You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233556795

CFD Study on Compound Impaction in a Jet-in-Well Impactor

Article  in  Aerosol Science and Technology · December 2007


DOI: 10.1080/02786820701777432

CITATIONS READS
5 540

3 authors, including:

Satya Seshadri
Indian Institute of Technology Madras
12 PUBLICATIONS   90 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Satya Seshadri on 20 November 2014.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [Ingenta Content Distribution (Publishing Technology)]
On: 15 October 2014, At: 04:06
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Aerosol Science and Technology


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/uast20

CFD Study on Compound Impaction in a Jet-in-Well


Impactor
a a a
Shishan Hu , Satyanarayanan Seshadri & Andrew R. McFarland
a
Aerosol Technology Laboratory, Department of Mechanical Engineering , Texas A&M
University, College Station , Texas, USA
Published online: 22 Jan 2008.

To cite this article: Shishan Hu , Satyanarayanan Seshadri & Andrew R. McFarland (2007) CFD Study on Compound Impaction
in a Jet-in-Well Impactor, Aerosol Science and Technology, 41:12, 1102-1109, DOI: 10.1080/02786820701777432

To link to this article: http://dx.doi.org/10.1080/02786820701777432

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Aerosol Science and Technology, 41:1102–1109, 2007
Copyright c American Association for Aerosol Research
ISSN: 0278-6826 print / 1521-7388 online
DOI: 10.1080/02786820701777432

CFD Study on Compound Impaction in a Jet-in-Well


Impactor

Shishan Hu, Satyanarayanan Seshadri, and Andrew R. McFarland


Aerosol Technology Laboratory, Department of Mechanical Engineering, Texas A&M University, College
Station, Texas, USA
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

conditions. Marple and Liu (1974) investigated the characteris-


In a classical jet impactor, the jet is assumed to impact on an tics of laminar jet, rectangular and round impactors using nu-
infinite plate; however, the typical construction of an impactor ne- merical simulations. Their studies involved analyzing the effects
cessitates that a wall surrounds the jet. If the wall is close to the on performance of the Reynolds number, the ratio of jet char-
jet, secondary impaction can take place on the wall. Also, for some
impactor designs, e.g., large particle scalping devices in air sam-
acteristic dimension to the distance between the jet exit plane
pling inlets, the jet is purposefully placed in a well and the overall and plate, and the Stokes number. Hari et al. (2005) conducted
performance of the impactor is the compound effect of the jet im- numerical studies on the sensitivities of the factors that affect
pacting on a collection plate and on the wall of the well. In this study the performance of a rectangular slit impactor including gravity
we primarily use CFD to examine the performance of compound and ultra-Stokesian drag. Their predictions agreed well with the
round jet impactors, where the ratio of well diameter to jet diame-
ter is from 3 to 15 and the particle size is in the range of 5 to 15 µm
experimental results. Burwash et al. (2006) studied turbulent
AD. The onset of the secondary impaction occurs at a ratio of well dispersion effects on 5 µm particles in an axisymmetric im-
diameter to jet diameter of about eight with the collection on the pinging jet at a Reynolds number of 104 and suggested that the
side wall increasing for smaller ratios. Compound impaction can turbulent dispersion for particle deposition may be many times
significantly enhance the overall particle collection and can result of that of particle inertial impaction. They used the Shear Stress
in a much smaller value of Stk0.5 . At a ratio of well diameter to jet
Transport (SST) turbulent model in CFX (CFX5.7.1, ANSYS,
diameter of 3, the value of Stk0.5 is 0.07, as compared with a classical
circular jet impactor where Stk0.5 is about 0.24 for a typical set of Inc., Canonsburg, PA) to simulate the jet and showed reasonable
operational and configuration parameters. Numerical predictions agreement with the experimental observations. John (1999) gave
for the jet-in-well impactors are compared with experimental re- a simple derivation for impactors and showed Stk0.5 values of
sults using liquid particles and good agreement is obtained. Photos about 0.49 and 0.25 for rectangular and circular jet, respectively.
taken over a range of experimental conditions verify the strong side
wall impaction for certain conditions.
Huang and Tsai (2002) numerically investigated the influence
of the ratio of the jet diameter to the diameter of the impaction
plate. They found that the collection efficiency increases when
the ratio decreases (same jet, larger plate) in a Reynolds number
INTRODUCTION range of 100 to 500 at a fixed Stokes number, and attributed the
The geometry of a classical impactor is simple and the de- phenomenon to a slower jet and gravitational effects.
vice is both easy to fabricate and operate. Impactors have been In this study, numerical techniques were used to predict the
studied extensively for different configurations and operational performance of jet-in-well impactors, which have the primary
application of scalping larger particles from the size distribu-
tions of sampled aerosols. Unlike classical jet impactors, the
jet-in-well impactors may exhibit compound aerosol collection,
Received 28 February 2007; accepted 30 October 2007. with part of the aerosol being deposited on a collection plate
This work was supported by the Edgewood Chemical Biological placed normal to the jet issuing from the acceleration nozzle,
Center of the U.S. Army Research, Development, and Engineering
Command under Contract DAAD13-03-C-0050 and by the Aerosol
and part of the collection occurring on the wall of the well,
Technology Laboratory of Texas A&M University. The authors wish which is normal to the jet that flows over the collection surface.
to express their appreciation to Drs. Edward W. Stuebing and Jerold R. The k–ω turbulence model was used to calculate the flow field
Bottiger of the Army for their guidance and support. We also wish to and a Discrete Random Walk Model (DRWM) was used with a
thank our colleagues Dr. John Haglund and Dr. Sridhar Hari for their Discrete Phase Model (DPM) to calculate particle trajectories.
help and suggestions.
Address correspondence to A. R. McFarland, Department of Me-
Several ratios of well-to-jet sizes were simulated to characterize
chanical Engineering, Texas A&M University, College Station, TX the influence of the ratio on the total collection of the impactor
77069, USA. E-mail: armcfarland@gmail.com and on the side wall impaction. Impactors with three different

