The pioneering work of Taylor on the turbulent dispersion of aerosols is one century old and provides an interesting way to introduce both diffusive processes and turbulence at an undergraduate level. Low mass particles transported by a turbulent flow exhibit a Brownian-like motion over time scales larger than the velocity correlation time. Aerosols and gases are, therefore, subjected to an effective turbulent diffusion at large length scales. However, the case of a source of pollutant much smaller than the integral scale is not completely understood. Here, we present experimental results obtained by undergraduate students in the context of the COVID-19 pandemic. The dispersion of a fog of oil droplets by a turbulent flow is studied in a wind tunnel designed for pedagogical purposes. It shows a ballistic-like regime at short distance, followed by Taylor's diffusive-like regime, suggesting that scale-free diffusion by the turbulent cascade process is bypassed. Measurements show that the dispersion of CO2 emitted when breathing in a natural, indoor air flow is not isotropic but rather along the flow axis. The transverse spread is ballistic-like, leading to the concentration decaying as the inverse-squared distance to the mouth. The experiment helps students understand the role of fluctuations in diffusive processes and in turbulence. A Langevin equation governing aerosol dispersion based on a single correlation time allows us to model the airborne transmission risk of pathogens, indoors and outdoors. The results obtained in this study have been used to provide public health policy recommendations to prevent transmission in shopping malls.

The challenge of offering experimental physics courses online during the COVID-19 pandemic motivated the following question: can students learn experimental techniques during lockdowns while simultaneously working to reduce viral transmission risk indoors? A group of students studying experimental physics at our university worked on several projects to reduce the risk of airborne transmission of SARS-CoV-2, including the filtration efficiency of face masks, the calibration of CO2 sensors using candles or dry ice in a closed transparent box, and the analysis of epidemic data. We report here the work on a project to understand the dispersion of viral particles both indoors and outdoors. This project has required a detailed understanding of both diffusion processes and turbulent flows.

However, turbulence is not an easy subject to teach. At the undergraduate level, topics in fluid mechanics mainly include hydrostatics, potential laminar flows, and viscous flows. The concept of turbulence appears when teaching the drag force exerted on a moving spherical solid as a function of the Reynolds number (see, e.g., Ref. 1). In this approach, one of the fundamental aspects of turbulence, namely, space and time fluctuations, is bypassed. Here, we show that turbulence can be taught by considering the dispersion of particles in a regime where concentration is low enough that it acts as a passive scalar field; that is, the concentration has no dynamical effect on the flow. We treat this dispersion theoretically in Sec. II and also present very simple experiments. In Sec. III, we make use of controlled experiments performed both in a small wind tunnel used for pedagogical purposes, using a fog of oil droplets, and in the large corridors of two shopping malls under various ventilation conditions using CO2. We show that the concentration decays faster than expected by the turbulent dispersion theory of Taylor,2 following instead a ballistic law. Indeed, introducing a single velocity decorrelation time provides a perfect fit to the data. This suggests, as shown previously by Villermaux et al.,3,4 that concentration fluctuations bypass the turbulent cascade: diffusion does not follow a continuous process starting from the injection scale and down to the dissipation scale.5 The Kolmogorov scaling law is not observed. The work presented here can be used to introduce students to the turbulent dispersion not only of pollutant molecules but also of aerosols. We finally discuss the relevance of the results to the airborne transmission of SARS-CoV-2, which has motivated this research. SARS-CoV-2 is mainly transmitted through the airborne route:6 infection happens when a susceptible person inhales viral particles emitted by another infected person. In between their respective respiratory tracts, the air carrying viral particles is gradually diluted by turbulent dispersion. Near an emitter, the concentration of viral particles is, therefore, higher than far away, which has implications on the airborne transmission risk.

We first consider the one-dimensional diffusive transport of a particle of diameter d, mass m, and density ρp in a gas of viscosity η without turbulence. The force exerted by the gas on the particle can be decomposed into two parts: the average hydrodynamic Stokes1 force 3 π η d v and a noise7  Γ ( t ), which describes the thermal fluctuations around this average. The particle dynamics is described by the Langevin equation
(1)
where the damping term is the Stokes force, rewritten after introducing the Stokes time τS, which is the particle velocity response time to a force. Γ ( t ) is only known statistically as it results from the random collisions of gas molecules with the particle, which are uncorrelated. Therefore, we define the ensemble average · ¯ as the average over many realizations of the motion with different values of the noise. The noise is Gaussian, has zero mean, and is δ-correlated in time7 
(2)
where δ stands for the Dirac delta function.
Equation (1) is a linear relaxation equation: the homogeneous solution decays to zero at times much longer than the relaxation time τS. Particle inertia, which drives the homogeneous solution, can, therefore, be neglected in the t τ S limit. In this overdamped case, the particle velocity is directly proportional to the noise force: v ( t ) = τ S Γ ( t ) / m. The noise does not induce any average motion: v ( t ) ¯ = 0. However, as the velocities are δ-correlated in time, the particle diffuses around its initial position x ( t = 0 ) = 0 with a mean squared displacement σ R 2 = x 2 ( t ) ¯ given by Taylor's theorem2,8,9
(3)
As the velocity autocorrelation function reads v ( t ) v ( t ) ¯ = q ( τ S / m ) 2 δ ( t t ), Eq. (3) integrates into
(4)
D = q τ S 2 / 2 m 2 is called the diffusion coefficient. The distance traveled σR typically increases as t, which is a key characteristic of diffusion.
We focus here on the motion of particles transported by a turbulent air flow of mean velocity u ¯ along the x axis in cylindrical coordinates ( r , θ , x ). L denotes the typical large-scale length, which is in our case set by the transverse dimensions of the space. The Reynolds number Re compares inertia with viscous effects
(5)
with ν being the kinematic viscosity of air. When Re is large enough, typically above 103, the fluid motion given by the Eulerian velocity field u ( r , t ) becomes unsteady, chaotic, and irregular. We, therefore, treat this field statistically by using the Reynolds decomposition: we split u into its ensemble average u ¯, which is a drift velocity, and its fluctuating part with zero mean u , i.e., u = u ¯ + u .10 Likewise, we decompose the particle velocities v into v ¯ and v . The typical particle fluctuating velocity σV is the root mean squared fluctuating velocity σ V 2 v 2 ¯.
It has been argued in a pioneering article by Taylor2 published one century ago that turbulent dispersion, induced by the random velocity fluctuations, is diffusive. The particle dynamical equation reads
(6)
The thermal noise Γ of Eq. (1) is negligible compared to turbulent fluctuations. Likewise, we neglect particle inertia for timescales longer than τS, which is the case if the particles are light enough. Particles, therefore, move with the fluid: v ¯ = u ¯ and v = u so that their velocity correlation functions are the same: v ( t ) v ( t + τ ) ¯ = u ( t ) u ( t + τ ) ¯. As the fluctuating velocity u acts as a noise, Taylor's picture of turbulent diffusion is a Brownian-like motion analogous to molecular diffusion. The turbulent kinetic energy plays the role of temperature.
We assume that longitudinal diffusion along the direction of the average air flow is negligible compared to advective transport, and we inject many particles at the origin with mass rate m ̇. Turbulent dispersion, therefore, only creates a transverse mean squared displacement σR with zero mean. As the individual particle probability distributions are identically distributed, the probability of finding a particle at transverse radial position r will be Gaussian, according to the central limit theorem, i.e., proportional to exp ( r 2 / ( 2 σ R 2 ) ). The average concentration C ¯ is proportional to this probability. The prefactor is set by conservation of mass as the total concentration along the transverse section must be equal to the injection rate m ̇ / v ¯
(7)
In the turbulent regime, the velocity u of an elementary volume of fluid, followed along its Lagrangian path, remains correlated at short times.5,11,12 The velocity of a transported particle is, therefore, correlated as well, which leads to an anomalous diffusion at short distances. Because the particles have negligible inertia, their velocities are the fluid velocities: velocity correlation results from the underlying flow structure driving their motion rather than from the particle dynamics. On a phenomenological ground, the simplest hypothesis would be to assume that the velocity fluctuation is a low-pass filtered noise associated with a typical relaxation time T. We assume that an elementary volume of fluid follows a Langevin equation of the type of Eq. (1) driven by a white noise with a linear friction term and including inertia. The particle velocity correlation, which is equal to the fluid velocity correlation, is not a Dirac δ as previously but decays exponentially7–9,13
(8)
The correlation time T is called the Lagrangian integral time scale, defined in the general case as
(9)
We apply Taylor's theorem (3) to Eq. (8) at time t = x / v ¯ to get the transverse turbulent spread8,9
(10)
At large distances ( x v ¯ T), we recover a diffusive regime
(11)
The concentration along the axis (r =0) is, therefore, expected to decay as 1 / x. Conversely, at short distances ( x v ¯ T), the formula predicts a ballistic-like regime of the form
(12)
which corresponds to a faster decay of the concentration along the axis as 1 / x 2.

We performed “field” experiments in the large corridors of two commercial malls and of our university during lockdown, while they were empty. The experimental setup is schematized in Fig. 1. A controlled CO2 source of constant mass rate m ̇ was obtained by sublimating dry ice sticks in an open cylindrical container of diameter 20 cm heated by a power controlled hot plate. The source was positioned at height 1.1 m above the ground. The imposed sublimation rate, measured using a scale, was equal to m ̇ = 1.5 g / s. This rate was chosen to measure CO2 concentrations with high relative accuracy without saturating the sensors.

Fig. 1.

Schematic of the experimental setup used to measure the turbulent dispersion of contaminants. Dry ice sticks stored in an open cylindrical container are sublimated using a hot plate. The gas produced is convected downstream by the small draft blowing in the corridor where the CO2 concentration is recorded. The scale measures the mass injection rate m ̇. (a) Top view of the corridor layout. The source size ( 20 cm) is much smaller than the typical corridor width (5– 10 m). (b) Zoomed-in view of the source. The sensors are placed at 1.1 m above ground.

Fig. 1.

Schematic of the experimental setup used to measure the turbulent dispersion of contaminants. Dry ice sticks stored in an open cylindrical container are sublimated using a hot plate. The gas produced is convected downstream by the small draft blowing in the corridor where the CO2 concentration is recorded. The scale measures the mass injection rate m ̇. (a) Top view of the corridor layout. The source size ( 20 cm) is much smaller than the typical corridor width (5– 10 m). (b) Zoomed-in view of the source. The sensors are placed at 1.1 m above ground.

Close modal

Our initial motivation was to characterize the dispersion of breath in public spaces. The source used generated CO2 at about 150 times the exhalation rate of an adult at rest. This increased generation rate does not change the dispersion equations as the measured CO2 concentration is still very small in the 10 5 10 4 volume fraction range. The experiment had been designed believing that the turbulent dispersion would be statistically isotropic, as expected for an effective diffusion caused by large scale incoherent turbulent motion.14 Instead, our preliminary observations using a source of micron-sized glycerol/water droplets showed that in most large public spaces, there are air drafts causing bulk horizontal transport and biased dispersion. After identification of the mean air flow direction, non-dispersive infrared (NDIR) CO2 sensors were placed downstream from the source. They recorded the CO2 concentration C ¯ over the duration of each experiment around 30 min. The initial and final values of C ¯ were used to determine the background CO2 level Ce.

After an initial short transient time, a concentration field in a statistically steady state was established. CO2 concentration in excess was measured in eight different locations. The draft wind velocity was measured using a hot-wire anemometer. It ranged from 0.1 to 2 m / s depending on the location. The flow Reynolds number Re = u ¯ L / ν was between 105 and 106 with L 1 m, the typical corridor scale. Measurements were done with different ventilation flow rates and recycled air fractions; the use of fire safety ventilation, namely, mechanical smoke extractors and smoke vents, was tested when available. Measurements were performed with entrance doors both open and closed as open doors create large drafts in some locations. Control sensors (not shown in Fig. 1) placed immediately upstream to the left and to the right of the source showed no concentration increase: convection dominated over turbulent diffusion.

The concentration measured by each sensor fluctuates over time during the experiment. The sensor response time, around a minute, is much longer than typical turbulent time scales. Noise in the concentration signal only comes from small fluctuations in the wind direction, which tilt around the dispersion cone. To remove these fluctuations, concentration signals for each sensor are averaged over the fraction of time during which the air flow is reasonably aligned with the sensor axis. We consider the two to be aligned when the concentration stays close to its maximum value for that particular sensor for a few minutes. This is the case for about 40 % of the duration of each sensor signal. Figure 2 shows that C ¯ decays spatially typically as x 2. The mass conservation equation provides the scaling law
(13)
where ρ CO 2 = 1.83 kg / m 3 is the density of sublimated CO2. In Fig. 2, the CO2 concentration is rescaled by m ̇ / ρ CO 2 v ¯ and plotted as a function of space x. The reasonable collapse between data shows that the CO2 concentration is proportional to the inverse wind velocity v ¯. The fraction of fresh air injected in the ventilation system had no effect on the measured concentrations: in practice, only the local airflow disperses CO2. In other words, the airflow induced to renew air is negligible in comparison to natural air drafts.
Fig. 2.

Time-averaged concentration excess profiles as a function of distance x to the source. The concentration is rescaled by m ̇ / ( ρ CO 2 v ¯ ), where m ̇ is the mass injection rate of sublimating dry ice sticks and v ¯ is the air velocity. The solid curve is the best fit in Eq. (14). The two free parameters are the source size a (here, 0.35 m) and the cone angle α (here, 0.10). The measurements were done at two different locations at the same mall. Red circles and dashed line: human breathing measurements ( m ̇ = 10 mg / s) and its best fit curve ( a = 0.27 m and α = 0.14).

Fig. 2.

Time-averaged concentration excess profiles as a function of distance x to the source. The concentration is rescaled by m ̇ / ( ρ CO 2 v ¯ ), where m ̇ is the mass injection rate of sublimating dry ice sticks and v ¯ is the air velocity. The solid curve is the best fit in Eq. (14). The two free parameters are the source size a (here, 0.35 m) and the cone angle α (here, 0.10). The measurements were done at two different locations at the same mall. Red circles and dashed line: human breathing measurements ( m ̇ = 10 mg / s) and its best fit curve ( a = 0.27 m and α = 0.14).

Close modal
The decay of C ¯ C e is consistent with the overall conical shape of the dispersion zone in the ballistic regime given in Eq. (12). We can parametrize the dilution by
(14)
where α is the dispersion cone angle, determined by the turbulent fluctuation rate σ V / v ¯. We find a = 0.35 m, which is close to the actual diameter of the dry ice container and α = 0.10. We have also included in Fig. 2 the CO2 concentration in the wake of a volunteer breathing through the mouth in a fan-induced wind of velocity v ¯ = 0.3 m / s. We find a 0.27 m and a slightly higher fluctuation rate α = 0.14 in that case. The fluctuation rate depends on the processes creating the mean flow (fans, natural indoor air drafts, wind outdoors) and the geometry, as obstacles upstream destabilize the flow and create larger fluctuations.

In order to better characterize the turbulent transport at play, we studied after lockdown the dispersion of a fog composed of micrometer-sized oil droplets in a small wind tunnel used for experimental physics student projects, schematized in Fig. 3. The 120 × 23 × 23 cm 3 test section is illuminated from above by a LED array with a diffuser. A turbulence generating grid (3 × 3 grid of 7 × 7 cm 2 separated by 1 cm wide bars) is inserted immediately after the air inlet contraction section. u ¯ is measured with a hot-wire anemometer (testo 405i). We estimate the flow Reynolds number u ¯ L / ν with L being the grid mesh size to be between 105 and 106—as in the field experiments. Oil vapor is injected at a controlled rate through a 6 mm nozzle heated by a resistor at a controlled power. A fog of droplets nucleates at few millimeters downstream. The fog is made dilute enough so that double scattering is negligible, and the light intensity scattered is proportional to the local drop concentration. We qualitatively check the dilution by putting the light behind the tunnel rather than above. High resolution pictures with a 1 or 2 s exposure time are taken with a Digital Single-Lens Reflex camera (Nikon D800). A series of ten pictures under the same conditions is averaged to achieve a satisfying statistical convergence. The mean green channel from the pictures in raw format is used to measure the light intensity. The image of the tunnel without fog is subtracted to remove residual background light. The light intensity profile I 0 ( x ) is then calibrated by placing a diffusive black sheet inside the tunnel. I 0 ( x ) is flat along the tunnel axis, except at the start and the end of the tunnel where it falls off due to greater distance to the illumination source.

Fig. 3.

(a) Schematic of the wind tunnel (EA600, Deltalab, Carcassonne, France). A grid generates turbulence at the entrance of the wind tunnel; it is removed to perform laminar experiments. Downstream, fog is injected and we photograph its dispersion. (b) Turbulent fog dispersion cone for u ¯ = 15.2 m / s. The injection nozzle can be seen on the left. The background intensity has been subtracted.

Fig. 3.

(a) Schematic of the wind tunnel (EA600, Deltalab, Carcassonne, France). A grid generates turbulence at the entrance of the wind tunnel; it is removed to perform laminar experiments. Downstream, fog is injected and we photograph its dispersion. (b) Turbulent fog dispersion cone for u ¯ = 15.2 m / s. The injection nozzle can be seen on the left. The background intensity has been subtracted.

Close modal
The intensity I ( x , y ) of a particular pixel is an integral over the z-axis of the scattered intensity. The concentration profile is expected to depend on the distance to the central axis in a Gaussian way: exp ( ( ( y y 0 ) 2 + ( z z 0 ) 2 ) / 2 σ R 2 ). Integrating it over z, the image intensity field I ( x , y ) remains Gaussian along y
(15)

Transverse intensity profiles at regularly spaced distances from the fog injection nozzle are shown in Fig. 4. As the droplets are formed by vaporizing oil, the fog is hot and slightly buoyant so the centerline y 0 ( x ) gets slightly shifted upwards. The centerline slope, which is the transverse-to-longitudinal mean particle velocity ratio, is at most 10 2, allowing any additional effects of this heating to be ignored. The profiles are nicely fitted by a Gaussian, which leads to the measurement of the three parameters of Eq. (15): the concentration on the axis, C, the radius σR, and the centerline y 0 ( x ). Assuming that v ¯ is constant along the flow axis, mass conservation imposes that C ( x ) σ R 2 ( x ) v ¯ must be constant proportional to the rate of emission by the source. Figure 5 shows typical profiles C ( x ) σ R 2 ( x ), which are indeed flat outside the injection zone. It means that the droplets, once nucleated, do not change much and keep the same light scattering properties; this validates the use of this measurement technique.

Fig. 4.

Transverse intensity profiles I at various distances from the injection nozzle. I is rescaled by the maximal intensity for x = 8 cm. The best fit by a Gaussian is superimposed, which allows us to extract the concentration C on the axis, the radius σ R ( x ), and the centerline position y 0 ( x ).

Fig. 4.

Transverse intensity profiles I at various distances from the injection nozzle. I is rescaled by the maximal intensity for x = 8 cm. The best fit by a Gaussian is superimposed, which allows us to extract the concentration C on the axis, the radius σ R ( x ), and the centerline position y 0 ( x ).

Close modal
Fig. 5.

Profiles of the fog flow rate C ( x ) σ R 2 ( x ) as a function of distance downstream x for u ¯ = 15 m / s. The fog flow rate is rescaled by its mean along x. Red curve: turbulent profile. Gray curve: laminar profile. The curves are slightly shifted vertically around 1 for readability. The two profiles collapse on the same curve except near the injection nozzle, where oil drops nucleate from their vapor. The nearly constant profile shows that the droplets do not change much in size during convection and dispersion. The variations, equal on both curves, are due to residual heterogeneities of light. This validates the measurement of the concentration field within a multiplicative constant from experimental data.

Fig. 5.

Profiles of the fog flow rate C ( x ) σ R 2 ( x ) as a function of distance downstream x for u ¯ = 15 m / s. The fog flow rate is rescaled by its mean along x. Red curve: turbulent profile. Gray curve: laminar profile. The curves are slightly shifted vertically around 1 for readability. The two profiles collapse on the same curve except near the injection nozzle, where oil drops nucleate from their vapor. The nearly constant profile shows that the droplets do not change much in size during convection and dispersion. The variations, equal on both curves, are due to residual heterogeneities of light. This validates the measurement of the concentration field within a multiplicative constant from experimental data.

Close modal

Figure 6 shows various profiles of the dispersion radius σ R ( x ). The dispersion of aerosols is significantly larger with a grid generating turbulence than without. In all cases, Eq. (10) provides an excellent fit to the data within the scatter. It constitutes a very striking result: a model based on a single correlation timescale is sufficient to adequately represent all the data. At short distance x < v ¯ T from the source, the dispersion takes place in a cone: σR is linear in x. This controversial regime has been explored in a series of recent papers.3,4,15–20 At large distance x > v ¯ T, conversely, the diffusion regime predicted by Taylor with a constant diffusivity is recovered, in which the radius increases as the square root of time, i.e., x here.

Fig. 6.

Radius σ R ( x ) profiles as a function of distance downstream x. The best fit in Eq. (10) is superimposed, which provides measurements of the fit parameters σ V / v ¯ and v ¯ T. The fit by a model based on a single correlation timescale is sufficient within the scatter of the data.

Fig. 6.

Radius σ R ( x ) profiles as a function of distance downstream x. The best fit in Eq. (10) is superimposed, which provides measurements of the fit parameters σ V / v ¯ and v ¯ T. The fit by a model based on a single correlation timescale is sufficient within the scatter of the data.

Close modal

From each profile σ R ( x ), two quantities are, therefore, extracted: the turbulent fluctuation rate σ V / v ¯ and the cross-over length v ¯ T, which is the cross-over distance between the ballistic and diffusive regimes. These quantities are plotted in Fig. 7 as a function of the mean flow velocity u ¯. The fluctuation rate σ V / v ¯ is independent of the wind velocity u ¯, as expected, and equals to 7.5 %. The cross-over length v ¯ T is very small when no grid is introduced at the entrance of the wind tunnel. It is much larger with a turbulence generating grid. It tends to a nonzero value as u ¯ 0 and increases with the fluid velocity u ¯. The Lagrangian correlation length σ V T increases from 0.7 to 2.5 cm in the range of velocities explored, which is significantly smaller than the grid mesh size: correlations come from intermediate scale flow structures. The slope of the curve gives a characteristic time around 9 ms, which is much larger than the Stokes time of oil droplets, around 0.3 ms. As shown in the discussion of Eq. (6), this is consistent with the dynamics of particles with negligible inertia. To the best of our knowledge, there is no simple mechanistic explanation to the observed dependence of T with u ¯. This dependence is a near-field effect of the grid, which is why we have used a single decorrelation time, independent of u ¯, to fit the field data of Fig. 2.

Fig. 7.

Dependence of the fluctuation rate σ V / v ¯ and of the cross-over length v ¯ T with the mean flow velocity u ¯. In panel (b), the squares are measurement performed in the quasi-laminar case in the absence of a grid at the entrance of the wind tunnel.

Fig. 7.

Dependence of the fluctuation rate σ V / v ¯ and of the cross-over length v ¯ T with the mean flow velocity u ¯. In panel (b), the squares are measurement performed in the quasi-laminar case in the absence of a grid at the entrance of the wind tunnel.

Close modal

SARS-CoV-2 is transmitted in the airborne route by respiratory aerosols: droplets of mucus, ranging between 200 nm and 200 μ m, which can contain viral particles.6,21,22 These droplets are produced in the respiratory tract during normal respiratory activity by various fluid instabilities.23–25 The air exhaled by an infected person contains active SARS-CoV-2 viral particles, which may start an infection when one of the virions is deposited on an ACE2 receptor of the epithelium and successfully bypasses the immune response. Infection may also start if virions deposit on the eyes. The probability of infection increases with the intake viral dose, defined as the amount of virus particles inhaled by a person, cumulated over time. It, therefore, increases with the exposure time and with the concentration of viral particles in the air.26,27

The viral transmission risk by a given infected patient is, therefore, determined by the concentration in viral particles in the air inhaled by another susceptible person. This concentration is given by the rate of emission of viral particles in exhaled air and by the dilution of the air between exhalation and inhalation. In closed rooms, viral particles accumulate, and their concentration eventually reaches a steady state between exhalation and ventilation, which is the replacement of contaminated air by fresh air. This explains the large transmission clusters in poorly ventilated indoor spaces.6 The average transmission risk, uniform in space, can be determined by assuming that viral particles are well-mixed by turbulence. In that case, dilution is due to the natural or forced air exchanges with outside air.26,27 Near an infected person, however, the concentration is much higher and decays with distance, and the dilution from the exhalation concentration of viral particles to the average concentration of viral particles is controlled by the local airflow. The total transmission risk is, therefore, the sum of an average risk, controlled by ventilation, and an additional risk in the dispersion cone of exhalations. Outdoors, in the absence of homogenizing flow, this short-range risk is the only transmission risk. Respiratory activity, such as unmasked coughs and sneezes, can create large airflows that dominate over the pre-existing flow patterns. In that case, air is exhaled in a buoyant jet that gets gradually mixed at its boundaries as it travels.23,25 Face masks strongly alter the airflow by blocking the jet; small jets can remain along the nose and the cheeks if the mask is not properly fitted.28,29 The stopping distance of a masked respiratory puff is around29  20 cm; the more filtering masks make the puff stop earlier. Beyond this very near field distance, dilution is fully controlled by the ambient airflow, which is what we have extensively developed here.

Our analysis applies to any public space where either face mask wearing is mandatory and there is a horizontal air draft or the air velocity is much larger than the exhalation velocity. We have found that in large indoor corridors, the airflow is a horizontal draft with u ¯ = 0.1 1 m / s. Outdoors, the wind is also horizontal with larger velocities. Understanding the dilution in these public spaces allows us to assess their transmission risk, given biological data on pathogen contagiousness.26,27 This is needed to ground infection prevention policies on a solid physical basis, as some (such as the “6 ft rule”) are not (see Refs. 22 and 30 for a historical perspective and critical discussion). Exhaled CO2 is a commonly used risk proxy in well-mixed rooms;26,27 its spatial decay allows for a simple translation of the short-range risk into an equivalent long-range risk. For the gentle draft of shopping mall corridors u ¯ 0.2 m / s, the increase in risk at 6 ft corresponds to a CO2 concentration excess of 50 ppm.

In this article, we have investigated experimentally and theoretically the dispersion of a passive scalar (the concentration of particles) by a turbulent flow. This constitutes an accessible problem to teach turbulence at an undergraduate level, combining possible experiments and theory. Using any source of fog, a fan and a camera, the Reynolds decomposition between average and fluctuating quantities can be illustrated qualitatively and quantitatively. The derivations based on physical reasoning and dimensional analysis remain linear and much simpler than attempting to solve the Navier–Stokes equation. While the Reynolds number dependence of the drag force exerted on a solid is useful to introduce dimensionless numbers, asymptotic regimes, viscosity, and inertial effects, as it is usually done in fluid dynamics textbooks (see, e.g., Ref. 1), the diffusion of a passive scalar provides a way to understand the role of fluctuations to enhance turbulent mixing. The applications to the dispersion of odors and animal olfactory search,31 the dispersion of pollutants, or airborne viral transmission are immediately appealing to students.

Unibail-Rodamco-Westfield has funded this work under CNRS Contract No. 217977 and provided access to Forum des Halles and Carré Sénart as well as technical assistance. This article involves seven undergraduate students who have worked on the problem in the context of their final-year experimental physics courses (“Phy Exp ”) at the Université de Paris Cité. The authors thank the “Phy Exp ” team and, in particular, the lathe-mill operator, Wladimir Toutain, and the technician, Thibaut Fraval De Coatparquet, for their assistance. The authors thank Alessandra Lanotte and Luca Biferale for fruitful discussions. The authors thank the referees for their insightful comments, which have helped them to improve the manuscript.

This work was funded by Unibail-Rodamco-Westfield on behalf of Conseil National des Centres Commerciaux (CNCC), who asked the authors to make recommendations for a health protocol aiming to reduce and quantify the transmission risk in shopping centers. The conclusions of the present article are, therefore, of direct interest for the funding company. The authors declare no financial competing interest. The funding company had no such involvement in study design, in the collection, analysis, and interpretation of data, nor in the writing of the article. The authors had the full responsibility in the decision to submit it for publication.

1.
G. K.
Batchelor
,
An Introduction to Fluid Dynamics
, 14th ed. (
Cambridge U. P., Cambridge
,
2010
), pp.
211
235
.
2.
G. I.
Taylor
, “
Diffusion by continuous movements
,”
Proc. London Math. Soc.
s2–20
,
196
212
(
1922
).
3.
E.
Villermaux
and
C.
Innocenti
, “
On the geometry of turbulent mixing
,”
J. Fluid Mech.
393
,
123
147
(
1999
).
4.
E.
Villermaux
,
C.
Innocenti
, and
J.
Duplat
, “
Short circuits in the Corrsin–Obukhov cascade
,”
Phys. Fluids
13
,
284
289
(
2001
).
5.
G. K.
Batchelor
, “
The application of the similarity theory of turbulence to atmospheric diffusion
,”
Q. J. R. Meteorol. Soc.
76
,
133
146
(
1950
).
6.
T.
Greenhalgh
et al, “
Ten scientific reasons in support of airborne transmission of SARS-CoV-2
,”
Lancet
397
,
1603
1605
(
2021
).
7.
H.
Risken
and
T.
Frank
,
The Fokker-Planck Equation: Methods of Solution and Applications
, 2nd ed. (
Springer
,
Berlin, Heidelberg
,
1989
), pp.
32
37
.
8.
J. P. L. C.
Salazar
and
L. R.
Collins
, “
Two-particle dispersion in isotropic turbulent flows
,”
Annu. Rev. Fluid Mech.
41
,
405
432
(
2009
).
9.
B.
Sawford
, “
Turbulent relative dispersion
,”
Annu. Rev. Fluid Mech.
33
,
289
317
(
2001
).
10.
M.
Lesieur
,
Turbulence in Fluids
, 4th ed. (
Springer
,
Dordrecht
,
2008
), pp.
187
208
.
11.
G. K.
Batchelor
, “
Diffusion in a field of homogeneous turbulence. I. Eulerian analysis
,”
Aust. J. Chem.
2
,
437
450
(
1949
).
12.
G. K.
Batchelor
, “
Diffusion in a field of homogeneous turbulence: II. The relative motion of particles
,”
Math. Proc. Cambridge Philos. Soc.
48
,
345
362
(
1952
).
13.
B. L.
Sawford
, “
Reynolds number effects in Lagrangian stochastic models of turbulent dispersion
,”
Phys. Fluids A
3
,
1577
1586
(
1991
).
14.
K.-C.
Cheng
et al, “
Modeling exposure close to air pollution sources in naturally ventilated residences: Association of turbulent diffusion coefficient with air change rate
,”
Environ. Sci. Technol.
45
,
4016
4022
(
2011
).
15.
J.
Bec
et al, “
Acceleration statistics of heavy particles in turbulence
,”
J. Fluid Mech.
550
,
349
358
(
2006
).
16.
L.
Biferale
et al, “
Lagrangian structure functions in turbulence: A quantitative comparison between experiment and direct numerical simulation
,”
Phys. Fluids
20
,
065103
(
2008
).
17.
F.
Toschi
and
E.
Bodenschatz
, “
Lagrangian properties of particles in turbulence
,”
Annu. Rev. Fluid Mech.
41
,
375
404
(
2009
).
18.
M.
Bourgoin
, “
Turbulent pair dispersion as a ballistic cascade phenomenology
,”
J. Fluid Mech.
772
,
678
704
(
2015
).
19.
J.
Bec
et al, “
Turbulent pair dispersion of inertial particles
,”
J. Fluid Mech.
645
,
497
528
(
2010
).
20.
M.
Bourgoin
et al, “
The role of pair dispersion in turbulent flow
,”
Science
311
,
835
838
(
2006
).
21.
L.
Morawska
et al, “
A paradigm shift to combat indoor respiratory infection
,”
Science
372
,
689
691
(
2021
).
22.
K.
Randall
et al, “
How did we get here: What are droplets and aerosols and how far do they go? A historical perspective on the transmission of respiratory infectious diseases
,”
Interface Focus
11
,
1
10
(
2021
).
23.
L.
Bourouiba
,
E.
Dehandschoewercker
, and
J. W. M.
Bush
, “
Violent expiratory events: On coughing and sneezing
,”
J. Fluid Mech.
745
,
537
563
(
2014
).
24.
R.
Mittal
,
R.
Ni
, and
J.-H.
Seo
, “
The flow physics of COVID-19
,”
J. Fluid Mech.
894
,
F2-1-14
(
2020
).
25.
L.
Bourouiba
, “
The fluid dynamics of disease transmission
,”
Annu. Rev. Fluid Mech.
53
,
473
508
(
2021
).
26.
S. N.
Rudnick
and
D. K.
Milton
, “
Risk of indoor airborne infection transmission estimated from carbon dioxide concentration
,”
Indoor Air
13
,
237
245
(
2003
).
27.
M. Z.
Bazant
and
J. W. M.
Bush
, “
A guideline to limit indoor airborne transmission of COVID-19
,”
Proc. Natl. Acad. Sci.
118
,
e2018995118
(
2021
).
28.
J. W.
Tang
et al, “
A schlieren optical study of the human cough with and without wearing masks for aerosol infection control
,”
J. R. Soc. Interface
6
,
S727
736
(
2009
).
29.
S.
Verma
,
M.
Dhanak
, and
J.
Frankenfield
, “
Visualizing the effectiveness of face masks in obstructing respiratory jets
,”
Phys. Fluids
32
,
061708-1-7
(
2020
).
30.
T.
Greenhalgh
,
M.
Ozbilgin
, and
D.
Contandriopoulos
, “
Orthodoxy, illusio, and playing the scientific game: A Bourdieusian analysis of infection control science in the COVID-19 pandemic
,”
Wellcome Open Res.
6
,
126-1-32
(
2021
).
31.
E.
Balkovsky
and
B. I.
Shraiman
, “
Olfactory search at high Reynolds number
,”
Proc. Natl. Acad. Sci.
99
,
12589
12593
(
2002
).