Você está na página 1de 12

Energy and Buildings 40 (2008) 2224–2235

Contents lists available at ScienceDirect

Energy and Buildings


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

Effect of roof solar reflectance on the building heat gain in a hot climate
Harry Suehrcke a,*, Eric L. Peterson b,c, Neville Selby d
a
James Cook University, School of Engineering, Townsville, Australia
b
School of Architectural, Civil and Mechanical Engineering, Victoria University, Australia
c
Institute of Sustainability and Innovation Melbourne, Victoria University, Australia
d
Townsville, Queensland, Australia

A R T I C L E I N F O A B S T R A C T

Article history: The effect of the roof solar reflectance on the thermal performance of a building is often ignored.
Received 20 December 2007 However, there are significant differences in heat gain from light and dark-coloured roof surfaces. In this
Received in revised form 6 June 2008 paper an equation for the average daily downward heat flow of a sunlit roof is derived. Using building
Accepted 18 June 2008
simulation, it is first shown that the thermal mass of the roof does not significantly affect the overall daily
heat gain (although it causes a time lag and reduction in peak heat flow). As a consequence the daily heat
Keywords:
gain from the roof may be estimated by integrating the equation for the steady-state downward heat
Roof solar absorptance
transfer over the day. For north Australia the derived equation suggests that a light-coloured roof has
Solar reflectance
Roof heat gain
about 30% lower total (air temperature difference and solar-driven) heat gain than a dark-coloured one.
Thermal resistance The effect of aging (change in solar reflectance with time) is considered in the calculations and a
Hot climate relationship between the solar absorptance of new and aged material is suggested. A classification of roof
Cool roof colours with respect to their solar absorptance (dark, medium, light and reflective) is proposed to enable a
quick and simple assessment of the effect of roof colour on the heat gain and R-value.
ß 2008 Elsevier B.V. All rights reserved.

1. Introduction downward heat flow generally is undesired, as it tends to overheat


the building or put extra load on an air-conditioning system.
The thermal performance of a building is affected by the solar Fig. 1 illustrates the difference between downward and upward
absorptance of the roof. During clear sky conditions up to about heat flow through a roof space. For downward heat flow,
1 kW/m2 of solar radiation can be incident on a roof surface, and schematically depicted in Fig. 1(a), the heat transfer between
between 20% and 95% of this radiation is typically absorbed. The the roof and the ceiling is predominately due to thermal radiation.
roof colour that is apparent from the reflected visible1 part of the For upward heat flow, on the other hand, where the air in the roof
solar radiation usually gives an indication of the value of solar space is heated from below, heat is transferred by both convection
absorptance (e.g. a black surface with low visible reflectance and radiation.
suggests a high solar absorptance). In this paper we analyse the The downward heat flow from the roof can be reduced through
building heat gain from a roof in a warm/hot climate. the use of a light roof colour, reflective foil and/or insulation. While
In locations close to the equator (tropical areas with latitude the installation of roof insulation and reflective foil has now
angle 23.58 or less) ambient temperatures and solar radiation become mandatory in many countries (e.g. Building Code of
levels are sufficiently high that even during winter buildings do not Australia 2005 [2]), the choice of an appropriate roof colour is not
require active heating. Daytime heat flow from a sun-exposed roof generally accepted as a contribution to the insulation.2 The
surface is essentially only in downward direction and the problem is that no ready and general assessment procedure for the
benefit of light roof colour appears to be available, although the
influence of roof colour on the building energy efficiency has
clearly been recognised. For example, the ‘‘Cool Roofs’’ website
* Corresponding author. Present address: 14 Allerton Way, Booragoon, W.A.
6154, Perth, Australia. Tel.: +61 8 9316 0540; fax: +61 8 9316 0540.
2
E-mail addresses: suehrcke@bigpond.com, harry.suehrcke@jcu.edu.au An exception is the Building Code of Australia (BCA) 2006 [3] for commercial
(H. Suehrcke). buildings occupied during daylight hours. The BCA now enables a reduction in
1
Nearly 50% of the solar radiation energy reaching the earth’s surface is in the ceiling insulation R-value of 1.0 m2 K/W to be made for a roof solar absorptance of
visible range (0.38–0.78 mm) [1]. 0.35, or less for particular hot climates regions.

0378-7788/$ – see front matter ß 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.enbuild.2008.06.015
H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235 2225

