Next Article in Journal
Effects of Adding Native Annual Seeds to South Korea Native Perennial Seed Mixture on Early Stage Vegetation Recovery, Soil Enzymes, and Nutrient Dynamics in Post-Fire Soils
Previous Article in Journal
Introgression as an Important Driver of Geographic Genetic Differentiation within European White Oaks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heat Mitigation Benefits of Urban Trees: A Review of Mechanisms, Modeling, Validation and Simulation

1
School of Water Conservancy and Civil Engineering, South China Agricultural University, Wushan Road 483, Guangzhou 510642, China
2
School of Civil and Transportation Engineering, Guangdong University of Technology, Outer Ring West Road 100, Guangzhou 510006, China
3
Department of City and Regional Planning, The Ohio State University, Columbus, OH 43210, USA
4
School of Architecture, South China University of Technology, Guangzhou 510641, China
5
State Key Laboratory of Subtropical Building and Urban Science, South China University of Technology, Guangzhou 510641, China
6
Energy Saving Technology Research Institute, South China University of Technology, Guangzhou 510641, China
*
Author to whom correspondence should be addressed.
Forests 2023, 14(12), 2280; https://doi.org/10.3390/f14122280
Submission received: 22 October 2023 / Revised: 14 November 2023 / Accepted: 17 November 2023 / Published: 21 November 2023

Abstract

:
Modeling, validating, and simulating are three essential parts in investigating the heat mitigation benefits of urban trees (BUT). Therefore, 81 relevant studies from the last ten years are reviewed, analyzed, and summarized in this study. Three main ways for urban trees to adjust the environment are summarized, including shade creation and radiation modification, cooling effects of transpiration, and airflow blocking and modification effects. Research works are analyzed with regard to four categories: (1) heat and moisture exchange mechanisms and their mathematical modeling; (2) verification of modeling predictions based on measurements; (3) thermal performance simulation and prediction; and (4) environmental assessment and human thermal comfort analyses. Future research opportunities are discussed: (1) conduct real-time and in-depth measurements to analyze the mechanisms of heat and moisture transfer of trees in different areas; (2) develop tree radiation attenuation, airflow resistance, and transpiration models to accurately describe heat and moisture transfer processes in the urban environment; and (3) establish a three-dimensional numerical simulation method that can accurately simulate the urban thermal environment with trees. This review provides researchers with an overview and potential research opportunities on the thermal effects of urban trees.

1. Introduction

The urban heat island (UHI) reduces the quality of life of urban dwellers [1], changes the local hydrology and subsurface structure [2], and causes several ecological problems [3,4]. Recent research indicates that a rational vegetation layout is an effective means for alleviating the UHI and preventing the mortality burden [5], reducing energy consumption, and providing better outdoor thermal comfort [6]. Wong et al. [7] evaluated the potential of green infrastructure as a mitigation strategy and found that greenery on the ground reduces the peak surface temperature by 2–9 °C, while green roofs and green walls reduce the surface temperature by 17 °C. Bowler et al. [8] used a systematic review methodology and meta-analysis to evaluate the available evidence on whether greening interventions affect the air temperature of an urban area and found that, on average, a park with trees was 0.94 °C cooler in the day. Thus, focusing on climatic adaptive design that combines elements of landscape with microclimatic factors is an important way to deal with urban ecological problems and to improve the urban thermal environment [9,10].
Trees, as one of the most important urban landscape elements, play a crucial role in mitigating the UHI [11]. Radiation attenuation and transpiration are important tree factors for the balance of surface energy and the water vapor cycle of the hydrosphere–atmosphere–biosphere [12]. Through canopy shading and transpiration [13], trees can adjust the outdoor radiant heat, affect the heat and moisture balance of the surrounding environment [14], and change the outdoor microclimate and urban environment. Trees can also reduce wind speed, enhance airflow turbulence [15], change outdoor wind flow characteristics, reduce convective heat and mass transfer coefficients [16], and further affect the heat and moisture balance between trees and the surrounding environment.
The physical properties of trees, such as diameter at breast height, crown width, and leaf area index (LAI), are paramount factors in regulating the microclimate and increasing user comfort [17,18,19]. Due to the significant differences in root depth, crown width, LAI, tree morphology, and leaf reflectance among different tree species, the cooling performance of different tree species varies greatly under different climate conditions. The complexity of the heat exchange process of trees and the climatic particularity of “wind, heat, rain, and humidity” in different areas ensure that the heat and mass transfer between trees and the environment is a coupled and complex process. How to accurately describe this process has become an urgent problem in disciplines such as architectural technology, landscape architecture, urban planning and design, and urban climatology.
Of particular importance to this review, the latest progress in urban climate modeling indicates that it is crucial to incorporate urban trees into urban canopy models (UCMs) so as to realistically capture the surface energy budget and achieve an accurate simulation of the urban thermal environment [20,21,22,23]. However, as the modeling of tree radiation attenuation, tree airflow resistance, and tree transpiration is still lacking [19,24,25], the current urban thermal environment assessment tools cannot achieve accurate simulations of tree heat and moisture transfer processes and, thus, cannot accurately predict the urban energy balance, natural ventilation, and outdoor thermal comfort in the presence of trees, nor can they predict the thermal environment performance of design plans. Numerical simulations under certain assumptions or simplifications cannot yield accurate and universal models and need to be further developed and improved.
Previous related reviews have mainly focused on urban green space cooling effects [26,27,28,29], such as threshold size [29], cooling potential [28], contribution to ambient air cooling [27], and enhancement of human comfort [26], with limited insights on trees, particularly their heat and moisture exchange mechanisms and their mathematical modeling. It is, therefore, the purpose of this paper to present a comprehensive review and analysis of recent developments involving the heat mitigation benefits of urban trees, including (1) impact mechanisms of urban trees on the urban thermal environment; (2) heat and moisture transfer mechanisms between trees and the environment and related mathematical models. This study focuses on the experimental testing of trees, theoretical methods, mathematical models, numerical simulations, and validations. The opportunities to expand research on the thermal effects of urban trees will also be presented.

2. Data and Methodology

To understand to what extent the available guidelines of tree heat and moisture exchange mechanisms and their mathematical modeling could fulfill researchers’ real needs, we attempted a condensed literature survey for the last ten years in the tropics. We developed a four-step workflow that serves as the analytical framework: (1) reviewing the mechanism of urban trees’ influence on the urban thermal environment; (2) reviewing the mechanism of heat and moisture exchange and its mathematical modeling; (3) reviewing and analyzing the research results of recent years on heat mitigation by trees; and (4) reviewing pending problems in the heat and moisture transfer mechanism and related modeling. The eligibility criterion “tropics” was defined by cities located at 23.5° S–23.5° N and the criterion “subtropical” was defined by cities located at 23.5–35° S and 23.5–35° N latitudes.
In the literature review, the Web of Science was chosen for the literature search. The keywords “tree”, “human thermal comfort”, “microclimate”, “urban heat island”, and “radiation modification” were used as the topics for the literature search. The literature was limited to the last 10 years. The total number of articles that emerged from the initial review was 136. All articles were subsequently manually reviewed, and only those field and modelling studies closely related to the impact of urban trees on urban microclimate were selected. All selected field measurements studies needed to include basic information, such as date, time, tree’s species country name, city name, number of stations, locations of stations, measurement type, and measured variables. Based on the topics discussed in the paper, more relevant keywords, such as “influence mechanism”, “exchange mechanism”, “mathematical modeling”, “simulation”, “ENVI-met”, and “modelling” were selected to limit the search. Finally, 81 papers were included in the in-depth analysis. A flowchart of Paper Retrieval Research is shown in Figure 1. The four-part framework analysis and article citations are described specifically in the following sections.

3. Impact Mechanisms of Urban Trees on the Urban Thermal Environment

The heat mitigation benefits of urban trees have complex internal and external causes. It is necessary to consider not only their physical characteristics, such as tree species, crown albedo, leaf reflectance, transpiration rate, and leaf area index, but also the impacts of different microclimatic parameters on a tree’s heat and moisture transfer processes.
There are three main ways for trees to adjust the environmental microclimate [29]: (1) due to the unique albedo and absorptivity of the tree crown, trees can reduce solar heat reaching the ground by absorbing and reflecting solar radiation; (2) the transpiration of trees can reduce surface and ambient temperatures; and (3) trees can guide and control the airflow by blocking and filtering it.

3.1. Shade Creation and Radiation Modification

The attenuation of solar radiation by the tree canopy is a crucial way to reduce the UHI, affecting the surface energy balance and improving the urban microclimate [28]. The canopy can reduce solar radiation, modify heat exchanges between buildings and the surrounding environment, and affect human outdoor thermal comfort [27]. Trees can also indirectly reduce surface temperatures by modifying ground radiation [30]. The ability of trees to attenuate the UHI is mainly related to a reduction in direct solar radiation, especially visible and near-infrared light [11,31,32,33]. Current research is mainly focused on tropical, arid, and temperate regions and their native tree species. Little is known in humid and hot climates [13,34].

