Next Article in Journal
Study of the Seismic Behavior of Simplified RCS Joints via Nonlinear Finite Element Analysis
Next Article in Special Issue
Exploring the Implementation Path of Passive Heat-Protection Design Heritage in Lingnan Buildings
Previous Article in Journal
Towards an Evaluation System of Disabled Individuals’ Friendly Communities from the Perspective of Inclusive Development—A Case Study in Jinan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CFD Analysis of Building Cross-Ventilation with Different Angled Gable Roofs and Opening Locations

School of Architecture and Urban Planning, Chongqing Jiaotong University, Chongqing 400074, China
*
Author to whom correspondence should be addressed.
Buildings 2023, 13(11), 2716; https://doi.org/10.3390/buildings13112716
Submission received: 21 September 2023 / Revised: 5 October 2023 / Accepted: 26 October 2023 / Published: 27 October 2023
(This article belongs to the Special Issue Indoor Climate and Energy Efficiency in Buildings)

Abstract

:
The geometric shape of the roof and the opening position are important parameters influencing the internal cross-ventilation of buildings. Although there has been extensive research on natural ventilation, most of it has focused on flat or sloping roofs with the same opening positions. There is still limited research on the impact of different opening positions and sloping roofs on natural ventilation. In this study, computational fluid dynamics (CFD) was used to investigate the air exchange efficiency (AEE) in general isolated buildings. These buildings encompassed three distinct opening configurations (top–top, top–bottom, and bottom–top) and six varying slope angles for gable roofs (0°, 9°, 18°, 27°, 36°, and 45°). Computational simulations were carried out using the SST k-omega turbulence model, and validation was performed against experimental data supplied by the Japanese AIJ Wind Tunnel Laboratory. Grid independence validation was also conducted to ensure the reliability of the CFD simulation results. The study revealed that the highest AEE was 48.1%, achieved with the top–bottom opening configuration and a gable roof slope angle of 45°. Conversely, the lowest AEE was 31.4%, attained with the bottom–top opening configuration and a gable roof slope angle of 27°. Furthermore, it was observed that when the opening configuration was set to top–top and bottom–top, the slope angle of the gable roof had minimal influence on AEE, with an average AEE of only around 33%. When the opening configuration was top–bottom, it was found that there was a positive correlation between the gable roof slope angle and AEE.

1. Introduction

The global demand for energy has been continually increasing, with the construction industry playing a significant role. Therefore, enhancing its energy efficiency becomes especially important. An overreliance on mechanical and electrical equipment to provide a comfortable environment has led to a disproportionate energy consumption in buildings, particularly from heating, ventilation, and air conditioning (HVAC) systems [1]. In China, between 2005 and 2020, the national total energy consumption from buildings increased from 930 million tce to 2.233 billion tce, with an average annual growth rate of 6.0%. By the end of 2020, the building stock in China had reached 69.6 billion square meters [2]. With such an immense inventory, reduction in building energy consumption has remained a key concern for both the government and society. With mild weather conditions, it is highly recommended to use natural ventilation to replace air conditioning for cooling [3]. Compared with air conditioning systems, the cooling ability of natural ventilation seems to be limited and its performance is also not quite stable. Therefore, many researchers have moved to studying how to maximize the performance of natural ventilation in cooling buildings.
Regarding opening positions, natural ventilation can be generally categorized into three types: (1) SSV—single-sided ventilation (one or more vents are placed on one side of the building); (2) CV—cross-ventilation (allowing natural airflow through the building by having windows positioned on opposite facades), and (3) SV—stack ventilation (utilizing temperature difference between indoor and outdoor environments, with vents located at different heights) [4]. Compared with SSV, CV generally gives higher ventilation efficiency [5], hence it was selected in this study.
In terms of opening positions in the vertical direction, according to Karava [6], apart from wall porosity, the relative locations of openings and exits on building facades are significant factors affecting cross-ventilation. Zhang [7] conducted a study on two different-sized external openings and found that the most effective configuration for enhancing cross-ventilation was to have a larger opening on the leeward side than on the windward side. The aforementioned studies primarily focused on the impact of opening size and position on cross-ventilation, with limited exploration of the building’s form. However, as the roof is an integral component of the building’s form, its significance is self-evident. Roof forms can be categorized into gable roofs [8,9,10], arched roofs [11], mono-pitched roofs [12], sawtooth roofs [13] and various other roof types [14,15]. Previous studies have focused on analyzing the surface wind-pressure coefficients of these different types of roofs and providing some information on the potential for natural ventilation, but they have lacked an analysis of airflow inside buildings.
In recent years, the effects of various types of roofs and openings on the natural ventilation of buildings have been studied. Vaishnani [16] conducted numerical simulations and used the PMV model to assess the impact of roof slope angles on cross-ventilation under winter, summer, and monsoon weather conditions in Delhi, India. The results showed that the PMV values decreased with increasing roof slope angles in winter, while in summer, the PMV values increased with greater roof slope angles. Esfeh [17] enhanced natural ventilation capacity by adjusting the geometric shape of a semi-circular arc-shaped roof. The research revealed that the ventilation performance of the arched roof is highly sensitive to the wind angle α, with the best performance at α equals 0 and the poorest performance at 75° and 90°. Kosutova [18] conducted a study on the impact of louvers with varying positions and fixed angles in buildings with flat roofs on natural ventilation. Building upon this foundation, Tai [19] conducted research on the impact of louvers with different positions and angles in buildings with flat roofs on air exchange efficiency. He also proposed optimization factors to measure the relationship between AEE and DFR. Starting from roof types and roof angles, Peren [20] conducted simulations using computational fluid dynamics (CFD) and found that, compared to buildings without eaves, the maximum increase in internal air volume flow occurred when the windward eave had a slope angle of 27°. Then Peren [21] investigated a single-sloped roof and several other concave and convex roofs, and the results showed that convex roofs could maximize the negative pressure of the wake near the building’s exhaust vents, thereby enhancing cross-ventilation inside the building. Leite [22], on the other hand, combined the slope angle of a single-sloped roof with the window opening position and found that the roof slope angle had a significant impact on ventilation airflow, while the vertical position of the outlet opening had a minor effect. Atmaca [23] conducted wind-tunnel experiments and computational fluid dynamics (CFD) simulations on three differently sloped gable roofs, and found that the lowest negative pressure at all roof slopes occurred when the wind direction angle was 90 degrees. Tominaga [24] conducted numerical simulations (CFD) to study the airflow around isolated gable-roofed buildings with different roof slopes and found significant differences in the flow fields for slope ratios of 3:10 and 5:10. Additionally, there are also studies related to wind towers, wind catchers, Mansard roofs, and other structures on rooftops [25,26,27,28,29,30,31,32,33,34].
A review of the literature indicates that there is almost no research on the combination of the slope angle of gable roofs and the vertical position of outlet openings. Therefore, the innovation of this study was to analyze the effects of gable roof slope angle and vertical opening position on indoor air exchange efficiency (AEE), wind-pressure coefficient, and wind speed of independent houses by using CFD method. In the remaining part of this paper, Section 2 introduces the research methods adopted in this study. Section 3 analyzes the results from the CFD simulation with relevant discussions. A conclusion can then be found in Section 4.

