Next Article in Journal
Effect of the Progressive Failure of Shear Connectors on the Behavior of Steel-Reinforced Concrete Composite Girders
Next Article in Special Issue
Multi-Task Learning for Concurrent Prediction of Thermal Comfort, Sensation and Preference in Winters
Previous Article in Journal
A Study on Daylighting Performance of Split Louver with Simplified Parametric Control
Previous Article in Special Issue
Data Reconstruction of Wireless Sensor Network and Zonal Demand Control in a Large-Scale Indoor Space Considering Thermal Coupling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Study of Micro-Thermal Environment in Block Based on Porous Media Model

1
School of Energy and Environment, Southeast University, Nanjing 210016, China
2
School of Architecture, Southeast University, Nanjing 210016, China
3
Jiangsu Provincial Key Laboratory of Solar Energy Science and Technology, Southeast University, Nanjing 210016, China
*
Author to whom correspondence should be addressed.
Buildings 2022, 12(5), 595; https://doi.org/10.3390/buildings12050595
Submission received: 7 April 2022 / Revised: 27 April 2022 / Accepted: 2 May 2022 / Published: 4 May 2022

Abstract

:
The mitigation of the heat island effect has become one of the most challenging issues with the rapid urbanization and increased human activities. A standard model and a porous media model were developed to simulate the microthermal environment in the presence of anthropogenic heat in Nanjing Xinjiekou block. The accuracy of the simulation results was verified by field measurement data. Compared with the standard CFD method, the porous media method reduces the number of meshes by 27.8% and the total computation time by 66.7%. By comparing and analyzing the thermal environment of the block with various porosities and heat intensities at different heights, calculations proved that the velocity is positively correlated with the porosity change, and the temperature is negatively correlated with it in contrast. The temperature increases linearly with the increase in anthropogenic heat intensity under the block height range, and the gradient is about 0.025 K/W at the height of 2 m. The porous media approach allows for effective prediction of the micro-thermal environment in the initial stages of urban design while increasing the porosity of the block and controlling the intensity of anthropogenic heat emissions can be a prominent means of improving the thermal environment.

1. Introduction

With the increasing process of urbanization in China, the phenomenon of the urban heat island becomes a serious problem due to the growing total number of buildings and human activities [1]. Computational Fluid Dynamics (CFD) simulation can accurately simulate heat transfer and diffusion in a block [2]; thus, providing the theoretical basis and reference for layout design and planning. Standard CFD simulation requires establishment of a detailed geometric model of the simulation area in large-scale simulations, which is not only complicated for modeling and pre-processing, but also has a long running time, a large amount of calculation, and a corresponding increase in simulation cost [3]. The porous media model can use key parameters to reflect the overall characteristics of the region, simplify the model, and save the amount of calculation [4,5]. Therefore, based on the similarity between building blocks and porous media, a new method for simplifying building groups into porous media has emerged. In 1997, Antohe [6] developed a porous turbulence model to study the flow of porous urban canopy, indicating that the porous media model has certain practical value in simulating large-scale urban areas. Hang [7] regarded the city as a porous media, and the air in the city as an incompressible fluid passing through the porous media. By dividing the whole mixed region into porous urban areas and fluid areas, it is found that this method can significantly reduce the workload of modeling, meshing, and calculation in simulation. Gu Zhaolin et al. [8] used the porous method and experiments to analyze the real urban area and found that the simulation results were in perfect agreement with the wind tunnel experimental results and verified the validity of the numerical model with the wind tunnel experimental results. Wang et al. [9] compared and analyzed the three simulation results of the actual model, the roughness height model, and the mixed model of porous media modeling in some areas of the building clusters, and found that the porous media model is more effective and accurate than the roughness height model, and requires much less calculation. Li et al. [10] validated the porous media simulation method and micro-scale CFD simulations for a relatively regular arrangement of settlements with a horizontal length of about 1 km in Xi’an, and the results showed that the wind velocity distribution and temperature distribution trends of both were in excellent agreement. Ming et al. [11] established a concentric city structure and obtained the temperature and flow fields over the city using porous media simulations, and analyzed the effects of anthropogenic heat, ambient wind velocity, and porosity in the central region on urban turbulence and heat transfer. It is found that the three-dimensional turbulent porous media model is suitable for estimating the UHI effect. These studies demonstrate the practical value of applying the porous media model to urban building clusters and provide a reference and theoretical basis for this study.
In recent years, researchers have found that anthropogenic heat affects the urban thermal environment at different time and space levels [12], which is one of the main driving forces that exacerbate global warming [13,14,15,16]. Especially in the city center area with high building density, the heat dissipation capacity is not ideal due to poor ventilation, and the impact of anthropogenic heat removal on the urban heat island effect cannot be ignored [17,18]. Anthropogenic heat emissions are mainly in the form of industrial heat emissions, traffic heat emissions, and household heat emissions into the atmospheric thermal environment [19,20]. The research on a large amount of urban anthropogenic heat by Du et al. [21] showed that anthropogenic heat played an obvious role in aggravating the urban heat island effect. Xie Min et al. [22] found that anthropogenic heat emissions have apparent regional characteristics, and the anthropogenic heat emissions in areas with a high degree of urbanization are significantly higher than those in other areas. Additionally, anthropogenic heat emissions vary with season and time [23]. Takane et al. [24] explored the impact of the feedback relationship between urban warming and air-conditioning use on atmospheric temperature in future urban climates and found that anthropogenic heat emissions from air-conditioning use would lead to a linear increase in the rate of temperature change. Ohashi et al. [25] found that the waste heat from air conditioners in Osaka City in summer may cause the air temperature to increase by 0.36~0.72 °C. Tong Hua [26] explored the effect of anthropogenic heat on the increase in urban temperature by estimating the intensity of anthropogenic heat emissions in Beijing, and evaluated the effect of reducing anthropogenic heat emissions on mitigating urban heat islands. Jiang Weimei [27] found that the anthropogenic heat sources in Nanjing and Hangzhou could increase the heat island intensity by 1~3 °C. However, the research on anthropogenic heat mostly focuses on the statistical analysis of different anthropogenic heat source intensity in cities in different regions, as well as the analysis of the impact of anthropogenic heat on the thermal environment on the urban and even global scale [28,29,30]. There is still a lack of research on the use of numerical simulation from the perspective of small and medium scales.
The simulation of predicted airflow around buildings has an essential impact on the outdoor microthermal environment, and the considerable simulation costs are one of the bottlenecks limiting the development of microclimates. There are significant difficulties in using ordinary computers to quickly implement thermal environment calculations in urban spaces, while porous media simulations can provide significant savings in computational costs and computational effort. Research on porous media for urban microthermal environments is still preliminary, and scholars are currently experimenting mainly on ideal building blocks and using wind tunnel experiments for verification. In this paper, an actual block is selected, and measured data are used to verify the applicability of the porous media model in the study of microthermal environments in the actual block. The effect of porosity and anthropogenic heat emissions on the microthermal environment is also explored, considering the inescapable role of anthropogenic heat emissions on the microthermal environment. The modeling results could be used to guide development of scientific urban development plans to mitigate the UHI effect.