3.2. Transpiration and Its Cooling Effects

The transpiration of trees is the main reason for the cooling and humidification effects [35,36]. The transpiration process of trees can convert heat energy into latent heat, increase the humidity, and reduce the temperature of the surrounding environment [37,38]. In a hot dry climate, trees can evaporate about 100 gallons of water per day [36]. On a typical summer weather day, because of the transpiration, approximately 33% of incoming solar radiation is converted into latent heat [39]. Quantifying the cooling and humidification effects caused by trees requires accurate tree transpiration rates [40].

3.3. Wind Flow Modification and Blocking Effects

Trees also have significant implications for wind flow, which significantly impacts outdoor thermal comfort [41,42], energy efficiency [3,43], urban pollutant dispersion [22], and urban heat island mitigation [44]. The insulation effect of trees in terms of windward and leeward wind speed reduction is related not only to their physiological characteristics, such as spacing, size, porosity, and orientation, but also to their location and surrounding environment. Shahidan [18] found that, in a typical urban area, a tree’s physical parameters (leaf area index, crown width, and branches, etc.) have significant implications for airflow and wind environment, and that the leaf area index (LAI) and crown width are the main factors affecting the outdoor wind environment. Zheng et al. [45] found that the control of wind speed and direction by trees further affects air temperature. In the mainstream wind direction, the distance from a tree influencing air temperature is about five-times its height, but only two-times in the case of a non-mainstream direction.

4. Heat and Moisture Exchange Mechanisms and Their Mathematical Modeling

Through canopy shading and transpiration, trees can reduce solar radiation, affect the heat and moisture balance of the surrounding environment, and change the outdoor microclimate. At the same time, trees can also reduce wind speed, enhance airflow turbulence, change outdoor flow field characteristics and wind field distribution, and further affect the heat and moisture balance between trees and the surrounding environment. In order to realistically capture and present these phenomena, the common method is to load the tree’s radiative heat transfer, convective heat transfer, and transpiration latent heat into the boundary layer’s energy balance equations when discussing the regional heat balance. The goal is to control the energy balance of the body [46,47].

4.1. Transpiration Mechanism and Model

There are three evapotranspiration models commonly used in the agriculture and hydrology fields: Priestley–Taylor, Penman–Monteith, and Shuttleworth–Wallace (S-W) [48,49,50]. Only the S-W model can be used for urban trees as it comprehensively considers the canopy and soil source evapotranspiration processes [51,52,53,54].
The S-W model calculations are as follows [50]:
λ ET = C c ET c × C s ET s
ET c = Δ R n G + ρ C P e s e a Δ r a c R n s G / r a a + r a c Δ + γ 1 + r s s r a a + r a c
ET s = Δ R n G + ρ C P e s e a Δ r a s R n s G / r a a + r a s Δ + γ 1 + r s s r a a + r a s
C C = 1 1 + R s R a / [ R s R c + R a ]
C s = 1 1 + R s R a / [ R c R s + R a ]
R a = Δ + γ r a a
R c = Δ + γ r a c + γ r s c
R s = Δ + γ r a s + γ r s s
The nomenclature of Equations (1)–(8) is presented in Table 1.
The heat transfer structure of the S-W model is present in Figure 2. The S-W model integrates two source terms of evapotranspiration from the soil and the plant canopy by introducing soil resistance and canopy resistance parameters. It has been widely used in recent years.

4.2. Mechanisms and Modeling of Shade Creation and Radiative Properties Modification

In order to calculate the shielding effect of the canopy on solar radiation, there are two common calculation methods: transmittance and numerical simulation. The transmittance method obtains the solar radiation transmittance ( σ ) of the canopy by measuring the solar radiation (S0) and incident solar radiation (St) under the tree, as shown in Figure 3. According to Zheng et al. [19] and Shahidan et al. [18], the solar radiation transmittance of single-layer leaves is quite different: 10% of visible radiation and 30% of infrared radiation (Figure 4).
Only a part of solar radiation (>0.38 µm) has thermal effects on trees [55,56]. From a biological viewpoint, solar radiation <4 µm can be divided into three parts: ultraviolet radiation (UV), photosynthetically active radiation (PAR), and infrared radiation. The thermal effects of this radiation are provided in Table 2, which shows that UV does not have a thermal effect on trees. Near-infrared radiation has mainly thermal effects and can be absorbed by water in the stems of the leaves. Far-infrared radiation can only warm the tree.

4.3. Mechanism and Modeling of Canopy Flow

The influence of trees on the airflow field is related to their enhancing turbulence and reducing wind speed. In numerical simulation, in order to facilitate research, a tree’s physical model is divided into two parts: trunk and crown (Figure 5). The crown is considered as a porous medium because of its air permeability. In order to reflect the obstruction effect of trees on airflow, current microclimatic models generally modify the three-dimensional momentum equation by adding source terms [45,57].
F d = 1 2 C d η au i u i 2 0.5
where Fd is the resistance source term caused by the tree, and Cd is the drag coefficient.
The key for this method is to obtain the drag coefficient. Because the resistance coefficient of trees is difficult to obtain, most of these models use empirical resistance coefficients to simplify the exchange of momentum between the trees and their environment. It is assumed that the empirical resistance coefficient is constant and independent of wind speed and direction. In temperate regions, the empirical resistance coefficient usually varies between 0.1 and 0.3. Therefore, the default drag coefficient in the ENVI-Met tree model was set at 0.2 [58]. However, the latest research [57] shows that the actual resistance coefficient of trees in different climatic regions differs greatly from the empirical resistance coefficient in the actual situation, and the drag coefficient decreases with increasing wind speed, as shown in Figure 5 [57]. In addition, the drag coefficient is quite different in different areas. The resistance coefficient of common trees is 0.6 in the Mediterranean climate [59] and 0.8 in tropical regions [60]. In order to accurately simulate the momentum exchange between the tree and its environment, it is necessary to obtain the real drag coefficient in a given area.
For a scaled tree model, the aerodynamic performances of real trees in the wind, such as reconfigurations and the change in the projection area of a tree crown against incoming wind, are difficult to simulate. Based on wind tunnel experiments, Manickathan et al. [59] and Cao et al. [60] found that the reconfiguration phenomenon is mainly affected by the branch stiffness and wind speed. This phenomenon affects the characteristics of tree forces and tree forms, especially the frontal area of trees (the area of the orthographic projection of the tree on a plane perpendicular to the wind direction) and drag coefficients. The frontal area of trees and drag coefficient decrease with increasing wind speed (U) (Table 3). The negative exponential relationships between drag coefficients and wind speed (U) can be expressed by the formulas Cd = a × U−b [60].

5. Review and Analysis of Recent Research Works on Heat Mitigation by Trees

In this section, recent works on the heat mitigation benefits of urban trees (HMBUT) are reviewed and analyzed in the order of the publication year. Works associated with HMBUT were found to be substantial and were analyzed along the following themes: (1) microclimate benefit performance evaluation through measurement; (2) thermal performance simulation and prediction; (3) verification of modeling prediction based on measurement; and (4) environmental assessment and human thermal comfort analyses.

