You are on page 1of 13

Aerosol Science and Technology

ISSN: 0278-6826 (Print) 1521-7388 (Online) Journal homepage: http://www.tandfonline.com/loi/uast20

Comparison of Three Approaches to Model


Particle Penetration Coefficient through a Single
Straight Crack in a Building Envelope

Bin Zhao , Chun Chen , Xinyan Yang & Alvin C. K. Lai

To cite this article: Bin Zhao , Chun Chen , Xinyan Yang & Alvin C. K. Lai (2010) Comparison
of Three Approaches to Model Particle Penetration Coefficient through a Single Straight
Crack in a Building Envelope, Aerosol Science and Technology, 44:6, 405-416, DOI:
10.1080/02786821003689937

To link to this article: https://doi.org/10.1080/02786821003689937

Published online: 20 Apr 2010.

Submit your article to this journal

Article views: 793

View related articles

Citing articles: 20 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=uast20
Aerosol Science and Technology, 44:405–416, 2010
Copyright © American Association for Aerosol Research
ISSN: 0278-6826 print / 1521-7388 online
DOI: 10.1080/02786821003689937

Comparison of Three Approaches to Model Particle


Penetration Coefficient through a Single Straight
Crack in a Building Envelope

Bin Zhao,1 Chun Chen,1 Xinyan Yang,2 and Alvin C. K. Lai3


1
Department of Building Science, School of Architecture, Tsinghua University, Beijing, China
2
Department of Building Services, School of Environmental Science and Engineering, Tianjin University,
Tianjin, China
3
Department of Building and Construction, City University of Hong Kong, Hong Kong

almost 100% of their time indoors. To accurately assess individ-


In this study, we present three approaches to predict particle ual indoor exposure, it is important to distinguish the contribu-
penetration coefficients through a single straight crack in build- tion from outdoor-origin fine particles, which penetrate indoors,
ing envelops. The three approaches are an analytical approach, an to that of indoor-generated fine particles. This issue may be more
Eulerian approach, and a Lagrangian approach, respectively. The
particle penetration coefficient through an idealized straight crack
important for those people who are living in high-rise residences
(smooth inner surfaces) and a strand board crack (rough inner in metropolitan cities. During summer time while air condi-
surfaces) were modeled by the three presented approaches. The tioning system is operating and winter time while the heating
calculated results were compared with the literature results. The system is operating, windows are closed in most of the time. It
comparison shows that for the idealized smooth crack, the modeled results in very low air exchange rate so that infiltration becomes
results by the Eulerian approach match the experiments best for the
entire range of particle sizes studied among the three approaches.
the most dominant pathways for ambient air entering residential
The predicted results by the analytical approach also match the places. The study by Zhu et al. (2005) also indicated that the
experiments reasonable well. Results modeled by the Lagrangian penetration of particles may contribute dominantly for indoor
approach are less satisfied for fine particles (d p < 0.1 µm). Overall, particle concentration level for those buildings near highway.
all the three approaches agree well with the experiments for par- There are mainly three methodologies to quantify the “ef-
ticle sizes ranging from 0.4–1.2 µm. For cracks with rough inner
fectiveness” of infiltration of particles: (1) indoor-to-outdoor
surfaces, the results agree better with the measurements for all the
three approaches by adjusting the boundary conditions to incor- concentration ratio (I/O); (2) infiltration factor; and (3) pene-
porate the “intercept” effect of roughness on particle deposition in tration coefficient. Among the three parameters, the most rel-
the cracks. evant parameter for the penetration mechanism through cracks
is the penetration coefficient (or penetration factor) which is
defined as a parameter to indicate the fractional penetration of
1. INTRODUCTION particles from outdoors to indoors that is associated with infil-
Recent epidemiological evidence has observed that ambient tration airflow rate. Unlike the I/O ratio and infiltration factor,
particle concentration is very closely associated with cardio- the third approach of penetration coefficient, avoids the interfer-
pulmonary morbidity and mortality and even premature death ential factors such as air exchange rate and particle decay rate,
(EPA 2005). Since most people spend much of their time in- and thus it is a better measure to characterize the infiltration rate
doors, human exposure to outdoor-originated particles may not of particles.
equal to outdoor particle concentration levels. The elderly and There are three major deposition mechanisms controlling
infant, which are the most susceptible population groups, spend particle deposition when particle penetrating through cracks,
i.e., gravitational setting, Brown diffusion, and inertial im-
paction. Based on the particle deposition mechanisms of gravi-
tational settling, Fuchs (1964) applied a flow equation to derive
Received 17 June 2009; accepted 3 February 2010. the penetration coefficient for a channel flow, which can be
This work was sponsored by the National Natural Science Founda-
expressed as:
tion of China (Grant No. 50908127) and partially supported by a grant
from City University of Hong Kong (7002378).
Address correspondence to Bin Zhao, Dept. of Building Science,
School of Architecture, Tsinghua University, Beijing, China. E-mail: z vs
Pg = 1 − , [1]
binzhao@tsinghua.edu.cn d um

405
406 B. ZHAO ET AL.

where z is the crack length, d is the crack height, vs is the


gravitational settling velocity of particles, and um is the average
velocity of air in the crack.
Licht (1980) proposed a similar method that calculated pen-
etration coefficient by:

P = 1 − ε, [2]
FIG. 1. The two-dimensional crack model.

where ε is the deposition ratio.


