Next Article in Journal
Assessment of Particulate Matters, Metals, and PAHs’ Air Pollution in Industrial vs. Animal Husbandry Areas
Previous Article in Journal
Projected Changes of Wind Energy Input to Surface Waves in the North Indian Ocean Based on CMIP6
Previous Article in Special Issue
Numerical Study on the Impact of Large Air Purifiers, Physical Distancing, and Mask Wearing in Classrooms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational Fluid Dynamics-Based Calculation of Aerosol Transport in a Classroom with Window Ventilation, Mechanical Ventilation and Mobile Air Purifier

E.ON Energy Research Center, Institute for Energy Efficient Buildings and Indoor Climate, RWTH Aachen University, 52074 Aachen, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Atmosphere 2024, 15(1), 140; https://doi.org/10.3390/atmos15010140
Submission received: 29 November 2023 / Revised: 10 January 2024 / Accepted: 18 January 2024 / Published: 22 January 2024

Abstract

:
During the SARS-CoV-2 pandemic, the air quality and infection risk in classrooms were the focus of many investigations. Despite general recommendations for sufficient ventilation, quantitative analyses were often lacking due to the large number of combinations of boundary conditions. Hence, in this paper, we describe a computational fluid dynamics model that predicts the time-resolved airflow for a typical 45 min classroom scenario. We model 28 students and a teacher, each emitting CO2 and an individual aerosol. We investigated 13 ventilation setups with window or mechanical ventilation and different positions and operating conditions of an additional air purifier. The ventilation performance is assessed by evaluating the ventilation effectiveness, aerosol removal effectiveness, local air exchange efficiency and overall inhaled aerosol mass of the occupants, which is a measure of the infection risk. If the window is opened according to the “20-5-20” recommendation, the incoming airflow reduces both the CO2 and aerosol concentration whilst decreasing the thermal comfort at low ambient temperatures. An active air purifier enhances aerosol removal, but, depending on the position, the local air exchange efficiency and individual aerosol inhalation vary. If mechanical ventilation with 700 m3/h is utilised, the CO2 concentration is kept below 1250 ppm while also effectively removing aerosol from the classroom.

1. Introduction

During the SARS-CoV-2 pandemic, the virus mainly spread within insufficiently ventilated indoor environments. The prominent case described by Lu et al. [1], where the virus was spread by the air conditioning system in a restaurant, emphasises not only the need for an overall sufficient ventilation rate but also the requirement of an overall good distribution of fresh air. Public spaces like restaurants or cinemas were widely shut down, but since schools had to stay open as long as possible, the indoor environment of classrooms became an intensively discussed topic. To enable a safe environment within classrooms that prevented the uncontrolled spread of the virus, several institutions and authors analysed case studies and developed models and/or tools to assess the infection risk and derive measures to allow the normal operation of schools.
One reported case of infection within a classroom setting was described by Lam-Hine et al. [2]. In an elementary school, half of the students were infected by the symptomatic teacher. The teacher and the students did not wear face masks, but a mobile air purifier was operated.
Many researchers developed models to predict the risk of infection [3,4,5,6,7]. Müller et al. [4] developed a simplified calculation approach to determine a relative infection risk from aerosol-borne viruses in indoor environments. They assume a reference scenario for which an absolute infection risk was derived based on various approaches as known from the present literature. Using this reference scenario as a baseline, a relative infection risk can be computed for various other scenarios. The calculation was compared to the results of other tools [7,8,9].
The German Federal Environment Agency (UBA) published a handout where they proposed natural ventilation as a measure to reduce the risk of infection, especially in the context of classrooms [10]. They also note that mobile air purifiers are viable add-on measures to further reduce the infection risk. The Center for Disease Control and Prevention (CDC) suggests a minimum of five air changes per hour [11] but does not further specify how this is best achieved. A similar recommendation is given by the Chinese Center for Disease Control and Prevention (CCDC), as they suggest a minimum fresh air rate of 30 m 3 / h per person [12]. If this lower limit can not be sustained, the number of occupants in the room has to be reduced.
Lowther et al. [13] rated the removal efficiency of selected air purifiers regarding different particle size distributions. Overall, they observed higher removal efficiencies at higher volume flow rates. Curtius et al. [14] investigated the filtration behaviour of mobile air purifiers in a closed classroom setting. They observed a significant reduction in inhaled particle doses when operating the mobile air purifier at high volume flow rates. In their study, Jain et al. [15] investigate a classroom setting where a large air purifier is operated at different positions and volume flow rates. Under the assumption of steady-state conditions, neglecting any transient effects, they calculate an infection risk originating from varying index patients. They observed an overall reduction in the infection risk in all investigated scenarios where the air purifier was active, although zones with high local concentrations were found. Assuming different heating settings in a small chamber, Sabanskis et al. [16] confirmed the strong sensitivity of the filtering efficiency of a mobile air purifier to the positions of both the aerosol source and the filtration device. Duill et al. [17] conducted a thorough study to investigate a classroom setting using different air purifier types, with one type operated at two different locations. The classroom was additionally ventilated following the guidelines from the UBA [10]. They were able to show that mobile air purifiers can significantly reduce aerosol concentrations. A developed flow model was able to predict the aerosol decay rate in good agreement with measured data, although it was only validated for periods without window ventilation. At the tested volume flow rates of about 1000 m 3 / h , the noise of the devices was reported to not negatively impact the learning environment. Fierce et al. [18] performed simulations of a quadrature-based model to predict the spread and inhalation of airborne pathogens. They investigated the influence of the use of face masks and increased ventilation rates. Both measures lead to a significant reduction in droplet inhalation for larger distances away from the droplet source.
Chirico et al. [19] reviewed several reported outbreaks in Asia that had in common that a central ventilation system recirculated at least a portion of the airflow into the rooms. Because the ventilation systems used recirculation air, they captured the infectious particles and distributed them over the whole ventilation system, leading to more infections. In their review, no studies on classrooms were included. Elsaid and Ahmed [20] reviewed current design methods of ventilation systems with a focus on reducing the spread of potentially infectious particles. They recommend operating air handling units such that as much outdoor air as possible is provided into the indoor environment. Furthermore, the air handling units should utilise filters that are capable of effectively removing particles below 0.3   μ m from the air. In his study, Melikov [21] reviewed a mixing, underfloor and displacement ventilation scenario by conduction tracer gas measurements. He concludes that increasing the ventilation rate utilising mixing ventilation is not always the best way to reduce the infection risk, since it potentially spreads the airborne infection in the room. Using an infection risk to derive the ventilation effectiveness, Kurnitski et al. [22] assess the local air quality using tracer gas measurements. They discuss different room scenarios with varying locations of the point source, which represents an infected person. Across all scenarios, the source location in combination with the position(s) of the extract air ducts strongly impacts the overall distribution of the air quality.
There seems to exist an overall consensus on the fact that a ventilation system utilising recirculation air poses an increased risk of infection. To reduce the infection risk, the same ventilation system can be used if only non-contaminated air, such as fresh outdoor air or cleaned air, is supplied. Since many public facilities, especially German school classrooms, are not equipped with a mechanical ventilation system, only natural ventilation or mobile air purifiers are available to lower the infection risk. As most air purifiers do not filter human CO2 emissions, their usage without additional ventilation is not advisable or purposeful regarding indoor air quality. Particularly during the winter season, window ventilation is only utilisable intermittently due to thermal comfort reasons [23]. In contrast, air purifiers can be operated continuously. This may result in transient progressions of pollutant concentrations during a school lesson. In addition, there are different thermal buoyancy effects to be considered inside a classroom due to heat sources, such as students and teachers, radiators and lamps, and further thermal effects due to the unconditioned outdoor air caused by window ventilation. Under these aspects, it can be assumed that inhomogeneous and transient pollutant concentrations may be present under the operation of air purifiers and window ventilation in classrooms. Under inhomogeneous concentration distributions, the position of the air purifier in the classroom can have an important impact on effectiveness [17]. Thus, to accurately evaluate the effectiveness of air purifiers in filtering potentially infectious particles, both transient and inhomogeneous effects in the room must be considered.
All the discussed studies have in common that they assume steady-state flow conditions. Window ventilation is a strongly transient flow, which is challenging for the descriptions commonly applied to comparable problems. Additional transient effects result from the intermittent use of window ventilation. We introduce a significantly higher number of tracers into our simulation setup, which allows us to reliably derive the uncertainty associated with each investigated scenario. By treating CO2 as an inert tracer that is not filtered, we can segregate flow behaviour and filtration effectiveness. By using detailed, transient flow simulation models, including window or mechanical ventilation and mobile air purifiers, we aim to derive metrics that describe various effects influencing the inhalation of a potentially infectious aerosol. Besides the ventilation strategy proposed by the UBA, we also assess the influence of different air purifier operation strategies by varying the volume flow rate and air purifier location. In this work, we further extend the investigations from Ostmann et al. [24] by also taking a simplified air handling unit into account.

2. Methods

This section is dedicated to the description of the developed flow model. We will first describe the geometry, the generated computational grid and the used flow models. A detailed discussion of our assumptions regarding energy and species transport is given before we introduce the developed metrics to rate the performance of the proposed ventilation concepts.
The whole model is developed using the software package ANSYS 2020 R2 [25]. We use the associated Design Modeler to design the geometry and then feed it into ANSYS Meshing to generate the computational grid. The setup of the flow model is implemented using the CFX package.

2.1. Description of the Geometry

The geometry of the classroom model is derived from a real scenario [26], and a summary of the geometric parameters is given in Table 1. An overview of the investigated classroom settings is shown in Figure 1. The 28 students and the teacher are modelled as manikins. The geometric volume of the classroom is reduced by the volume of the manikins and the air purifiers to an effective air volume of roughly 200 m 3 . We treat the desks as solid, adiabatic surfaces. The students are seated in a typical manner, with the mouths of the students being oriented either towards the opposite wall or towards the wall behind the teacher Atmosphere 15 00140 i013 (see Figure 1). The teacher is assumed to be standing facing towards the back wall of the classroom and therefore in the direction of the students. We denote the students using an index in the range i , j = 1 , 28 , and the teacher is denoted by either the index i , j = 29 or “T”.
All present occupants are simplified using manikins consisting of a cylindrical base shape with a half-sphere on top according to Menchaca-B. and Leon R. Glicksman [27]. To yield roughly equal surface areas, the seated students feature a height of 1.3   m with a diameter of 0.43   m , and the teacher is 1.75   m tall with a diameter of 0.32   m . Each manikin features a mouth area of A Mouth = 3   cm · 1   cm = 3   cm 2 , which is located 0.13   m below the tip.
Four radiators are located below the window sills. We simplify the radiators using a cuboid volume with a volumetric heat source. We assume a total of ten ceiling lights by dedicating two long surfaces where each surface approximates five lights. The ceiling lights are aligned with the ceiling itself to decrease the mesh complexity.
We propose four possible locations for a mobile air purifier (see Figure 1a) in the corners of the classroom. The air purifier (AP) model is based on a simplified version of a commercially available air purifier of the type “TAP” from the company TROX GmbH TROX TAP (see [28]). We approximate the inlet of the air purifier using a rectangular surface and the outlet using a circular area. For practical reasons regarding the operability of the model, the same geometric setup is used for each air cleaner scenario, and the respective air cleaners are switched on or off only. In consequence, the geometry of all four possible air purifiers is present in all investigated scenarios. However, we estimate the influence of an inactive air purifier to be insignificant if the respective surfaces are treated as adiabatic boundaries. A more detailed geometric description of the air purifier can be found in Ostmann et al. [29].
The classroom features four windows, each representing a two-winged window as per the referenced classroom [26]. The ratio of window area to classroom floor area is roughly 0.157. The window sills protrude into the room by 0.35   m and cover the windows from the radiators below.
In addition to purely natural convective ventilation, we also investigate the case of mechanical ventilation with a simplified air handling unit (AHU) (see Figure 1b). The supply air is provided through a long rectangular slot with a flow cross-section of A SUP = 0.33   m 2 , which covers nearly the whole width of the device. The extract air is drawn through a rectangular opening in the bottom of the device with a cross-section of A ETA = 0.56   m 2 .