2. Methodology

2.1. Wind-Tunnel Experiment

To calibrate the CFD model used in this study, data of wind loads on low-rise buildings, which were collected by wind-tunnel experiments at the Tokyo Institute of Technology in Japan [35], were used. Due to the prevalence of low-rise buildings in suburban areas, we selected suburban terrain, categorized as Terrain Category III in accordance with AIJ (2004), as the test wind field [36]. This particular category features a mean wind velocity profile exponent of 0.20 and a gradient height of 450 m. To replicate it, we utilized turbulence-generating spires, roughness elements, and a carpet on the upstream floor of the wind tunnel’s test section. The wind velocity profile and turbulence intensity profile of the simulated wind field are presented in Figure 1. At a height of 10 cm, the turbulence density measured approximately 0.25. The wind velocity during testing at this height reached approximately 7.4 m per second, which corresponds to roughly 22 m per second at a height of 10 m at full scale. The baseline model selected for this study was the flat-roof model retrieved from the database. Wind pressure measurement taps were uniformly distributed across the surfaces of the tested models, as illustrated in Figure 2. The spacing between the taps was set at 20 mm, equivalent to 2 m at full scale. However, due to limitations in the capacity of wind pressure measurement scanivalve to measure a large number of taps simultaneously, certain inner points on models with larger surfaces were not measured. Each tap was connected to a pressure measurement scanivalve via synthetic resin tubes, measuring 80 cm in length and 1.2 mm in internal diameter. These scanivalves were capable of synchronously recording fluctuating wind pressures at nearly 384 points.

2.2. Model Geometry, Computational Domain, and Building Configuration

The dimensions of the building were 160 mm × 160 mm × 160 mm (L × W × H). The distance from the front of the building model to the entrance was 3H, from the side of the building model to the side and top wall of the fluid domain it was 5H, and from the back of the building model to the exit it was 15H, where H represents the height of the building model. The blockage ratio of the fluid domain was less than 3%. The above settings complied with the design specifications for the fluid domain [37].
The front view of the fluid domain is shown in Figure 3a, the longitudinal section in Figure 3b, and the transverse section in Figure 3c. ANSYS 2022 R1 was used for grid partitioning and CFD simulation, and local refinement of BOI was required, as shown in Figure 3b,c. The inlet and incident profile are shown in Figure 3e and the 3D mesh is shown in Figure 3f. Poly-hexcore cells were used in the meshing and 20 prism layers with a growth rate of 1.2 were applied on the ground and buildings, as shown in Figure 3d. The first cell height of the building was 35 μm, with a corresponding Y+ value of less than 0.60. Skewness of the grid was controlled to be no more than 0.65 in all numerical cases. Fine element sizes were applied at the model edges to achieve the desired grid quality and control the maximum Y+ value. The final grid had a maximum Y+ value of 1.26 for the walls and 1.13 for the ground. The respective average values of Y+ for the walls and ground were 0.48235 and 0.41380.

2.3. Governing Equations

The shears-stress transport (SST) k-ω model transport equations are as follows:
t   ( ρ k ) + x i   ( ρ k u i ) = x j [ Γ k k x j ] + G k Y k + S k
and
t   ( ρ ω ) + x j   ( ρ ω u j ) = x j [ Γ ω ω x j ] + G ω Y ω + D ω + S ω
where G k represents the production of turbulence kinetic energy; G ω represents the generation of ω ;   Γ k and Γ ω represent the effective diffusivity of k and ω , respectively; Y k and Y ω represent the dissipation of k and ω due to turbulence; D ω represents the cross-diffusion term; and S k and S ω are user-defined source terms. The calculation process is as follows:

2.3.1. Production of k

According to the exact equation of the transport of k , the term can be defined as
G k = ρ u i u j u j x i
To evaluate G k   in a manner consistent with the Boussinesq hypothesis,
G k = μ t S 2
where S is the modulus of the mean rate-of-strain tensor, defined as
S 2 S i j S i j

