You are on page 1of 28

This article was downloaded by: [Ira Irawati]

On: 27 January 2015, At: 19:23


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office:
Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Advances in Building Energy Research


Publication details, including instructions for authors and subscription
information:
http://www.tandfonline.com/loi/taer20

The Application of Urban Climate Research in


the Design of Cities
Evyatar Erell

The Desert Architecture and Urban Planning Group , The Jacob Blaustein
Institutes for Desert Research, Ben-Gurion University of the Negev , Israel
Phone: +972-8-6596875 E-mail: www.bgu.ac.il/CDAUP
Published online: 06 Jun 2011.

To cite this article: Evyatar Erell (2008) The Application of Urban Climate Research in the Design of Cities,
Advances in Building Energy Research, 2:1, 95-121
To link to this article: http://dx.doi.org/10.3763/aber.2008.0204

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content)
contained in the publications on our platform. However, Taylor & Francis, our agents, and our
licensors make no representations or warranties whatsoever as to the accuracy, completeness, or
suitability for any purpose of the Content. Any opinions and views expressed in this publication
are the opinions and views of the authors, and are not the views of or endorsed by Taylor &
Francis. The accuracy of the Content should not be relied upon and should be independently
verified with primary sources of information. Taylor and Francis shall not be liable for any
losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities
whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any substantial
or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or
distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use
can be found at http://www.tandfonline.com/page/terms-and-conditions

publishing for a sustainable future

The Application of Urban Climate


Research in the Design of Cities

Downloaded by [Ira Irawati] at 19:23 27 January 2015

Evyatar Erell

Abstract
In spite of the growing body of research on urban climatology and the increasing demand
for architects and urban planners to practise climate-conscious design, there is too little
evidence of the application of urban climatology in practice. This chapter explores the
relationship between climatologists and urban planners, seeking to establish some of
the reasons for this state of affairs. It then sets out a methodological framework for the
application of climatology in the planning process, outlining possible goals for such
intervention, as well as its limitations. The chapter then attempts to establish the effects
on the urban microclimate of a broad range of decisions taken routinely by architects
and planners, based on an extensive survey of applied research in the field.

Keywords urban microclimate; urban planning; architectural design

INTRODUCTION
The quality of life of millions of people living in cities can be improved if the factors
that affect the urban microclimate are understood and the form of the city responds to
them in a manner that is appropriate to its environment. Underlying this approach is
the idea that climatically responsive urban design is vital to any future notion of
sustainability: it enables individual buildings to make use of natural energy, it
enhances pedestrian comfort and activity in outdoor spaces, and it may encourage city
dwellers to moderate their dependence upon air conditioning in buildings and upon
private vehicles.
It has been known for well over a century that cities generate their own climate
(Howard, 1833). However, urban climate research has evolved as a specialist discipline
within meteorology and climatology only in the past 50 years or so. During this period, the
focus of research has shifted from descriptive studies of the properties of the urban wind
field, temperature and humidity, to an experimental approach designed to investigate the
physical processes responsible for the unique meteorological conditions found in cities.
Breaking down complex urban forms into basic components, notably the urban canyon,
has allowed researchers to isolate the effects of individual factors. Comprehensive

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121


doi:10.3763/aber.2008.0204 2008 Earthscan ISSN 1751-2549 (Print), 1756-2201 (Online) www.earthscanjournals.com

Downloaded by [Ira Irawati] at 19:23 27 January 2015

96

EVYATAR ERELL

reviews of progress in research on urban climatology, in general, may be found in Arnfield


(2003) and Grimmond (2006).
Architecture (the design of buildings) and urban planning (the design of the urban
context within which individual buildings are constructed) have traditionally responded to
climatic conditions, as, indeed, they have to other aspects of the location of towns. The
importance of environmental factors in the design of settlements was recognized at least
as early as Roman times, as is recorded in the writings of Vitruvius (1999), who warned
planners, for example, that if the streets run straight in the direction of the winds then
their constant blasts rush in and sweep the streets with great violence. More recently,
there have been numerous manuals describing modern methods of bio-climatic
architecture, many of which owe much to the pioneering work of Olgyay (1963).
Two modern inventions have diverted the attention of architects and allowed them,
for a time, to ignore the effects of climate: first, the automobile, which replaced human
beings as the focus of attention in the design of outdoor space; and, second,
advanced space heating and cooling systems, which provided comfortable and stable
interiors regardless of the vagaries of climate, and allowed architects to avoid some of
the consequences of designing buildings that showed no regard to local conditions.
Climatic considerations have nonetheless had a major effect on the plans of a number
of new neighbourhoods or towns during recent years (Gotz, 1982; Etzion, 1990; Evans and
de Schiller, 19901991; El-Shakhs, 1994). In all of these cases, planning decisions were
made primarily on the basis of an intuitive understanding of local conditions, and less on
the basis of scientific analysis of the meteorological conditions and likely urban effects. The
only variable treated in a quantitative manner in some modern neighbourhood plans is solar
access rights, which are based on geometric rules derived from apparent solar position.
Unfortunately, most publications on the application of urban climatology to the
planning process fall into one of two categories: they are either cases studies of conditions
in existing settlements (Potchter, 1988); or they provide only general recommendations, but
not detailed design tools (Landsberg, 1968; Golany, 1996). Borve (1982), Westerberg and
Glaumann (1990/1991) and Pressman (1996) offered general guidelines for so-called winter
cities; Aynsley and Gulson (1999) proposed design strategies for humid tropical cities, in
general; while Emmanuel (1995) suggested strategies specific to Colombo, Sri Lanka. Herz
(1988) drew up recommendations for the Sahel region.
The awareness of the importance of climatological inputs in the process of urban
planning is growing in recent years. However, even where environmental concerns are the
subject of much public debate and where urban planners are interested in climatic
aspects of design, the use of climatic information is not systematic and climatology has
little impact on the planning process (Eliasson, 2000). The reasons for this are not only
conceptual and knowledge based, but are also related to technical matters, policy issues,
organizational aspects and the market.

ON THE DIALOGUE BETWEEN URBAN PLANNERS


AND URBAN CLIMATOLOGISTS
Mills (1999) has suggested that the difficulty in applying research in urban climatology to
architectural design problems may be explained by the fact that despite the common

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

The Application of Urban Climate Research in the Design of Cities

97

interest in the urban climate, these fields pursue different research interests, employ
contrasting methodologies and present results differently. Each of these contrasts
deserves further elaboration.
Urban climatology now encompasses issues as diverse as the effect of street design
on pollution dispersion and the effect of cities on the hydrological balance; but a large
proportion of the studies has been devoted to the phenomenon of the urban heat island.
Although it has been demonstrated that the urban heat island occurs mostly on clear
nights with no wind conditions that are not necessarily frequent in many cities it has
been the focus of considerable research since it was proposed that its maximum intensity
was related to the population size of cities (Oke, 1973). The success of models devoted
to the urban street canyon (Nunez and Oke, 1977; Oke, 1981; Nakamura and Oke, 1988)
has led to its widespread use as the preferred unit of study irrespective of the fact that
most urban spaces do not conform to this simplified generic form. Architectural research,
meanwhile, has tended to focus mostly on daytime phenomena and on issues of human
thermal comfort.
Research by architects on urban climate is typically concerned with observing the
climatic behaviour of urban spaces, with the underlying assumption that successful
examples may then be examined to elucidate the fundamental physical characteristics that
are most responsible for creating the desirable conditions. It often relies on studies of
vernacular architecture, seeking to apply the distilled experience of previous generations to
modern-day situations (see, for example, Krishan, 1996). The emphasis on physical form
and material is in marked contrast to the interest of climatologists in studying the
processes and the fundamental principles that drive them, which may require isolation of
a process and its presentation in abstract terms. Meteorological models necessarily involve
simplification of the real world, and applying the insights gained by such methods to
planning in the complex reality of a city may therefore be quite difficult.
Climatic research by architects generally has a strong focus on practical application.
As Mills (1999) noted, the results of architectural research are often formulated in terms
of guidelines or methodologies for other designers. These are frequently demonstrated
with examples from the real world, which are presented as proof that the underlying
principles discovered through the research may be applied in practice.

A PARADIGM FOR CLIMATE-CONSCIOUS URBAN PLANNING


DESIGN GOALS
Oke (1988a) with mid-latitude cities in mind, proposed the following objectives for urban
planners considering an appropriate response to climate:

maximize
maximize
maximize
maximize

shelter;
dispersal of pollutants;
urban warmth;
solar access.

Considering the aspect ratio of streets as the only parameter to be modified in order to
achieve these objectives, Oke (1988a) suggested that a height-to-width ratio (H/W) of 0.4

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

98

EVYATAR ERELL

was a compromise that would lead to satisfactory performance with respect to all of
them. This implies a very low density compared to what is found in many existing urban
centres. However, several studies (Manins et al, 1998; Mills, 2006) have shown that a
compact city performs best on a number of measures, especially reduction of energy use
in transport implying a much higher density is desirable.
Okes (1998a) objectives were framed in a very narrow and specific framework. Page
(1972) discussed the effects of microclimate climate on a very broad range of issues
encompassed in the field of urban planning and design. These include optimization of
land-use patterns in relation to different activities to be carried out in the town;
identification and development of suitable microclimates for various activities, such as
parks or recreation; identification of adverse microclimatic factors likely to affect the
detailed design of urban systems, such as high local winds; optimization of building form
in relation to external climatic inputs, such as solar radiation and wind; optimization of
building form in relation to microclimatic modification of the immediate exterior domain
of the building, such as the high winds induced near ground level by tall buildings;
constructional safety, especially with respect to high winds; selection of appropriate
building materials; planning of the construction process itself in view of climatic
constraints; control of water runoff; assessment of building running costs (heating,
ventilation and air conditioning (HVAC), lighting, etc.) in advance of construction;
optimization of the operating environment of transport systems (e.g. avoidance of ice
hazards); and control of the environmental impact of a transport system on its adjacent
urban systems (e.g. with respect to air pollution by vehicles).
As Page (1968) demonstrated, urban microclimate may affect our lives in diverse
ways. To summarize this section, it may be useful to organize these effects into two main
categories:
1 The effect of microclimate on human activity, especially pedestrian, in the spaces
between buildings. The urban fabric consists of buildings and open space,
which may, in turn, be classified according to intended patterns of use. Where
pedestrian access is considered valuable, design of the outdoor spaces intended
for humans should provide optimal conditions, as appropriate in the local climatic
conditions.
2 The effect of microclimate on the performance of buildings, especially with respect
to energy conservation. The magnitude of modifications to the microclimate
resulting from the effects of the urban fabric has drawn the attention of researchers
to the need for tools to predict them and to devise design strategies to respond to
them (Taha, 1978; El-Sioufi, 1987; El Nahas, 1996). The impetus for some of the
research has been the proliferation of computer software for building thermal
analysis, which relies on meteorological data to predict interior conditions. While
many building simulation codes are now considered to be quite accurate, significant
errors may be introduced to the simulations as a result of weather data, which is
based on regional averages but which may not be representative of site-specific
conditions (Even and Williamson, 2006).