5.1. Microclimate Benefit Performance Evaluation through Measurement

The experimental research on trees in the field of urban microclimate mainly concerns microclimatic data (air temperature, humidity, solar radiation, wind direction, and wind speed, etc.) at measurement points and compares them with a tree’s physiological parameters (three-dimensional green quantity, leaf area index, canopy cover, canopy closure, plant coverage, average leaf inclination, etc.), as presented in Figure 6 and Table 4 [61,62]. Some scholars have also combined thermal environment simulation software, such as Envi-met 4.2 and Airpak 3.0, to analyze the impact of landscape design methods on the microclimate [63,64,65].
There are three main ways in which trees adjust the environmental microclimate: solar radiation modification, transpiration, and blocking effect on airflow.
The solar radiation attenuation by the canopy is mainly affected by physical factors, such as branches and leaves, which differ somewhat across tree species. Therefore, the solar radiation attenuation performance of different tree species varies greatly, especially in different climate regions. Kotzen et al. [17] tested this attenuation effect for street trees common in tropical regions, and they analyzed the effects of solar radiation intensity, incident angle characteristics, and canopy leaf area density on solar radiation attenuation. Based on measured data, Akbari [26] discussed radiation occlusion and transmission mechanisms, indicating that planting design needs to take into account tree canopy density, tree height, canopy transmittance in different seasons, and canopy structure levels.
The transpiration of trees is an important influencing factor on the surface energy balance and cooling effects [66,67]. In order to quantify the cooling and humidification caused by trees, we need to accurately obtain a tree’s transpiration rate. At present, there are two main methods for measuring tree transpiration rates [26,68]: measuring the convective mass transfer coefficient (a) and air humidity on the blade surface. However, it is difficult to obtain the convective mass transfer coefficient in heterogeneous urban environments. Another is the trunk runoff method, which can use the trunk runoff meter to measure the liquid flow for a long time, but the instrument is usually very expensive and causes significant damage to the tree. Due to the limitations of the above methods, there are few experimental studies on tree transpiration rates. Akbari [26] found that trees can evaporate about 100 gallons of water per day in dry and hot climates. If evapotranspiration is combined with proper layout and shade, the temperature drop caused by nearby trees can reach 9 °C. Chen et al. [69] established a regression model for calculating the biomass of garden plants by measuring the daily transpiration rates of common trees in Beijing. Han [70] tested the transpiration rate, ecological effect, and utilization of light energy of common tree species in severely cold areas. By simplifying the calculation of the transpiration heat transfer, the cooling effect of different tree species in different months was obtained.
A tree’s blocking effects on airflow are not only related to its location and surrounding environment but also to the tree’s characteristics, such as size, orientation, porosity, and canopy density. Many researchers at home and abroad have conducted studies on the effects of trees on the near-surface wind environment. Shahidan et al. [18] found that, in a typical urban area, the physical parameters (leaf area index, crown width, and branches, etc.) of different trees yield large differences in impacts on airflow, and they also have a great impact on the wind environment, especially the leaf area index (LAI) and crown width. Heisler [71] found that the canopy blocking effect on wind speed in residential areas depends on the density of the canopy. Increasing the density by 10% can reduce wind speed by 10% to 20%, and increasing it by 30% can reduce wind speed by 15% to 35%.
In addition, the control of wind pressure and direction by trees will further affect the urban microclimate. Zheng [45] found that, if a site is located in the downwind direction of the plant coverage area, trees can play a role in reducing wind speed and wind pressure. Planting a dense row of trees can concentrate and strengthen the airflow under the canopy and improve ventilation conditions at the ground level under the trees. Dimoudi and Nikolopoulou [46] found that in the mainstream wind direction, the influencing distance of a tree on temperature is about five-times its height, and a rational tree layout can effectively improve the thermal comfort of pedestrians around a building. However, when the tree is in the non-mainstream wind direction, the impact is not obvious. Therefore, many studies have suggested that pedestrian comfort should be improved by combining urban greening with the main ventilation channels of urban areas.

5.2. Simulation and Prediction of Thermal Performance

In recent years, with the continuous improvements in computer performance, numerical simulation has become the main research method for the quantitative prediction and evaluation of urban thermal environments. It is necessary to include urban trees in numerical simulations to accurately calculate the urban surface energy balance and achieve accurate simulation of the UHI [19,36]. Because of the complexity of trees’ heat and moisture transfer processes and the diversity of their geometric shapes and spatial locations, it is very difficult to create 3D tree models in a given urban street environment, accounting for their spatial locations and sizes [36]. In order to meet this challenge, the commonly used approach is to use existing simulation software.
ENVI-met and ANSYS Fluent are CFD models that are widely used in microclimate simulation and outdoor thermal comfort studies. ENVI-met is a three-dimensional urban microclimate simulation software developed by Bruse and Fleer in 1998. It is based on heat transfer and computational fluid dynamics, and it is mainly used to simulate at the urban block scale, across ground, buildings, vegetation, and the atmosphere [72,73]. ANSYS Fluent, a general-purpose CFD platform based on the Finite Volume, provides comprehensive modelling of fluid flows under steady or transient conditions [74]. Since ANSYS Fluent requires the user to formulate a specific problem via user-defined functions, it requires a high level of physics expertise [75]. Until now, most numerical simulations of the impact of trees on the outdoor thermal environment have been carried out using ENVI-met. Zhang [67] used ENVI-met to simulate the arrangement of eight tree species in residential areas in summer and winter and found that the tree spacing ratio is essential to improve the outdoor thermal environment. Duarte [66] used ENVI-met to explore the influence of trees on air temperature and found that densely planted street trees are cooler than central and pocket parks. Chen [76] studied the impact of common tree species in humid and hot areas by coupling the energy consumption simulation software EQUEST 2.0 with ENVI-met 4.2.
However, because of the complexity of trees’ impact on the outdoor environment, ENVI-met simplifies the tree model as follows [19,57,58]: (1) in terms of solar radiation, ENVI-met only considers the attenuation of direct solar radiation by the tree canopy and does not consider the influence of trees on long-wave radiation and heat transfer between trees and the surrounding environment; (2) ENVI-met adopts an empirical resistance coefficient (0.2), which cannot be modified according to the actual species. These simplifications may cause ENVI-met to inaccurately simulate heat and mass exchanges between a tree and its surrounding environment.

5.3. Verification of Modeling Prediction Based on Measurement

Many researchers have evaluated the ENVI-met tree model in their own climate zones [72,73,77]. Zheng et al. [19,57] verified ENVI-met accuracy in hot and humid areas, showing that ENVI-met greatly simplifies the calculation processes of radiation, convection, and transpiration between trees and the environment, with large deviations in the simulation of radiation attenuation, wind speed, and transpiration rate. The root mean square error between the simulated and measured values of solar radiation under a tree reaches 256 W/m2.
With the development of computational fluid dynamics (CFD) technology, some studies have represented the effect of trees on airflow and heat and moisture transfer by adding source terms to the Navier–Stokes (N-S) equation [78]. Upreti [79] studied trees’ radiation attenuation, canopy flow, and heat and mass transfer, using the Monte Carlo method to calculate solar radiation and long-wave radiation attenuation by the tree canopy with a structured grid, simplifying the calculation of canopy radiation transmission by the method of spherical crown envelope surface. Gao and Long [80] coupled the CFD simulation of airflow with outdoor radiation calculation but did not calculate long- and short-wave radiation, nor did they involve the coupling of the tree canopy energy equation with the convection, heat transfer, and radiation equations of the surrounding environment. Argiro [46] simplified the microclimate model of trees by using fixed solar transmittance and transpiration rates. Using this model, they analyzed the microclimatic benefits of trees in the urban environment and conducted a parameter sensitivity analysis of the simplified tree model, which assumes that the sunlight transmittance of a tree canopy is constant and does not consider long-wave radiation.

6. Outlines of Further Research

Tree canopy shading, transpiration, and airflow obstruction have important effects on solar radiation [81], water circulation [82], air temperature [83], and wind environment, as widely recognized by scholars around the world [84,85,86,87]. However, there are still several pending problems related to heat and moisture transfer mechanisms and related modeling, as detailed below.

6.1. Conducting Comprehensive and In-Depth Measurements to Analyze the Mechanisms of Tree Heat and Moisture Transfer in Different Areas

Although recent research on the impact of trees on the urban thermal environment has achieved fruitful qualitative results [88,89,90,91], there are few studies on tree heat and mass transfer under the coupling of radiation, convection, and transpiration. Also, current relevant research is mainly concentrated on tropical, dry, and temperate regions, with little known about common trees in humid and hot areas [92,93,94,95]. The native tree species growing in humid and hot areas have completely different tree shapes and philological characteristics, as compared with trees in other areas. More research is needed to determine their own heat and moisture transfer laws and mechanisms.

6.2. Developing Tree Radiation Attenuation, Airflow Resistance, and Transpiration Models to Accurately Represent Heat and Moisture Transfer Processes in Urban Environments

6.2.1. Radiation Attenuation Model

Previous radiation attenuation models have not included long- and short-wave radiation calculations, or they have only calculated tree canopy radiation in the one-dimensional case [96,97,98]. As a result, it is impossible to accurately simulate the occlusion, reflection, transmission, and absorption of short-wave solar radiation by the tree canopy and the long-wave radiation heat transfer with the surrounding environment. This hinders further research on the heat mitigation benefits of urban trees.

6.2.2. Airflow Resistance Model

The obstruction of airflow by urban trees affects the urban wind environment and ventilation, which, in turn, affects urban pollutant diffusion, energy distribution, and outdoor thermal comfort [99,100,101,102]. In order to reflect the obstruction effect of trees on airflow, current microclimatic models generally modify the three-dimensional momentum equation by adding source terms [103,104,105,106]. The key to this method is to obtain the drag coefficient, which varies across different trees. It is urgent to obtain the actual resistance coefficient of the tree species in a given area to specify a flow resistance model and then accurately simulate the resistance characteristics of the trees and the momentum exchange with the environment.

