Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Energy & Buildings 284 (2023) 112824

Contents lists available at ScienceDirect

Energy & Buildings


journal homepage: www.elsevier.com/locate/enb

Building form and outdoor thermal comfort: Inverse design the


microclimate of outdoor space for a kindergarten
Rui Sun a, Junjie Liu b, Dayi Lai a,⇑, Wei Liu c,⇑
a
Department of Architecture, School of Design, Shanghai Jiao Tong University, Shanghai 200240, China
b
Tianjin Key Laboratory of Indoor Air Environmental Quality Control, School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, China
c
Division of Sustainable Buildings, Department of Civil and Architectural Engineering, KTH Royal Institute of Technology, Brinellvägen 23, Stockholm 100 44, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: Thermally comfortable mircoclimate is essential for creating high-quality outdoor spaces that attract cit-
Received 13 October 2022 izens and boost city vitality. Previous design efforts to improve outdoor thermal comfort were usually
Revised 29 December 2022 conducted at large scales, such as city scale, neighborhood scale, urban block scale. Few researchers
Accepted 23 January 2023
focused on the building scale. This study proposes an optimization framework based on genetic algorithm
Available online 2 February 2023
to determine the building shape, orientation, and location during early design stage that reduces the
overall thermal stress in the target outdoor space. Solar radiation and wind fields were simulated to
Keywords:
obtain the outdoor Universal Thermal Climate Index (UTCI) as the performance indicator. The simulations
Genetic algorithm
Radiation
were validated against the experimental data. This investigation applied the proposed optimization
Wind framework to design the outdoor space for a kindergarten under the climate of Tianjin and Shanghai,
CFD respectively. The results showed that optimization reduced the overall thermal stress. The most favour-
UTCI able kindergarten forms were suggested through optimization. Moreover, solar radiation has been proved
Thermal stress to contribute more to outdoor thermal comfort than wind field and heat stress is more important than
cold stress during optimization. This study supplements the inverse design of outdoor thermal comfort
at building scale and provides suggestions to create comfortable urban outdoor spaces.
Ó 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).

1. Introduction influence human thermal comfort in outdoor spaces [41]. While air
temperature and humidity are greatly determined by the diurnal
Cities, as important centers for human social and economic and seasonal changes of local weather and usually have limited
activities [62], are expected to accommodate almost 70% of the glo- spatial variation, the distributions of solar radiation and wind
bal population by 2050. Under rapid urbanization, it is crucial to speed can be largely affected by the urban geometry. For example,
vitalize urban outdoor spaces because these places provide essen- buildings alter shortwave solar radiation and change radiative heat
tial places for the everyday lives of citizens. Encouraging citizens to gain on human body [46]. In addition to solar radiation, wind can
participate in outdoor activities also reduces the usages of air con- be accelerated at the windward region and decelerated at the wake
ditioners and other electronic equipment [40], increases revenue region of a building [10]. The effects of urban geometry on micro-
from outdoor commercial and recreational activities [28], and pro- climate can be conducted at different scales, as Table 1 illustrates
motes their physical and mental health [53]. Thus, it is essential to [46,3].
design high-quality outdoor spaces to attract people to the out- Among all the scales listed in Table 1, outdoor thermal comfort
doors. A thermally comfortable microclimate is fundamental to a studies were always conducted at large scales like city scale, neigh-
high-quality outdoor space [34], as increased vitality was usually bourhood scale, urban block scale and street canyon scale. Few
observed under better microclimate [22]. Thus, it is necessary to studies paid attention to the building scale. In fact, the current
take microclimate into consideration when designing urban out- building scale researches were mainly focused on indoor thermal
door spaces. comfort [63,64]but not outdoor microclimate, and the few remain-
Physically speaking, air temperature, solar radiation, wind ing outdoor thermal comfort researches yield obvious limitations.
speed, and humidity are four basic parameters of microclimate that Anisha et al. [36] investigated effects of building forms on microcli-
mate in Dhaka, Bangladesh, but considered just four simple and
⇑ Corresponding authors. fundamental forms. Beta et al. [47] compared the microclimate of
E-mail addresses: dayi_lai@sjtu.edu.cn (D. Lai), weiliu2@kth.se (W. Liu). five low-cost apartments with different building form and config-

https://doi.org/10.1016/j.enbuild.2023.112824
0378-7788/Ó 2023 The Author(s). Published by Elsevier B.V.
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Table 1
Analysis scales in outdoor thermal comfort studies.

Scales Research objects Research topics


City scale Whole city The difference of climatic phenomena between cities and their rural environment
[38,51,39]
Neighbour- hood scale Urban landscape with a degree of homogeneity (e.g. The impact of urban geometry features,
heavy industrial, high-density residential areas, etc.) such as building density, configuration,
aspect ratio, sky view factor and so on,
on outdoor thermal comfort
[55,29]
[15,1]
[4]
Urban block scale Urban unit delineated by the principal road network
(Several urban blocks form neighbourhoods.)
Street canyon scale Geometric form created by a street with its three facets The effect of street canyon features like height-to-width ratio, canyon orientation,
etc., on local microclimate [23,2,14]
Building scale One or several buildings located on a target site The impact of building form, including shape, height, orientation and location of
the target building, etc., on the microclimate of its outdoor space [36,47]