2. METHODOLOGY
Based on the particle deposition mechanisms of gravitational
setting and Brown diffusion, Taulbee and Yu (1975) solved To compare the calculated results with the measured data by
the particle transport equation for steady-state laminar flow in Liu and Nazaroff (2003), the studied case was selected as the
cracks to calculate the particle penetration coefficient: two dimensional straight crack which shown in Figure 1.

∂C ∂C ∂ 2C 2.1. Analytical Model


u + vs =D 2, [3] 2.1.1. Particle Deposition in the Cracks
∂x ∂y ∂y
The airflow through the building envelope is usually lami-
where C is the particle concentration, u is the airflow velocity, nar due to the small length scale and low air velocity (Liu and
D is the particle Brownian diffusivity, x and y denote the hor- Nazaroff 2003). Hence the particle deposition rate in cracks
izontal and vertical axis (or horizontal and vertical distance), with smooth inner surfaces is mainly governed by two mecha-
respectively. nisms: Brownian diffusion and gravitational settling. The ana-
Based on all the three deposition mechanisms, Liu and lytical model is based on the understanding that there is a very
Nazaroff (2001) developed a mathematical model to calculate thin particle concentration boundary layer submerged inside the
the particle penetration coefficient through building cracks, as airflow (momentum) boundary layer, and throughout the con-
shown below: centration boundary layer the particle flux, J , is constant and
can be described by a modified form of Fick’s law:
P = Pg × Pd × Pi . [4] ∂C
J = −D − ivs C, [5]
∂n
Pg , Pd , and Pi represent the particle deposition rate due to
mechanisms of gravitational settling, Brownian diffusion and where n is the normal distance to the crack inner surfaces, and
inertial impaction respectively. The latest study by Tian et al. i is a parameter to describe the direction of the wall surfaces,
(2009) presented a similar model by Liu and Nazaroff (2001) to i.e., for an upward-facing horizontal surface (floor), i = 1; for
calculate particle penetration coefficient. a downward-facing horizontal surface (ceiling), i = −1; for a
Reviewing the particle penetration through building cracks, vertical surface, i = 0.
one can recognize that particle penetration can be determined D, the Brownian diffusivity of the particle, can be calculated
directly by calculating the particle deposition onto the inner sur- by:
faces of the cracks of the building envelope. Experimental mea-
surements are very scarce and the geometry studied were very k B T Cc
simple (Liu and Nazaroff 2003). Due to many plausible depo- D= , [6]
3π µdp
sition mechanisms are involved, numerical modeling approach
is always preferred to study particle penetration coefficient. In where kB is Boltzmann constant, and it is equal to 1.38 ×
literature there are three popular mathematical approaches to 10–23 J/K; T is the absolute temperature of the air; µ is the
calculate the particle penetration coefficient. In this work we molecular dynamic viscosity of air; dp is the particle diameter;
presented and compared the results obtained from the three ap- Cc is Cunningham coefficient caused by slippage (Hinds 1982):
proaches based on an analytical, an Eulerian and a Lagrangian
model.   
λ dp
The experiments of particle penetration coefficient through Cc = 1 + 2.514 + 0.8 × exp − 0.55 , [7]
a two-dimensional crack measured by Liu and Nazaroff (2003) dp λ
under different conditions were employed to compare with the
modeling results by the three approaches. The accuracy and λ is the mean free length of the air molecule which is equal to
performance of the three approaches were also discussed in 0.0667 µm at a temperature of 20◦ C and atmospheric pressure
details. (Hinds 1982).
COMPARISON OF MODELS FOR PARTICLE PENETRATION 407

The gravitational deposition velocity can be calculated by: Hence the solution of Equation (11) based on the boundary
conditions shown in Equation (13) is:
vs = τp g, [8]
1.5vs
For floor surface (i = 1), vd =  vs  dp  ,
1 − exp D 2
−h
where τ p is the particle relaxation time. For the particle sizes
studied in this work, it can be calculated by: [14]

−1.5vs
Cc ρp dp2 For ceiling surface (i = −1), vd =  d  .
τp = , [9] 1 − exp − vDs 2p − h
18µ [15]

where ρP is particle density.