6.2.3. Transpiration Model

The commonly used transpiration models (P-M, P-T and S-W) have all been established and widely used in agriculture. A transpiration model is also urgently needed to accurately represent the urban tree transpiration process and its impact on the heat island effect [107,108,109,110]. However, due to differences in plant and soil properties between urban trees and crops, the transpiration model established in the agricultural field may not accurately represent the transpiration process of urban trees. It is an urgent problem to establish a transpiration model suitable for urban trees.

6.3. Establishing a Three-Dimensional Numerical Simulation Method That Can Accurately Simulate the Urban Thermal Environment with Trees

At present, the existing urban thermal environment evaluation tools either do not fully consider radiation attenuation, airflow resistance, and transpiration of trees or cannot accurately simulate the thermal environment with trees [111,112,113,114,115]. Zheng et al. [8,56,59] assessed the commonly used simulation software ENVI-met 4.2 and found that it cannot accurately simulate trees’ solar radiation, flow resistance, and transpiration rate. Therefore, in order to accurately simulate the heat and moisture transfer of trees in urban environments and to improve the predictability of the energy balance, natural ventilation, and outdoor thermal comfort [116,117,118,119,120,121,122], a three-dimensional numerical simulation model that can accurately simulate the urban thermal environment with trees is urgently needed.

7. Conclusions

Urban trees’ thermal effects play a key role in mitigating the UHI, reducing residential energy consumption, and improving outdoor thermal comfort. A critical review of urban trees’ thermal effects was carried out, and the influencing mechanisms of urban trees’ thermal effects were described, as well as their classifications into (1) shade creation and radiation modification, (2) transpiration and its cooling effects, and (3) wind flow modification and its blocking effects, based on urban trees’ heat and moisture exchange mechanisms. Mathematical equations and models for radiation modification, transpiration, and wind flow modification were presented.
Research opportunities related to urban trees’ thermal effects are numerous, but the major points lie in the following: (1) heat and moisture exchange mechanisms and their mathematical modeling; (2) verification of modeling predictions based on measurements; (3) thermal performance simulation and prediction; and (4) environmental assessment and human thermal comfort analyses. In view of the current research status and outstanding problems, further research opportunities are outlined as follows: (1) conducting comprehensive and in-depth measurements to analyze the mechanisms of heat and moisture transfer of trees in different areas; (2) developing tree radiation attenuation, flow resistance, and transpiration models to accurately describe the heat and moisture transfer processes of trees in the urban environment; and (3) establishing a three-dimensional numerical simulation method that can accurately simulate the urban thermal environment with trees.
This review focused on urban trees. Less relevant types of green spaces, such as green roofs and green walls, could be considered in a future review.

