You are on page 1of 20

Sustainable and Resilient Infrastructure

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tsri20

Urban Heat Implications from Parking, Roads, and


Cars: a Case Study of Metro Phoenix

Christopher G. Hoehne , Mikhail V. Chester , David J. Sailor & David A. King

To cite this article: Christopher G. Hoehne , Mikhail V. Chester , David J. Sailor & David A. King
(2020): Urban Heat Implications from Parking, Roads, and Cars: a Case Study of Metro Phoenix,
Sustainable and Resilient Infrastructure, DOI: 10.1080/23789689.2020.1773013

To link to this article: https://doi.org/10.1080/23789689.2020.1773013

View supplementary material

Published online: 06 Jul 2020.

Submit your article to this journal

Article views: 61

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tsri20
SUSTAINABLE AND RESILIENT INFRASTRUCTURE
https://doi.org/10.1080/23789689.2020.1773013

Urban Heat Implications from Parking, Roads, and Cars: a Case Study of Metro
Phoenix
a a b b
Christopher G. Hoehne , Mikhail V. Chester , David J. Sailor and David A. King
a
Civil, Environmental, and Sustainable Engineering, Arizona State University, Tempe, Arizona, USA; bSchool of Geographical Sciences and
Urban Planning, Arizona State University, Tempe, Arizona, USA

ABSTRACT ARTICLE HISTORY


To understand the transportation sector’s role on influencing and mitigating heat in cities, this Received 28 October 2019
research quantifies added heat from pavement infrastructure and vehicle travel in the hot and Accepted 15 May 2020
automobile dependent metropolitan Phoenix, Arizona. Construction of a one-dimensional heat KEYWORDS
transfer model for local weather conditions and pavement design is combined with vehicle travel Heat; pavement;
densities to simulate spatiotemporal sensible heat flux magnitudes. In metro Phoenix, sensible transportation; parking;
heat from pavements and vehicles is comprised of 67% from roadways, 29% from parking, and Phoenix
3.9% vehicles. Concrete and asphalt pavement emit 15% and 37% more sensible heat compared to
the bare ground, respectively. Added sensible heat from pavement peaks during summer after­
noons when heat emissions relative to the ground are 26% and 46% greater for concrete and
asphalt. Results indicate pavement infrastructure contributes significantly to Phoenix’s urban heat
balance, and areas surrounding busy vehicle corridors may be undesirable for outdoor activities
during summer rush hours.

1. Introduction
As global urbanization and climate change persists, of an urban surface (such as asphalt and concrete) have
cities are becoming gradually warmer. One consequence a greater ability to absorb and store heat compared to the
of urbanization is Urban Heat Island (UHI), Earth’s natural terrain due to different intensive proper­
a phenomenon where urban areas are warmer than ties. Increased coverage of built infrastructure and
rural areas. Increasing urban heat from UHI and climate decreased vegetation in urban areas also reduce the poten­
change threatens urban vitality and prosperity by poten­ tial for evaporative cooling. Asaeda et al. (1996) found
tially reducing productivity and economic development asphalt pavement can contribute significant amounts of
(Graff Zivin & Neidell, 2014; Kjellstrom et al., 2009), heat relative to bare ground; afternoon heat absorption
increasing demand for energy (Burillo et al., 2019; from asphalt pavement within 30 m of the ground
Miller et al., 2008; Reyna & Chester, 2017), increasing accounted for four times the daily averaged total anthro­
urban infrastructure vulnerability (Bondank et al., 2018; pogenic heat across the Tokyo metropolitan area. Wasted
Gasser et al., 2019; Markolf et al., 2017; Schaeffer et al., heat from buildings, vehicles, and human metabolism also
2012), dissuading outdoor activity and travel (Karner contributes to urban warming. While it is well established
et al., 2015; Obradovich & Fowler, 2017; Stamatakis that internal combustion engine waste significant
et al., 2013), and causing increased heat-related injury amounts of energy as heat (Rajoo et al., 2014; Reddy
or death (CDC, 2012; Eisenman et al., 2016; Gasparrini et al., 2015), the influence of urban vehicle travel on the
et al., 2015; Kovats & Hajat, 2008). Given the breadth of urban energy balance has been less rigorously studied
externalities, there is a great desire to fully understand compared to other heat sources. This is due in part due
and mitigate urban heat. to poor quality transportation sector heat emissions
The transportation system is naturally embedded in (Smith et al., 2009). Hart and Sailor (2009) found up to
the issue of urban heat through infrastructure design and 2°C warmer air masses above urban roads during the
transportation planning choices. UHI is caused by weekday compared to the weekend in Portland, Oregon,
anthropogenic infrastructure and activity, and previous indicating that increased weekday vehicle travel may be
research has established heating from vehicles and pave­ the primary cause. Sailor and Lu (2004) found that heat­
ments as significant. Impervious and engineered materials ing from vehicles dominated the summer anthropogenic

CONTACT Christopher G. Hoehne chris.hoehne@asu.edu


Supplementary data for this article can be accessed here.
© 2020 Informa UK Limited, trading as Taylor & Francis Group
2 C. G. HOEHNE ET AL.

heating in six US cities, accounting for 47% to 62% of the to urban heat from vehicle travel and pavement infra­
total. structure. We focus this study on the Phoenix metropo­
Heat transfer models have been used frequently for litan region for three primary reasons: first, Phoenix has
many purposes, and have proven a viable tool to estimate a very auto-centric urban design with high automobile
surface temperatures and surface heat transfer in materi­ dependence and supporting infrastructure; second,
als including pavements. Even before the wide availabil­ Phoenix may suffer significant consequences of urban
ity of computer programing, heat transfer models were heat due to urban heat island, climate change, and
constructed and validated to assess pavement thermal rapid urban growth (Clark et al., 2019); and third, the
performance (Dempsey & Thompson, 1970). With an arid climate in Phoenix makes it desirable for modeling
increased interest in UHI in the late twentieth century sensible heat transfer. This study aims to answer two
(Arnfield, 2003; Oke, 1982), more research emerged that research questions: (1) What aspects of urban pavements
focused on explicitly modeling paved surfaces to under­ are most or least influential to sensible heat flux magni­
stand their influence on UHI; Asaeda et al. (1996) were tudes? and (2) How do pavement infrastructure and
the first to model and assess the effects of paved surfaces vehicle travel contribute to the urban heat balance?
on the near-surface urban climate. One-dimensional
heat transfer models using finite difference solving
2. Methodology
schemes are among the most popular due to straight-
forward implementation and ability to achieve reason­ Two approaches are used to quantify spatial and temporal
able predictions of pavement surface temperatures (Gui urban heat flux magnitudes from pavements and vehicles
et al., 2007; Hall et al., 2012; Hermansson, 2004; Wang & in metropolitan Phoenix, Arizona. First, a one-
Roesler, 2012). Despite the common use of heat transfer dimensional (1D) model based on fundamental heat
modeling in the study of pavements, results of such transfer is developed to approximate diurnal sensible
analyses have generally not been used to assess the effects heat fluxes from various types of pavements. The model
of paved infrastructure on UHI. Instead, the urban sur­ is validated by comparing simulated material surface tem­
face’s influence on UHI is more commonly assessed by peratures to remotely sensed land surface temperatures at
relating UHI intensities to spatial variability in land use various sites in metro Phoenix that are dominantly bare
and albedo (Carnielo & Zinzi, 2013; Dai et al., 2018; ground or covered by pavement. Additional pavement
Golden & Kaloush, 2006; Hart & Sailor, 2009; Minjun designs are then simulated to represent the various regio­
Kim et al., 2017; Sun et al., 2018; Wicki et al., 2018). nal urban roadway and parking pavement designs. Next,
Numerous studies have quantified heat fluxes from regional vehicle travel data for a typical day are combined
pavements and vehicles independently within their scope with internal combustion engine (ICE) vehicle efficiency
(Allen et al., 2011; Arnfield & Grimmond, 1998; Golden estimates from literature to estimate rates of wasted heat
& Kaloush, 2006; Ichinose et al., 1999; Smith et al., 2009); during vehicle travel. Heat from vehicle travel and pave­
however, only one study has quantified both simulta­ ments is then estimated at 250 m2 resolution across the
neously. Fujimoto et al. (2015) investigated the influence Phoenix metropolitan area to allow comparison to pre­
of vehicle travel on the heat balance surrounding an vious estimates of anthropogenic heating in cities.
urban intersection in Fukui, Japan. The authors found
that vehicle-related heat fluxes accounted for 3% to 12%
2.1. One-dimensional Heat Transfer Model
of the total winter heat balance depending on traffic
Overview
density and time of day. As a result of increased vehicle
travel, they predicted increased pavement surface tem­ Following extensive previous research on modeling fun­
peratures of 1.5°C to 4°C compared to measured pave­ damental heat transfer, a 1D model is developed that
ment surface temperature increases of 3.5°C. predicts temperatures and sensible heat flux of
Urban automobile travel is pervasive and often dom­ a delineated material according to its thermophysical
inatesmode share and land use in cities (Kenworthy & properties and surrounding environmental conditions.
Laube, 1999). Pavement infrastructure can make up 30% Only sensible heat transfer is considered as it is the
to 66% of the urban land cover (Akbari et al., 1999, dominant term affecting warming in arid climates
2003), and parking infrastructure alone may account such as Phoenix. The 1D transient heat conduction
for as much as 10% to 14% of incorporated urban land model ignores lateral heat transfer since the horizontal
in Los Angeles and Phoenix, respectively (Chester et al., dimensions of paving are two orders of magnitude lar­
2015; Hoehne et al., 2019). To help understand how city ger than the depth. The model balances surface energy
planning and the transportation sector can influence transfer from convection, incoming solar radiation, and
urban heat, this research aims to quantify contributions outgoing infrared radiation as well subsurface energy
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 3