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

99

Downloaded by [Ira Irawati] at 19:23 27 January 2015

SUBSIDIARITY
Architects and planners must deal with a multitude of factors. Often, the demands of
different consultant experts introduce conflicting requirements upon the architect, so the
design of urban space frequently involves a process of optimization. It is thus of great value
to be able to establish the benefits of a particular approach in general terms without resorting
to a unique policy required to achieve the desired goal. Furthermore, if there is more than one
solution that may yield the required result, the preferred solution is one that may be applied
as late as possible in the planning process, and which thus has the least impact upon other
aspects of the design. This approach, of seeking the solution for a particular issue at the
lowest possible level of the planning process, may be termed subsidiarity.
The following example may serve to illustrate this principle. Solar access, especially
with respect to direct radiation, has generated much research interest (Gupta, 1984;
Littlefair, 1998; Bourbia and Awbi, 2004; Robinson and Stone, 2004; Bozonnet et al, 2005;
Ali-Toudert and Mayer, 2006). The reasons for the profusion of studies into this particular
aspect of the urban microclimate are self-evident: solar radiation is the driving force of all
climatic systems; it may be studied through the application of simple rules of geometry;
and its recommendations are likewise formulated in terms of geometrical restrictions on
building volumes or the proportions of streets. Guptas (1984) work is instructive: using
the rather basic computing tools of the time, he produced an excellent analysis of the
effects of solar radiation on urban geometry in hot climates. Yet, while acknowledging the
fact that external shading devices such as pergolas, awnings, etc. may be used to limit
the solar exposure of building openings thus undermining much of the rationale for
the analysis the study then proceeds to study the effects of structural shading (i.e. the
shadows cast by the basic form of the building) as the primary criterion for assessing the
relative merits of the different options. These include the choice of building form, building
height, street orientation and street height-to-width ratios all of which could be decided
upon prior to the study of window shading.

COMPLEXITY
The previous example illustrates another common shortcoming of some attempts to
apply the scientific method to urban planning. In order to analyse a particular question,
researchers often simplify the issue, studying its effects in isolation from other factors
that may be involved. This has clear advantages as far as analysis of the physical
processes is concerned; yet, great care must be used in the synthesis of research results
into an overall planning strategy that may be applied to a specific urban location. Thus,
deriving an optimal urban form on the basis of exposure to solar radiation risks
overlooking the effects of other factors, such as energy emitted by long-wave radiation or
heat exchange by convection. The formulation of design recommendations on the basis
of such research must also take care to very carefully define the goals to be achieved:
design for pedestrian comfort and for building energy savings may lead to contradictory
requirements.
The role of climatologists in a real-life design team also includes responsibility for
analysing climatic conditions in the urban area in question in order to identify the critical

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

100

EVYATAR ERELL

Downloaded by [Ira Irawati] at 19:23 27 January 2015

issues to be addressed by the proposed plan. Some cities may have only one clearly
defined period in which meteorological conditions introduce significant stress that
deserves the planners attention. Yet, many cities, such as New York, for instance,
experience cold winters and hot-humid summers, both conditions requiring careful
analysis with respect to possible planning intervention. Givoni (1989) presents such an
analysis for a number of different climates. The report deals with the definitions of human
thermal comfort, provides a general description of the relevant characteristics of the
urban climate, and discusses the effects of various planning features, housing types and
vegetation on the urban microclimate. Its broad scope is an indication of the complex
nature of the issue and the multifaceted response required in order to be of value in the
context of city planning.

ECONOMIC VIABILITY
The recommendations of urban climatologists with respect to city planning may often
have significant economic implications. Urban development is driven to a great degree by
economic considerations, and zoning regulations often reflect the desire of city planners
to attract investment by real-estate developers. Street width, for instance, is generally
determined by the requirements of vehicular access, while building height reflects the
desire to maximize the value of land. Thus, any recommendation concerning the heightto-width ratio of streets, which has a major influence on the canopy layer climate of cities,
may also have considerable economic implications. Any explanation for the relative lack
of success in implementing climate-related strategies in urban planning must therefore
consider the lack of a practical framework to assess their economic effects too.

CLEAR AND IMMEDIATE BENEFITS


In order for a particular recommendation based on urban climate considerations to be
adopted by planners, the proposed strategy should have clear and immediate benefits.
For example, the effect of urban climate modification on building energy consumption
could be estimated, taking into account the urban heat island, as has been demonstrated
from measured temperature data for Athens (Santamouris et al, 2001) and London
(Kolokotroni et al, 2006). To realize this aim, urban climatology is required to develop
sufficiently accurate and reliable predictive tools. In the absence of quantitative studies on
the effect of proposed designs upon climate, and on the basis of well-documented
evidence from other planning professions, decision-makers, in general, tend to
downgrade the importance of climatic considerations in urban planning.

COMPREHENSIVE APPROACH TO PROBLEM-SOLVING


In order to apply urban climatology effectively in the process of town planning, a
comprehensive approach must be adopted. Recommendations must be based on
analysis of all factors influencing the urban microclimate. Conclusions based upon the
study of one or even several factors in conjunction may be misleading if they fail to take
into account the effect of other significant processes. With the increase in computing
power, computerized modelling may be capable of providing accurate and comprehensive
analysis of the urban microclimate. Once such models become reliable and accurate

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

101

enough, they should be applied wherever possible to inform the decision-makers of the
microclimatic implications of urban planning strategies under consideration.
For the models to be useful, they must allow the study of the particular issues that are
foremost in the architects mind. In other words, they must be formulated so that the inputs
include parameters related directly to the architects decision-making process. The
following section illustrates the effect of planning decisions on the resulting microclimate.

Downloaded by [Ira Irawati] at 19:23 27 January 2015

THE EFFECT OF DESIGN DECISIONS ON THE URBAN MICROCLIMATE


Modern urban planning is a complex process, generally carried out by teams of
professionals from a number of fields. While the architect or town planner may have
overall responsibility for producing the plan and for coordinating the project, crucial inputs
are provided by consultants from other disciplines. The process is generally driven
primarily by economic forces in response to market demand for housing, retail space, etc.
The input of the urban climatologist is therefore but one of many that compete not only
for the architects attention, but also for appropriate consideration in the process of
optimization that takes place in urban planning.
Mills (2006) noted that while the meteorologically ideal settlement serves a useful
pedagogical purpose, it does not recognize planning realities where climate issues are
rarely a dominant concern. So, while a statement of desirable outcomes and the means
of achieving them may be a logical means of applying urban climatology, climatologists
seeking to inform the decision-making process are rarely in the position of generating a
plan. Rather, they may be required to respond to proposals made by other members of
the design team. In order to do so effectively, urban climatologists must be aware of the
effect that climatic considerations may have on the issues that the planners must resolve
when preparing a town plan. This section is therefore structured to deal with the effects
of various features of the urban form on microclimate from the perspective of an urban
planner.

URBAN DENSITY
The density of a city is generally determined by economic considerations, reflected in the
price and availability of land. It is also influenced, and in turn has an effect upon, the
overall form of the town. Architects and planners typically measure urban density by
means of the number of dwellings per unit area of the site (Knowles, 2003) or by the ratio
of the total built floor area to the area of the site, an index more suited to non-residential
development. Urban climatologists, on the other hand, refer to density by different
measures: the plan area density, which is the ratio of the buildings footprint to the total
area of the site; or the frontal area density, which is the ratio of its (windward) elevation
to the site area (Grimmond and Oke, 1999).
Density has a direct effect on the exposure of urban surfaces to direct solar radiation.
Such exposure may be considered beneficial in cold climates or detrimental in hot
conditions, where shade is considered desirable. However, whereas the latter objective is
fairly easy to achieve for example, by the addition of specialized shading elements such
as pergolas and blinds the former imposes stringent limitations on the overall built
volume that can be constructed in a given site, as well as upon its geometry. The so-called

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

102

EVYATAR ERELL

solar envelope has been the subject of numerous studies by architects, with the aim of
ensuring appropriate solar exposure of buildings (Knowles, 1981) or of open space
(Yezioro et al, 2006).
Solar exposure of buildings has been a primary concern of many planners in climates
with cold winters to allow for passive solar heating. It has sometimes been investigated
in isolation from other aspects of urban climatology because it only requires knowledge
of geometry. Littlefair (1998) surveyed a variety of graphic methods to establish solar
exposure, though a substantial number of computer-aided design (CAD) programmes
now perform the task automatically, given a geographical location and time of year.
However, the study raises the intriguing question of how to establish the criteria for solar
exposure. It proposes that rather than mandate solar access on the winter solstice, when
the sun is lowest, it may be preferable to define a heating season and aim to maximize
gains over the whole period. Meeting the first requirement in high-latitude locations
requires exceedingly large distances between adjacent buildings in order to provide what
may be marginal benefits because sunshine hours are short and insolation is limited.
Guaranteeing exposure when insolation levels are higher may fill a proportion of heating
requirements that is only slightly smaller, yet requires far less stringent geometric
limitations on building height.
In tropical climates, exposure to solar radiation is generally undesirable. Planners in
such locations are therefore concerned with creating urban geometries that maximize
shade. Narrow streets and a dense urban matrix sometimes are often recommended for
desert locations (Mazouz and Zerouala, 1998), although they restrict ventilation; but
designs for warm-humid locations must maximize airflow while providing shade in public
spaces too (Emmanuel, 1993), so tropical cities may be less dense.
In temperate climates, design for solar access reflects a desire to accommodate
sometimes contradicting requirements in response to seasonal weather and solar
exposure. Swaid (1992) proposed that operable screens be installed on building rooftops
to restrict solar access into street canyons when radiation is excessive, yet which are
capable of being folded away when full exposure is desirable. He simulated the effect of
adjustable screens on canyon air temperature, and reported significant differences among
the various configurations tested. Knowles (2003), too, suggested that the solar envelope
of a building should be adjustable, terming the zone that lies between the summer and
winter extremes the interstitium.
The effect of urban thermal mass on air temperature may be seen as being analogous
to the effect of thermal mass on interior temperatures of buildings (Ratti et al, 2003). Erell
and Williamson (2007) suggest that since the three-dimensional geometry of the city
results in much larger surface area compared with a flat rural site with the same plan area,
effective thermal mass is greater in the former, which tends to reduce the diurnal
temperature swing. Field experiments carried out in Adelaide, Australia, have shown that
the diurnal amplitude of air temperature in a deep street canyon was 60 to 70 per cent of
the rural one, and was the result not only of elevated night-time minima, but also of
reduced daytime maxima. This suggests that the dense urban structure often found in
traditional quarters of many desert cities reduces the diurnal temperature extremes
characteristic of such climates.