Author Contributions

Conceptualization: S.Z. and X.L.; methodology and formal analysis: S.Z. and X.L.; software, resources, and data curation: S.Z., C.H. and H.X.; validation: S.Z., X.L., C.H. and H.X.; investigation and visualization: S.Z., X.L., C.H. and H.X.; writing—original draft preparation and writing—review and editing: S.Z., X.L. and J.-M.G.; funding acquisition: S.Z. and X.L.; supervision: S.Z., X.L. and J.-M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (grant nos. 52008115 and 52108011); the Fundamental Research Funds for the Central Universities (grant no. QNMS202211); Guangdong Basic and Applied Basic Research Foundation (grant no. 2023A1515011137); the 2022 Guangdong Philosophy and Social Science Foundation (grant no. GD22XGL02); Guangzhou Philosophy and Social Science Planning 2022 Annual Project (grant no. 2022GZQN14); Department of Housing and Urban-Rural Development of Guangdong Province (grant no. 2021-K2-305243); Department of Education of Guangdong Province (grant no. 2021KTSCX004); Guangzhou Basic and Applied Basic Research Foundation (grant no. SL2024A04J00890).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brysse, K.; Oreskes, N.; O’Reilly, J.; Oppenheimer, M. Climate change prediction: Erring on the side of least drama? Glob. Environ. Change-Hum. Policy Dimens. 2013, 23, 327–337. [Google Scholar] [CrossRef]
  2. Vailshery, L.S.; Jaganmohan, M.; Nagendra, H. Effect of street trees on microclimate and air pollution in a tropical city. Urban For. Urban Green. 2013, 12, 408–415. [Google Scholar] [CrossRef]
  3. Wang, J.; Meng, Q.; Tan, K.; Zhang, L.; Zhang, Y. Experimental investigation on the influence of evaporative cooling of permeable pavements on outdoor thermal environment. Build. Environ. 2018, 140, 184–193. [Google Scholar] [CrossRef]
  4. Grimmond, C.S.B.; Blackett, M.; Best, M.J.; Baik, J.; Belcher, S.E.; Beringer, J.; Bohnenstengel, S.I.; Calmet, I.; Chen, F.; Coutts, A.; et al. Initial results from Phase 2 of the international urban energy balance model comparison. Int. J. Clim. 2011, 31, 244–272. [Google Scholar] [CrossRef]
  5. Iungman, T.; Cirach, M.; Marando, F.; Barboza, E.P.; Khomenko, S.; Masselot, P.; Quijal-Zamorano, M.; Mueller, N.; Gasparrini, A.; Urquiza, J.; et al. Cooling cities through urban green infrastructure: A health impact assessment of European cities. Lancet 2023, 401, 577–589. [Google Scholar] [CrossRef]
  6. Fu, J.; Dupre, K.; Tavares, S.; King, D.; Banhalmi-Zakar, Z. Optimized greenery configuration to mitigate urban heat: A decade systematic review. Front. Arch. Res. 2022, 11, 466–491. [Google Scholar] [CrossRef]
  7. Wong, N.H.; Tan, C.L.; Kolokotsa, D.D.; Takebayashi, K. Greenery as a mitigation and adaptation strategy to urban heat. Nat. Rev. Earth Environ. 2021, 2, 166–181. [Google Scholar] [CrossRef]
  8. Bowler, D.E.; Buyung-Ali, L.; Knight, T.M.; Pullin, A.S. Urban greening to cool towns and cities: A systematic review of the empirical evidence. Landsc. Urban Plan. 2010, 97, 147–155. [Google Scholar] [CrossRef]
  9. Tamaskani Esfehankalateh, A.; Ngarambe, J.; Yun, G.Y. Influence of tree canopy coverage and leaf area density on urban heat island mitigation. Sustainability 2021, 13, 7496. [Google Scholar] [CrossRef]
  10. Morgan, G.R.; Fulham, A.; Farmer, T.G. Machine Learning in Urban Tree Canopy Mapping: A Columbia, SC Case Study for Urban Heat Island Analysis. Geographies 2023, 3, 359–374. [Google Scholar] [CrossRef]
  11. Frank, S.D.; Backe, K.M. Effects of Urban Heat Islands on Temperate Forest Trees and Arthropods. Curr. For. Rep. 2023, 9, 48–57. [Google Scholar] [CrossRef]
  12. Ribeiro, A.P.; Bollmann, H.A.; de Oliveira, A.; Rakauskas, F.; Cortese, T.T.P.; Rodrigues, M.S.C.; Quaresma, C.C.; Ferreira, M.L. The role of tree landscape to reduce effects of urban heat islands: A study in two Brazilian cities. Trees 2023, 37, 17–30. [Google Scholar] [CrossRef]
  13. Shashua-Bar, L.; Rahman, M.A.; Moser-Reischl, A.; Peeters, A.; Franceschi, E.; Pretzsch, H.; Rötzer, T.; Pauleit, S.; Winters, G.; Groner, E.; et al. Do urban tree hydraulics limit their transpirational cooling? A comparison between temperate and hot arid climates. Urban Clim. 2023, 49, 101554. [Google Scholar] [CrossRef]
  14. Huang, J.; Kong, F.; Yin, H.; Middel, A.; Liu, H.; Zheng, X.; Wen, Z.; Wang, D. Transpirational cooling and physiological responses of trees to heat. Agric. For. Meteorol. 2022, 320, 108940. [Google Scholar] [CrossRef]
  15. Zhao, Y.; Li, H.; Bardhan, R.; Kubilay, A.; Li, Q.; Carmeliet, J. The time-evolving impact of tree size on nighttime street canyon microclimate: Wind tunnel modeling of aerodynamic effects and heat removal. Urban Clim. 2023, 49, 101528. [Google Scholar] [CrossRef]
  16. Lai, D.; Liu, Y.; Liao, M.; Yu, B. Effects of different tree layouts on outdoor thermal comfort of green space in summer Shanghai. Urban Clim. 2023, 47, 101398. [Google Scholar] [CrossRef]
  17. Kotzen, B. An investigation of shade under six different tree species of the Negev desert towards their potential use for enhancing micro-climatic conditions in landscape architectural development. J. Arid Environ. 2003, 55, 231–274. [Google Scholar] [CrossRef]
  18. Shahidan, M.F.; Shariff, M.K.; Jones, P.; Salleh, E.; Abdullah, A.M. A comparison of Mesua ferrea L. and Hura crepitans L. for shade creation and radiation modification in improving thermal comfort. Landsc. Urban Plan. 2010, 97, 168–181. [Google Scholar] [CrossRef]
  19. Zheng, S.; Guldmann, J.-M.; Liu, Z.; Zhao, L.; Wang, J.; Pan, X. Modeling of shade creation and radiation modification by four tree species in hot and humid areas: Case study of Guangzhou, China. Urban For. Urban Green. 2020, 47, 126545. [Google Scholar] [CrossRef]
  20. De Abreu-Harbicha, L.V.; Labakia, L.C.; Matzarakis, A. Effect of tree planting design and tree species on human thermal comfort in the tropics. Landsc. Urban Plan. 2015, 138, 99–109. [Google Scholar] [CrossRef]
  21. Rahman, M.A.; Moser, A.; Gold, A.; Rötzer, T.; Pauleit, S. Vertical air temperature gradients under the shade of two contrasting urban tree species during different types of summer days. Sci. Total Environ. 2018, 633, 100–111. [Google Scholar] [CrossRef]
  22. Shirzadi, M.; Naghashzadegan, M.; Mirzaei, P.A. Improving the CFD modelling of cross-ventilation in highly-packed urban areas. Sustain. Cities Soc. 2018, 37, 451–465. [Google Scholar] [CrossRef]
  23. Wang, Z.-H. Monte Carlo simulations of radiative heat exchange in a street canyon with trees. Sol. Energy 2014, 110, 704–713. [Google Scholar] [CrossRef]
  24. Li, R.; Zeng, F.; Zhao, Y.; Wu, Y.; Niu, J.; Wang, L.; Gao, N.; Shi, X. CFD simulations of the tree effect on the outdoor microclimate by coupling the canopy energy balance model. Build. Environ. 2023, 230, 109995. [Google Scholar] [CrossRef]
  25. Wang, L.; Su, J.; Gu, Z.; Tang, L. Numerical study on flow field and pollutant dispersion in an ideal street canyon within a real tree model at different wind velocities. Comput. Math. Appl. 2021, 81, 679–692. [Google Scholar] [CrossRef]
  26. Akbari, H.; Davis, S.; Huang, J.; Dorsano, S.; Winnett, S. Cooling Our Communities: A Guidebook on Tree Planting and Light-Colored Surfacing; Lab. (LBNL): Berkeley, CA, USA; Environmental Protection Agency: Washington, DC, USA; Climate Change Div.: Washington, DC, USA, 1992.
  27. Saaroni, H.; Amorim, J.; Hiemstra, J.; Pearlmutter, D. Urban Green Infrastructure as a tool for urban heat mitigation: Survey of research methodologies and findings across different climatic regions. Urban Clim. 2018, 24, 94–110. [Google Scholar] [CrossRef]
  28. Santamouris, M.; Ban-Weiss, G.; Osmond, P.; Paolini, R.; Synnefa, A.; Cartalis, C.; Muscio, A.; Zinzi, M.; Morakinyo, T.E.; Ng, E.; et al. Progress in urban greenery mitigation science—Assesment methodologies advanced technologies and impact on cities. J. Civ. Eng. Manag. 2018, 24, 638–671. [Google Scholar] [CrossRef]
  29. Yu, Z.; Yang, G.; Zuo, S.; Jørgensen, G.; Koga, M.; Vejre, H. Critical review on the cooling effect of urban blue-green space: A threshold-size perspective. Urban For. Urban Green. 2020, 49, 126630. [Google Scholar] [CrossRef]
  30. Zheng, S.; Guldmann, J.-M.; Liu, Z.; Zhao, L. Influence of trees on the outdoor thermal environment in subtropical areas: An experimental study in Guangzhou, China. Sustain. Cities Soc. 2018, 42, 482–497. [Google Scholar] [CrossRef]
  31. Jones, H.G. Plants and Microclimate: A Quantitative Approach to Environmental Plant Physiology; Cambridge University Press: Cambridge, UK, 2013; 407p. [Google Scholar]
