Sei sulla pagina 1di 9

Energy and Buildings 131 (2016) 123–131

Contents lists available at ScienceDirect

Energy and Buildings


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

Steady-state and transient thermal measurements of green roof


substrates
Andrea Pianella a,d,∗ , Robin E. Clarke b , Nicholas S.G. Williams a , Zhengdong Chen c ,
Lu Aye d
a
School of Ecosystem and Forest Sciences, Faculty of Science, The University of Melbourne, Burnley 3121, Australia
b
CSIRO Infrastructure Technologies, Clayton South 3169, Australia
c
CSIRO Land and Water Flagship, Clayton South 3169, Australia
d
Renewable Energy and Energy Efficiency Group, Department of Infrastructure Engineering, Melbourne School of Engineering, The University of Melbourne,
Parkville 3010, Australia

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

Article history: There has been growing interest in using extensive green roofs for commercial and residential buildings
Received 13 August 2016 in urban areas. Green roofs provide many benefits, including adding an additional insulation layer. The
Received in revised form potential of this benefit depends on many factors, including the thermal properties of the green roof sub-
12 September 2016
strate. Thermal conductivity values of three substrates comprised primarily of scoria, crushed roof tile
Accepted 13 September 2016
Available online 15 September 2016
and bottom-ash were measured with steady-state and transient techniques under three moisture condi-
tions. Specific heat capacities of the green roof substrates were also measured with a transient technique.
Steady-state measurements were performed with a “k-Matic” apparatus while transient measurements
Keywords:
Green roof substrate with KD2 Pro needles. In general, the steady-state measurements showed more consistency than tran-
Soil moisture measurements sient measurements. Thermal conductivity differed among the three substrates: crushed roof tile had
Thermal conductivity the highest conductivity values across all moisture contents. Substrate moisture content consistently
Transient and steady-state methods increased thermal conductivity across all substrates, but this was significantly greater for the crushed
roof tile substrate. Steady-state thermal conductivity curves were fitted using the thermal conductivity
model for green roof substrates adopted by Sailor (2011). The coefficients obtained are presented and
can be used in green roof models to quantify the thermal performance of green roofs and building energy
savings.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction not the only potential driver of their uptake. Reduction of building
heating and cooling loads, and the resulting building energy sav-
Green roofs are becoming more common in cities and towns ings, can help promote green roof installations [15], as buildings
due to the multiple ecosystem services they provide [1,2]. Green consume about 20% of the total Australian annual energy con-
roofs can limit some of the environmental and societal problems sumption [16]. However, many studies have shown that building
urban areas are facing. They can reduce storm water runoff [3], energy savings from green roofs are strongly dependant on climate
increase urban biodiversity [4,5], mitigate the urban heat island and location [17], as well as building characteristics and existing
effect [6], improve air quality [7], reduce building energy consump- insulation [18,19]. Given that Australia is conventionally divided
tion [8] and may even improve worker attention and productivity into eight different climate zones [20], the thermal performance
[9,10] helping to sustain their health and wellbeing [11]. How- and consequent building energy savings would be different across
ever, in new and emerging green roof markets, such as Australia, the country, and specific recommendations on the best green roof
there is often a lack of quantifiable evidence for these benefits in a build-up should be tested and provided locally [21].
local context [12]. While evidence of the benefits of green roofs to A number of energy balance models have been developed and
urban stormwater retention in Australia is mounting [13,14], this is refined to estimate heat fluxes through green roofs and then esti-
mate building energy savings [22–26]. Energy savings from green
roofs can be predicted from models if we know specific features
∗ Corresponding author at: School of Ecosystem and Forest Sciences, Faculty of of the two main layers which form a green roof: the substrate (or
Science, The University of Melbourne, Burnley 3121, Australia. growing medium) and the vegetation.
E-mail addresses: andrea.pianella@unimelb.edu.au, pianella@yahoo.com
(A. Pianella).

http://dx.doi.org/10.1016/j.enbuild.2016.09.024
0378-7788/© 2016 Elsevier B.V. All rights reserved.
124 A. Pianella et al. / Energy and Buildings 131 (2016) 123–131

Table 1
Sample standard deviations of the thermal conductivity of the green roof substrates
[W m−1 K−1 ].

Green roof substrate Transient SH-1 Transient TR-1 Steady-state

Scoria dry 0.010 0.006 0.002


Scoria moist 0.020 0.057 0.007
Scoria wet 0.032 0.027 0.009
Crushed roof tile dry 0.012 0.007 0.007
Crushed roof tile moist 0.060 0.013 0.036
Crushed roof tile wet 0.071 0.070 0.025
Bottom-ash dry 0.007 0.008 0.003
Bottom-ash moist 0.011 0.019 0.005 Fig. 1. The green roof substrates tested. From left to right: crushed roof tile, bottom-
Bottom-ash wet 0.035 0.044 0.027 ash and scoria mix.

quently, quantification of substrate thermal conductivity values at


different substrate moisture contents is needed.
Although many studies have investigated the thermal perfor- Studies of green roof thermal properties conducted so far have
mance of vegetated green roofs, few have focused specifically on predominantly used transient techniques to measure the thermal
the thermal performance of the substrates. We do this here to iso- conductivity of green roof substrates [32,34,42,43]. They are often
late the relative effects of substrates and the vegetation (paper in made using either single or dual needles that measure thermal
preparation) and gain a better understanding of the heat trans- conductivity along the needle itself or between the two needles.
fer across the substrate. Physical properties such as composition, Transient measurements, which imply the temperature of the sam-
water holding capacity, moisture content and porosity are impor- ples varies with time [41], may be inconsistent because they are
tant parameters which dictate heat transfer through green roof affected by external factors such as ambient air temperature and
substrates. Generally, substrates with high porosity are good insu- relative humidity. Green roof substrates are also normally quite
lators [27]. This is because, in general, heat flows more easily porous with many voids that allow water to drain freely. Transient
through solid particles (by conduction) than through pores, and measurements for wet substrates are therefore strongly influenced
higher porosity implies a lower effective thermal conductivity (or by the point where the measurement is taken because it can con-
higher thermal resistivity). The lower the porosity, the tighter the tain uncertain moisture content. These factors can cause transient
particles are compressed together and the more contact points exist measurements with thermal needles to be inaccurate and may also
among particles. The contact points then facilitate the conduction explain widely varying thermal conductivity values reported for
heat transfer [28–30]. wet green roof substrates.
Thermal conductivity data is available for a number of green In contrast to transient techniques, measurements in steady-
roof substrates, including those whose main aggregate materials state require that the thermal properties of the substrates are
are pumice (volcanic porous rock), porous silica, expanded clay, recorded without any external influence and after the sample has
expanded shale (sedimentary rock) and expanded slate (metamor- reached the steady-state conditions. Using thermal conductivity
phic rock) [31–34]. Presumably, these substrate components were values measured under steady-state conditions should therefore
chosen due to their high porosity, which is natural or, in the case of make green roof thermal models more realistic. However, existing
the expanded components, derived after heating at very high tem- steady-state methods such as ASTM C518 [44] may produce results
peratures, their lightness, availability and cost. Their use varies. with high uncertainties due to the nature of green roof substrates
For example, expanded shale is a very common and cheap sub- which are highly conducting but have varying contact resistance
strate component in western U.S.A., while expanded clay is more due to their porous nature. This study employed a new steady-state
commonly used on green roofs in mid-western and eastern U.S.A. technique specifically developed to better measure the thermal
[32]. These components have different thermal properties. Shale conductivity of porous substrates [45] and therefore improve the
has a higher thermal conductivity than many other substrates, validity and accuracy of steady-state measurements.
such as pumice [32] because of its mineral composition, particle The main objectives of this research were to quantify the ther-
size and porosity (Table 1 [28]). In contrast, expanded clay, due to mal conductivity values of three green roof substrates that are
its engineered porosity obtained by heating clay to about 1200 ◦ C, commonly used on green roofs in south-eastern Australia but
has a lower conductivity than other green roof substrates (Table 1 whose major components are also readily available elsewhere.
[28,33]). Thermal conductivities obtained using the steady-state and tran-
In south-eastern Australia scoria, crushed terracotta roof tile sient methods were compared and correlations were developed
and bottom-ash (non-combustible residue of burning black coal between the thermal conductivities and substrate moisture con-
also known as Horticultural Ash) have been used as the primary tent. In addition, we discuss the relative merits of the methods used
components of green roof substrates [13,14,35,36]. These materials and the specific heat capacities of the green roof substrates tested
are inexpensive and readily available but their thermal properties are reported.
have not been previously assessed. However, this information is
required to predict the thermal performance of Australian green 2. Materials and methods
roofs, estimate potential energy savings and facilitate the inclusion
of green roofs in building energy rating systems such as the Green The thermal properties of three green roof substrates commonly
Star Energy scheme [37]. used in south-eastern Australia (Fig. 1) were assessed. These sub-
The moisture content of a substrate is one of the key variables strates were developed by The University of Melbourne following
that determines the thermal resistance of a green roof. Kotsiris et al. international standards and using materials readily available in
[26] stated that the heat transfer in a substrate is linearly dependent Australia. Their major component was scoria mix, crushed roof tile
on the substrate’s moisture content. However, studies of soil ther- and bottom-ash respectively. Each substrate also contained 20%
mal conductivity [38–41] show that a linear dependency between composted coir, which provided the organic matter necessary to
soil moisture content and soil thermal conductivity is not realistic. support plant growth. Their chemical and physical properties have
These findings also apply to green roof substrates [31]. Conse- been previously assessed (Table 2 [13]) and the substrates have
A. Pianella et al. / Energy and Buildings 131 (2016) 123–131 125

been used on green roofs in Melbourne such as in the Growing In the TR-1 probe, the entire length of the needle contains a
Up roof (132 Queen St), the Venny Childcare Centre, the Monash heater, and there is a temperature sensor at the centre of the nee-
Civic Centre, Minifie Park Kindergarten and the Pixel Building [36] dle. As the needle heats, the thermistor detects the heating time
amongst others. Australian green roof companies now use very sim- and change in temperature, and also measures the cooling once
ilar proprietary substrates containing the same major components. the heating stops. The temperature changes are used to calculate
For our experiment, each substrate was prepared at the Univer- the thermal conductivity.
sity of Melbourne’s Burnley Campus using a small concrete mixer. The SH-1 dual-needle probe has one needle containing a heater
The materials proportions were determined by volume. and the other containing the temperature sensor. They are at a
fixed 6 mm distance from each other, so knowing the travel time
2.1. Sample preparation of the heat pulse and the resulting temperature change, thermal
conductivity, resistivity, diffusivity and specific heat capacity can
To determine how moisture content affects substrate thermal be measured.
conductivity, the three substrates were tested under three different The SH-1 and the TR-1 are both used in solid media, such as soil.
moisture content conditions: 1) Dry; 2) Moist (18–22% moisture However, with both sensors, care must be taken to avoid air gaps
content by volume); and 3) Wet – substrate water holding capacity as far as possible, as this creates an air boundary layer which may
in the holding frame (30–48% moisture content by volume). Three introduce contact resistance and lead to inaccuracies.
replicates of each substrate moisture condition were prepared. Due to the differences in heat transmission measurement, the
Dry substrate samples were prepared by drying them in labora- durations of measurement are also different between the two nee-
tory ovens (Steridium d500) at 105–110 ◦ C for 48–72 h. We aimed dle probes. Single-needle TR-1 requires five minutes to accurately
for moist samples to have a volumetric moisture content of approx- measure moist (or wet) substrates, and 10 min for dry substrates,
imately 20%. We chose this value as it was the average volumetric while the dual-needle probe requires two minutes for every sub-
moisture content of three experimental non-vegetated green roofs strate.
at the University of Melbourne’s Burnley Campus over a period of Five measurements were taken for each substrate sample using
two months in spring. Based on the dry weight and by comparing the SH-1 probe (Fig. 2): one at the crossing point of the two diag-
it with the weight of the substrates before drying them out, it was onals, and four about 80 mm from the central point along the
possible to calculate the moisture content before the drying pro- diagonals of the square filling frame. The area formed by the four
cess and consequently to estimate a priori the content of volumetric outer points was approximately equivalent of the area measured by
moisture by adding extra water at the substrates. the steady-state k-Matic apparatus (100 mm × 100 mm) and it also
To produce wet samples at water holding capacity, the sub- avoided any edge interference. Finally, as recommended by the KD2
strates were completely submerged in water and then drained until Pro manual, consecutive measurements were taken 15 min apart.
they stopped dripping. The samples were then placed into the hold- Three measurements were taken for each sample using the TR-
ing frames (Fig. 2). This procedure prevented water accumulating 1 needle. Because the cylinders had a cross section area of only
at the bottom of the filling frames as it was retained within the 25 cm2 and the single-needle is thick and multiple points would
substrate. The wet samples had different moisture contents, water have significantly disturbed the samples, only one single point was
holding capacity, and porosity due to the different characteristics used which is the centre of the cylinder.
of aggregate materials (Table 2 [13]).
For testing, samples were placed in pine-wood holding frames 2.3. Steady-state measurements
(300 mm × 300 mm × 60 mm) (Fig. 2) which had been coated with
three layers of oil based polyurethane lacquer (Cobot’s Cabothane Steady-state measurements were taken using a commercial k-
Clear) to prevent them absorbing water. Holding frames were filled Matic apparatus. The k-Matic apparatus follows the ASTM C518
using a wood filling frame placed on top of the holding frames. The standard [44] for measurements and has cold and hot plates whose
substrates were poured loosely into the frame until they reached temperatures were pre-set at 13 ◦ C and 33 ◦ C respectively. Sam-
a height of 100 mm above the holding frames. As per standard AS ples were put into the apparatus wrapped in plastic film to prevent
3473-2003, to settle the substrate, both the frames were then lifted change in moisture content, and with two silicone sponge sheets
50 mm off the table and dropped five times. No further compaction of 4 mm and 6 mm placed at the bottom and on top of the holding
was performed; however, additional settlement may have occurred frames respectively (Fig. 3). The k-Matic apparatus incorporates
when transporting the samples by car. After settling, the upper fill- a single HFM (heat flow meter) sensor to measure an area of
ing frame was removed and the excess mix was sliced off using 100 × 100 mm2 . Data were collected by a Doric 245 data logger.
a metal ruler. The samples were then sealed with plastic wrap to Once the measurement starts, data are saved at 10-min intervals.
maintain the desired moisture content. Each datum is the average over the previous 40 min. After reaching
equilibrium when the heat flow through the sample was constant,
2.2. Transient measurements the final datum was recorded and the total R-value of the sample
was obtained. The R-value of the substrate can be obtained by sub-
The substrates’ thermal conductivity (k-value) was first mea- tracting the R-value of the two silicone sponge sheets, the phenolic
sured through a transient approach using the KD2 Pro thermal board and the plastic film wrapping the sample. By knowing the
needle set (Decagon). Thermal probes can provide high quality thickness of each sample, the thermal conductivity of the substrate
measurements [46], and have been utilized in a number of green can be calculated by Eq. (1).
roof studies [26,28,33]. It should be noted that there are commonly
t
two types of thermal needle probe: single and dual-needle [47], k= (1)
which utilize different techniques to measure thermal conductiv- R
ity [48]. In this study, we have used both probe types: a 100 mm Where: t = thickness [m]
single-needle probe (TR-1) and a 30 mm dual-needle probe (SH-
R = thermal resistance [K m2 W−1 ]
1). Due to the depth of our holding frames (60 mm), we employed
120 mm-high plastic cylinders to take measurements with the TR-1 k = thermal conductivity [W m−1 K−1 ]
needle probe. Cylinders were filled with the same substrates using Clarke et al. [45] provides further details on this technique and
the same procedure used for the holding frames. how it was adapted to measure granular loose-fill materials such
126 A. Pianella et al. / Energy and Buildings 131 (2016) 123–131

Fig. 2. Loose-fill holding frame for the green roof substrates (R. E. Clarke, CSIRO) and points measured by the dual needle probe.

Fig. 3. Cross section diagram of k-Matic apparatus for steady-state sample measurement (R. E. Clarke, CSIRO).

as green roof substrates. After both the transient and steady-state 3.1. Effect of green roof substrate composition and physical
measurements were performed, all the moist and the wet sam- properties on thermal conductivity
ples were completely oven-dried to determine their volumetric
moisture content. Thermal conductivity (k) is defined as the rate of heat transfer
per unit across sectional area by conduction for a unit temperature
difference and a unit thickness. Different materials have different
3. Results and discussion thermal conductivity values. Good thermal conductor such as cop-
per has a higher thermal conductivity than good thermal insulator
Thermal conductivities of scoria mix, crushed roof tile and such as polystyrene foam.
bottom-ash substrates at dry, moist and wet conditions under tran- The three green roof substrates we tested, had different major
sient and steady-state measurements are shown in Figs. 4–6, while components and different air filled porosities which influenced
the sample standard deviations (n < 30) are shown in Table 1. From their thermal conductivity (Table 2 [13]). The crushed roof tile
Table 1, we can see that the steady-state measurements always substrate (Fig. 5) had the highest thermal conductivity of all the
show the smallest standard deviation in every treatment, with the tested substrates using both transient and steady-state methods.
exception of the crushed roof tile substrate at moist condition. All The crushed roof tile had the highest density and the lowest air
the data were validated with Student’s t-test at 95% confidence filled porosity, which allowed more conductive heat transfer which
interval, and showed no significant differences.
A. Pianella et al. / Energy and Buildings 131 (2016) 123–131 127

Fig. 4. Transient and steady-state thermal conductivities of scoria mix substrate.

Fig. 5. Transient and steady-state thermal conductivities of the crushed roof tile substrate.

Fig. 6. Transient and steady-state thermal conductivities of the bottom-ash substrate.


128 A. Pianella et al. / Energy and Buildings 131 (2016) 123–131

Fig. 7. Specific heat capacity of green roof substrates as a function of moisture saturation level.

was facilitated by the compaction. The crushed roof tile substrate mal conductivities. Local transient heating around the probe may
had the highest heat capacity (Fig. 7) under any given moisture con- cause moisture to move away, so that the region around the
tent. In contrast, the scoria mix substrate had the highest air filled probe becomes drier and therefore has lower thermal conductiv-
porosity, the lowest moisture content at wet state and, overall, the ity. Steady-state measurement avoids this local heating effect since
lowest thermal conductivity. Due to this, the scoria mix was the it imposes a stable overall temperature gradient during measure-
most insulating substrate under any moisture content condition, ment.
while the crushed roof tile substrate was the least insulating. Other observations using the probes are below:

3.2. Effect of green roof substrate moisture content on thermal a) Previous research has suggested that the thermal conductivity
conductivity measured by a probe should include a correction term to take
into account the presence of the probe [48]. This is because the
As expected, substrate moisture content was consistently asso- probes may facilitate heat transfer through the needle itself and
ciated with increased thermal conductivity across all substrates. therefore make heat conduction more significant than in the k-
However, the thermal conductivities of the three green roof Matic apparatus due to the contact of the probe.
substrates increased at different rates (Figs. 4–6). The thermal con- b) While the steady-state measurements take into account heat
ductivity of the crushed roof tile substrate increased at a greater transfer through the entire substrate profile (60 mm) and
rate with increasing moisture content compared to scoria mix and account for a larger volume of the sample, the transient mea-
bottom-ash substrates. Generally, the thermal conductivity values surements are mostly limited to the sampling point.
at steady-state condition are congruous results with little fluctua- c) For the dual-needle measurements, the region tested by the
tion. In contrast, transient measurements of thermal conductivities probe was closer to the surface (top 15 mm) and was potentially
have wider fluctuation in particular for moist and wet samples even drier, and therefore less thermally conductive, than the bot-
within the same sample (see Section 3.3) which is reflected in their tom region which was likely to be wetter due to water dynamic
standard deviation (Table 1). nature to move down.
Transient thermal conductivities of moist and wet samples were d) Five points for each sample were measured with the dual-needle
more challenging to measure than the thermal conductivity of probe, so the moisture content might have been slightly differ-
dry samples due to the dynamic nature of the water distribution ent in each measured point.
and transportation in the sample. Heat transfer is not purely one e) The KD2 Pro manual acknowledges the possibility of a slight
dimensional, and this is particularly true for moist and wet samples decrease in thermal conductivity due to the probe contact resis-
when the heat transfer is coupled with the water mass transfer. tance that may occur when an air gap is created.
The moist and wet samples do not have perfectly homogeneous f) The dual-needle probe has a large heat pulse which causes free
moisture contents due to the size of the substrate holding frames. convection in wet or liquid samples and thus slightly increases
This was particularly evident when measuring thermal conductiv- thermal conductivity result.
ity with the dual-needle thermal probe at five different points of g) The KD2 Pro manual also acknowledges that measurements
the same sample. Smaller fluctuation in the thermal conductivity are less accurate (±10%) when measuring thermal conductiv-
for moist and wet samples was measured with the single probe as ities higher than 0.2 W m−1 K−1 , conditions which occurred in
the sampling point was the same for all the replicated measure- the moist and wet samples. In contrast, steady-state measure-
ments, therefore the localised differences in moisture content did ments reduced the inaccuracies to ±7% for highly conducting
not affect the final result. wet crushed roof tile samples and to ±4% for dry samples [45].
h) Finally, the penetration of the probe slightly alters the density
3.3. Comparison between steady-state and transient and compaction of the sample and may potentially create more
measurements air gaps.

The transient measurements with single and dual-needle probes According to the KD2 Pro manual, each probe is designed to
produced less consistent thermal conductivities with higher stan- test different samples with different properties. The 100 mm single-
dard deviation than the steady-state measurements (Figs. 4–6). In needle probe (TR-1) is recommended for moist or dry soils, powders
general, we found that steady-state thermal conductivities were and granular materials and not liquid. The 30 mm dual-needle
higher for moist and wet samples compared to transient ther- probe (SH-1) is designed mostly for solid and granular materials
A. Pianella et al. / Energy and Buildings 131 (2016) 123–131 129

and not for liquids, but, according to the manual, it is still suitable concavity, the roof tile shows a slight upward concavity, highlight-
for moist or dry soils, powders and granular materials. ing how this material is sensitive to water content. This may be
For dry substrates we observed that data collected by SH-1 bet- explained by the physical properties of the crushed roof tile, which
ter matched the results obtained using the steady-state approach, is the substrate aggregate with the highest bulk density and the
while the results collected by TR-1 were usually closer to the lowest air filled porosity (Table 2 [13]). In the wet condition, the
steady-state thermal conductivities of the substrate samples with bulk density of the crushed roof tile was 30% higher than the scoria
higher moisture content. Differences are due to the design of the mix and bottom-ash substrates, making the substrate more com-
probes and measurement techniques. The single-needle probe has pact and less porous.
a diameter of 2.4 mm, while the diameter of the dual-needle probe The single-needle probe measurements tended to overestimate
is 1.3 mm. This implies that the TR-1 probe is more invasive and the normalized thermal conductivity for wet scoria and crushed
generates a greater contact local disturbance than the SH-1 probe. roof tile substrates calculated with our coefficients, while the
TR-1 and SH-1 probes measure thermal conductivity in two dual-needle probe tended to underestimate these values. For the
different ways: the single-needle probe measures the thermal bottom-ash, the single and dual-needle probes gave similar results
conductivity through the probe itself where both the heater and for the wet samples.
temperature sensors are located, while the dual-needle probe Sailor’s coefficients were calculated using the thermal conduc-
measures thermal conductivity between the needles. In fact, the tivity measured with the dual-needle SH-1 probe. Indeed, for the
dual-needle probe is formed of two different probes: one gener- scoria mix and crushed roof tile substrates (Figs. 8 and 9), we found
ates the heat pulse, while the other measures the temperature. The that Sailor’s curve fits the thermal conductivities that we measured
longer measurement time for dry sample with the single probe may with the same type of probe (SH-1) well. However, as we believe
also be a reason why the results differ. The single probe requires a the steady-state thermal conductivities to be more accurate, we
higher measurement time: 10 min for dry samples, 5 min for moist needed to calculate new coefficients for Eq. (2).
and wet, while the dual probe requires only 2 min for every sample. It is noted that the volumetric saturation level for green roof
Although the k-Matic apparatus provides the most consistent substrates is complex to measure as it is dependent on the sub-
and reliable thermal conductivity values, we found that both probes strate density and on the substrate container (i.e. a pot for plants
provided useable values for our substrates under different moisture would have a different volumetric saturation level than a green
contents. However, the dual-needle probe proved to have rela- roof module due to different bulk density and compaction). For
tively more accurate results for the dry samples, while the single future studies, the water content by weight might be a more feasible
probe was better for moist and wet samples. The exception was the alternative to set up standard models or equations.
bottom-ash substrate where both probes had comparable results
perhaps due to the finer particles of which the substrate is com-
posed. 3.5. Implications for buildings

Existing green roof thermal models [22,24] use the thermal con-
3.4. Thermal conductivity correlations for green roof substrates
ductivity of substrates to calculate heat transfers through the roof
profile and estimate building energy savings. It is crucial to feed
The thermal conductivity of green roof substrates changes with
models with accurate values for all the required inputs, includ-
water content. After analysing three soil thermal conductivity mod-
ing the conductivity of dry soil. A parametric analysis for a 120 m2
els [38–40], Sailor and Hagos [31] adopted and optimized a model
single-storey building located in the Melbourne suburb of Tulla-
for the thermal conductivities of their green roof substrates by
marine was simulated with EnergyPlus 8.5.0 using Sailor’s 2008
plotting the ratio of the thermal conductivities (k) on the ther-
[22] green roof model. Weather data used for the simulations were
mal conductivity of the dry substrate (kdry ) as a function of the
collected at Melbourne airport, located in the same suburb of Tul-
volumetric degree of saturation (Sr ). The resulting Eq. (2) was:
lamarine, from 1st April 2014 through 31st March 2015. Default
k ␣ exp(␤Sr ) data were used for all the green roof inputs, except the conductiv-
= (2)
kdry [1 + (␣ − 1) exp(␤Sr )] ity of dry soil (CDS). We performed two simulations: i) the first with
CDS as 0.35 W m−1 K−1 (default value); ii) the second time with CDS
Where ␣ = 1.45 and ␤ = 4.411. We plotted the thermal conductivi- as 0.13 W m−1 K−1 (average thermal conductivity of the scoria mix
ties of our substrates over their dry steady-state values (normalized substrate at dry condition measured using our steady-state tech-
thermal conductivity) using Sailor’s ␣ and ␤ coefficients. With Eq. nique). In comparison with the first simulation, it was found that
(2) we also calculated our own ␣ and ␤ coefficients to minimise the heating load for one year of the second simulation decreased
the root mean square error (RMSE) using the Microsoft Excel solver significantly (∼32%), while the cooling load increased (∼7%).
function. The following coefficients were calculated: Furthermore, our study highlights how higher substrate mois-
ture content is consistently associated with significantly higher
• For scoria mix substrate, ␣ = 1.339 and ␤ = 3.681 with thermal conductivity. Existing green roof thermal models might
RMSE = 0.0036 compared to RMSE = 0.0205 with Sailor coef- be more accurate if different substrate thermal conductivity values
ficients (Fig. 8). reflecting different moisture contents are used. They might also be
• For crushed roof tile substrate, ␣ = 1.155 and ␤ = 2.24 with able to predict seasonal differences and calculate yearly benefit if
RMSE = 0.0061 compared to RMSE = 0.0612 with Sailor coeffi- soil moisture sensors are installed in the green roofs.
cients (Fig. 9). Of the three green roof substrates we tested, scoria was found
• For bottom-ash substrate, ␣ = 1.305 and ␤ = 3.221 with to have the lowest thermal conductivity under different moisture
RMSE = 0.0009 compared to RMSE = 0.0192 with Sailor coef- contents. Consequently, if all other design considerations are equal,
ficients (Fig. 10). it should be chosen if insulation is to be maximised.
However, we recognise that further studies are required to
Figs. 8–10 show the normalized thermal conductivity as defined model the thermal performance of buildings using these substrates
by Eq. (2) for scoria mix, crushed roof tile and bottom-ash substrates across Australia climate zones [21]. We also understand that the
against their volumetric saturation levels. While the relationship green roofs should be judged on the multiple benefits they provide,
for scoria and bottom-ash is a curve with a slight downward and not only for their insulating properties.
130 A. Pianella et al. / Energy and Buildings 131 (2016) 123–131

Fig. 8. Normalized thermal conductivity curves for scoria mix substrate as a function of moisture saturation level.

Fig. 9. Normalized thermal conductivity curves for crushed roof tile substrate as a function of moisture saturation level.

Fig. 10. Normalized thermal conductivity curves for bottom-ash substrate as a function of moisture saturation level.
A. Pianella et al. / Energy and Buildings 131 (2016) 123–131 131

4. Conclusions [16] Department of Industry, I.a.S, Australian Energy Statistics, 2015 [cited 2016
24/05/2016]; Available from: http://www.industry.gov.au/Office-of-the-
Chief-Economist/Publications/Pages/Australian-energy-statistics.aspx#.
The thermal conductivity values of three green roof substrates [17] P. La Roche, U. Berardi, Comfort and energy savings with active green roofs,
based on scoria mix, crushed roof tile, and bottom-ash respectively Energy Build. 82 (2014) 492–504.
were measured using steady-state and transient techniques. All the [18] H.F. Castleton, V. Stovin, S.B.M. Beck, J.B. Davison, Green roofs; building energy
savings and the potential for retrofit, Energy Build. 42 (10) (2010) 1582–1591.
substrates were tested at three different moisture contents: dry, [19] U. Berardi, The outdoor microclimate benefits and energy saving resulting
moist and wet. Results showed that steady-state measurements from green roofs retrofits, Energy Build. 121 (2016) 217–229.
are more consistent within replicates than transient measurements [20] ABCB, NCC 2016 Building Code of Australia – Volume Two. NCC 2016 Building
Code of Australia. Vol. Two. 2016: The Australian Building Codes Board.
taken by the needle probes. Substrates have different thermal con-
[21] A. Pianella, J. Bush, Z. Dong, N.S.G. Williams, L. Aye, Green roofs in Australia:
ductivity due to their main aggregate material and the moisture review of thermal performance and associated policy development, in
content. For the three substrates tested, their thermal conductiv- Architectural Science Association Conference 2016. In review: Adelaide,
Australia.
ity increases with substrate moisture content. Steady-state thermal
[22] D.J. Sailor, A green roof model for building energy simulation programs,
conductivities are generally higher than transient thermal conduc- Energy Build. 40 (8) (2008) 1466–1478.
tivities for moist and wet samples due to longer measurement [23] E. Palomo Del Barrio, Analysis of the green roofs cooling potential in
time across a larger volume of sample, and the technical limi- buildings, Energy Build. 27 (2) (1998) 179–193.
[24] P.C. Tabares-Velasco, J. Srebric, A heat transfer model for assessment of plant
tations of the probes. Although transient measurements provide based roofing systems in summer conditions, Build. Environ. 49 (1) (2012)
useable thermal conductivity values under different moisture con- 310–323.
tents, steady-state measurements provide more accurate results to [25] F. Olivieri, C. Di Perna, M. D”Orazio, L. Olivieri, J. Neila, Experimental
measurements and numerical model for the summer performance
feed existing green roof thermal models. To quantify thermal con- assessment of extensive green roofs in a Mediterranean coastal climate,
ductivity for energy rating schemes and similar applications, it is Energy Build. 63 (2013) 1–14.
important that results are comparable and therefore tests should [26] G. Kotsiris, A. Androutsopoulos, E. Polychroni, P.A. Nektarios, Dynamic
U-value estimation and energy simulation for green roofs, Energy Build. 45
be conducted under similar steady-state conditions. (2012) 240–249.
[27] Y.J. Lin, H.T. Lin, C.Y. Chou, Physical properties and thermal performance of
calcined sludge as planting substrate in green roofs. 2011: Shangri-La. pp.
Acknowledgments 3–11.
[28] T.S. Yun, J.C. Santamarina, Fundamental study of thermal conduction in dry
This research was funded by ARC Linkage grant LP130100731 soils, Granular Matter 10 (3) (2008) 197–207.
[29] W. Woodside, J.H. Messmer, Thermal conductivity of porous media. I.
and conducted in collaboration with CSIRO Infrastructure Tech- Unconsolidated sands, J. Appl. Phys. 32 (9) (1961) 1688–1699.
nologies. Andrea Pianella is supported by the Melbourne [30] G.A. Narsilio, J. Kress, T.S. Yun, Characterisation of conduction phenomena in
International Research Scholarship (MIRS) and Melbourne Inter- soils at the particle-scale: finite element analyses in conjunction with
synthetic 3D imaging, Comput. Geotech. 37 (7–8) (2010) 828–836.
national Fee Remission Scholarship (MIFRS). Authors would like to [31] D.J. Sailor, M. Hagos, An updated and expanded set of thermal property data
thank Jörg Werdin for his help preparing the substrate samples and for green roof growing media, Energy Build. 43 (9) (2011) 2298–2303.
building the filling frame, Associate Professor Barbara Ozarska for [32] D.J. Sailor, D. Hutchinson, L. Bokovoy, Thermal property measurements for
ecoroof soils common in the western U.S, Energy Build. 40 (7) (2008)
her advice on waterproofing the timber frames and John Rayner for
1246–1251.
discussion of substrate components. [33] P.C. Tabares-Velasco, J. Srebric, Experimental quantification of heat and mass
transfer process through vegetated roof samples in a new laboratory setup,
Int. J. Heat Mass Transfer 54 (25–26) (2011) 5149–5162.
References [34] M. Zhao, P.C. Tabares-Velasco, J. Srebric, S. Komarneni, R. Berghage, Effects of
plant and substrate selection on thermal performance of green roofs during
[1] E. Oberndorfer, J. Lundholm, B. Bass, et al., Green roofs as urban ecosystems: the summer, Build. Environ. (2014) (0).
ecological structures, functions, and services, Bioscience 57 (10) (2007) [35] A.M. Coutts, E. Daly, J. Beringer, N.J. Tapper, Assessing practical measures to
823–833. reduce urban heat: green and cool roofs, Build. Environ. 70 (2013) 266–276.
[2] U. Berardi, A. GhaffarianHoseini, A. GhaffarianHoseini, State-of-the-art [36] DEPI, Growing Green Guide: a Guide to Green Roofs, Walls and Facades in
analysis of the environmental benefits of green roofs, Appl. Energy 115 (2014) Melbourne and Victoria, Australia, Victorian Department of Environment and
411–428. Primary Industries, Melbourne, 2014.
[3] Y. Li, R.W. Babcock Jr, Green roof hydrologic performance and modeling: a [37] GBCA, Green Star Design & As Built v1.1, Submission Guidelines, Green
review, Water Sci. Technol. 69 (4) (2014) 727–738. Building Council of Australia, Sydney, 2015.
[4] N.S.G. Williams, J. Lundholm, J. Scott Macivor, Do green roofs help urban [38] O. Johansen, Thermal Conductivity of Soils, Norwegian University of Science
biodiversity conservation? J. Appl. Ecol. 51 (6) (2014) 1643–1649. and Technology, Trondheim, 1977.
[5] J.T. Lundholm, Green roof plant species diversity improves ecosystem [39] J. Côté, J.M. Konrad, A generalized thermal conductivity model for soils and
multifunctionality, J. Appl. Ecol. 52 (3) (2015) 726–734. construction materials, Can. Geotech. J. 42 (2) (2005) 443–458.
[6] D. Kolokotsa, M. Santamouris, S.C. Zerefos, Green and cool roofs’ urban heat [40] S. Lu, T. Ren, Y. Gong, R. Horton, An improved model for predicting soil
island mitigation potential in European climates for office buildings under thermal conductivity from water content at room temperature, Soil Sci. Soc.
free floating conditions, Sol. Energy 95 (2013) 118–130. Am. J. 71 (1) (2007) 8–14.
[7] J.J. Baik, K.H. Kwak, S.B. Park, Y.H. Ryu, Effects of building roof greening on air [41] O.T. Farouki, Thermal Properties of Soils, Trans Tech Publications, 1986.
quality in street canyons, Atmos. Environ. 61 (2012) 48–55. [42] J. Darkwa, G. Kokogiannakis, G. Suba, Effectiveness of an intensive green roof
[8] A. Niachou, K. Papakonstantinou, M. Santamouris, A. Tsangrassoulis, G. in a sub-tropical region, Build. Serv. Eng. Res. Technol. 34 (4) (2013) 417–432.
Mihalakakou, Analysis of the green roof thermal properties and investigation [43] S.E. Ouldboukhitine, R. Belarbi, R. Djedjig, Characterization of green roof
of its energy performance, Energy Build. 33 (7) (2001) 719–729. components: measurements of thermal and hydrological properties, Build.
[9] K.E. Lee, K.J.H. Williams, L.D. Sargent, N.S.G. Williams, K.A. Johnson, 40-second Environ. 56 (2012) 78–85.
green roof views sustain attention: the role of micro-breaks in attention [44] ASTM, ASTM C518-10. Test method for steady-state thermal transmission
restoration, J. Environ. Psychol. 42 (2015) 182–189. properties by means of the heat flow meter apparatus Annual Book of ASTM
[10] K.E. Lee, Gazing at nature makes you more productive, in: N. Torres (Ed.), Standards, vol. 14.02, ASTM International, 2010.
Defend Your Research, Harvard Business Review, 2015, pp. 32–33. [45] R.E. Clarke, A. Pianella, B. Shabani, G. Rosengarten, Steady-state thermal
[11] T. Hartig, R. Mitchell, S. De Vries, H. Frumkin, Nature and health, Annu. Rev. measurement of moist granular earthen materials, J. Build. Phys. (2016),
Public Health (2014) 207–228. http://jen.sagepub.com/content/early/2016/03/12/1744259116637864.
[12] N.S.G. Williams, J.P. Rayner, K.J. Raynor, Green roofs for a wide brown land: abstract.
opportunities and barriers for rooftop greening in Australia, Urban For. Urban [46] K.L. Bristow, G.J. Kluitenberg, C.J. Goding, T.S. Fitzgerald, A small multi-needle
Green. 9 (3) (2010) 245–251. probe for measuring soil thermal properties, water content and electrical
[13] C. Farrell, R.E. Mitchell, C. Szota, J.P. Rayner, N.S.G. Williams, Green roofs for conductivity, Comput. Electron. Agric. 31 (3) (2001) 265–280.
hot and dry climates: interacting effects of plant water use, succulence and [47] G. Liu, B.C. Si, Single- and dual-probe heat pulse probe for determining
substrate, Ecol. Eng. 49 (2012) 270–276. thermal properties of dry soils, Soil Sci. Soc. Am. J. 75 (3) (2011) 787–794.
[14] C. Farrell, X.Q. Ang, J.P. Rayner, Water-retention additives increase plant [48] K.L. Bristow, R.D. White, Comparison of single and dual probes for measuring
available water in green roof substrates, Ecol. Eng. 52 (2013) 112–118. soil thermal properties with transient heating, Aust. J. Soil Res. 32 (3) (1994)
[15] O. Saadatian, K. Sopian, E. Salleh, A review of energy aspects of green roofs, 447–464.
Renew. Sustain. Energy Rev. 23 (2013) 155–168.

Potrebbero piacerti anche