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

103

STREET ORIENTATION

Downloaded by [Ira Irawati] at 19:23 27 January 2015

The planning of streets is often the primary input in the master plan of a city. While it is
true that it is the architecture of the buildings that give a city character, it is the street
layout that determines its structure. In modern town planning, streets are generally
planned in response to the requirements of transport systems. Transport planners draw
up the appropriate road links in response to the density and type of expected land use in
accordance with national standards or accepted best practice.
The orientation of streets with respect to the path of the sun or to the prevailing winds
is now rarely considered during the design process, although its effect on microclimate
was recognized even in historical times: Vitruvius (1999), discussing orientation in
approximately 20 BC, wrote:
. . . the orientation of streets and lanes according to the regions of the heavens. This
process will be properly accomplished if, with foresight, the lanes are kept from facing
into the path of the prevailing winds. For if the winds are cold, they injure; if hot, they
corrupt; if moist, they are noxious.
Kenworthy (1985), discussing the regular street pattern of some ancient cities such as
Miletus (in Asia Minor), proposed that the opposite was true: promoting exposure to
regional winds on an urban scale has been an aim of city planners from ancient times.
What, then, is the orientation (relative to wind direction) that best promotes
ventilation? Kenworthy (1985) found that while streets oriented parallel to the prevailing
winds would appear to offer the least aerodynamic resistance, scale-model tests of
orthogonal block grids showed that the maximum wind speed at street level was, in fact,
measured when the wind blew at a small angle to the main axis. Bottema (1999) also
found that parallel flow resulted in lower roughness (z0) than normal flow, but noted that
long buildings aligned with the wind created flow channelling that actually impaired
ventilation and removal of pollutants because of reduced vertical mixing. If the street
width is less than twice building height, shelter is enhanced, but ventilation is reduced.
Several studies on the orientation of streets, usually in a grid scheme, have
recommended various orientations on the basis of exposure of buildings or of street
surfaces to direct solar radiation. Gupta (1984) found that in composite climates, an
eastwest street with continuous wall surfaces was the optimum configuration (from
building energy considerations), but in low-latitude locations, a northsouth street axis
gave buildings equal solar protection to an eastwest oriented street. A similar study
(Mills, 1997) compared the effect of building group configuration on the thermal stresses
affecting individual buildings on the basis of two measures: solar exposure (controlling
heating) and sky view factor (controlling cooling by long-wave radiation). The study
provides a useful insight into the effects of urban geometry; but recommendations for
different climate zones (defined by latitude) are, of necessity, too simplistic, not least
since they ignore the effects of convection on building energy needs.
Considering solar exposure of streets, rather than buildings, Knowles (1981) arrived at
different conclusions: a Spanish grid (in which streets are oriented at 45 to the cardinal
points of the compass) was found to be preferable to the so-called Jeffersonian grid

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

104

EVYATAR ERELL

(in which streets are oriented eastwest and northsouth). This is because in the Spanish
grid, part of the street is shaded while the opposite sidewalk is exposed to the sun all year
round, allowing pedestrians to choose between different conditions. The Jeffersonian
grid, in contrast, maximizes midday exposure on the northsouth streets in both summer
and winter, while eastwest streets will be in shadow most of the time in winter, unless
they are very wide, and exposed to the sun during summer. The effects of street
orientation on solar access in low-latitude desert cities were also modelled by Bourbia and
Awbi (2004) and Ali-Toudert and Mayer (2006), the latter emphasizing that the effects of
street orientation on thermal comfort should be considered with respect to the daily
patterns of use by pedestrians since solar access varies with time of day. Shashua-Bar and
Hoffman (2003) pointed out that extensive planting of trees minimizes the effect of street
orientation with respect to solar access. However, the question of street orientation
should be considered not only in relation to solar access (or protection), but also with
respect to wind direction, especially where lack of ventilation is a major problem (Ahmed,
2000).

STREET ASPECT RATIO


Practising architects typically refer to the height of building faades facing a street or to
the width of the street; but rarely, if at all, do they refer to the ratio between them, except
in a qualitative sense. Climatologists, on the other hand, have found it useful to study the
city by means of a simplified form referred to as an urban canyon: a semi-infinite street
with a rectangular cross-section bounded on both sides by continuous buildings of
equal height. This prototypical urban space is characterized by means of its aspect ratio,
or the ratio of building height (H) to the width of the street (W). The aspect ratio of a street
is one of the most important controls on microclimatic conditions in the street, affecting
the transfer of energy, mass and momentum occurring in the space between buildings.

Radiant exchange
The canyon aspect ratio affects radiant exchange in different ways during the daytime and
at night. Deep streets have a small sky view factor and therefore lose less heat at night
by long-wave radiation than shallower ones. Oke (1981) demonstrated how canyon
geometry could explain the formation of the nocturnal urban heat island (UHI), previously
thought to be linked to the population size of the city (Oke, 1973). Okes (1981) experiment
is supported by numerous field studies (Barring et al, 1985; Yamashita et al, 1986;
Runnalls and Oke, 1998; Goh and Chang, 1999; Livada et al, 2002; Chow and Roth, 2006;
Erell and Williamson, 2006) and has formed the basis for much subsequent research on
the UHI.
The effect of the canyon aspect ratio on air temperature at night is well understood:
the city cools down more slowly than rural areas and has higher minimum temperatures.
However, the processes that occur during the daytime are more complex. Deep street
canyons restrict access to direct solar radiation and provide shade at street level, but they
also act as radiation traps: solar radiation that might otherwise have been reflected back
to the atmosphere from high-albedo surfaces experiences multiple-reflections among
canyon surfaces. The overall balance of these opposing tendencies depends inter alia on

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

The Application of Urban Climate Research in the Design of Cities

105

the exact proportions of the street and on the thermal properties of the surfaces. Several
studies of air temperature in urban streets show that they are warmer than rural areas in
the daytime as well as at night (Santamouris, 1998; Giridharan et al, 2004), while others
demonstrate that air temperature in city streets may, in fact, be slightly cooler during the
day (Steinecke, 1999; Runnalls and Oke, 2000; Erell and Williamson, 2007).
The canyon aspect ratio also has an effect on the overall albedo of a city with a regular
geometry. Aida and Gotoh (1982), using a numerical model, found that the maximum solar
absorption in a mid-latitude city such as Tokyo occurs when the canyon width is
approximately twice the block width. Soler and Ruiz (1994) confirmed this finding in an
empirical study comparing the intensity of reflected radiation in satellite images of urban
and rural areas near Barcelona with terrestrial measurements. It should be noted that
existing high-albedo cities are characterized by either high density (where roof albedo
dominates) or very low density (where the albedo of the ground is most important). In
medium-density cities, multiple reflections among canyon facets emphasize the
contribution of walls in addition to the horizontal surfaces.
The effect of street canyon geometry on radiant exchange may have a great effect on
human thermal comfort, an issue often overlooked in modern street design. In hot
climates with high radiant loads, net radiant balance may be more important than
convective exchange, and the benefit to pedestrians of shade outweighs minor
modifications to air temperature that might be measured in the street (Pearlmutter et al,
1999, 2006).

Airflow
Although airflow in cities takes place in open spaces of all types and sizes for example,
backyards, parks and plazas, in addition to streets it is streets that have received the
most attention from both climatologists and designers addressing this issue (although for
different reasons).
Climatologists have found the two-dimensional street canyon a useful generic urban
form to model airflow, using the aspect ratio (H/W) as one of its primary geometric
descriptors. Oke (1987) classified urban airflow into three basic regimes isolated
roughness flow, wake interference flow and skimming flow identifying flow patterns
associated with certain ranges of street aspect ratios. Airflow in street canyons with
uniform building heights and an aspect ratio of approximately unity is characterized by a
lee vortex generated by the above-roof wind, and may display a corkscrew pattern as the
cross-canyon circular motion is combined with an along-canyon component (Hotchkiss
and Harlow, 1973; Yamartino and Wiegand, 1986). The effect of the canyon aspect ratio
on airflow patterns in street canyons that are either much shallower or much deeper, or
which have buildings of unequal height on either side (step up or step down with respect
to the wind), has been the subject of several recent studies (Baik and Kim, 1999; Baik et al,
2000). In deep canyons, two major counter-rotating vortices may occur in some conditions,
the lower one driven by the upper (So et al, 2005). The thermal flows generated by
unequal heating of canyon surfaces further complicate the picture (Kim and Baik, 2001;
Xie et al, 2005, 2006). Assimakopoulos et al (2006) report on field studies that show the
existence of intermittent vortices that cannot be modelled by quasi-steady state models.

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

106

EVYATAR ERELL

They note that existing computational fluid dynamic (CFD) models are still incapable of
fully reproducing the turbulent nature of airflow in reality, and cannot be used to predict
the evolution of airflow at a particular urban location for any length of time. The
agreement between predicted and actual airflow is especially poor when wind speed is
low less than 2m/s when ventilation might be difficult to provide (Dixon et al, 2006).
Large eddy simulation (LES), requiring very substantial computing resources, though
much more powerful than CFD, is also still not sufficiently developed to be applied as a
tool for generating design strategies for even simplified generic forms of the urban
structure.
The geometry of street canyons is of primary importance in the study of pollutant
dispersion, especially automobile exhaust fumes. Studies have been carried out using
scale models in wind tunnels (Cermak, 1995; Kastner-Klein and Plate, 1999; Al-Sallal et al,
2001), as well as numerical simulation, once computing power had developed sufficiently
to allow numerical simulation of airflow in the presence of arrays of bluff objects
representing buildings (Kim and Baik, 1998, 2004; Chan et al, 2001, 2003). The flow
regime in a canyon (and, hence, the effectiveness of flushing out pollutants) depends not
only upon the wind speed but also, to a great degree, upon the canyon aspect ratio
(So et al, 2005). As might be expected, deeper canyons are more likely to suffer from high
concentrations of pollutants than shallower ones, while irregular building height promotes
turbulence and therefore improves pollutant removal. The effect of roof geometry is also
important (Rafailidis, 1997; Kastner-Klein and Plate, 1999). Much of the pollutant removal
occurs at intersections, which are characterized by unique flow regimes (Chan et al, 2003;
Dobre et al, 2005).