Let C∞ as the particle concentration outside the boundary 2.1.2. Calculation of Penetration Coefficient
layer, the particle deposition velocity, vd , is calculated by the
After calculating the particle deposition velocity in the
following equation:
cracks, the penetration coefficient can be computed readily by
using the following equation based on the previous analysis by
Jw Wu and Zhao (2007):
vd = , [10]
C∞  
vd z
P = exp − , [16]
where Jw means the particle flux at the wall surfaces, for this um d
study it is the particle flux towards the upward-facing (floor) or
downward-facing (ceiling) surfaces of the crack. Substituting With known particle deposition velocity and using Equation
Equation (10) for Equation (5), we got an ordinary differential (16), we could compute the particle penetration coefficient for
equation (ODE) of particle concentration in the boundary layer: different particle diameters. For practical cases where the in-
ner surfaces of the cracks are rough, the boundary conditions
dC vd C∞ ivs C of Equation (11) must be adjusted to incorporate the “inter-
=− − . [11] cept” effect of the surface roughness on particle deposition.
dn D D
According to the previous study by Zhao and Wu (2006), the
boundary conditions shown in Equation (12) should be modified
There is no vertical surface for the two dimensional crack. as:
For the idealized condition in which the inner wall surfaces of
the crack are smooth, the boundary conditions for Equation (11)
n = y0 , C = 0
are (referring to Figure 1):
d
n = = h, C = Cmax [17]
2
n = 0, C = 0
d where y0 is the shifted distance of the particle concentration
n = = h, C = Cmax [12]
2 boundary layer, and it can be estimated by the three types of
flow regime over the rough surfaces (Zhao and Wu 2006).
It should be noted that n denotes the absolute normal dis-
tance from the ceiling or floor surface of the crack in Equation 2.2. Eulerian Model
(12). Cmax is the particle concentration at the centerline of the 2.2.1. Modeling of Particle Mass Conservation Equation
straight crack. The exact relationship between Cmax and C∞ is
The Eulerian model solves the particle mass conserva-
difficult to determine, as it depends strongly on the particle size
tion/transport equation numerically to predict the particle pene-
as well as the position in the axial direction of crack length.
tration coefficient directly. Zhao et al. (2009) concluded that in
For simplifying the analytical calculation, we assume that the
normal cases, (1) the drift flux caused by the momentum change
relationship between Cmax and C∞ is similar to the relationship
rate per unit volume of particle phase has little effect on particle
between umax and um , i.e., umax = 32 um (Taulbee and Yu 1975).
distribution for ultrafine particles (dp < 0.1 µm); (2) the ther-
Thus Equation (12) can be rewritten as:
mophoresis force may be larger than the gravitational force for
particles with a diameter in the range of 0.01 to 0.1 µm, but both
n = 0, C = 0 forces are relatively small to affect the particle dispersion. Thus
d 3 [13] only the slip effect due to gravitational settling is included in the
n = = h, C = C∞
2 2 “drift flux” term. The particle mass transport equation models
408 B. ZHAO ET AL.

the settling effect by incorporating the “drift flux” term in the For the ideal case in which the inner surfaces of the cracks are
equation, as shown in following equation: smooth, the boundary conditions are:
 
∂C ∂C ∂C ∂C ∂C C(0, y) = C∞
+ u +v +w + vs
∂t ∂x ∂y ∂z ∂y C(x, ±h) = 0, [22]
 
∂ 2C ∂ 2C ∂ 2C where h represents the vertical distance from the midline of the
=D 2
+ 2
+ 2 , [18]
∂x ∂y ∂z crack to the ceiling and floor, respectively.
To facilitate the model development, Equation (21) is nor-
where u, v, and w are the airflow velocity in x, y, z direction, malized by the following dimensionless parameters which are
respectively. For two dimensional steady-state laminar flow in defined as:
cracks, the equation can be simplified as:
2x
  X= , [23]
2 2 (d/2)P e
∂C ∂C ∂ C ∂ C
u + vs =D + . [19] 2y
∂x ∂y ∂x 2 ∂y 2 Y = , [24]
d
dvs
Here the term vs ∂C is the drift flux term due to gravitational σy = , [25]
∂y 2D
settling. c
According to previous study by Taulbee and Yu (1975), the C= , [26]
c∞
effect of Brownian diffusion is negligible for Peclet number u
(Pe) > 100, where Pe denotes the ratio of the convection and U = , [27]
um
diffusion and it is defined as:
For a Poiseuille flow that works for the airflow through the
lum straight cracks, the air velocity along the cracks can be calculated
Pe = , [20]
D from an analytical solution of laminar airflow (Taulbee and Yu
1975).
Here two orientations are considered separately. In y-
direction, l represents the crack height d. For cases considered  2

u 3 y
in this study, the crack height is of the order of magnitude of = 1−4 . [28]
10–4 m (see Table 1), thus combining particle diameters studied um 2 d
in the work, Pe in y direction is in the order of 102 or less and
hence we need to consider Brownian diffusion and gravitational Equation (20) could be rewritten as the following dimension-
settling simultaneously. On the other hand, in the direction along less form:
the horizontal length of the crack, l indicates the crack length z.
In this case Pe is large enough (in the order of 104 ) that we could ∂C ∂C ∂ 2C
U + σy = , [29]
ignore the effect of Brownian diffusion, and hence the equation ∂X ∂Y ∂Y 2
above can be defined as
and the dimensionless boundary conditions are:

∂C ∂C ∂ 2C C(0, Y ) = 1
u + vs =D 2. [21]
∂x ∂y ∂y C(X, ±1) = 0 [30]

TABLE 1
The five studied cases
Pressure difference Crack height Crack length Particle
Case no. Inner surfaces of the cracks P (Pa) d (mm) z (cm) diameter dp (µm)
1 Aluminum, smooth inner surface 4 0.25 4.3 0.01-10
2 Aluminum, smooth inner surface 4 1 9.4 0.01-10
3 Aluminum, smooth inner surface 10 0.25 4.3 0.01-10
4 Aluminum, smooth inner surface 10 1 9.4 0.01-10
5 Strand board, rough inner surface 4 0.25 4.5 0.01-10
COMPARISON OF MODELS FOR PARTICLE PENETRATION 409

