Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

International Journal of Ventilation

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tjov20

Numerical simulation of droplet aerosol


transmission in an enclosed space

Zhenhai Wu, Qing Xu & Changming Ling

To cite this article: Zhenhai Wu, Qing Xu & Changming Ling (2021): Numerical simulation of
droplet aerosol transmission in an enclosed space, International Journal of Ventilation, DOI:
10.1080/14733315.2021.1971872

To link to this article: https://doi.org/10.1080/14733315.2021.1971872

© 2021 The Author(s). Published by Informa


UK Limited, trading as Taylor & Francis
Group

Published online: 09 Sep 2021.

Submit your article to this journal

Article views: 767

View related articles

View Crossmark data

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


https://www.tandfonline.com/action/journalInformation?journalCode=tjov20
INTERNATIONAL JOURNAL OF VENTILATION
https://doi.org/10.1080/14733315.2021.1971872

Numerical simulation of droplet aerosol transmission in an


enclosed space
Zhenhai Wu, Qing Xu and Changming Ling
School of Mechanical and Power Engineering, Guangdong Ocean University, Zhanjiang, China

ABSTRACT ARTICLE HISTORY


A three-dimensional model of an enclosed air-conditioned space was Received 12 January 2021
established. Numerical simulation was conducted for the temperature Accepted 19 August 2021
distribution in the enclosed space with the air conditioner turned on
KEYWORDS
using computational fluid dynamics, and the movement track of aerosol
Enclosed air-conditioned
droplets exhaled by a patient in the enclosed space was traced using space; aerosol droplet
the discrete phase model based on the Eulerian–Lagrangian method. transmission; CFD
The analysis shows that, after the air conditioner is turned on, the vel-
ocity and temperature of outflowing air from the air conditioner greatly
influence the temperature field and the movement track of aerosol in
the enclosed space. With the change of intake air temperature, there is
little difference in the absorption of aerosol by air conditioning; When
the air velocity of outflowing air from the air conditioner is 0.8 m/s-2m/
s, the amount of aerosol absorbed by the air conditioner increases with
the increase of air velocity of outflowing air from the air condition.
Equipping the air conditioner with an aerosol filter can prevent the
aerosol droplets from exiting of the air conditioner again and decrease
the quantity of aerosol particles in the enclosed space, thus reducing
the risk of airborne viruses infecting susceptible people. In some public
places where central air conditioning is used, aerosol filters can also
avoid transmitting aerosols to other spaces through the central air con-
ditioning system.

1. Introduction
The diameters of viruses range from 0.08 to 0.160 lm (Monto, 1974), and a virus can survive for
3 h in a general indoor environment (Ai et al., 2019) Meanwhile, the respiratory tract of an
infected person may produce virus-containing droplets (Almstrand et al., 2009; Papineni &
Rosenthal, 1997), whose diameters are in the range from 0.05 to 500 lm (Graton et al., 2010;
Chao et al., 2009). After being exhaled from the person’s body, these droplets, which are sus-
pended in the air, rapidly evaporate into aerosol particles within an extremely short time
(Ji et al., 2018). Therefore, the main transmission routes of viruses of this type are droplet
transmission, aerosol transmission, and contact transmission, which are also the transmission
routes of COVID-19 (Yen et al, 2020; Yen et al., 2011; Yen et al., 2014). When the air condi-
tioner is turned on indoors, the aerosol may be inhaled by the air conditioner and propa-
gated again with the ventilation system (Michael, et al., 2020; Philip et al., 2020). In

CONTACT Changming Ling ling-cm@163.com


ß 2021 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group
This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives License (http://
creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided
the original work is properly cited, and is not altered, transformed, or built upon in any way.
2 Z. WU ET AL

Nomenclature
U velocity vector (m/s) FT thermophoretic force
t kinematic viscosity (m2/s) FS Saffman force
T temperature (K) l viscosity of air
cp specific heat capacity (J/(kgK)) dp droplet diameter
ST source term (in W/m3) Re Reynolds number
Sh internal heat source (in W/m3) DT, p thermophoretic coefficient
U dissipation coefficient dij deformation tensor
Gk generation of turbulence kinetic energy u friction velocity
due to the mean velocity gradients y distance between the first grid node
Gb generation of turbulence kinetic energy and the wall
due to buoyancy C dimensionless concentration of the
k turbulence kinetic energy aerosol particles
Ɛ rate of dissipation C0 initial concentration of the particles
mp mass of the particle of patients
Up velocity of the particle Cm concentration of the particles in
q density of air the outlet
qp density of the aerosol droplet
sr relaxation time of the aerosol droplet