urations in Bandung, Indonesia, but only three building forms were 2.1. Framework for outdoor microclimate optimization
studied. To fully explore the effect of building forms on outdoor
thermal comfort, a series of design scenarios for various building Fig. 1 illustrates the overall framework of the outdoor microcli-
forms should be considered. mate optimization, which consists of four parts: background input,
In order to consider more design scenarios, computational opti- microclimate simulation, performance evaluation, and genetic
mization that features automatically generating a series of design algorithm. The background input drives the microclimate simula-
solutions is a useful and effective tool. In the last few years, several tion, the performance is evaluated according to the simulation
attempts have been made to apply computational optimization to results, and the optimization process efficiently searches for a set
the design of outdoor microclimate for realizing optimum thermal of design variables that achieve the best performance. The simula-
comfort, these attempts are very effective but still have a lot of lim- tion and optimization were conducted in the Rhino – Grasshopper
itations. Taleb et al. [54] optimized the form and orientation of a parametric modeling platform with the use of plug-ins such as
building cluster to achieve the optimum solar radiation and urban Ladybug, Butterfly, and Galapagos [50]. Details of the simulation
ventilation, respectively. Kaseb et al. [37] proposed a CFD-based and optimization processes are covered in the following sections.
evolutionary framework to optimize building heights and plan area
densities to improve ventilation in complex real urban areas. These 2.1.1. Background input
studies only considered the effect of solar radiation or wind. To The background information can be categorized into morpho-
account for the comprehensive effect of the four environmental logical and meteorological inputs. The morphological inputs
parameters on outdoor thermal comfort, Bajsanski et al. [7] and include the geometric features of the target building and surround-
Ibrahim et al. [33] used an integrated index Universal Thermal Cli- ings. The morphological features of the target building, including
mate Index (UTCI) in the objective function when optimizing lay- building location, shape, height, orientation, etc., can be refined
outs of urban blocks. But they just simulated radiation field and into design variables. While setting up the design variables, con-
the assumption that the wind speed of the urban area was uniform straints from building codes or actual demands are required to pro-
may bring unknown inaccuracy. Xu et al. [61,60] simulated both vide reasonable restrictions. To consider the influence of
solar radiation and wind field but failed to precisely consider the surrounding buildings on local microclimate, appropriate propor-
impact of surrounding buildings. In addition, the above investiga- tion needs to be suggested to get reliable results without compro-
tions only focused on either summer or winter and did not con- mising the computational accuracy. According to Liu et al. [43],
sider overall seasonal variations. Considering the seasonal detailed building structures around the target space within a
changes throughout a year, a design that improves thermal com- radius of at least 3L (L is the maximum dimension of the target
fort in summer may cause discomfort in winter and vice versa. space) are modeled. Meteorological data are also necessary inputs
Also, none of these optimization studies was conducted at building because they provide boundary conditions to drive the microcli-
scale. Therefore, the research aim and objectives are as follows: mate simulation.

 Propose a framework that inversely design the form of buildings


with the objective to optimize outdoor microclimate at building 2.1.2. Microclimate simulation
scale. This study simulates the wind speed and radiation fields. Then
 Consider the effect of both surrounding buildings and seasonal the wind speed and radiation fields are combined with air temper-
changes in this framework. ature and humidity from weather data file for the evaluation of
 Validate the numerical simulations of solar radiation and wind outdoor thermal comfort. The followings show the details of solar
field in the framework against experimental data. radiation and wind speed calculations.
 Apply the proposed inverse design framework to design the
outdoor space of a kindergarten for demonstration. 2.1.2.1. Calculation of solar radiation field. This study uses mean
radiant temperature (T mrt ) to denote the solar radiation field. T mrt
2. Methods is not only a comprehensive variable integrating the long wave
and short wave radiant components, but also a significant param-
This section first introduces the framework for outdoor micro- eters governing the energy balance and the thermal comfort of the
climate optimization. Then, the simulations of solar radiation and human body [16]. Ladybug plug-in in Grasshopper calculates T mrt
wind speed were validated by comparing with experimental data. as addition of both long wave and short wave heat exchange. The
2
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 1. Framework to optimize the outdoor microclimate.

long wave part (T mrt;long wav e ) is estimated using the MENEX model U 2ref
[8] as follows: kðzÞ ¼ pffiffiffiffiffiffi ð4Þ
Cl
0:25
Ihor
T mrt;long wav e ¼ f svv ð Þ  273:15 ð1Þ U 3ref
raLW eðzÞ ¼ ð5Þ
where f svv represents the sky view factor from a certain point unob-
jðz þ z0 Þ
structed by opaque surfaces. Ihor (W/mm2) is the horizontal infrared where z0 is the aerodynamic roughness length and different for var-
radiation intensity obtained from the weather file. r is the Stephan ious terrains [25]. In this study, z0 was set to 1.0 to represent den-
Boltzmann constant (5.667108 W/m2k4), aLW is the long wave sely build-up area. z is the height coordinate, Uref is the reference
absorptivity. As for the short wave solar radiation, the Solarcal wind speed at z = 10 m, j = 0.42 is the von Karman constant,
model proposed by Arns et al. [5] is employed to calculate short and C l =0.09 is the turbulence model constant.
wave part of T mrt (T mrt;short wav e ) as follows: For outdoor wind speed simulation, the zero static gauge pres-
aSW sure is imposed at the outlet and symmetry conditions are applied
T mrt;short wav e ¼ ½ð0:5f eff f svv ÞðIhor Rfloor þ Idiff Þ at the top and lateral sides. The standard wall functions and the
aLW f eff hr
roughness modification [21] are used for the ground surface and
þ ðf p f eff Idir Þ þ T a ð2Þ the building walls. Then, BlockMesh and SnappyHexMesh tools in
OpenFOAM are utilized together to generate mesh. The k-e turbu-
where f eff is the fraction of the body exposed to solar radiation and
lence model [10] based on the Reynolds-Averaged Navier–Stokes
hr is the radiation heat transfer coefficient (W/m2K). aSW is short
(RANS) Equations is used for CFD calculation and convergence is
wave absorptivity. Idiff and Idir are respectively the diffused and
obtained when all the scaled residuals reach a minimum of 106 .
direct normal radiation. Rfloor is the floor reflectance and f p is the
projected area factor of the body. The detailed explanation of Eq.
2.1.3. Performance evaluation
2 can be found in Ref. [5]. According to ASHRAE standard 55 [6],
This study employs UTCI that based on Fiala multi-node human
f eff was set to 0.725 to indicate a standing person, hr was
thermoregulation model [19] to evaluate the outdoor thermal
6.012 W/m2 K, aSW was 0.7 for brown skin and medium clothing,
comfort. UTCI is claimed to be applicable to a wide range of cli-
aLW was 0.95, and Rfloor was 0.25 in this study. mates and locations [9] and the calculation of UTCI requires four
input parameters: T a ; T mrt , relative humidity(RH) and U ref . Among
2.1.2.2. Calculation of wind speed. This study uses the Grasshopper
the four parameters, T a and RH can be obtained from Energy Plus
plug-in Butterfly that relies on the open-source computational
Weather (EPW) weather file while the distributions of T mrt and U
fluid dynamics (CFD) engine OpenFOAM [58] to simulate the wind
at z meters are calculated according to Section 2.1.2. U can be con-
field. According to the guidelines for CFD simulation of urban wind
verted to U ref using Eq. 6:
conditions [20,56], the height of the domain, the upstream domain
10
size, and the lateral size were 6H (H is the maximum height of the log 0:01
U ref ¼ U ð6Þ
involved buildings), and the downwind domain size was 15H. The z
log 0:01
Atmospheric Boundary Layer (ABL) conditions of mean wind speed
U, turbulent kinetic energy k, and turbulent dissipation rate e are The resulted UTCI is categorized into 10 thermal stress levels and
prescribed at the inlet as [49]: the level of 9–26 °C is considered as no thermal stress. The stronger
the thermal stress, the further UTCI is away from 9–26 °C. The ther-
U ref z þ z0 mal stress at time point m and spatial grid n is defined as the dis-
UðzÞ ¼ ln ð3Þ
j z0 tance from the no thermal stress range 9–26 °C:
3
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Table 2
Daily average of input meteorological parameters.

