Accepted Manuscript CFD modeling of particle dispersion and deposition coupled with particle dynamical models in a ventilated room Guangping Xu, Jiasong Wang PII:
S1352-2310(17)30468-5
DOI:
10.1016/j.atmosenv.2017.07.027
Reference:
AEA 15441
To appear in:
Atmospheric Environment
Received Date: 5 December 2016 Revised Date:
14 July 2017
Accepted Date: 16 July 2017
Please cite this article as: Xu, G., Wang, J., CFD modeling of particle dispersion and deposition coupled with particle dynamical models in a ventilated room, Atmospheric Environment (2017), doi: 10.1016/ j.atmosenv.2017.07.027. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
1
ACCEPTED MANUSCRIPT 1
CFD modeling of particle dispersion and deposition coupled with particle dynamical
2
models in a ventilated room
3 Guangping Xu, Jiasong Wang*
5 6
School of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China
RI PT
4
7
Abstract
9
Two dynamical models, the traditional method of moments coupled model (MCM) and Taylor-series
10
expansion method of moments coupled model (TECM) for particle dispersion distribution and
11
gravitation deposition are developed in three-dimensional ventilated environments. The turbulent
12
airflow field is modeled with the renormalization group (RNG) k–ε turbulence model. The particle
13
number concentration distribution in a ventilated room is obtained by solving the population balance
14
equation coupled with the airflow field. The coupled dynamical models are validated using
15
experimental data. A good agreement between the numerical and experimental results can be
16
achieved. Both models have a similar characteristic for the spatial distribution of particle
17
concentration. Relative to the MCM model, the TECM model presents a more close result to the
18
experimental data. The vortex structure existed in the air flow makes a relative large concentration
19
difference at the center region and results in a spatial non-uniformity of concentration field. With
20
larger inlet velocity, the mixing level of particles in the room is more uniform. In general, the new
21
dynamical models coupled with computational fluid dynamics (CFD) in the current study provide a
22
reasonable and accurate method for the temporal and spatial evolution of particles effected by the
23
deposition and dispersion behaviors. In addition, two ventilation modes with different inlet velocities
24
are proceeded to study the effect on the particle evolution. The results show that with the ceiling
25
ventilation mode (CVM), the particles can be better mixed and the concentration level is also higher.
26
On the contrast, with the side ceiling ventilation mode (SVM), the particle concentration has an
27
obvious stratified distribution with a relative lower level and it makes a much better environment
28
condition to the human exposure.
AC C
EP
TE D
M AN U
SC
8
*Corresponding author. Tel.: +86 21 34205311; fax: +86 21 34205311 E-mail address:
[email protected] (J.S.Wang).
2
ACCEPTED MANUSCRIPT 29
Keywords: Computational fluid dynamics; dynamical model; indoor particles; deposition;
30
ventilation
31
1. Introduction In recent years, air pollution is one of the most social concerns especially in developing countries.
33
When people in daily life spend most of their time in different indoor environments, the air quality
34
becomes the focus of our attention (Steinleet al. 2013; Matz et al. 2014). Aerosol particles are one of
35
the main components in indoor air pollutants. These suspended particles in the air, such as dust,
36
smoke, and fumes strongly impact the human respiratory system and cause kinds of certain sick
37
syndromes. Epidemiological studies have shown that the ultrafine particle number concentration
38
associates strongly with morbidity and mortality (Lippmann et al. 2014; Mataloni et al. 2016).
39
Therefore, the understanding of evolution rule of the particulate matters appears to be particularly
40
important.
M AN U
SC
RI PT
32
The evolution of particles is effected by various dynamical behaviors such as deposition,
42
coagulation, and nucleation. Generally in indoor environments with the existence of relative low
43
particle concentration, deposition and dispersion are the main dynamical behavior which dominate
44
the evolution of particles. Many studies on particle deposition have been conducted in the last few
45
decades. In the work of Lai and Nazaroff (2000), the particle deposition velocities at vertical and
46
horizontal surfaces were acquired, which have been applied in many studies about particle deposition
47
as a good reference. Nazaroff (2004) then concentrated on summarizing different modeling
48
techniques to deal with indoor particle dynamics including deposition, indoor source emission,
49
dispersion, and filtration. More recently, the researchers studied the characteristics of fine and
50
ultrafine particles deposition in different real indoor environments (El et al. 2008; Laiman et al.
51
2014). However, these reviews and researches are based on the assumption that the air condition of
52
room and particles is well-mixed, which is hardly satisfied in real conditions. The particles
53
dispersion and deposition are always effected by turbulent flow significantly and result in spatial
54
non-uniformity. For this reason, it is essential to couple with air flow field to consider the spatial
55
distribution of particles.
AC C
EP
TE D
41
3
ACCEPTED MANUSCRIPT Computational fluid dynamics (CFD) is a strong useful tool applied to predict various flow fields.
57
It has been used for studying particle dispersion and spatial distribution with different models
58
controlling particle evolution. Theoretically, there are two approaches used extensively to estimate
59
the process of particle transport coupled with CFD models: the Lagrangian and the Eulerian methods.
60
In the Lagrangian method, the particles are treated as discrete phase and the dynamics of single
61
particle is focused by trajectory approach. The motion equations on account of various forces
62
exerting on an individual particle is solved to acquire the history information of every single particle.
63
On the other hand, the Eulerian method treats particles as a continuum. The particle concentration
64
field is obtained by solving governing equations derived from the species conservation condition.
SC
RI PT
56
In particle dynamics studies, each method has its own applicable and appropriate situations.
66
From one perspective, the Eulerian method can easily incorporate the particle diffusion effect since
67
the randomness of the particle phase is included in the diffusion term of governing equation. From
68
the aspect of numerical simulation cost, the computational requirement of the Eulerian method is
69
obviously lower than that required by the Lagrangian method. Among the Eulerian approaches, the
70
method of moment (MOM) is extensively adopted in literatures studying particle dynamics.
71
Settumba and Garrick (2004) coupled MOM with CFD code to study the coagulation behavior in
72
temporal mixing layers. The same method was also applied to study the planar jet flow (Yu et al.
73
2006) and the poiseuille flow (Song and Lin 2008). More recently, Yu et al (2008) extended a new
74
moment method called Taylor-expansion moment method (TEMOM), which could give more
75
reasonable and accurate simulation results. This new method was then applied on some particle
76
dynamics behaviors including coagulation (Lin et al 2010; 2012), nucleation (Yu et al 2009) and
77
breakage (Gan et al 2010) coupled with different flow fields. However, there is little research about
78
the two moment methods on the study of particle deposition in indoor environments as we have
79
known.