Nomenclature
aweathered solar absorptance of weathered roofing material
e surface emissivity for long-wave thermal radiation
A area (m2)
s Stefan–Boltzmann constant (5.67  108 W/
F = q*/q Factor of increase in heat flow due to solar
m K 4)
illumination
G solar radiation per unit area (beam + diffuse)
(W/m2)
Akbari et al. [4] provides extensive experimental information
Ḡ average radiation incident on a (horizontal) roof
about roof materials. Building simulation (e.g. TRNSYS [5] as used,
surface, Ḡ ¼ H=Dt (W/m2)
for example, by [6]) can be used to quantify the relative effect of
hc convective heat transfer coefficient between roof roof and wall colour on the energy flow of each building element.
and ambient air (W/m2 K) A quantitative assessment of the effect of roof colour (more
hi heat transfer coefficient between roof upper sur- correctly roof solar absorptance and thermal emittance) is
face and building inside air, including the sum of complicated by the fact that
thermal resistances through roof construction and
inside air film (W/m2 K) (1) For locations that require both heating and cooling the benefits
ho outside total heat transfer coefficient between from a light roof colour are not always positive and the
roof and ambient air (W/m2 K). Note that ho is downward and upward heat flow cases require separate
treatment.
linearized and includes the radiative cooling to the
(2) The heat flow due to solar absorption on the roof surface
sky, ho = hc + hr(T  Tsky)/(T  Ta), see Appendix A.1
combines with that due to air temperature differences between
for details the outside and inside. Both heat flows need to be evaluated in
hr radiative heat transfer coefficient between roof order to judge the relative effect of roof colour.
surface and sky hr ¼ seðT 2 þ Tsky2
ÞðT þ T sky Þ (3) The heat flows due to solar absorption and outside to inside air
(W/m2 K) temperature difference are variable and influenced by the
R
H daily horizontal surface irradiation, H ¼ G dt thermal mass of the roof.
(MJ/m2) (4) The solar absorptance of a roof will change with time due to
* *
q = Q /A rate of heat flow per unit area (flux) from an dust and aging.
illuminated roof to the building inside (W/m2)
q = Q/A rate of heat flow per unit area (flux) from an
However, restricting the analysis to locations where there are
unilluminated (shaded) roof to the building inside
no heating requirements (cooling only) simplifies the problem.
(W/m2) Therefore this paper only deals with the downward heat transfer
R thermal resistance (R = 1/U) (m2 K/W) case, as it is typical, for example, in northern parts of Australia
*
R effective thermal resistance for an illuminated wall (Zone 1 in the Building Code of Australia 2006 [3]). An effective R-
(m2 K/W) value for cooling that takes into account the solar absorption (roof
Ta ambient (outside) temperature (8C) colour) is derived.
Ti inside temperature (average internal space air) (8C)
Tsky sky temperature (K) (typically 2–20 K below 2. Review of some existing literature
ambient temperature)
The benefits of light roof colour have been noted many times,
Tsol–air sol–air temperature (Ta + aG/ho) (8C)
particularly in locations where the sun is almost directly overhead
Dt day length in seconds (=24  3600 s) (s)
and for single story buildings. For example, Parker et al. [7]
DT temperature difference between outside and inside
reported that the space cooling requirements, after application of
air (K) reflective roof coatings on nine Florida homes, decreased by 19%
DT* Effective temperature difference for a solar radia- and the staff of a Florida school found that interior comfort
tion-illuminated surface (DT* = DTsol–air) (K) noticeably improved after the grey bitumen roof surface was
DT̄ average daily temperature difference between painted white. Suehrcke [8], on the other hand, using a numerical
outside and inside (K) simulation of a building, suggested that the peak values of heat

DT̄solair average daily sol–air temperature difference DT̄ ¼ flow from a roof could reduce by as much as 60% when a white
DT̄solair (K) surface replaces a corroded galvanised one.
U overall heat transfer coefficient between the Simpson and McPherson [9] found from measurements of 1/4-
scale models in Arizona, that ceiling insulation is more effective in
ambient and inside (W/m2 K), Note: 1/U = 1/
reducing daytime heat gain than increased roof albedo. However,
ho + 1/hi
Simpson and McPherson also made the interesting observation
that on a 24-h basis an increased albedo was about as effective as
Greek symbols addition of ceiling insulation in reducing building heat gain. This is
a absorptance for solar radiation (fraction of solar explained in the paper due to the enhanced nighttime heat loss
radiation energy absorbed from all wavelengths through the uninsulated ceiling of the high albedo roof.
and directions) Griggs et al. [10] provide a comprehensive and very useful study
aave average roof solar absorptance (approx. 0.7) on the effect of roof colour. Their study includes a ‘‘work sheet’’ to
adust solar absorptance of dust and dirt (assumed calculate the energy cost savings as a result of a roof reflectance
change and its use is demonstrated with two examples. However,
here 0.8)
the report by Griggs et al. unfortunately does not contain explicit
2226 H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235

Fig. 1. Downward and upward heat transfer through a roof space: (a) upward heat transfer; (b) downward heat transfer. In the above figure Ta denotes the ambient
temperature, Tsky the radiative temperature of the sky and Ti the inside building temperature (in this paper assumed to be the air temperature below ceiling).