transfer via conduction (Figure 1). Simulated materials where qsw;in is incoming shortwave (solar) radiation,
are idealized as a series of stacked nodes starting at the qsw;ref is reflected shortwave radiation, qconv is convec­
surface at continuing downward to a defined depth. tion, qlw;in is incoming longwave radiation, and qlw;out is
Heat transfer is first balanced between the nodes at an outgoing longwave (infrared) radiation. Outgoing long­
initial condition, then solved by stepping forward in wave radiation is assumed to obey the Stefan-Boltzmann
time using an explicit finite difference scheme. While law such that a surface emits longwave radiation as
many similar 1D models have been implemented and a black body. Therefore, net longwave radiation at the
validated in literature, this methodology most closely surface is defined as
replicates the implementation and some assumptions � �
of Gui et al. (2007) because it was also implemented qlw;net ¼ qlw;out qlw;in ¼ εσ Ts4 Tsky
4
(2)
and validated for conditions in Phoenix, Arizona.
To simulate heat transfer of a pavement or bare where ε is emissivity of the surface, σ is the Stefan-
ground, a number of input variables are utilized. Boltzmann constant, Ts is the surface temperature, and
Uniform or composite materials may be simulated, Tsky is the sky temperature. To approximate longwave
and for each unique layer of material, various material radiation between pavements and their surroundings,
properties are required: albedo (surface layer only), previous 1D models have assumed surrounding condi­
emissivity (surface layer only), layer thickness, thermal tions form a hypothetical sky temperature such that
conductivity, and volumetric heat capacity. Other para­ Tsky ¼ T1 ð0:004Tdew þ 0:8Þ0:25 (3)
meters that must be defined include sky view factor
(SVF), characteristic length of the surface, initial start­ where T1 is the atmospheric dry-bulb temperature and
ing temperature profile, nodal spacing, and time step Tdew is the dew point temperature (Chiasson et al., 2000;
length. The model is forced using hourly or sub-hourly Gui et al., 2007; Hall et al., 2012). Convective heat
measured solar radiation, air temperature, humidity, transfer at the surface is defined as
and wind velocity data. qconv ¼ h1 ðTs T1 Þ (4)
Temperature and sensible heat transfer is estimated
by transient energy balance of surface convection, where h1 is the convective heat coefficient of air.
incoming surface solar radiation, outgoing surface Assuming laminar flow over a flat plate, Çengel (2002)
infrared radiation, and subsurface conduction. The gen­ defines the convective heat coefficient as
eralized equation for net heat transfer (in W m−2) at the 0:3 0:5
h1 ¼ 0:664k1 Pr1 v1 L 0:5 0:5
U1 (5)
surface is defined as
where k1 , Pr1 , and v1 are the thermal conductivity,
qnet ¼ qsw;in þ qsw;ref þ qconv qlw;in þ qlw;out (1) Prandtl number, and kinematic viscosity of air,

Figure 1. One-dimensional Heat Transfer Diagram for a Typical Pavement. The boundary condition at the bottom of the subgrade
layer is assumed to be a constant temperature; a sufficient depth is chosen such that it is reasonable to assume there are no
temperature fluctuations with respect to time (in this case, 1.5 m).
4 C. G. HOEHNE ET AL.

respectively; L is the characteristic length of pavement conditions, after which the tolerance for convergence is
assumed to be the ratio of the pavement length in the relaxed to 1.0 K.
direction of wind to the perimeter; and U1 is the hor­ This outlined heat transfer model is implemented
izontal wind velocity. Net shortwave radiation at the using R statistical software, and documentation for
surface is defined as application in this analysis is maintained in a GitHub
repository (Hoehne, 2019).
qsw;net ¼ qsw;in Ψsky ð1 αÞ (6)

where qsw;in is the incoming shortwave (solar) radiation,


the sky view factor (SVF), and α is the albedo of the 2.2. Selected Pavement Designs and Model
surface. The generalized equation for subsurface sensi­ Validation for Phoenix Sites
ble heat transfer (conduction) is defined as Validation is performed by comparing various modeled
pavements and bare ground surface temperatures to
@T k @2T
¼ (7) remotely sensed land surface temperatures from the
@t ρc @x2
Advanced Spaceborne Thermal Emission and
where T is temperature, t is time, k is thermal diffusivity, Reflection Radiometer (ASTER) onboard the Terra
ρ is density, c is specific heat capacity, and x is depth. satellite (NASA, 2019). ASTER On Demand Surface
Where multiple layers of differing materials are present Kinetic Temperature measurements are generated
(as is common in pavements), boundary conditions are from five thermal infrared bands and are atmospheri­
also implemented. Due to a lack of information, it is cally corrected (USGS, 2018). Validation sites are
assumed that thermal contact resistance between layers selected across the Phoenix metro region such that the
is zero. While this assumption will affect subsurface materials of interest uniformly and completely cover
pavement temperatures, a similar pervious model did a 90 m2 ASTER raster pixel. Major roadways in
not find significant impacts on near-surface tempera­ Phoenix do not reach 90 m in width, and as a result,
tures with zero thermal contact resistance between the sites selected are asphalt or concrete parking lots and
layers (Gui et al., 2007). As a result, the upper and airport tarmacs. Figure 2 displays a sample of the
lower interface temperatures can be assumed to be selected sites highlighted on an ASTER surface tempera­
equal, and are idealized as a single node. Therefore, ture image. For a detailed list of the selected sites used in
the boundary condition at the interface must obey validation, see Appendix S.1.
A variety of pavement designs are categorized into
Tb 1 Tb Tb Tbþ1 three classes and developed by referencing relevant lit­
kb 1 ¼ kbþ1 (8) erature and pavement engineering design recommenda­
Δxb;b 1 Δxb;bþ1
tions or requirements. A summary of the range of
where subscriptb refers to the node at the boundary and material parameters for various pavements and bare
b 1 and b þ 1 refer to the conditions at nodes imme­ ground is displayed in Table 1. Pavement designs are
diately above and below the boundary node. classified in one of three pavement classes: (1) asphalt
To ensure feasible results from the explicit finite pavement (primarily hot-mixed asphalt) which utilizes
difference scheme, all calculations are required to satisfy a bitumen binder with aggregate; (2) concrete pavement
the Courant-Friedrichs-Lewy (CFL) condition for stabi­ (primarily Portland cement concrete) which utilizes
lity (Gui et al., 2007; Heath, 2002). In order to satisfy the a cement binder with aggregate; and (3) composite
CFL stability condition, sufficiently small time and asphalt-concrete pavements that combine distinct bitu­
nodal spacing are required. To achieve stable solutions men-bound and cement-bound layers in a single pave­
such that simulation times are reasonable, a nodal spa­ ment design such as whitetopping overlaid on an
cing of 10 mm and time step spacing of 30 seconds are asphalt pavement or rubberized asphalt overlaid on
chosen. Therefore, linear interpolation between weather Portland cement concrete (PCC). Whitetopping is
observations is required to achieve matching temporal a common method where a thin PCC layer is overlaid
frequency. on top of an existing asphalt pavement. Whitetopping
To ensure initial conditions begin at an equilibrium, has become increasingly popular for existing pavement
the initial conditions are iteratively simulated until con­ rehabilitation and to increase the surface layer albedo
vergence occurs between the initial (t0 ) and first time for potential cooling benefits. Lastly, a fourth material
step (t1 ). The tolerance for convergence is defined such class emulating desert soil is created to serve as
that each nodal temperature at t1 is within 0.1 K of its a reference for undeveloped natural land that would be
t0 temperature for the first 100 iterations of the initial found in an arid region such as Phoenix.
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 5

Figure 2. ASTER Nighttime Land Surface Temperature across Phoenix Metro on March 22nd, 2014 with Selected Validation Sites
Highlighted. Satellite imagery of four validation sites are shown where the inner box is the 90 m2 pixel location of the ASTER cell.

Various pavement designs and bare ground composi­ 2.3. Estimating Pavement and Vehicle Heat at a
tions are simulated to emulate the expected materials City-wide Scale
observed at each site. Simulation time periods are chosen
by identifying ASTER observation dates that have less Profiles of pavement and vehicle heat transfer are applied
than 10% cloud cover, no precipitation, and all data to regional pavement and traffic flow data to approximate
passing quality control checks. ASTER observation spatial and diurnal heat flux magnitudes across the
dates occurring during the day and night as well as Phoenix metropolitan region. A pavement infrastructure
occurring across all four seasons are selected to ensure inventory for metro Phoenix is developed by combining
a variety of dry weather conditions. Historical weather OpenStreetMap (OSM) roadway data (OpenStreetMap
data for the same periods as the ASTER observation dates contributors, 2019) and Phoenix parking inventory data
are retrieved for all stations across the entire metro from Hoehne et al. (2019). Average annual daily traffic
Phoenix region from the MesoWest weather data net­ estimates are obtained from Maricopa County origin-
work (University of Utah, 2019). All weather stations destination travel demand data simulated in MATSim
with consistent weather observations for each simulation travel modeling software. Simulation outputs for pavement
date are chosen. Consistent weather observations are designs by roadway and parking functional classes are
defined as (1) having an average of at least one weather combined with waste heat flux estimates from vehicle
observation per hour for all relevant variables during the travel and linked to the roadway and parking inventory
desired dates (solar radiation, air temperature, humidity, data. Spatial and temporal mean daily and hourly sensible
and wind speed); (2) having no gaps in observations for heat fluxes for a typical clear spring or fall day are esti­
greater than 2 hours for all relevant variables; (3) no mated at a 250 m2 spatial resolution for all of metro
rainfall during analysis period; and (4) 95% of intra- Phoenix.
station observations fall within two standard deviations Utilizing Phoenix parking inventory data at the indi­
of the intra-hour mean across all relevant variables and vidual property (parcel) level from Hoehne et al. (2019)
all stations in the region. Each validation site is assumed and OSM roadway network data, fractional pavement
to have a SVF of 0.947, the mean of the Phoenix metro area is estimated across the region. Fractional areas of
area (Middel et al., 2018). Pavement and bare ground pavements are estimated by different functional classes
compositions are then simulated for a period of 3 days corresponding to expected variations in functional
such that the final day corresponds to a desired ASTER design. Roadway pavements are split into four major
measurement using the hourly mean weather variables classes: highway, major arterial, minor arterial, and local
from all selected stations with specified material design roads. Parking pavements are split into two major classes:
parameters emulating the validation site materials. residential parking, and non-residential parking. Each
6 C. G. HOEHNE ET AL.

Table 1. Ranges of Material Parameters Utilized for Pavement Design and Bare Ground from Literature. ‘Ground’ refers to bare, native,
and uncompacted material consistent with materials found in the Phoenix region (i.e., desert soil) that would represent undeveloped
land. ‘Subbase’ refers to the aggregate supporting layer between the pavement wearing course and compacted ground. ‘Subgrade’
refers to compacted ground underneath the pavement.
Parameter Units Asphalt Concrete Subbase Ground or Subgrade
Albedo,~
α dimension-less 0.05–0.15 a 0.18–0.29 b NA 0.30 e
0.08–0.09 b 0.20–0.40 a 0.40–0.50 f
0.17 c 0.31–0.43 c
0.12–0.20 d 0.42–0.46 d
Emissivity, dimension-less 0.85 c, g 0.90 j NA 0.90–0.97 e
ε 0.90 h 0.92–0.96 k
0.90–0.95i
Thermal conductivity,k W
m�K
1.2 c 1.2 o 1.5 h 1.0 c
1.4–1.8 l 1.2–1.4 k 3.0 m 1.2 m
1.5 h 1.5 j 1.8 h
1.6 m 2.2 m
1.9–2.2 n
Density,ρ kg 2200 c 1800–2100 k, s 2400 h 1500 c
m3
2300 n 2300 o 2200 h
2300–2500 l 2400 j
2400–2600 h
1800–2500 o
Specific heat capacity,c J
kg�K
810–960 k 840–1050 k 800 h 1100 h
850–860 h 1000 j 1900 c
900 o
920 k
1200–1900 l
Layer thickness mm 40–200 o, p, q 100–300 p, r 100–300 o, p, q NA
a
(Qin, 2015)
b
(Li et al., 2013a)
c
(Gui et al., 2007)
d
(Golden & Kaloush, 2006)
e
(Monteith & Unsworth, 2013)
f
(Dobos, 2011)
g
(Hermansson, 2004)
h
(Minhoto et al., 2006)
i
(Tan & Fwa, 1992)
j
(Bentz & Turpin, 2007)
k
(Hu et al., 2017)
l
(Luca & Mrawira, 2005)
m
(Wang & Roesler, 2012)
n
(Im et al., 2015)
o
(Hall et al., 2012)
p
(ADOT, 2017b)
q
(FAA, 2016)
r
(USACE, 2018)
s
(Rafi & Aziz, 2019)

parking space (residential or non-residential) is assumed specifications by local municipalities. The majority of
to occupy approximately 31 m2 of space consistent with urban pavements in metro Phoenix fully or partially utilize
previous research (Hoehne et al., 2019; Holland, 2014; asphalt with over 80% of Arizona highways utilizing rub­
Manville & Shoup, 2005). On-street parking is ignored as berized asphalt pavement (EPA, 2016). Commonly, regio­
the roadway inventory accounts for parking space on nal city streets and highways are paved or resurfaced with
roadway shoulders and metered on-street parking in rubberized asphalt to improve durability, reduce traffic
Phoenix is insignificant. Links from the OSM road net­ noise, and improve ride smoothness (ADOT, 2017a,
work are spatially buffered by the mean expected roadway 2019). Asphalt pavements are often preferred for pavement
widths and rasterized at a 250 m2 resolution by functional design due to their viscoelastic properties that can provide
class. To ameliorate the issue of parking inventory data improved long-term performance under thermal and load-
not explicitly spatially locating spaces, parking area is bearing stress in contrast to rigid concrete pavement
spatially assigned at each property by buffering around designs (Hall et al., 2012). Pavements made only of con­
the property centroid to create an area of parking cen­ crete are primarily found in single family residential park­
tered on the property. ing (e.g., driveways). As a result, asphalt is assumed to be
Each functional class of pavement is assigned pavement the dominant pavement type. Table 2 overviews the
designs such that it corresponds to the expected in-situ assumed pavement designs assigned by pavement func­
pavement and complies with required engineering design tional class following guidelines from the Arizona
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 7

Department of Transportation (ADOT, 2017b) and typical Therefore, for properties requiring greater than 100
pavement designs from literature in Table 1. spaces, we assume a conservative approach where park­
The selected pavement designs to estimate roadway ing area visible to the sky follows a non-linear decay
and parking heat fluxes are simulated using the same under such that properties at 100 spaces have an unad­
specifications as the validation phase with two alterations: justed parking SVF of 1.0 which decays to a mean SVF of
(1) measured solar radiation is replaced with estimated 0.325 at 100,000 spaces. While these assumptions for high
solar radiation using the ‘insol’ R package (Corripio, densities of parking are difficult to validate given the lack
2019) with inputs of local latitude, observed relative of research and data on parking and SVFs, these are edge
humidity, observed air temperature, Julian day, time cases; only 497 properties of the 1.55 million in the
of day, and an ozone thickness of 2.75 mm; and (2) all urbanized area have greater than 1,000 parking spaces.
pavement designs for all dates are simulated under 1.0 However, if left unadjusted, numerous adjacent 250 m2
SVF and 0.1 SVF to capture shaded and unshaded pave­ are otherwise found to be entirely covered by parking
ment scenarios. Estimated insolation is used to represent pavement. For details on the specific parking SVF decay
a clear day and avoid impacts of sporadic cloud cover. For functions assumed, see Appendix S.2.
full details on all parameters simulated, see Appendix S.1. Metropolitan-wide vehicle travel data are combined
As pavements in the region are typically not comple­ with vehicle efficiencies to estimate vehicle waste heat
tely visible to the sky, the heat transfer of partially shaded during operation from energy consumption by OSM
pavements is incorporated by utilizing SVF data along roadway link. The partial amount of consumed energy
the Phoenix roadway network from Middel et al. (2018) wasted as heat is uniformly attributed to the traversed
to calibrate heat transfer from pavements under direct roadway link. Vehicle travel data are obtained from
solar radiation exposure versus pavements under shade. a MATSim regional travel demand model that utilizes
Moise and Aynsley (1999) found that incoming daylight travel and population data provided by the Maricopa
shaded-to-unshaded radiation had a median ratio of 0.09 County Association of Governments. This obtained
for horizontal shading and 0.11 for vertical shading. data represents all personal light duty vehicle trips across
Therefore, we assume pavement in the shade receives the regional OSM road network for a typical spring or
0.10 of estimated unshaded incoming radiation. This is fall day. This travel data excludes heavy duty vehicle
implemented in modelling such that any areas where travel such as freight and public transit, and due to
partial shade is present (SVF < 1.0), the portion of shaded a lack of similar high fidelity data, heavy vehicle traffic
pavement is treated as though it has 0.10 SVF and the on links is not considered. A typical passenger vehicle
unshaded portion has 1.0 SVF. For example, may lose 60–64% of energy to heat during city driving
a neighborhood with 0.80 SVF would have 80% of pave­ and 56–60% of energy to heat during highway driving
ment modeled as unshaded (SVF = 1.0), and 20% of the (DOE, 2019). Other academic literature cites ranges of
pavement that is shaded is modeled with a SVF of 0.10. 30% to 80% of fuel energy wasted as heat during vehicle
All pavements (parking and roadway) are are assigned a operation (Hsiao et al., 2010; Shiho Kim et al., 2011; Orr
measured SVF along the roadway in the 250 m2 cell with et al., 2016; Rajoo et al., 2014; Yang & Stabler, 2009).
exceptions for extremely high cases of parking. Properties Given the most commonly cited factor is ‘nearly two
that require large amounts of parking are increasingly thirds,’ this analysis assumes a static mean of 65% of
likely to be underground or inside a parking structure. fuel energy lost as heat to the surrounding environment

Table 2. Assumptions for Pavement Design and Vehicle Travel Applied to the Phoenix Metropolitan Area. Vehicle energy released per
kilometer is estimated from urban city and highway driving efficiencies from Davis and Boundy (2019) using a mean of 31.7 MJ per
liter of gasoline or gasoline-equivalent fuel.
Generalized functional Assumed mean Assumed two-way road width Assumed coverage of asphalt Assumed energy released
class description pavement or parking space area vs. concrete pavement from traversed vehicles
thickness
Highway or freeway 280 mm 43 m 95% Asphalt 356 Wh/km
5% Concrete (40 MPGe)
Major or minor arterial 210 mm 28 m 90% Asphalt 561 Wh/km
road 10% Concrete (25 MPGe)
Major or minor collector 170 mm 18 m 90% Asphalt 718 Wh/km
road 10% Concrete (20 MPGe)
Minor collector or local 140 mm 11 m 90% Asphalt 718 Wh/km
road 10% Concrete (20 MPGe)
Commercial parking 140 mm 31 m2 85% Asphalt NA
15% Concrete
Residential Parking 140 mm 31 m2 10% Asphalt NA
90% Concrete
8 C. G. HOEHNE ET AL.

for all vehicles during travel. For this analysis, we focus pavements and 18 W m–2 from concrete-surfaced pave­
on the heat lost during driving (65%) as a lower bound ments. During the winter, added sensible heat flux magni­
for vehicle heat emissions. When factoring in friction tudes decline with daytime magnitudes of 110 W m–2 for
(e.g., braking, drag), it should be noted that as much as asphalt surfaced pavements and 66 W m–2 for concrete-
93–97% of fuel energy is likely converted to heat with the surfaced pavements. Nighttime winter added sensible heat
remainder being stored in unburned hydrocarbons due flux magnitudes are 27 W m–2 asphalt-surfaced pavements
to inefficient fuel combustion (Baglione et al., 2007; and 10 W m–2 concrete-surfaced pavements.
Wallington et al., 2006). Total energy consumed by vehi­ A pavement’s daytime maximum outgoing sensible heat
cles for each link is calculated by multiplying total vehicle flux is most influenced by its albedo, while its nighttime
kilometers traveled (VKT) by the assumed traversed minimum outgoing sensible heat flux is most influenced by
vehicle efficiency. As vehicle driving efficiencies depend its emissivity. An increase in albedo of 0.01 resulted in
on vehicle characteristics and driving patterns, different a decrease of maximum afternoon outgoing sensible heat
efficiencies are assigned by roadway functional class. fluxes by 5.5 W m–2 (95% confidence interval: 4.7 to 6.2
Recent studies indicate light duty vehicle efficiencies W m–2; R2 = 0.96; p < 0.001). A decrease in emissivity of
can range widely (Davis & Boundy, 2019). In 2017, the 0.01 resulted in a decrease of minimum nighttime outgoing
estimated real-world fuel economy for US light duty sensible heat fluxes by 1.4 W m–2 (95% confidence interval:
vehicles was 10.6 km/L (24.9 mi/gal) (EPA, 2017). 0.73 to 2.0 W m–2; R2 = 0.69; p < 0.001). Albedo more
Vehicles are assumed to have the highest efficiency on strongly impacts maximum (daytime) sensible heat fluxes
highways and lowest efficiency on local roads. For details because albedo impacts the fraction of incoming solar
on the applied vehicle efficiencies, see Table 2. Vehicles radiation which occurs only during sunlight hours.
dominantly emit waste heat as thermal radiation from In addition to albedo and emissivity, altering
the engine and brakes, and convection from the exhaust; a pavement’s thermal inertia properties has noticeable
impacts from tire friction and convective heat transfer impacts on the diurnal sensible heat flux magnitudes.
from vehicle motion have been found to be insignificant, Thermal inertia describes the slowness of material to
accounting for 1% or less of total balance near the road approach thermal equilibrium (e.g., high thermal inertia
surface (Fujimoto et al., 2015). As a result, we consider materials are slower to reach thermal equilibrium) and is
only sensible heat emitted via exhaust and the engine. equivalent to the square-root of the product of the thermal
conductivity (k), density (ρ), and specific heat capacity (c)
with SI units of J m−2 K−1 s−1/2. An increase in a pavement
3. Results surface layer thermal inertia by 100 J m−2 K−1 s−1/2 resulted
3.1. Evaluating Factors Influencing the Thermal in a decrease of maximum afternoon outgoingsensible
Performance of Pavements heat fluxes 8.6 W m–2 (95% confidence interval: 1.1 to
16 W m–2; R2 = 0.57; p = 0.031) and an increase in
Across all seasons in Phoenix, asphalt and concrete pave­ minimum nighttime outgoing sensible heat fluxes by
ments have greater diurnal outgoing sensible heat fluxes 1.7 W m–2 (95% confidence interval: 1.1 to 2.3 W m–2;
relative to the natural bare ground. The increase of out­ R2 = 0.88; p < 0.001). Thermal conductivity was the most
going sensible heat flux from pavements relative to the bare influential thermal inertia factor influencing minimum
ground is defined as the added sensible heat from pave­ and maximum sensible heat fluxes, while specific heat
ments. Asphalt surfaced pavements have mean daily added capacity was the least impactful. With the exception of
sensible heat fluxes of 70 W m–2 relative to the bare ground subsurface thermal conductivity, subsurface layer thermal
(37% increase), and concrete-surfaced pavements have inertia properties were insignificant in influencing the
mean daily sensible heat fluxes of 33 for W m–2relative to diurnal outgoing sensible heat fluxes at the surface.
the bare ground (15% increase). Figure 3 displays the To further explore the impact thermal inertia properties
summer and winter mean diurnal outgoing sensible heat have on sensible heat flux magnitudes, the highest and
fluxes for simulated asphalt pavements, concrete pave­ lowest literature values of thermal conductivity, density,
ments, and bare ground (desert soil). The largest added and specific heat capacity (Table 1) are compared with all
sensible heat flux magnitudes from pavements occur in other parameters constant to test a material’s dirurnal sen­
summer around 3 pm when asphalt surfaced pavements sible heat flux sensitivity to its thermal inertia properties.
contribute 143 W m–2(46%) more than the natural ground, Figure 4 displays diurnal outgoing heat fluxes from four
and concrete-surfaced pavements contribute 80 W m–2 different types of pavements with variations only to the
(26%) more than the natural ground. During summer thermal inertia properties. Overall, high thermal inertia
nights, added sensible heat from pavements is still signifi­ pavements reduced the mean daily outgoing sensible heat
cant with magnitudes of 44 W m–2 from asphalt surfaced flux across all seasons by 23 W m–2 compared to low
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 9

Figure 3. Mean Diurnal Outgoing Sensible Heat Flux for Simulated Asphalt Pavements, Concrete Pavements, and Bare Ground (Desert
Soil) during Summer and Winter Periods. Outgoing sensible heat flux is defined as outgoing convection plus outgoing infrared
radiation. All pavements are assumed as completely unshaded. For these simulations, asphalt surfaces had a mean albedo of 0.15,
concrete surfaces had a mean albedo of 0.30, and the bare ground had a mean albedo of 0.40.

thermal inertia pavements (asphalt only: 28 W m–2; asphalt pavement infrastructure and vehicle travel of 13 W m–2
overlays on PCC: 26 W m–2; concrete only: 21 W m–2; and across metro Phoenix for; roadway pavement contributes
white-topped asphalt 19 W m–2). Low thermal inertia 8.5 W m–2, parking pavement contributes 3.6 W m–2, and
pavements increased maximum daytime outgoing sensible vehicles contribute 0.49 W m–2. For areas with 10% or
heat fluxes by 86 to 134 W m–2 relative to high thermal greater coverage of pavement infrastructure, the total
inertia pavements. During nighttime, low thermal inertia mean daily heat flux rises to 19 W m–2. In more dense
pavements decrease the minimum outgoing sensible heat regions with high pavement coverage and vehicle travel,
fluxes by 15 to 23 W m–2 relative to high thermal inertia heat fluxes from vehicles and pavements may reach as high
pavements. High thermal inertia pavements were found to as 73 W m–2.
have delayed maximum sensible heat flux magnitudes by Pavement infrastructure typically dominates contribu­
up to 45 minutes for asphalt surfaced pavements and up to tions to the urban heat balance relative to waste heat from
60 minutes for concrete-surfaced pavements. Only the vehicle travel both spatially and temporally in metro
thermal inertia properties of a pavement’s surface layer Phoenix. This is true even given that ignoring vehicle
were found to significantly affect a pavement’s thermal braking heat losses might result in an underestimation
response. It should be noted that these results reflect of vehicle heat losses by up to 30%. Figure 5 displays the
a shorter timescale of these pavements’ thermal behavior, spatial variation in mean daily added sensible heat fluxes
and over longer periods with fluctuating weather, from roadways and pavements in metro Phoenix. Figure
a pavement’s thermal performance may differ. 6 displays the temporal variation in mean daily sensible
heat fluxes from roadways and vehicles in metro Phoenix.
Total heat from pavements and vehicles is comprised of
3.2. Spatiotemporal Heat Fluxes from Pavements
67% from roadway pavements, 29% from parking pave­
and Vehicles in Phoenix
ments, and 3.9% from light duty vehicles. However, dur­
Spatiotemporal heat fluxes from pavements and vehicles ing peak daytime travel periods, total heat from vehicles
are assessed for a typical (seasonal) day at a resolution of makes up 30% in the morning rush hour (8 am) and 18%
250 m2 for areas with at least 1% coverage of pavement. in the evening rush hour (5 pm). These results agree with
We find a mean daily added sensible heat flux from Fujimoto et al. (2015) which found vehicle heat fluxes
10 C. G. HOEHNE ET AL.

Figure 4. Comparison of Outgoing Sensible Heat Fluxes for High and Low Thermal Inertia Properties across Four Simulated Pavement
Types. The first layer (L1) and second layer (L2) thermal inertias (TI) are displayed for each simulated case in J m−2 K−1 s−1/2. All
nonthermal inertia parameters were held constant. Composite pavement design in (b) and (d) are identical to pavements (a) and (b)
respectively with only the additional asphalt and concrete overlay. All pavements are assumed as completely unshaded.

accounted for 3–12% of total heat flux across a road sur­ 3.3. Pavement Heat Transfer Model Validation
face with constant traffic. Across all validation simulations, modeled surface tem­
Sensible heat flux magnitudes from vehicles can peratures of various pavement designs were compared
reach as high as 132 W m–2 over the highest traf­ to the measured ASTER satellite land surface tempera­
ficked highways during rush hour, making up 74% tures at the validation sites. Root Mean Square Error
of the pavement-vehicle heat balance. However, the (RMSE) across all sites and pavements was 5.2°C and
mean daily heat flux magnitudes from vehicles Mean Absolute Error (MAE) was 4.4°C. Figure 7 shows
across all highway and arterial roads is much the modeled versus observed surface temperatures by
lower at 22 W m–2 and 17 W m–2, respectively. season and time of day. Seasonality had little effect on
Across low-trafficked collector and local roads, errors with spring and summer having slightly higher
vehicles contribute a daily average of 2.7 W m–2 MAE of 4.3°C and 4.8°C than winter and fall MAE of
and 0.64 W m–2 respectively while the pavement 4.2°C and 4.0°C. Daytime predicted surface tempera­
contributed a mean of 66 W m–2 (relative to tures had an MAE of 4.6°C versus 4.1°C for night-time
unpaved natural ground). This indicates that areas predictions. The most accurate concrete surfaced pave­
surrounding major arterials and highways with high ment design was a 200 mm thick PCC pavement with
vehicle traffic are the only areas that would see low albedo (2.7°C MAE), and a 140 mm asphalt pave­
measurable impacts to local climate as a result of ment (3.1°C MAE). Asphalt pavements typically had
vehicle use. higher predicted surface temperatures than the ASTER
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 11

Figure 5. Mean daily added sensible heat flux from roadway pavements, parking pavements, and vehicles in metropolitan Phoenix, AZ
(urbanized Maricopa County) at a 250 m2 resolution. Only light duty vehicle travel is included. Cells with less than 1.0 W m–2 are
ignored.

Figure 6. Mean Diurnal Sensible Heat Flux over Roadways from Pavements and Vehicles. Heat fluxes are averaged across the roadway
area only. Pavement heat fluxes are for an unshaded pavement. Vehicle travel before 4:30am is not present in the travel data and
therefore vehicle heat fluxes before this time are not estimated and shown as zero.
12 C. G. HOEHNE ET AL.

Figure 7. Modeled Versus Observed Surface Temperatures for Four Material Classes by Season and Time of Day. Note that not all sites
had ASTER observations for every simulation date.

observed surface temperatures (Figure 7a). There are pavements is common especially in arid climates (Gui
a number of reasons that may cause this discrepancy, et al., 2007; Hermansson, 2004), the heat balance of soils
but the most likely factor is differing in-situ albedos may be more impacted by latent heat transfer. Days with
because albedo is the strongest single parameter to pre­ precipitation were ignored to minimize the potential
dict pavement surface temperature. While sites selected impact of latent heat transfer at validation sites. In arid
are nearly covered by a uniform material, small amounts climates like Phoenix, dry soils typically have latent heat
of non-asphalt materials may alter the average ASTER fluxes well below an order of magnitude than sensible
pixel albedo and thermal properties, causing less heat (Albertson et al., 1995), but may peak as high as
absorbed and retained heat over time. For example, 50 W m–2 at the surface during morning hours
one site of an asphalt parking lot had a small amount (Heusinkveld et al., 2004). A previous study found dou­
of concrete, vegetation, and white stripping paint. bling or halving the latent heat flux in a dry soil caused
+0.4 and – 1.4 C or less change to the maximum soil
surface temperatures (Nobel & Geller, 1987). Given
3.4. Sensitivity and Uncertainty these findings, it is expected that ignoring latent heat
Given limitations of the imposed methodology, predic­ transfer in bare soils may have marginal impacts on
tion errors observed in validation are most likely attrib­ estimations of total added heat from pavements but
uted to: (a) omission of latent heat fluxes; (b) differing would not significantly alter the trends of added sensible
in-situ pavement design and soil compositions than heat presented in the results.
assumed; (c) limitations of the available weather data A number of other factors of urban vehicle travel and
in explaining the site microclimate; and pixel obstruc­ pavement infrastructure design may influence the sen­
tions such as from shade, cars (on parking lots), or other sitivity and uncertainty of sensible heat flux magnitudes
objects that would influence the land surface tempera­ such as roadway design widths, vehicle driving efficien­
tures detected by ASTER. While omitting latent heat cies, asphalt versus concrete pavement coverage, and the
flux transfer in simulating the surface heat balance of intensive properties of a pavement. The sensitivity and
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 13

uncertainty of the results are evaluated by simulating City had peaks less than 15 W m–2. This study finds
high and low estimates for many of these factors. heat fluxes from pavements in Phoenix relative to the
Assuming smaller roadway widths, increased vehicle native ground reach as high as 70 W m–2 at 250 m2
efficiencies, increased use of concrete pavements relative resolution and 143 W m–2 directly over the pavement
to asphalt, and smaller heat flux magnitudes from pave­ during summer afternoons. This indicates pavement
ments relative to the ground, mean daily added sensible infrastructure may make up a significant portion of
heat flux of from pavement and anthropogenic heat urban heat fluxes, especially during summers and in
from vehicles could decrease from 13 W m–2 to 6.1 more sprawled urban areas. However, more research is
W m–2 across metro Phoenix. Conversely, mean daily needed that compares pavements to other anthropogenic
added sensible heat flux from pavement infrastructure sources for the same region, scale, and time period.
and anthropogenic heat from vehicle travel could Many studies quantify urban anthropogenic heat
increase to 19 W m–2 across the urban area with the from buildings, vehicles, and metabolic processes, but
opposite of aforementioned assumptions (e.g., wider none considers the added heat from pavement infra­
roads, increased vehicle waste heat during travel, etc.). structure within their scope. As a result, no research has
In both of the extreme cases, contributions from vehi­ quantified heat flux from pavements simultaneously
cles are marginal across the whole urbanized area, with other anthropogenic heat sources. This may be
accounting for 0.38 to 0.60 W m–2. Roadway pavement because pavements do not waste heat through mechan­
heat contributions have a total sensitivity of 3.2 to ical or metabolic processes, but heat from pavement
13.7 W m–2, and parking pavements a sensitivity of 2.5 infrastructure is undoubtedly a consequence of urban
to 4.8 W m–2. anthropogenic activity. While these outcomes for metro
Phoenix may not be generalizable due to its climate,
natural geology, and auto-centric urban design, pave­
4. Discussion
ment infrastructure contributes significantly to urban
Previous estimates of anthropogenic heating from build­ heating. Future research that aims to holistically quan­
ings, vehicles, and metabolism are similar in magnitude tify urban heat flux magnitudes should include esti­
to this analysis of only pavement and vehicle heating in mates of added heating from pavement infrastructure
metro Phoenix. Anthropogenic heat fluxes in cities and other unnatural surface materials in combination
(excluding pavements) commonly range from 2 to with typical anthropogenic sources. This is important at
60 W m–2 during the summer to 4 to 210 W m–2 in the the city-wide and microclimate scale. Fully understand­
winter, with buildings contributing the highest propor­ ing the opportunities for related heat mitigation strate­
tions, followed by contributions from vehicles and mar­ gies (e.g., the impact of shading or changing SVF) may
ginal contributions from metabolic activity (Allen et al., require including pavements in the local heat balance
2011; Sailor & Lu, 2004; Taha, 1997). Allen et al. (2011) through microclimate, fluid, and coupling modeling
found mean urban anthropogenic heat fluxes from build­ approaches (Krayenhoff & Voogt, 2010; Wang et al.,
ings, vehicles, and metabolic processes to be 20 W m–2 2016). The approach presented herein should be adap­
across London and 60 W m–2 across Tokyo. Smith et al. table for use in other urban domains and allow for
(2009) quantified heat fluxes in greater Manchester, UK future estimates of heat fluxes in cities to include
from buildings, traffic, and metabolism at the same spatial added heat from pavement infrastructure.
resolution of this study (250 m2) and found mean heat Planning for urban density over urban sprawl may
emission of 6.12 W m–2, reaching as high as 23 W m–2 in reduce urban heat contributions from the transporta­
city center areas. In greater Manchester (1,960 people tion sector, but it is unclear if it would provide a net
per km2), buildings accounted for approximately benefit to mitigating urban heat. Auto-centric urban
3.67 W m–2 and vehicles accounted for 1.96 W m–2 design inhibits urban density and can lead to high
while in greater Phoenix (1,210 people per km2), pave­ coverage of pavement infrastructure supporting auto­
ment contributions alone accounted for 12.1 W m–2 but mobile dependence. In metro Phoenix, total roadway
a smaller amount from vehicles of 0.49 W m–2. Given that pavement coverage is dominated by low-trafficked local
higher anthropogenic heating typically occurs in the win­ and collector roads, often in residential neighborhoods,
ter due to increased building energy use, relative contri­ contributing to 56% of the total mean daily heat balance
butions from pavement infrastructure may be much from all pavements and vehicles despite accounting for
more significant in the summer. Sailor and Lu (2004) only 22% of the total daily VKT. An analysis of Atlanta
estimated peak summer anthropogenic heat fluxes of found lower density residential developments contri­
30–60 W m–2 in Chicago, San Francisco, and bute more radiant heat energy than higher density
Philadelphia, but the less dense Atlanta and Salt Lake developments to surface heat island (Stone & Rodgers,
14 C. G. HOEHNE ET AL.

2001). While parking pavements contribute a smaller pavement infrastructure. While increasing pavement
fraction to the heat balance than roads, it is still signifi­ albedo can significantly reduce the total heat stored
cant and dominantly from paved parking lots. This and emitted, it also comes at the sacrifice of increasing
sprawled urban design may be problematic for cities the incident reflected solar radiation. As a result, high
concerned with issues of urban heat and climate change; albedo pavements may compromise the thermal com­
more sprawled urban metros have a higher prevalence fort of nearby pedestrians (Erell et al., 2014; Li et al.,
and increased rate of extreme heat events after control­ 2016), and may increase mean radiant temperatures
ling for climate and population growth (Stone et al., experienced by 7.8°C (Taleghani et al., 2016). To avoid
2010). Increasing urban density could reduce automo­ this drawback but still improve the thermal environ­
bile VKT and pavement infrastructure needs due to ment through pavement design, increasing pavement
closer destinations, mixed-use planning, and more thermal inertia may be a viable alternative. The feasi­
effective public transit, thus reducing the transportation bility of increasing pavement thermal inertia has rarely
sectors' influence on anthropogenic and pavement- been discussed, but Yun et al. (2014) found using sur­
added urban heat. Additionally, densification contri­ rogate aggregates practical for reducing concrete ther­
butes to increased prevalence of urban canyons which mal conductivities in building applications and noted
improve human thermal comfort (Andreou & Axarli, that aggregate size does not appear to affect thermal
2012; Johansson, 2006; Middel et al., 2014). However, an behaviors. Increasing a pavement’s thermal inertia will
issue still exists: cities with higher population densities slow its ability to warm and reach thermal equilibrium,
consistently have higher estimates of total urban anthro­ resulting in an average decrease in daytime heat fluxes
pogenic heat (Allen et al., 2011; Sailor & Lu, 2004). Yet but an average increase in nighttime heat fluxes. During
these analyses exclude pavement infrastructure heat periods of extended heating or cooling, high thermal
fluxes, so the implications of increased urban density inertia pavements will more slowly warm up or cool off.
on urban heat are unclear. As urban areas grow and As a result, the primary benefit of altering thermal
tackle issues associated with urban heat and climate inertia in pavements is likely reducing the extreme
change, moving towards auto independence has path­ magnitudes of outgoing heat by offsetting the release
ways to reducing urban heat, but more research and of energy to nighttime or generally cooler periods. For
strategic planning are necessary to ensure desirable example, this behavior could be beneficial in reducing
outcomes. the local heat severity during heat waves by increasing
For a typical day across metro Phoenix, pavements the pavement energy storage capacity and delaying heat
contribute nearly 25 times as much heat to the urban emissions until less severe periods. Overall, the potential
heat balance compared to vehicles, but in some cases for high thermal inertia properties in pavements should
such as during rush hour in a densely traveled corridor, be more deeply explored to flatten diurnal urban heat
vehicles can contribute nearly three times as much as fluxes and potentially mitigate impacts of increasingly
pavements to the local heat balance. When vehicle travel severe weather under climate change.
density is at its peak during rush hour, heat flux magni­ Planning strategies to reduce urban heat contributions
tudes can reach 132 W m–2 directly over the roadway, from pavements could include increasing pavement
while pavement can reach 143 W m–2. This indicates shading, reducing surface parking lots through parking
that during warmer months in hot climates, areas sur­ reform, or considering road diets. Previous research has
rounding high-trafficked roads may be increasingly established significant amounts of surface parking cover­
undesirable for outdoor travel or activities due to high age in many cities (Chester et al., 2015; Hoehne et al.,
amounts of added sensible heat from pavements and 2019), and reforming parking standards can have many
anthropogenic heating from dense vehicle travel. As benefits (Shoup, 2017). The contribution of parking pave­
a result, urban planning strategies to improve ment to urban heat is not insignificant, and cities, espe­
a community’s net thermal comfort during hot periods cially in sprawling hot climates like Phoenix, may
(especially late afternoons during the summer) should consider this another argument for reforming parking
be cognizant of these issues and should consider target­ standards and redesigning urban space to deemphasize
ing active transportation developments away from cor­ pavement coverage. Other options such as requiring
ridors with high pavement coverage and vehicle traffic. shade structures in parking lots or planting trees around
Many strategies to mitigate urban heat through pave­ travel corridors may also promote cooler urban environ­
ment design focus heavily on altering pavement albedo ments. However, as pavements elevate surrounding soil
(Li et al., 2013b; Santamouris et al., 2012), but this study temperatures, vaporation is accelerated, and nearby trees
indicates there may be potential to mitigate the severity consequently need more water than usual. This is espe­
of urban heat by increasing the thermal inertia of cially problematic for hot climates since this trend
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 15

worsens as temperatures rise, and Phoenix is already estimate. Despite this, vehicle heat emissions are often
constrained on water use. Another common suggestion significantly overshadowed by the added sensible heat
is to cover parking lots with solar panels for shade in from pavements. More accurate vehicle drive cycle model­
addition to their primary purpose of capturing solar ing paired with microtraffic simulations could be valuable
energy. However, a study of solar installations over to improve the accuracy of estimating vehicle’s contribu­
asphalt parking lots in Phoenix actually found they may tions to heat flux magnitudes at local scales.
warm the local environment (Pham et al., 2019). This
leads to pavement whitetopping and solar panel shading
5. Conclusion
being more context-specific solutions to alleviating pave­
ment-related heat, while reducing urban pavement cover­ An analysis was conducted to quantify contributions to
age is likely a more ubiquitous strategy. urban heat from vehicle travel and pavement infrastructure
in metropolitan Phoenix, Arizona to help understand how
city planning and the transportation sector can influence
4.1. Limitations
urban heat. Pavement infrastructure typically dominates
Some limitations of this analysis exist as a consequence of contributions to the urban heat balance relative to waste
the analysis region, selected methodologies, and available heat from vehicle travel both spatially and temporally in
data. Only sensible heat fluxes are estimated for clear and metro Phoenix. Relative to the natural ground, pavement
dry conditions in metro Phoenix. Some aspects of heat infrastructure contributes the most to the Phoenix urban
transfer between vehicles and pavements are not consid­ heat balance during summer afternoons. Vehicles may
ered. Vehicles traversing or parked over pavement will contribute significant amounts of heat but only in high
provide transient shading, blocking marginal amounts of travel corridors during rush hours. Urban densification
incoming solar radiation during the daytime. Some wasted could mitigate urban heat contributions from the transpor­
heat from vehicles may also affect the surface temperature tation sector by promoting less auto-dependent infrastruc­
of the pavement through friction and downward heat flux ture, mixed use, and higher density transit. To promote
from the bottom of the vehicle. Vehicles traveling over pedestrian thermal comfort, active transportation plans
a roadway will also induce convection at the roadway sur­ could separate active transit corridors from high-
face by generating airflow from their motion. Additionally, trafficked roadways and incorporate targets to reduce
the impact of vehicles shading the roadways is not consid­ nearby pavement coverage and traffic density. Altering
ered. We suspect these heat transfer mechanisms between pavement design to achieve high thermal inertia properties
roads and vehicles to be marginal; previous research has in pavements should be more deeply explored as a method
established that friction warming and convective cooling to mitigate impacts of increasingly severe daytime heat in
from vehicle motion to be insignificant to the overall pave­ urban areas. Future research should consider quantifying
ment surface temperature fluctuations (Fujimoto et al., added heat from pavement infrastructure in addition to
2015). While this analysis assesses vehicle heat generated anthropogenic heating for a more holistic understanding of
during vehicle operation, some vehicle heat will also be urban heat flux magnitudes.
emitted shortly after vehicle operation which is not con­
sidered. This additional amount of wasted heat from travel
will primarily at controlled intersections and parking des­ Acknowledgments
tinations, creating additional heat surrounding highly Funding for the research was made possible by a Dwight David
paved areas (e.g., shopping centers). In addition to weather, Eisenhower Transportation Fellowship (#693JJ31845020) and the
some pavement heat transfer model inputs differ by time following National Science Foundation awards: A Simulation
of day (albedo, SVF) affecting diurnal heat flux magnitudes. Platform to Enhance Infrastructure and Community Resilience
to Extreme Heat Events (#1635490); and Urban Resilience to
Waste heat from heavy vehicle traffic (e.g., freight, transit Extremes Sustainability Research Network (#1444755).
vehicles) is not assessed, likely leading to under estimation
magnitude of heat fluxes on and near highways. Limited
availability of travel data and established vehicle waste Disclosure statement
modeling methods made it difficult to assess how changing The authors declare no competing interests.
traffic patterns could spatiotemporally affect heating.
When considering city driving and increased prevalence
of braking, it is likely that more than 97% of fuel energy is Funding
converted to heat into calipers, tires, and the road surface This work was supported by the Federal Highway Administration
through friction. Therefore, heat emissions from vehicle [693JJ31845020]; National Science Foundation [1444755];
travel presented in this study are likely a conservative National Science Foundation [1635490].
16 C. G. HOEHNE ET AL.

Notes on contributors Andreou, E., & Axarli, K. (2012). Investigation of urban can­
yon microclimate in traditional and contemporary envir­
Christopher G. Hoehne, Ph.D is a research scientist focused onment. Experimental investigation and parametric
on understanding the energy and socioeconomic impacts of analysis. Renewable Energy, 43, 354–363. doi:10.1016/j.
urban transportation systems to embolden pathways towards renene.2011.11.038
sustainable, equitable, and productive mobility. Arnfield, A. J. (2003). Two decades of urban climate research:
Mikhail V. Chester, Ph.D. is Director of the Metis Center for A review of turbulence, exchanges of energy and water, and
Infrastructure and Sustainable Engineering at ASU and the urban heat island. International Journal of Climatology,
focuses on preparing infrastructure and their institutions for 23(1), 1–26. doi:10.1002/joc.859
the challenges of the coming century. Arnfield, A. J., & Grimmond, C. S. B. (1998). An urban canyon
energy budget model and its application to urban storage
David J. Sailor, Ph.D. is a Professor in Geographical Sciences heat flux modeling. Energy and Buildings, 27(1), 61–68.
and Urban Planning and Director of the Urban Climate doi:10.1016/S0378-7788(97)00026-1
Research Center at ASU who researches the intersection of Asaeda, T., Ca, V. T., & Wake, A. (1996). Heat storage of
climate with the built environment. pavement and its effect on the lower atmosphere.
David A. King, Ph.D. is an Assistant Professor in Atmospheric Environment, 30(3), 413–427. doi:10.1016/
Geographical Sciences and Urban Planning at ASU who 1352-2310(94)00140-5
focuses on intersections between transportation, land use, Baglione, M., Duty, M., & Pannone, G. (2007). April). Vehicle
and economics. System Energy Analysis Methodology and Tool for
Determining Vehicle Subsystem Energy Supply and
Demand. doi:10.4271/2007-01-0398
Bentz, D. P., & Turpin, R. (2007). Potential applications of
ORCID phase change materials in concrete technology. Cement and
Christopher G. Hoehne http://orcid.org/0000-0002-9904- Concrete Composites, 29(7), 527–532. doi:10.1016/j.
2256 cemconcomp.2007.04.007
Mikhail V. Chester http://orcid.org/0000-0002-9354-2102 Bondank, E. N., Chester, M. V., & Ruddell, B. L. (2018). Water
David J. Sailor http://orcid.org/0000-0003-1720-8214 Distribution System Failure Risks with Increasing
David A. King http://orcid.org/0000-0002-8401-6514 Temperatures [Research-article]. Environmental Science &
Technology, 52(17), 9605–9614. doi:10.1021/acs.est.7b01591
Burillo, D., Chester, M. V., Pincetl, S., Fournier, E. D., &
References Reyna, J. (2019). Forecasting peak electricity demand for
Los Angeles considering higher air temperatures due to
ADOT, (Arizona Department of Transportation). (2017a). climate change. Applied Energy, 23(6), 1–9.
ADOT’s use of rubberized asphalt gives new life to recycled (November2018). doi:10.1016/j.apenergy.2018.11.039
tires. Retrieved from March 26, 2019 https://www.azdot. Carnielo, E., & Zinzi, M. (2013). Optical and thermal char­
gov/mobile/media/news/2017/05/11/adot-s-use-of- acterisation of cool asphalts to mitigate urban temperatures
rubberized-asphalt-gives-new-life-to-recycled-tires and building cooling demand. Building and Environment,
ADOT, (Arizona Department of Transportation). (2017b). 60, 56–65. doi:10.1016/j.buildenv.2012.11.004
Pavement Design Manual. Retrieved from https://apps. CDC, (Centers for Disease Control and Prevention). (2012).
azdot.gov/files/materials-manuals/Preliminary- Number of Heat-Related Deaths, by Sex, United States,
Engineering-Design/PavementDesignManual.pdf 1999–2010. National Vital Statistics System, 61(36), 729.
ADOT, (Arizona Department of Transportation). (2019). Çengel, Y. (2002). Heat Transfer: A Practical Approach (2nd
Quiet Pavement Program. Retrieved from https://www. ed.). New York: Mcgraw-Hill.
azdot.gov/business/environmental-planning/programs/ Chester, M., Fraser, A., Matute, J., Flower, C., & Pendyala, R.
quiet-pavement-program (2015). Parking Infrastructure: A Constraint on or
Akbari, H., Rose, L. S., & Taha, H. (1999). Characterizing the Opportunity for Urban Redevelopment? A Study of Los
fabric of the urban environment: A case study of Angeles County Parking Supply and Growth. Journal of
Sacramento, California. Lawrence Berkeley National the American Planning Association, 81(4), 268–286.
Laboratory. Berkeley, CA. doi:10.2172/764362 doi:10.1080/01944363.2015.1092879
Akbari, H., Shea Rose, L., & Taha, H. (2003). Analyzing the Chiasson, A. D., Spitler, J. D., Rees, S. J., & Smith, M. D. (2000).
land cover of an urban environment using high-resolution A Model for Simulating the Performance of a Pavement
orthophotos. Landscape and Urban Planning, 63(1), 1–14. Heating System as a Supplemental Heat Rejecter with
doi:10.1016/S0169-2046(02)00165-2 Closed-Loop Ground-Source Heat Pump Systems. Journal of
Albertson, J. D., Parlange, M. B., Katul, G. G., Chu, C. R., Solar Energy Engineering, 122(4), 183. doi:10.1115/1.1330725
Stricker, H., & Tyler, S. (1995). Sensible Heat Flux From Clark, S. S., Chester, M. V., Seager, T. P., &
Arid Regions: A Simple FluxVariance Method. Water Eisenberg, D. A. (2019). The vulnerability of interde­
Resources Research, 31(4), 969–973. doi:10.1029/ pendent urban infrastructure systems to climate change:
94WR02978 Could Phoenix experience a Katrina of extreme heat?
Allen, L., Lindberg, F., & Grimmond, C. S. B. (2011). Global to Sustainable and Resilient Infrastructure, 4(1), 21–35.
city scale urban anthropogenic heat flux: Model and doi:10.1080/23789689.2018.1448668
variability. International Journal of Climatology, 31(13), Corripio, J. G. (2019). Insol: Solar Radiation. Retrieved from
1990–2005. doi:10.1002/joc.2210 https://meteoexploration.com/R/insol/index.html
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 17

Dai, Z., Guldmann, J. M., & Hu, Y. (2018). Spatial regression of Pavement Materials on the Evolution of Temperature
models of park and land-use impacts on the urban heat Depth Profiles in Different Climatic Regions. Journal of
island in central Beijing. Science of the Total Environment, Materials in Civil Engineering, 24(1), 32–47. doi:10.1061/
626, 1136–1147. doi:10.1016/j.scitotenv.2018.01.165 (ASCE)MT.1943-5533.0000357
Davis, S. C., & Boundy, R. G. (2019). Transportation Energy Data Hart, M. A., & Sailor, D. J. (2009). Quantifying the influence of
Book: Edition 37. Retrieved from http://cta.ornl.gov/data land-use and surface characteristics on spatial variability in
Dempsey, B. J., & Thompson, M. R. (1970). A Heat Transfer the urban heat island. Theoretical and Applied Climatology,
Model for Evaluating Frost Action and 95(3–4), 397–406. doi:10.1007/s00704-008-0017-5
Temperature-Related Effects in Multilayered Pavement Heath, M. T. (2002). Scientific computing: An introductory
Systems. Highway Research Record, (342) Retrieved from survey (2nd ed.). Boston, MA: McGraw-Hill.
https://trid.trb.org/view/102331 Hermansson, Å. (2004). Mathematical model for paved
Dobos, E. (2011). Encyclopedia of Soil Science, 2nd. surface summer and winter temperature: Comparison
doi:10.1081/E-ESS of calculated and measured temperatures. Cold Regions
DOE, (U.S. Department of Energy). (2019). Where the Energy Science and Technology, 40(1–2), 1–17. doi:10.1016/j.
Goes: Gasoline Vehicles. Retrieved from March 28, 2019 coldregions.2004.03.001
https://www.fueleconomy.gov/feg/atv.shtml Heusinkveld, B. G., Jacobs, A. F. G., Holtslag, A. A. M., &
Eisenman, D. P., Wilhalme, H., Tseng, C., Chester, M., Berkowicz, S. M. (2004). Surface energy balance closure
English, P., Pincetl, S., . . . Dhaliwal, S. K. (2016). Heat Death in an arid region: Role of soil heat flux. Agricultural and
Associations with the built environment, social vulnerability Forest Meteorology, 122(1–2), 21–37. doi:10.1016/j.
and their interactions with rising temperature. Health & Place, agrformet.2003.09.005
41, 89–99. doi:10.1016/j.healthplace.2016.08.007 Hoehne, C. G. (2019). Heat Transfer Modeling for Phoenix
EPA, (Environmental Protection Agency). (2016). Scrap Tires. Urban Transportation-Heat Analysis. Retrieved from
Retrieved from March 26, 2019https://archive.epa.gov/epa https://github.com/cghoehne/transport-uhi-phx
waste/conserve/materials/tires/web/html/live.html Hoehne, C. G., Chester, M. V., Fraser, A. M., &
EPA, (Environmental Protection Agency). (2017). King, D. A. (2019). Valley of the sun-drenched parking
Automotive Trends Report 2018. EPA-420-R-19-002. space: The growth, extent, and implications of parking
Retrieved from https://www.epa.gov/automotive-trends infrastructure in Phoenix. Cities, 89, 186–198.
/download-data-automotive-trends-report doi:10.1016/j.cities.2019.02.007
Erell, E., Pearlmutter, D., Boneh, D., & Kutiel, P. B. (2014). Holland, R. (2014). Estimating the Number of Parking Spaces
Effect of high-albedo materials on pedestrian heat stress in Per Acre. Center for Profitable Agriculture, Institute of
urban street canyons. Urban Climate, 10(P2), 367–386. Agriculture, University of Tennessee website: https://ag.
doi:10.1016/j.uclim.2013.10.005 tennessee.edu/cpa/InformationSheets/CPA222.pdf
FAA, (Federal Aviation Administration). (2016). Airport Hsiao, Y. Y., Chang, W. C., & Chen, S. L. (2010).
Pavement Design and Evaluation. U.S. Department of A mathematic model of thermoelectric module with appli­
Transportation, Washington D.C. doi:org/150/5320-6F cations on waste heat recovery from automobile engine.
Fujimoto, A., Saida, A., Fukuhara, T., & Futagami, T. (2015). Heat Energy, 35(3), 1447–1454. doi:10.1016/j.energy.2009.11.030
Transfer Analysis on Road Surface. Retrieved from https:// Hu, L., Li, Y., Zou, X., Du, S., Liu, Z., & Huang, H. (2017).
www.researchgate.net/publication/268351732%0AHEAT Temperature Characteristics of Porous Portland Cement
Gasparrini, A., Guo, Y., Hashizume, M., Lavigne, E., Concrete during the Hot Summer Session. Advances in
Zanobetti, A., Schwartz, J., . . . Armstrong, B. (2015). Materials Science and Engineering. doi:10.1155/2017/2058034
Mortality risk attributable to high and low ambient tempera­ Ichinose, T., Shimodozono, K., & Hanaki, K. (1999). Impact of
ture: A multicountry observational study. The Lancet, 386 anthropogenic heat on urban climate in Tokyo. Atmospheric
(9991), 369–375. doi:10.1016/S0140-6736(14)62114-0 Environment, 33(24–25), 3897–3909. doi:10.1016/S1352-
Gasser, P., Lustenberger, P., Cinelli, M., Kim, W., 2310(99)00132-6
Spada, M., Burgherr, P., . . . Sun, T. Y. (2019). Im, J. S., Dessouky, S., Bai, B. C., Vo, H. V., & Park, D.-W.
A review on resilience assessment of energy systems. (2015). Thermal Properties of Asphalt Mixtures Modified
Sustainable and Resilient Infrastructure, 00(00), 1–27. with Conductive Fillers. Journal of Nanomaterials, (2015,
doi:10.1080/23789689.2019.1610600 1–6. doi:10.1155/2015/926809
Golden, J. S., & Kaloush, K. E. (2006). Mesoscale and micro­ Johansson, E. (2006). Influence of urban geometry on outdoor
scale evaluation of surface pavement impacts on the urban thermal comfort in a hot dry climate: A study in Fez,
heat island effects. International Journal of Pavement Morocco. Building and Environment, 41(10), 1326–1338.
Engineering, 7(1), 37–52. doi:10.1080/10298430500505325 doi:10.1016/j.buildenv.2005.05.022
Graff Zivin, J., & Neidell, M. (2014). Temperature and the Karner, A., Hondula, D. M., & Vanos, J. K. (2015). Heat exposure
Allocation of Time: Implications for Climate Change. during non-motorized travel: Implications for transportation
Journal of Labor Economics, 32(1), 1–26. doi:10.1086/671766 policy under climate change. Journal of Transport & Health, 2
Gui, J., Phelan, P. E., Kaloush, K. E., & Golden, J. S. (4), 451–459. doi:10.1016/j.jth.2015.10.001
(2007). Impact of Pavement Thermophysical Properties Kenworthy, J. R., & Laube, F. B. (1999). Patterns of automo­
on Surface Temperatures. Journal of Materials in Civil bile dependence in cities: An international overview of key
Engineering, 19(8), 683–690. doi:10.1061/(ASCE)0899- physical and economic dimensions with some implications
1561(2007)19:8(683) for urban policy. Transportation Research Part A: Policy
Hall, M. R., Dehdezi, P. K., Dawson, A. R., Grenfell, J., & and Practice, 33(7–8), 691–723. doi:10.1016/S0965-
Isola, R. (2012). Influence of the Thermophysical Properties 8564(99)00006-3
18 C. G. HOEHNE ET AL.

Kim, M., Lee, K., & Cho, G. H. (2017). Temporal and spatial Minhoto, M. J. C., Pais, J. C., & Pereira, P. A. (2006).
variability of urban heat island by geographical location: Asphalt Pavement Temperature Prediction. Asphalt
A case study of Ulsan, Korea. Building and Environment, Rubber 2006: Proceedings, Los Angeles, CA, 193–207.
126(August), 471–482. doi:10.1016/j.buildenv.2017.10.023 10.1061/9780784413005.029
Kim, S., Park, S., Kim, S., & Rhi, S.-H. (2011). A Thermoelectric Moise, A. F., & Aynsley, R. (1999). Ambient ultraviolet
Generator Using Engine Coolant for Light-Duty Internal radiation levels in public shade settings. International
Combustion Engine-Powered Vehicles. Journal of Journal of Biometeorology, 43(3), 128–138. doi:10.1007/
Electronic Materials, 40(5), 812–816. doi:10.1007/s11664- s004840050128
011-1580-6 Monteith, J. L., & Unsworth, M. H. (2013). Principles of
Kjellstrom, T., Holmer, I., & Lemke, B. (2009). Workplace heat Environmental Physics: Plants, animals, and the atmosphere
stress, health and productivity – An increasing challenge for (4th ed.). Amsterdam: Academic Press.
low and middle-income countries during climate change. NASA, (National Aeronautics and Space Administration).
Global Health Action, 2(1), 2047. doi:10.3402/gha.v2i0.2047 (2019). ASTER. Retrieved from March 14, 2019 https://
Kovats, R. S., & Hajat, S. (2008). Heat Stress and Public Health: terra.nasa.gov/about/terra-instruments/aster
A Critical Review. Annual Review of Public Health, 29(1), Nobel, P. S., & Geller, G. N. (1987). Temperature Modelling of
41–55. doi:10.1146/annurev.publhealth.29.020907.090843 Wet and Dry Desert Soils. The Journal of Ecology, 75(1),
Krayenhoff, E. S., & Voogt, J. A. (2010). Impacts of urban 247. doi:10.2307/2260549
albedo increase on local air temperature at daily-annual Obradovich, N., & Fowler, J. H. (2017). Climate change may
time scales: Model results and synthesis of previous work. alter human physical activity patterns. Nature Human
Journal of Applied Meteorology and Climatology, 49(8), Behaviour, 1(April), 0097. doi:10.1038/s41562-017-0097
1634–1648. doi:10.1175/2010JAMC2356.1 Oke, T. R. (1982). The energetic basis of the urban heat island.
Li, H., Harvey, J., & Kendall, A. (2013a). Field measurement of Quarterly Journal of the Royal Meteorological Society, 108
albedo for different land cover materials and effects on (455), 1–24. doi:10.1002/qj.49710845502
thermal performance. Building and Environment, 59, OpenStreetMap contributors. (2019). Planet dump Retrieved
536–546. doi:10.1016/j.buildenv.2012.10.014 from https://www.openstreetmap.org/
Li, H., Harvey, J. T., Holland, T. J., & Kayhanian, M. (2013b). Orr, B., Akbarzadeh, A., Mochizuki, M., & Singh, R. (2016).
The use of reflective and permeable pavements as A review of car waste heat recovery systems utilising
a potential practice for heat island mitigation and storm­ thermoelectric generators and heat pipes. Applied
water management. Environmental Research Letters, 8(1), Thermal Engineering, 101, 490–495. doi:10.1016/j.
015023. doi:10.1088/1748-9326/8/1/015023 applthermaleng.2015.10.081
Li, H., He, Y., & Harvey, J. (2016). Human Thermal Comfort: Pham, J. V., Baniassadi, A., Brown, K. E., Heusinger, J., &
Modeling the Impact of Different Cool Pavement Strategies. Sailor, D. J. (2019). Comparing photovoltaic and reflective
Transportation Research Record: Journal of the Transportation shade surfaces in the urban environment: Effects on surface
Research Board, 2575(1), 92–102. doi:10.3141/2575-10 sensible heat flux and pedestrian thermal comfort. Urban
Luca, J., & Mrawira, D. (2005). New Measurement of Thermal Climate, 29, 100500. doi:10.1016/j.uclim.2019.100500
Properties of Superpave Asphalt Concrete. Journal of Qin, Y. (2015). Urban canyon albedo and its implication on
Materials in Civil Engineering, 17(1), 72–79. doi:10.1061/ the use of reflective cool pavements. Energy and Buildings,
(ASCE)0899-1561(2005)17:1(72) 96, 86–94. doi:10.1016/j.enbuild.2015.03.005
Manville, M., & Shoup, D. (2005). Parking, People, and Cities. Rafi, M. M., & Aziz, T. (2019). Experimental testing of fly ash
Journal of Urban Planning and Development, 131(4), containing recycled aggregate concrete exposed to high
233–245. doi:10.1061/(ASCE)0733-9488(2005)131:4(233) temperatures. Sustainable and Resilient Infrastructure, 00
Markolf, S. A., Hoehne, C. G., Fraser, A., & Chester, M. (00), 1–16. doi:10.1080/23789689.2019.1681820
(2017). Maintaining reliability of transportation systems Rajoo, S., Romagnoli, A., Martinez-Botas, R., Pesiridis, A.,
and interconnected infrastructure under climate change. Copeland, C., & Bin Mamat, A. M. I. (2014). Automotive
International Conference on Sustainable Infrastructure exhaust power and waste heat recovery technologies.
2017: Methodology - Proceedings of the International Automotive Exhaust Emissions and Energy Recovery (pp.
Conference on Sustainable Infrastructure 2017, New York, 265–281). Hauppauge, NY: NOVA Science Publishers.
NY. 10.1061/9780784481196.020 doi:10.13140/2.1.4809.0565
Middel, A., Häb, K., Brazel, A. J., Martin, C. A., & Reddy, G. N., Venkatesan, V., & Maniyar, U. (2015).
Guhathakurta, S. (2014). Impact of urban form and design Estimation of harvestable energy from vehicle waste heat.
on mid-afternoon microclimate in Phoenix Local Climate 2015 International Conference on Renewable Energy
Zones. Landscape and Urban Planning, 122, 16–28. Research and Applications (ICRERA), Palermo, 618–625.
doi:10.1016/j.landurbplan.2013.11.004 https://doi.10.1109/ICRERA.2015.7418487
Middel, A., Lukasczyk, J., Maciejewski, R., Demuzere, M., & Reyna, J. L., & Chester, M. V. (2017). Energy efficiency to
Roth, M. (2018). Sky View Factor footprints for urban reduce residential electricity and natural gas use under
climate modeling. Urban Climate, 25(April), 120–134. climate change. Nature Communications, 8(1), 14916.
doi:10.1016/j.uclim.2018.05.004 doi:10.1038/ncomms14916
Miller, N. L., Hayhoe, K., Jin, J., & Auffhammer, M. (2008). Sailor, D. J., & Lu, L. (2004). A top–down methodology for
Climate, Extreme Heat, and Electricity Demand in developing diurnal and seasonal anthropogenic heating
California. Journal of Applied Meteorology and Climatology, profiles for urban areas. Atmospheric Environment, 38(17),
47(6), 1834–1844. doi:10.1175/2007JAMC1480.1 2737–2748. doi:10.1016/j.atmosenv.2004.01.034
SUSTAINABLE AND RESILIENT INFRASTRUCTURE 19

Santamouris, M., Gaitani, N., Spanou, A., Saliari, M., heat mitigation strategies on pedestrian thermal comfort in
Giannopoulou, K., Vasilakopoulou, K., & Kardomateas, T. a Los Angeles neighborhood. Environmental Research
(2012). Using cool paving materials to improve microcli­ Letters, 11(2), 2. doi:10.1088/1748-9326/11/2/024003
mate of urban areas – Design realization and results of the Tan, S. A., & Fwa, T. F. (1992). Influence of pavement materi­
flisvos project. Building and Environment, 53, 128–136. als on the thermal environment of outdoor spaces. Building
doi:10.1016/j.buildenv.2012.01.022 and Environment, 27(3), 289–295. doi:10.1016/0360-
Schaeffer, R., Salem, A., Frossard, A., Lucena De, P., Soares, B., 1323(92)90030-S
Cesar, M., . . . Sadeck, M. (2012). Energy sector vulnerability University of Utah. (2019). MesoWest Data. Retrieved from
to climate change: A review. Energy, 38(1), 1–12. March 14, 2019 https://mesowest.utah.edu/
doi:10.1016/j.energy.2011.11.056 USACE, (United States Army Corps of Engineers). (2018).
Shoup, D. (2017). The high cost of free parking: Updated Unified Facilities Criteria. Washington D.C: U.S.
edition. New York, NY: Routledge. Department of the Army. doi:org/UFC3-201-01
Smith, C., Lindley, S., & Levermore, G. (2009). Estimating USGS, (United States Geological Survey). (2018). On
spatial and temporal patterns of urban anthropogenic Demand Surface Kinetic Temperature. Washington D.C:
heat fluxes for UK cities: The case of Manchester. U.S. Department of the Interior. doi:10.5067/ASTER/
Theoretical and Applied Climatology, 98(1–2), 19–35. AST_08.003
doi:10.1007/s00704-008-0086-5 Wallington, T. J., Kaiser, E. W., & Farrell, J. T. (2006).
Stamatakis, E., Nnoaham, K., Foster, C., & Scarborough, P. Automotive fuels and internal combustion engines:
(2013). The influence of global heating on discretionary A chemical perspective. Chemical Society Reviews, 35(4),
physical activity: An important and overlooked conse­ 335. doi:10.1039/b410469m
quence of climate change. Journal of Physical Activity & Wang, D., & Roesler, J. R. (2012). One-Dimensional
Health, 10(6), 765–768. 10.1123/jpah.10.6.765 Rigid Pavement Temperature Prediction Using
Stone, B., Hess, J. J., & Frumkin, H. (2010). Urban form and Laplace Transformation. Journal of Transportation
extreme heat events: Are sprawling cities more vulnerable to Engineering, 138(9), 1171–1177. doi:10.1061/(ASCE)
climate change than compact cities? Environmental Health TE.1943-5436.000041
Perspectives, 118(10), 1425–1428. doi:10.1289/ehp.0901879 Wang, Y., Berardi, U., & Akbari, H. (2016). Comparing the
Stone, B., & Rodgers, M. O. (2001). Urban Form and Thermal effects of urban heat island mitigation strategies for
Efficiency: How the Design of Cities Influences the Urban Toronto, Canada. Energy and Buildings, 114, 2–19.
Heat Island Effect. Journal of the American Planning doi:10.1016/j.enbuild.2015.06.046
Association, 67(2), 186–198. doi:10.1080/01944360108976228 Wicki, A., Parlow, E., & Feigenwinter, C. (2018). Evaluation
Sun, R., Xie, W., & Chen, L. (2018). A landscape connectivity and Modeling of Urban Heat Island Intensity in Basel,
model to quantify contributions of heat sources and sinks in Switzerland. Climate, 6(3), 55. doi:10.3390/cli6030055
urban regions. Landscape and Urban Planning, 178, Yang, J., & Stabler, F. R. (2009). Automotive Applications of
(September2017) 43–50. 10.1016/j.landurbplan.2018.05.015 Thermoelectric Materials. Journal of Electronic Materials,
Taha, H. (1997). Urban climates and heat islands: Albedo, 38(7), 1245–1251. doi:10.1007/s11664-009-0680-z
evapotranspiration, and anthropogenic heat. Energy and Yun, T. S., Jeong, Y. J., & Youm, K.-S. (2014). Effect of
Buildings, 25(2), 99–103. doi:10.1016/S0378-7788(96) Surrogate Aggregates on the Thermal Conductivity of
00999-1 Concrete at Ambient and Elevated Temperatures. The
Taleghani, M., Sailor, D., & Ban-Weiss, G. A. (2016). Scientific World Journal, (2014, 1–9. doi:10.1155/2014/
Micrometeorological simulations to predict the impacts of 939632

You might also like