You are on page 1of 68

Journal Pre-proof

Recent research on expiratory particles in respiratory viral infection


and control strategies: A review

Yunchen Bu , Ryozo Ooka , Hideki Kikumoto , Wonseok Oh

PII: S2210-6707(21)00389-9
DOI: https://doi.org/10.1016/j.scs.2021.103106
Reference: SCS 103106

To appear in: Sustainable Cities and Society

Received date: 27 October 2020


Revised date: 7 June 2021
Accepted date: 15 June 2021

Please cite this article as: Yunchen Bu , Ryozo Ooka , Hideki Kikumoto , Wonseok Oh , Recent re-
search on expiratory particles in respiratory viral infection and control strategies: A review, Sustainable
Cities and Society (2021), doi: https://doi.org/10.1016/j.scs.2021.103106

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. 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.

© 2021 Published by Elsevier Ltd.


Highlights

 Expiratory particle and airflow features of different respiratory activities

 Environmental impact on particle evaporation and SARS-CoV-2 viability

 Infection risks of SARS-CoV-2 and determination of safe physical distance

 Influence of masks on expiratory particles and flow; wearing suggestions

 Advanced air distribution, air filtration, and disinfection methods


Recent research on expiratory particles in respiratory viral infection and control

strategies: A review

Title: Recent research on expiratory particles in respiratory viral infection and control

strategies: A review

Yunchen Bu a,*, Ryozo Ooka b, Hideki Kikumoto b, Wonseok Oh b

a
Graduate School of Engineering, The University of Tokyo, 4-6-1 Komaba, Meguro-ku,

Tokyo 153-8505, Japan.


b
Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku,

Tokyo 153-8505, Japan.

E-mail address

Yunchen Bu: bychen@iis.u-tokyo.ac.jp.

Ryozo Ooka: ooka@iis.u-tokyo.ac.jp

Hideki Kikumoto: kkmt@iis.u-tokyo.ac.jp

Wonseok Oh: oh-ws@iis.u-tokyo.ac.jp

Declarations of interest

None

2
Abstract

The global spread of coronavirus disease 2019 poses a significant threat to human health.

In this study, recent research on the characteristics of expiratory particles and flow is

reviewed, with a special focus on different respiratory activities, to provide guidance for

reducing the viral infection risk in the built environment. Furthermore, environmental

influence on particle evaporation, dispersion, and virus viability after exhalation and the

current methods for infection risk assessment are reviewed. Finally, we summarize promising

control strategies against infectious expiratory particles. The results show that airborne

transmission is a significant viral transmission route, both in short and long ranges, from

infected individuals. Relative humidity affects the evaporation of middle-sized droplets most,

and temperature accelerates the inactivation of SARS-CoV-2 both on surfaces and in aerosols.

Future research is needed to improve infection risk models to better predict the infection

potential of different transmission routes. Moreover, further quantitative studies on the

expiratory flow features after wearing a mask are needed. Systematic investigations and the

design of advanced air distribution methods, portable air cleaners, and ultraviolet germicidal

irradiation systems, which have shown high efficacy in removing contaminants, are required

to better control indoor viral infection.

Keywords: air distribution; air cleaner; airborne transmission; environmental influence;

expiratory particle; infection risk; masks; physical distancing; SARS-CoV-2

3
1 Introduction

1.1 Background and scope of discussion

The worldwide spread of coronavirus disease 2019 (COVID-19), which is a severe

respiratory disease caused by the novel severe acute respiratory syndrome coronavirus 2

(SARS-CoV-2), has infected more than 110 million people and resulted in more than 2.4

million deaths as the end of February 2021 (World Health Organization, 2021). This poses a

significant threat to human health, following other well-known pandemics that spread as a

respiratory viral infection during this century, including the 2003 severe acute respiratory

syndrome, 2009 H1N1 influenza pandemic, and 2013 Middle East respiratory syndrome. As

a substantially more severe pandemic than has previously occurred in the world, COVID-19

also considerably hinders the development of cities and society. Therefore, to promote the

construction of sustainable cities and society equipped with epidemic prevention functions

(Megahed & Ghoneim, 2020; Rahmani & Mirmahaleh, 2021), there is an urgent need to seek

out effective precaution and control methods against respiratory viral infection in the built

environment. The success of this is critically dependent on understanding the transmission

characteristics of SARS-CoV-2.

To date, plenty of evidence has shown that SARS-CoV-2 RNA or viable virus can be

detected in air samples collected in the room with COVID-19 patients (Ge et al., 2020;

Lednicky et al., 2020; Y. Mao et al., 2020), indicating that the virus is highly likely to be

spread by expiratory particles (Fig. 1). However, the distribution of viral concentration in

different-sized expiatory particles is still unclear, and is generally assumed to be uniformly

distributed (N. Mao et al., 2020). Generated from different respiratory activities such as

coughing, sneezing, speaking and breathing, expiratory particles generally have a size

between < 0.1 μm and 500 μm (Gralton et al., 2011). The World Health Organization (WHO)

4
defines expiratory particles having diameters > 5 μm as droplets and < 5 μm as aerosols

(World Health Organization, 2014). Aerosols can remain suspended in the air and are

considered airborne. By contrast, according to Tang et al. (2021), there is no definite cut-off

diameter for airborne particles because factors such as the expiratory momentum and ambient

airflow could also affect the ability of a particle to remain suspended. Therefore, ―droplets‖

are defined rather as particles that fall to the ground due to gravity and/or expiratory

momentum, while ―aerosols are particles that remain suspended due to size and/or

environmental conditions. The WHO also regard ―droplet nuclei‖ as a synonym for ―aerosol‖

(World Health Organization, 2020b). However, because ―droplet nuclei‖ explicitly refers to

the residue of dried expiratory particles that result from evaporation of droplets or

aerosolization of infective material (World Health Organization, 2014), they should be

considered as belonging to aerosols (Tang et al., 2021).

Carried by expiratory particles, potential transmission routes for respiratory virus include

contact transmission, droplet transmission, and airborne transmission (World Health

Organization, 2020b). According to the World Health Organization (2014), contact

transmission refers to ―the spread of an infectious agent caused by physical contact of a

susceptible host with people or objects.‖ Droplet transmission refers to ―the spread of an

infectious agent caused by the dissemination of droplets.‖ Airborne transmission refers to

―the spread of an infectious agent caused by the dissemination of aerosols.‖ On the other

hand, despite the rough classification of viral transmission routes mentioned above, the

transmission characteristics of the expiratory virus from the infector to the susceptible can be

very complex. Depending on different respiratory activities, the initial characteristics of

expiratory particles and airflow can be quite different from each other. Moreover, after

exhalation, those particles are exposed to ambient environmental conditions such as

temperature, humidity, and airflow. All these factors can significantly affect the evaporation

5
and dispersion of expiratory particles, and the viability of the virus that is contained.

Therefore, to understand the transmission characteristics of SARS-CoV-2, evaluate the

infection risks caused by infectious expiratory particles, and find effective control strategies,

it is essential to learn the initial properties of expiratory particles and the influence of

environmental factors on the virus transmission.

Meanwhile, during this COVID-19 pandemic, the detection of live viruses in feces (Wang

et al., 2020) has also been reported. In addition, a simulation showed that the toilet flushing

process could induce turbulent flow, which expels aerosol particles out of the bowl (Y. Y. Li

et al., 2020), and viral RNA has been detected in aerosols and on surfaces in hospital toilets

(Y. Liu et al., 2020; Ong et al., 2020), indicating that the fecal-oral transmission route is

highly likely (Hindson, 2020). To date, the fecal-oral transmission mechanism of

SARS-CoV-2 has rarely been discussed; however, this transmission route should not be

ignored. This review only presents relevant topics on viral infection caused by respiratory

activities, considering the numerous existing studies conducted in this field.

According to the transmission routes of expiratory particles, control strategies against

respiratory viral infection may be classified into close-contact (≤ 1.5 m) and long-distance (>

1.5 m) prevention methods. The distance of 1.5 m can be considered as the average threshold

distance for droplet transmission (the maximum traveling distance of large droplets) during

coughing and breathing (Zhang et al., 2020). General strategies such as surveillance, social

distancing, quarantine, and mechanistically based strategies are examples of close-contact

prevention methods. Zhang et al. (2020) classified mechanistically based strategies for

close-contact transmission into three categories according to the following transmission

routes: droplet transmission (surgical masks and face shield), short-range airborne

transmission (N95 respirators, personalized ventilation systems, and push-pull systems), and

immediate body-surface transmission (surface disinfection, hand hygiene, and oral hygiene).

6
By contrast, control strategies for long-range airborne transmission routes mainly include

humidity and temperature control, total volume mechanical ventilation, air filtration (C. Xu,

Luo, et al., 2020), and distant fomite sterilization. This review focuses on common infection

control strategies, such as masks, air distribution methods, air filtration, and disinfection

methods, and suggests their potential in preventing the spread of expiratory particles and

SARS-CoV-2.

1.2 Objective and paper structure

After the outbreak of COVID-19, a significant number of review papers concerning

respiratory viral infections have been published. Among them, several studies set their

research object on expiratory particles, such as the possible role that aerosols play in the

transmission of COVID-19 (Asadi et al., 2020; Wilson et al., 2020), transmission risks of

infectious droplets (N. Mao et al., 2020), and experimental techniques of characterizing

expiratory particles (Mahjoub Mohammed Merghani et al., 2021). Compared to previous

review work, this paper aims to focus on several important issues concerning the

transmission of expiratory particles in the air, which have not been discussed in detail yet.

Selected topics include initial characteristics of expiratory particles and airflow from different

respiratory activities (Section 2), environmental influences on particle evaporation, dispersion,

and viability of SARS-CoV-2 (Section 3), methods used for infection risk assessment of

SARS-CoV-2, and perspective on physical distancing (Section 4), and common infection

control strategies against expiratory particles or SARS-CoV-2 (Section 5). The paper

structure is shown in Fig. 1.

7
Fig. 1 Expiratory particles in respiratory viral infections and the associated control strategies

1.3 Methodology and screening criteria

In this paper, we present a selective review of recent research on expiratory particles in

respiratory viral infections and the associated control strategies. We utilized Web of Science

and Google Scholar to retrieve the literature and prioritized highly cited English, refereed

journal articles, and focused on SARS-CoV-2 regarding the virus type analyzed in this study.

Keywords used for searching papers include ―expiratory particles/droplets/aerosols‖,

―cough/sneeze/talk/breath‖, ―SARS-CoV-2‖, ―environment/temperature/humidity/flow‖,

―evaporation/dispersion/viability‖, ―infection risks‖, ―physical distancing‖ and ―control

strategies/masks/air distribution/filtration/disinfection‖. The transmission process of

expiratory particles discussed in this study starts from the exhalation of the infector, ends in

the exposure by the susceptible; phases of infectious particle generation in the respiratory

tract or inhalation by the susceptible are excluded. For the time range, we have concerned

mainly but not exclusively on research since January 2011 until January 2021, although

8
studies relevant to SARS-CoV-2 are confined to years after 2019. We reviewed 117 papers,

and the number of papers for each topic is shown in Fig. 1.

2 Expiratory particle and flow features

2.1 Size and number of expiratory particles

Table 1 presents previous experimental measurements on expiratory particles and flow

features from different respiratory activities. Significant discrepancies remain in previous

measurements of the size range of particles generated from coughing, sneezing, speaking, and

breathing (Gralton et al., 2011) (Table 1). These contradictions may potentially result from

individual physiological differences, various measurement technologies, and the influence of

evaporation and condensation (Wei & Li, 2016). The size of these expiratory particles range

0.01–500 μm in healthy people and 0.05–500 μm in infected patients; therefore, it may be

concluded that these respiratory activities contain droplets and airborne transmission (Gralton

et al., 2011). This indicates that infection control precautions should consider both

transmission routes. Moreover, the number of small particles (< 1 μm) accounted for the

largest proportion of exhaled particles by coughing and breathing (Fabian et al., 2008; Zayas

et al., 2012) (Table 1).

In terms of the particle numbers produced in different respiratory activities, some

experiments counted the total particles emitted ―per cough‖ or ―when counting 1–100 by

speaking‖ (Chao et al., 2009; Lindsley et al., 2012), whereas others set uniform standards and

measured particle numbers per liter for each respiratory activity (Wei & Li, 2016) (Table 1).

However, to date, there has been no exact value for each activity. Particle numbers of each

size are usually required in simulation boundary conditions or theoretical models. However,

based on the uncertainty and difficulty in counting the exact particle number using the

interferometric Mie imaging technique, there is doubt as to whether such direct measurement

9
is necessary. Compared to the detailed number of total expiratory particles, the experimental

value of number distribution should be more important. The number distribution and total

mass of expiratory saliva (which are much easier to measure) can be used to indirectly

calculate the number for each particle size. The average number of particles per cough from

influenza-infected people was found to be higher (75,400) than the number when they had

recovered (52,200); most of these particles could be inhaled easily because they were in the

respirable size fractions (Lindsley et al., 2012). In another study (Asadi et al., 2019), the

particle emission rate (1–50 particles per second) during speech was measured, and a positive

correlation with loudness was found. However, speech super-emission was not attributable to

loudness and possibly resulted from more individual physiological factors.

2.2 Expiratory flow velocity

It is commonly accepted that coughing, speaking, and breathing represent the descending

order of maximum airflow velocity magnitude ( ) due to different respiratory activities

(H. Zhang et al., 2015). However, documented sneezing velocities vary widely from 4.5 to

100 m/s (Nishimura et al., 2013; O’Hara, 1956; Tang et al., 2013) (Table 1). The

long-standing 100 m/s (O’Hara, 1956) was later challenged by Tang et al. (2013) because it

was inferred from some basic physical principles instead of measurements from actual

sneezes, and the estimated droplet velocity did not necessarily equal the airflow velocity. The

inconsistency of sneezing velocities may also reflect the difficulty in measuring the sneeze at

a required time, because it is not easy to create a sneeze on command compared to other

respiratory activities. Therefore, more methods and experiments are needed to accurately

measure sneeze velocity. In terms of velocity differences between sexes, it was found that the

of coughing and speaking of males was higher than that of females (Han et al., 2021;

Kwon et al., 2012) (Table 1). Researchers found a positive linear correlation between the

10
height of a person and their coughing and speaking velocity (Kwon et al., 2012). A

comparison of the particle transmission characteristics due to different showed that

more droplet nuclei were suspended in air at a slower jet velocity. In addition, the maximum

horizontal distance that a droplet could travel increased with the maximum velocity of

expiratory flow ( ) (X. Xie et al., 2007). was also found to affect the dispersion of

expiratory particles, whereby particles emitted from coughing were dispersed more widely

than breathing (Y. Zhang et al., 2019).