Date 08/13/2018 10/24/2018 02/28/2019 05/13/2019


T a (°C) 33.75 17.44 11.41 23.00
Idir (W/m2) 548.45 641.92 328.61 383.24
Idiff (W/m2) 79.48 54.20 28.96 54.84
Ihor (W/m2) 447.12 340.71 280.3 381.50
Sky condition Cloudy Sunny Sunny Sunny

Fig. 2. Comparison between simulated and measured T mrt on the four measured dates: (a) A time-series plot, and (b) A scatter plot.

Fig. 3. (a) Building configuration and (b) Measurement points of the experiment.

4
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 4. Comparisons between simulated and measured wind velocity along vertical lines at: (a) x=H ¼ 1:4, (b) x=H ¼ 1:0, (c) x=H ¼ 0:6, (d) x=H ¼ 1:4, (e) x=H ¼ 1:0, (f)
x=H ¼ 0:6 at y=H ¼ 0.

8 m
< UTCIn  26; if UTCI > 26
> generations or a maximum of 15 generations are performed. In this
TSm ¼ 0; if 9 6 UTCI 6 26 ð7Þ study, the size of the initial population was 30 individuals and the
n
>
: size of other generations was 10 individuals. In every evolutionary
9  UTCIm
n; if UTCI < 9
period, 20% of all individuals were selected. Then, Blend Crossover
The objective of outdoor microclimate optimization is to reduce the [17] and Point Mutation[26] were employed to produce new
overall thermal stress in the target outdoor space at specified times. individuals.
With the thermal stress in Eq. 7, the objective function (OF) is
defined as the mean of the spatial and temporal thermal stress in 2.2. Validation of the solar radiation and wind speed simulations
Eq. 8:
Although numerical modelling offers flexibility, it entails uncer-
X
m X
n
TSji tainties due to many approximations used. Therefore, it is neces-
j¼1 i¼1 sary to validate the numerical simulations. The following details
OF ¼ ð8Þ
mn the validation of solar radiation and wind simulations by compar-
ing the simulated results with the experimental data.
2.1.4. Genetic algorithm
2.2.1. Solar radiation
This investigation uses genetic algorithm based on Galapagos
In order to provide experimental data for the validation of solar
plug-in in Grasshopper to identify the design variables of mini-
radiation simulation, this study measured T mrt in the center of a
mum OF. Genetic algorithm is effective for finding the global opti-
mal condition and can efficiently reduce the total number of
iterations needed to reach one or more optimal solutions [44]. As Table 3
shown in Fig. 1, the first step of genetic algorithm generates a set The layout and surroundings of investigated kindergartens
of random initial design variables for the individuals in the first Number of buildings Percentage (%)
population. Then, with background inputs, simulations are con-
One 66.1
ducted to evaluate the performance by determining the OF or so- Two 18.2
called fitness value for each individual. The genetic algorithm con- Three 8.3
ducts selection, crossover, and mutation on the design variables of Four and more 7.4
individuals by using the objective function in the current genera- Surroundings Percentage (%)
High-density residential area 85.8
tion to generate a new population. The optimization process stops
Low-density open crossroads 14.2
when the optimal individual stays the same for five consecutive
5
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

six-story courtyard of a building in Tianjin University, China. In Fig. 2 compares the simulated and measured T mrt and the sim-
order to cover different seasons and meteorological conditions, ulation accurately reflected the variation of the measurement with
measurements were conducted on August 13, 2018, October 24, a R2 of 0.979. Also, compared with previous studies [18,33], the
2018, February 28, 2019 and May 13, 2019. To avoid the rainy peri- error between the simulated and measured data was acceptable
ods, measurements on August 13, 2018 was conducted from 10:00 with RMSE (root mean square error) of 3.098 K and MBE (mean
a.m. to 15:00 p.m. The rest three measurements were performed bias error) of 2.090 K. Hence the accuracy of Ladybug tool to
from 8:00 a.m. to 15:00 p.m. The six-directional (upward, down- model the T mrt was validated.
ward, east, south, west, north) short wave and long wave solar
radiations were measured by short wave radiometer Kipp&Zonen
and long wave radiometer CNR4 at the height of 1.1 m. With the
measured short and long wave solar radiation intensities from 2.2.2. Wind speed
six directions, the T mrt was calculated via the following equation To validate the wind speed simulation, we used the experimen-
[30]: tal data from Case 1H by Tominaga et al. [57]. As shown in Fig. 3,
sffiffiffiffiffiffiffiffi Tominaga et al. [57] provided a dataset of the three-dimensional
4 Sstr (3D) turbulent flow over nine identical cubic buildings in an ABL
T mrt ¼  273:15 ð9Þ
ep r wind tunnel. The width, depth and height of each building and
the distance between all the buildings were all equal to H
X
6 X
6 = 0.1 m. The reference inlet velocity measured at H was
Sstr ¼ ak K i F i þ ep Li F i ð10Þ U H = 3.1 m/s. The simulated and measured wind velocity was com-
i¼1 i¼1
pared along the y/H = 0 vertical line at x/H = 1.4, x/H = 1.0, x/H
where Sstr (W/m2) is the mean radiant flux intensity of the human = 0.6, x/H = 1.4, x/H = 1.0, x/H = 0.6, as shown in Fig. 4. The loca-
body, ep = 0.97 is the emissivity of the human body, and ak = 0.7 is tions of the measurement points were illustrated in Fig. 3. A good
short wave absorptivity. K i and Li are the short and long wave radi- agreement between the simulated and measured data can be seen
ation intensities, respectively. The angle factors F i are set to 0.22 for from Fig. 4 with average R2 of 0.991, average RMSE of 0.074 and
radiant fluxes from the east, west, south and north, and 0.06 for average MEB of 0.016. The maximum discrepancy of 0.26 m/s
upward and downward radiant fluxes. The meteorological data appeared at the position of x/H = 0.6, z/H = 0.3, because steady
from the nearest meteorological station, including the T a ; Idir ; Idiff , RANS simulation is not good at predicting the separation and recir-
and Ihor , was used to drive the simulation. The time-averaged mete- culation in the downstream of buildings [12] [11]. With reasonable
orological parameters are shown in Table 2. overall accuracy, we validated the standard k-e model in the But-

Fig. 5. Classification of the investigated kindergartens based on building forms.

