You are on page 1of 15

Building and Environment 87 (2015) 154e168

Contents lists available at ScienceDirect

Building and Environment


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

Thermal performance characteristics of unshaded courtyards in hot


and humid climates
Amirhosein Ghaffarianhoseini a, Umberto Berardi b, *, Ali Ghaffarianhoseini c
a

Department of Geography, Faculty of Arts and Social Science, University of Malaya (UM), Kuala Lumpur, Malaysia
Department of Architectural Science, Ryerson University, 350 Victoria Street, Toronto, Ontario M5B 2K3, Canada
c
School of Engineering, Faculty of Design and Creative Technologies, AUT University, Auckland, New Zealand
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 10 November 2014
Received in revised form
17 January 2015
Accepted 1 February 2015
Available online 7 February 2015

In recent years, there has been a growing interest in the design of courtyards for the microclimatic
enhancement of outdoor spaces. However, there is still little knowledge regarding the thermal performance characteristics of courtyards, particularly in hot and humid climates. This study evaluates the
ability of unshaded courtyards for providing thermally comfortable outdoor spaces according to different
design congurations and scenarios, including the orientations, height and albedo of wall enclosure, and
use of vegetation. The software ENVI-met was used as a tool for simulating the thermal performance of
courtyards in the hot and humid climate of Kuala Lumpur, Malaysia. The PMV and the number of hours
per day that a courtyard could be enjoyed once the proposed design suggestions were implemented are
assessed. Likewise, the Physiologically Equivalent Temperature (PET) index allowed to further explore
the thermal comfort conditions of courtyards. As a result, guidelines are proposed in order to optimize
the design of courtyards towards enhancing their thermal performance characteristics. In particular, the
study shows that according to design parameters such as the building height ratio, an abundance in the
amount vegetation the courtyard can achieve an acceptable level of thermal comfort for the tropics and
may be enjoyed by its users for a long duration of daytime even during the noontime. Finally, this paper
stresses that only well designed courtyards may represent a valid option for sustainable built
environments.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Outdoor thermal comfort
Urban microclimate
Hot and humid climate
Courtyard
Urban design

1. Introduction
Modeling the relationship between buildings and the surrounding outdoor environment is a multidisciplinary imperative
for urban climate and outdoor thermal comfort [1e3]. In view of
the negative impacts of the urban heat island effect, particularly on
energy use, air quality and human health [4] and its signicant
inuence on urban comfort [5], meteorological studies which
previously focused primarily on the meso-scale (10e40 km) have
recently started to focus on the micro-scale (less than 1 km). This is
due to the importance of the microclimate of outdoor spaces and
urban canopy layers as signicant elements of contemporary urban
areas [3e5].
Given the growing interest in outdoor thermal comfort and
urban life [6], various attempts have been made to study the impacts of courtyards on natural ventilation and thermal comfort

* Corresponding author. Tel.: 1 416 979 5000x3263; fax: 1 416 979 5153.
E-mail address: uberardi@ryerson.ca (U. Berardi).
http://dx.doi.org/10.1016/j.buildenv.2015.02.001
0360-1323/ 2015 Elsevier Ltd. All rights reserved.

[7e9]. In fact, several potential benets can be achieved by controlling the micro-scale characteristics of outdoor spaces through
courtyards.
The impact of courtyards in some climates has been assessed
qualitatively and quantitatively by using eld measurements and
computer modeling [1,10e14] However, there have been very few
studies [2,15,16] that focus on the tropical climate where, due to
high temperatures and relative humidity levels, the utilization of
courtyards merits detailed investigations. In the context of the
tropical climate, cooling effects in outdoor spaces can be enhanced
by reducing the solar radiation received by the ground [17].
This study aims to evaluate quantitatively the thermal effects of
a courtyard in Malaysia and to suggest guidelines to design more
sustainable built environments in this climate zone.
2. Thermal effects of courtyards
A courtyard is an enclosed outdoor or semi-outdoor space surrounded by buildings and open to the sky. Courtyards were primarily adopted in vernacular buildings in parts of Asia, the Middle

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

East, South America, and the Mediterranean countries [18e20].


Their function was to improve comfort conditions by modifying the
microclimate around the building and by enhancing ventilation.
Different archetypes have been adopted for courtyards design in
different countries (and weathers) through the centuries. Romans
and Arabs often included colonnades, especially in convents and
important palaces. The history of using courtyards in Malaysia goes
back to the era of indigenous architecture where Malay traditional
houses in Melaka, Malaysia's oldest city, incorporated enclosed
inner courtyards based on the inuence of traditional Chinese
houses [2,21]. Courtyards were similarly observed in the design of
Chinese traditional [22]. The growing interest in the use of courtyards in various types of contemporary architectural projects
including residential, educational and healthcare is evident in the
Malay context (Fig. 1). Nevertheless, it can be argued that despite
the importance of courtyard design in Malaysia, considerably less
attention has been paid to the circumstances of enhancing their
thermal performance and the development of design guidelines for
improving their effectiveness towards achieving outdoor thermal
comfort.
Literature related to the performance of courtyards mainly examines inter-courtyard air movements, sun-shadow relations, and
formal congurations [24]. One of the rst attempts towards
analyzing the thermal performance of courtyards is reported by
Dunham [25]. Subsequently, the potential of courtyards for
ensuring sufcient ventilation and maximized airow was
conrmed through CFD analysis and wind tunnel studies while
airow patterns within a courtyard were illustrated [25e27].
Likewise, shading inherent in courtyards has been reported to be

155

highly contributive to thermal comfort [26]. Meanwhile, the signicant roles of courtyards in bringing in daylight, ensuring natural
ventilation and optimizing the thermal behavior have been reported by Khan et al. [18] and Acosta et al. [13].
The study by Aldawoud reveals that courtyards are more energy
efcient in hot climates than in temperate or cold climates [25]. AlMasri and Abu-Hijleh identify that a building integrated with
courtyard in the hot and humid climate of Dubai consumes 6.9%
less energy per year in comparison to a typical building [19].
Courtyards are claimed to be highly efcient in enhancing the
ventilation and decreasing the humidity level, as shown by Rajapaksha et al. who also illustrated a strong correlation between
courtyard wall surface temperatures and indoor air temperatures
[15].
Ernest ascertains that the application of bio-climatic features
such as the use of vegetation is highly recommended for improving
the performance of courtyards [28]. For instance, in Israel, the
utilization of trees and grass in courtyards led to enhanced comfort
through their daytime cooling which ranges a PMV between 1.5 and
2.5 based on different landscape treatments [12]. In addition,
experimental studies in Saudi Arabia indicated that covering
courtyards during the daytime while opening it to the sky during
night reduces the outdoor air temperature by 4  C [29].
The study by Safarzadeh and Bahador show that courtyards
alone cannot ensure a high level of thermal comfort in the hot
summer hours in Tehran, Iran, although they can decrease the
cooling energy load [10]. Muhaisen analyzes the impact of different
design congurations of courtyards based on shading simulations
[7]. This study found that shading conditions of courtyards are

Fig. 1. Samples of courtyards in different building types in Malaysia; (a) Courtyard in Melaka town houses; (b) Courtyard in a renovated terrace house, Bangsar; (c) Courtyard in a
restored 18th century Melaka shop house; (d) Courtyard in British council complex, Kuala Lumpur; (e) Courtyard in the University Putra Malaysia, Serdang [23].

