang, C. S. (1975). Inertial deposition of particles in a bend. Journal of Aerosol Science, 6, 139145.
Cheng, Y. S., & Wang, C. S. (1981). Motion of particles in bends of circular pipes. Atmospheric Environment, 15,
301306.
Cheng, Y. S., Zhou, Y., & Chen, B. T. (1999). Particle deposition in a cast of human oral airways. Aerosol Science and
Technology, 31, 286300.
DeHaan, W. H., & Finlay, W. H. (2001). In vitro monodisperse aerosol deposition in a mouth and throat with six diKerent
inhalation devices. Journal of Aerosol Medicine, 14(3), 361367.
DeHaan, W. H., & Finlay, W. H. (2003). Predicting extrathoracic deposition from dry powder inhaler. Journal of Aerosol
Science, submitted for publication.
Emmett, P. C., Aitken, R. J., & Hannan, W. J. (1982). Measurements of the total and regional deposition of inhaled
particles in the human respiratory tract. Journal of Aerosol Science, 13, 549560.
Finlay, W. H. (2001). The mechanics of inhaled pharmaceutical aerosols: An Introduction. London: Academic Press.
Foord, N., Black, A., & Walsh, M. (1978). Regional deposition of 2.5 7:5 m diameter inhaled particles in healthy male
non-smokers. Journal of Aerosol Science, 9, 343357.
Gosman, A. D., & Ioannides, E. (1981). Aspects of computer simulation of liquid-fueled combustors. In Paper
AIAA-81-0323 American Institute of Aeronautics and Astronautics (AIAA) 19th Aerospace Science Meeting,
St. Louis, MO.
Grgic, B., Finlay, W. H., & Heenan, A. F. (2004). Regional aerosol deposition and 8ow measurements in an idealized
mouth and throat. Journal of Aerosol Science, 35, 2132.
Grgic, B., Finlay, W. H., Burnell, P. K. P., & Heenan, A. F. (2004a). In vitro intersubject and intrasubject deposition
measurements in realistic mouththroat geometries. Journal of Aerosol Science, submitted for publication.
Heenan, A. F., Matida, E. A., Pollard, A., & Finlay, W. H. (2003). Experimental measurements and computational
modeling of the 8ow in an idealized extrathoracic airway. Experiments in Fluids, 35, 7084.
Johnston, J. R., Isles, K. D., & Muir, D. C. F. (1977). In W. H. Walton, (Ed.), Inhaled particles, Oxford: Pergamon
Press, Vol. 4, pp. 6172.
Kato, M., & Launder, B. E. (1993). The modeling of turbulent 8ow around stationary and vibrating square cylinders.
Proceedings of the Ninth symposium on turbulent shear :ows, pp. 10-4-110-4-6.
Landahl, H. D., & Herrmann, R. G. (1949). Sampling of liquid aerosols by wires, cylinders and slides and the e;ciency
of impaction of the droplets. Journal of Colloid Science, 4, 103136.
Matida, E. A., DeHaan, W. H., Finlay, W. H., & Lange, C. F. (2003). Simulation of particle deposition in an idealized
mouth with diKerent small diameter inlets. Aerosol Science and Technology, 37, 924932.
Matida, E. A., Finlay, W. H., Lange, C. F., & Grgic, B. (2004). Improved numerical simulation of aerosol deposition in
an idealized mouththroat. Journal of Aerosol Science, 35, 119.
Matida, E. A., Nishino, K., & Torii, K. (2000). Statistical simulation of particle deposition on the wall from turbulent
dispersed pipe 8ow. International Journal of Heat and Fluid Flow, 21, 389402.
Y. Zhang et al. / Aerosol Science 35 (2004) 789 803 803

McFarland, A. R., Gong, H., Muyshond, A., Wente, W. B., & Anand, N. K. (1997). Aerosol deposition in bends with
turbulent 8ow. Environmental Science and Technology, 31(12), 33713377.
Pui, D. Y. H., Novas, F. R., & Liu, B. Y. H. (1987). Experimental study of particle deposition in bends of circular cross
section. Aerosol Science and Technology, 7, 301315.
Schuen, J. S., Chen, L. D., & Faeth, G. M. (1983). Evaluation of a stochastic model of particle dispersion in a turbulent
round jet. American Institute of Chemical Engineers (AIChE) Journal, 29(1), 167170.
Srichana, T., Martin, G., & Marriott, C. (2000). A human oral-throat cast integrated with a twin-stage impinger for
evaluation of dry powder inhalers. The Journal of Pharmacy and Pharmacology, 52, 771778.
Stahlhofen, W., Gebhart, J., & Heyder, J. (1980). Experimental determination of the regional deposition of aerosol particles
in the human respiratory tract. American Industrial Hygiene Association Journal, 41(6), 385398a.
Stahlhofen, W., Gebhard, J., & Heyder, J. (1981). Biological variability of regional deposition of aerosol particles in the
human respiratory tract. American Industrial Hygiene Association Journal, 42, 348352.
Stahlhofen, W., Gebhard, J., Heyder, J., & Scheuch, G. (1983). New regional deposition data of the human respiratory
tract. Journal of Aerosol Science, 14, 186188.
Stahlhofen, W., Gebhard, J., Heyder, J., Scheuch, G., & Juraske, P. (1984). Particle deposition in extrathoracic airways
of healthy subjects and of patients with early stages of laringeal carcinoma. Journal of Aerosol Science, 15, 115117.
Stahlhofen, W., Rudolf, G., & James, A. C. (1989). Intercomparison of experimental regional aerosol deposition data.
Journal of Aerosol Medicine, 2(3), 285308.
Stapleton, K. W., Guentsch, E., Hoskinson, M. K., & Finlay, W. H. (2000). On the suitability of k-turbulence modeling
for aerosol deposition in the mouth and throat: a comparison with experiment. Journal of Aerosol Science, 31(6),
739749.
United States Pharmacopeia, (1995). USP (United States Pharmacopeia), Vol. 23 (pp. 1765 1766).
United States Pharmacopeia, (2000). USP (United States Pharmacopeia), Vol. 24 (pp. 26782679).
Wilcox, D. C. (1988). Reassessment of the scale determining equation for advanced turbulence models. American Institute
of Aeronautics and Astronautics (AIAA) Journal, 26(11), 12991310.
Yeh, H. C. (1974). Experimental study of particle deposition in bends of circular cross section. Bulletin of Mathematical
Biology, 36, 105116.
Yu, C. P., Diu, C. K., & Soong, T. T. (1981). Statistical analysis of aerosol deposition in nose and mouth. American
Industrial Hygiene Association Journal, 42, 726733.

S-ar putea să vă placă și