Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272160951

Assessment of urban heat island using satellite remotely sensed


imagery: A review

Article  in  The South African geographical journal, being a record of the proceedings of the South African Geographical Society · July 2014
DOI: 10.1080/03736245.2014.924864

CITATIONS READS

25 1,852

5 authors, including:

Adeline Ngie Khaled Abubakr Ali Abutaleb


North-West University National Authority for Remote Sensing and Space Sciences
18 PUBLICATIONS   105 CITATIONS    43 PUBLICATIONS   249 CITATIONS   

SEE PROFILE SEE PROFILE

Fethi Ahmed Ahmed Darwish


University of the Witwatersrand National Research Center, Egypt
70 PUBLICATIONS   991 CITATIONS    3 PUBLICATIONS   73 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

M..Sc. View project

Retrieval of vegetation physiological parameters using remote sensing data View project

All content following this page was uploaded by Adeline Ngie on 12 February 2015.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [Adeline Ngie]
On: 20 June 2014, At: 06:45
Publisher: Routledge
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

South African Geographical Journal


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/rsag20

Assessment of urban heat island using


satellite remotely sensed imagery: a
review
a ab a b
Adeline Ngie , Khaled Abutaleb , Fethi Ahmed , Ahmed Darwish
b
& Mahmoud Ahmed
a
University of Johannesburg, P.O. Box 524, Auckland Park 2006,
South Africa
b
National Authority for Remote Sensing and Space Sciences, P.O.
Box 1564, Alf Maskan, Cairo, Egypt
Published online: 17 Jun 2014.

To cite this article: Adeline Ngie, Khaled Abutaleb, Fethi Ahmed, Ahmed Darwish & Mahmoud
Ahmed (2014): Assessment of urban heat island using satellite remotely sensed imagery: a review,
South African Geographical Journal, DOI: 10.1080/03736245.2014.924864

To link to this article: http://dx.doi.org/10.1080/03736245.2014.924864

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [Adeline Ngie] at 06:45 20 June 2014
South African Geographical Journal, 2014
http://dx.doi.org/10.1080/03736245.2014.924864

Assessment of urban heat island using satellite remotely sensed


imagery: a review
Adeline Ngiea*, Khaled Abutaleba,b, Fethi Ahmeda, Ahmed Darwishb and
Mahmoud Ahmedb
a
University of Johannesburg, P.O. Box 524, Auckland Park 2006, South Africa; bNational Authority
for Remote Sensing and Space Sciences, P.O. Box 1564, Alf Maskan, Cairo, Egypt

Urban heat island (UHI) is an important phenomenon given its direct and indirect
impacts on human populations in urban environments. The UHI phenomenon has been
studied extensively using both directly measured and remotely sensed temperature data
sets. This review presents an overview of the UHI background concepts and provides
details of satellite remote sensing data and processing techniques applied by various
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

studies to retrieve land surface temperatures (LSTs) in order to establish the existence
of this phenomenon. The review reveals that various factors influence the utility of
remote sensing in UHI studies. However, remote sensing has the potential to provide
fairly accurate LST measurements that could be used to characterize UHI over large
areas. UHI studies have been conducted mostly in the Americas, Europe and Asia with
very few studies in Africa. Recommendations on means to improve the use of remote
sensing in UHI studies have been made. Accurate and up-to-date remotely sensed
assessments of UHIs will inform city planners and managers and assist in addressing
challenges related to the phenomenon.
Keywords: urban heat island; remote sensing; land cover

1. Introduction
1.1 Description of urban heat island and scope of this paper
One of the most typical phenomena of urban climates that challenges sustainable
livelihoods is urban heat island (UHI), which describes the excess temperature near the
ground (canopy layer) of the central urban locations as being higher than those of nearby
or surrounding areas of similar elevation (Voogt & Oke, 2003). UHI has been regarded as
the most well-documented example of anthropogenic climate modification within the field
of urban climate (Arnfield, 2003). Usually, UHI was considered as the difference between
urban and rural areas but with recent developments in cities, there is no distinct borderline
between ‘urban’ and ‘rural’ areas as a result of urban growth. Therefore, the UHI can be
considered in terms of the difference between the central parts of the city and its
surrounding areas.
UHI effect impacts on the development of meteorological events such as increased
precipitation, boosts energy demands, poses threats to environmental quality and long-
term sustainability of localities, and potentially contributes to global warming (Kikegawa,
Genchi, Kondo, & Hanaki, 2006; Yang & Liu, n.d.). Another impact of UHI is economical
on the energy usage by local residents. There is a high risk that UHI may induce unchecked
increases in fossil-fuel consumption for cooling and its consequent increases in the
anthropogenic carbon dioxide emission, which could contribute to global warming

*Corresponding author. Email: adelinengie@gmail.com

q 2014 Society of South African Geographers


2 A. Ngie et al.
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

Figure 1. A UHI profile over two cities – Baltimore regulated by vegetation cover and Las Vegas
in a desert region. Adapted from Imhoff, Zhang, Wolfe, and Bounoua (2010).

(Huang, Li, Zhao, & Zhu, 2008). There are also devastating human health impacts as has
been shown by the excess mortality of 15,000 people in France during the heat wave of the
2003 summer, especially in the largest agglomerations (Hémon & Jougla, 2003).
Mapping the UHI produces an isothermal map with closed contours on the urban area
and wider contours over surrounding areas. The UHI thermal profile is represented
graphically by the isothermic curve rise throughout the urban area, which contrasts with
the characteristic low flattened curve of the surrounding areas. A typical UHI thermal
profile shows undulations of cliff, plateau, peak and depressions (Figure 1). The
surrounding area’s thermal field shows interruptions by a steep temperature gradient at the
surrounding/urban central boundaries (i.e. cliff), and thereafter a steady but weaker
horizontal gradient of increasing temperature (i.e. plateau) is prolonged until reaching the
highest temperature point at the urban core or city centre (i.e. peak).
The uniformity of this island profile’s shape pattern generally indicates some depressions
due to the presence of particular heat points (micro UHIs) associated with features such
as parking lots, malls, industrial facilities, etc., and some rises due to the presence of
particular heat sinks associated with features such as parks, fields, water bodies, etc. (Imhoff
et al., 2010). The difference between the warmest urban point and the least surrounding area
temperatures defines the intensity or magnitude of the UHI.
There is a growing base of studies on the UHI phenomenon in other parts of the globe
with the Asian continent being the most researched and the African continent the least.
In the African continent, few studies were reported: Adebayo (1987) and Goldreich
(1992). Investigating UHI over Ibadan, Adebayo (1987) made use of statistical
comparisons of air temperature values of the central urban area and its surroundings.
Goldreich (1992) measured temperatures at rooftops and ground levels in Johannesburg,
and reported differences in UHI by height.
There are few UHI studies done around the African continent that applied recent
remote-sensing techniques to establish and map the phenomenon. This review therefore
seeks to understand how a more up-to-date study could be done within the continent, given
South African Geographical Journal 3

the growth of megacities with increasing challenges on energy and sustainable living
standards by its urban population.
It will be confined to scientific literature relevant to the phenomenon and assessing
advances in the study of UHI investigation using remote sensing. Working through three
different sections, this review commences by describing the UHI phenomenon with some
of the factors influencing it and some relevant concepts. Some measurement techniques of
earth’s temperature (UHI) were examined from which a research gap around the African
continent was discovered. Section 2 then dwells on the remotely sensed data used this far
to investigate the UHI with a focus on satellite imagery. Section 3 addresses the processing
techniques applied from pre-processing to retrieving the land surface temperature (LST).
The review concludes with a summary of progress so far and comments on future
prospects.

2. Factors affecting UHI


The causative factors include urbanization and anthropogenic activities, because they
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

induce changes in the physical characteristics of the surface (albedo, thermal capacity,
heat conductivity, moisture) and changes in radiative fluxes and the near surface flow.
These two factors account for a huge replacement of soil and vegetation with pavements,
building structures and dark surfaces with urban material (concrete, asphalt and metal)
leading to the growth in thermal radiation which alters surface energy balance, with a
consequent increase in LST (Basar, Kaya, & Karaca, 2008; Lo & Quattrochi, 2003; Voogt,
2002; Zhang, Ji, Shu, Deng, & Wu, 2008). This leads to the increase in sensible heat flux at
the expense of latent heat flux and eventually air temperature (Wang et al., 2007).
Urban geometry is also another leading factor to UHI with canyon geometry as a result
of narrow streets and tall buildings which intercept air movement. The geometry also
contributes to this effect through wind flow (Sailor & Fan, 2002; Weng, Liu, & Lu, 2007).
Two primary weather characteristics that affect UHI development are wind and cloud
cover. In general, UHIs form during periods of calm winds and clear skies, because these
conditions maximize the amount of solar energy reaching urban surfaces and minimize the
amount of heat that can be convected away. Conversely, strong winds and cloud cover
suppress UHIs by convecting the warm air away or reflecting it, respectively (Kim & Baik,
2005).
Climate and topography, which are in part determined by a city’s geographic location,
influence UHI formation. For example, large bodies of water moderate temperatures and
can generate winds that convect heat away from cities. Nearby mountain ranges can either
block wind from reaching a city, or create wind patterns that pass through a city. Local
terrain has a greater significance for heat island formation when larger scale effects, such
as prevailing wind patterns, are relatively weak. Proximity to large water bodies and
mountainous terrain can influence local wind patterns and UHI formation (Oke, 1973).
The changes in land use/cover (LULC) pattern in relation to urbanization also impact
on the intensity and spatial pattern of the UHI effect (Chen, Zhao, Li, & Yin, 2006). This
has been affirmed with different land uses influencing UHI differently as a result of the
surface properties (Hung, Uchihama, Ochi, & Yasuok, 2006). Huang et al. (2008) used
four types of land covers in Nanjing (China) namely urban bare concrete cover, urban
woods or the shade of trees, urban water areas and urban lawn with their microclimates
examined, and the UHI analysed using air temperature data measured at four fixed
observation spots. The results revealed among others that the microclimate of these four
types of land covers had significantly different impacts on UHI in the order of bare
4 A. Ngie et al.

concrete cover . lawn . water areas . woods or the shade of trees during daytime with
reversed order during night-time.
A similar study has been done recently in Delhi (India), which confirmed the
difference in influence of land covers on the UHI phenomenon. Mallick, Rahman, Hoa,
and Joshi (2009) investigated how the phenomenon corresponds with characteristics of the
urban LULC. The study recommended that proper urban planning could avert the effects
of this phenomenon. It is vital in studying UHI to understand the LULC of the study area
which will in turn open up avenues to mitigate the effect.

2.1 Mitigation methods


Green areas are actually the ecological measure to combat the problems of concreted
environments. Any surface planted with vegetation has a different Bowen ratio1 than a
bare or mineral surface, since the incoming solar radiation is converted into energy for
transpiration and photosynthesis through plants and the sensible heat flux is consequently
lower. At night, the energy of the outgoing net radiation from a green surface is fed from
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

the thermal heat flux and the latent heat flux (Wong & Yu, 2005). Various studies point to
the importance of green vegetation in reducing the severity of UHI effect and cooling
impacts of green areas at a macro-level.
Different approaches have been used to illustrate the cooling effect over green areas
ranging from numerical models (Honjo & Takakura, 1991) to mesoscale atmospheric
models (Avissar, 1996). Remotely sensed Normalized Difference Vegetation Index
(NDVI) has been used to estimate surface temperatures (Li & Yu, 2008; Lo & Quattrochi,
2003; Yang & Liu, n.d.). Other studies used the NDVI to show that the presence of
vegetation and quantity is important in influencing LST. For example, Weng, Lu, and
Schubring (2004) suggested that the areal measure of vegetation abundance by unmixed
vegetation fraction has a more direct correspondence with the radiative, thermal and
moisture properties of the earth’s surface that determine LST.
In a recent study in Birmingham city (England), Tomlinson, Chapman, Thornes, and
Baker (2012) reported a significant cold spot in the conurbation being a city park with
recorded surface temperatures up to 78C lower than the city centre. Therefore, the
temperature over a green area is lower than that in a built environment, and the UHI effect
is aggravated mainly due to the loss of green areas in the urban environment. To this effect
it has been proposed that the vegetative fraction increase on the side walls of buildings in
residential canopies can result in daily and spatially averaged decreases in near-ground
summer air temperature of 0.2– 1.28C (Kikegawa et al., 2006).

3. UHI measurement approaches


UHI characteristics vary in different geographical locations as are meteorological conditions.
The UHI phenomenon has been studied in different parts of the world using various
approaches to measure its magnitude and possibly its extent. The phenomenon has long been
studied by ground-based observations taken from fixed thermometer networks or by traverses
with thermometers mounted on vehicles from which empirical data are generated and
analysed. For the sake of spatial mapping over larger areas and not just measuring the
magnitude of UHI, satellite remote sensing has recently been used (Li & Yu, 2008; Mirzaei &
Haghighat, 2010; Rhinane, Hilali, Bahi, & Berrada, 2012). This has provided new avenues
for the observation of UHI and the study of its causation through the combination of thermal
remote sensing and urban micrometeorology (Voogt & Oke, 2003).
South African Geographical Journal 5

UHIs may be identified by measuring surface or air temperatures. Surface


temperatures have an indirect but significant influence on air temperatures. For example,
parks and vegetated areas, which typically have cooler surface temperatures, contribute to
cooler air temperatures (Wong & Yu, 2005). Dense built-up areas, on the other hand,
typically lead to warmer air temperatures. Because air mixes within the atmosphere,
though, the relationship between surface and air temperatures is not constant (Nichol,
2005). Therefore, the measurement of UHI is done at various levels and by different tools.
According to Voogt (2002) UHI is best observed at clear and calm nights since maximal
differences in radiative cooling between central urban areas and surroundings occur.

3.1 LST measurements


All surfaces give off thermal energy that is emitted in wavelengths. Instruments on
satellites and other forms of remote sensing can identify and measure these wavelengths,
providing an indication of temperature. Satellite imagery can therefore provide
measurements that record energy reflected and emitted from the land surface, including
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

roofs, pavements, vegetation, bare ground and water. By using radiometers mounted on
aircraft or a satellite, many surface observations can easily be collected (Mirzaei &
Haghighat, 2010). For example, Landsat satellite data [Enhanced Thematic Mapper plus
(ETM þ ) bands 2, 3 and 4 for classification and band 6 for LST] can be used to classify
land cover and identify heat islands, respectively (Qian, Hai-Shan, & Chang, 2006).
Surface measurements taken by remote sensing have some limitations: first, they do
not fully capture radiant emissions from vertical surfaces, such as a building’s wall,
because the equipment mostly observes emissions from horizontal surfaces such as streets,
rooftops and treetops (Goldreich, 2006). Second, remotely sensed data represent radiation
that has travelled through the atmosphere twice, as wavelengths travel from the sun to the
earth as well as from the earth to the atmosphere (Mirzaei & Haghighat, 2010). Thus, the
data must be corrected to accurately estimate surface properties including solar reflectance
and temperature. However, remotely sensed temperature measurements provide wider
spatial coverage and higher temporal resolution. In addition, the relatively low cost of
imagery and the recent developments in image processing techniques make remote
sensing attractive for estimating temperature values over urban environments.

3.2 Air temperature measurements


Air temperatures are usually measured at about 5 ft (1.5 m) above the ground, where
standard weather observations are taken. A variety of platforms can be used to take these
measurements, which involve using hand-held measurement devices or mounting
measurement equipment on cars or aircraft (Wong & Yu, 2005).
Like surface temperature measurements, air temperature measurements have
limitations. The correlation between surface and air temperatures decreases as altitude
increases. Thus, the ability of air temperatures to provide a reliable indicator of the thermal
properties of various surfaces depends on the height of the measurement (Goldreich,
2006). Using air temperature measurements to evaluate a heat island is also complicated
by several factors:
(a) Data are needed from weather stations both within a city and in non-urban areas
close to the city.
(b) Changes in instrumentation, sampling, data recording methods and station
microclimate must be considered.
6 A. Ngie et al.

(c) Comparisons between central urban areas and surrounding areas’ temperature
records become less valid as areas around airports (often used to situate weather
stations) become urbanized (Mirzaei & Haghighat, 2010).
It is essential that the nature of the measurements used to define the heat island be
reported in a manner that is interpretable to the urban climate community. Likewise, great
care must be exercised in comparing UHIs if the medium sensed and the methods
employed in sensing it differ. There has been a prolific development in the study of UHIs
under urban climates with varying methods of measurements. The remotely sensed
approach has been prominent and has been improved over the years with the availability of
the data and improved techniques to both qualitatively and quantitatively investigate this
phenomenon.

4. Satellite remotely sensed data used


Meteorological data from weather stations or thermometers mounted on vehicles have
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

generated statistical data used to investigate UHI (Adebayo, 1987; Changnon, 1999; Huang
et al., 2008; Wilby, 2003; Wong & Yu, 2005). This has been considered as limited in space
and biased, since the readings are made on the spot where the equipment is situated. These
limitations have been overcome by the use of satellite data which provide more uniform
spatial sets of observations and wider coverage to the UHI. However, satellite imagery is
also guided by sensor characteristics and image resolution. Image resolution is usually
considered through three aspects which include spectral, spatial and radiometric
resolutions. The temporal resolution which is an indication of the revisiting frequency of a
sensor over a particular place is also an important guide for satellite imagery.
Remote sensing is extremely useful for understanding the spatio-temporal land cover
change in relation to the basic physical property of materials used on urban structures.
These properties related with the surface radiance and emissivity data (Joshi & Bhatt,
2012). The electro-magnetic wavelength region of 3 –35 mm is called thermal-infrared
(TIR) region. The useful spectral bands in this range are limited by the intensity of
radiation emitted and the atmospheric windows. An excellent atmospheric window lies
between 8 and 14 mm, in which most remote sensors are set up to detect the thermal
radiative properties of the ground materials (Gupta, 1991).
Currently available satellite TIR sensors provide different spatial resolution and
temporal coverage data that can be used to estimate LST. The Geostationary Operational
Environmental Satellite has a 4-km resolution in the TIR, while National Oceanic and
Atmospheric Administration (NOAA), Advanced Very High Resolution Radiometer
(AVHRR) and Terra and Aqua Moderate Resolution Imaging Spectroradiometer
(MODIS) have 1-km spatial resolutions. Significantly, better-resolution data come from
the Terra-Advanced Space borne Thermal Emission and Reflection Radiometer (ASTER),
which has a 90-m spatial resolution, meanwhile the Landsat 5 Thematic Mapper (TM) has
a 120-m resolution, and Landsat 7 ETM þ has a 60-m resolution in the TIR bands. Better
resolution and less frequent TIR observations from Landsat might be used to understand
the spatial variation within coarser-resolution observations made by MODIS and AVHRR,
which provide more frequent measurements. Recent satellite systems, e.g. MODIS and
ASTER, include features to allow easier calibration and provide LST as standard products.
This is not true for the Landsat data (Li et al., 2004).
Satellite-derived surface temperature data have been utilized for urban climate
analyses in several studies (Carlson, Augustine, & Boland, 1977; Carnahan & Larson,
South African Geographical Journal 7

1990; Kidder & Wu, 1987; Lo, Quattrochi, & Luvall, 1997; Matson, McClain, McGinnis,
& Pritchard, 1978; Price, 1979, 1984; Rao, 1972; Roth, Oke, & Emery, 1989; Voogt &
Oke, 2003). The first research conducted on urban heat island investigation using satellite
remote sensing was performed by Rao (1972). VHRR thermal data acquired at night were
used to examine urban and rural surface temperature differences (Matson et al., 1978).
UHI is defined along space (urban/rural boundaries) and it would be worthwhile to
consider spatial resolution for classification of sensor types in this review paper.
Coarseness or fineness of a raster grid is what describes spatial resolution (Lillesand,
Kiefer, & Chipman, 2008). The grid cells correspond to ground areas where the fine spatial
resolution will refer to smaller area size, such as those produced by IKONOS at 4 m,
medium spatial resolution such as Landsat ETM þ sensor being 30 m and the low spatial
resolution such as MODIS being 50 m (or even NOAA’s AVHRR sensor being 1 km).
The finer the spatial resolution of a digital image, the more detail it contains. Detail is
valuable for some applications such as LULC classification, but it is also costly. Hence,
most research on the UHI phenomenon applied medium spatial resolution with the
TIR band.
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

4.1 Fine resolution


With a spatial resolution of 0.6– 4 m, satellite sensors are classified as fine and examples
include IKONOS, Spot-5, QuickBird, EarlyBird, GeoEye-1 and others (Lillesand et al.,
2008). Sensors operating at this spatial resolution provide very detailed images but lack
the TIR band which is useful in UHI. Hence, they can only be used in any UHI project to
enable a finer landuse/cover classification (Chen et al., 2006).
Day and night airborne-TIR image data at 5 m spatial resolution acquired with the
15-channel (0.45 –12.2 mm) Advanced Thermal and Land Applications Sensor over
Alabama, Huntsville were used to study changes in the thermal signatures of urban land
cover types between day and night. Thermal channel number 13 (9.60 – 10.2 mm) data with
the best noise-equivalent temperature change (Lo et al., 1997).

4.2 Medium resolution


The medium category is considered of spatial resolution range between 4 and 30 m.
Landsat is one of the most used sensors in this category for UHI studies. This is as a result
of its high-quality multispectral data with worldwide coverage for historical and multi-
temporal data that are easily accessible. It also has the TIR on band 6 which is suitable for
estimating temperature even though the spatial resolution of this band is coarse [120 m for
Multispectral Scanner (MSS) and TM and 60 m for the ETM þ ; Lillesand et al., 2008].
Another important sensor in use within this category is the ASTER sensor with a visible
and near-infrared (VNIR) data at 15 m like the ETM þ panchromatic band but currently
considered the best multispectral satellite data available commercially with the exception
of very high resolution data such as IKONOS or QuickBird (Lillesand et al., 2008). At a
15-m spatial resolution, these data are also good for landuse/cover classification. Even
though ASTER is spatially and spectrally better than ETM þ , there is evidence that more
studies used ETM þ only (Bechtel, 2011; Weng et al., 2004, 2007; Yang & Liu, n.d.;
Zhang et al., 2008) than ASTER only (Cai, Du, & Xue, 2008; Zhihao, Li, Gao, & Zhang,
2006) which may be partly due to insufficient historical data.
Investigation of the relationship between surface urban heat islands (SUHIs) and the
per cent impervious surface area was carried out in Shanghai, China. The per cent
8 A. Ngie et al.

impervious surface area was characterized from a Landsat ETM þ multispectral data set
using the linear mixture spectral analysis. LST derived from the ETM þ spectral data
was proved to be a good surrogate for SUHI. Image-induced LST can evaluate urban
surface temperature not only in quantity but also in spatial patterns. However, LST cannot
replace atmosphere temperature; each one has its own meteorological functions. For
this reason, the UHI identified from remote sensing imagery can only be called SUHI
(Zhang et al., 2008).
The 120-m spatial resolution Landsat TM TIR data has also been processed to derive
surface temperature. Landsat TM TIR data were used to observe meso-scale temperature
differences between the urban and rural areas in Indianapolis (Carnahan & Larson, 1990;
Larson & Carnahan, 1997), while a similar phenomenon was studied in Washington, DC
(Kim, 1992). A significant attempt was carried out to utilize TM thermal data to monitor
microclimate for housing estates in Singapore (Nichol, 1994). It was concluded that there
was a high correlation between satellite-derived temperature and biomass indices, as well
as similarity with air temperature data, but satellite-derived temperatures were not good
for obtaining absolute quantitative values for ambient air temperature.
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

4.3 Coarse resolution


Coarse resolution sensors have a spatial resolution of . 30 m. MODIS is the common
example in this category generating data at several bands. MODIS data are suitable for
continuous monitoring and mapping of surface clouds and has global coverage. For UHI
studies over large scene areas, MODIS is very suitable given its spatial resolution that
ranges from 250 to 1000 m. MODIS has a hyperspectral range of 36 bands of which 16 are
thermal bands and the atmospheric temperature is covered by bands 24 (4.433 – 4.498 mm)
and 25 (4.482 – 4.549 mm) (Lillesand et al., 2008). The AVHRR on board NOAA is
another sensor being used for UHI investigation since it simultaneously records data in
four bands at a spatial resolution of 1.1 km in the VNIR and TIR portions of the spectrum
(Streutker, 2002).
Data from NOAA and AVHRR were used to determine urban/rural differences in
surface temperature and vegetation index for 37 US cities. It was found that the difference
in vegetation index showed a positive correlation with the differences in both surface and
air temperature (Gallo et al., 1993).
The AVHRR thermal data can be utilized to assess the UHI intensities for several cities
as was the case in North America (Roth et al., 1989). Daytime thermal patterns of surface
temperature were associated with land use and higher surface temperatures were observed
in industrial areas than in vegetated regions. Night-time observations revealed higher
surface temperatures of urban central areas than their surroundings. This suggests that the
sides of buildings (rather than roofs) and the characteristics of other urban features may
contribute to the observed higher temperatures. The greatest air temperature differences
are observed at night, while the greatest differences in surface radiant temperature are
observed during midday (Roth et al., 1989).

5. Analysis techniques
The estimated temperature of the surface requires processing and knowledge of several
parameters. In this process, an atmospheric correction (noise removal) is useful to
eliminate the influence of the atmosphere and isolate the spectral signatures of terrestrial
objects (Rhinane et al., 2012). These atmospheric effects include absorption, upward
South African Geographical Journal 9

emissions and downward irradiance reflected from the surface. Identifying landuse/cover
with satellite images is of great importance before calculating the UHI (Cai et al., 2008).

5.1 Pre-processing
Generally pre-processing involves correction of remotely sensed data through
radiometric and geometric rectifications. Some data providers are able to make available
radiometrically calibrated and geometrically co-registered data such as ASTER, Landsat,
MODIS and others. With such data provided, pre-processing would involve classification
of the non-thermal bands to show landuse/cover or urbanization levels in study areas
(Bechtel, 2011; Chen et al., 2006; Li & Yu, 2008; Lo & Quattrochi, 2003; Mallick et al.,
2009).
In terms of atmospheric correction, imagedigital numbers are converted to normalized
atmospheric reflectance using the provided formulae by sensor handbooks as well as
calibration parameters retrieved from the image header files (Zhang et al., 2008). Some
authors reported that atmospheric correction using the NASA Atmospheric Correction
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

Parameter Calculator was possible to retrieve LST (Mallick et al., 2009; Yang & Liu, n.d.).
There are periodic noise such as bands and stripes in the Landsat TM/band 6 and non-
periodic noise (speckles) that may affect the retrieval of brightness temperature or LST.
Noise removal techniques such as self-adaptive filter can be used for non-periodic noise
and the fast Fourier transform method to automatically remove periodic noise (Chen et al.,
2006).
In a case where images from multiple sensors with varying spatial resolutions are used,
there has to be some unifying process beginning with georeferencing to a common
coordinate system, outputting to a single-scale spatial resolution for all bands (Chen et al.,
2006; Hung et al., 2006; Lo & Quattrochi, 2003). A radiometric normalization to register
minimal radiometric differences caused by sensor variations and changes in sensor-target-
illumination geometry can also be conducted when using data from different generations
of sensors such as landsat MSS and TM (Lo & Quattrochi, 2003; Zhang et al., 2008).

5.2 Processing methods to compute LST


Methods used to analyse the data include the reference channel method (Kahle, Madura, &
Soha, 1980), the emissivity normalization method (Gillespie, 1985; Realmuto, 1990), the
spectral ratio method (Watson, 1992), the alpha residuals method (Kealy & Gabell, 1990)
and the temperature and emissivity separation algorithm (Gillespie et al., 1998;
Schmugge, French, Ritchie, Rango, & Pelgrum, 2002). However, for satellites with a
single thermal band, such as Landsat TM and ETM þ , obtaining LST is more difficult.
In addition to an accurate radiative transfer model and some knowledge of the atmospheric
profile, emissivity information is also required (Qin, Karnieli, & Berliner, 2001).
Two approaches have been developed to recover LST from multispectral TIR imagery.
The first approach uses a radiative transfer equation to correct the at-sensor radiance to
surface radiance, followed by an emissivity model to separate the surface radiance into
temperature and emissivity (Schmugge, Hook, & Coll, 1998). The second approach
applies the split window technique for sea surfaces to land surfaces, assuming that the
emissivity in the channels used for split window is similar. Land surface brightness
temperatures are then calculated as a linear combination of the two channels (Table 1).
A disadvantage of this approach is that the coefficients are only valid for the data sets used
to derive those coefficients (Dash, Gottsche, Olesen, & Fischer, 2002).
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

10

Table 1. Summary of LST retrieval methods using satellite remote sensing.

References Method Sensor RMSE Study area


Price (1984), Sobrino and Split-window algorithm AVHRR Acceptable, satisfactory, Valencian region, semi-arid region
A. Ngie et al.

Caselles (1991) and 0.4 K of northwestern Victoria in Australia


Sobrino, Li, Stoll, and and mountainous region of northeast
Becker (1994) of France
Wan and Dozier (1996) Generalized split-window algorithm AVHRR/MODIS Favourable accuracy
Wan (2008) and Wan et al. Generalized split-window algorithm MODIS Better than 1 K Lake Tahoe in California and
(2002) Namco lake in Tibet; Lake Titicaca
in Bolivia, Mono Lake, Bridgeport
grassland, and a rice field in Chico,
CA and Walker Lake, Nevada
Sobrino et al. (2004) Qin et al.’s mono-window ASTER 2K Requena-Utiel, Valencia, Spain
Sobrino and Jiménez-Muñoz’s 0.9 K
single channel algorithm 2003
The radiative transfer equation 0.6 K
Cai et al. (2008) Split-window algorithm Landsat TM/ETM þ Beijing
Zhang, Wang, and Li Qin et al.’s mono-window 0.5– 1.5 K Guangzhou city, Guangdong, China
(2006)
Sobrino and Jiménez-Muñoz’s Landsat TM/ETM þ 0.7– 1.5 K
single channel algorithm 2003
Qin et al. (2001) Qin et al.’s mono-window TM/ASTER 1.18C Israel – Egypt border region
Liu and Zhang (2011) 0.7/0.8 K Hong Kong
Mao, Shi, Li, and Tang Radiative transfer model-neural MODIS 0.36 K Xiaotangshan
(2007) network (RM-NN)
South African Geographical Journal 11

As far as the Landsat TM and ETM þ are concerned, three different single-channel
methods have been proposed to retrieve LST from their thermal bands. These three
methods are: (i) the radiative transfer equation, (ii) Qin et al.’s (2001) mono-window
algorithm and (iii) Jiménez-Muñoz and Sobrino’s algorithm (Sobrino, Jiménez-Muñoz, &
Leonardo, 2004). The first method requires an in situ atmospheric profile launched
simultaneously with the satellite pass and this is a major constraint for using this method.
Usually, the second and third methods are used when the ground truth data are not
available. The at-sensor brightness temperature values derived from Landsat 5 thermal
data are less than what is actually measured at ground level (weather stations). Therefore,
ancillary data such as ground emissivity and atmospheric correction should be used to
calibrate at-sensor brightness temperature to much accurate LST from TM/ETM data
(Fuqin et al., 2004). One of the techniques for estimating emissivity that can be applied is
the fractional cover mixture model. Soil background and the vegetation have specific
known emissivity and that they ‘mix’ according to the fractional cover (Sobrino,
Raissouni, & Li, 2001). Hence, fractional vegetation cover could be estimated from the
NDVI. Other empirical models correlating NDVI and emissivity have been established
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

(Valor & Caselles, 1996; Van de Griend & Owe, 1993).


There are three commonly used methods to retrieve LST, depending mostly on how
the sensor thermal bands were designed. These are the split window method, such as is
used for NOAA –AVHRR data (Becker & Li, 1990; Price, 1984), the day –night pairs of
TIR data in several bands, such as that used for MODIS (Wan, Zhang, Zhang, & Li, 2002)
and single band correction (Price, 1983). Some sensors have multi-thermal bands, which
allow simultaneous estimation of surface emissivity and LST. A number of methods have
been developed to separate surface emissivity and LST using multi-thermal (e.g. ASTER
and MODIS) band data (Li, Becker, Stoll, & Wan, 1999).

6. Summary and future research needs


Interpreting thermal data and images of temperature distribution over an area is often not
straightforward due to many complex controlling factors. The most important factors
controlling the urban canopy layer heat island are the distribution of surface cover
characteristics, and urban morphology, such as building materials, geometry and density
(Oke, 1982). It would be interesting to relate the diversity in building structures and layout
within African cities to the UHI effect.
Several important factors to future development in satellite remote sensing of UHIs
can be identified based on the examination of the current literature: first, the determination
of surface radiant temperature can be more complicated than the methods presently used.
Major technical and theoretical difficulties in using TIR remotely sensed data obtained at
different spatial and temporal scales related to the scale-dependent nature of landscape
characteristics, physiography, emissivity, atmospheric effects and sensor-to-target noise
(Quattrochi & Goel, 1995). The effects of surface roughness on surface temperature shall
be taken into account. It is important to examine the temperatures of each part of the
vegetation-ground system (such as shaded ground, sunny ground, shade vegetation and
sunny vegetation) and to examine the effects of different canopy structures. Effective LST
can only be derived after its relationship to the component temperatures has been
mathematically modelled. Effective measurement of surface temperatures further requires
an analysis of the influence of the nature of surface and its roughness on emissivity. The
emissivity values for urban areas are highly variable and need to be refined in future
studies (Cassels, Sobrino, & Coll, 1992; Kimes, 1983).
12 A. Ngie et al.

Second, the relationship between LST and NDVI needs to be further calibrated.
The nonlinearity of the relationship and the platform dependency of NDVI suggest that it
may not be a good indicator for quantitative analyses of vegetation. More quantitative,
physically based measures of vegetation abundance are needed for applications that
require biophysical measures (Small, 2001). The importance of spatial resolution for
detecting landscape patterns and changes should also be emphasized (Frohn, 1998), and
the relationship between NDVI variability and pixel size should be further investigated
(Jasinski, 1990).
Third, the effect of urban morphology on UHI should be further investigated. More
research is desirable to apply height measurements and elevation data to the UHI studies.
Radar-generated topography data (e.g. Shuttle Radar Topography Mission (SRTM), Light
Detection and Ranging (LIDAR), Interferometric Synthetic Aperture Radar (IFSAR) data)
are especially desirable to relate urban morphology and topography to UHI studies. As the
high accuracy of elevation data becomes available, the benefit of data fusion for UHI
studies will become obvious. Information on building heights is not always obtainable.
The ASTER DEM data set contains topographic information derived from the along-track,
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

15-m ASTER optical stereo data acquired in near-infrared bands 3N (nadir-viewing) and
3B (backward-viewing). These high spatial resolution digital elevation models (DEM) can
be used to derive urban topography, slope and aspect over horizontal distances of more
than 100 m (Welch, Jordan, Lang, & Murakami, 1998). After the data quality of ASTER
derived topography has been examined, the fusion of the topography data with the ASTER
VNIR and SWIR data sets can be undertaken to interpret urban geometries at relevant
resolution levels, so that urban structures can be related to the infrared measured LSTs.
Finally, remotely sensed data from different sensors should be acquired to examine the
effects of various spectral and spatial resolutions on the resultant LST and texture
measurements. For example, application of fractals to data from the NASA Earth
Observation System sensors (e.g. Landsat ETM þ , ASTER, MODIS) holds great
potential for the analysis and characterization of urban thermal landscape, as well as for
the planning and development of more advanced TIR platforms. It should be noted that
most of the studies used multi-sensor satellite data comprising of a coarse resolution
usually for mapping and analysing temporal spatial variations of UHI and a medium
resolution to study the relation of UHI to urban surface characteristics (Hung et al., 2006).
With the advancements in remote sensing applications over the years, satellite images
have been used in support of other data inputs to study UHI. These include geographical
information system (GIS) layers and meteorological data. Commonly used GIS data layers
include cadastral data, topographic maps, LULC maps and NDVI. The LULC and NDVI
are also created at times from the satellite data. The meteorological support data or
calibration includes air temperature, relative humidity, wind speed, emissivity, daily
sunshine hours, precipitation, evaporation and pressure.

7. Conclusion
This review illustrates that remotely sensed data has been widely used to investigate the
UHI phenomenon. With the continual innovations in this technology, its advantages in
both methods and processing techniques have increasingly widened. Data sources have
also expanded with better spatial and spectral resolutions, and frequent global coverage.
According to this review, UHI studies have been conducted mostly in the Americas,
Europe and Asia. The African continent also has a need for such studies in order to
inform its city planners to cope with the challenges of urban climates (socio-economic and
South African Geographical Journal 13

human impacts). Such studies will help the planners and decision-makers to be proactive
rather than reactive in urban planning and management.

Acknowledgements
The authors wish to express their appreciation to the National Research Foundation (NRF, South
Africa) and the Ministry of State for Scientific Research, Egypt, for providing funds. We also
acknowledge the input of the anonymous reviewers.

Funding
This work was funded by the NRF (South Africa) and the Ministry of State for Scientific Research,
Egypt, through the Joint Research grant under the South African/Egypt Research Partnership
Programme Bilateral Agreement [grant number UID78639].

Note
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

1. Bowen ratio is the proportion of sensible heat to latent heat leaving a surface (Campbell, 1977))

References
Adebayo, Y. R. (1987). A note on the effect of urbanization on temperature in Ibadan. Journal of
Climatology, 7, 185– 192.
Arnfield, 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 – 26.
Avissar, B. (1996). Potential effects of vegetation on the urban thermal environment. Atmospheric
Environment, 30, 437– 448.
Basar, U. G., Kaya, S., & Karaca, M. (2008). Evaluation of urban heat island in Istanbul using remote
sensing techniques. Commission VII, WG VII/5. The international archives of the
photogrammetry, remote sensing and spatial information sciences (Vol. XXXVII, Part B7).
Beijing: International Society for Photogrammetry and Remote Sensing (ISPRS).
Becker, F., & Li, Z. L. (1990). Temperature-independent spectral indices in thermal infrared bands.
Remote Sensing of Environment, 32, 17 – 33.
Bechtel, B. (2011). Multitemporal Landsat data for Urban Heat Island assessment and classification
of local climate zones. In U. Stilla, P. Gamba, C. Juergens, & D. Maktav (Eds.), JURSE 2011 –
Joint Urban Remote Sensing Event, Munich, Germany, April 11 – 13 (pp. 129– 132). Institute of
Electrical and Electronics Engineers (IEEE).
Cai, G., Du, M., & Xue, Y. (2008). Monitoring of seasonal change of urban heat island effect in
Beijing using ASTER data. In Deren Li, Jianya Gong, & Huayi Wu (Eds.), International
Conference on Earth Observation Data Processing and Analysis (ICEODPA), Proceedings of
SPIE, (Vol. 7285, 72855B), q 2008 SPIE, CCC code: 0277-786X/08/$18. Bellingham, WA:
The International Society for Optical Engineering. doi:10.1117/12.812413
Campbell, G. S. (1977). An introduction to environmental biophysics. New York, NY: Springer.
Carlson, T. N., Augustine, J. A., & Boland, F. E. (1977). Potential application of satellite temperature
measurements in the analysis of land use over urban areas. Bulletin of the American
Meteorological Society, 58, 1301–1303.
Carnahan, W. H., & Larson, R. C. (1990). An analysis of an urban heat sink. Remote Sensing of
Environment, 33, 65 – 71.
Cassels, V., Sobrino, J. A., & Coll, C. (1992). A physical model for interpreting the land surface
temperature obtained by remote sensors over incomplete canopies. Remote Sensing of
Environment, 39, 203– 211.
Changnon, S. A. (1999). A rare long record of deep soil temperatures defines temporal temperature
changes and an urban heat island. Climatic Change, 42, 531– 538.
Chen, X. L., Zhao, H. M., Li, P. X., & Yin, Z. Y. (2006). Remote sensing image-based analysis of the
relationship between urban heat island and land use/cover changes. Remote Sensing of
Environment, 104, 133– 146.
14 A. Ngie et al.

Dash, P., Gottsche, F. M., Olesen, F. S., & Fischer, H. (2002). Land surface temperature and
emissivity estimation from passive sensor data: Theory and practice – Current trends.
International Journal of Remote Sensing, 23, 2563– 2594.
Frohn, R. C. (1998). Remote sensing for landscape ecology. Boca Raton, FL: Lewis.
Fuqin, L., Thomas, J. J., Williams, P. K., Thomas, J. S., Andrew, N. F., Michael, H. C., & Rajat, B.
(2004). Deriving land surface temperature from Landsat 5 and 7 during SMEX02/SMACEX.
Remote Sensing of Environment, 92, 521– 534.
Gallo, K. P., McNab, A. L., Karl, T. R., Brown, J. F., Hood, J. J., & Tarpley, J. D. (1993). The use of
NOAA AVHRR data for assessment of the urban heat island effect. Journal of Applied
Meteorology, 32, 899–908.
Gillespie, A. R. (1985). Lithologic mapping of silicate rocks using TIMS. Proceedings of the thermal
infrared multispectral scanner (TIMS) Data User’s Workshop, 86-38 (pp. 29 – 44). Pasadena,
CA: Jet Propulsion Laboratory.
Gillespie, A. R., Rokugawa, S., Matsunaga, T., Cothern, J. S., Hook, S. J., & Kahle, A. B. (1998).
A temperature and emissivity separation algorithm for advanced spaceborne thermal emission
and reflection radiometer (ASTER) images. IEEE Transactions on Geoscience and Remote
Sensing, 36, 1113– 1126.
Goldreich, Y. (2006). Ground and top of canopy layer urban heat island partitioning on an airborne
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

image. Remote Sensing of Environment, 104, 247– 255.


Goldreich, Y. (1992). Urban climate studies in Johannesburg, a sub-tropical city located on a ridge –
A review. Atmospheric Environment, 26B, 407–420.
Gupta, R. P. (1991). Remote sensing geology. Germany: Springer-Verlag Berlin Heidelberg.
Hémon, D., & Jougla, E. (2003). Estimation de la surmortalité et principales caractéristiques
épidémiologiques (Evaluation of comparatively high death rate and epidemiologic main
features). Technical report, INSERM. Retrieved May 8, 2012, from http://www.sante.gouv.fr/
htm/actu/surmort_canicule/rapport_complet.pdf
Honjo, T., & Takakura, T. (1991). Simulation of thermal effects of urban green areas on their
surrounding areas. Energy and Buildings, 15, 443– 446.
Huang, L., Li, J., Zhao, D., & Zhu, J. (2008). A fieldwork study on the diurnal changes of urban
microclimate in four types of ground cover and urban heat island of Nanjing, China. Building
and Environment, 43, 7 – 17.
Hung, T., Uchihama, D., Ochi, S., & Yasuok, Y. (2006). Assessment with satellite data of the urban
heat island effects in Asian mega cities. International Journal of Applied Earth Observation and
Geoinformation, 8, 34 – 48.
Imhoff, M. L., Zhang, P., Wolfe, R. E., & Bounoua, L. (2010). Remote sensing of the urban heat
island effect across biomes in the continental USA. Remote Sensing of Environment, 114,
504– 513.
Jasinski, M. F. (1990). Sensitivity of the Normalized Difference Vegetation Index to subpixel
canopy cover, soil albedo, and pixel scale. Remote Sensing of Environment, 32, 169– 187.
Joshi, J. P., & Bhatt, B. (2012). Estimating temporal land surface temperature using remote sensing:
A study of Vadodara urban area, Gujarat. International Journal of Geology, Earth and
Environmental Sciences, 2, 123–130. ISSN: 2277-2081.
Kahle, A. B., Madura, D. P., & Soha, J. M. (1980). Middle infrared multispectral aircraft scanner
data: Analysis for geological application. Applied Optics, 19, 2279– 2290.
Kealy, P. S., & Gabell, A. R. (1990). Estimation of emissivity and temperature using alpha
coefficients. Proceedings of Second TIMS Workshop, 90-55 (pp. 11 – 15). Pasadena, CA: Jet
Propulsion Laboratory.
Kidder, S. Q., & Wu, H. T. (1987). A multispectral study of the St. Louis area under snow-covered
conditions using NOAA-7 AVHRR data. Remote Sensing of Environment, 22, 159– 172.
Kikegawa, Y., Genchi, Y., Kondo, H., & Hanaki, K. (2006). Impacts of city-block-scale
countermeasures against urban heat-island phenomena upon a building’s energy-consumption
for air-conditioning. Applied Energy, 83, 649– 668.
Kim, H. H. (1992). Urban heat island. International Journal of Remote Sensing, 13, 2319– 2336.
Kim, Y. -H., & Baik, J. -J. (2005). Spatial and temporal structure of the urban heat island in Seoul.
American Meteorological Society, 44, 591– 605.
Kimes, D. J. (1983). Remote sensing of row crop structure and component temperatures using
directional radiometric temperatures and inversion techniques. Remote Sensing of Environment,
13, 33 – 55.
South African Geographical Journal 15

Larson, R. C., & Carnahan, W. H. (1997). The influence of surface characteristics on urban radiant
temperatures. Geocarto International, 12, 5 – 16.
Li, F., Jackson, T., Kustas, W., Schmugge, T., French, A., Cosh, M., & Bindlish, R. (2004). Deriving
land surface temperature from Landsat 5 and 7 during SMEX02/SMACEX. Remote Sensing
of Environment, 92, 521– 534.
Li, K., & Yu, Z. (2008). Comparative and combinative study of urban heat island in Wuhan City with
remote sensing and CFD simulation. Sensors, 8, 6692– 6703.
Li, Z., Becker, F., Stoll, M., & Wan, Z. (1999). Evaluation of six methods for extracting relative
emissivity spectra from thermal infrared images. Remote Sensing of Environment, 69, 197–214.
Lillesand, T. M., Kiefer, R. W., & Chipman, J. W. (2008). Remote sensing and image interpretation
(6th ed.). Hoboken, NJ: Wiley.
Lo, C. P., & Quattrochi, D. A. (2003). Land-use and land-cover change, urban heat island
phenomenon, and health implications: A remote sensing approach. Photogrammetric
Engineering and Remote Sensing, 69, 1053– 1063.
Lo, C. P., Quattrochi, D. A., & Luvall, J. C. (1997). Application of high-resolution thermal infrared
remote sensing and GIS to assess the urban heat island effect. International Journal of Remote
Sensing, 18, 287– 304.
Liu, L., & Zhang, Y. (2011). Urban heat island analysis using the Landsat TM data and ASTER data:
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

A case study in Hong Kong. Remote Sensing, 3, 1535– 1552.


Mallick, J., Rahman, A., Hoa, P. V., & Joshi, P. K. (2009). Assessment of night-time urban surface
temperature – LULC relationship for thermal urban environment studies using optical and
thermal satellite data. 7th FIG Regional Conference Spatial Data Serving People: Land
Governance and the Environment – Building the Capacity Hanoi, Vietnam, 19 – 22 October
2009.
Mao, K., Shi, J., Li, Z. -L., & Tang, H. (2007). An RM-NN algorithm for retrieving land surface
temperature and emissivity from EOS/MODIS data. Journal of Geophysical Research, 112
(D21102), 1 – 17.
Matson, M., McClain, E. P., McGinnis, D. F., & Pritchard, J. A. (1978). Satellite detection of urban
heat islands. Monthly Weather Review, 106, 1725– 1734.
Mirzaei, P. A., & Haghighat, F. (2010). Approaches to study Urban Heat Island – Abilities and
limitations. Building and Environment, 45, 2192– 2201.
Nichol, J. (1994). A GIS-based approach to microclimate monitoring in Singapore’s high-rise
housing estates. Photogrammetric Engineering and Remote Sensing, 60, 1225– 1232.
Nichol, J. (2005). Remote sensing of urban heat islands by day and night. Photogrammetric
Engineering and Remote Sensing, 71, 613– 621.
Oke, T. R. (1973). City size and the urban heat island. Atmospheric Environment, 7, 769– 779.
Oke, T. R. (1982). The energetic basis of the Urban Heat Island. Journal of Royal Meteorological
Society, 108, 1 –24.
Price, J. C. (1979). Assessment of the urban heat island effect through the use of satellite data.
Monthly Weather Review, 107, 1554– 1557.
Price, J. C. (1983). Estimating surface temperature from satellite thermal infrared data – A simple
formulation for the atmospheric effect. Remote Sensing of Environment, 13, 353– 361.
Price, J. C. (1984). Land surface temperature measurements from the split window channels of the
NOAA 7 Advanced Very High Resolution Radiometer. Journal of Geophysical Research, 89,
7231– 7237.
Qian, L. -X., Hai-Shan, C., & Chang, J. (2006). Impacts of land use and cover change on land surface
temperature in the Zhujiang Delta. Pedosphere, 16, 681– 689.
Qin, Z., Karnieli, A., & Berliner, P. (2001). A mono-window algorithm for retrieving land surface
temperature from Landsat TM data and its application to the Israel-Egypt border region.
International Journal of Remote Sensing, 18, 3719– 3746.
Quattrochi, D. A., & Goel, N. S. (1995). Spatial and temporal scaling of thermal remote sensing data.
Remote Sensing Review, 12, 255– 286.
Rao, P. K. (1972). Remote sensing of urban heat islands from an environmental satellite. Bulletin of
the American Meteorological Society, 53, 647– 648.
Realmuto, V. J. (1990). Separating the effects of temperature and emissivity: Emissivity spectrum
normalization. Proceedings of the 2nd TIMS workshop, 90-55 (pp. 31 – 37). Pasadena, CA: Jet
Propulsion Laboratory.
16 A. Ngie et al.

Rhinane, H., Hilali, A., Bahi, H., & Berrada, A. (2012). Contribution of Landsat TM data for the
detection of urban heat islands areas case of Casablanca. Journal of Geographic Information
System, 4, 20 – 26.
Roth, M., Oke, T. R., & Emery, W. J. (1989). Satellite derived urban heat islands from three coastal
cities and the utilisation of such data in urban climatology. International Journal of Remote
Sensing, 10, 1699– 1720.
Sailor, D. J., & Fan, H. (2002). Modeling the diurnal variability of effective albedo for cities.
Atmospheric Environment, 36, 713– 725.
Schmugge, T., French, A., Ritchie, J. C., Rango, A., & Pelgrum, H. (2002). Temperature and
emissivity separation from multispectral thermal infrared observations. Remote Sensing of
Environment, 79, 189– 198.
Schmugge, T., Hook, S. J., & Coll, C. (1998). Recovering surface temperature and emissivity from
thermal infrared multispectral data. Remote Sensing of Environment, 65, 121– 131.
Small, C. (2001). Estimation of urban vegetation abundance by spectral mixture analysis.
International Journal of Remote Sensing, 22, 1305– 1334.
Sobrino, J. A., & Caselles, V. (1991). A methodology for obtaining the crop temperature from
NOAA-9 AVHRR data. International Journal of Remote Sensing, I2, 2461– 2475.
Sobrino, J. A., Jiménez-Muñoz, J. C., & Leonardo, P. (2004). Land surface temperature retrieval
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

from LANDAT TM5. Remote Sensing of Environment, 90, 434– 440.


Sobrino, J. A., Li, Z. L., Stoll, M. P., & Becker, F. (1994). Improvements in the split window
technique for land surface temperature determination. IEEE Transactions on Geoscience and
Remote Sensing, 32, 243– 253.
Sobrino, J. A., Raissouni, N., & Li, Z. (2001). A comparative study of land surface emissivity
retrieval from NOAA data. Remote Sensing of Environment, 75, 256– 266.
Streutker, D. R. (2002). A remote sensing study of urban heat island of Houston, Texas. International
Journal of Remote Sensing, 23, 2595– 2608.
Tomlinson, C. J., Chapman, L., Thornes, J. E., & Baker, C. J. (2012). Derivation of Birmingham’s
summer surface urban heat island from MODIS satellite images. International Journal of
Climatology, 32, 214– 324.
Valor, E., & Caselles, V. (1996). Mapping land surface emissivity from NDVI: Application to
European, African, and South American areas. Remote Sensing of Environment, 57, 167– 184.
Van de Griend, A. A., & Owe, M. (1993). On the relationship between thermal emissivity and the
Normalized Difference Vegetation Index for natural surfaces. International Journal of Remote
Sensing, 14, 1119– 1131.
Voogt, J. (2002). Urban heat island. In T. Munn (Ed.), Encyclopedia of global environmental change
(Vol. 3, 660– 666). New York, NY: Wiley.
Voogt, J. A., & Oke, T. R. (2003). Thermal remote sensing of urban climates. Remote Sensing of
Environment, 86, 370– 384.
Wan, Z. (2008). New refinements and validation of the MODIS land-surface temperature/emissivity
products. Remote Sensing of Environment, 112, 59 – 74.
Wan, Z., & Dozier, J. (1996). A generalized split-window algorithm for retrieving land-surface
temperature from space. IEEE Transactions on Geoscience and Remote Sensing, 34, 892–905.
Wan, Z., Zhang, Y., Zhang, Q., & Li, Z. (2002). Validation of the landsurface temperature products
retrieved from Terra Moderate Resolution Imaging Spectroradiometer data. Remote Sensing of
Environment, 83, 163– 180.
Wang, K., Wang, J., Wang, P., Sparrow, M., Yang, J., & Chen, H. (2007). Influences of urbanization
on surface characteristics as derived from the moderate-resolution imaging spectroradiometer:
A case study for the Beijing metropolitan area. Journal of Geophysical Research, 112, 1 – 12.
Watson, K. (1992). Spectral ratio method for measuring emissivity. Remote Sensing of Environment,
42, 113– 116.
Welch, R., Jordan, T., Lang, H., & Murakami, H. (1998). ASTER as a source for topographic data in
the late 1990’s. IEEE Transactions on Geoscience and Remote Sensing, 36, 1282– 1289.
Weng, Q., Liu, H., & Lu, D. (2007). Assessing the effects of LULC patterns on thermal conditions
using landscape metrics in city of Indianapolis. United States. Urban Ecosystems, 10, 203–219.
Weng, Q., Lu, D., & Schubring, J. (2004). Estimation of land-surface temperature-vegetation
abundance relationship for urban heat island studies. Remote Sensing of Environment, 89,
467– 483.
South African Geographical Journal 17

Wong, N. H., & Yu, C. (2005). Study of green areas and urban heat island in a tropical city. Habitat
International, 29, 547– 558.
Wilby, R. L. (2003). Past and projected trends in London’s urban heat island. Weather, 58, 251–260.
Yang, H., & Liu, Y. (n.d.). A satellite remote sensing based assessment of urban heat island in Lanzhou
city, northwest China. Retrieved May 17, 2012, from http://www.isprs.org/proceedings/XXXVI/
8-W27/yang_lui.pdf
Zhang, J., Wang, Y., & Li, Y. (2006). A Cþþ program for retrieving land surface temperature from
the data of Landsat TM/ETM Band 6. Computers Geosciences, 32, 1796– 1805.
Zhang, Z., Ji, M., Shu, J., Deng, Z., & Wu, Y. (2008). Surface urban heat island in Shanghai, China:
Examining the relationship between land surface temperature and impervious surface fractions
derived from Landsat ETM+ imagery. The International Archives of the Photogrammetry,
Remote Sensing and Spatial Information Sciences. Part B8 (Vol. XXXVII, pp. 601– 606).
Beijing.
Zhihao, Q., Li, W., Gao, M., & Zhang, H. (2006). An algorithm to retrieve land surface temperature
from ASTER thermal band data for agricultural drought monitoring. In M. Owe, G. D’Urso,
C. M. U. Neale, & B. T. Gouweleeuw (Eds.), Remote sensing for agriculture, ecosystems, and
hydrology VIII. Proceedings of SPIE (Vol. 6359, 63591F). The International Society for Optical
Engineering.
Downloaded by [Adeline Ngie] at 06:45 20 June 2014

View publication stats

You might also like