particular, most air conditioning systems are all-air systems, so it is very likely that infection
occurring in one room will lead to infection occurring in the whole building (Guo et al.,
2020). Relevant building environment organisations, such as REHVA (European Federation of
heating, ventilation and air conditioning associations) and ASHRAE (American Association of
heating, ventilation and Air Conditioning Engineers) believe that aerosols have potential air
transmission hazards and recommend ventilation control measures accordingly; For example,
in case of human infection, all air conditioning systems of public buildings (such as office
buildings, schools, shopping areas, etc.) are prohibited from circulating (ASHRAE, 2020;
REHVA, 2020; Shen & Liu, 2020).
Many scholars have done a lot of research on the mechanism of droplet evaporation and dif-
fusion. Redrow et al. (2011) simulated the evaporation and diffusion processes of the sputum
droplets produced when a person coughed or sneezed. Their research results show that buoyant
jets and the environment have important effects on the transport and diffusion of virus-carrying
droplets. By solving the Langevin equation, Das et al. (2020) found that droplets (aerosol par-
ticles) with small diameter can be transmitted over a relatively long distance and can be sus-
pended for a relatively long time under the action of air flow. Liu et al. (2020) conducted water
tank experiments based on the similarity protocol to simulate the movement of exhaled air flow
and particles in uniform and stratified environments. Their results indicate that the aerosol drop-
let particles are “locked” by the indoor thermally stratified layers and are dispersed over a very
long distance with the heat flow; however, large droplets may be deposited over a short dis-
tance and are barely influenced by the thermally stratified layers. Yu & Marchal (2020) studied
the influence of environmental relative humidity on the transmission of droplets in air using
computational fluid dynamics. Their results show that the droplets can be suspended for a rela-
tively long time in an environment with relatively low relative humidity. Yan et al. (2019) ana-
lyzed the influence of the heat flow from a human body model on the evaporation of cough
droplets using a multi-component Eulerian–Lagrangian method. Their results demonstrate that
the heat emitted from a human body strongly influences the droplet mass fraction and the local
distribution of air velocity.
Other scholars simulated the diffusion of particles in various environments. George et al.
(2015) used CFD method to simulated the UFP emitted by vehicles inhaled by people on
the street. Through experiments and CFD simulation, Qian and Li (2010) found that ceiling
INTERNATIONAL JOURNAL OF VENTILATION 3

exhaust in the ward can remove gaseous and fine particles more effectively than floor
exhaust. Prathemesh et al. (2021) made a comprehensive calculation and analysis of the
ventilation system of a typical intercontinental aircraft. Li et al. (2020) and Ji-Xiang et al.
(2020) found that strong turbulence will be generated when the urinal and toilet
are flushed, and the turbulence will discharge the aerosol from the urinal and toilet. In add-
ition, some scholars have studied the dynamics of masks (K€ahler & Hain, 2020; Dbouk
et al. 2020).
Closed space was more favourable for the propagation of aerosol, especially in a small closed
space (Asadi et al., 2020; Liu et al., 2020; Yun Yun et al., 2020). At present, most of the research
on indoor aerosol focuses on the first indoor propagation. The influence of air conditioning sys-
tem on enclosed space is not taken into account. Turning on the air conditioning will absorb
aerosols, which will be sprayed out of the air conditioning again and circulate indoors or enter
other rooms through the air conditioning system. In this paper, the trajectory of aerosol after
opening the air conditioner in a closed space is numerically simulated. The changes of indoor
temperature field and velocity field caused by different working states of the air conditioner are
studied, and the absorption of aerosol by the air conditioner under different flow fields is ana-
lyzed; this paper analyzes the risk of reducing the secondary transmission of aerosol after filter-
ing measures are taken in the air conditioner in the closed space.

2. Methods
2.1. Basic equations (Quan, 2001, pp.1-6)
If the air in an enclosed air-conditioned space is considered as an incompressible ideal gas of
constant density, the continuity equation can be expressed as
rU ¼0 (1)
where U is the velocity vector.
For an incompressible fluid with viscosity as a constant, the momentum conservation equa-
tions are
@u 1 @p
þ r  ðuUÞ ¼ r  ðruÞ
@t q @x
@v 1 @p
þ r  ðvUÞ ¼ r  ðrvÞ (2)
@t q @y
@w 1 @p
þ r  ðwUÞ ¼ r  ðrwÞ
@t q @z
where, u, v, and w are the velocity components in x, y, and z directions, respectively, and t is the
kinematic viscosity of the fluid.
For an incompressible ideal gas, the energy conservation equation is
 
@T k ST
þ r  ðUTÞ ¼ r  rT þ (4)
t qcp q
with
p ¼ qRT
ST ¼ Sh þU
where T is the temperature (in kelvins), cp is the specific heat capacity at constant pressure (in J/
(kgK)), ST is the source term (in W/m3), Sh is the internal heat source (in W/m3), and U is the dis-
sipation coefficient.
K-Ɛ model is selected as the turbulence model. The turbulence kinetic energy, k, and its rate
of dissipation, Ɛ, are obtained from the following transport equations:
4 Z. WU ET AL

"  #
@ @ @ lt @k
ðqkÞ þ ðqkui Þ ¼ lþ þ Gk þ Gb qeYM
@t @xi @xj rk @xj
"  #
and (5)
@ @ @ lt @e e e2
ðqeÞ þ ðqeui Þ ¼ lþ þ C1e ðGk þ C3e Gb ÞC2e q
@t @xi @xj re @xj k k

Where, Gk represents the generation of turbulence kinetic energy due to the mean velocity
gradients; Gb is the generation of turbulence kinetic energy due to buoyancy; C1e , C2e and C3e are
constants. rk andre are the turbulent Prandtl numbers for k and e, respectively.