1102
JET-IN-WELL IMPACTOR 1103

ratios were tested with benchmark physical experiments using TABLE 1


liquid particles, which like the particles used in the numerical Design parameters of artifacts used in physical experiments
simulations, are assumed to adhere to surfaces that they contact.
There are practical applications of compound impactors. Parameter Case 1 Case 2 Case 3
Westech Inc. (Atlanta, GA) provides a jet-in-well 10 µm pre- Jet diameter, D j (mm) 30.5 12.7 12.7
separator for its Andersen eight-stage impactor. ESM Andersen Well diameter DW (mm) 91.4 50.8 91.4
GmbH. (Erlangen, Germany) distributes a similar device for a Well-to-Jet ratio, χ 3 4 7.2
stack sampling system. Peters et al. (2001) developed the EPA Inlet tube diameter, Di (mm) 45.3 45.3 45.3
WINS impactor, which uses a jet-in-well impactor operated at a Gap between well and housing, 12.7 33.0 12.7
flow rate of 16.7 L/min to remove particles larger than 2.5 µm. δ (mm)
The effect of the size of the collection well on the performance Height of housing, Hc (mm) 87.9 87.9 87.9
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

of the WINS was discussed, but no details were provided. The Jet-to-plate spacing, S (mm) 45.1 27.9 27.9
Reynolds number in the WINS is about 6,000 and Stk0.5 is 0.238. Jet-to-plate spacing ratio, S/W 1.5 2.2 2.2
Exhaust tube diameter, Do (mm) 40.1 40.1 40.1
SYSTEM DESCRIPTION
With reference to Figure 1, the jet-in-well impactor con- enced to jet outlet conditions (Stk), and the ratio of well-to-jet
sidered herein is comprised of an acceleration nozzle that dis- diameters (χ ). Symbolically, these parameters are:
charges air into a well, where the flow impacts on the collection
plate (base of the well), and then is directed horizontally toward π 2
Q̇ = D U
the side wall of the well. The flow then makes another turn and 4 j
is directed upwards where it is vented from the well. Primary ρ U Dj
Re =
and secondary inertial impaction of aerosol particles take place µ
on the collection plate and the side wall, and this process will be Cc ρ P U D 2P
referred to as compound impaction. Parameters used to charac- Stk = [1]
9µ D j
terize the jet-in-well impactors include the volumetric flow rate
DW
( Q̇), the flow Reynolds number (Re), the Stokes number refer- χ =
Dj

Here, U is the spatial mean velocity at the acceleration nozzle


exit plane, D j is the diameter of the acceleration jet, DW is
the diameter of the well, ρ is the air density, µ is the dynamic
viscosity of the air, Cc is Cunningham’s correction, ρ p is the
particle density, and D p is the particle diameter.
The dimensions of the test articles used in the physical exper-
iments using monodisperse liquid particles with the jet-in-well
impactors are listed in Table 1. Three test cases were examined,
which involved use of different parameters including the ratio
of well-to-jet diameters, χ .