  32. Scharenbroch, B.C.; Lloyd, J.E.; Johnson-Maynard, J.L. Distinguishing urban soils with physical, chemical, and biological properties. Pedobiologia 2005, 49, 283–296. [Google Scholar] [CrossRef]
  33. Shahidan, M.F.; Salleh, E.; Mustafa, K.M.S. (Eds.) Effects of tree canopies on solar radiation filtration in a tropical microclimatic environment. In Proceedings of the 24th International Conference on Passive and Low Energy Architecture, PLEA 2007, Singapore, 22–24 November 2007. [Google Scholar]
  34. Potchter, O.; Cohen, P.; Bitan, A. Climatic behavior of various urban parks during hot and humid summer in the mediterranean city of Tel Aviv, Israel. Int. J. Clim. 2006, 26, 1695–1711. [Google Scholar] [CrossRef]
  35. Johnson, M.S.; Lathuillière, M.J.; Tooke, T.R.; Coops, N.C. Attenuation of urban agricultural production potential and crop water footprint due to shading from buildings and trees. Environ. Res. Lett. 2015, 10, 064007. [Google Scholar] [CrossRef]
  36. Qiu, G.Y.; Yu, X.; Wen, H.; Yan, C. An advanced approach for measuring the transpiration rate of individual urban trees by the 3D three-temperature model and thermal infrared remote sensing. J. Hydrol. 2020, 587, 125034. [Google Scholar] [CrossRef]
  37. Rahman, M.A.; Moser, A.; Rötzer, T.; Pauleit, S. Microclimatic differences and their influence on transpirational cooling of Tilia cordata in two contrasting street canyons in Munich, Germany. Agric. For. Meteorol. 2017, 232, 443–456. [Google Scholar] [CrossRef]
  38. Wang, H.; Wang, X.; Zhao, P.; Zheng, H.; Ren, Y.; Gao, F.; Ouyang, Z. Transpiration rates of urban trees, Aesculus chinensis. J. Environ. Sci. 2012, 24, 1278–1287. [Google Scholar] [CrossRef]
  39. Liu, X.; Li, X.-X.; Harshan, S.; Roth, M.; Velasco, E. Evaluation of an urban canopy model in a tropical city: The role of tree evapotranspiration. Environ. Res. Lett. 2017, 12, 094008. [Google Scholar] [CrossRef]
  40. Kong, F.; Yan, W.; Zheng, G.; Yin, H.; Cavan, G.; Zhan, W.; Zhang, N.; Cheng, L. Retrieval of three-dimensional tree canopy and shade using terrestrial laser scanning (TLS) data to analyze the cooling effect of vegetation. Agric. For. Meteorol. 2016, 217, 22–34. [Google Scholar] [CrossRef]
  41. Hang, J.; Li, Y.; Buccolieri, R.; Sandberg, M.; DI Sabatino, S. On the contribution of mean flow and turbulence to city breathability: The case of long streets with tall buildings. Sci. Total Environ. 2012, 416, 362–373. [Google Scholar] [CrossRef]
  42. He, B.-J.; Ding, L.; Prasad, D. Enhancing urban ventilation performance through the development of precinct ventilation zones: A case study based on the Greater Sydney, Australia. Sustain. Cities Soc. 2019, 47, 101472. [Google Scholar] [CrossRef]
  43. Niu, J.; Liu, J.; Lee, T.-C.; Lin, Z.; Mak, C.; Tse, K.-T.; Tang, B.-S.; Kwok, K.C. A new method to assess spatial variations of outdoor thermal comfort: Onsite monitoring results and implications for precinct planning. Build. Environ. 2015, 91, 263–270. [Google Scholar] [CrossRef]
  44. He, B.-J. Potentials of meteorological characteristics and synoptic conditions to mitigate urban heat island effects. Urban Clim. 2018, 24, 26–33. [Google Scholar] [CrossRef]
  45. Zheng, S.; Zhao, L.; Li, Q. Numerical simulation of the impact of different vegetation species on the outdoor thermal environment. Urban For. Urban Green. 2016, 18, 138–150. [Google Scholar] [CrossRef]
  46. Dimoudi, A.; Nikolopoulou, M. Vegetation in the urban environment: Microclimatic analysis and benefits. Energy Build. 2003, 35, 69–76. [Google Scholar] [CrossRef]
  47. Ng, E. Policies and technical guidelines for urban planning of high-density cities—Air ventilation assessment (AVA) of Hong Kong. Build. Environ. 2009, 44, 1478–1488. [Google Scholar] [CrossRef]
  48. Allen, S.; Grime, V. Measurements of transpiration from savannah shrubs using sap flow gauges. Agric. For. Meteorol. 1995, 75, 23–41. [Google Scholar] [CrossRef]
  49. Scott, R.L.; Huxman, T.E.; Cable, W.L.; Emmerich, W.E. Partitioning of evapotranspiration and its relation to carbon dioxide exchange in a Chihuahuan Desert shrubland. Hydrol. Process. 2006, 20, 3227–3243. [Google Scholar] [CrossRef]
  50. Zhou, M.C.; Ishidaira, H.; Hapuarachchi, H.P.; Magome, J.; Kiem, A.S.; Takeuchi, K. Estimating potential evapotranspiration using Shuttleworth–Wallace model and NOAA-AVHRR NDVI data to feed a distributed hydrological model over the Mekong River basin. J. Hydrol. 2006, 327, 151–173. [Google Scholar] [CrossRef]
  51. Cheng, Y.F.; Wang, G.; Xi, H.Y. Variations of land evapotranspiration in the plain of the middle reaches of Heihe River in the recent 35 years. J. Glaciol. Geocryol. 2007, 29, 406–412. [Google Scholar]
  52. Hagishima, A.; Tanimoto, J.; Katayama, T. Numerical analysis of air temperature increases in urban area using the building-urban-soil simultaneous simulation model. Proc. Build. 2023, 16, 32–39. [Google Scholar]
  53. Jia, H.; Hu, J.; Zhang, J. Estimation of evapotranspiration during the maize growing season using the shuttleworth-wallace model. J. Irrig. Drain. 2008, 27, 77–80. [Google Scholar]
  54. Myneni, R.B.; Hoffman, S.; Knyazikhin, Y.; Privette, J.L.; Glassy, J.; Tian, Y.; Wang, Y.; Song, X.; Zhang, Y.; Smith, G.R.; et al. Global products of vegetation leaf area and fraction absorbed PAR from year one of MODIS data. Remote Sens. Environ. 2002, 83, 214–231. [Google Scholar] [CrossRef]
  55. Lin, B.-R. Studies of Greening’s Effects on Outdoor Thermal Environment; Tsinghua University: Beijing, China, 2004. [Google Scholar]
  56. Pedruzo-Bagazgoitia, X.; Ouwersloot, H.G.; Sikma, M.; Van Heerwaarden, C.C.; Jacobs, C.M.J.; De Arellano, J.V.G. Direct and Diffuse Radiation in the Shallow Cumulus–Vegetation System: Enhanced and Decreased Evapotranspiration Regimes. J. Hydrometeorol. 2017, 18, 1731–1748. [Google Scholar] [CrossRef]
  57. Zheng, S.; Guldmann, J.-M.; Liu, Z.; Zhao, L.; Wang, J.; Pan, X.; Zhao, D. Predicting the influence of subtropical trees on urban wind through wind tunnel tests and numerical simulations. Sustain. Cities Soc. 2020, 57, 102116. [Google Scholar] [CrossRef]
  58. Liu, Z.; Zheng, S.; Zhao, L. Evaluation of the ENVI-Met Vegetation Model of Four Common Tree Species in a Subtropical Hot-Humid Area. Atmosphere 2018, 9, 198. [Google Scholar] [CrossRef]
  59. Manickathan, L.; Defraeye, T.; Allegrini, J.; Derome, D.; Carmeliet, J. Aerodynamic characterisation of model vegetation by wind tunnel experiments. In Proceedings of the 4th International Conference on Countermeasures to Urban Heat Island, Singapore, 30 May–1 June 2016. [Google Scholar]
  60. Cao, J.; Tamura, Y.; Yoshida, A. Wind tunnel study on aerodynamic characteristics of shrubby specimens of three tree species. Urban For. Urban Green. 2012, 11, 465–476. [Google Scholar] [CrossRef]
  61. Quattrochi, D.; Luvall, J.; Rickman, D.; Estes, M., Jr.; Laymon, C.; Howell, B. A Decision Support Information System for Urban Landscape Management Using Thermal Infrared Data. Photogramm. Eng. Remote Sens. 2000, 66, 1195–1207. [Google Scholar]
  62. Sanusi, R.; Johnstone, D.; May, P.; Livesley, S.J. Microclimate benefits that different street tree species provide to sidewalk pedestrians relate to differences in Plant Area Index. Landsc. Urban Plan. 2017, 157, 502–511. [Google Scholar] [CrossRef]
  63. Gromke, C.; Ruck, B. Aerodynamic modelling of trees for small-scale wind tunnel studies. For. Int. J. For. Res. 2008, 81, 243–258. [Google Scholar] [CrossRef]
  64. Parker, J.H. Landscaping to reduce the energy used in cooling buildings. J. For. 1983, 81, 82–84. [Google Scholar]
  65. Shinzato, P.; Duarte, D. (Eds.) Microclimatic effect of vegetation for different leaf area index—LAI. In Proceedings of the 28th International PLEA Conference on Sustainable Architecture + Urban Design: Opportunities, Limits and Needs—Towards an Environmentally Responsible Architecture, PLEA 2012, Lima, Peru, 7–9 November 2012. [Google Scholar]
  66. Duarte, M.C.; Pimenta, D.C.; Menezes-Souza, D.; Magalhães, R.D.M.; Diniz, J.L.C.P.; Costa, L.E.; Chávez-Fumagalli, M.A.; Lage, P.S.; Bartholomeu, D.C.; Alves, M.J.M.; et al. Proteins Selected in Leishmania (Viannia) braziliensis by an Immunoproteomic Approach with Potential Serodiagnosis Applications for Tegumentary Leishmaniasis. Clin. Vaccine Immunol. 2015, 22, 1187–1196. [Google Scholar] [CrossRef]
  67. Zhang, L.; Zhan, Q.; Lan, Y. Effects of the tree distribution and species on outdoor environment conditions in a hot summer and cold winter zone: A case study in Wuhan residential quarters. Build. Environ. 2018, 130, 27–39. [Google Scholar] [CrossRef]