We adjusted the boundary conditions for the actual cases where The particle penetration coefficient can be calculated by:
the inner surfaces of the cracks are rough. The dimensionless
form is: Ccm
P = . [36]
C∞
C(0, Y ) = 1
  It is more complicated to solve Equation (29) of this model
d − 2y0
C X, ± =0 [31] compared with the analytical one (Equation [11]), as Equation
d (29) is a PDE and Equation (11) is an ODE. However, we can get
the detailed information of particle concentration distribution
Equation (29) was solved numerically by finite difference along the crack with this model. Besides, this model has the
method (FDM). The numerical solution was solved with the feasibility to incorporate other mechanisms by adjusting the
assistance of the partial differential equations (PDE) function of drift flux term in Equations (21) and (29), e.g., the drift flux
the software MATLAB (Mathworks Inc. 2007). We performed caused by thermophoresis. It is also flexible for unsteady airflow
the grid independence test by using the same mode with finer conditions and cracks with more complex geometry by keeping
grids until the calculated results yielded small changes of the the time variation term with the three dimensional form as shown
particle concentration. Grid convergence index (GCI), which is by Equation (18).
based on Richardson extrapolation method (Richardson, 1910)
and has been suggested by Roache (1994), was calculated to
2.3. Lagrangian Model
show the relative error of grid independent test.
2.3.1. Model for Fluid Phase
εrms Particle motion results from various forces exerted by the
GCI(C) = Fs , [32] airflow, so it is important to simulate the airflow field accurately.
rp − 1
For this study, laminar airflow was modeled with computational
where Fs = 3, p = 2. r is the ratio of amounts of fine gird to fluid dynamics (CFD) tool. Air exchange/infiltration is driven
that of coarse grid. εrms is defined as: by pressure difference between indoors and outdoors, thus the
boundary conditions for the CFD simulation in this study are:
 n  12 defining the pressure difference at the inlet and outflow condition
2
i=1 εi,C for the outlet of the cracks, respectively. The inner surfaces are
εrms = , [33]
n stationary adiabatic walls.
We used FLUENT 6.2 CFD program to solve the governing
equations for fluid flow (Fluent Inc., 2005). We conducted a
where εi,C is defined as:
grid independence test by calculating the same mode with finer
grids until calculated results yielded only small changes during
Ci,coarse − Ci,fine
εi,C = , [34] simulations. Again, the similar grid convergence index defined
Ci,fine by Equations (32)–(34), GCI(u), was employed. The tested grid
densities are 50 × 500 and 100 × 1000. The values of GCI(u)
We tested three groups of grids (number of X grids × number were all less than 5%, which show that the grids are fine enough.
of Y grids = 30 × 401, 40 × 501, 10 × 801, respectively). The Thus the grid 50 × 500 was used for all simulations.
values of GCI(C) were all less than 0.1% for the simulated
cases, which shows that the grids are fine enough. Thus the grid 2.3.2. Particle Equation of Motion
30 × 401 was selected. Other cases and particle sizes followed
The Lagrangian approach calculates the trajectory of each
the similar rule and thus the procedure was not repeated here.
particle by integrating the force balance on each particle, which
is written as
2.2.2. Calculation of the Penetration Coefficient
To calculate the particle concentration, we used the flow p
du gx (ρp − ρ)
= FD (
u−u
p ) + + Fx , [37]
rate-weighted average concentration as the exit concentration as dt ρp
recommended by Middleman (1997). The concentration, Ccm ,
is estimated as: where, u  is the fluid phase velocity, u p is the particle velocity, ρ is
the fluid density and Fx is an additional acceleration (force/unit
u(y)C(Xz , Y )dA particle mass) term.
Ccm = , [35]
u(y)dA It is very flexible to incorporate different forces in Equation
(37). For the case studied, the drag force plays an important
where C(Xz , Y ) is the particle concentration distribution at the role according to the analysis by Zhao et al. (2004). The drag
outlet of the cracks. force follows the Stokes drag law for small Reynolds number
410 B. ZHAO ET AL.