2. Models and Methods

2.1. Macroscopic Turbulence Model

Based on Antohe’s [6] derivation, Hang and Li [7] provided a single domain method in the transport equations of the macroscopic k–ε porous model to simulate the idealized building cluster. This approach performed reasonably well in predicting the reduction in flow velocity through the building cluster. The transport equations of the incompressible flow in the urban canopy are as follows.
Continuity equation:
( ρ ϕ u i f ) x i = 0
The momentum conservation equation:
( ρ ϕ u i f u j f ) x j = x j [ ( μ l J 1 + μ t ) ( ϕ u i f ) x j ] ( ( ϕ p f ) x i + 2 3 ρ ( ϕ k f ) x i ) ϕ 2 μ K u i f ϕ 3 ρ C F K Q f u i f + δ i 3 ρ [ β ( T T i n ) 1 ]
Equation for the turbulent kinetic energy (THK):
( ρ ϕ u i f k f ) x j = x j [ ( μ l J 1 + μ t σ k ) ( ϕ k f ) x j ] ρ ϕ ε f + ϕ G k 2 ϕ 2 μ l K k f 8 3 ϕ 3 ρ C F K Q f k f + 2 ϕ 3 C F K F k
Equation for the energy dissipation:
( ρ ϕ u i f ε f ) x j = x j [ ( μ l J 1 + μ t σ ε ) ( ϕ ε f ) x j ] C ε 1 ϕ ε f k f G k J 1 ρ C ε 2 ϕ ε f k f ε f 2 ϕ 2 μ l K ε f 8 3 ϕ 3 ρ C F K Q f ε f 8 μ l 3 ϕ 3 C F K k f x r Q f x r + 2 ϕ 3 C F K [ μ l v t x r ( u i f u j f Q f ) 2 u i f x r x j + 2 μ l v t u i f u j f Q f 2 x r 2 ( u i f x j ) ]
where
Q f = u i f u j f
μ t = C μ ρ ( k f ) 2 ε f
F k = μ t u i f u j f Q f ( ϕ u j f x i )
G k = μ t u i f x j ( u i f x j + u j f x i )
where ρ is density, u i f , Q f , p f , k f , and ε f are intrinsic averages of time-averaged velocity components, representing the velocity, pressure turbulence kinetic energy and its dissipation rate, respectively. μ l and μ t are the dynamic and turbulent viscosities. J 1 is the viscosity ratio which can be assumed to be 1 for most applications in porous media. σ k and σ ε are constant, and the default values are 1.0 and 1.3, respectively.

2.2. Parameterization of Buildings as Porous Media

The porosity is the percentage of pores in the porous media, and the calculation is as follows:
ϕ = U v U = U U s U
where ϕ is the porosity, U is the volume of the porous media, U v is the pore volume of the porous media, and U s is the volume of the skeleton of the porous media.
On the basis of on-site investigation, the block was divided into 7 building blocks, as shown in Figure 1. Block zone porosity can be determined by calculating the following:
ϕ = V V i V = 1 S i × h i S × H
where V is the volume of the block, V i is the volume of each building, h i is the height of each building, S i is the floor area of each building, S is the area of the block, and H is the height of the block.
The permeability K and Forchheimer coefficient C F for the form drag can be calculated as follows:
K = ϕ 3 d p 2 150 ( 1 ϕ 2 )
C F = 1.75 150 ϕ 3
where d p is the characteristic size of the solid particles in the porous media.