2.3 Duration and maximum direct reach

Gupta et al. (2009) measured the variation in flow rate at the mouth using a

pneumotachograph-based spirometer, to determine the duration of a cough or sneeze ejection

flow. They determined that the duration was approximately 0.3–0.8 s for a cough. Bourouiba

(2016), Bourouiba et al. (2014), and Scharfman et al. (2016) used high-speed imagery to

directly observe the propagation of expelled gas and liquid phases. The measured duration

was approximately 0.3 s for a cough and 0.15–0.25 s for a sneeze (Table 1). These results

show that coughing has a longer duration than sneezing.

Nishimura et al. (2013) defined the maximum direct reach as the distance from the mouth

when the front-line particles lose momentum (equivalent to the natural convention current) or

reaches an apparent plateau velocity. This reach was measured by using a digital high-vision

and high-speed video system combining vector analysis by particle image velocimetry (PIV);

the maximum direct reach was 0.3 m for a cough and 0.84 m for a sneeze (Table 1). Tang et

al. (2012, 2013) used the shadowgraph imaging technique to measure the maximum visible

distance traveled by ejecting puffs. The measured maximum direct reach was 0.7 m for a

cough and 0.6 m for a sneeze (Table 1). The results obtained from these different methods

show apparent disagreements. Therefore, this highlights the need for further experiments to

11
measure the maximum direct reach of particles emitted by different respiratory activities and

a comparison between experimental methods.

Table 1. Expiratory particle and flow features of different respiratory activities


Coughing Sneezing Speaking Breathing
Size range < 0.1–500 < 1.0–125 0.1–125 0.01–100
(μm) (Gralton et al., 2011) (Gralton et al., (Gralton et al., (Gralton et al.,
2011) 2011) 2011)
Number 97% < 1 μm - - 87% < 1 μm
distribution (Zayas et al., 2012) (Fabian et al.,
2008)
Particle 947–2085 - 112–6720 -
number (per cough) (counting 1–100)
(total) (Chao et al., 2009) (Chao et al., 2009)
Particle 24–23,600 - 4–600 1–320
concentration (Wei & Li, 2016) (Wei & Li, 2016) (Wei & Li, 2016)
(per liter)
Mass of saliva1.1–6.7 - 18.7–79.4 -
(mg, total) (per cough) (Xiaojian. (counting 1–100)
Xie et al., 2009; Zhu (Xiaojian. Xie et
et al., 2006) al., 2009)
Maximum 5–22 4.5–100 Female: 2.31 Nasal: 1.4
flow velocity Female: 10.6–13.07 (Nishimura et al., Male: 4.07 Mouth: 1.3
(m/s) Male: 15.2–15.3 2013; O’Hara, (Kwon et al., (Tang et al., 2013)
(Han et al., 2021; 1956; Tang et al., 2012)
Kwon et al., 2012; 2013)
Nishimura et al.,
2013; Zhu et al.,
2006)
Duration of 0.3–0.8 0.15–0.25 - 3.3–5
flow (per cough) (per sneeze) (tidal breath)
expulsion (s) (Bourouiba et al., (Bourouiba, 2016; (Barrett &
2014; Gupta et al., Bourouiba et al., Ganong, 2010)
2009; Scharfman et 2014; Scharfman
al., 2016) et al., 2016)
The maximum 0.3–0.7 0.6–0.84 - Nasal: 0.6
direct reach (Nishimura et al., (Nishimura et al., Mouth: 0.8
(m) 2013; Tang et al., 2013; Tang et al., (Tang et al., 2013)
2012, 2013) 2013)

2.4 Turbulent flow

Wei and Li (2015) discovered that turbulent cough airflow could facilitate the widespread

dispersion of exhaled particles. Small particles (10 µm) traveled with airflow closely over a

12
long distance, and small droplets (30 µm) were quickly transformed into droplet nuclei after

exhalation and behaved similar to small particles. The medium droplets (50 µm) were

sensitive to relative humidity and could still be carried by the jet up to a certain distance even

when settling. The traveling distance of large droplets (100 µm) was mostly dependent on jet

outlet velocity and diameter, and Wei and Li, (2015) utilized the jet model to study cough

turbulence in further detail. However, Bourouiba et al. (2014) found that violent respiratory

flows of coughing and sneezing are turbulent multiphase puffs rather than jets, and puff

buoyancy may be impacted by the ambient conditions. VanSciver et al. (2011) demonstrated

that PIV can be used to map the flow of human coughs in vitro. However, their experimental

results indicated that the human cough cannot be simplified as a typical flow field when

conducting numerical simulations because of the large variance in velocity data among

different subjects and trials.

3 Environmental influence

3.1 Environmental influence on particle evaporation and dispersion

The evaporation process of expiratory particles is mainly affected by the initial properties

of expiratory particles and flow and the ambient environmental conditions. This indicates that

when the respiratory activity type is fixed, the environmental impact factors play a significant

role in droplet evaporation.

Researchers often use different indices to evaluate the effect of particle evaporation and

dispersion. These indices include time-dependent/equilibrium droplet size, droplet

evaporation/settling time, and droplet traveling distance. Some researchers (Wei & Li, 2015)

have developed indices such as the ―reach probability,‖ which aims to characterize the

traveling distance of droplets within a turbulent jet. Some researchers have adopted the

―droplet lifetime‖ (Chaudhuri et al., 2020), which refers to the smaller of the complete

13
evaporation time and settling time; this index can describe the maximum duration a droplet

can exist prior to evaporation to droplet nuclei or settlement.

Table 2 summarizes recent studies on the environmental influence of droplet evaporation.

The most studied environmental factor is relative humidity (RH) (Ji et al., 2018; L. Liu, Wei,

et al., 2017; Wei & Li, 2015). For a single droplet, the time required for evaporation at high

RH is greater than that at low RH for droplets of different sizes (Wei & Li, 2015). For

droplets spread in turbulent buoyant jets caused by coughing, both small (20–30 μm) and

large-sized (100 μm) droplets are insensitive to RH. Small droplets evaporate into droplet

nuclei soon after expiration. Large droplets are deposited on the ground, in which spreading

distance is mainly determined by the jet outlet boundary. By contrast, RH has a significant

effect on medium-sized (50–60 μm) droplets. At low RH, evaporation is complete prior to

falling out of the jet. At high RH, some droplets are deposited out of the jet, whereas some

travel long distances (L. Liu, Wei, et al., 2017; Wei & Li, 2015). Therefore, further analysis

of droplet transmission by turbulent buoyant jets rather than a single droplet is important.

Compared to RH, the influence of ambient temperature and flow has been studied less. Lea

Der Chen (2020) found that when RH is below 37% and remains constant, the temperature

increase is inversely related to droplet evaporation time. Conversely, when RH exceeds 37%,

the temperature increase at a constant RH causes a longer evaporation time. Chaudhuri et al.

(2020) conducted a parametric study and found that droplet lifetime is short at high

temperatures and low RH and high at low temperatures and high RH. Ji et al. (2018)

conducted a computational fluid dynamics (CFD) simulation to compare the impact of

mixing and displacement ventilation. They reported that both ventilation systems can

accelerate evaporation compared to stagnant ambient flow. Mixing ventilation accelerates

droplet evaporation more because its violent turbulence causes more rapid mass transfer.

14
Further studies are required to determine the influence of ambient temperature and flow on

droplet evaporation.

Table 2 shows that theoretical or numerical analysis was conducted in most recent studies.

To verify the feasibility of theoretical models, L. Liu, Wei, et al. (2017) conducted an

experiment using a droplet deposited on a Teflon-printed slide. To avoid inaccuracy caused

by shape change, Chaudhuri et al. (2020) adopted acoustic levitation to suspend a droplet;

this is similar to the state of real droplets in the atmosphere. Due to the many simplifications

in theoretical or numerical studies, there may be inaccuracies generated in the results. For

example, some researchers simplified the model by not considering crystallization and

assuming that partial pressure at the droplet surface was equivalent to that of ambient air (L.

Liu, Wei, et al., 2017). The composition of exhaled droplets was simplified to some extent in

all studies, and this might have played a significant role in droplet evaporation. Therefore,

more efforts are required to refine theoretical models in the future.

Table 2. Summary of recent studies concerning environmental factors on droplet evaporation


Ref. Research Initial properties of expiratory Environmental factors
Method droplets and flow
Diamet Compositi Expiratory R Temperatu Ambient
er (μm) on flow H re (°C) flow
(%
)
Wei & Theoretic 10, 30, Solid + Cough 0, 25 Stagnant
Li al model 50, 100 NaCl = 10 90
(2015) solution m/s
L. Liu, Theoretic 20, 60, Solid + Cough 0, 25 Stagnant
Wei, et al model 100 NaCl = 10 90
al. + solution m/s
(2017) Experime
nt

15
Ji et al. CFD 10, 50, Solid + Breath 0, - Ventilati
(2018) 100, Water 30, on
200 60,
90
Lea Der Theoretic 1, 10, Water Sneeze 0– 20, 30 Stagnant
Chen al model 100, 10
(2020) 1000 0
Chaudhu Theoretic 300– NaCl Sneeze/cou 20 5–30 Stagnant
ri et al. al model 1000 solution gh –
(2020) + = 10 80
Experime m/s
nt

3.2 Influence of flow field on particle dispersion

In addition to the influence of ambient temperature, humidity, and airflow on evaporation

of particles and thus on their dispersion, the ambient flow field also has a direct influence on

particle dispersion. For example, thermal stratification may extend the scope of short-range

airborne infections (Qian & Zheng, 2018). Zhou et al. (2017) developed a non-dimensional

theoretical buoyant jet dispersion model and observed that exhaled flow could freely float

upward within a thermally uniform environment, while fluctuating at a certain height in a

thermally stratified environment. Based on this work, F. Liu et al. (2019) developed a jet

integral model and reported that the lock-up phenomenon within a thermally stratified

environment significantly impeded concentration decay. They also observed that the longest

distance of direct exposure occurred when the breathing height of the susceptible person was

at the lock-up layer in distribution ventilation. Conversely, Ai et al. (2019) found that the

disturbance in horizontal air flow supply could strengthen the mixture of exhaled air and

ambient air, thus reducing the exposure risk of close contact and making monotonically

decreasing relationship between exposure risk and separate distance unobvious. Indoor

16
airflow patterns are typically adjusted by applying engineering control strategies; based on

the complexity, the influence of different air distribution methods on the transmission of

expiratory particles is specifically discussed in Section 5.2.

3.3 Environmental influence on virus viability

Based on the transmission routes of expiratory particles, studies regarding the

environmental influence on virus viability are categorized into two types: those regarding

droplets on the surface and those regarding suspended aerosols. SARS-CoV-2 is more stable

on smooth surfaces such as plastic, stainless steel, glass, banknotes, and the outer layers of

surgical masks, but less stable on printing and tissue papers, treated wood, cardboard, and

copper (Chin et al., 2020; van Doremalen et al., 2020). Under common room temperature and

RH conditions, SARS-CoV-2 was found to survive for up to 72 h on plastic (21–23 °C, 40%)

(van Doremalen et al., 2020), and infectivity was retained on respirable-sized aerosols for up

to 16 h (23 °C, 53%) (Fears et al., 2020). These measurements indicate that airborne and

contact transmission of SARS-CoV-2 is plausible, and these risks may persist over a long

period if proper infection control strategies are not implemented.

For surface-deposited droplets of SARS-CoV-2, higher temperatures were found to cause

more rapid inactivation of the virus (Chin et al., 2020; Riddell et al., 2020). However, there

remains a discrepancy in the influence of RH on viral stability; Biryukov et al. (2020) and

Matson et al. (2020) claimed that a higher RH facilitates the decay of SARS-CoV-2, whereas

Morris et al. (2020) stated that the virus has greater survivability at extremely low or high RH.

This discrepancy may be caused by the limited cases and the lack of rigorous control of

variables in some studies, requiring further confirmation experiments and discussions.

For SARS-CoV-2 in suspended aerosols, Smither et al. (2020) considered that aerosolized

media containing the virus may influence viral stability under different RH. The virus was

17
more stable at medium RH than in higher RH in tissue culture media, whereas the opposite

result was observed in artificial saliva. Dabisch et al. (2021) found that the effect of

temperature or ultraviolet (UV) radiation was much higher than that of humidity, with higher

temperature and UV increasing the viral decay rate. In summary, these results also highlight

the poor understanding regarding the influence of humidity on the infectivity of

SARS-CoV-2 in aerosols. To date, no studies have been conducted on the infectivity

variation of size-fractioned particles under different environmental conditions; such research

may aid the identification of differences in viral viability in aerosols and large droplets.

Although evidence has shown that the stability of SARS-CoV-2 could be a function of

environmental factors such as temperature, humidity, and UV, the joint effect of these factors

remains unclear. Based on experimental data, Morris et al. (2020) proposed a mechanistic

model to estimate the joint effect of temperature and humidity on viral viability, including

SARS-CoV-2 and other enveloped viruses. This type of model can be improved in the future

after addressing the discrepancies and difficulty in interpreting the existing experimental

phenomena. These improvements would be significantly beneficial in the prediction of viral

viability under conditions that cannot be observed.

Furthermore, in many studies, meteorological and virus epidemiological data have been

used to determine the relationship between climatic conditions and the seasonality of

confirmed COVID-19 cases. For example, Yao et al. (2020) focused on major Chinese cities

from January to March and indicated that the communicability of COVID-19 is negatively

correlated with high ozone concentrations, high temperatures, and low RH. V et al. (2020)

compared data from March to April in various Indian states and found that a decrease in RH

is positively correlated with the number of confirmed cases. The dispersion characteristics of

expiratory particles and the survivability of SARS-CoV-2 in different climate conditions may

potentially influence the seasonality of confirmed COVID-19 cases. However, these findings

18
should be interpreted with caution because these influences are confounded by many other

complicating factors, such as human behavior and circannual variation in immunity (Morris

et al., 2020).

4 Infection risk assessment

4.1 Exposure risks

If the virus content of expiratory particles is not considered, the quantification of infection

risk can be simplified to the risk of exposure to expiratory particles. Exposure risks can be

assessed more easily than infection risks where there is limited knowledge of virus

distribution on expiratory particles as well as variation over time of virus viability according

to environmental conditions.

As mentioned in introduction section, droplets can only remain in the air for a short time,

and thus they transmit over short distances before depositing, while aerosols can disperse

over long distances and remain airborne for longer periods. The short-range transmission

routes of respiratory particles include both droplet and airborne transmission, whereas the

long-range transmission route of expiratory particles generally refers to airborne transmission.

Therefore, the cut-off distance for ―short-range‖ and ―long-range‖ depends on the maximum

horizontal distance that the droplets can reach. This threshold distance is determined by many

factors, including the initial properties of expiratory particles and flow, as well as ambient

environmental conditions. For example, a study by X. Xie et al. (2007) revisited the Wells

evaporation-falling curve (Wells, 1934, 1955). They found that expelled large droplets could

be carried by expiratory airflow more than 3‒6 m away at a velocity of 20‒50 m/s (sneezing),

more than 2 m away at a velocity of 10 m/s (coughing), and less than 1 m away at a velocity

of 1 m/s (breathing) at room temperature and humidity conditions (20 ℃, 50%).

19
Therefore, when a susceptible and an infected person are in close contact, with both

droplet and airborne transmission, the exposure risks to expiratory particles are expected to

be highest. However, consensus on the relative contribution of each transmission route to

close-contact exposure has changed several times. Traditionally, close-contact exposure is

believed to be mainly due to droplet transmission (Brankston et al., 2007). Later, L. Liu, Li,

et al. (2017) pointed out that both droplet and airborne transmission can be important for

close-contact (< 1.5 m) exposure to expiratory particles. By contrast, a recent study by

Wenzhao Chen et al. (2020) found that when a susceptible person is closer (< 2 m) to an

infected person, the short-range airborne transmission contributed most to the exposure risk

of expiratory particles, while the droplet transmission only dominated when the droplets were

over 100 μm at a distance within 0.2 m when talking, or 0.5 m when coughing, which is

fairly negligible.

4.2 Infection risk model

The Wells-Riley and dose-response models are two common approaches for risk

assessment of infectious respiratory diseases. The Wells-Riley model has been extensively

used. It allows simple and quick evaluation, based on the quantum concept that implicitly

considers the infectivity, infectious source strength, and biological decay of pathogens. By

contrast, the dose-response model is less frequently used because it requires information that

is costly to obtain through experimental and on-site studies (Sze To & Chao, 2010; S. Zhang

& Lin, 2020).

As for COVID-19, most infection risk models have been proposed based on the

Wells-Riley model, which is shown in Equation (1) (Riley et al., 1978; Wells, 1955).

(1)

20
where is the probability of infection, is the number of infectors, is the quanta

emission rate by one infector (quanta/h), is the breathing rate of each susceptible person

(m3/h), is the exposure time interval (h), and is the room ventilation rate (m3/h).

Quanta refers to the number of infectious airborne particles required to infect a person

(Wells, 1955). The Wells-Riley model assumes a spatially and temporally uniform

distribution of pathogen-laden aerosols. As shown in Table 3, Buonanno et al. (2020) and

Kriegel et al. (2020) integrated the time-varying quanta concentration (quanta/m3) to assess

infection risk under transient conditions. Kriegel et al. (2020) also considered the

time-varying viability of viruses. Zhang & Lin (2020) proposed a dilution-based evaluation

method. This expansion of the Wells-Riley model can assess the airborne infection risk for

both spatial and temporal resolutions. Furthermore, considering the simplification of the

Wells-Riley model, the social distance index and ventilation factor (Sun & Zhai, 2020), as

well as the filtration effect of masks (Dai & Zhao, 2020) were introduced to improve the

prediction accuracy of the model.

To apply the Wells-Riley model to infection risk assessment effectively, many efforts have

been made to determine the quanta emission rate ( ) (quanta/h) value of COVID-19 (Table 3).

Buonanno et al. (2020) estimated values based on the viral load in expiratory particles and

the quanta-RNA copies correction factor, which are influenced by different respiratory

activities, as well as different activity levels (such as resting, standing, light exercise,

moderate exercise, and heavy exercise). Sun & Zhai (2020) and Miller et al. (2021)

back-calculated using other known parameters from actual pandemic cases. Dai & Zhao

(2020) estimated with the basic reproductive number of COVID-19 based on the fitted

equation from other airborne transmitted infectious diseases. Kriegel et al. (2020) combined

the calculation methods of Dai and Zhao (2020) and Buonanno et al. (2020). They estimated

using the basic reproductive number of COVID-19 and modified it according to human

21
activities. As a result, there are great divergences among values from these studies.

Among them, the largest explicit value for airborne transmission of SARS-CoV-2 has

been estimated is 970 during singing and vocalization (Miller et al., 2021). The > 100

calculated by Buonanno et al. (2020) can be higher than 1000 if the viral load in sputum is

extremely high. Therefore, we can only obtain an approximate upper limit value of

approximately 1000, which is back-calculated from common cases, without considering the

super-spreaders. Comparing the three above-mentioned calculation methods for the value

of SARS-CoV-2, inferring from basic productive number involves data on many other virus

types, estimating from viral load and quanta-RNA copies correction factor involves the data

on SARS-CoV-1, while the back-calculation method only involves the data on the target

virus. Considering the large errors introduced in the data collection of other virus types, the

quanta value of SARS-CoV-2 obtained from the back-calculation method may be more

credible. Moreover, because the value is influenced by many factors (such as different

respiratory activities and different activity levels), as Buonanno et al. (2020) claimed, further

studies are needed to classify different situations and examine the appropriate values.

The Wells-Riley model was originally proposed to assess the infection risk of airborne

transmission routes. Some studies have modified this model to evaluate the infection risks of

other transmission routes. Sun & Zhai (2020) predicted the risk of droplet transmission

infection by inserting a social distancing index. Gao et al. (2021) combined the Wells-Riley

and dose-response model to examine the relative contributions of different transmission

routes, including long-range airborne transmission, respiratory close contact transmission,

and contact transmission. These trials are crucial for future infection risk prediction, although

the appropriateness of the parameters integrated into the expanded Wells-Riley models,

which are derived from Poisson distribution, should be carefully assessed.

22
Table 3. Summary of recent studies on expansions of the Wells-Riley model for COVID-19
Ref. Modification Quanta emission rate Calculating method for value
( ) value (quanta/h)
(Buonanno Integrate > 100 Estimate based on the viral load in
et al., time-varying (asymptomatic, expiratory particles and the
2020) quanta vocalization, light quanta-RNA copies correction factor,
concentration activities) which are influenced by different
< 1 (symptomatic, respiratory activities, and different
resting condition) activity levels
(Sun & Introduce 856.8 (speaking, Back-calculate with other known
Zhai, social distance droplet transmission) parameters from actual pandemic
2020) index and cases
ventilation
factor
(Dai & Introduce 14–48 Estimate with the basic
Zhao, filtration effect (asymptomatic, light reproductive number of COVID-19
2020) of masks exercise) based on the fitted equation
(S. Zhang 1. - -
& Lin, Dilution-based
2020) 2. Consider
spatially and
temporally
non-uniform
(Kriegel et 1. Integrate 139 (breathing/ Estimate with the basic
al., 2020) time-varying speaking, sitting/ reproductive number of COVID-19
quanta standing) based on the fitted equation, and
concentration modify according to human activity
2. Consider levels
time-dependent
viability of the
virus

23
(Miller et 1. Integrate 970 ± 390 (singing/ Back-calculate with other known
al., 2021) time-average vocalization) parameters from actual pandemic
quanta cases
concentration
2. Consider
deposition and
virus decay

4.3 Perspective on physical distancing

Safe physical distancing is difficult to determine as many factors, such as initial properties

of expiratory droplets and flow, viral load, infectious dose, environmental influence on

expiratory particles, and virus viability, need to be considered.

Table 4 summarizes recent studies concerning physical distancing applied to COVID-19.

Feng et al. (2020) and Dbouk & Drikakis (2020a) define safe physical distance as the furthest

distance that coughing droplets can reach. Both studies conducted numerical simulation and

examined the significant impact of ambient wind on physical distancing. Sun & Zhai (2020)

considered the infection risks from respiratory droplets. They modified the Wells-Riley

model and indicated the minimum physical distance to be 1.6‒3.0 m when only breathing or

speaking.

Other studies calculate physical distancing by considering the infection risks from both

droplets and droplet nuclei. F. Yang et al. (2020) utilized a dose-response model and showed

that physical distancing is insufficient to control infection probability and that decreasing

contact time is also important. They suggested that the contact time should be less than 8 min

for a 1 m distance and 16 min for a 2 m distance from the speaking or breathing infector.

Mittal et al. (2020) adopted a protection factor based on concentration decay to compare the

infection risks at different physical distances. They found that when the velocity of the

crossflow is much larger than that of the exhaled jet, the protection factor increases faster as

24
the distance increases. Furthermore, buoyancy also influences the rate of increase in the

protection factor. Rosti et al. (2020) conducted a numerical simulation, systematically

comparing eight coughing scenarios selected from past research. They found that different

initial conditions influenced the results. Therefore, further experiments on the initial

properties of expiratory particles are needed in the future.

Due to strong evidence of the infection potential of droplet nuclei, it is inadequate to

estimate a safe physical distance using only the maximum reaching distance of expiratory

droplets, although this could specify the lower limit of physical distancing. Although some

studies considering infection risks of expiratory particles suggested detailed physical

distancing values, the viral loads assumed in these models are questionable, requiring further

credible real cases or experiments on infection capability of expiratory particles carrying

SARS-CoV-2.

Table 4. Summary of recent studies concerning physical distancing applied to COVID-19


Ref. Methodology Respiratory Environment Eligibility Physical
activity condition standard distancing
recommendations
Feng et Numerical Cough/ Temperature: Little Over 1.83 m due
al. simulation sneeze 27 °C droplet to wind
(2020) RH = 40%, reach convection
99.5%
Dbouk Numerical Mild cough Temperature: Little Under 2 m with
& simulation 20 °C droplet no wind;
Drikakis RH = 50% reach Over 2 m with
(2020a) parallel wind
existing
Sun & Theoretical Speaking/ Temperature: Infection 1.6–3.0 m
Zhai model breathing 25 °C probability
(2020) (Modified under 2%
Wells-Riley

25
model)
Mittal et Theoretical Breathing Temperature: High Decreases with
al. model 0 °C, 42 °C protection strong crossflow
(2020) factor & light plume;
increases with
heavy plume
F. Yang Theoretical Speaking & RH = 50% Infection 8 min for 1 m;
et al. model Breathing risk under 16 min for 2 m
(2020) (Dose-response 0.63
model)
Rosti et Numerical Cough RH = 40%, Low Conflicting results
al. simulation 60% cumulative based on current
(2020) viral load knowledge

5 Infection control strategies

5.1 Masks

According to the World Health Organization (2020a), masks used in healthcare settings

include surgical masks and filtering facepiece respirators (FFR), or respirators. Surgical

masks can filter 3 μm droplets, while respirators must filter 0.075 μm solid particles.

Moreover, surgical masks have ―an open shape and a potentially leaking structure.‖ By

contrast, the outer edges of the FFR should seal around the wearer’s face to guarantee

claimed filtration. Exhalation valves for respirators are discouraged because they ―bypass the

filtration function of exhaled air‖ (World Health Organization, 2020a).

Many studies have confirmed the importance and efficacy of masks against viral

respiratory infections. Zhai (2020) claimed that wearing a facial mask is necessary to beat

COVID-19. Feng et al. (2020) analyzed a computational fluid-particle dynamics model and

revealed that wearing facial masks, even loosely while coughing, could significantly reduce

26
the suspension of small particles in the air. Leung et al. (2020) indicated that surgical masks

could efficiently reduce the release of influenza virus particles in droplets rather than aerosols

but could reduce the coronavirus both in large droplets and aerosols. Moreover, depending on

the wearer’s activity level, filtration efficiency of surgical masks can be 44% (at rest) or 10%

(in moderate activity) for 0.3‒6 µm aerosols, if considering the influence of air leakage

through the mask fitting gap (Konda et al., 2020).

In addition to filtering expiratory particles, masks also influence the characteristics of the

expiratory flow field. As shown in Table 5, most studies have qualitatively described the

features of the expiratory flow field under the influence of different masks. All types of

masks can reduce the front throughflow (Dbouk & Drikakis, 2020b; Kähler & Hain, 2020;

Tang et al., 2009). However, surgical and handmade masks can generate significant leakage

jets, although the jet has been redirected, which causes less harm in face-to-face

circumstances (Kähler & Hain, 2020; Tang et al., 2009; Viola et al., 2020). However,

quantitative studies on the expiratory flow field when wearing a mask are limited. Only the

distance of the filtered front throughflow has been measured using the Schlieren optical

technique or PIV or simulated using CFD (Dbouk & Drikakis, 2020b; Kähler & Hain, 2020;

Viola et al., 2020). Neither the velocity of the front throughflow nor the leakage flow in other

directions has been measured. Considering that an increasing number of people choose to

wear masks in public, the characteristics of expiratory particles and airflow when wearing a

mask should be quantified.

Despite the limitations of relevant studies in this area, we can still acquire some knowledge

on the usage of facial masks. A mask will not completely prevent the front dispersion of the

respiratory flow; therefore, physical distancing is still essential. A greater distance is required

for masks with a lower filtering efficacy. Nevertheless, when the environment is

contaminated with a known source of infection, respirators are necessary. Moreover, due to

27
the mask fitting gap, when sneezing or coughing it is advisable to use the elbow to prevent air

leakage.

Table 5. Recent studies on expiratory flow field when wearing a mask


Ref. Method Respiratory Mask type Influence on expiratory flow
activities field
Tang et al. Schlieren optical Coughing Surgical and Surgical mask can redirect
(2009) technique N95 masks the cough jet to reduce harm;
N95 mask can block the
formation of the jet
Dbouk & Multiphase Mild coughing Surgical mask The bulk of expiratory
Drikakis computational particles will travel about
(2020b) fluid dynamics 70 cm without a mask, but
35 cm with a mask;
Mask efficiency will drop
about 8% after 10 cough
cycles
Kähler & Particle Image Coughing Mouth-and-nose The flow resistance of masks
Hain Velocimetry cover, surgical can prevent the spread of
(2020) mask, respirator exhaled air;
Expiratory flow can leak
through the edge gaps of
mouth-and-nose cover
Viola et al. Background Quiet and FFP2 and FFP1 All masks can reduce the
(2020) oriented heavy masks, a front throughflow by over
schlieren breathing, respirator, a 63%;
technique coughing surgical, Surgical and handmade
a handmade masks generate significant
mask leakage jets

5.2 Air distribution methods

According to Qian & Zheng (2018), there are three key elements of ventilation that affect

particle transmission: flow direction, ventilation rate, and airflow pattern. A pressure

difference can be used to control the flow direction to prevent cross-infection between zones.

Increasing the ventilation rate is useful for reducing long-range airborne infection but might

not help control short-range droplet transmission when the droplets are more influenced by

gravity and exhalation velocity. The effect of ventilation rate on short-range airborne

infections requires further investigation. Dai and Zhao (2020) estimated the necessary
28
ventilation rate based on the Wells-Riley model and the reproductive number of COVID-19,

as mentioned before. To control the infection probability at less than 1%, ventilation rates

greater than 100‒350 m3/h per infector and 1200‒4000 m3/h per infector for 0.25 h and 3 h of

exposure, respectively, are required. However, wearing an ordinary surgical mask could help

reduce the required ventilation rate to a quarter, which is more achievable using normal

ventilation modes. Considering the limitations of the Wells-Riley model, the accuracy of the

suggested ventilation rate needs further discussion. Moreover, when the ventilation rate

reaches a certain value, increasing the ventilation rate has no significant effect on reducing

the transmission of aerosols (Y. Zhang et al., 2019).

Ventilation systems can influence airflow patterns through different arrangements of

diffusers and exhausts. Designing ventilation systems is complex, especially for large and

complicated spaces. Sometimes, a combination of several air distribution methods is adopted

to achieve a satisfactory environment. Therefore, here, we only summarize the basic

characteristics of selected traditional and advanced air distribution methods.

(1) Total volume air distribution (TVAD)

Total volume air distribution (TVAD) methods refer to ventilation strategies at a

whole-room scale. Mixing ventilation (MV) and displacement ventilation (DV) are two

traditional and principal ventilation types. MV aims to supply air at a high velocity to

facilitate the mixing of the entire indoor air and dilution of contaminants, while DV aims to

utilize the buoyancy force from the heat source to remove the contaminant in the breathing

zone. In recent years, advanced air distribution methods have been developed and expanded

based on the mechanisms of MV and DV (Cao, Awbi, et al., 2014; B. Yang et al., 2019). The

characteristics of these air distribution methods for removing airborne contaminants are

summarized in Table 6.

29
Concerning the physical mechanisms of these ventilation strategies, downward ventilation

(DWV), underfloor ventilation (UFV), impinging jet ventilation (IJV), confluent jet

ventilation (CJV), and wall-attached ventilation (WAV), all utilize the principle of DV in the

latter process of air distribution. Except for DWV, the other four systems are all low-level air

supply systems with high-velocity supplied air initially. CJV and WAV were both developed

from the IJV. They deliver downstream flows with high momentum that impinge the floor,

form a very thin shear layer along the floor, and reach a region further than DV. Meanwhile,

all three ventilation systems can operate in heating mode. Stratum ventilation (SV) delivers

clean air horizontally into the breathing zone, while Piston ventilation (or laminar airflow)

can supply clean air vertically or horizontally across the whole room and swipe airborne virus

away in a ―washing effect.‖ Strictly speaking, Protected occupied zone ventilation (POV) is

not a type of TVAD, as it separates the indoor space into several subzones through

downward-plane jets with low turbulence, which can effectively reduce short-range

cross-infection (Cao, Sirén, et al., 2014). Moreover, MV or DV can be combined with the

POV in each subzone to achieve better ventilation efficacy.

30
Table 6. Comparison of the efficacy of air distribution methods in removing airborne
contaminants
Ventilation type Mechanism Advantages & Limitations
Applications
Mixing ventilation Ceiling-based Can create a uniform Can enhance the
(MV) diffuser introduces indoor environment dispersion of airborne
air at high velocity to achieve thermal contaminants,
to facilitate mixing comfort (B. Yang et Occupants may be
of the indoor air al., 2019) exposed to infectious
(Bolashikov & particles regardless of
Melikov, 2009) source location;
Low energy
efficiency
Displacement Slightly cooler air Can minimize indoor Cannot be used in
ventilation (DV) (cooler than mixing, ideally keeps heating mode to
ambient air) is contaminant away avoid full mixing;
delivered at floor from the breathing With intensive heat
level at low zone; sources, the supplied
velocity, moving Natural or air penetrates limited
upwards and mechanical DV can distance;
entrained by flows be alternatives of the Thermal stratification
generated from negative pressure lock up phenomena
heat sources, ventilation for can increase the
finally extracted at makeshift hospitals dispersion distance of
ceiling height (Bhagat & Linden, exhaled jet (Qian &
(Bolashikov & 2020) Zheng, 2018)
Melikov, 2009)
Downward ventilation Cooler air is When the downward When the downward
(DWV) supplied from a flow is supplied to air is supplied
ceiling diffuser areas outside the towards the occupied
with low velocity, occupied zone with zone with thermal
giving a downward exhausts at a high buoyancy, this system
flow. The local location, this system will operate like MV,
flow pattern can be can remove warm and facilitate the
similar to DV or aerosols as per DV, dispersion of airborne
MV depending on and handle a high contaminants
the distribution flow rate without throughout the space
position of causing high (Qian et al., 2008)
downward flow velocity, which is
(Nielsen et al., suitable for hospital
2010) wards (Nielsen et al.,
2010)

31
Underfloor ventilation Cooler clean air is Can facilitate rapid The high-speed
(UFV) delivered at floor mixing of air in the supply jet promotes
level with higher occupied zone, with resuspension of
velocity through the space above the particles from the
many diffusers. occupied zone floor and into the
The air undergoes stratified to prevent breathing zone
good mixing within the return of (Bolashikov &
the occupied zone, contaminants; Melikov, 2009)
then starts to lift Close to occupants
upwards similar to for easy control
DV (B. Yang et al.,
2019)
Impinging jet A jet of air is Fresh air can be Because of the
ventilation (IJV) supplied delivered directly to discomfort of the
downwards with the occupied zone draught, the
high momentum at due to thermal application is
a certain height, stratification; suggested in
then strikes and Heated air can be scenarios where
spreads over the supplied in winter, occupants remain in
floor, forming a which has sufficient fixed positions
very thin shear momentum to (Haghshenaskashani
layer with a far overcome the & Sajadi, 2018)
reach. The exhaust buoyancy force
is generally at the generated by heat
ceiling level sources (B. Yang et
(Karimipanah & al., 2019)
Awbi, 2002; B.
Yang et al., 2019)
Confluent jet Circular jets are Confluent jets have
ventilation (CJV) delivered in slower velocity
parallel directions decay than other jet
in the same plane, forms due to less
coalescing at a entrainment of the
certain distance ambient air, therefore
downstream and the momentum is
moving as a single conserved better and
jet (Cho et al., the confluent jets can
2008) penetrate further in
the occupied zone
(Andersson et al.,
2018)
Wall attached Fresh air is Can be used in both
ventilation (WAV) delivered from the cooling and heating
linear slot inlet, mode;
attached to the Simple to install (A.
sidewall or column Li et al., 2019)
surface due to the
Coanda effect,
32
moving
downwards,
impinging and
spreading over the
floor (A. Li et al.,
2019)

Stratum ventilation Clean air is Forms a fresh air Supplied air


(SV) delivered layer and reduces temperature and
horizontally, contaminant contaminant source
forming the lowest concentration in the position can affect the
air temperature and breathing level (Lu et performance of SV
highest air velocity al., 2020) (Tian et al., 2010)
at the breathing
zone (Lu et al.,
2020)
Piston ventilation / Air is delivered The laminar airflow Large energy
Laminar airflow vertically or can swipe airborne consumption due to
horizontally across virus away in a high air change rate;
the whole room at ―washing effect‖ (B. Lack of consensus on
low velocity and Yang et al., 2019) efficacy in reducing
turbulence to create surgical site infection
a ―piston‖ type (B. Yang et al., 2019)
flow (Cao, Awbi,
et al., 2014)
Protected occupied Indoor space is Can protect May cause noise
zone ventilation (POV) separated into a susceptible people issues, thermal
few subzones by from respiratory viral discomfort, and
downward low contaminants, reduce require high energy
turbulence plane the risk of consumption (Cao,
jets (Cao, Sirén, et cross-infection at Awbi, et al., 2014)
al., 2014) short-range (Cao,
Sirén, et al., 2014)

(2) Personalized ventilation and personalized exhaust

To control the microenvironment around each occupant, personalized ventilation (PV) and

personalized exhaust (PE) are commonly used. Compared with TVAD, these systems are

effective in cleaning polluted air locally and from the source, can be controlled by occupants

individually or self-respond to physiological signals, be quickly modified to suit requirements,

and reduce energy consumption (Melikov, 2016).

33
PV delivers fresh air to the breathing zone of occupants directly, and therefore, the

interactions between PV airflow and this microenvironment should be carefully considered

(C. Xu, Wei, et al., 2020). The distance between the exposed individual and PV terminal, as

well as the relative position of the PV terminal, infected individual, and susceptible

individual, are both important factors affecting infection exposure risks (J. Xu et al., 2020). It

has been widely documented that PV could reduce cross-infection risks when used by the

exposed individual; however, the use of PV by the infected individual should be avoided

owing to its further dispersion of exhaled infectious particles (Ai et al., 2019). Another

important factor that influences the cross-infection risk is the air volume of the PV. When

occupants are at a close distance, PV only protects against infection with a high ventilation

efficacy in the inhalation zone by reducing the entrainment of infectious flow and by being

operated with a higher clean air volume (C. Xu, Wei, et al., 2020). As higher velocity might

cause discomfort, an intermittent PV system that balances thermal comfort and good air

quality is proposed (Assaad et al., 2018). Dynamic PV airflow could cause a higher exposure

risk than constant PV airflow due to higher turbulence intensity (J. Xu et al., 2020).

Furthermore, the penetration of the PV jet into the thermal boundary layer (TBL) was not

found as challenging as expected. The interaction between the PV airflow and the TBL alters

the airflow distribution in the breathing zone and affects the inhaled air quality, and this

should be carefully monitored (C. Xu et al., 2018).

PE works by extracting air locally (J. Yang et al., 2014). Either top-PE or shoulder-PE

shows excellent performance in reducing infection risks of airborne transmission when used

by an infected individual (J. Yang et al., 2015a). However, PE can cause the under-pressure

problem, if the contaminated air is expelled from the room without being balanced with the

supply airflow (J. Yang et al., 2014).

34
When comparing the efficacy of PV and PE, the single use of a PE system for the infected

individual performed better than the single use of a PV system for the susceptible individual

(J. Yang et al., 2015b). When combining these two solutions, the additional installation of a

PE system for the susceptible individual can facilitate the clean PV flow to reach the

breathing zone; this improvement of inhaled air quality can be better achieved under the

background ventilation of DV than MV (J. Yang et al., 2014). On the other hand, when the

infected individual uses the PE system, the PV used by the susceptible individual can

increase or reduce the exposure risks, depending on many influential factors (J. Yang et al.,

2015b). According to the above studies, layouts and operating modes of the PV and PE,

positions of people, and ambient environmental factors, might influence the efficacy of the

PV and PE systems. Therefore, as suggested by Ai & Melikov (2018) as well, further

systematic investigation of different scenarios is needed.

Because PV and PE affect the air around individuals, both droplet and airborne

transmission of the virus should be considered. Past studies utilized tracer gas to simulate the

expiratory particles to investigate the airborne transmission performance (J. Yang et al., 2014,

2015a, 2015b). Some recent studies have utilized nebulizers to generate particles in the size

range of breathing or talking (C. Xu, Wei, et al., 2020; J. Xu et al., 2020). Further studies are

required to assess the control of droplet transmission by PV and PE during violent respiratory

activities, such as coughing and sneezing, which generate larger droplets with different size

distributions.

(3) Occupancy-aided ventilation

Most existing ventilation systems are designed to be energy-efficient under non-pandemic

conditions. When the airborne transmissibility of a certain type of virus is extremely high,

like the SARS-CoV-2, the ventilation rate of the system may be insufficient to dilute the

35
concentration of viral aerosols. In this case, reducing occupancy can reduce the risk of

infection by reducing virus from the source. Based on this principle, Melikov et al. (2020)

proposed an improved control strategy in which occupants were asked to leave the ventilated

room at set intervals. The ventilation should be in operation before any occupant enters the

room. Short room occupation times with long breaks are recommended. However, this

occupancy-aided ventilation may reduce work productivity. Therefore, Zhang et al. (2021)

further optimized the occupancy schedule to maximize the total duration of normal

occupancy, balancing low airborne infection risks and work productivity. Despite the

innovation of occupancy-aided ventilation, there are limitations such as potential corridor

blockage, simplification of work productivity evaluation, assumption of uniform virus

distribution, and difficulty in obtaining infector numbers (S. Zhang et al., 2021), and further

research is required.

5.3 Air filtration

Air filtration is another method that can assist in reducing the indoor virus concentration

when the existing ventilation system cannot provide an adequate air change rate. A temporary

anteroom with air filtration devices could help remove virus particles and convert a general

patient room into an isolation space (Mousavi et al., 2020). Air filtration methods are divided

into in-duct devices (filters) and portable (stand-alone) air cleaners. In-duct devices work for

the entire house but function only during the operation of an air-handling system where they

are installed (Wenhao Chen et al., 2005). Portable air cleaners (PACs) can be operated and

positioned in a room with the flexibility to target the problem areas and have received

increasing attention in recent years.

The particle-removal performance of air filtration devices is evaluated by the clean air

delivery rate (CADR), which is the product of the single-pass efficiency and airflow rate.

36
According to particle filtration efficiency, air filters are divided into four types, which is pre

filter, medium filter, high-efficiency particulate air (HEPA) filter and ultra-low particulate air

filter (G. Liu et al., 2017). Among these, HEPA filters have shown excellent performance in

removing expiratory aerosols and are most widely adopted in the market. As shown in Fig. 2,

they have at least 99.97% efficiency for removing particles in the range of 0.15 to 0.2 μm,

which is the most penetrating particle size (MPPS) due to the combined effect of different

mechanisms of particle capture (Christopherson et al., 2020). Experiments by Qian et al.

(2010) found that a common HEPA-filter PAC could reach an effective air change rate

(dividing CADR by room volume) from 2.7 to 5.6 1/h in the ward with a size of 109 m3.

However, regarding the efficacy of PAC on removing viral particles, several assumptions

have been made in studies to date, such as neglecting the influence of particle evaporation

and loss of particle infectiousness, assuming a constant rate of influenza emission, and using

NaCl particles as an analog to the viral particles (Zuraimi et al., 2011). Thus, further research

on the suitability of these assumptions is required. Moreover, the purification performance of

an air cleaner can degrade over time. Pei et al. (2020) showed that the standard cumulate

clean mass test in a laboratory using cigarette smoke did not represent the actual dust loading

capacity and that using the initial CADR value could overestimate the daily purification

capability. Therefore, the authors proposed a new PAC service life evaluation method that

considers decay in actual applications.

PACs usually discharge cleaned air with strong momentum, and it was found that the

HEPA filter PAC could produce global air mixing indoors at high speeds (Qian et al., 2010).

Moreover, flow interaction between other momentum sources (e.g., ventilation systems) and

PAC could play a significant role in indoor air cleaning. Thus, the parameters and positioning

of a variety of momentum sources should ensure a good air circulation cycle to enhance air

purification efficacy (T. Zhang et al., 2010). When airborne particles were distributed

37
uniformly indoors, the flow rate had the greatest impact on the indoor particle concentration,

and the position of the air cleaner affected the particle concentration at low volumetric flow

rates. A minor difference was found between the effects of horizontal and upward ejection

(Jin et al., 2016). By contrast, under non-uniform aerosol distributions, caused by smoking

and coughing, the position of the air cleaner significantly impacted the particle concentration,

which was also affected by the expiratory airflow velocity and the orientation of the air

cleaner (Lin Chen et al., 2017). When the PACs were applied in wards, Mousavi et al. (2020)

found that if only one machine was used, the optimal placement was close to the patient’s bed,

and if two machines were used, the optimal placement was on either side of the plastic

division wall between the isolation space and the anteroom. From previous studies, we can

learn that the contaminant sources, sinks and airflow can significantly affect the PAC

efficacy in removing indoor viral contaminants. Thus, further systematic studies concerning

the influence of the position, orientation, and flow rate of air cleaners are needed. Moreover,

the influence of expiration characteristics under different respiratory activities, purifying

efficacy under different ventilation systems, and purifying efficacy with the interaction of

multiple air cleaners also warrant further study.

Fig. 2 HEPA efficiency as a function of particle size and filtration mechanism

38
(Christopherson et al., 2020)

5.4 Disinfection

Compared to filtration, which refers to the removal of virus particles from airstreams

passing through the filter, disinfection methods focus on the inactivation of viruses

circulating in the air or depositing on surfaces. Possible COVID-19 disinfection technologies

include ultraviolet germicidal irradiation (UVGI), electrostatic spraying, disinfectant fogging,

ozone generators, plasma, photocatalytic disinfection, and heat sterilization (T. Chen &

O’Keeffe, 2020; Tysiąc-Miśta et al., 2020).

Among the disinfection methods mentioned above, the chemical-free technology of UVGI

is most widely used. UV radiation is classified into UV-A (315–400 nm), UV-B (280–315

nm), and UV-C (100–280 nm) based on wavelength, among which UV-C shows optimum

germicidal effects, and disinfects microbes by damaging their nuclei acid (Reed, 2010). To

achieve a valid inactivation efficiency, the UV-C dose (the irradiance delivered to microbial

cells multiplied by the exposure time), intrinsic microbial characteristics, and target medium

(air or surface) characteristics should be considered (Raeiszadeh & Adeli, 2020). A recent

study showed that, similar to other human coronaviruses, SARS-CoV-2 can effectively be

inactivated by UV-C radiation as well (Heilingloh et al., 2020). On the other hand, excessive

exposure to UV-C can also damage human skin and eyes, as well as degrade some materials

(Raeiszadeh & Adeli, 2020). Therefore, in the UV-C dose, a balance between disinfection

efficiency and the potential harm to human health or surfaces is needed. Further research,

standardized protocols of optimum UV-C doses, and optimization of the system design in

different settings are required.

UVGI technology has been applied in many scenarios for different purposes, such as

in-duct and cooling coil systems, upper-room, UV barriers, lower-room, recirculating or air

39
cleaning units, area disinfection systems, disinfection chambers, and UV-C wands (T. Chen

& O’Keeffe, 2020). Thus far, upper-room UVGI may provide the most practical option

owing to its safe distance from occupants. As shown in Fig. 3, this system utilizes the entire

upper room as a disinfection chamber, as well as the rising warm air from the human body,

with the ceiling fan assuring good mixing, and the shielding sufficiently protecting occupants

below (Nardell, 2016). The upper-room UVGI system indicates the potential of combining

UVGI and mixing ventilation in the future.

Fig. 3 Schematic of an upper-room UVGI system (Nardell, 2016)

6 Discussion

6.1 Characteristics of expiratory particles and flow from different respiratory activities

Expiratory particle and flow features of different respiratory activities are summarized, to

provide quick references for further experiments and boundary conditions of numerical

analysis. Due to many influential factors (e.g., individual physiological differences, various

measurement technologies, particle evaporation and condensation, health status),

experimental measurements so far cannot determine explicit size range and number of
40
expiratory particles for each respiratory activity. However, the number of particles under 1

μm was found to occupy the largest proportion, indicating a significant threat of airborne

transmission. The maximum velocity of expiratory airflow can also be influenced by

individual physiological differences such as gender and height. It can affect the number of

aerosols suspending in the air and the maximum horizontal distance that a droplet can travel

(X. Xie et al., 2007). Among different respiratory activities, the data of sneezing

characteristics are scarce and inconsistent with each other, because of the difficulty of

creating a sneeze on command during experiments. In the future, to improve the database on

characteristics of expiratory particles and airflow from different respiratory activities, there is

still a lot of work to be done.

6.2 Environmental influence on evaporation and dispersion of expiratory particles

Environmental factors such as humidity, temperature and airflow can influence the

dispersion of expiratory particles by influencing their evaporation. RH has different

influences on the evaporation of a single droplet and droplets in the turbulent flow. For a

single droplet of different sizes, the evaporation time at a high RH was greater than that at a

low RH (Wei & Li, 2015). For droplets spread in turbulent buoyant flow, small (20–30 μm)

and large (100 μm) sized droplets were insensitive to RH; RH significantly affects

medium-sized (50–60 μm) droplets (L. Liu, Wei, et al., 2017; Wei & Li, 2015). Therefore,

further research should consider the turbulent buoyant flow when discussing the evaporation

features of expiratory droplets. The influence of temperature on droplet evaporation was

found to be affected by RH. More violent turbulence in the ambient flow field can facilitate

evaporation (Ji et al., 2018). In the future, more studies are needed on the influence of

ambient temperature and flow on droplet evaporation. Considering the current methodologies

used to study the impact of environmental factors, experimental studies should attempt to

41
suspend droplets in the air. More efforts should be made to refine theoretical models, such as

considering the composition of expiratory droplets.

The flow field in the ambient environment can also directly influence the dispersion of

expiratory particles. The lock-up phenomena in thermal stratification may extend the scope of

short-range airborne transmission (F. Liu et al., 2019; Zhou et al., 2017). Disturbance in the

horizontal airflow supply can strengthen the mixture of expiratory flow and ambient air, thus

reducing the exposure risks of close contact (Ai et al., 2019). Engineering control strategies,

such as ventilation and air cleaners can reduce long-range airborne transmission through

positive intervention with indoor airflow patterns.

6.3 Environmental influence on the viability of SARS-CoV-2 and seasonality of COVID-19

At room temperature and RH, SARS-CoV-2 can survive for up to 72 h on plastic (21–

23 °C, 40%) (van Doremalen et al., 2020), up to 16 h in aerosols (23 °C, 53 %) (Fears et al.,

2020). This indicates that airborne and contact transmission of SARS-CoV-2 is plausible, and

risks can last for long periods. For droplets deposited on surfaces, higher temperatures were

found to cause faster inactivation of SARS-CoV-2, but a discrepancy still exists in the

influence of RH on virus stability. For different-sized aerosols, the influence of humidity on

the infectivity of SARS-CoV-2 is also unclear. Therefore, further studies with rigorously

controlled variables regarding the influence of humidity on viral viability are required. In

addition, further improvements can be made to models that show the joint effect of

environmental factors, to improve the prediction of virus viability under varying conditions.

The seasonality of confirmed COVID-19 cases may be affected by the evaporation and

dispersion characteristics of expiratory particles, as well as the survival ability of

SARS-CoV-2, under different climate conditions. However, other factors such as human

42
behavior and circannual variation in immunity can also influence the seasonality of

COVID-19.

6.4 Exposure risks of expiratory particles and infection risks of SARS-CoV-2

The cut-off distance for ―short-range‖ and ―long-range‖ transmission routes of expiratory

particles depends on the maximum horizontal distance that the droplets can reach. This

threshold distance is influenced by many factors, including the initial properties of expiratory

particles and flow, as well as ambient environmental conditions. The maximum horizontal

distance can be up to 3‒6 m at a velocity of 20‒50 m/s (sneezing) (X. Xie et al., 2007). The

long-range exposure risk of expiratory particles generally results from airborne transmission.

Moreover, a recent study showed that airborne transmission also contributes the most to the

short-range exposure risk of expiratory particles (Wenzhao Chen et al., 2020). This shows

that airborne transmission route plays a critical role in both short-range and long-range

transmission of expiratory particles.

Most infection risk assessment methods for airborne transmission of SARS-CoV-2 have

been proposed and expanded based on the classic Wells-Riley model, in which the

determination of the quanta emission rate ( ) (quanta/h) value is significant. The value is

influenced by many factors, such as different respiratory activities and activity levels

(Buonanno et al., 2020). Moreover, considering the different methods adopted and different

scenarios considered in relevant studies, the value varies considerably in different studies.

So far, we can only obtain an upper limit value of approximately 1000 under common

speaking and singing scenarios using the back-calculated method. In the future, more studies

concerning the determination of q values under different scenarios are needed. Moreover,

further studies on the infection risk assessment methods for droplet and contact transmission

are also required.

43
6.5 Determination of safe physical distance and consideration of safe contact time

Many influential factors, such as initial properties of expiratory droplets and flow, viral

load, infectious dose, and environmental influence on expiratory particles and virus viability,

make it difficult to determine safe physical distancing. It is inadequate to estimate a safe

physical distance based only on the maximum reaching distance of expiratory droplets

because strong evidence has shown the potential infectiousness of aerosols. Meanwhile,

considering the short-range airborne transmission, a safe contact time is also significant (F.

Yang et al., 2020). In the future, to obtain a more reasonable safe physical distance and

contact time, the relationship between exposure risks and infection risks must be established,

by determining the infection capability of expiratory particles carrying SARS-CoV-2.

Therefore, further studies on dose-response models are needed.

6.6 Influence of masks on expiratory particles and flow, and associated wearing suggestions

All types of masks can reduce the front flow, but air leakage in other directions can occur

when wearing surgical and handmade masks. If air leakage is considered, the aerosol

filtration efficiency of surgical masks can be 44% when the wearer is at rest, or 10% when

the wearer is in moderate activity (Konda et al., 2020). Greater physical distance is required

for masks with larger fitting gap and lower filtering efficiency, and respirators are necessary

in the presence of a known infection source. To prevent airflow leakage, the use the elbow is

recommended during violent respiratory activities. In the future, further quantitative studies

on the expiratory flow field when wearing a mask are needed.

6.7 Air distribution methods and occupancy-aided ventilation

Ventilation, which usually refers to total volume air distribution, is the most used

engineering control strategies against airborne transmission of respiratory virus. Ventilation

rate is an important index to evaluate the effectiveness of contaminant removal through


44
ventilation systems. However, the appropriate ventilation rate depends on the development of

suitable infection risk assessment models. The efficacy of ventilation is also influenced by

the airflow pattern, especially in large and complex spaces. Mixing ventilation (MV) and

displacement ventilation (DV) are two traditional ventilation types creating different airflow

patterns. Most advanced air distribution methods are based on the mechanism of DV for

expansion, given the high effectiveness of DV in removing viral contaminants. In the future,

with the development of CFD technology, designing airflow patterns will become more

convenient.

Personalized ventilation (PV) and personalized exhaust (PE) are two air distribution

methods targeting at the microenvironment around human body. Generally, PV is used by the

susceptible, while PE is used by the infector to prevent cross-infection (Ai et al., 2019; J.

Yang et al., 2015a). The combination of PV and PE for the susceptible individual can further

enhance the penetration of PV airflow to the breathing zone (J. Yang et al., 2014). In the

future, a more systematic investigation of PV and PE effectiveness in different conditions, as

well as the study of their influence on controlling droplet transmission, is needed.

Occupancy-aided ventilation, which is a type of source control, aims to reduce the airborne

infection by managing the indoor occupancy (Melikov et al., 2020). The adopted occupancy

schedule should balance short occupancy duration and work productivity. In the future,

studies should address problems such as potential corridor blockage, simplified evaluation of

work productivity, assumption of homogeneous virus distribution, and difficulty in obtaining

infector number (S. Zhang et al., 2021).

45
6.8 Supplementary infection control strategies for total volume air distribution

When total volume air distribution methods cannot provide sufficient air change rate due to

heating or cooling requirements, potential air purifying strategies such as air filtration and

disinfection can be implemented for infection control.

Air filtration aims to remove expiratory particles from airstreams passing through the filter.

The HEPA filters are most widely used among common air filters, with the particle filtration

efficiency of up to 99.97% (Christopherson et al., 2020). Flow interaction between other

momentum sources and HEPA-filter portable air cleaners (PACs) can significantly influence

indoor air purifying efficacy (T. Zhang et al., 2010). In the future, further systematic studies

on the influence of the layouts and operation parameters of PACs are needed. Moreover, the

expiration characteristics under different respiratory activities and purifying efficacy under

different airflow patterns need to be discussed.

Disinfection aims to directly inactivate virus circulating in the air, among which the

chemical-free ultraviolet germicidal irradiation (UVGI) is now the most widely used. The

UV-C dose should balance disinfection efficiency and the potential harm to human health or

surfaces. The upper-room UVGI system could present a safe option (Nardell, 2016), which

also show a potential of combination with mixing ventilation. In the future, standardized

protocols of appropriate UV-C dose, as well as the system design for SARS-CoV-2

disinfection, are needed.

7 Conclusion

In this study, the characteristics of expiratory particles and airflow, environmental

influence on particle transmission and virus viability, infection risk assessment methods and

common control strategies are critically reviewed. The primary conclusions of this review are

as follows:

46
1) Expiratory particle and flow features of different respiratory activities obtained from

experiments are summarized. Discrepancies still exist in some measurement results, and data

on some respiratory activities are limited. Therefore, further research is required in terms of

measuring methodologies.

2) Ambient environmental factors influence the transmission of expiratory particles by

affecting their evaporation and dispersion. The RH significantly influences the evaporation of

medium-sized (50–60 μm) droplets. A lower RH contributes to faster evaporation. However,

there is limited research on the influence of ambient temperature and flow on expiratory

particle evaporation. Higher temperatures accelerate the inactivation of SARS-CoV-2 both on

surfaces and in aerosols. However, the influence of RH on virus viability requires further

research.

3) Airborne transmission is found to be a significant route in both short- and long-range

virus transmission. The Wells-Riley model and its expansions are widely used to predict the

infection risks of airborne SARS-CoV-2 transmission. Further efforts to calculate a more

accurate quanta emission rate ( ), as well as to assess the suitability of applying the expanded

Wells-Riley model in other transmission routes, are needed. For a safe physical distance, the

maximum reaching distance of expiratory droplets can determine the lower limit, but the

infection potential of aerosols should be considered in future studies.

4) Based on scientific analysis, suggestions for mask usage during the pandemic are offered.

Further studies on expiratory particle and airflow features when wearing a mask are needed.

As for engineering control strategies, traditional and advanced air distribution methods, from

the total volume scale to microenvironment scale, are introduced and compared. Moreover, as

promising complementary of ventilation to control the spread of COVID-19, air filtration and

disinfection methods are summarized and discussed.

47
Nomenclature
CADR Clean air delivery rates (m3/h) PPE Personal protection equipment
CFD Computational fluid dynamics PV Personalized ventilation
CJV Confluent jet ventilation Quanta emission rate (quanta/h)
COVID-1 Coronavirus disease 2019 RH Relative humidity
9
DV Displacement ventilation RNA Ribonucleic acid
DWV Downward ventilation SARS-CoV- Severe acute respiratory syndrome
2 coronavirus 2
FFR Filtering facepiece respirator SV Stratum ventilation
HEPA High-efficiency particulate air TBL Thermal boundary layer
IJV Impinging jet ventilation TVAD Total volume air distribution
MPPS Most penetrating particle size Maximum velocity of expiratory
flow (m/s)
MV Mixing ventilation UFV Underfloor ventilation
NaCl Sodium chloride UV Ultraviolet radiation
PAC Portable air cleaners UVGI Ultraviolet germicidal irradiation
PE Personalized exhaust WAV Wall attached ventilation
PIV Particle image velocimetry WHO World Health Organization
POV Protected occupied zone
ventilation

Declaration of Competing Interest

The authors report no declarations of interest.

Acknowledgments

This research did not receive any specific grant from funding agencies in the public,

commercial, or not-for-profit sectors.

48
Reference

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(March), 106328. https://doi.org/10.1016/j.buildenv.2019.106328

Ai, Z. T., & Melikov, A. K. (2018). Airborne spread of expiratory droplet nuclei between the

occupants of indoor environments: A review. Indoor Air, 28(4), 500–524.

https://doi.org/10.1111/ina.12465

Andersson, H., Cehlin, M., & Moshfegh, B. (2018). Experimental and numerical

investigations of a new ventilation supply device based on confluent jets. Building and

Environment, 137(March), 18–33. https://doi.org/10.1016/j.buildenv.2018.03.038

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, 54(6), 635–638. https://doi.org/10.1080/02786826.2020.1749229

Asadi, S., Wexler, A. S., Cappa, C. D., Barreda, S., Bouvier, N. M., & Ristenpart, W. D.

(2019). Aerosol emission and superemission during human speech increase with voice

loudness. Scientific Reports, 9(1), 1–10. https://doi.org/10.1038/s41598-019-38808-z

Assaad, D. Al, Habchi, C., Ghali, K., & Ghaddar, N. (2018). Effectiveness of intermittent

personalized ventilation in protecting occupant from indoor particles. Building and

Environment, 128(November 2017), 22–32.

https://doi.org/10.1016/j.buildenv.2017.11.027

Barrett, K. E., & Ganong, W. F. (2010). Ganong’s review of medical physiology.

Bhagat, R. K., & Linden, P. F. (2020). Displacement ventilation: a viable ventilation strategy

for makeshift hospitals and public buildings to contain COVID-19 and other airborne

diseases. Royal Society Open Science, 7(9), 200680. https://doi.org/10.1098/rsos.200680

49
Biryukov, J., Boydston, J. A., Dunning, R. A., Yeager, J. J., Wood, S., Reese, A. L., Ferris,

A., Miller, D., Weaver, W., Zeitouni, N. E., Phillips, A., Freeburger, D., Hooper, I.,

Ratnesar-Shumate, S., Yolitz, J., Krause, M., Williams, G., Dawson, D. G., Herzog, A.,

… Altamura, L. A. (2020). Increasing Temperature and Relative Humidity Accelerates

Inactivation of SARS-CoV-2 on Surfaces. MSphere, 5(4), 1–9.

https://doi.org/10.1128/msphere.00441-20

Bolashikov, Z. D., & Melikov, A. K. (2009). Methods for air cleaning and protection of

building occupants from airborne pathogens. Building and Environment, 44(7), 1378–

1385. https://doi.org/10.1016/j.buildenv.2008.09.001

Bourouiba, L. (2016). A sneeze. New England Journal of Medicine, 375(8), e15.

https://doi.org/10.1056/NEJMicm1501197

Bourouiba, L., Dehandschoewercker, E., & Bush, J. W. M. (2014). Violent expiratory events:

On coughing and sneezing. Journal of Fluid Mechanics, 745, 537–563.

https://doi.org/10.1017/jfm.2014.88

Brankston, G., Gitterman, L., Hirji, Z., Lemieux, C., & Gardam, M. (2007). Transmission of

influenza A in human beings. Lancet Infectious Diseases, 7(4), 257–265.

https://doi.org/10.1016/S1473-3099(07)70029-4

Buonanno, G., Stabile, L., & Morawska, L. (2020). Estimation of airborne viral emission:

Quanta emission rate of SARS-CoV-2 for infection risk assessment. Environment

International, 141(May), 105794. https://doi.org/10.1016/j.envint.2020.105794

Cao, G., Awbi, H., Yao, R., Fan, Y., Sirén, K., Kosonen, R., & Zhang, J. (Jensen). (2014). A

review of the performance of different ventilation and airflow distribution systems in

buildings. Building and Environment, 73, 171–186.

https://doi.org/10.1016/j.buildenv.2013.12.009

50
Cao, G., Sirén, K., & Kilpeläinen, S. (2014). Modelling and experimental study of

performance of the protected occupied zone ventilation. Energy and Buildings,

68(PARTA), 515–531. https://doi.org/10.1016/j.enbuild.2013.10.008

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 immediately at the mouth opening.

Journal of Aerosol Science, 40(2), 122–133.

https://doi.org/10.1016/j.jaerosci.2008.10.003

Chaudhuri, S., Basu, S., Kabi, P., Unni, V. R., & Saha, A. (2020). Modeling the role of

respiratory droplets in Covid-19 type pandemics. Physics of Fluids, 32(6), 063309.

https://doi.org/10.1063/5.0015984

Chen, Lea Der. (2020). Effects of ambient temperature and humidity on droplet lifetime – A

perspective of exhalation sneeze droplets with COVID-19 virus transmission.

International Journal of Hygiene and Environmental Health, 229(April), 113568.

https://doi.org/10.1016/j.ijheh.2020.113568

Chen, Lin, Jin, X., Yang, L., Du, X., & Yang, Y. (2017). Particle transport characteristics in

indoor environment with an air cleaner: The effect of nonuniform particle distributions.

Building Simulation, 10(1), 123–133. https://doi.org/10.1007/s12273-016-0310-7

Chen, T., & O’Keeffe, J. (2020). COVID-19 in indoor environments — Air and surface

disinfection measures. National Collaborating Centre for Environmental Health., 1–25.

https://ncceh.ca/documents/guide/covid-19-indoor-environments-air-and-surface-disinfe

ction-measures

Chen, Wenhao, Zhang, J. S., & Zhang, Z. (2005). Performance of air cleaners for removing

multiple volatile organic compounds in indoor air. ASHRAE Transactions, 111 PART

1(January), 1101–1114.

51
Chen, Wenzhao, Zhang, N., Wei, J., Yen, H., & Li, Y. (2020). Short-range airborne route

dominates exposure of respiratory infection during close contact. Building and

Environment, 176(April), 106859. https://doi.org/10.1016/j.buildenv.2020.106859

Chin, A. W. H., Chu, J. T. S., Perera, M. R. A., Hui, K. P. Y., Yen, H.-L., Chan, M. C. W.,

Peiris, J. S. M., & Poon, L. L. M. (2020). Stability of SARS-CoV-2 in different

environmental conditions. The Lancet Microbe, 1(1), e10.

https://doi.org/10.1016/s2666-5247(20)30003-3

Cho, Y., Awbi, H. B., & Karimipanah, T. (2008). Theoretical and experimental investigation

of wall confluent jets ventilation and comparison with wall displacement ventilation.

Building and Environment, 43(6), 1091–1100.

https://doi.org/10.1016/j.buildenv.2007.02.006

Christopherson, D. A., Yao, W. C., Lu, M., Vijayakumar, R., & Sedaghat, A. R. (2020).

High-Efficiency Particulate Air Filters in the Era of COVID-19: Function and Efficacy.

Otolaryngology - Head and Neck Surgery (United States), 163(6), 1153–1155.

https://doi.org/10.1177/0194599820941838

Dabisch, P., Schuit, M., Herzog, A., Beck, K., Wood, S., Krause, M., Miller, D., Weaver, W.,

Freeburger, D., Hooper, I., Green, B., Williams, G., Holland, B., Bohannon, J., Wahl, V.,

Yolitz, J., Hevey, M., & Ratnesar-Shumate, S. (2021). The influence of temperature,

humidity, and simulated sunlight on the infectivity of SARS-CoV-2 in aerosols. Aerosol

Science and Technology, 55(2), 142–153.

https://doi.org/10.1080/02786826.2020.1829536

Dai, H., & Zhao, B. (2020). Association of the infection probability of COVID-19 with

ventilation rates in confined spaces. Building Simulation.

https://doi.org/10.1007/s12273-020-0703-5

52
Dbouk, T., & Drikakis, D. (2020a). On coughing and airborne droplet transmission to

humans. Physics of Fluids, 32(5). https://doi.org/10.1063/5.0011960

Dbouk, T., & Drikakis, D. (2020b). On respiratory droplets and face masks. Physics of Fluids,

32(6). https://doi.org/10.1063/5.0015044

Fabian, P., McDevitt, J. J., DeHaan, W. H., Fung, R. O. P., Cowling, B. J., Chan, K. H.,

Leung, G. M., & Milton, D. K. (2008). Influenza virus in human exhaled breath: An

observational study. PLoS ONE, 3(7), 5–10.

https://doi.org/10.1371/journal.pone.0002691

Fears, A. C., Klimstra, W. B., Duprex, P., Hartman, A., Weaver, S. C., Plante, K. S.,

Mirchandani, D., Plante, J. A., Aguilar, P. V., Fernández, D., Nalca, A., Totura, A.,

Dyer, D., Kearney, B., Lackemeyer, M., Bohannon, J. K., Johnson, R., Garry, R. F.,

Reed, D. S., & Roy, C. J. (2020). Persistence of Severe Acute Respiratory Syndrome

Coronavirus 2 in Aerosol Suspensions. Emerging Infectious Diseases, 26(9).

https://doi.org/10.3201/eid2609.201806

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(April), 105585.

https://doi.org/10.1016/j.jaerosci.2020.105585

Gao, C. X., Li, Y., Wei, J., Cotton, S., Hamilton, M., Wang, L., & Cowling, B. J. (2021).

Multi-route respiratory infection: When a transmission route may dominate. Science of

the Total Environment, 752, 141856. https://doi.org/10.1016/j.scitotenv.2020.141856

Ge, X. Y., Pu, Y., Liao, C. H., Huang, W. F., Zeng, Q., Zhou, H., Yi, B., Wang, A. M., Dou,

Q. Y., Zhou, P. C., Chen, H. L., Liu, H. X., Xu, D. M., Chen, X., & Huang, X. (2020).

Evaluation of the exposure risk of SARS-CoV-2 in different hospital environment.

53
Sustainable Cities and Society, 61(May), 102413.

https://doi.org/10.1016/j.scs.2020.102413

Gralton, J., Tovey, E., McLaws, M. L., & Rawlinson, W. D. (2011). The role of particle size

in aerosolised pathogen transmission: A review. Journal of Infection, 62(1), 1–13.

https://doi.org/10.1016/j.jinf.2010.11.010

Gupta, J. K., Lin, C. H., & Chen, Q. (2009). Flow dynamics and characterization of a cough.

Indoor Air, 19(6), 517–525. https://doi.org/10.1111/j.1600-0668.2009.00619.x

Haghshenaskashani, S., & Sajadi, B. (2018). Evaluation of thermal comfort, IAQ and energy

consumption in an impinging jet ventilation (IJV) system with/without ceiling exhaust.

Journal of Building Engineering, 18(November 2017), 142–153.

https://doi.org/10.1016/j.jobe.2018.03.011

Han, M., Ooka, R., Kikumoto, H., Oh, W., Bu, Y., & Hu, S. (2021). Measurements of

Exhaled Airflow Velocity through Human Coughs using Particle Image Velocimetry.

Building and Environment, (Submitted).

Heilingloh, C. S., Aufderhorst, U. W., Schipper, L., Dittmer, U., Witzke, O., Yang, D., Zheng,

X., Sutter, K., Trilling, M., Alt, M., Steinmann, E., & Krawczyk, A. (2020).

Susceptibility of SARS-CoV-2 to UV irradiation. American Journal of Infection Control,

48(10), 1273–1275. https://doi.org/10.1016/j.ajic.2020.07.031

Hindson, J. (2020). COVID-19: faecal–oral transmission? Nature Reviews Gastroenterology

& Hepatology, 17(5), 259–259. https://doi.org/10.1038/s41575-020-0295-7

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(October 2017), 164–172.

https://doi.org/10.1016/j.jaerosci.2017.10.009

54
Jin, X., Yang, L., Du, X., & Yang, Y. (2016). Particle transport characteristics in indoor

environment with an air cleaner. Indoor and Built Environment, 25(6), 987–996.

https://doi.org/10.1177/1420326X15592253

Kähler, C. J., & Hain, R. (2020). Fundamental protective mechanisms of face masks against

droplet infections. Journal of Aerosol Science, 148(May).

https://doi.org/10.1016/j.jaerosci.2020.105617

Karimipanah, T., & Awbi, H. B. (2002). Theoretical and experimental investigation of

impinging jet ventilation and comparison with wall displacement ventilation. Building

and Environment, 37(12), 1329–1342. https://doi.org/10.1016/S0360-1323(01)00117-2

Konda, A., Prakash, A., Moss, G. A., Schmoldt, M., Grant, G. D., & Guha, S. (2020).

Aerosol Filtration Efficiency of Common Fabrics Used in Respiratory Cloth Masks.

ACS Nano, 14(5), 6339–6347. https://doi.org/10.1021/acsnano.0c03252

Kriegel, M., Buchholz, U., Gastmeier, P., Bischoff, P., Abdelgawad, I., Hartmann, A., &

Kriegel, I. M. (2020). Predicted Infection Risk for Aerosol Transmission of Sars-Cov-2.

MedRxiv, 2020.10.08.20209106. https://doi.org/10.1101/2020.10.08.20209106

Kwon, S.-B., Park, J., Jang, J., Cho, Y., Park, D.-S., Kim, C., Bae, G.-N., & Jang, A. (2012).

Study on the initial velocity distribution of exhaled air from coughing and speaking.

Chemosphere, 87(11), 1260–1264. https://doi.org/10.1016/j.chemosphere.2012.01.032

Lednicky, J. A., Lauzard, M., Fan, Z. H., Jutla, A., Tilly, T. B., Gangwar, M., Usmani, M.,

Shankar, S. N., Mohamed, K., Eiguren-Fernandez, A., Stephenson, C. J., Alam, M. M.,

Elbadry, M. A., Loeb, J. C., Subramaniam, K., Waltzek, T. B., Cherabuddi, K., Morris, J.

G., & Wu, C. Y. (2020). Viable SARS-CoV-2 in the air of a hospital room with

COVID-19 patients. International Journal of Infectious Diseases, 100, 476–482.

https://doi.org/10.1016/j.ijid.2020.09.025

55
Leung, N. H. L., Chu, D. K. W., Shiu, E. Y. C., Chan, K.-H., McDevitt, J. J., Hau, B. J. P.,

Yen, H.-L., Li, Y., Ip, D. K. M., Peiris, J. S. M., Seto, W.-H., Leung, G. M., Milton, D.

K., & Cowling, B. J. (2020). Respiratory virus shedding in exhaled breath and efficacy

of face masks. Nature Medicine. https://doi.org/10.1038/s41591-020-0843-2

Li, A., Hou, Y., & Yang, J. (2019). Attached ventilation based on a curved surface wall.

Building Simulation, 12(3), 505–515. https://doi.org/10.1007/s12273-019-0505-9

Li, Y. Y., Wang, J. X., & Chen, X. (2020). Can a toilet promote virus transmission? From a

fluid dynamics perspective. Physics of Fluids, 32(6). https://doi.org/10.1063/5.0013318

Lindsley, W. G., Pearce, T. A., Hudnall, J. B., Davis, K. A., Davis, S. M., Fisher, M. A.,

Khakoo, R., Palmer, J. E., Clark, K. E., Celik, I., Coffey, C. C., Blachere, F. M., &

Beezhold, D. H. (2012). Quantity and Size Distribution of Cough-Generated Aerosol

Particles Produced by Influenza Patients During and After Illness. Journal of

Occupational and Environmental Hygiene, 9(7), 443–449.

https://doi.org/10.1080/15459624.2012.684582

Liu, F., Zhang, C., Qian, H., Zheng, X., & Nielsen, P. V. (2019). Direct or indirect exposure

of exhaled contaminants in stratified environments using an integral model of an

expiratory jet. Indoor Air, 29(4), 591–603. https://doi.org/10.1111/ina.12563

Liu, G., Xiao, M., Zhang, X., Gal, C., Chen, X., Liu, L., Pan, S., Wu, J., Tang, L., &

Clements-Croome, D. (2017). A review of air filtration technologies for sustainable and

healthy building ventilation. Sustainable Cities and Society, 32(April), 375–396.

https://doi.org/10.1016/j.scs.2017.04.011

Liu, L., Li, Y., Nielsen, P. V., Wei, J., & Jensen, R. L. (2017). Short-range airborne

transmission of expiratory droplets between two people. Indoor Air, 27(2), 452–462.

https://doi.org/10.1111/ina.12314

56
Liu, L., Wei, J., Li, Y., & Ooi, A. (2017). Evaporation and dispersion of respiratory droplets

from coughing. Indoor Air, 27(1), 179–190. https://doi.org/10.1111/ina.12297

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. fai, Kan, H., Fu, Q., & Lan, K. (2020).

Aerodynamic analysis of SARS-CoV-2 in two Wuhan hospitals. Nature, 582(7813),

557–560. https://doi.org/10.1038/s41586-020-2271-3

Lu, Y., Oladokun, M., & Lin, Z. (2020). Reducing the exposure risk in hospital wards by

applying stratum ventilation system. Building and Environment, 183(August), 107204.

https://doi.org/10.1016/j.buildenv.2020.107204

Mahjoub Mohammed Merghani, K., Sagot, B., Gehin, E., Da, G., & Motzkus, C. (2021). A

review on the applied techniques of exhaled airflow and droplets characterization.

Indoor Air, 31(1), 7–25. https://doi.org/10.1111/ina.12770

Mao, N., An, C. K., Guo, L. Y., Wang, M., Guo, L., Guo, S. R., & Long, E. S. (2020).

Transmission risk of infectious droplets in physical spreading process at different times:

A review. Building and Environment, 185(June), 107307.

https://doi.org/10.1016/j.buildenv.2020.107307

Mao, Y., Jones, R. M., Tan, Q., Ji, J. S., Li, N., Shen, J., Lv, Y., Pan, L., Ding, P., Wang, X.,

Wang, Y., & Macintyre, C. R. (2020). Aerosol transmission of SARS-CoV-2? Evidence,

prevention and control Song. Environment International, 144(January), 1–10.

Matson, M. J., Yinda, C. K., Seifert, S. N., Bushmaker, T., Fischer, R. J., van Doremalen, N.,

Lloyd-Smith, J. O., & Munster, V. J. (2020). Effect of Environmental Conditions on

SARS-CoV-2 Stability in Human Nasal Mucus and Sputum. Emerging Infectious

Diseases, 26(9), 2276–2278. https://doi.org/10.3201/eid2609.202267

57
Megahed, N. A., & Ghoneim, E. M. (2020). Antivirus-built environment: Lessons learned

from Covid-19 pandemic. Sustainable Cities and Society, 61(June), 102350.

https://doi.org/10.1016/j.scs.2020.102350

Melikov, A. K. (2016). Advanced air distribution: Improving health and comfort while

reducing energy use. Indoor Air, 26(1), 112–124. https://doi.org/10.1111/ina.12206

Melikov, A. K., Ai, Z. T., & Markov, D. G. (2020). Intermittent occupancy combined with

ventilation: An efficient strategy for the reduction of airborne transmission indoors.

Science of the Total Environment, 744(2).

https://doi.org/10.1016/j.scitotenv.2020.140908

Miller, S. L., Nazaroff, W. W., Jimenez, J. L., Boerstra, A., Buonanno, G., Dancer, S. J.,

Kurnitski, J., Marr, L. C., Morawska, L., & Noakes, C. (2021). Transmission of SARS‐

CoV‐ 2 by inhalation of respiratory aerosol in the Skagit Valley Chorale superspreading

event. Indoor Air, 31(2), 314–323. https://doi.org/10.1111/ina.12751

Mittal, R., Meneveau, C., & Wu, W. (2020). A mathematical framework for estimating risk

of airborne transmission of COVID-19 with application to face mask use and social

distancing. Physics of Fluids, 32(10). https://doi.org/10.1063/5.0025476

Morris, D. H., Yinda, K. C., Gamble, A., Rossine, F. W., Huang, Q., Bushmaker, T., Fischer,

R. J., Matson, M. J., van Doremalen, N., Vikesland, P. J., Marr, L. C., Munster, V. J., &

Lloyd-Smith, J. O. (2020). The effect of temperature and humidity on the stability of

SARS-CoV-2 and other enveloped viruses. BioRxiv : The Preprint Server for Biology.

https://doi.org/10.1101/2020.10.16.341883

Mousavi, E. S., Godri Pollitt, K. J., Sherman, J., & Martinello, R. A. (2020). Performance

analysis of portable HEPA filters and temporary plastic anterooms on the spread of

surrogate coronavirus. Building and Environment, 183(August), 107186.

https://doi.org/10.1016/j.buildenv.2020.107186

58
Nardell, E. A. (2016). Indoor environmental control of tuberculosis and other airborne

infections. Indoor Air, 26(1), 79–87. https://doi.org/10.1111/ina.12232

Nielsen, P. V., Li, Y., Buus, M., & Winther, F. V. (2010). Risk of cross-infection in a

hospital ward with downward ventilation. Building and Environment, 45(9), 2008–2014.

https://doi.org/10.1016/j.buildenv.2010.02.017

Nishimura, H., Sakata, S., & Kaga, A. (2013). A new methodology for studying dynamics of

aerosol particles in sneeze and cough using a digital high-vision, high-speed video

system and vector analyses. PLoS ONE, 8(11).

https://doi.org/10.1371/journal.pone.0080244

O’Hara, D. (1956). Airborne Contagion and Air Hygiene: An ecological study of droplet

infections. New England Journal of Medicine, 254(8), 396–397.

https://doi.org/10.1056/nejm195602232540823

Ong, S. W. X., Tan, Y. K., Chia, P. Y., Lee, T. H., Ng, O. T., Wong, M. S. Y., & Marimuthu,

K. (2020). Air, Surface Environmental, and Personal Protective Equipment

Contamination by Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2)

from a Symptomatic Patient. JAMA - Journal of the American Medical Association,

323(16), 1610–1612. https://doi.org/10.1001/jama.2020.3227

Pei, J., Dai, W., Li, H., & Liu, J. (2020). Laboratory and field investigation of portable air

cleaners’ long-term performance for particle removal to be published in: Building and

environment. Building and Environment, 181(July), 107100.

https://doi.org/10.1016/j.buildenv.2020.107100

Qian, H., Li, Y., Nielsen, P. V., & Hyldgaard, C. E. (2008). Dispersion of exhalation

pollutants in a two-bed hospital ward with a downward ventilation system. Building and

Environment, 43(3), 344–354. https://doi.org/10.1016/j.buildenv.2006.03.025

59
Qian, H., Li, Y., Sun, H., Nielsen, P. V., Huang, X., & Zheng, X. (2010). Particle removal

efficiency of the portable HEPA air cleaner in a simulated hospital ward. Building

Simulation, 3(3), 215–224. https://doi.org/10.1007/s12273-010-0005-4

Qian, H., & Zheng, X. (2018). Ventilation control for airborne transmission of human

exhaled bio-aerosols in buildings. Journal of Thoracic Disease, 10(Suppl 19), S2295–

S2304. https://doi.org/10.21037/jtd.2018.01.24

Raeiszadeh, M., & Adeli, B. (2020). A Critical Review on Ultraviolet Disinfection Systems

against COVID-19 Outbreak: Applicability, Validation, and Safety Considerations. ACS

Photonics, 7(11), 2941–2951. https://doi.org/10.1021/acsphotonics.0c01245

Rahmani, A. M., & Mirmahaleh, S. Y. H. (2021). Coronavirus disease (COVID-19)

prevention and treatment methods and effective parameters: A systematic literature

review. Sustainable Cities and Society, 64(October 2020), 102568.

https://doi.org/10.1016/j.scs.2020.102568

Reed, N. G. (2010). The history of ultraviolet germicidal irradiation for air disinfection.

Public Health Reports, 125(1), 15–27. https://doi.org/10.1177/003335491012500105

Riddell, S., Goldie, S., Hill, A., Eagles, D., & Drew, T. W. (2020). The effect of temperature

on persistence of SARS-CoV-2 on common surfaces. Virology Journal, 17(1), 1–7.

https://doi.org/10.1186/s12985-020-01418-7

Riley, E. C., Murphy, G., & Riley, R. L. (1978). Airborne Spread Of Measles in a Suburban

Elementary School. American Journal of Epidemiology, 107(5), 421–432.

https://doi.org/10.1093/oxfordjournals.aje.a112560

Rosti, M. E., Olivieri, S., Cavaiola, M., Seminara, A., & Mazzino, A. (2020). Fluid dynamics

of COVID-19 airborne infection suggests urgent data for a scientific design of social

distancing. Scientific Reports, 10(1), 1–9. https://doi.org/10.1038/s41598-020-80078-7

60
Scharfman, B. E., Techet, A. H., Bush, J. W. M., & Bourouiba, L. (2016). Visualization of

sneeze ejecta: steps of fluid fragmentation leading to respiratory droplets. Experiments

in Fluids, 57(2), 1–9. https://doi.org/10.1007/s00348-015-2078-4

Smither, S. J., Eastaugh, L. S., Findlay, J. S., & Lever, M. S. (2020). Experimental aerosol

survival of SARS-CoV-2 in artificial saliva and tissue culture media at medium and high

humidity. Emerging Microbes and Infections, 9(1), 1415–1417.

https://doi.org/10.1080/22221751.2020.1777906

Sun, C., & Zhai, Z. (2020). The efficacy of social distance and ventilation effectiveness in

preventing COVID-19 transmission. Sustainable Cities and Society, 62(July), 102390.

https://doi.org/10.1016/j.scs.2020.102390

Sze To, G. N., & Chao, C. Y. H. (2010). Review and comparison between the Wells-Riley

and dose-response approaches to risk assessment of infectious respiratory diseases.

Indoor Air, 20(1), 2–16. https://doi.org/10.1111/j.1600-0668.2009.00621.x

Tang, J. W., Bahnfleth, W. P., Bluyssen, P. M., Buonanno, G., Jimenez, J. L., Kurnitski, J.,

Li, Y., Miller, S., Sekhar, C., Morawska, L., Marr, L. C., Melikov, A. K., Nazaroff, W.

W., Nielsen, P. V., Tellier, R., Wargocki, P., & Dancer, S. J. (2021). Dismantling myths

on the airborne transmission of severe acute respiratory syndrome coronavirus

(SARS-CoV-2). Journal of Hospital Infection.

https://doi.org/10.1016/j.jhin.2020.12.022

Tang, J. W., Liebner, T. J., Craven, B. A., & Settles, G. S. (2009). A schlieren optical study

of the human cough with and without wearing masks for aerosol infection control.

Journal of the Royal Society Interface, 6(SUPPL. 6), 727–736.

https://doi.org/10.1098/rsif.2009.0295.focus

Tang, J. W., Nicolle, A. D., Klettner, C. A., Pantelic, J., Wang, L., Suhaimi, A. Bin, Tan, A.

Y. L., Ong, G. W. X., Su, R., Sekhar, C., Cheong, D. D. W., & Tham, K. W. (2013).

61
Airflow Dynamics of Human Jets: Sneezing and Breathing - Potential Sources of

Infectious Aerosols. PLoS ONE, 8(4), 1–7.

https://doi.org/10.1371/journal.pone.0059970

Tang, J. W., Nicolle, A., Pantelic, J., Koh, G. C., de Wang, L., Amin, M., Klettner, C. A.,

Cheong, D. K. W., Sekhar, C., & Tham, K. W. (2012). Airflow dynamics of coughing in

healthy human volunteers by shadowgraph imaging: An aid to aerosol infection control.

PLoS ONE, 7(4). https://doi.org/10.1371/journal.pone.0034818

Tian, L., Lin, Z., & Wang, Q. (2010). Comparison of gaseous contaminant diffusion under

stratum ventilation and under displacement ventilation. Building and Environment, 45(9),

2035–2046. https://doi.org/10.1016/j.buildenv.2010.01.002

Tysiąc-Miśta, M., Dubiel, A., Brzoza, K., Burek, M., & Pałkiewicz, K. (2020). Air

disinfection procedures in the dental office during the COVID-19 pandemic. Medycyna

Pracy, 72(1), 39–48. https://doi.org/10.13075/mp.5893.01005

V, A. A. R., R, V., & Haghighat, F. (2020). The contribution of dry indoor built environment

on the spread of Coronavirus: Data from various Indian states. Sustainable Cities and

Society, 62(May), 102371. https://doi.org/10.1016/j.scs.2020.102371

van Doremalen, N., Bushmaker, T., Morris, D. H., Holbrook, M. G., Gamble, A., Williamson,

B. N., Tamin, A., Harcourt, J. L., Thornburg, N. J., Gerber, S. I., Lloyd-Smith, J. O., de

Wit, E., & Munster, V. J. (2020). Aerosol and Surface Stability of SARS-CoV-2 as

Compared with SARS-CoV-1. New England Journal of Medicine, 382(16), 1564–1567.

https://doi.org/10.1056/NEJMc2004973

VanSciver, M., Miller, S., & Hertzberg, J. (2011). Particle Image Velocimetry of Human

Cough. Aerosol Science and Technology, 45(3), 415–422.

https://doi.org/10.1080/02786826.2010.542785

62
Viola, I. M., Peterson, B., Pisetta, G., Pavar, G., Akhtar, H., Menoloascina, F., Mangano, E.,

Dunn, K. E., Gabl, R., Nila, A., Molinari, E., Cummins, C., Thompson, G., McDougall,

C. M., Lo, T. Y. M., Denison, F. C., Digard, P., Malik, O., Dunn, M. J. G., & Mehendale,

F. V. (2020). Face coverings, aerosol dispersion and mitigation of virus transmission

risk. ArXiv. https://doi.org/10.1109/OJEMB.2021.3053215

Wang, W., Xu, Y., Gao, R., Lu, R., Han, K., Wu, G., & Tan, W. (2020). Detection of

SARS-CoV-2 in Different Types of Clinical Specimens. JAMA, 323(11), 1061–1069.

https://doi.org/10.1001/jama.2020.3786

Wei, J., & Li, Y. (2015). Enhanced spread of expiratory droplets by turbulence in a cough jet.

Building and Environment, 93(P2), 86–96.

https://doi.org/10.1016/j.buildenv.2015.06.018

Wei, J., & Li, Y. (2016). Airborne spread of infectious agents in the indoor environment.

American Journal of Infection Control, 44(9), S102–S108.

https://doi.org/10.1016/j.ajic.2016.06.003

Wells, W. F. (1934). On air-borne infection: Study II. Droplets and droplet nuclei. American

Journal of Epidemiology, 20(3), 611–618.

https://doi.org/10.1093/oxfordjournals.aje.a118097

Wells, W. F. (1955). Airborne contagion and air hygiene: An ecological study of droplet

infections. In Cambridge, MA: Harvard University Press. Cambridge, MA: Harvard

University Press. https://doi.org/10.1093/ajcp/25.11.1301

Wilson, N. M., Norton, A., Young, F. P., & Collins, D. W. (2020). Airborne transmission of

severe acute respiratory syndrome coronavirus-2 to healthcare workers: a narrative

review. Anaesthesia, 75(8), 1086–1095. https://doi.org/10.1111/anae.15093

World Health Organization. (2014). Infection prevention and control of epidemic- and

pandemic-prone acute respiratory infections in health care.

63
https://apps.who.int/iris/bitstream/handle/10665/112656/9789241507134_eng.pdf;jsessi

onid=41AA684FB64571CE8D8A453C4F2B2096?sequence=1

World Health Organization. (2020a). Mask use in the context of COVID-19 (Issue December).

https://www.who.int/publications/i/item/advice-on-the-use-of-masks-in-the-community-

during-home-care-and-in-healthcare-settings-in-the-context-of-the-novel-coronavirus-(2

019-ncov)-outbreak

World Health Organization. (2020b). Transmission of SARS-CoV-2: implications for

infection prevention precautions.

https://www.who.int/news-room/commentaries/detail/transmission-of-sars-cov-2-implic

ations-for-infection-prevention-precautions

World Health Organization. (2021). COVID-19 Weekly Epidemiological Update, 21

Feburary 2021.

https://www.who.int/docs/default-source/coronaviruse/situation-reports/weekly_epidemi

ological_update_22.pdf

Xie, X., Li, Y., Chwang, A. T. Y., Ho, P. L., & Seto, W. H. (2007). How far droplets can

move in indoor environments ? revisiting the Wells evaporation falling curve. Indoor Air,

17(3), 211–225. https://doi.org/10.1111/j.1600-0668.2007.00469.x

Xie, Xiaojian., Li, Y., Sun, H., & Liu, L. (2009). Exhaled droplets due to talking and

coughing. Journal of The Royal Society Interface, 6(suppl_6).

https://doi.org/10.1098/rsif.2009.0388.focus

Xu, C., Luo, X., Yu, C., & Cao, S. J. (2020). The 2019-nCoV epidemic control strategies and

future challenges of building healthy smart cities. Indoor and Built Environment, 29(5),

639–644. https://doi.org/10.1177/1420326X20910408

Xu, C., Nielsen, P. V., Liu, L., Jensen, R. L., & Gong, G. (2018). Impacts of airflow

interactions with thermal boundary layer on performance of personalized ventilation.

64
Building and Environment, 135(March), 31–41.

https://doi.org/10.1016/j.buildenv.2018.02.048

Xu, C., Wei, X., Liu, L., Su, L., Liu, W., Wang, Y., & Nielsen, P. V. (2020). Effects of

personalized ventilation interventions on airborne infection risk and transmission

between occupants. Building and Environment, 180(April), 107008.

https://doi.org/10.1016/j.buildenv.2020.107008

Xu, J., Fu, S., & Chao, C. Y. H. (2020). Performance of airflow distance from personalized

ventilation on personal exposure to airborne droplets from different orientations. Indoor

and Built Environment, 0(0). https://doi.org/10.1177/1420326X20951245

Yang, B., Melikov, A. K., Kabanshi, A., Zhang, C., Bauman, F. S., Cao, G., Awbi, H., Wigö,

H., Niu, J., Cheong, K. W. D., Tham, K. W., Sandberg, M., Nielsen, P. V., Kosonen, R.,

Yao, R., Kato, S., Sekhar, S. C., Schiavon, S., Karimipanah, T., … Lin, Z. (2019). A

review of advanced air distribution methods - theory, practice, limitations and solutions.

Energy and Buildings, 202. https://doi.org/10.1016/j.enbuild.2019.109359

Yang, F., Pahlavan, A. A., Mendez, S., Abkarian, M., & Stone, H. A. (2020). Towards

improved social distancing guidelines: Space and time dependence of virus transmission

from speech-driven aerosol transport between two individuals. Physical Review Fluids,

5(12), 1–10. https://doi.org/10.1103/PhysRevFluids.5.122501

Yang, J., Sekhar, C., Cheong, D. K. W., & Raphael, B. (2015a). A time-based analysis of the

personalized exhaust system for airborne infection control in healthcare settings. Science

and Technology for the Built Environment, 21(2), 172–178.

https://doi.org/10.1080/10789669.2014.976511

Yang, J., Sekhar, C., Cheong, D., & Raphael, B. (2014). Performance evaluation of an

integrated Personalized Ventilation-Personalized Exhaust system in conjunction with

65
two background ventilation systems. Building and Environment, 78, 103–110.

https://doi.org/10.1016/j.buildenv.2014.04.015

Yang, J., Sekhar, S. C., Cheong, K. W. D., & Raphael, B. (2015b). Performance evaluation of

a novel personalized ventilation-personalized exhaust system for airborne infection

control. Indoor Air, 25(2), 176–187. https://doi.org/10.1111/ina.12127

Yao, M., Zhang, L., Ma, J., & Zhou, L. (2020). On airborne transmission and control of

SARS-Cov-2. The Science of the Total Environment, 731, 139178.

https://doi.org/10.1016/j.scitotenv.2020.139178

Zayas, G., Chiang, M. C., Wong, E., MacDonald, F., Lange, C. F., Senthilselvan, A., & King,

M. (2012). Cough aerosol in healthy participants: fundamental knowledge to optimize

droplet-spread infectious respiratory disease management. BMC Pulmonary Medicine,

12(1), 11. https://doi.org/10.1186/1471-2466-12-11

Zhai, Z. (2020). Facial mask: A necessity to beat COVID-19. Building and Environment, 175,

106827. https://doi.org/10.1016/j.buildenv.2020.106827

Zhang, H., Li, D., Xie, L., & Xiao, Y. (2015). Documentary Research of Human Respiratory

Droplet Characteristics. Procedia Engineering, 121, 1365–1374.

https://doi.org/10.1016/j.proeng.2015.09.023

Zhang, N., Chen, W., Chan, P. T., Yen, H. L., Tang, J. W. T., & Li, Y. (2020). Close contact

behavior in indoor environment and transmission of respiratory infection. Indoor Air,

February, 1–17. https://doi.org/10.1111/ina.12673

Zhang, S., Ai, Z., & Lin, Z. (2021). Occupancy-aided ventilation for both airborne infection

risk control and work productivity. Building and Environment, 188(October 2020),

107506. https://doi.org/10.1016/j.buildenv.2020.107506

66
Zhang, S., & Lin, Z. (2020). Dilution-based Evaluation of Airborne Infection Risk -

Thorough Expansion of Wells-Riley Model. MedRxiv, 2020.10.03.20206391.

http://medrxiv.org/content/early/2020/10/06/2020.10.03.20206391.abstract

Zhang, T., Wang, S., Sun, G., Xu, L., & Takaoka, D. (2010). Flow impact of an air

conditioner to portable air cleaning. Building and Environment, 45(9), 2047–2056.

https://doi.org/10.1016/j.buildenv.2009.11.006

Zhang, Y., Feng, G., Bi, Y., Cai, Y., Zhang, Z., & Cao, G. (2019). Distribution of droplet

aerosols generated by mouth coughing and nose breathing in an air-conditioned room.

Sustainable Cities and Society, 51(July), 101721.

https://doi.org/10.1016/j.scs.2019.101721

Zhou, Q., Qian, H., Ren, H., Li, Y., & Nielsen, P. V. (2017). The lock-up phenomenon of

exhaled flow in a stable thermally-stratified indoor environment. Building and

Environment, 116, 246–256. https://doi.org/10.1016/j.buildenv.2017.02.010

Zhu, S. W., Kato, S., & Yang, J. H. (2006). Study on transport characteristics of saliva

droplets produced by coughing in a calm indoor environment. Building and

Environment, 41(12), 1691–1702. https://doi.org/10.1016/j.buildenv.2005.06.024

Zuraimi, M. S., Nilsson, G. J., & Magee, R. J. (2011). Removing indoor particles using

portable air cleaners: Implications for residential infection transmission. Building and

Environment, 46(12), 2512–2519. https://doi.org/10.1016/j.buildenv.2011.06.008

67

You might also like