2.2. Discrete phase model


Laboratory experiments show that 90% of large-size droplets (with droplet diameter > 50 lm)
settle rapidly at places 1–2 m from the emitting point (Redrow et al., 2011). Only 4%–10% of
droplets remain suspended in the indoor air (with droplet diameter < 10 lm), and the time for
large-diameter droplets to evaporate from 5–10 lm to 2 lm is <0.1 s (Morawska, 2006). The
simulation was performed to study the movement track of aerosol droplets. Large-diameter
droplets would evaporate within a short time after being exhaled and become relatively small
droplet cores (aerosol droplet particles), so the aerosol droplet particles with particle size of
2.5 lm exhaled by patients were selected in the simulation, and the evaporation process of aero-
sol droplets was neglected. The thermophoretic force and Saffman force acting on the aerosol
droplets were taken into account, but the pressure gradient force and the virtual mass force
were assumed to be negligible. The forces on the aerosol droplets were analyzed, and the
movement track of aerosol droplets could be calculated based on Newton’s second law using
the formula
dUp UUp gðqp  qÞ
mp ¼ mp þ mp þ FT þ FS (6)
dt sr qp
where mp is the mass of the particle, U is the velocity of air, Up is the velocity of the particle, q
UU
is the density of air, qp is the density of the aerosol droplet, mp sr p is the resistance, sr is the
relaxation time of the aerosol droplet, FT is the thermophoretic force, and FS is the Saffman
force. The expressions for sr , FT , and FS are as follows:
qp dp2 24
sr ¼
18l Cd Re
where l is the viscosity of air, dp is the droplet diameter, and Re is the Reynolds number;
1
FT ¼ DT , p rT,
T
where DT , p is the thermophoretic coefficient; and
2Kv1=2 qdij
FS ¼ mp ð u  up Þ
qp dp ðdlk dkl Þ1=4
where K ¼ 2.594 and dij is the deformation tensor.

3. Calculation model and boundary conditions


3.1. Calculation model
The physical model given in Figure 1 is an enclosed air-conditioned space with dimensions of
5 m  5 m  4.5 m (length  width  height) (where the side with a door is the left wall and the
INTERNATIONAL JOURNAL OF VENTILATION 5

Figure 1. Physical model of enclosed air-conditioned space.

side opposite to it is the right wall, and where the air outlet and return air outlet of the air con-
ditioner are located on the front wall and the opposite side is the rear wall.) The cold air is
blown out of the air conditioner, performing convective heat exchange with the indoor hot air
(with radiant heat transfer being neglected). The cold air sinks, and the hot air floats upward
and then leaves through the return air outlet. The geometric parameters of the physical model
were set as follows: Each simplified human body model was 1.7 m high and 1.7 m from the left
wall, and these two human body models were 1.8 m from the front and rear walls, respectively.
The computational domain was discretised with unstructured grids. The grids were densified
for the air outlet, return air outlet, patient’s expiratory mouth, and susceptible person’s inspira-
tory mouth, and 10 boundary expansion layers were set up for the patient and susceptible per-
son. The thickness of the inflation layer is obtained from the yþ number, which is given by the
following formula:
qu y
yþ ¼ (7)
l
Where, q Is the air density, u is friction velocity, y is the distance between the first grid node
and the wall, l Is air viscosity.
Take the thickness of the first layer of grid as 2  104m,yþ<10。According to Abuhegazy’s
research, grids can be used (Abuhegazy et al., 2020). The grid is shown in Figure 2. Grid inde-
pendence was verified with seven sets of grids (i.e. 0.263, 0.462, 0.659, 0.861, 1.061, 1.262, and
1.464 million grids, respectively), with the temperature and pressure at the observation point
selected as assessment indexes. The results are shown in Figure 2. With the assessment results of
temperature and pressure being taken into account comprehensively, the grid quantity of 1.061
million was selected for calculation, and the quality of 99.8% of the grids reached 0.79.

3.2. Model validation


Hanhui et al. (2009) studied the concentration change of aerosol particles from an outdoor envir-
onment into the experimental device; The experimental setup consists of an
width  depth  height ¼ 1 m  1 m  0.3 m test chamber with one inlet and multiple outlets of
uniform rectangular shapes of width  height ¼ 0.2 m  0.3 m; In addition, a 1 m long rectangular
tube is installed at the inlet to stabilise the incoming gas (Figure 3). Discrete particles with a
density of about 2100 kg/m3 are mixed with the supply gas flow at speeds of 1.5 m/s and 0.5 m/
s. The 9 measuring points are evenly distributed on the central plane of Z, and the direction is
6 Z. WU ET AL

Figure 2. Grid independence assessment with (a) temperature as the assessment index and (b) pressure as the assessment
index.(c) is grid model.
INTERNATIONAL JOURNAL OF VENTILATION 7

Figure 3. Geometry of the ventilated chamber.

Figure 4. Comparisons of simulated and experimental particle concentrations at the measuring points.

3  3 arrangement. They used a laser particle counter (CLJ-BII) to measure particle concentration.
The aerosol particle size was divided into five groups: 0.3-0.5 l m, 0.5-1.0, l m, 1.0-3.0 l m, 3.0-
5.0 lm and 5.0-10.0 lm.
According to the particle diameters distribution curve, two peak diameters (0.9 lm and 5 lm)
were selected to validate the numerical model. The simulation verifies that when the air velocity
is 0.5 m/s and the outlet O1 is open, the concentration change is shown in Figure 4.
The results show that the numerical results are agreement with the experimental data. The
average relative error of 0.9 lm particles is 4.51%, and the maximum error is 10.69%; The aver-
age relative error of 5.0 lm particles is 9.25%, and the maximum error is 18.46%; Since the
experimental measurement error is unknown, the simulation results with an average relative
error of less than 10% are acceptable. It shows that the numerical method can reasonably pre-
dict the indoor particle concentration.

3.3. Boundary conditions


After droplets are exhaled, they quickly evaporate into aerosol droplets. If the air conditioner is
not equipped with a filter, these aerosol droplets would enter the space again through the air
outlet of the air conditioner. Therefore, simulations were conducted for two cases: one in which
8 Z. WU ET AL

Table 1. Boundary types and parameter settings.


Boundary Boundary type qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Parameter setting
Air outlet velocity-inlet v ¼ vx2 þ vy2 þ vz2 , where vz ¼ 0, vx ¼ 2vy
Return air outlet pressure-outlet Gage pressure ¼ 0 Pa
Ceiling and floor wall Adiabatic, wall thickness ¼ 0.2 m
Front and rear walls wall Adiabatic, wall thickness ¼ 0.2 m
Left and right walls and door wall Constant wall temperature T; wall thickness ¼ 0.2 m
Human body surface wall Constant heat transfer coefficient k ¼ 3.4 W/(m2K) (de Dear et al., 1997).
Patient’s mouth velocity-inlet vA ¼ 2:14 m/s
Susceptible person’s mouth velocity-inlet vB ¼ 2:14 m/s

Table 2. Settings of operating conditions for calculation.


Velocity of Velocity of
outflowing air Temperature of outflowing air Temperature of
Operating condition (m/s) outflowing air (K) Operating condition (m/s) outflowing air (K)
1 0.8 295 14 2.0 297
2 1.2 295 15 2.4 297
3 1.6 295 16 0.8 298
4 2.0 295 17 1.2 298
5 2.4 295 18 1.6 298
6 0.8 296 19 2.0 298
7 1.2 296 20 2.4 298
8 1.6 296 21 0.8 299
9 2.0 296 22 1.2 299
10 2.4 296 23 1.6 299
11 0.8 297 24 2.0 299
12 1.2 297 25 2.4 299
13 1.6 297

the air conditioner was equipped with a droplet filter and the other in which the air conditioner
was not equipped with a droplet filter.
In the case in which a filter was provided, the air outlet of the air conditioner was a velocity-
type inlet, where the horizontal velocity was vx , the downward vertical velocity was vy, and the
temperature of outflowing air was T: The return air outlet was a pressure-type outlet. The move-
ment track of aerosol droplet was calculated, and a steady-state calculation was used. To simplify
the calculation, the patient was assumed to only exhale, and the susceptible person inhales all
the time. However, because of the steady-state calculation, the respiration rate of the calculation
model was taken as a constant. For a normal person, the respiratory flux is 8.36 L/min, and the
velocity of airflow generated by respiration is 2.14 m/s (Liu et al., 2020; Shimer et al., 1995). The
patient’s expiratory temperature was taken as 306 K. The left and right walls were heat exchange
walls, whose temperature was 303 K and whose thickness was 0.2 m. It was assumed that there
were rooms at the top, bottom, front, and rear, where the air conditioners were turned on.
Accordingly, all of the top, bottom, front, and rear walls were adiabatic. The specific boundary
conditions and parameter settings are listed in Table 1.
For the discrete phase boundary conditions, the inlet and outlet of the air conditioner were
set as escape boundaries, the susceptible person’s inspiratory mouth was set as a trap boundary,
the patient’s expiratory mouth was set as a wall-jet boundary, and other surfaces were set as
trap boundaries. There were 50 droplets in total emitted from the patient’s expiratory mouth. To
simulate the movement track of aerosol droplets under different indoor temperatures and veloc-
ities, 25 operating conditions for calculation were set according to the temperature and velocity
at the air outlet of the air conditioner, as given in Table 2.
In addition, in the case in which the air conditioner was not equipped with a droplet filter,
the aerosol droplet, after being sucked in by the air conditioner, would return to the space from
the air outlet of the air conditioner. At this moment, the air outlet of the air conditioner was an
INTERNATIONAL JOURNAL OF VENTILATION 9

Table 3. Solver settings.


Name Method
turbulence model Standard k- e Modeland enhanced wall treatment
algorithm SIMPLE
Discretization of pressure gradient term PRESTO!
Convergence criteria The solution is assumed to be converged when all the
scaled residuals stabilise and approach a minimum of
105 for k, e, x, y, and z momentum equations as well
as 104 for the continuity equation (Abuhegazy et
al., 2020).

Figure 5. Temperature distribution in the section through the center of the human body with the velocity of outflowing air
being 0.8 m/s and the temperature of outflowing air being 295–299 K.

aerosol particle jet (wall jet). Compared with case 1, the operating conditions for calculation
remained unchanged.
A pressure-based solver was used to find the solution. Owing to the influence of the tempera-
ture change in an enclosed air-conditioned space on the indoor air, the thermal buoyancy flow
field was solved by using the incompressible Navier–Stokes equations and the Boussinesq
approximation method, and the simulation was conducted with the standard k-e turbulence
model. The solver settings are shown in Table 3.

4. Analysis and discussion


4.1. Analysis of the temperature field in space
The temperature distribution was viewed on a section through the center of the human body
(y ¼ 1.925 m). Figure 5 shows the temperature distribution on the interface through the center of
the human body in the case of constant velocity of outflowing air at the air outlet (0.8 m/s) and
varying temperature. The air with low temperature exits the air outlet and convectively
exchanges heat with the air with high temperature in the space. The hot air flows to the upper
part of the space and the cold air sinks to the lower part of the space, generating a temperature
gradient in the vertical direction. The lower the temperature of the outflowing air is, the more
evident the temperature stratification caused by the convection with the hot air in the space will
be. In contrast, when the temperature of outflowing air is higher, convection with the hot air in
10 Z. WU ET AL

Figure 6. Temperature distribution in the section through the center of the human body with the temperature of outflowing
air being 295 K at different outflowing air velocity values.

the space is weaker, and the stratification is not evident. Owing to the temperature gradient, the
air flow exhaled from the human body may exhibit an upward deflection phenomenon.
As shown in Figure 6, the temperature of outflowing air at the air outlet is 295 K, and varied
temperature may also lead to the change in temperature gradient. When the velocity of outflow-
ing air is 0.8 and 1.2 m/s, temperature stratification in the section is evident and the temperature
gradient is relatively large. When the velocity of outflowing air is 1.6, 2.0, and 2.4 m/s, the tem-
perature values in the section are mostly 295–296 K, and the temperature gradient in the section
is small. When the velocity of outflowing air is low, the hot air and cold air in the space form air
masses, which are not dispersed by blowing, so it is easy to form cold–hot air stratification and a
large temperature gradient. Increasing the velocity of outflowing air to a high value makes it dif-
ficult for the hot air and cold air in the space to form air masses, and the hot air is cooled by
the cold air, so the formed temperature gradient is small.

4.2. Influence of a droplet-filter-equipped air conditioner on aerosol droplets in space


The quantity of droplet particles injected under each operating condition was 50. The droplets
were emitted from the patient’s mouth. The movement tracks of aerosol droplets under operat-
ing condition 3 are shown in Figure 7, which gives the movement tracks of 1, 10, 20, 30, 40, and
50 aerosol droplet particles, respectively. Because the temperature of the air flow exhaled from
the human body is higher than that of the environment, the aerosol droplet flow is deflected
upward under the action of the exhaled air flow with relatively high temperature and the ther-
mal flow field on the human body surface so that the droplets exhaled by the patient are less
inhaled by the susceptible person.
The aerosol droplet particles not inhaled by the susceptible person are influenced by the
movement of the indoor air flow. Some of these aerosol droplet particles move to the area near
the return air outlet of the air conditioner and then are sucked in by the return air outlet of the
air conditioner; others collide with the walls during movement and then adhere to the walls or
human body models. The distributions of aerosol droplet particles after cessation of movement
under each operating condition are given in Table 4. Figure 8 shows the statistical results of the
aerosol droplet particles at the positions when movement stops within various time periods
when the temperature of outflowing air is 296 K for five velocities of outflowing air. As can be
INTERNATIONAL JOURNAL OF VENTILATION 11

Figure 7. Movement track of aerosol droplets under operating condition 3.

seen from Figure 8, the time from exhalation to cessation of movement is <100 s for 64.4% of
aerosol droplet particles, indicating that, after being emitted, the aerosol droplet particles will be
sucked in or adhere to the walls within a short time. The aerosol droplet particles adhering to
the wall will increase the risk of contact transmission, and, in this case, disinfection can be per-
formed by spraying a medicine solution. Moreover, the percentage of aerosol droplet particles
sucked in by the air conditioner of the total aerosol droplet particles varies with operating condi-
tion, as listed in Table 4 and shown in Figure 8. The quantity of aerosol droplet particles sucked
in by the air conditioner is the least, being only 7 (14%), under operating condition 6, and the
most, being 32 (64%), under operating condition 9.

Suction of aerosol droplets by air conditioning increases the risk of aerosol transmission.
In this case, if the air conditioner is equipped with a droplet filter, the droplets will be fil-
tered off by the filter after being sucked in by the air conditioner. The more the droplets
are sucked in by the air conditioner, the more they will be filtered off by the filter. This is
evidently a way to reduce the quantity of indoor droplets. It can also be discerned from
Figure 8 that the quantity of aerosol droplet particles sucked in by the air conditioner varies
with operating condition. In the time range of <100 s, the aerosol droplet particles are more
easily sucked in by the air conditioner with the increase in velocity of outflowing air.
However, when the velocity of outflowing air is >2.4 m/s, the quantity of aerosol droplet par-
ticles sucked in by the air conditioner decreases. Because the increased velocity of outflow-
ing air enhances the indoor air flow, the aerosol droplet particles move more easily with the
air flow to the suction area of the air conditioner and are then sucked in by the air condi-
tioner. When the velocity of outflowing air is too high, the aerosol droplet particles more
easily collide with the wall under the action of the air flow than move to the suction area of
the air conditioner, so the quantity of aerosol droplet particles sucked in by the air condi-
tioner decreases. In the time range of 100–300 s, the quantity of aerosol droplet particles
sucked in by the air conditioner is less than that in the time range of <100 s. However, the
quantity of aerosol droplet particles at low velocity of outflowing air sucked in by the air
conditioner is more than that of aerosol droplet particles at high velocity of outflowing air.
Because, in the case of low-velocity outflowing air, indoor air convection is weak and the
12
Z. WU ET AL

Table 4. Quantity of aerosol droplet particles at positions when movement stops under various operating conditions.
Operating Sucked in by air Inhaled by Operating Sucked in by air Inhaled by
condition conditioner susceptible person Wall Human body condition conditioner susceptible person Wall Human body
1 13 1 30 6 14 18 2 27 3
2 13 1 36 0 15 11 2 33 4
3 14 0 36 0 16 14 0 34 2
4 15 0 33 2 17 13 4 26 7
5 14 2 31 3 18 18 7 23 2
6 7 6 33 4 19 22 0 25 3
7 16 0 32 2 20 15 0 34 1
8 10 4 32 4 21 11 2 30 7
9 32 0 18 0 22 19 2 29 0
10 13 0 36 1 23 19 2 28 1
11 6 9 33 2 24 19 2 28 1
12 10 1 36 3 25 11 2 30 7
13 15 0 34 1
INTERNATIONAL JOURNAL OF VENTILATION 13

Figure 8. Distribution of aerosol droplet particles at positions when movement stops in various time periods at a temperature
of 296 K.

influence of air flow on the aerosol droplet is weak, and the time from when the aerosol
droplet is emitted to when it is sucked in by the air conditioner is long.

4.3. Movement of aerosol droplets in space with an air conditioner not equipped with a
droplet filter
An air conditioner not equipped with a droplet filter cannot filter off the aerosol droplet, so the
aerosol droplet, after being sucked in by the air conditioner, will enter the space again from
the air outlet together with the cold air, as shown in Figure 9. Figure 9 gives the movement of
the aerosol droplets returning to the space after being sucked by the air conditioner under oper-
ating condition 9, and it shows the movement tracks of 1, 10, 20, and 30 aerosol droplet par-
ticles exiting the air conditioner. It can be seen from the figure that the movement of aerosol
droplets is mainly influenced by the outflowing air from the air conditioner. The aerosol droplets
are sucked by the air conditioner and then exit the air conditioner. Table 5 gives the quantity
distribution of aerosol droplet particles at the positions when movement stops. As can be seen
from the table, most of the aerosol droplet particles adhere to the walls, ceiling, and floor, and a
few adhere to the human body. Figure 10 shows the distribution of the aerosol droplet particles
returning to the space from the air conditioner after being sucked by the air conditioner within
various time periods at an outflowing air temperature of 296 K. As can be discerned from the fig-
ure, the aerosol droplets cannot be filtered off in the air conditioner, and the aerosol droplets
exiting the air conditioner again may increase the risk of aerosol transmission. Under operating
14 Z. WU ET AL

Figure 9. Movement track of aerosol droplets returning to space after being sucked in by the air conditioner under operating
condition 9.

condition 9, the susceptible person inhaled one aerosol droplet particle. The movement track of
an aerosol droplet exiting the air conditioner and adhering to the human body is shown in
Figure 11. The movement tracks of aerosol droplets returning from the air conditioner and
inhaled by the susceptible person are shown in Figure 12. Most of the aerosol droplet particles
adhere to the walls after exiting the air conditioner, and relatively few aerosol droplet particles
are inhaled by or adhere to the human body. The aerosol droplet particles exiting from the air
outlet, different from those emitted from the mouth, are more greatly influenced by the air flow,
and they may collide with the walls along with the air flow when the velocity of the air flow is
high. Consequently, there are more aerosol droplet particles adhering to the walls. Because the
air conditioner has no filtering measures, 44% of the aerosol is left indoors, increasing the risk of
transmitting infection.

4.4. Discussion on aerosol concentration in closed space


When the temperature of the air flowing out of the air conditioner is 296k,set the velocity of air
flowing out of the air conditioner to 0.8 m/s, 1.2m/s, 1.6m/s, 2.0m/s, 2.4m/s. Convert the concen-
tration to dimensionless concentration, which is defined as:
Table 5. Quantity of aerosol droplet particles at positions when movement stops after exiting the air conditioner under various operating conditions.
Operating Sucked in by air Operating Sucked in by air
condition conditioner Wall Inhaled by patient Human body condition conditioner Wall Inhaled by patient Human body
1 4 8 0 1 14 1 17 0 0
2 4 11 0 0 15 5 6 0 0
3 0 13 0 0 16 2 10 0 0
4 4 15 0 0 17 0 14 0 0
5 3 14 0 0 18 1 12 0 0
6 1 7 0 0 19 3 14 0 1
7 1 14 0 1 20 6 15 0 0
8 3 26 0 0 21 4 13 0 0
9 8 21 1 1 22 3 16 0 0
10 0 14 0 0 23 3 16 0 0
11 2 8 0 0 24 2 13 0 0
12 3 5 0 2 25 4 9 0 0
13 6 8 0 1
INTERNATIONAL JOURNAL OF VENTILATION
15
16 Z. WU ET AL

Figure 10. Distribution of aerosol droplet particles at positions when movement stops after returning to space from the air
conditioner in various time periods at a temperature of 296 K.

Figure 11. Movement track of an aerosol droplet adhering to the human body model after exiting the air conditioner.

Cm
C¼ (8)
C0
where: C is the dimensionless concentration of the aerosol particles; C0 is the concentration of
the particles of patients exhaling; and Cm is the measured concentration of the particles in
the outlet.
INTERNATIONAL JOURNAL OF VENTILATION 17

Figure 12. Movement track of aerosol droplets inhaled by a susceptible person after they are return from the air conditioner.

Figure 13. Aerosol concentration at outlet.

As shown in Figure 13, when the inlet speed is lower than 2 m/s, the aerosol concentration at
the outlet increases with the increase of inlet speed. The results show that the increase of air
conditioning wind speed can destroy the human stratification formed in the space, and make
the aerosol easier to move with the wind and be inhaled by the air conditioning.

5. Conclusions
For the problem of aerosol droplet transmission in enclosed air-conditioned space in public pla-
ces, a three-dimensional steady-state model was established, the movement track of aerosol
droplet particles was traced using a discrete phase model based on the Eulerian–Lagrangian
method, and the influence of the air conditioner on the indoor temperature distribution and the
movement track of aerosol droplets in air-conditioned space under 25 operating conditions were
analyzed. The following conclusions were drawn:
18 Z. WU ET AL

1. In an enclosed air-conditioned space, the velocity and temperature of outflowing air from
the air conditioner influence the air flow and temperature change. With the velocity of out-
flowing air remaining unchanged, the lower the temperature of outflowing air, the more evi-
dent the temperature stratification in the space, and vice versa. With the temperature of
outflowing air remaining unchanged, the lower the velocity of outflowing air, the more evi-
dent the temperature stratification in the space. When thermal stratification occurs, it is
more difficult for aerosols to cross thermal stratification.
2. When the velocity of outflowing air is in the range of 0.8–2.0 m/s, When the outlet velocity
of droplet aerosol is 0.8  2.0m/s, the greater velocity can easily destroy the thermal stratifi-
cation, and the aerosol is more vulnerable to the action of air flow and circulates in the air
conditioning system with the air flow, the more aerosol droplets the air conditioner will
suck in even more easily, and the shorter the time required is to do so. When the velocity is
2.4 m/s, the aerosol is more easily influenced by the air velocity, thus facilitating droplet col-
lisions with the wall. The proportion of aerosol inhaled by air conditioning is reduced
3. The aerosols inhaled by the air conditioner return indoors again, and 44% of the aerosols
are left indoors, increasing the risk of secondary transmission and infection.
4. The air conditioner has a great influence on the aerosol droplet transmission. Deploying a
droplet filter in the air conditioner can effectively decrease the quantity of aerosol droplet
particles in the enclosed air-conditioned space by 14%–64%, reducing the risk that the sus-
ceptible person inhales more aerosol droplets.
5. The case studied in this paper is listed as theoretical research, and the research model is
only based on it, which is different from the actual environment. Next, we will study the
actual environment, such as the carriage of high-speed railway, aircraft cabin, hotels and
shopping malls with all air conditioning system.

Disclosure statement
The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection,
analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Funding
This research was funded by the Shenzhen Science and Technology Innovation Committee (Grant Number
JCYJ20170818111558146) and Special Funds for Industrial Development of the Shenzhen Dapeng New District
(Grant Number KY20180113).

References
Abuhegazy, M., Talaat, K., Anderoglu, O., & Poroseva, S. V. (2020). Numerical investigation of aerosol transport in a
classroom with relevance to covid-19. Physics of Fluids (Woodbury, N.Y. : 1994), 32(10), 103311. https://doi.org/10.
1063/5.0029118
Ai, Z. T., Huang, T., & Melikov, A. K. (2019). Airborne transmission of exhaled droplet nuclei between occupants in a
room with horizontal air distribution. Building and Environment, 163, 106328. https://doi.org/10.1016/j.buildenv.
2019.106328
Almstrand, A. C., Ljungstro €m, E., Lausmaa, J., Bake, B., Sjo
€vall, P., & Olin, A. C. (2009). Airway monitoring by collec-
tion and mass spectrometric analysis of exhaled particles. Analytical Chemistry, 81(2), 662–668. https://doi.org/10.
1021/ac802055k
Asadi, S., Bouvier, N., Wexler, A. S., & Ristenpart, W. D. (2020). The coronavirus pandemic and aerosols: Does COVID-
19 transmit via expiratory particles? Aerosol Science and Technology: The Journal of the American Association for
Aerosol Research, 54 (6), 635–638. https://doi.org/10.1080/02786826.2020.1749229
ASHRAE (2020). Guide to the COVID-19 Pages. American Society of Heating, Refrigerating and Air-Conditioning
Engineers.
INTERNATIONAL JOURNAL OF VENTILATION 19

Chao, C. Y. H., Wan, M. P., Morawska, L., Johnson, G. R., Ristovski, Z. D., Hargreaves, M., Mengersen, K., Corbett, S.,
Li, Y., Xie, X., & Katoshevski, D. (2009). Characterization of expiration air jets and droplet size distributions imme-
diately at the mouth opening. Journal of Aerosol Science, 40(2), 122–133. https://doi.org/10.1016/j.jaerosci.2008.
10.003
Das, S. K., Alam, J., Plumari, S., & Greco, V. (2020). Transmission of airborne virus through sneezed and coughed
droplets. Phys Fluids (1994), 32(9), 97102.
de Dear, R. J., Arens, E., Hui, Z., & Oguro, M. (1997). Convective and radiative heat transfer coefficients for individual
human body segments. International Journal of Biometeorology, 40(3), 141–156. https://doi.org/10.1007/
s004840050035
Desai, P. S., Sawant, N., & Keene, A. (2021). On COVID-19-safety ranking of seats in intercontinental commercial air-
crafts: A preliminary multiphysics computational perspective. Building Simulation, pp. 1–12.
Feng, Y., Marchal, T., Sperry, T., & Yi, H. (2020). Influence of wind and relative humidity on the social distancing
effectiveness to prevent COVID-19 airborne transmission: A numerical study. Journal of Aerosol Science, 147,
105585. https://doi.org/10.1016/j.jaerosci.2020.105585
Gralton, J., Tovey, E., McLaws, M.-L., & Rawlinson, W. D. (2011). The role of particle size in aerosolised pathogen
transmission: A review. The Journal of Infection, 62(1), 1–13. https://doi.org/10.1016/j.jinf.2010.11.010
Guo, M. Y., Xu, P., Xiao, T., He, R. K., & Dai, M. K. (2020). Comparison of guidelines related to HV for coping with
COVID-19 at home and abroad. Journal of HV&AC, 2020, 1–12.
Habilomatis, G., & Chaloulakou, A. (2015). A CFD modeling study in an urban street canyon for ultrafine particles
and population exposure: The intake fraction approach. The Science of the Total Environment, 530-531, 227–232.
https://doi.org/10.1016/j.scitotenv.2015.03.089
Hanhui, Q., Li, L., Chen, J., Fan, L. & Lu, L., (2009). Experimental analysis of particle concentration heterogeneity in a
ventilated scale chamber. Atmospheric Environment, 43(28), 4311–4318.
Ji, Y., Qian, H., & Zheng, X. (2018). Interrupting COVID-19 transmission by implementing enhanced traffic control
bundling: Implications for global prevention and control efforts. Journal of Microbiology, Immunology and
Infection, 53(3), 377–380.
Ji, Y., Qian, H., Ye, J., & Zheng, X. (2018). The impact of ambient humidity on the evaporation and dispersion of
exhaled breathing droplets: A numerical investigation. Journal of Aerosol Science, 115, 164–172. https://doi.org/
10.1016/j.jaerosci.2017.10.009
Ji-Xiang, W., Yun-Yun, L., Xiang-Dong, L., & Xiang, C. (2020). Virus transmission from urinals. Physics of Fluids
(Woodbury, N.Y. : 1994), 32(8), 081703. https://doi.org/10.1063/5.0021450
K€ahler, C. J., & Hain, R. (2020). Fundamental protective mechanisms of face masks against droplet infections.
Journal of Aerosol Science, 148, 105617. https://doi.org/10.1016/j.jaerosci.2020.105617
Kohanski, M. A., James Lo, L., & Waring, M. S. (2020). Review of indoor aerosol generation, transport, and control in
the context of COVID-19. International Forum of Allergy & Rhinology, 10(10), 1173–1179. https://doi.org/10.1002/
alr.22661
Liu, L., Zhang, Y., Fu, L. Z., & Wang, Y. (2020). Study on the risk of and countermeasures for interpersonal droplet
infection in thermally stratified environment. Journal of HV&AC, 50(6), 19–25.
Liu, Y., Ning, Z., Chen, Y., Guo, M., Liu, Y., Gali, N. K., Sun, L., Duan, Y., Cai, J., Westerdahl, D., Liu, X., Xu, K., Ho, K-f.,
Kan, H., Fu, Q., & Lan, K. (2020). Aerodynamic analysis of SARS-CoV-2 in two Wuhan hospitals. Nature:
International Weekly Journal of Science, 582(7813), 557–560. https://doi.org/10.1038/s41586-020-2271-3
Liu, F., Qian, H., Luo, Z., Wang, S., & Zheng, X. (2020). A laboratory study of the expiratory airflow and particle dis-
persion in the stratified indoor environment. Building and Environment, 180, 106988. https://doi.org/10.1016/j.
buildenv.2020.106988
Monto, A. S. (1974). Medical reviews. Coronaviruses. Yale Journal of Biology & Medicine, 47(4), 234–251.
Morawska, L. (2006). Droplet fate in indoor environments, or can we prevent the spread of infection?. Indoor Air,
16(5), 335–347. https://doi.org/10.1111/j.1600-0668.2006.00432.x
Papineni, R. S., & Rosenthal, F. S. (1997). The size distribution of droplets in the exhaled breath of healthy human
subjects. Journal of Aerosol Medicine : The Official Journal of the International Society for Aerosols in Medicine,
10(2), 105–116. https://doi.org/10.1089/jam.1997.10.105
Philip, A., Valentyn, S., Christina E, B., & Adriaan, B. (2020). Visualizing Speech-Generated Oral Fluid Droplets with
Laser Light Scattering. The New England Journal of Medicine, 382(21), 2061–2063.
Qian, H., & Li, Y. (2010). Removal of exhaled particles by ventilation and deposition in a multibed airborne infection
isolation room. Indoor Air, 20(4), 284–297. https://doi.org/10.1111/j.1600-0668.2010.00653.x
Quan, T. W. (2001). Numerical heat transfer (2nd ed.). Xi’an Jiaotong University Press.
Redrow, J., Mao, S., Celik, I., Posada, J. A., & Feng, Z-g. (2011). Modeling the evaporation and dispersion of airborne
sputum droplets expelled from a human cough. Building and Environment, 46(10), 2042–2051. https://doi.org/10.
1016/j.buildenv.2011.04.011
REHVA (2020). REHVA COVID-19 Guidance. Federation of European Heating, Ventilation and Air Conditioning
Associations.
20 Z. WU ET AL

Shen, J. M., & Liu, Y. M. (2020). The countermeasures for prevention and control of COVID-19 should be reasonable,
suitable and compliant. Journal of HV&AC, 50(6), 59.
Shimer, D. A., Jenkins, P. L., Hui, S. P., & Adams, W. C. (1995). 132 Measurement of breathing rate and volume in
routinely performed daily activities. Epidemiology, 6(2), S30.
Talib, D., & Dimitris, D. (2020). On respiratory droplets and face masks. Physics of Fluids (Woodbury, N.Y. : 1994),
32(6), 063303.
Yan, Y., Li, X., & Tu, J. (2019). Thermal effect of human body on cough droplets evaporation and dispersion in an
enclosed space. Building and Environment, 148, 96–106. https://doi.org/10.1016/j.buildenv.2018.10.039
Yen, M.-Y., Chiu, A. W.-H., Schwartz, J., King, C.-C., Lin, Y. E., Chang, S.-C., Armstrong, D., & Hsueh, P.-R. (2014). From
SARS in 2003 to H1N1 in 2009: Lessons learned from Taiwan in preparation for the next pandemic. Journal of
Hospital Infection, 87(4), 185–193. https://doi.org/10.1016/j.jhin.2014.05.005
Yen, M.-Y., Lin, Y.-E., Lee, C.-H., Ho, M.-S., Huang, F.-Y., Chang, S.-C., & Liu, Y.-C. (2011). Taiwan’s traffic control bun-
dle and the elimination of nosocomial severe acute respiratory syndrome among healthcare workers. Journal of
Hospital Infection, 77(4), 332–337. https://doi.org/10.1016/j.jhin.2010.12.002
Yen, Muh-Yong et al. (2020). Interrupting COVID-19 transmission by implementing enhanced traffic control bun-
dling: Implications for global prevention and control efforts. Journal of Microbiology, Immunology and Infection,
53(3), 377–380. DOI:10.1016/j.jmii.2020.03.011
Yun-Yun, L., Ji-Xiang, W., & Chen, X. (2020). Can a toilet promote virus transmission? From a fluid dynamics perspec-
tive. Physics of Fluids (Woodbury, N.Y.: 1994), 32(6), 065107.

You might also like