139
Urban Planning and Urban Design
Coordinating Lead Author
Jeffrey Raven (New York)
Lead Authors
Brian Stone (Atlanta), Gerald Mills (Dublin), Joel Towers (New York), Lutz Katzschner (Kassel), Mattia Federico Leone (Naples), Pascaline
Gaborit (Brussels), Matei Georgescu (Tempe), Maryam Hariri (New York)
Contributing Authors
James Lee (Shanghai/Boston), Jeffrey LeJava (White Plains), Ayyoob Shari (Tsukuba/Paveh), Cristina Visconti (Naples), Andrew Rudd
(Nairobi/New York)
5
This chapter should be cited as
Raven, J., Stone, B., Mills, G., Towers, J., Katzschner, L., Leone, M., Gaborit, P., Georgescu, M., and Hariri, M. (2018). Urban planning and
design. In Rosenzweig, C., W. Solecki, P. Romero-Lankao, S. Mehrotra, S. Dhakal, and S. Ali Ibrahim (eds.), Climate Change and Cities:
Second Assessment Report of the Urban Climate Change Research Network. Cambridge University Press. New York. 139172
ARC3.2 Climate Change and Cities
140
Embedding Climate Change in Urban
Planning and Urban Design
Urban planning and urban design have a critical role to play
in the global response to climate change. Actions that simul-
taneously reduce greenhouse gas (GHG) emissions and build
resilience to climate risks should be prioritized at all urban
scales – metropolitan region, city, district/neighborhood, block,
and building. This needs to be done in ways that are responsive
to and appropriate for local conditions.
Major Findings
Urban planners and urban designers have a portfolio of cli-
mate change strategies that guide decisions on urban form and
function:
Urban waste heat and GHG emissions from infrastructure –
including buildings, transportation, and industry – can be reduced
through improvements in the efciency of urban systems.
Modifying the form and layout of buildings and urban dis-
tricts can provide cooling and ventilation that reduces energy
use and allow citizens to cope with higher temperatures and
more intense runoff.
Selecting low heat capacity construction materials and reec-
tive coatings can improve building performance by managing
heat exchange at the surface.
Increasing the vegetative cover in a city can simultaneously
lower outdoor temperatures, building cooling demand, run-
off, and pollution, while sequestering carbon.
Key Messages
Integrated climate change mitigation and adaptation strategies
should form a core element in urban planning and urban design,
taking into account local conditions. This is because decisions
on urban form have long-term (>50 years) consequences and
thus strongly affect a city’s capacity to reduce GHG emissions
and to respond to climate hazards over time. Investing in miti-
gation strategies that yield concurrent adaptation benets should
be prioritized in order to achieve the transformations necessary
to respond effectively to climate change.
Consideration needs to be given to how regional decisions
may affect neighborhoods or individual parcels and vice versa,
and tools are needed that assess conditions in the urban environ-
ment at city block and/or neighborhood scales.
There is a growing consensus around integrating urban plan-
ning and urban design, climate science, and policy to bring
about desirable microclimates within compact, pedestrian-
friendly built environments that address both mitigation and
adaptation.
Urban planning and urban design should incorporate long-
range mitigation and adaptation strategies for climate change
that reach across physical scales, jurisdictions, and electoral
timeframes. These activities need to deliver a high quality of life
for urban citizens as the key performance outcome, as well as
climate change benets.
Chapter 5 Urban Planning and Urban Design
141
5.1 Introduction
Key concepts, challenges, and pathways for adaptation
and mitigation of climate change through recent advances in
the planning and design of cities are reviewed in this chapter.
Section 5.2 presents the concept of integrated mitigation and
adaptation as a framework and introduces the factors of urban-
scale form and function as inuences on urban climate. Section
5.3 explains how urban microclimates are embedded in zones
of human occupation and links metropolitan-scale urbanization
with heat and storm-water impacts. Section 5.4 focuses on
planning and design innovations that can be applied to achieve
integrated mitigation and adaptation. Section 5.5 describes a
process for implementing climate-responsive urban planning
and urban design. Section 5.6 identies key climate-resilient
urban planning and urban design stakeholders and a set of
value propositions to engage a broader constituency. Section
5.7 describes the challenges in cross-sector linkages between
the scientic, design, and policy-making communities. Section
5.8 identies knowledge gaps and future research opportunities,
and Section 5.9 presents conclusions and recommendations for
practitioners and policy-makers. Case Studies are distributed
throughout the chapter to illustrate on-the-ground, effective
implementations of the planning and design strategies
presented.
5.2 Framework for Sustainable and
Resilient Cities
Urban planning and urban design encompasses multiple dis-
ciplines, providing critical input to inform systems, manage-
ment, and governance for sustainability and resilience to climate
change (see Box 5.1). They congure spatial outcomes that yield
consequences for and constitute responses to climate change
(see Figure 5.1). The spatial form of a city – from the scale of the
metropolitan region to the neighborhood block – strongly prede-
termines per capita greenhouse gas (GHG) emissions. With each
10% reduction in urban sprawl, per capita emissions are reduced
by 6% (Laidley, 2015). Although compact urban form generally
contributes positively to mitigation, it can paradoxically exacer-
bate local climate effects, requiring creative forms of adaptation.
Research in this area is expanding, and, as a result, planning and
design strategies are increasingly providing win-win solutions
for compact urban morphology.
However, not all existing urban areas are compact (see
Figure 5.2). Low-density areas continue to contribute dis-
proportionately to emissions because of the excess mobility
required by long distances, few alternatives to the private car,
and scant possibilities for shared building envelopes. Whether
such patterns are the result of planning or a lack thereof, it is
1. Efficiency of Urban Systems2. Form and Layout
3. Heat-Resistant
Construction Material
4. Vegetative Cover
Urban Water
Drainage
Urban
Farm
Green
Roof
Hot
Roof
Cool
Roof
Solar
Energy
Natural
Ventilation
Green
Path
Transit -
Oriented
Zone
Transit
Rail
500 M
Figure 5.1 Strategies used by urban planners and urban designers to facilitate integrated mitigation and adaptation in cities: (1) reducing waste heat and greenhouse gas
emissions through energy efficiency, transit access, and walkability; (2) modifying form and layout of buildings and urban districts; (3) use of heat-resistant construction
materials and reflective surface coatings; and (4) increasing vegetative cover.
Source: Jeffrey Raven, 2016
ARC3.2 Climate Change and Cities
142
clear that the planning and design disciplines will increasingly
need to prioritize the retrotting of these areas for greater land-
use efciency (UN-Habitat, 2012).
A high proportion of urban areas that will need to minimize
GHG emissions and adapt to climate change have not yet been
built. Beyond aiming for appropriate levels of compactness, new
urban development can and must be strategic about location
(avoiding, e.g., areas particularly vulnerable to heat, ooding,
or landslides).
5.2.1 Integrated Mitigation and Adaptation
Urban planning and urban design can be critical platforms
for integrated mitigation and adaptation responses to the chal-
lenges of climate change. They have the opportunity to expand
on the traditional inuence and capabilities of practitioners and
policy-makers and integrate climate science, natural systems,
and urban form – particularly compact urban form – to congure
dynamic, desirable, and healthy communities.
Traditionally, urban planning and urban design have focused
on settlement patterns, optimized land use, maximized proxim-
ity, community engagement, place-making, quality of life, and
urban vitality. Their focus is increasingly expanding to include
principles such as resilience, comfort, resource efciency,
and ecosystem services (see Chapter 8, Urban Ecosystems).
Applying these principles to urban policy helps to identify and
strengthen prescriptive measures and performance standards
and broaden urban performance indicators.
If future cities are to be sustainable and resilient, they must
develop the physical and institutional capacities to respond to
constant change and uncertainty. This will require strategies for
long-term commitments across multiple electoral cycles and often
among many political jurisdictions that constitute functional met-
ropolitan areas (see Chapter 16, Governance and Policy).
Global climate risk is accumulated in urban areas because
people, private and public assets, and economic activities
become more concentrated in cities (Mehrotra et al., 2011;
Revi et al., 2014). Recognition of the growing vulnerability of
Box 5.1 Key Definitions for Urban Planning and Urban Design
Urban planning: A eld of practice that helps city leaders
to transform a sustainable development vision into reality
using space as a key resource for development and engag-
ing a wide variety stakeholders in the process. It generally
takes place at the scale of the city or metropolitan region
whose overall spatial pattern it sets. Good urban planning
formulates medium- and long-term objectives that recon-
cile a collective vision with the rational organization of the
resources needed to achieve it. It makes the most of munic-
ipal budgets by informing infrastructure and services invest-
ments and balancing demands for growth with the need to
protect the environment. And it ideally distributes economic
development within a given urban area to reach wider social
objectives (UN-Habitat 2013).
Urban design: Urban design involves the arrangement and
design of buildings, public spaces, transport systems, ser-
vices, and amenities. Urban design is the process of giving
form, shape, and character to groups of buildings, to whole
neighborhoods, and to a city. It is a framework that orders
the elements into a network of streets, squares, and blocks.
Urban design blends architecture, landscape architecture,
and city planning together to make urban areas functional
and attractive.
Urban design is about making connections between peo-
ple and places, movement and urban form, nature and the
built fabric. Urban design draws together the many strands
of place-making, environmental stewardship, social equity,
and economic viability into the creation of places with dis-
tinct beauty and identity. Urban design is derived from but
transcends planning, transportation policy, architectural
design, development economics, engineering, and land-
scape. It draws together create a vision for an urban area
and then deploys the resources and skills needed to bring
the vision to life (urbandesign.org).
Figure 5.2 Each bar represents an entire metropolitan area (i.e., the city and
the continuous urban footprint surrounding it), including often much lower-density
suburbs.
Source: A. L. Brenkert, Oak Ridge National Laboratory. Maps created by Andreas Christen, UBC
Chapter 5 Urban Planning and Urban Design
143
Figure 5.3 “Green and blue fingers” in Thanh Hoa City, Vietnam, planned for 2020:
Contiguous green corridors and canal circulation networks aligned with prevailing
summer breezes, punctuated by stormwater retention bodies as urban design
amenities.
Source: Jeffrey Raven, Louis Berger Group, 2008
urban populations to climate-related health threats requires that
the climate management activities of municipal governments
be broadened (see Chapter 10, Urban Health).
One means of doing so is to prioritize investments in mitiga-
tion strategies that yield concurrent adaptive benets over those
that do not (see Chapter 4, Mitigation and Adaptation). At pres-
ent, nonintegrated mitigation and adaptation is most commonly
pursued, with the majority of mitigation funds directed to
energy projects that produce no secondary benets for local
populations in the form of heat management and enhanced
ood protection or reduced damage to private property
and public infrastructure. For example, mitigation strate-
gies involving the substitution of a lower carbon-intensive
fuel, such as natural gas, for a higher carbon-intensive fuel,
such as coal, are an effective means of lowering CO
2
emissions yet provide few benets related to climate
adaptation.
5.2.2 Form and Function
Forward-thinking cities are beginning to exploit the positive
potential of built and natural systems – including green infra-
structure, urban ventilation, and solar orientation – to “future-
proof” the built environment in response to changing conditions
(see Figure 5.3 and Box 5.2). These passive urban design strate-
gies “lock in” long-term resilience and sustainability, protecting
Box 5.2 Urban Form and Function
The physical character of cities can be described by
three aspects of form: land cover, urban materials, and
morphology. On the other hand, the ow of materials
through a city describes its metabolism, the character
of which is regulated by its functions (see, e.g., Decker
et al., 2000).
Surface cover (Form): The replacement of natural land
covers by impermeable materials limits the inltration
of precipitation into the substrate, increases runoff, and
decreases evapotranspiration.
Construction materials and surface coating (Form):
Common urban materials such as concrete have high con-
ductivity and heat capacity values that can store heat ef-
ciently. Also, many urban materials are dark colored and
reect poorly (e.g., asphalt).
Morphology (Form): The conguration and orientation of
the built environment, from regional settlement patterns to
buildings, create a corrugated surface that slows and redi-
rects near-surface airow and traps radiation.
Urban activities (Function): Cities concentrate material,
water, and energy use that must be acquired from a
much larger area. Some is used to build the city (chang-
ing its form), but most is employed to sustain its econ-
omy and society. Once used, the wastes and emissions
are deposited into the wider environment, degrading soil,
water, and air quality, and increasing heat through the
exacerbated greenhouse effect.
ARC3.2 Climate Change and Cities
144
the city from future decisions that could undermine its adapt-
ability. They also remove the risk of relying on bolted-on,
applied technologies that may require expensive maintenance
or become obsolete in a short time. These form-based, contextu-
ally specic urban planning and urban design strategies are the
ultimate guarantors of successful life cycle costs, payback, and
liveability.
Integrated mitigation and adaptation in cities can assume
many forms across spatial scales, urban systems, and physical
networks (see Figure 5.4); a wide range of strategies adopted in
service of urban sustainability already advances this objective.
Enhanced urban transit, for example, has the effect of reducing
both carbon emissions from single-occupant vehicle use and
waste heat emissions that contribute to the urban heat island
(UHI) effect. Investments in pedestrian and cycling corridors,
particularly when integrated with parks and other green spaces
in cities, can reduce carbon emissions, enhance carbon seques-
tration, and, perhaps most effectively, cool cities through
evapotranspiration and shading. Sustainability strategies across
urban systems can contribute to climate management goals
under the umbrella of integrated mitigation and adaptation (see
Figure 5.5).
National
Region
Sub-Region
City
Neighborhood
Site
Building
Desired Outcome
s
Figure 5.4 Spatial scales relevant to urban planning and urban design for climate
change mitigation and adaptation.
Source: Jeffrey Raven, 2008
Figure 5.5 L’ Ecosystème Urbain (Urban Ecosystem).
Source: Duvigneaud, P. and Denayer-de Smet, S., 1975
Chapter 5 Urban Planning and Urban Design
145
5.3 Climate in Cities
Urban areas occupy a small percentage (perhaps less than 3%)
of the planet’s land area, but this area is intensively modied
(Miller and Small, 2003; Schneider et al., 2009). The landscape
changes that accompany urbanization modify climate across a
spectrum of scales, from the micro-scale (e.g., street), city-scale,
and regional scales (see Chapter 2, Urban Climate Science).
The magnitude of the modication is evaluated by comparing
the urban climate with its background climate, which is taken
to be the “natural” climate. Because each city has a unique geo-
graphical region (latitude and topography primarily), the natu-
ral climate is assessed in the same region but over a non-urban
surface (Lowry, 1977). One of the challenges posed by global
climate change is that the background climate is itself changing
and that cities contribute signicantly to this change through the
emission of GHGs.
The most profound changes occur in the layer of air below
roof height. Here, access to sunlight is restricted, wind is slowed
and diverted, and energy exchanges between buildings are the
norm. The spatial heterogeneity of the urban landscape creates
a myriad of microclimates associated with individual buildings
and their relative disposition, streets, and parks (Errel et al.,
2012). This is also the layer of intense human occupation, where
building heating and cooling demand is met, emissions of waste
heat and pollution from trafc are concentrated, and humans are
exposed to a great variety of indoor and outdoor urban climates.