2.3.2. Production of ω

The production of ω is given by
G ω = α α * V t G k
where G k is given by Equation (3) and α is given by
α = F 1 α , 1 + ( 1 F 1 )   α , 2
α , 1 = β i , 1 β * κ 2 σ ω , 1 β *
α , 2 = β i 2 β * κ 2 σ ω 2 β *
κ = 0.41
The coefficient is given by
α = α α * [ α 0 + R e t / R ω 1 + R e t / R ω ]
α * = α * [ α 0 * + R e t / R k 1 + R e t / R k ]
R e t = ρ k μ ω
R k = 6
α 0 * = β i 3
β i = 0.072

2.3.3. Dissipation of k

The dissipation of k is given by
Y k = ρ β * f β * k ω
where
f β * = 1
and
β * = β i * 1 + ζ * F M t
β i * = β * [ 4 / 15 +   ( R e t / R β )   4 1 +   ( R e t / R β )   4 ]
ζ * = 1.5
R β = 8
β * = 0.09
where R e t is given by Equation (13).

2.3.4. Dissipation of ω

The dissipation of ω is given by
Y ω = ρ β f β ω 2
where
f β = 1
β i = F , β i , 1 + ( 1 F 1 )   β i , 2
β = β i [ 1 β i * β i ζ * F   ( M t )   ]
F 1 is obtained from Equation (38), β i * is obtained from Equation (20), and F   ( M t )   is obtained from Equation (29).

2.3.5. Cross-Diffusion Modification

D ω   is defined as
D ω = 2   ( 1 F   1 )   ρ 1 ω σ ω , 2 k x j ω x j

2.3.6. Compressibility Effects

The compressibility function, F   ( M t ) , is given by
F   ( M t ) = { 0 M t M t 0 M t 2 M t 0 2 M t > M t 0
where
M t 2 2 k a 2
M t 0 = 0.25
a = γ R T
In the high-Reynolds number form of the k ω model, β i *   = β * . In the incompressible form, β *  =   β i * .

2.3.7. Modeling the Effective Diffusivity

The effective diffusivities for the SST k ω model are given by
Γ k = μ + μ t σ k
Γ ω = μ + μ t σ ω
where σ ω and μ t are the turbulent Prandtl numbers for k and ω , respectively. The turbulent viscosity is μ t .
σ k = 1 F 1 / σ k , 1 + ( 1 F 1 )   / σ k , 2
σ ω = 1 F 1 / σ ω 1 + ( 1 F 1 )   / σ ω , 2
μ t = ρ k ω 1 max [ 1 α * , S F 2 a 1 ω ]
where S is the strain rate magnitude and α *   is defined in Equation (12).
The blending function     F   1 is given by
F   1 = t a n h   ( ϕ 1 4 )  
ϕ   1 = m i n [ m a x [ k 0.09 ω y , 500 μ ρ y 2 ω ] , 4 ρ k σ ω 2 D ω y 2 ]
D ω + = m a x [ 2 ρ 1 σ ω 2 1 ω k x j ω x j , 1 0 10 ]
The blending function   F 2 is given by
F   2 = t a n h   ( ϕ 2 2 )  
ϕ   2 = m a x [ 2 k 0.09 ω y , 500 μ ρ y 2 ω ]
where Y is the distance to the next surface.

2.3.8. Continuity Equation

p / t +   ( p u ) = 0
where p represents the velocity of the fluid and u represents the velocity of the fluid.

2.3.9. Momentum Equation

  ( p u )   / t +   ( p u u ) = p +   ( μ e f f u ) + ρ g + F
where ρ represents the pressure of the fluid, μ e f f represents the effective viscosity of the fluid, g represents the acceleration of gravity, and F represents the external force.

2.3.10. Turbulence Equation

k / t +   ( k u ) = P k ε +   ( μ e f f t k )  
ω / t +   ( ω u ) = P ω β ω ω +   ( μ e f f t ω )  
where k represents turbulent flow energy, ω represents turbulent dissipation rate, P k represents the generation term of turbulent flow energy, ε represents the dissipation term of turbulent flow energy, μ e f f t represents the effective viscosity of turbulence, P ω represents the generation term of turbulent dissipation rate, and β ω represents the dissipation coefficient of turbulent dissipation rate. The constant values used in Equations (1)–(46) are shown in Table 1.

2.4. Atmospheric Boundary Layer (ABL)

The ABL is the velocity profile that needs to be created from the inlet. The vertical velocity profile U on flat terrain is typically given by Equation (47), where Uref is the wind speed at the reference height of 7.4 m/s, Yref is the reference height of 0.1m, and α is the roughness coefficient of 0.2 [36].
U = U r e f y y r e f α
Turbulent kinetic energy K is determined by Equation (48), in which, Iu represents turbulent intensity, equal to 0.25, and A is equal to 1.
k = A   ( I U U )   2
The value of turbulent dissipation rate ε is typically estimated using the assumption of local equilibrium [38].
ε = C μ 1 / 2 k d U d y
where Cμ is the model constant equal to 0.09. Equation (50) can be obtained by rearranging the above equation.
ε = C μ 1 / 2 k U r e f y r e f α   ( y y r e f )   α 1
Therefore, the specific turbulent dissipation rate ω is defined by Equation (51) [39]. The specific values of U, k, ε, ω are shown in Table 2.
ω = ε C μ k = C μ 1 / 2 U r e f y r e f α   ( y y r e f )   α 1

2.5. Fluent Solver Settings

All CFD simulations were performed using ANSYS FLUENT 2022 R1. The SST K-omega turbulence model was employed, and a production limiter was chosen to obtain more accurate wind-pressure coefficients. The COUPLE scheme was used for pressure–velocity coupling. Pressure was selected with a second-order interpolation scheme, while momentum, turbulent kinetic energy, and turbulent dissipation rate were discretized using a second-order upwind scheme; user-defined scalars (UDS) were chosen as first-order upwind to generate more accurate results. Standard initialization was used in this numerical simulation. Convergence criteria for x, y, and z velocities; turbulent kinetic energy (k); and turbulent dissipation rate (ε) were set to 1 × 10−4. For user-defined scalars (UDS), the convergence criteria were set to 1 × 10−5. The computation method employed was pseudo-transient, with a time step of 0.02 s and 900 timesteps. All solutions converged after 500 iterations.