156

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

highly inuenced by formal proportions, location latitude and climatic conditions. Furthermore, the study suggests that in the hot
and humid climate of Kuala Lumpur, courtyards with three story
levels and a long axis oriented towards north-east/south-west
result in an optimized performance. In another related study,
Muhaisen and Gadi show that the proportion and geometry of
courtyards play inuential roles in improving the shading performances, hence, deep courtyard forms with any geometry in sumn
mer and shallow forms in winter are recommended [11]. Canto
et al. demonstrate that by comparing the maximum temperatures
in pre-elementary schools in the semi-dry continental climate of
Mendoza, Argentina, an open courtyard is 2  C warmer than a
boxed-type compact courtyard [30]. Yasa and Ok suggest several
wall height ratios that in order to achieve the most optimized form
of courtyards [24].
Makaremi et al. propose the use of trees and plants in order to
provide high shading levels for improving the outdoor thermal
comfort in courtyards in hot and humid climates [2]. The ndings of
that study reveal that solar radiation has a more signicant inuence on physiological equivalent temperature (PET) than the wind
speed, and so radiation needs to be controlled [2].
The recent study by Yahia and Johansson demonstrates that
through the use of greeneries and landscape elements in the hot
and dry climate of Syria, thermal comfort could be achieved even in
the critical hot hours of summer [31]. Likewise, the experimental
analysis of the thermal behavior of courtyards in France demonstrates that the use of trees and water ponds increases the local
thermal comfort, with the maximum PMV at 14:00 of 3.4 for the
empty square, and 0.54 for the model with trees and water pond
[32]. A theoretical and experimental investigation of courtyards in
Greece shows that the use of soil and grass instead of concrete
pavements, in addition to the application of greeneries and water
bodies, lead to cooling effects in summer [33]. Simulations in
summer conrm that water pools lead to a decrease in air

temperature by 0.2  C and the use of soil and grass results in


0.2e0.6  C air temperature reductions [33]. Specically focused on
the advantages of greeneries are the recent studies by Berardi et al.
[34] and La Roche and Berardi [35].
The study by Hisarligil simulates the thermal performance of
courtyards in the temperate climate of Turkey and shows the great
inuence of shading on thermal comfort [36]; the ndings show
that Tmrt is less than 26.4  C in shading areas during the critical
noon time while in the central part of the courtyard which is
exposed to direct solar radiation is 66.5  C. Another recent study
ascertains that 50% coverage of trees or 30% coverage of grass plus
modied building design can decrease the average PET value by
0.4  C in the subtropical climate of Hong Kong [37]. Recently, an
extensive study by Taleghani et al. examines the inuence of the
application of greeneries, ponds and high albedo surfaces in
reducing the heat gain in Portland, US [14]: ndings present 1.6  C
and 1.1  C air temperature reductions for courtyards with vegetation and water ponds respectively, while increasing the albedo of
pavement from 0.37 to 0.91 resulted in a 1.3  C air temperature
reductions and a 2.9  C increase in the Tmrt. Taleghani et al. have
also examined the consequences of applying heat mitigation strategies on urban courtyards in the Netherlands and concluded that
higher albedo of facades leads to higher Tmrt with a maximum increase of 20  C at 12:00, while a water pool in the courtyard
considerably decreases the level of mean radiant temperature with
a maximum decrease of 18  C at 15:00, and greeneries similarly
reduce the Tmrt to a maximum of 17  C [8].
Table 1 summarizes the results of recent studies about strategies
for improving the outdoor thermal comfort in urban environments
especially in courtyards.
 n et al. stress that improper design choices can
Finally, Canto
result in the courtyards being less thermally comfortable than
surrounded open environments [30]. This last paper and the
controversial results of a eld measurement campaign in Kuala

Table 1
Key design strategies for improving the thermal performance characteristics of urban spaces.
Proposed design strategies

Area of investigation

Climate

Key reference

Use of water and sprays


Incorporation of larger surface area and high thermal mass, shallow
plan form and narrow spaces for shade
Use of soil and grass as well as addition of water pools

Al-Oyyena, Riyadh, Saudi Arabia

Hotearid climate
Hot and arid climate

[29]
[38]

The historic centre of Thessaloniki,


Greece
Near the campus of Sharif University
of technology, Tehran, Iran

Temperate climate with high


seasonal variations
Semi-arid, continental climate

[33]

[7]
[32]

Kuala Lumpur, Malaysia

Summer and winter


Typical hot clear sunny day
of summer
Hot and humid climate

Stockholm

Cold climate

Rome

Temperate climate

Cairo

Hot and dry climate

Terrace houses in Malaysia


Israeli desert area

[16]
[1]

Beijing, China
University Putra Malaysia campus
The business district of Hong Kong

Hot and humid climate


Mediterranean summer in
hot and arid climate
Summer and winter
Hot and humid climate
Subtropical climate

Inhabited rural area Damascus, Syria


De Bilt, Netherlands

Hot and dry climate


Temperate climate

[31]
[8]

Portland State University, USA

Summer period in
temperate climate

[14]

Use of shading, water ponds and trees as well as conservation


measures such as insulation, double-glazed windows,
Persian Blinds and sealing tapes
Use of deep courtyard forms in summer and shallow forms in winter
Use of trees and water ponds
Increasing the height to three levels and orienting the long axis
of courtyards towards northeast-southwest
Decreasing the height to one level and orienting the form of
courtyards along north-south
Increasing the height to two levels and orienting the form of
courtyards along north-south
Increasing the height to two levels and with an orientation between
northeast-southwest and north-south
Adding shading roof
Addition of trees or/and galleries
Increasing thermal mass, surface albedo and conductivity
Use of trees and vegetation plus providing high shading levels
Coverage of trees and grass plus strategic building design
modications
Use of greeneries and landscape elements
Use of water pool and urban greeneries plus decreasing the
albedo of facades
Application of greeneries, ponds and low albedo surfaces for
reducing the heat gain

Fleuriot square, Nantes, France

[10]

[7]

[39]
[2]
[37]