(Re < 1) cases: where Nescape is the escaped particles at the outlet, and Ntotal is
the total particles released at the crack inlet.
18µ
Fdrag = FD (
ua − u
p ) = ua − u
(  p ). [38]
ρp dp2 CC
3. CASES STUDY
For fine particles that can penetrate through building en-
velopes, other forces such as Brownian motion should be con- 3.1. Cases Description
sidered, especially for ultrafine particles (Zhao et al. 2004). The modeling configuration of crack is a single rectangular
Therefore, the final form of the trajectory equation is: slot of uniform geometry as a surrogate of a leakage path in a
building envelope. The measurement of particle penetration co-
p
du gx (ρp − ρ) efficient by Liu and Nazaroff was based on this configuration.
= FD (
u−u
p ) + + Fb , [39] The smallest dimension of the crack (known here as “crack
dt ρp
height”) is denoted as d and the crack length is denoted as z.
where Fb represents Brownian motion. The expression of drag The crack pattern is illustrated in Figure 1. Airflow can be rep-
force and Brownian force can be found in Fluent 6.2 user’s guide resented as two dimensions, as the crack width is much larger
(Fluent Inc., 2005). than crack height. The airflow is driven by the pressure differ-
When particles strike the wall, they will be trapped since they ence (P) which was fixed at one of two values: 4 or 10 Pa.
usually cannot accumulate enough rebound energy to overcome The crack length, the dimension parallel to the airflow direc-
the adhesion force (Hinds 1982). When particles reach the crack tion, was 4.3 and 9.4 cm for idealized cracks, and 4.5 cm for
outlet, they will escape and the trajectories calculation terminate. the roughness cases. Various crack heights, perpendicular to the
The boundary conditions for the particle phase modeling are: flow direction, were considered (d = 0.25 and 1.0 mm). Parti-
particles will be trapped when they reach the inner surfaces of cles were assumed to be spherical with a density of 1 g/cm3 and
the cracks or they will escape from the outlet. The particles are with a broad range of particle diameters, 0.01 to 10 µm. Table 1
released at the crack inlet uniformly following the assumption shows the details of the five studied cases of different condi-
by Fuchs (1964) and the initial velocity of the particles is set as tions, which includes four cases with smooth inner surfaces and
zero as the inertia of studied particles is small enough to neglect one case with rough inner surfaces of cracks. As it is the sim-
the influence of initial velocity on the later trajectory. plest among all, the analytical method in Equation (4) (Liu and
We used FLUENT program to track the particle trajecto- Nazaroff 2001) was also included in the comparison.
ries (Fluent Inc., 2005). We conducted a particle number inde-
pendence test with different particle numbers: 1000, 5000, and 3.2. Results
10,000. The differences between the results (penetration coef- Figures 2–5 present the prediction results of the three mod-
ficient) by different particle numbers are less than 2% for each eling approaches as well as the measured results and modeling
test, which shows that the particle numbers are sufficient. Thus, results by Liu and Nazaroff (2001) under four different condi-
the particle number was set as 1000 for tracking. tions (i.e., Case 1, Case 2, Case 3, and Case 4).
It is a bit difficult to model the cases of cracks with rough
inner surfaces by the Lagrangian model. Here we assumed that
the particles trapped by the roughness of the inner surface and
thus the particle tracks were terminated once they reached the
effective roughness over the surfaces. By defining this intercept
effect of the wall roughness, particle deposition was modeled.
We set two imaginary planes as the virtual boundaries of the
effective roughness, which were shifted a distance of the height
of effective roughness from the ceiling and floor surface, respec-
tively. When particles flow into the area limited by the imaginary
planes, they were assumed to be trapped and the trajectories were
terminated.

2.3.3. Calculation of Penetration Coefficient Based on


Trajectories
It is easy to calculate the particle penetration coefficient based
on the tracked trajectories:

Nescape
P = , [40] FIG. 2. Comparison of model predictions with experimental data for alu-
Ntotal minum cracks for case 1.
COMPARISON OF MODELS FOR PARTICLE PENETRATION 411

For a given pressure difference, we observe that the penetra-


tion coefficient is almost unity for particles of 0.1–0.5 µm when
the crack height is 0.25 mm, and for particle of 0.02–7 µm when
the crack height is 1 mm. Particles outside of this size range
present lower penetration coefficient as they are expected to
deposit on crack surfaces by means of Brownian diffusion (for
finer particles) or gravitational settling (for coarser particles).
For crack height d = 1 mm, most of the particle penetration co-
efficients are greater than 0.9, while for d is smaller than 1 mm
(d = 0.25 mm), penetration varies significantly. Compared the
results under different crack lengths, we can see that penetration
is significantly reduced in the longer cracks for many particle
sizes. Figures 2 and 4 present the predicted penetration for the
straight-through cracks under various pressure difference, P
= 4 and 10 Pa, respectively. The results indicate that penetra-
FIG. 3. Comparison of model predictions with experimental data for alu- tion coefficient increases under larger pressure difference. For
minum cracks for case 2. instance, the penetration coefficient of 0.03 µm particles with
P = 10 Pa is about 0.81, but only 0.65 with P = 4 Pa.
Figures 3 and 5 also show similar observation that the pene-
tration coefficient of 5 µm particles with P = 10 Pa is about
0.82, but only 0.63 with P = 4 Pa. These results show that
the influence of the key parameters, such as particle diameters,
crack height and length and pressure difference on particle
penetration coefficient were captured appropriately by the three
approaches.
Inferring from the results, it could be found that the results
predicted by the three approaches generally agree well with
the experimental results within a certain range of particle sizes,
while some of them show relative larger errors for particle sizes
outside the range. Eulerian model satisfied best for all studied
range of particle diameters among the three models. The results
of the analytical method presented in this paper and the results of
the simplest analytical method in Equation (4) (Liu and Nazaroff
2001) are very close to those of the Eulerian model. For coarse
FIG. 4. Comparison of model predictions with experimental data for alu-
minum cracks for case 3.
particles (dp > 1 µm), gravitational settling is dominant for
deposition; while Brownian diffusion controls the deposition
for fine particles (dp < 0.1 µm). The similarly in results are
not surprising as these approaches incorporated the main depo-
sition mechanisms when particles penetrating through a single
straight crack. Nevertheless, we consider that the Eulerian ap-
proach is more flexible as it can incorporate different deposition
mechanisms in cracks for different conditions. Figures 2 and
4 indicate that the predicted results by the Lagrangian model
agree worst for fine particles (dp < 0.1 µm). The most plau-
sible reason for the poor performance is that the Lagrangian
approach models a weaker effect of the Brownian diffusion.
To prove this, two typical analytical methods for modeling the
effect of Brownian diffusion (De Marcus and Thomas 1952;
Martonen et al. 1996) were included in the comparison for fine
particles, as shown in Figure 6. The results show that only those
modeled by the Lagrangian approach do not match the results of
the two analytical methods. The Brownian force function in the
FIG. 5. Comparison of model predictions with experimental data for alu- original Lagrangian model (Li and Ahmadi 1992) was replaced
minum cracks for case 4. by the function suggested by De Marcus and Thomas (1952)
412 B. ZHAO ET AL.