2.6. Grid Sensitivity Analysis

In the same computational domain, the grid sensitivity analysis was conducted using the aforementioned boundary conditions and solution methods. By varying the grid size of the computational domain, grid sensitivity analysis was performed on three different grid schemes with 707,430 cells (Grid 1); 1,913,049 cells (Grid 2); and 2,665,529 cells (Grid 3). Twelve points were placed on the windward, top, and leeward sides of the building to measure the wind-pressure coefficients. As shown in Figure 4, there was a significant deviation between mesh 1 and mesh 2, while there was a slight deviation between mesh 2 and mesh 3. The average percentage deviation of wind-pressure coefficients between mesh 1 and mesh 2 was 7.47%, while between mesh 2 and mesh 3, it was 0.96%. From the grid sensitivity analysis, the simulation results were not sensitive to mesh 2 and mesh 3. Therefore, mesh 2 was used for all other CFD simulations.

2.7. Model Calibration

The results from the CFD simulation were compared with data from the Japanese wind-tunnel laboratory. Twelve measurement points with a radius of 1.2 mm were selected on the windward, roof, and leeward sides of the building for comparing pressure coefficients. The pressure coefficients at various locations were calculated using Equation (6). Figure 5 compares the pressure coefficients on the building surfaces, showing overall good consistency despite some differences in negative pressure areas.
C P =   ( P P 0 )     ( 0.5 ρ U r e f 2 )  
where P represents static pressure, P0 is the reference static pressure, ρ is the air density, and Uref is the reference velocity.

3. Results and Discussion

To assess the impact of the roof inclination angle (RIA) and vertical opening position on ventilation airflow, a study was conducted on 18 different building models. These models encompassed three different opening positions (top–top, top–bottom, and bottom–top) and six different roof inclination angles (0°, 9°, 18°, 27°, 36°, and 45°). To facilitate a smooth comparison, the above models had the following characteristics: (1) identical internal volumes, (2) identical inlet and outlet opening sizes, and (3) different exit and entrance locations. Figure 6 and Figure 7 illustrate three different opening positions and various angles of gable roofs.

3.1. Pressure Coefficient, Cp

The pressure coefficient Cp is defined as the dimensionless ratio of static pressure at a specific location to the freestream static pressure, as shown in Equation (48). The contour lines of pressure coefficients are presented in Table 3. It can be observed that, with the same opening configuration, as the slope increased, the pressure stagnation region at the top of the building gradually shifted towards the highest point of the roof, and the leeward-side openings typically exhibited negative Cp values. In various building cases, the negative pressure difference at the building’s windward outlets was consistently greater than the positive pressure difference at the windward inlets, indicating that the roof slope angle is a crucial parameter for enhancing wind-driven natural cross-ventilation.

3.2. Dimensionless U/Uref

The dimensionless quantity U/Uref was obtained by dividing the average streamwise velocity U by the reference velocity Uref (7.4 m/s). The contour lines of the dimensionless quantity U/Uref are depicted in Table 4. From Table 4 and Table 5, it can be observed that there was a notable increase in flow velocity in the opening region, a phenomenon evident across all models. Through observation, it was noted that when the opening configuration was top–top or top–bottom, the flow velocity in the exhaust opening area increased with changes in the roof slope, reaching its maximum at 45°, with an acceleration rate 1.2 times that of a flat roof. Conversely, in the case of a bottom–top opening configuration, due to the Coanda effect, the airflow remained attached to the building’s bottom, resulting in a lack of sensitivity of flow velocity in the exhaust opening area to changes in roof angle, with flow velocity remaining relatively constant at around 0.6. Overall, the airflow velocity in the exhaust opening area was more sensitive to the vertical opening position than to the slope angle of the gable roof.

3.3. Air Exchange Efficiency (AEE)

Air exchange efficiency represents the efficiency of external air flushing the interior of a ventilated building [40]. The AEE is defined by Equation (53), where τr represents the AOA at the leeward openings, and τav is the average volume-averaged AOA within the building.
ε A = τ r 2 τ a ν × 100 [ % ]
AOA is calculated using the following scalar transport equation [41].
t ρ φ +   ( ρ U φ )   ( Γ φ ) = S φ
where φ represents the scalar being solved for, i.e., AOA, with Sφ = 1 denoting the source term. The diffusion coefficient Γ is determined using a specific formula [41].
Γ = ρ D + μ t S c t
where D = 2.88 × 10−5m−2/s represents the laminar viscosity of air at a working temperature of 20 °C, μt is the local turbulent viscosity, and Sct = 0.7 is the turbulent Schmidt number.
Under steady-state conditions, the term (ρφ)/∂t in Equation (54) equals zero. The transport equation was implemented in FLUENT using user-defined scalars (UDS) to compute AOA within the building. Equation (55) was employed to determine boundary conditions, setting zero values at the inlet and zero gradients at the walls and exit surfaces.
Table 6 presents the AOA distribution for all operating conditions, Figure 8 illustrates the percentage of AEE for all conditions, and Figure 9 displays the average volume AOA within the building. From Table 4, it can be observed that in the top–top opening configuration, at 0°, the AOA between the windward-facing window and the roof was relatively high, gradually decreasing with an increase in roof angle, reaching an optimum at 36°. In the top–bottom configuration, the AOA between the windward-facing window and the roof was lower but higher than the ground as the roof angle increased, with the ground portion of AOA gradually decreasing. In the bottom–top configuration, the formation of eddies within the building led to excessively high AOA in the central region of the structure, and it was insensitive to changes in roof slope.
From Figure 8, it can be observed that the AEE was highest at 45° top–bottom, reaching 48.1%, and lowest at 27° bottom–top, at only 31.4%. In the top–top opening configuration, the AEE for the flat roof was highest at 42.5%, while the AEE for the other five angles was around 32%. This indicates that in this configuration, AEE is not highly sensitive to changes in the gable roof angle. In the top–bottom opening configuration, AEE increased with an increase in roof slope angle, reaching its peak at 45°, at 48.1%. This suggests that in this configuration, the angle of the gable roof effectively promotes indoor airflow, thereby enhancing AEE.
In the bottom–top opening configuration, the AEE for roofs with different slope angles was quite similar, averaging around 32%. This suggests that AEE is not highly sensitive to changes in the gable roof angle in this configuration. Due to the presence of low–high windows, it is easy for vortices to form within the building, leading to a relatively low AEE.
Figure 9 reveals that in the top–bottom opening configuration with a gable roof slope angle of 45°, τav was at its lowest, measuring 0.33 s. In the bottom–top opening configuration with a roof slope angle of 18°, τav was highest, at 0.6 s. When considering Figure 8 and Figure 9 together, it becomes evident that τav maintained a strong negative correlation with AEE. When τav was low, AEE tended to be high, and conversely, when AEE was high, τav was low.

4. Conclusions

The impact of gable roofs with varying slope angles and different vertical opening positions was investigated. The opening configurations included: (1) top–top, (2) top–bottom, and (3) bottom–top, with gable roof angles of 0°, 9°, 18°, 27°, 36°, and 45°, hence a total number of 18 combinations. The design of the computational domain adhered to best practices outlined in the literature and was simulated numerically with the inclusion of an atmospheric boundary layer (ABL). Reference grids underwent grid sensitivity analysis, and numerical simulations were carried out using the SST k-omega model, with results demonstrating good consistency. Subsequently, an analysis of internal velocities, pressure coefficients, AOA, AEE, and other parameters was conducted, contributing to the advancements made in this study.
(1)
In all configurations in this paper, the slope angle of the roof and the vertical opening positions did appear to have a significant impact on the wind-pressure coefficients.
(2)
In terms of flow velocity, favorable opening positions were more conducive to an acceleration in velocity. With the top–bottom opening configuration, the wind speed at the windward opening was greater than that in the other two opening configurations.
(3)
The highest AEE was 48.1%, obtained with the opening configuration of top–bottom and a roof slope angle of 45°, while the lowest AEE was 31.4%, achieved with the opening configuration of bottom–top and a roof slope angle of 27°.
(4)
When the opening configuration was top–bottom, the slope angle of gable roofs had a significant impact on AEE, with a 5.5% relative increase in AEE for a 45° gable roof compared to a flat roof. This configuration is encouraged for use in buildings.
(5)
When the opening configuration was bottom–top, the slope angle of the gable roof had a very limited impact on the profile wind speed and AEE. Therefore, it is not recommended to use this opening configuration in buildings.
In conclusion, this research has shown that in cross-ventilation, the slope angle of gable roofs and the vertical opening positions play crucial roles in internal airflow, pressure coefficients, AEE, AOA, and other parameters. It is hoped that through this study, buildings’ reliance on mechanical cooling can be reduced. Future work should include, but not be limited to, researching wind speed profiles and pressure coefficients under different wind directions or introducing various types of guiding windows and more complex ventilated roofs on the basis of existing openings and studying their ventilation performance.

Author Contributions

Conceptualization, J.S., C.Z. and Y.L.; methodology, J.S.; software, C.Z.; validation, J.S.; formal analysis, J.S., C.Z. and Y.L.; investigation, C.Z.; resources, J.S.; data curation, C.Z.; writing—original draft preparation, J.S.; writing—review and editing, J.S., C.Z. and Y.L.; supervision J.S., C.Z. and Y.L.; project administration, J.S., C.Z. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the support of the National Natural Science Foundation of China Youth Fund Project (52208095), the Chongqing Municipal Science and Technology Commission “Doctoral Direct Train” Project (CSTB2022BSXM-JCX0162), the Chongqing Municipal Education Commission Science and Technology Project (KJQN202100736), and Research on the construction of Genetic Map of Traditional Villages in Chongqing Drieven by Digital Intelligence (23SKGH130).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chenari, B.; Carrilho, J.D.; da Silva, M.G. Towards sustainable, energy-efficient and healthy ventilation strategies in buildings: A review. Renew. Sustain. Energy Rev. 2016, 59, 1426–1447. [Google Scholar] [CrossRef]
  2. China Building Energy Consumption and Carbon Emissions Research Report 2022. Available online: https://www.cabee.org/site/content/24420.html (accessed on 29 July 2023).
  3. Schulze, T.; Eicker, U. Controlled natural ventilation for energy efficient buildings. Energy Build. 2013, 56, 221–232. [Google Scholar] [CrossRef]
  4. Jiang, Z.; Kobayashi, T.; Yamanaka, T.; Sandberg, M. A literature review of cross ventilation in buildings. Energy Build. 2023, 291, 113143. [Google Scholar] [CrossRef]
  5. Omrani, S.; Garcia-Hansen, V.; Capra, B.R.; Drogemuller, R. Effect of natural ventilation mode on thermal comfort and ventilation performance: Full-scale measurement. Energy Build. 2017, 156, 1–16. [Google Scholar] [CrossRef]
  6. Karava, P.; Stathopoulos, T.; Athienitis, A.K. Airflow assessment in cross-ventilated buildings with operable façade elements. Build. Environ. 2011, 46, 266–279. [Google Scholar] [CrossRef]
  7. Zhang, X.; Weerasuriya, A.U.; Wang, J.; Li, C.Y.; Chen, Z.; Tse, K.T.; Hang, J. Cross-ventilation of a generic building with various configurations of external and internal openings. Build. Environ. 2022, 207, 108447. [Google Scholar] [CrossRef]
  8. Pierre, L.S.; Kopp, G.A.; Surry, D.; Ho, T.C.E. The UWO contribution to the NIST aerodynamic database for wind loads on low buildings: Part 2. Comparison of data with wind load provisions. J. Wind. Eng. Ind. Aerodyn. 2005, 93, 31–59. [Google Scholar] [CrossRef]
  9. Quan, Y.; Tamura, Y.; Matsui, M. Mean Wind Pressure Coefficients on Surfaces of Gable-Roofed Low-Rise Buildings. Adv. Struct. Eng. 2007, 10, 259–271. [Google Scholar] [CrossRef]
  10. Holmes, J.D. Wind pressures on tropical housing. J. Wind Eng. Ind. Aerodyn. 1994, 53, 105–123. [Google Scholar] [CrossRef]
  11. Holmes, J.D.; Paterson, D.A. Mean wind pressures on arched-roof buildings by computation. J. Wind. Eng. Ind. Aerodyn. 1993, 50, 235–242. [Google Scholar] [CrossRef]
  12. Stathopoulos, T.; Mohammadian, A.R. Modelling of wind pressures on monoslope roofs. Eng. Struct. 1991, 13, 281–292. [Google Scholar] [CrossRef]
  13. Cui, B. Wind Effects on Monosloped and Sawtooth Roofs. Ph.D. Thesis, Clemson University, Clemson, SC, USA, 2007. [Google Scholar]
  14. Liu, Z.; Yu, Z.; Chen, X.; Cao, R.; Zhu, F. An investigation on external airflow around low-rise building with various roof types: PIV measurements and LES simulations. Build. Environ. 2020, 169, 106583. [Google Scholar] [CrossRef]
  15. Prasad, D.; Uliate, T.; Ahmed, M.R. Wind Loads on Low-Rise Building Models with Different Roof Configurations. Int. J. Fluid Mech. Res. 2009, 36, 231–243. [Google Scholar] [CrossRef]
  16. Vaishnani, Y.; Ali, S.F.; Joshi, A.; Rakshit, D.; Wang, F. Thermal performance analysis of a naturally ventilated system using PMV models for different roof inclinations in composite climatic conditions. J. Braz. Soc. Mech. Sci. Eng. 2020, 42, 124. [Google Scholar] [CrossRef]
  17. Esfeh, M.K.; Sohankar, A.; Shahsavari, A.R.; Rastan, M.R.; Ghodrat, M.; Nili, M. Experimental and numerical evaluation of wind-driven natural ventilation of a curved roof for various wind angles. Build. Environ. 2021, 205, 108275. [Google Scholar] [CrossRef]
  18. Kosutova, K.; van Hooff, T.; Vanderwel, C.; Blocken, B.; Hensen, J. Cross-ventilation in a generic isolated building equipped with louvers: Wind-tunnel experiments and CFD simulations. Build. Environ. 2019, 154, 263–280. [Google Scholar] [CrossRef]
  19. Tai, V.C.; Kai-Seun, J.W.; Mathew, P.R.; Moey, L.K.; Cheng, X.; Baglee, D. Investigation of varying louver angles and positions on cross ventilation in a generic isolated building using CFD simulation. J. Wind. Eng. Ind. Aerodyn. 2022, 229, 105172. [Google Scholar] [CrossRef]
  20. Perén, J.I.; van Hooff, T.; Leite, B.C.C.; Blocken, B. Impact of eaves on cross-ventilation of a generic isolated leeward sawtooth roof building: Windward eaves, leeward eaves and eaves inclination. Build. Environ. 2015, 92, 578–590. [Google Scholar] [CrossRef]
  21. Peren, J.I.; van Hooff, T.; Ramponi, R.; Blocken, B.; Leite, B.C.C. Impact of roof geometry of an isolated leeward sawtooth roof building on cross-ventilation: Straight, concave, hybrid or convex? J. Wind. Eng. Ind. Aerodyn. 2015, 145, 102–114. [Google Scholar] [CrossRef]
  22. Perén, J.I.; van Hooff, T.; Leite, B.C.C.; Blocken, B. CFD analysis of cross-ventilation of a generic isolated building with asymmetric opening positions: Impact of roof angle and opening location. Build. Environ. 2015, 85, 263–276. [Google Scholar] [CrossRef]
  23. Atmaca, M. Wind tunnel experiments and CFD simulations for gable-roof buildings with different roof slopes. Acta Phys. Pol. A 2019, 135, 690–693. [Google Scholar] [CrossRef]
  24. Tominaga, Y.; Akabayashi, S.-I.; Kitahara, T.; Arinami, Y. Air flow around isolated gable-roof buildings with different roof pitches: Wind tunnel experiments and CFD simulations. Build. Environ. 2015, 84, 204–213. [Google Scholar] [CrossRef]
  25. Heidari, A.; Sahebzadeh, S.; Dalvand, Z. Natural Ventilation in Vernacular Architecture of Sistan, Iran; Classification and CFD Study of Compound Rooms. Sustainability 2017, 9, 1048. [Google Scholar] [CrossRef]
  26. Wu, Y.; Gao, N.; Niu, J.; Zang, J.; Cao, Q. Numerical study on natural ventilation of the wind tower: Effects of combining with different window configurations in a low-rise house. Build. Environ. 2021, 188, 107450. [Google Scholar] [CrossRef]
  27. Cruz-Salas, M.V.; Castillo, J.A.; Huelsz, G. Experimental study on natural ventilation of a room with a windward window and different windexchangers. Energy Build. 2014, 84, 458–465. [Google Scholar] [CrossRef]
  28. Horan, J.M.; Finn, D.P. Sensitivity of air change rates in a naturally ventilated atrium space subject to variations in external wind speed and direction. Energy Build. 2008, 40, 1577–1585. [Google Scholar] [CrossRef]
  29. Asfour, O.S.; Gadi, M.B. Using CFD to investigate ventilation characteristics of vaults as wind-inducing devices in buildings. Appl. Energy 2008, 85, 1126–1140. [Google Scholar] [CrossRef]
  30. Montazeri, H.; Azizian, R. Experimental study on natural ventilation performance of one-sided wind catcher. Build. Environ. 2008, 43, 2193–2202. [Google Scholar] [CrossRef]
  31. Alghaseb, M.A.; Hassen, W.; Mesloub, A.; Kolsi, L. Effect of Heat Source Position in Fluid Flow, Heat Transfer and Entropy Generation in a Naturally Ventilated Room. Mathematics 2022, 10, 178. [Google Scholar] [CrossRef]
  32. Hnaien, N.; Hassen, W.; Kolsi, L.; Mesloub, A.; Alghaseb, M.A.; Elkhayat, K.; Abdelhafez, M.H.H. CFD Analysis of Wind Distribution around Buildings in Low-Density Urban Community. Mathematics 2022, 10, 1118. [Google Scholar] [CrossRef]
  33. Van Hooff, T.; Blocken, B.; Aanen, L.; Bronsema, B. A venturi-shaped roof for wind-induced natural ventilation of buildings: Wind tunnel and CFD evaluation of different design configurations. Build. Environ. 2011, 46, 1797–1807. [Google Scholar] [CrossRef]
  34. Blocken, B.; van Hooff, T.; Aanen, L.; Bronsema, B. Computational analysis of the performance of a venturi-shaped roof for natural ventilation: Venturi-effect versus wind-blocking effect. Comput. Fluids 2011, 48, 202–213. [Google Scholar] [CrossRef]
  35. Available online: http://wind.arch.t-kougei.ac.jp/system/eng/contents/code/tpu (accessed on 29 July 2023).
  36. AIJ. Recommendations for Loads on Buildings; Architectural Institute of Japan: Tokyo, Japan, 2004. [Google Scholar]
  37. Ramponi, R.; Blocken, B. CFD simulation of cross-ventilation for a generic isolated building: Impact of computational parameters. Build. Environ. 2012, 53, 34–48. [Google Scholar] [CrossRef]
  38. Yoshie, R.; Mochida, A.; Tominaga, Y.; Kataoka, H.; Harimoto, K.; Nozu, T.; Shirasawa, T. Cooperative project for CFD prediction of pedestrian wind environment in the Architectural Institute of Japan. J. Wind Eng. Ind. Aerodyn. 2007, 95, 1551–1578. [Google Scholar] [CrossRef]
  39. Wilcox, D.C. Turbulence Modeling for CFD; DCW Industries: La Canada, CA, USA, 1998; Volume 2, pp. 103–217. [Google Scholar]
  40. Hang, J.; Li, Y. Age of air and air exchange efficiency in high-rise urban areas and its link to pollutant dilution. Atmos. Environ. 2011, 45, 5572–5585. [Google Scholar] [CrossRef]
  41. Chanteloup, V.; Mirade, P.-S. Computational fluid dynamics (CFD) modelling of local mean age of air distribution in forced-ventilation food plants. J. Food Eng. 2009, 90, 90–103. [Google Scholar] [CrossRef]
Figure 1. Simulated wind field of suburban terrain.
Figure 1. Simulated wind field of suburban terrain.
Buildings 13 02716 g001
Figure 2. Arrangement of wind pressure measurement taps.
Figure 2. Arrangement of wind pressure measurement taps.
Buildings 13 02716 g002
Figure 3. (a) The building in its computational domain, (b) longitudinal section, (c) transverse section, (d) prism layer around the house and ground, (e) inlet and incident profiles, and (f) 3D mesh.
Figure 3. (a) The building in its computational domain, (b) longitudinal section, (c) transverse section, (d) prism layer around the house and ground, (e) inlet and incident profiles, and (f) 3D mesh.
Buildings 13 02716 g003
Figure 4. Results of grid sensitivity analysis.
Figure 4. Results of grid sensitivity analysis.
Buildings 13 02716 g004
Figure 5. Comparison of CFD result with AIJ test.
Figure 5. Comparison of CFD result with AIJ test.
Buildings 13 02716 g005
Figure 6. Opening positions and sizes. (a) The location and size of the top and bottom openings, (b) Top-Top opening configuration, (c) Top-Bottom opening configuration, (d) Bottom-Top opening configuration.
Figure 6. Opening positions and sizes. (a) The location and size of the top and bottom openings, (b) Top-Top opening configuration, (c) Top-Bottom opening configuration, (d) Bottom-Top opening configuration.
Buildings 13 02716 g006
Figure 7. Gable roof inclination angles.
Figure 7. Gable roof inclination angles.
Buildings 13 02716 g007
Figure 8. Air exchange efficiency (AEE) with different angles of gable roofs and different opening positions.
Figure 8. Air exchange efficiency (AEE) with different angles of gable roofs and different opening positions.
Buildings 13 02716 g008
Figure 9. Average volume-averaged age of air (AOA) with different angles of gable roofs and different opening positions.
Figure 9. Average volume-averaged age of air (AOA) with different angles of gable roofs and different opening positions.
Buildings 13 02716 g009
Table 1. Values of the constants used in Equations (1)–(46).
Table 1. Values of the constants used in Equations (1)–(46).
σ k , 1 σ ω , 1 σ k , 2 σ ω , 2 α 1 β i , 1 β i , 2 α * α α β * R β R k R ω ζ * M t 0
1.1762.01.01.1680.310.0750.082810.52 1 / 9 0.09862.951.50.25
Table 2. U,   k ,   ε ,   ω values at the inlet boundary.
Table 2. U,   k ,   ε ,   ω values at the inlet boundary.
Y (m)U (m/s) k  (-) ε (-) ω (-)
0.025.3631.79828.924178.799
0.046.1602.37121.917102.682
0.068.8164.85832.45574.237
0.089.3385.44028.92758.975
Table 3. Pressure coefficient, Cp, suitable for different open configurations with gable roofs.
Table 3. Pressure coefficient, Cp, suitable for different open configurations with gable roofs.
ConfigurationLegend0918273645
Top–TopBuildings 13 02716 i001Buildings 13 02716 i002Buildings 13 02716 i003Buildings 13 02716 i004Buildings 13 02716 i005Buildings 13 02716 i006Buildings 13 02716 i007
Top–BottomBuildings 13 02716 i008Buildings 13 02716 i009Buildings 13 02716 i010Buildings 13 02716 i011Buildings 13 02716 i012Buildings 13 02716 i013
Bottom–TopBuildings 13 02716 i014Buildings 13 02716 i015Buildings 13 02716 i016Buildings 13 02716 i017Buildings 13 02716 i018Buildings 13 02716 i019
Table 4. U/Uref suitable for different open configurations with gable roofs.
Table 4. U/Uref suitable for different open configurations with gable roofs.
ConfigurationLegend0918273645
Top–TopBuildings 13 02716 i020Buildings 13 02716 i021Buildings 13 02716 i022Buildings 13 02716 i023Buildings 13 02716 i024Buildings 13 02716 i025Buildings 13 02716 i026
Top–BottomBuildings 13 02716 i027Buildings 13 02716 i028Buildings 13 02716 i029Buildings 13 02716 i030Buildings 13 02716 i031Buildings 13 02716 i032
Bottom–TopBuildings 13 02716 i033Buildings 13 02716 i034Buildings 13 02716 i035Buildings 13 02716 i036Buildings 13 02716 i037Buildings 13 02716 i038
Table 5. U suitable for different open configurations with gable roofs.
Table 5. U suitable for different open configurations with gable roofs.
ConfigurationLegend0918273645
Top–TopBuildings 13 02716 i039Buildings 13 02716 i040Buildings 13 02716 i041Buildings 13 02716 i042Buildings 13 02716 i043Buildings 13 02716 i044Buildings 13 02716 i045
Top–BottomBuildings 13 02716 i046Buildings 13 02716 i047Buildings 13 02716 i048Buildings 13 02716 i049Buildings 13 02716 i050Buildings 13 02716 i051
Bottom–TopBuildings 13 02716 i052Buildings 13 02716 i053Buildings 13 02716 i054Buildings 13 02716 i055Buildings 13 02716 i056Buildings 13 02716 i057
Table 6. Age of air (AOA) suitable for different open configurations with gable roofs.
Table 6. Age of air (AOA) suitable for different open configurations with gable roofs.
ConfigurationLegend0918273645
Top–TopBuildings 13 02716 i058Buildings 13 02716 i059Buildings 13 02716 i060Buildings 13 02716 i061Buildings 13 02716 i062Buildings 13 02716 i063Buildings 13 02716 i064
Top–BottomBuildings 13 02716 i065Buildings 13 02716 i066Buildings 13 02716 i067Buildings 13 02716 i068Buildings 13 02716 i069Buildings 13 02716 i070
Bottom–TopBuildings 13 02716 i071Buildings 13 02716 i072Buildings 13 02716 i073Buildings 13 02716 i074Buildings 13 02716 i075Buildings 13 02716 i076
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shi, J.; Zhao, C.; Liu, Y. CFD Analysis of Building Cross-Ventilation with Different Angled Gable Roofs and Opening Locations. Buildings 2023, 13, 2716. https://doi.org/10.3390/buildings13112716

AMA Style

Shi J, Zhao C, Liu Y. CFD Analysis of Building Cross-Ventilation with Different Angled Gable Roofs and Opening Locations. Buildings. 2023; 13(11):2716. https://doi.org/10.3390/buildings13112716

Chicago/Turabian Style

Shi, Jingyuan, Changkai Zhao, and Yanan Liu. 2023. "CFD Analysis of Building Cross-Ventilation with Different Angled Gable Roofs and Opening Locations" Buildings 13, no. 11: 2716. https://doi.org/10.3390/buildings13112716

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

Article Metrics

Back to TopTop