6
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

terfly plugin in Rhino - Grasshopper for subsequent simulation of


outdoor flow.

3. Case study

According to Chinese Code for Design of Nursery and Kinder-


garten Buildings [27], the general layout of kindergartens should
include well-designed outdoor space in line with children’s physi-
ological and psychological characteristics. The geometric form and
location of a kindergarten would significantly alter the microcli-
mate of its outdoor spaces. Therefore, our optimization framework
was applied to the inverse design of a kindergarten for
demonstration.

3.1. General forms of kindergartens

In order to reasonably setup the design variables and con-


straints, we collected a list of kindergartens from a school informa-
tion database in China [32] and randomly selected 120 samples of
kindergarten to study their general forms, including the building
height, building shape, number of buildings, and the surrounding
environment. First, it was found that most of the main building
of these kindergartens have three floors, which is in accordance
with the Chinese Code for Design of Nursery and Kindergarten
Buildings [27]. Then, as Table 3 shows, most of the selected kinder-
gartens were single-building layout built within high-density res-
idential areas. The building forms can be summarized into strip
form, L form, C form, ladder form, T form, branch form, H form, O
form and courtyard form. The demonstration and proportion of
each form are shown in Fig. 5. The above analysis of the building
forms and surroundings provided a foundation for the subsequent
setup of a typical design case.

3.2. Surroundings, constraints, and design variables

According to the analysis, we decided to set the design in a


high-density residential community. A rectangular site of
42 m  38 m located in the middle of a typical community was
Fig. 6. Residential buildings surrounding the property line of the kindergarten
selected as the construction site for the kindergarten. The site looking from (a) Top view and (b) Perspective view.
was surrounded by high-density residential buildings with height
from 30 to 40 meters. It is worth mentioning that such kind of
high-density residential community is typical in China, and it is and y1 , was set and the area of the middle box (S1 ) was then deter-
common to have kindergarten within residential communities to mined. Because the whole building has limited floor area, the area
offer convenient care services to preschool children from sur- of the second (S2 ) and third box (S3 ) could be distributed by spec-
rounding communities. Fig. 6 shows the top and perspective views ifying the area ratio (S2 /(S2 +S3 )). Then the shape of the second and
of the kindergarten site with surrounding residential buildings. third box could be determined by their lengths, x2 and x3 . Finally,
While setting up the geometric model of the kindergarten, the the relative position of the three boxes is determined by the posi-
plot ratio was limited to 0.6 and the number of floors was set to tions of the second and third boxes from the middle box, y12 ; y13 .
three with the floor height of 4 m, according to Standard for Kin- The ranges and allowed change steps for the design variables are
dergarten Construction [35]. shown in Table 4.
After the surroundings and constrains were determined, we
started to generate the building on the construction site. Analysis
shows that the majority of kindergartens only have one building, 3.3. Studied cities and their climate
so we placed a single building on the site. For the building shape,
we used the combination of three boxes with variable length, In order to examine the effect of climate on microclimate opti-
width, and orientation to represent the building since three boxes mization, we chose to design the kindergarten under the climate of
can easily create most building forms investigated (Fig. 7(c)). To Tianjin and Shanghai. Tianjin (38°33’N–40°15’N, 116°42’E–118°0
start with, four values (0°, 90°, 180°and 360°) were provided for 4’E) is located in the northern part of China with an annual average
the rotation angle a of the building. When a = 0°or 180°(Figure 7 T a of 12.9 °C. The hottest month July has a monthly average T a of
(a)), the position of the building could only be moved along the 26.1 °C and coldest month January has a monthly average T a of
vertical central axis and L1 was the distance from the center of 2.4 °C. Shanghai (30°40’N - 31°53’N, 120°51’E - 122°12’E) is
the middle box to the center of the site. When a = 90°or located in the southern part of China. The annually mean T a of
270°(Fig. 7(b)), the position could be changed along the horizontal Shanghai is about 16.9 °C, monthly mean T a of the hottest month
central axis and L2 was the distance. After determining the building July is about 27.5 °C and monthly mean T a of the coldest month
orientation and position, the length and width of the middle box, x1 January is about 4.5 °C.
7
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 7. Explanation for the building design variables: (a) A case where the building is parallel to the length of the site, (b) A case where the building is parallel to the width of
the site, and (c) Examples of generated buildings of strip form, L form, C form, latter form, H form, and T form.

Table 4 meteorological parameters in the two months were selected for


Ranges and step sizes for the design variables boundary conditions of microclimate simulations, as shown in
Table 5.
Variables Range Step size
a (°) [0, 360] 90
L1 (m) [6.5, 6.5] 0.1
L2 (m) [8, 8] 0.1 4. Results
S2 /(S2 +S3 ) [0.35, 0.65] 0.01
x1 (m) [7, 12.5] 0.1 This section first presents the variation of OF to illustrate the
y1 (m) [7, 12.5] 0.1 effect of optimization. Then, the worst solution, the best solution
x2 (m) [7, 12.5] 0.1
x3 (m) [7, 12.5] 0.1
in summer, the best solution in winter and the overall best solution
y12 [y1 /2, y1 /2] 0.1 are shown to provide guidance for building design.
y13 [y1 /2, y1 /2] 0.1

4.1. Evolution of OF

Fig. 9 shows the optimization process by demonstrating OFs at


The outdoor activity of children in kindergarten is usually con- each generation under the climates of Tianjin. A decreased trend of
ducted at 15:00 according to our investigation. Thus, this study OF was demonstrated in Fig. 9, indicating that the overall thermal
used 15:00 as the time for microclimate evaluation and lasted stress was reduced. The maximum and minimum values of OF in
for an hour. In addition, to comprehensively consider the effect Fig. 9 were 6.53 °C and 5.37 °C occurred at generation 9 and 15,
of hot and cold conditions, this study used climate data from both respectively, with a difference of 1.16 K. Fig. 10 shows that under
summer and winter as inputs. To avoid vacations in July and Jan- the Shanghai climate, the OF also showed a overall decreasing
uary, June and December were selected as the time of study. trend with the optimization. The maximum and minimum values
Fig. 8 shows the T a , wind speed and wind roses at 15:00 in Tian- of OFs were 3.57 °C and 2.87 °C at Generation 2 and 15, respec-
jin and Shanghai in June and December. The median values of tively, with a reduction of 0.70 K. Therefore, this framework was
8
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 8. Meteorological parameters at 15:00 in June and December in Tianjin and Shanghai: (a) air temperature, (b) wind speed, and (c) wind roses.