equations for the effect of roof colour and they appear to be and wind start removing some of the deposits). The paper
‘‘hidden’’ within the work sheets. mentions a study from Sacramento, California that indicates a
There are analytical methods for the calculation of roof heat 20% reduction of the first year energy savings for all subsequent
gain (e.g. see Granja and Labaki [11]). However, many of these years (2–10 years).
methods appear to be limited to simple thermal conduction A detailed recent study on the aging and weathering of cool
problems (e.g. a solid concrete roof with insulation) and do not roofing membranes (Akbari et al. [18]), which included 13 white
provide general answers on the effect of roof colour. Perhaps the roof material samples that had been exposed for 5–8 years in eight
best calculation method for the roof heat gain accompanies a paper different locations, found that their solar reflectivity had dropped
on non-white roof coatings that reflect the invisible parts of solar from 0.8 to nearly 0.5. However, in the study it was also found that
radiation (Levinson et al. [12]). The method recognises that washing the samples could almost completely restore the original
changes in solar roof absorption approximately cause proportional reflectivity. See also Levinson et al. [19].
changes in ceiling heat flux. However, the issue of the roof thermal In summary, it is widely recognised that a reflective (e.g. white)
mass is not fully resolved and the ceiling heat flux due to outside to roof surface in place of a dark one can be of great benefit in hot and
inside air temperature differences is not explicitly included in the sunny climates (increase human comfort and/or reduce the cooling
equations. load). However, the benefits are variable and are not easily
The Australian/New Zealand standard on thermal insulation AS accounted for (unlike the thermal resistance of solid material and
4859.1:2002 [13] acknowledges (Appendix G) that external sunlit bulk insulation).
surfaces with low solar absorptance stay cooler and help reduce
the heat flow. Specifically, in Appendix G it also says ‘‘. . ., similar 3. Temperature profile in roof space for downward heat flow
thermal benefits may be derived by adding a certain thermal
insulation or by changing solar reflectance by a certain amount’’. Downward heat flow from the roof may be caused by
However, the notion of assigning an insulation value was rejected
because of the variability of conditions and because ‘‘solar (1) A temperature difference between the outside and inside
reflectance offers no benefit when the sun is not shining’’. temperature.
The CSIRO advisory note (Clarke and Delsante [14]) with the title: (2) Solar radiation that is being absorbed on the roof surface.
‘‘Do insulating paints work’’ points out that the thermal resistance of
a thin paint film is effectively zero and that reflective paint cannot The exposure of the roof surface to solar radiation increases the
provide insulation for a building, but keep it cool. This statement is of roof surface temperature and sets up a heat flow in parallel to the
course correct, though it is noted that for hot and sunny locations one set-up by temperature difference. Fig. 2 illustrates the
bulk insulation or a reflective roof coating can both achieve the same temperature profile of a roof with and without solar radiation
(daytime) effect—reduce the downward heat flow.3 exposure, suggesting the effect of solar radiation absorption
The recent ANSI/ASHRAE Standard 90.2-2004 [15] now usually is the most important factor for the daytime downward
recognises the importance of roof reflectance in low-rise residen- heat flow from a roof surface in a hot climate. In tropical regions
tial buildings (see also Akbari et al. [16] for more information). the daytime heat flow due to the outside–inside air temperature
According to the ANSI/ASHRAE 90.2 standard, the resistance (U- difference is typically in the order of 5 K, while the heat flow due to
value) of the ceiling insulation is increased by up to a factor of 1.5 solar radiation absorption on the roof surface is of the order of 20 K
when the roof reflectance and thermal emittance values have (see also Appendix A.1).
minimum values of 0.65 and 0.75, respectively. In other words, a It should be noted that the solar absorptance of the roof surface
reflective roof surface increases the effective thickness of the is not the only variable influencing the roof temperature. The
insulation by 50% (e.g. in parts of Florida), but no increase is thermal emittance from the roof surface, which determines the
provided for a climate zone such as Wisconsin. The benefit factors radiative heat exchange with the sky, also plays an important role
for a high roof albedo apparently were obtained from building (particularly in low wind conditions). Measurements suggest that
simulation (Akbari et al. [16]). roofs in tropical climates cool 1–6 8C below ambient temperature
The cooling benefit of a roof surface with high solar reflectance during nighttime and that the temperature depression may be
decreases with time as the surface accumulates dust and deposits. limited by the formation of dew (Khedari et al. [20]). Typically the
According to Eilert [17] the reduction in solar reflection of white sky temperature depression below ambient is 2–20 K and strongly
roofs is in the order of 10–30% with most of the reduction occurring depends on the level of cloudiness (Cooper et al. [21], Oliveti et al.
in the first year (presumably the degradation usually slows as rain [22]). For non-metallic surfaces (e.g. painted steel and roof tiles)
the thermal radiation emittance to the sky is typically about 0.9
3
A roof surface with high solar reflectance is similar to metal foil reflecting
while for bare metal surfaces it varies (for zinc/aluminium and
thermal radiation, only that it is active for approximately 12 h/day instead of 24 h/ galvanised steel the thermal emittance is approximately 0.3
day. [23,24]).
H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235 2227

Fig. 2. Approximate daytime roof temperature profile: (a) without solar radiation; (b) with solar radiation. See Nomenclature and Section 5 for definition of symbols (note
DT* = Tsol–air  Ti = DT + aG/ho).

4. An illustrative case study significant time lag in the roof temperature response. However,
knowledge of the steady-state heat transfer rate is useful in
The effect of a change in roof colour is illustrated in Fig. 3 for a analysing the daily heat gain from a roof.
residential building in Townsville, Australia. The high set building For steady state the solar radiation absorption on the roof
is of timber frame and clad construction and has a corrugated steel surface in Fig. 5(a) changes the heat flow to the outside and the
roof (see Fig. 4). Before the reflective white paint (Solacoat, with inside environment as shown in Fig. 5(d). Using the thermal
a = 0.20 in new condition) was applied the roof had been painted network of Fig. 5(b) and its reduced form in Fig. 5(c), one can find
with an aluminium type paint, but this was in poor condition with the steady-state heat flux that flows into the building, q*. The result
severe flaking. from Stoecker and Jones [25], which is derived in Appendix A.2 for
From Fig. 3 it is seen that the application of the reflective white completeness, is
paint (on the afternoon of the second day) has reduced the roof  
temperatures by about 20 K. More importantly however, the Q aG
q ¼ ¼ U ðT a  T i Þ þ (1)
increased solar roof reflectance has reduced the interior tempera- A ho
tures relative to the ambient temperature!
More specifically, for the 2 days before the application of the where U = (1/ho + 1/hi)1 is the overall heat transfer coefficient
reflective white paint the average ambient and interior air between the ambient and inside environment, Ta  Ti the ambient
temperatures were 21.8 and 23.9 8C, respectively, while for the to outside temperature difference, a the roof surface
2 days after the application of the white paint the ambient and solar absorptance, G the solar irradiance and ho the roof to
house interior temperatures were 23.2 and 24.0 8C. Before painting ambient heat transfer coefficient. Note in particular that
the interior temperature was 2.1 K above ambient and after the ho = hc + hr(T  Tsky)/(T  Ta) includes both the convective heat
interior temperature was only 0.8 K above ambient. This means the transfer with the outside air and radiative heat transfer to the
room temperature, relative to ambient, had been lowered by 1.3 K. sky—see Nomenclature and Appendix A.1 for further explanation.
This reduction in internal room temperature, although small in Eq. (1) gives the downward heat transfer rate for steady-state
magnitude, delivered a noticeable improvement in human comfort conditions. Collecting the term Ta + (aG/ho) in Eq. (1), which is
and reduced the desire for air-conditioning. called the sol–air temperature [25], the downward steady-state
heat flux can also be expressed as
5. Steady-state heat transfer rate
q ¼ UðT solair  T i Þ (2)
A roof does not experience a steady state while temperatures
and radiation are changing because the thermal mass causes a

Fig. 3. Temperature measurements of four successive days (4–7 September 2005)


from a house in Townsville, Australia before and after the application of a white Fig. 4. House in Townsville, Australia used for roof and interior temperature
reflective paint (Solacoat with a = 0.20 in new condition). measurements (after application of the reflective white paint).
2228 H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235

Fig. 5. Steady-state roof heat flow: (a) roof construction, (b) thermal network including separate emissivity circuit to sky temperature, (c) equivalent thermal network and (d)
heat flow.

The sol–air temperature is a hypothetical outdoor air tempera- for Ta  Ti > 0 the effective thermal resistance, R*, is smaller
ture that would cause the same heat flow for a shaded (roof) than R.
surface as that experienced by the illuminated one.
Eq. (1) shows the heat flow due to temperature difference and
7. Effect of thermal mass on the downward roof heat transfer
radiation can be obtained from the superposition of the heat flows
due to temperature difference, Ta  Ti, and radiation, G. It should be
Eq. (1) gives the downward steady-state heat transfer rate per
stressed that Eq. (1) or (2) cannot be used to calculate the
unit area from the roof into the building, q*. However, as already
instantaneous heat flow for a roof under non-steady state
discussed heat flow predicted by Eq. (1) usually is not reached
conditions.
because of the thermal mass of the roof and ongoing changes in
environmental conditions.
6. Definition of thermal resistance (R-value)
The effect of thermal mass on the heat transfer through the
building envelope has been extensively studied (e.g. see [26] for
For the subsequent development, a clear understanding of the
an early example and [27] for a review of methods). However,
definition of the thermal resistance value (R-value) is important.
for this study a specific aspect of the effect of thermal mass is of
The steady-state heat flow per unit area (heat flux), Q/A, through a
interest—the roof thermal mass influence on the integrated daily
wall or roof across which a fixed temperature difference, DT, is
heat gain for a fixed R-value. We have investigated this effect
applied can be expressed as
for a 1 mm steel (7.85 kg/m2), clay tile (50.0 kg/m2) and 10 cm
Q DT thick concrete (240 kg/m2) roof during a clear sky day. All three
q¼ ¼ U DT ¼ (3) roof constructions were assumed to have an unventilated air
A R
space and a suspended plasterboard ceiling of 10 mm thickness
where U is the overall heat transfer coefficient and R = 1/U is the (8 kg/m2) below the roof. Roof insulation was added so that
thermal resistance for steady state. The problem with Eq. (3) is that all three roofs had the same overall thermal resistance of
is does not apply for an illuminated surface. To make Eq. (3) 2.2 m2 K/W.
applicable to a sun-illuminated surface one needs to use the sol–air The results from a numerical simulation of the three roof
temperature difference (see also Section 5): constructions types are shown in Table 1 for three sky
temperatures and plots of diurnal roof temperatures can be found
Q DT solair in Appendix A.3. Table 1 suggests that the integrated daily heat
q ¼ ¼ U DT solair ¼ (4)
A R gain from the roof into the building is nearly independent of
where DTsol–air = Ta + (aG/ho)  Ti. Dividing Eq. (4) by (3) gives: the roof thermal mass. More specifically for the same value of Tsky
the daily (24 h) net heat gain differs less than 0.7% between the
q Q  DT solair different roof construction types.
¼ ¼ ¼F (5) This is an important result for this study as it suggests that
q Q DT
the integrated daily heat flow from an unventilated roof (not the
where F is the factor of increase in heat flow due to the solar instantaneous heat flow) to good approximation can be
illumination (in case of heat loss from the building, the solar calculated from Eq. (1) without considering heat capacitance
heating may decrease the heat loss). The latter factor allows one to effects! The result appears physically reasonable as the amount
express the roof heat flow, which includes the effect of solar of solar radiation absorbed by the roof does not depend on
illumination as
Table 1
 DT solair DT F DT DT Daily heat gain for three roof constructions (steel, tile and concrete roof) with
q ¼ ¼ ¼ ¼ (6) suspended plasterboard ceiling and R = 2.2 m2 K/W. The heat gain was calculated
R R R=F R
for three values of sky temperature. Ambient and inside building air temperatures
were assumed fixed at 24 8C
where R* = R/F may be interpreted as the ‘‘effective thermal
resistance’’ that a (sun-illuminated) roof or wall offers to inward Sky temperature Daily heat gain (kJ/m2)
heat flow (note R* ! R as the solar radiation G ! 0). The concept
Steel roof Clay tiled roof Concrete roof
of effective thermal resistance for cooling is useful as it enables
a direct comparison between illuminated roof surfaces of Tsky = Ta 349.8 350.0 351.0
Tsky = Ta  10 K 266.4 266.5 267.6
different surface colour (solar absorptance), but experiencing
Tsky = Ta  20 K 192.6 192.7 193.8
the same outdoor to indoor air temperature difference DT. Note
H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235 2229

the roof’s thermal capacitance and for a fixed R-value the 9.2. Average outside to inside temperature difference for air-
absorbed solar radiation must be dissipated at about the same conditioned buildings
ratio to the ambient and to the inside of the building. Moreover,
the result is consistent with the observation of Burch et al. [28] In a hot climate, where active heating is not necessary, most5
that in a hot climate with continuous positive heat gain even the benefit from thermal insulation and solar reflective coatings is
daily space cooling load can be predicted from steady-state gained when the ambient temperature rises above 24 8C. We
theory.4 therefore restrict the period of analysis to those times when the
ambient temperature rises above 24 8C and evaluate the tempera-
8. Daily heat gain from the roof ture difference from the number of cooling degrees hours CDH
(base 24 8C):
In order to judge the benefit of light roof colour on the heat gain
P
from a roof it is required to calculate the heat flow due to both all hours of the year ðT a  24Þþ
temperature difference and solar radiation over a whole day. As
DT̄ ¼ (9)
24  365
was suggested in the previous section, the daily quantity of heat
transferred from the roof into the building can be calculated where the + sign indicates that only temperature differences above
without consideration of the thermal mass by integrating Eq. (1) 24 8C are included in the summation. For locations where no
over a whole day: hourly temperature data is available DT̄ may be estimated from the
cooling degree days (CDD) with Ta = (Ta,max  Ta,min)/2.
fdaily heat gain from roof into buildingg We have evaluated Eq. (9) for several north Australian locations
Z   using the US EnergyPlus Weather Data Site [25] and the results are
aG
¼ U ðT a  T i Þ þ dt presented in Table 2. The population weighted average form the
day ho
Z Z    data of Table 2, using the north Australian conditions as an
a a illustrative example, is
¼U ðT a  T i Þdt þ G dt ¼ U DT̄ Dt þ H
day ho day ho
 
a DT̄  2:2 K (10)
¼ U Dt DT̄ þ Ḡ ¼ U DT̄solair Dt (7)
ho

where DT̄ is the average daily temperature difference between 9.3. Average outside to inside temperature difference for passive
outside and inside, H the daily horizontal surface irradiation, buildings
R
H ¼ day G dt, and Ḡ ¼ H=Dt the average daily irradiance with Dt
denoting the day length in seconds (=24  3600 s). The above For non-air-conditioned (passive) buildings the inner tempera-
analysis assumes that the heat gain per unit area is based on the ture is floating and the temperature difference DT̄ ¼ T̄a  T̄i tends
projected horizontal roof area or building floor area. In Eq. (7) it to be negative, i.e. the inside average air temperature tends to be
has been assumed that U and a/ho are both constant, in higher than the ambient one (e.g. DT̄ ¼ 2:1 K for the house of
particular that a is (approximately) independent of the Section 4 before application of the reflective paint). The inside
incidence angle. temperature elevation above ambient can be caused by the solar
heating of the building envelope and internal heat sources (e.g.
9. Average solar absorptance and environmental conditions electric appliances).
For a non-air-conditioned building a negative DT̄ means the
For the evaluation of Eq. (7) various average quantities, such as roof insulation loses some of it benefits (e.g. it may restrict the
the average building outside to inside temperature difference, desired nighttime cooling) while the benefits of solar reflection
DT̄ ¼ T̄a  T̄i , are required. remain unchanged. We make the conservative assumption that the
outside to inside temperature difference for non-air-conditioned
9.1. Average roof solar absorptance buildings is about zero, i.e.

The average roof solar absorptance depends on the type of roof DT̄  0 K (11)
construction used in a particular geographical location. A rough
estimate for the Australian use of roof colours is that 1/3 are dark
colours (e.g. grey and black), 1/3 medium colours (e.g. red and 9.4. Average solar radiation
green) and 1/3 light colours (e.g. white and cream). Using known
solar absorptance values for the above roof surfaces (see also The average radiation can be estimated for many locations on
Table A1), suggest the average roof solar absorptance approxi- earth from available solar radiation data (e.g. see Appendix G of
mately is Ref. [1]). Here horizontal surface averages of daily solar radiation
data by Morrison and Litvak [30] are used for north Australian
1 1 1 conditions (see Table 3). It is noted that close to the equator the
aave  0:9 þ 0:7 þ 0:5 ¼ 0:7: (8)
3 3 3 summer to winter solar radiation variation is small (less than 10%
for Darwin and 20% for Port Headland and Halls Creek).
The average solar absorptance of Eq. (8) is only a crude estimate Assuming that the average annual daily radiation on a
and does not include bare metal coatings (e.g. galvanised and zinc/ horizontal (roof) surface H̄  21:5 MJ=m2 for tropical regions in
aluminium coatings). Note that a roof solar absorptance of 0.7 has Australia6 with latitude less than 23.58 (approx. Zone 1 as defined
also been specified by the Building Code of Australia 2006 [3] for in BCA, 2005 [2]), we can estimate the average radiation during the
energy simulations as a nominal value upon which to compare
variations. 5
Due to internal loads some active cooling of a building may be required before
the ambient temperature reaches 24 8C.
4 6
Note the daily the space-cooling load is more likely to be affected by the The chosen value is supported by the contour map shown in the Australian
building’s thermal mass than the daily heat gain. Standard AS/NZS 4234:2008 (Fig. A1).
2230 H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235

Table 2 Using the population averaged DT̄air ¼ DT̄ ¼ 2:2 K from Eq. (10)
Average temperature difference between ambient and 24 8C for several North
for an air-conditioned building in northern Australia and the
Australian locations from Eq. (8). Temperature raw data from US EnergyPlus
Weather Data Site [29] approximate average values ā ¼ 0:7, Ḡ ¼ 250 W=m2 and
P ho = 25 W/m2 K from Eqs. (8), (12) and (13) in Eq. (14), suggests
City/town/place State Est. DT̄ ¼ CDH=8760 ½K
that in DTsol–air = 2.2 + 7 K, i.e. the temperature difference due to
population
solar radiation is more than three times as large as that due to air
Townsville/ Queensland 165,059 1.7 temperature difference (Eq. (5) gives F  4). Note the result for the
Thuringowa
temperature differences was calculated for a zero thermal mass
Cairns Queensland 132,765 1.7
Darwin Northern Territory 111,300 3.7
roof. However, as the daily heat gain shows little dependence on
Mackay Queensland 73,091 1.1 the roof thermal mass of the roof, the factor of increase in heat
Mt Isa Queensland 21,000 3.3 transfer due to solar radiation should also approximately apply for
Broome Western Australia 13,858 3.6 a roof with non-zero thermal mass.
Port Hedland Western Australia 12,697 3.4
Similarly, from Eq. (7) the average daily rate of heat flow per
Weipa (Cook Shire) Queensland 9,409 2.8
Tennant Creek Northern Territory 3,500 3.7 unit area (flux) for a roof with zero heat capacitance can be
Halls Creek Western Australia 1,289 4.0 expressed as
Wyndham Western Australia 800 5.8
  
Willis Island Coral Sea Territory 4 2.2 Q̄ a
q̄ ¼ ¼ U DT̄ þ Ḡ (15)
Sum/average 544,772 3.1 A ho

Evaluating Eq. (15) for an air-conditioned building under


average north Australian conditions for a total R-value (=1/U) of
Table 3 2.2 m2 K/W, gives an average daily heat flux of approximately
Annual average values of daily solar radiation H̄ for selected north Australian
4.2 W/m2 (=1/2.2(2.2 + 7.0) W/m2) for a (horizontal) roof with 0.7
locations from Morrison and Litvak [30]
solar absorptance.
Location Daily average solar On a daily average (24 h) basis this results in a roof heat gain of
radiation H̄ (MJ/m2)
approximately 361 kJ/m2 (=24  3600 s  4.2 W/m2). Note the
Port Headland, WA (20.388S latitude) 23.1 heat gain is different from that shown in Table 1 due to the air
Halls Creek, WA (18.238S latitude) 22.3 temperature difference driven component DT̄, which was pre-
Darwin, NT (12.428S latitude) 21.7
viously set to zero, and due to the fact that the radiation estimate Ḡ
Rockhampton (23.388S latitude) 19.0
is based on average rather than clear sky radiation.
Average 21.5
11. Effect of colour on roof downward heat transfer

The solar absorptance of a roof surface has been identified as a


day by dividing the radiation total by the day length (24 h):
substantial influence on the roof heat flow, and we now ask what is
H̄ 21:5  106 the objective benefit a light roof colour relative to an average one?
Ḡ ¼ ¼  250 W=m2 (12) Using Eq. (15) for the average rate of heat gain from a light-
Dt 24  3600
coloured roof, and dividing it by the corresponding equation for an
average roof gives:
9.5. Heat transfer coefficient from roof surface to ambient  
q̄ U DT̄ þ ða=ho ÞḠ
¼   (16)
Eq. (1) and the subsequent equations for the roof heat gain q̄ave U ref DT̄ þ ðaave =ho ÞḠ
depend on the heat loss coefficient to ambient, ho, which in turn
depends on the wind speed and the sky temperature. The higher In Eq. (16) aave is the average roof solar absorptance (here assumed
the heat loss coefficient, ho, and the smaller the effect of roof colour to be 0.7) and U and Uref denote the overall heat transfer coefficient
(solar absorptance). between the outside and inside for the coloured and reference roof,
For a roof, the heat transfer coefficient, ho, due to wind and sky
radiation typically is 15–25 W/m2 K [1], where the latter value is
suggested by the BCA, 2005 [2] for 3 m/s wind speed. Urban
environments offer some wind shielding (e.g. due to buildings and
trees) which reduces wind speeds at roof level. However, for the
subsequent work we conservatively adopt:

ho ¼ 25 W=m2 K (13)

10. Daily heat gain from roof due to solar radiation and
temperature difference

The sol–air temperature difference in Eq. (7) can be split into a


component due to real outside to inside temperature difference
and effective extra temperature difference due to solar radiation:

DT solair ¼ DT air þ DT solar


(14) Fig. 6. Roof average downward heat transfer as a function of solar absorptance
aG relative to an average roof with aave = 0.7 for DT = 2.2 K (north Queensland air-
¼ ðT a  T i Þ þ
ho conditioned building) and DT = 0 K (no air-conditioning) with ho = 25.0 W/m2 K.
H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235 2231

Table 4
Solar absorptance values form measurements of solar reflectance for weathered roof/coating materials. Note for opaque materials solar absorptance = (1  solar reflectance)
and that NSW stands for the Australian state New South Wales

Roof/coating Age (years) Solar absorptance Comments

Refl. white NSW home 8 0.36 Roof dirty and paint applied too thin
Refl. white NSW factory 8 0.45 Roof very dirty from cement plant
Refl. white NSW resort 4 0.33 Paint professionally applied
Standard white NSW resort Unknown 0.52 Front section of building
Refl. white painted flat sheet Unknown 0.26 Painted cover
Weathered galvanised flat sheet Unknown 0.66 Unpainted cover

respectively. The heat transfer coefficients can be expressed in 11.3. Effect of roof thermal emittance
terms of their corresponding thermal resistance as: 1/U = R and
1/Uref = Rref, where Rref is the reference thermal resistance deemed Eqs. (17) and (18) consider the effect of solar absorptance on the
to satisfy building regulations for minimum insulation. heat gain, but do not explicitly consider the effect of the roof
thermal emittance for long wave radiation exchange with the sky.
11.1. Equal insulation but different roof colours The roof emittance to the sky affects the overall resistance to the
ambient (=1/ho), but ho has been assumed the same for both the
One way to assess the effect of roof colour is to compare the coloured and average surface. However, for the same environ-
reduction in downward heat flow when the bulk insulation for the mental conditions a bare metal surface, say with zinc/aluminium
coloured roof is equal to that of the average one (U = Uref). Using coating, will have a lower heat loss coefficient, ho, than a painted
Eq. (16) this yields: surface. Using different values of ho for the coloured and average
surface in Eqs. (17) and (18), the effect of roof emittance on the
q̄ DT̄ þ ða=ho ÞḠ DT̄solair heat gain can be considered (see Appendix A.1 on how to calculate
¼ ¼ (17) the effect of surface emissivity on ho). A lower roof surface thermal
q̄ave DT̄ þ ðaave =ho ÞḠ DT̄solair;ref
emissivity increases the influence of the solar absorptance value
for bare metal coatings, and explains why, for example, a corroded
Eq. (17) gives ratio of downward heat gain between a coloured and
bare metal surface exposed to sunlight can get nearly as hot as a
an average coloured roof surface, both having the same insulation
black one.
below the roof surface.
Eq. (17) is readily evaluated for different values of solar
absorptance and air temperature differences, and the results are 12. Effect of dust and aging on the roof solar absorptance
shown in Fig. 6. Fig. 6 shows that reducing the solar absorptance
from 0.7 to 0.4 reduces the average downward heat transfer from Two studies dealing with changes in solar roof absorptance over
the roof by about 33% for an air-conditioned building and 43% for a time have already been discussed in Section 2, and it is clear that
non-air-conditioned building in north Australia. Note for an air- white coatings do not retain their original reflectivity (e.g. because
conditioned building with positive DT̄ð¼ T̄a  T i Þ the insulation dust and dirt build up [17]).
offers benefits even when the roof solar absorptance a approaches Cool Roof Rating Council website [31] lists 3-year aged-
zero. However, for a negative DT̄ (denoting a passive building with performance ratings for roofing products, where samples are
internal load) the roof insulation can restrict the cooling of a weathered for 3 years at test farms in the U.S. The tested 13
building (for a < 0.22 in Fig. 6). reflective products with initial absorptance ranging 0.18–0.21
exhibited an average 0.12 increase in absorptance after 3 years.
11.2. Same downward average heat flow Table 4 shows reflectance measurements of aged reflective
white paint, standard white, and galvanised steel, which were
Another way of investigating the effect of roof colour is to find carried out by the authors of this paper using ASTM E1918-97. It is
the thickness of insulation below a light-coloured roof that gives apparent from Table 4 that the value and deterioration of solar
the same downward heat flow as it is achieved for a roof with reflectance is dependent on the environment. However, Table 4
average solar absorptance and standard insulation thickness. The
R-value7 for a roof of a particular colour that gives the same
downward heat flow as that for the reference or nominal colour
can be calculated by assuming qi ¼ qi;ref and using Eq. (16) with 1/
U = R and 1/Uref = Rref:

R DT̄ þ ða=ho ÞḠ


¼ (18)
Rref DT̄ þ ðaave =ho ÞḠ

Eq. (18) gives the fraction of R-value that is necessary for a coloured
roof to provide the same average heat gain or insulation. The RHS of
Eq. (18) is the same as the RHS for Eq. (17) and therefore the plot in
Fig. 6, originally for the heat gain ratio q̄ =q̄ref , also can be used for
the R/Rref ratio.

7 Fig. 7. Average increase in solar absorptance due to dust and weathering.


Based on the outside to inside air temperature difference DT.
2232 H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235

Table 5
Roof colour contribution to the R-value and heat gain in a hot and sunny climate for different surfaces in the weathered state. The effects of roof colour have been calculated
from Eqs. (22) and (18), respectively, using Ḡ ¼ 250 W=m2 , ho = 25 W/m2 K (bracketed values are for ho = 15 W/m2 K). Note for DT̄ ¼ 0 K ho cancels out of Eqs. (22) and (17)

Roof colour Solar absorptance Nominal a Rcolour/Rref DT̄ ¼ 2:2 K Rcolour/Rref DT̄ ¼ 0 K qi =qi;ref DT̄ ¼ 2:2 K qi =qi;ref DT̄ ¼ 0 K
a value range value

Reflective white (bright white) 0.45 > aweathered 0.35 0.38 (0.42) 0.50 0.62 (0.58) 0.50
Light colour (e.g. greyish 0.6 > aweathered  0.45 0.55 0.16 (0.18) 0.21 0.84 (0.82) 0.79
white, light yellow)
Medium colour (e.g. green, red) 0.8 > aweathered  0.6 0.7 0.00 0.00 1.00 1.00
Dark colour (e.g. grey, black) aweathered  0.8 0.85 0.16 (0.18) 0.21 1.16 (1.18) 1.21

also suggests that exposed reflective paint or white steel roofing than predicted by Eq. (19) as they assume a visibly duller
maintains a significantly higher reflectance than galvanised steel appearance due to weathering.
and standard white.
The paper by Griggs et al. [10], more generally states that: 13. Equivalent R-value due to roof colour
‘‘weathering tends to reduce the reflectance of a light roof and
increases the reflectance of a dark roof’’. If one assumes that the With an estimate for the reduction in roof reflectance due to
weathering tends to change the solar reflectance towards a weathering in place, the effective R-value for a roof colour can now
particular reflectance value,8 the data from Table 4 may be used to be determined. The difference between the reference Rref and R in
estimate the weathered solar absorptance value. Eq. (18) may be credited in a hot climate as a heat transfer
The reflective white paint used for this study has a solar resistance that is provided by the particular roof colour, i.e.
absorptance of approximately 0.20 (Solacoat) when new and when
it is applied with the recommended coverage. The average solar Rcolour ¼ Rref  R (21)
absorptance of the reflective white weathered 8 years according to
Table 4 is about 0.35 (=(0.36 + 0.45 + 0.33 + 0.26)/4) and therefore It is instructive to calculate the Rcolour credit for cooling for a
the increase in absorptance is 0.15 (=0.35  0.20). If it is now reflective white roof surface that has a (weathered) solar
assumed that the absorptance increase (loss in reflectance) is the absorptance of a = 0.35. Using the previously determined values
proportional to the difference anew  adust, the increase (or for northern Australia (Rref = 2.2 m2 K/W, ho = 25 W/m2 K,
decrease) in absorptance may be estimated as DT̄ ¼ 0:0 K, aref = 0.7, and Ḡ ¼ 250 W=m2 ) in Eq. (18) gives:
R = 1.1 m2 K/W, i.e. a house roof with a solar absorptance of 0.35
adust  anew only requires a R-value of 1.1 m2 K/W to achieve the same
Da ¼ 0:15 (19)
adust  0:20 ‘‘insulation’’ as a nominal roof with 0.7 solar absorptance and R-
value of 2.2 m2 K/W. Hence, for this particular case one may assign
so that the weathered value for the solar absorptance may be an effective R-value for cooling of 1.1 m2 K/W (=2.2  1.1 m2 K/W)
expressed as to the reflective roof paint.
Dividing Eq. (21) by Rref allows the effect of roof colour to be
aweathered ