You are on page 1of 10

Landscape and Urban Planning 135 (2015) 1–10

Contents lists available at ScienceDirect

Landscape and Urban Planning


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

Research paper

Impacts of urban biophysical composition on land surface


temperature in urban heat island clusters
Guanhua Guo a,b,c , Zhifeng Wu d,∗ , Rongbo Xiao e , Yingbiao Chen d ,
Xiaonan Liu a,b,c , Xiaoshi Zhang d
a
Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, Guangzhou 510640, China
b
University of Chinese Academy of Sciences, Beijing 100049, China
c
Guangdong Key Laboratory of Agricultural Environment Pollution Integrated Control, Guangdong Institute of Eco-environment and Soil Sciences,
Guangzhou 510650, China
d
School of Geographical Sciences, Guangzhou University, Guangzhou 510006, China
e
Guangdong Provincial Academy of Environment Science, Guangzhou 510045, China

h i g h l i g h t s

• We extracted urban heat island (UHI) clusters using an innovative approach.


• We examined the nonlinear relationships between the LST and urban biophysical composition in UHI clusters.
• The spatial heterogeneity of the landscape within UHI results in the complexity of LST.
• Both NDVI and NDBI are great indicators for predicting the variations of LST in UHI clusters.

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

Article history: The spatio-temporal pattern of biophysical composition significantly affects land surface temperature
Received 11 February 2014 (LST). Previous studies, however, mostly characterized urban heat island (UHI) clusters being spatially
Received in revised form 28 October 2014 homogeneous. The landscape spatial heterogeneity in urban across UHI clusters challenges us to more
Accepted 11 November 2014
accurately characterize the relationships between LST and corresponding urban biophysical composition.
In this study, we introduced an innovative integrated approach that combined object-oriented image
Keywords:
segmentation with local indicators of spatial autocorrelations (LISA) to extract UHI clusters from an LST
UHI clusters
image. We used a regression tree model to examine the nonlinear relationships between LST and each
Object-oriented segmentation
Nonlinear relationship
of three satellite-based indices within the UHI clusters: normalized differential vegetation index (NDVI),
Urban biophysical composition normalized differential build-up index (NDBI), and normalized difference bareness index (NDBaI). We
Land surface temperature found that both NDVI and NDBI are strongly correlated with the variations of LST whereas NDBaI has a
weaker correlation with LST. We also found that the regression tree model built in this study enabled us
to effectively detect the nonlinear relationship between LST and biophysical composition. Furthermore,
based on a set of rules derived from a regression tree analysis, we found that urban landscapes strongly
affect LST and its spatial heterogeneity within a UHI. These rules were used to detect the nonlinear impacts
of complex urban biophysical composition on LST. The results of this study provided insights into how
LST within UHI varies with urban surface characteristics at fine spatial scale and also a new method for
investigating effects of land surface composition on LST in urbanized areas.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction

Urbanization is considered to be one of the most evident aspects


of human impacts on the earth in terms of greenhouse gas emis-
sions and land use changes (e.g., road and building construction),
∗ Corresponding author. Tel.: +86 13922448970.
both of which can in turn affect daily mean surface temperature
E-mail addresses: gghgiser@qq.com (G. Guo), gzuwzf@163.com (Z. Wu),
ecoxiaorb@163.com (R. Xiao), 1064553130@qq.com (Y. Chen), 42447481@qq.com (Gallo & Owen, 1999; Owen, 1998). It is well documented that
(X. Liu), 285265827@qq.com (X. Zhang). urban centers experience higher temperature than surrounding

http://dx.doi.org/10.1016/j.landurbplan.2014.11.007
0169-2046/© 2014 Elsevier B.V. All rights reserved.
2 G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10

suburban/rural areas, and this phenomenon is known as the urban variations through space and time. Given the spatial heterogeneity
heat island (UHI) (Voogt & Oke, 2003). Traditionally, UHI was often and nonlinear interactions among urban ecological systems (Wu &
identified using in-situ weather station networks or mobile mea- David, 2002) and their significance for urban landscape manage-
surements of the urban canopy layer (Li et al., 2011). With the ment and planning, it is essential to improve our understanding of
development of remote sensing (RS) technology, RS-based data the relationship between urban thermal behavior and its surface
– such as surface radiative temperatures derived from remote biophysical composition at fine spatial scales.
imagery thermal bands – are increasingly used to examine UHI. Therefore, the objectives of this study are to (i) introduce an
RS-based technology enables us to thoroughly characterize LST dis- integrated and object-oriented segmentation approach for extract-
tribution in urban areas and thus contribute us to explicit revealing ing the UHI clusters from LST image, (ii) quantify the relationship
spatial patterns and variations of UHI although LST only provides a between LST and each of three urban biophysical compositions,
snapshot of temperature during the day (Li et al., 2011). As a result, and (iii) explore the spatial heterogeneity of LST in UHI clusters. To
RS-based LST have been widely used to examine the relationship achieve these ends, we selected the center of Guangzhou in China,
between UHI and surface biophysical parameters (e.g., Weng, 2009; an area characterized by strong spatio-temporal variability in UHI,
Arnfield, 2003). as the study region. Specifically, we focused on UHI clusters in the
Previous studies that focused on the relationship between urban Central Urban Districts of Guangzhou City to investigate impacts
thermal behavior and landscape patterns indicated that LST is of three selected urban biophysical compositions (i.e., NDVI, NDBI,
strongly subject to both the composition and configuration of land and NDBaI) on LST in UHI clusters. The results from this study
cover/use (Buyantuyev & Wu, 2010; Li et al., 2011; Weng, Liu, & Lu, will improve our understanding of how LST varies across urban
2007; Zhou, Huang, & Cadenasso, 2011). RS-based urban surface landscapes at fine spatial scale characterized by different surface
biophysical composition – such as vegetation abundance, impervi- conditions.
ous surface and soil fractions – were found to be the best indicators
of LST variations across space (Deng & Wu, 2013; Xiao et al., 2007; 2. The study site
Zhang, Odeh, & Han, 2009). The relationship between biophysi-
cal composition and LST also varies with seasons and is thought Guangzhou is the capital of Guangdong province of China. It is
to be nonlinear (Chen, Zhao, Li, & Yin, 2006; Owen, Carlson, & situated on the broad flat alluvial plain of the Pearl River Delta of
Gillies, 1998). Based on landscape metrics derived from different southern China, and has a total area of approximately 7434.40 km2
LST zones, Liu and Weng (2008) found that the seasonal variations (Fig. 1). The climate condition in Guangzhou is characterized by
in landscape pattern was correlated with LST. Buyantuyev and Wu the typical subtropical monsoon maritime climate of Southern Asia,
(2010) linked LST variations with changes in land covers as well with annual mean temperature about 23.2 ◦ C and maximum tem-
as socioeconomic features (i.e., family income, population density perature about 38.7 ◦ C (based on long-term climate records from
and age). Nevertheless, the results from these studies were based on the Wushan weather station that is located at 23.17◦ N and 113.33◦ E
urban-wide scale and ignored the spatial heterogeneity of compo- and operated by Guangzhou Meteorological Bureau). Guangzhou
sitions and configurations of urban landscapes at fine spatial scales. has been one of the fastest urbanizing cities in China since the
The neglect of spatial heterogeneity of urban landscapes across 1980s and its GDP reached 1.35 trillion RMB in 2012, a 10.5%
space is likely to disable these studies to accurately predict LST increase compared to that in 2011. The population in the core

Fig. 1. Geographic location of the Guangzhou Central Urban Districts.


G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10 3

2001a; Qin, Dall’Olmo, Karnieli, & Berliner, 2001b). The derived


LST was treated as an indicator of urban heat condition. Observed
surface air temperatures from the Wushan weather station dur-
ing the period when the TM image was acquired were collected to
determine the LST (Fig. 5A).

3.2. Object-oriented segmentation of the LST

Previous studies have attempted to carry out segmentation


process by simple density slicing based on pixel-by-pixel image
analysis, and assumed that one pixel is not related to its nearest
neighbors. Blaschke and Hay (2001) and Blaschke (2003,2010) rec-
ommended a method of “beyond pixels” during image analysis.
To analyze the spatial structure of LST, We used the object-
oriented image segmentation approach to build LST objects for
the study area. The segmentation processing of LST images
Fig. 2. The variations in monthly mean air temperature and humidity in 2005. Cli- was accomplished using software known as eCognition Devel-
mate data shown here were collected from Wushan weather station in Guangzhou.
oper (http://www.ecognition.com/). Fig. 3 shows a comparison
Error bars indicate the standard deviation.
of density-slicing and object-oriented-segmentation-derived LST
images. It is clear that the results generated with a density slicing
urban area of Guangzhou has increased from 4.09 million in 2000 method will result in a so-called ‘salt and pepper effect’ phe-
to 7.72 million in 2010. Guangzhou and its Central Urban Districts, nomenon (Fig. 3A) because this approach classifies LST levels based
therefore, have become the engine of urbanization in southern on pixel-by-pixel values. In contrast, the LST based on objected-
China’s economy. The Central Urban Districts of Guangzhou con- oriented segmentation is spatially more homogeneous compared
sists of 126 census tracts that cover approximately 1461.66 km2 and that based on density-slicing. (Fig. 3B). The LST objects with geo-
encompasses diverse landscapes such as forested hills, farmland, graphic boundaries are relatively homogeneous and composed of
under-developing built-up areas, urban fringes, and high-density distinct groups of LST pixels. They can be recognized and trans-
built-up areas. The diversity and spatial heterogeneity of land sur- ferred into meaningful objects for UHI management and mitigation
face conditions make it an ideal study area for UHI research. as well as for urban design and planning.
The modification of the segmented scale parameter will influ-
3. Methodology ence the consequent results in a way that higher values lead to
larger LST objects whereas smaller values yield smaller LST objects.
3.1. Data and LST retrieval It is, however, often unable to determine the absolute correct scale
of segmentation in advance because different LST image charac-
A cloud-free Landsat TM image, acquired at approximately teristics (i.e., ASTER and Landsat ETM+ thermal bands with various
10:43 am (Beijing time) on July 18, 2005 was utilized for this study. spatial resolutions) require different segmentation scales. In many
This daytime image was selected because UHI are most remark- cases, UHI processes emerge at different segmentation scales on
able during the daytime of warm season when air temperatures the same image due to the hierarchical network of image objects
and humidity are both higher in this study area (Fig. 2). In addition, and the urban landscape’s spatial structure (Burnett & Blaschke,
UHI is generally stronger and spatially more variations during the 2003; Wu, 1999; Wu & Loucks, 1995). A visual comparison was
daytime (Arnfield, 2003; Nichol, Fung, Lam, & Wong, 2009; Roth, used to determine the appropriate scale parameter, and the results
Oke, & Emery, 1989). are described in Section 4.1.
The Landsat TM image used in this study has been pre-processed
using radiometric correction and geometric rectification, and thus 3.3. UHI clusters extraction
is of good quality. However, given that the spatial extent of the
study area is much smaller than that of the TM image, we further As a measurement of spatial associations, the Local Indicators of
processed the TM image as follows. First, the raw TM image was Spatial Autocorrelations (LISA) statistics allow us to detect clusters
masked with a boundary shapefile of the Central Urban Districts, on the assumption that a spatial pattern is a non-random distribu-
and then rectified with a universal transverse mercator (UTM) pro- tion of geospatial parameters (Anselin, 1995; Anselin, Syabri, & Kho,
jection using a WGS84 datum and zone 49 UTM. Second, IKONOS 2006). The LST objects that are located close to one another are more
fusion imagery with a 1-m spatial resolution acquired in 2005 were likely to have similar characteristics and geographic homogene-
used to georeference the TM image based on 60 ground control ity, compared to those located farther from each other. Therefore,
points. A nearest neighbor algorithm was used to resample the the application of LISA can be extended to UHI studies in urban
rectified and georeferenced image of 120 m spatial resolution onto areas. This study adopted a local Moran’s I to determine the location
30 m by 30 m grid points and this approach was applied to all bands clusters of UHI after the segmentation of LST image, as mentioned
of the TM image. The re-gridding of the TM image aims to match the in Section 3.2. These potential UHI clusters allow for the evalua-
spatial resolutions of RS indices derived from different bands. Third, tion of the relationship between LST and related urban biophysical
we calibrated the TM image through converting the digital num- composition at fine spatial scale.
bers of (DNs) of all bands to at-satellite reflectance. The resultant The local Moran’s I is calculated as follows (Anselin, 1995;
root mean squared errors were smaller than 0.5 pixel (15 m). Anselin et al., 2006):
We calculated NDVI, NDBI and NDBaI for the study area based on (xi − x)  zi 
the TM image. These variables have proven useful in establishing Ii = n ×  2
× wij (xi − x) =
mo
× wij zj , with :
i (xi
− x)
meaningful statistical associations with the UHI (Bechtel, Zakšek, & j j
Hoshyaripour, 2012; Chen et al., 2006; Li et al., 2011; Weng, Lu, & Z
i
Liang, 2006). A thermal band from Landsat TM was used to retrieve m0 = ,
n
the LST using a mono-windows algorithm (Qin, Karnieli, & Berliner, i
4 G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10

Fig. 3. Comparison of (A) density slicing and (B) object-oriented segmentation for a LST image.

where xi is the observation x, and x is the average across the units; n the regression tree model, the spatial distribution of rules in the
is the total number of observation x; The spatial weights matrix Wij study area was also mapped. Therefore, a corresponding landscape
was standardized by row to avoid scale dependence and to facilitate of these rules was identified from the higher spatial resolution of
the interpretation of the statistics (Anselin, 1995), and defined as IKONOS imagery, as well as from field surveys.
a local neighborhood with a threshold distance of 919 m around
each geographic unit and 9999 permutation tests by setting the 4. Results
significance level as 0.01; Zi and ZZ are deviations from the mean. mo
is the average of total Zi deviations. The UHI clusters were obtained 4.1. Determination of the scale parameter for LST image
using GeoDa software environment (Anselin et al., 2006) and shown segmentation
in Fig. 5B.
Four types of spatial patterns were defined on the basis of pre- According to the characteristics of the LST image, five layers of
specified significances: high–high (H–H), low–low (L–L), low–high LST objects were generated using the scale parameters of 2, 4, 6,
(L–H), and high–low (H–L), a combination of the Local Moran’s I 8 and 10. The resulting segment outputs were visually analyzed
(Anselin, 1995; Anselin et al., 2006) statistics calculated for the clus- by overlaying them with the high spatial resolution IKONOS image
ter map. The specific meanings of the four patterns defined in this (Fig. 4A–C list scale parameters of 6, 8 and 10). As shown in Fig. 4,
study are described as follows: the scale parameter of 6 results in smaller LST objects. For exam-
H–H (hot spots): the LST object and its surroundings have a tem- ple, LiWan Lake Park was over-segmented into several LST objects
perature concentration level greatly higher than the average level, (Fig. 4A, the arrow). When the scale parameter increased to 10, the
suggesting UHI clusters from a central place. boundary of built-up areas tended to be composed of various build-
L–L (cold spots): the temperature levels of both the LST object ing densities with greater LST heterogeneity although LiWan Lake
and its surroundings are relatively lower than the average level of Park keeps its integrity (Fig. 4C, the arrow). These results indicate
all objects. that the scale parameter of eight is the best option (Fig. 4B) because
H–L (island): the LST object exhibits an obviously high con- it resulted in meaningful LST objects with ‘natural’ geographic
centration whereas the concentration of its neighboring objects is boundaries and avoided the production of scattered classification
lower than average. resulting from over-segmentation. The total number of LST objects
L–H (Atoll): the LST object has a significantly low-level concen- in the study area amounted to 6776, with a mean temperature of
tration, while the concentration levels of its surrounding objects approximately 312.52 K and standard deviation of 2.56 K.
are relatively high.
Furthermore, “non-significance” is also a type of pattern, indi-
4.2. Correlation analysis of the LST with the NDVI, NDBI and
cating no significant local spatial associations.
NDBaI

3.4. Statistical methods


The Pearson’s correlation coefficients indicate that all of the
biophysical variables were significantly correlated with the LST
A zonal analysis method with 250 × 250 m blocks was used to
(Table 1). The NDBI and NDVI had stronger relationship with the
evaluate the mean of LST and other three biophysical compositions.
Pearson’s correlation coefficients were first conducted to exam-
ine the strength of relationship between LST and each of the other Table 1
variables within UHI clusters. We further evaluated the mean LST Pearson’s correlation coefficients between the LST and the NDVI, NDBI and NDBaI.

at 0.01 increments from the minimum and maximum NDVI, NDBI LST NDVI NDBI NDBaI
and NDBaI, and plotted the relationships between them. Regression
LST 1.00
tree was used to develop a model to investigate the nonlinear rela- NDVI −0.67** 1.00
tionship between the LST and each of the biophysical composition. NDBI 0.66** 0.69** 1.00
Furthermore, a traditional multiple regression method was also NDBaI −0.11* 0.36** 0.29** 1.00
developed to exhibit the superiority of regression tree analysis dur- *
Correlation is significant at the 0.05 level (two-tailed).
**
ing LST prediction modeling. Based on various rules developed from Correlation is significant at the 0.01 level (two-tailed).
G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10 5

Fig. 4. The segmentation results of an LST image at various scales (from panel A to C, the scale was 6, 8, and 10, respectively).

LST than the NDBaI. The NDVI was negatively correlated with the to the scattergram of NDVI and NDBI, the shape of NDBaI’s scat-
LST, whereas the NDBI had a positive effect on the LST. The relation- tergram formed like a circle (Fig. 6C). This pattern may indicate a
ship between the LST and NDVI we observed in this study agreed random relationship between LST and NDBaI.
well with those reported by a number of previous studies (Gillies, Although there is a negative relationship between NDVI and LST
Kustas, & Humes, 1997; Weng, Lu, & Schubring, 2004; Yue, Xu, and a positive relationship between NDBI and LST, the relation-
Tan, & Xu, 2007). The NDBaI that represents soil distribution in ship is subject to NDVI and NDBI levels, respectively. For example,
urban areas, also tended to have a negative relationship with LST Figs. 6D and E indicate that the linear relationships between the
as reported by Chen et al. (2006). The unexpected negative rela- LST and NDVI/NDBI are strong but the relationships are not mono-
tionship between LST and NDBaI may result from the complexity tonic across all NDVI values. For NDVI, the strong negative linear
of landscape in urban. relationship only exists when the NDVI values are approximately
Previous studies have documented a “temperature edge” phe- greater than −0.06 (Fig. 6D), suggesting that the LST has a linear
nomenon when investigating the relationship between the LST and relationship with NDVI when the urban surface is covered by veg-
NDVI (Carlson, Capehart, & Gillies, 1995; Carlson & Traci Arthur, etation. This finding was similar to that reported by Li et al. (2011)
2000; Gillies et al., 1997; Li et al., 2011). In this study, we found and Yuan and Bauer (2007). Interestingly, we found that the lin-
that the scattergram of the LST and NDBI also clearly displayed ear relationship between the LST and NDBI no longer exist when
both a “warm edge” and “cool edge”, locating at the upper edge the NDBI values are greater than 0.2 or less than −0.5 (Fig. 6E).
and lower edge of the scattergram. This phenomenon also applies A similar phenomena also applies to NDBaI (Fig. 6F), which is not
to NDVI (Fig. 6A and B). The pixels with the same NDBI values are strongly correlated with the mean LST when its values are greater
warmer and drier near the warm edge compared to those near the than −0.2 or less than −0.45. This suggests that the relationship
cool edge. Therefore, the potential impacts of the NDBI on the LST between NDVI and LST or between NDBI and LST is not always linear
needed further emphasizing during future UHI studies. Compared or can be nonlinear.

Fig. 5. The LST distribution (A) and spatial evolution of UHI clusters (B).
6 G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10

Fig. 6. The LST (A–C) scattergrams and linear plots of mean LSTs (D–F) versus the NDVI, NDBI and NDBaI within the UHI clusters.

4.3. LST nonlinear modeling of UHI clusters the regression tree predicted LST variations better than the mul-
tiple linear regression as suggested by the higher R2 adj (0.61), the
Regression tree analysis was used to construct the LST nonlinear lower root mean square error (RMSE), and the higher correlation
model based on NDVI, NDBI and NDBaI. Table 2 shows the compari- coefficients (0.81, significant at the 0.01 level).
son of the LST predictions based on the regression tree analysis and The calculated R2 adj based on the regression tree analysis sug-
the traditional multiple linear regression method. It indicates that gested that approximately 61% of variations in the LST can be
explained jointly by the three biophysical compositions (Table 2).
Table 2 In addition, the regression tree analysis enabled us to differentiate
A comparison of the result from a regression tree analysis and a linear regression. the impacts of each of the three biophysical compositions on the
LST. The nonlinear associations between LST and each of the three
RTa LRb
biophysical compositions can be grouped into four categories as
R2 adj 0.61 0.54 follows:
RMSE 1.05 1.12
Correlation 0.81 0.73
a,b
Note that the RT and LR represent a regression tree and linear regression, respec- Rule 1: [108 cases, mean 316.275 K, range 310.481 K to 320.813 K, est
tively. err 1.3188 K]
G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10 7

Fig. 7. The predicted LST data based on different rules against the actual satellite-derived LST. The vertical black line represents a 1:1 relationship between the predicted and
actual LST values. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Rule: NDVI > −0.0592 and NDBI > 0.0216 YueXiu districts, and the western part of the Haizhu district, and
Expression: LST = 313.8386 − 11.3NDBaI + 2.6NDBI − 0.6NDVI the eastern part of the Fangcun district along the Pearl River. Rule 4
Rule 2: [3885 cases, mean 314.8267 K, range 309.18 K to 320.36 K, est accounted for a smaller area compared to Rules 2 and 3, and it was
err 0.8025 K] distributed mainly in the Fangcun and Huangpu districts near the
Rule: NDBI ≤ 0.2106 and NDBaI > −0.3422 Pearl River.
Expression: LST = 310.1869 + 21.7NDBI − 7.4NDBaI − 4.3NDVI The contribution of built-up areas to the LST within UHI clus-
Rule 3: [2136 cases, mean 315.1658 K, range 309.907 K to 320.185 K, ters strongly depends on its morphology and vegetation conditions;
est err 0.7864 K] the former influences the aerodynamics, and the latter affects the
Rule: NDVI > −0.0592 and −0.4258 < NDBaI ≤ −0.3422 transpiration (Li et al., 2011; Nichol, 1996, 1998). The landscape
Expression: LST = 311.832 + 21.2NDBI − 3.7NDVI − 1.8NDBaI characteristics in Rule 2 (with lower LST values; average tempera-
Rule 4: [192 cases, mean 315.3750 K, range 310.555 K to 319.929 K, ture of 314.8267 K) mainly consist of newly developed commercial
est err 1.2039 K] buildings and institutions with relatively lower building density
Rule: NDBI ≤ 0.2106 and NDBaI ≤ −0.4258 and inner courtyards composed of grass or trees (Fig. 8B). Wind
Expression: LST = 291.8386 + 50.2NDBI − 42.9NDBaI + 16.2NDVI speeds in these open spaces around tall buildings in the lower
boundary layer increases and thus cools the surrounding environ-
The relationships between the satellite-derived LST and the esti- ment. In a densely-built low-rise environment (i.e., urban villages,
mated LST based on above four rules were examined (Fig. 7). The Rule 3), although the closeness of low residential blocks can pre-
results indicate that the variation in satellite-derived LST is strongly clude solar radiation from heating the ground or streets directly,
correlated with each of the estimated LST based on the four rules, the roofing on these residential blocks is warmer due to the pri-
respectively, as suggested by the corresponding coefficients that marily dark concrete material that it is made of (Fig. 8A). As a
range from 0.31 to 0.66. Among these four rules, the predictive result, a remote sensor can only “see” this part rather than the
model based on Rule 2 was the best with an R2 adj of 0.66 (Fig. 7 narrow alleyways (Nichol, 1996). Furthermore, urban villages in
A). This model can predict 61.47% of the total cases with a variance Guangzhou’s core areas tend to be surrounded by high buildings
of 0.8025 K. The number of cases in Rule 3 accounts for the second and are formed like a “basin” (Fig. 9), which could also lower the
proportion of the UHI clusters area size, with an R2 adj of 0.56 and a wind speed and weaken the environmental cooling effect, espe-
variance of only 0.7864 K. It suggests a high level of homogeneity cially at night. Compared to Rules 2 and 3, the landscape features
existing within Rule 3 (Fig. 7B). Both Rules 1 and 4 can only correctly in Rule 4 are largely factories or industrial-residential complexes,
predict relatively smaller cases. Rule 1 shows higher temperature encompassing very low buildings. For instance, warehouses and
than the average value. However, the formula of Rule 1 has a weaker heavy industry buildings have very dark iron tiles and asphalt
capability to capture variations in the LST (R2 adj is 0.31). Although rooftops, and little vegetation along the streets (Fig. 8C).
Rule 4 produces relatively stronger heterogeneity when compared
to the other three rules with a variance of 1.2039 K, it only explains 5. Discussions
approximately 55% of the variations in the LST within UHI clusters
(Fig. 7B). 5.1. Implications of UHI clusters
To further determine if different rules represent different types
of urban surface characteristics, we mapped the spatial distribution Compared to previous UHI-related studies that selected a whole
of rules using the same colors as in Fig. 7, and identified the specific city or region as the study area (Buyantuyev & Wu, 2010; Cao,
landscape structure characteristics of each rule using high spatial Onishi, Chen, & Imura, 2010; Nichol, 1996; Zhou et al., 2011), we
resolution IKONOS imagery (Fig. 8). We found that the spatial pat- specifically focused on UHI clusters and investigated the impacts
tern of Rule 2 (yellow) tends to concentrate on the outlying areas of urban biophysical composition on the magnitude of LST at rel-
of the urban villages. For instance, the Central Business District atively fine spatial scale. We found that the combination of the
(CBD) of Zhujiang New Town in the Tianhe district was associated integrated method of object-oriented segmentation and LISA sig-
with relatively abundant vegetation and open space. In contrast, nificantly improved the efficiency in differentiating the LST levels
the distribution of Rule 3 was clearly similar to the old downtown and keeping the integrity of LST objects (Fig. 2). The UHI clusters
of Guangzhou’s core area, such as the whole areas of the LiWan and we derived for the Guangzhou city included a group of meaningful
8 G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10

Fig. 8. The spatial distribution of various rules developed by a regression tree analysis and corresponding urban surface landscape characteristics identified by IKONOS
imagery. Panels A to C represent urban villages, CBD and heavy industry, respectively.

real objects with generally higher LST temperature values dur- and natural resource managers. It also is necessary for spatial plan-
ing daytime of warm season. The ‘natural’ UHI boundaries were ning of the urban environment as well as for many monitoring and
an antithesis to the view that the ecological variables in urban simulation UHI study programmers.
areas are continuum of mosaics. It provides important insights into The UHI is a typical phenomenon in urbanized areas and
urban ecology design and management. Therefore, our studies of can occur across various spatial scales or along a scale contin-
UHI clusters may have important implications for urban planners uum similar to visible landscape mosaics. Many studies suggested

Fig. 9. The morphology of typical urban villages in Guangzhou. They tend to be surrounded by high buildings and formed like a “basin”.
G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10 9

that object-oriented segmentation could build a critical bridge 2007; Zhang et al., 2009; Zhou, Qian, Li, Li, & Han, 2014). The ecolog-
between the spatial concepts applied to multi-scale landscape ical context and different climatic conditions also can significantly
analyses and Geographic Information Science (Blaschke, 2010; influence the amplitude of summer daytime UHI (Imhoff, Zhang,
Hay, Marceau, Dube, & Bouchard, 2001; Wu, 1999). Burnett and Wolfe, & Bounoua, 2010). Therefore, further studies that address
Blaschke (2003) considered objected-oriented segmentation as “A the temporal and spatial variation of the relationships between UHI
multi-scale segmentation/object relationship modeling methodol- and the corresponding land cover features are necessary.
ogy for landscape analysis.” This helps explain why object-oriented
segmentation has recently been widely used for explaining the 6. Conclusions
complexity and heterogeneity of multi-scale and hierarchical UHI.
The LST plays an essential role in affecting the exchange and
5.2. Relationship between urban compositions and LST interaction of energy fluxes between the land surface and atmo-
sphere. Understanding the linkage between LST and urban surface
The LST-NDVI relationship has been a heated research topic characteristics is important for designing effective measures to mit-
because of the physical significance of vegetation to the LST igate the amplitude of UHI. In this study, an innovative approach
(Carlson et al., 1995; Gutman & Ignatov, 1998). Although there is a that combined object-oriented segmentation and LISA statistics
strong negative relationship between NDVI and LST with a “temper- was used to extract UHI clusters. Then, we quantified the nonlin-
ature edge” (Table 1 and Fig. 6A), LST is positively related to NDVI ear relationships between LST and each of NDVI, NDBI and NDBaI
at regional scales (Fig. 6D). A similar finding has been reported by Li in UHI clusters, respectively. We found that both NDVI and NDBaI
et al. (2011), who studied the whole city of Shanghai. Interestingly, are significantly correlated with the LST: the former mitigates the
we found the LST - NDBI relationship also has a “temperature edge” UHI whiles the latter strengthens it. As a result, both NDVI and
with an association almost as strong as that between LST and NDVI NDBaI are good indicators of UHI. In addition, we found that the
(Table 1 and Fig. 6B). Considering the seasonal variability of NDVI, NDBI also has a “temperature edge” relationship to the LST, as does
NDBI can provide a complementary metric for LST studies. the NDVI. Comparatively, the NDBaI has a weak negative relation-
The nonlinearity and seasonal variations in the urban LST- ship to the LST in UHI clusters. Compared to traditional multiple
characteristic (i.e., LST–NDVI, LST–NDBI) relationship have been linear regressions, the tree regression models resulted in advanced
documented (Li et al., 2011; Weng et al., 2004; Yuan & Bauer, improvement for exploring nonlinear relationships between LST
2007; Zhang et al., 2009). This challenges us to predict LST varia- and landscape characteristics. Furthermore, a set of rules derived
tions, especially in spatially heterogeneous landscapes as our study from regression tree analyses demonstrated the spatial hetero-
area. The use of regression trees to predict LST has proven to be geneity of landscapes within UHI clusters. These rules were linked
a greater improvement with an overall accuracy of 61%, higher to special landscape characteristics, such as urban villages and CBD.
than 54% from multiple regressions (Table 2). Furthermore, differ- Based on the type of landscape characteristics in urban areas, we
ent rules obtained from regression trees enable us to interpret the build a particular model that is able to predict the LST variations
relationship between LST and landscape characteristics in a more efficiently and can be applied to other urban planning and UHI
meaningful way. For example, higher temperatures in urban areas mitigation efforts.
are usually associated with large amounts of built-up areas with
higher NDBIs or impervious surface area, as well as little vegeta-
Acknowledgements
tion (Chen et al., 2006; Xiao et al., 2007). Oke (1976) reported that
the density of UHI has a closer relationship than a city’s size (as
This study was supported by the National Natural Science Foun-
measured by its population) to surface structure characteristics. In
dation of China (41171446, 41201543 and 31170445), National
this study, we found that landscape characteristics with various
Science and Technology support project plan (2012BAC13B01). The
rules varies significantly, indicating that the LST and landscapes in
authors would like to thank the editor and the anonymous review-
UHI clusters are of high spatial heterogeneity. Different rules tende
ers for their helpful comments and suggestions. Especially, we are
to link to a particular landscape structure in urban areas (Fig. 8).
grateful to Dr. Guoping Tang for his comments on the early version
This further suggests that UHI mitigation and management poli-
of this manuscript.
cies should consider explicitly the particular landscape structure
(i.e., CBD or urban villages).
References
5.3. Limitations
Anselin, L. (1995). Local indicators of spatial association—LISA. Geographical Analysis,
27(2), 93–115.
The results from this study are likely to be subject to several limi- Anselin, L., Syabri, I., & Kho, Y. (2006). GeoDa: An introduction to spatial data analysis.
tations. First, the scale of segmentation is one of the most significant Geographical Analysis, 38(1), 5–22.
Arnfield, A. J. (2003). Two decades of urban climate research: A review of turbulence,
parameters for segmentation processing. In this study, the scale of exchanges of energy and water, and the urban heat island. International Journal
segmentation was empirically determined based on visual analysis of Climatology, 23(1), 1–26.
and comparison. Therefore, study that can optimize the selection of Bechtel, B., Zakšek, K., & Hoshyaripour, G. (2012). Downscaling land surface tem-
perature in an urban area: A case study for Hamburg, Germany. Remote Sensing,
the segmented scale for LST image is necessary for future studies. 4(10), 3184–3200.
Second, the timing of Landsat TM (morning) is not ideal for study- Blaschke, T. (2003). Object-based contextual image classification built on image seg-
ing the UHI because UHI often peaks later during a day. However, mentation. In Advances in techniques for analysis of remotely sensed data 2003 IEEE
workshop on (pp. 113–119). IEEE.
since the limitations of remote sensing technology, the paucity of Blaschke, T. (2010). Object based image analysis for remote sensing. ISPRS Journal of
thermal sensors that can produce thermal images with high spatial Photogrammetry and Remote Sensing, 65(1), 2–16.
resolution hinders us to detect LST variations at fine spatial scales Blaschke, T., & Hay, G. J. (2001). Object-oriented image analysis and scale-space: The-
ory and methods for modeling and evaluating multiscale landscape structure.
among different time of a day in urban (Chun & Guldmann, 2014).
International Archives of Photogrammetry and Remote Sensing, 34(4), 22–29.
Third, because only one set of horizontal daytime LST data were Burnett, C., & Blaschke, T. (2003). A multi-scale segmentation/object relationship
used for this study, our study thus ignored the variation of LST in modelling methodology for landscape analysis. Ecological Modelling, 168(3),
the vertical direction within a UHI. Previous studies have demon- 233–249.
Buyantuyev, A., & Wu, J. (2010). Urban heat islands and landscape heterogeneity:
strated that diurnal and seasonal variations exist between LST and Linking spatiotemporal variations in surface temperatures to land-cover and
land surface characteristics (Buyantuyev & Wu, 2010; Yuan & Bauer, socioeconomic patterns. Landscape Ecology, 25(1), 17–33.
10 G. Guo et al. / Landscape and Urban Planning 135 (2015) 1–10

Cao, X., Onishi, A., Chen, J., & Imura, H. (2010). Quantifying the cool island intensity Qin, Z.-h., Karnieli, A., & Berliner, P. (2001). A mono-window algorithm for retriev-
of urban parks using ASTER and IKONOS data. Landscape and Urban Planning, ing land surface temperature from Landsat TM data and its application to
96(4), 224–231. the Israel–Egypt border region. International Journal of Remote Sensing, 22(18),
Carlson, T. N., Capehart, W. J., & Gillies, R. R. (1995). A new look at the simpli- 3719–3746.
fied method for remote sensing of daily evapotranspiration. Remote Sensing of Qin, Z., Dall’Olmo, G., Karnieli, A., & Berliner, P. (2001). Derivation of split window
Environment, 54(2), 161–167. algorithm and its sensitivity analysis for retrieving land surface temperature
Carlson, T. N., & Traci Arthur, S. (2000). The impact of land use—Land cover changes from NOAA-advanced very high resolution radiometer data. Journal of Geophys-
due to urbanization on surface microclimate and hydrology: A satellite perspec- ical Research: Atmospheres (1984–2012), 106(D19), 22655–22670.
tive. Global and Planetary Change, 25(1), 49–65. Roth, M., Oke, T., & Emery, W. (1989). Satellite-derived urban heat islands from three
Chen, X.-L., Zhao, H.-M., Li, P.-X., & Yin, Z.-Y. (2006). Remote sensing image-based coastal cities and the utilization of such data in urban climatology. International
analysis of the relationship between urban heat island and land use/cover Journal of Remote Sensing, 10(11), 1699–1720.
changes. Remote Sensing of Environment, 104(2), 133–146. Voogt, J. A., & Oke, T. R. (2003). Thermal remote sensing of urban climates. Remote
Chun, B., & Guldmann, J.-M. (2014). Spatial statistical analysis and simulation of the Sensing of Environment, 86(3), 370–384.
urban heat island in high-density central cities. Landscape and Urban Planning, Weng, Q. (2009). Thermal infrared remote sensing for urban climate and
125, 76–88. environmental studies: Methods, applications, and trends. ISPRS Journal of Pho-
Deng, C., & Wu, C. (2013). Examining the impacts of urban biophysical compositions togrammetry and Remote Sensing, 64(4), 335–344.
on surface urban heat island: A spectral unmixing and thermal mixing approach. Weng, Q., Liu, H., & Lu, D. (2007). Assessing the effects of land use and land cover
Remote Sensing of Environment, 131, 262–274. patterns on thermal conditions using landscape metrics in city of Indianapolis,
Gallo, K. P., & Owen, T. W. (1999). Satellite-based adjustments for the urban heat United States. Urban Ecosystems, 10(2), 203–219.
island temperature bias. Journal of Applied Meteorology, 38(6), 806–813. Weng, Q., Lu, D., & Liang, B. (2006). Urban surface biophysical descriptors and land
Gillies, R., Kustas, W., & Humes, K. (1997). A verification of the ‘triangle’ method surface temperature variations. Photogrammetric Engineering & Remote Sensing,
for obtaining surface soil water content and energy fluxes from remote mea- 72(11), 1275–1286.
surements of the Normalized Difference Vegetation Index (NDVI) and surface e. Weng, Q., Lu, D., & Schubring, J. (2004). Estimation of land surface
International Journal of Remote Sensing, 18(15), 3145–3166. temperature–vegetation abundance relationship for urban heat island
Gutman, G., & Ignatov, A. (1998). The derivation of the green vegetation fraction from studies. Remote sensing of Environment, 89(4), 467–483.
NOAA/AVHRR data for use in numerical weather prediction models. International Wu, J. (1999). Hierarchy and scaling: Extrapolating information along a scaling lad-
Journal of Remote Sensing, 19(8), 1533–1543. der. Canadian Journal of Remote Sensing, 25(4), 367–380.
Hay, G., Marceau, D., Dube, P., & Bouchard, A. (2001). A multiscale framework for Wu, J., & David, J. L. (2002). A spatially explicit hierarchical approach to model-
landscape analysis: Object-specific analysis and upscaling. Landscape Ecology, ing complex ecological systems: Theory and applications. Ecological Modelling,
16(6), 471–490. 153(1), 7–26.
Imhoff, M. L., Zhang, P., Wolfe, R. E., & Bounoua, L. (2010). Remote sensing of the Wu, J., & Loucks, O. L. (1995). From balance of nature to hierarchical patch
urban heat island effect across biomes in the continental USA. Remote Sensing of dynamics: A paradigm shift in ecology. Quarterly Review of Biology, 70(4),
Environment, 114(3), 504–513. 439–466.
Li, J., Song, C., Cao, L., Zhu, F., Meng, X., & Wu, J. (2011). Impacts of landscape structure Xiao, R.-b., Ouyang, Z.-y., Zheng, H., Li, W.-f., Schienke, E. W., & Wang, X.-k.
on surface urban heat islands: A case study of Shanghai, China. Remote Sensing (2007). Spatial pattern of impervious surfaces and their impacts on land sur-
of Environment, 115(12), 3249–3263. face temperature in Beijing, China. Journal of Environmental Sciences, 19(2),
Liu, H., & Weng, Q. (2008). Seasonal variations in the relationship between land- 250–256.
scape pattern and land surface temperature in Indianapolis, USA. Environmental Yuan, F., & Bauer, M. E. (2007). Comparison of impervious surface area and
Monitoring and Assessment, 144(1–3), 199–219. normalized difference vegetation index as indicators of surface urban heat
Nichol, J. (1998). Visualisation of urban surface temperatures derived from satellite island effects in Landsat imagery. Remote Sensing of Environment, 106(3),
images. International Journal of Remote Sensing, 19(9), 1639–1649. 375–386.
Nichol, J. E. (1996). High-resolution surface temperature patterns related to urban Yue, W., Xu, J., Tan, W., & Xu, L. (2007). The relationship between land surface temper-
morphology in a tropical city: A satellite-based study. Journal of Applied Meteo- ature and NDVI with remote sensing: Application to Shanghai Landsat 7 ETM+
rology, 35(1), 135–146. data. International Journal of Remote Sensing, 28(15), 3205–3226.
Nichol, J. E., Fung, W. Y., Lam, K.-s., & Wong, M. S. (2009). Urban heat island diagnosis Zhang, Y., Odeh, I. O., & Han, C. (2009). Bi-temporal characterization of land surface
using ASTER satellite images and ‘in situ’ air temperature. Atmospheric Research, temperature in relation to impervious surface area, NDVI and NDBI, using a
94(2), 276–284. sub-pixel image analysis. International Journal of Applied Earth Observation and
Oke, T. (1976). The distinction between canopy and boundary-layer urban heat Geoinformation, 11(4), 256–264.
islands. Atmosphere, 14(4), 268–277. Zhou, W., Huang, G., & Cadenasso, M. L. (2011). Does spatial configuration matter?
Owen, T. (1998). Using DMSP-OLS light frequency data to categorize urban envi- Understanding the effects of land cover pattern on land surface temperature in
ronments associated with US climate observing stations. International Journal of urban landscapes. Landscape and Urban Planning, 102(1), 54–63.
Remote Sensing, 19(17), 3451–3456. Zhou, W., Qian, Y., Li, X., Li, W., & Han, L. (2014). Relationships between land cover
Owen, T., Carlson, T., & Gillies, R. (1998). An assessment of satellite remotely-sensed and the surface urban heat island: Seasonal variability and effects of spatial and
land cover parameters in quantitatively describing the climatic effect of urban- thematic resolution of land cover data on predicting land surface temperatures.
ization. International Journal of Remote Sensing, 19(9), 1663–1681. Landscape Ecology, 29(1), 153–167.

You might also like