Table 5
Boundary conditions for the numerical simulations.

City Shanghai Tianjin


Season Summer Winter Summer Winter
T a (°C) 26.5 9 27 2.7
Relative Humidity (%) 76 48 70 63
U ref (m/s) 4 3 3 3
Wind Direction East North South North
Idir (Wh/m2) 683 302 476 305
Idiff (Wh/m2) 190 98 214 164
Ihor (Wh/m2) 398 301 415 268

Fig. 9. Violin plots of objective function of each generation in Tianjin.


Fig. 10. Violin plots of objective function of each generation in Shanghai.

effective for improving outdoor thermal comfort conditions in both


Tianjin and Shanghai.
optimized kindergarten building was ladder form, located in the
middle of the site with north–south orientation. The best solution
4.2. The comparison of the worst and best solutions in summer was branch form, located in the middle of the site and
north–south orientation. Comparing the best solution with the best
Figs. 11–14 shows the wind speed, T mrt and UTCI distributions solution in summer, we could see that although the best solution
of the worst, the best in winter, the best in summer and overall cast less shadows in summer (Fig. 11(b)), it also block less sunlight
best solutions during the investigated time in Tianjin and Shang- in winter (Fig. 12(b)). Also, the wind from the north in winter was
hai. Besides, the wind speed, T mrt and UTCI distributions of the accelerated by the north edges of the best solution in summer
overall site without the kindergarten are also illustrated in (Fig. 12(a)), causing more cold stress than the best solution. The
Fig. 11–14 to demonstrate the impact of surrounding buildings best solution in winter was a east-west oriented C form building,
on the target site. located in the north of the site. It obstructed the strong airflow
In Tianjin, heat stress accounts for 60% of the OF and cold stress from the north (Fig. 12)) and cast little shadow on the site
account for 40% of the OF, so the heat stress in summer was given a (Fig. 12(b)) in winter. The least recommended solution was ladder
little priority over cold stress in winter during optimization. The form, located in the south of the site and east–west orientation. It
9
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 11. Distributions of (a) wind speed, (b) T mrt and (c) UTCI in summer for the overall site without kindergarten, the worst, the best in winter, the best in summer and the
overall best solutions at 1 m in Tianjin.

10
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 12. Distributions of (a) wind speed, (b) T mrt and (c) UTCI in winter for the overall site without kindergarten, the worst, the best in winter, the best in summer, and the
overall best solution at 1 m in Tianjin.

11
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

cast large shadows in winter (Fig. 12)) and small shadows in sum- from the east and accelerated in the canyon (Fig. 13(c)). Also, the
mer (Fig. 11(b)). Not only could it not block the airflow from the flow from the southeast was further accelerated by the south edge
north in winter, it accelerated it at its northern edges (Fig. 12(a)). of the target building. The high wind speed together with large
In Shanghai, cold stress accounts for just 13% of OF while heat shadows created a strong cooling effect in summer. The best solu-
stress account for 87% of OF, so the optimization basically just con- tion in winter was a branch form building parallel to the length of
sidered the cooling effect in summer. For the same reason, the best the site, located in the middle of the site with low wind speed in
solution was almost identical to the best solution in summer, winter (Fig. 14(c)). The least recommended solution was strip
which was ladder-shaped, located in the middle of the site and par- form, located in the south of the site and parallel to the length of
allel to the width of the site (Fig. 13, Fig. 14). In summer, the solu- the site(Fig. 13, Fig. 14). It could block little sunshine in summer
tion cast large shadows (Fig. 13)) and formed a canyon with the (Fig. 13)) and accelerate the northern flow of winter with the long
surrounding buildings on the north (Fig. 13(c)). The flow entered windward wall (Fig. 14(c)).

Fig. 13. Distributions of (a) wind speed, (b) T mrt and (c) UTCI in summer for the overall site without kindergarten, the worst, the best in winter, the best in summer, and the
overall best solutions at 1 m in Shanghai.

12
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 14. Distributions of (a) wind speed, (b) T mrt and (c) UTCI in winter for the overall site without kindergarten, the worst, the best in winter, the best in summer, and the
overall best solution at 1 m in Shanghai.

5. Discussion and winter. The data used in the correlation analyses were from
different individuals during the optimization.
In addition to thermal stress, we also compared the effect of Fig. 16 and Fig. 16(a) show that T mrt had stronger correlation
wind speed and solar radiation on outdoor thermal comfort during with thermal stress than wind speed in summer in both Tianjin
the optimization since they are two key factors that lead to the and Shanghai. That’s because wind speed on the target site was
spatial variations of thermal stress. Then, limitations and future generally low and the soalr radiation was very strong in summer,
studies are discussed for future improvement. so changes in wind speed had less impact on UTCI than T mrt .
Fig. 15(b) illustrates that T mrt had stronger correlation with ther-
mal stress than wind speed in winter Tianjin. Comparing Fig. 15
5.1. Comparing the effect of solar radiation and wind (b) and Fig. 15(b), it can be found that wind speed had higher cor-
relation with thermal stress in winter than in summer. One expla-
Fig. 15 and Fig. 16 further identified the correlations of T mrt and nation is that the wind speed in winter was higher than summer in
wind speed with thermal stress on featured scenarios in summer Tinajin. Another possible reason is that UTCI has been found to be

13
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

Fig. 15. T mrt and wind speed vs. (a) average heat stress and (b) average cold stress in Tianjin.