NEIGHBOURHOOD AND BUILDING TYPOLOGY


A feature of urban development clearly visible to all city dwellers is building typology.
Many streets conform to the urban canyon model that has proved such a useful tool for
urban climatologists. Yet, many more do not. Gupta (1984) identified three basic building
forms, which may translate into quite different urban morphologies: pavilion, court and
street. Steemers et al (1997) found six generic urban forms: pavilions; slabs; terraces;
terrace courts; pavilion courts; and courts different combinations of which produced
distinct wind patterns in wind tunnel experiments.
The complexity of real cities, as opposed to simplified conceptual models composed
of generic building types and regular street patterns, makes analysis of the effect of
urban form on micro-scale environmental behaviour very difficult. Computerized imageprocessing techniques offer a method of dealing with such conditions by analysing urban
texture using a digital elevation model (DEM) of the site in question (Ratti and Richens,
1999, 2004). The DEM is a compact way of storing urban 3D information using a 2D
matrix of elevation values: each pixel represents the height of a building (or part of it) and
can be displayed in shades of grey as a digital image. The computer model may then be
analysed for any geometric characteristic desired, such as the sky view factor from the
ground at any point, the extent of shadowing or sunlight availability. The technique was
used in an analysis of building energy consumption and urban texture (Ratti et al, 2005)
using the LT model1 to calculate energy requirements and the DEM to differentiate

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

The Application of Urban Climate Research in the Design of Cities

107

between core sections of buildings, which require more energy to service, and perimeter
areas, which are more amenable to daylighting, ventilation and passive heating. A further
study using this technique suggested that in a hot-arid climate, courtyard buildings (of
certain proportions) were a more appropriate building type than the pavilion type on the
basis of their surface-to-volume ratio, shadow density, daylight distribution and sky view
factor (Ratti et al, 2003). The same study also suggests that courtyard buildings would
be less suitable for tropical climates, where a narrow temperature variation would not
benefit from the high thermal mass and reduced ventilation characteristic of this
configuration. Texture analysis with image processing can also be used in the study of
winds and dispersal of pollutants in the city (Ratti et al, 2006). The variation of wind
speed above the ground is affected by the morphology of the surface, which can be
described by means of the roughness length (z0) and the displacement height (zd). The
roughness length is, in turn, a function of the average height of the buildings and a
measure of their density, such as the frontal area density (for a given wind direction) or
the plan area density (Grimmond and Oke, 1999). These can be obtained easily from a
DEM and the aerodynamic roughness derived for the site being studied. Using this tool,
analysis of hypothetical arrangements of long buildings arranged in parallel rows showed
that counter to design intuition and common sense, aligning the streets parallel to the
wind resulted in poorer ventilation and slower pollutant removal than aligning them
perpendicular to the wind.

SIZE, TYPE AND LOCATION OF URBAN PARKS


Vegetation affects conditions in the city in a variety of ways. For instance, several studies
found that vegetation dry-precipitates and adsorbs pollutants and, by doing so, decreases
the mass of airborne gases and particulate matter (Raza et al, 19901991; Taha et al,
1997). However, the following section deals only with the effects of vegetation on the
energy balance and upon air temperature.
The effect of local parks in a non-homogeneous urban area has been the subject of
intense study, especially once it became clear that the microclimate of built-up areas
differed substantially from that of rural areas. The so-called park cool island (PCI), a
manifestation of the more general oasis effect, is the converse of the urban heat island
(UHI): empirical findings show that air temperature in moderate- to large-sized parks may
be substantially lower than temperature in surrounding built-up areas (Jauregi,
19901991; Kanda and Moriwaki, 1998), although there are significant variations among
different types of parks.
Landscaping, specifically the incorporation of planted areas within the urban fabric,
may reduce differences between natural terrain and the urban surface. Bonan (2000)
demonstrated in an empirical study of a town that the availability of water and the
resulting increase in evaporation was the main factor responsible for lower surface
temperature and air temperature above the lawns, rather than the mere presence of
plants. This helps to explain the finding that increasing the proportion of planted areas in
a city tends to reduce daytime maximum temperatures, but may often have little or no
effect on night-time minima (Urano et al, 1995). Emmanuel (1997) noted that the main
effect of trees is to reduce radiant exchange at the ground surface: this may decrease

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

108

EVYATAR ERELL

daytime maximum temperature, but would also restrict nocturnal cooling, actually leading
to higher minima.

Downloaded by [Ira Irawati] at 19:23 27 January 2015

Classification of urban parks


Spronken-Smith and Oke (1998) distinguished between park cool islands that are defined
by surface temperature, which may be quite large, and cool islands defined by air
temperature, where the effects of surface temperature variation are diluted by nearsurface turbulent mixing and advection by wind. During daytime, surface temperatures
are affected by the presence or absence of shade, by surface albedo, by water availability
and by the thermal properties of the soil. These properties govern the receipt of solar
radiation, its absorption and the role of evaporative cooling. At night the thermal
properties of surfaces and the radiative geometry are the major controls on cooling.
Urban parks vary substantially with respect to the above factors, and may be classified
according to the arrangement of vegetation (Spronken-Smith and Oke, 1998): grass;
grass with tree border; savannah (grass with isolated trees); garden; forest; and multi-use.
Park cool islands may develop either during the daytime or at night. However, a given
urban park will display a regular diurnal pattern, indicating that the formation of PCIs may
be the result of a number of mutually exclusive factors. Spronken-Smith and Oke (1999)
found that daytime PCIs formed as a result of the combined effects of soil moisture and
shading: trees shade the surface, while grass is typically cooler than most solid surfaces
during the daytime if it is well irrigated. The relative coolness of irrigated parks therefore
peaks in the afternoon (forest type) or early evening (garden, savannah and multi-use
types). However, trees also inhibit nocturnal long-wave radiative cooling by blocking off
part of the sky, while excess moisture increases the thermal capacity of the soil and slows
down surface cooling. Night-time PCIs therefore typically form in relatively dry urban parks
with a sparse tree cover. They are driven by long-wave radiative cooling (since the sky view
factor is close to unity), and since evaporative fluxes are generally weak at night,
evaporation does not play a significant role in the formation of this type of PCI. In such
parks, daytime temperatures may sometimes be higher than in neighbouring urban areas.
However, an edge effect exists that applies within distances of about 2.2 to 3.5 times the
height of the park border, and which results in weaker radiative cooling where the sky view
factor is reduced by the obstructing features, such as perimeter trees or buildings.

Effects of size
The extent of vegetated area required to produce measurable effects on air temperature
is of great interest to urban planners. Upmanis et al (1998) found that the magnitude of
the intra-urban temperature difference between parks and their urban surroundings at
night increased with increasing park size, although large differences were also found
within the parks and in the built-up areas. These were attributed to the degree of exposure
to the sky and, hence, to the intensity of radiant cooling. Saito et al (19901991), in
contrast, reported cool islands in clumps of vegetation less than 200m across, although
the effect of vegetation was limited to the planted area itself and was not felt at distances
as little as 20m away from the park edge. Ca et al (1998) reported on the basis of
measured data that an urban park 0.6km2 in area can reduce air temperature in a commercial

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

109

Downloaded by [Ira Irawati] at 19:23 27 January 2015

area 1km downwind by as much as 1.5C; but Shashua-Bar and Hoffman (2000), who
reported differences of up to 3C between air temperature in tree-shaded urban avenues
and nearby non-shaded reference points when wind speed was very low, noted that the
cooling effect declined at an exponential rate with increasing distance from the border of
the planted area, and vanished less than 100m away from the sites studied. Numerical
modelling (Bruse and Fleer, 1998) indicates that small parks of only tens of metres across
may create temperature differentials of 2C or more. However, the horizontal gradients of
air temperature represented by the models may be quite large, and their spatial patterns
shift constantly with wind speed and direction, as demonstrated by Upmanis and Chen
(1999) in Goteborg. A pronounced difference in air temperature is almost always observed
very close to the edge of the park, a phenomenon also noted by Santamouris (2001) in a
study carried out in Athens.

The effect of trees on urban air temperature


The presence of trees in the urban matrix may affect air temperature at a variety of spatial
scales, from individual streets (Shashua-Bar and Hoffman, 2000) to city-scale
modifications (Huang et al, 1987). However, the magnitude of this effect may depend
upon a variety of factors because the interaction between trees and other constituents of
the urban environment is so complex (Oke, 1989). Trees intercept not only solar radiation,
but also long-wave radiation from the ground, building surfaces and the sky. The
dissipation of this heat load depends upon the water balance and wind climate of the tree.
In the presence of unrestricted water, transpiration will cause substantial cooling.
However, water supply to the root system may be restricted; stomata may be physically
blocked by particulates; or the heat load may be excessively high, leading to closure of
the stomata. Furthermore, at night (in the absence of sunlight) there is no photosynthesis,
so the stomata are closed and the tree is not cooled by transpiration. The response of
trees to increased energy loading, which may occur when individual trees are planted in
extensive paved areas such as parking lots, for example, will vary with species, humidity
of the atmosphere and how much of the crown is exposed (Kjelgren and Montague,
1998). Species from hot or arid habitats may be tolerant of high temperatures or able to
dissipate heat with small leaves but the evapotranspiration rates from such trees may
accordingly be lower than those of broadleaf trees, and thus they have a smaller effect on
air temperature in their surroundings.
Grimmond et al (1996) found that in a neighbourhood with a relatively dense tree cover,
the latent heat flux increased as a fraction of available energy compared to an otherwise
similar neighbourhood with a sparse tree cover, as did the storage flux, whereas the
sensible heat flux decreased. However, in absolute terms, all fluxes, including the sensible
heat flux, were enhanced at this neighbourhood. Trees and shrubs lowered the albedo and
surface temperatures, thereby reducing the loss of solar and long-wave radiation,
respectively resulting in an increase in the overall amount of energy to be dissipated
compared to the sparsely planted neighbourhood. As a result of the difference in fluxes,
the maximum air temperature above the canopy was about 1C higher in the densely
planted neighbourhood, while temperatures below the trees at noon and in the early
afternoon were similar or up to 0.5C higher than in the sparsely planted neighbourhood.

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

110

EVYATAR ERELL

In contrast to Grimmond et al (1996), most researchers report that trees reduce urban
air temperature. This is usually attributed to evapotranspiration; but Shashua-Bar and
Hoffman (2003) suggested that the cooling is due almost entirely to shading, which
more than offsets the exchange of sensible heat between the tree canopy and the air.
The actual reduction in air temperature might be overstated in many cases due to the
difficulty of measuring air temperature accurately in the presence of strong radiant loads
(Erell et al, 2005).
A model specifically designed to predict the general microclimatic effect on energy
consumption of augmenting urban vegetation was proposed by Sailor (1998). The
meso-scale model (2 x 2km grid) predicts that by increasing the vegetated fraction of the
core of a hypothetical city by 0.065, annual cooling loads could be reduced by 3 to 5 per
cent simply by lowering the background air temperature. Shashua-Bar and Hoffman (2002)
proposed a quantitative model for predicting the cooling effect of trees in an urban
setting; but their method incorporates site-specific empirical factors to account for
convective exchange, the magnitude of which is not systematic and for which there is no
method of calculation. The method is, nonetheless, demonstrated in an empirical study
of the effect of trees on air temperature in the streets of Tel Aviv, Israel, which found that
the overall cooling effect of trees could be as high as 3K, depending, in addition to the
shade coverage, upon the depth of the urban canyon, the albedo of canyon walls and
street orientation (Shashua-Bar and Hoffman, 2004).

The effects of vegetation on building energy consumption


Landscaping and careful planting of vegetation near buildings have been credited with
energy savings of anything up to 80 per cent in hot, dry climates (Meier, 19901991).
Several studies have been carried out to quantify the effect of vegetation, especially trees,
on the energy consumption of buildings (Simpson and McPherson, 1998; Simpson,
2002); but much of the evidence remains anecdotal.
The mechanisms by which vegetation affects the energy exchange between buildings
and the environment may be summarized as follows:
Vegetation can reduce energy consumption in buildings in hot climates if air

temperature is reduced near the planted area. However, it should be noted in this
context that heat transfer through building walls is driven by differences in surface
temperature, rather than by air temperature. Furthermore, the reduction in air
temperature resulting from evapotranspiration is accompanied by an increase in the
vapour content of the air. Therefore, the air-conditioning system must deal with an
increased latent heat load, offsetting, to some extent, any gains from a lower
sensible heat load.
Plants may shade building surfaces, reducing the radiant load on the envelope. This
may be beneficial in hot climates, but detrimental in cold ones. In temperate climates
with distinct heating and cooling seasons, deciduous trees are often planted, and
vine-covered trellises are common in many Mediterranean areas. However, the timing
of defoliation and the permeability of the bare trees vary widely from species to
species (Canton et al, 1994) and may not match the desired pattern of exposure to

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

The Application of Urban Climate Research in the Design of Cities

111

the sun. McPherson et al (1988) found that in the middle latitudes, cooling loads
were most sensitive to shading on the roof and on the west wall, while heating loads
were affected most by exposure of the south (equator-facing) and east walls. The
direct effect of shading building surfaces by plants was studied by Papadakis et al
(2001). Thick foliage producing a full shade effect resulted in a reduction of the
surface temperature of a light-coloured concrete wall by up to 8C, with concomitant
reductions in heat flux through the surface. When wind speed was negligible, air
moisture content in the planted canopy was up to 2g/m3 greater than in the
surrounding air. Likewise, a dense growth of ivy can block radiant exchange at the
wall surface almost entirely (Hoyano, 1988). However, several studies show that
where the shaded area has a limited spatial dimension for instance, beneath a
pergola (Hoyano, 1988) or in the shade of a liman (small clump of trees in an artificial
floodplain in the desert) (Schiller and Karschon, 1974) the effect on air temperature
at a height of 1m above the ground is negligible.
Plants may reduce wind speed near buildings, limiting unwanted infiltration, but also
restricting ventilation and reducing convective exchange at building surfaces. The
first two mechanisms are self-explanatory; but the third has less well-known
consequences. For instance, in hot climates, wind is an asset for unshaded houses
because it helps to remove radiant heat at the external building surfaces (McPherson
et al, 1988), reducing temperature differentials between the interior and exterior.
However, in poorly insulated houses, especially in cold climates, increased
convective exchange at the building envelope results in increased loads on building
heating or cooling systems.
Plants in warm climates may reduce temperatures of ground surfaces through
evapotranspiration (although planted surfaces may be warmer than bare soil in cold
climates), with two effects: cooler surfaces emit less infrared radiation, thus reducing
the radiant load on building surfaces; and they release less sensible heat to the
adjacent air, so that buildings are exposed to cooler ambient air.
Roof gardens (or planted roofs) are perhaps the most obvious example of the use of
plants to control building energy performance, and are sometimes credited with
improving the urban microclimate as well (Wong et al, 2003). The shading and
evapotranspiration of the plants contribute to lower surface temperatures and thus to
lower heat gains through the roof. Experiments show that the surface temperature of
an exposed roof can be reduced substantially by the addition of an irrigated lawn on
a fabric matrix (Onmura et al, 2001). A complete thermal model of a planted roof
(Palomo Del Barrio, 1998) showed that the contribution of the planting to the thermal
performance of the roof depends mainly upon the density of the foliage, the
composition, density and thickness of the substrate, and its water content. However,
unless the thermal conductivity of the soil is particularly low or the thickness of the
substrate is considerable, the thermal resistance provided by the planting and
substrate is usually insufficient during the cold season, even in mid-latitude countries
with relatively mild winters. However, much of the work in this field remains
empirical and lacks a comprehensive theoretical framework at the urban scale
(Niachou et al, 2001).

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

112

EVYATAR ERELL

Downloaded by [Ira Irawati] at 19:23 27 January 2015

BUILDING AND PAVING MATERIALS


Roofing materials, wall finishes and paving blocks are typically specified by architects for
a variety of reasons, such as cost, durability and appearance. Their thermal properties,
especially insulation, are usually assessed in the context of the energy budget of the
building. However, the properties of the materials that form the surfaces that make up the
urban canopy also have an effect on urban microclimate, which may be modified through
the selection of suitable finishes.
The role of materials in the formation of the urban heat island has been the subject of
some conjecture. Oke (1981) demonstrated that the nocturnal urban heat island could be
explained not only by the effects of street canyon geometry on long-wave radiant
exchange, but also by differences in the thermal properties of urban and rural surfaces.
However, he also noted that, in reality, differences in thermal admittance between urban
and rural materials were too small and were, in any case, affected to a great degree by
rural soil moisture. A later study (Oke, 1982) suggested that differences in thermal inertia
between rural and urban sites could be accounted for by considering the role of increased
surface area in the city, as well as differences in moisture availability. However, the
difficulty of measuring the storage flux directly in situ on a neighbourhood scale has thus
far limited efforts to explain this mechanism satisfactorily.
The effect of the thermal properties of the substrate on the surface temperature and
on the magnitude of the energy fluxes was investigated by Asaeda et al (1996), who
demonstrated a substantial difference in the behaviour of asphalt and concrete (painted
black to create similar albedo). The former has a lower thermal conductivity, and the
radiant energy it absorbs at the surface during the daytime results in a higher temperature
and therefore greater emission of long-wave radiation and higher rate of sensible heat
loss. Conversely, an asphalt surface also cools down more quickly than concrete at night.
Buildings in the warmer parts of the Mediterranean are often whitewashed to reduce
the external radiant load on the buildings and, hence, their surface temperature during the
daytime. The same strategy has been proposed to reduce convective heat transfer from
pavements and buildings to the air, although the temperature of air near the surface is also
affected by several additional factors. Paving materials with a variety of colours and
textures have been investigated by several researchers to evaluate the effect of exposure
to intense sunlight on surface temperature (Bretz et al, 1998; Doulos et al, 2004; Synnefa
et al, 2006). The studies found that during the daytime, all paving materials had a mean
temperature higher than that of the ambient air, while at night all surfaces were cooler, in
some cases by almost 6C (Synnefa et al, 2006). However, differences among the samples
are illuminating: compared to an exposed concrete tile serving as the reference, Synnefa
et al (2006) found that some of the coatings tested resulted in mean daytime
temperatures lower by approximately 4C to 5C, with a maximum difference in excess of
7C, while Doulos et al (2004) found that the difference between the coolest surface
(white marble) and the warmest one (asphalt) was over 22C.
The albedo of roof surfaces may have less effect on air temperature in the canopy
layer than the properties of the pavement or the ground because it affects heat transfer
above roof level. However, since it affects the energy performance of a house directly, it
has been the subject of much research. The obvious strategy with respect to roofing

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

Downloaded by [Ira Irawati] at 19:23 27 January 2015

The Application of Urban Climate Research in the Design of Cities

113

materials in hot climates is to employ light-coloured materials. However, progress in the


production of wavelength-selective paints now allows other colours to be specified,
which although visually fairly dark, nevertheless have relatively high albedo (Levinson et al,
2007; Synnefa et al, 2007). The so-called cool coatings can reduce the temperature of
roof tiles having the same appearance (i.e. colour) by between 1.5C and 10C (for green
and black tiles, respectively). Some of the coatings have been tested to evaluate the
long-term effects of exposure to ultraviolet (UV) radiation and mechanical damage, and
show relatively little deterioration over extended periods (Bretz and Akbari, 1997; Synnefa
et al, 2007). Some of the reduction in reflectance is attributed to soiling and can therefore
be reversed by rinsing.
High surface temperatures contribute directly to pedestrian discomfort by imposing a
long-wave radiant load (Pearlmutter et al, 2006, 2007), and may also affect air temperature
in close proximity. However, the thermal properties of urban surfaces may also have an
effect on the climate of a city as a whole: if roof albedo is modified on a large proportion
of city roofs, the cumulative effect on the climate of a city may be significant. Using a onepoint model of Sacramento, California, Taha et al (1988) suggested that maximum air
temperature (in the early afternoon on a summer day) could be reduced by about 4C if
urban albedo was increased from 0.25 to 0.4, while reducing urban albedo to only 0.1
would result in an increase of more than 4C relative to the base case. Taha et al (1997)
found, using the Colorado State University Mesoscale Model (CSUMM), that feasible
albedo changes could lower daytime maximum temperatures in central Los Angeles by
up to 2C, with smaller reductions predicted for the suburbs. The actual albedo of most
urban areas varies from about 0.1 to about 0.2 (Taha, 1997; Sailor and Fan, 2002), with a
mean value of about 0.14 for urban centres (Oke, 1988b), so there appears to be
substantial potential for cooling if the large-scale application of high-albedo surfaces can,
in fact, be attained (Bretz et al, 1998). The scale of modification possible depends upon
the area of solid surfaces, such as rooftops, streets and other paved areas as a proportion
of the total urban area. However, the micro-scale intra-urban variation of air temperature
in the urban canopy, observed in several field studies (Eliasson, 1996; Erell and
Williamson, 2007), means that the benefits of such a strategy may not be felt uniformly
over the entire city.
Extensive glazed areas also affect radiant exchange in a street canyon. While the use
of mirror glass results in extremely high albedo, sunlight transmitted to building interiors
through transparent windows creates urban surfaces that have extremely low albedo.
Tsangrassoulis and Santamouris (2003) showed that increasing the window-to-wall ratio
resulted in a lower canyon albedo because sunlight is transmitted by the glass to building
interiors. In very deep canyons (H/W > 2), the predicted albedo for all sun angles as well
as for uniform diffuse radiation was between 0.01 to 0.03 meaning that practically all of
the incident solar radiation is absorbed. Deep streets with extensive glazed areas in
adjacent buildings may therefore be viewed as a very efficient trap for solar radiation.

DISCUSSION AND CONCLUSIONS


As the above survey has demonstrated, there is a substantial body of knowledge on the
effects upon microclimate of a wide variety of planning decisions. However, it is equally

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

114

EVYATAR ERELL

evident that very little of this knowledge is being applied in practice, except in a
haphazard, piecemeal way.
Page (1968) identified three reasons why scientific information available to
researchers might, nonetheless, be rejected by practitioners:

Downloaded by [Ira Irawati] at 19:23 27 January 2015

1 The potential user considers the information irrelevant.


2 The user considers the information provided inapplicable in the form presented.
3 The user considers the information provided incomprehensible.
Page (1968) also noted that in order to be of use to urban planners, urban climatology
must first be predictive descriptive science is not in itself sufficient. Furthermore, the
problem is not to produce an idealized climatological plan, but to produce a workable
evolutionary plan that is economically viable and accepts that the planner must consider
other factors, such as the requirements of transportation systems.
Bitan (1988) noted that climatology affected urban design at a variety of spatial scales,
and that failure to address the effects at each of them, in turn, would necessarily restrict
the effectiveness of any intervention. However, in reality, there are very few projects
where climatologists take part in ongoing design processes over many years, beginning
with environmental impact assessments at the stage of site selection for large urban
developments and continuing throughout the entire planning process. The number and
hierarchy of authorities that are involved in the planning and regulatory approval of such
projects require a commitment that is rarely found in practice.
Oke (1984) suggested that what climatologists require is an ability to demonstrate the
importance of climatic information in the design of a settlement, and the predictive power
to foretell the climatic impact of alternative design strategies. According to Oke (1984),
several shortcomings in microclimate research at that time prevented it from becoming
an applied science, including a lack of quantitative techniques and relationships; lack of
standardization, generality and transferability; and the absence of clear guidelines for
those wishing to learn and use climatological principles in settlement planning.
Advances in addressing any of the following issues would represent real progress in
applying urban climatology in practical urban design:
What are the urban forms that will provide the best microclimate for any given

density, taking into consideration meso-scale climate, latitude, etc.?


Can street orientation be used as a policy guideline to achieve desirable microclimatic

conditions in a city?
Can the aspect ratio of streets be used as a policy guideline to achieve desirable

microclimatic conditions in a city? What are the quantitative limits that may be
usefully applied in realistic planning situations in high-density cities with regard to
specific goals of solar access and airflow?
What are the most appropriate neighbourhood layouts for every climate type? Are
some building typologies better at providing good microclimates than others?
What are the real effects of vegetation in various urban contexts? There seems to be
a consensus that plants improve the microclimate of cities; but numerous

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

115

Downloaded by [Ira Irawati] at 19:23 27 January 2015

quantitative studies have produced a wide range of estimates regarding the


magnitude and spatial extent of the effect of urban parks.
What are the city-scale modifications to microclimate that can be achieved by
implementing particular types of building and paving materials? The benefits to
individual buildings may be clear; but although models suggest that the compound
effect of enforcing their use all across an urban area may be substantial, the
magnitude of even local-scale effects is still open to some question.
Can urban climatology provide detailed site-specific weather data modified from
regional weather stations to allow more accurate simulation of building energy
performance, including not only air temperature, but also humidity and wind
characteristics?
The answers to these questions will probably only come about through research
collaboration between planners and urban climatologists.

AUTHOR CONTACT DETAILS


Associate Professor Evyatar Erell: The Desert Architecture and Urban Planning Group, The Jacob Blaustein
Institutes for Desert Research, Ben-Gurion University of the Negev, Israel; tel +972-8-6596875; www.bgu.ac.il/
CDAUP; E-mail: erell@bgu.ac.il

NOTE
1 LT is the name Baker and Steemers (2000) gave to a computer model dealing with lighting and
thermal properties of buildings.

REFERENCES
Ahmed, K. S. (2000) Problems and prospects of bioclimatic urban design: An analytical study of shading and ventilation
needs of the urban spaces in the humid tropics, PLEA 2000: Architecture City Environment, Cambridge, UK
Aida, M. and K. Gotoh (1982) Urban albedo as a function of the urban structure a two-dimensional numerical simulation,
Boundary Layer Meteorology, vol 23, pp415424
Al-Sallal, K. A., M. M. AboulNaga, and A. M. Alteraifi (2001) Impact of urban spaces and building height on airflow
distribution: Wind tunnel testing of an urban setting prototype in Abu Dhabi City, Architectural Science Review vol 44,
no 3, pp227232
Ali-Toudert, F. and H. Mayer (2006) Numerical study on the effects of aspect ratio and orientation of an urban street canyon
on outdoor thermal comfort in hot and dry climate, Building and Environment, vol 41, pp94108
Arnfield, J. (2003) Two decades of urban climate research: A review of turbulence, exchanges of energy and water, and the
urban heat island, International Journal of Climatology, vol 23, pp126
Asaeda, T., V. T. Ca and A. Wake (1996) Heat storage of pavement and its effect on the lower atmosphere, Atmospheric
Environment, vol 30, no 3, pp413427
Assimakopoulos, D. N., C. Georgakis and M. Santamouris (2006) Experimental validation of a computational fluid dynamics
code to predict the wind speed in street canyons for passive cooling purposes, Solar Energy, vol 80, pp423434

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

116

EVYATAR ERELL

Aynsley, R. and L. Gulson (1999) Microclimate and urban planning in the humid tropics, Planning in the Hothouse 27th
National Congress of the Royal Australian Planning Institute, Darwin, Australia
Baik, J. and J. Kim (1999) A numerical study of flow and pollutant dispersion characteristics in urban street canyons, Journal
of Applied Meteorology, vol 38, no 11, pp15761589
Baik, J., R, Park, H. Chun and J. Kim, J. (2000) A laboratory model of urban street-canyon flows, Journal of Applied
Meteorology, vol 39, no 9, pp15921600
Baker, N. and Steemers, K. (2000) Energy and Environment in Architecture: A Technical Design Guide, E. and F. Spon, London
Barring, L., J. O. Mattsson and S. Lindqvist (1985) Canyon geometry, street temperatures and urban heat island in Malmo,
Sweden, Journal of Climatology, vol 5, pp433444
Bitan, A. (1988) The methodology of applied climatology in planning and building, Energy and Buildings, vol 11, pp111
Bonan, G. B. (2000) The microclimates of a suburban Colorado (USA) landscape and implications for planning and design,

Downloaded by [Ira Irawati] at 19:23 27 January 2015

Landscape and Urban Planning, vol 49, pp97114


Borve, A. B. (1982) Settlement planning under Arctic conditions in northern Norway, Energy and Buildings, vol 4, pp6770
Bottema, M. (1999) Towards rules of thumb for wind comfort and air quality, Atmospheric Environment, vol 33,
pp40094017
Bourbia, F. and H. B. Awbi (2004) Building cluster and shading in urban canyon for hot dry climate, Part 2: Shading
simulations, Renewable Energy, vol 29, pp291301
Bozonnet, E., R. Belarbi and F. Allard (2005) Modelling solar effects on the heat and mass transfer in a street canyon, a
simplified approach, Solar Energy, vol 79, pp1024
Bretz, S. and H. Akbari (1997) Long-term performance of high-albedo roof coatings, Energy and Buildings, vol 25, pp159167
Bretz, S., H. Akbari and A. Rosenfeld (1998) Practical issues for using solar-reflective materials to mitigate urban heat
islands, Atmospheric Environment, vol 32, no 1, pp95101
Bruse, M. and H. Fleer (1998) Simulating surfaceplantair interactions inside urban environments with a three dimensional
numerical model, Environmental Modelling and Software, vol 13, pp373384
Ca, V. T., T. Asaeda and E. M. Abu (1998) Reductions in air conditioning energy caused by a nearby park, Energy and
Buildings, vol 29, pp8392
Canton, M. A., J. L. Cortegoso and C. de Rosa (1994) Solar permeability of urban trees in cities of western Argentina,
Energy and Buildings, vol 20, pp219230
Cermak, J. E. (1995) Physical modelling of flow and dispersion over urban areas, in Cermak, J. E., Davenport, A. G., Plate, E. J.
and Viegas, D. X. (eds) Wind Climate in Cities, Kluwer Academic Publishers, Dordrecht/Boston/London, pp383403
Chan, A., E. So and S. Samad (2001) Strategic guidelines for street canyon geometry to achieve sustainable street air
quality, Atmospheric Environment, vol 35, pp56815691
Chan, A., W. Au and E. So (2003) Strategic guidelines for street canyon geometry to achieve sustainable street air
quality part II: Multiple canopies and canyons, Atmospheric Environment, vol 37, pp27612772
Chow, W. and M. Roth (2006) Temporal dynamics of the urban heat island of Singapore, International Journal of Climatology,
vol 26, pp22432260
Dixon, N. S., J. W. D. Boddy, R. J. Smalley and A. S. Tomlin (2006) Evaluation of a turbulent flow and dispersion model in a
typical street canyon in York, UK, Atmospheric Environment, vol 40, pp958972
Dobre, A., S. Arnold, R. J. Smalley, J. W. D. Boddy, J. Barlow, A. S. Tomlin and S. Belcher (2005) Flow field measurements in
the proximity of an urban intersection in London, UK, Atmospheric Environment, vol 39, pp46474657
Doulos, L., M. Santamouris and I. Livada (2004) Passive cooling of outdoor spaces: The role of materials, Solar Energy,
vol 77, pp231249
El Nahas, M. M. (1996) Energy Saving Through Urban Design A Microclimatic Approach, PhD thesis, Department of
Architecture, University of Adelaide, Adelaide

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

117

El-Shakhs, S. (1994) Sadat City, Egypt, and the role of new town planning in the developing world, Journal of Architectural
and Planning Research, vol 11, no 3, pp239259
El-Sioufi, M. (1987) Urban Patterns for Improved Thermal Performance, PhD thesis, Department of Architecture, University of
Michigan, MI
Eliasson, I. (1996) Intra-urban nocturnal temperature differences: A multivariate approach, Climate Research, vol 7, pp2130
Eliasson, I. (2000) The use of climate knowledge in urban planning, Landscape and Urban Planning, vol 48, pp3144
Emmanuel, R. (1993) A hypothetical shadow umbrella for thermal comfort enhancement in the equatorial urban outdoors,
Architectural Science Review, vol 36, pp173184
Emmanuel, R. (1995) Energy efficient urban design guidelines for warm-humid cities: Strategies for Colombo, Sri Lanka,
Journal of Architectural and Planning Research, vol 12, no 1, pp5879
Emmanuel, R. (1997) Summertime urban heat island mitigation: Propositions based on an investigation of intra-urban air

Downloaded by [Ira Irawati] at 19:23 27 January 2015

temperature variations, Architectural Science Review, vol 40, no 4, pp155164


Erell, E. and T. Williamson (2006) Simulating air temperature in an urban street canyon in all weather conditions using
measured data from a reference meteorological station, International Journal of Climatology, vol 26, no 12,
pp16711694
Erell, E. and Williamson, T. (2007) Intra-urban differences in canopy layer air temperature at a mid-latitude city, International
Journal of Climatology, vol 27, pp12431255
Erell, E., V. Leal and E. Maldonado (2005) Measurement of air temperature in the presence of a large radiant flux: An
assessment of passively ventilated thermometer screens, Boundary Layer Meteorology, vol 114, pp205231
Etzion, Y. (1990) A desert solar neighbourhood in Sde-Boqer, Israel, Architectural Science Review, vol 33, pp105111
Evans, J. M. and S. de Schiller (19901991) Climate and urban planning: The example of the planning code for Vicente
Lopez, Buenos Aires, Energy and Buildings, vol 1516, pp3541
Giridharan, R., S. Ganesan and S. Lau (2004) Daytime urban heat island effect in high-rise and high-density residential
developments in Hong Kong, Energy and Buildings, vol 36, pp525534
Givoni, B. (1989) Urban Design in Different Climates, WMO/TD, no 346
Goh, K. and C. Chang (1999) The relationship between height to width ratios and the heat island intensity at 22:00 for
Singapore, International Journal of Climatology, vol 19, pp10111023
Golany, G. S. (1996) Urban design morphology and thermal performance, Atmospheric Environment, vol 30, no 3, pp455465
Gotz, L. (1982) Integration of climate in planning and building illustrated in a case of extreme climatic conditions, Energy and
Buildings, vol 4, pp5165
Grimmond, C. S. B. (2006) Progress in measuring and observing the urban atmosphere, Theoretical and Applied Climatology,
vol 84, pp322
Grimmond, C. S. B. and T. R. Oke (1999) Aerodynamic properties of urban areas derived from analysis of surface form,
Journal of Applied Meteorology, vol 38, pp12621292
Grimmond, C. S. B., C. Souch and M. D. Hubble (1996) Influence of tree cover on summertime surface energy balance fluxes,
San Gabriel Valley, Los Angeles, Climate Research, vol 6, pp4557
Gupta, V. K. (1984) Solar radiation and urban design for hot climates, Environment and Planning B: Planning and Design,
vol 11, pp435454
Herz, R. K. (1988) Considering climatic factors for urban land-use planning in the Sahelian zone, Energy and Buildings, vol 11,
pp91102
Hotchkiss, R. S. and F. H. Harlow (1973) Air Pollution Transport in Street Canyons, US EPA Report (EPAR473029)
Howard, L. (1833) The Climate of London, Harvey and Darton, London
Hoyano, A. (1988) Climatological uses of plants for solar control and the effects on the thermal environment of a building,
Energy and Buildings, vol 11, pp181200

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

118

EVYATAR ERELL

Huang, Y. J., H. Akbari, H. Taha and A. H. Rosenfeld (1987) The potential of vegetation in reducing summer cooling loads in
residential buildings, Journal of Climate and Applied Meteorology, vol 26, pp11031116
Jauregi, E. (19901991) Influence of a large urban park on temperature and convective precipitation in a tropical city, Energy
and Buildings, vol 1516, pp457463
Kanda, M. and R. Moriwaki (1998) Environmental effect of Meiji-shrine forest as a sink of energy and pollutants, Second
Urban Climate Symposium, Albuquerque, New Mexico
Kastner-Klein, P. and E. J. Plate (1999) Wind-tunnel study of concentration fields in street canyons, Atmospheric
Environment, vol 33, pp39733979
Kenworthy, A. T. (1985) Wind as an influential factor in the orientation of the orthogonal street grid, Building and
Environment, vol 20, no 1, pp3338
Kim, J. and J. Baik (1998) A numerical study of thermal effects on flow and pollutant dispersion in urban street canyons,

Downloaded by [Ira Irawati] at 19:23 27 January 2015

Second Urban Environment Symposium, Albuquerque, New Mexico


Kim, J. and J. Baik (2001) Urban street canyon flows with bottom heating, Atmospheric Environment, vol 35, pp33953404
Kim, J. and J. Baik (2004) A numerical study of the effects of ambient wind direction on flow and dispersion in urban street
canyons using the RNG k- turbulence model, Atmospheric Environment, vol 38, pp30393048
Kjelgren, R. and T. Montague (1998) Urban tree transpiration over turf and asphalt surfaces, Atmospheric Environment,
vol 32, no 1, pp3541
Knowles, R. (1981) Sun, Rhythm, Form, The MIT Press, Cambridge, MA
Knowles, R. (2003) The solar envelope: Its meaning for energy and buildings, Energy and Buildings, vol 35, pp1525
Kolokotroni, M., I. Giannitsaris and R. Watkins (2006) The effect of the London urban heat island on building summer cooling
demand and night ventilation strategies, Solar Energy, vol 80, pp383392
Krishan, A. (1996) The habitat of two deserts in India: Hot-dry desert of Jaisalmer (Rajasthan) and the cold-dry high altitude
mountainous desert of Leh (Ladakh), Energy and Buildings, vol 23, no 3, pp217229
Landsberg, H. E. (1968) Climates and urban planning, Symposium on Urban Climates and Building Climatology, Brussels,
Belgium
Levinson, R., H. Akbari and J. Reilly (2007) Cooler tile-roofed buildings with near-infrared reflective non-white coatings,
Building and Environment, vol 42, pp25912605
Littlefair, P. (1998) Passive solar urban design: Ensuring the penetration of solar energy into the city, Renewable and
Sustainable Energy Reviews, vol 2, pp303326
Livada, I., M. Santamouris, K. Niachou, N. Papanikolaou and G. Mihalakakou (2002) Determination of places in the great
Athens area where the heat island effect is observed, Theoretical and Applied Climatology, vol 71, pp219230
Manins, P. C., M. E. Cope, P. J. Hurley, P. W. Newton, N. C. Smith and M. Lo (1998) The impact of urban development on air
quality and energy use, 14th International Clean Air & Environment Conference, Melbourne, Australia, 1822 October 1998
Mazouz, S. and M. S. Zerouala (1998) Shading as a modulator for the design of urban layouts based on vernacular
experiences, Energy and Buildings, vol 29, pp1115
McPherson, E. G., L. P. Herrington and G. M. Heisler (1988) Impacts of vegetation on residential heating and cooling, Energy
and Buildings, vol 12, pp4151
Meier, A. K. (19901991) Strategic landscaping and air-conditioning savings: A literature review, Energy and Buildings,
vol 1516, pp479486
Mills, G. (1997) The radiative effects of building groups on single structures, Energy and Buildings, vol 25, pp5161
Mills, G. (1999) Urban climatology and urban design, 15th International Congress of Biometeorology and International
Conference on Urban Climatology, Sydney, Australia
Mills, G. (2006) Progress toward sustainable settlements: A role for urban climatology, Theoretical and Applied Climatology,
vol 84, pp6976

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

119

Nakamura, Y. and T. R. Oke (1988) Wind, temperature and stability conditions in an eastwest oriented urban canyon,
Atmospheric Environment, vol 22, no 12, pp26912700
Niachou, A., K. Papakonstantinou, M. Santamouris, A. Tsangrassoulis and G. Mihalakakou (2001) Analysis of the green roof
thermal properties and investigation of its energy performance, Energy and Buildings, vol 33, pp719729
Nunez, M. and T. R. Oke (1977) The energy balance of an urban canyon, Journal of Applied Meteorology, vol 16, pp1119
Oke, T. R. (1973) City size and the urban heat island, Atmospheric Environment, vol 7, no 8, pp769779
Oke, T. R. (1981) Canyon geometry and the nocturnal urban heat island: Comparison of scale model and field observations,
Journal of Climatology, vol 1, no 3, pp237254
Oke, T. R. (1982) The energetic basis of the urban heat island, Quarterly Journal of the Royal Meteorological Society, vol 108,
no 455, pp124
Oke, T. R. (1984) Towards a prescription for the greater use of climatic principles in settlement planning, Energy and

Downloaded by [Ira Irawati] at 19:23 27 January 2015

Buildings, vol 7, pp110


Oke, T. R. (1987) Boundary Layer Climates, Methuen, London and New York
Oke, T. R. (1988a) Street design and urban canopy layer climate, Energy and Buildings, vol 11, pp103113
Oke, T. R. (1988b) The urban energy balance, Progress in Physical Geography, vol 12, no 4, pp471508
Oke, T. R. (1989) The micrometeorology of the urban forest, Philosophical Transactions of the Royal Society London B,
vol 324, pp335349
Olgyay, V. (1963) Design with Climate, Princeton University Press, Princeton, NJ
Onmura, S., M. Matsumoto and S. Hokoi (2001) Study on the evaporative cooling effect of roof lawn gardens, Energy and
Buildings, vol 33, pp653666
Page, J. K. (1968) The fundamental problems of building climatology considered from the point of view of decision-making by
the architect and urban designer, Symposium on Urban Climates and Building Climatology, Brussels, Belgium
Page, J. K. (1972) The problem of forecasting the properties of the built environment from the climatological properties of
the green-field site, in Taylor, J. A. (ed) Weather Forecasting for Agriculture and Industry, David & Charles, London
Palomo Del Barrio, E. (1998) Analysis of the green roofs cooling potential in buildings, Energy and Buildings, vol 27,
pp179193
Papadakis, G., P. Tsamis and S. Kyritsis (2001) An experimental investigation of the effect of shading with plants for solar
control of buildings, Energy and Buildings, vol 33, pp831836
Pearlmutter, D., A. Bitan and P. Berliner (1999) Microclimatic analysis of compact urban canyons in an arid zone,
Atmospheric Environment, vol 33, pp41434150
Pearlmutter, D., P. Berliner and E. Shaviv (2006) Physical modeling of the pedestrian energy exchange within the urban
canopy, Building and Environment, vol 41, pp783795
Pearlmutter, D., P. Berliner and E. Shaviv (2007) Integrated modeling of pedestrian energy exchange and thermal comfort in
urban street canyons, Building and Environment, vol 42, pp23962409
Potchter, O. (1988) Climatic aspects in rural settlement development in hot, arid zones: A case study of the Central Jordan
Valley, Energy and Buildings, vol 11, pp7390
Pressman, N. (1996) Sustainable winter cities: Future directions for planning, policy and design, Atmospheric Environment,
vol 30, no 3, pp521529
Rafailidis, S. (1997) Influence of building areal density and roof shape on the wind characteristics above a town, Boundary
Layer Meteorology, vol 85, pp225271
Ratti, C. and P. Richens (1999) Urban texture analysis with image processing techniques, Computers in Building: Proceedings
of the CAADfutures 1999 Conference, Atlanta, Georgia, 78 June 1999
Ratti, C. and P. Richens (2004) Raster analysis of urban form, Environment and Planning B: Planning and Design, vol 31,
pp297309

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

120

EVYATAR ERELL

Ratti, C., D. Raydan and K. Steemers (2003) Building form and environmental performance: Archetypes, analysis and an arid
climate, Energy and Buildings, vol 35, pp4959
Ratti, C., N. Baker and K. Steemers (2005) Energy consumption and urban texture, Energy and Buildings, vol 37,
pp762776
Ratti, C., S. Di Sabatino and R. Britter (2006) Urban texture analysis with image processing techniques: Winds and
dispersion, Theoretical and Applied Climatology, vol 84, pp7790
Raza, S. H., M. S. R. Murthy, O. B. Lakshmi and G. Shylaja (19901991) Effect of vegetation on urban climate and healthy
urban colonies, Energy and Buildings, vol 1516, pp487491
Robinson, D. and A. Stone (2004) Solar radiation modelling in the urban context, Solar Energy, vol 77, pp295309
Runnalls, K. and T. R. Oke (1998) The urban heat island of Vancouver, BC, Second Urban Environment Symposium,
Albuquerque, New Mexico, 26 November 1998

Downloaded by [Ira Irawati] at 19:23 27 January 2015

Runnalls, K. and T. R. Oke (2000) Dynamics and controls of the near-surface heat island of Vancouver, British Columbia,
Physical Geography, vol 21, no 4, pp283304
Sailor, D. J. (1998) Simulations of annual degree day impacts of urban vegetative augmentation, Atmospheric Environment,
vol 32, no 1, pp4352
Sailor, D. and H. Fan (2002) Modeling the diurnal variability of effective albedo for cities, Atmospheric Environment, vol 36,
pp713725
Saito, I., O. Ishihara and T. Katayama (19901991) Study of the effect of green areas on the thermal environment in an urban
area, Energy and Buildings, vol 1516, pp493498
Santamouris, M. (1998) The Athens urban climate experiment, PLEA 1998: Environmentally Friendly Cities, Lisbon, Portugal,
14 June 1998
Santamouris, M. (ed) (2001) Energy and Climate in the Urban Built Environment, James & James, London
Santamouris, M., N. Papanikolaou, I. Livada, I. Koronakis, C. Georgakis, A. Argiriou and D. N. Assimakopoulos (2001) On the
impact of urban climate on the energy consumption of buildings, Solar Energy, vol 70, no 3, pp201216
Schiller, G. and R. Karschon (1974) Microclimate and recreational value of tree plantings in deserts, Landscape Planning,
vol 1, pp329337
Shashua-Bar, L. and M. Hoffman (2000) Vegetation as a climatic component in the design of an urban street: An empirical
model for predicting the cooling effect of urban green areas with trees, Energy and Buildings, vol 31, no 3, pp221235
Shashua-Bar, L. and M. Hoffman (2002) The Green CTTC model for predicting the air temperature in small urban wooded
sites, Building and Environment, vol 37, pp12791288
Shashua-Bar, L. and M. Hoffman (2003) Geometry and orientation aspects in passive cooling of canyon streets with trees,
Energy and Buildings, vol 35, pp6168
Shashua-Bar, L. and M. Hoffman (2004) Quantitative evaluation of passive cooling of the UCL microclimate in hot regions in
summer, case study: Urban streets and courtyards with trees, Building and Environment, vol 39, pp10871099
Simpson, J. R. (2002) Improved estimates of tree-shade effects on residential energy use, Energy and Buildings, vol 34,
no 10, pp10671076
Simpson, J. R. and E. G. McPherson (1998) Simulation of tree shade impacts on residential energy use for space conditioning
in Sacramento, Atmospheric Environment, vol 32, no 1, pp6974
So, E., A. Chan and A. Wong (2005) Large-eddy simulations of wind flow and pollutant dispersion in a street canyon,
Atmospheric Environment, vol 39, pp35733582
Soler, M. R. and C. Ruiz (1994) Urban albedo derived from direct measurements and LANDSAT 4 TM satellite data,
International Journal of Climatology, vol 14, pp925931
Spronken-Smith, R. A. and T. R. Oke (1998) The thermal regime of urban parks in two cities with different summer climates,
International Journal of Remote Sensing, vol 19, no 11, pp20852104

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

The Application of Urban Climate Research in the Design of Cities

121

Spronken-Smith, R. and T. R. Oke (1999) Scale modelling of nocturnal cooling in urban parks, Boundary Layer Meteorology,
vol 93, pp287312
Steemers, K., N. Baker, D. Crowther, J. Dubiel, M.-H. Nikolopoulou and C. Ratti (1997) City texture and microclimate, Urban
Design Studies, vol 3, pp2550
Steinecke, K. (1999) Urban climatological studies in the Reykjavik subarctic environment, Iceland, Atmospheric Environment,
vol 33, pp41574162
Swaid, H. (1992) Intelligent urban forms (IUF): A new climate-concerned urban planning strategy, Theoretical and Applied
Climatology, vol 46, pp179191
Synnefa, A., M. Santamouris and I. Livada (2006) A study of the thermal performance of reflective coatings for the urban
environment, Solar Energy, vol 80, pp968981
Synnefa, A., M. Santamouris and K. Apostolakis (2007) On the development, optical properties and thermal performance of

Downloaded by [Ira Irawati] at 19:23 27 January 2015

cool colored coatings for the urban environment, Solar Energy, vol 81, no 4, pp488497
Taha, H. G. (1978) An Urban Micro-Climate Model for Site-Specific Building Energy Simulation: Boundary Layers, Urban
Canyon and Building Conditions, PhD thesis, Department of Architecture, University of California, Berkeley, CA
Taha, H. (1997) Urban climates and heat islands: Albedo, evapotranspiration and anthropogenic heat, Energy and Buildings,
vol 25, pp99103
Taha, H., H. Akbari, A. Rosenfeld and J. Huang (1988) Residential cooling loads and the urban heat island the effects of
albedo, Building and Environment, vol 23, no 4, pp271283
Taha, H., S. Douglas and J. Haney (1997) Mesoscale meteorological and air quality impacts of increased urban albedo and
vegetation, Energy and Buildings, vol 25, pp169177
Tsangrassoulis, A. and M. Santamouris (2003) Numerical estimation of street canyon albedo consisting of vertical coated
glazed facades, Energy and Buildings, vol 35, pp527531
Upmanis, H. and D. Chen (1999) Influence of geographical factors and meteorological influences on nocturnal urban-park
temperature differences a case study of summer 1995 in Goteborg, Sweden, Climate Research, vol 13, pp125139
Upmanis, H., I. Eliasson and S. Lindqvist (1998) The influence of green areas on nocturnal temperatures in a high latitude city
(Goteborg, Sweden), International Journal of Climatology, vol 18, pp681700
Urano, A., Y. Morikawa and M. Nishimura (1995) Urban environment analysis of the Kanto Plain with 1-dimensional and
3-dimensional numerical simulations, Pan Pacific Symposium on Building and Urban Environmental Conditioning in Asia,
Nagoya, Japan, March 1995
Vitruvius (1999) Ten Books on Architecture, Cambridge University Press, Cambridge, UK
Westerberg, U. and M. Glaumann (1990/1991) Design criteria for solar access and wind shelter in the outdoor environment,
Energy and Buildings, vol 1516, pp425431
Wong, N. H., Y. Chen, C. L. Ong and A. Sia (2003) Investigation of the thermal benefits of rooftop garden in the tropical
environment, Building and Environment, vol 38, pp261270
Xie, X., Z. Huang, J. Wang and Z. Xie (2005) Thermal effects on vehicle emission dispersion in an urban street,
Transportation Research Part D, vol 10, pp197212
Xie, X., C. Liu, D. Leung and M. Leung (2006) Characteristics of air exchange in a street canyon with ground heating,
Atmospheric Environment, vol 40, pp63966409
Yamartino, R. J. and G. Wiegand (1986) Development and evaluation of simple models for the flow, turbulence and pollutant
concentration fields within an urban street canyon, Atmospheric Environment, vol 20, pp21372156
Yamashita, S., K. Sekine, M. Shoda, K. Yamashita and Y. Hara (1986) On relationships between heat island and sky view
factor in the cities of Tama River basin, Japan, Atmospheric Environment, vol 20, no 4, pp681686
Yezioro, A., I. G. Capeluto and E. Shaviv (2006) Design guidelines for appropriate insolation of urban squares, Renewable
Energy, vol 31, pp10111023

ADVANCES IN BUILDING ENERGY RESEARCH 2008 VOLUME 2 PAGES 95121

You might also like