AC C
EP
TE D
M AN U
65
80
The approaches proposed in the previous studies could only be valid for the specific issue such as
81
the deposition mechanism. However, these methods are not suitable for relative high concentration
82
environments that the particles are dominated not only by deposition but also by coagulation
83
behavior. Therefore, it is very limited to predict the particles evolution especially in more
84
complicated indoor environments. For this reason, it is essential to find another way with more
4
ACCEPTED MANUSCRIPT 85
extensive applicability that could be applied in more real-world conditions. The moment method
86
used in this study has the ability to predict particles evolution under multiple dynamic behaviors
87
coupled with flow fields. Compared with the previous Eulerian methods used in the study of deposition, the method of
89
moment could easily extend to other particle dynamics, as well as provide more evolution laws about
90
not only particle concentration but also geometric diameter and polydispersity as concerned in
91
real-world conditions that existing more than one dynamic behavior simultaneously. In many indoor
92
environments as described in the previous studies such as Elghobashi (1994) and Zhao et al. (2004),
93
with the assumption that the particle sources are all from the outdoor with lower volume fraction of
94
particles, as well as there are no particles in the initial indoor environment condition, the deposition
95
and dispersion are the main influence factors on indoor particles evolution. When the particle
96
concentration in an indoor environment is much higher, the coagulation behavior greatly affect the
97
particles evolution, and both the coagulation and deposition behaviors should be considered together.
98
However, in other more common indoor environment cases that there exist some particle emission
99
sources indoors or lower ventilation rates, the particles evolution is not only dominated by deposition
100
but also by other dynamical mechanisms such as coagulation and condensation. For example,
101
Glytsos et al. (2010) monitored ultrafine particles (UFP) number concentrations due to several indoor
102
sources and indicated that strong coagulation effects shift the particle size distribution to larger sizes
103
during candle burning, smoking, onion frying, and hair dryer operation. Similarly, Rim et al. (2016)
104
also validated that the coagulation effect can be significant during the source emission due to high
105
concentration and high mobility of nano-size particles when there exist indoor UFP sources: an
106
electric stove, a natural gas burner, and a paraffin wax candle. In addition, in the study of Rivas et al.
107
(2015), the authors had given a result that the coagulation and condensation processes might affect
108
UFP indoors when investigating the correlation of indoor and outdoor particle concentrations. The
109
higher concentrations of indoor-generated particles were observed in school rooms when closed
110
windows hindered dispersion (cold season). This means the low ventilation rate condition indoors
111
could also make the particle concentration higher with time elapses and the effect of coagulation
112
needs to be considered simultaneously. In addition, Furthermore, when existing in a flow field with
113
high turbulence intensity, the particles are dominated by both coagulation and breakage. Besides,
AC C
EP
TE D
M AN U
SC
RI PT
88
5
ACCEPTED MANUSCRIPT when emitted from the tailpipe of motor vehicles, the particles are dominated by deposition,
115
coagulation, condensation, and nucleation together. These complicated conditions could not be
116
solved by the common Eulerian or Lagrangian model. On the contrary, the method for deducing the
117
new models is easy to extend to other mechanisms and has the ability to deal with these situations.
118
Previous studies have done some work with respect to coagulation or nucleation coupled with flow
119
field. However, there is no related studies, as we have known, on the interaction between the
120
deposition and flow field by using this method. This study could provide a basis for studying the
121
particles evolution under multiple dynamics behaviors in different type of flow fields.
RI PT
114
In this study, it is aimed to develop new dynamical models coupled with CFD by using the two
123
moment methods to the study particle transport and distribution in ventilated indoor environments.
124
The main work and the challenge is to transform the mathematical equation describing the deposition
125
behavior by the two moment methods into the proper form that could be coupled with turbulence
126
models and be solved numerically. The two dynamical models are argued for their applicability and
127
accuracy when simulating particle indoor dispersion. A three dimensional ventilated room with
128
detailed measured data reported in the literature is selected as the case for comparisons, and the
129
simulated results by the two models with measurement are discussed. Moreover, the particle
130
concentration in a simulated room with some lay outs and human body is studied to see the effects of
131
different ventilation modes and inlet velocities on the particles temporal and spatial evolution.
132
2. CFD modeling
133
2.1 The turbulent air flow modeling
AC C
EP
TE D
M AN U
SC
122
134
In ventilated indoor air conditions, airflow is often considered to be turbulent and particles are
135
much easier to transport and disperse affected by turbulence. In order to analyze the particle transport
136
process, it is essential to know the airflow field in which particles are suspended. The air phase here
137
is treated as a continuum and the Eulerian method is used in the numerical analysis. The interaction
138
between fluid phase and particle phase is assumed to be one way, and the impact from particle phase
139
to fluid phase is negligible. This assumption is feasible since particle volume fraction is sufficiently
140
small. Among Eulerian methods, the k–ε two-equation turbulence models are computationally
6
ACCEPTED MANUSCRIPT efficient and stable compared to more complicated Reynolds stress models. Compared to the
142
standard k–ε model, the renormalization group (RNG) k–ε turbulence model has its own advantage
143
on simulating turbulent flows. The coefficients of the RNG model are derived theoretically rather
144
than experimental fitting adopted in the standard model. The dissipation rate equation has an
145
additional strain term. The strong anisotropy at regions occurring large shear can also be considered
146
in the RNG model. Wang (2009) has also showed that the RNG k–ε turbulence model has much
147
better ability of predicting the vortex structure and flow separation characteristics than the standard
148
k–ε model. It is more proper to use RNG k–ε model for simulating indoor airflow and the previous
149
researches have also showed better agreements between simulated results and experiment data
150
compared to other turbulence models (Posner et al 2003; Kuznik et al 2007). For these reasons, the
151
RNG k–ε model is selected to predict the incompressible turbulent airflow in a room.
M AN U
SC
RI PT
141
The general form of the governing equations can be generalized as:
152
∂ + ∂
=
+
, 1
represents the independent variables: time-averaged velocity components u, v, w, turbulent
153
where
154
kinetic energy k, and dissipation rate of turbulent kinetic energy ε. When
155
represents the conservation of mass.
156
and
157
source term of the general equation. The modeled equations for the turbulent kinetic energy and its
158
dissipation rate are written as:
TE D
is the velocity component in the coordinate
is the effective diffusion coefficient for each dependent variable
EP
is time.
AC C
+
+
159
where
160
for
=
∙
=
∙
is the effective kinetic viscosity,
+ and
+ !
'
+
where ν% is the turbulent viscosity and is written as:
'
direction, .
is the
− , 2 −
"
"
−
, 3
are the inverse effective Prandtl numbers
and , respectively. The generation term of turbulence kinetic energy = ν% &
161
is unity, the equation
(
'
is expressed as:
, 4
7
ACCEPTED MANUSCRIPT ν% = is given by = 163
*-
1 − -⁄-/ 1 + 1-.
.
"
, 6
RI PT
And the expression for the
. 5
where -=
3&
'
*,
+
'
,
,
(
' !
4
!⁄"
"
164
The constants in the RNG k–ε model,
165
1.393, 1.393, 1.42, 1.68, 4.38 and 0.012 (Yakhot et al. 1992).
, -/ , and 1, are specified as 0.0845,
M AN U
,
. 7
SC
162
*
"
The air is assumed to be incompressible and isothermal. The turbulent model is solved by the
167
form of discretized algebraic equations. The discretization scheme used for all variables except the
168
pressure term is the second-order upwind scheme, and for pressure the standard scheme is selected.
169
The transient formulation is solved by the first-order implicit scheme. Finally, this study used the
170
SIMPLE algorithm to couple pressure and velocity fields.