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

Lumpur by one of the authors [2], suggested the investigations


proposed in this paper whose aims are to understand the real potential of courtyards in hot and humid climates, and to indicate
design characteristics that can actually reduce the outdoor thermal
comfort.
3. Parametric simulations of courtyards
Parametric simulations for analyzing the thermal effects of
courtyards in Kuala Lumpur have been carried out. Kuala Lumpur is
the capital city of Malaysia, a South-East Asian maritime country
encompassing a tropical area. The city is situated near the equator
(3 80 N, 101420 E) and has a hot and humid climate and abundant
solar radiation. As such, when designing outdoor spaces in this area
special attention needs to be paid to the impacts of solar radiation
and possible ventilation [2,16]. Yearly mean air temperature for
Kuala Lumpur is approximately 27  C, whereas the monthly mean
of maximum air temperatures ranges from 33.5  C (March and
April) to 31.9  C (December), and the monthly mean of minimum
temperatures ranges from 23.1  C in January to 24.3  C in May
[16,40,41].
The relative mean humidity ranges between 70% and 90% with
maximum average daily levels as high as 94% [40]. The monthly
mean solar radiation ranges from 14 to 16 MJ m2 d1. The wind is
regularly light, but it becomes strong with monsoon seasons. In
essence, Malaysia's microclimatic conditions are generally consistent with little variations except for rainfall quantities and wind
directions Overall, generally high temperature and humidity
values, long hours of sunshine and overcast cloud cover plus heavy
rainfalls and relatively weak and variable wind velocity characterize the microclimate of Kuala Lumpur.
3.1. Overview of the methodology
The thermal performance characteristics of courtyards were
analyzed according to the following design parameters: location
and orientation, dimensions and albedo of wall enclosures, and
presence of greeneries. ENVI-met 3.1, a three-dimensional uid
dynamics microclimate software was used to carry out the parametric studies. This software provides the platform to model the
surfaceeplanteair interactions in urban spaces with a typical resolution from 0.5 to 10 m in space and 10 s in time [42]. ENVI-met
requires detailed inputs related to the meteorological data, building, vegetation and soil characteristics.
One of the main advantages of ENVI-met is the capability to
calculate the Tmrt and the predicted mean vote (PMV). Tmrt is
calculated using the ISO 7726 [43] denition as the uniform
temperature of an imaginary enclosure in which the radiant heat
transfer from the human body is equal to the radiant heat transfer
in the actual non-uniform enclosure. Recent studies indicate that
Tmrt has a strong correlation to the human energy balance and
thermal comfort and it is often a parameter better than the air
temperature to assess thermal comfort [44e47]. PMV index was
originally formulated for indoor thermal comfort; however, it has
recently been applied to outdoor conditions too. This parameter is
calculated using the Fanger equation modied for outdoor thermal
comfort conditions [1]. PMV generally ranges between 4 and 4,
with lower and upper bound values denoting the very cold and very
hot sensations respectively; however, these values do not represent
the theoretical limits of PMV.
In order to interpret properly the simulation outputs, it is
essential to consider the assumptions behind ENVI-met as shown
in Table 2.
Subsequently, the study used RayMan 1.2 software to evaluate
the outdoor thermal comfort conditions of courtyards based on

157

Table 2
Assumptions in the simulation.
Assumptions in ENVI-met
Flat ground
Box shaped buildings
Cubic grid with horizontal resolution of 1 m. Higher resolution is enabled only
for the vertical axis
Empirical initial boundary conditions, found by trial and error, in order to get
good agreement with average measurement data
Constant wind profile during all simulation times
Buildings have constant indoor temperature and no heat storage
1D soil model considering a ve level profile of humidity and temperature
Vegetation model considering the photosynthesis rate, the CO2 demand, and the
state of the stomata, the interaction of humidity and radiation in soil and air

different design congurations according to the Physiologically


Equivalent Temperature (PET) thermal comfort index [45]. PET has
been repeatedly considered an appropriate thermal index for
evaluation of outdoor thermal comfort [48]. Accordingly, thermal
comfort was analyzed based on the thermal performance classication (TPC) for (sub) tropical regions (Table 3). Accordingly, thermal comfort range for (sub) tropics lies in the PET values between
26  C and 30  C representing the neutral perception, while a wider
range of thermal comfort is dened for PET values between 22 and
34  C.
3.2. Selection of a typical day
Based on the Malaysian Meteorological Department report [40],
Kuala Lumpur is generally exposed to variable winds, with four
monsoon seasons throughout the year: north-east monsoon season, south-west monsoon season and two intermediate monsoon
seasons. Since the impact of wind on the thermal performance of
outdoor spaces can be signicant, especially in humid climates, a
typical day is selected from the main monsoon season in the present study.
In Kuala Lumpur, the highest monthly global radiation can be
found between March and June [49]. Looking into the 2013 climate
data [40], the month of March had the highest records of the global
radiation, direct radiation, diffuse radiation, and dry-bulb temperatures. This month was also selected in previous studies in the
tropical context of Malaysia [2,16,50]. Accordingly, this study
selected the day of March 05, which had an averagely hot and
sunny day with low temperature of 23.8  C at 20:00 and high
temperature of 31.8  C at 14:00.
3.3. Characteristics of the courtyard model
The fundamental input parameters including the building, soil
and meteorological data incorporated into the ENVI-met model are
given in Table 4.
Table 3
Thermal perceptions classication for temperate region and (sub) tropical region (a:
[48]; b: [45]).
Thermal perception

PET (sub)tropical
regiona ( C)

PET for temperate


regionb ( C)

Very cold
Cold
Cool
Slightly cool
Neutral
Slightly warm
Warm
Hot
Very hot

<14
14e18
18e22
22e26
26e30
30e34
34e38
38e42
<42

<4
4e8
8e13
13e18
18e23
23e29
29e35
35e41
<41

158

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

simulation are subsequently compared with PET values obtained


from RayMan calculations to ensure consistency in the ndings.

Table 4
ENVI-met parameters as specied in the conguration le.
Simulations input parameters
Location

Simulation day
Simulation duration
Grid size
Soil data
Initial temperature, upper layer (0e20 cm)
Initial temperature, middle layer (20e50 cm)
Initial temperature, deep layer (>50 cm)
Relative humidity, upper layer (0e20 cm)
Relative humidity, middle layer (20e50 cm)
Relative humidity, deep layer (>50 cm)
Building data
Inside temperaturea
Heat transmission coefcient of walls
Heat transmission coefcient of roofs
Albedo walls
Albedo roofs
Meteorological data
Wind speed, 10 m above ground
Wind direction (0:N, 90:E, 180:S, 270:W)
Roughness length
Initial atmospheric temperature
Absolute humidity at 2500 m
Relative humidity at 2 m
Cloud cover
Physiological data
Walking speed
Mechanical factor
Heat transfer resistance cloths

Kuala Lumpur, Malaysia


(latitude 3 70 N and
longitude 101 330 E)
5th March (hot day)
14 h, from 6:00 am to
8:00 pm
80  80  30
[ C]
[ C]
[ C]
[%]
[%]
[%]

28
26
24
88
90
93

[ C]
[W m2 K-1]
[W m2 K-1]

20
1.7
2.2
0.3
0.15

[m s1]
[ ]
[m]
[K]
[g kg1]
[%]

1.1
60
0.1
302
8
87
0

[m s1]
[met]
[clo]

0
0
0.6

a
While the indoor temperature may seem low for a building in a hot and humid
climate, this represents a typical set point of air conditioning systems in Malaysia
[41].