The climate effect of cities extends well beyond the urban-
ized area. As air ows over the urban surface, a boundary layer
forms that deepens with distance from the upwind edge. This
envelope may be 1–2 kilometers thick by mid-afternoon and
is distinguishable as a warm and turbulent atmosphere that
is enriched with contaminants, including GHGs. The extent
of urban inuence depends on the character of the city (e.g.,
its area, built density, and the intensity of its emissions) and
on the background climate, which regulates the spread and
dilution of the urban envelope. As a result, cities contribute
signicantly to regional and global air pollution (Guttikunda
et al., 2003; Monks et al., 2009). Moreover, meeting urban
energy demand accounts for up to three-quarters of CO
2
emis-
sions from global energy use and thus represents a signicant
driver of global climate change (IPCC, 2014) (see Chapter 12,
Urban Energy).
The magnitude of the urban climate effect is linked to both
the form and function of cities (Box 5.2). The former refers to
aspects of the physical character of cities, including the extent
of paving and the density of buildings. The latter describes the
nature of urban occupancy including the energy used in build-
ings, transport, and industry. Integrated mitigation and adap-
tation strategies focus on managing urban form and function
together to moderate and respond to climate changes at urban,
regional, and global scales.
5.3.1 Urban Climate Zones
The changes that accompany urbanization have profound
impacts on the local environment and are clearly seen in aspects
of climate and hydrology (Hough, 1989) (see Chapter 2 Urban
Climate Science). The magnitude of these urban effects depends
on both the form and functions of individual cities. However,
cities are highly heterogeneous landscapes, and impacts vary
across the urbanized area as well. Detailed mapping of urban
layout, including aspects of form (e.g., impervious land cover)
and of function (e.g., commercial land use), provides a basis for
examining climate at a local scale.
For example, Stewart and Oke (2012) have developed a simple
scheme that classies urban neighborhoods mainly by form into
local climate zones (LCZ) (see Figure 5.6). Each LCZ is char-
acterized by typical building heights, street widths, vegetative
cover, and paved area. Not surprisingly, the most intense local
climate impacts are found where building density is greatest,
streets are narrowest, and there is little vegetation (e.g., compact
high-rises or dense slums). In many of these areas, the population
is highly vulnerable due to poverty or age (see Chapter 6, Equity
and Environmental Justice). Cities comprise many LCZ types
that occupy varying proportions of the urbanized landscape. This
Figure 5.6 Local climate zone type. Admittance, or thermal admittance, is a
measure of a material’s ability to absorb heat from, and release it to, a space over
time. Albedo is the proportion of the incident light or radiation that is reflected by a
surface back into space.
Source: Stewart and Oke, 2012
ARC3.2 Climate Change and Cities
146
chapter describes win-win form/function strategies to mitigate
local climate impacts in compact districts.
5.3.2 Urbanization as Amplifier of Global Climate
Change
Global climate change is modifying the background climate
within which cities are situated, altering the frequency and
intensity of extreme weather experienced (see Chapter 2, Urban
Climate Science). The most recent Intergovernmental Panel on
Climate Change (IPCC) assessment (IPCC, 2014a) concludes
that global climate change has already resulted in warming both
days and nights over most land areas and will cause more fre-
quent hot days and nights in the future.
One of the most widely recognized climate impacts of urban-
ization is the UHI effect (e.g., Arneld, 2003; Roth, 2007) (see
Chapter 2, Urban Climate Science). The magnitude of the UHI
is measured as the difference in air and surface temperatures
between the city and proximate rural areas; these differences
increase from the edge of the city to the center, where it is usu-
ally at a maximum. It is strongest during calm and clear weather
but exhibits different impacts on surface and air temperatures.
When measured as differences in air temperature between urban
and non-urban surfaces, the UHI is strongest at night (due to
heat retention), whereas differences in surface temperatures are
largest during daytime (due to solar absorption).
Both types of UHI show a clear correlation with the amount
of impervious surface cover and building density, whereas parks
and green areas appear as cooler spots. The maximum value of
the UHI as measured by air temperatures is likely to be between
2°C and 10°C, depending on the size and built density of the
city, with largest values occurring in densely built and impervi-
ous neighborhoods (Oke, 1981). The magnitude of the surface
temperature UHI depends greatly on the material characteristics
of the surface, especially its albedo (i.e., reectivity) and mois-
ture status (see, e.g., Doulos et al., 2004).
In urban areas, the UHI adds to current warming trends
due to global climate change contributes to poor air quality,
increases energy demand for cooling, and elevates the inci-
dence of heat stress (Akbari, et al., 2001; Grimmond, 2007;
Oleson et al., 2015).
Once built, many aspects of the urban form are difcult to
change (overall layout and morphology especially), so imme-
diate emphasis must focus on altering aspects of surface cover
and construction materials in the short term. At the same time,
the role of urban planning to shape the potential doubling of cit-
ies’ total physical footprint within the next 15 years must not
be ignored because it represents a signicant opportunity for
mitigating future climate change at the global scale. New urban
development – particularly since much future urban develop-
ment will occur in warmer climate zones – can lower its emis-
sions drastically by pursuing compact development that employs
mixed-use zoning and public transit (Zhao et al., 2017; Resch et
al., 2016).
Projections of climate change show that there will be distinct
urban impacts. The locations of cities tend to be at low elevation,
close to coasts and in river valleys/basins, which exposes urban
areas to hazards such as high winds and ooding (McGranahan et
al., 2007; Miller and Small, 2003). The concentration of popula-
tion and infrastructure in cities makes them especially vulnerable
to the impacts of natural hazards. Land-use and land-cover strate-
gies designed to regulate these urban effects (such as urban green-
ing to mitigate urban ooding and heating) can complement global
climate change adaptation strategies that emphasize resilience.
Urbanization also has a dramatic impact on local hydrological
processes and water quality (see, e.g., Brabec et al., 2002; Paul
and Meyer, 2001). Impervious surface cover reduces the rate of
inltration to the underlying soil, thus limiting storage. While
sewers and channelized rivers improve the hydraulic efciency
of drainage networks, the net effect is to increase the risk of ood-
ing by increasing the volume and intensity of runoff (Kravík et
al., 2007; Konrad, 2003). In addition, the water that washes off
impermeable urban surfaces during rain events adds warm and
polluted water to river courses, further degrading water quality.
Moderating the magnitude of these urban effects in cities
requires altering aspects of urban form, especially surface
cover and materials. Vegetation in particular has an important
role to play as a versatile tool that can cool surfaces through
shading and cool air via evaporation (Shashua-Bar et al., 2010).
Green areas can also play a key role in water management by
delaying urban runoff and using soils as a lter to improve
water quality. Where the landscape is densely built, green roofs
can both insulate buildings, moderate air temperature above the
urban surface, and slow urban runoff (Mentens et al., 2006).
Similarly, changing surface albedo by applying surface coat-
ings or replacing impervious surfaces with permeable materials
can moderate urban effects (Gafn et al., 2012; Santamouris,
2014). Altering urban morphology once in place is a more chal-
lenging prospect. However, where change is possible, design
goals are to ensure access to the sun and provide shade, pro-
tection from wind, or ventilation by breezes (Bottema, 1999;
Knowles, 2003; Emmanuel et al., 2007; Chen et al., 2010).
Urban areas not yet built, particularly those in the developing
world, have the advantage of being able to design along these
parameters in advance of construction.
The role of urban design is critical because the urban climate
impact is a product of both its physical character and the back-
ground climate. The best solution in a city where the climate is
cool and wet will not be the same for a city in an arid and warm
climate. Similarly, where cities are already substantially built,
the opportunities for change will differ for each neighborhood.
Nevertheless, managing the outdoor climate can have multiple
benets including reduced demand for indoor cooling/heating
and increased use of outdoor spaces for health and improved air
quality (Akbari et al., 2001) (see Case Study 5.4).
Chapter 5 Urban Planning and Urban Design
147
In cities, emissions of GHGs arise mainly from buildings
(residential and commercial), transportation, and industries,
but the proportions vary based on the character of the urban
economy and the source of energy (Kennedy et al., 2009) (see
Chapter 12, Urban Energy). The contributions of buildings
and transport have received the most attention because each is
amenable to management at the urban scale using a variety of
measures, including building energy codes, public transit sys-
tems, and land-use management (ARUP, 2014). Much of the
current evidence indicates that densely occupied cities are more
efcient in their use of energy (and generate less waste heat as
a consequence) (Resch et al., 2016). The evidence is especially
strong for transport energy, which is largely based on cities
where there are mass transit systems (Newman and Kenworthy,
1989). Increasing urban population density through policies that
co-manage land-use and transport networks is an important strat-
egy for reducing urban GHG emissions (Dulal et al., 2011).
Good urban planning and urban design are critical to achiev-
ing climate change objectives at city, regional, and global scales.
Compact and densely occupied cities do not have to feature imper-
meable, densely built, and high-rise neighborhoods associated
with unwanted urban effects. In a study of urban spatial structure
and the occurrence of heat wave days across more than 50 large
U.S. cities, for example, Stone et al. (2010) found the annual fre-
quency of extreme heat events to be rising more slowly in compact
cities than in sprawling cities. A wealth of studies nd the enhance-
ment of vegetation and surface reectivity in dense urban environ-
ments to measurably reduce urban temperatures at the urban and
regional scale (see Taha et al., 1999; USEPA, 2008; Gafn et al.
2012; Stone et al., 2014). Further, tall buildings in cities impede
direct sunlight from reaching the ground (see Figure 5.7).
Figure 5.7 The sky view from street level in Berlin, Germany. The hemispheric
image shows the extent to which the sky is obscured by the surrounding buildings.
The path of the sun at different times of the year is plotted to show the loss of direct
sunlight at street level.
Source: F. Meier, TU Berlin
5.4 Innovations
In this section of the chapter, we explore the implementation
of strategies that utilize the four urban climate factors – urban
function, form, construction materials, and surface cover – to
achieve integrated mitigation and adaptation (see Figure 5.1).
5.4.1 Transportation, Energy, and Density
Since the middle of the 20th century, built environments the
world over have tended to increase outward from central cities,
consuming great swaths of previously undeveloped land while
reinvestment in city centers falters. The infrastructure network
needed to maintain this sprawling development pattern, par-
ticularly roads, has resulted in development that is land and
infrastructure inefcient. It has also led to increased reliance
on motor vehicles to get from one place to another. This reli-
ance on motor vehicles has consequently led to a signicant
increase in vehicle miles (or kilometers) traveled (VMT), a
concomitant increase in GHG emissions, and an amplication
of the UHI effect through increased imperviousness, reduced
green cover, and enhanced waste heat emissions (Stone et al.,
2010). By developing in a denser, more compact form that
mixes land use and supports mass transit use, cities may begin
to reverse these trends (see Figure 5.8) (see Chapter 13, Urban
Transportation).
If cities are to reduce VMT, then they must change their
sprawling development pattern into one that relies on compact
development. This focuses on regional accessibility through
multiple transportation modes, including walking and bicycling,
and clustered land-use patterns incentivized with vehicle distance
traveled–based fees. At the core of this strategy is transit-oriented
development (TOD) (Zheng and Peeta, 2015). TOD is compact,
pedestrian-friendly development that incorporates housing,
retail, and commercial growth within walking distance of public
transportation, including commuter rail, light rail, ferry, and
bus terminals (see Figure 5.8). It has become an essential and
sustainable economic development strategy that responds to
changing demographics and the need to reduce GHG emissions
and health-related impacts.
Changing the built form from conventional suburban
sprawl to compact, walkable, mixed-use and transit-oriented
neighborhoods reduces travel distances (VMT). The meta-
study Effects of the Built Environment on Transportation:
Energy Use, Greenhouse Gas Emissions, and Other Factors,
prepared by the National Renewable Energy Laboratory
and Cambridge Systematics, Inc. (March 2013), notes that
residents of compact, walkable neighborhoods have about
20–40% fewer VMTs per capita, on average, than residents
of less-dense neighborhoods. Other studies have found a dou-
bling of residential densities in U.S. cities to be associated
with a 5–30% reduction in VMT (Gomez-Ibanez et al., 2009;
Stone et al., 2010).
ARC3.2 Climate Change and Cities
148
Not only does compact, walkable TOD lower VMT, but it also
requires less energy (see Chapter 12, Urban Energy). As Nolon
(2012) reports, residential and commercial buildings used an
extraordinary amount of electricity and energy in the past genera-
tion. In 2008, U.S. residential and commercial buildings consumed
29.29 quadrillion BTUs, which represented 73.2% of all electric-
ity produced in the United States (Nolon, 2012). By 2035, the U.S.
Department of Energy estimates that residential and commercial
buildings will use 76.5% of the total electricity in the United States
(Nolon, 2012). This energy consumption is also highly inefcient
due to the systems used to produce and transmit it.
Two-thirds of the energy used to produce electricity in the
United States is vented as waste heat that escapes into the atmo-
sphere during generation and contributes to UHI formation
(Nolon, 2012). Additionally, up to 15–20% of the net energy
produced at these plants is then lost during electricity trans-
mission (Nolon, 2012). By increasing the density of the built
environment and reducing the distances that both electricity
and people must travel, energy efciency is notably increased
in compact, transit-centered development (see Figure 5.8). As
discussed in Jonathan Rose Companies (2011), a single-family
home located in a compact, transit-oriented neighborhood uses
38% less energy than the same size home in a conventional sub-
urban development (149 million BTU/year versus 240 million
BTU/year).
Because compact development reduces VMTs and is more
energy efcient, it also lessens GHG and waste heat emis-
sions (see Figure 5.9). In a 2010 report to Congress, the U.S.
Department of Transportation concluded that land-use strategies
relying on compact, walkable, TOD could reduce U.S. GHG
emissions by 28–84 million metric tons carbon dioxide equiva-
lent (CO
2
-eq) by the year 2030. Benets would grow over time to
possibly double that amount annually in 2050 (U.S. DOE, 2010).
Integrated mitigation and adaptation in urban planning and
urban design can be successfully implemented through the con-
guration of low-carbon compact settlements congured for
local microclimates. The mitigation perspective focuses on com-
pact TOD prototypes recongured as low-carbon ecodistricts
(see Box 5.3). The integrated mitigation and adaptation para-
digm also addresses stormwater runoff and the UHI in high-den-
sity zones through material composition, urban morphology, and
ecosystem services. This paradigm effectively “locks-in” long-
term resilience.
The efcient use and recycling of energy and resources is
a cornerstone of a resilient city and should be integral to the
concept of integrated mitigation and adaptation, along with
other climate-management strategies. This suggests the need
for two levels of integrated mitigation and adaptation: pas-
sive and active. Passive Integrated Mitigation and Adaptation
Figure 5.8 Efficiency of urban systems.
Source: Jeffrey Raven, 2016
Chapter 5 Urban Planning and Urban Design
149
Figure 5.9 Metropolitan region containment index (1995–2005).