2.3. Physical Model

Model validation is the key to determine the accuracy and reliability of the numerical simulation. In urban simulations, the sensing size of the measuring instruments is too small relative to the study area. It is necessary to arrange numerous measuring points to obtain data representing the entire area, which is difficult to achieve in practice. Therefore, it is more practical to validate porous media models at the microscale. The representative high-density block buildings in the central area of Xinjiekou, Nanjing, in hot summer and cold winter areas, were selected as the research object. Streets surround the selected block, and the buildings in the block are almost always north–south orientation. According to Ma Jian’s [31] simulation results of streets with different aspect ratios, when B/H = 0~1 and B/W = 0~1, the whole building group can be integrated into an integral building, as shown in Figure 2. Based on the site investigation, the simulated areas are divided into 7 building blocks with different porosities, as shown in Figure 1.
The models established by standard CFD modeling and the porous media model method are shown in Figure 3 and Figure 4. Due to the limitation of computer power and the complexity of the research object, the local model of the building group is appropriately simplified in the standard CFD modeling. When the porous model is established, the research method proposed by Wang et al. [9] is used to perform additional processing on the block. In order to fully develop the airflow and ensure the accuracy of the calculation, the computational domain was extended by 3H in the vertical direction and by more than 5H in the other four directions (where H is the height of the highest building in the block).
When the standard CFD method divides the model grid, the building group was the key research object, and the grid independence test was carried out. After setting the same boundary conditions for the four grids with grid numbers of 1,789,673, 2,185,209, 3,268,935, and 4,195,998, the simulations were performed separately, and the average wind speed at the outlet of the block was recorded. The simulation results were 0.341, 0.384, 0.430, and 0.449 m/s, respectively, indicating that when the number of grids exceeds 3 million, the simulation results tend to be stable (the difference is about 4.4%). Therefore, subsequent calculations of standard CFD method would use grids with a grid number of about 3 million for calculation. The porous media model uses the same method to evaluate the meshes with the number of meshes of 797,651, 1,847,621, 2,767,710, and 3,952,174 and four meshes, namely 2,767,710. The meshes of the two models are shown in Figure 5 and Figure 6.

2.4. Boundary Conditions and Simulation Settings

When simulating the micro-thermal environment of the block, setting reasonable boundary condition parameters can make the simulation result closer to the actual situation.
The inlet was set as the velocity inlet, and the wind speed varied with the height. For the distribution of the incoming wind speed, the most commonly used exponential law [32] was used, such as below:
u i n u z e f = ( z i n z r e f ) α
where uzef is the average wind speed of incoming flow at the height of Zref; uin is the average wind speed at Zin at any height, α is the ground roughness coefficient, which is taken as 0.22 in the urban center according to China’s specifications.
The inlet air temperature was set as 32 °C, and the wind direction was set as south. The outlet boundary was set as the outflow boundary condition.
The profiles of turbulent kinetic energy ( κ i n z ) and its dissipation ( ε i n z ) [7] at the domain inlet were calculated as described in Equations (14) and (15).
κ i n ( z ) = 0.01 u z e f 2
ε i n ( z ) = c μ 3 / 4 κ i n ( z ) 3 / 2 k v z
where c μ = 0.09 , k v is von Karman’s constant, taken as 0.4.
Other boundaries, such as the top and sides of the domain, were set to symmetry boundaries. The anthropogenic heat was simplified to the heat source intensity based on the surface area for the thermal boundary conditions, and the surface heat flux was given. In order to compare the simulation results of the porous media method and the standard CFD method sufficiently, the standard CFD method and the porous media method use the same size calculation domain and consistent boundary conditions. The boundary conditions of the model using both methods are detailed in Table 1.
When simulating calculations, the standard CFD method and the porous media method use the same solution settings. The control volume method was used to discretize the equation in the simulation calculation. The RNG k–ε turbulence model is employed. The coupling relationship between velocity and pressure was the SIMPLEC scheme, and the residual was less than 10−5. The convection term and the diffusion term were simulated by the second-order upwind style, which had higher accuracy and slower convergence speed. After several attempts, the under-relaxation factor for speed and pressure was set to 0.5 and the other factor below 0.5.

3. Actual Measurement Verification

3.1. Measured Plan

Model validation is the key to determining the accuracy and reliability of numerical simulations. In this study, the standard CFD modeling method and the porous media method were used to simulate the micro-thermal environment of the block, and the measurement points were measured in the block to verify the accuracy of the simulation results. In this paper, the “virtual synchronous measurement method” was selected for measurement on a typical summer day in July. The measurement points were mainly selected along the main road, the corner of the building, and the open space around the building, and the measurement height was about 1.5 m. The measuring device was used for each measurement point test time of 10 min, every interval of 30 s to record the data, the average value was taken as the result, and the instrument parameters tested are shown in Table 2. The selection of measuring points is shown in Figure 7, and the field measurement is shown in Figure 8.