171
2.2 The particle phase modeling
TE D
166
The transport of ultrafine particles dispersed through the fluid is governed by the general
173
dynamics equations (GDE). The GDE describes the particle dynamics under the effect of different
174
physical and chemical processes: advection, diffusion, deposition, coagulation and other dynamical
175
behaviors (Friedlander, 1977). In this study, only the deposition behavior is considered, and the
176
dynamics equation is written as:
AC C
EP
172
∂6 + ∂
6
=
∙
78 + 7%
6
+
96 , 8 9
177
where 6 is the particle number concentration,78 is the Brownian diffusion coefficient, which is
178
given by Stokes-Einstein expression: 78 =
;< = , 9 3>?9@
8
ACCEPTED MANUSCRIPT 179 180
in which ;< is the Boltzmann constant, = is the air absolute temperature, ? is the gas viscosity, 9@ is the particle diameter, and
is the slip correction factor, which is represented as follows:
= 1+
0.43359@ 2B 31.264 + 0.418exp &− (4, 10 9@ B
where B is the mean free path length of the gas molecules.7% is the turbulent diffusivity, which is
182
given by: 7% =
G
?
H%
RI PT
181
, 11
where ?
184
Schmidt number, which is selected to be 0.7 (Li and Stathopoulos 1997;Tominaga and Stathopoulos 2007). And the source term, 96⁄9 is given by
H%
is the turbulent
M AN U
185
is the dynamic viscosity of fluid, G is the fluid density, and
SC
183
96 = −1 I ∙ 6 I, , 12 9
186
where I is the particle volume, and 1 I is the deposition rate or the wall loss rate, which is given
187
by:
K! ⁄ .
TE D
1 I = J ∙ I K"
188
where J = 4>⁄3
189
the mean free path length of the gas molecules;
1.664B
8 = ⁄6>?
K! ⁄
EP
" K! ⁄.
8
+ L ∙ I "⁄. , 13
Q MN ∙ ON6 >⁄N ∙ ∙ P
⁄>R S, in which B is
is the Boltzmann constant; = is the absolute
190
temperature; ? is the gas viscosity; N is an empirical value which is selected to be 2.6 (Cheng 1997;
191
Park et al. 2001);
AC C
192
is coefficient of the eddy diffusivity;
is the total surface area of the enclosed
chamber; R is the volume of the chamber. And L = 3⁄4>
"⁄.
2GT⁄9?U , where G is the
193
particle density; U is the height of the chamber. In Eq. (13), the two terms on the right hand refer to
194
the diffusion term and the gravitation term respectively. Eq. (12) is not able to be solved directly by
195
numerical methods, and it has to be changed into other forms. Here, a moment method is introduced
196
to describe the particle field in time and space. The kth order moment V of the particle distribution
197
is defined as: X
V = W I 6 I 9I . 14 /
9
ACCEPTED MANUSCRIPT
199
By multiplying both sides of Eq. (2) with I
applying the definition of moment, the transport equation for the moment V can be expressed as: ∂V + ∂
200
after substituting Eq. (13) into Eq. (12) and
V
=
V
& 78 + 7%
where 9V ⁄9 = −MJ ∙ VK"
(+
9V , 9
= 0,1,2 15
+ L ∙ V"⁄. S is the source term due to the deposition effect.
K! ⁄ .
RI PT
198
The first three moment variables V/ , V! , and V" are used to describe the particle properties,
202
representing the particle number concentration, the volume concentration and the scattered level,
203
respectively. Here, the two methods of dealing with the source term are given as below.
205
The first method proposed by Park and Lee (2000) is making an assumption that the initial particle number concentration follows lognormal form as: 6 I,
206
in which Y
=
M AN U
204
SC
201
Y
3I√2>[6\]
^ _ 3−
[6" MI⁄I] 18[6" \]
S
4, 16
is the total particle number concentration at time , I]
is the geometric mean
207
volume, and \]
208
in the moment equations can get closed. According to the prior proposed method of moment (MOM),
209
the source term can be converted into the first three moments form and the final expression of the
210
traditional method of moments coupled model (MCM) can be expressed as:
V!
∂V" + ∂t
TE D
∂V! + ∂t
=
a 78 + 7%
EP
V/
AC C
∂V/ + ∂t
is the standard deviation. Consequently, the fractional moment variables existed
V"
=
a 78 + 7%
=
a 78 + 7%
V/
V!
⁄
⁄
⁄d
b − J ∙ V/!c d V!Ke d V"" ⁄
⁄
⁄
⁄
⁄
⁄
⁄
⁄
⁄d
− L ∙ V/K! d V!g d V"g d , 18
⁄d
− L ∙ V/g d V!K!h d V""/ d . 19
b − J ∙ V/" d V!f d V"K!
V"
⁄
− L ∙ V/" d V!f d V"K! d , 17
b − J ∙ V/K! d V!g d V"g
⁄
⁄
⁄
⁄
211
Another approach of treating the fractional moment variables is using the Taylor-series expansion
212
method of moment, which is proposed by Yu et al. (2008) for studying the coagulation behavior. This
213
approach has no prior requirement for the particle size spectrum, and the limitation inherent in the
214
lognormal distribution theory automatically disappears. This advantage has greatly expanded the
10
ACCEPTED MANUSCRIPT 215
feasibility of the moment method in real environment conditions. However, this new approach has
216
not been used in studying the deposition behavior of fine particles as far as we have known. Here, the
217
coupled dynamics model is reestablished to study the distribution of particles in a ventilated room
218
condition. In the source term, the first three moment equations are written as:
RI PT
219
9V/ = −MJ ∙ VK!h⁄.d + L ∙ V"⁄. S, 20 9 9V! = −MJ ∙ VK".⁄.d + L ∙ Vg⁄. S, 21 9
SC
9V" = −MJ ∙ VKh"⁄.d + L ∙ Vf⁄. S, 22 9
221
in which the fractional variables are treated by Taylor-series expansion. In Eq. (14), I expanded about point I = _ as follows: I = _ + _
222
"
−
I−_ +
K!
2
_
I
K" "
_
+ −_
K"
K! "
TE D
=
M AN U
220
−1 I−_ 2
+ 2_
K!
Then substituting Eq. (17) into Eq. (8), it is obtained as V =
"
2
2−3 +
"
V/ + _
K!
2 −
"
V! +
"
I+
_
K"
2
is first to be
"
−3 2
"
−
_
. 23
V" , 24
and the expansion point _ is defined as the mean particle size V! ⁄V/ (Yu et al 2008). By applying
224
Eq. (24), the fractional moment variables in Eqs. (20) to (22) can be eliminated and substituted by the first three moments V/ , V! , and V" as follows:
AC C
225
EP
223
VK!h⁄.d = VK".⁄.d = VKh"⁄.d =
1081 gg⁄.d K!h⁄.d 440 dc⁄.d Kdc⁄.d V V! + V V! V" , 25 1521 / 1521 / 1705 !h⁄.d ".⁄.d 184 gg⁄.d Kgg⁄.d V V! − V V! V" , 26 1521 / 1521 /
2234 K".⁄.d Kh"⁄.d 713 !h⁄.d K!h⁄.d V V! + V V! V" , 27 1521 / 1521 /
1 ⁄ 10 ⁄ ⁄ ⁄ V"⁄. = − V/c . V!Kc . V" + V/! . V!" . , 28 9 9
11
ACCEPTED MANUSCRIPT Vg⁄. = Vf⁄. =
5 !⁄. K!⁄. 4 ⁄ ⁄ V/ V! V" + V/K" . V!g . , 29 9 9
20 K"⁄. "⁄. 11 ⁄ ⁄ V/ V! V" − V/Kg . V!f . . 30 9 9
Then the final form of the Taylor-series method of moments coupled models (TECM) can be written
227
as:
∂V! + ∂t
V!
a 78 + 7%
V/
1081 gg⁄.d K!h⁄.d 440 dc⁄.d Kdc⁄.d V V! + V V! V" 1521 / 1521 /
b−J∙
1 ⁄ 10 !⁄. "⁄. ⁄ −L ∙ − V/c . V!Kc . V" + V V! , 31 9 9 / =
a 78 + 7% −L ∙
V" =
a 78 + 7%
V!