3.4. Reliability of the software


Reliability of ENVI-met for simulating the thermal performance
of outdoor spaces has been repeatedly proven [14,31,41,51e55].
These studies demonstrated an agreement between measured
(from eld measurements or observed data at local meteorological
stations) and simulated air temperatures.
In order to calibrate the model used in this study, the measured
hourly average meteorological data of dry-bulb temperature in
Kuala Lumpur are compared with the predicted hourly data from
ENVI-met for the built conguration studied in Makaremi et al. [2].
This comparison conrmed the accuracy of simulation data and of
the weather le (Fig. 2). After minor adjustments, the comparison
shows high levels of correlation demonstrating an acceptable
agreement between the predicted values and real data of meteorological stations. Likewise, the PMV values derived from ENVI-met

4. Results and analysis


The study looked at different courtyard congurations. The
courtyards had an area of 576 m2, being 24 m width by 24 m length,
initially surrounded by a single story building with the dimension
of 60  60 m which was modeled according to the typical ofce
characteristics. This results presented in this section are divided
into the following subsections: courtyard orientation (Section 4.1),
height of wall enclosures (4.2), reectance of wall enclosures (4.3),
and presence of vegetation (4.4).
4.1. Effect of the orientation
It is important to take into consideration the orientation and
location during the early design stage of courtyards. Fig. 3 shows
ve courtyard congurations that have been studied: facing North,
South, East, West and having the courtyard central.
The impact of location and orientation of courtyards towards
inuencing the microclimate parameters including ambient temperature, relative humidity, wind speed, mean radiant temperature,
and change of temperature with time was considered mapping the
different parameters or focusing on the values assumed by previous
parameters in the center of the courtyard spaces.
Results in Fig. 4 indicate that in all ve congurations, the level
of temperature is considerably high, and temperature variations
follow similar patterns among the different orientations. However,
courtyards facing North and East have slightly lower temperatures,
particularly from 10:00 to 17:00, with approximately 0.5 of difference, while the highest temperatures were recorded in the
courtyard facing West, as this orientation has the lowest level of
wind speed. Courtyards facing North and East also have a higher
level of humidity compared to the other courtyards.
According to the distribution of air temperature through the
simulated models, and focusing on two critical hours (12:00 and
14:00), it is explicitly shown that the courtyards facing North and
East have lower temperatures also far from the center of the
courtyards (Figs. 5 and 6). This result is due to the fact that the
courtyards are open to the northeast wind directions and receive
slightly more shading (Fig. 7). Nevertheless, the variation pattern
and hourly value of Tmrt are similar in all cases.
Looking into the change of air temperature relative to time of
day, in all courtyard models, the temperature is constantly
increasing from 9:00 to 15:00, while it is decreasing from 16:00 to
20:00. The highest change in air temperature occurs at 10:00 for all
ve courtyards with an approximate value of 1.8  C.
Considering the outdoor thermal comfort and looking into the
PMV values at critical hours (12:00 and 14:00), it is evident that

Fig. 2. Comparison between measurements and simulated data in the context of Kuala Lumpur.

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

Fig. 3. Square-shaped courtyard models according to different orientations.

Fig. 4. Comparison of average of hourly ambient temperature (left) and relative humidity (right).

Fig. 5. Simulated spatial distributions of air temperature in the courtyard models at 12:00 at 2 m height.

159

160

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

Fig. 6. Simulated spatial distributions of air temperature in the courtyard models at 14:00 at 2 m height.

almost all courtyard spaces are thermally uncomfortable with PMV


values above 4 due to the high solar radiation intensity, long
duration of direct solar radiation, high level of temperature and
humidity, and lack of shading. Fig. 8 shows that the PMV values for
courtyards facing North and East are slightly lower; nevertheless,
also these courtyard spaces do not provide thermally comfortable
environments due to the excessive solar exposure. Focusing on the
courtyard facing North, which has a relatively better performance,
the PMV values indicated that people can only enjoy this courtyard
during morning hours before 10:00 and evening hours after 19:00.
Findings point out that without properly equipping courtyard
spaces with cooling strategies and if direct solar radiation is not
avoided, courtyards surrounded by short buildings are not
comfortable for hot and humid climates. Actually, the thermal
comfort condition could be even worse than the open areas

surrounding the building. This is mainly due to the fact that


courtyards lack the cooling effect from free breezes.
Although location and orientation of a courtyard lead to
different microclimate results, these differences are small and the
outdoor thermal comfort condition of courtyards in all ve scenarios is generally poor.

4.2. Increasing the height of wall enclosures


Three different heights of wall enclosures were simulated to
investigate changes in the ground-to-wall ratio of courtyards. The
height variations were made on the courtyard model facing North,
as it was the courtyards with slightly better thermal performance.
Three different design scenarios with heights of 4 m (1-storey e

Fig. 7. Comparison of average of wind speed (left) and mean radiant temperature (right).

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

161

Fig. 8. Simulated distributions of PMV in the courtyard models at 14:00, at 2 m height.

model 6), 12 m (3-storey e model 7) and 24 m (6-storey e model 8)


were developed in order to test the ratios 6:1, 2:1 and 1:1 (Fig. 9).
It is evident in Fig. 10 that due to having a clear sky (cloud
cover 0), solar radiation is very high during the noon time,
approximately 950 W/m2, as similarly reported by Berkovic et al.
[1] and Taleghani et al. [8] and therefore, blocking the direct radiation can be highly inuential in elevating comfort due to achieving
larger shaded areas. The simulation results indicate that increasing
the height of wall enclosures in courtyards decreases the air temperature. Accordingly, the temperature of the courtyard with the
highest wall enclosure (model 8, 6-storey) is 1 cooler compared to
the courtyard with the lowest height (model 6, 1-storey). Similarly,
the courtyard with the highest level of wall enclosure receives
signicantly lower direct solar radiation from 11:00 to 17:00
(Fig. 10). Overall, it is inferred that through increasing the heights,
sky view factor (SVF) is decreased and as a result, the longest
duration of receiving direct solar radiation is limited to the courtyard with the lowest height (and highest SVF).