  68. Santamouri, M.; Asimakopoulos, D.N.; Assimakopoulos, V.D.; Chrisomallidou, N.; Klitsikas, N.; Mangold, D.; Michel, P.; Tsangrassoulis, A. Energy and Climate in the Urban Built Environment; Routledge: London, UK, 2013; 402p. [Google Scholar]
  69. Chen, Z.; Su, X.; Liu, S. Study on the Ecological Benefits of Urban Greening in Beijing; China Environmental Science Press: Beijing, China, 1998; Volume 10, pp. 117–121. [Google Scholar]
  70. Han, H. Study on the Ecological Function of Urban Greening Tree Species; Northeast Forestry University: Harbin, China, 2007. [Google Scholar]
  71. Heisler, G.M. Effects of individual trees on the solar radiation climate of small buildings. Urban Ecol. 1986, 9, 337–359. [Google Scholar] [CrossRef]
  72. Helge, S. Modeling Urban Microclimate: Development, Implementation and Evaluation of New and Improved Calculation Methods for the Urban Microclimate Model ENVI-Met; Johannes Gutenberg-University Mainz: Mainz, Germany, 2016. [Google Scholar]
  73. Shinzato, P.; Yoshida, D.; Duarte, D. Parametrization of Tropical Plants Using ENVI-Met V.4 and Its Impact on Urban Microclimates—Sao Paulo Case Study. In Proceedings of the 4th International Conference on Countermeasures to Urban Heat Islands, Singapore, 30 May–1 June 2016. [Google Scholar]
  74. Lam, C.K.C.; Lee, H.; Yang, S.-R.; Park, S. A review on the significance and perspective of the numerical simulations of outdoor thermal environment. Sustain. Cities Soc. 2021, 71, 102971. [Google Scholar] [CrossRef]
  75. Stavrakakis, G.M.; Katsaprakakis, D.A.; Damasiotis, M. Basic Principles, Most Common Computational Tools, and Capabilities for Building Energy and Urban Microclimate Simulations. Energies 2021, 14, 6707. [Google Scholar] [CrossRef]
  76. Chen, Z. Study on the Impact of Greening System on the Outdoor Thermal Environment of Building Clusters in Humid and Hot Areas; South China University of Technology: Guangzhou, China, 2010. [Google Scholar]
  77. Lindén, J.; Simon, H.; Fonti, P.; Esper, J.; Bruse, M. Observed and Modeled Transpiration Cooling from Urban Trees in Mainz, Germany. In Proceedings of the 9th International Conference on Urban Climate, Toulouse, France, 24 July 2015. [Google Scholar]
  78. Krayenhoff, E.S.; Christen, A.; Martilli, A.; Oke, T.R. A Multi-layer Radiation Model for Urban Neighbourhoods with Trees. Bound.-Layer Meteorol. 2014, 151, 139–178. [Google Scholar] [CrossRef]
  79. Upreti, R.; Wang, Z.-H.; Yang, J. Radiative shading effect of urban trees on cooling the regional built environment. Urban For. Urban Green. 2017, 26, 18–24. [Google Scholar] [CrossRef]
  80. Gao, Y.; Long, D. Progress in model for evapotranspiration estimation using remotely sensed data. Hydrol. Process. 2008, 20, 3227–3247. [Google Scholar]
  81. Liu, Z. Study on Microclimate Effects of Trees in Humid and Hot Regions Based on Evidence-Based Design; South China University of Technology: Guangzhou, China, 2020. [Google Scholar]
  82. Kántor, N.; Kovács, A.; Takács, A. Small-scale human-biometeorological impacts of shading by a large tree. Open Geosci. 2016, 8, 231–245. [Google Scholar] [CrossRef]
  83. El-Bardisy, W.M.; Fahmy, M.; El-Gohary, G.F. Climatic Sensitive Landscape Design: Towards a Better Microclimate through Plantation in Public Schools, Cairo, Egypt. Procedia-Soc. Behav. Sci. 2016, 216, 206–216. [Google Scholar] [CrossRef]
  84. Cohen, P.; Potchter, O.; Matzarakis, A. Daily and seasonal climatic conditions of green urban open spaces in the Mediterranean climate and their impact on human comfort. Build. Environ. 2012, 51, 285–295. [Google Scholar] [CrossRef]
  85. Zhao, Q.; Sailor, D.J.; Wentz, E.A. Impact of tree locations and arrangements on outdoor microclimates and human thermal comfort in an urban residential environment. Urban For. Urban Green. 2018, 32, 81–91. [Google Scholar] [CrossRef]
  86. Duarte, D.H.; Shinzato, P.; Gusson, C.d.S.; Alves, C.A. The impact of vegetation on urban microclimate to counterbalance built density in a subtropical changing climate. Urban Clim. 2015, 14, 224–239. [Google Scholar] [CrossRef]
  87. Lee, H.; Mayer, H. Maximum extent of human heat stress reduction on building areas due to urban greening. Urban For. Urban Green. 2018, 32, 154–167. [Google Scholar] [CrossRef]
  88. Wong, N.H.; Jusuf, S.K. Study on the microclimate condition along a green pedestrian canyon in Singapore. Arch. Sci. Rev. 2010, 53, 196–212. [Google Scholar] [CrossRef]
  89. Sun, S.; Xu, X.; Lao, Z.; Liu, W.; Li, Z.; García, E.H.; He, L.; Zhu, J. Evaluating the impact of urban green space and landscape design parameters on thermal comfort in hot summer by numerical simulation. Build. Environ. 2017, 123, 277–288. [Google Scholar] [CrossRef]
  90. Wang, Y.; Ni, Z.; Peng, Y.; Xia, B. Local variation of outdoor thermal comfort in different urban green spaces in Guangzhou, a subtropical city in South China. Urban For. Urban Green. 2018, 32, 99–112. [Google Scholar] [CrossRef]
  91. Chen, L.; Ng, E. Simulation of the effect of downtown greenery on thermal comfort in subtropical climate using PET index: A case study in Hong Kong. Arch. Sci. Rev. 2013, 56, 297–305. [Google Scholar] [CrossRef]
  92. Kong, L.; Lau, K.K.-L.; Yuan, C.; Chen, Y.; Xu, Y.; Ren, C.; Ng, E. Regulation of outdoor thermal comfort by trees in Hong Kong. Sustain. Cities Soc. 2017, 31, 12–25. [Google Scholar] [CrossRef]
  93. Morakinyo, T.E.; Lam, Y.F. Simulation study on the impact of tree-configuration, planting pattern and wind condition on street-canyon’s micro-climate and thermal comfort. Build. Environ. 2016, 103, 262–275. [Google Scholar] [CrossRef]
  94. Morakinyo, T.E.; Lau, K.K.-L.; Ren, C.; Ng, E. Performance of Hong Kong’s common trees species for outdoor temperature regulation, thermal comfort and energy saving. Build. Environ. 2018, 137, 157–170. [Google Scholar] [CrossRef]
  95. Tan, Z.; Lau, K.K.-L.; Ng, E. Urban tree design approaches for mitigating daytime urban heat island effects in a high-density urban environment. Energy Build. 2016, 114, 265–274. [Google Scholar] [CrossRef]
  96. Tan, Z.; Lau, K.K.-L.; Ng, E. Planning strategies for roadside tree planting and outdoor comfort enhancement in subtropical high-density urban areas. Build. Environ. 2017, 120, 93–109. [Google Scholar] [CrossRef]
  97. Correa, E.; Ruiz, M.A.; Canton, A.; Lesino, G. Thermal comfort in forested urban canyons of low building density. An assessment for the city of Mendoza, Argentina. Build. Environ. 2012, 58, 219–230. [Google Scholar] [CrossRef]
  98. Stocco, S.; Alicia Canton, M.; Norma Correa, E. Design of urban green square in dry areas: Thermal performance and comfort. Urban For. Urban Green. 2015, 14, 323–335. [Google Scholar] [CrossRef]
  99. Picot, X. Thermal comfort in urban spaces: Impact of vegetation growth: Case study: Piazza della Scienza, Milan, Italy. Energy Build. 2004, 36, 329–334. [Google Scholar] [CrossRef]
  100. Gómez, F.; Montero, L.; De Vicente, V.; Sequí, A.; Castilla, N. Vegetation influences on the human thermal comfort in outdoor spaces: Criteria for urban planning. WIT Trans. Ecol. Environ. 2008, 117, 151–163. [Google Scholar]
  101. Vanos, J.K.; Warland, J.S.; Gillespie, T.J.; Slater, G.A.; Brown, R.D.; Kenny, N.A. Human Energy Budget Modeling in Urban Parks in Toronto and Applications to Emergency Heat Stress Preparedness. J. Appl. Meteorol. Clim. 2012, 51, 1639–1653. [Google Scholar] [CrossRef]
  102. Shashua-Bar, L.; Pearlmutter, D.; Erell, E. The influence of trees and grass on outdoor thermal comfort in a hot-arid environment. Int. J. Clim. 2011, 31, 1498–1506. [Google Scholar] [CrossRef]
  103. Snir, K.; Pearlmutter, D.; Erell, E. The moderating effect of water-efficient ground cover vegetation on pedestrian thermal stress. Landsc. Urban Plan. 2016, 152, 1–12. [Google Scholar] [CrossRef]
  104. Massetti, L.; Petralli, M.; Napoli, M.; Brandani, G.; Orlandini, S.; Pearlmutter, D. Effects of deciduous shade trees on surface temperature and pedestrian thermal stress during summer and autumn. Int. J. Biometeorol. 2019, 63, 467–479. [Google Scholar] [CrossRef]
  105. Coccolo, S.; Pearlmutter, D.; Kaempf, J.; Scartezzini, J.-L. Thermal Comfort Maps to estimate the impact of urban greening on the outdoor human comfort. Urban For. Urban Green. 2018, 35, 91–105. [Google Scholar] [CrossRef]
  106. Coutts, A.M.; White, E.C.; Tapper, N.J.; Beringer, J.; Livesley, S.J. Temperature and human thermal comfort effects of street trees across three contrasting street canyon environments. Theor. Appl. Climatol. 2016, 124, 55–68. [Google Scholar] [CrossRef]
  107. Hu, H. The Research Based on the Sealing Arch Border Tree’s Shade Which Impacted on the Urban Street Canyon’s Micro-Climate in Xi’an City; Xi’an University of Architecture and Technology: Xi’an, China, 2017. [Google Scholar]
  108. Wu, Z.; Kong, F.; Wang, Y.; Sun, R.; Chen, L. The Impact of Greenspace on Thermal Comfort in a Residential Quarter of Beijing, China. Int. J. Environ. Res. Public Health 2016, 13, 1217, Erratum in Int. J. Environ. Res. Public Health 2017, 14, 156. [Google Scholar] [CrossRef]
  109. Lin, B.; Li, X.; Zhu, Y.; Qin, Y. Numerical simulation studies of the different vegetation patterns’ effects on outdoor pedestrian thermal comfort. J. Wind Eng. Ind. Aerodyn. 2008, 96, 1707–1718. [Google Scholar] [CrossRef]
  110. Hsieh, C.-M.; Jan, F.-C.; Zhang, L. A simplified assessment of how tree allocation, wind environment, and shading affect human comfort. Urban For. Urban Green. 2016, 18, 126–137. [Google Scholar] [CrossRef]
  111. Xia, S. Analysis of the Plant Configuration Mode’s Impact on the Microclimate in the Enclosed Residential Area during Summer. Master’s Thesis, Shenyang Agricultural University, Shenyang, China, 2017. [Google Scholar]
  112. Morakinyo, T.E.; Adegun, O.B.; Balogun, A.A. The effect of vegetation on indoor and outdoor thermal comfort conditions: Evidence from a microscale study of two similar urban buildings in Akure, Nigeria. Indoor Built Environ. 2016, 25, 603–617. [Google Scholar] [CrossRef]
  113. Xu, N. Microclimate Effect and Its Influence on Touristbehavior in Heilongjiang Forest Botanical Garden; Northeast Forestry University: Harbin, China, 2018. [Google Scholar]
  114. Gao, J.; Wang, C.; Wu, Z. Microclimate Conditions of Different Urban Land Types and Their lmpacts on Pleasing Feeling of Human Body. Urban For. China 2004, 32, 41–49+51. [Google Scholar]
  115. Wang, L. Study on Microclimate Environment Effects of Campus Green Space—As the Example of Northwest A&F University; Northwest A&F University: Xianyang, China, 2016. [Google Scholar]
  116. Wang, J. Comparative study on the effects of summer plant community on human comfort in Beijing. Shanxi Archit. 2018, 44, 198–199. [Google Scholar]
  117. Wang, X. Rural Residental Forests Evaluation of Microclimate and Air Anion Effects in Mountainous; Shandong Agricultural University: Jinan, China, 2011. [Google Scholar]
  118. Marando, F.; Heris, M.P.; Zulian, G.; Udías, A.; Mentaschi, L.; Chrysoulakis, N.; Parastatidis, D.; Maes, J. Urban heat island mitigation by green infrastructure in European Functional Urban Areas. Sustain. Cities Soc. 2022, 77, 103564. [Google Scholar] [CrossRef]
  119. Konijnendijk, C.C. Evidence-based guidelines for greener, healthier, more resilient neighbourhoods: Introducing the 3–30–300 rule. J. For. Res. 2022, 34, 821–830. [Google Scholar] [CrossRef]
  120. Liu, X.; He, J.; Xiong, K.; Liu, S.; He, B.-J. Identification of factors affecting public willingness to pay for heat mitigation and adaptation: Evidence from Guangzhou, China. Urban Clim. 2023, 48, 101405. [Google Scholar] [CrossRef]
  121. Liu, S.; Wang, Y.; Liu, X.; Yang, L.; Zhang, Y.; He, J. How does future climatic uncertainty affect multi-objective building energy retrofit decisions? Evidence from residential buildings in subtropical Hong Kong. Sustain. Cities Soc. 2023, 92, 104482. [Google Scholar] [CrossRef]
  122. Li, K.; Liu, X.; Zhang, H.; Ma, J.; He, B.-J. Evaluating and improving the adaptability of commonly used indices for predicting outdoor thermal sensation in hot and humid residential areas of China. Dev. Built Environ. 2023, 100278, in press. [Google Scholar] [CrossRef]
Figure 1. Flowchart of Paper Retrieval Research.
Figure 1. Flowchart of Paper Retrieval Research.
Forests 14 02280 g001
Figure 2. Schematic diagram of the S-W model [50].
Figure 2. Schematic diagram of the S-W model [50].
Forests 14 02280 g002
Figure 3. Calculation of the solar radiation transmittance of trees.
Figure 3. Calculation of the solar radiation transmittance of trees.
Forests 14 02280 g003
Figure 4. Optical properties of leaves (transmission, absorption, and reflection of visible and infrared radiation, respectively) [17,18].
Figure 4. Optical properties of leaves (transmission, absorption, and reflection of visible and infrared radiation, respectively) [17,18].
Forests 14 02280 g004
Figure 5. Physical models of trees [57].
Figure 5. Physical models of trees [57].
Forests 14 02280 g005
Figure 6. Layout of measuring points and instruments to obtain trees’ cooling effects. (a) is a schematic diagram; (b) is field measurement chart. A: weather station to obtain meteorological parameters in open areas; B and C: sensors to obtain meteorological parameters in shaded areas [30].
Figure 6. Layout of measuring points and instruments to obtain trees’ cooling effects. (a) is a schematic diagram; (b) is field measurement chart. A: weather station to obtain meteorological parameters in open areas; B and C: sensors to obtain meteorological parameters in shaded areas [30].
Forests 14 02280 g006
Table 1. Nomenclature of Equations (1)–(8) of the S-W model.
Table 1. Nomenclature of Equations (1)–(8) of the S-W model.
SymbolFull NameUnit
ETTotal evapotranspirationJ/m2·s
λ Latent heat of water vaporizationJ/kg
ET C Canopy transpirationJ/m2·s
ET C Bare surface evaporationJ/m2·s
C C and C S Weight coefficients of ET C and ET s -
R n Canopy net radiationsJ/m2·s
R n s Soil surface net radiationsJ/m2·s
GSoil heat fluxJ/m2·s
γAir humidity constantkPa/°C
e s Saturated water vapor pressurekPa
e a Actual water vapor pressurekPa
r s c Canopy stomatal resistances/m
r a c Canopy boundary layer resistances/m
r s s Soil surface resistances/m
Δ Water vapor pressure-temperature curveJ/kg
Table 2. Thermal effect of solar radiation on the canopy at different wavelengths [55].
Table 2. Thermal effect of solar radiation on the canopy at different wavelengths [55].
Type of RadiationWavelength (μm)Proportion (%)
UV0.29~0.380~4
PAC0.38~0.7121~46
Near-infrared0.71~4.050~79
Far-infrared>4.0-
Table 3. Variations in the drag coefficients (Cd) of four tree species with wind speed (U) [57].
Table 3. Variations in the drag coefficients (Cd) of four tree species with wind speed (U) [57].
SpeciesLeaf Area Density (m2/m2)Drag Coefficient (Cd)
Ficus microcarpa4.97Cd = 1.07 × U−0.075
Mangifera indica4.79Cd = 1.0 × U−0.05
Michelia alba2.88Cd = 1.0 × U−0.19
Bauhinia blakeana4.27Cd = 0.89 × U−0.069
Table 4. Commonly used microclimatic measuring parameters and instruments [58].
Table 4. Commonly used microclimatic measuring parameters and instruments [58].
Test ParameterTest EquipmentFactory OwnersAccurateTest Range
Air temperature
Relative humidity
HOBO pro v2 data logger
(U23-001)
Onset Computer Corporation, Bourne, MA, USA±0.2 °C
(0~50 h)
−40~70 °C
Wind Speed, Wind Direction
Black sphere temperature
Ultrasonic anemometer
sensor (Model 81000)
M. Young Company,
Traverse, MI, USA
±1% ± 0.05 m/s0~40 m/s
Meteorological parametersDavis Vantage Pro2Davis Company,
Boston, MA, USA
±0.5 °C (Ta)
±5% (v)
−40~65 °C (Ta)
0–1800 W/m2 (S)
Transpiration rate
Leaf surface temperature
Photosynthesis apparatus
Li-6400
Decagon Company,
Pullman, WA, USA
±0.007 mmol/mol0~75 mol
Solar radiation
Long-wave radiation
4-component net radiation sensor NR01Hukseflux Company,
Delft, The Netherlands
7–25 μV/W/m20~2000 W/m2
Soil temperatureT type thermocoupleSensors Company,
Wuxi, China
±0.05 °C−200~260 °C
Thermal imagingThermal infrared imagerKaise Company,
Ueda, Japan
±2 °C−40~500 °C
Leaf reflectanceSpectrophotometer (U-4100)Hitachi Company
Tokyo, Japan
/175~2600 nm
Root depth, root width and root densityTree Radar (TRU-100)Tree Radar Company,
Silver Spring, MD, USA
1 cm/
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

Zheng, S.; He, C.; Guldmann, J.-M.; Xu, H.; Liu, X. Heat Mitigation Benefits of Urban Trees: A Review of Mechanisms, Modeling, Validation and Simulation. Forests 2023, 14, 2280. https://doi.org/10.3390/f14122280

AMA Style

Zheng S, He C, Guldmann J-M, Xu H, Liu X. Heat Mitigation Benefits of Urban Trees: A Review of Mechanisms, Modeling, Validation and Simulation. Forests. 2023; 14(12):2280. https://doi.org/10.3390/f14122280

Chicago/Turabian Style

Zheng, Senlin, Caiwei He, Jean-Michel Guldmann, Haodong Xu, and Xiao Liu. 2023. "Heat Mitigation Benefits of Urban Trees: A Review of Mechanisms, Modeling, Validation and Simulation" Forests 14, no. 12: 2280. https://doi.org/10.3390/f14122280

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