NUMERICAL SIMULATION
Fluent 6.2.14 (Fluent Inc., Lebanon, NH) was employed to
perform the simulations using the appropriate models to calcu-
late the flow field and particle trajectories. The turbulent models
together with higher-order discretization schemes were used to
simulate the air flow because the Reynolds number was about
3,400 and turbulent flow was assumed. After calculating the
flow field, the DRWM together with the Lagrangian Discrete
Phase model, DPM, was used to calculate the particle collection
efficiency in the impactor.
For the calculations of aerosol particle collection, the particle
concentration was assumed to be dilute so the problem can be
treated as a one-way coupling simulation. The air phase acts
on the particles and determines their motion but air flow is not
FIG. 1. Schematic diagram of a jet-in-well impactor. affected by the particles.
1104 S. HU ET AL.

In the calculation of the turbulent flow, the Reynolds number particle Reynolds number, Re p . The factor, C D Re p /24, in Equa-
is important for understanding flow conditions, creating the com- tion (4) accounts for non-Stokesian behavior as the Reynolds
putational mesh, and deciding the boundary conditions. The k–ω number increases.
model used in this study considers the low-Reynolds-number ef- For the range of Reynolds numbers under consideration, tur-
fects and it calculates the flow throughout the field including the bulent dispersion can affect the particle motion and it may even
core region and boundary area. When k–ω is used, it is important be dominant in some particle trajectory calculations. Turbulent
that the distance between the wall and the centroid of the first flow is characterized by fluctuation velocities and eddies, which
cell adjacent to the wall, is chosen to properly satisfy a require- have various length and time scales. The principle of the parti-
ment that y + should be approximately unity. The dimensionless cle motion in a turbulent flow is still not fully understood and
distance, y + , is defined as (Fox et al. 2004): there is no completely accurate solution. But for engineering
applications a stochastic tracking approach, the DRWM, is used
yu ∗
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

by Fluent to predict the particle motion. In Equation (4), the


y+ = [2]
υ instantaneous fluid velocity is required for calculation of par-
ticle motion, where the instantaneous velocity is the sum of
Here, y is the distance from the wall, υ is the kinematic viscosity the mean velocity, which can be obtained from the flow field
of air, and u ∗ is the friction velocity. calculations, and the fluctuating component, which needs to be
With respect to the solver and solution control, a second- modeled. When the particle resides in one computational cell,
order upwind scheme was selected for discretization of variables generally, the mean velocity can be obtained using the center
such as momentum, turbulent kinetic energy production, and value of this cell and the random velocity fluctuations (u  , v  ,
dissipation. SIMPLEC (Van Doormaal and Raithby 1984) was and w ) can be calculated using the local RMS value of the ve-
used for pressure-velocity coupling because it provides a rapid locity fluctuations or turbulent kinetic energy, which in turn are
convergence of the solution. Boundary conditions were based obtained from the flow field calculations, e.g.:
on turbulence intensity and the inside diameter of the jet. The
turbulence intensity was calculated from the Reynolds number   
  
of the jet (Fluent Documentation 2005): u = ζ u 2 , v = ζ v 2 , w = ζ w 2 [8]
0.16
i= [3] Here, ζ is a normally distributed random number used to
Re1/8
generate the fluctuation velocity. If the turbulent flow is isotropic,
The trajectory of a particle is calculated by the integration the turbulent kinetic energy, k, can be used in the calculation,
of the force balance on the particle in the Lagrangian reference viz:
frame to determine its location in each computational cell and
  