1705 !h⁄.d ".⁄.d 184 gg⁄.d Kgg⁄.d V V! − V V! V" 1521 / 1521 /
b−J∙
5 !⁄. K!⁄. 4 ⁄ ⁄ V/ V! V" + V/K" . V!g . , 32 9 9 V"
b−J∙
2234 K".⁄.d Kh"⁄.d 713 !h⁄.d K!h⁄.d V V! + V V! V" 1521 / 1521 /
TE D
∂V" + ∂t
=
SC
V/
M AN U
∂V/ + ∂t
RI PT
226
20 K"⁄. "⁄. 11 ⁄ ⁄ V/ V! V" − V/Kg . V!f . . 33 9 9
EP
−L ∙
Now, we can use the new MCM model of Eqs. (17-19) and the new TECM model of Eqs.
229
(31-33) coupled with the RANS equations of (1-8) to obtain the particle dispersion distribution and
230
gravitation deposition under indoor ventilated environments.
231
2.3 Boundary condition
AC C
228
232
In this study, the semi-empirical particle deposition model developed by Lai and Nazaroff (2000)
233
is selected to account for the boundary condition. This method is widely employed in studying
234
indoor particle deposition performance (Zhao et al. 2008; Guichard et al. 2014). The particle mass
235
flux at wall surface, ij , is calculated by
ij = Ik ∙
8
∙ G@ ∙ Jj , 34
12
ACCEPTED MANUSCRIPT 236
where Ik is the deposition velocity of particles;
8
is defined as the particle concentration at the
237
first near-wall cell, and the concentration in a small cell is considered to be uniform; and Jj is the
238
total area of each boundary surface.
241 242
surfaces were used in this study.
RI PT
240
In the study of Lai and Nazaroff (2000), the detailed expressions of Ik at different wall The deposition velocity at the vertical surface (IkX ), the upward horizontal surface (Ikl ), and
the downward surface (Ikk ) is expressed respectively as follows: ∗
IkX =
Ikk = ∗
; 35
Ip
X s
1 − exp q− lr∗ t
; 36
M AN U
Ikl =
n
SC
239
Ip
X s
exp q− lr∗ t − 1
, 37
is friction velocity, Ip is the gravitational settling velocity of particle, and I is an
in which
244
integral term which is shown as:
TE D
243
n = u3.64
H
"⁄.
.
v − w + 39x, 38
EP
M10.92 H K!⁄. + 4.3S 1 8.6 − 10.92 v = ln { | + √3 tanK! 3 K! 2 + 0.0609 √3 × 10.92 H M10.92
AC C
1 w = ln { 2
H
K!
H
K!⁄.
+ •€S
.
+ 7.669 × 10Kc • €
.
| + √3 tan
• € = 9@
∗⁄
K!
H
H
K!⁄.
K!⁄.
4, 39
2• € − 10.92 H K!⁄. 3 4, 40 ⁄ √3 × 10.92 H K! .
2 , 41
is the kinematic viscosity of air, 9@ is particle diameter. And the total deposition velocity
245
where
246
is expressed as: Ik =
IkX JX + Ikl Jl + Ikk Jk . 42 J
13
ACCEPTED MANUSCRIPT 247
The accuracy of the calculated deposition velocity by this equation has been validated by the study of
248
Zhao et al. (2008) that compared the result with the experimental data. Also, this method has been
249
widely used in indoor environment conditions.
250
The inlet profile of velocity is set to be uniform, which is consistent with the real condition of the
251
experiment of Chen et al (2006). The parameter k in the RNG k–ε turbulence model is set to 3.828×
252
10-8 m2/s2, which is determined by
253
the turbulent intensity, which is defined as n% = 0.16 •^‚
254
inlet). The parameter epsilon ( *
.⁄c .⁄"
„ 0.07ƒ , where
*
K!⁄f
RI PT
=
n% " ⁄2, in which u is the mean inlet velocity, and n% is
(•^‚ is the Reynolds number at the
) is set to 2.119 × 10-10 m2/s3, which is defined as
is an empirical constant specified in the turbulence model
SC
255
=3
(approximately 0.085), and ƒ is the relevant dimension of the inlet. The outlet boundary condition is
257
set to be outflow, which is usually applied in the uncompressible flow. And the particle concentration
258
at the outlet is set to zero according to the experimental setup.
M AN U
256
259
3. Results and discussion
261
3.1 Model validation for coarse particles
262
3.1.1 Modeling setup based on the experiment
EP
TE D
260
In order to validate the numerical model for particle transport and deposition in a room, the
264
experiment data including airflow and particle concentration measured by Chen et al (2006) is
265
selected as the comparison case. The model room geometry is 0.8 m length, 0.4 m width and 0.4 m
266
height. The inlet and outlet are both 0.04 m × 0.04 m at the center plane Y=0.2 m as shown in Fig. 1.
AC C
263
14
SC
RI PT
ACCEPTED MANUSCRIPT
267
Fig. 1 Schematic of the ventilated chamber used for experiments by Chen et al. (2006)
M AN U
268 269
The particle density is 1400 kg/m3, and its diameter is tested to be 10 µm. The particle
271
concentration is normalized by the inlet concentration (a lower concentration level that the particle
272
coagulation effect could be neglected) and set to be 1.0. Simulations were conducted for two inlet
273
velocities (Vinlet), 0.225 m/s and 0.45 m/s, respectively. A monitoring point at the center of outlet is
274
set up to estimate the time for the steady state of concentration field. Three sets of uniform structured
275
grids 16,000, 128,000 and 512,000 are tested for the transient simulation. The 128,000 (40×40×80)
276
grids for the velocity profile, the turbulent intensity and the particle concentration along the
277
x-direction of inlet center at the center plane shows a quite similar result to those of the 512,000 grids
278
as shown in Fig. 2. Hence, all the following simulations are based on the 128,000 grids to save the
279
simulation time. All the simulations are carried out by the ANSYS Fluent (version 17.0) software.
EP
AC C
280
TE D
270
281 (a) 282 283 284
(b)
15
ACCEPTED MANUSCRIPT 285
RI PT
(c)
287
SC
286
Fig. 2 Grids number independence test; (a) velocity; (b) turbulent intensity; (c) particle concentration
289
M AN U
288
3.1.2 Comparison of simulation results with the experiment of Chen et al. (2006) After simulated for about 400 s with the two inlet velocities as shown in Fig. 3, the concentration
291
value at the monitoring point has no longer changed, which could be considered to reach a steady
292
state for both flow and concentration fields to present the final results of the simulation. The mixing
293
time for inlet velocity 0.225 m/s is about 386 s and for inlet velocity 0.45 m/s is about 323 s. The
294
mixing time of higher inlet velocity case is significantly smaller. It is evident that higher ventilation
295
rate could provide better mixing efficiency.