2.2. Computational Flow Model

We use ANSYS Meshing to generate an unstructured tetrahedral volume mesh. To resolve the boundary layer on relevant surfaces, we generate so-called “prism layers”. A section through the whole computational grid of the classroom is shown in Figure 2, where the most relevant features of the flow model are highlighted. We set an average grid resolution of Δ x = 0.1   m but implement certain refinements to capture the relevant flow structures. A summary of the implemented refinements is given in Table 2. Most of the refinements are implemented as surface refinements, where the resulting surface mesh is refined to Δ x Refinement . To increase the influence of the refined surface mesh into the volume mesh, a certain influence radius r Influence is specified. As already described, the radiators are simplified as volumes, which is why we use volumetric refinement to properly resolve the heat sources. The resulting mesh consists of roughly 2.4 × 106 cells.
To properly capture the heat transfer from heated surfaces, it is crucial to resolve the airflow boundary layer by implementing several prism layers. Figure 3 shows the prism layer distribution on the surface of a manikin.
To fully describe the prism layers, three parameters (the first layer height, number of layers and growth ratio) are needed. The first layer height h 1 describes the height of the first prism layer on the surface and is required by the turbulence model to assess the wall regime and select the appropriate wall function. Based on the first layer height, the remaining prism layers are constructed using the height of the preceding layer h i and a growth rate factor f Growth according to Equation (1).
h i + 1 = h i · f Growth
In Table 3, the parameters of the prism layers are given for all surfaces, where we resolve the boundary layer.
We assess the quality of our mesh by conducting a grid independency study according to the procedure described by Celik et al. [30]. To reduce the required computational effort, we compute a steady-state solution with the AP active at location Atmosphere 15 00140 i001 with a volume flow rate of 1400 m 3 / h . Furthermore, we only consider student Atmosphere 15 00140 i011 to be present. The geometries of all other occupants are included but treated as passive and adiabatic. We iterate the solution for 12,000 steps, which is enough to achieve sufficient convergence when averaging the steady-state solutions over the last 1000 iterations. Apart from the base mesh configuration, we also investigate two additional meshes where we reduce and increase all mesh parameters, including the number and first layer height of prism layers, by a factor of 1.5. This yields a coarser ( 1.0 × 10 6 cells) and a finer ( 6.9 × 10 6 cells) computational grid. We evaluate both grid convergence indices GCI base / fine for the average classroom temperature T ¯ Room and the aerosol concentration that is captured by the AP x Aerosol , 20 , AP 1 , in . The average classroom temperature features convergent behaviour as indicated by GCI base = 3.86 × 10 4 and GCI fine = 2.95 × 10 6 and a ratio of ε 21 / ε 32 = 0.13 (see Figure A1). The deviation ε ij is defined as the difference between the values on the finer grid i and the coarser grid j. Although the captured aerosol mass also features good grid convergence with GCI base = 4.74 × 10 2 and GCI fine = 4.49 × 10 2 , the associated order p is significantly lower. The ratio ε 21 / ε 32 = 2.30 indicates oscillatory and diverging behaviour (see Figure A2). The overall deviations across the three investigated grids are below 4% compared to the base value. Therefore, we evaluate the resolution of the base grid to be a reasonable tradeoff between accuracy and computational effort.
An overview of the models used is given in Table 4. We model the fluid as air with a constant density of ρ Air = 1.185   k g / m 3 . Due to the lower density of warmer air, buoyancy forces arise, which are modelled using the Boussinesq assumption with a reference temperature of 20 °C, assuming sufficiently small temperature variations. The reference temperature matches the surface temperature of the room walls (see Section 2.5.3) and is, therefore, a consistent baseline temperature across all scenarios. Since scale-resolved modelling of the occurring turbulence is very resource-intensive, the turbulence is represented in a simplified way using a two-equation model. In this work, we use the k- ω -SST model, as it performs better at describing the boundary layer very close to walls than the k- ε model [31,32]. We can therefore use the fine grid discretisation of our mesh to model the heat transfer process described in Section 2.5.

2.3. Implementation of Window Ventilation

To accurately model the ventilation behaviour, the classroom’s flow model needs to be extended by an additional volume representing the outdoor environment. Although more simple empirical models describing the flow through open windows exist, they are not suitable to represent the strongly transient and, regarding the window opening area, not uniform behaviour during the investigated window ventilation phase [36]. We model the outdoor environment as a cuboid with the dimensions 10 m × 16 m × 6 m (L × W × H), embodying a total volume of 960 m 3 , as shown in Figure 4.
All scenarios that feature window ventilation use the same (20/5/20) recommendation strategy proposed by the German Federal Environment Agency [10]. They recommend fully opening the windows for at least 3–5 min after every 20 min during a lecture. Therefore, the windows are closed for the first 20 min of the simulation. Within one simulation step, the windows are fully opened, representing the maximum available cross-section for window ventilation. When opening the windows, we change the window surfaces from a wall boundary condition to an interface connected to the outdoor environment. The outdoor environment is initialised with an air temperature T amb as specified by the respective scenario and a homogeneous CO2 concentration equal to c CO 2 , ref . We further assume no air movement like wind or gusts in the environment, which could further enhance the air exchange. When closing the windows after an additional 5 min , the window’s surfaces are changed back to wall boundaries and the outdoor environment is neglected again. For the periods where the windows are closed, they are modelled with a constant heat transfer coefficient. This represents a heat loss to the (colder) ambient environment.
To evaluate the effective volume flow rate entering the classroom, we first define our x-direction to be positive in the direction from the outdoor environment into the classroom. Therefore, we define a surface where the velocity in the x-direction is positive ( v x > 0 ), representing the section of the window cross-section where air enters the classroom. Using this condition, we develop the expression in Equation (2) that yields the volume flow rate entering the classroom V ˙ Window , in . A similar expression can be developed for the volume flow rate exiting the classroom V ˙ Window , out by switching the condition for the k direction .
V ˙ Window , in = A Window v · k direction d A with k direction = 1 v x > 0 0 v x 0
To calculate the average volume flow entering the classroom during the window ventilation period over the windows, V ˙ ¯ Window , in we average the V ˙ Window , in from Equation (2) and relate it to the duration of one lesson according to Equation (3).
V ˙ ¯ Window , in = τ = 20 min 25 min V ˙ Window , in ( τ ) d t 45 min

2.4. Species Transport

Each occupant is assumed to emit CO2 and an individual aerosol Φ i . The aerosol represents potentially infectious airborne particles and is used to evaluate the effectiveness of ventilation designs or air cleaners. Typically, airborne infections are caused by so-called droplet nuclei, i.e., residues of vaporised droplets and particles from the human respiratory tract. The WHO classifies droplet nuclei as particles with a diameter of less than 5 μ m [37]. Laboratory studies show that particles in the sizes from 70 n m to 5 μ m disperse similarly to N2O tracer gas in indoor air under mixed ventilation conditions [38]. In another study, particles in sizes from 30 n m to 3.2   μ m  were shown to disperse similarly to SF6 tracer gas in indoor air under both mixed ventilation conditions and natural convection [39]. Thus, it can be assumed that particles in the relevant size range up to about 5 μ m ideally follow the room airflow similarly to dilute gas. Therefore, we model both species with a passive scalar transport approach, assuming both the CO2 and aerosol follow the airflow perfectly.
To ensure a realistic dispersion from the source into the room, we implement a breathing cycle for every occupant. During this cycle, a specified volume flow rate is exhaled according to the emission concentrations. Afterwards, the same volume flow rate is inhaled. Figure 5 shows the two periods of the breathing cycle.
One breathing cycle takes Δ t Breath = 6   s , where both the exhalation and inhalation take Δ t Breath , ex = Δ t Breath , in = 2   s with a gap of 1 s between. The profiles’ amplitude is taken from Buonanno et al. [3] with V ˙ Breath = 0.54 m3/h. Using the mouth area A Mouth , we calculate the necessary flow velocity and prescribe the boundary condition on the mouth surfaces, neglecting breathing through the nose. Since Buonanno et al. [3] report only an average volume flow rate and not the peak volume flow rate, our model underestimates the species emission associated with the breath. To compensate for this, we introduce a correction factor k Breath according to Equation (4) and apply it to the emissions during post-processing (see Equations (10) and (14)).
k Breath = Δ t Breath Δ t Breath , ex = 6   s 2   s = 3
Although we also underestimate the introduced momentum during exhalation, the effects on the airflow pattern in the room are negligible. According to Regenscheit [40], we can estimate that even with a higher flow velocity, the resulting jet velocity decays to below 0.1 m/s after 80 mm. This distance is small compared to the overall length scales of our flow model.
All emission-related parameters are summarised in Table 5. We will further discuss the sources and sinks for each species in the following two sections.

2.4.1. CO2 Transport

We model CO2 as an additional passive scalar m CO 2 emitted equally by all occupants. To convert between the mass and volume fraction of CO2, we use the ideal gas law with a constant temperature of 300 K . By further assuming a constant pressure, the density of CO2 is also constant according to Equation (5). We obtain the relation between the volume fraction c CO 2 and the mass fraction x CO 2 according to Equation (6).
ρ CO 2 = p R s · T const .
x CO 2 = c CO 2 · ρ CO 2
Our temperature assumption leads to an overestimation of the volume fraction since the observed classroom temperatures are generally below 300 K .
Neglecting the occupants’ inhalation of CO2, we establish the mass balance according to Equation (7) to describe the transport of CO2 assuming a perfectly air-tight building façade. If we assume the CO2 concentration in the inhaled air to be equal to the average room concentration (less than 2000 ppm ), the CO2 contribution of the exhaled air dominates by a factor of more than 20.
m CO 2 t = i = 1 28 + 1 V ˙ Breath , i ( t ) · c CO 2 , i · ρ CO 2 + m ˙ CO 2 , Window , in m ˙ CO 2 , Window , out + m ˙ CO 2 , AHU , in m ˙ CO 2 , AHU , out
Assuming equal entering and exiting volume flow rates through the windows and air handling unit, Equation (7) reduces to the mass balance in Equation (8).
m CO 2 t = i = 1 28 + 1 V ˙ Breath , i ( t ) · x CO 2 , A , i + V ˙ Window ( t ) x CO 2 , W , in x CO 2 , W , out + V ˙ AHU x CO 2 , AHU , in x CO 2 , AHU , out
In the case that the air handling unit is operated without recirculation, the CO2 concentration of both the supply air and ambient air can be considered equal ( c CO 2 , W , in = c CO 2 , AHU , in = c CO 2 , ref ). The total CO2 mass is then evaluated by integrating the expression in Equation (8) in time according to Equation (9).
m CO 2 = i = 1 28 + 1 0 τ V ˙ Breath , i ( t ) · x CO 2 , A , i d t + 0 τ V ˙ Window ( t ) x CO 2 , W , in x CO 2 , W , out d t + V ˙ AHU · t · x CO 2 , AHU , in x CO 2 , AHU , out
The assumption of the exhaled CO2 concentration is based on a continuous emission. Hence, we have to compensate for the intermitting breathing cycle by applying the correction factor k Breath according to Equation (10), since the aerosol emission is directly coupled to underpredict the breathing volume flow rate.
m CO 2 , corr = m CO 2 + m CO 2 m CO 2 , ref · k Breath

2.4.2. Aerosol Transport

Every person emits an individual aerosol Φ i that is modelled as a passive scalar representing the mass of the aerosol. For the total mass of one individual aerosol Φ i , tot , we establish the mass balance according to Equation (11).
Φ i , tot t = V ˙ Breath , i ( t ) · x Aerosol , i Φ ˙ i , AP , in ( t ) Φ ˙ i , AP , out ( t ) Φ ˙ i , Window ( t ) Φ ˙ i , AHU ( t ) j = 1 , i j 28 + 1 Φ ˙ i , j ( t )
The difference of the terms Φ ˙ i , AP , in / out represent the aerosol mass that is filtered by the air purifier with the filter efficiency η AP according to Equation (12).
Φ ˙ i , AP , out ( t ) = 1 η AP Φ ˙ i , AP , in ( t )
Φ ˙ i , Window represents the aerosol exiting the room via window ventilation. Φ ˙ i , AHU defines the aerosol removed by the AHU. We assume that the air supplied by the AHU is not contaminated by the aerosol. Finally, Φ ˙ i , j captures the aerosol inhaled by the other occupants j.
We assume the teacher to be speaking while the students are listening only. Therefore, the teacher emits a higher aerosol concentration compared to the students’ emissions ( c Aerosol , T > c Aerosol , S ). We derive the aerosol’s density from the assumption that the exhaled droplets mainly consist of water. In analogy to CO2, a similar expression correlates the mass and volume fraction (see Equation (13)).
x Aerosol , i = c Aerosol , i · ρ Aerosol
We calculate the values of the volume fraction c Aerosol , i according to Buonanno et al. [3,43]. They reference the measurements of Adams [41], who observed an average droplet concentration of 6 × 10 2   m L / m 3 . Depending on the respiratory activity, Morawska et al. [42] reports a factor of 0.03 for oral breathing and 0.16 for loud speaking, which we use to determine the respective volume fractions reported in Table 5. To compensate for the breathing cycle, we have to apply the correction factor f Breath for the aerosol according to Equation (14).
Φ i , corr = Φ i , tot · k Breath

2.5. Heat Sources and Sinks

The undisturbed airflow pattern is dominated by natural convection due to various heat sources and sinks. To accurately predict the airflow and therefore the spread of the emitted aerosol throughout the classroom, the modelling of the heat sources needs careful consideration.

2.5.1. Occupants

The main heat sources are the occupants, consisting of their surface heat flux q ˙ i and the exhaled air. We approximate the surface heat flux from the basal thermal power, where we model the students P Basal , i and the teacher P Basal , T according to Equations (15) and (16) based on their mass m i and age [44,45].
P Basal , i W = 0.065 · m i k g + 2.826 · 11.574
P Basal , T W = 0.048 · m T k g + 3.653 · 11.574
We assume the students’ ages to be in the range of 16–18 y and assume them to be half male and half female, which is why the factors in Equation (15) represent their respective average values according to Schofield [45]. The teacher’s age is assumed to be in the range of 30–60 y . We approximate the occupants’ weights according to Pharmacia [46,47], who base their models on the works of Reinken et al. [48], Brandt [49,50], Brandt and Reinken [51], Reinken and v. Oost [52]. The values for the students align well with more recent data as published by Gao et al. [53]. All parameters necessary to describe the occupants’ heat emissions are summarised in Table 6.
We calculate the surface heat flux according to Equation (17), where we assume a certain activity factor k Activity and an approximated body surface area according to Equation (18) as proposed by Du Bois and Du Bois [54].
q ˙ i = P Basal , i A DuBois , i · k Activity
A DuBois , i m 2 = 0.007184 · h i c m 0.725 · m i k g 0.425
Although insignificant on the overall thermal balance of the classroom, the exhaled air significantly impacts the transport of the aerosol, due to the exhaled air being warmer and therefore experiencing thermal buoyancy.

2.5.2. Air Purifiers

An air purifier is typically equipped with a fan delivering the necessary airflow V ˙ AP . Due to electrical, mechanical and aerodynamic losses, we establish a simplified energy balance of the consumed electrical power P el , AP according to Equation (19). The relation between the airflow rate and electrical power demand is linear according to Equation (20) [28].
P el , AP = Q ˙ AP + P Aero = Q ˙ Case + Q ˙ Air , out + P Aero
P el , AP = V ˙ AP · 0.095 W m 3 / h
We assume an electric fan efficiency of the air purifier of η el , AP = 60 % , which determines the portion of the electrical power that is converted into heat according to Equation (21).
Q ˙ AP = 1 η el , AP · P el , AP
To determine the waste heat’s Q ˙ AP portion that is transfered to the air, we define the factor f Q , AP = 70 % . The rest is emitted over the case surface. Using Equation (21), we can develop an expression that represents the heat portion introduced into the exiting airflow Q ˙ Air , out (see Equation (22)), which determines the temperature of the airflow exiting the air purifier T AP , out according to Equation (23).
Q ˙ Air , out = P el , AP · 1 η el , AP · f Q , AP
T AP , out = T AP , in + Q ˙ Air , out V ˙ AP , out · ρ AP , out · c p , Air
Similarly, we can describe the heat portion emitted by the air purifier’s case surfaces Q ˙ Case (see Equation (24)). With the surface area A AP , we can derive the case’s surface heat flux q ˙ Case according to Equation (25).
Q ˙ Case = P el , AP · 1 η el , AP · ( 1 f Q , AP )
q ˙ Case = Q ˙ Case A AP

2.5.3. Room and Window Surfaces

The room facade is split into three areas. The first area consists of the window frames and the wall below the window sills. The second area represents the windows, and the third consists of the window sills. Each area features an individual heat transfer coefficient k HTC , i (see Table 7) used to calculate the heat transfer where the reference temperature is equal to the ambient air temperature defined by the respective scenario.
The values resemble a weakly insulated facade with single-layered windows. All other room surfaces are modelled with a constant wall temperature of 20 °C except the wall opposite of the facade, where we apply a constant wall temperature of 18 °C to account for a slightly cooler hallway.

2.5.4. Simple Heat Sources

The air handling unit is assumed to provide a constant supply air temperature of T AHU , SUP = 21   C = const . in all investigated scenarios. This represents a device equipped with heat recovery and heating and cooling coils. The room-facing surfaces of the device are modelled with a constant wall temperature of 20 °C.
We model the radiators as simple cuboid volumes below the window sills. With their dimensions being 1.5   m × 0.25   m × 0.65   m (L × W × H), they resemble typical radiators [26]. We apply a volumetric heat source to the radiators’ volumes, such that each radiator delivers 100 W of thermal power, which adds up to a total of 400 W across all four radiators. We chose a rather low value to prevent an excessive heat-up of the classroom. For the scenarios (see Section 2.6) with an ambient air temperature of T amb = 20   C , the radiators are disabled.
The ceiling lights are modelled by two rectangular surfaces in the ceiling. The total area of the ceiling lights adds up to A Lights = 2 × 7   m × 0.3   m = 4.2   m 2 , equaling the area of ten typical ceiling lights. Each ceiling light is assumed to emit 50 W of heating power, which results in 500 W of heat power being introduced into the room by the lighting.

2.6. Investigated Scenarios

We investigate 13 scenarios in total, which can be grouped into five different sets, which are summarised in Table 8. The first set (“WV”) represents a baseline with no active air purifiers where we compute the window ventilation behaviour for three different ambient air temperatures T amb = ( 5 ,   15   and   20 )   C . For the second set (“WV + AP”), we analyse the simultaneous operation of window ventilation and air cleaner Atmosphere 15 00140 i001, which operates at a volume flow rate of V ˙ AP , 1 = 1400   m 3 / h , for the same variation in ambient air temperature.
The ambient air temperature is set to T amb = 5   C for all other scenario sets. The third (“1400 m 3 / h ”) and fourth (“700 m 3 / h ”) scenario sets investigate the influence of the different air purifier locations and the air purifier volume flow rate. All scenario sets described so far share the same geometry as shown in Figure 1a. A detailed description of the transient solution scheme applied for the cases with window ventilation can be found in Ostmann et al. [29].
The last scenario set (“AHU”) uses the geometry shown in Figure 1b. We do not consider window ventilation in this scenario set; therefore, the ambient air temperature is mainly relevant for the heat convection through the facade wall. We set the AHU’s volume flow V ˙ AHU to 700 m 3 / h and compare two scenarios with and without the air purifier.
The calculation of one scenario required about 12,000 core-h on the CLAIX-18 high-performance computing cluster system of the RWTH Aachen University. Therefore, we required roughly 160,000 core-h to calculate the presented results.

2.7. Evaluation Metrics

We use three metrics to rate the effectiveness of the different ventilation strategies. They are derived from metrics recognised in the literature and adapted to the case of transient phenomena. Due to our setup, we can not only capture spatially averaged phenomena but are also able to track the transmission between individuals. Figure 6 shows an overview of the seating arrangement to identify individual occupants i and the positions of the air purifier. Besides the metrics discussed in this paper, we also evaluated the thermal comfort in another publication (see Ostmann et al. [24]). The metrics described in the following section are also summarised in the Appendix B (see Table A1).
To rate the ability to remove contaminants from an indoor environment, the REHVA Guidebook proposes the contaminant removal effectivess (CRE) ε c [55,56]. It is defined by the ratio of the contaminant concentration in the extracted air x e to the spatially averaged concentration of the contaminant in the room x according to Equation (26). Furthermore, the definition assumes a steady-state flow.
ε c = x e x
As shown in Figure 7, the contaminant removal effectiveness can be divided into three ranges.
If the contaminant is extracted out of the room quickly without significant mixing with the surrounding air, x e will be greater than x , leading to ε c > 1 . This can be achieved by placing the extract air opening near the contaminant source. If the contaminant source is located further away from the extract air opening or in a recirculation region, the relation is inversed. The contaminant accumulates near its source and is not removed effectively, leading to ε c < 1 . In the special case of ideal mixing ventilation where the contaminant is homogeneously mixed in the air, the CRE approaches ε c = 1 .
While Mundt [55] developed the contaminant removal effectiveness assuming an AHU, we propose a similar classification if a mobile air purifier is present instead of an AHU, as shown in Figure 8. In this case, the relevant extracted contaminant concentration is the one at the inlet of the air purifier.
The air change efficiency ε a by Mundt [55] rates the time that is required to replace the air inside a room in comparison to the theoretically shortest time. Since the shortest time is achieved if a piston flow pattern is established, the air change efficiency is 100% for this case. The more practical displacement flow achieves values in the range of 50 % < ε a < 100 % , where the lower value of 50% resembles a fully mixed flow. If the air change efficiency drops below 50%, it is an indication of a short-circuit flow.
A simplified classification of both the contaminant removal effectiveness and the air change efficiency as defined by Mundt [55] in the style of Wildeboer and Müller [57] is shown in Figure 9. Four domains can be distinguished by different value ranges of both metrics, with all four domains touching at an air exchange efficiency of 50% and a ventilation effectiveness of one. The northeast quadrant indicates piston-flow behaviour with either displacement ventilation or at least a strong stratification of air layers. In the southeast quadrant, the airflow follows a displacement ventilation behaviour, but the contaminant source is not simultaneously the driving heat source. The two western quadrants indicate a short-circuit flow, whereas the northwest quadrant includes the case of a contaminant source near the extract position.

2.7.1. Ventilation Effectiveness

For steady-state conditions, Mundt [55], Brouns and Waters [56] propose various indices to assess the ventilation effectiveness under the assumption of steady-state flow conditions. In the scope of ventilation effectiveness, we treat CO2 as the investigated contaminant. Since window ventilation (see Section 2.3) features strongly transient behaviour, we develop a time-resolved definition of the ventilation effectiveness for window ventilation according to Equation (27). We relate the CO2 concentration of an idealised mixing ventilation behaviour x CO 2 , MV to the actual computed CO2 concentration in the modelled classroom x CO 2 .
ζ c ( t ) = x CO 2 , MV ( t ) x CO 2 ( t )
To calculate the idealised mixing ventilation, we use the simplified CO2 mass balance according to Equation (8) but approximate the breathing processes of the 29 occupants using a constant CO2 emission of m ˙ CO 2 according to Equation (28).
i = 1 28 + 1 V ˙ Breath , i ( t ) · x CO 2 , A , i 29 · V ˙ Breath · c CO 2 , i · ρ CO 2 = m ˙ CO 2
Using the simple explicit numerical scheme shown in Equation (29), we propagate Equation (8) in time with the constant timestep Δ t .
x CO 2 , MV ( t ) = x CO 2 , MV ( t Δ t ) + m ˙ CO 2 , P · Δ t V Room + V ˙ Window ( t Δ t ) · Δ t V Room · x CO 2 , ref x CO 2 , MV ( t Δ t ) + V ˙ AHU ( t Δ t ) · Δ t V Room · x CO 2 , ref x CO 2 , MV ( t Δ t )
The window volume flow rate V ˙ Window is extracted from the respective flow results. This leads to a different window volume flow rate used for each simplified ideal mixing ventilation calculation, as we calculate this for every scenario individually. The ventilation effectiveness is evaluated at t = 45 min when the lecture is over.
For ζ c = 1 , the window ventilation performs as well as a perfect mixing ventilation. In the case of ζ c > 1 , the ventilation behaviour approaches the so-called displacement or piston-flow behaviour. We give a more thorough discussion of this method in Ostmann et al. [36].

2.7.2. Aerosol Removal Effectiveness

During the duration of the investigated lesson, it is not possible to achieve a steady-state flow. Therefore, we introduce a slightly adapted definition of the contaminant removal effectiveness. The time-averaged room concentration is expressed by the ratio of the emitted aerosol mass m Aerosol and a mixing volume V mix (see Equation (30)).
x ¯ = m Aerosol V mix
Both the emitted aerosol mass and the mixing volume are time integrals up to the investigated point in time τ according to Equations (31) and (32).
m Aerosol = i = 1 28 + 1 0 τ m ˙ i , Aerosol d t
V mix = V Room + 0 τ V ˙ Window , in d t V mix , Window + 0 τ V ˙ AHU d t V mix , AHU + 0 τ η AP · V ˙ AP d t V mix , AP
We define the time-averaged extracted concentration in the same fashion by relating the extracted aerosol mass m e to the mixing volume V mix (see Equation (33)).
x ¯ e = m e , AHU / AP V mix , AHU / AP
The extracted aerosol mass is an integral until τ according to Equation (34) using the extract contaminant concentration and the volume flow rate respective to the ventilation configuration.
m e , AHU / AP = 0 τ V ˙ AHU / AP · x e d t
Using the definitions from Equations (30) and (33), we finally derive the aerosol removal effectiveness ε ¯ A c according to Equation (35).
ε ¯ A c = x ¯ e x ¯
The classification by Mundt [55] also applies to the aerosol removal effectiveness. If ε ¯ A c > 1 , either the operation of an air purifier or AHU effectively removes aerosol from the classroom, because across the investigated time frame the extracted aerosol concentration is higher than the average aerosol concentration inside the room.

2.7.3. Local Air Exchange Efficiency

To rate the air quality in the breathing zones of the occupants, we adopt the concept of the “age of air”, which was introduced by Sandberg [58]. Here, only a summary of the whole derivation is given. A detailed discussion can be found in Ostmann et al. [23]. In general, the local air exchange efficiency ε ¯ P , i a can be evaluated for every occupant i according to Equation (36) by relating the average room concentration (see Equation (30)) to the inhaled aerosol mass concentration emitted from all other occupants x ¯ P , i (see Equation (37)).
ε ¯ P , i a = x ¯ x ¯ P , i
The inhaled aerosol mass concentration x ¯ P , i is defined by the ratio of the total inhaled aerosol mass originating from other occupants m i , Aerosol , in (see Equation (38)) to the inhaled air volume of occupant V Breath , i (see Equation (39)).
x ¯ P , i = m i , Aerosol , in V Breath , i
When computing the imission of the aerosol, we only take the aerosol that is is emitted by the other occupants i j and transported to occupant i into account.
m i , Aerosol , in = i = 1 ; i j n 0 τ m ˙ i , j , Aerosol , in d t
V Breath , i = 0 τ V ˙ Breath , i d t
A simple classification of the local air exchange efficiency can be made as follows. For ε ¯ P , i a > 1 , the local concentration of aerosol emitted by other occupants is lower than the average room concentration, indicating an airflow pattern that is beneficial for the investigated occupant i. In the case of an ideally mixed room volume, the local air exchange efficiency converges towards ε ¯ P , i a = 1 . It is important to note that our definition yields different value ranges than the definition of the air change efficiency by Mundt [55].
Besides using the inhaled aerosol mass to calculate the local air exchange efficiency, it is also an indicator of the infection risk as it represents the potential dose of infectious aerosol that a person experiences. We therefore evaluate the inhaled aerosol mass during the lecture for each occupant individually for every scenario.

3. Results

The dilution of the contaminants (CO2 and individual aerosol) is mainly driven by the volume flow rate of supply or filtered air that is introduced into the classroom. While the volume flow rates of the mobile air purifiers or the AHU are constant during the lesson, the volume flow entering the classroom via window ventilation is subject to various influences. These influences are described in more detail in Ostmann et al. [36]. Table 9 summarises the average window volume flow rate for all scenarios that utilise window ventilation. In general, the ambient air temperature has the biggest impact on the volume flow rate through the windows. Between the different AP locations and operating points, the window volume flow rate varies by up to 8.4%.
For selected scenarios, Figure 10 shows the average CO2 concentration in the classroom. The grey area for 20–25 min indicates the period of window ventilation. The overall behaviour differs significantly between the scenarios utilising window ventilation and those with an AHU.
In the case of window ventilation, the CO2 levels rise above 1500 ppm before opening the windows. Depending on the scenario, the window ventilation for 5 min is able to reduce the CO2 levels to 591 ppm at 5 °C, 672 ppm at 15 °C and 753 ppm at 20 °C. If the additional air purifier is operated at location Atmosphere 15 00140 i001 with 1400 m 3 / h , the CO2 level after the window ventilation phase reaches 591 ppm for 5 °C, 692 ppm for 15 °C and 987 ppm for 20 °C, respectively. Over the whole lecture duration, the average CO2 level is in the range of 1047–1142 ppm for the investigated scenarios utilising window ventilation. Towards the end of the lesson, the levels continue to increase to nearly 2000 ppm . Equipping the room with an AHU and keeping the windows shut leads to an asymptotic increase in the CO2 levels. Although the stationary level is not reached during the lesson, the CO2 level is below 1250 ppm at the end of the lesson, with average levels during the lecture of 974–989 ppm .
The calculated values of the average aerosol concentration for all scenarios, as defined by Equation (30), are shown in Figure 11. While the ambient air temperature has a strong influence on the aerosol concentration (from 3.53 × 10 8   k g / m 3 at 5 °C to 5.26 × 10 8   k g / m 3 at 20 °C), its impact is weakened as soon as an AP is active. Operating an AP with 1400 m 3 / h reduces the average aerosol concentration by 62.6 % at 5 °C, 67.5 % at 15 °C and 70.5 % at 20 °C in the investigated scenarios. Compared to the scenarios with window ventilation at 5 °C where each air purifier Atmosphere 15 00140 i001 and Atmosphere 15 00140 i002 is operated with 1400 m 3 / h , a 50% reduction in the AP’s volume flow rate increases the aerosol concentration by 46.97% at location Atmosphere 15 00140 i001; and 47.01% at location Atmosphere 15 00140 i002. The scenario with the AHU achieves a slightly lower aerosol concentration of 3.06 × 10 8   k g / m 3 , which is a reduction of 13.3%, than the case where only window ventilation is utilised. By activating the air purifier Atmosphere 15 00140 i001; in addition to the AHU, the aerosol concentration is further reduced by 42.2%.
The values of the aerosol mass inhaled by the occupants are shown in Figure 12 in a boxplot diagram. The boxes represent the range of values among all occupants in the room. The median of the total mass of inhaled aerosol is 3.18 × 10 9   k g for 5 °C, 3.09 × 10 9   k g for 15 °C and 3.21 × 10 9   k g for 20 °C for the scenarios with only window ventilation. The scatter between the individual occupants is (0.91–1.16) × 10−9  k g , indicating the similar airflow behaviour of these scenarios.
Operating an AP with 1400 m 3 / h reduces the median to values of (1.96–2.11) × 10−9  k g . Although the median is within a small interval, the scatter between the individual occupants is highly dependent on the scenario. The minimum scatter is found for the scenario “WV + AP1 5 °C” with 0.25 × 10 9   k g . The maximum scatter of 2.32 × 10 9   k g is found in the scenario “ 1400 m 3 / h AP 3”.
Reducing the AP volume flow rate to 700 m 3 / h increases the median of the inhaled aerosol mass to (2.51–2.55) × 10−9  k g , with a scatter in the range of (0.95–1.15) × 10−9  k g .
The scenarios with an AHU achieve scatter intervals of (1.14–1.78) × 10−9  k g . Due to the additional filtered volume flow rate provided by the AP, scenario “AHU + AP1” achieves a lower median inhaled aerosol mass of 2.20 × 10 9   k g compared to scenario “AHU” with a median of 3.26 × 10 9   k g .
A more detailed distribution of the inhaled aerosol mass is shown in Figure A3, Figure A4, Figure A5, Figure A6 and Figure A7 for all investigated scenarios.

3.1. Ventilation Effectiveness

Figure 13 demonstrates the achieved values of the ventilation effectiveness at t = 45   min for all investigated scenarios. In all investigated scenarios, the ventilation effectiveness ζ c > 1 indicates airflow patterns that are more beneficial in terms of pollutant removal than ideal mixing ventilation.
In the case of window ventilation (“WV”), we evaluate the ventilation effectiveness to 1.065 for 5 °C, 1.100 for 15 °C and 1.134 for 20 °C, indicating a significant impact of the ambient air temperature. If the additional AP at location Atmosphere 15 00140 i001 is operated with window ventilation, the change in ventilation effectiveness is negligible for an ambient air temperature of 5 °C. The results at location Atmosphere 15 00140 i001 indicate no clear correlation between the ambient air temperature and ventilation effectiveness. Increasing the ambient air temperature from 5 °C to 15 °C leads to an increase in the ventilation effectiveness from 1.067 to 1.099. Nevertheless, an additional temperature increase to 20 °C leads to a decrease in the ventilation effectiveness to 1.060, which is below the value of 5 °C. We assume this to be due to the enhanced air mixing that is caused by the AP operation.
The AP operation leads to a higher ventilation effectiveness for both locations near the teacher (Atmosphere 15 00140 i001 and Atmosphere 15 00140 i002) compared to the ones in the back of the classroom (Atmosphere 15 00140 i003 and Atmosphere 15 00140 i004) (see Figure 6). At location Atmosphere 15 00140 i004; the AP impacts the natural stratification such that there is more mixing. Since the teacher produces less thermal buoyancy than the closely seated students, the additional mixing of the AP at either location Atmosphere 15 00140 i001 or Atmosphere 15 00140 i002 impacts the natural stratification less. Reducing the AP volume flow rate does not seem to affect the ventilation effectiveness significantly.
Regarding the ventilation effectiveness, the AHU outperforms nearly all scenarios at the end of the lesson where window ventilation was employed. The introduced constant volume flow with a slightly lower temperature of 20 °C compared to the average room temperature in the range of 21–23 °C and moderate momentum affects the natural stratification only locally. The AHU produces a big recirculation flow pattern that, once established, effectively captures the pollutant-contaminated air inside the classroom. The ventilation effectiveness increases when an AP operates at location Atmosphere 15 00140 i001 since the additional transport towards the ceiling supports the airflow pattern generated by the AHU. If the AHU was designed in a way that enables a displacement flow pattern, we assume this interaction to be even more prominent [55].

3.2. Aerosol Removal Effectiveness

Figure 14 presents the calculated aerosol removal effectiveness values at the end of the lesson for all scenarios where at least one AP or the AHU was active. Both the window volume flow rate (see Table 9) and the average aerosol concentration (see Figure 11) are closely related to the aerosol removal effectiveness. A more detailed discussion of the underlying aerosol transport effects can be found in Ostmann et al. [24].
Overall, all APs operated with 1400 m3/h achieve values between 0.937 and 1.025. The corresponding airflow patterns where AP 1 and AP 3 are active are shown in Figure A8 and Figure A9, respectively. In these figures, we extracted the flow field at t = 23   min (3 min after the windows were opened) and t = 45   min on two plane sections cutting through the active AP. A lower volume flow rate (700 m 3 / h ) leads to a significant reduction of 0.088–0.148 to ε ¯ A c 0.85 for both investigated scenarios. In contrast, the ambient air temperature impacts the aerosol removal effectiveness only to a small extent, which leads to a range of ε ¯ A c of 0.937–0.988. By varying the position, the removal effectiveness can be increased by up to 0.087, while an increase in the ambient air temperature leads to an increase of up to 0.050.
Equipping the classroom with an AHU can achieve an aerosol removal effectiveness of 1.045, which is an improvement over the best scenario with window ventilation where air purifier Atmosphere 15 00140 i003 was operated at the higher volume flow rate. Activating air purifier Atmosphere 15 00140 i001 parallel to the AHU increases the aerosol removal effectiveness of the AHU to 1.407 (indicated by Atmosphere 15 00140 i005; see Figure 14). However, the AP achieves a significantly lower removal effectiveness of 0.746 in the same scenario (indicated by Atmosphere 15 00140 i006; see Figure 14). For both scenarios utilising the AHU, we extracted the airflow patterns at t = 45   min on four different plane sections, where two planes are centre cuts through the classroom and the other two planes cut through the AHU and AP 1. The extracted airflow patterns are shown in Figure A10 and Figure A11, respectively.

3.3. Local Air Exchange Efficiency

The resulting air quality in the breathing zones expressed by the local air exchange efficiency is shown in Figure 15 for all investigated scenarios. The cases with only window ventilation do achieve the highest local air exchange efficiencies but also experience the highest average aerosol concentrations (see Figure 11). Furthermore, for these cases, the most aerosol mass is inhaled (see Figure 12), with a clear dependency on the ambient air temperature. The scenario with an ambient air temperature of 20 °C achieves a median local air exchange efficiency of approximately 2.25.
For the scenarios with an active AP, the local air exchange efficiency’s median is close to the ideal mixing assumption ( ε ¯ P a = 1 ). Depending on the AP location and volume flow rate, the scatter between the occupants increases. The AP at location Atmosphere 15 00140 i003 produces a very inhomogeneous local air quality, where students Atmosphere 15 00140 i007 and Atmosphere 15 00140 i008 experience high doses of the aerosol (see Figure A5b). The scenarios with the lower AP volume flow rate produce slightly higher median local air exchange efficiencies with significantly increased scatter between the occupants.
Equipping the classroom with an AHU increases the local air exchange efficiency but increases the inhaled aerosol mass as well for both scenarios. Although the operation of an additional AP slightly reduces the local air exchange efficiency, it significantly reduces the average aerosol concentration as well (see Figure 11) and, therefore, the inhaled aerosol mass. The occupants sitting near the extract air opening of the AHU experience higher aerosol concentrations. In consequence, they inhale more aerosol, as shown in Figure A7.

4. Discussion

Based on the transient observations presented so far, we can derive several correlations and interactions between various effects discussed in this section. Our results mainly confirm the findings of investigations conducted by other authors (see [14,15,16,17,18,21,22]), as we also show that an air purifier leads to a significant reduction in the aerosol concentration in the room air. We extend the existing knowledge by highlighting the additional impact of transient, intermittent window ventilation. Our results furthermore allow for a more robust design and operation of air purifiers in classroom environments. While the developed metrics give a first implication on how well a certain ventilation configuration performs under the given boundary conditions, the absolute values of the contaminant concentrations need to be considered.
The results of the ventilation effectiveness show how the ambient air temperature impacts the flow behaviour when utilising window ventilation only. A higher ambient air temperature tends to achieve a higher ventilation effectiveness due to the lower momentum of the ambient air entering the classroom. This leads to a lower mixing of the fresh air with the classroom air and more displacement-like ventilation behaviour. The AP activation weakens this effect because the AP introduces more mixing to the overall airflow. In the case of the low momentum airflow at a 20 °C ambient air temperature, we can see a clear reduction in the ventilation efficiency towards a value of 1, which shows that the AP changes the airflow pattern in the classroom more toward mixing ventilation.
The aerosol transport is described by both the aerosol removal effectiveness and local air exchange efficiency. The aerosol removal effectiveness (see Figure 14) indicates that the AP’s aerosol capture ability depends on the location and the volume flow rate. The AHU is investigated only in one setting without varying the position of the extract air opening. A different position might improve the associated aerosol removal effectiveness, but the AHU outperforms any AP under the investigated boundary conditions. The AHU extracts the air from the air layers near the ceiling with a high contaminant concentration due to natural convection. Since the AP transports more air toward the ceiling, the combination of the AHU and AP supports the airflow pattern that is generated by natural convection caused by thermal buoyancy.
While the natural convection is also responsible for the high local air exchange efficiencies for the “WV” scenarios, this only indicates a better air quality concerning the rest of the classroom, since the overall inhaled aerosol mass is higher by about 50% when compared to the other scenarios (see Figure 12) due to the lower flow rates.
Using an AP leads to comparable median values of the local air exchange efficiency within the scenario sets “WV + AP1” and “1400 m 3 / h ” and reduces the influence of the ambient air temperature. Since the enhanced mixing due to the AP is more dominant in the breathing zones than any natural convection flow, the median values are gathered slightly below ε ¯ P a < 1 . Across all scenarios with a combination of window ventilation and an AP, the main distinction between the scenarios is the extent of the scatter between the occupants. The AP in the back corner of the classroom opposite the windows (location Atmosphere 15 00140 i003) produces an airflow pattern that benefits some occupants (Atmosphere 15 00140 i009, Atmosphere 15 00140 i010, Atmosphere 15 00140 i011 and Atmosphere 15 00140 i012) but also drastically reduces the air quality for others (Atmosphere 15 00140 i007 and Atmosphere 15 00140 i008 in this particular scenario), as shown in Figure A5b. This is probably due to the local flow situation that is heavily influenced by the students’ desks and the high density of occupants in that area. Furthermore, the AP at location Atmosphere 15 00140 i003 is transporting the air in the opposite direction compared to the overall recirculation flow caused by natural convection. If an AP is operated at the other locations, the scatter is reduced, leading to a more consistent protection from aerosol for all occupants (see Figure A5).
As discussed in Ostmann et al. [24], all scenarios relying on window ventilation will cause significant thermal discomfort. This can only be remedied by utilising an AHU that can control the supply air conditions within certain bounds.
Our model underpredicts the turbulence inside the classroom since we assume all occupants to be simplified, stationary manikins. Therefore, they do not move, and the direction in which they are breathing is not changing. During window ventilation, we neglect the impact of wind or strong gusts, which could potentially influence the associated ventilation behaviour of window ventilation. We further neglect natural infiltration through the building envelope and solar radiation through the windows. Our aerosol is assumed to be small particles that perfectly follow the airflow, neglecting larger particles that may deposit on surfaces. As larger particles are transported up to several meters, we underpredict the related aerosol immission of close occupants.

5. Conclusions

In this research, we present a numerical flow model capable of predicting the immission of an aerosol by students in a typical classroom setting. The classroom is ventilated by a combination of natural or mechanical ventilation and a mobile purifier at different locations and volume flow rate settings.
We adapt the concept of the contaminant removal effectiveness and the age of air to express the aerosol transport using the aerosol removal effectiveness ε ¯ A c and the local air exchange efficiency ε ¯ P , i a . Both metrics indicate how effective a specific ventilation concept performs for the investigated boundary conditions. To estimate an infection risk, the most important value is the inhaled aerosol mass m i , Aerosol , in . This information can be derived only if, in addition to the local air exchange efficiency, the average aerosol concentration x ¯ is known.
We show that an air purifier is a viable solution to significantly reduce the aerosol mass to which a student is exposed. Depending on the air purifier location and volume flow rate, the reduction is more or less homogeneous across all students. If the AP is operated at location Atmosphere 15 00140 i003, a few students experience significantly higher amounts of the aerosol, while a decrease in the AP volume flow rate increases the overall level of inhaled aerosol mass by roughly 20%.
An air handling unit can locally reduce the inhaled aerosol mass. However, the students near the extract air position experience higher doses of the aerosol. By operating an air handling unit and a mobile air purifier at the same time, low amounts of inhaled aerosol can be achieved.
While mobile air purifiers can effectively reduce aerosol exposure, they are not able to enhance the indoor air quality further in regards to thermal comfort or the amount of “fresh” air (low CO2 levels). Only by equipping the classroom with a mechanical ventilation device with heat recovery is an energy-efficient conditioning of the indoor air environment possible.
To extend this study, additional scenarios could cover cases where not all windows are used for window ventilation. We also think it is feasible to investigate an AHU configuration where we achieve displacement ventilation or adjust the flow direction of the air purifier(s).

Author Contributions

Conceptualisation, P.O., D.D., K.R., M.K. and D.M.; methodology, P.O., D.D., K.R. and D.M.; software, P.O. and D.D.; validation, P.O., D.D., K.R., M.K. and D.M.; formal analysis, P.O. and D.D.; investigation, P.O. and D.D.; resources, D.M.; data curation, P.O. and D.D.; writing—original draft preparation, P.O. and D.D.; writing—review and editing, P.O., D.D., K.R., M.K. and D.M.; visualisation, P.O.; supervision, K.R., M.K. and D.M.; project administration, K.R., M.K. and D.M.; funding acquisition, D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Heinz Trox Wissenschafts gGmbH (Grant number HTx0020). Simulations were performed with computing resources granted by RWTH Aachen University under projects rwth0703 and eoncov.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to the large size of the associated files.

Conflicts of Interest

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

Abbreviations

The following abbreviations are used in this manuscript:
AeroAerodynamic
AHUAir handling unit
APAir purifier
CFDComputational fluid dynamics
corrCorrected
elElectrical
ETAExtract air
GCI Grid Convergence Index
refReference
SUPSupply air
totTotal

Appendix A. Results of the Grid Convergence Study

Figure A1. Results of the grid convergence study of the average classroom temperature for the three investigated computational grid resolutions.
Figure A1. Results of the grid convergence study of the average classroom temperature for the three investigated computational grid resolutions.
Atmosphere 15 00140 g0a1
Figure A2. Results of the grid convergence study of the captured aerosol concentration for the three investigated computational grid resolutions.
Figure A2. Results of the grid convergence study of the captured aerosol concentration for the three investigated computational grid resolutions.
Atmosphere 15 00140 g0a2

Appendix B. Summary of the Developed Metrics

Table A1. Overview of the developed metrics and intermediate values with their scope and dependent variables.
Table A1. Overview of the developed metrics and intermediate values with their scope and dependent variables.
Metric/VariablePollutantDefinitionDescription
Average room concentrationAerosol x ¯ = m Aerosol V mix Time-averaged room concentration, using total emitted aerosol mass and delivered mixing air volume provided by window or mechanical ventilation
Average extracted concentrationAerosol x ¯ e = m e , AHU / AP V mix , AHU / AP Time-averaged extracted concentration, using the device-specific extracted aerosol mass and delivered mixing air volume
Average inhaled concentrationAerosol x ¯ P , i = m i , Aerosol , in V Breath , i Time-averaged inhaled concentration, using the occupant-specific inhaled aerosol mass and inhaled breath volume
Ventilation effectivenessCO2 ζ c ( t ) = x CO 2 , MV ( t ) x CO 2 ( t ) Rates the performance of the ventilation behaviour by using idealised mixing ventilation as a benchmark
ζ < 1 : short circuit
ζ = 1 : mixing ventilation
ζ > 1 : displacement ventilation
Aerosol removal effectivenessAerosol ε ¯ A c = x ¯ e x ¯ Describes the ability of the air cleaner and/or AHU to remove aerosol from the classroom
ε ¯ A c < 1 : device position suboptimal
ε ¯ A c > 1 : devices are well positioned
Local air exchange efficiencyAerosol ε ¯ P , i a = x ¯ x ¯ P , i Rates the local air quality in relation to the average room air quality
ε ¯ P , i a < 1 : bad relative air quality
ε ¯ P , i a > 1 : good relative air quality

Appendix C. Distributions of Inhaled Aerosol Mass by Individual Occupants

Appendix C.1. Comparison of Different Mobile Air Purifier Locations

Figure A3. Inhaled aerosol mass for all individuals for the “WV” scenarios at t = 45 min .
Figure A3. Inhaled aerosol mass for all individuals for the “WV” scenarios at t = 45 min .
Atmosphere 15 00140 g0a3
Figure A4. Inhaled aerosol mass for all individuals for the “WV + AP1” scenarios at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure A4. Inhaled aerosol mass for all individuals for the “WV + AP1” scenarios at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g0a4
Figure A5. Inhaled aerosol mass for all individuals for the “1400 m 3 / h ” scenarios at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure A5. Inhaled aerosol mass for all individuals for the “1400 m 3 / h ” scenarios at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g0a5
Figure A6. Inhaled aerosol mass for all individuals for the “700 m 3 / h ” scenarios at t = 45 min .
Figure A6. Inhaled aerosol mass for all individuals for the “700 m 3 / h ” scenarios at t = 45 min .
Atmosphere 15 00140 g0a6

Appendix C.2. Influence of an Additional Mobile Air Purifier When Equipping the Classroom with an AHU

Figure A7. Inhaled aerosol mass for all individuals for the scenarios with AHU at t = 45 min .
Figure A7. Inhaled aerosol mass for all individuals for the scenarios with AHU at t = 45 min .
Atmosphere 15 00140 g0a7

Appendix D. Flow Fields Inside the Classroom

Scenarios with a Combination of Air Purifier and Window Ventilation

Figure A8. Flow field on two plane sections where AP 1 is active with 1400 m 3 / h at different points in time.
Figure A8. Flow field on two plane sections where AP 1 is active with 1400 m 3 / h at different points in time.
Atmosphere 15 00140 g0a8
Figure A9. Flow field on two plane sections where AP 3 is active with 1400 m 3 / h at different points in time.
Figure A9. Flow field on two plane sections where AP 3 is active with 1400 m 3 / h at different points in time.
Atmosphere 15 00140 g0a9
Figure A10. Flow field on four plane sections cutting through the centre of the classroom, the AHU and AP 1, where only the AHU is operated with 700 m 3 / h at t = 45 min .
Figure A10. Flow field on four plane sections cutting through the centre of the classroom, the AHU and AP 1, where only the AHU is operated with 700 m 3 / h at t = 45 min .
Atmosphere 15 00140 g0a10
Figure A11. Flow field on four plane sections cutting through the centre of the classroom, the AHU and AP 1, where the AHU and AP 1 are operated with 700 m 3 / h each at t = 45 min .
Figure A11. Flow field on four plane sections cutting through the centre of the classroom, the AHU and AP 1, where the AHU and AP 1 are operated with 700 m 3 / h each at t = 45 min .
Atmosphere 15 00140 g0a11