Source: Philipp Rode, 2012
Denver
Washington
Minneapolis
Dallas
Baltimore
Philadelphia
San Francisco
Houston
Chicago
Prague
Frankfurt
Portland
Hamburg
Berlin
Paris
Oslo
Stockholm
R square = 0.503
Helsinki
London
Brussels
0
5
10
15
20
25
Greenhouse Gas emissions per capita (Mt CO
2
eq)
–3% –2% –1% 0% 1%
Metropolitan region containment index (1995–
2005)
(difference in population growth rates between core and belt)
Box 5.3 Multisectoral Synergies for Transit-Oriented, Low-Carbon Districts
Multisectoral approaches that integrate land use, mass tran-
sit, green buildings, and green districts to promote healthy,
climate-resilient cities can be described as transit-synergized
development (TSD). TSD leverages the greater scale, density,
and economic value of transit-oriented development (TOD;
nominal 1 km
2
urban districts around transit nodes) to create
compact, vibrant, mixed-use communities that increase urban
efciency and reduce transport-related energy use, conges-
tion, pollution, and greenhouse gas (GHG) emissions. As a
co-benet, the more compact development at the heart of TSD
reduces pressures on interstitial spaces between transit corri-
dors that can provide critical green infrastructure for managing
climate impacts within dense urban zones. In China, promot-
ing mass transit could generate up to 4 QBTUs (4.2 Exajoules)
in energy savings per year (McKinsey Global Institute, March
2009). A number of cities in Canada and the United States
are beginning to retrot their urban fabric through TOD, and
others in Latin America – most notably Curitiba, Brazil – have
successfully used bus rapid transit (BRT) as a centerpiece of
wider urban revitalization (Lindau et al., 2010).
TSD is a “node and network” model of sustainable urbanism.
At each transit node, TSD combines passive and active green
building design (high-performance envelope and mechanical,
electrical, and plumbing [MEP] systems) with passive and
active green district design (integrated urban design and
advanced district infrastructure). District infrastructure pro-
vides a platform for the reuse and recycling of energy and
resources among the buildings within the district. It is also
a platform for innovation and “’forward integration’” of new
technologies, important attributes of a robust, resilient, and
adaptive community (Lee, 2012).
At the network level, the transit nodes collectively provide
diversity, redundancy, and synergy, effectively transforming
the transit network into a framework for a robust, resilient,
and adaptive city (Walker and Salt, 2006).
Box Figure 5.3 Figure 1 Transit-synergized development concept.
Source: iContinuum Group
ARC3.2 Climate Change and Cities
150
Figure 5.10 Urban form and layout.
Source: Jeffrey Raven, 2016
(PIMA) includes climate-responsive designs such as green
cover, reective ground surface, natural ventilation, and solar
orientation. PIMA represents good design practice and should
be the basic design strategy for all buildings and urban areas.
In high- density urban districts, however, Active Integrated
Mitigation and Adaptation (AIMA) may be required. AIMA
deploys advanced building systems and district infrastructure
such as integrated building energy management, renewable
energy, energy storage, district energy systems, water recy-
cling, and on-site wastewater treatment that actively reduce
energy and climate impacts.
District-scale AIMA infrastructure can be inherently more
cost-effective to “upgrade” than individual building systems
and can therefore better “climate-proof” the built environment
against changing conditions. An example of this is Singapore’s
Marina Bay District Cooling System, which is envisioned as an
“energy platform that enables forward integration of new energy
technologies” (Tey Peng Kee, Managing Director, Singapore
District Cooling Pte Ltd.; Interview, 2012).
5.4.2 Climate-Resilient Urban Form
Urban morphology is dened as the three-dimensional form
and layout of the built environment and settlement pattern. From
regional, urban, and district scales to ner-grained street grids
that promote walkability and social cohesion, climate-resilient
planning and design strategies include conguration of urban
morphology inuenced by solar design, urban ventilation, and
enhanced vegetation (see Figure 5.10). There are almost innite
combinations of different climate contexts, urban geometries,
climate variables, and design objectives. A starting point in any
project is to assess the micro- and macroclimatic characteristics
of the site, an exercise that will indicate appropriate bioclimatic
design strategies (Brophy et al., 2000). As the climate heats up,
compact communities offer attractive alternatives to suburban
sprawl by featuring comfortable, healthy microclimates with
comparable natural amenities.
Wind velocities in cities are generally lower than those in the
surrounding countryside due to the obstruction to air ow caused
by buildings. In dense, compact communities, natural ventilation
is challenging during warm months, often leading to increased
cooling demand. This is partly because natural ventilation sys-
tems require very little energy but may need more space to
accommodate low-resistance air paths (Thomas, 2003). Built-up
areas with tall buildings may lead to complex air movement
through a combination of wind channeling and resistance, often
resulting in wind turbulence in some areas and concentrated
pollution where there are wind shadows (Brophy et al., 2000).
In general, denser developments result in a greater reduction in
wind speeds but proportionally increased turbulence. Compact
developments have less heat loss because there is generally less
surface area for the volume enclosed due to shared wall space
(Thomas, 2003).
Chapter 5 Urban Planning and Urban Design
151
Street-level ventilation for warm and humid climates is key,
using approaches that do not necessarily require changing mor-
phology. A simulation exercise of the likely urban warming
effects of the planned urban growth trajectories in the warm,
humid city of Colombo, Sri Lanka, indicates that there are sig-
nicant differences in the likely warming rates between different
urban growth trajectories (Emmanuel et al., 2007). A moderate
increase in built cover (at an LCZ class of compact midrise)
appears to lead to the least amount of warming. At the neighbor-
hood scale, streets oriented to the prevailing wind directions with
staggered building arrangements together with street trees appear
to offer the best possibility to deal with urban warming in warm
humid cities. The combined approach could eliminate the warm-
ing effect due to the heat island phenomenon. The Hong Kong
example (see Case Study 5.3) illustrates how existing high-rise
districts can be retrotted to exploit passive urban ventilation.
Exploiting prevailing breezes is a key factor in implementing
district-wide passive cooling strategies (see Figure 5.10). Wind
affects temperature, rates of evaporative cooling, and plant tran-
spiration and is thus an important factor at a microclimatic level
(Brophy et al., 2000). Urban morphology is responsible for vary-
ing the “porosity” of the city and the extent of airow through
it, and it is a lynchpin for using passive cooling to reduce energy
loads in the built environment (Smith et al., 2008). Wind ow
across evapotranspiring surfaces and water bodies provide cool-
ing benets. The morphology and surface roughness of the built
environment has signicant impacts on the effectiveness of
urban ventilation.
Passive methods to increase comfort and reduce energy loads
through solar design include orienting street and public space
layout to reduce solar gain during hot months, shading through
the conguration of adjacent vegetation, orienting neighborhood
congurations to the sun’s path to maximize daylight in ground
oor living rooms, placing tall buildings to the north edges of a
neighborhood to preserve solar potential for photovoltaic arrays,
varying building heights and breaks in the building line to reduce
shadowing and increase solar access during cold months, and
maximizing use of cool surfaces and reective roofs in hot
climates. Figures 5.10, 5.11, and the Masdar example (Case
Study 5.4) illustrate these approaches.
5.4.3 Construction Materials
Increasing the surface reectivity or albedo of urban materials
is a well-established urban heat management strategy. Due to the
darkly hued paving and roong materials distributed throughout
cities, a larger quantity of solar energy is often absorbed in cities
than in adjacent rural areas with higher surface reectivity, thus
contributing to a lower albedo (see Chapter 2, Urban Climate
Science). Unable to compensate for an enhanced absorption of
solar energy through an increase in evapotranspiration, a larger
percentage of this absorbed energy is returned to the atmosphere
as sensible heat and longwave radiation, raising temperatures
(see Figure 5.11).
Recognition of the potential to measurably cool cities through
the application of highly reective coatings to roong surfaces
Figure 5.11 Surface reflectivity.
Source: Jeffrey Raven, 2016
ARC3.2 Climate Change and Cities
152
Case Study 5.1 Green Infrastructure as a Climate Change Adaptation Option for
Overheating in Glasgow, UK
Rohinton Emmanuel
Glasgow Caledonian University, Glasgow
From its medieval ecclesiastical origins, Glasgow (originally Glaschu
– ‘dear green place’) expanded into a major port in the 18th century,
and, with the advent of the Industrial Revolution, added a massive
industrial base to its already well- developed built fabric. However,
the success of its industrial base could not withstand the pressures
of globalization, and, by the early 20th century, the city had begun to
lose population. This decline appears to have been arrested in recent
years. The long history of growth, decline, and regrowth provides
Glasgow a historic opportunity to recreate its “green” past.
Emmanuel and Kruger (2012) showed that even when urban growth
had subsided, Glasgow’s local warming that results from urban
morphology (increased built cover, lack of vegetation, pollution,
anthropogenic heat generation) continues to generate local heat
islands. Such heat islands are of the same order of magnitude as
the predicted warming due to climate change by 2050. And the
microscale variations are strongly related to local land cover/land-
use patterns.
MITIGATING URBAN OVERHEATING
An option analysis exploring the role of green infrastructure (land-
scape strategies) in and around the Glasgow (Glasgow Clyde Valley,
GCV) Region revealed the following (Emmanuel and Loconsole,
2015):
1. Green infrastructure could play a signicant role in mitigating
the urban overheating expected under a warming climate in the
GCV Region.
2. A green cover increase of approximately 20% over the present
level could eliminate a third to a half of the expected extra urban
heat island (UHI) effect in 2050.
3. This level of increase in green cover could also lead to local
reductions in surface temperature by up to 2°C.
4. More than half of street users would consider a 20% increase in
green cover in the city center to be thermally acceptable, even
under a warm 2050 scenario.
ACHIEVING GREEN COVER
Not all green areas contribute equally to local cooling, nor are
they equal in their other environmental and sustainability benets.
Recognizing this, planners have begun to develop weighting systems
that capture the relative environmental performance of different types
of green cover. The most widely used among these is the Green Area
Ratio (GAR) method (Keeley, 2011). GAR is currently implemented
in Berlin and has been adapted in Malmo (Sweden), several cities in
South Korea, and Seattle (USA). Elements of GAR include:
Impermeable surfaces (i.e., surfaces that do not allow the inltration
of water)
Includes roof surfaces, concrete, asphalt and pavers set upon
impermeable surfaces or with sealed joints). = 0.0
Impermeable surfaces from which all storm water is infiltrated
on property
Includes surfaces that are disconnected from the sewer
system. Collected water is instead allowed to inltrate on site in
Keywords Urban overheating, green
infrastructure, green area ratio
method, planning and design
Population
(Metropolitan Region)
a
606,340 (National Records of
Scotland, 2016)
Area
(Metropolitan Region)
b
3,345.97 km² (Ofce for National
Statistics, 2012)
Income per capita US$42,390 (World Bank, 2017)
Climate zone Cfb – Temperate, without dry
season, warm summer
(Peel et al., 2007)
a
Counting the following Local Authority areas: East Dunbartonshire; East Renfrewshire;
Glasgow City; Inverclyde; North Lanarkshire; Renfrewshire; South Lanarkshire; West
Dunbartonshire
b
Counting the following Local Authority areas: East Dunbartonshire; East Renfrewshire;
Glasgow City; Inverclyde; North Lanarkshire; Renfrewshire; South Lanarkshire; West
Dunbartonshire
Case Study 5.1 Figure 1 Summer daytime temperature.
y=0.0007x
2
– 0.0295x – 0.0143
R
2
= 0.9212
–0.45
–0.40
–0.35
–0.30
–0.25
–0.20
–0.15
–0.10
–0.05
0.00
0.05
0.10
–5 0510 15 20
25
Change in Air Temperature (
o
C)
Green cover (%)
Case Study 5.1 Figure 2 Extra green cover needed in reduction against
green cover in Glasgow City Center to mitigate overheating.
152
Chapter 5 Urban Planning and Urban Design
153
a swale or rain garden. Guidelines for preventing groundwater
and soil contamination must be followed. = 0.2
Nonvegetated, semi-permeable surfaces
Includes cover types that allow water inltration, but do not
support plant growth. Example include brick, pavers and
crushed stone. = 0.3
Vegetated, semi-permeable surfaces
Includes cover types that allow water inltration and integrate
vegetation such as grass. Examples include wide-set pavers
with grass joints, grass pavers, and gravel-reinforced grassy
areas. = 0.5
Green façades
Includes vines or climbing plants growing (often from ground)
on training structures such as trellises that are attached to a
building. The façade’s area is measured as the vertical area
the selected species could cover after 10 years of growth up
to a height of 10 meters; window areas are subtracted from the
calculation. = 0.5
Extensive green roofs
Includes green roofs with substrate/soil depths of less than 80
centimeters. However, Berlin excludes green roofs constructed
on high-rise buildings. = 0.5
Intensive green roofs and areas underlain by shallow subterranean
structures
Includes green roofs with substrate/soil depths of greater than
80 centimeters. This category includes subterranean garages.
= 0.7
Vegetated areas
Any area that allows unobstructed inltration of water without
evaluation of the quality or type of vegetation present. Examples
range from lawns to gardens and naturalistic wooded areas. =
1.0
A system of target setting is initially required that takes into
account the severity of the environmental risk faced by a partic-
ular urban neighborhood. Once the target green cover is deter-
mined, the above-indicated weighting is used to develop alternate
green infrastructure scenarios.
Case Study 5.1 Table 1 shows alternate approaches to a 20%
increase in green cover in Glasgow city center. These employ an
urban park, street trees, roof gardens, façade greening, or combi-
nations of these.
Case Study 5.1 Table 1 Alternative approaches to increasing green cover by 20% in Glasgow City Center.
Scenario
Permeable
vegetated
area (m
2
)
Street trees
(Nos.)
Intensive
roof gardens
(m
2
)
Extensive
roof gardens
(m
2
)
Green
façades
1. A single large park 1,056
2. Street trees only 528
3. 50% of additional greenery in street trees, balance
intensive roof gardens
264 755
4. 50% of additional greenery in street trees, balance
extensive roof gardens
264 1,056
5. Mix of intensive (50%) and extensive (50%) roof gardens 755 1,056
6. 50% of all “sun facing (i.e., South & West) façade covered
by green façades
1,268
and streets has led to the development of new product lines
known as “cool” roong and paving materials. For roong sur-
faces concealed from ground view, such as atop a at indus-
trial building, very high-albedo, cool material coatings can be
applied to reect away a substantial percentage of incoming
solar radiation. Industry analyses of these materials have found
that the surface temperature of roong materials can be reduced
by as much as 50°F/10°C during periods of intense solar gain
(Gafn et al., 2012).
To explore the extent to which cool materials could reduce
temperatures not only within the treated buildings themselves but
also throughout the ambient urban environment, scientists at the
Lawrence Berkeley Labs and the Columbia Center for Climate
Systems Research have modeled extensive albedo enhancement
strategies (Rosenzweig et al., 2014). Measured on a scale of
0 to 1, average surface albedos in U.S. cities tend to range from
0.10 to 0.20, much lower than the albedos of 0.6 to 0.8 associ-
ated with cool roong and paving materials. In densely settled
districts such as Manhattan, the potential to raise average albe-
dos is great, but all cities can enhance their reectivity through
the use of higher albedo materials in routine resurfacing over
time. Finding optimal values of reectivity adjustment, rather
than an all-out pursuit of maximum attainable values over large
swaths of surface areas, will limit potential hydroclimatic trade-
offs while still attaining temperature reduction goals (Georgescu
et al., 2014; Jacobson and Ten Hoeve, 2012).
An important advantage of albedo enhancement over other
urban climate management strategies is its relatively low cost.
Cool roong treatments can be applied to low-sloping roofs
for a cost premium of between US$0.05 to US$0.10 per square
ARC3.2 Climate Change and Cities
154
foot, raising the cost of a 1,000 square foot roong project by
as little as US$100. Balanced against this low initial cost are
annual energy savings estimated by the U.S. Environmental
Protection Agency (EPA) to be about U.S. $0.50 per square foot,
an estimate accounting for potentially greater winter heating
costs (U.S. EPA, 2008). Also advantageous is the immediacy
of benecial returns from cool materials strategies, especially
in semi-arid areas where the use of water for vegetation is not
sustainable. In contrast to tree planting and other vegetative pro-
grams through which maximum cooling benets are not realized
until plants reach maturity, high-albedo coatings yield maximum
benets upon installation, with benets diminishing somewhat
thereafter with weathering, and as roofs become soiled.
5.4.4 Green and Blue Infrastructure
The interaction of green and blue components in the urban
environment links together integrated mitigation and adaptation
strategies at different scales – from buildings and open spaces
design to landscape design and metropolitan region planning –
and can yield many co-benets (see Figure 5.12) (see Chapter 8,
Urban Ecosystems). A comprehensive climate-based design sup-
ports developing and maintaining a network of green and blue
infrastructure integrated with the built environment to conserve
ecosystem functions and provide associated benets to human
populations (STAR Communities, 2014). Urban planning and
urban design strategies focusing on green infrastructure and
Case Study 5.2 Adapting to Summer Overheating in Light Construction with
Phase-Change Materials in Melbourne, Australia
Jun Han
1,2
, Xiaoming Wang
1
, Dong Chen
1
1
Commonwealth Scientific and
Industrial Research Organization (CSIRO), Melbourne
2
Heriot-Watt University, Dubai Campus
Keywords Thermal comfort, residential
buildings, passive cooling, phase
change materials, mitigation and
adaptation, planning and design
Population
(Metropolitan Region)
4,258,000 (UN, 2016)
Area
(Metropolitan Region)
9,999.5 km² (Australian Bureau of
Statistics, 2013)
Income per capita US$54,420 (World Bank, 2017)
Climate zone Cfb – Warm temperate, fully humid,
warm summer (Peel et al., 2007)
Melbourne, ranked one of the most livable cities around the world
since 2011, is the capital city in the state of Victoria and the second
most populous city in Australia (2006 Census QuickStats). It has a
population of 3.99 million living in the greater metropolis (Australian
Bureau of Statistics, 2013).
As a result of recent rapid urbanization, the city has undergone
an outward expansion. The recent construction boom in both the
Central Business District and nearby suburbs has led to a signi-
cant change in land use. This may imply that more buildings will be
constructed in the near future and, consequently, more greenhouse
gas (GHG) emissions from building operations. Change in land use
such as replacement of green space by construction and increas-
ing concrete or paved roads can be anticipated if appropriate urban
planning for climate adaptation is lacking.
Existing studies have recognized the challenges facing current
urbanization posed by the urban heat island (UHI) effect and global
warming. Exacerbated thermal conditions in the urban built environ-
ment and increasing human health issues can be expected without
proper intervention. In this regard, we now face challenges not only
in designing low-energy buildings to reduce GHG emissions for mit-
igating global warming but also to meet thermal comfort require-
ments without sacricing indoor environment quality (IEQ) to actively
adapt to climate change.
However, modern construction methods introduce more lightweight
buildings. These methods employ offsite, prefabrication strategies to
reduce construction time. Consequently, there is a potential risk of
overheating and deteriorated thermal comfort conditions with light-
weight construction products, which are likely to be exacerbated in
a warming climate. Occupant thermal comfort in lightweight build-
ings therefore is receiving increasing attention among architects and
designers.
In this Case Study, phase-change materials (PCMs) are used as a
heat sink to absorb heat from the sun during the day and to reduce
rapid room temperature rise due to added thermal stability. To
examine the effectiveness of a lightweight building using PCMs, a
one-dimensional numerical model was developed and solved by
an enthalpy
1
method with an explicit scheme. The performance of
PCMs for cooling a lightweight building with a brick veneer residen-
tial wall during the hot summer of 2009 in Melbourne was predicted
numerically. The study reveals that application of PCMs in light-
weight buildings could achieve better thermal comfort and energy
savings in summer.
Bio-based PCM, applied instead of conventional wall insulation is
made with a mix of soy-based chemicals that change from liquid to
solid and vice versa at specic melting or solidication temperatures.
The advantage of using PCMs is increased thermal storage capacity.
It is estimated that about 30% of heating and cooling costs would
be reduced.
Case Study 5.2 Figure 1 depicts the geometric conguration of the
thermally enhanced brick veneer wall with PCMs and its original con-
guration without PCMs before modication. The thermally enhanced
PCM wall is composed of 110-millimeter brickwork, 40-millimeter air
gap, 20-millimeter PCMs, and 10-millimeter plasterboard.
1 Enthalpy is the thermodynamic quantity equivalent to the total heat content of a system. It is equal to the internal energy of the system plus the product of pressure
and volume.
Chapter 5 Urban Planning and Urban Design
155
To understand the occupancy comfort level, the peak wall tem-
peratures were considered when evaluating the performance of
the PCM wall. The interior surface temperatures for the west-fac-
ing walls are compared for the PCM brick veneer wall and for the
reference wall without PCM, as shown in Case Study 5.2 Figure
2. It was found that the interior surface temperature of the PCM
wall in the day is lower than the conventional wall due to the
presence of PCM heat storage during the daytime. The max-
imum peak temperature of the conventional brick veneer wall
reached almost 48°C, whereas the maximum of the PCM wall
was around 32°C. It is generally believed that lower surface tem-
peratures result in greater occupant thermal comfort and energy
savings in summer. A significant peak cooling load reduction
therefore would be expected.
The successful testing of PCMs in lightweight building materials in
the weather conditions of Melbourne demonstrated the effectiveness
and validity of both mitigation and adaptation strategies. Combined
with other sustainable energy technologies as an integrated approach
for climate change mitigation and adaptation, PCMs could be useful
for other cities with similar conditions.
Case Study 5.2 Figure 2 Surface temperature profiles of two west-facing
walls with and without phase-change materials.
36 40 44 48 52 56 60 64 68 72 76
16
20
24
28
32
36
40
44
48
52
56
liquidus temperature 26
o
C
phase change
West facing wall
Without PCM
With PCM
TEMPEARATURE (
o
C)
TIME (HOUR)
without PCM
with PCM
Case Study 5.2 Figure 1 Schematic of lightweight brick veneer wall with integrated phase-change materials (left), and without (right).
Roof membrane
Roof membrane
Cant strip
Brickwork 110 mm
Air gap 40 mm
wall without PCMswall with PCMs
Foundation
Foundation
Brickwork 110 mm
Air gap 40 mm
Cant strip
Plasterboard 10 mm
Plasterboard 10 mm
PCMs 20 mm
Fascia
Fascia
sustainable water management help restore interactions between
built and ecological environments. This is necessary to improve
the resilience of urban systems, reduce the vulnerability of socio-
economic systems, and preserve biodiversity (UNEP, 2010).
Integration of water management with urban planning and
urban design represents an effective opportunity for climate
change adaptation (UNEP, 2014) (see Figure 5.13). This has been
demonstrated by the emerging Water Sensitive Urban Design
(WSUD) approach (Ciria, 2013; ARUP, 2011; Flörke et al., 2011;
Hoyer et al., 2011; BMT WBM, 2009). All the elements of the
water cycle and their interconnections are considered to achieve
together an outcome that sustains a healthy natural environment
while addressing societal needs and reducing climate-related
risks (Ciria, 2013). The implementation of integrated water cycle
management as adaptive design strategy should be based on a
ARC3.2 Climate Change and Cities
156
Figure 5.12 Surface cover.
Source: Jeffrey Raven, 2016
Building greening
Vegetated soils
Draining surfaces
Raingardens Rainwater harvesting
and recycling
Grey water collection and
sedimentation systems
Phytoremediation systems
Figure 5.13 Green and blue infrastructure design: building/open space scale, Naples, Italy.
Source: Cristina Visconti and Mattia Leone
Chapter 5 Urban Planning and Urban Design
157
Case Study 5.3 Application of Urban Climatic Map to Urban Planning of High-Density
Cities: An Experience from Hong Kong
Edward Ng and Chao Ren
Chinese University of Hong Kong
Lutz Katzschner
University of Kassel
Hong Kong is located on China’s south coast and situated in a sub-
tropical climate region with hot and humid summers. As a high-den-
sity city with a population of 7.3 million living on 25 square kilometers
of land, Hong Kong has a hilly topography and 40% of the territory is
classied as country-park, where development is prohibited; hence
only about 25% is built-up. Due to limited land area and increas-
ing land prices, taller and bulkier buildings with higher building
plot ratios, very limited open space, large podium structures, and
high building-height-to-street ratios have been built. These tall and
wall-like buildings in the urban areas block the incoming wind and
sea breezes. This leads to a worsening of urban air ventilation and
exacerbates the city’s urban heat island (UHI) intensity. The num-
ber of very hot days (maximum air temperature greater than 33°C)
and very hot nights (maximum air temperature greater than 28°C)
has increased, whereas the mean wind speeds recorded in urban
areas over the past 10 years have decreased. This intensies uncom-
fortable urban living, heat stress, and related health problems and
increases energy consumption.
The Hong Kong Observatory has conducted studies that note
that Hong Kong’s urban temperature has been increasing over the
decades (Leung et al., 2004). Good urban air ventilation is an effective
adaptation measure for the UHI effect and rising temperatures under
climate change. However, Hong Kong’s urban wind environment is
deteriorating due to intensive urban development that increases the
surface roughness and blocks the free ow of air, leading to weaker
urban air ventilation and higher urban thermal heat stress. Higher air
temperatures and a higher occurrence and longer duration of heat
waves will have a severe impact on urban living; therefore, there is
a need to plan and design the city to optimize urban climatic con-
ditions and urban air ventilation based on a better understanding of
the UHI phenomenon and the urban climate to reduce the impact of
urban climate and climate change.
The Planning Department of the Hong Kong SAR Government pro-
duces the Hong Kong Urban Climatic Map System (PlanD, 2012) to
provide an evidence-based tool for planning and decision making.
Keywords Heat island effect, high density,
urban climatic map, ventilation
corridors, planning and design
Population
(Metropolitan Region)
7,310,000 (GovHK, 2016)
Area
(Metropolitan Region)
1,105.7 km
2
(GovHK, 2016)
Income per capita US$60,530 (World Bank, 2017)
Climate zone Cwa – Monsoon-inuenced humid
subtropical, hot summer
(Peel et al., 2007)
Case Study 5.3 Figure 1 The Urban Climatic (Planning Recommendation) Map of Hong Kong.
Source: Hong Kong Planning Department, HKSAR Government
ARC3.2 Climate Change and Cities
158
The Urban Climatic (Planning Recommendation) Map classies
Hong Kong’s urban and rural areas into ve planning recommenda-
tion zones. General planning advice is given for each zone. Detailed
advice is contained in the map’s accompanying notes.
Based on a scientic understanding of the Hong Kong Urban
Climatic Maps, future planning scenarios may be tested and effec-
tive adaptation measures (including advice on building density, site
coverage, building height, building permeability, and greening) may
be developed (PlanD, 2012). Prescriptive guidelines and perfor-
mance-based methodologies in the Hong Kong Urban Climatic Map
System provide further quantication. With a better understanding
of urban climate, planners can balance various planning needs and
requirements when making their nal decisions.
Based on an understanding of the Urban Climatic Maps, the follow-
ing planning and design measures should thus be taken into account
in project planning and in the formulation of development param-
eters. They could help improve the urban climate and reduce the
impact of climate change:
The UC-ReMap provides a strategic and comprehensive urban cli-
matic planning framework and information platform for Hong Kong
that can be also applied to other high-density cities. It helps to
clarify and identify appropriate planning and design measures for
the formulation of planning guidelines on matters related to urban
climate and climate change, and it provides a strategic urban plan-
ning and development process for future development (e.g., max-
imizing the adaptation opportunities within urban climate planning
zones (UCPZs) 3, 4, and 5) and accommodating comprehensive
new development areas in UCPZ 2 with prudent planning and
building design measures (PlanD, 2012). It also provides an urban
climatic planning framework for reviewing outline zoning plans and
formulating suitable planning parameters.
Case Study 5.3 Table 1 Planning and design measures to be taken into
account in project planning.
Planning
parameters Recommendations
Building volume Site plot ratio of 5 or less. Higher plot area
must be adapted using other planning
parameters.
Building
permeability
2533% of the project site’s frontal
elevation. Lower permeability must be
adapted using other planning parameters.
Building site
coverage
70% of the site area. Higher site coverage
must be adapted using other planning
parameters.
Air paths and
breezeways
Open spaces must be linked with
landscaped pedestrianized streets from
one end of the city to the other end in the
direction of the prevailing wind.
Building heights Vary building heights so that there is a
mixture of building heights in the area with
an average aggregated differential of 50%.
Greenery 2030% of tree planting preferably at
grade, or essentially in a position less than
20m from the ground level. Trees with large
canopy and a leaf area index of more than 6
are preferred.
(a) (b)
Case Study 5.3 Figure 2 Building volume density study of the area (left); open spaces (blue) and air paths (red lines) suggested for the area (right),
Hong Kong.
Chapter 5 Urban Planning and Urban Design
159
Case Study 5.4 An Emerging Clean-Technology City: Masdar, Abu Dhabi,
United Arab Emirates
Gerard Evenden, David Nelson,
Irene Gallou
Foster + Partners, London
Keywords Carbon-free technologies, nearly-
zero energy buildings, microclimate
comfort, planning and design
Population
(Metropolitan Region)
1,179,000 (UN, 2016)
Area
(Metropolitan Region)
803 km
2
(Demographia, 2016)
Income per capita US$72,850 (World Bank, 2017)
Climate zone Bwh – Arid, desert, hot
(Peel et al., 2007)
Masdar City is an emerging clean-technology cluster located in what
aims to be one of the world’s most sustainable urban developments
powered by renewable energy. The project continues to be a work in
progress. Located about 17 kilometers from downtown Abu Dhabi, the
area is intended to host companies, researchers, and academics from
across the globe, creating an international hub focused on renewable
energy and clean technologies. The master plan is designed to be
highly exible, to benet from emergent technologies, and to respond
to lessons learned during the implementation of the initial phases.
Expansion has been anticipated from the outset, allowing for growth
while avoiding the sprawl that besets so many cities (Bullis, 2009;
Manghnani and Bajaj, 2014).
The aim of a new development settlement characterized by
a comfortable living environment in such an extreme desert
climate required the implementation of adaptive design strategies to
effectively respond to issues related to scarcity of precipitation, sea-
sonal drought, high temperatures, and wide daily temperature range.
The carbon-free new development sets new standards for climate
change mitigation in arid countries through the adoption of nearly
zero energy standards and building-integrated energy production
from renewable sources.
The design concept explores the adoption of sustainable technolo-
gies and planning principles of traditional Arab settlements combined
with contemporary city spatial-functional needs and state-of-the-art
technological solutions to develop a carbon-neutral and zero-waste
community despite the extreme climatic conditions. The quest for
a mixed-use, low-rise, and high-density development, entirely car-
free, with a combination of personal and public transit systems and
pedestrian areas, is achieved through an extensive use of traditional
solutions such as narrow streets and optimal orientation; shaded
windows; exterior walls and walkways to control solar radiation;
thick-walled buildings to maximize thermal mass and reduce energy
consumption; courtyards and wind towers for natural ventilation; and
Case Study 5.4 Figure 1 Masdar, Carbon-Neutral Development Case Study: Abu Dhabi, UAE.
Source: Foster + Partners
ARC3.2 Climate Change and Cities
160
vegetation design with optimized water management to improve the
local micro-climatic conditions of open spaces.
The northeast-southwest orientation of the city makes best use
of the cooling night breezes and lessens the effect of hot daytime
winds. Green parks separate built-up areas, not only to capture and
direct cool breezes into the heart of the city, but also to reduce solar
gain and provide cool pleasant oases throughout the city. The intelli-
gent design of residential and commercial spaces, based on building
standards currently set by internationally recognized organizations,
reduces demand for articial lighting and air conditioning. Such
standards, adapted to the local climatic context, contributed to the
development of Abu Dhabi’s “Estidama” rating system for sustain-
able building (Abu Dhabi Estidama Program, 2008).
Carefully planned landscape and water features lower ambient tem-
peratures while enhancing the quality of the street. The elimination
of cars and trucks at street level not only makes the air cleaner for
pedestrians, but also allows buildings to be closer together, provid-
ing more shade but allowing maximum natural light. The placement
of residential, recreational, civic, leisure, retail, commercial, and light
industrial areas across the master plan, along with the public trans-
portation networks, ensures that the city is pedestrian friendly and a
pleasant and convenient place in which to live and work.
Case Study 5.4 Figure 2 Masdar wind tower.
Source: Foster + Partners
(a) (b)
Case Study 5.4 Figure 3 Thermal imaging comparing streetscapes in Abu Dhabi and Masdar City.
Source: Foster + Partners
Chapter 5 Urban Planning and Urban Design
161
dual denition of water as both resource and hazard (see Chapter
9, Coastal Zones). In the framework of urban design and plan-
ning, managing water as a resource addresses environmental
quality and microclimate conditions of urban spaces, availability
of water, rebalancing of ecosystem exchange, and the hydrolog-
ical cycle in buildings and open spaces (ARUP, 2011). When
considering water as a hazard, design should focus on the control
of water discharge through runoff management and inltration
measures able to achieve wastewater retention and employing a
decentralized sewage system.
Best practices of adaptation-driven urban policies worldwide
provide signicant examples of how the paradigm shift toward
water-sensitive and water-resilient cities allows for the implemen-
tation of an integrated approach that combines risk prevention with
a regeneration of urban fabric driven by adaptive design solutions
(Kazmierczak and Carter, 2010). The Sydney Water Sensitive
Urban Design (WSUD) Program and the post-Sandy “Rebuild by
Design” initiatives in New York demonstrate such practices.
Recirculation of water on site is among the main concepts
of a water-sensitive approach to urban design and planning.
It represents a key priority for enhancing water resilience and
requires an integrated set of complementary measures, including
decentralization of water discharge, harvesting and recycling,
draining, and vegetated surfaces, thus improving urban micro-
climate and ood prevention.
Evaporative cooling processes, fostered by the development
of green spaces in cities, allow for sustainable management
of the water cycle and a reduction of the UHI effect. Building
greening measures (sustainable roofs and vegetative facades)
reduce the amount of water owing into sewage systems, mit-
igate temperature extremes, provide thermal insulation, and
increase biodiversity in urban areas (Ciria, 2007b; Schimdt et
al., 2009; Nolde et al., 2007, Steffan et al., 2010; UNEP, 2012).
Green street scapes, including the use of permeable paving, pro-
vide shade and reduce thermal radiation.
Greening and permeable paving, as elements of storm- water
management, have the potential to retain water that is then
evaporated while delaying and reducing runoff (Scholz and
Grabowiecki, 2007). Such a design approach for urban open
spaces is strengthened by the integration of sustainable drainage
Wind towers, which can be found on both sides of the Arabic Gulf,
are a traditional form of enhancing thermal comfort within the court-
yard houses of the region. Masdar is exploring the possibility of using
the principle to passively ventilate undercroft spaces and the pub-
lic realm, thereby reducing the demand for mechanical ventilation
systems.
Urban devices from local vernacular architecture – such as colon-
nades, wind towers, green canopies, and fountains – can bring the
felt temperature down by 20 degrees compared to open desert.
Cumulatively, all of these design principles have the effect of pro-
longing the moderate season in the city.
The land surrounding the city will contain wind and photovol-
taic farms, as well as research elds and plantations, allowing the
community to be entirely energy self-sufcient (Foster+Partners,
2009).
The design process has been based on studies and simulations
of energy and thermal comfort, solar and wind analysis, material
heat gain, and thermal imaging. Field measurement studies (Case
Study 5.4 Figures 3 and 4) have been conducted to assess the
microclimatic performance of spaces in Masdar City. Infrared ther-
mal imaging was used in addition to hand-held equipment to track
variations in temperature in urban spaces and then compared
them to similar urban spaces in central Abu Dhabi and the des-
ert. The comparisons of images show the superior performance of
Masdar City due to the shade provided by built form, correct use of
materials, natural ventilation, and evaporative cooling strategies.
(a) (b)
Case Study 5.4 Figure 4 Masdar field studies in the desert.
Source: Foster + Partners
ARC3.2 Climate Change and Cities
162
systems (SuDS). This is a set of measures aimed at retaining
and inltrating storm water (bio-swales, rain gardens, retention
basins, bio-lakes, wetlands, rainwater harvesting systems). This
allows for the control of water discharge and reduces ood risk
(Ciria, 2007a, 2010; Charlesworth, 2010; Poleto, 2012) (see
Chapter 14, Urban Ecosystems).
The location and form of green infrastructure should be deter-
mined in relation to the built environment and aligned in relation
to natural systems, including water bodies, solar impacts, and
prevailing winds. A network of local microclimates can com-
prise small green spaces, planted courtyards, shaded areas, and
“urban forests” to moderate temperature, as demonstrated in the
Manchester, United Kingdom, Case Study. Vegetation should be
sited to maximize the absorption rate of solar radiation. Localized
water bodies can moderate temperature extremes through their
high thermal storage capacity and through evaporative cooling.
5.5 Steps to Implementation
A planning and design approach to urban climate intervention
should follow a four-phase approach: climate analysis mapping,
public space evaluation, planning and design intervention, and
post-intervention evaluation (see Figure 5.14).
5.5.1 Climate Analysis and Mapping
Considering climate in urban planning and urban design, the
rst step is to understand large-scale climatic conditions and
individual inner-city local climates, including their reciprocal
interactions (see Figure 5.14). Considerations include:
Regional occurrence and frequency of air masses exchange
(ventilation) and their frequencies;
Seasonal occurrence of the thermal and air quality effects of
urban climate (stress areas, insolation rates, shading conditions);
Regional presentation and evaluation of the impact area and
stress areas; and
Energy optimization of location based on urban climate anal-
ysis with regard to areas with heat load, cooler air areas, and
building density.
One also has to address sectoral planning (see Table 5.1).
Climate analyses and maps provide a critical rst step in iden-
tifying urban zones subject to the greatest impacts associated
with rising temperatures, increasing precipitation, and extreme
weather events (see Figure 5.14). A climate analysis map may
be developed in consecutive steps on the basis of spatial ref-
erence data. The spatial resolution is tailored to the planning
level (see Table 5.1). Commonly employed climate analysis
maps include urban heat hotspot and ood zone maps rou-
tinely employed in urban planning applications. Geographical
Information System (GIS) layers include topographical infor-
mation, buildings, roughness and greenery needed to create an
urban climate map (see Figure 5.14).
5.5.2 Evaluation of Public Space
Urban climate is an essential part of urban planning eval-
uation. Urban climate maps are increasingly used in the plan-
ning process for urban development as well as for open space
design. The public should be involved at all stages through the
Table 5.1 Meteorological scales for planning. Source: Lutz Katzschner
Instruments and Plans
Scale, Spatial
Resolution of Maps
Climate Analysis Components
(Air Quality and Human Biometeorology)
Regional
Planning
Regional Land-Use Plan 1:50,000 to 1:
100,000 100m
Meso-scale climate
Comprehensive pollution control maps
Thermal stress areas (overhead areas)
Ventilation lanes
Cool air production areas
Planning recommendation map
Urban
Planning
Land-Use
Preliminary Urban Land
Planning:
Land-use plan
1:5000 to 1:25,000,
25m to 100m
Meso-scale climate
Area-related ambient air quality maps
Air exchange
Thermal stress areas (overheated areas)
Planning recommendation map
Mandatory Urban Land
Planning:
Local development plan
Planning permission and
procedure
(1:1000), 2m to
10m
Micro-scale climate
Local ambient quality calculations for “most severely affected areas”
Neighborhood considerations
Air exchange
Human bio-meteorological suitability tests for “highly relevant areas”
Planning recommendation map
Chapter 5 Urban Planning and Urban Design
163
use of interactive geographical information systems and/or sur-
veys that help citizens foresee potential land-use changes (see
Figure 5.14). This is enabled by climatic evaluation through
spatial and temporal quantitative descriptions and specications.
At the regional level, areas worth protecting by virtue of their
climatic functions, e.g., areas of heat load, fresh air supply, and
ventilation pathways, are identied as key targets for planning
measures.
5.5.3 Planning and Design Interventions
The task of planning and designing interventions relevant to
urban climatology is to improve thermal conditions and air qual-
ity (see Figure 5.14):
Reduction of UHIs (heat islands being an indication of ther-
mal comfort/discomfort) through open space planning
Optimization of urban ventilation via air exchange and wind
corridors
Prevention of stagnating air in stationary temperature inver-
sion conditions by eliminating barriers to air exchange
Maintenance and promotion of fresh air or cool air genera-
tion areas to further air exchange and improve air quality
Regional specications may include sustaining cool air or
fresh air generation areas (slopes) or ventilation lanes and tak-
ing into consideration building orientation, building height,
and density of development. Such specications may be imple-
mented according to building codes in urban land-use planning.
In addition, mandatory regulations for areas designated as open
spaces due to urban climate analyses are possible in the zoning
plan. Regulations at the regional level should also be reviewed.
Climatic concerns are thus considered in regional as well as city
planning.
5.5.4 Post-Intervention Evaluation
Field measurement studies (see Case Study 5.4) should be
conducted to assess the microclimatic performance of the urban
design and planning intervention (see Figure 5.14). Infrared
thermal imaging and/or population surveys can be undertaken
to assess temperature variations compared to conditions prior to
intervention. Climate-resilient strategies can have the effect of
prolonging moderate temperatures with associated benets to
public health and energy savings.
5.6 Stakeholders and Public Engagement
Building resilient urban environments through integrated mit-
igation and adaptation approaches requires a strong framework
for ongoing engagement with a broad range of stakeholders –
households, regional communities, and local and national tiers
of government, as well as academia, private businesses, health
care providers, and civil society organizations – that make up the
urban landscape.
Figure 5.14 Urban climate planning and design process.
Source: Jeffrey Raven, 2016
ARC3.2 Climate Change and Cities
164
Figure 5.15 Urban design intervention – future climate scenario.
Source: Urban Climate Lab, Graduate Program in Urban & Regional Design, New York Institute of Technology with Klimaat Consulting, 2014
Public engagement is as much about discourse as it is about
design. From an urban planning and urban design perspective,
community engagement and public participation are essential
for operationalizing any policy, program, or intervention and
offer many tangible and intangible benets. Despite the poten-
tial obstacles for developing and carrying out a successful
public outreach process, the absence of a robust stakeholder
engagement process has many higher costs and ramica-
tions. For instance, it can result in the development of locally
inappropriate solutions, increased conict and tension between
community groups, absence of a shared vision for the future,
“rebound effect” (whereby appropriate solutions are not appro-
priately used due to lack of community awareness), and, ulti-
mately, lack of preparedness and resiliency. Alternatively,
structured conversations can facilitate participation, knowledge
exchanges, shared decision-making, and ultimately, resiliency
actions. While the process can often be messy and contentious
because each group is pursuing different goals and interests,
Chapter 5 Urban Planning and Urban Design
165
a robust public engagement process with genuine stakeholder
participation and partnerships can ensure the sustained, inclu-
sive, and meaningful transformation of regulations, built envi-
ronments, and society (Kloprogge and Van Der Sluijs, 2006).
Community-led and place-based initiatives recognize and
leverage local knowledge and expertise. Top-down or expert-
driven outreach processes that lack genuine grassroots organiz-
ing and leave little opportunity for community-led initiatives can
act as a barrier to an effective public outreach campaign. Civic
engagement is often driven in an expert-dominated and exclu-
sive manner (Velazquez et al., 2005). Participatory actions on
the ground can be undermined or even neutralized by govern-
ing institutions’ lack of willingness to change (Warburton and
Yoshimura, 2005). These attitudes may lead to local resistance
to implementing identied solutions. More functionally, simple,
clear, and comprehensive resources must be readily available to
the public in order to increase knowledge and awareness.
Even if decision-makers are actively seeking to develop gen-
uine engagement and partnerships with stakeholders, an abun-
dance of confusing and contradictory resources can serve as
a barrier to robust engagement. Stakeholders and inhabitants
must be provided tailored resources with clear and easy steps on
how best to contribute to developing integrated mitigation and
adaptation solutions for their communities. This not only helps
ensure that integrated mitigation and adaptation solutions are
appropriate for the local context (i.e., a good “t” for the com-
munity and/or city), but also serves to empower communities
and strengthen the relationships among government, the private
sector, and citizens.
Public engagement is a means of ensuring the diverse needs
of communities, particularly those of the most vulnerable
groups, including populations with disabilities or chronic health
conditions; seniors and children; those socially isolated, histor-
ically underrepresented, or otherwise marginalized; and people
living below the poverty line. It is vital that these populations are
integrated into the decision-making process and solutions (see
Chapter 6, Equity and Environmental Justice).
Genuine and sustained stakeholder participation (as opposed
to simply assessing opinions or asking for rubber-stamp approval
on already-made plans) can draw out disagreements early, pro-
vide opportunities to work through different scenarios, and move
plans toward a shared vision, thereby actually saving time, money,
and political will. In order for stakeholders to be more commit-
ted, decision-makers and experts need to identify joint solutions
that break down institutional and disciplinary silos. Urban gover-
nance can facilitate a robust civic participation process that cre-
ates mechanisms for systematic learning and capacity-building in
communities as well as a transparent and open system in which
responsibilities and accountabilities are clearly dened at the local
level (see Chapter 6, Urban Governance). The dynamic and vari-
able conditions that climate change introduces call for a robust
stakeholder engagement process to help ensure integrated miti-
gation and adaptation responses are not simply implemented as
one-off, discrete protection measures, but rather are incrementally
adjusted and become part of the mechanisms for systematic learn-
ing, engagement, and transformation in a community. Community
engagement offers decision-makers an opportunity to prototype
a wider range of innovative solutions by creatively brainstorm-
ing, testing, and iterating. Despite focus by decision-makers on
formal planned responses, most adaptation responses are actually
carried out informally by individuals, households, and organiza-
tions. Therefore, communities that have been provided with infor-
mation and resources and that possess increased knowledge about
complex climate challenges and the integrated mitigation and
adaptation solutions needed to adequately address them will more
likely be willing to adopt resilient practices and policies (Yohe and
Leichenko, 2010).
Although physical interventions often provide protection from
only a single hazard or risk, communities that are integrated into the
mitigation and adaptation planning process increase their capacity
to prepare for, withstand, and recover from a wider range of cli-
mate-related disasters (not just a single hazard) as well as every-
day challenges that span health, income, and equity considerations.
Given that most cities have pre-existing vulnerabilities and that the
potential for institutional and/or systems failures is ever-present,
iterative, exible, and redundant responses are needed to build
ongoing capacity for adaptation. In other words, a robust stake-
holder outreach process can yield benets beyond simply helping
to implement policies and instead offer opportunities for improved
quality of life, public health, and equity.
The role of local authorities in the public engagement pro-
cess is evolving as integrated mitigation and adaptation and
sustainable urban development goals become a more signi-
cant priority for urban residents (Kloprogge and Van Der Sluijs,
2006). Local authorities must help facilitate and negotiate com-
peting interests between urban challenges (sprawl, fragmenta-
tion of spaces, complexity of scales), social necessities (health,
education, employment, culture, access to basic services), and
environmental concerns (GHG emissions, ecosystems protec-
tion, resource management, and conservation) (Broto, 2017).
Local decision-makers can use a range of urban planning and
urban design tools to overcome the obstacles to engaging the
public. There are different levels of local participation (Donzelot,
2009). The lowest level is the simple distribution of information
to inhabitants, whereas the highest level is direct engagement of
communities to share decision-making and prioritize the results
of consultations (Donzelot, 2009). For example, charrettes, an
intense, multiday design exercise with the community, are com-
monly used in urban planning and design projects to engage the
community to help create a plan for a particular site, area, or
neighborhood.
The private sector is also at the forefront of eco-innova-
tion systems to develop more sustainable cities (see Chapter 7,
Economics, Finance, and the Private Sector). Examples include
smarter solutions for energy efciency and energy provision,
development of renewable energy like geothermal systems,
ARC3.2 Climate Change and Cities
166
braking energy recovery from light rail systems, energy smart
grids, and solutions to improve intermodal transport.
Although public engagement is a critical and necessary step
for scaling integrated mitigation and adaptation solutions, there
are real challenges that can inhibit ongoing engagement and, ulti-
mately, action. As Bai et al. (2010) suggest, there is frequently an
inherent temporal (“not in my term”), spatial (“not in my patch”),
and institutional (“not my business”)-scale mismatch between
urban decision-making and global environmental concerns. Also,
the involvement of stakeholders requires a general level of trust and
cooperation (Tilly, 2005). A complication lies in the fact that deci-
sion-makers must often act as referees to solve potential conicts
and facilitate negotiations. However, to create consensus, promote
cooperation, and move toward an equitable environmental man-
agement process, an intimate understanding of the motivations and
drivers of the different stakeholder groups is needed (Schaltegger,
2003). Lack of transparency between the public and the private
sectors can be a key obstacle to reaching consensus among differ-
ent stakeholders. Stakeholders need to know who is accountable
for the local decisions on sustainability and climate solutions and
how they can contribute to the decision-making process.
5.7 Sector Linkages
There is a growing consensus around integrating urban planning
and urban design, climate science, and policy to bring about
desirable microclimates within compact, pedestrian-friendly built
environments. However, there remains much work to do to bridge
the gaps in tools, methods, and language between the scientic,
design, and policy-making communities. Conveying a compelling
investment/payback narrative also remains a challenge for setting
priorities with stakeholder groups. Ad hoc, disconnected approaches
fail to exploit synergies between professional practitioners,
and departments within government administrations are often
insufciently coordinated to capitalize on cross-disciplinary
actions. Silos of expertise are difcult to harness over the long-term
due to different departmental missions. A central challenge remains
the poor interdisciplinary connections between the various policy
experts, technical specialists, and urban planners/urban designers.
The absence of objective evaluation methodologies in the practice
of climate-resilient urban design illustrates this divergence, as does
the lack of a common interdisciplinary methodology for addressing
various spatial scales (Odeleye et al., 2008). For example, urban
climatology research produces sophisticated but theoretical results
that resist easy integration with empirical, design-oriented ndings
of urban design (Ali-Toudert et al., 2005).
Although experts in sustainable buildings and ecological foot-
prints are familiar with sustainability metrics to measure prog-
ress, the absence of objective evaluation of urban morphology
across spatial scales is a challenge in the urban design profes-
sion. In a recent survey on sustainability methods and indicators
in three built-environment professions in the United Kingdom
(planning, urban design, and architecture), sustainability guid-
ance documents had not led to a range of formal, systemic, or
morphological sustainability indicators in the urban design sec-
tor – instead continuing to emphasize “place-making” and the
quality/vitality of the public realm (Odeleye et al., 2008).
Although there has been extensive research on the UHI effect,
the relationship between urban geometry and thermal comfort is
by far less well understood and the numbers of studies are very
few (Ali-Toudert et al., 2005). Traditional urban design tools
must be rened and expanded to serve climate-resilient urban
design.
From parcel and neighborhood scale to municipal and regional
scale, a discontinuity of policy between scales challenges the pub-
lic’s understanding of holistic urban form. Consideration needs to
be given to how regional decisions may affect neighborhoods or
individual parcels and vice versa, and few tools have been devel-
oped to assess conditions in the urban environment at city block
or neighborhood scale (Brophy et al., 2000; Miller et al., 2008).
For urban areas, the key is a coordinated response that
addresses issues simultaneously rather than individually. For
example, an important concern underlying natural ventilation is
air quality, which means that transport management and building
microclimate need to be linked. Whereas one of the most effec-
tive passive ventilation strategies is to introduce cooler night air
to the perimeters of buildings, excessive pollution from nearby
vehicles and industry often undermines this strategy in compact
communities. This anthropogenic pollution is exacerbated by
lack of space for air movement, resulting in insufcient ground-
level air movement to disperse pollutants (Odeleye, 2008). The
totality of the environment – including noise, activity, climate,
and pollution – affects human health and well-being.
For compact development, proximity is the principal goal.
Proximity requires integrating infrastructure, housing, and sus-
tainable development into land-use planning to reduce the car-
bon footprint through compact development patterns. At the
urban scale, a comprehensive approach to transportation from
the perspective of sustainable development requires a holistic
view of planning (see Chapter 13, Urban Transportation). The
challenge remains to establish national frameworks and policies
for integrated mitigation and adaptation to address sustainable
development that encourages all sectors to coordinate and inte-
grate their activities (Hall, 2006) (see Chapter 16, Governance
and Policy).
5.8 Knowledge Gaps and Future Research
Urban areas alter their regional climates by adjusting the
overlying airshed. A substantial number of observational stud-
ies across the world have illustrated the prevalence of warmer
and drier conditions within cities, degraded air quality regimes,
and altered hydrological patterns resulting from impacts on pre-
cipitation and changes in drainage associated with increased
impervious cover. Advances in the physical understanding of
the urban climate system together with progress in computing
Chapter 5 Urban Planning and Urban Design
167
technologies has enabled the development and renement of
complex process-based models that characterize urban areas and
their interaction with the overlying atmosphere in a mathemat-
ical framework (Chen et al., 2011). Such process-based models
have considerable planning and design utility and are increas-
ingly applied to examine the impact of urban expansion and of
commonly proposed urban adaptation strategies to a long-term
globally changing climate (Georgescu et al., 2014).
To support utility for the planning and design process of cities,
such process-based models require two important improvements.
The spatial extent and morphology of urban areas remain sim-
plistic in contemporary modeling approaches (Chen et al., 2011).
The nature of this representation is assumed to vary by urban land
use and land cover, which is conditional on the density of urban
structures. However, implied in this assumption is that a diver-
sity of key morphological characteristics (e.g., sky view factor)
within a particular urban cover (e.g., high-density residential)
is nonexistent, an important condition that presents limitations
for place-based planning and design decisions. Therefore, the
realistically heterogeneous representation of cities within cur-
rent modeling frameworks remains an important but as-yet unre-
alized objective. How effective landscape conguration can be
as an adaptive strategy has only recently become a ripe area of
research in climate modeling (Connors et al., 2013; Rosenzweig
et al., 2014).
Waste heat resulting from energy use within cities (primarily
from building heating, ventilation, and air conditioning [HVAC]
systems) is also only crudely accounted for in current representa-
tions (Sailor, 2011). There does not yet exist a database of waste
heat proles for modeling applications for a diverse set of cities,
and although efforts are under way to develop spatially explicit
and time-varying heating proles, such datasets remain absent
for most urban areas (Chow et al., 2014). In regard to energy use,
suburbs in the United States account for roughly 50% of the total
domestic household carbon footprint due to longer commutes
(Jones and Kammen 2014).
Such concerns highlight the importance of future cooperation
between urban climatologists, planners, urban designers and
architects. A key asset of process-based urban climate model-
ing frameworks is their ability to offer insights by examining
the adaptive capacity of “what-if” growth scenarios and growth
management strategies to inform the planning and design pro-
cess prior to incipient stages of development.
Other major knowledge gaps concern the GHG emissions of
different cities and the association between GHG emissions and
different urban forms. The IPCC AR5 highlights the importance
of the next several decades for inuencing low-carbon urbaniza-
tion. It is essential to develop urban GHG emission inventories
and experiment with alternative urbanization patterns that facil-
itate low-carbon urban development. This is crucially related
to medium-sized cities in developing counties such as China
and India, which are expected to accommodate the majority of
expected urban population growth.
The impacts of medium-sized cities on the climate at urban,
regional, and global scales is a topic of considerable debate, but
their comparatively small size poses a conundrum for research-
ers: how do we acquire and incorporate the relevant information
into global understanding? Much research has focused on map-
ping these urban centers using demographic and administrative
information often supplemented by remote sensing. However,
these data provide little information on the internal makeup of
cities, which is crucially important for understanding their GHG
emissions and vulnerabilities. The absence of such information
inhibits international comparisons, knowledge transfer, and
effective integrated mitigation and adaptation.
5.9 Conclusions
This chapter has endorsed the concept of integrated mitiga-
tion and adaptation: climate management activities designed to
reduce global GHG emissions while producing regional benets
related to urban heat, ooding, and other extremes.
Cities shaped by integrated mitigation and adaptation
principles can reduce energy consumption in the built
environment, strengthen community adaptability to climate
change, and enhance the quality of the public realm.
Through energy-efcient planning and urban design,
compact morphology can work synergistically with high-
performance construction and landscape conguration to create
interconnected, protective, and attractive microclimates. The
long-term benets are also signicant, ranging from economic
savings through lowered energy consumption to the improved
ability of communities to thrive despite climate-related impacts
(Raven, 2011). A community’s capacity to cope with adversity,
adapt to future challenges, and transform in anticipation of
future crises yields greater social resilience with particularly
positive benets for poor and marginalized populations (Keck
and Sakdapolrak, 2013).
Annex 5.1 Stakeholder Engagement
The contributors to the ARC3.2 chapter on Urban Planning
and Urban Design engaged stakeholder groups and experts
throughout the chapter production process, with specic forums
in Asia, Europe, and the United States. The International
Conference on Urban Climate (ICUC9), held in Toulouse,
France, in July 2015, is an international forum for global urban
climatologists. Chapter Coordinating Lead Authors (CLAs),
Lead Authors, Contributing Authors, and Case Study Authors
participated in the conference, including Gerald Mills, Lutz
Katzschner, Matei Georgescu, and Jeffrey Raven. The Chapter
Key Findings and Major Messages were presented by the chap-
ter CLA at the launch of the Urban Climate Change Research
Network (UCCRN) European Hub, in partnership with the
Centre National de la Recherche Scientique, the Pierre and
Marie Curie University, and l’Atelier International du Grand
Paris launched in Paris, July 2015.
ARC3.2 Climate Change and Cities
168
At the ARC3.2 Midterm Authors Workshop in London, in
September 2014, which was attended by key chapter stakehold-
ers, the Coordinating Lead Author and Lead Authors cong-
ured the scientic basis for the chapters Major Messages and
Key Findings. The chapter CLA presented Major Messages and
Key Findings at Beijing University for Civil Engineering and
Architecture (BUCEA) and Tongji University in Shanghai, in
October 2014.
The 4th China–Europa Forum in Paris was held in
December 2014, focusing on the theme “Facing Climate
Change: Rethinking Our Global Development Model,” in
preparation for the UN FCCC Conference of the Parties
(COP21) in Paris 2015. It brought together more than 300
participants through two plenary sessions and three round
tables as well as twelve thematic workshops. As part of the
conference, at the Project EAST workshop with lead author
Pascaline Gaborit in Brussels, the Chapter CLA presented
ARC3.2 key scientic ndings and other research ndings.
At the closing plenary held on December 5 at the Town Hall
of the 4th Arrondissement of Paris, ARC3.2 CLA Raven pre-
sented the ARC3.2 draft Major Messages. This included the
most relevant points in the ARC3.2 chapter that will guide
urban decision-makers.
At the New York City Department of City Planning, the
ARC3 draft chapter Key Findings and Major Messages were
presented to experts in the city government. Co-presenters
included ARC3.2 lead author Gerald Mills, CLA Jeffrey
Raven, and the President of the American Institute of Architects
(AIANY). The ARC3.2-based presentation provided an oper-
ational framework, Case Studies, and a policy framework for
NYC municipal government. The chapter CLA and editor
introduced draft chapter Key Findings in discussions during the
National Science Foundation Research Coordination Network
project. The chapter CLA presented chapter Key Findings to
climate and health experts at the Center for Disease Control
in Atlanta, July 2014. The chapter CLA presented chapter Key
Findings at the American Institute of Architects Dialogues on
the Edge of Practice, February 2015, and presented ARC3.2
chapter Key Findings and Major Messages at the Center for
Architecture in NYC, April 2015. Lead author Brian Stone and
CLA Raven presented climate- resilient urban planning and
urban design research related to the ARC3.2 chapter at the New
York conference Extreme Heat: Hot Cities, November 2015.
The chapters Key Findings and Major Messages was pre-
sented by the CLA at the New York Institute of Technology-
Peking University Sustainable Megacities conference in Beijing,
China in October 2015. A Critical Climate Change Debrief:
COP21 Paris Conference was held at the Center for Architecture
in New York, January 2016, where chapter CLA Jeffrey Raven
presented how ARC3.2 strategies were tested in a Paris district
during COP21, in collaboration with lead authors Gerald Mills
and Mattia Leone.
The ARC3.2 Urban Planning and Urban Design chapter
abstract was presented at the Design Solutions for Climate Change
in Urban Areas conference in Naples, Italy, in July 2016. Jeffrey
Raven (CLA), and Lead Author Mattia Leone each led sessions
during the conference. Joining them in presenting ARC3.2’s
integrated cross-disciplinary framework was Chantal Pacteau
who is Coordinating Lead Author of the ARC3.2 Mitigation and
Adaptation chapter and Co-Director of the UCCRN European
Hub. Jeffrey Raven (CLA) presented ARC3.2 urban plan-
ning and urban design research-action at the National Science
Foundation Workshop, International Conference on Sustainable
Infrastructure (ICSI), Shenzhen, China, in October 2016.
5 Urban Planning and Urban Design
References
Akbari, H., Pomerantz, M., and Taha, H. (2001). Cool surfaces and shade
trees to reduce energy use and improve air quality in urban areas. Solar
Energy, 70(3), 295–310.
Ali-Toudert, F., and Mayer, H. (2005). Effects of street design on outdoor
thermal comfort. Meteorological Institute, University of Freiburg.
Alley, R., Kwok, K., Lam, D., Lau, W., Watts, M., and Whyte, F. (2011).
Water resilience for cities. ARUP. http://publications.arup.com/
publications/u/urban_life_water_resilience_for_cities
American Planning Association (APA). (2015). What is planning?.
American Planning Association. Accessed March 20, 2016: https://
www.planning.org/ aboutplanning/whatisplanning.htm
Arneld, A. 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 23(1), 1–26.
ARUP. (2014). C40 Climate action in megacities: A quantitative study of
efforts to reduce GHG emissions and improve urban resilience to
climate change in C40 cities. Accessed July 30, 2015: C40.org.
Australian Sustainable Built Environment Council (ASBEC). (2013). What
is urban design? Accessed April 11, 2014: http://www.urbandesign.org
.au/whatis/ index.aspx
BMT WBM Pty Ltd. (2009). Evaluating options for water sensitive urban
design: A national guide. Joint Steering Committee for Water Sensitive
Cities (JSCWSC), Australia.
Bottema, M. (1999). Towards rules of thumb for wind comfort and air
quality. Atmospheric Environment 33(24),4009–4017.
Brophy, V., O’Dowd, C., Bannon, R., Goulding, J., and Lewis, J. O. (2000).
Sustainable urban design. Energy Research Group, University College
Dublin.
Broto, V.C. (2017). Urban Governance and the Politics of Climate Change.
World Development. 93:1–15.
Charlesworth, S. (2010). A review of the adaptation and mitigation of
Global Climate Change using Sustainable Drainage in cities. Journal
of Water and Climate Change 1(3), 165–180
Chen, F., Kusaka, H., Bornstein, R., Ching, J., Grimmond, C. S. B.,
Grossman-Clarke, S., and Zhang, C. (2011). The integrated WRF/
urban modelling system: Development, evaluation, and applications to
urban environmental problems. International Journal of Climatology
31(2), 273–288.
Chen, L., Ng, E., An, X. P., Ren, C., He, J., Lee, M. Wang, U., and He, J.
(2010). Sky view factor analysis of street canyons and its implications
for intra-urban air temperature differentials in high-rise, high-density
urban areas of Hong Kong: A GIS-based simulation approach. Interna-
tional Journal of Climatology 32(1), 121–136. doi: 10.1002/joc. 2243
.(EdNg)
Chapter 5 Urban Planning and Urban Design
169
Chow, W. T., Salamanca, F., Georgescu, M., Mahalov, A., Milne, J. M., and
Ruddell, B. L. (2014). A multi-method and multi-scale approach for
estimating city-wide anthropogenic heat uxes. Atmospheric Environ-
ment 99, 64–76.
Connors, J. P., Galletti, C. S., and Chow, W. T. (2013). Landscape congu-
ration and urban heat island effects: Assessing the relationship between
landscape characteristics and land surface temperature in Phoenix,
Arizona. Landscape Ecology 28, 271–283. doi:10.1007/s10980-012-
9833-1
Decker, E., Smith, B., and Rowland, F. (2000). Energy and material ow
through the urban ecosystem. Annual Review of Energy and the Envi-
ronment 25, 685–740.
Dickie, S., Ions, L., McKay, G., and Shaffer, P. (2010). Planning for SuDS
Making it Happen. Ciria.
Donzelot, J. (2009). La ville à trois vitesses, Éditions de la Villette, 2009.
Éditions de la rue d’Ulm, 2009 and presentation in Sénart 2009
Energy-Cités Info and newsletters.
Doulos, L., Santamouris, M., and Livada, I. (2004). Passive cooling of out-
door urban spaces. The role of materials. Solar Energy 77(2), 231–249.
Duvigneaud, P., and Denayer-de Smet, S. (eds.). (1975). L’Ecosystème
Urbain—Application à l’Agglomération bruxelloise. Publication de
Colloque International organisé par L’Agglomération de Bruxelles 14
et 15 septembre 1974, Bruxelles.
Early P., Gedge D., and Wilson S. (2007b). Building Greener: Guidance on
the Use of Green Roofs, Green Walls and Complementary Features on
Buildings. Ciria.
Erell, E., Pearlmutter, D., and Williamson, T. (2012). Urban Microclimate:
Designing the Spaces Between Buildings. Routledge.
Flörke, M., Wimmer, F., Laaser, C., Vidaurre, F., Tröltzsch, J., Dworak, T.,
Stein, U., Marinova, N., Jaspers, F., Ludwig, F., Swart, R., Long, H.,
Giupponi, G., Bosello, F., and Mysiak, J. (2011). Final Report for the
Project Climate Adaptation – Modeling Water Scenarios and Sectoral
Impacts. CESR – Center for Environmental Systems Research.
Gafn, S.R., M. Imhoff, C. Rosenzweig, R. Khanbilvardi, A. Pasqualini, A.Y.Y.
Kong, D. Grillo, A. Freed, D. Hillel, and E. Hartung, 2012: Bright is the
new black – Multi-year performance of high-albedo roofs in an urban cli-
mate. Environ. Res. Lett., 7, 014029, doi:10.1088/1748-9326/7/1/014029.
Georgescu, M., Moreeld, P. E., Bierwagen, B. G., and Weaver, C. P. (2014).
Urban adaptation can roll back warming of emerging megapolitan regions.
Proceedings of the National Academy of Sciences 111(8), 2909–2914.
Gomez-Ibanez, D. J., Boarnet, M. G., Brake, D. R., Cervero, R. B.,
Cotugno, A., Downs, A., Hanson, S., Kockelman, K. M., Mokhtarian,
P. L., Pendall, R. J., Santini, D. J., and Southworth, F. (2009). Driv-
ing and the built environment: The effects of compact development on
motorized travel, energy use, and CO2 emissions. Oak Ridge National
Laboratory (ORNL).
Grimmond, S. (2007). Urbanization and global environmental change:
Local effects of urban warming. Geography Journal 173, 83–88.
Guttikunda, S. K., Carmichael, G. R., Calori, G., Eck, C., and Woo, J. H.
(2003). The contribution of megacities to regional sulfur pollution in
Asia. Atmospheric Environment 37(1), 11–22.
Hall, R. (2006). Understanding and Applying the Concept of Sustainable
Development to Transportation Planning and Decision-Making in the
U.S. Doctoral dissertation, MIT.
Hoyer, J., Dickhaut, W., Kronawitter, L., and Weber, B. (2011). Water sensi-
tive urban design. Jovis Verlag GmbH.
Hough, M. (1989). City Form and Natural Process: Toward a New Urban
Vernacular. Routledge.
International Panel on Climate Change (IPCC). (2014a). Climate Change
2014: Synthesis Report. Contribution of Working Groups I, II and III
to the Fifth Assessment Report of the Intergovernmental Panel on Cli-
mate Change [Core Writing Team, Pachauri, R. K., and Meyer, L. A.
(eds.)]. IPCC.
Jacobson, M. Z., and Ten Hoeve, J. E. (2012). Effects of urban surfaces and
white roofs on global and regional climate. Journal of Climate 25(3),
1028–1044.
Jonathan Rose Companies. (2011). Location efciency and housing type: Boil-
ing it down to BTUs. Accessed December 8, 2014: https://www.epa.gov/
sites/ production/les/2014-03/documents/location_efciency_btu.pdf
Jones, C., and Kammen, D. M. (2014). Spatial distribution of U.S. house-
hold carbon footprints reveals suburbanization undermines greenhouse
gas benets of urban population density. Environmental Science and
Technology 48(2), 895–902.
Kazmierczak, A., and Carter, J. (2010). Adaptation to climate change using
green and blue infrastructure: A database of case studies. University
of Manchester, Interreg IVC Green and blue space adaptation for
urban areas and eco-towns (GRaBS). Accessed December 20, 2015:
https://orca.cf.ac.uk/64906/1/Database_Final_no_hyperlinks.pdf
Keck, M., and Sakdapolrak, P. (2013). What is social resilience? Lessons
learned and ways forward. Erdkunde 67(1), 5–19. Accessed August 20,
2014: http://www.academia.edu/3110553/What_is_Social_Resilience_
Lessons_Learned_and_Ways_Forward
Kennedy, C., Steinberger, J., Gasson, B., Hansen, Y., Hillman, T., Havranek,
M., Pataki, D., Phdungsilp, A., Ramaswami, A., and Mendez, G. V.
(2009). Greenhouse gas emissions from global cities. Environmental
Science & Technology 43(19), 7297–7302.
Kloprogge, P. and Van Der Sluijs, J.P. (2006). The inclusion of stakehold-
er knowledge and perspectives in integrated assessment of climate
change. Climatic Change. 75:359–389.
Knowles, R. (2003). The solar envelope: Its meaning for energy and
buildings. Energy and Buildings 35, 15–25.
Konrad, C. P. (2003). Effects of urban development on oods. U.S.
Geological Survey Fact Sheet 076–03.
Kravík, M., Pokorný, J., Kohutiar, J., Ková, M., and Tóth, E. (2007).
Water for the Recovery of the Climate A New Water Paradigm.
Municipality of Tory.
Laidley, T. (2015). Measuring sprawl: A new index, recent trends, and future re-
search. Urban Affairs Review 1078087414568812. Accessed April 6, 2016:
http://uar.sagepub.com/citmgr?gca=spuar%3B1078087414568812v1
Lee, J. (2012). Transit Synergized Development Framework for a Smart,
Low-Carbon, Eco-City. Accessed August 20, 2014: http://www
.academia.edu/6286908/Transit_Synergized_Development_Frame-
work_for_a_Smart_Low-Carbon_Eco-City_Executive_Summary_
Li, W., Chen, S., Chen, G., Sha, W., Luo, C., Feng, Y., and Wang, B. (2011).
Urbanization signatures in strong versus weak precipitation over
the Pearl River Delta metropolitan regions of China. Environmental
Research Letters 6(3), 034020.
Lindau, L.A., Hidalgo, D., and Facchini, D. (2010). Curitiba, the Cradle of
Bus Rapid Transit. Build Environment (1978). 35(3):274–282.
Lowry, W. P. (1977). Empirical estimation of urban effects on climate: A
problem analysis. Journal of Applied Meteorology 16(2), 129–135.
Luhmann N. (2006). La Conance un mécanisme de réduction de la
complexité sociale., 2–111.
McGranahan, G., et al. (2007). The rising tide: Assessing the risks of climate
change and human settlements in low elevation coastal zones. Environ-
ment and Urbanisation 19, 17–37.
Mehrotra, S., Lefevre, B., Zimmerman, R., Gercek, H., Jacob, K., and Srini-
vasan, S. (2011). Climate change and urban transportation systems. In
Rosenzweig, C., Solecki, W. D., Hammer, S. A., Mehrotra, S. (eds.), Cli-
mate Change and Cities: First Assessment Report of the Urban Climate
Change Research Network (145–177).Cambridge University Press.
Mentens, J., Raes, D., and Hemy, M. (2006). Green roofs as a tool for
solving the rainwater runoff problem in the urbanized 21st century?
Landscape and Urban Planning 77, 217–226.
Miller, N., Cavens, D., Condon, P., Kellett, N., and Carbonell, A. (2008).
Policy, urban form and tools for measuring and managing greenhouse
gas emissions: The North American problem. Climate change and
urban design: The third annual congress of the Council for European
Urbanism, Oslo.
Miller, R. B., and Small, C. (2003). Cities from space: Potential applications
of remote sensing in urban environmental research and policy. Envi-
ronmental Science & Policy 6(2), 129–137.
ARC3.2 Climate Change and Cities
170
Mills, G. (2005). Urban form, function and climate. Dept. of Geography,
University of California, Davis. Accessed August 30, 2014: epa.gov/
heatisland/resources/pdf/GMills4.pdf
Mills, G. (2004). The urban canopy layer heat island. IAUC Teach-
ing Resources. Accessed June 18, 2015: www.epa.gov/heatislands/
resources/…/ HeatIslandTeachingResource.pdf
Monks, P. S., Granier, C., Fuzzi, S., Stohl, A., Williams, M. L., Akimoto, H.,
Amann, M., Baklanov, A., Baltensperger, U., Bey, I., Blake, N., Blake,
R. S., Carslaw, K., Cooper, O. R., Dentener, F., Fowler, D., Fragkou,
E., Frost, G. J., Generoso, S., Ginoux, P., Grewe, V., Guenther, A.,
Hansson, H. C., Henne, S., Hjorth, J., Hofzumahaus, A., Huntrieser,
H., Isaksen, I. S. A., Jenkin, M. E., Kaiser, J., Kanakidou, M., Klimont,
Z., Kulmala, M., Laj, P., Lawrence, M. G., Lee, J. D., Liousse, C., Mai-
one, M., McFiggans, G., Metzger, A., Mieville, A., Moussiopoulos, N.,
Orlando, J. J., O’Dowd, C. D., Palmer, P. I., Parrish, D. D., Petzold,
A., Platt, U., Poeschl, U., Prevot, A. S. H., Reeves, C. E., Reimann,
S., Rudich, Y., Sellegri, K., Steinbrecher, R., Simpson, D., ten Brink,
H., Theloke, J., van der Werf, G. R., Vautard, R., Vestreng, V., Vla-
chokostas, C., and von Glasow, R. (2009). Atmospheric composition
change – global and regional air quality. Atmospheric Environment 43,
5268–5350. doi:10.1016/j.atmosenv.2009.08.021
Morgan, C., Bevington, C., Levin, D., Robinson, P., Davis, P., Abbott, J.,
and Simkins, P. (2013). Water Sensitive Urban Design in the UK
Ideas for Built Environment Practitioners. Ciria.
Newman, P. G., and Kenworthy, J. R. (1989). Cities and Automobile Depen-
dence: An International Sourcebook. Gower Technical.
Nolde, E., Vansbotter, B., Rüden, H., König, K.W. (2007). Innovative water
concepts. Service water utilization in buildings. Berlin Senate Depart-
ment for Urban Development. Accessed January 17, 2015: http://www
.stadtentwicklung .berlin.de/internationales_eu/stadtplanung/down-
load/betriebswasser_englisch_2007.pdf
Nolon, J. R. (2012). Land use for energy conservation and sustainable
development: A new path toward climate change mitigation. Journal
of Land Use & Environmental Law 27. Accessed June 21, 2014: http://
digitalcommons.pace.edu/lawfaculty/793/
Odeleye, D., and Maguire, M. (2008). Walking the talk? Climate change and
UK spatial design policy. Climate change and urban design: The Third
Annual Congress of the Council for European Urbanism, Oslo.
Oke, Tim R. (1981). Canyon geometry and the nocturnal urban heat island:
Comparison of scale model and eld observations. Journal of Clima-
tology 1 (3), 237–254.
Oleson, K. W., Monaghan, A., Wilhelmi, O., Barlage, M., Brunsell, N., Fed-
dema, J., Hu, L., Steinhoff, D. F. (2015). Interactions between urbaniza-
tion, heat stress, and climate change. Climatic Change. 129:525–541.
Ong, B. L. (2003). Green plot ratio: An ecological measure for architecture
and urban planning. Landscape and Urban Planning 63(4), 197–211.
Paul, M. J., and Meyer, J. L. (2001). Streams in the urban landscape. Annual
Review of Ecological Systems 32, 333–365.
Poleto, C., and Tassi, R. (2012). Sustainable urban drainage systems. In
Javaid, M. S. (ed.), Drainage Systems (55–72). InTech.
Raven, J. (2011). Cooling the public realm: Climate-resilient urban de-
sign · resilient cities. In Otto-Zimmermann, K. (ed.), Cities and Ad-
aptation to Climate Change: Local Sustainability (Vol. 1, 451–463),
Springer.
Resch, E., Bohne, R.A., Kvamsdal, T., and Lohne, J. (2016). Impact of
urban density and building height on energy use in cities. Energy Pro-
cedia. 96:800–814.
Revi, A., Satterthwaite, D. E., Aragón-Durand, F., Corfee-Morlot,
J., Kiunsi, R. B. R., Pelling, M., Roberts, D. C., and Solecki, W.
(2014). Urban areas. In Field, C. B., Barros, V. R., Dokken, D.
J., Mach, K. J., Mastrandrea, M. D., Bilir, T. E., Chatterjee, M.,
Ebi, K. L., Estrada, Y. O., Genova, R. C., Girma, B., Kissel, E.
S., Levy, A. N., MacCracken, S., Mastrandrea, P. R., and White,
L. L. (eds.), Climate Change 2014: Impacts, Adaptation, and
Vulnerability. Part A: Global and Sectoral Aspects. Contribution
of Working Group II to the Fifth Assessment Report of the Inter-
governmental Panel on Climate Change (535–612). Cambridge
University Press. Accessed October 20, 2015: http://www.ipcc
.ch/pdf/ assessment-report/ar5/wg2/WGIIAR5-Chap8_FINAL.pdf
Rosenzweig, C., Horton, R. M. Bader, D. A. Brown, M. E. DeYoung, R.
Dominguez, O. Fellows, M. Friedl, L. Graham, W. Hall, C. Higuchi,
S. Iraci, L. Jedlovec, G. Kaye, J. Loewenstein, M. Mace, T. Milesi,
C. Patzert, W. Stackhouse, P. W., and Toufectis, K. (2014). Enhancing
climate resilience at NASA centers: A collaboration between science
and stewardship. Bulletin of the American Meteorological Society
95(9), 1351–1363. doi:10.1175/BAMS-D-12–00169.1
Rosenzweig, C., Solecki, W., Pope, G., Chopping, M., Goldberg, R., and
Polissar, A. (2004). Urban heat island and climate change: An assess-
ment of interacting and possible adaptations in the Camden, New
Jersey region. New Jersey Department of Environmental Protection.
Roth, M. (2007). Review of urban climate research in (sub) tropical regions.
International Journal of Climatology 27(14), 1859–1873.
Sailor, D. J. (2011). A review of methods for estimating anthropogenic heat
and moisture emissions in the urban environment. International Jour-
nal of Climatology 31(2), 189–199.
Santamouris, M. (2014). Cooling the cities A review of reective and
green roof mitigation technologies to ght heat island and improve
comfort in urban environments. Solar Energy 103, 682–703.
Schaltegger, S., Burit, R., and Petersen, H. (2003). An Introduction to
Corporate Environmental Management- Striving for Sustainability.
Greenleaf Publishing.
Schmidt M. (2009). Rainwater harvesting for mitigating local and global
warming. In Proceedings Fifth Urban Research Symposium: “Cities
and Climate Change: Responding to an Urgent Agenda. World Bank.
Schneider, A., Friedl, M. A., and Potere, D. (2009). A new map of global
urban extent from MODIS satellite data. Environmental Research
Letters 4(4), 044003.
Scholz, M., and Grabowiecki, P. (2007). Review of permeable pavement
systems. Building and Environment 42(11), 3830–3836.
Schuler, M. (2009). The Masdar Development – showcase with global effect.
Urban Futures 2030. Visionen künftigen Städtebaus und urbaner Lebens-
weisen, Band 5 der Reihe Ökologie, Herausgegeben von der Heinrich-
Böll-Stiftung. Accessed March 27, 2014: http://www.boell.de/downloads/
Urban-Future-i.pdf
Shashua-Bar, L., Potchter, O., Bitan, A., Boltansky, D., and Yaakov, Y.
(2010). Microclimate modelling of street tree species effects within the
varied urban morphology in the Mediterranean city of Tel Aviv, Israel.
International Journal of Climatology 30(1), 44–57.
Smith, C., and Levermore, G. (2008). Designing urban spaces and buildings
to improve sustainability and quality of life in a warmer world. Energy
Policy. doi:10.1016/j.enpol.2008.09.011
Steffan C., Schmidt M., Köhler M., Hübner I., Reichmann B. (2010). Rain
water management concepts, greening buildings, cooling buildings
planning, construction, operation and maintenance guidelines. Berlin
Senate Department for Urban Development. Accessed June 12, 2014:
http://www.stadtentwicklung.berlin.de/bauen/oekologisches_bauen/
download/SenStadt_Regenwasser_engl_bfrei_nal.pdf
Stewart, I. D., and Oke, T. R. (2012). Local climate zones for urban tem-
perature studies. Bulletin of the American Meteorological Society
93(12),1879–1900.
Stone, B., Hess, J., and Frumkin, H. (2010). Urban form and extreme heat
events: Are sprawling cities more vulnerable to climate change? Envi-
ronmental Health Perspectives 118, 1425–1428.
Stone, B., Vargo, J., Liu, P., Habeeb, D., DeLucia, A., Trail, M., Hu, Y.,
Russel, A. (2014). Avoided heat-related mortality through climate
adaptation strategies in three U.S. cities. Plos One 9, e100852.
Taha, H., Konopacki, S., and Gabersek, S. (1999). Impacts of large scale surface
modications on meteorological conditions and energy use: A 10-region
modeling study. Theoretical and Applied Climatology 62, 175–185.
Thomas, R., and Ritchie, A. (eds.). (2003). Sustainable Urban Design: An
Environmental Approach. Spon Press.
Tilly C. (2005). Trust and Rule. Cambridge Studies in Comparative Politics.
Cambridge University Press
United Nations Environment Programme (UNEP). (2010). Green Economy
Developing Countries Success Stories. UNEP.
United Nations Environment Programme (UNEP). (2012). Using ecosys-
tems to address climate change: Ecosystem based adaptation Regional
seas program Report. UNEP.
Chapter 5 Urban Planning and Urban Design
171
World Bank. (2017). 2016 GNI per capita, Atlas method (current US$). Ac-
cessed August 9, 2017: http://data.worldbank.org/indicator/NY.GNP
.PCAP.CD
Case Study 5.2 Adapting to Summer Overheating
in Light Construction with Phase-
Change Materials in Melbourne,
Australia
Australian Bureau of Statistics. (2013). 2011 Census Community Proles,
Greater Capital City Statistical Areas, Greater Melbourne. Accessed
January 16, 2015: http://www.censusdata.abs.gov.au/census_
services/getproduct/ census/2011/communityprole/2GMEL?open-
document&navpos=230
Peel, M. C., Finlayson, B. L., and McMahon, T. A. (2007). Updated world map
of the Köppen-Geiger climate classication. Hydrology and Earth Sys-
tem Sciences Discussions 4(2), 462. Accessed July 31, 2015: http://www
.hydrol-earth-syst-sci-discuss.net/4/439/2007/hessd-4-439-2007.pdf
World Bank. (2017). 2016 GNI per capita, Atlas method (current US$).
Accessed August 9, 2017: http://data.worldbank.org/indicator/NY.GNP
.PCAP.CD
Case Study 5.3 Application of Urban Climatic Map
to Urban Planning of High-Density
Cities: An Experience from Hong
Kong
Census SAR. (2011). Census and Statistics Department, Government of
Hong Kong Special Administrative Region (SAR). Accessed June 12,
2014: http://www.census2011.gov.hk/en/main-table/A202.html
Census SAR. (2013). Information Services Department, Census and Statistics
Department, Hong Kong Special Administrative Region Government.
Accessed March 27, 2014: http://www.gov.hk/en/about/abouthk/facts
heets/docs/population.pdf
GovHK. (2015). Hong Kong – the facts. Accessed: http://www.gov.hk/en/
about/abouthk/facts.htm
Leung, Y. K., Yeung, K. H., Ginn, E. W. L., and Leung, W. M. (2004).
Climate change in Hong Kong. Technical Note No. 107, Hong Kong
Observatory.
Peel, M. C. Finlayson, B. L. , and McMahon, T. A. (2007). Updated world
map of the Köppen-Geiger climate classication. Hydrology and Earth
System Sciences Discussions 4(2), 462.
PlanD. (2012). Final Report, Urban Climatic Map and Standards for Wind En-
vironment – Feasibility Study. Planning Department, Hong Kong Govern-
ment. Accessed January 16, 2015: http://www.pland.gov.hk/pland_en/p_
study/prog_s/ucmapweb/ucmap_project/content/reports/nal_report.pdf
World Bank. (2017). 2016 GNI per capita, Atlas method (current US$).
Accessed August 9, 2017: http://data.worldbank.org/indicator/NY
.GNP.PCAP.CD
Case Study 5.4 An Emerging Clean-Technology
City: Masdar, Abu Dhabi, United
Arab Emirates
Bullis, K. (2009). A zero-emissions city in the desert. MIT Technology
Review 56–63.
Demographia. (2016). Demographia World Urban Areas 12th Annual Edi-
tion: 2016:04. Accessed August 9, 2017: http://www.demographia
.com/db-worldua.pdf
Clarion Associates. (2008). Abu Dhabi estidama program: Interim estidama
community guidelines assessment system for commercial, residential,
and institutional development. Prepared for the Emirate of Abu Dhabi
Urban Planning Council.
United Nations Environment Programme (UNEP). (2014). Green Infra-
structure Guide for Water Management: Ecosystem-based manage-
ment approaches for water-related infrastructure projects. UNEP.
UN-Habitat. (2012). Urban Patterns for a Green Economy: Leveraging
Density. UN-Habitat.
UN-Habitat. (2013). Urban Planning for City Leaders. UN-Habitat.
U.S. Department of Energy. (2013). Effects of the Built Environment on
Transportation: Energy Use, Greenhouse Gas Emissions, and Other
Factors. U.S. DOE.
U.S. Environmental Protection Agency (EPA). (2008). Reducing Urban
Heat Islands: Compendium of Strategies. U.S. EPA.
Velazquez, J., Yashiro, M., Yoshimura, S., and Ono, I. (2005). Innova-
tive Communities: People-centred Approaches to Environmental
Management in the Asia-Pacic Region. United Nations University
Press.
Walker, B., and Salt, D. (2006). Resilience Thinking. Island Press.
Warburton, D., and Yoshimura, S. (2005). Local to global connections. In
Velasquez, J., Yashiro, M., Yoshimura, S., and Ono, I. (eds.), Inno-
vative Communities: People-centred Approaches to Environmental
Management in the Asia-Pacic Region, United Nations University
Press.
Woods-Ballard, B., Kellagher, R., Martin, P., Jefferies, C., Bray, R., and
Shaffer, P. (2007). The SuDS Manual. Ciria.
Yohe, G. and Leichenko, R. (2010), Chapter 2: Adopting a risk-based ap-
proach. Annals of the New York Academy of Sciences, 1196: 29–40.
doi:10.1111/j.1749-6632.2009.05310.x
Zhao, S.X., Guo, N.S., Li, C.L.K., Smith, C. (2017). Megacities, the World’s
Largest Cities Unleashed: Major Trends and Dynamics in Contempo-
rary Global Urban Development. World Development. 98:257–289.
Zheng, H. and Peeta, S. (2015). Network design for personal rapid transit
under transit-oriented development. Transportation Research Part C.
55:351–362.
Chapter 5 Case Study References
Case Study 5.1 Green Infrastructure as a Climate
Change Adaptation Option for
Overheating in Glasgow, UK
Emmanuel, R., and Krüger, E. (2012). Urban heat islands and its impact on cli-
mate change resilience in a shrinking city: The case of Glasgow, UK. Build-
ing and Environment 53, 137–149. Accessed June 21, 2014: http://dx.doi
.org/10.1016/j.buildenv.2012.01.020
Emmanuel R., and Loconsole, A. (2015). Green infrastructure as an adapta-
tion approach to tackle urban overheating in the Glasgow Clyde Valley
Region. Landscape and Urban Planning 138, 71–86. Accessed April 4,
2016: http://dx.doi.org/10.1016/j.buildenv.2012.01.020
Glasgow City Council. (2010). Climate change strategy & action plan.
Accessed June 21, 2014: https://www.glasgow.gov.uk/CHttpHandler
.ashx?id=7609andp=0
Keeley, M. (2011). The green area ratio: An urban site sustainability metric.
Journal of Environmental Planning and Management 54, 937–958,
http://dx.doi.org/10.1080/09640568.2010.547681
Ofce for National Statistics. (2012). 2011 Census: Population and household
estimates for the United Kingdom Table 2. Accessed October 12, 2015:
http://www.ons.gov.uk/ons/rel/census/2011-census/population-and-house-
hold-estimates-for-the-united-kingdom/rft-table-2-census-2011.xls
Peel, M. C., Finlayson, B. L., and McMahon, T. A. (2007). Updated world map
of the Köppen-Geiger climate classication. Hydrology and Earth Sys-
tem Sciences Discussions 4(2), 462. Accessed July 31, 2015: http://www
. hydrol-earth-syst-sci-discuss.net/4/439/2007/hessd-4-439-2007.pdf
United Nations Development Programme. (2014). Human Development
Index (HDI). Accessed November 4, 2015: http://hdr.undp.org/sites/
default/les/hdr14_statisticaltables.xls
ARC3.2 Climate Change and Cities
172
Foster & Partners. (2009). Masdar development. Accessed February 25,
2014: http://www.fosterandpartners.com/projects/masdar-development
Manghnani, N., and Bajaj, K. (2014). Masdar City: A model of urban
environmental sustainability. International Journal of Engineering
Research and Applications 4(10) 38–42.
Peel, M. C., Finlayson, B. L., and McMahon, T. A. 2007. Updated world
map of the Köppen-Geiger climate classication. Hydrology and Earth
System Sciences Discussions, 4 (2), 462.
SCAD. (2011). Statistic Centre of Abu Dhabi, United Emirates, Accessed
February 25, 2015: http://www.scad.ae/en/pages/default.aspx
United Nations, Department of Economic and Social Affairs, Population
Division (2016). The World’s Cities in 2016 – Data Booklet (ST/ESA/
SER.A/392).
World Bank. (2017). 2016 GNI per capita, Atlas method (current US$).
Accessed August 9, 2017: http://data.worldbank.org/indicator/NY
.GNP.PCAP.CD