EP
TE D
290
AC C
296
Vinlet=0.225 m/s
Vinlet=0.45 m/s
297 298
Fig. 3 Particle concentration variation at the monitoring points with time elapses for two inlet velocities 0.225 m/s
299
and 0.45 m/s
16
ACCEPTED MANUSCRIPT 300
The numerical results of the three-dimensional ventilation flow field are first presented. Fig. 4
302
shows the comparison of simulated x direction velocity by RNG k–ε model with the experimental
303
data for the case with Vinlet=0.225 m/s. The agreement between predicted and measured results is very
304
good although there exists a little difference. The result shows that the selected turbulent model is
305
appropriate to calculate the air flow field in a ventilated chamber. The particle concentration is
306
solved simultaneously based on the airflow field calculated by RNG k–ε turbulent model.
SC
EP
307
TE D
M AN U
Experiment Simulation
RI PT
301
Fig. 4 Comparison of measured and predicted x-direction velocities at three different locations of center plane
309
x=0.2m, 0.4 m, and 0.6 m (inlet velocity 0.225 m/s).
310
AC C
308
311
Fig. 5 shows the comparisons of simulated particle concentration by the MCM and TECM
312
models with the experimental data under two inlet velocities 0.225 m/s and 0.45 m/s at 400 s. In
313
contrast to the measured results obtained by Chen et al (2006), the simulated concentration
314
distribution shows reasonable agreement at three locations. In most positions, the simulated data are
315
close to the experiment data within the uncertainty range except at locations near the ceiling area.
316
The relative error to the measurement is acceptable. The discrepancy means that the simulation
17
ACCEPTED MANUSCRIPT 317
results overestimate the particle flux to the ceiling wall. It is supposed that this may be mainly caused
318
by the complicated turbulence effect on the deposition which is not accurately estimated. The MCM model essentially captured the characteristics of particle and in a reasonable
320
agreement with the experiment. It gives good predictions that the data is in the measured range of
321
deviation at most positions. Relative large discrepancies occur at the ceiling wall area, where the
322
particle concentration are overestimated. This may result from the complicated air flow condition
323
near the wall surface which has strongly effect on the particle concentration. And the concentration
324
on the boundary surface are also acquired by the semi-empirical model, which may also a possible
325
reason for the difference.
SC
RI PT
319
Compared with the MCM model, the TECM model gives a relative smaller simulation values at
327
almost all the locations. When the inlet velocity is 0.225 m/s, the difference is very close below
328
0.35m, and presents a little larger as well as more close to the measured data when the velocity is
329
0.45 m/s. At the locations above 0.35m, it also presents more accurate simulation results although
330
some discrepancies existed. The difference of results between MCM and TECM models may come
331
from the derivation of the moment equations. The MCM model is based on the assumption that the
332
particle size distribution is taken as lognormal, which is not consistent with this measurement setup.
333
This could make some inevitable mathematical errors during calculation process. On the other hand,
334
the TECM model is not included any assumption and this difference makes the discrepancy of the
335
final simulation results. On the whole, the two moment equation models could give a reasonable and
336
accurate simulation result about the particle concentration evolving with time and space. Compared
337
with the MCM model, the results calculated by TECM model are more close to the experiment data
338
although there still exists some discrepancies in several locations near the ceiling wall.
TE D
EP
AC C
339
M AN U
326
18
ACCEPTED MANUSCRIPT
SC
RI PT
(a)
M AN U
340
AC C
341
EP
TE D
(b)
342
Fig.5 Comparison of measured and simulated particle concentrations at three different locations of center plane; (a)
343
inlet velocity 0.225 m/s; (b) inlet velocity 0.45 m/s.
344
In order to make a quantitative comparison, the relative errors, defined as |
p
−
|/ , where
p
345
is the simulated value and
is the experimental value, at the nine monitoring points with two inlet
346
velocities are shown in Fig. 6. The numbers are set from 1 to 9, representing the points from the
347
bottom to the top. The simulation results with the two methods are compared with that of Chen et al.
348
(2006). When the inlet velocity is 0.225 m/s, the relative errors of MCM and TECM are smaller at
19
ACCEPTED MANUSCRIPT point 1 to 4. The mean error of MCM is 13.26%, which is a little smaller than the 16.27% of Chen et
350
al. (2006), and the mean error of TECM is 16.82%, which is almost the same with that of Chen et al.
351
(2006). Meanwhile, when the inlet velocity is 0.45 m/s, the relative errors of MCM and TECM are
352
smaller at most locations except at the point 9. The mean relative errors of MCM, TECM and Chen
353
et al. (2006) are 16.16%, 12.15% and 15.99%, respectively as shown in Table 1. This result shows
354
that the TECM performs better on simulating the deposition behavior with lager inlet velocity, and
355
the MCM model gives the same level of relative errors compared with the results of Chen et al.
356
(2006). It can be considered that the TECM model performs a more accurate result than the MCM
357
model, which gives a result with almost the same level of relative error compared with Chen et al.
358
(2006).
TE D
M AN U
SC
RI PT
349
359 360
Fig. 6 Relative errors of different models for the simulation
361
Table 1 Mean relative error between the simulated values and the experimental data
EP
362
Inlet velocity
Mean relative error (%) MCM
TECM
0.225 m/s
16.27
13.26
16.82
0.45 m/s
15.99
16.16
12.15
AC C
Chen et al. (2006)
363 364 365
Fig. 7 shows the concentration evolution profiles at the center plane simulated by the two
366
moment methods when the inlet velocity is 0.225 m/s. For both cases, the highest concentrations are
20
ACCEPTED MANUSCRIPT around the inlet area, as is treated to be the source of particles. Due to the effect of dispersion and
368
gravitation, the concentration gradually decreases along the x-direction at the centerline of inlet. The
369
concentration at the above part of the center plane is always higher than that at the below, where the
370
gravitational settling effect is more obvious than the turbulent dispersion. The particle number
371
concentrations have the similar distribution law with the two models. In the middle region, the MCM
372
model gives a little higher concentration distribution result compared with the TECM model. In
373
addition, the lowest particle concentrations are both at the left bottom of the center plane.
RI PT
367
Fig. 8 shows the same concentration condition when the inlet velocity is 0.45 m/s. Like the case
375
in Fig.5, the result calculated by MCM is relative larger than that by TECM especially at the region
376
below Y=0.3 m. Compared with the case when the inlet velocity is 0.225 m/s, it also presents that
377
with the inlet velocity increases or larger turbulent intensity exists, the particle mixing level is
378
enhanced, and the particle concentration distribution is more uniform.
TE D
M AN U
SC
374
379
Fig.7 Simulation result of particle concentration distribution at the center surface when inlet velocity is 0.225 m/s.
381
(a) MCM model; (b) TECM model.