Looking into the change of the air temperature with time, it is


found that temperature continuously increases from 9:00 to 15:00
for all courtyard models irrespective of the height of wall enclosures. Subsequently, the temperature constantly decreases from
16:00 onwards for all courtyard models. In the courtyard with the
highest height of wall enclosure (model 8), the hourly increased air
temperature reduces to approximately half with respect to the
other models. However, from 12:00 to 14:00 when the sun is above
the courtyard, the temperature change in courtyard ambient with
higher heights is larger. Comparing the wind speeds, it is found that
through increasing the height of wall enclosures in courtyards,
wind speed considerably decreases.
Calculating the average daily temperature and Tmrt reveals that
the courtyard with the highest height resulting in the ratio of 1:1
(model 8) has noticeably better thermal performance followed by
courtyards by ratios of 2:1 (model 7) and 6:1 (model 6). According
to the comparison with the new scenarios, changing patterns of
Tmrt represent the cooling benets of increasing height of wall

Fig. 9. Square-shaped courtyard models (models 6e8) according to height of wall enclosure.

162

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

Fig. 10. Comparison of average of hourly ambient temperature (above) and direct radiation (below) for different heights of the courtyard.

enclosure according to reduced Tmrt values particularly from 8:00


to 12:00 and from 15:00 to 19:00. This is primarily due to the potential of increased height towards expanding the shaded area
(Fig. 11).
In the courtyards with higher heights (3-storey and 6-storey),
high values of Tmrt only were obtained in the critical period from
12:00 to 15:00. High value can only be seen for the 1-storey
courtyard. This is due to its extremely poor performance as a result
of its ratio (1:6); in this low courtyard, surrounding walls do not
block the sun at all, but they do block the wind. Hence, due to its
continuous exposure to the sun with no possibility of shading plus
the reduced wind speed in the courtyard ambient, the high level of
Tmrt does not drop until late afternoon. Moreover, all these values
are obtained at the receptor located in the center of courtyard,
hence, even if the surrounding walls with low height partially block
the sun (walls facing east during morning and walls facing west
during afternoon), the center-point of courtyard is less affected.

Fig. 11. Mean radiant temperature for courtyards with different heights of wall
enclosure.

Taleghani et al. (2014) also looked at the thermal performance


characteristics of courtyards on a hot day in the Netherlands and
reported Tmrt value of 70e75  C for courtyards at 15:00 to 18:00.
Similarly, the study by Berkovic et al. [1] examined the thermal
performance characteristics of courtyards on a typical hot day in
Israel, and the ndings demonstrated high Tmrt values of 62  C at
15:00. Similarly, Shahidan [56] investigated the thermal conditions
in Malaysia and showed high average value of 50  C for mean
radiant temperature and 30.8e67.8  C for the maximum Tmrt range.
Resulted by Hisarligil showed a high Tmrt value of 66  C for the
center of a courtyard directly exposed to the sun in the temperate
climate of Turkey [36].
As a result, the PMV distribution in courtyards during the
morning time clearly shows that the courtyards with the lowest
height (model 6) is thermally uncomfortable even at 9 am, and its
thermal comfort condition is worse than the outdoor areas surrounding the building. On the other hand, the other two courtyards
(models 7 and 8) are thermally comfortable in the morning time
and their thermal comfort is signicantly improved compared to
the outdoor context. After 10 am, once the PMV distribution of
outdoor context is above 4, the courtyard with the height of 12 m
(model 7) is partially comfortable at its east side, while the courtyard with the height of 24 m (model 8) shows a high level of
thermal comfort (Fig. 12).
Fig. 12 shows that the PMV could be below 1 in models 7 and 8,
whereas it has high values in the courtyard model 6. Considering
the size of unshaded courtyard ambient for model 6 compared to its
surrounded walls with extremely low heights which are unable to
provide shading, high values of PMV, beyond the standard
maximum value of 4, are justied.
In a recent study, Makaremi et al. [2] showed that although high
values of PMV and PET were achieved for courtyards in Malaysia,
many of the responders still feel thermally comfortable given their
psychological adaptation. This is due to the difference between
human subjective assessment and objective measurements [57]. It
is, hence, important to clarify that respondents can be more
tolerant of the warm thermal environment due to their expectation
about the tropics, especially in extreme climates such as the
monsoon region in Malaysia.
Exploring the thermal performance of courtyards in critical
times of the day (12:00 and 14:00), it emerged that all courtyards are thermally uncomfortable during this time, since the
sun is high in the sky and hence, less blockage of direct solar
radiation could be carried out through wall enclosures. Nevertheless, according to PMV distributions, at 12:00, the east side of
courtyards with higher heights, particularly the courtyard with
the height of 24 m (model 8) has a better thermal comfort
condition while at 14:00, only a small portion of the west side of
the model with the highest height (model 8) has a better thermal
comfort condition.
This analysis reveals that increasing the height of the wall
enclosure is insufcient for ensuring improved thermal comfort
during the critical times of the day in the tropics.
Findings conrm the distinct effect of increasing the height of
wall enclosures in courtyards and decreasing the duration of
receiving direct excessive solar radiations. Nonetheless, according
to the PMV distributions in critical times of the day, thermal performance of courtyards is extremely poor. This highlights the need
to look into other strategies for providing comfortable spaces, such
as the application of shading devices or greeneries. Supporting
these ndings, the PET graphs explicitly indicate that increasing the
height of wall enclosures reduces the critical period of thermal
discomfort to 12:00 to 15:00, while the courtyard with the lowest
height continuously has a thermally uncomfortable ambient from
9:00 to 18:00 (Fig. 13).

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

163

Fig. 12. Simulated distributions of PMV in courtyard models 6 to 8 at 9:00 and 10:00, at 2 m height.

4.3. Increasing the albedo of wall enclosures


Increasing the albedo of facades has been proposed for reducing
the surface temperature and for cooling both indoor and outdoor
spaces. Cotana et al. [57] investigate albedo control strategies using
highly reective materials for improving the energy efciency of
buildings. Facades with higher albedo absorb less solar radiation
and reect the majority of it.
The study by Taleghani et al. [8] shows that increasing the albedo of facades of courtyards on relatively hot days in the
Netherlands increases the outdoor ambient temperature. This is
due to more solar radiation being reected to the center area of
courtyards, and a lower chance to dissipate the solar energy

received from building facades. Three scenarios were developed


based on increasing the albedo of wall enclosures: albedo values of
0.3 (model 9 e brick), 0.55 (model 10 e marble) and 0.93 (model 11
e highly reective plaster).
The results with different values of albedo show that from 7:00
to 8:00, a courtyard with a higher albedo has slightly higher air
temperature (approximately 0.25  C) and considerably higher
Tmrt (Fig. 14). Looking at the thermal performance at 13:00, it
emerges that although the courtyard ambient in all three cases
demonstrates a relatively high level of thermal discomfort according to the PMV distribution (Fig. 15), the thermal condition is
worsened in the courtyards with higher albedo. The PMV value of
the courtyard with an albedo of 0.3 is approximately 4 indicating a
discomfort, while an albedo of 0.55 results in an increased PMV
value of 5, whereas an albedo of 0.93 leads to a PMV value of 6.
Looking at a less critical time of the day (16:00), although direct
solar radiation is decreased in comparison to the noon time,
increasing the albedo still signicantly reduces thermal comfort. In
fact, the PMV values for albedo of 0.30, 0.55 and 0.93 were 2.5, 3.5
and 4.5 respectively. Similarly, the calculated PET values demonstrate that increasing the albedo values of wall surfaces results in a
signicant reduction of thermal comfort conditions. It is evident
that the PET values are above the wider range of thermal comfort
(22e34  C) in all courtyards during the critical period of 12:00 to
15:00, while a better thermal performance of courtyards with
lower albedo may be appreciated (Fig. 16).
4.4. Utilization of vegetation

Fig. 13. Variations of calculated PET values in courtyard models 6 to 8.

The study then considered the utilization of ve different congurations of grass and trees in the selected courtyard models. It is

164

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

Fig. 14. Hourly ambient air temperature (left) and mean radiant temperature (right).

Fig. 15. Simulated distributions of PMV in courtyard models 9 to 11 at 13:00 and 16:00, at 2 m height.

Fig. 16. Variations of calculated PET values in courtyard models 9 to 11.

known from previous studies that adding greeneries, particularly


trees, will improve the thermal comfort level in courtyards in
temperate according to larger shaded areas.
The thermal performance of courtyards with no greeneries
(model 12), 100% covered by grass (model 13), 25% covered by trees
(model 14), 50% covered by trees (model 15), and 75% covered by
trees (model 16) were hence evaluated (Fig. 17).
The results clearly show that increasing the use of trees in
courtyards directly results in an increased level of relative humidity. Using trees considerably reduces the level of ambient air temperature from 7:00 to 11:00 and from 17:00 onwards. The reduction
rate can reach up to three degrees at particular times of the day
such as 8:00 and 20:00. Nevertheless, it is evident that during the
critical time of the day (12:00 to 15:00) when the sun is almost
above the building and insufcient shading is achieved, the adoption of trees increases the ambient air temperature (Fig. 18). Indeed,
despite the undeniable benets of trees, there could be negative

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

165

Fig. 17. Square-shaped courtyard models (models 12e16) according to the utilization of greeneries (grass 50 cm aver, dense e tree 20 m aver, dense, no distinct crown).

Fig. 18. Comparison of average hourly air temperature (left) and relative humidity (right).

impacts in hot climates due to different reasons: trees can reduce


the wind speed [48]; and their canopies can block the net outgoing
long-wave radiations [44]. Looking at the thermal comfort conditions, it is inferred that increasing the coverage of trees generally
improves the thermal comfort zones and levels. However, coverage
of courtyard ground with grass does not have a signicant impact
on the thermal comfort. As shown in Fig. 19, the bare courtyard has
the highest PMV, and in the courtyard with the highest coverage of
trees, the area of thermal comfort zone is considerably increased.

Furthermore, looking into the PET graphs, it is clearly shown that


the courtyard with highest coverage of trees (Model 16e75%
coverage of trees) signicantly reduces the critical period of thermal discomfort to 13:00 to 14:00 (Fig. 20). Moreover, the represented PET values refer to receptors located at the center of the
courtyard where less shading was achieved through the surrounded walls and trees; on the contrary, once referring to other
zones of the courtyards, it was obtained a thermal comfort condition over the entire day.

Fig. 19. Simulated spatial distributions of PMV in courtyard models 12e16 at 12:00 and 15:00, at 2 m height.

166

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

Fig. 20. Variations of calculated PET values in courtyard models 12 to 16.

Fig. 21. Comparison of PET values derived from eld measurements and simulations
for the courtyard model 16.

Comparing the measured PET values in two areas of a courtyard


(space I extremely near to a multi-story building, and space II in the
central point of courtyard) [2] with the simulated PET values, it
emerged a good agreement. The variation of values follow a relatively similar slope, although courtyard model 16 generally overestimates the thermal comfort (Fig. 21).
Overall, ndings show that the number of hours per day that a
courtyard can generally be enjoyed by the users could be extended
to the entire day except the critical period of 12:00 to 15:00 (this
period is reduced to 12:00 to 13:00 in courtyard model 16). Likewise, during this critical period, particular areas of courtyard
adjacent to the surrounded buildings are still more comfortable
than the center area which is fully exposed to the sun.
Finally, it is evident that the evaluation of the thermal performance characteristics of courtyards during the early design stage
can ensure creating comfortable outdoor spaces throughout the
entire day, including the extremely hot hours of noontime.

5. Conclusions
Outdoor thermal comfort plays an important role in enhancing
urban life. Literature reports that courtyards may improve the
thermal comfort of outdoor spaces if they are properly designed.
Scientic studies focusing on the thermal behavior of courtyards
are predominantly observed in arid or temperate areas. Nevertheless, evaluating the micro-scale thermal environment in tropical
contexts was considered important. Accordingly, the thermal
comfort during typical sunny day conditions in the hot and humid
tropical context of Malaysia was studied in this research.

Findings present new insights to ameliorate the thermal comfort conditions of the built environment in tropical areas by
designing efcient courtyards. According to numerous simulations,
effective design options for the integration of courtyards are proposed. Despite recording some very hot and thermally uncomfortable conditions for unshaded courtyards, ndings present
evident examples of more acceptable comfort levels and cooling
potential based on the use of heat mitigation strategies in courtyards. As a result, the outdoor thermal comfort in enclosed or semienclosed courtyards in the hot humid climate of Malaysia could
improve through adequate attention to the design congurations.
Hence, although many studies present that courtyards are predominantly operative in hot and arid areas, they can also perform in
hot and humid climates.
A comparison between the models and their outdoor thermal
comfort situations helps clarify guidelines for future optimization
of the thermal performance characteristics of courtyards. Based on
the interpretations derived from simulations:
 It is generally very crucial to control the amount of incoming
solar radiation received by the ground in tropical contexts.
Meanwhile, it is important to express that in this study, simulations were run in clear sky conditions (cloud cover), and as a
result, SW direct solar radiations are very high, particularly
during the noon time (over 1000 W/m2 in March). Hence, it
should be noted that the high values of PMV and PET belong to a
sunny day with clear sky and high solar radiation intensity,
while in reality, sky in the tropics can be cloudy, which can
result in a lower level of PMV and PET;
 Proper selection of the location and orientation of courtyards
can lead to receiving maximized wind with higher speed plus an
increased amount of shade during the daytime, hence, it is an
important factor for ameliorating the outdoor thermal comfort.
Simulations reveal that the courtyard facing North has slightly
better thermal performance with minimum air temperature of
300 K at 8:00 and maximum air temperature of 305 K at 15:00.
Nevertheless, it is essential to consider the relatively weak and
variable wind velocity and direction in the context of analysis;
 Increasing the height of wall enclosures in courtyards signicantly improves the outdoor thermal comfort by blocking the
intense solar radiations and providing more shaded areas.
Findings conclude that increasing the height of courtyards
considerably decreases the level of Tmrt during the daytime with
an average decrease rate of 30  C, however, during the critical
period of 12:00 to 15:00, the Tmrt in the courtyard with the
lowest height is 2 lower than the rest. Furthermore, PET values
reveal that increasing the height of courtyard signicantly reduces the duration of thermal discomfort from 9 h (9:00 to
18:00) to 3 h (12:00 to 15:00);
 Increasing the albedo of wall enclosures in courtyards considerably reduces the outdoor thermal comfort, hence, the mean
radiant temperature in the courtyard with high albedo of 0.93 is
12 higher at 12:00 than the courtyard with albedo of 0.3. As a
result, higher level of PMV and PET belongs to the courtyard
with highest albedo;
 Use of vegetation such as grass for covering the courtyard provides a limited inuence towards improving the thermal comfort with a maximum air temperature decrease rate of 0.13 at
7:00 compared to the courtyard with bare ground;
 Use of trees in courtyards can enhance the overall thermal
comfort and it can reduce the unshaded areas that have a high
level of discomfort. Covering the courtyard with 75% trees leads
to the highest air temperature decrease rates of 3.3 and 2.5 at
20:00 and 8:00 respectively and an average increase rate of
10.5% for the relative humidity, compared to the courtyard with

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

bare ground. Nevertheless, use of trees in warm climates can


have negative effects on air temperature at particular times
through blocking wind, reducing the wind velocity, and
decreasing the nocturnal cooling. Accordingly, a courtyard
covered with 75% trees is approximately 1 warmer than the
bare courtyard during the period of 13:00 to 14:00. Nonetheless,
PMV distribution shows that during the critical period of 12:00
to 15:00 and despite the increased temperature of courtyards
with more greeneries, courtyards with the highest ratio of
greeneries have a higher thermal comfort than bare courtyards.
More importantly, it can be concluded that although courtyards
covered with greeneries from 0 to 50% provide thermally
comfortable conditions in all daytime except the period from
12:00 to 15:00, courtyards with highest coverage of greeneries
(75%) are capable of reducing the critical period of thermal
discomfort to 13:00 to 14:00;
 Comparing the ndings with previous studies, it is indicated
that with the proposed design congurations and heat mitigation strategies of this research, higher level of thermal comfort
may be achieved also in tropical climate;
 Courtyards in the hot and humid climate of Malaysia can provide
a thermally comfortable environment which can be enjoyed by
people only during the early morning period (7:00 to 10:00) and
the evening time (19:00 onward). This duration could be substantially increased once the courtyard's performance is optimized through proper analysis of design variants. It is important
to state that the focus of this research was on the unshaded
outdoor courtyards which are directly open to the sky and fully
exposed to solar radiations, and therefore providing high thermal comfort is hard to achieve.
It is also useful to remind some of the limits of the ENVI-met
software used: building temperature is xed and identical in all
the points of the building for all buildings; similarly, the transmittance and albedo values are xed and identical for all the supercies of the buildings; nally, anthropogenic heat and mass
(included water vapor) uxes such as trafc or air conditioning are
not considered.
Future research will look at the thermal conditions of shaded
courtyards based on exploring the variety of shading design options
and examining their effectiveness for further improving the outdoor thermal performance characteristics even during critical
hours of the day. This study was limited to the thermal performance
characteristics of courtyards; nevertheless, future research is
needed to look into the impacts of courtyard on the indoor thermal
conditions and energy performance of surrounding buildings.

References
[1] Berkovic S, Yezioro A, Bitan A. Study of thermal comfort in courtyards in a hot
arid climate. Sol Energy 2012;86:1173e86.
[2] Makaremi N, Salleh E, Jaafar MZ, GhaffarianHoseini A. Thermal comfort conditions of shaded outdoor spaces in hot and humid climate of Malaysia. Build
Environ 2012;48:7e14.
[3] Taleb H, Taleb D. Enhancing the thermal comfort on urban level in a desert
area: case study of Dubai. Urban For Urban Green 2014;13(2):253e60.
[4] Sailor DJ. A holistic view of the effects of urban heat island mitigation. In:
Lehmann S, editor. Low carbon cities: transforming urban systems. New York:
Routledge; 2014. p. 270e81.
[5] Wang Y, Akbari H. Development and application of thermal radiative power
for urban environmental evaluation. Sustain Cities Soc 2015;14:316e22.
[6] Lai D, Guo D, Hou Y, Lin C, Chen Q. Studies of outdoor thermal comfort in
northern China. Build Environ 2014;77:110e8.
[7] Muhaisen AS. Shading simulation of the courtyard form in different climatic
regions. Build Environ 2006;41(12):1731e41.
[8] Taleghani M, Tenpierik M, van den Dobbelsteen A, Sailor DJ. Heat in courtyards: a validated and calibrated parametric study of heat mitigation strategies for urban courtyards in the Netherlands. Sol Energy 2014b;103:108e24.

167

[9] Yang X, Li Y, Yang L. Predicting and understanding temporal 3D exterior


surface temperature distribution in an ideal courtyard. Build Environ 2012;57:
38e48.
[10] Safarzadeh H, Bahadori MN. Passive cooling effects of courtyards. Build Environ 2005;40(1):89e104.
[11] Muhaisen AS, Gadi MB. Shading performance of polygonal courtyard forms.
Build Environ 2006;41(8):1050e9.
[12] Shashua-Bar L, Pearlmutter D, Erell E. Microscale vegetation effects on outdoor thermal comfort in a hot-arid environment. In: Proc. of the seventh international conference on urban climate, Yokohama, Japan; 2009.
[13] Acosta I, Navarro J, Sendra JJ. Predictive method of the sky component in a
courtyard under overcast sky conditions. Sol Energy 2013;89:89e99.
[14] Taleghani M, Sailor DJ, Tenpierik M, van den Dobbelsteen A. Thermal
assessment of heat mitigation strategies: the case of Portland State University,
Oregon, USA. Build Environ 2014;73:138e50.
[15] Rajapaksha I, Nagai H, Okumiya M. A ventilated courtyard as a passive cooling
strategy in the warm humid tropics. Renew Energy 2003;28(11):1755e78.
[16] Sada N, Salleh E, Haw LC, Jaafar Z. Evaluating thermal effects of internal
courtyard in a tropical terrace house by computational simulation. Energy
Build 2011;43(4):887e93.
[17] Shahidan M. The potential optimum cooling effect of vegetation with ground
surface physical properties modication in mitigating the urban heat island
effect in Malaysia [Doctoral dissertation]. UK: Cardiff University; 2011.
[18] Khan N, Su Y, Riffat SB. A review on wind driven ventilation techniques. Energy Build 2008;40(8):1586e604.
[19] Al-Masri N, Abu-Hijleh B. Courtyard housing in midrise buildings: an environmental assessment in hot-arid climate. Renew Sustain Energy Rev
2012;16(4):1892e8.
[20] Berardi U, Wang T. Daylight in an atrium type high performance house. Build
Environ 2014;76:92e104.
[21] GhaffarianHoseini A, Berardi U, Dahlan ND, GhaffarianHoseini A. What can we
learn from Malay vernacular houses? Sustain Cities Soc 2014;13:157e70.
[22] Kubota T, Toe DHC, Ossen DR. Field investigation of indoor thermal environments in traditional Chinese shophouses with courtyards in Malacca.
J Asian Archit Build Eng 2014;13(1):247e54.
[23] Malacca town house. 2014. Available from: http://malaccatownhouse.
blogspot.com/p/melaka-town-house_2.html.
[24] Yasa E, Ok V. Evaluation of the effects of courtyard building shapes on solar
heat gains and energy efciency according to different climatic regions. Energy Build 2014;73:192e9.
[25] Aldawoud A. Thermal performance of courtyard buildings. Energy Build
2008;40(5):906e10.
[26] Meir IA, Pearlmutter D, Etzion Y. On the microclimatic behavior of two semienclosed attached courtyards in a hot dry region. Build Environ 1995;30:563e72.
[27] Sharples S, Bensalem R. Airow in courtyard and atrium buildings in the urban environment: a wind tunnel study. Sol Energy 2001;70(3):237e44.
[28] Ernest R. The role of multiple courtyards in the promotion of convective
cooling [PhD thesis]. UK: University of Nottingham; 2011.
[29] Al-Hemiddi NA, Al-Saud KAM. The effect of a ventilated interior courtyard on
the thermal performance of a house in hot-arid region. Renew Energy
2001;24:581e95.
n MA, Ganem C, Barea G, Llano JF. Courtyards as a passive strategy in
[30] Canto
semi dry areas. Assessment of summer energy and thermal conditions in a
refurbished school building. Renew Energy 2014;69:437e46.
[31] Yahia MW, Johansson E. Landscape interventions in improving thermal
comfort in the hot dry city of Damascus, Syria e the example of residential
spaces with detached buildings. Landsc Urban Plan 2014;125:1e16.
[32] Robitu M, Musy M, Inard C, Groleau D. Modeling the inuence of vegetation
and water pond on urban microclimate. Sol Energy 2006;80:435e47.
[33] Chatzidimitriou A, Yannas S. Microclimatic studies of urban open spaces in
Northern Greece. In: Proc. PLEA Eindhoven, 1; 2004. p. 83e8.
[34] Berardi U, GhaffarianHoseini A, GhaffarianHoseini A. State-of-the-art analysis
of the environmental benets of green roofs. Appl Energy 2014;115:411e28.
[35] La Roche P, Berardi U. Comfort and energy saving with active green roofs.
Energy Build 2014;82:492e504.
[36] Hisarligil H. Exploring the courtyard microclimate through an example of
Anatolian Seljuk Architecture: the thirteenth-century Sahabiye Madrassa in
jer M, Howlett RJ, Jain LC, editors. Sustainability
Kayseri. In: Hakansson A, Ho
in energy and buildings. Springer Berlin Heidelberg; 2013. p. 59e69.
[37] 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.
Archit Sci Rev 2013;56(4):297e305.
[38] Ratti C, Raydan D, Steemers K. Building form and environmental performance:
archetypes, analysis and an arid climate. Energy Build 2003;35:49e59.
[39] Yang X, Zhao L, Bruse M, Meng Q. Evaluation of a microclimate model for
predicting the thermal behavior of different ground surfaces. Build Environ
2013;60:93e104.
[40] Malaysian Meteorological Department (MMD). 2014. Available from: http://
www.met.gov.my.
[41] Qaid A, Ossen DR. Effect of asymmetrical street aspect ratios on microclimates
in hot, humid regions. Int J Biometeorol 2014 [in press].
[42] Bruse M. ENVI-met 3.1 beta. 2014. Available from: http://www.envi-met.com.
[43] ISO 7726. Ergonomics of the thermal environment e instrument for
measuring physical quantities. Geneva: ISO; 1998.

168

A. Ghaffarianhoseini et al. / Building and Environment 87 (2015) 154e168

[44] Ali-Toudert F, Mayer H. Numerical study on the effects of aspect ratio and
orientation of an urban street canyon on outdoor thermal comfort in hot and
dry climate. Build Environ 2006;41(2):94e108.
[45] Matzarakis A, Rutz F, Mayer H. Modelling radiation uxes in simple and
complex environmentsdapplication of the RayMan model. Int J Biometeorol
2007;51(4):323e34.
[46] Nikolopoulou M, Baker N, Steemers K. Thermal comfort in outdoor urban
spaces: understanding the human parameter. Sol energy 2001;70(3):227e35.
[47] Halawa E, van Hoof J, Soebarto V. The impacts of the thermal radiation eld on
thermal comfort, energy consumption and control e a critical overview.
Renew Sustain Energy Rev 2014;37:907e18.
[48] Lin TP, Matzarakis A, Hwang RL. Shading effect on long-term outdoor thermal
comfort. Build Environ 2010;45:213e21.
[49] Shahidan MF, Salleh E, Shariff KM. Effects of tree canopies on solar radiation
ltration in a tropical microclimatic environment. In: Proc. PLEA 2007, 18;
2007. p. 400e6.
[50] Nugroho AM, Ahmad MH, Ossen DR. A preliminary study of thermal comfort
in Malaysia's single storey terraced houses. J Asian Archit Build Eng 2007;6(1):
175e82.

[51] Johansson E. Inuence of urban geometry on outdoor thermal comfort in a hot


dry climate: a study in Fez, Morocco. Build Environ 2006;41:1326e38.
[52] Krger EL, Minella FO, Rasia F. Impact of urban geometry on outdoor thermal
comfort and air quality from eld measurements in Curitiba, Brazil. Build
Environ 2011;46(3):621e34.
[53] Chow WT, Brazel AJ. Assessing xeriscaping as a sustainable heat island mitigation approach for a desert city. Build Environ 2012;47:170e81.
[54] Srivanit M, Hokao K. Evaluating the cooling effects of greening for improving
the outdoor thermal environment at an institutional campus in the summer.
Build Environ 2013;66:158e72.
[55] Middel A, H
ab K, Brazel AJ, Martin CA, Guhathakurta S. Impact of urban form
and design on mid-afternoon microclimate in Phoenix Local Climate Zones.
Landsc Urban Plan 2014;122:16e28.
[56] Shahidan MF, Jones PJ, Gwilliam J, Salleh E. An evaluation of outdoor and
building environment cooling achieved through combination modication of
trees with ground materials. Build Environ 2012;58:245e57.
[57] Cotana F, Rossi F, Filipponi M, Coccia V, Pisello AL, Bonamente E, et al. Albedo
control as an effective strategy to tackle global warming: a case study. Appl
Energy 2014;130:641e7.

You might also like