You are on page 1of 63

Journal Pre-proof

Towards a performance-based approach for multifunctional green roofs: An


interdisciplinary review

Lauren M. Cook, Tove A. Larsen

PII: S0360-1323(20)30856-8
DOI: https://doi.org/10.1016/j.buildenv.2020.107489
Reference: BAE 107489

To appear in: Building and Environment

Received Date: 17 September 2020


Revised Date: 23 November 2020
Accepted Date: 26 November 2020

Please cite this article as: Cook LM, Larsen TA, Towards a performance-based approach for
multifunctional green roofs: An interdisciplinary review, Building and Environment (2021), doi: https://
doi.org/10.1016/j.buildenv.2020.107489.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


1 Towards a performance-based approach for multifunctional green roofs: an

2 interdisciplinary review

3 Lauren M. Cook1*, Tove A. Larsen1

1
4 Department of Urban Water Management, Swiss Federal Institute for Aquatic Research,

5 Dübendorf, Switzerland

6 *Corresponding author. Postal address: Eawag, Überland Str. 133, 8600 Dübendorf, Switzerland

of
7 E-mail address: Lauren.cook@eawag.ch

ro
8

9 ABSTRACT:
-p
re
lP

10 Green roofs have the potential to offer numerous ecosystem services; however, they are rarely

11 designed to achieve them. Instead, design is restricted by perceived structural and maintenance
na

12 constraints, which consequently diminish the achievable benefits. For green roofs to improve
ur

13 sustainability and resilience of cities, their design should match their promised multi-functional
Jo

14 application using performance-based design. The first step towards a comprehensive performance

15 model is to synthesize design recommendations across disciplines to identify synergies and trade-

16 offs in design objectives for multiple benefits. This study discusses design strategies that could

17 alter the energy and water balance in the green roof in order to attenuate urban stormwater,

18 increase building energy performance, mitigate urban heat, and improve the output of solar

19 panels placed on top of green roofs. These benefits are mathematically linked to quantifiable

20 processes (discharge rate, water content, evapotranspiration, sensible heat, net radiation,

21 insulation, and thermal mass), forming the foundation for a performance-based design model.

22 Design recommendations are then summarized for each process, followed by a discussion of

1
23 synergies, trade-offs, and research needs that arise when green roofs are designed to achieve

24 multiple functions. Selecting vegetation with high leaf area and albedo improves multiple

25 benefits without affecting structural constraints, whereas choosing plants with low stomatal

26 resistance leads to trade-offs between higher evapotranspiration and higher irrigation

27 requirements. Trade-offs in substrate depth and properties including organic matter and moisture

28 are also apparent. Interdisciplinary collaborations are needed to simulate and optimize design

29 parameters based on stakeholder preferences related to co-benefits and constraints.

of
30

ro
31 -p
KEYWORDS: green roofs, co-benefits, design, performance objectives, interdisciplinary review
re
32
lP

33
na
ur
Jo

2
34 Nomenclature

TERM DESCRIPTION UNITS


As Surface area of green roof m2
AMC Antecedent moisture content %
Albedo or reflectivity of the surface -
cp Specific heat of air kJ kg-1 K-1
Heat capacity of substance J m-3 K-1
Psychrometric constant kPa K-1
dz Incremental depth between two points m
dt Incremental change in time (or Δ as delta) s; min

of
dT Incremental change in temperature K

ro
Vapor pressure of the air kPa
Vapor pressure at the surface kPa
Ea
ET
Energy produced per time period -p
Evapotranspiration rate as a depth / volume per unit area
kWh
m s-1
re
Emissivity of the surface -
Hp Peak sensible heat flux W m-2
lP

η Fractional vegetation coverage -


VMC at wilting point %
na

K Unsaturated hydraulic conductivity of substrate m s-1


Ks Hydraulic conductivity at saturation m s-1
ur

LAI Leaf area index -


LWin/out Incoming/outgoing long-wave radiation W m-2
Jo

Thermal conductivity of substrate W m-1 K-1


Λ Latent heat of vaporization J kg-1
p Rate of incoming precipitation across green roof area m3 s-1
Peak power under standard conditions kW
Mean air density at constant pressure kg m-3
qd Rate of water discharged below the green roof m3 s-1
qr Rate of surface runoff from top of green roof m3 s-1
Qe Latent heat flux W m-2
QH Sensible heat flux entering air as convection W m-2
QG Soil heat flux as conduction W m-2
(Δ ) Heat/thermal storage (change over time) J m-2 (W m-2)
rav Aerodynamic resistance to vapor transfer s m-1
rah Aerodynamic resistance to heat transfer s m-1

3
rI Stomatal resistance s m-1
rs Surface resistance s m-1
Rin/out Incoming/outgoing radiation W m-2
Rn Net radiation flux W m-2
RET Relative amount of water retained -
S (ΔS) Water storage (change over time) m3 (m3 s-1)
Smax Maximum retention capacity %
SWin/out Incoming/outgoing short-wave radiation W m-2
Stefan-Boltzmann constant (5.67 x 10-8) W m-2 K-4
Lag time s; min

of
Time at start of precipitation or irrigation event s; min

!" Time at start of discharge s; min

ro
Ta Ambient air temperature K
Ts
#
Surface temperature
Transmittance of vegetation canopy
-p K
-
re
υ wind speed m s-1
Vr Volume water retained over time m3
lP

Vin Volume incoming water over time m3


VMC Volumetric moisture content (or as soil moisture) %
na

VPD Vapor pressure deficit KPa


WHCmax Maximum water holding capacity (or $% as VMC at field capacity) %
ur

Y Specific energy yield kWh kW-1


&' Substrate matric potential Pa
Jo

z Depth of green roof substrate layer m


35

4
36 1. Introduction

37 From urban heat island effect to pluvial flooding, cities are faced with a growing list of

38 environmental challenges. This list includes addressing a considerable energy demand that must

39 eventually be met using sustainable technologies, like rooftop solar panels, as well as adapting to

40 the effects of climate change, such as more intense rainfall, drought, and extreme heat [1,2]. Due

41 to the scale and urgency of these problems, innovative and integrated solutions that can resolve

of
42 multiple urbanization challenges at once are required [3].

ro
43 Green roofs, which are engineered rooftops that sustain vegetation, have been suggested as

44 -p
one such “nature based” solution [4] to improve urban sustainability and resilience. They can
re
45 provide multiple ecosystem services [5–8], including reducing urban stormwater runoff [9],

46 decreasing the temperature of cities [10], diminishing the energy consumption of buildings [11],
lP

47 and more recently, increasing the efficiency of rooftop solar panels due to their cooling effect
na

48 [12]. Green roofs also have the capacity to improve biodiversity [13,14], sequester CO2 [15],
ur

49 reduce air pollution [16,17], and dampen noise [18,19]. These co-benefits are relevant decision
Jo

50 factors that could increase the potential to disseminate green roofs throughout cities [20].

51 Despite the numerous co-benefits attributed to green roofs, they are rarely designed to

52 achieve them [5]. Existing green roof design guidelines [5,21] favor design properties (e.g.,

53 vegetation and substrate characteristics) that would minimize static loading, maintenance, and

54 cost [21] and rarely consider whether these recommendations would lead to the promised co-

55 benefits. As a result, most green roofs are extensive (soil depth less than 20 cm) [22], not

56 irrigated, and intended for aesthetic appeal. Substrate depths can even be restricted to a few

57 centimeters and plant species are then selected for their ability to survive in shallow depths with

58 low water and nutrient availability [23–27]. Sedums, a type of succulent plant, are the most

5
59 common green roof vegetation for these reasons [26]; however, they have been shown to retain

60 nearly the same amount of runoff as bare soil [28,29] and cool rooftops less than other species

61 [30].

62 In order for green roofs to systematically improve urban sustainability, many recognize that

63 the design process should evaluate environmental performance objectives [25,31,32] alongside

64 structural and maintenance constraints. Several studies have reviewed modeling and field

of
65 experiments that assess individual performance objectives at the building scale (e.g., stormwater

66 attenuation [9,33,34], heat mitigation and energy savings [6,10,11,35–37], and pollution and

ro
67 climate change [38]) or the processes (e.g., evapotranspiration [39]) and the design properties

68
-p
(e.g., materials [40] and vegetation [41,42]) that influence these objectives. However, few studies
re
69 have reviewed multiple green roof co-benefits [5,43], and these studies did not consider trade-
lP

70 offs or preferences for design properties that maximize these co-benefits. Further efforts are still
na

71 needed to continue progress towards performance-based design of green roofs, which ensures

72 design properties are evaluated holistically with respect to multiple performance objectives.
ur
Jo

73 The present work advances this goal by providing the foundation for a multi-objective,

74 performance-based design model for green roofs. This is accomplished first by associating co-

75 benefits to specific performance objectives and to the physical processes that influence them

76 (Section 2). Second, the mathematical formulation is presented that links these physical

77 processes to each other and to the green roof design properties (Section 3). Finally, the literature

78 is summarized to evaluate how changes in design properties would alter different performance

79 objectives (Section 4) and to identify synergies, trade-offs, and future research needs (Section 5)

80 that arise when green roofs are designed to achieve multiple functions. As a first step, this review

81 is limited primarily to roof scale studies of extensive green roofs that evaluate the effect of design

6
82 parameters on four co-benefits: stormwater runoff attenuation, building thermal performance,

83 urban heat mitigation, and rooftop solar panel output improvement.

84 2. Background

85 2.1. Current green roof design characteristics

86 Green roofs are primarily composed of three layers: vegetation, engineered media (substrate),

87 and a waterproof membrane (impermeable liner) that prevents water from entering the building.

of
88 Between the impermeable liner and media, some roofs also include a root barrier, which prevents

ro
89 roots from piercing the liner; a filter layer, that separates media particles; and/or a water retention

90
-p
layer, which retains additional water available for plant uptake [44–46]. Water drains from the
re
91 bottom layer along the roof slope (typically less than 1%) into roof gutters. Figure 1 presents a
lP

92 schematic of typical extensive green roof layers and design characteristics, discussed in the
na

93 following paragraphs.
ur

Green roof Vertical Typical


layer dimensions properties
Jo

Shallow roots; low water,


nutrient & maintenance
Vegetation 3 – 25 cm
requirements
Mixed composition; low
Substrate 7 – 15 cm density; low nutrients
Water Higher retention than
retention layer 2 – 5 cm substrate; low density
Physical or chemical
Root barrier < 2 cm prevention of root growth
Concrete slab Filter layer < 3 cm High tensile strength
Impermeable < 1 cm Waterproof
Drain liner
94

95 Figure 1. Schematic and typical properties of common extensive green roof layers

96 Green roof media and vegetation properties differ from similar systems on the ground. Media

97 composition is engineered to contain a majority of porous aggregate, less than 25% organic

7
98 material, and a small percentage of fines, lightweight additives, and minerals [46]. The bulk (dry)

99 density of the media, ranging from 0.34 to 1.31 g cm-3 [25,28,47,48], is thus low compared to

100 natural soils [49]. Vegetation that survives in thin substrates with limited nutrients and water

101 availability are typically used [23–27], including those that are drought and heat tolerant, with

102 low biomass, minimal height (less than 20 cm), and short roots [5,26]. Green roofs are intended

103 to be fully covered in vegetation (a fractional vegetation coverage (η) of 100%); however, in

104 reality, 80% coverage is considered high because roofs are rarely irrigated or fertilized to increase

of
105 coverage. These design properties are ideal to minimize structural loading and maintenance;

ro
106 however, they could also limit the benefits achievable by the green roof, as discussed in Section

107 4.
-p
re
lP

108 2.2. Description of co-benefits and associated physical processes

109 Before assessing the influence of the design properties on the achievable benefits,
na

110 quantifiable performance metrics must be designated that relate these benefits to the physical
ur

111 processes and properties governing them. This section describes the four co-benefits discussed in
Jo

112 this study (stormwater attenuation, heat mitigation, solar panel output improvement, and building

113 thermal performance), which metrics are useful to characterize them, and how they relate to the

114 water and energy balance. Defined in Table 1 and mathematically formulated in Section 3, the

115 relevant processes and properties include: discharge rate, volumetric water content,

116 evapotranspiration, sensible heat flux, net radiation flux, insulation, and thermal mass.

117 Table 1. Definitions of important physical processes and properties related to green roofs

Associated
Term Symbol Definition equations

Discharge rate qd Rate of water flowing from the bottom the green roof Eqn. 5

8
Percentage of water stored relative to the total volume of
Volumetric water VMC or Eqn. A1,
the wet material (water, air, substrate); the maximum is
content referred to as the maximum water holding capacity Eqn. 6
(WHCmax) or the water content at field capacity (θfc)
Rate of water evaporating from soil and transpiring from Eqn. 7, Eqn.
Evapotranspiration ET
plants A6
Energy transferred into the atmosphere as heat that can be
"sensed;" although several forms of heat can by sensed, in
Sensible heat flux QH Eqn. 8
this study, sensible heat refers only to convection (see
Section 3.3)
Difference between incoming and outgoing short-wave Eqn. 9
Net radiation flux Rn (SW) and long-wave (LW) radiation. SW radiation is solar
radiation, while LW is infrared. Eqn. 10

of
Physical material that slows the transfer of heat into and
Insulation R-value Eqn. 12
out of the building.

ro
Thermal mass Cs The ability of a material to absorb and store heat Eqn. 11

118
-p
re
119 Discussed throughout the following subsections, Table 2 summarizes which processes
lP

120 and properties influence the different co-benefits (and their evaluation metrics).
na

121 Table 2. Important processes and properties (rows) and co-benefits that they influence (columns). The [+/-] symbol refers to the
122 direction of change needed in the metric for improvement (e.g., a decrease [-] in the conductive heat flux would lead to higher
123 energy performance). Markers represent seasonality (●all year, ◌warm months only)
ur

124
125
Jo

Benefits → Stormwater Urban Solar panel Building


attenuation heat output thermal
mitigation improvement performance
Detention Retention
Evaluation metric → Lag time Relative Peak Specific yield Conductive
(tlag) [+] retention sensible (Y) [+] heat flux (QG)
(RET) heat flux [-]
Processes ↓ [+] (Hp) [-]
Discharge rate (qd) ● ●
Soil water content ● ●
(VMC)
Evapotranspiration (ET) ● ● ◌ ◌
Sensible heat (QH) ● ◌

9
Net radiation flux (Rn) ● ● ◌
Insulation (R-value) and ●
thermal mass (Cs)
126

127 2.2.1. Stormwater attenuation and retention

128 Urban stormwater is rainwater (and melted snow) that collects on impervious surfaces and

129 usually runs off into collection systems. When the flowrate of runoff exceeds the inflow capacity

of
130 of the collection system, water can back up into streets and cause localized flooding or, in cities

ro
131 with combined sewers, trigger combined sewer overflows (CSOs) of untreated wastewater into

132 surface waters [50].


-p
re
133 Green roofs can diminish these problems by attenuating the magnitude and timing of the
lP

134 maximum runoff flowrate per storm event (referred to as peak discharge). This is accomplished
na

135 through temporary stormwater detention in soil or on leaves [22,51], which is influenced by the

136 discharge rate (qd) from the engineered media and the volumetric water content (*+ ) of the
ur

137 media (Table 2). Detention can be evaluated with metrics related to changes in magnitude or
Jo

138 timing of discharge [51], including the lag time (tlag). Usually calculated in minutes [52], lag

139 time is the time between the start of a rainstorm ( ) and the start of discharge ( !" ) from the

140 bottom layer (Eqn. 1). Detention is improved with a longer lag time or slower discharge.

141 = !" − (1)

142 In addition to peak attenuation through detention, green roofs can also reduce the total

143 runoff entering the sewer [9]. Volume reductions are sometimes used to calculate water quality

144 targets [53,54]; however, attenuation is the more important stormwater service because it

145 diminishes flooding and CSOs, whereas reducing total stormwater has a minimal influence on

10
146 treatment plant operation. On the other hand, retention of water inside the engineered media is

147 important to ensure plant survival during dry periods [55]. Retention is quantified in percent as

148 the relative retention (RET; Eqn. 2). RET is the volume of water retained (Vr) in the soil after

149 losses over a period of time (t), relative to the volume of water that entered the soil during this

150 period (Vin), usually as precipitation (p). As shown in Table 2, Vr depends on VMC and the water

151 lost through discharge (qd) and evapotranspiration (ET). This relationship is defined

152 mathematically in Section 3.1.

of
12
./0 =

ro
153 (2)
134

154
-p
Green roofs have been shown to annually retain an average of 55 to 88% of runoff [56];
re
155 whereas retention per rainfall event varies more widely, from 15 to 72% for median event based
lP

156 retention [57]. Median peak discharge attenuation ranges from 43 to more than 90% [58,59].

157 However, event based performance metrics depend on the magnitude of the rainstorm [60] (see
na

158 Section 3.5) and the green roof design properties (see Section 4.1).
ur

159 2.2.2. Urban heat mitigation


Jo

160 Urban heat mitigation refers to strategies that aim to reduce the negative impacts of urban

161 heat, especially those that lower temperatures at the hottest time of day [61]. Green roofs can help

162 mitigate urban heat [10,62] because they reduce the sensible heat flux (QH) above the roof

163 through increases in evapotranspiration and alterations in the radiation flux (Rn) (as shown in

164 Table 2).

165 The sensible heat flux can be used as an evaluation index for urban heat mitigation, as

166 defined by Takebayashi and Moriyama (2007) [63] and Sailor (2008) [64]. Scherba et al. (2011)

167 went further to suggest that the peak sensible heat flux (Hp) (W m-2) and the total daily sensible

11
168 heat flux (MJm-2 day-1) are also relevant [65] because they influence the maximum daytime

169 temperature and nighttime cooling, respectively. The peak sensible heat flux (Hp) is presented in

170 Table 2 as an evaluation metric for urban heat mitigation; however, quantifying reductions in

171 sensible heat flux are not straightforward because direct measurements are difficult in practice

172 [66] and modeling is complex. Urban scale microclimate models [67–70] are needed to link the

173 sensible heat from the green roof directly to urban heat mitigation. Some experimental studies

174 instead use the surface or canopy air temperature as a proxy to evaluate urban cooling from green

of
175 roofs [30,44,71–75] or try to quantify reductions in space cooling [76]. However, none of these

ro
176 proxies are satisfactory because building scale factors are difficult to experimentally link to the

177
-p
urban scale without modeling the microclimate [67–69,77].
re
178 Green roofs can reduce peak surface temperatures by as much as 25°C relative to
lP

179 conventional roofs [29,78,79], whereas air temperatures above the roof may only change by one
na

180 degree or less [80,81]. Scherba et al. (2011) [65] found that a green roof has the potential to

181 reduce the total daily sensible heat flux by 52% compared to a black roof, even though white
ur

182 roofs reduced this flux by 30% more than green roofs. However, reductions in heat will vary
Jo

183 depending on climate, soil moisture, diurnal patterns, and urban canyon configuration (walls,

184 roads, etc.). Reductions can be small [71] or even exhibit a negative penalty [25,71,82].

185 2.2.3. Solar panel output improvement

186 Solar photovoltaic (PV) panels produce distributed, renewable energy by converting solar

187 radiation into electricity. Solar panel power output depends on the available solar radiation and

188 the conversion efficiency, which is reduced as temperatures rise above the reference temperature

189 (usually 25ºC) by approximately 0.45% per 1ºC for crystalline silicon panels [83,84]. Shown in

190 Table 2, green roofs can improve conversion efficiency by keeping operating temperatures low

12
191 through their ability to decrease the sensible heat (by increasing ET) and LW radiation absorbed

192 by the panels (by decreasing surface temperature) [12,85,86]. Depending on the green roof

193 configuration, outgoing LW radiation could also increase, which could be converted to heat on

194 the panel, decreasing panel efficiency. However, some PV panels can convert LW radiation to

195 electricity, which would increase solar panel output. Light colored green roof vegetation and

196 substrate may also reflect SW radiation back into the atmosphere, which can increase the solar

197 radiation available for panels to convert to electricity in some cases, including when they are

of
198 vertically mounted [87].

ro
199 Various performance metrics have been used to quantify the effect of green roofs on solar

200
-p
panel output, including comparisons of: instantaneous power output over time [12], the average
re
201 energy yield for a fixed time period [88], and the specific energy yield (Y), which is the energy
lP

202 per unit peak power for a fixed time period [86]. The latter, shown in Eqn. 3 and Table 2, is
na

203 useful for comparison of PV systems of different sizes and nominal efficiency ratings on

204 vegetated and non-vegetated roofs because it describes the amount of energy produced (Ea)
ur

205 relative to the maximum capacity of the system under standard conditions (Ppeak), irrespective of
Jo

206 panel surface area.

67
207 5 = (3)
89:7;

208 Compared to traditional roofs, experimental and modeling results show that green roofs

209 could increase annual PV electricity yield from 0.5% to 8.5% [12,37,89–93]; however, results

210 vary depending on the climate and the green roof design properties. In cooler climates, the

211 increases may not be as significant [89].

13
212 2.2.4. Building thermal performance

213 Efforts aimed at increasing the ability of buildings to maintain indoor temperature, despite

214 external weather fluctuations, is a concept referred to as improving the thermal performance of a

215 building [94–96]. Green roofs increase building thermal performance in summer by altering the

216 radiation flux (through shading and reflectivity) and increasing evapotranspiration (shown in

217 Table 2). Green roofs also improve it throughout the year by adding insulation, which can be

of
218 approximated as the thermal resistance (R-value) (shown in Table 2) and by increasing thermal

219 mass (approximated by the heat capacity), which decreases the rate the roof will heat up and cool,

ro
220 stabilizing building temperatures [11,29,97]. The combination of the available energy and

221
-p
thermal resistance controls the amount of heat flowing into and out of the building, which can
re
222 save energy in both warm and cool months.
lP

223 To quantify the influence of the green roof on building thermal performance, a few studies
na

224 consider the conductive heat flux (QG) from the green roof to the building [98–103], which
ur

225 could be used an evaluation metric. To quantify building energy load (in MJ or kWh), other
Jo

226 metrics are used, such as the monthly energy usage difference between a building with a

227 conventional roof and one with a green roof [64] or the ratio between the two, referred to as the

228 Dynamic Benefit of Green Roofs (DBGR)[104]. In this case, the conductive flux is used as input

229 to whole building energy simulation models, such as EnergyPlusTM, or in urban scale climate

230 models that represent the built environment (e.g., [67,77]).

231 Experimental and modeling results show that green roofs can reduce annual building

232 electricity consumption by as much as 4% [64,81,104,105]; however, this can be seasonally

233 higher or lower depending on the climate and green roof design properties [106–108]. Due to the

14
234 complex energy balance (described in the next section), a green roof will not always lead to

235 building energy savings [104].

236 3. Conceptual water and energy balance in a green roof

237 In this section, we show how the processes listed in Table 2 are linked to each other through

238 the water and energy balances, increasing the complexity of understanding green roof co-

239 benefits. Water and energy cycle in and out of the green roof in three main ways: (i) as water

of
240 interactions within the substrate layer, (ii) as a release of water, heat and radiation at the green

ro
241 roof surface, and (iii) as heat stored and transferred through the substrate.

242 3.1. Substrate-water interactions


-p
re
243 Generalized in Eqn. 4 and shown in Figure 2, water entering the green roof as
lP

244 precipitation (p) is equal to the change in water stored in the substrate (Δ<) plus water leaving as
na

245 ET, surface runoff (qr), or vertical drainage (qd). Not all processes happen simultaneously,
ur

246 however. Discharge takes place primarily during rain events, whereas ET is released in between

247 these events, decreasing the amount of water stored [109,110]. Furthermore, surface runoff is
Jo

248 usually negligible because water infiltrates quickly into the highly porous media [48,111].

249 = = ET ∙ As + CD + CE + Δ< (4)

250 When water rapidly enters the green roof, the substrate media is likely unsaturated and the

251 water in non-steady state, meaning that all pores are not saturated before drainage occurs

252 [48,111,112]. Unsaturated flow through a vertical soil column (1-D) can be described using the

253 Darcy-Buckingham Law (Eqn. 5) [67,113] or Richard’s equation under transient flow [114] (see

254 Appendix A.3.).

15
IJK
255 CF = Κ H + 1N (5)
IL

256

of
ro
-p
re
lP

257
na

258 Figure 2. Surface and subsurface water interactions in a green roof

259 Shown in Figure 2, vertical drainage is a function of the unsaturated hydraulic


ur

260 conductivity (Κ) and the change in soil matric potential (&' ) with respect to the change in depth
Jo

261 (z). K, which describes the permeability, or how well water flows through connected pore space

262 (see Appendix A.4.), is a function of the hydraulic conductivity at saturation (KS), related to soil

263 properties [115], and the soil moisture ( ) or the volumetric water content (VMC; defined in

264 Table 1 and Appendix A.1.). Soil matric potential, which describes the ability of soil to attract

265 water using cohesion and surface tension, increases as decreases [117] and also relates to soil

266 properties.

267 The maximum amount of water that can be stored in the substrate before drainage by gravity

268 is referred to as the maximum water holding capacity (WHCmax) or the water content at field

16
269 capacity ( $% ) [118]. At a certain point, referred to as the wilting point ( ), plants can no longer

270 uptake water due to high matric potential [119]. The maximum retention capacity (Smax) (Eqn. 6)

271 is the difference between the water content at field capacity and wilting point.

272 <' O = $% − (6)

273 3.2. Water and heat released at the surface

274 The water balance is linked to the energy balance through ET, which is converted to an

of
275 energy flux (called latent heat, Qe) by multiplying it by the latent heat of vaporization (Λ)

ro
276 [39,120]. ET (Eqn. 7 [121]) is a dynamic process governed by the vapor pressure deficit (VPD)

277 between the surface and air ( −


-p
), as well as, the resistance to water loss from the surface (D )
re
278 and into the air (D P ). The VPD, a function of ambient air temperature, Ta, is the amount of
lP

279 moisture that can be added to the air before saturation [39,99,122]. For practical reasons, Eqn. 7
na

280 is often replaced by the Penman-Monteith (P-M) reference equation [120], discussed in

281 Appendix A.5.


ur

Q7 R9
Λ/0 = = ( − )
Jo

282 (7)
S(TU VT7W )

283 where variables are defined in the nomenclature.

284 The surface resistance (rs; Eqn. A7 Appendix A.6. [120]), represents the resistance due to

285 physiological properties of the soil and vegetation, including the stomatal resistance (rI), the

286 force opposing the release of water vapor through leaf stomata (pores), and the leaf area index

287 (LAI), the projected leaf area per unit ground area (m2/m2)[39]. Both of these variables can be

288 influenced by the choice of vegetation, while rI also changes over time as a particular plant

289 responds to changes in environmental conditions, including solar radiation, vapor pressure

17
290 deficit, soil water content, air temperature [123] and even soil type [74]. The aerodynamic

291 resistance to vapor transfer (rav; Eqn. A8 Appendix A.7.), is related to atmospheric turbulence

292 (e.g., friction from air flowing over the top of the vegetation) [124]. It is a function of surface

293 roughness (related to plant height), wind speed (u) and atmospheric stability [78,120,125] (see

294 details in Appendix A.7.).

295 The available energy at the surface is partitioned between the latent and sensible heat [126].

of
296 The sensible heat flux (QH; Eqn. 8 [97,127]), is a function of the temperature differential between

ro
297 the surface and the air, as well as the aerodynamic resistance to heat transfer (rah) which is

298 -p
related to rav, but describes resistance to heat instead of vapor. The two terms are often
re
299 interchangeable since differences are minimal [125], and are sometimes referred to collectively as
lP

300 aerodynamic resistance (ra) to the transfer of heat and vapor (as in [124]; see Appendix A.7.).

Z[ \]
301 QY = (T − T` ) (8)
na

^[_
ur

302 3.3. Radiation interactions at the surface


Jo

303 ET and sensible heat are part of a larger energy balance, generalized for green roofs in Eqn. 9

304 [125,128]. Net radiation (Rn), the difference between incoming (.ab ) and outgoing (.cde )

305 radiation, is equal to convective heat released into the air as latent (Qe) or sensible heat (QH),

306 conductive heat (QG) transferred through solids, or energy transiently stored by the substrate or

307 vegetation (Δ ).

308 .b = .ab − .cde = + Y + f +Δ (9)

309 Outgoing radiation (Eqn. 10) is the sum of outgoing short (<gcde ) and longwave (hgcde )

310 radiation. <gcde , governed by the surface albedo ( ), is the fraction of incoming short-wave

18
311 radiation (<gab ) that is reflected from the surface. Often assumed constant, albedo changes

312 depending on the substrate moisture content [129], fractional vegetation coverage (η), and

313 vegetation height [97,130]. hgcde from the green roof to the atmosphere is a function of surface

314 temperature (Ts) to the fourth power, surface emissivity ( ), as well as, incoming LW radiation

315 [99].

316 .cde = <gcde + hgcde = <gab + ϵ 0 j + (1 − )hgab (10)

of
317 SW and LW radiation fluxes within plant canopies are more complex. They can be

ro
318 represented using the radiation fluxes at the top of the canopy and the transmittance (#) through

319 -p
the plant canopy (see Figure 3). Transmittance, which quantifies how much radiation is scattered
re
320 by the vegetation [97], relates to the vegetation properties, including vegetation coverage (η),
lP

321 LAI and leaf angle distribution, as discussed in Appendix A.8. [97,131,132].
na
ur
Jo

322

19
323 Figure 3. Schematic of energy balance taking place between the lower atmosphere and the green roof canopy and engineered

324 media layers.

325 3.4. Heat stored and transferred through the substrate

326 Just as water is stored in the substrate, energy can also be stored in green roofs, although

327 this may be negligible in thin soils [128] and vegetation. Thermal storage (Δ ; Eqn. 11

328 [125,128]), the rate of change in heat content of the roof per unit volume, is represented as the

329 rate of change in roof temperature (dT/dt) times the heat capacity ( , with units J m-3 K-1),

of
330 which is defined as the amount of heat (J) needed to raise a unit volume (m3) by 1 degree (K).

ro
331 Specifically for substrate, the heat capacity ( ) depends on the relative fraction of solid particles,

332
-p
air, and water and the thermal properties of each material [125].
re
Fk
Δ = El
lP

333 (11)
Fe

334 Heat flowing through the substrate is referred to as the conductive heat flux (QG; see
na

335 Figure 3). Shown in Eqn. 12 [99,133]), the magnitude of the flux is proportional to the
ur

336 temperature differential between two depths that define a layer, as well as, the thermal
Jo

337 conductivity ( ) of this layer. As with water, heat entering the green roof surface is not equal to

338 heat leaving from the subsurface. Thus, the depths and temperatures to consider will change

339 depending on whether the top or bottom of the green roof is evaluated (see Appendix A.9.).

mn
340 f = FL
(12)

341 Thermal conductivity ( ) the rate at which heat is transferred across a material, is the inverse

342 of thermal resistance (R-value), which is related to insulation (see Section 2.2.4). It varies over

343 time with the depth of substrate and moisture content [99,125] and is a function of substrate

20
344 porosity (density) and composition, and thermal capacity of substrate particles [134] (see

345 Appendix A.9. for specific equations).

346 3.5. Intersection of surface, sub-surface, and boundary conditions

347 Figure 4 summarizes the relationships between the exogenous weather variables known as

348 boundary conditions (white boxes at top), the water (blue circles) and energy (black circles)

349 balance terms (Eqn. 4 and Eqn. 9 respectively) and the parameters that dictate them, which are

of
350 related to green roof properties for vegetation (green squares) and substrate (brown squares).

ro
-p
re
lP
na
ur
Jo

351

352 Figure 4. Relationships between exogenous weather variables, energy and water balance terms, and their parameters that are

353 linked to green roof soil and vegetation design properties.

354 From Figure 4, we see that the local climate patterns (i.e., the boundary conditions: rainfall,

355 wind, vapor pressure deficit, air temperature, and incoming radiation) influence the co-benefits

21
356 both directly and indirectly. Of particular importance is rainfall intensity, which directly

357 influences soil moisture (i.e., retention of stormwater) and discharge rate [52,60,135,136]. The

358 magnitude of the rain storm dictates the percent of rainfall that is retained and the time between

359 storms dictates the antecedent moisture conditions (AMC). Based on an analysis in three

360 Canadian cities, small events (< 3 mm) were fully retained, large events (> 15 mm) were not, and

361 retention of medium storms (3 – 15 mm) was dependent on the AMC [60]. When antecedent soil

362 moisture is high and infiltration capacity is reduced, surface runoff (or ponding) could also

of
363 increase [137].

ro
364 In addition to stormwater discharge, soil moisture also influences many of the other processes

365
-p
and parameters, including thermal conductivity, which increases with moisture content
re
366 [108,138,139] (see Section 4.5). Evapotranspiration also depends on VMC. ET is usually low
lP

367 when VMC is low, high when it is high, and is approximately linear to increases in moisture in
na

368 between [140] (see Section 4.3). When ET is not limited by soil moisture, like in a wet,

369 temperate climate, it uses 70% to 80% of energy received from solar radiation, leaving only 16 –
ur

370 30% for sensible heat, and 3 – 15 % for conductive heat through the green roof [99]. However, in
Jo

371 drier climates, sensible heat can account for nearly 70% of energy, while ET only accounts for

372 25%. Thus, the sensible heat is indirectly linked to soil moisture through its dependence on ET.

373 Some studies have found that irrigating the green roof during times of low rainfall increases ET

374 [25,110], which would also reduce sensible heat.

375 Processes relying on ET for cooling, however, are dependent on the relative humidity (linked

376 to VPD), which determines the amount of water the air can accept. Irrigation is more effective in

377 drier and more temperate climates, and less so in more humid and tropical climates where air is

378 nearly saturated [141]. The following recommendations are provided for a dry or temperate

22
379 climate where ET could be a key driver for cooling and where soil moisture could influence

380 retention for medium sized precipitation storms.

381 4. Altering processes through changes in green roof design parameters

382 In the previous section, we presented how the processes identified in Table 2 are linked to

383 each other and to the substrate and vegetation parameters of a green roof. In this section, we

384 review the literature on each process to describe how green roof design properties can influence

of
385 co-benefits achieved by the green roof. Figure 5 presents the 13 design properties discussed in

ro
386 this section by category (left most column): irrigation, substrate, vegetation, and surface, and (a)

387 -p
how an increase in these properties would influence the physical processes, or the direction of
re
388 change needed in these properties to (b) maximize the co-benefits, or (c) minimize the
lP

389 constraints.
na
ur
Jo

23
of
ro
-p
re
lP
na
ur
Jo

390

391 Figure 5. (a) Direction of change in the physical processes if the design parameters (rows) were increased. Colors show the
392 direction of change in the process (blue: increase, orange: decrease). The width and shade of the color represents how
393 much influence the design property has on the process (considerable ---/ +++) to limited (-/+). (b) Direction of change
394 in the design property that would maximize the co-benefit or (c) minimize the constraint. The arrow size represents how
395 significant the change in the benefit or constraint would be. *Below the vegetation canopy.

396 4.1. Discharge rate

397 As shown in Figure 5, the discharge rate primarily influences stormwater detention.

398 Decreasing the discharge rate will improve detention; however, the rate must still be high enough

399 to avoid water ponding on the surface of the green roof [55]. As discussed in Section 3.5,

24
400 discharge increases with higher precipitation intensity and antecedent soil moisture. Discharge

401 can be decreased by changing the substrate depth and composition, the properties of other green

402 roof layers [51], and in some cases, vegetation coverage. For example, Yio et al (2013) showed

403 that discharge decreases as substrate depth increases, but found that the relationship was not

404 linear [142]. Peak discharge attenuation has been shown to be higher on roofs with vegetation

405 than bare soil [57], although it remains unclear whether the type of vegetation (and the leaf area)

406 can also influence discharge.

of
407 Discharge can be also decreased by altering the physical properties of the substrate, which

ro
408 influence the unsaturated hydraulic conductivity (K) and the matric potential (&' ) (see Eqn. 5).

409
-p
For example, Bollman et al. (2019) found that peat moss detained more water than other
re
410 substrates that had a higher mean grain size (e.g., lower density) and higher hydraulic
lP

411 conductivity than peat moss. However, a lower hydraulic conductivity may not guarantee higher
na

412 detention. A different sample of sand that had the lowest hydraulic conductivity, but higher

413 density than the peat moss, released more water after a minute [55]. The material’s surface
ur

414 tension, or capillary pressure, which was not evaluated in this study, also plays a role, although
Jo

415 this characteristic is difficult to attribute to a particular soil.

416 Theoretically, the capacity of a medium to convey water (related to K) depends on the pore

417 size distribution (abundance and size of soil particles), connectivity (how many flow paths), and

418 tortuosity (number of turns per path) [143,144]. Larger pores drain water faster, but the presence

419 of smaller pores alongside larger ones also slows water down. For instance, several studies found

420 that increasing the amount of organic matter (even by 5%), which increases the number of fine

421 particles and decreases K, can considerably reduce the discharge rate [55,142]. The microporosity

422 and particle texture also plays a role in water drainage. This is why sand (round particles), which

25
423 has a lower hydraulic conductivity than other common green roof media (e.g., perlite or tuff), has

424 been shown to contain less water at the same matric potential than other substrates with higher K

425 and more jagged particles [55,143,145]. The relationship between unsaturated flow and substrate

426 composition is complex and the literature is far from exhaustive on this topic, especially for green

427 roof media, which behaves differently from natural soils.

428 Ultimately, to reduce the discharge rate and improve detention, substrate depth, organic

of
429 content, particle roughness, and vegetation coverage should be increased when possible, while

430 antecedent soil moisture should be decreased (Figure 5 – first column) by increasing ET (see

ro
431 Section 4.3) or limiting irrigation.
-p
re
432 4.2. Volumetric water content
lP

433 As discussed in Section 3, the substrate water content can alter each of the energy and water

434 balance processes (to varying degrees) and thus influences all of the co-benefits. However,
na

435 depending on the benefit, it may be preferential to increase or decrease soil moisture (see Figure
ur

436 5). This section discusses ways to retain more moisture in the substrate after a rainfall or
Jo

437 irrigation event, thus increasing the time it takes the soil to reach the wilting point. Although

438 decreasing ET from plants would technically increase soil moisture, this aspect is excluded from

439 this discussion, as is depression storage on leaves.

440 VMC after rainfall (or irrigation) depends on the maximum retention capacity (Smax, Eqn. 6),

441 substrate depth and substrate properties. Substrate depth is shown to be significantly positively

442 correlated to moisture retention [22], although there is an inflection point where additional

443 increases in substrate depth will only marginally change the retention capacity. For instance,

444 Metselaar (2012) found that increasing substrate depth from 5 to 15 cm increased the retention

26
445 fraction by 10%, yet beyond 60 cm, increases in retention were minimal [146]. This is promising

446 because larger moisture retention could be achieved with minimal increases in substrate depth.

447 As with discharge, the ability of substrate to retain moisture also depends on organic content

448 and particle distribution, however, the relationship is not straightforward. Nektarios et al [147]

449 observed higher water content in loamy media during drought than in media without organics

450 [111]. However, Bollman et al. (2019) found the highest water content after 14 days in perlite

of
451 media, which had a lower percentage of fine particles (< 2 mm) than other materials, including

452 peat moss and sand [55]. Graceson et al (2013) did not find a significant difference in short-term

ro
453 retention in substrates with varying percentages of fines (2.5% to 57%) [148]. As shown in

454
-p
Figure 5, it is still unclear how changes in particle distribution (ratio of small particles) influence
re
455 the retention of moisture. It is likely also linked to particle texture and thus capillary forces of
lP

456 each media type; however, this has not been evaluated.
na

457 By mixing different substrate materials or adding additional layers to the green roof, moisture
ur

458 content can be increased without considerably decreasing detention. For example, Bollman et al.
Jo

459 (2019) mixed 80% perlite (highest 14-day moisture content) with 20% peat moss (highest short-

460 term detention) and found that the percent of water retained doubled while detention only

461 decreased by one third [55]. Moreover, Farrell, Ang, and Rayner (2013) found that adding water

462 retaining materials to green roof soils, such as silicate granules, improved water retention by as

463 much as 9.4% [149]. Finally, another approach could be to place an engineered water retention

464 layer below the substrate, similar to those offered by Nophadrain [39], in order to retain moisture

465 for plants. Overall, soil moisture is retained longer by increasing substrate depth and the

466 percentage of organics and fines (although not in every case).

27
467 4.3. Evapotranspiration

468 As shown in Figure 5, increasing ET would improve all of the co-benefits. Apart from

469 ensuring there is ample soil moisture for ET to take place, one of the primary ways to increase ET

470 is to decrease the resistance to water loss from the surface to the air (see Eqn. 7). Aerodynamic

471 resistance (rav, equation shown in Appendix A.7.) could be lowered by increasing surface

472 roughness, which could be accomplished by increasing plant height (or height variation). Nagase

of
473 and Dunnett (2012) [28] found that for every 1 cm increase in plant height, cumulative runoff

474 volume reduced by 0.4%. Since substrate characteristics did not change, this likely means that ET

ro
475 increased with plant height; however, it could also be due to increased interception. This

476
-p
relationship was not linear (R2 = 0.465), implying that there are other factors at play. Most
re
477 notably, these findings could also be due to changes in ET through the surface resistance term,
lP

478 which is related to the stomatal resistance (rI) and leaf area index (LAI) of the vegetation (shown
na

479 in Eqn. A7).


ur

480 Surface resistance decreases as LAI increases, which is one of the most important plant
Jo

481 characteristics related to ET [39], because there is more surface area available to release water. It

482 is difficult to decouple LAI from stomatal resistance in terms of ET, especially since LAI

483 changes seasonally, while stomatal resistance changes daily. However, the role of LAI is non

484 negligible. One agricultural study showed the ET rate of two crops diverged at the same point in

485 the season when LAI of one crop started to decline while the other continued to increase [150].

486 Moreover, when sparse vegetation is compared to dense (i.e., more surface area of leaves), ET is

487 more intense in the former case [151] and less stormwater leaves the roof [152]. These increases

488 in ET could also be due to differences in structural complexity (e.g., variation in plant height,

28
489 branching, and leafiness); however, one study found that vegetation structure did not significantly

490 affect green roof cooling [153].

491 Surface resistance also decreases with lower stomatal resistance (rI), which varies among

492 plant species [154]. Stomatal resistance is lowest in C4 plants, a vegetation group consisting

493 mainly of grasses and herbs, with rI values of ~50 to 60 sm-1 for these species [155,156]. This is

494 one of the reasons why grasses have been shown to be more effective at reducing stormwater

of
495 (and thus releasing water through ET) than succulent and forbs species [28,155], which have

496 higher stomatal resistance. Crassulacean acid metabolism (CAM) plants (mostly succulents) have

ro
497 an rI ranging from 170 [156] to1000 s/m [157]. Of particular importance is the relationship

498
-p
between stomatal resistance and VMC because a lack of moisture limits stomatal opening and
re
499 ET, but ample moisture increases them [39,74,99,101,110,138]. However, an exception to this
lP

500 pattern would occur if deep-rooted plants access moisture at lower substrate depths. In this case,
na

501 ET could still increase even as the top of the soil decreases in moisture content.
ur

502 Ensuring soil remains moist (i.e., through irrigation) is especially important for survival of
Jo

503 taller plant species with lower stomatal resistance (like grasses). These plants will continue to

504 release water even when soil moisture decreases and will wilt once soil moisture gets too low. It

505 follows logically that some studies have found that irrigating the green roof during times of low

506 rainfall also increases ET [25,110] and reduces air temperatures [153]. As an alternative to

507 increasing irrigation, substrate depth and plant root depth could be increased so that plants could

508 access moisture at lower soil levels, even when the soil surface is dry.

509 Finally, it may be possible to increase ET by reducing vegetation coverage, thus excluding

510 the resistance linked to vegetation. Several field studies have shown that bare (or sparsely

511 covered) substrate loses more water to ET than well covered green roofs [25,158,159]; although

29
512 this is often only the case when soil is moist after irrigation. Over dry soil before irrigation,

513 Coutts et al (2013) found that ET was lower and sensible heat higher over soil than vegetation

514 [25]. Stovin et al (2015) suggest that ET is only increased by plants as soil moisture declines

515 [57]. However, some studies have also shown that ET is higher from vegetated roofs than from

516 bare soil [57,110,160] even under well-watered conditions, as in a greenhouse experiment

517 conducted by Voyde et al (2013) [110]. Many factors may play a role in whether soil or

518 vegetation contribute more to latent heat, including differences in shading, albedo, substrate

of
519 temperature and moisture, stomatal resistance [25,135,155], root depth, and atmospheric

ro
520 conditions.

521
-p
While it is still unclear whether vegetation encourages or suppresses ET, it is clear that
re
522 without it, a green roof would lose the ecological function to promote biodiversity as well as the
lP

523 shading function that decreases energy sent to the building (see Section 4.4). Moreover, a dry
na

524 green roof with little to no vegetation could increase heat at the surface due to low thermal

525 admittance (the square root of the product of thermal conductivity and heat capacity) [125]. Thus,
ur

526 the preferred way to increase ET and surface cooling could be to select vegetation with low
Jo

527 surface resistance or to mix these plants with higher surface resistance vegetation (e.g., Sedums)

528 so more water is available to the plants that need it [24,153]. Overall, as shown in Figure 5, when

529 plants are present, ET increases with lower resistance (e.g., taller plants, higher LAI) and higher

530 moisture availability (e.g., increasing irrigation or root depth). Many of these species grow in

531 substrate depths of 15 cm or less [21].

532 4.4. Outgoing radiation and shading

533 Limiting the incoming radiation below the vegetation canopy (i.e., absorbed radiation) will

534 positively influence all of the heat related benefits. As shown in Figure 5, absorbed radiation is

30
535 decreased if more shortwave radiation is sent back to the atmosphere by altering the albedo (Eqn.

536 10) and if radiation is blocked from reaching the soil surface through shading. Green roof

537 surfaces below dense canopies (with high shading) absorb as little as 2% of net radiation, while

538 surfaces below sparse canopies absorb as much as 30% of Rn [154]. It may also be possible to

539 release more absorbed radiation (reducing stored energy) by increasing surface emissivity.

540 Albedo is increased by selecting light or silver colored plant species, including: many grasses

of
541 (albedo ranging from 0.25 to 0.30 [161]), light colored Sedums such as Sedum tomentosum (0.23)

542 or sexangulare (0.22) [162], Stachys byzantia [30], or others found in [26] and [163]. Ensuring

ro
543 high vegetation coverage can increase albedo since plants generally have higher albedo than

544
-p
substrate (ranging from 0.08 to 0.13 [162]). However, albedo decreases with higher soil moisture
re
545 content [164] and with increased plant height (taller plants can trap radiation)[97,130]. When
lP

546 vegetation coverage is sparse, higher albedo substrates can increase reflectivity, including those
na

547 containing perlite [161], white gravel [25], or mineral aggregates (e.g., rooflite extensive media

548 [161] or Volkatech™ Extensiv plus with magnesium oxide). Similarly, green roof emissivity can
ur

549 be increased by selecting high emissivity plants (e.g., grass and cacti [97,130,165]) or substrate
Jo

550 materials (e.g., wood, brick [166,167], tile [166,168], calcite, and feldspar [169]).

551 To increase shading, which limits exposure to radiation, the transmittance of the plant canopy

552 [97] should decrease by selecting plant species with large and/or horizontal leaves [134] or

553 increasing vegetation coverage [153]. For example, in a study conducted in Singapore, a type of

554 palm reduced annual energy consumption more than trees and low growing, turfing plants [108].

555 Shading from high LAI or dense coverage has been shown to be more important at reducing

556 available energy in the soil than increasing ET through changes in stomatal resistance [99]. This

557 is likely why MacIvor et al (2016) found that Sedums were significantly more effective at cooling

31
558 than meadow plants [153]. Shading (and subsequent reductions in absorbed short wave radiation)

559 reduces more energy than increases in ET can provide through decreases in stomatal resistance.

560 One plant that has been shown to cool better than others, likely as a combination of higher

561 LAI and albedo, is Stachys byzantina, a large leafed perennial herb covered in silver white hairs.

562 Blanusa et al. (2013) found the substrate below the Stachys canopy was up to 12 °C cooler than

563 the other plant species (including other broad leafed perennials and Sedums), the leaf surface was

564 cooler by up to 5°C, and canopy temperature up to 1°C cooler [30]. Overall, as shown in Figure

of
565 5, leaf area, vegetation coverage and albedo will play a larger role in reducing absorbed energy

ro
566 than emissivity.
-p
re
567 4.5. Sensible heat
lP

568 Shown in Figure 5, decreasing the sensible heat flux improves heat mitigation and solar

569 panel output improvement. Based on the energy and radiation balance (Eqns. 9 and 10), sensible
na

570 heat is primarily decreased by increasing latent heat (ET), reflected short-wave radiation, and
ur

571 shading. Following recommendations in the previous subsections (shown in Figure 5), this can
Jo

572 be done by increasing soil moisture and organic content and selecting taller plant species with

573 low stomatal resistance, higher LAI (low transmissivity) and high albedo. Coutts et al (2013)

574 suggested that green roofs would only be able to provide a comparable ability to mitigate urban

575 heat as a high albedo rooftop (e.g., a white roof) if they are planted with a dense mixture of plants

576 that actively transpire (low stomatal resistance) and are regularly irrigated [25]. However,

577 increasing vegetation coverage is more important for cooling than increasing plant height,

578 variation [153] and stomatal resistance [99]. Finally, it may also be possible to reduce sensible

579 heat by reducing stored energy through increases in surface emissivity; however, this may only

580 lead to cooler surfaces at night.

32
581 4.6. Insulation and conduction

582 Building thermal performance is improved by decreasing the downward conductive heat from

583 the bottom of the green roof. As shown in Eqn. 12, heat transfer in substrate is decreased with

584 lower thermal conductivity, , which means that more energy is stored in the substrate given the

585 same energy input from the surface. Thermal conductivity depends on moisture content (VMC),

586 as well as, the substrate density and the thermal capacity [134] of substrate particles (see

of
587 Appendix A.9.). Specifically, heat transfer has been shown to increase in higher density

ro
588 substrates [26], meaning that lower porosity soils with high field capacity [134] and a lower

589 percentage of small particles are preferred to reduce conduction and improve building thermal

590
-p
performance (see Figure 5). However, in this case, surface temperatures may increase. Due to
re
591 lower thermal diffusivity, low density soils (e.g., peat) diffuse heat slower than higher density
lP

592 soils, meaning more heat is available at the surface [125] and higher organic content tends to
na

593 increase surface cooling [153].


ur

594 As discussed in Section 3.5, conductive heat increases through wet substrates [108,138,139].
Jo

595 Specifically, thermal conductivity of saturated soils has been shown to be double that of dry soils

596 [129] because air is a better insulator than water [11]. Dry, highly porous substrates will be most

597 effective at decreasing the conductive heat flux through the green roof [134]. However, this does

598 not account for the fact that heat capacity of saturated soil is higher than for dry soils [125], thus

599 energy will be stored longer inside moist substrate with more energy released at night (usually at

600 the surface). Lower soil moisture also decreases ET [39,125], which would increase the surface

601 temperature and sensible heat flux. Contributions from ET and energy stored in the soil could be

602 why Sailor (2008) found that irrigating during summer could decrease monthly building energy

33
603 consumption, although marginally. One way to increase the insulating properties of the green

604 roof without reducing moisture could be to add glass beads filled with air into the substrate [131].

605 Heat released below the green roof also depends on the substrate depth [99,125].

606 Mathematically, the energy released over time is related to the change in temperature per change

607 in depth (see Eqn. 12) and the energy storage (see Eqn. 11). Deeper substrates have been

608 attributed to lower heat gains and losses, a higher degree of insulation [106,131], and reduced

of
609 building energy consumption [108]. However, improvements in building energy consumption

610 due to increases in depth are smaller in moist substrates [108]. Sailor (2008) found that the

ro
611 substrate depth was a larger contributor to thermal performance than the fractional vegetation

612
-p
coverage (related to shading, transmissivity) [129], although Tabares-Velasco and Srebric (2012)
re
613 found that the LAI had a larger influence than the substrate depth on the conductive heat flux
lP

614 through the green roof [99]. The difference may depend on the magnitude of feasible changes to
na

615 design parameters. For example, increasing the LAI by 2 or 3 units may be easier than increasing

616 the substrate depth by more than 10 cm, as was suggested by Sailor (2008). Ultimately, the
ur

617 relative importance of each also depends on local climate, season, and the baseline comparison
Jo

618 values.

619 The ability of a green roof to reduce heat flows into the building also depends on the

620 insulation (e.g., wool or fiber glass) that lies between the green roof and building. Wong et al.

621 (2003) found that a green roof on an uninsulated building decreased energy consumption for

622 cooling by 15%, but a green roof over an insulated roof decreased consumption by less than 2%

623 [108]. Thicker insulation layers dampen the benefit of a green roof to reduce energy flows into

624 the building [108,162], especially in winter. In summer, Vera et al. (2017) found that ET and

625 shading from the vegetation could be more effective at reducing heat fluxes into the building than

34
626 increasing roof insulation [170]. Shown in Figure 5, to reduce conduction and improve building

627 thermal performance, one should increase its insulation power (by increasing substrate depth,

628 decreasing small particles and decreasing soil moisture) and reduce available radiation (see

629 Section 4.4).

630 4.7. Trade-offs and synergies between co-benefits and constraints

631 Figure 5c summarizes how the green roof design properties influence all the co-benefits and

of
632 constraints (which include structural loading, maintenance, and cost). The constraints are

ro
633 minimized primarily by reducing irrigation, substrate depth, substrate density, and plant height.

634 -p
Relaxing these constraints could improve the achievable benefits or lead to trade-offs. For
re
635 instance, deeper substrates (e.g., ~ 15 cm) with more organics detain more water, reduce surface
lP

636 heat, and allow for a wider variety of plants to grow, including deeper rooted grasses and forbs

637 species with higher ET and cooling [25]. However, increasing substrate density could increase
na

638 conductive heat and reduce building thermal performance. There are other design properties that
ur

639 could be altered to maximize co-benefits without trade-offs, including substrate particle
Jo

640 roughness or engineered substrate additives, leaf area index, and surface properties like albedo

641 and emissivity. Increasing leaf area and vegetation coverage can considerably cool rooftop

642 temperatures and increase building thermal performance [6,9,36,39]. Using performance-based

643 modeling to determine the optimal combinations of design properties is a challenging, yet crucial

644 step towards boosting synergies and managing trade-offs.

645 Performance-based modeling is also needed to evaluate trade-offs and synergies related to

646 irrigation systems, which increase cost and loading, remain unused when rainfall is sufficient,

647 and may increase demand for drinking water during periods of water stress [40]. Dry soil could

648 also increase stormwater detention (more room to store water) and energy savings (air insulates

35
649 better than water). However, irrigation may be needed during hot, dry spells (expected to increase

650 under climate change [171,172]) for survival of plant species with low stomatal resistance that

651 encourage ET (and cooling). These factors are compounded by the fact the irrigation water could

652 be supplied from harvested stormwater or grey water and could also be used to clean solar panels.

653 Ultimately, the decision to irrigate depends on these trade-offs as well as the importance of

654 cooling relative to the other benefits, especially in a more extreme climate.

of
655 As integrated solar panel green roof systems become more established, performance-based

656 modeling for green roofs will become more critical to balance synergies and trade-offs in design

ro
657 properties with other co-benefits, including the fact that PV modules can decrease surface albedo

658
-p
[173]. When solar panels are present, the choice of vegetation and surface properties is more
re
659 critical, yet more constrained [163]. For instance, vegetation should be short in order to avoid
lP

660 shading the panels from above [21] and large leafed, silvery plants that increase albedo, shade,
na

661 and cool the rooftop [30] should be prioritized since they have been shown to increase electricity

662 production [86]. Finally, adding solar panels to the roof also increases shading on the roof [173],
ur

663 which would improve building thermal performance. As communities continue to expand rooftop
Jo

664 solar energy, these interactions will be important to examine in future performance-based design

665 models.

666 5. Outlook and future research needs

667 Performance-based design models for green roofs are needed in order to optimize green roof

668 design parameters based on multiple objectives, including environmental co-benefits and

669 structural and maintenance constraints. Given the trade-offs discussed previously, the optimal

670 balance of design properties will depend on the importance of each benefit (or constraint) to

671 stakeholders, which can be investigated using decision analysis (e.g., [174,175]. A decision

36
672 analysis model first requires a performance analysis that uses simulation models to capture the

673 relationships between co-benefits and design properties. This study generalized the complex

674 physical concepts to provide the foundation for this performance analysis; however, more work is

675 needed to use this foundation to build a comprehensive performance-based design model. This

676 requires the integration of simulation models and solvers [35,64,176–179] that can portray

677 complex relationships into a simplified framework so the performance analysis can be linked to

678 optimization and/or decision analysis. Consortia is needed among researchers conducting green

of
679 roof simulation modeling in different disciplines, as well as, with those working on multi-

ro
680 objective decision making.

681
-p
In addition to model integration across disciplines, more research is needed within disciplines
re
682 to standardize modeling parameters for different green roof designs and link these parameters
lP

683 mathematically to the performance objectives. For instance, combining various materials into the
na

684 substrate, as suggested in this study, would change the thermal and hydraulic properties of the

685 green roof. The parameters of these substrate mixtures, however, are not well known.
ur

686 Coefficients that depict transpiration (e.g., stomatal resistance) are limited for many landscape
Jo

687 plants, as are canopy values for combinations of plant species. Laboratory and field studies are

688 required to classify and standardize these properties. In order for practitioners to use this

689 information, a universal database where these values can be reported is also vital. Finally,

690 additional research is needed in some cases to show how certain design properties influence the

691 physical processes. For instance, it is still unclear how LAI and substrate organic content relate to

692 discharge timing or how solar panels above green roofs affect air and building temperatures. As

693 practitioners gather more information about system interactions, more intricate green roof designs

694 can be evaluated using performance-based design.

37
695 While this study only focused on the four co-benefits linked to the energy and water balance,

696 future performance-based design models should also consider other green roof co-benefits, such

697 as biodiversity enhancement [13,14], carbon dioxide sequestration [15], air pollution abatement

698 [16,17], and improvements to urban livability. For instance, a minimum soil depth could be

699 instrumental for invertebrates to hibernate in substrate over winter months, and pollinating

700 species may be more prone to certain flowering plants. Moreover, carbon dioxide sequestration is

701 increased as stomatal resistance is decreased (the amount water vapor exiting pores is

of
702 proportional to CO2 entering), which could be another reason to select plant species with low

ro
703 stomatal resistance. Interdisciplinary collaborations among green roof researchers are needed to

704 incorporate multiple perspectives.


-p
re
705
lP

706 6. Acknowledgements
na

707 The authors would like thank the following people for useful discussions and feedback that
ur

708 considerably contributed to the quality and content of the manuscript: Erfan Haghighi, Simone
Jo

709 Fatichi, Juan Pablo Carbajal, Dieter Ramseier, Peter Bach, Joao Leitao, Jennifer Jacobs, Erin

710 Dawson, and Eberhard Morgenroth.

711 Funding: This work was supported by the 2019 Eawag Post-doctoral Fellowship.

712 7. References

713 [1] J. Sheffield, E.F. Wood, Projected changes in drought occurrence under future global
714 warming from multi-model, multi-scenario, IPCC AR4 simulations., Clim Dyn. 31 (2008)
715 79–105.
716 [2] R.P. Allan, B.J. Soden, Atmospheric Warming and the Amplification of Precipitation
717 Extremes, Science. 321 (2008) 1481. https://doi.org/10.1126/science.1160787.

38
718 [3] T.A. Larsen, S. Hoffmann, C. Lüthi, B. Truffer, M. Maurer, Emerging solutions to the
719 water challenges of an urbanizing world, Science. 352 (2016) 928.
720 https://doi.org/10.1126/science.aad8641.
721 [4] H. Eggermont, E. Balian, J.M.N. Azevedo, V. Beumer, T. Brodin, J. Claudet, B. Fady, M.
722 Grube, H. Keune, P. Lamarque, Nature-based solutions: new influence for environmental
723 management and research in Europe, GAIA-Ecol. Perspect. Sci. Soc. 24 (2015) 243–248.
724 [5] K. Vijayaraghavan, Green roofs: A critical review on the role of components, benefits,
725 limitations and trends, Renew. Sustain. Energy Rev. 57 (2016) 740–752.
726 [6] O. Saadatian, K. Sopian, E. Salleh, C. Lim, S. Riffat, E. Saadatian, A. Toudeshki, M.
727 Sulaiman, A review of energy aspects of green roofs, Renew. Sustain. Energy Rev. 23
728 (2013) 155–168.
729 [7] N.S. Williams, J.P. Rayner, K.J. Raynor, Green roofs for a wide brown land: Opportunities
730 and barriers for rooftop greening in Australia, Urban For. Urban Green. 9 (2010) 245–251.

of
731 [8] F. Bianchini, K. Hewage, Probabilistic social cost-benefit analysis for green roofs: a
732 lifecycle approach, Build. Environ. 58 (2012) 152–162.

ro
733 [9] J.C. Berndtsson, Green roof performance towards management of runoff water quantity
734 and quality: A review, Ecol. Eng. 36 (2010) 351–360.
735
736
[10] -p
M. Santamouris, Cooling the cities – A review of reflective and green roof mitigation
technologies to fight heat island and improve comfort in urban environments, Sol. Energy.
re
737 103 (2014) 682–703. https://doi.org/10.1016/j.solener.2012.07.003.
738 [11] H.F. Castleton, V. Stovin, S.B. Beck, J.B. Davison, Green roofs; building energy savings
739 and the potential for retrofit, Energy Build. 42 (2010) 1582–1591.
lP

740 [12] D. Chemisana, Chr. Lamnatou, Photovoltaic-green roofs: An experimental evaluation of


741 system performance, Appl. Energy. 119 (2014) 246–256.
na

742 https://doi.org/10.1016/j.apenergy.2013.12.027.
743 [13] S. Brenneisen, Space for urban wildlife: designing green roofs as habitats in Switzerland,
744 Urban Habitats. 4 (2006) 27–36.
ur

745 [14] C.A. Lepczyk, M.F.J. Aronson, K.L. Evans, M.A. Goddard, S.B. Lerman, J.S. MacIvor,
746 Biodiversity in the City: Fundamental Questions for Understanding the Ecology of Urban
Jo

747 Green Spaces for Biodiversity Conservation, BioScience. 67 (2017) 799–807.


748 https://doi.org/10.1093/biosci/bix079.
749 [15] J. Li, O.W. Wai, Y. Li, J. Zhan, Y.A. Ho, J. Li, E. Lam, Effect of green roof on ambient
750 CO2 concentration, Build. Environ. 45 (2010) 2644–2651.
751 [16] D.B. Rowe, Green roofs as a means of pollution abatement, Environ. Pollut. 159 (2011)
752 2100–2110.
753 [17] J. Yang, Q. Yu, P. Gong, Quantifying air pollution removal by green roofs in Chicago,
754 Atmos. Environ. 42 (2008) 7266–7273.
755 [18] H.S. Yang, J. Kang, M.S. Choi, Acoustic effects of green roof systems on a low-profiled
756 structure at street level, Build. Environ. 50 (2012) 44–55.
757 [19] T. Van Renterghem, D. Botteldooren, Numerical evaluation of sound propagating over
758 green roofs, J. Sound Vib. 317 (2008) 781–799.
759 [20] T. Brudermann, T. Sangkakool, Green roofs in temperate climate cities in Europe – An
760 analysis of key decision factors, Urban For. Urban Green. 21 (2017) 224–234.
761 https://doi.org/10.1016/j.ufug.2016.12.008.
762 [21] FLL - Landscape Development and Landscaping Research Society e.V., Gren Roof
763 Guidelines - Guidelines for the Planning, Construction and Maintenance of Green Roofs;,
764 Bonn, Germany, 2018.

39
765 [22] J. Mentens, D. Raes, M. Hermy, Green roofs as a tool for solving the rainwater runoff
766 problem in the urbanized 21st century?, Landsc. Urban Plan. 77 (2006) 217–226.
767 [23] A. Nagase, N. Dunnett, Drought tolerance in different vegetation types for extensive green
768 roofs: effects of watering and diversity, Landsc. Urban Plan. 97 (2010) 318–327.
769 [24] C.E. Thuring, R.D. Berghage, D.J. Beattie, Green roof plant responses to different
770 substrate types and depths under various drought conditions, HortTechnology. 20 (2010)
771 395–401.
772 [25] A.M. Coutts, E. Daly, J. Beringer, N.J. Tapper, Assessing practical measures to reduce
773 urban heat: Green and cool roofs, Build. Environ. 70 (2013) 266–276.
774 [26] E.C. Snodgrass, L.L. Snodgrass, Green roof plants: a resource and planting guide, Timber
775 Press Portland, 2006.
776 [27] C. Berretta, S. Poë, V. Stovin, Moisture content behaviour in extensive green roofs during
777 dry periods: The influence of vegetation and substrate characteristics, J. Hydrol. 511

of
778 (2014) 374–386.
779 [28] A. Nagase, N. Dunnett, Amount of water runoff from different vegetation types on

ro
780 extensive green roofs: Effects of plant species, diversity and plant structure, Landsc. Urban
781 Plan. 104 (2012) 356–363. https://doi.org/10.1016/j.landurbplan.2011.11.001.
782
783
-p
[29] S.-E. Ouldboukhitine, R. Belarbi, R. Djedjig, Characterization of green roof components:
measurements of thermal and hydrological properties, Build. Environ. 56 (2012) 78–85.
re
784 [30] T. Blanusa, M.M.V. Monteiro, F. Fantozzi, E. Vysini, Y. Li, R.W. Cameron, Alternatives
785 to Sedum on green roofs: can broad leaf perennial plants offer better ‘cooling service’?,
786 Build. Environ. 59 (2013) 99–106.
lP

787 [31] N.S. Williams, J. Lundholm, J. Scott MacIvor, Do green roofs help urban biodiversity
788 conservation?, J. Appl. Ecol. 51 (2014) 1643–1649.
na

789 [32] B. Minsker, L. Baldwin, J. Crittenden, K. Kabbes, M. Karamouz, K. Lansey, P.


790 Malinowski, E. Nzewi, Pandit Arka, Parker John, Rivera Samuel, Surbeck Cristiane,
791 Wallace William A., Williams John, Progress and Recommendations for Advancing
ur

792 Performance-Based Sustainable and Resilient Infrastructure Design, J. Water Resour. Plan.
793 Manag. 141 (2015) A4015006. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000521.
Jo

794 [33] Y. Li, R.W. Babcock, Green roof hydrologic performance and modeling: a review, Water
795 Sci. Technol. 69 (2014) 727–738.
796 [34] M. Akther, J. He, A. Chu, J. Huang, B. van Duin, A review of green roof applications for
797 managing urban stormwater in different climatic zones, Sustainability. 10 (2018) 2864.
798 [35] S. Vera, C. Pinto, P.C. Tabares-Velasco, W. Bustamante, A critical review of heat and
799 mass transfer in vegetative roof models used in building energy and urban enviroment
800 simulation tools, Appl. Energy. (2018).
801 [36] D.E. Bowler, L. Buyung-Ali, T.M. Knight, A.S. Pullin, Urban greening to cool towns and
802 cities: A systematic review of the empirical evidence, Landsc. Urban Plan. 97 (2010) 147–
803 155.
804 [37] B.Y. Schindler, L. Blank, S. Levy, G. Kadas, D. Pearlmutter, L. Blaustein, Integration of
805 photovoltaic panels and green roofs: review and predictions of effects on electricity
806 production and plant communities, Isr. J. Ecol. Evol. 62 (2016) 68–73.
807 https://doi.org/10.1080/15659801.2015.1048617.
808 [38] Y. Li, R.W. Babcock, Green roofs against pollution and climate change. A review, Agron.
809 Sustain. Dev. 34 (2014) 695–705. https://doi.org/10.1007/s13593-014-0230-9.

40
810 [39] S. Cascone, J. Coma, A. Gagliano, G. Pérez, The evapotranspiration process in green
811 roofs: A review, Build. Environ. 147 (2019) 337–355.
812 https://doi.org/10.1016/j.buildenv.2018.10.024.
813 [40] S. Cascone, Green Roof Design: State of the Art on Technology and Materials,
814 Sustainability. 11 (2019) 3020.
815 [41] B. Dvorak, A. Volder, Green roof vegetation for North American ecoregions: A literature
816 review, Landsc. Urban Plan. 96 (2010) 197–213.
817 https://doi.org/10.1016/j.landurbplan.2010.04.009.
818 [42] S.C. Cook-Patton, T.L. Bauerle, Potential benefits of plant diversity on vegetated roofs: A
819 literature review, J. Environ. Manage. 106 (2012) 85–92.
820 https://doi.org/10.1016/j.jenvman.2012.04.003.
821 [43] U. Berardi, A. GhaffarianHoseini, A. GhaffarianHoseini, State-of-the-art analysis of the
822 environmental benefits of green roofs, Appl. Energy. 115 (2014) 411–428.

of
823 [44] C.Y. Jim, Effect of vegetation biomass structure on thermal performance of tropical green
824 roof, Landsc. Ecol. Eng. 8 (2012) 173–187.

ro
825 [45] S.W. Peck, C. Callaghan, M.E. Kuhn, B. Bass, Greenbacks from green roofs: forging a
826 new industry in Canada, Citeseer, 1999.
827
828
-p
[46] I.R. Baciu, M.L. Lupu, S.G. Maxineasa, Principles of Green Roofs Design, Bull. Polytech.
Inst. Jassy Constr. Archit. Sect. 65 (2019) 63–76.
re
829 [47] A. Jahanfar, J. Drake, B. Sleep, B. Gharabaghi, A modified FAO evapotranspiration model
830 for refined water budget analysis for Green Roof systems, Ecol. Eng. 119 (2018) 45–53.
831 [48] Z. Peng, C. Smith, V. Stovin, Internal fluctuations in green roof substrate moisture content
lP

832 during storm events: Monitored data and model simulations, J. Hydrol. 573 (2019) 872–
833 884.
na

834 [49] M. Raviv, J.H. Lieth, M. Raviv, Significance of soilless culture in agriculture, Soil. Cult.
835 Theory Pract. (2008) 1–11.
836 [50] R. Field, D. Sullivan, A.N. Tafuri, Management of combined sewer overflows, CRC Press,
ur

837 2003.
838 [51] V. Stovin, G. Vesuviano, S. De-Ville, Defining green roof detention performance, Urban
Jo

839 Water J. 14 (2017) 574–588.


840 [52] T.L. Carter, T.C. Rasmussen, HYDROLOGIC BEHAVIOR OF VEGETATED ROOFS,
841 JAWRA J. Am. Water Resour. Assoc. 42 (2006) 1261–1274.
842 [53] T.D. Fletcher, Vg. Mitchell, A. Deletic, T.R. Ladson, A. Seven, Is stormwater harvesting
843 beneficial to urban waterway environmental flows?, Water Sci. Technol. 55 (2007) 265–
844 272.
845 [54] C.J. Walsh, T.D. Fletcher, A.R. Ladson, Stream restoration in urban catchments through
846 redesigning stormwater systems: looking to the catchment to save the stream, J. North Am.
847 Benthol. Soc. 24 (2005) 690–705.
848 [55] M.A. Bollman, G.E. DeSantis, R.M. DuChanois, M. Etten-Bohm, D.M. Olszyk, J.G.
849 Lambrinos, P.M. Mayer, A framework for optimizing hydrologic performance of green
850 roof media, Ecol. Eng. 140 (2019) 105589.
851 [56] M. Shafique, R. Kim, M. Rafiq, Green roof benefits, opportunities and challenges–A
852 review, Renew. Sustain. Energy Rev. 90 (2018) 757–773.
853 [57] V. Stovin, S. Poë, S. De-Ville, C. Berretta, The influence of substrate and vegetation
854 configuration on green roof hydrological performance, Ecol. Eng. 85 (2015) 159–172.
855 [58] B.G. Johannessen, T.M. Muthanna, B.C. Braskerud, Detention and retention behavior of
856 four extensive green roofs in three nordic climate zones, Water. 10 (2018) 671.

41
857 [59] L. Arias, R. Bournique, J.-C. Grimard, J.-L. Bertrand-Krajewski, Performances
858 hydrologiques de trois toitures végétalisées différentes, (2017).
859 [60] A.W. Sims, C.E. Robinson, C.C. Smart, J.A. Voogt, G.J. Hay, J.T. Lundholm, B. Powers,
860 D.M. O’Carroll, Retention performance of green roofs in three different climate regions, J.
861 Hydrol. 542 (2016) 115–124.
862 [61] A. Martilli, E.S. Krayenhoff, N. Nazarian, Is the urban heat island intensity relevant for
863 heat mitigation studies?, Urban Clim. 31 (2020) 100541.
864 [62] J. Dong, M. Lin, J. Zuo, T. Lin, J. Liu, C. Sun, J. Luo, Quantitative study on the cooling
865 effect of green roofs in a high-density urban Area—A case study of Xiamen, China, J.
866 Clean. Prod. (2020) 120152. https://doi.org/10.1016/j.jclepro.2020.120152.
867 [63] H. Takebayashi, M. Moriyama, Surface heat budget on green roof and high reflection roof
868 for mitigation of urban heat island, Build. Environ. 42 (2007) 2971–2979.
869 [64] D.J. Sailor, A green roof model for building energy simulation programs, Energy Build. 40

of
870 (2008) 1466–1478.
871 [65] A. Scherba, D.J. Sailor, T.N. Rosenstiel, C.C. Wamser, Modeling impacts of roof

ro
872 reflectivity, integrated photovoltaic panels and green roof systems on sensible heat flux
873 into the urban environment, Build. Environ. 46 (2011) 2542–2551.
874
875
-p
https://doi.org/10.1016/j.buildenv.2011.06.012.
[66] J. Heusinger, S. Weber, Surface energy balance of an extensive green roof as quantified by
re
876 full year eddy-covariance measurements, Sci. Total Environ. 577 (2017) 220–230.
877 [67] N. Meili, G. Manoli, P. Burlando, E. Bou-Zeid, W.T. Chow, A.M. Coutts, E. Daly, K.A.
878 Nice, M. Roth, N.J. Tapper, An urban ecohydrological model to quantify the effect of
lP

879 vegetation on urban climate and hydrology (UT&C v1. 0), Geosci. Model Dev. Discuss.
880 (2019).
na

881 [68] Y.-H. Ryu, E. Bou-Zeid, Z.-H. Wang, J.A. Smith, Realistic representation of trees in an
882 urban canopy model, Bound.-Layer Meteorol. 159 (2016) 193–220.
883 [69] Z. Wang, E. Bou‐Zeid, J.A. Smith, A coupled energy transport and hydrological model
ur

884 for urban canopies evaluated using a wireless sensor network, Q. J. R. Meteorol. Soc. 139
885 (2013) 1643–1657.
Jo

886 [70] C. De Munck, A. Lemonsu, R. Bouzouidja, V. Masson, R. Claverie, The GREENROOF


887 module (v7. 3) for modelling green roof hydrological and energetic performances within
888 TEB, Geosci. Model Dev. 6 (2013) 1941–1960.
889 [71] A. Solcerova, F. van de Ven, M. Wang, M. Rijsdijk, N. van de Giesen, Do green roofs cool
890 the air?, Build. Environ. 111 (2017) 249–255.
891 [72] O. Schweitzer, E. Erell, Evaluation of the energy performance and irrigation requirements
892 of extensive green roofs in a water-scarce Mediterranean climate, Energy Build. 68 (2014)
893 25–32.
894 [73] M. Köhler, M. Schmidt, F. Wilhelm Grimme, M. Laar, V. Lúcia de Assunção Paiva, S.
895 Tavares, Green roofs in temperate climates and in the hot-humid tropics–far beyond the
896 aesthetics, Environ. Manag. Health. 13 (2002) 382–391.
897 [74] C.L. Tan, P.Y. Tan, N.H. Wong, H. Takasuna, T. Kudo, Y. Takemasa, C.V.J. Lim, H.X.V.
898 Chua, Impact of soil and water retention characteristics on green roof thermal
899 performance, Energy Build. 152 (2017) 830–842.
900 [75] C.Y. Jim, H. He, Coupling heat flux dynamics with meteorological conditions in the green
901 roof ecosystem, Ecol. Eng. 36 (2010) 1052–1063.

42
902 [76] L.L. Peng, C. Jim, Economic evaluation of green-roof environmental benefits in the
903 context of climate change: The case of Hong Kong, Urban For. Urban Green. 14 (2015)
904 554–561.
905 [77] A. Dwivedi, B.K. Mohan, Impact of green roof on micro climate to reduce Urban Heat
906 Island, Remote Sens. Appl. Soc. Environ. 10 (2018) 56–69.
907 [78] R. Djedjig, S.-E. Ouldboukhitine, R. Belarbi, E. Bozonnet, Development and validation of
908 a coupled heat and mass transfer model for green roofs, Int. Commun. Heat Mass Transf.
909 39 (2012) 752–761. https://doi.org/10.1016/j.icheatmasstransfer.2012.03.024.
910 [79] J. Sonne, Evaluating green roof energy performance, ASHRAE J. 48 (2006) 59.
911 [80] C.Y. Jim, Effect of vegetation biomass structure on thermal performance of tropical green
912 roof, Landsc. Ecol. Eng. 8 (2012) 173–187.
913 [81] U. Berardi, The outdoor microclimate benefits and energy saving resulting from green
914 roofs retrofits, Energy Build. 121 (2016) 217–229.

of
915 https://doi.org/10.1016/j.enbuild.2016.03.021.
916 [82] A.M. Broadbent, A.M. Coutts, N.J. Tapper, M. Demuzere, J. Beringer, The microscale

ro
917 cooling effects of water sensitive urban design and irrigation in a suburban environment,
918 Theor. Appl. Climatol. 134 (2018) 1–23.
919
920
-p
[83] D. Evans, Simplified method for predicting photovoltaic array output, Sol. Energy. 27
(1981) 555–560.
re
921 [84] G. Makrides, B. Zinsser, G.E. Georghiou, M. Schubert, J.H. Werner, Temperature
922 behaviour of different photovoltaic systems installed in Cyprus and Germany, Sol. Energy
923 Mater. Sol. Cells. 93 (2009) 1095–1099.
lP

924 [85] T. Baumann, D. Schär, F. Carigiet, A. Dreisiebner, F. Baumgartner, Performance Analysis


925 of PV Green Roof Systems, in: 2016.
na

926 [86] T. Baumann, F. Carigiet, R. Knecht, M. Klenk, A. Dreisiebner, H. Nussbaumer, F.


927 Baumgartner, Performance analysis of vertically mounted bifacial PV modules on green
928 roof system, in: WIP, 2018.
ur

929 [87] T. Baumann, H. Nussbaumer, M. Klenk, A. Dreisiebner, F. Carigiet, F. Baumgartner,


930 Photovoltaic systems with vertically mounted bifacial PV modules in combination with
Jo

931 green roofs, Sol. Energy. 190 (2019) 139–146.


932 https://doi.org/10.1016/j.solener.2019.08.014.
933 [88] M. Perez, N. Wight, V. Fthenakis, C. Ho, Green-roof integrated PV canopies–an empirical
934 study and teaching tool for low income students in the South Bronx, in: 2012: p. 6.
935 [89] A. Nagengast, C. Hendrickson, H.S. Matthews, Variations in photovoltaic performance
936 due to climate and low-slope roof choice, Energy Build. 64 (2013) 493–502.
937 [90] S.C. Hui, S. Chan, Integration of green roof and solar photovoltaic systems, in: 2011: pp.
938 1–12.
939 [91] M. Köhler, W. Wiartalla, R. Feige, Interaction between PV-systems and extensive green
940 roofs, in: 2007: pp. 1–10.
941 [92] H. Ogaili, D.J. Sailor, Measuring the effect of vegetated roofs on the performance of
942 photovoltaic panels in a combined system, J. Sol. Energy Eng. 138 (2016) 061009.
943 [93] G. Osma-Pinto, G. Ordóñez-Plata, Measuring factors influencing performance of rooftop
944 PV panels in warm tropical climates, Sol. Energy. 185 (2019) 112–123.
945 [94] V.A. Dakwale, R.V. Ralegaonkar, S. Mandavgane, Improving environmental performance
946 of building through increased energy efficiency: A review, Sustain. Cities Soc. 1 (2011)
947 211–218.

43
948 [95] M. Song, F. Niu, N. Mao, Y. Hu, S. Deng, Review on building energy performance
949 improvement using phase change materials, Energy Build. 158 (2018) 776–793.
950 [96] R. Pacheco, J. Ordóñez, G. Martínez, Energy efficient design of building: A review,
951 Renew. Sustain. Energy Rev. 16 (2012) 3559–3573.
952 [97] P.C. Tabares-Velasco, Predictive heat and mass transfer model of plant-based roofing
953 materials for assessment of energy savings (Ph. D. thesis), Ph.D. Thesis, Dept. of
954 Architectural Engineering, Pennsylvania State University, 2009.
955 [98] P.C. Tabares-Velasco, M. Zhao, N. Peterson, J. Srebric, R. Berghage, Validation of
956 predictive heat and mass transfer green roof model with extensive green roof field data,
957 Ecol. Eng. 47 (2012) 165–173.
958 [99] P.C. Tabares-Velasco, J. Srebric, A heat transfer model for assessment of plant based
959 roofing systems in summer conditions, Build. Environ. 49 (2012) 310–323.
960 [100] P.-Y. Chen, Y.-H. Li, W.-H. Lo, C. Tung, Toward the practicability of a heat transfer

of
961 model for green roofs, Ecol. Eng. 74 (2015) 266–273.
962 https://doi.org/10.1016/j.ecoleng.2014.09.114.

ro
963 [101] R.M. Lazzarin, F. Castellotti, F. Busato, Experimental measurements and numerical
964 modelling of a green roof, Energy Build. 37 (2005) 1260–1267.
965
966
-p
[102] H. He, C.Y. Jim, Simulation of thermodynamic transmission in green roof ecosystem,
Ecol. Model. 221 (2010) 2949–2958.
re
967 [103] M. Squier, C.I. Davidson, Heat flux and seasonal thermal performance of an extensive
968 green roof, Build. Environ. 107 (2016) 235–244.
969 [104] S.S. Moody, D.J. Sailor, Development and application of a building energy performance
lP

970 metric for green roof systems, Energy Build. 60 (2013) 262–269.
971 [105] C. Guo, Research on the impact of sedum lineare planted roof on the indoor and outdoor
na

972 thermal environment in Guangzhou, Univ. South China Univ. Technol. Press Guangzhou.
973 (2008).
974 [106] K. Liu, J. Minor, Performance evaluation of an extensive green roof, Present. Green
ur

975 Rooftops Sustain. Communities Wash. DC. (2005) 1–11.


976 [107] A. Niachou, K. Papakonstantinou, M. Santamouris, A. Tsangrassoulis, G. Mihalakakou,
Jo

977 Analysis of the green roof thermal properties and investigation of its energy performance,
978 Energy Build. 33 (2001) 719–729.
979 [108] N.H. Wong, D.W. Cheong, H. Yan, J. Soh, C. Ong, A. Sia, The effects of rooftop garden
980 on energy consumption of a commercial building in Singapore, Energy Build. 35 (2003)
981 353–364.
982 [109] H. Kasmin, V. Stovin, E. Hathway, Towards a generic rainfall-runoff model for green
983 roofs, Water Sci. Technol. 62 (2010) 898–905.
984 [110] E. Voyde, E. Fassman, R. Simcock, J. Wells, Quantifying evapotranspiration rates for New
985 Zealand green roofs, J. Hydrol. Eng. 15 (2010) 395–403.
986 [111] R. Liu, E. Fassman-Beck, Pore structure and unsaturated hydraulic conductivity of
987 engineered media for living roofs and bioretention based on water retention data, J.
988 Hydrol. Eng. 23 (2018) 04017065.
989 [112] A. Palla, I. Gnecco, L.G. Lanza, Unsaturated 2D modelling of subsurface water flow in the
990 coarse-grained porous matrix of a green roof, J. Hydrol. 379 (2009) 193–204.
991 [113] H.-H. Liu, Generalization of the Darcy-Buckingham Law: Optimality and Water Flow in
992 Unsaturated Media, in: H.-H. Liu (Ed.), Fluid Flow Subsurf. Hist. Gen. Appl. Phys. Laws,
993 Springer International Publishing, Cham, 2017: pp. 45–102. https://doi.org/10.1007/978-3-
994 319-43449-0_2.

44
995 [114] L.A. Richards, Capillary conduction of liquids through porous mediums, Physics. 1 (1931)
996 318–333.
997 [115] R. Singh, J.B. Franzini, Unsteady flow in unsaturated soils from a cylindrical source of
998 finite radius, J. Geophys. Res. 72 (1967) 1207–1215.
999 [116] E. Buckingham, Studies on the movement of soil moisture, US Dept Agic Bur Soils Bull.
1000 38 (1907).
1001 [117] D. Schneider, B.M. Wadzuk, R.G. Traver, Using a weighing lysimeter to determine a crop
1002 coefficient for a green roof to predict evapotranspiration with the FAO standardized
1003 Penman-Monteith equation, in: 2011: pp. 3629–3638.
1004 [118] N. She, J. Pang, Physically based green roof model, J. Hydrol. Eng. 15 (2010) 458–464.
1005 [119] V. Stovin, S. Poë, C. Berretta, A modelling study of long term green roof retention
1006 performance, J. Environ. Manage. 131 (2013) 206–215.
1007 [120] I.A. Walter, R.G. Allen, R. Elliott, M. Jensen, D. Itenfisu, B. Mecham, T. Howell, R.

of
1008 Snyder, P. Brown, S. Echings, ASCE’s standardized reference evapotranspiration equation,
1009 in: Watershed Manag. Oper. Manag. 2000, 2000: pp. 1–11.

ro
1010 [121] P.C. Tabares-Velasco, J. Srebric, Experimental quantification of heat and mass transfer
1011 process through vegetated roof samples in a new laboratory setup, Int. J. Heat Mass Transf.
1012
1013
54 (2011) 5149–5162. -p
[122] P.G. Jarvis, K.G. McNaughton, Stomatal control of transpiration: scaling up from leaf to
re
1014 region, in: Adv. Ecol. Res., Elsevier, 1986: pp. 1–49.
1015 [123] D. Hillel, Environmental soil physics. Academic Press, San Diego., Environ. Soil Phys.
1016 Acad. Press San Diego. (1998).
lP

1017 [124] R.G. Allen, L.S. Pereira, D. Raes, M. Smith, Crop evapotranspiration-Guidelines for
1018 computing crop water requirements-FAO Irrigation and drainage paper 56, Fao Rome. 300
na

1019 (1998) D05109.


1020 [125] T.R. Oke, Boundary layer climates, Routledge, 2002.
1021 [126] J.G. Lockwood, Causes of climate, Wiley, 1979.
ur

1022 [127] S. Verma, Aerodynamic resistances to transfers of heat, mass and momentum, Estim. Areal
1023 Evapotranspiration. 177 (1989) 13–20.
Jo

1024 [128] S. Gaffin, C. Rosenzweig, L. Parshall, D. Beattie, R. Berghage, G. O’Keeffe, D. Braman,


1025 Energy balance modeling applied to a comparison of white and green roof cooling
1026 efficiency, Green Roofs N. Y. Metrop. Reg. Res. Rep. 7 (2010).
1027 [129] D.J. Sailor, D. Hutchinson, L. Bokovoy, Thermal property measurements for ecoroof soils
1028 common in the western US, Energy Build. 40 (2008) 1246–1251.
1029 [130] R.A. Pielke Sr, Mesoscale meteorological modeling, Academic press, 2013.
1030 [131] E.P. Del Barrio, Analysis of the green roofs cooling potential in buildings, Energy Build.
1031 27 (1998) 179–193.
1032 [132] J.L. Monteith, M.H. Unsworth, Chapter 6 - Microclimatology of Radiation: (i) Radiative
1033 Properties of Natural Materials, in: J.L. Monteith, M.H. Unsworth (Eds.), Princ. Environ.
1034 Phys. Fourth Ed., Academic Press, Boston, 2013: pp. 81–93. https://doi.org/10.1016/B978-
1035 0-12-386910-4.00006-8.
1036 [133] E. Haghighi, J.W. Kirchner, Near‐surface turbulence as a missing link in modeling
1037 evapotranspiration‐soil moisture relationships, Water Resour. Res. 53 (2017) 5320–5344.
1038 [134] E.P. Del Barrio, Analysis of the green roofs cooling potential in buildings, Energy Build.
1039 27 (1998) 179–193.

45
1040 [135] N.D. VanWoert, D.B. Rowe, J.A. Andresen, C.L. Rugh, R.T. Fernandez, L. Xiao, Green
1041 roof stormwater retention: effects of roof surface, slope, and media depth, J. Environ. Qual.
1042 34 (2005) 1036–1044.
1043 [136] E. Fassman-Beck, E. Voyde, R. Simcock, Y.S. Hong, 4 Living roofs in 3 locations: Does
1044 configuration affect runoff mitigation?, J. Hydrol. 490 (2013) 11–20.
1045 [137] C.Y. Jim, L.L. Peng, Substrate moisture effect on water balance and thermal regime of a
1046 tropical extensive green roof, Ecol. Eng. 47 (2012) 9–23.
1047 [138] Y. He, H. Yu, A. Ozaki, N. Dong, S. Zheng, Influence of plant and soil layer on energy
1048 balance and thermal performance of green roof system, Energy. 141 (2017) 1285–1299.
1049 [139] V. Azeñas, J. Cuxart, R. Picos, H. Medrano, G. Simó, A. López-Grifol, J. Gulías, Thermal
1050 regulation capacity of a green roof system in the mediterranean region: The effects of
1051 vegetation and irrigation level, Energy Build. 164 (2018) 226–238.
1052 [140] C. Feng, Q. Meng, Y. Zhang, Theoretical and experimental analysis of the energy balance

of
1053 of extensive green roofs, Energy Build. 42 (2010) 959–965.
1054 [141] G. Manoli, S. Fatichi, M. Schläpfer, K. Yu, T.W. Crowther, N. Meili, P. Burlando, G.G.

ro
1055 Katul, E. Bou-Zeid, Magnitude of urban heat islands largely explained by climate and
1056 population, Nature. 573 (2019) 55–60. https://doi.org/10.1038/s41586-019-1512-9.
1057
1058
-p
[142] M.H. Yio, V. Stovin, J. Werdin, G. Vesuviano, Experimental analysis of green roof
substrate detention characteristics, Water Sci. Technol. 68 (2013) 1477–1486.
re
1059 [143] R. Wallach, Physical characteristics of soilless media, in: Soil. Cult., Elsevier, 2019: pp.
1060 33–112.
1061 [144] S.E. Allaire, J. Caron, I. Duchesne, L.-É. Parent, J.-A. Rioux, Air-filled porosity, gas
lP

1062 relative diffusivity, and tortuosity: Indices of Prunus× cistena sp. growth in peat substrates,
1063 J. Am. Soc. Hortic. Sci. 121 (1996) 236–242.
na

1064 [145] R. Wallach, F. Da Silva, Y. Chen, Unsaturated hydraulic characteristics of composted


1065 agricultural wastes, tuff, and their mixtures, Soil Sci. 153 (1992) 434–441.
1066 [146] K. Metselaar, Water retention and evapotranspiration of green roofs and possible natural
ur

1067 vegetation types, Resour. Conserv. Recycl. 64 (2012) 49–55.


1068 [147] P.A. Nektarios, I. Amountzias, I. Kokkinou, N. Ntoulas, Green roof substrate type and
Jo

1069 depth affect the growth of the native species Dianthus fruticosus under reduced irrigation
1070 regimens, HortScience. 46 (2011) 1208–1216.
1071 [148] A. Graceson, M. Hare, J. Monaghan, N. Hall, The water retention capabilities of growing
1072 media for green roofs, Ecol. Eng. 61 (2013) 328–334.
1073 [149] C. Farrell, X.Q. Ang, J.P. Rayner, Water-retention additives increase plant available water
1074 in green roof substrates, Ecol. Eng. 52 (2013) 112–118.
1075 [150] I. Teare, E. Kanemasu, W. Powers, H. Jacobs, Water‐Use Efficiency and Its Relation to
1076 Crop Canopy Area, Stomatal Regulation, and Root Distribution 1, Agron. J. 65 (1973)
1077 207–211.
1078 [151] S. Hodo-Abalo, M. Banna, B. Zeghmati, Performance analysis of a planted roof as a
1079 passive cooling technique in hot-humid tropics, Renew. Energy. 39 (2012) 140–148.
1080 [152] A. Teemusk, Ü. Mander, Rainwater runoff quantity and quality performance from a
1081 greenroof: The effects of short-term events, Ecol. Eng. 30 (2007) 271–277.
1082 [153] J.S. MacIvor, L. Margolis, M. Perotto, J.A. Drake, Air temperature cooling by extensive
1083 green roofs in Toronto Canada, Ecol. Eng. 95 (2016) 36–42.
1084 [154] H.G. Jones, Plants and microclimate: a quantitative approach to environmental plant
1085 physiology, Cambridge university press, 2013.

46
1086 [155] J. Lundholm, J.S. MacIvor, Z. MacDougall, M. Ranalli, Plant species and functional group
1087 combinations affect green roof ecosystem functions, PloS One. 5 (2010) e9677.
1088 [156] D.G. Cirkel, B.R. Voortman, T. Van Veen, R.P. Bartholomeus, Evaporation from (blue-)
1089 green roofs: Assessing the benefits of a storage and capillary irrigation system based on
1090 measurements and modeling, Water. 10 (2018) 1253.
1091 [157] W. Larcher, Physiological plant ecology: ecophysiology and stress physiology of
1092 functional groups, Springer Science & Business Media, 2003.
1093 [158] J.S. MacIvor, M.A. Ranalli, J.T. Lundholm, Performance of dryland and wetland plant
1094 species on extensive green roofs, Ann. Bot. 107 (2011) 671–679.
1095 [159] C. Loiola, W. Mary, L.P. da Silva, Hydrological performance of modular-tray green roof
1096 systems for increasing the resilience of mega-cities to climate change, J. Hydrol. 573
1097 (2019) 1057–1066.
1098 [160] R. Berghage, A. Jarrett, D. Beattie, K. Kelley, S. Husain, F. Rezai, B. Long, A. Negassi, R.

of
1099 Cameron, W. Hunt, Quantifying evaporation and transpirational water losses from green
1100 roofs and green roof media capacity for neutralizing acid rain, Natl. Decentralized Water

ro
1101 Resour. Capacity Dev. Proj. (2007).
1102 [161] M.E. Hulley, 5 - The urban heat island effect: causes and potential solutions, in: F. Zeman
1103
1104
-p
(Ed.), Metrop. Sustain., Woodhead Publishing, 2012: pp. 79–98.
https://doi.org/10.1533/9780857096463.1.79.
re
1105 [162] M. Zhao, P.C. Tabares-Velasco, J. Srebric, S. Komarneni, R. Berghage, Effects of plant
1106 and substrate selection on thermal performance of green roofs during the summer, Build.
1107 Environ. 78 (2014) 199–211.
lP

1108 [163] Chr. Lamnatou, D. Chemisana, A critical analysis of factors affecting photovoltaic-green
1109 roof performance, Renew. Sustain. Energy Rev. 43 (2015) 264–280.
na

1110 https://doi.org/10.1016/j.rser.2014.11.048.
1111 [164] N. Sugathan, V. Biju, G. Renuka, Influence of soil moisture content on surface albedo and
1112 soil thermal parameters at a tropical station, J. Earth Syst. Sci. 123 (2014) 1115–1128.
ur

1113 [165] J. Monteith, M. Unsworth, Principles of environmental physics: plants, animals, and the
1114 atmosphere, Academic Press, 2013.
Jo

1115 [166] S. Kotthaus, T. Smith, M. Wooster, C. Grimmond, Derivation of an urban materials


1116 spectral library through emittance and reflectance spectroscopy, ISPRS J. Photogramm.
1117 Remote Sens. (2014) 194–212.
1118 http://www.met.reading.ac.uk/micromet/LUMA/SLUM.html (accessed October 19, 2020).
1119 [167] Y. Zhang, MODIS UCSB Emissivity Library, MODIS. (1999).
1120 https://icess.eri.ucsb.edu/modis/EMIS/html/em.html (accessed October 19, 2020).
1121 [168] CIBSE, Heat transfer, in: Ref. Data, Elsevier, London, 2001: pp. 3–11.
1122 [169] J.A. Sobrino, C. Mattar, P. Pardo, J.C. Jiménez-Muñoz, S.J. Hook, A. Baldridge, R.
1123 Ibañez, Soil emissivity and reflectance spectra measurements, Appl. Opt. 48 (2009) 3664–
1124 3670.
1125 [170] S. Vera, C. Pinto, P.C. Tabares-Velasco, W. Bustamante, F. Victorero, J. Gironás, C.A.
1126 Bonilla, Influence of vegetation, substrate, and thermal insulation of an extensive
1127 vegetated roof on the thermal performance of retail stores in semiarid and marine climates,
1128 Energy Build. 146 (2017) 312–321.
1129 [171] A. Dai, Drought under global warming: A review, Wiley Interdiscip. Rev. Clim. Change. 2
1130 (2011) 45–65. http://onlinelibrary.wiley.com/doi/10.1002/wcc.81/full.

47
1131 [172] E.M. Fischer, C. Schar, Future changes in daily summer temperature variability: driving
1132 processes and role for temperature extremes, Clim. Dyn. 33 (2009) 917–935.
1133 https://doi.org/10.1007/S00382-008-0473-8.
1134 [173] A.M. Broadbent, E.S. Krayenhoff, M. Georgescu, D.J. Sailor, The observed effects of
1135 utility-scale photovoltaics on near-surface air temperature and energy balance, J. Appl.
1136 Meteorol. Climatol. 58 (2019) 989–1006.
1137 [174] M. Convertino, K. Baker, J. Vogel, C. Lu, B. Suedel, I. Linkov, Multi-criteria decision
1138 analysis to select metrics for design and monitoring of sustainable ecosystem restorations,
1139 Ecol. Indic. 26 (2013) 76–86.
1140 [175] V. Belton, T. Stewart, Multiple criteria decision analysis: an integrated approach, Springer
1141 Science & Business Media, 2002.
1142 [176] H. Qin, Y. Peng, Q. Tang, S.-L. Yu, A HYDRUS model for irrigation management of
1143 green roofs with a water storage layer, Ecol. Eng. 95 (2016) 399–408.

of
1144 [177] B. Hörnschemeyer, M. Henrichs, M. Uhl, Setting up a SWMM-integrated Model for the
1145 Evapotranspiration of Urban Vegetation, in: 2019: pp. 1–4.

ro
1146 [178] N. Blair, A.P. Dobos, J. Freeman, T. Neises, M. Wagner, System Advisor Model, SAM
1147 2014.1.14: General Description, National Renewable Energy Laboratory, Golden, CO,
1148
1149
-p
USA, 2014. https://www.nrel.gov/docs/fy14osti/61019.pdf.
[179] USEPA, EPA Storm Water Management Model, 2015. http://www.epa.gov/water-
re
1150 research/storm-water-management-model-swmm.
1151
lP
na
ur
Jo

48
1 Appendix

2 Towards a performance based approach for multifunctional green roofs: an

3 interdisciplinary review

4 Lauren M. Cook1*, Tove A. Larsen1

1
5 Department of Urban Water Management, Swiss Federal Institute for Aquatic Research,

6 Dübendorf, Switzerland

7 *Corresponding author. Postal address: Eawag, Überland Str. 133, 8600 Dübendorf,

of
8 Switzerland

ro
9 E-mail address: Lauren.cook@eawag.ch

10
-p
re
11
lP

12
na
ur
Jo

1
13 A.1. Volumetric water content

14 The volumetric water content (VMC) of the green roof layer, Θ, referred to as soil

15 moisture in the substrate, is the percent of stored water (Vw) filling the total volume of pore

16 space (Vp), as shown in Eqn. A1.

17 = / (A1)

18 When pores are completed filled (Vw=Vp), the media is considered saturated.

19 However, since green roof substrate is highly porous and water infiltrates quickly, saturation

of
20 is never expected to occur [1]. This is why surface runoff and ponding are also not expected to

ro
21 take place [2].

22
-p
re
23 A.2. Soil properties and Porosity
lP

24 The particle size distribution (PSD), which is the abundance and sizes of soil particles,
na

25 defines the soil texture and determines internal geometry, porosity, and interactions with
ur

26 fluids[3]. The PSD is used to estimate hydraulic properties, including the saturated and
Jo

27 unsaturated hydraulic conductivity. However, these relationships are lacking for soilless

28 media (often found in green roofs) [3].

29 Large pores (macropores) allow roots, gas, and water to enter the substrate.

30 Connected, or continuous macropores (related to the pore distribution) directly affect the rate

31 of infiltration of water and benefits root growth[4]. The connectedness of pore space

32 influences the transport of water (hydraulic conductivity). The capacity of a medium to store

33 and convey water depends on the pore size distribution, tortuosity, and connectivity[5].

34 Tortuosity is the ratio of the “effective average path in the porous medium to the shortest

35 distance measured along the direction of the pore” (Jury et al., 1991). It expresses how

36 complex the pore path is and indicates resistance to flow. Connectivity measures the number

37 of independent paths between two points. Tortuosity is an average measure of how many
2
38 turns are in each path. The water retention curve is nearly equivalent to the pore-size

39 distribution curve.

40 Substrates have different abilities to retain water depending on the physical and

41 chemical properties of the media. The affinity to retain water is defined as the total potential

42 or the pressure head (if it is divided by the bulk density) and relates to capillary pressure

43 exerted by pores. Texture of media used to grow plants is coarse with low air-entry values that

44 enable fast free drainage. This fast drainage enables air to replace draining water and prevents

45 clogging. The coarseness of the media should be lower than the container height.

of
46 Porosity, which related to coarseness of the media, is calculated as the fraction of the

ro
47 volume of void space relative to the total volume of soil (Eqn. A2).

48 =
-p ( )
(A2)
( )
re
lP

49 It depends on the material structure, which is a function of the texture and composition

50 (distribution of the grain size) of particles present in the material [6]. Porosity is another way
na

51 to describe the density of soil, although this relationship is complex in real soil systems that
ur

52 contain a variety of grain sizes and textures. For an ideal systems of balls, density is related to
Jo

53 the porosity as follows [7]:

54 = (1 − ρ/ρ ) (A3)

55 Where ρ is the density and ρ is the specific gravity of the particles.

56

57 A.3. Richards Equation

58 The rate of water drainage through the substrate can be evaluated in terms of the

59 change in volumetric water content over time. Due to its ability to represent moisture content

60 in variably saturated media, Richards’ equation has been employed to represent change in

61 moisture in green roof substrate [1, 8, 9]. The 2-D Richard’s equation (encompassing both

62 horizontal and vertical flow) and the 1-D form (vertical flow only), shown in Eqn. A4, have

3
63 both been employed in green roof studies; it is unclear whether the added complexity of the 2-

64 D equation is needed for green roofs because no comparison studies have been performed.

65 The 1-D change in moisture depends on the vertical pressure gradient as well as how well soil

66 pore spaces are connected, characterized using the unsaturated hydraulic conductivity, K (also

67 known as permeability).

! (
68 = #$% (ℎ) ∙ ( − 1)) (A4)
" "

69 where * is the volumetric water content (VMC), K(h) is the unsaturated hydraulic

of
70 conductivity at suction head h, and dz is the elevation of the point relative to the reference

ro
71 level.

72 -p
73 A.4. Hydraulic conductivity and Matric Potential
re
74 The hydraulic conductivity at saturation (Ks) depends on the viscosity (+) and specific
lP

75 weight (,) of the fluid, as well as, the pore size and other properties of the media. Since we
na

76 are only interested in one fluid, water, we can separate out the fluid properties to define the
ur

77 term, specific permeability (ks), which is dependent on the pore size and how well the pores
Jo

78 are connected.

79 Using a representative grain diameter (dg) to substitute for pore size that cannot be

80 directly measured, specific permeability is defined as follows:

81 -. = /01 (A5)

82 Where C is a dimensionless proportionality constant that indicates other properties of the

83 porous medium like particle distribution (230 ) and soil structure (4).

84 For a number of natural soils, a strong correlation between specific permeability and the

85 representative grain diameter can be shown (Figure A1, [6]).

86

4
of
ro
87
88 Figure A1. Relationship between specific permeability and representation grain size of substrates
-p
re
89 Under transient flow, unsaturated hydraulic conductivity is a function of these

90 properties as well as the total soil potential (5). Total potential is made of up several
lP

91 contributions exerting pressure, including the matric potential and osmotic potential (related
na

92 to suction from roots). Matric potential relates to adsorptive and capillary forces of the
ur

93 medium substrate properties and is dynamic. The rate that matric potential decreases
Jo

94 (increasing tension) depends on how fast water is depleted through evapotranspiration. The

95 relationship between matric potential and volumetric water content is described by the soil-

96 water characteristic (SWC) curve or moisture retention curve (RC). SWC relates to the pore

97 space distribution and organic matter content [3]. SWCs are non-linear and difficult to obtain,

98 but are important to estimate water available to plants and to model water flow.

99 As water is drained or suction is applied to substrate, larger pores are emptied first

100 because they cannot retain water against the suction. As suction increases, progressively

101 smaller pores are emptied until only the smallest retain water. The water retained as

102 equilibrium is a function of the sizes and volume of water filled pores. Soilless-growing

103 media notably have a high value of saturated water content. The rate that moisture content

5
104 decrease per unit of increase in suction is very steep compared to soils, meaning that water is

105 lost very quickly [3]. [3] summarizes sources of curves for SWC and unsaturated hydraulic

106 conductivity for different types of soilless media. This difference is not only due to

107 differences in particle size, but also due to porosity. This difference in porosity is linked to the

108 inner microporosity and particle shapes that are not present in soils. For instance, sand

109 particles (that represent typical soil) are regular and smooth, but some soilless-media particles

110 are rough and irregular. At the same matric potential, sand has been shown to have a lower

111 moisture content than tuff (also known scoria), a soilless-media [3, 10].

of
112

ro
113 A.5. Evapotranspiration using Penman-Monteith

114
-p
In addition to the vapor pressure differential and wind speed, many factors also
re
115 contribute to evapotranspiration, including diurnal variability, meteorological conditions, and
lP

116 vegetation properties. Over time, many researchers, including Penman and Monteith, have
na

117 contributed to the ET equation in order to account for these factors. Due to its relatively

118 accurate performance across North America and Europe, the most widely used equation for
ur

119 modeling ET is the FAO Penman-Monteith (P-M) reference equation [11] shown in Eqn. A6.
Jo

(BC DB@ )
9(:; <=)>?@ A
120 678 = E
E@
(A6)
9>F(G> C )
E@

121 where: Rn is net radiation, G is the soil heat flux, (es - ea) is the VPD, H is mean air density

122 at constant pressure, cp is the specific heat of air, Δ is the slope of the saturation vapor

123 pressure temperature relationship, , is the psychrometric constant, rs is the surface resistance

124 (a function of vegetation and moisture), ra is the aerodynamic resistances to mass transfer

125 (related to the vegetation and wind), and 6 is the latent heat of vaporization.

126 Of particular importance is the relationship between stomatal resistance and soil

127 moisture (VMC). A lack of moisture limits stomatal opening; however, this is not explicitly

128 accounted for in the P-M model because adequate moisture is assumed to be available for ET

6
129 to take place. Disregard of fluctuating soil moisture conditions could result in systematic

130 errors of estimated values[12]. The only way to account for a soil moisture deficit in P-M is to

131 incorporate it into the stomatal resistance term.

132 P-M assumes that vegetation densely covers the surface and, subsequently, that

133 transpiration accounts for 100% of ET. However, when bare soil is exposed to the

134 atmosphere, which can be the case in green roofs, evaporation from soil also contributes to

135 overall ET. Thus, a parameter that must be incorporated into the Penman-Monteith equation is

136 the fractional vegetation coverage (η). To account for it, the calculation of ET is separated

of
137 into two steps: (1) calculation of transpiration from plants using a surface resistance (rs) for

ro
138 vegetation, and (2) calculation of evaporation from the soil using a surface resistance for soil.

139
-p
ET is then calculated as the weighted sum of these two values, where the weights are the
re
140 fractional vegetation coverage (η) and (1 - η), respectively, for transpiration and evaporation.
lP

141 To avoid the simplifications and assumptions of the P-M model, other models exist
na

142 that simulate the physical processes related to stomata opening and closing (e.g., Haghighi

143 and Kirchner (2017)[13]); however, these complexities may not be necessary in order to
ur

144 estimate the energy and water balance of a green roof.


Jo

145 A.6. Surface resistance

146 The surface resistance, rs (Eqn. AA77. [11]) describes the resistance due to the bulk

147 physiologic properties of the vegetation, which include the stomatal resistance, rI, and the leaf

148 area index, LAI [14]. The surface resistance generally decreases with increasing solar

149 radiation, increases as soil water content decreases, and is independent from wind speed[15].

150 J. = JK /LMN A (A7)

151 LAI, which is a function of vegetation type and season, is the projected leaf area per

152 unit ground area (m2/m2), e.g., a value of 3 if a unit area of roof is covered by a plant stem

153 with three leaves of the same unit area. Active LAI is the amount of leaf area that is actively

154 contributing to the surface heat and vapor transfer and is typically calculated as one half of
7
155 LAI [11]. Stomatal resistance, rI, is the force opposing the release of water vapor through leaf

156 stomata. It is also a function of vegetation type and varies over time as a particular plant

157 responds to changes in environmental conditions [16] and soil type [17]. Equations for

158 stomatal resistance that include these functions can be found in the literature (e.g., [18–20]

159 and references therein).

160

161 A.7. Aerodynamic resistance

162 Aerodynamic resistance, ra, is broadly used to describe the resistance to turbulent

of
163 fluxes between a surface and the ambient air [21]. The resistance form changes depending on

ro
164 the type of flux that is transferred: vapor (rav), heat (rah), and momentum (ram). The

165
-p
aerodynamic resistance for latent heat (evapotranspiration), where water is exchanged,
re
166 corresponds to the aerodynamic resistance to vapor or mass transport (rav). The resistance to
lP

167 sensible heat exchange corresponds to rah. Both are a function of wind speed (u) and
na

168 Karman’s constant (ka, 0.41)[22]. The Penman-Monteith (P-M) equation assumes that ra = rav

169 = rah [21], which is usually applicable [15]. Aerodynamic resistance to vapor and heat
ur

170 transport, ra, for use in the P-M equation [11], is shown in Eqn. A8.
Jo

P DR P DR
O# Q S ) OV W S X
TUQ TUW
171 J = YZ@
(A8)

172 where zm is the height of the wind measurements; zh is the height of humidity or temperature

173 measurements; d0 is the zero-plane displacement height; zom is the roughness length governing

174 momentum transfer; zoh is the roughness length governing heat and vapor transfer; ka is

175 Karman’s constant, 0.41, and u is wind speed at height zm.

176 The resistance is determined by the aerodynamic roughness lengths (zo) governing

177 momentum (zom) and heat/humidity transfer (zoh), which are a function of plant height (h).

178 The roughness length of grass ranges from 0.003 – 0.01 for short grass and 0.04 – 0.20 for

179 taller grass [11]. The FAO reference P-M [11] assumes momentum transfer roughness (zom) is

8
180 equal to 0.123 of plant height and roughness length for humidity transfer (zoh) is 0.1 of

181 momentum transfer roughness (zom).

182 The height where the measurements are taken also influences the results. Although

183 both temperature and humidity measurements can be used to determine ra, the right-hand term

184 in Eqn. A8 usually represents resistance to heat, thus zh is the height of air temperature

185 measurements [15]. zm is typically the height of wind measurements.

186 The zero plane displacement height (d0) represents the height above the surface where

187 fluctuations in the wind profile approach zero (i.e., the level of the apparent momentum sink)

of
188 [15, 23]. This typically occurs at a height around two thirds of plant height [11, 15]; although

ro
189 formulations vary throughout the literature [18]. Values for d0 range from ≤ 0.07 for short

190
-p
grass and ≤ 0.66 for taller grass (0.25 – 1.0 m) [15].
re
191 Eqn. A8 is only applicable under neutral atmospheric stability. In cases of a strongly
lP

192 stable or unstable atmosphere, which may be the case in shorter time periods, a correction for
na

193 stability must be applied, as in [11, 18].

194
ur

195 A.8. Transmittance


Jo

196 The canopy transmittance, [, describes the amount of short and long wave radiation

197 that is intercepted by the leaves [19, 24]. A lower transmittance means less radiation passes

198 through the leaves (i.e., higher shading), which is desirable for green roofs. As shown in Eqn.

199 A9 [19], transmittance relates to the LAI and leaf angle distribution, which is often described

200 using an extinction coefficient, -\ [25, 26]. The extinction coefficients, which differ for

201 longwave and shortwave radiation, are higher for flat or cone shaped leaves [25].

202 [ = ] <^_`aK (A9)

203 where [ is the transmittance, -\ is the extinction coefficient, and LAI is the leaf area index.

204

205 A.9. Thermal conductivity

9
206 Thermal conductivity, 6, is the rate at which heat is transferred across a material; a

207 lower value means heat is transferred relatively slowly. QG coming from the surface relates to

208 the effective thermal thickness (ZT), which is the depth where temperature fluctuations

209 approach zero. This depth is typically 10 to 30 mm below the soil surface [13]. Thermal

210 conductivity from the bottom of the roof depends on the soil temperature, the interior building

211 temperature, and the thermal conductivity of the material in between. Thermal conductivity of

212 dry extensive green roof substrate typically ranges from 0.18 to 0.4 W/mK, but this value

213 increases in saturated substrate (from 0.5 to 1.0 W/mK) [12]. This is because thermal

of
214 conductivity depends on moisture content, as well as, the density and composition of the

ro
215 substrate, as follows [19]:

216
-p
6 = bG + b1 (A10)
re
217 where a1 and a2 are coefficients that vary depending on the substrate’s specific heat (cp,substrate),
lP

218 density, porosity, and composition (e.g., organic, mineral, and quartz content) [12, 27, 28].
na

219 Based on laboratory analysis of sand, Chen (2008) [28] proposed the following

220 equation relating thermal conductivity to porosity and soil moisture:


ur

G<e e
221 6 = 6d 6 f(1 − g)h + giAe (A11)
Jo

222 where Sr is the degree of saturation of the sand, 6w is the thermal conductivity of water, 60 is

223 thermal conductivity of the solid grain of sand, is the porosity, and b and c are empirical

224 coefficients.

225 Thermal resistance, commonly referred to as an R-value, is the inverse of thermal

226 conductivity.

227

10
228 A.10. Appendix References

229 [1] Peng Z, Smith C, Stovin V. Internal fluctuations in green roof substrate moisture
230 content during storm events: Monitored data and model simulations. J Hydrol 2019;
231 573: 872–884.

232 [2] Fassman E, Simcock R. Moisture measurements as performance criteria for extensive
233 living roof substrates. J Environ Eng 2012; 138: 841–851.

234 [3] Wallach R. Physical characteristics of soilless media. In: Soilless Culture. Elsevier,
235 2019, pp. 33–112.

236 [4] Bennie A. Growth and mechanical impedance. p. 393–414. Y. Waisel et al.(ed.) Plant
237 roots: The hidden half. Marcel Dekker, New York. Growth Mech Impedance P 393–
238 414 Waisel Aled Plant Roots Hidden Half Marcel Dekker N Y.

of
239 [5] Allaire SE, Caron J, Duchesne I, et al. Air-filled porosity, gas relative diffusivity, and
240 tortuosity: Indices of Prunus× cistena sp. growth in peat substrates. J Am Soc Hortic Sci

ro
241 1996; 121: 236–242.

242 [6] -p
McCuen RH. Hydrologic Analysis and Design. 3rd ed. Pearson Prentice Hall, 2005.
re
243 [7] Del Barrio EP. Analysis of the green roofs cooling potential in buildings. Energy Build
244 1998; 27: 179–193.
lP

245 [8] Palla A, Gnecco I, Lanza LG. Unsaturated 2D modelling of subsurface water flow in
246 the coarse-grained porous matrix of a green roof. J Hydrol 2009; 379: 193–204.
na

247 [9] Vesuviano G, Sonnenwald F, Stovin V. A two-stage storage routing model for green
248 roof runoff detention. Water Sci Technol 2013; 69: 1191–1197.
ur

249 [10] Wallach R, Da Silva F, Chen Y. Unsaturated hydraulic characteristics of composted


Jo

250 agricultural wastes, tuff, and their mixtures. Soil Sci 1992; 153: 434–441.

251 [11] Walter IA, Allen RG, Elliott R, et al. ASCE’s standardized reference
252 evapotranspiration equation. In: Watershed management and operations management
253 2000. 2000, pp. 1–11.

254 [12] Vera S, Pinto C, Tabares-Velasco PC, et al. A critical review of heat and mass transfer
255 in vegetative roof models used in building energy and urban enviroment simulation
256 tools. Appl Energy.

257 [13] Haghighi E, Kirchner JW. Near‐surface turbulence as a missing link in modeling
258 evapotranspiration‐soil moisture relationships. Water Resour Res 2017; 53: 5320–
259 5344.

260 [14] Cascone S, Coma J, Gagliano A, et al. The evapotranspiration process in green roofs: A
261 review. Build Environ 2019; 147: 337–355.

262 [15] Oke TR. Boundary layer climates. Routledge, 2002.

263 [16] Hillel D. Environmental soil physics. Academic Press, San Diego. Environ Soil Phys
264 Acad Press San Diego.

11
265 [17] Tan CL, Tan PY, Wong NH, et al. Impact of soil and water retention characteristics on
266 green roof thermal performance. Energy Build 2017; 152: 830–842.

267 [18] Djedjig R, Ouldboukhitine S-E, Belarbi R, et al. Development and validation of a
268 coupled heat and mass transfer model for green roofs. Int Commun Heat Mass Transf
269 2012; 39: 752–761.

270 [19] Tabares-Velasco PC, Srebric J. A heat transfer model for assessment of plant based
271 roofing systems in summer conditions. Build Environ 2012; 49: 310–323.

272 [20] Sailor DJ. A green roof model for building energy simulation programs. Energy Build
273 2008; 40: 1466–1478.

274 [21] Oke TR, Mills G, Voogt J. Urban climates. Cambridge University Press, 2017.

275 [22] Verma S. Aerodynamic resistances to transfers of heat, mass and momentum. Estim

of
276 Areal Evapotranspiration 1989; 177: 13–20.

ro
277 [23] Crago R, Hervol N, Crowley R. A complementary evaporation approach to the scalar
278 roughness length. Water Resour Res; 41.

279
-p
[24] Monteith JL, Unsworth MH. Chapter 6 - Microclimatology of Radiation: (i) Radiative
re
280 Properties of Natural Materials. In: Monteith JL, Unsworth MH (eds) Principles of
281 Environmental Physics (Fourth Edition). Boston: Academic Press, pp. 81–93.
lP

282 [25] Tabares-Velasco PC. Predictive heat and mass transfer model of plant-based roofing
283 materials for assessment of energy savings (Ph. D. thesis). Ph.D. Thesis, Dept. of
na

284 Architectural Engineering, Pennsylvania State University, 2009.

285 [26] Del Barrio EP. Analysis of the green roofs cooling potential in buildings. Energy Build
ur

286 1998; 27: 179–193.


Jo

287 [27] Pielke Sr RA. Mesoscale meteorological modeling. Academic press, 2013.

288 [28] Chen SX. Thermal conductivity of sands. Heat Mass Transf 2008; 44: 1241.

289

12
HIGHLIGHTS

• Green roofs provide benefits for stormwater discharge and climate change mitigation
• Practical design of extensive green roofs do not align with expected co-benefits
• We quantify co-benefits by linking them to energy and water balance phenomena
• When optimizing for benefits, trade-offs and building constraints are important
• Stakeholder preferences must be considered when deciding on the benefits to pursue

of
ro
-p
re
lP
na
ur
Jo
Declaration of interests

☐ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☒The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

The authors have received support in the form of green roof construction materials from the company
NophaDrain.

of
ro
-p
re
lP
na
ur
Jo

You might also like