3.2. Comparison of Measured and Simulated Results

The measurement results of each measuring point in the block were obtained from the above measurement method, and the average values were calculated respectively. The results show that the wind speed changes violently at each time in a short time, while the temperature changes relatively gently, so the temperature standard deviation is of more research significance. The results are shown in Table 3.
Since the monitoring points in the simulation may be different from the field measurement points, the average values of 5~6 points near the simulation’s monitoring points were used to compare the simulated values and the measured values. Figure 9 shows the comparison results of wind velocity and temperature between measured and simulated values at different points.
The comparison results show that the measured results using the selected measurement points are in agreement with the simulation results of both the standard CFD model and the porous media model to a high degree.
In order to further judge the degree of agreement between the simulation results and the measured values, the coefficient of determination R 2 was used to measure the degree of fitting between the simulation results and the measured values. The closer R 2 is to 1, the better the fitting is. The calculation formula is as follows:
R 2 = [ i ( y i y ¯ ) ( y i ^ y ^ ¯ ) ] 2 [ i ( y i y ¯ ) 2 ] [ i ( y i ^ y ^ ¯ ) 2 ]
The R 2 for each of the two methods is shown in the Table 4.
The R 2 values of wind velocity and temperature between the standard CFD simulation method and the porous media simulation are 0.08 and 0.16, respectively. It is indicated that the accuracy of the standard CFD simulation method is slightly higher in comparison. Nevertheless, the R 2 obtained by either the standard CFD method simulation or the porous media model method is above 0.65, which indicates that the fitting effect is better and both methods can effectively reflect the thermal environment of the block.

4. Discussion

4.1. Comparison between Porous Media Modeling Method and Standard Modeling Method

4.1.1. Comparative Analysis of Calculation Results

The simulation results of the standard model and the porous media model are compared to analyze further the feasibility of using the porous media modeling method instead of the standard CFD modeling method in studying the microthermal environment of the block. Figure 10 shows the wind velocity at the pedestrian height of the block using two methods, respectively, and Figure 11 shows the cloud diagram of temperature distribution. The velocity and temperature distribution in the block are similar to different simulation methods, and the porous media model method can characterize the micro-thermal environment of each block zone well. In the main road between buildings, especially in the south–north direction of the main street, the wind speed shows a trend of attenuation along the direction of the street, which promotes ventilation to improve the micro-thermal environment of the block. The velocity of the main street calculated by both methods is between 0.5 and 0.62 m/s, and the temperature is between 307 and 309 K.
Because the porous media method simplifies the geometric shape of the actual building, the overall thermal environment characteristics of the region are more conducive to comparison and verification than the interior points of the building cluster. According to the simulation results, each region’s average wind velocity and temperature are calculated, respectively, and the results are shown in Figure 12. The distribution trends of wind speed and temperature in different regions are the same. The maximum value of wind speed and the minimum value of temperature are obtained in area 6, and the minimum value of wind velocity and the maximum value of temperature are obtained in area 4. The maximum variation in wind velocity between the regions using the standard CFD model and the porous media model is 0.51 and 0.60 m/s, respectively, and the maximum variation in temperature is 1.6 and 1.2 K. The maximum variation in average wind speed of each region obtained by different methods is only 0.07 m/s, the maximum variation in temperature is 1.4 K, and the maximum error is less than 10%. Thus, it can be seen that the mean temperature and wind velocity of each region are in good agreement with each other.
Comparing the overall characteristics of the microenvironment within each zoning area and along the central axis when the microthermal environment was modeled using both standard CFD and porous media methods, it was observed that the wind velocity distributions and temperature distributions were similar between the standard CFD modeling approach and the porous media approach. Additionally, porous media are effective in reflecting the wind and thermal environments of the building complex as a whole and of the nearby main street.

4.1.2. Comparative Analysis of Computational Cost

When the computer with the same performance was used for simulation, the two simulation methods’ running time and iterative step size are shown in Table 5. The results show that the porous media model method reduces the number of grids by about 15.3% compared with the standard method. The porous media model method takes only 6 s to complete one iteration, while the standard method requires 14.2 s. The porous media model method and the standard method converge after 854 and 1516 steps, respectively, and the total number of iterations of the porous media model method is reduced by 43.6%. The standard CFD method takes about 480 min to achieve convergence, while the porous media model method takes only about 140 min, about 70.8% faster than the standard method. Compared with the standard CFD modeling method, the porous media model method has the advantages of simple implementation, low cost of computing resources, and can effectively simulate the overall micro-thermal environment of the block. Therefore, the porous media method can reduce the computational cost of microclimate research and practice and is of great value in the initial micro-environmental assessment.
When designing a comprehensive city, the two methods must be combined or alternated according to the type of project and its requirements. A porous media model should be used for determining the main comprehensive partitions and ventilation corridors, whereas a standard model should be used for determining the surrounding areas of important structures. Consequently, the problems associated with the standard methods taking a long time to simulate as well as those associated with the porous media model not being able to accurately reflect the micro-thermal environment within a specific building can be simultaneously solved.

4.2. Effect of Porous Media Parameters on Micro-Thermal Environment

When it comes to porous media blocks, porosity indicates the density of the building layout within the block, and its size has an important influence on the ventilation capacity of the block. In order to explore the influence of porosity on the microthermal environment of the actual block, the zones with different porosity, except zones 6 and 1, where the average building height is much higher than that of the other zones, were selected to study the corresponding wind velocity and temperature at the heights of 2, 10, and 50 m (Figure 13).
The results show a positive correlation between wind velocity and porosity at different heights and a negative correlation between temperature and porosity. The greater the porosity in the dominant wind direction, the higher the average wind velocity, and the better the ventilation environment of the block. At the height of 2 m, the change in wind speed and porosity in each zone is the most obvious. The wind velocity in the area with porosity of 0.6 is about 0.32 m/s higher than that in the area with porosity of 0.4, and the temperature is about 0.9 K lower. The wind velocity in different porosity zones at 10 m height is about 0.25 m/s higher than that at 2 m height, and the temperature is about 0.9 K lower, but the variation law of wind velocity and temperature is basically the same as that at 2 m height. As the wind velocity at the block entrance changes exponentially with height, the higher the height, the greater the wind velocity, and the higher wind velocity can quickly transport and diffuse the heat. The wind velocity at 50 m height will still be affected by the airflow disturbance on the edge of the block, but the temperature hardly changes with porosity change.

4.3. Influence of Anthropogenic Heat on Micro-Thermal Environment

Based on Luo et al.’s [33,34] research on urban anthropogenic heat emission, four anthropogenic heat intensities with values of 0, 20, 40, and 60 W/m2 were selected in this study to investigate the effects of anthropogenic heat intensity on the microthermal environment of the block. They represent non-anthropogenic heat emission, anthropogenic heat intensity in other cities, anthropogenic heat intensity in the study area, and anthropogenic heat intensity in the central city of Nanjing, respectively. Figure 14 shows the effect of anthropogenic heat on the block at different heights.
The results show that the mean temperature of each area at different heights increases significantly with the increase in the anthropogenic heat intensity in the block. Furthermore, since the anthropogenic heat will diffuse downstream with the wind direction, the upstream heat source intensity will impact the downstream streets and areas, which is the main reason for the significantly higher temperature in Block 1. The distribution of wind speed and temperature in each area is also in line with the law that the block with high porosity has high wind speed and low temperature. When the airflow enters the block, the wind velocity in the area with large porosity is relatively large, and the high-speed wind increases the heat. The exchange flux makes the anthropogenic heat dissipate rapidly, so the anthropogenic heat has relatively little effect on the areas with high porosity. At the heights of 2 and 10 m, the temperature distribution in each region is basically the same. At the height of 2 m, the variation gradient of anthropogenic heat is about 0.025 K/W. However, at the height of 10 m, due to the increase in wind speed and the effect of lateral wind, the heat diffuses relatively faster, and the influence of anthropogenic heat on the thermal environment at this height is weakened. Since the average building height is lower than 50 m, the anthropogenic heat intensity at the height of 50 m has little effect on the thermal environment.
Accordingly, the intensity of anthropogenic heat emission plays a crucial role in the formation of the microthermal environment of the block. In order to alleviate the urban heat island effect in summer, the first consideration should be given to reducing the intensity of urban anthropogenic heat emission. Although the anthropogenic heat intensity has almost no effect on the wind velocity of the block, the wind velocity determines the speed of heat diffusion. The more significant wind velocity allows for rapid transport of heat generated by the block. Various methods, such as changing the layout and increasing the porosity of the block, can be used to promote airflow and improve the microthermal environment of the block.

5. Conclusions

In this paper, the standard CFD modeling and porous media method simulation were verified by actual measurement and comparison. The influence of porous media parameters and anthropogenic heat intensity on the thermal environment of the block was analyzed. The main findings can be summarized as follows:
(1)
The actual measurement shows that both the porous media method and the standard model method can accurately characterize the overall regional ventilation capacity of the building complex. The wind speed and temperature of each area in the block with different simulation methods have a similar distribution. Porous media also improves the efficiency of modeling and calculation. The porous media model method reduces the number of meshes by about 15.3% compared with the traditional method, and the calculation rate is about 70.8% faster than the traditional method.
(2)
As an essential parameter of porous media, porosity, reflects the density of the building layout in the block. From the building ground to the average height of the block, the wind speed is positively correlated with the porosity change, and the temperature is negatively correlated with the porosity change. Above the block height, say 50 m, the wind speed is still affected by the porosity due to airflow disturbances over the block edge, but the temperature is hardly affected by the porosity at this height.
(3)
The anthropogenic heat emission is one of the critical factors leading to increased temperature in the block. The intensity of anthropogenic heat has little effect on the wind speed of the block, but the regional temperature at the same height increases linearly with the intensity of anthropogenic heat. The change gradient is about 0.025 K/W.
Accordingly, the porous media method can save a lot of time and cost required for research on microthermal environments compared with the standard CFD modeling calculation. Using this method can significantly reduce the time and workload of urban planners for numerical simulation of microclimates and provide a particular reference value for comprehensive urban design and planning based on climate improvement. At the same time, increasing the porosity of the block and controlling the intensity of anthropogenic heat emission can improve the micro-thermal environment from two crucial aspects: enhancing ventilation and reducing heat gain, which is of great significance for alleviating the heat island effect. Since the practical application of the porous media method is still in its early stages, in the future, parameterizing the porous media as a function of horizontal or vertical changes may be considered to address the challenges posed by uneven building heights and inaccuracies caused by complex building environments.

Author Contributions

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

Funding

This research was funded by The Renewable Energy Multi-Energy Complementary Heating Key Technology Research and Demonstration, the funding number is 2021ZD26.

Data Availability Statement

Not applicable.

Acknowledgments

The author wishes to express his appreciation and gratitude to the anonymous reviewers and editors for their insightful comments and suggestions to improve the paper’s quality.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

u i f intrinsic averages of time-averaged velocity components (m·s−1)
x i Cartesian coordinates
T temperature (°C)
p f intrinsic averages of time-averaged pressure (Pa)
k f intrinsic averages of time-averaged turbulent kinetic energy (m2·s−2)
J 1 viscosity ratio
kvon Karman’s constant, 0.4
Q f intrinsic averages of time-averaged velocity (m·s−1)
C μ constant value: 0.09
Kpermeability (m2)
C F Forchheimer coefficient
d p characteristic size of the solid particles (m)
U volume of the porous media (m3)
V volume of the block (m3)
h i height of each building (m)
S i floor area of each building (m2)
S area of the block (m2)
H height of the block (m)
uaverage wind speed (m/s)
zheight (m)
R 2 coefficient of determination
y i measured value
y ¯ mean measured value
y i ^ simulation value
Greek symbols
ϕ porosity
ρ density (kg·m−3)
β thermal expansion coefficient (K−1)
μ l dynamic viscosities (kg·m−1·s−2)
μ t turbulent viscosities (kg·m−1·s−2)
δ i j Kronecker delta operator
ε f intrinsic averages of time-averaged dissipation (m2·s−3)
αground roughness coefficient
ĸ turbulent kinetic energy (m2·s−2)
ε dissipation   of   ĸ (m2·s−3)
ϕ porosity
Subscripts
ininlet
refreference

References

  1. Zhong, S.; Qian, Y.; Zhao, C.; Leung, R.; Yang, X.Q. A case study of urbanization impact on summer precipitation in the greater Beijing metropolitan area: Urban heat island versus aerosol effects. J. Geophys. Res.-Atmos. 2015, 120, 10903–10914. [Google Scholar] [CrossRef]
  2. Toparlar, Y.; Blocken, B.; Maiheu, B.; Van Heijst, G.J.F. A review on the CFD analysis of urban microclimate. Renew. Sustain. Energy Rev. 2017, 80, 1613–1640. [Google Scholar] [CrossRef]
  3. Ashie, Y.; Kono, T. Urban-scale CFD analysis in support of a climate-sensitive design for the Tokyo Bay area. Int. J. Climatol. 2011, 31, 174–188. [Google Scholar] [CrossRef]
  4. Mahgoub, A.O.; Ghani, S. Numerical and experimental investigation of utilizing the porous media model for windbreaks CFD simulation. Sustain. Cities Soc. 2021, 65, 102648. [Google Scholar] [CrossRef]
  5. Skote, M.; Sandberg, M.; Westerberg, U.; Claesson, L.; Johansson, A.V. Numerical and experimental studies of wind environment in an urban morphology. Atmos. Environ. 2005, 39, 6147–6158. [Google Scholar] [CrossRef]
  6. Antohe, B.V.; Lage, J.L. A general two-equation macroscopic turbulence model for incompressible flow in porous media. Int. J. Heat Mass Transf. 1997, 40, 3013–3024. [Google Scholar] [CrossRef]
  7. Hang, J.; Li, Y. Wind conditions in idealized building clusters: Macroscopic simulations using a porous turbulence model. Bound.-Layer Meteorol. 2010, 136, 129–159. [Google Scholar] [CrossRef] [Green Version]
  8. Gu, Z.; Zhang, Y. Large Eddy Simulation of Urban and Building Wind Environment and Application; Science Press: Beijing, China, 2014. (In Chinese) [Google Scholar]
  9. Wang, X.; Li, Y.; Hang, J. A combined fully-resolved and porous approach for building cluster wind flows. Build. Simul. 2017, 10, 97–109. [Google Scholar] [CrossRef]
  10. Tiantian, L.; Bingfeng, Y.; Zhangbao, H.; Zhi, C. Numerical Simulation on Impact of Building Density on Urban Heat Island with an Urban Porous Media Model. J. Xi’an Jiaotong Univ. 2012, 46, 134–138. [Google Scholar]
  11. Ming, T.; Lian, S.; Wu, Y.; Shi, T.; Peng, C.; Fang, Y.; de Richter, R.; Wong, N.H. Numerical investigation on the urban heat island effect by using a porous media model. Energies 2021, 14, 4681. [Google Scholar] [CrossRef]
  12. Sun, R.; Wang, Y.; Chen, T. Impacts of anthropogenic heat emissions on urban thermal environment: A review. Acta Ecol. Sin. 2017, 37, 3991–3997. (In Chinese) [Google Scholar]
  13. Kim, G.; Cha, D.H.; Song, C.K.; Kim, H. Impacts of Anthropogenic Heat and Building Height on Urban Precipitation Over the Seoul Metropolitan area in Regional Climate Modeling. J. Geophys. Res.-Atmos. 2021, 126, e2021JD035348. [Google Scholar] [CrossRef]
  14. Fung, K.Y.; Tam, C.Y.; Lee, T.C.; Wang, Z. Comparing the Influence of Global Warming and Urban Anthropogenic Heat on Extreme Precipitation in Urbanized Pearl River Delta Area Based on Dynamical Downscaling. J. Geophys. Res.-Atmos. 2021, 126, e2021JD035047. [Google Scholar] [CrossRef]
  15. Falasca, S.; Catalano, F.; Moroni, M. Numerical Study of the Daytime Planetary Boundary Layer over an Idealized Urban Area: Influence of Surface Properties, Anthropogenic Heat Flux, and Geostrophic Wind Intensity. J. Appl. Meteorol. Climatol. 2016, 55, 1021–1039. [Google Scholar] [CrossRef]
  16. Holst, C.C.; Tam, C.; Chan, J.C.L. Sensitivity of urban rainfall to anthropogenic heat flux: A numerical experiment. Geophys. Res. Lett. 2016, 43, 2240–2248. [Google Scholar] [CrossRef] [Green Version]
  17. Wen, Y.; Lian, Z. Influence of air conditioners utilization on urban thermal environment. Appl. Therm. Eng. 2009, 29, 670–675. [Google Scholar] [CrossRef]
  18. Quah, A.K.L.; Roth, M. Diurnal and weekly variation of anthropogenic heat emissions in a tropical city, Singapore. Atmos. Environ. 2012, 46, 92–103. [Google Scholar] [CrossRef]
  19. Flanner, M.G. Integrating anthropogenic heat flux with global climate models. Geophysical Research Letters 2009, 36. [Google Scholar] [CrossRef] [Green Version]
  20. Liu, P.; Li, L.P.; Liao, J.B.; Guo, X.M.; Zhang, Y.R. A study of the effect of anthropogenic heat on meteorological factors in the summer of Shanghai City. J. Lanzhou Univ. (Nat. Sci.) 2019, 55, 267–273. (In Chinese) [Google Scholar]
  21. Du, H.; Wang, D.; Wang, Y.; Zhao, X.; Qin, F.; Jiang, H.; Cai, Y. Influences of land cover types, meteorological conditions, anthropogenic heat and urban area on surface urban heat island in the Yangtze River Delta Urban Agglomeration. Sci. Total Environ. 2016, 571, 461–470. [Google Scholar] [CrossRef]
  22. Xie, M.; Zhu, K.; Wang, T.; Feng, W.; Zhu, X.; Chen, F.; Ouyang, Y.; Liu, Z. Study on the distribution of anthropogenic heat flux over China. China Environ. Sci. 2015, 35, 728–734. (In Chinese) [Google Scholar]
  23. Lu, Y.; Wang, Q.G.; Zhai, Y.R.; Song, Y.Y.; Zhang, Y.Y.; Sun, P. Anthropogenic heat emissions in the Yangtze River Delta region. China Environ. Sci. 2014, 34, 295–301. (In Chinese) [Google Scholar]
  24. Takane, Y.; Kikegawa, Y.; Hara, M.; Grimmond, C.S.B. Urban warming and future air-conditioning use in an Asian megacity: Importance of positive feedback. NPJ Clim. Atmos. Sci. 2019, 2, 39. [Google Scholar] [CrossRef] [Green Version]
  25. Ohashi, Y.; Suido, M.; Kikegawa, Y.; Ihara, T.; Shigeta, Y.; Nabeshima, M. Impact of seasonal variations in weekday electricity use on urban air temperature observed in Osaka, Japan. Q. J. R. Meteorol. Soc. 2016, 142, 971–982. [Google Scholar] [CrossRef]
  26. Tong, H.; Liu, H.Z.; Sang, J.G.; Hu, F. The Impact of Urban Anthropogenic Heat on Beijing Heat Environment. Clim. Environ. Res. 2004, 9, 409–421. (In Chinese) [Google Scholar]
  27. Jiang, W.; Chen, Y. The Impact of Anthropogenic Heat on Urban Boundary Layer Structure. Chin. J. Atmos. Sci. 2007, 31, 37–47. (In Chinese) [Google Scholar]
  28. Mei, S.; Yuan, C. Analytical and numerical study on transient urban street air warming induced by anthropogenic heat emission. Energy Build. 2021, 231, 110613. [Google Scholar] [CrossRef]
  29. Cao, Z.; Wen, Y.; Song, S.; Hung, C.H.; Sun, H. Spatiotemporal Variations and Controls on Anthropogenic Heat Fluxes in 12 Selected Cities in the Eastern China. Chin. Geogr. Sci. 2021, 31, 444–458. [Google Scholar] [CrossRef]
  30. Yuan, C.; Adelia, A.S.; Mei, S.; He, W.; Li, X.X.; Norford, L. Mitigating intensity of urban heat island by better understanding on urban morphology and anthropogenic heat dispersion. Build. Environ. 2020, 176, 106876. [Google Scholar] [CrossRef]
  31. Ma, J. Numerical Study of Wind Environment around Building Complexes; Zhejiang University: Zhejiang, China, 2006. (In Chinese) [Google Scholar]
  32. Leung, K.K.; Liu, C.; Wong, C.C.C.; Lo, J.C.; Ng, G.C. On the study of ventilation and pollutant removal over idealized two-dimensional urban street canyons. Build. Simul. 2012, 5, 359–369. [Google Scholar] [CrossRef]
  33. Luo, Y.H.; Zhu, S.Y.; Zhang, G.X.; Liu, Y.; Xiang, J.M.; Zhou, Y. Estimation and Analysis of Spatial Distribution of Urban Population During the Daytime and Nighttime in Qinhuai District of Nanjing. Resour. Environ. Yangtze Basin 2018, 27, 1020–1030. (In Chinese) [Google Scholar]
  34. Yang, H.L. Estimation on Temporal—Spatial Characteristics of Anthropogenic Heat in Qinhuai District of Nanjing Based on the Inventory Survey Method; Nanjing University of Information Science and Technology: Nanjing, China, 2018. [Google Scholar]
Figure 1. Research block zoning.
Figure 1. Research block zoning.
Buildings 12 00595 g001
Figure 2. Modeling method of porous media.
Figure 2. Modeling method of porous media.
Buildings 12 00595 g002
Figure 3. Modeling for the standard CFD method.
Figure 3. Modeling for the standard CFD method.
Buildings 12 00595 g003
Figure 4. Method for modeling porous media.
Figure 4. Method for modeling porous media.
Buildings 12 00595 g004
Figure 5. Mesh for the standard CFD method.
Figure 5. Mesh for the standard CFD method.
Buildings 12 00595 g005
Figure 6. Mesh for modeling porous media.
Figure 6. Mesh for modeling porous media.
Buildings 12 00595 g006
Figure 7. The field measurement points.
Figure 7. The field measurement points.
Buildings 12 00595 g007
Figure 8. Photo of field measurement.
Figure 8. Photo of field measurement.
Buildings 12 00595 g008
Figure 9. Comparison of measured and simulated (a) wind velocity and (b) temperature at different points.
Figure 9. Comparison of measured and simulated (a) wind velocity and (b) temperature at different points.
Buildings 12 00595 g009
Figure 10. Diagram of wind velocity distribution modeled by the (a) standard CFD method and (b) porous media method.
Figure 10. Diagram of wind velocity distribution modeled by the (a) standard CFD method and (b) porous media method.
Buildings 12 00595 g010
Figure 11. Diagram of temperature distribution modeled by the (a) standard CFD method and (b) porous media method.
Figure 11. Diagram of temperature distribution modeled by the (a) standard CFD method and (b) porous media method.
Buildings 12 00595 g011
Figure 12. (a) Mean wind velocity and (b) mean temperature in different block zones.
Figure 12. (a) Mean wind velocity and (b) mean temperature in different block zones.
Buildings 12 00595 g012
Figure 13. Effect of porosity on (a) wind velocity and (b) temperature at different heights.
Figure 13. Effect of porosity on (a) wind velocity and (b) temperature at different heights.
Buildings 12 00595 g013
Figure 14. Temperature of the area corresponding to the heights of 2 m (a), 10 m (b), and 50 m (c) under different intensities of anthropogenic heat.
Figure 14. Temperature of the area corresponding to the heights of 2 m (a), 10 m (b), and 50 m (c) under different intensities of anthropogenic heat.
Buildings 12 00595 g014
Table 1. Setting of boundary conditions.
Table 1. Setting of boundary conditions.
SurfacesStandard CFD ModelPorous Media Model
InletVelocity inletVelocity inlet
OutletOutflowOutflow
Top and sidesSymmetrySymmetry
Underlying surfaceNon-slipping wallNon-slipping wall
Building wallNon-slipping wall
Porous zoneInterior
Table 2. Measurement device parameters.
Table 2. Measurement device parameters.
NameMeasuring ParametersRangeResolutionPrecision
GM1361 split type of thermometerTemperature−10–50 °C0.1 °C±1.0 °C
AnemometerVelocity0~45 m/s0.1 m/s±0.3 m/s
Table 3. Measurement results and standard deviation of temperature.
Table 3. Measurement results and standard deviation of temperature.
ParametersVelocity (m/s)Temperature (K)Standard Deviation of Temperature
10.284310.10.116
20.345309.40.136
30.39309.20.196
40.317308.70.129
50.26308.10.143
60.22309.60.123
70.396309.10.219
80.54308.80.224
Table 4. Coefficient of determination.
Table 4. Coefficient of determination.
Simulation MethodVelocityTemperature
Porous media method0.860.67
Standard CFD simulation method0.940.83
Table 5. Simulation cost comparison.
Table 5. Simulation cost comparison.
Standard CFD Modeling MethodPorous Media Model Method
Number of grids3,268,9352,767,710
Time per iteration/s14~206~12
Number of iterations1516854
Simulation time/min480140
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lei, J.; Wang, D.; Chen, Z. Numerical Study of Micro-Thermal Environment in Block Based on Porous Media Model. Buildings 2022, 12, 595. https://doi.org/10.3390/buildings12050595

AMA Style

Lei J, Wang D, Chen Z. Numerical Study of Micro-Thermal Environment in Block Based on Porous Media Model. Buildings. 2022; 12(5):595. https://doi.org/10.3390/buildings12050595

Chicago/Turabian Style

Lei, Jie, Dengyun Wang, and Zhenqian Chen. 2022. "Numerical Study of Micro-Thermal Environment in Block Based on Porous Media Model" Buildings 12, no. 5: 595. https://doi.org/10.3390/buildings12050595

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