the migration between the cells. Drag force and gravity were the 
external forces considered in calculating the particle trajectories, u = v = w 2 = 2k/3
2 2 [9]
viz:
In the DRW model, the turbulent dispersion effect is calcu-
d V p  a − V p )
C D Re p (U lated as an interaction between the particles and a succession
= + g [4]
dt 24 τ of turbulent fluid eddies. An eddy is described by a Gaussian
distribution random velocity fluctuation, a time scale (lifetime),
where:
and a length scale (size). The eddy lifetime can be calculated
Cc ρ P D 2P from (Shirolkar et al. 1996; Fluent Documentation 2005):
τ= [5]
18µ
C D = f (Re p ) [6] τe = 2TL or τe = −TL log(r ) [10]

and: In Equation (10), the eddy lifetime could be based on a con-


stant of 2 for isotropic flow or use of a Random Eddy Lifetime by
 a − V p |
ρ D p |U
Re p = [7] involving a uniform random number r , where r ∈ [0, 1]. Here,
µ TL is the fluid Lagrangian integral time scale which is deter-
mined by the particle inertia and describes a time during which
Here, V p is particle velocity, τ is the aerodynamic relaxation the particle could maintain its original velocity before it changes
time in the low Reynolds number limit, U  a is undisturbed fluid velocity and migrates to another eddy. For small particles,
velocity at the location of the particle, g is the gravitational ac- which are assumed to have shorter relaxation time than the time
celeration, and C D is the drag coefficient, which is based on the scales of all eddies in the turbulent flow, TL can be calculated
JET-IN-WELL IMPACTOR 1105

from: from:

k ṁ sw
TL = C L [11] ηsw = [15]
ε ṁ ref

where C L has a value of about 0.15 for the k–ω model. Here, ṁ sw is the rate of aerosol particle collection at the side
The interaction time τinteraction of a particle and an eddy is wall of the well in the impactor.
determined from the eddy life time, τe , or the particle crossing
time,τcross , i.e.:
RESULTS AND DISCUSSION
τinteraction = min(τe , τcross ) [12] Visual Evidence of Compound Impaction
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

Figure 2 shows four sets of photos that were taken after ex-
where: posing jet-in-well impactors for an extended period of time (typ-
  ically one hour) to an aerosol formed from a solution to which
Le
τcross = −τ ln 1 − [13] dye was added to increase contrast of the particles deposited on
τ | u − vp|
filter substrates. The glass fiber filters were only used to help
qualitatively visualize the impaction of particles on the bottom
Here, L e is the eddy length scale and τ is the particle relax-
surface and side wall. All experimental tests for quantitative
ation time. After the interaction time is reached, the instanta-
analysis of particle collection were based on smooth surfaces
neous velocity is calculated using a new ζ (random number).
as described earlier. In each set of photos, the upper picture
shows the deposition on the collection plate and the lower pic-
EXPERIMENTAL METHODS ture shows deposition on a small section of a fibrous filter that
The jet-in-well impactor was experimentally tested with had been placed on the side wall of a well. When the filter placed
monodisperse liquid particles to validate the numerical predic- on the side wall is flattened, the deposition pattern becomes a
tions. For aerosol generation, a vibrating orifice aerosol genera- strip. It can be seen that the particle deposition patterns are quite
tor, VOAG (Model 3450, TSI Inc, Shoreview, MN) was used to different in the impactors that have different combinations of
generate liquid droplets from a solution containing ethanol, oleic well-to-jet ratios and particle Stokes numbers.
acid, and a fluorescent tracer. The consistency of aerosol output With reference to Figure 2a, there is a circular pattern on
by the VOAG was monitored with an aerodynamic particle sizer, the collection plate with a lower density of deposition in the
APS (Model 3321, TSI Inc, Shoreview, MN). Here the aerosol- center, which may be observed with unbounded circular jet im-
to-aerosol fractional efficiency is calculated from the amounts pactors where the less dense deposition in the center can occur
of particulate matter that enter and exhaust from the impactor, due to air stagnation effects; however, there is also a narrow
without taking into consideration where particles are deposited strip of deposition on the side wall, which is about 6 mm wide
in the impactor. Aerosol samples were alternately collected on a and is elevated about 6 mm above the primary impaction col-
reference filter, which sampled the aerosol transmitted through lection plate. These photos represent the Case 2 impactor listed
the impactor with no well in the system, and filters placed at the in Table 1, where the well-to-jet ratio is 4.0, and the particle
outlet of the impactor with the well in place. Glass fiber filters Stokes number is 0.12 (9.7 µm AD aerosol particles). The side-
(Type A/D, Pall Corp., East Hills, NY) were placed at the outlet wall strip is caused by secondary impaction, because if it were
port of the impactor to collect the aerosol particles. The exposed caused by turbulent deposition, it would exhibit a more diffuse
glass filters were placed in solutions of isopropyl and water to pattern. When the velocity contours are simulated, for situa-
elute the fluorescent tracer. Solutions were analyzed with a fluo- tions where secondary impaction is observed, e.g., Figure 3,
rometer (Model FM109535, Quantech Barnstead International, they show the secondary turn of the air flow impacts the side
Dubuque, IA). The fractional efficiency, η, of the impactor, for wall at a location that starts about 5 mm above the bottom plate.
the particle size being tested, was calculated from: The curvature of the abrupt turn of the airstream at the sidewall is
similar to the curvature of the airstream in the primary impaction
ṁ impactor process, and thus should result in somewhat similar impaction
η =1− [14]
ṁ ref characteristics.
Figure 2a and Figure 3 can be used as the basis for describ-
Here, ṁ impactor is the rate of aerosol particle collection at the ing the compound impaction phenomenon. During the first turn,
outlet of the impactor with the well in place and ṁ ref is the rate when the air impacts the collection plate the particles will strike
of aerosol particle collection by the reference filter. the plate if their inertia is sufficient, and if not they will fol-
Smooth thin plastic film was used as a substrate for collection low the airstream towards the side wall. However, during the
of particles on the side wall. The fractional efficiency ηsw of first turn, particles that do not reach the collection plate can
the side wall, for the particle size being tested, was calculated move a short distance closer to the collection plate. When the
1106 S. HU ET AL.
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

FIG. 2. Photographs of deposition patterns observed on collection plate (upper pictures) and side walls (lower pictures) for different combinations of well-to-jet
ratio and particle Stokes numbers. Aerosol particles were collected on glass fiber filters to obtain contrast for these photographs.

air flow approaches the side wall in a small-sized well, its ve- unit sampled aerosol for a period that was twice the sampling
locity does not decrease significantly and the particles having period used for the other samples shown in Figure 2, deposition
benefited from the first turn are closer to the collection surface, is not evident on the side wall.
so they can more easily be collected during the second turn of the In contrast to the results of the test conditions shown in
air. Figures 2a and 2b, for small values of χ and Stk, the impaction
If the Stokes number for the first turn is sufficiently high, the may take place predominantly on the side wall, Figure 2c. For
particle deposition will primarily take place on the collection this case χ = 3 and Stk = 0.06.
surface. Figure 2b shows the patterns for particle collection with As should be expected, as the well-to-jet ratio increases, the
χ = 4, and Stk = 0.27, where the value of χ is the same as that impaction process tends to that of the classical jet impactor. Fig-
for Figure 2a, but the Stokes number is higher. Even though the ure 2d shows the deposition patterns for an arrangement with

FIG. 3. Velocity contours for the jet-in-well impactors with well-to-jet ratios of 3 and 4.
JET-IN-WELL IMPACTOR 1107

χ = 7.2 and Stk = 0.12 (the same value used to obtain the
photographs shown in Figure 2a). Here, the deposition is pre-
dominantly on the collection plate with no visible evidence of
deposition on the side wall.

Characterization of the Compound Impaction


Phenomenon: Total Collection
It can be expected that the total collection of a jet-in-well
impactor could increase when the well size decreases due to
the secondary impaction effect. Figure 4 shows the numerical
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

predictions for the collection efficiency as a function of Stokes


number for jet-in-well impactors with well-to-jet ratios from 3
to 15. The geometry of the jet-in-well impactors is based on
the physical configuration shown in Figure 1. When the ratio of
well-to-jet increases from 3 to 15, the total collection decreases FIG. 5. The cutpoint Stokes number, Stk0.5 ,as a function of well-to-jet ratio χ .
and the collection curves shift toward the right. For well-to-jet
ratios in a small range from 3 to 8, the total efficiency decreases
rapidly with an increasing ratio. The effect of the well-to-jet ratio on the value of Stk0.5 can
The value of the cutpoint Stokes number, Stk0.5 , decreases be represented by an expression, such as that given in Equation
significantly as the well-to-jet ratio decreases from 8 to 3. For (15) and shown plotted in Figure 5. In particular:
example, at a value of χ = 3, the value of Stk0.5 = 0.07, whereas
at a value of χ = 8, the value of Stk0.5 = 0.23. However, when Stk0.5 = 0.0314e0.2486χ for: 3≤χ ≤8
the well-to-jet ratio falls in a range larger than 8, the jet-to-well [16]
Stk0.5 = 0.23 for: χ ≥8
ratio has little effect on the total collection, which means that
the secondary side wall impaction is weak. This is probably
as a result of the attenuation of velocity in the radial direction When the ratio is less than 8, Stk0.5 will increase with the
along the collection plate. increasing ratio and when it is larger than 8, Stk0.5 remains at
Figure 4 also shows the experimental results for the three about 0.23.
cases of jet-in-well impactors with well-to-jet ratios of 3, 4, Figure 6 shows the numerically predicted collection for the
and 7.2. Numerical predictions agree well with the experimental bottom plate and side wall with the well-to-jet ratios of 4 and
tests for the particles having larger Stokes numbers. However, 7.2. The collection curves for the bottom plate are similar for
the numerical results over-predict particle collection for Stokes the two ratios because the acceleration jets are operated at the
numbers smaller than 0.04. same conditions so the primary impaction is similar; however,
the collections for the side walls are significantly different. When
the well-to-jet ratio is 4 and the secondary impaction is strong,
e.g., at a Stokes number of about 0.1, the secondary impaction is
twice the primary impaction; however, the secondary impaction
drops rapidly for higher values of Stokes number. The high level
of secondary side wall impaction may be caused by two factors.
First, the particles that are not collected on the bottom plate
could get closer to the bottom plate as discussed earlier. Second,
gravity could enhance this movement of the particles towards
the bottom plate since the first impaction and gravity have the
same direction. These two effects could concentrate the particles
in the region closer to the bottom plate and make the particles
easier to impact onto the side wall.
In Figure 6, the side wall particle collection is significantly
less for the well-to-jet ratio of 7.2 compared with the ratio of 4.
At the ratio of 7.2, the collection on the side wall is less than that
on the bottom plate for both low and high inertia particles. When
the jet is decelerated in a large well, the velocity parallel to the
FIG. 4. Simulated and experimentally determined total collection as a func-
tion of Stokes number for various well-to-jet ratios. Here EXP refers to experi- collection plate, in the region of the well wall, is lower than that
mental results. in a small well. The maximum side wall collection is only about
1108 S. HU ET AL.
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

FIG. 7. Secondary side wall collection efficiency as a function of the well-to-


jet ratio, χ . Here EXP refers to experimental results.

in reduced values of Stk0.5 , e.g., the cutpoint Stokes number


is about 0.07 when the well-to-jet ratio is 3. Simulations were
conducted for a series of jet-in-well impactors with different
size ratios, which showed that the side wall and the total collec-
tion will decrease when the ratio increases in the range of 3–8,
beyond which the secondary side wall impaction can be ignored.
The side wall impaction is demonstrated photographically and
supported by experimental data.
Matida et al. (2004) reported that turbulent dispersion model
could cause an over-prediction of the particle deposition. For
FIG. 6. Total bottom plate and side wall collection of the jet-in-well impactor small Stokes number particles, turbulent dispersion may be sev-
different well-to-jet ratios. (a) χ =4. (b) χ = 7.2. eral times higher than the inertial impaction (Burwash et al.
2006). However, for particles having larger Stokes numbers,
10%, which is associated with particles having a Stokes number inertia should be dominant and predictions of the deposition
of about 0.2 when the ratio is 7.2. caused by inertia could be assumed to be accurate. In our exper-
Computational results for side wall deposition over the range
iments, we noted that the model and physical experiments agree
of 3 ≤ χ ≤ 15 are shown in Figure 7. The peak value of side wall
well at a flow Reynolds number of about 3,400 except for the
deposition decreases from about 43% for χ = 3 to about 3% for
small Stokes numbers.
χ = 8. For larger values of the well-to-jet ratio, the secondary
When the Jet-to-Plate spacing ratio (S/W ) is small it may
deposition could be ignored.
affect the total collection efficiency. Marple (1974) and Hari
An unusual feature of the compound impactor occurs when
(2003) investigated this influence for a laminar round jet and they
the well-to-jet ratio is about 4. With reference to Figure 3(b),
reported that when the ratio is less than 1, Stk0.5 will increase
which shows the velocity contours for χ = 4, the reflected flow
rapidly as the ratio increases. But it has little influence on the
from the side wall interacts with the initial jet from the acceler-
Stk0.5 when the ratio is in the range of 1–3. In our study, the ratio
ation nozzle, which could cause the total flow to fluctuate. We
of the Jet-to-Plate spacing was 2.2 except for Case 1 where the
suggest that if a jet-in-well impactor is to be used, χ should
Jet-to-Plate spacing ratio we 1.5. It was assumed in our study that
be in the range of 6–8 to achieve a suitable flow field, and yet
the Jet-to-Plate spacing ratio did not affect the total collection
avoid having a large-sized well. However, the smaller ratio of 3,
significantly.
Figure 3a, will also provide a stable flow field, and can provide
The WINS (Peters et al. 2001) has a well-to-jet ratio of about
a given cutpoint with a lower jet velocity.
9.45 which falls in a large ratio range so the secondary impaction
in WINS can be ignored. The Stk0.5 of WINS is about 0.238
DISCUSSION AND SUMMARY which is close to our numerically predicted result of 0.23.
In this study, the compound impaction effect is character-
ized for circular jet-in-well inertial impactors. The compound
REFERENCES
impaction includes primary impaction under the acceleration Burwash, W., Finlay, W., and Matida, E. (2006). Deposition of Particles by
nozzle and secondary impaction at the side wall of the well, a Confined Impinging Jet onto a Flat Surface at Re = 104 , Aerosol Sci.
which can significantly enhance particle collection and result Technol. 40:147–156.
JET-IN-WELL IMPACTOR 1109

Fluent 6.2 Documentation (2005), Fluent Inc. Lebanon, New Hampshire, John, W. (1999). A Simple Derivation of the Cutpoint of an Impactor, J. Aerosol
USA. Sci. 30(10): 1317–1320.
Fox, R. W., McDonald, A. T., and Pritchard, P. J. (2004). Introduction to Fluid Matida, E. A., Finlay, W. H., Lange, C. F., and Grgic, B. (2004). Improved
Mechanics, 6th ed., John Wiley & Sons Inc., Hoboken, NJ. Numerical Simulation of Aerosol Deposition in an Idealized Mouth-Throat,
Fuchs, N. A. (1964). The Mechanics of Aerosols. Pergamon Press, New J. Aerosol Sci. 35:1–19.
York. Marple, V. A. and Liu, B. Y. H. (1974). Characteristics of Laminar Jet Impactors,
Hari, S., Hassan, Y. A., and McFarland, A. R. (2005). Computational Fluid Environ. Sci. Technol. 8:648–654.
Dynamics Simulation of a Rectangular Slit Real Impactor’s Performance, Peters, T. M., Vanderpool, R. W., and Wiener, R. W. (2001), Design and
Nuclear Engineering and Design. 235: 1015–1028. Calibration of EPA PM-2.5 Well Impactor Ninety-Six (WINS), Aerosol Sci.
Hari, S. (2003). Computational Fluid Dynamics (CFD) Simulations of Dilute Technol. 34: 389–397.
Fluid-Particle Flows in Aerosol Concentrators. Ph.D. Dissertation. Depart- Shirolkar, J. S., Coimbra, C. F. M., and McQuay, M. Q. (1996), Fundamental
ment of Mechanical Engineering, Texas A&M University, College Station, Aspects of Modeling Turbulent Particles Dispersion in Dilute Flows, Prog.
TX. Energy Combust. Sci. 22:363–399.
Huang, C. H., and Tsai, C. J. (2002). Influence of Impaction Plate Diameter Van Doormaal, J. P., and Raithby, G. D. (1984). Enhancements of the SIMPLE
Downloaded by [Ingenta Content Distribution (Publishing Technology)] at 04:06 15 October 2014

and Particle Density on the Collection Efficiency of Round-Nozzle Inertial Method for Predicting Incompressible Fluid Flows, Numerical Heat Transfer,
Impactors, Aerosol Sci. Technol, 36:714–720. Part 1, 7:147–163.

View publication stats

You might also like