AC C
382
EP
380
383 384
Fig.8 Simulation result of particle concentration distribution at the center surface when inlet velocity is 0.45 m/s. (a)
385
MCM model; (b) TECM model.
21
ACCEPTED MANUSCRIPT Fig. 9 shows the particle number concentration distribution in the whole flow field with two inlet
387
velocities by the TECM model. It is evident that the concentration at the right part is much higher
388
than that at the left part due to the air flow. The lowest concentration is at the corner near the floor
389
down the inlet. Meanwhile, the streamlines from the inlet are presented to show the vortex structures
390
in the simulated room. From the distribution of streamlines, it can be seen that a large vortex
391
structure exists in the center part of the room, which makes the particle concentration lower at the
392
center region of the vortex than that at the place around. This indicates that this type of flow structure
393
strongly affect the particle concentration distribution and produce a lower concentration region at the
394
center of the vortex. When the inlet velocity is 0.45 m/s, the effect of the vortices presents a similar
395
vortex structure, and a lower concentration region also appears in the center region. However, due to
396
the stronger turbulent dispersion, the concentration mixing level is promoted.
M AN U
SC
RI PT
386
AC C
398
EP
TE D
397
399
Fig. 9 Spatial distribution of particle concentration and streamlines (dashed lines with arrows) from the inlet in the
400
ventilated room. (a) Vinlet =0.225 m/s; (b) Vinlet =0.45 m/s.
401 402
3.2 Model validation for fine particles
403
For further validate the accuracy of the new models for fine particles, another experiment by Zhao
404
et al. (2009) is selected to make the comparison as shown in Fig. 10. In this experiment, the NaCl
405
particles with diameter of 101.8 nm are injected into the room at the inlet. Three grids sets of 3,6550,
22
ACCEPTED MANUSCRIPT 14,6200, and 58,4800 are tested and the set of 14,6200 grids is selected , which gives a relative good
407
result considering the numerical simulation cost. The simulated particle concentration at the 27
408
locations of the lines across the 9 places vertically as shown in Fig. 11 are compared with the
409
experimental data, and both the MCM and the TECM give good agreements. Similar situation with
410
the above experiment, the TECM performs much better at the location A, C, D and F. The mean
411
relative errors of MCM, TECM and Zhao et al. (2009) are also calculated, which are 16.16%, 12.15%
412
and 15.99%, respectively. It can be considered that the accuracy of the MOM and the model in the
413
reference are at the same level, and the TECM could give a result with relative less error. Therefore,
414
the two models could better simulate the particle deposition behavior in the flow field.
M AN U
SC
RI PT
406
415
AC C
EP
417
Fig. 10 Experiment set up by Zhao et al. (2009)
TE D
416
418 419
Fig. 11 Comparison of particle concentration between the simulation and the experimental data
23
ACCEPTED MANUSCRIPT 420 421
3.3 Characteristic of particle dispersion in different ventilation systems
422
3.3.1 Case description After the validations above, the present developed dynamical model TECM can be taken as a
424
reliable model used to study the particle dispersion and deposition characteristics under different
425
ventilation conditions in a hypothetic room, which is representative in daily life. The dimension
426
room is 5.0 m length, 3.5 m width, and 2.5 m height as shown in Fig. 12. Two ventilation modes
427
named ceiling ventilation mode (CVM) and side ventilation mode (SVM) are designed to supply air
428
flow by two inlets located at the ceiling wall and the side wall with particles and an exhaust at the
429
floor level. Two inlet velocities 0.3 m/s and 1.2 m/s are selected to provide low and high levels of air
430
flow. The particle diameter is set to 10 µm, and its density is 1400 kg/m3, which are the same set as
431
in the experiment above. In addition, the other sets for the turbulent flow field and the particle phase
432
are also the same as that in the above simulation. A monitoring point is set at the location (X=4.0,
433
Y=1.8, Z=1.2) near the head of the human body to simulate human exposure and evaluate the particle
434
concentration level. The wall surfaces are treated to be adiabatic and the temperature difference is
435
also neglected. Three grids sets of 3,1800, 12,7100, and 50,8400 are tested, and the set of 12,7000
436
could give a relative good result with consideration of numerical simulation cost.
SC
M AN U
TE D
EP AC C
437
RI PT
423
438
24
ACCEPTED MANUSCRIPT 439
Fig. 12 Configuration of the simulated office room with 5.0 m length (X), 3.5 m width (Y), and 2.5 m height (Z); the
440
dimension of the SVM inlet is 0.4 m × 0.4 m, the CVM inlet is 0.4 m × 0.4 m, and the outlet is 0.4 m × 0.3 m
441
respectively. The human body is simplified by a squre cylinder with 1.2 m height.
443
3.3.2 Particle distribution with CVM and SVM
RI PT
442
The temporal evolution of particle concentration at the monitoring point with two inlet velocities
445
under CVM and SVM conditions are shown in Fig. 13. With CVM, it takes about 11 minutes to
446
reach the final steady concentration for the inlet velocity of 0.3 m/s, and 19 minutes for the inlet
447
velocity of 1.2 m/s. It is evident that with a larger inlet velocity, the concentration increases more
448
quickly to the ultimate stable value. The required time to reach 90% of the final concentration for the
449
two inlet velocities are about 9.1 minutes and 5.2 minutes respectively, and the terminal
450
concentrations reach to the 79% and 65% levels of the supplied inlet concentration. This implies that
451
with CVM, the particles in the room are much easy to disperse to other places by the effect of
452
turbulent flow. When the inlet velocity increases, it changes the increase rate of particle
453
concentration, and makes a little lower level of the final concentration to the human exposure. It can
454
be considered that the CVM is not obviously reduce the particle concentration around the occupant
455
with an arbitrary inlet velocity. On the contrast, with SVM, the time to reach the final steady
456
concentration when the inlet velocity is 0.3 m/s is 7.9 minutes, and 2.9 minutes when the inlet
457
velocity is 1.2 m/s. Meanwhile, the final stable concentrations are 0.186 and 0.468 respectively.
458
Under the same inlet velocity, the time values for the two cases may depend on the inlet location.
459
However, the final stable concentrations under the two inlet velocities decrease by more than half,
460
and there also exists a large discrepancy between the two inlet velocities, which is very different with
461
the case of SVM. It implies that when the inlet velocity is larger, the particles are easier to disperse
462
than the deposit to the wall surfaces so that the particle concentration decreases effectively.
AC C
EP
TE D
M AN U
SC
444
25
RI PT
ACCEPTED MANUSCRIPT
463
Fig. 13 Temporal particle concentration at the monitoring point in the CVM and SVM cases with two inlet
465
velocities 0.3 m/s and 1.2 m/s.
M AN U
SC
464
Fig. 14 shows the particle concentrations at different height levels at one location at front of the
467
human body (X=4.1 m, Y=1.7 m). With CVM as shown in Fig. 11(a), the concentration varies in a
468
small range with different height levels. With 1.2 m/s inlet velocity, this difference is very slight.
469
This implies that the particles are in a well-mixed state in the ventilated room, and the effect of
470
dispersion dominated by the turbulent flow makes particles suspend in the air. The higher
471
concentration at Y=1.1 m height is resulted from the particle deposition on the desk where the
472
particles are not easy to disperse. Compared with the case of CVM, the vertical concentration with
473
SVM as shown in Fig. 11(b) presents obvious stratification, which may result from the particles
474
brought to the upper zone by the upward airflow, as well as the particles at the lower places deposit
475
to the floor due to the gravitational settling. For this reason, the particle concentration to the human
476
body (Y=1.2 m) is much lower than that with the CVM. With larger inlet velocity, the turbulent
477
airflow is more likely to make particles stay in the air and make a relative higher concentration level
478
as well.
EP
AC C
479
TE D
466
26
RI PT
ACCEPTED MANUSCRIPT
SC
480
Fig. 14 Vertical distribution of particle concentration at the seated human body under two ventilation modes. (a)
482
CVM and (b) SVM.
AC C
EP
TE D
M AN U
481
483 484
Fig. 15 Particle concentration distribution at the pla ne (Y=1.8) across the human body. (a) CVM when Vinlet is 0.3
485
m/s; (b) CVM when Vinlet is 1.2 m/s; (c) SVM when Vinlet is 0.3 m/s; (d) SVM when Vinlet is 1.2 m/s.
27
ACCEPTED MANUSCRIPT 486
Fig. 15 shows the particle concentration distribution at the plane across the human body with
488
CVM and SVM. With CVM (Fig. 15(a) and Fig. (15b)), the particle concentration at the places with
489
human body is much lower which is caused by the particles deposition on more surfaces that less
490
particles are dispersd by the air flow. On the contrast, with SVM (Fig. 15(c) and Fig. (15d)), the
491
concentration around the human body is much higher compared with the other places at the plane.
492
Some particles are dispersed and suspended to the places far away from the particle inlet by the air
493
flow, and the concentration level is dependent on the inlet velocity. Generally, due to the deposition
494
effect, the particle concentration near the floor is also higher than that at the upper places. Meanwhile,
495
there also exist some lower concentration regions at the middle places due to the vortex structure
496
induced by the turbulent flow. In addition, the average particle concentration at the outlet with SVM
497
is much lower than that with CVM as shown in Fig. 16. It is also proved that the particle deposition
498
effect on the evolution of number concentration is more evident with the SVM.
M AN U
SC
RI PT
487
0.8 0.7 0.5 0.4
Inlet velocity is 1.2 m/s
EP
0.3
TE D
0.6
Inlet velocity is 0.3 m/s
0.2
AC C
0.1 0
499 500
CVM
SVM
Fig. 16 Average particle concentration at outlet with CVM and SVM
501 502
Fig. 17 shows the spatial distribution of particle concentraiton with CVM and SVM under two
503
inlet velocities 0.3 m/s and 1.2 m/s. With CVM as shown in Fig. 17(a) and Fig. 17(b), the particle
504
concentration at the places around the human body is at a relative lower level than that at the other
505
half of the room. This difference is supposed to be caused by the exsistence of canbinet, desk and
28
ACCEPTED MANUSCRIPT human body which could increase the particle deposition chance so as to reduce the concentration,
507
which can also be illustrated in Fig. 15(a) and Fig. 15(b). In the core region of the vortex structures
508
presented by the streamlines in the figures, the particle concentration is relative lower than that at the
509
places around due to the turbulence agglomeration effect. With SVM as shown in Fig. 17(c) and Fig.
510
17(d), the concentration presents stratified obviously. At the places far away from the inlet, the
511
turbulent flow gets more weak that the deposition effect dominates the evolution of particles to make
512
the concentraiton decrease greatly. Meanwhile, the turbulent air flow takes less particles to the places
513
near the ceiling surfaces due to the flow field structure. As a result, the concentration level near the
514
floor surfaces is much higher than that at the upper region. In addition, the larger supplied air
515
velocity can imporve the mixing level with both the ventilation situations. It can be concluded that
516
under the same inlet velocity, compared with CVM, the SVM could get a better environment for the
517
seated person in the room with a relative lower particle concentration.
M AN U
SC
RI PT
506
519 520
AC C
EP
TE D
518
29
RI PT
ACCEPTED MANUSCRIPT
SC
521
Fig. 17 Spatial distribution of particle concentration with CVM and SVM. (a) CVM when Vinlet is 0.3 m/s; (b) CVM
523
when Vinlet is 1.2 m/s; (c) SVM when Vinlet is 0.3 m/s; (d) SVM when Vinlet is 1.2 m/s.
M AN U
522
524
525
4. Conclusions
In this paper, two moment based coupled dynamical models, MCM and TECM, coupled with
527
flow field are established to simulate spatial and temporal distribution of particle number
528
concentration dominated by deposition in a ventilated indoor environment. Both the effects of
529
dispersion and gravitation settling are considered during the process of particle evolution. The air
530
flow field is simulated by RNG k–ε turbulent model, and the result shows a good agreement with the
531
experiments. The numerical simulation of particle number concentration is also validated by the
532
measured data under two inlet velocities, 0.225 m/s and 0.45 m/s, as well as two diameters of 10 µm
533
and 0.1018 µm.
AC C
EP
TE D
526
534
According to the comparison with the two experimental results, the new models could better
535
simulate the particles evolution dominated by deposition and air flow with different particle sizes.
536
Compared with the MCM model, the TECM model gives relative smaller concentration values at
537
most locations, and it is shown that the results agree much better with the measured data. The
538
particle concentration calculated by both two models have a similar spatial distribution, and with a
539
large vortex structure, the particle concentration at the center region is much lower than that at the
30
ACCEPTED MANUSCRIPT 540
places around. Moreover, with larger inlet velocity or higher turbulence intensity, the mixing level is
541
more uniform. The discrepancy between the results acquired by the two moment models may result from the
543
different derivation of the moment equations, and the TECM model seems to give a better simulation
544
result. In general, the two moment dynamical models are found to be effective methods to study the
545
indoor particle transport and concentration distribution, although there still exist some errors at some
546
locations. Furthermore, another good feature of the moment method is the expansibility to deal with
547
other dynamical behaviors, like coagulation or nucleation, which may be considered simultaneously
548
for further applications.
SC
RI PT
542
The simulation results of concentration comparison between the two ventilated modes, CVM and
550
SVM, indicate that the supplied air with particle pollutants from the ceiling wall surface can produce
551
a more uniform and higher concentration field. Compared with CVM, the SVM produces a more
552
stratified, as well as a lower concentration level that the human exposure could be less serious. The
553
mixing level gets also higher with the inlet velocity increases for both CVM and SVM.
M AN U
549
TE D
554
References
556 557
Chen F, Simon C M, Lai A C K. Modeling particle distribution and deposition in indoor environments with a new drift–flux model[J]. Atmospheric Environment, 2006, 40(2): 357-367.
558 559
Cheng Y S. Wall deposition of radon progeny and particles in a spherical chamber[J]. Aerosol Science and Technology, 1997, 27(2): 131-146.
560 561
El Hamdani S, Limam K, Abadie M O, et al. Deposition of fine particles on building internal surfaces[J]. Atmospheric environment, 2008, 42(39): 8893-8901.
562
Elghobashi S. On predicting particle-laden turbulent flows. Applied Scientific Research[J], 1994, 52(4):309-329.