References

  1. Lu, J.; Gu, J.; Li, K.; Xu, C.; Su, W.; Lai, Z.; Zhou, D.; Yu, C.; Xu, B.; Yang, Z. COVID-19 Outbreak Associated with Air Conditioning in Restaurant, Guangzhou, China, 2020. Emerg. Infect. Dis. 2020, 26, 1628. [Google Scholar] [CrossRef] [PubMed]
  2. Lam-Hine, T.; McCurde, S.A.; Santora, L.; Duncan, L.; Corbett-Detig, R.; Kapusinszky, B.; Willis, M. Outbreak Associated with SARS-CoV-2 B.1.617.2 (Delta) Variant in an Elementary School—Marin County, California, May–June 2021. Morb. Mortal. Wkly. Rep. (MMWR) 2021, 70, 1214–1219. [Google Scholar] [CrossRef] [PubMed]
  3. Buonanno, G.; Morawska, L.; Stabile, L. Quantitative assessment of the risk of airborne transmission of SARS-CoV-2 infection: Prospective and retrospective applications. Environ. Int. 2020, 145, 106112. [Google Scholar] [CrossRef] [PubMed]
  4. Müller, D.; Rewitz, K.; Derwein, D.; Burgholz, T.M.; Schweiker, M.; Bardey, J.; Tappler, P. Estimation of the Risk of Infection by Aerosol-Borne Viruses in Ventilated Rooms: White Paper, RWTH-EBC 2021-002; Technical Report; RWTH Aachen University: Aachen, Germany, 2021. [Google Scholar] [CrossRef]
  5. Aganovic, A.; Cao, G.; Kurnitski, J.; Wargocki, P. New dose-response model and SARS-CoV-2 quanta emission rates for calculating the long-range airborne infection risk. Build. Environ. 2023, 228, 109924. [Google Scholar] [CrossRef] [PubMed]
  6. Kriegel, M.; Buchholz, U.; Gastmeier, P.; Bischoff, P.; Abdelgawad, I.; Hartmann, A. Predicted Infection Risk for Aerosol Transmission of SARS-CoV-2; Technical University of Berlin: Berlin, Germany, 2020. [Google Scholar] [CrossRef]
  7. Lelieveld, J.; Helleis, F.; Borrmann, S.; Cheng, Y.; Drewnick, F.; Haug, G.; Klimach, T.; Sciare, J.; Su, H.; Pöschl, U. Model Calculations of Aerosol Transmission and Infection Risk of COVID-19 in Indoor Environments. Int. J. Environ. Res. Public Health 2020, 17, 8114. [Google Scholar] [CrossRef] [PubMed]
  8. Dinklage, F.; Ehmann, A.; Erdmann, E.; Klack, M.; Mast, M.; Stahnke, J. So schnell verbreitet sich das Coronavirus in Innenräumen. Zeit Online 2020, 26, 2020. [Google Scholar]
  9. Max Planck Institute for Chemistry. COVID 19 Aerosol Transmission Risk Calculator; Max Planck Institute for Chemistry: Mainz, Germany, 2020. [Google Scholar]
  10. German Federal Environment Agency. Richtig Lüften in Schulen; German Federal Environment Agency: Dessau-Roßlau, Germany, 2021.
  11. CDC. Ventilation in Buildings; CDC: Atlanta, GA, USA, 2023.
  12. CCDC. Guideline for Operation Management and Use of Air Conditioners in Summer; CCDC: Cambridge, UK, 2020. [Google Scholar]
  13. Lowther, S.D.; Deng, W.; Fang, Z.; Booker, D.; Whyatt, D.J.; Wild, O.; Wang, X.; Jones, K.C. How efficiently can HEPA purifiers remove priority fine and ultrafine particles from indoor air? Environ. Int. 2020, 144, 106001. [Google Scholar] [CrossRef]
  14. Curtius, J.; Granzin, M.; Schrod, J. Testing mobile air purifiers in a school classroom: Reducing the airborne transmission risk for SARS-CoV-2. Aerosol Sci. Technol. 2021, 55, 586–599. [Google Scholar] [CrossRef]
  15. Jain, A.; Duill, F.F.; Schulz, F.; Beyrau, F.; van Wachem, B. Numerical Study on the Impact of Large Air Purifiers, Physical Distancing, and Mask Wearing in Classrooms. Atmosphere 2023, 14, 716. [Google Scholar] [CrossRef]
  16. Sabanskis, A.; Vidulejs, D.D.; Teličko, J.; Virbulis, J.; Jakovičs, A. Numerical Evaluation of the Efficiency of an Indoor Air Cleaner under Different Heating Conditions. Atmosphere 2023, 14, 1706. [Google Scholar] [CrossRef]
  17. Duill, F.F.; Schulz, F.; Jain, A.; Krieger, L.; van Wachem, B.; Beyrau, F. The Impact of Large Mobile Air Purifiers on Aerosol Concentration in Classrooms and the Reduction of Airborne Transmission of SARS-CoV-2. Int. J. Environ. Res. Public Health 2021, 18, 11523. [Google Scholar] [CrossRef] [PubMed]
  18. Fierce, L.; Robey, A.J.; Hamilton, C. High efficacy of layered controls for reducing exposure to airborne pathogens. Indoor Air 2022, 32, e12989. [Google Scholar] [CrossRef] [PubMed]
  19. Chirico, F.; Sacco, A.; Bragazzi, N.L.; Magnavita, N. Can Air-Conditioning Systems Contribute to the Spread of SARS/MERS/COVID-19 Infection? Insights from a Rapid Review of the Literature. Int. J. Environ. Res. Public Health 2020, 17, 6052. [Google Scholar] [CrossRef] [PubMed]
  20. Elsaid, A.M.; Ahmed, M.S. Indoor Air Quality Strategies for Air-Conditioning and Ventilation Systems with the Spread of the Global Coronavirus (COVID-19) Epidemic: Improvements and Recommendations. Environ. Res. 2021, 199, 111314. [Google Scholar] [CrossRef] [PubMed]
  21. Melikov, A. Mitigation of airborne transmission of respiratory viruses by ventilation—Past, present and future. In Proceedings of the Ventilation, IEQ and Health in Sustainable Buildings, Copenhagen, Denmark, 4–5 October 2023; pp. 33–34. [Google Scholar]
  22. Kurnitski, J.; Kiil, M.; Mikola, A.; Võsa, K.V. Point source ventilation effectiveness in infection risk-based post-COVID ventilation design. In Proceedings of the Ventilation, IEQ and Health in Sustainable Buildings, Copenhagen, Denmark, 4–5 October 2023; pp. 35–44. [Google Scholar]
  23. Ostmann, P.; Derwein, D.; Kremer, M.; Rewitz, K.; Müller, D. Wirksamkeit der Einseitigen Fensterlüftung in Klassenräumen—Stoßlüfung: 2nd, Reviewed and Updated Edition; White Paper, RWTH-EBC 2023-003; RWTH Aachen University: Aachen, Germany, 2023. [Google Scholar] [CrossRef]
  24. Ostmann, P.; Derwein, D.; Kremer, M.; Rewitz, K.; Müller, D. Rating the Effectiveness of Air Purifiers on Aerosol Removal in a Classroom with Numerical Flow Simulations (Hg.) 2023—Healthy Buildings 2023 Europe. In Proceedings of the Healthy Buildings 2023 Europe, Aachen, Germany, 11–14 June 2023; pp. 654–662. [Google Scholar]
  25. ANSYS Inc. CFX-Solver Modeling Guide 2020 R2; ANSYS Inc.: Canonsburg, PA, USA, 2020. [Google Scholar]
  26. Burgholz, T.M.; Müller, D. CO2 levels in differently ventilated classrooms with regard to occupational ventilation behavior under spring and summer conditions. In Proceedings of the 16th Conference of the International Society of Indoor Air Quality & Climate, Indoor Air, Virtual, 20–24 July 2020. [Google Scholar]
  27. Menchaca-Brandan, M.A.; Glicksman, L.R. How To Replicate Computationally a Human Plume with a Simple Geometry, and How Not To (Hg.) 2011—Roomvent 2011. In Proceedings of the Roomvent 2011, Trondheim, Norway, 19–21 June 2011. [Google Scholar]
  28. TROX GmbH. Produktdatenblatt Luftreiniger TAP; TROX GmbH: Neukirchen-Vluyn, Germany, 2020. [Google Scholar]
  29. Ostmann, P.; Derwein, D.; Kremer, M.; Rewitz, K.; Müller, D. Bewertung der Lüftungswirksamkeit eines Luftreinigers Hinsichtlich Aerosolausbreitung; White Paper, RWTH-EBC 2023-004; RWTH Aachen University: Aachen, Germany, 2023. [Google Scholar] [CrossRef]
  30. Celik, I.B.; Urmila, G.; Roache, P.J.; Freitas, C.J.; Coleman, H.W.; Raad, P.E. Procedure for Estimation and Reporting of Uncertainty Due to Discretization in CFD Applications. J. Fluids Eng. 2008, 130, 078001. [Google Scholar] [CrossRef]
  31. Kriegel, M. Experimentelle Untersuchung und Numerische Simulation eines Quellluftsystems. Ph.D. Thesis, Technische Universität Berlin, Berlin, Germany, 2005. [Google Scholar]
  32. Streblow, R. Thermal Sensation and Comfort Model for Inhomogeneous Indoor Environments. Ph.D. Thesis, RWTH Aachen University, Aachen, Germany, 2010. [Google Scholar]
  33. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [Google Scholar] [CrossRef]
  34. Shah, N.G. The Computation of Radiation Heat Transfer. Ph.D. Thesis, University of London, London, UK, 1979. [Google Scholar]
  35. Boussinesq, J. Théorie de l’écoulement Tourbillonnant et Tumultueux des Liquides; Gauthier-Villars et fils: Paris, France, 1879. [Google Scholar]
  36. Ostmann, P.; Derwein, D.; Rewitz, K.; Kremer, M.; Müller, D. Entwicklung eines Numerischen Strömungssimulationsmodells zur Berechnung der Aerosolausbreitung in Klassenräumen; White Paper, RWTH-EBC 2023-002; RWTH Aachen University: Aachen, Germany, 2023. [Google Scholar] [CrossRef]
  37. World Health Organization. Infection Prevention and Control of Epidemic- and Pandemic-Prone Acute Respiratory Infections in Health Care Health Care: WHO Guidelines; Pandemic and Epidemic Diseases; World Health Organization: Geneva, Switzerland, 2014. [Google Scholar]
  38. Bivolarova, M.; Ondráček, J.; Melikov, A.; Ždímal, V. A comparison between tracer gas and aerosol particles distribution indoors: The impact of ventilation rate, interaction of airflows, and presence of objects. Indoor Air 2017, 27, 1201–1212. [Google Scholar] [CrossRef]
  39. Rim, D.; Novoselac, A. Transport of particulate and gaseous pollutants in the vicinity of a human body. Build. Environ. 2009, 44, 1840–1849. [Google Scholar] [CrossRef]
  40. Regenscheit, B. Luftbewegung in klimatisierten Räumen. Klimatechnik 1959, 11, 3–8. [Google Scholar]
  41. Adams, W.C. Measurement of Breathing Rate and Volume in Routinely Performed Daily Activities: Final Report; Technical Report; Human Performance Laboratory Physical Education Department: Davis, CA, USA, 1993. [Google Scholar]
  42. Morawska, L.; Johnson, G.R.; Ristovski, Z.D.; Hargreaves, M.; Mengersen, K.; Corbett, S.; Chao, C.; Li, Y.; Katoshevski, D. Size distribution and sites of origin of droplets expelled from the human respiratory tract during expiratory activities. J. Aerosol Sci. 2009, 40, 256–269. [Google Scholar] [CrossRef]
  43. Buonanno, G.; Stabile, L.; Morawska, L. Estimation of airborne viral emission: Quanta emission rate of SARS-CoV-2 for infection risk assessment. Environ. Int. 2020, 141, 105794. [Google Scholar] [CrossRef] [PubMed]
  44. FAO; WHO; UNU. Human Energy Requirements: Report of a Joint FAO/WHO/UNU Expert Consultation; Technical Report; FAO: Rome, Italy, 2001. [Google Scholar]
  45. Schofield, W.N. Predicting basal metabolic rate, new standards and review of previous work. Hum. Nutr. Clin. Nutr. 1985, 39 (Suppl. 1), 5–41. [Google Scholar] [PubMed]
  46. Pharmacia GmbH. Wachstums- und Gewichtskurven in Perzentilen—Jungen; Pharmacia GmbH: Berlin, Germany, 2009. [Google Scholar]
  47. Pharmacia GmbH. Wachstums- und Gewichtskurven in Perzentilen—Mädchen; Pharmacia GmbH: Berlin, Germany, 2009. [Google Scholar]
  48. Reinken, L.; Stolley, H.; Droese, W.; van Oost, G. Longitudinale Körperentwicklung gesunder Kinder. II. Größe, Gewicht, Hautfettfalten von Kindern im Alter von 1,5 bis 16 Jahren. Klin. Pädiatr. 1980, 192, 25–33. [Google Scholar] [CrossRef] [PubMed]
  49. Brandt, I. Perzentilenkurven für das längenwachstum bei Früh-nd Reifgeborenen in den ersten 5 lebensjahren. Der Kinderarzt 1980, 11, 43–51. [Google Scholar]
  50. Brandt, I. Human Growth: A Comprehensive Treatise, 2nd ed.; Springer: New York, NY, USA, 1986; Volume 1. [Google Scholar] [CrossRef]
  51. Brandt, I.; Reinken, L. Die Wachstumsgeschwindigkeit gesunder Kinder in den ersten 16 Lebensjahren: Longitudinale Entwicklungsstudie Bonn-Dortmund. Klin. Pädiatr. 1988, 200, 451–456. [Google Scholar] [CrossRef] [PubMed]
  52. Reinken, L.; van Oost, G. Longitudinale Körperentwicklung gesunder Kinder von 0 bis 18 Jahren. Körperlänge/-höhe, Körpergewicht und Wachstumsgeschwindigkeit. Klin. Pädiatr. 1992, 204, 129–133. [Google Scholar] [CrossRef] [PubMed]
  53. Gao, S.; Yang, L.; Li, Y.; Liu, S.; Zhang, H.; Arens, E.; Zhai, Y. Metabolic rate in children and adolescents: Tabulate values for common activities and comparisons with standards and adult values. Build. Environ. 2023, 244, 110804. [Google Scholar] [CrossRef]
  54. Du Bois, D.; Du Bois, E.F. Clinical Calorimetry: Tenth Paper—A Formula to Estimate the Approximate Surface Area if Height and Weight Be Known. Arch. Intern. Med. 1916, XVII, 863–871. [Google Scholar] [CrossRef]
  55. Mundt, E. Ventilation Effectiveness; REHVA Guidebook; Rehva: Brussels, Belgium, 2004; Volume 2. [Google Scholar]
  56. Brouns, C.; Waters, B. A Guide to Contaminant Removal Effectiveness; Technical Report AIC-TN-28.2-1991; Air Infiltration and Ventilation Centre: Coventry, UK, 1991. [Google Scholar]
  57. Wildeboer, J.; Müller, D. Lüftungseffektivität als Qualitätskriterium für Quellluftsysteme—Die Auswirkung des Luftaustauschwirkungsgrads auf die Höhe der Frischluftschicht. KI—Kälte Luft Klimatech. 2006, 42, 439–443. [Google Scholar]
  58. Sandberg, M. What is ventilation efficiency? Build. Environ. 1981, 16, 123–135. [Google Scholar] [CrossRef]
Figure 1. Overview of the investigated classroom models.
Figure 1. Overview of the investigated classroom models.
Atmosphere 15 00140 g001
Figure 2. Section of the computational grid along the classroom width.
Figure 2. Section of the computational grid along the classroom width.
Atmosphere 15 00140 g002
Figure 3. Resolution of the boundary layer around the head of a manikin.
Figure 3. Resolution of the boundary layer around the head of a manikin.
Atmosphere 15 00140 g003
Figure 4. Overview of the flow model with the attached outdoor environment.
Figure 4. Overview of the flow model with the attached outdoor environment.
Atmosphere 15 00140 g004
Figure 5. Volume flow rate during the occupant breathing cycle.
Figure 5. Volume flow rate during the occupant breathing cycle.
Atmosphere 15 00140 g005
Figure 6. Seating arrangement of the occupants and positions of the air purifiers inside the classroom.
Figure 6. Seating arrangement of the occupants and positions of the air purifiers inside the classroom.
Atmosphere 15 00140 g006
Figure 7. Different ranges of the contaminant removal effectiveness for a room equipped with an air handling unit according to Mundt [55]: (a) contaminant source near extract opening; (b) contaminant source within a recirculation zone; (c) fully mixed flow pattern (mixing ventilation).
Figure 7. Different ranges of the contaminant removal effectiveness for a room equipped with an air handling unit according to Mundt [55]: (a) contaminant source near extract opening; (b) contaminant source within a recirculation zone; (c) fully mixed flow pattern (mixing ventilation).
Atmosphere 15 00140 g007
Figure 8. Different ranges of the contaminant removal effectiveness for a room equipped with a mobile air purifier in the style of Mundt [55]: (a) contaminant source near the air purifier; (b) contaminant source within the recirculation zone produced by the air purifier.
Figure 8. Different ranges of the contaminant removal effectiveness for a room equipped with a mobile air purifier in the style of Mundt [55]: (a) contaminant source near the air purifier; (b) contaminant source within the recirculation zone produced by the air purifier.
Atmosphere 15 00140 g008
Figure 9. Classification of the contaminant removal effectiveness and the air change efficiency in the style of Wildeboer and Müller [57].
Figure 9. Classification of the contaminant removal effectiveness and the air change efficiency in the style of Wildeboer and Müller [57].
Atmosphere 15 00140 g009
Figure 10. Average CO2 concentration in the classroom during the lesson for selected scenarios.
Figure 10. Average CO2 concentration in the classroom during the lesson for selected scenarios.
Atmosphere 15 00140 g010
Figure 11. Averaged aerosol concentration x ¯ evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure 11. Averaged aerosol concentration x ¯ evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g011
Figure 12. Inhaled mass of aerosol m i , Aerosol , in across all occupants for all investigated scenarios evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure 12. Inhaled mass of aerosol m i , Aerosol , in across all occupants for all investigated scenarios evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g012
Figure 13. Ventilation effectiveness ζ c evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure 13. Ventilation effectiveness ζ c evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g013
Figure 14. Aerosol removal effectiveness ε ¯ A c evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure 14. Aerosol removal effectiveness ε ¯ A c evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g014
Figure 15. Local air exchange efficiency ε ¯ P , i a evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Figure 15. Local air exchange efficiency ε ¯ P , i a evaluated at t = 45 min (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Atmosphere 15 00140 g015
Table 1. Geometric dimensions of the investigated classroom.
Table 1. Geometric dimensions of the investigated classroom.
QuantityValueUnitDescription
b Room 7.35 m Width of the classroom
h Room 3.00 m Height of the classroom
l Room 9.52 m Length of the classroom
V Room , geom ≈210 m 3 Geometric volume of the classroom
V Room 199.92 m 3 Effective air volume of the classroom
b Window 1.86 m Width of the windows
h Window 1.48 m Height of the windows
A Window 2.75 m 2 Area of a window
n Window 4-No. of windows
A AP 4.70 m 2 Surface area of an air cleaner
Table 2. Parameters that describe the mesh refinements.
Table 2. Parameters that describe the mesh refinements.
Surface/Volume Δ x Refinement r Influence
Manikins 2.0 × 10 2   m -
Mouth areas 2.5 × 10 3   m 1 × 10 2   m
Ceiling lights 5.0 × 10 2   m 1 × 10 1   m
Air purifiers 5.0 × 10 2   m 4 × 10 1   m
Radiators 5.0 × 10 2   m -
Table 3. Parameters of the generated prism layers on the respective surfaces.
Table 3. Parameters of the generated prism layers on the respective surfaces.
Surface h 1 n Layers f Growth
Manikins 2 × 10 3   m 71.2
Windows 4 × 10 3   m 51.2
Air purifiers 4 × 10 3   m 51.2
Table 4. Models used in the flow simulation.
Table 4. Models used in the flow simulation.
ModelSelected
Turbulencek- ω -SST [33]
RadiationDiscrete transfer radiation [34] with 8 rays (surface to surface)
FluidAir as a single-component gas with Boussinesq assumption [35] and gravity
EnergyThermal energy
TimeTransient
Table 5. Parameters that describe the species transport.
Table 5. Parameters that describe the species transport.
QuantityValueUnitDescription
f Breath 10 1 / min Breathing frequency
T Breath 36.5 °C Temperature of the exhaled air
V ˙ Breath 0.54 m 3 / h Breathing flow rate during moderate activity when seated [3]
c Aerosol , S 1.8 × 10−3 m L / m 3 Exhaled aerosol concentration of students [3,41,42]
c Aerosol , T 9.6 × 10−3 m L / m 3 Exhaled aerosol concentration of the teacher [3,41,42]
ρ Aerosol 1000 k g / m 3 Density of the aerosol
c CO 2 , ref 410 ppm CO2 concentration of the ambient air
c CO 2 , i 40 × 103 m L / m 3 exhaled CO2 concentration
ρ CO 2 1.98 k g / m 3 Density of CO2
η AP 0.95-Filter efficiency of the air purifiers
Table 6. Parameters that describe the occupants’ heat emissions.
Table 6. Parameters that describe the occupants’ heat emissions.
QuantityStudentTeacher
Weight m i 67 k g 84 k g
Height h i 175 cm180 cm
Activity k Activity 1.11.6
Breathing temperature T Breath 36.5   C 36.5   C
Table 7. Heat transfer coefficients that describe the heat transfer through the room facade.
Table 7. Heat transfer coefficients that describe the heat transfer through the room facade.
AreaHeat Transfer Coefficient k HTC , i
Window frames4 W /( m 2   K )
Windows6 W /( m 2   K )
Window sills1 W /( m 2   K )
Table 8. Overview of the investigated scenarios (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Table 8. Overview of the investigated scenarios (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
ParameterScenario Sets
WV WV + AP 1400 m3/h700 m3/h AHU
T amb /°C515205152055555555
V ˙ AP , 1 / ( m 3 / h ) ---1400140014001400---700--700
V ˙ AP , 2 -------1400---700--
V ˙ AP , 3 --------1400-----
V ˙ AP , 4 ---------1400----
V ˙ AHU / ( m 3 / h ) ------------700700
Table 9. Averaged volume flows entering the classroom through the windows for all scenarios utilising window ventilation. (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
Table 9. Averaged volume flows entering the classroom through the windows for all scenarios utilising window ventilation. (“WV + AP1 5 °C” is the same as “1400 m 3 / h AP1”).
WVWV + AP11400 m3/h700 m3/h
[m3/h] 5 °C 15 °C 20 °C 5 °C 15 °C 20 °C AP1 AP2 AP3 AP4 AP1 AP2
V ˙ ¯ Window , in 573.6403.5297.8571.4388.9242.8571.4534.0578.9546.1562.2536.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ostmann, P.; Derwein, D.; Rewitz, K.; Kremer, M.; Müller, D. Computational Fluid Dynamics-Based Calculation of Aerosol Transport in a Classroom with Window Ventilation, Mechanical Ventilation and Mobile Air Purifier. Atmosphere 2024, 15, 140. https://doi.org/10.3390/atmos15010140

AMA Style

Ostmann P, Derwein D, Rewitz K, Kremer M, Müller D. Computational Fluid Dynamics-Based Calculation of Aerosol Transport in a Classroom with Window Ventilation, Mechanical Ventilation and Mobile Air Purifier. Atmosphere. 2024; 15(1):140. https://doi.org/10.3390/atmos15010140

Chicago/Turabian Style

Ostmann, Philipp, Dennis Derwein, Kai Rewitz, Martin Kremer, and Dirk Müller. 2024. "Computational Fluid Dynamics-Based Calculation of Aerosol Transport in a Classroom with Window Ventilation, Mechanical Ventilation and Mobile Air Purifier" Atmosphere 15, no. 1: 140. https://doi.org/10.3390/atmos15010140

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop