Hazardous pollutant containment zones should be maintained at a pressure lower than the outdoor atmospheric pressure to prevent pollutants from escaping to the outdoor environment. However, atmospheric wind conditions can cause breaching of the containment zone that is established through mechanical ventilation. This paper combines external wind pressure (Pe) time series on an internally depressurized building with a carefully designed ventilation network to analyze indoor pressure (Pi) and containment breach duration. The Pe data are obtained by wind-tunnel (WT) tests and computational fluid dynamics (CFD) simulations of large-eddy simulation (LES) and scale-adaptive simulation (SAS). The objectives of the paper are (1) comparing Pe results by WT and CFD; (2) assessing the impact of Pe uncertainties in CFD on the resulting Pi and breach duration; and (3) estimating Pi and breach duration for an initial case study with indoor depressurization of -40 Pa and reference wind speed (Uref) of 12.65 m/s at building height. The results are discussed in terms of dimensionless pressure coefficients (Cpe and Cpi). It is shown that LES and especially SAS Cpe data deviate substantially from the WT values but that the impact of these CFD uncertainties on Cpi and breach duration remains fairly limited. Deviations for Cpi statistics fall within the experimental uncertainty, and the CFD breach duration deviates generally less than 10% from the WT result. Estimated breach durations for Uref = 12.65 m/s, however, can go up to 80% in spite of the −40 Pa depressurization, stressing the importance of this type of studies.

Indoor pollutant-containing zones require utmost care in sealing the containment particularly if the pollutant is hazardous to public and occupational health, as is the case for nuclear industries, asbestos abatement zones, and others. It should be ensured that no pollutant particles escape to the outdoor environment. However, these particles could escape through unintended openings in building envelopes such as leaks. Hence, it is advised to ensure a negative pressure within the containment zone with respect to the outer environment, so that any uncontrolled flow is directed always into the containment zone [e.g., Great Britain, Health and Safety Executive (HSE), 2006; KTA-Geschäftsstelle, 2017]. For example, asbestos abatement from building interiors has the risk of suspended fibers being airborne and escaping into the outdoor atmosphere in the case of reverse flow from the containment zone to the outdoors. In order to minimize such situations, different government authorities (e.g., World Health Organization Regional Office for Europe, 2012; Legifrance, 2013; Ministerie van Sociale Zaken en Werkgelegenheid (SZW), 2016; and Occupational Safety and Health Administration (OSHA), 2019 in the case of an asbestos abatement worksite) have laid out regulatory measures. This includes maintaining a negative pressure within the containment with respect to the outdoor environment by means of a mechanical ventilation system (D'Angelo , 1987). This mechanical ventilation system should ensure that the flow out of the containment occurs only through a negative pressure unit equipped with high-efficiency particulate air (HEPA) filters. The inflow of fresh air into the containment is achieved by means of specifically designed air inlets and airlocks with check valves. The depressurization requirement established by the different authorities is, however, laid out without considering explicitly the potential effects of atmospheric winds and its fluctuations, as evident from the lack of scientific publications, standards, and guidelines on this issue. However, as clearly demonstrated in the scientific literature, atmospheric wind can create high suction pressures on the building envelopes (e.g., Cermak, 1975; Simiu and Scanlan, 1996; ASHRAE, 2001; Stathopoulos, 2002; and Holmes, 2004) and, therefore, it is expected that they can be responsible for breaches during containment asbestos removal.

Ventilation network models (VNMs) are valuable tools for analyzing the airflow in mechanically ventilated buildings (e.g., Axley, 2007; Chen, 2009). Their advantages include fast simulations with minimal computational resource requirements, compared to other numerical methods such as computational fluid dynamics (CFD). Over the past few decades, several VNMs have been developed, such as COMIS (Feustel, 1999), CONTAM (Dols and Polidoro, 2015), and SYLVIA (IRSN, 2022). VNMs are zone-based airflow models, where the building equipped with a ventilation system is represented by nodes and branches. Nodes represent rooms, junctions, or an exterior environment (boundary condition) in which properties like pressure, temperature, and pollutant concentration are assumed to be uniform. The equations of conservation of mass, momentum, and energy are applied for the nodes. Branches represent components of the ventilation system, such as air inlets, airlocks, leaks, filters, and fans in the case of asbestos abatement worksites. The mass flow rate between nodes is calculated with the generalized Bernoulli's equation for the branches.

Using external wind pressure data as input, VNMs can effectively analyze and investigate the effects of wind on indoor airflow through mechanically ventilated buildings or compartments within buildings (Cóstola , 2009; Le Roux , 2013). In this research study, the code SYLVIA has been used. SYLVIA is a VNM developed by the Institut de Radioprotection et de Sûreté Nucléaire of France (IRSN, 2022) and it has been validated in the past years for ventilation, fire and airborne contamination in nuclear facilities equipped with mechanical ventilation systems to establish internal depressurization (Le Roux , 2013, 2012; Vaux and Prétrel, 2013; Coutin, 2015; and Le Dez , 2021). Additionally, Jayakumari (2025) validated the model's effectiveness in predicting internal pressure for mechanically ventilated asbestos containment zones, which operate at a lower depressurization range compared to nuclear facilities (e.g., −5 to −40 Pa). SYLVIA has the capability to perform transient simulations using time series of external pressures as input, to predict its influence on an initial depressurization established within a containment volume. The user-friendly interface includes data import/export and checks, minimizing the risk of setup errors made by users and enabling the design and simulation of complex systems.

However, reliable data of external wind pressures are imperative for an accurate indoor airflow prediction with VNMs. As discussed by Cóstola (2009), there are different sources of external wind pressures available for use in VNMs. This includes analytical models and provisions in codes/standards, field measurements, WT tests, and CFD simulations. Analytical models and provisions in codes and standards (e.g., ASHRAE, 2001; ASCE, 2002) typically focus on isolated buildings with simplified geometries. Furthermore, they lack the provision of time-series data for external wind pressures, which is essential for input in transient simulations. Field measurements are the most realistic and consequently also the most representative source of wind pressures on buildings (Levitan and Mehta, 1992; Richards , 2001; and Li , 2021). However, recording external pressure time series at different locations on building facades over a sufficiently long period as input for SYLVIA is challenging and very cumbersome and expensive. In addition, a field measurement campaign comes with the practical challenges, such as obtaining approvals for measurements in existing buildings, particularly when hazardous materials like asbestos particles are involved, making them not always feasible. Such measurements are typically used to gain insights into reality and for validation purposes (e.g., Blocken, 2014, 2015; Mateus , 2016). Atmospheric boundary layer wind tunnel (ABLWT) tests are still the most common source to obtain wind pressures for VNMs (Castro and Robins, 1977; Cheung , 1997; Stathopoulos, 2002; Cermak, 2003; Blocken, 2018; and Lamberti , 2020). The tests are typically performed on reduced-scale models, following the required similarity criteria to attain high-quality data under controlled conditions (e.g., Castro and Robins, 1977; Hunt, 1982; Lin , 1995; Richards , 2001; Tominaga and Blocken, 2015; 2016; and Jayakumari , 2023; 2024). In addition to WT tests, the rapid explosion of computational capacities over the last decades led to an improved prediction of wind pressures on buildings by CFD (e.g., Nozawa and Tamura, 2002; Ono , 2008; Montazeri and Blocken, 2013; Ricci , 2018; Guichard, 2019; Lamberti , 2020; Xu 2023; Xing , 2023; and Ricci, 2024). However, the reliability of CFD (as an alternative to WT) as source of external wind pressure (Pe) for VNM to predict indoor pressure (Pi) is not known yet. Despite the great efforts of the computational wind engineering community in recent decades to improve the prediction of external wind pressures by CFD, the accurate prediction of pressure fluctuation and, in particular, peak pressures is still challenging (e.g., Potsis , 2023).

This paper combines external wind pressure (Pe) time series on an internally depressurized building with a carefully designed ventilation network to analyze indoor pressure (Pi) and containment breach duration. The Pe time series and statistical parameters are obtained by reduced-scale WT tests performed on an idealized isolated building (Jayakumari , 2023), on the one hand, and by CFD simulations of large-eddy simulation (LES) and scale-adaptive simulation (SAS) (Menter and Egorov, 2010) on the other hand. SAS is a hybrid RANS-LES approach recently gaining widespread applicability in various wind engineering applications (e.g., García , 2015; Jadidi , 2018; Rezaeiha , 2019; van Druenen and Blocken, 2023; Žužul , 2023; and Qin , 2024). The SAS approach integrates aspects from both LES and RANS to perform transient simulations, by resolving large-scale turbulent structures of the flow (similar to LES) rather than modeling the impact of these on the mean flow like the RANS approach. Its advantage lies in its ability to perform transient simulations with a lower resolution of grid and time step size near the walls (compared to LES), while still capturing a large part of the energy turbulence spectrum. Additionally, unlike other hybrid models such as DES, SAS can dynamically switch from LES-like to RANS-like approach when temporal or spatial resolution is inadequate for achieving LES-like solutions. For this study, SAS is particularly useful as it enables the generation of Pe time series required for the ventilation network model SYLVIA (IRSN, 2022). The Pe time series from WT, LES, and SAS are used as input to the ventilation network model SYLVIA (IRSN, 2022). The depressurization value used is -40 Pa, and the reference wind speed (Uref) at building height is 12.65 m/s. The goal of the paper is threefold: (1) comparing the Pe data as obtained by WT and CFD in terms of time series and statistics; (2) assessing the impact of uncertainties in the Pe as generated by CFD on the resulting Pi and on the breach duration percentage; (3) estimating the Pi statistics and breach duration percentage for an initial case study with the indoor depressurization of −40 Pa and a fixed Uref of 12.65 m/s at building height. The objective (2) aims to determine whether, despite the inherent uncertainties in CFD results, the combinations LES-SYLVIA (LES-S) or SAS-SYLVIA (SAS-S) could be used as alternatives to the WT-SYLVIA (WT-S) combination in future studies.

The manuscript is structured as follows. Section II describes the methodology and the case study and introduces the experimental and numerical techniques used in the study. Section III discusses the results from the experiments and numerical simulations. Section IV closes the paper with summary, limitations, and conclusions.

The methodology used in this study is illustrated in the workflow of Fig. 1. In the first stage, WT tests and CFD (i.e., SAS and LES) simulations are performed on an idealized isolated cubical building model (see Sec. II B) to generate the time series of external pressures (Pe). The Pe results from SAS and LES simulations are validated by means of the WT results. In the second stage, the corresponding Pe results from the WT tests and CFD (i.e., SAS and LES) simulations are used as input for SYLVIA to predict the internal pressure (Pi) variation over time within the asbestos containment zone. In the third stage, the Pi results from SAS-S and LES-S are compared to those obtained from WT-S, to assess the effect of different methods for obtaining Pe on the prediction of Pi for a mechanically ventilated asbestos containment zone. The results are analyzed and presented in terms of external and internal pressure coefficients (Cpe and Cpi). The percentage containment breach predicted by SYLVIA is also compared among WT-S, SAS-S, and LES-S at multiple locations on the building facades.

FIG. 1.

Methodology overview and indication of the workflow.

FIG. 1.

Methodology overview and indication of the workflow.

Close modal

An idealized cubical building of dimensions 18 × 18 × 18 m3 in full scale is considered for the case study. The asbestos containment zone constitutes the upper half of the building, corresponding to a volume of 18 × 18 × 9 m3. The ventilation system for asbestos abatement consists of air inlets, negative pressure units (NPU) with HEPA filters, an airlock for materials, an airlock for people, and air inlets for tuning. The ventilation system for the considered containment volume is designed following a well-established methodology (Dubernet , 2018) to define an internal depressurization of −40 Pa for the reduced-scale building. This ventilation system comprises 40 air inlets (A1–A40), 4 NPUs (N1–N4), an airlock for materials, an airlock for people, and an air inlet for tuning. A schematic diagram of the reduced-scale building with ventilation components and photographs of the ventilation components from a real worksite are reported in Fig. 2. In addition to these ventilation components, four leakages (L1–L4) positioned close to the edges are considered. The WT tests and CFD simulations are performed on a 1:40 scaled model of the case study building to generate the Pe time series at the locations of each ventilation component (including the leakages). Two wind directions, θ = 0° and θ = 45°, are investigated. The exact positioning of the ventilation components (i.e., where Pe is generated) on the building facades and their labels for reference are shown in Fig. 3.

FIG. 2.

Schematic diagram of the reduced-scale building with the ventilation components.

FIG. 2.

Schematic diagram of the reduced-scale building with the ventilation components.

Close modal
FIG. 3.

Location of ventilation components on the building model where external pressure time series are sampled in WT tests and CFD simulations: (a) facade 0–1, (b) facade 1–2, (c) facade 2–3, (d) facade 3–0, and (e) isometric view of the building model.

FIG. 3.

Location of ventilation components on the building model where external pressure time series are sampled in WT tests and CFD simulations: (a) facade 0–1, (b) facade 1–2, (c) facade 2–3, (d) facade 3–0, and (e) isometric view of the building model.

Close modal
WT tests are performed in the ABLWT facility of the Eindhoven University of Technology, in the Netherlands. The setup and results related to these experimental tests are reported in detail in Jayakumari (2023) and are only briefly summarized in the present paper. Figure 4 shows the photograph of the building model (0.45 × 0.45× 0.45 m3) in the WT test section positioned at the center of the turntable. A neutral ABL representing the flow over moderately rough terrain (VDI-3783, 2000) is generated using roughness elements and vortex generators as shown in Fig. 4. The characteristics of the ABL wind at the building location, also called incident profile (e.g., Blocken , 2008), including the stream-wise mean velocity (U) and turbulence intensity (Iu), are presented in Fig. 7 and discussed in Sec. II D 2, along with the inflow conditions for the CFD simulations. The building model is instrumented with 48 external pressure taps, and the locations are shown in Fig. 3. Scanivalve MPS4264 pressure scanners are used to record Pe time series for a duration of 180 s at a frequency of 800 Hz. However, only a duration of 34.4 s is considered for the analysis in this paper, to be consistent with the duration of the CFD simulations (Sec. II D 2). The results of the WT tests are discussed in Sec. III for θ = 0° and θ = 45° and are presented in terms of Cpe calculated with the following equation:
(1)
where Pe is the measured pressure at a given location on the building surface, P0 is the static reference pressure, Uref = 12.65 m/s is the reference mean velocity at building height, and ρ is the air density. The Reynolds number based on building height is 3.8 × 104. The peak pressure coefficients from the external pressures from WT and CFD as well as the internal pressures from SYLVIA are determined based on Cook and Mayne (1979) using the Gumbel distribution, also known as generalized extreme value distribution type I.
FIG. 4.

Photograph of the WT setup with the reduced-scale building model (1:40).

FIG. 4.

Photograph of the WT setup with the reduced-scale building model (1:40).

Close modal

1. Computational domain and grid

The computational domain for the 1:40 scaled building is created following the best practice guidelines for CFD simulations in wind engineering and urban and building aerodynamics (e.g., Franke , 2007; Tominaga , 2008) with an upstream fetch of 5 h, a downstream fetch of 15 h, and a height of 6 h, where h is the building height in reduced scale equal to 0.45 m (Fig. 5). The blockage ratio is 0.8%, which is well below the maximum of 3% (Franke , 2007; Tominaga , 2008; and Blocken, 2015). The Gambit 2.4.6 software is used for grid generation (Fluent Inc., 2005). A grid-sensitivity analysis is carried out on three grids by means of a factor of √2 in each of the three Cartesian directions (x, y, z): the coarse grid (4 193 128 cells), the basic grid (17 357 985 cells), and the fine grid (32 208 108 cells) (Fig. 6). For the basic grid, the cells close to the building facades for a distance of h around the building are cubical with a size of 0.0125 m, equivalent to 0.5 m at full scale. Farther away from the building, cells are stretched with a growth rate of 1.02 to limit the overall number of cells while ensuring a close-to-cubical cell shape suitable for LES simulations. The coarse and fine grids are derived from the basic grid by applying the above-mentioned factor for coarsening or refining, respectively. All simulations for the grid-sensitivity analysis are performed with the LES approach.

FIG. 5.

Perspective view of the computational domain and boundary conditions for θ = 0° and θ = 45°.

FIG. 5.

Perspective view of the computational domain and boundary conditions for θ = 0° and θ = 45°.

Close modal
FIG. 6.

Computational grids used for the grid-sensitivity analysis: (a) coarse (4 193 128 cells), (b) basic (17 357 985 cells), and (c) fine (32 208 108 cells).

FIG. 6.

Computational grids used for the grid-sensitivity analysis: (a) coarse (4 193 128 cells), (b) basic (17 357 985 cells), and (c) fine (32 208 108 cells).

Close modal

2. Boundary conditions and other settings

The boundary conditions for the SAS and LES simulations are reported in Fig. 5, for the wind directions θ = 0° and θ = 45°.

For the SAS simulations, the velocity inlet is imposed at plane-1 for θ = 0°, and at plane-1 and plane-2 for θ = 45°. In particular, the incident vertical mean wind velocity (U) and turbulence intensity (Iu) profiles from the WT tests are used (Fig. 7). The vertical U profile measured in the WT and the corresponding fit to the power law is shown in Fig. 7(a). The fitted U profile is given by the following equation (Davenport, 1960):
(2)
where U(z) is the mean wind velocity, Uh is the reference mean velocity (equal to 12.65 m/s) at building height h = 0.45 m, and α is the power law coefficient (equal to 0.17) corresponding to moderately rough terrain, in adherence with the standard considered in the present study (i.e., VDI-3783, 2000). The Iu(z) values measured in the WT [Fig. 7(b)] are fit to an empirical Eq. (3) (Guichard, 2019), and the results are used in Eq. (4) to calculate the turbulent kinetic energy (k). The turbulence dissipation rate (ε) and the specific dissipation rate (ω) are calculated with Eq. (5) (Richards and Hoxey, 1993) and Eq. (6) (Wilcox, 1994), respectively,
(3)
(4)
(5)
(6)
where Iu is the bulk turbulence intensity value (equal to 0.1144), Iug is the ground turbulence intensity value (equal to 0.355), a is the curvature parameter (equal to 4.984), u* is the friction velocity (equal to 0.69 m/s), z0 is the aerodynamic roughness length (equal to 0.0003 m) at reduced scale (1:40) obtained from the logarithmic fit of U(z) data from the WT, κ is the von Karman constant assumed equal to 0.40, and Cμ is the empirical constant equal to 0.09. The spectral synthesizer method is adopted to perturb the flow field at the inlet (Smirnov , 2001; Vasaturo 2018; and Guichard, 2019). For θ  =  0°, zero static gauge pressure is applied at plane-3 for the outlet. Symmetry is imposed at plane-2 and plane-4 as well as at the top. For θ  =  45°, a zero static gauge pressure outlet is imposed at plane-3 and plane-4, while symmetry is applied to the top. At the ground and building surfaces, the no-slip wall condition with the standard wall functions proposed by Launder and Spalding (1974) is used. The wall functions are implemented with the roughness modification by Cebeci and Bradshaw (1977). The equivalent sand-grain roughness height (ks) is calculated with the following equation in accordance with Blocken (2007):
(7)
where the roughness constant (Cs) is taken equal to 0.5. At the ground, kS = 0.0059 m is imposed, while kS = 0 is chosen for the building surfaces.
FIG. 7.

Comparison of vertical incident flow profiles measured in the WT and corresponding fitted curves used for the LES and SAS simulations: (a) normalized mean wind velocity (U) and (b) streamwise turbulence intensity (Iu).

FIG. 7.

Comparison of vertical incident flow profiles measured in the WT and corresponding fitted curves used for the LES and SAS simulations: (a) normalized mean wind velocity (U) and (b) streamwise turbulence intensity (Iu).

Close modal

For the LES simulations, similar boundary conditions as for SAS are used. The U(z), k(z), and ε(z) profiles at the inlet are obtained from the WT data using Eqs. (2), (4), and (5), respectively. The Wall-Adapting Local Eddy-Viscosity (WALE) sub-grid scale model proposed by Nicoud and Ducros (1999) is adopted for modeling the sub-grid scale stresses. As in SAS, the spectral synthesizer method is adopted to perturbate the flow field at the inlet.

Both LES and SAS simulations are performed with ANSYS Fluent 2021 R2 (Ansys Inc., 2021), with the SAS approach coupled to the shear-stress transport (SST) k–ω turbulence model (Menter , 2003). The fractional step method is used for pressure–velocity coupling with non-iterative time advancement (NITA). For the spatial discretization, the least squares cell-based method is used for gradient computation, the second-order scheme is used for pressure, a bounded-central-differencing scheme is used for momentum, and the second-order upwind scheme is used for k and ω. For the temporal discretization, the bounded second-order implicit scheme is used. A time step of 6.25 × 10−4 s is used to keep the Courant–Friedrichs–Lewy number (CFL) smaller than 1. Ten flow-through-time is executed to initialize the flow field and 35 flow-through times for the sampling of the flow statistics. Thus, the simulations are performed for a total duration of 34.4 s (at reduced scale) generating 54998 data samples. All simulations are carried out with the Dutch national supercomputer Snellius (SURFsara), with 2 nodes (128 cores each) and 256 GB RAM. The results of the external pressures from the SAS and LES simulations are presented and discussed in terms of Cpe in Sec. III.

The ventilation system for the containment zone considered in this study is modeled with SYLVIA release 11.5 (IRSN, 2022,2023). A view of the network is presented in Fig. 8. As previously mentioned, the ventilation network is modeled as a combination of nodes and ventilation branches. The airflow through the ventilation components—assuming the component is horizontally disposed and the flow isothermal, incompressible, and one-dimensional—is solved using the kinetic energy theorem derived from the generalized Bernoulli equation of Eq. (8),
(8)
where L is the branch length (i.e., the length of a ventilation component), S is the branch section area, R is the airflow resistance, Q is the volumetric flow rate, ΔP is the static pressure difference, and the exponent n depends on the branch type and is equal to 2 for the airlocks and air inlets (treated as ducts), 1 for the NPUs with HEPA filter and 1.67 for the leaks (ASHRAE, 2001). For the nodes, assuming an isothermal flow, the mass balance is given by the following equation:
(9)
where V is the volume of the node, M is the molecular mass of air, T is the ambient temperature, r is the perfect gas constant, and i and j are the subscripts for inlet and outlet flows, respectively.
FIG. 8.

Ventilation network of the containment zone as modeled by means of SYLVIA.

FIG. 8.

Ventilation network of the containment zone as modeled by means of SYLVIA.

Close modal

The backward differential formula well suited for stiff equations (Asher and Petzold, 1998) is used for time discretization. Each time step is resolved by an iterative Newton–Raphson algorithm. The procedure described as follows is adopted for building the network and performing the simulations with SYLVIA:

  • Step 1: The containment zone is created as a room.

  • Step 2: All the ventilation components (air inlets, airlocks, leaks, and NPUs with filter) are modeled connecting the containment zone with a node representing the boundary with the exterior environment.

  • Step 3: The properties of the fluid (air in this case) and the characteristic properties of the containment zone (the initial depressurization and the volume) are set up. An initial depressurization value of −40 Pa is imposed in the containment zone.

  • Step 4: The geometrical properties (L and S) and the aeraulic characteristics (Q vs ΔP) of each ventilation component as well as the fan operating curve are input in SYLVIA.

  • Step 5: Volumetric flow into and out of the containment zone is checked to assure there is flow balance satisfying the continuity equation.

  • Step 6: Q and ΔP are used to establish the corresponding airflow resistances (R) and the flow exponent (n) for each component within SYLVIA using the following equation:
    (10)
  • Step 7: Initial conditions, in the absence of external wind effects, are established based on the inputs from steps 3 and 4, with the time (t) equal to zero in Eqs. (8) and (9).

  • Step 8: The Pe time series at the locations of the different ventilation components are used as a boundary condition and the resulting corresponding Pi time series solution is obtained.

Step 8 is performed three times, based on the Pe time series from WT and CFD (i.e., SAS and LES). A time step size of 1.25 × 10−3 s is used for the WT-S case corresponding to the frequency of 800 Hz (Sec. III C), while a time step size of 6.25 × 10−4 s is used for SAS-S and LES-S cases. This latter time step size is consistent with that used to run the SAS and LES simulations (see Sec. II D 2). Note that an additional LES-S simulation was done with the same time step as the WT-S and that this provided identical results to LES-S with the time step of 6.25 × 10−4 s. The simulations with SYLVIA are performed for a duration of 34.4 s (corresponding to 35 flow-through times in the CFD simulations (see Sec. II D 2). The results are presented in terms of the internal pressure coefficient (Cpi):
(11)
where Pi is the predicted internal pressure in the containment zone and Uref is the reference wind velocity from the WT and CFD simulations.

A grid-sensitivity analysis is performed with the LES approach. The results in terms of mean (Cpe,mean), root mean square (Cpe,rms), peak positive and negative (Cpe,max and Cpe,min) external pressure coefficients obtained from the coarse, basic and fine grids, for θ = 0° and 28 pressure locations (at a height of 0.7 h) are presented in Fig. 9. Then, the average absolute difference between (δc) coarse and basic and (δf) fine and basic is calculated for comparison and reported in Table I. It can be observed that Cpe,mean and Cpe,rms predicted by all the three grids are in good agreement with each other as evident from δc and δf of 0.02. A relatively good agreement in terms of Cpe,max between the basic and the fine grids (δf = 0.05) is also found, while there is an increased average absolute difference of δc = 0.07 between the coarse and the basic grids. However, notable deviations are observed for Cpe,min with increased differences of δc = 0.14 and δf = 0.12. Although the difference in Cpe,min between the fine grid and the basic grid is not negligible, the basic grid can be considered a good compromise between computational accuracy and computational costs. Therefore, the basic grid is used for further SAS and LES simulations.