563 564
Friedlander S K. Smoke, dust and haze: Fundamentals of aerosol behavior[J]. New York, Wiley-Interscience, 1977. 333 p., 1977, 1.
565 566
Gan F J, Lin J Z, Yu M Z. Particle Size Distribution in a Planar Jet Flow Under-Going Shear-Induced Coagulation and Breakage[J]. Journal of Hydrodynamics, Ser. B, 2010, 22(4): 445-455.
567 568
Glytsos T, Ondráček J, Džumbová L, et al. Characterization of particulate matter concentrations during controlled indoor activities. Atmospheric Environment[J], 2010, 44(12):1539-1549.
569 570
Guichard R, Belut E, Nimbert N, et al. Evaluation of a moments-based formulation for the transport and deposition of small inertia aerosols[J]. The Journal of Computational Multiphase Flows, 2014, 6(4): 407-418.
AC C
EP
555
31
ACCEPTED MANUSCRIPT Jianzhong L I N, Peifeng L I N, Mingzhou Y U, et al. Nanoparticle transport and coagulation in bends of circular cross section via a new moment method[J]. Chinese Journal of Chemical Engineering, 2010, 18(1): 1-9.
573 574
Jianzhong L I N, Song L I U, Tatleung C. Nanoparticle migration in a fully developed turbulent pipe flow considering the particle coagulation[J]. Chinese Journal of Chemical Engineering, 2012, 20(4): 679-685.
575 576
Kuznik F, Rusaouen G, Brau J. Experimental and numerical study of a full scale ventilated enclosure: Comparison of four two equations closure turbulence models[J]. Building and Environment, 2007, 42(3): 1043-1053.
577 578
Lai A C K, Nazaroff W W. Modeling indoor particle deposition from turbulent flow onto smooth surfaces[J]. Journal of aerosol science, 2000, 31(4): 463-476.
579 580
Laiman R, He C, Mazaheri M, et al. Characteristics of ultrafine particle sources and deposition rates in primary school classrooms[J]. Atmospheric Environment, 2014, 94: 28-35.
581
Launder B E, Spalding D B. Lectures in mathematical models of turbulence[J]. 1972.
582 583
Li Y, Stathopoulos T. Numerical evaluation of wind-induced dispersion of pollutants around a building[J]. Journal of Wind Engineering and Industrial Aerodynamics, 1997, 67: 757-766.
584 585 586
Lippmann M, Chen L C, Gordon T, et al. National Particle Component Toxicity (NPACT) Initiative: integrated epidemiologic and toxicologic studies of the health effects of particulate matter components[J]. Research Report (Health Effects Institute), 2013 (177): 5-13.
587 588
Mataloni F, Badaloni C, Golini M N, et al. Morbidity and mortality of people who live close to municipal waste landfills: a multisite cohort study[J]. International Journal of Epidemiology, 2016: dyw052.
589 590 591
Matz C J, Stieb D M, Davis K, et al. Effects of age, season, gender and urban-rural status on time-activity: Canadian Human Activity Pattern Survey 2 (CHAPS 2)[J]. International journal of environmental research and public health, 2014, 11(2): 2108-2124.
592
Nazaroff W W. Indoor particle dynamics[J]. Indoor air, 2004, 14(s7): 175-183.
593 594
Park S H, Kim H O, Han Y T, et al. Wall loss rate of polydispersed aerosols[J]. Aerosol Science & Technology, 2001, 35(3): 710-717.
595 596
Park S H, Lee K W. Lognormal size distribution theory for deposition of polydisperse aerosol particles[J]. Nuclear science and engineering, 2000, 135(3): 288-295.
597 598
Posner J D, Buchanan C R, Dunn-Rankin D. Measurement and prediction of indoor air flow in a model room[J]. Energy and buildings, 2003, 35(5): 515-526.
599 600
Rim D, Choi J I, Wallace L A. Size-Resolved Source Emission Rates of Indoor Ultrafine Particles Considering Coagulation. Environmental Science & Technology[J], 2016, 50(18):10031.
601 602
Rivas I, Viana M, Moreno T, et al. Outdoor infiltration and indoor contribution of UFP and BC, OC, secondary inorganic ions and metals in PM 2.5 in schools. Atmospheric Environment[J], 2015, 106:129–138.
603 604
Settumba N, Garrick S C. A comparison of diffusive transport in a moment method for nanoparticle coagulation[J]. Journal of Aerosol Science, 2004, 35(1): 93-101.
605 606
Song Liu, Lin J. Numerical Simulation of Nanoparticle Coagulation in a Poiseuille Flow Via a Moment Method[J]. Journal of Hydrodynamics, Ser. B, 2008, 20(1): 1-9.
AC C
EP
TE D
M AN U
SC
RI PT
571 572
32
ACCEPTED MANUSCRIPT Steinle S, Reis S, Sabel C E. Quantifying human exposure to air pollution—Moving from static monitoring to spatio-temporally resolved personal exposure assessment[J]. Science of the Total Environment, 2013, 443: 184-193.
610 611
Tominaga Y, Stathopoulos T. Turbulent Schmidt numbers for CFD analysis with various types of flowfield[J]. Atmospheric Environment, 2007, 41(37): 8091-8099.
612 613
Wang J. A two-dimensional TVD scheme for incompressible turbulent flows with pseudo-compressibility method[J]. Progress in Computational Fluid Dynamics An International Journal, 2009, 9(2).
614 615
Yakhot V, Orszag S A, Thangam S, et al. Development of turbulence models for shear flows by a double expansion technique[J]. Physics of Fluids A Fluid Dynamics, 1992, 4(7):1510-1520.
616 617
Yu M, Lin J, Chan T. A new moment method for solving the coagulation equation for particles in Brownian motion[J]. Aerosol Science and Technology, 2008, 42(9): 705-713.
618 619
Yu M, Lin J, Chan T. Numerical simulation for nucleated vehicle exhaust particulate matters via the TEMOM/LES method[J]. International Journal of Modern Physics C, 2009, 20(03): 399-421.
620 621
Yu M, Lin J, Chen L, et al. Large eddy simulation of a planar jet flow with nanoparticle coagulation[J]. Acta Mechanica Sinica, 2006, 22(4): 293-300.
622 623
Zhao B, Chen C, Tan Z. Modeling of ultrafine particle dispersion in indoor environments with an improved drift flux model[J]. Journal of Aerosol Science, 2009, 40(1):29-43.
624 625
Zhao B, Li X, Zhang Z. Numerical Study of Particle Deposition in Two Differently Ventilated Rooms. Indoor & Built Environment[J], 2004, 13(6):443-451.
626 627
Zhao B, Yang C, Yang X, et al. Particle dispersion and deposition in ventilated rooms: Testing and evaluation of different Eulerian and Lagrangian models[J]. Building & Environment, 2008, 43(4):388-397.
AC C
EP
TE D
M AN U
SC
RI PT
607 608 609
ACCEPTED MANUSCRIPT Highlights Two moment based coupled dynamical models, MCM and TECM, coupled with flow field are established to simulate particles deposition in ventilated rooms.
RI PT
Both models could give good results on particles evolution in the indoor air flow, and the TECM performs relative better compared with MCM.
AC C
EP
TE D
M AN U
SC
Compared with CVM, the SVM could produce a more stratified and lower concentration level.