Fig. 16. T mrt and wind speed vs. (a) average heat stress and (b) average cold stress in Shanghai.

sensitive to identify severe weather caused by excessive wind low average wind speeds, as the wind speed increases, the effect of
speed in winter [48]. Fig. 16(b) shows that wind speed had strong wind may be equal to that of solar radiation and even exceed it. For
correlation with cold stress while T mrt had almost no correlation example, Brozovsky et al. [13] found that wind and sun can
with cold stress in winter Shanghai. The explanation can be found become equally effective in increasing outdoor thermal comfort
in Fig. 14, where we can see that the shaded places with wind when increasing inlet wind speed to 8 m/s. Besides, the influence
speed lower than 0.6 m/s were still comfortable, so excessive wind of radiation and wind on thermal comfort is also related to other
speed was the main cause of cold stress in winter Shanghai. In gen- parameters. For example, in winter Shanghai with a relatively
eral, on the simulated sunny days in Tianjin and Shanghai with a warm T a , wind sheltering was the more effective way to improve
comparatively low wind speed of 3 m/s and 4 m/s as an inlet con- outdoor thermal comfort than solar access.
dition, solar radiation contributed more to outdoor thermal com-
fort than wind speed. 5.2. Limitations and future studies
Our result corroborates the findings from other ‘‘radiation over
wind” studies. Liu et al. [42] in Tianjin concluded that the effect of Limitations in this study and some possible directions for fur-
sun on outdoor thermal comfort was more than two times greater ther improvement are discussed below. Firstly, this study simu-
that of wind. Liu et al. [45] in Changsha found that solar radiation lated wind and solar radiation while neglected the spatial
contributed more than wind speed to human outdoor thermal sen- variation of air temperature and humidity. On one hand, this is
sation. Xu et al. [59] investigated thermal comfort during winter in due to the less impact of urban morphology on air temperature
Xi’an and found that solar radiation had the most significant influ- and humidity than wind and solar radiation. For example, the air
ence on the outdoor overall comfort among the microclimatic temperature is changed by 2 °C on average for different climate
parameters. Moreover, wind speed simulation is more computa- on the hottest time of a day, and the humidity change was also lim-
tionally intensive than estimation of solar radiation. Therefore, ited [41]. On the contrary, the changes of the solar radiation and
when computation resources are limited, the optimization of solar wind by the urban morphology are huge. On the other hand, the
radiation should be given priority over wind. time consumed in the co-simulation of the entire microclimate is
However, the above ‘‘radiation over wind” principle cannot be prohibitive since the simulation is conducted for many times for
applied universally. Although wind is less effective than solar radi- a optimization study. Secondly, outdoor thermal comfort was used
ation to improve outdoor thermal comfort in locations with rather as a single assessment standard in this study and the indoor com-
14
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

fort, the day-lighting and energy consumption of the buildings Acknowledgments


have not been considered, so multi-objective optimization could
be carried out in the future. Thirdly, this study just considered This work was partially supported by the National Natural
the impact of urban geometry on microclimate, however, microcli- Science Foundation of China (Grant No. 52008242 and 51808487)
mate variability into the urban texture also depends on vegetation, and Stiftelsen für internationalisering av hügre utbild-ning och
buildings and pavements materials, anthropogenic activities and forskning (STINT), Sweden (Dnr: CH2020-8665).The computations
so on. Therefore, evapotranspiration phenomena of plants, charac- were partially enabled by resources provided by the Swedish
terization of buildings and pavements materials, thermal mass of National Infrastructure for Computing (SNIC) at PDC partially
surrounding buildings and grounds and the simulation of anthro- funded by the Swedish Research Council through grant agreement
pogenic heat should be introduced to this framework to enhance No. 2018-05973. We thank undergraduate students Yuanyuan Wu,
the accuracy of simulations in the future. Fourthly, the UTCI index Yaxi Zuo, Di Chen, and Yulun Xu for their participation and discus-
used in our study does not particularly consider the range of age. sions in the project.
Since we modelled a kindergarten in our case study, it would be
more rigorous if we adjusted human thermal comfort model to
children. We do notice that some researchers had studied adaptive
References
outdoor thermal comfort models for children, but their conclusions
can be only applied to the climates that they studied and universal [1] J.A. Acero, E.J. Koh, L.A. Ruefenacht, L.K. Norford, Modelling the influence of
conclusion hasn’t been reached yet at present [31,52,24]. Relevant high-rise urban geometry on outdoor thermal comfort in singapore, Urban
studies can be carried out in the future. Climate 36 (2021).
[2] S. Achour-Younsi, F. Kharrat, Outdoor thermal comfort: impact of the geometry
of an urban street canyon in a mediterranean subtropical climate–case study
tunis, tunisia, Proc.-Soc. Behav. Sci. 216 (2016) 689–700.
6. Conclusions
[3] R. Aghamolaei, M.M. Azizi, B. Aminzadeh, J. O’Donnell, A comprehensive
review of outdoor thermal comfort in urban areas: Effective parameters and
The contributions of this study can be summarized by two approaches, Energy Environ. 0958305X221116176 (2022).
parts: Theoretically, a framework to optimize the outdoor thermal [4] S. Ahmadi, M. Yeganeh, M.B. Motie, A. Gilandoust, The role of neighborhood
morphology in enhancing thermal comfort and resident’s satisfaction, Energy
comfort during early building design stage was proposed. The Reports 8 (2022) 9046–9056.
influence of building form and location on the thermal comfort [5] E. Arens, T. Hoyt, X. Zhou, L. Huang, H. Zhang, S. Schiavon, Modeling the
of its outdoor space was evaluated by normalizing the spatially comfort effects of short-wave solar radiation indoors, Build. Environ. 88 (2015)
3–9.
and temporally averaged thermal stress. After validating the wind [6] ASHRAE, 2020. Standard 55-2020. Thermal environmental conditions for
speed and solar radiation simulations, the proposed framework human occupancy, ASHRAE.
was applied to optimize the outdoor thermal comfort of a kinder- [7] I.V. Bajšanski, D.D. Milošević, S.M. Savić, Evaluation and improvement of
outdoor thermal comfort in urban areas on extreme temperature days:
garten located in a high-density residential community under the Applications of automatic algorithms, Build. Environ. 94 (2015) 632–643.
climates of Tianjin and Shanghai. The framework has been proven [8] Blazejczyk, K., 1992. Menexme man-environment heat exchange model and
to be able to improve the time-averaged outdoor thermal comfort its applications in bioclimatology, in: Proceedings of the fifth international
conference on environmental ergonomics, pp. 142–143.
at building scale. In Tianjin, optimization reduced the overall ther- [9] K. Blazejczyk, Y. Epstein, G. Jendritzky, H. Staiger, B. Tinz, Comparison of utci to
mal stress from 6.53 °C to 5.37 °C. In Shanghai, optimization selected thermal indices, Int. J. Biometeorol. 56 (2012) 515–535.
reduced the overall thermal stress from 3.57 °C to 2.87 °C. It indi- [10] B. Blocken, Les over rans in building simulation for outdoor and indoor
applications: A foregone conclusion?, Building Simulation, Springer (2018)
cates that this study expanded the spatial and temporal scale of
821–870
outdoor thermal comfort inverse design. [11] B. Blocken, T. Stathopoulos, J. Carmeliet, J.L. Hensen, Application of
Practically, this study provides some guidelines for kinder- computational fluid dynamics in building performance simulation for the
garten design considering outdoor thermal comfort in Tianjin and outdoor environment: an overview, J. Build. Performance Simul. 4 (2011) 157–
184.
Shanghai: Both heat and cold stresses should be considered in [12] B. Blocken, T. Stathopoulos, P. Saathoff, X. Wang, Numerical evaluation of
Tianjin, while in Shanghai, heat stress in summer was given prior- pollutant dispersion in the built environment: comparisons between models
ity over cold stress in winter. That’s because cold stress and heat and experiments, J. Wind Eng. Ind. Aerodyn. 96 (2008) 1817–1831.
[13] J. Brozovsky, S. Corio, N. Gaitani, A. Gustavsen, Evaluation of sustainable
stress were comparable in Tianjin while heat stress was seven strategies and design solutions at high-latitude urban settlements to enhance
times stronger than cold stress in Shanghai under the simulated outdoor thermal comfort, Energy Build. 244 (2021).
weather conditions. In Tianjin, the optimized kindergarten build- [14] A. Chatzidimitriou, S. Yannas, Street canyon design and improvement potential
for urban open spaces; the influence of canyon aspect ratio and orientation on
ing was ladder form, located in the middle of the site with microclimate and outdoor comfort, Sustainable Cities Soc. 33 (2017) 85–101.
north–south orientation. This form casts big shadows in summer [15] L. Chen, B. Yu, F. Yang, H. Mayer, Intra-urban differences of mean radiant
without blocking too much sunlight in winter. In Shanghai, the temperature in different urban settings in shanghai and implications for heat
stress under heat waves: A gis-based approach, Energy Build. 130 (2016) 829–
best solution was ladder-shaped, located in the middle of the site 842.
and parallel to the width of the site. This form casts large shadows [16] J. Du, C. Sun, L. Liu, X. Chen, J. Liu, Comparison and modification of
and accelerates wind speed in summer. Wind demonstrated more measurement and simulation techniques for estimating tmrt in summer and
winter in a severely cold region, Build. Environ. 199 (2021).
limited contributions than solar radiation during optimization
[17] Eshelman, L.J., Schaffer, J.D., 1993. Real-coded genetic algorithms and interval-
because the wind speed distributions were relatively low and uni- schemata, in: Foundations of genetic algorithms. Elsevier. vol. 2, pp. 187–202.
form with in high-density urban area. [18] G. Evola, V. Costanzo, C. Magrı̀, G. Margani, L. Marletta, E. Naboni, A novel
comprehensive workflow for modelling outdoor thermal comfort and energy
demand in urban canyons: Results and critical issues, Energy Build. 216
Data availability (2020).
[19] D. Fiala, G. Havenith, P. Bröde, B. Kampmann, G. Jendritzky, Utci-fiala multi-
node model of human heat transfer and temperature regulation, Int. J.
Data will be made available on request. Biometeorol. 56 (2012) 429–441.
[20] Franke, J., Hellsten, A., Schlünzen, K., Carissimo, B., 2007. Best practice
guideline for the cfd simulation of flows in the urban environment-a
Declaration of Competing Interest summary, in: 11th Conference on Harmonisation within Atmospheric
Dispersion Modelling for Regulatory Purposes, Cambridge, UK, July 2007,
The authors declare that they have no known competing finan- Cambridge Environmental Research Consultants.
[21] J. Franke, A. Hellsten, K.H. Schlunzen, B. Carissimo, The cost 732 best practice
cial interests or personal relationships that could have appeared guideline for cfd simulation of flows in the urban environment: a summary,
to influence the work reported in this paper. Int. J. Environ. Pollut. 44 (2011) 419–427.

15
R. Sun, J. Liu, D. Lai et al. Energy & Buildings 284 (2023) 112824

[22] A. Ghaffarianhoseini, U. Berardi, A. Ghaffarianhoseini, Thermal performance [45] W. Liu, Y. Zhang, Q. Deng, The effects of urban microclimate on outdoor
characteristics of unshaded courtyards in hot and humid climates, Build. thermal sensation and neutral temperature in hot-summer and cold-winter
Environ. 87 (2015) 154–168. climate, Energy Build. 128 (2016) 190–197.
[23] K. Giannopoulou, M. Santamouris, I. Livada, C. Georgakis, Y. Caouris, The [46] T.R. Oke, G. Mills, A. Christen, J.A. Voogt, Urban climates, Cambridge University
impact of canyon geometry on intra urban and urban: suburban night Press, 2017.
temperature differences under warm weather conditions, Pure Appl. Geophys. [47] B. Paramita, H. Fukuda, R. Perdana Khidmat, A. Matzarakis, Building
167 (2010) 1433–1449. configuration of low-cost apartments in bandung–its contribution to the
[24] H. Gu, Q. Hu, D. Zhu, J. Diao, Y. Liu, M. Fang, Research on outdoor thermal microclimate and outdoor thermal comfort, Buildings 8 (2018) 123.
comfort of children’s activity space in high-density urban residential areas of [48] S. Provençal, O. Bergeron, R. Leduc, N. Barrette, Thermal comfort in quebec city,
chongqing in summer, Atmosphere 13 (2022) 2016. canada: sensitivity analysis of the utci and other popular thermal comfort
[25] D. Hammond, L. Chapman, J. Thornes, Roughness length estimation along road indices in a mid-latitude continental city, Int. J. Biometeorol. 60 (2016) 591–
transects using airborne lidar data, Meteorol. Appl. 19 (2012) 420–426. 603.
[26] A.W. Hayes, T.A. Loomis, Loomis’s essentials of toxicology, Elsevier, 1996. [49] P. Richards, R. Hoxey, Appropriate boundary conditions for computational
[27] Heilongjiang Institute of Architectural Design, 2016. Chinese code for design of wind engineering models using the k-) turbulence model, J. Wind Eng. Ind.
nursery and kindergarten buildings. Aerodyn. 46 (1993) 145–153.
[28] T. Highfill, C. Franks, Measuring the us outdoor recreation economy, 2012– [50] Roudsari, M.S., Pak, M., Smith, A., et al., 2013. Ladybug: a parametric
2016, J. Outdoor Recreat. Tourism 27 (2019). environmental plugin for grasshopper to help designers create an
[29] B. Hong, B. Lin, Numerical studies of the outdoor wind environment and environmentally-conscious design, in: Proceedings of the 13th international
thermal comfort at pedestrian level in housing blocks with different building IBPSA conference held in Lyon, France Aug, pp. 3128–3135.
layout patterns and trees arrangement, Renewable Energy 73 (2015) 18–27. [51] A. Salvati, H.C. Roura, C. Cecere, Assessing the urban heat island and its energy
[30] P. Höppe, Ein neues verfahren zur bestimmung der mittleren impact on residential buildings in mediterranean climate: Barcelona case
strahlungstemperatur im freien, Wetter und Leben 44 (1992) 147–151. study, Energy Build. 146 (2017) 38–54.
[31] B. Huang, B. Hong, Y. Tian, T. Yuan, M. Su, Outdoor thermal benchmarks and [52] L. Shao, X. He, Y. Tang, S. Wu, Outdoor cold stress and cold risk for children
thermal safety for children: A study in china’s cold region, Sci. Total Environ. during winter: A study in china’s severe cold regions, Buildings 12 (2022) 936.
787 (2021). [53] J. Spagnolo, R. De Dear, A field study of thermal comfort in outdoor and semi-
[32] Huangjing, 2015. 51sxue. http://xuexiao.51sxue.com/schoolByArea/t_1.html. outdoor environments in subtropical sydney australia, Build. Environ. 38
[33] Y. Ibrahim, T. Kershaw, P. Shepherd, I. Elwy, A parametric optimisation study (2003) 721–738.
of urban geometry design to assess outdoor thermal comfort, Sustain. Cities [54] H. Taleb, M.A. Musleh, Applying urban parametric design optimisation
Soc. 75 (2021). processes to a hot climate: Case study of the UAE, Sustain. Cities Soc. 14
[34] E. Jamei, P. Rajagopalan, M. Seyedmahmoudian, Y. Jamei, Review on the impact (2015) 236–253.
of urban geometry and pedestrian level greening on outdoor thermal comfort, [55] M. Taleghani, L. Kleerekoper, M. Tenpierik, A. Van Den Dobbelsteen, Outdoor
Renew. Sustain. Energy Rev. 54 (2016) 1002–1017. thermal comfort within five different urban forms in the netherlands, Build.
[35] Jiangsu Education Department, 2016. Standard for kindergarten construction. Environ. 83 (2015) 65–78.
[36] A.N. Kakon, N. Mishima, The effects of building form on microclimate and [56] Y. Tominaga, A. Mochida, R. Yoshie, H. Kataoka, T. Nozu, M. Yoshikawa, T.
outdoor thermal comfort in a tropical city, J. Civil Eng. Architect. 6 (2012) Shirasawa, Aij guidelines for practical applications of cfd to pedestrian wind
1492. environment around buildings, J. Wind Eng. Industr. Aerodyn. 96 (2008) 1749–
[37] Z. Kaseb, M. Hafezi, M. Tahbaz, S. Delfani, A framework for pedestrian-level 1761.
wind conditions improvement in urban areas: Cfd simulation and [57] Y. Tominaga, M. Shirzadi, Wind tunnel measurement dataset of 3d turbulent
optimization, Build. Environ. 184 (2020). flow around a group of generic buildings with and without a high-rise
[38] N. Kikon, P. Singh, S.K. Singh, A. Vyas, Assessment of urban heat islands (uhi) of building, Data in Brief 39 (2021).
noida city, india using multi-temporal satellite data, Sustain. Cities Soc. 22 [58] H.G. Weller, G. Tabor, H. Jasak, C. Fureby, A tensorial approach to
(2016) 19–28. computational continuum mechanics using object-oriented techniques,
[39] S.W. Kim, R.D. Brown, Urban heat island (uhi) variations within a city Comput. Phys. 12 (1998) 620–631.
boundary: A systematic literature review, Renew. Sustain. Energy Rev. 148 [59] M. Xu, B. Hong, J. Mi, S. Yan, Outdoor thermal comfort in an urban park during
(2021). winter in cold regions of china, Sustain. Cities Soc. 43 (2018) 208–220.
[40] D. Lai, D. Guo, Y. Hou, C. Lin, Q. Chen, Studies of outdoor thermal comfort in [60] X. Xu, Y. Wu, W. Wang, T. Hong, N. Xu, Performance-driven optimization of
northern china, Build. Environ. 77 (2014) 110–118. urban open space configuration in the cold-winter and hot-summer region of
[41] D. Lai, W. Liu, T. Gan, K. Liu, Q. Chen, A review of mitigating strategies to china, Building Simulation, Springer (2019) 411–424.
improve the thermal environment and thermal comfort in urban outdoor [61] X. Xu, C. Yin, W. Wang, N. Xu, T. Hong, Q. Li, Revealing urban morphology and
spaces, Sci. Total Environ. 661 (2019) 337–353. outdoor comfort through genetic algorithm-driven urban block design in dry
[42] K. Liu, Z. Lian, X. Dai, D. Lai, Comparing the effects of sun and wind on outdoor and hot regions of china, Sustainability 11 (2019) 3683.
thermal comfort: A case study based on longitudinal subject tests in cold [62] X. Zhang, L. Han, H. Wei, X. Tan, W. Zhou, W. Li, Y. Qian, Linking urbanization
climate region, Sci. Total Environ. 825 (2022). and air quality together: A review and a perspective on the future sustainable
[43] S. Liu, W. Pan, X. Zhao, H. Zhang, X. Cheng, Z. Long, Q. Chen, Influence of urban development, J. Clean. Prod. 130988 (2022).
surrounding buildings on wind flow around a building predicted by cfd [63] T. Cao, Z. Lian, S. Ma, J. Bao, Thermal comfort and sleep quality under
simulations, Build. Environ. 140 (2018) 1–10. temperature, relative humidity and illuminance in sleep environment, Journal
[44] W. Liu, R. Duan, C. Chen, C.H. Lin, Q. Chen, Inverse design of the thermal of Building Engineering 43 (2021) 102575.
environment in an airliner cabin by use of the cfd-based adjoint method, [64] X. Xu, Z. Lian, J. Shen, L. Lan, Environmental factors affecting sleep quality in
Energy Build. 104 (2015) 147–155. summer: A field study in Shanghai, China, Journal of Thermal Biology 99
(2021) 102977.

16

You might also like