FIG. 6. Comparison of different methods for modeling the effect of Brownian FIG. 7. Comparison of the results of modified Lagrangian model with other
force for aluminum cracks for case 1 (fine particles). data for aluminum cracks for case 1 (fine particles).

in order to investigate the accuracy of the Brownian force in


the original Lagrangian model. The function suggested by De
Marcus and Thomas (1952) which represents the relationship
between penetration factor and diffusion was transformed into
the relationship between Brownian force and diffusion. Such
modification was implemented in FLUENT. As can be seen in
Figure 7, the results of modified Lagrangian model also do not
match the results of the measurements and the other models. The
problem of Lagrangian model lies on the approach itself instead
of the Brown force function. According to Fick’s law, Brown
diffusion depends on particle concentration gradient. However,
the information of particle concentration gradient cannot be
obtained when calculating the Brown force by Lagrangian ap-
proach. It should be pointed out that the calculation of particle
trajectory does NOT depend on the particle concentration gra-
dient, which may make the simulation of Brownian diffusion FIG. 8. Comparison of analytical model predictions with experimental data
for strand board cracks for case 5.
inaccurate. Similar conclusion was also reported by Robinson
et al. (1997) and they suggested that Fluent DPM was unable to
model molecular diffusion. This preliminary analysis deserves
to be further extended and studied.
Table 2 summarizes the comparisons of the three approaches
for particles in different range of sizes, and it includes all the
cases studied except for the case 5 where the inner surfaces
of the crack are rough, and this case will be analyzed sepa-
rately in the next paragraph. The analytical model costs mini-
mum computing resources as it is an analytical expression of
solving an ODE, while both the Eulerian and Lagrangian mod-
els need numerical simulation by discretizing the governing
equations into algebraic equations. The Lagrangian model costs
more computing resources than the Eulerian model as it re-
solves the air velocity distribution numerically, and the tracking
of a large number of particle trajectories will also cost many
resources.
Figures 8–10 show the comparison of the presented three FIG. 9. Comparison of Eulerian model predictions with experimental data for
approaches with the measurement of strand board case (Case 5), strand board cracks for case 5.
TABLE 2
Summary of the performance of the presented three approaches for the eight smooth inner surfaces cases
Standard Standard Averaged Standard Averaged Standard
Averaged deviation of Averaged deviation of the error deviation of error deviation of the
Computing error R̄(%)b the error σ c error R̄(%)b error σ c R̄(%)b the error σ c R̄(%)b error σ c
Approaches timea (dp < 0.4 µm) (dp < 0.4 µm) (0.4< dp < 1.2 µm) (0.4< dp <1.2 µm) (dp > 1.2 µm) (dp >1.2 µm) (All sizes) (All sizes)

Analytical Least (About 1s) 9.45 0.10 3.80 0.03 15.74 0.40 10.33 0.28
Eulerian Moderate (About 5 min for 9.07 0.11 2.58 0.02 10.30 0.18 3.85 0.14
30*401 grids)
Lagrangian Most (About 20 min for 9.82 0.12 6.68 0.04 18.36 0.40 12.51 0.28
50*500 grids)
a R

The computing time is based on the same personal computer with Intel CoreTM Duo CPU T2450 @ 2.00 GHz and RAM 1.0 GB DDR2.
b
Averaged
error is defined as:
R
R̄ = Np p ,
where Np is the number of studied particle sizes/diameters of all the cases; and Rp = (Ps − Pm )/Pm , where Ps and Pm are simulated and measured penetration coefficient for
certain particle diameter, respectively.
c
Standard
deviation of the error is defined as:

(Rp −R̄)2
σ = Np

413
414
TABLE 3
Comparison of the relative errors of the presented three approaches by treating the inner surfaces as smooth or rough for case 5
Standard Standard Averaged Standard Averaged Standard
Averaged deviation of Averaged deviation of the error deviation of error deviation of the
Rough or error R̄(%)b the error σ c error R̄(%)b error σ c R̄(%)b the error σ c R̄(%)b error σ c
Approaches smooth (dp < 0.4µm) (dp < 0.4 µm) (0.4< dp < 1.2µm) (0.4< dp < 1.2 µm) (dp > 1.2 µm) (dp > 1.2 µm) (All sizes) (All sizes)
Analytical Rough 22.75 0.13 2.53 0.02 260.93 3.37 101.04 2.36
Smooth 24.91 0.17 2.53 0.02 368.67 5.11 140.96 3.52
Eulerian Rough 36.94 0.35 3.48 0.03 71.81 0.33 37.45 0.40
Smooth 21.38 0.17 11.59 0.02 147.11 2.24 63.25 1.49
Lagrangian Rough 19.32 0.09 6.50 0.02 215.60 2.66 85.57 1.88
Smooth 28.00 0.29 3.85 0.03 319.23 4.87 124.45 3.28
b,c
The footnotes b and c are the same as shown in Table 2.
COMPARISON OF MODELS FOR PARTICLE PENETRATION 415

In reality the airflow from outdoor to indoor through the


building cracks fluctuates due to the unsteady of the pressure
difference created by the external wind pressure. The tempera-
ture gradient along the cracks may influence particle penetration,
which may be expected a dominant factor during summer and
winter seasons. The applicability of the presented approaches
for such complicated cases needs further exploration.
Another limitation of the analysis is related to the assump-
tion that particles are immediately removed once they deposit on
crack surfaces. However, in real situations, deposited particles
will accumulate inside cracks. The deposited particles are sub-
ject to resuspension when a burst of high differential pressure
occurs. Particle resuspension is a more complicated scenario and
it demands further study to advance our capability in predicting
pollutant penetration.
FIG. 10. Comparison of Lagrangian model predictions with experimental
data for strand board cracks for case 5.
5. CONCLUSION
We predicted the particle penetration coefficient through a
respectively. As there is no information about the height of straight crack in a building envelope by presenting three differ-
the effective roughness of the strand board (Liu and Nazaroff ent approaches, an analytical approach, an Eulerian approach
2003), we assumed this value instead of assigning to the mod- and a Lagrangian approach. The results were compared with
els. After checking the relative error calculated by the equation the published measured data by Liu and Nazaroff (2003). The
shown in Table 2 for different values of shifted distance, we conclusions drawn based on the presented results are as follows:
found that shifted distance of 32 µm for particle concentration
boundary layer yielded satisfactory results. The calculation re- (1) All the three models predicted particle penetration co-
sults of relative errors for using both the rough and smooth inner efficient well for particles in the range of 0.4–1.2 µm
surfaces treatments for case 5 are listed in Table 3. From the re- for the cases where the inner surfaces of the cracks
sults, we could find that after adjusting the boundary conditions are smooth, while the Eulerian model shows the best
to incorporate the “intercept” effect of roughness on particle agreement with the measurements.
deposition in the cracks, the results agree better with the mea- (2) Both the results of the analytical method presented in
surements for all the three approaches. Similar adjustment was this paper and the results of the simplest analytical
also reported in a previous study to account for the deposition method in Equation (4) (Liu and Nazaroff 2001) are
boundary for sandpapers (Lai 2005). Thus, we could predict very close to those prediction by Eulerian model.
penetration coefficient better by including the approaches pre- (3) The Lagrangian model cannot predict very well for par-
sented above. However, as the wall roughness is always irregular ticles smaller than 0.1 µm as the Lagrangian approach
with complex texture in reality, it is hard to deduce a universal models a weaker effect of the Brownian diffusion. To
rule for the shifted distance of velocity boundary layer theoret- accurately model Brownian diffusion is critical and
ically. Furthermore, the presence of roughness not only contri- challenging for Lagrangian model.
butes the “intercept” effect, but also changes the boundary layer (4) For cracks with rough inner surfaces, the results agree
of the airflow over the surfaces. Therefore, the relative errors for better with the measurement data after adjusting the
the rough inner surfaces cases are still relative large, as shown boundary conditions by shifting the inner surfaces a
in Table 3. certain distance virtually, which is expected to incor-
porate the intercept effect of wall roughness on particle
deposition.
4. DISCUSSION
All of the three approaches assume that the crack configura-
tion is straight-through. For real cracks, the irregularity of the NOMENCLATURE
crack shape may have a significant influence on the penetra- a air exchange rate (h−1 )
tion coefficient. Thus, we have to set up a much more practical C concentration of the particle (number/cm3 )
model, like a L-shaped model or a double-bend model measured Cc Cunningham correction factor
by Liu and Nazaroff (2003), as surrogates of leakage paths for Ccm flow rate-weighted average concentration at the outlet
building envelopes. In fact more experimental/measured data in of the crack (number/cm3 )
practical conditions are needed to validate the models to cover Cmax particle concentration at the centerline of the straight
a wider range of environmental conditions/parameters. crack (number/cm3 )
416 B. ZHAO ET AL.

C(Xz , Y ) particle concentration distribution at the outlet of the λ the mean free length of the air molecule; λ =
cracks (number/cm3 ) 0.0667 µm at a temperature of 20◦ C and atmospheric
C∞ particle concentration outside the boundary layer pressure
(number/cm3 ) µ molecular dynamic viscosity of the fluid (g cm–1 s–1 )
d height of the crack (cm) ρ fluid density (g cm–1 )
dp particle diameter (µm) ρp density of the particle (g cm–1 )
D Brownian diffusivity of the particle (m2 s–1 ) τp particle relaxation time (s)
Fb Brownian motion (N)
Fin infiltration factor
Fx additional acceleration (force/unit particle mass) REFERENCES
De Marcus, W., and Thomas, J. W. (1952). Theory of a Diffusion Battery. US
term (N)
Atomic Energy Commission, Oak Ridge National Laboratory, Oak Ridge,
h vertical distance from the midline of the crack to the TN.
ceiling and floor (mm) EPA. (2005). Review of the National Ambient Air Quality Standards for Par-
i a parameter to describe the direction of the wall sur- ticulate Matter: Policy Assessment of Scientific and Technical Information,
faces, i.e., for an upward facing horizontal surface OAQPS Staff Paper.
FLUENT Inc. (2005). Fluent 6.2 User’s Guide. Fluent Inc., Lebanon, NH.
(floor), i = 1; for a downward facing horizontal sur-
Fuchs, N.A. (1964). The Mechanics of Aerosols. Pergamon, New York.
face (ceiling), i = −1; for a vertical surface, i = Hinds, W. C. (1982). Aerosol Technology: Properties, Behavior, and Measure-
0 ment of Airborne Particles. Wiley, New York.
k particle decay rates (h–1 ) Lai, A. C. K. (2005). Modeling Indoor Coarse Particle Deposition onto Smooth
kB Boltzmann constant; kB = 1.38 × 10–23 J/K and Rough Vertical Surfaces. Atmos. Environ. 39:3823–3830.
Li, A., and Ahmadi, G. (1992). Dispersion and Deposition of Spherical Parti-
n vertical distance from the particle to the crack inner
cles from Point Sources in a Turbulent Channel Flow. Aerosol Sci. Technol.
surfaces 16:209–226.
Nescape number of the escaped particles at the outlet Licht, W. (1980. Air Pollution Control Engineering. Marcel Dekker, Inc., New
Ntotal number of the total particles released at the crack York.
inlet Liu, D. L., and Nazaroff, W. W. (2001). Modeling Pollutant Penetration Across
Building Envelopes. Atmos. Environ. 35:4451–4462.
P penetration coefficient
Liu, D. L., and Nazaroff, W. W. (2003). Particle Penetration through Building
Pd particle penetration coefficient due to Brownian dif- Cracks. Aerosol Sci. Technol. 37:565–573.
fusion alone Martonen, T., Zhang, Z., and Yang, Y. (1996). Particle Diffusion with Entrance
Pe Peclet number Effects in a Smooth-Walled Cylinder. J. Aerosol Sci. 27:139–150.
Pg particle penetration coefficient due to gravitational Mathworks Inc. (2007). MATLAB R2007a User’s Guide. Natick, MA.
Middleman S. (1997). An Introduction to Mass and Heat Transfer: Principles
settling alone
of Analysis and Design. John Wiley & Sons, New York.
Pi particle penetration coefficient due to impaction Richardson, L. F. (1910). The Approximate Arithmetical Solution by Finite
alone Differences of Physical Problems Involving Differential Equations, with an
T absolute temperature of the air (K) Application to the Stresses in a Masonry Dam. Phil. Trans. Royal Soc. London
u airflow velocity in x direction (m s–1 ) 210(A):307–357.
um air average velocity (m s–1 ) Roache, P. J. (1994). Perspective: A Method for Uniform Reporting of Grid
Refinement Studies. ASME J. Fluids Engineer. 116:405–413.
umax air velocity at the centerline of the straight crack (m Robinson, R. J., Snyder, P., and Oldham, M. J. (1997). Comparison of Parti-
s–1 ) cle Tracking Algorithms in Commercial CFD Packages: Sedimentation and

u velocity vector of fluid phase (m s–1 ) Diffusion. Inhal. Toxicol. 19:517–531.
p
u velocity vector of the particle (m s–1 ) Taubee, D. B., and Yu, C. P. (1975). Simultaneous Diffusion and Sedimentation
v airflow velocity in y direction (m s–1 ) of Aerosols in Channel Flows. J. Aerosol Sci. 6:433–441.
Tian, L., Zhang, G., Lin, Y., Yu, J., Zhou, J., and Zhang, Q. (2009). Mathemati-
vd particle deposition velocity (m s–1 ) cal Model of Particle Penetration Through Smooth/Rough Building Envelop
vs gravitational settling velocity of particles (m s–1 ) Leakages. Build. Environ. 44:1144–1149.
w airflow velocity in z direction (m s–1 ) Wu, J., and Zhao, B. (2007). Effect of Ventilation Duct as a Particle Filter. Build.
x horizontal axis Environ. 42:2523–2529.
Zhao, B., Chen, C., and Tan, Z. (2009). Modeling of Ultrafine Particle Dispersion
y absolute vertical distance from the ceiling or floor
in Indoor Environments with an Improved Drift Flux Model. J. Aerosol Sci.
surface of the crack (mm) 40:29–43.
y0 height of the rough layer (µm) Zhao, B., and Wu, J. (2006). Modeling Particle Deposition onto Rough Walls
z flow path distance (or length) along the leakage path in Ventilation Duct. Atmos. Environ. 40:6918–6927.
(cm) Zhao, B., Zhang, Y., Li, X., Yang, X., and Huang, D. (2004). Comparison of
Indoor Aerosol Particle Concentration and Deposition in Different Ventilated
Rooms by Numerical Method. Build. Environ. 39:1–8.
Zhu, Y., Hinds, W. C., Krudysz, M., Kuhn, T., Froines, J., and Sioutas, C.
Greek Symbols (2005). Penetration of Freeway Ultrafine Particles into Indoor Environments.
ε deposition ratio J. Aerosol Sci. 36:303–322.

You might also like