FIG. 9.

Results of grid-sensitivity analysis for θ = 0° at 28 pressure locations at height 0.7 h: comparison of external pressure coefficients (Cpe) in terms of mean (Cpe,mean), root mean square (Cpe,rms), positive and negative peaks (Cpe,max and Cpe,min) for coarse, basic, and fine grids.

FIG. 9.

Results of grid-sensitivity analysis for θ = 0° at 28 pressure locations at height 0.7 h: comparison of external pressure coefficients (Cpe) in terms of mean (Cpe,mean), root mean square (Cpe,rms), positive and negative peaks (Cpe,max and Cpe,min) for coarse, basic, and fine grids.

Close modal
TABLE I.

Average absolute differences between (δc) coarse and basic and (δf) fine and basic grids, in terms of the statistics of the external pressure coefficient (Cpe): mean (Cpe,mean), root-mean square (Cpe,rms), and positive and negative peak values (Cpe,max and Cpe,min).

Cpe,mean Cpe,rms Cpe,max Cpe,min
δc  0.02  0.02  0.07  0.14 
δf  0.02  0.02  0.05  0.12 
Cpe,mean Cpe,rms Cpe,max Cpe,min
δc  0.02  0.02  0.07  0.14 
δf  0.02  0.02  0.05  0.12 
The LES index of resolution quality (LES_IQ) commonly used to measure the quality of the grid in LES simulations is computed for the chosen basic grid. LES_IQ by Celik (2005) is defined as the ratio of resolved turbulent kinetic energy (kresolved) to total kinetic energy (ktotal), as reported in the following equation:
(12)
where μ is the molecular dynamic viscosity and μresolved is the turbulent dynamic viscosity. Figure 10 shows contours of LES_IQ on the building surfaces and the ground. In the proximity of the building, its value is higher than 0.98. The minimum value of LES_IQ obtained throughout the whole domain is equal to 0.94, thus larger than the recommended value of 0.8 (Pope, 2000; Celik , 2005). Overall, the basic grid can be considered of sufficiently high resolution for performing the LES and SAS simulations.
FIG. 10.

Contours of LES_IQ for the basic grid: (a) in the vertical center plane through the building in the direction of flow and (b) in the wall-adjacent cell layer on the building and ground surfaces.

FIG. 10.

Contours of LES_IQ for the basic grid: (a) in the vertical center plane through the building in the direction of flow and (b) in the wall-adjacent cell layer on the building and ground surfaces.

Close modal

A visual comparison of the coherent turbulent structures in the flow by LES and SAS is made by means of the Q-criterion (Jeong and Hussain, 1995). The Q-criterion is defined as the second invariant of velocity gradient tensor, i.e., Q=12Ω2S2, where Ω2 represents the rotation rate and S2 represents the strain rate. It implies that a positive Q iso-surface is a representation of the areas where the strength of rotation overcomes the strain and can, thus, be considered as vortex envelopes (Dubief and Delcayre, 2000). Figure 11 shows the Q iso-surfaces (colored by instantaneous Cpe) corresponding to a Q value of 1.0 × 104 s−2 for both wind directions (θ = 0° and θ = 45°) for LES and SAS, for the time instance t = 34.4 s. It is evident that the small-scale turbulent structures of the flow field are not as well resolved in SAS as in LES, with major turbulence structures developing only in the proximity of the building. The possible effect of this on the prediction of Cpe is discussed in Sec. III C.

FIG. 11.

Coherent turbulent structures in the flow, shown by means of Q iso-surfaces (Q value of 1.0 × 104 s-2) colored by instantaneous Cpe: (a) LES for θ = 0°, (b) SAS for θ = 0°, (c) LES for θ = 45°, and (d) SAS for θ = 45°.

FIG. 11.

Coherent turbulent structures in the flow, shown by means of Q iso-surfaces (Q value of 1.0 × 104 s-2) colored by instantaneous Cpe: (a) LES for θ = 0°, (b) SAS for θ = 0°, (c) LES for θ = 45°, and (d) SAS for θ = 45°.

Close modal

This subsection discusses the validation of the LES and SAS results of the external pressure statistical parameters with the WT data. The comparison in terms of Cpe time series for both wind directions is presented in Fig. 12. Only a duration of 15 s is displayed for clarity. For θ = 0°, Figs. 12(a) and 12(c) show the results at positions S1 and S3 in a positive pressure region (windward facade) for θ = 0° and θ = 45°, respectively. Figures 12(b) and 12(d) show the results at S2 and S4 positioned in a negative pressure region, i.e., side facade for θ = 0° and leeward facade for θ = 45°. The CFD and WT time series show some similarities but also some clear discrepancies. The fluctuations at the windward facades (S1 and S3) are not well reproduced by LES and SAS for both θ = 0° and 45° [Figs. 12(a) and 12(c)]. The Cpe magnitude is significantly underpredicted by SAS when compared to WT, with a similar (though less pronounced) discrepancies observed in LES, particularly for θ = 45°. As observed in Fig. 11, this may be due to the missing turbulent structures in the incident flow and the small-scale turbulent structures not being resolved by SAS. The agreement between LES and SAS with WT is considerably better for S2 and S4 positioned in negative pressure zones [Figs. 12(b) and 12(d)]. However, an underprediction of the Cpe with SAS can still be observed for S4 from the upward shift of SAS time series in Fig. 12(d).

FIG. 12.

Comparison between CFD (SAS and LES) and WT in terms of Cpe time series at positions: (a) S1 for θ = 0°, (b) S2 for θ = 0°, (c) S3 for θ = 45°, and (d) S4 for θ = 45°.

FIG. 12.

Comparison between CFD (SAS and LES) and WT in terms of Cpe time series at positions: (a) S1 for θ = 0°, (b) S2 for θ = 0°, (c) S3 for θ = 45°, and (d) S4 for θ = 45°.

Close modal

Figures 13 and 14 allow a quantitative comparison between CFD and WT, in terms of mean (Cpe,mean), root mean square (Cpe,rms), positive peak (Cpe,max), and negative peak (Cpe,min) pressure coefficients, for θ = 0° and θ = 45°, respectively. The Cpe,mean and Cpe,rms [Figs. 13(a), 13(b), 14(a), and 14(b)] predicted by CFD show a good agreement with WT, with the majority of locations within a deviation of 15% from WT. On the contrary, the Cpe,max and Cpe,min [Figs. 13(c), 13(d), 14(c), and 14(d)] predicted by CFD show higher deviations from the WT results, at certain locations even larger than 30%. Such deviations in peak pressures were also reported by other researchers (e.g., Kataoka , 2020; Nozawa and Tamura, 2002; Selvam, 1997; and Wang and Chen, 2022), indicating that accurate prediction of peak Cpe is still a quite challenging task even when using more sophisticated CFD approaches as LES. One key reason for these discrepancies is the sensitivity of peak pressures to the turbulence characteristics of the CFD incident wind flow, including turbulence intensities, turbulent length scales, and power spectra of turbulence. Discrepancies in these characteristics between CFD and WT can significantly impact peak pressure predictions, as also observed by Lamberti and Gorlé (2020). These discrepancies arise primarily due to the inflow generation method used. The recent past has seen increased research aimed at improving inflow generation methods (e.g., Yan and Li, 2015; Lamberti , 2018; Vasaturo , 2018; Guichard, 2019; Patruno and De Miranda, 2020; Melaku and Bitsuamlak, 2021; and Potsis , 2024) to better replicate incident flow conditions at the building location. While the precursor method has demonstrated better representation of incident wind flow characteristics (Vasaturo , 2018), its high computational cost limits its practicality for industrial applications. Additionally, while the vortex method predicts the mean pressures relatively well, it can lead to unrealistic pressure fluctuations. Therefore, in this research study, the in-built spectral synthesizer inflow generation method in ANSYS Fluent 2021 R2 (Ansys Inc., 2021) has been used without any additional code manipulation, in order to investigate its performance for the intended industrial application. The subsequent section discusses the significance (or potential insignificance) of accurate peak Pe predictions for input in SYLVIA in determining the resulting Pi.

FIG. 13.

Comparison of (a) mean (Cpe,mean), (b) root mean square (Cpe,rms), (c) peak positive (Cpe,max), and (d) peak negative (Cpe,min) pressure coefficients at all locations of ventilation components between CFD (SAS and LES) and WT for θ = 0°.

FIG. 13.

Comparison of (a) mean (Cpe,mean), (b) root mean square (Cpe,rms), (c) peak positive (Cpe,max), and (d) peak negative (Cpe,min) pressure coefficients at all locations of ventilation components between CFD (SAS and LES) and WT for θ = 0°.

Close modal
FIG. 14.

Comparison of (a) mean (Cpe,mean), (b) root mean square (Cpe,rms), (c) peak positive (Cpe,max), and (d) peak negative (Cpe,min) pressure coefficients at all locations of ventilation components between CFD (SAS and LES) and WT, for θ = 45°.

FIG. 14.

Comparison of (a) mean (Cpe,mean), (b) root mean square (Cpe,rms), (c) peak positive (Cpe,max), and (d) peak negative (Cpe,min) pressure coefficients at all locations of ventilation components between CFD (SAS and LES) and WT, for θ = 45°.

Close modal

Overall, it can be noted that LES outperforms SAS in the prediction of Cpe statistics. Furthermore, the SAS approach proves to be less efficient in this case requiring more computational resources than the corresponding LES simulations, despite using similar computational settings (see Sec. II D 2) and the same grid (i.e., the basic), as also observed by Jadidi (2018). Specifically, the SAS simulation took 42 h to complete compared to the 29 h required for the LES simulation using the same computational resources already mentioned in Sec. II D.

This subsection compares the temporal variation in Pi within the containment volume as predicted by SYLVIA using the Pe time series from WT vs CFD as input. Figure 15 shows the Cpi time series for θ = 0° [Figs. 15(a) and 15(b)] and θ = 45° [Figs. 15(c) and 15(d)]. The comparison for θ = 0° shows that LES-S and SAS-S are in fairly good agreement with WT-S, despite the discrepancies observed in terms of Cpe (see Sec. III C). However, for θ = 45° the Cpi is underpredicted by both CFD-S cases, with the SAS-S showing a larger underprediction than LES-S. This trend is consistent with the Cpe results commented in Sec. III C. This is evident from Figs. 15(c) and 15(d) in the upward shift of the LES-S and especially the SAS-S time series.

FIG. 15.

Comparison of Cpi time series between: (a) LES-S and WT-S for θ = 0°, (b) SAS-S and WT-S for θ = 0° and (c) LES-S and WT-S for θ = 45°, (d) SAS-S and WT-S for θ = 45°.

FIG. 15.

Comparison of Cpi time series between: (a) LES-S and WT-S for θ = 0°, (b) SAS-S and WT-S for θ = 0° and (c) LES-S and WT-S for θ = 45°, (d) SAS-S and WT-S for θ = 45°.

Close modal

Figure 16 shows the comparison in terms of Cpi statistics—i.e., mean, root mean square (rms), peak maximum (max), and peak minimum (min)—for the WT and CFD cases coupled with SYLVIA. The experimental uncertainty of the WT data are also reported in the figure. This uncertainty is estimated based on Cpe, by comparing the magnitudes of Cpe,mean, Cpe,rms, Cpe,max, and Cpe,min at symmetric locations on the cube. The 90th percentile of all absolute differences in magnitude is then selected as the uncertainty. In general, the Cpi statistics from LES-S and SAS-S show a satisfactory agreement with WT-S data for both θ = 0° [Fig. 16(a)] and θ = 45° [Fig. 16(b)]. The mean and rms of Cpi by LES-S and SAS-S show a difference as small as 0.01 from the WT-S data. The Cpi values from CFD-S show a larger deviation in terms of the max and min with respect to the WT-S. However, the absolute differences in terms of Cpi mean, rms, and peak between WT-S and CFD-S cases fall within the uncertainties associated with WT data (see Fig. 16) in spite of some clear deficiencies in reproducing the WT Cpe data by LES and especially SAS, as discussed in Subsection III C. This indicates that the prediction of the internal pressure statistics is not always very sensitive to the aforementioned deviations in the Cpe data used as input. This is because the Cpi value at any time step results from the combined effect of all Cpe inputs and the instantaneous differences (i.e., the differences in fluctuations and individual peaks) tend to be compensated rather than accumulate. Moreover, as discussed in Sec. III C, significant discrepancies in Cpe are observed only at positions within the positive pressure region (i.e., the windward facade), while in all other regions, the results remain comparable.

FIG. 16.

Internal pressure statistics in terms of mean, rms, max, and min of Cpi predicted by LES-S and SAS-S compared with WT-S for (a) θ = 0° and (b) θ = 45°.

FIG. 16.

Internal pressure statistics in terms of mean, rms, max, and min of Cpi predicted by LES-S and SAS-S compared with WT-S for (a) θ = 0° and (b) θ = 45°.

Close modal

The primary objective of using SYLVIA for a pollutant containment zone is to predict the possible breaches in containment due to wind effects. This breach occurs when the local external pressure is lower than the internal pressure. In this subsection, the containment breach is quantified for all the locations of the ventilation components as the percentage duration for which Cpi exceeds Cpe with respect to the total duration of the SYLVIA simulations. Figure 17 compares the duration of breach by WT-S, LES-S, and SAS-S for θ = 0° and θ = 45°. The locations of the breaches (in red) on the building are also shown in Fig. 17. The breaches occur in regions of flow separation characterized by pronounced negative external pressure, as on facades 1–2 and 3–0 for θ = 0° and on facades 2-3 and 3-0 for θ = 45° (Jayakumari , 2023). This observation is consistent for all three cases (WT-S, SAS-S, and LES-S), with breach occurrence at all locations along facades 1-2 and 3-0 for θ = 0° and on facades 2-3 and 3-0 for θ = 45° as marked in Fig. 17. In general, the percentage duration of the breaches predicted by LES-S and SAS-S is comparable to WT-S with an average difference in breach duration of 4% and 5%, respectively. The largest deviation in percentage breach duration between LES-S and WT-S is 10% for θ = 0° (at A36) and 7% for θ = 45° (at A35 and A37), while for the SAS-S, this is about 22% for θ = 0° (at L3) and 14% for θ = 45° (at A35). For θ = 45°, CFD-S is found to consistently underpredict the breach percentage, with LES-S predicting values closer to WT-S than SAS-S. This is a reflection of the underprediction of Cpi as discussed in Subsection III D. Thus, LES-S outperforms SAS-S for the prediction of percentage breach duration.

FIG. 17.

Percentage duration of containment breach at the location of the ventilation components predicted by LES-S and SAS-S compared with WT-S for (a) θ = 0° and (b) θ = 45°. The locations of containment breach occurrence are also marked in red in the isometric view of the building.

FIG. 17.

Percentage duration of containment breach at the location of the ventilation components predicted by LES-S and SAS-S compared with WT-S for (a) θ = 0° and (b) θ = 45°. The locations of containment breach occurrence are also marked in red in the isometric view of the building.

Close modal

Overall, the agreement between CFD-S and WT-S is acceptable for industrial applications when considering the possible uncertainties of measurements as well as simulations.

Hazardous pollutant containment zones should be maintained at a pressure lower than the outdoor atmospheric pressure to prevent pollutants from escaping to the outdoor environment. However, atmospheric wind conditions can cause breaching of the containment zone that is established through mechanical ventilation. This paper combines external wind pressure (Pe) time series on an internally depressurized building with a carefully designed ventilation network to analyze indoor pressure (Pi) and containment breach duration. The Pe time series and statistical parameters are obtained by reduced-scale wind tunnel (WT) tests performed on an idealized isolated building on the one hand, and by computational fluid dynamics (CFD simulations) by large-eddy simulation (LES) and scale-adaptive simulation (SAS) on the other hand. These Pe time series are used as input to the ventilation network model SYLVIA. The depressurization value used is −40 Pa, and the reference wind speed (Uref) at building height is 12.65 m/s. The goal of the paper is threefold: (1) comparing the Pe data as obtained by WT and CFD in terms of time series and statistics; (2) assessing the impact of uncertainties in the Pe as generated by CFD on the resulting Pi and on the breach duration percentage; and (3) estimating the Pi statistics and breach duration percentage for an initial case study with the indoor depressurization of −40 Pa and a fixed Uref of 12.65 m/s at building height. The objective (2) aims to determine whether, despite the inherent uncertainties in CFD results, the combinations LES-SYLVIA (LES-S) or SAS-SYLVIA (SAS-S) could be used as alternatives to the WT-SYLVIA (WT-S) combination in future studies. The results are discussed in terms of dimensionless pressure coefficients (Cpe and Cpi). It is worth noting that the study has the following limitations:

  1. Only two wind directions (i.e., θ = 0° and θ = 45°) have been investigated. While these are two representative wind directions for the present building, for more geometrically complex buildings with less symmetry, additional wind directions should be considered.

  2. The SAS simulations have been performed with the same settings as LES. However, a fine grid near the walls (as for LES) might not be needed for SAS, since RANS is applied in these regions. Therefore, the observations that SAS requires a longer computational time than LES should not be generalized. Further simulations could be performed in order to optimize the SAS grid by decreasing the computational time and improving the results.

  3. This study has been performed for an idealized building, for a fixed depressurization value and for a fixed reference wind speed of 12.65 m/s.

Despite the above listed limitations, the study has reached its objectives based on which future studies for more complex building geometries, different depressurization values, different reference wind speeds, and wind directions can be considered. The following conclusions can be drawn from the study.

  1. Mean and rms Cpe predicted by LES and SAS simulations are comparable to those predicted by WT tests; however, substantial deviations are observed for the maximum and minimum peak Cpe by LES and especially by SAS compared to the WT results.

  2. In spite of the deviations in Cpe, the Cpi time series predicted by LES-S and SAS-S matches fairly well with that by WT-S for θ = 0°. Larger discrepancies, however, are identified for SAS-S for θ = 45°, consistent with those detected for Cpe.

  3. Nevertheless, statistics of the Cpi predicted by LES-S and SAS-S show a rather good agreement with that predicted by WT-S. The absolute differences in terms of Cpi mean, rms, and peak between WT-S and CFD-S cases fall within the uncertainties associated with WT measurements. The largest observed absolute difference in Cpi statistics is 0.1, corresponding to the peak Cpi prediction when θ = 45°. A main source of uncertainty in the LES and SAS simulations is the use of the spectral synthesizer as inflow generation method that has been used as is, i.e., without any additional code manipulation. The rather good agreement in terms of Cpi indicates that its prediction is not very sensitive to the aforementioned deviations in the Cpe data used as input.

  4. The percentage duration of a possible containment breach predicted by LES-S and SAS-S shows an overall fair agreement with that predicted by WT-S, with an average difference of 4% and 5%, respectively, from WT-S. LES-S is found to slightly outperform SAS-S in the prediction of containment breach probability, as higher differences between SAS-S and WT-S are observed at localized regions for θ = 0° and at all locations for θ = 45° compared to that between LES-S and WT-S. This suggests that in future studies, if needed and if more convenient, LES-S can be used instead of WT-S.

  5. For this initial case study with the −40 Pa depressurization and Uref = 12.65 m/s at the building height, the breach duration percentage can go up to 80%, which is a reason for concern and which stresses the importance of this type of studies.

The project was undertaken in collaboration with Institut National de Recherche et de Sécurité (INRS), France, and their support throughout this project is greatly appreciated. The authors acknowledge the partnership with ANSYS CFD. They also acknowledge the support received from the Institut de radioprotection et de sûreté nucléaire (IRSN) in relation to SYLVIA. The CFD simulations were carried out on the Dutch national e-infrastructure with the support of SURF Cooperative (Grant Nos. EINF-816 and NWO-2021.027/L1), and their support is greatly acknowledged.

The authors have no conflicts to disclose.

A. K. R. Jayakumari: Data curation (lead); Formal analysis (lead); Visualization (lead); Writing – original draft (lead). A. Ricci: Conceptualization (equal); Investigation (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal). R. Guichard: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal). S. Gillmeier: Conceptualization (equal); Investigation (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal). B. Blocken: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
Axley
,
J.
, “
Multizone airflow modeling in buildings: History and theory
,”
HVAC&R Res.
13
,
907
928
(
2007
).
2.
American Society of Civil Engineers (ASCE)
, “
ASCE standard, minimum design loads for buildings and other structures
” Standard No. ASCE 7-02 (
ASCE
,
Reston, VA
,
2002
).
3.
Ansys Inc.
,
Ansys Fluent User's Guide
(
Ansys Inc
.,
Canonsburg, PA
,
2021
).
4.
ASHRAE
,
ASHRAE Handbook: Fundamentals
, SIed. (
American Society of Heating, Refrigerating and Air-Conditioning Engineers Inc.
,
2001
), p.
26
.
5.
Asher
,
U. M.
and
Petzold
,
L. R.
,
Computer Methods for Ordinary Differential Equations and Differential-Algebraic Equations
, 1st ed. (
SIAM
,
Philadelphia, PA
,
1998
).
6.
Blocken
,
B.
, “
50 years of computational wind engineering: Past, present and future
,”
J. Wind Eng. Ind. Aerodyn.
129
,
69
102
(
2014
).
7.
Blocken
,
B.
, “
Computational fluid dynamics for urban physics: Importance, scales, possibilities, limitations, and ten tips and tricks towards accurate and reliable simulations
,”
Build. Environ.
91
,
219
245
(
2015
).
8.
Blocken
,
B.
, “
LES over RANS in building simulation for outdoor and indoor applications: A foregone conclusion?
,”
Build. Simul.
11
(
5
),
821
870
(
2018
).
9.
Blocken
,
B.
,
Stathopoulos
,
T.
, and
Carmeliet
,
J.
, “
CFD simulation of the atmospheric boundary layer: Wall function problems
,”
Atmos. Environ.
41
,
238
252
(
2007
).
10.
Blocken
,
B.
,
Stathopoulos
,
T.
, and
Carmeliet
,
J.
, “
Wind environmental conditions in passages between two long narrow perpendicular buildings
,”
J. Aerosp. Eng.
21
(
4
),
280
287
(
2008
).
11.
Castro
,
I. P.
and
Robins
,
A. G.
, “
The flow around a surface-mounted cube in uniform and turbulent streams
,”
J. Fluid Mech.
79
,
307
335
(
1977
).
12.
Celik
,
I. B.
,
Cehreli
,
Z. N.
, and
Yavuz
,
I.
, “
Index of resolution quality for large eddy simulations
,”
J. Fluids Eng.
127
,
949
958
(
2005
).
13.
Cermak
,
J. E.
, “
Applications of fluid mechanics to wind engineering—A Freeman scholar lecture
,”
J. Fluids Eng.
97
,
9
(
1975
).
14.
D'Angelo
,
W. C.
,
Spicer
,
R. C.
, and
Mease
,
M. J.
,
Asbestos Removal in Occupied Buildings: Sophisticated Procedures for Structures With Operating HVAC Systems
, 1st ed. (
SourceFinders Information Corporation
,
New Jersey
,
1987
).
15.
Dols
,
W. S.
and
Polidoro
,
B.
,
CONTAM User Guide and Program Documentation: Version 3.2
,
US Department of Commerce, National Institute of Standards and Technology
(
2015
).
16.
Dubief
,
Y.
and
Delcayre
,
F.
, “
On coherent-vortex identification in turbulence
,”
J. Turbul.
1
,
N11
(
2000
).
17.
Cebeci
,
T.
and
Bradshaw
,
P.
,
Momentum Transfer in Boundary Layers
(
Hemisphere Publishing
,
New York
,
1997
).
18.
Cermak
,
J. E.
, “
Wind-tunnel development and trends in applications to civil engineering
,”
J. Wind Eng. Ind. Aerodyn.
91
,
355
(
2003
).
19.
Chen
,
Q.
, “
Ventilation performance prediction for buildings: A method overview and recent applications
,”
Build. Environ.
44
,
848
858
(
2009
).
20.
Cheung
,
J. C. K.
,
Holmes
,
J. D.
,
Melbourne
,
W. H.
,
Lakshmanan
,
N.
, and
Bowditch
,
P.
, “
Pressures on a 110 scale model of the Texas Tech building
,”
J. Wind Eng. Ind. Aerodyn.
69–71
,
529
538
(
1997
).
21.
Cook
,
N. J.
and
Mayne
,
J. R.
, “
A novel working approach to the assessment of wind loads for equivalent static design
,”
J. Wind Eng. Ind. Aerodyn.
4
,
149
164
(
1979
).
22.
Cóstola
,
D.
,
Blocken
,
B.
, and
Hensen
,
J. L. M.
, “
Overview of pressure coefficient data in building energy simulation and airflow network programs
,”
Build. Environ.
44
,
2027
2036
(
2009
).
23.
Coutin
,
M.
et al, “
Characterisation of open-door electrical cabinet fires in compartments
,”
Nucl. Eng. Des.
286
,
104
(
2015
).
24.
Davenport
,
A. G.
, “
Rationale for determining design wind velocities
,”
J. Struct. Div.
86
,
39
68
(
1960
).
25.
Dubernet
,
F.
,
Guichard
,
R.
, and
Romero-Hariot
,
A.
,
Amiante. Aéraulique des Chantiers Sous Confinement, Guide Pratique de Ventilation
, 1st ed. (
Institut National de Recherche et de Sécurité
,
2018
).
26.
Feustel
,
H. E.
, “
COMIS—An international multizone air-flow and contaminant transport model
,”
Build. Environ.
30
,
3
18
(
1999
).
27.
Franke
,
J.
,
Hellsten
,
A.
,
Schlünzen
,
H.
, and
Carissimo
,
B.
,
Best Practice Guideline for the CFD Simulation of Flows in the Urban Environment
(
COST Office
,
Brussels
,
2007
).
28.
Fluent Inc.
,
Fluent 6.2 User's Guide
(
Fluent Inc
.,
Lebanon, NH
,
2005
).
29.
García
,
J.
,
Muñoz-Paniagua
,
J.
,
Jiménez
,
A.
,
Migoya
,
E.
, and
Crespo
,
A.
, “
Numerical study of the influence of synthetic turbulent inflow conditions on the aerodynamics of a train
,”
J. Fluids Struct.
56
,
134
151
(
2015
).
30.
Great Britain, Health and Safety Executive (HSE)
,
Asbestos: The Licensed Contractors Guide
, 1st ed. (
Great Britain, Health and Safety Executive
,
2006
).
31.
Guichard
,
R.
, “
Assessment of an improved Random Flow Generation method to predict unsteady wind pressures on an isolated building using Large-Eddy simulation
,”
J. Wind Eng. Ind. Aerodyn.
189
,
304
313
(
2019
).
32.
Hunt
,
A.
, “
Wind-tunnel measurements of surface pressures on cubic building models at several scales
,”
J. Wind Eng. Ind. Aerodyn.
10
,
137
163
(
1982
).
33.
Holmes
,
J. D.
,
Wind Loading of Structures
, 3rd ed. (
Taylor & Francis
,
2004
).
34.
Institut de Radioprotection et de Sûreté Nucléaire (IRSN)
, see https://en.irsn.fr/en/research/scientific-tools/computer-codes/pages/calculation-software-system-sylvia-4577.aspx for “
Calculation Software System SYLVIA
” (
2023
) (accessed July 17, 2023).
35.
Institut de Radioprotection et de Sûreté Nucléaire (IRSN)
,
Physical Modelling of the SYLVIA V11 Software
(
IRSN
,
Paris, France
,
2022
).
36.
Jadidi
,
M.
,
Bazdidi-Tehrani
,
F.
, and
Kiamansouri
,
M.
, “
Scale-adaptive simulation of unsteady flow and dispersion around a model building: Spectral and POD analyses
,”
J. Build. Perform. Simul.
11
,
241
260
(
2018
).
37.
Jayakumari
,
A. K. R.
,
Guichard
,
R.
,
Ricci
,
A.
,
Gillmeier
,
S.
, and
Blocken
,
B.
, “
Towards an improved wind effect assessment for asbestos abatement: A methodology for reduced-scale experiments
,”
J. Environ. Manage.
370
,
122698
(
2024
).
38.
Jayakumari
,
A. K. R.
,
Guichard
,
R.
,
Gillmeier
,
S.
,
Ricci
,
A.
, and
Blocken
,
B.
, “
Atmospheric wind effects on depressurization for indoor asbestos pollution containment: Experimental analysis and a ventilation network model validation
,”
Build. Environ.
272
,
112664
(
2025
).
39.
Jayakumari
,
A. K. R.
,
Gillmeier
,
S.
,
Ricci
,
A.
,
Blocken
,
B.
, and
Guichard
,
R.
, “
Scaling effects on experimentally obtained pressures on an idealized building: Possible implications towards asbestos containment
,”
J. Wind Eng. Ind. Aerodyn.
239
,
105442
(
2023
).
40.
Jeong
,
J.
and
Hussain
,
F.
, “
On the identification of a vortex
,”
J. Fluid Mech.
285
,
69
94
(
1995
).
41.
Kataoka
,
H.
,
Ono
,
Y.
, and
Enoki
,
K.
, “
Applications and prospects of CFD for wind engineering fields
,”
J. Wind Eng. Ind. Aerodyn.
205
,
104310
(
2020
).
42.
KTA-Geschäftsstelle
,
Lüftungstechnische Anlagen in Kernkraftwerken
(
Sicherheitstechnische Regel des KTA
,
2017
), p.
7
.
43.
Lamberti
,
G.
and
Gorlé
,
C.
, “
Sensitivity of LES predictions of wind loading on a high-rise building to the inflow boundary condition
,”
J. Wind Eng. Ind. Aerodyn.
206
,
104370
(
2020
).
44.
Lamberti
,
G.
,
García-Sánchez
,
C.
,
Sousa
,
J.
, and
Gorlé
,
C.
, “
Optimizing turbulent inflow conditions for large-eddy simulations of the atmospheric boundary layer
,”
J. Wind Eng. Ind. Aerodyn.
177
,
32
44
(
2018
).
45.
Lamberti
,
G.
,
Amerio
,
L.
,
Pomaranzi
,
G.
,
Zasso
,
A.
, and
Gorlé
,
C.
, “
Comparison of high-resolution pressure measurements on a high-rise building in a closed and open-section wind tunnel
,”
J. Wind Eng. Ind. Aerodyn.
204
,
104247
(
2020
).
46.
Launder
,
B. E.
, and
Spalding
,
D. B.
, “
The numerical computation of turbulent flows
,”
Comput. Methods. Appl. Mech. Eng.
3
,
269
289
(
1974
).
47.
Le Dez
,
T.
,
Richard
,
J.
,
Inard
,
C.
,
Le Roux
,
N.
,
Demouge
,
F.
,
Faure
,
X.
, and
Ricciardi
,
L.
, “
Reduced-scale study of the coupling between thermal and wind effects on the ventilation systems of nuclear facilities
,”
Int. J. Vent.
20
,
1
19
(
2021
).
48.
Le Roux
,
N.
,
Faure
,
X.
,
Inard
,
C.
,
Soares
,
S.
, and
Ricciardi
,
L.
, “
Reduced-scale study of transient flows inside mechanically ventilated buildings subjected to wind and internal overpressure effects
,”
Build. Environ.
62
,
18
32
(
2013
).
49.
Le Roux
,
N.
,
Faure
,
X.
,
Inard
,
C.
,
Soares
,
S.
, and
Ricciardi
,
L.
, “
Reduced-scale study of wind influence on mean airflows inside buildings equipped with ventilation systems
,”
Build. Environ.
58
,
231
(
2012
).
50.
Legifrance
, see https://www.legifrance.gouv.fr/loda/id/JORFTEXT000027324535/ for “
Arrêté du 8 Avril 2013 Relatif aux Règles Techniques, Aux Mesures de Prévention et Aux Moyens de Protection Collective à Mettre en OEuvre par les Entreprises lors D'opérations Comportant un Risque D'Exposition à L'amiante
” (
2013
) (accessed October 9, 2021).
51.
Levitan
,
M. L.
and
Mehta
,
K. C.
, “
Texas Tech field experiments for wind loads. Part I. Building and pressure measuring system
,”
J. Wind Eng. Ind. Aerodyn.
43
,
1565
1576
(
1992
).
52.
Li
,
J. C.
,
Hu
,
S. Y.
, and
Li
,
Q. S.
, “
Comparative study of full-scale and model-scale wind pressure measurements on a gable roof low-rise building
,”
J. Wind Eng. Ind. Aerodyn.
208
,
104448
(
2021
).
53.
Lin
,
J. X.
,
Surry
,
D.
, and
Tieleman
,
H. W.
, “
The distribution of pressure near roof corners of flat roof low buildings
,”
J. Wind Eng. Ind. Aerodyn.
56
,
235
265
(
1995
).
54.
Mateus
,
N. M.
,
Simões
,
G. N.
,
Lúcio
,
C.
, and
da Graça
,
G. C.
, “
Comparison of measured and simulated performance of natural displacement ventilation systems for classrooms
,”
Energy Build.
133
,
185
196
(
2016
).
55.
Melaku
,
A. F.
and
Bitsuamlak
,
G. T.
, “
A divergence-free inflow turbulence generator using spectral representation method for large-eddy simulation of ABL flows
,”
J. Wind Eng. Ind. Aerodyn.
212
,
104580
(
2021
).
56.
Menter
,
F. R.
and
Egorov
,
Y.
, “
The scale-adaptive simulation method for unsteady turbulent flow predictions. Part 1: Theory and model description
,”
Flow, Turbul. Combust.
85
,
113
138
(
2010
).
57.
Menter
,
F. R.
,
Kuntz
,
M.
, and
Langtry
,
R.
, “
Ten years of industrial experience with the SST turbulence model
,” in
Proceedings of the Fourth International Symposium on Turbulence, Heat and Mass Transfer
(
Begell House, Inc.
,
Antalya, Turkey
,
2003
), pp.
625
632
.
58.
Ministerie van Sociale Zaken en Werkgelegenheid (SZW)
,
Artikel 43 (Asbestverwijdering, Punt 3) van Bijlage XIIIa (Werkveldspecifiek Certificatieschema voor de Procescertificaten Asbestinventarisatie en Asbestverwijdering
(
Ministerie van Sociale Zaken en Werkgelegenheid
,
2016
).
59.
Montazeri
,
H.
and
Blocken
,
B.
, “
CFD simulation of wind-induced pressure coefficients on buildings with and without balconies: Validation and sensitivity analysis
,”
Build. Environ.
60
,
137
149
(
2013
).
60.
Nicoud
,
F.
and
Ducros
,
F.
, “
Subgrid-scale stress modelling based on the square of the velocity gradient tensor
,”
Flow, Turbul. Combust.
62
,
183
200
(
1999
).
61.
Nozawa
,
K.
and
Tamura
,
T.
, “
Large eddy simulation of the flow around a low-rise building immersed in a rough-wall turbulent boundary layer
,”
J. Wind Eng. Ind. Aerodyn.
90
,
1151
1162
(
2002
).
62.
Occupational Safety and Health Administration (OSHA)
,
Asbestos, Code of Federal Regulations, 29 CFR 1910.1001 and 29 CFR 1926.58
(
Occupational Safety and Health Administration
,
2019
).
63.
Ono.
Y.
,
Tamura.
T.
, and
Kataoka.
H.
, “
LES analysis of unsteady characteristics of conical vortex on a flat roof
,”
J. Wind Eng. Ind. Aerodyn.
96
(10–11),
2007
2018
(
2008
).
64.
Patruno
,
L.
and
de Miranda
,
S.
, “
Unsteady inflow conditions: A variationally based solution to the insurgence of pressure fluctuations
,”
Comput. Methods Appl. Mech. Eng.
363
,
112894
(
2020
).
65.
Pope
,
S.
,
Turbulent Flows
(
Cambridge University Press
,
Cambridge
,
2000
), p.
802
.
66.
Potsis
,
T.
,
Ricci
,
A.
, and
Stathopoulos
,
T.
, “
On the reliability of the dynamic terrain method to generate ABL winds for environmental applications
,”
Meccanica
2024
,
1
.
67.
Potsis
,
T.
,
Tominaga
,
Y.
, and
Stathopoulos
,
T.
, “
Computational wind engineering: 30 years of research progress in building structures and environment
,”
J. Wind Eng. Ind. Aerodyn.
234
,
105346
(
2023
).
68.
Qin
,
P.
,
Ricci
,
A.
, and
Blocken
,
B.
, “
Modeling traffic pollutants in a street canyon by CFD: Idealized line sources versus multiple realistic car sources
,”
Sci. Total Environ.
955
,
177099
(
2024
).
69.
Rezaeiha
,
A.
,
Montazeri
,
H.
, and
Blocken
,
B.
, “
CFD analysis of dynamic stall on vertical axis wind turbines using Scale-Adaptive Simulation (SAS): Comparison against URANS and hybrid RANS/LES
,”
Energy Convers. Manage.
196
,
1282
1298
(
2019
).
70.
Ricci
,
A.
, “
Review of OpenFOAM applications in the computational wind engineering: From wind environment to wind structural engineering
,”
Meccanica
2024
,
1
.
71.
Ricci
,
M.
,
Patruno
,
L.
,
Kalkman
,
I.
,
de Miranda
,
S.
, and
Blocken
,
B.
, “
Towards LES as a design tool: Wind loads assessment on a high-rise building
,”
J. Wind Eng. Ind. Aerodyn.
180
,
1
18
(
2018
).
72.
Richards
,
P. J.
and
Hoxey
,
R. P.
, “
Appropriate boundary conditions for computational wind engineering models using the k-ε turbulence model
,”
J. Wind Eng. Ind. Aerodyn.
46-47
,
145
153
(
1993
).
73.
Richards
,
P. J.
,
Hoxey
,
R. P.
, and
Short
,
L. J.
, “
Wind pressures on a 6m cube
,”
J. Wind Eng. Ind. Aerodyn.
89
,
1553
1564
(
2001
).
74.
Selvam
,
R. P.
, “
Computation of pressures on Texas Tech University building using large eddy simulation
,”
J. Wind Eng. Ind. Aerodyn.
67–68
,
647
657
(
1997
).
75.
Simiu
,
E.
, and
Scanlan
,
R. H.
,
Wind Effects on Structures: Fundamentals and Applications to Design
, 3rd ed. (
A Wiley-Interscience Publication
,
1996
).
76.
Smirnov
,
A.
,
Shi
,
S.
, and
Celik
,
I.
, “
Random flow generation technique for large eddy simulations and particle-dynamics modeling
,”
J. Fluids Eng.
123
,
359
371
(
2001
).
77.
Stathopoulos
,
T.
, “
The numerical wind tunnel for industrial aerodynamics: Real or virtual in the new millennium?
,”
Wind Struct.
5
,
193
208
(
2002
).
78.
Tominaga
,
Y.
and
Blocken
,
B.
, “
Wind tunnel analysis of flow and dispersion in cross-ventilated isolated buildings: Impact of opening positions
,”
J. Wind Eng. Ind. Aerodyn.
155
,
74
88
(
2016
).
79.
Tominaga
,
Y.
and
Blocken
,
B.
, “
Wind tunnel experiments on cross-ventilation flow of a generic building with contaminant dispersion in unsheltered and sheltered conditions
,”
Build. Environ.
92
,
452
461
(
2015
).
80.
Tominaga
,
Y.
,
Mochida
,
A.
,
Yoshie
,
R.
,
Kataoka
,
H.
,
Nozu
,
T.
,
Yoshikawa
,
M.
, and
Shirasawa
,
T.
, “
AIJ guidelines for practical applications of CFD to pedestrian wind environment around buildings
,”
J. Wind Eng. Ind. Aerodyn.
96
,
1749
1761
(
2008
).
81.
van Druenen
,
T.
and
Blocken
,
B.
, “
Aerodynamic impact of cycling postures on drafting in single paceline configurations
,”
Comput. Fluids
257
,
105863
(
2023
).
82.
Vasaturo
,
R.
,
Kalkman
,
I.
,
Blocken
,
B.
, and
van Wesemael
,
P. J. V.
, “
Large eddy simulation of the neutral atmospheric boundary layer: Performance evaluation of three inflow methods for terrains with different roughness
,”
J. Wind Eng. Ind. Aerodyn.
173
,
241
261
(
2018
).
83.
Vaux
,
S.
and
Prétrel
,
H.
, “
Relative effects of inertia and buoyancy on smoke propagation in confined and forced ventilated enclosure fire scenarios
,”
Fire Saf. J.
62
,
206
220
(
2013
).
84.
VDI-3783
, “
Environmental meteorology—Physical modeling of flow and dispersion processes in the atmospheric boundary layer application of wind tunnels
,” (
Engl. VDI/DIN-Kommission Reinhaltung der Luft (KRdL) - Normenausschuss
,
2000
).
85.
Wang
,
Y.
and
Chen
,
X.
, “
Evaluation of wind loads on high-rise buildings at various angles of attack by wall-modeled large-eddy simulation
,”
J. Wind Eng. Ind. Aerodyn.
229
,
105160
(
2022
).
86.
Wilcox
,
D. C.
,
Turbulence Modeling for CFD
(
DCW Industries, Inc.
,
La Canada, California
,
1994
).
87.
World Health Organization Regional Office for Europe
,
National Programmes for Elimination of Asbestos-Related Diseases: Review and Assessment
(
WHO Regional Office for Europe
,
Copenhagen
,
2012
).
88.
Xing
,
J.
,
Patruno
,
L.
,
De Miranda
,
S.
,
Pinardi
,
S.
,
Majowiecki
,
M.
, and
Ubertini
,
F.
, “
Early stages wind load assessment using Computational Fluid Dynamics: The new Bologna Stadium roof
,”
Structures
47
,
1912
1926
(
2023
).
89.
Xu
,
M.
,
Patruno
,
L.
, and
de Miranda
,
S.
, “
A pressure–velocity jump approach for the CFD modelling of permeable surfaces
,”
J. Wind Eng. Ind. Aerodyn.
233
,
105317
(
2023
).
90.
Yan
,
B. W.
and
Li
,
Q. S.
, “
Inflow turbulence generation methods with large eddy simulation for wind effects on tall buildings
,”
Comput. Fluids
116
,
158
175
(
2015
).
91.
Žužul
,
J.
,
Ricci
,
A.
,
Burlando
,
M.
,
Blocken
,
B.
, and
Solari
,
G.
, “
CFD analysis of the WindEEE dome produced downburst-like winds
,”
J. Wind Eng. Ind. Aerodyn.
232
,
105268
(
2023
).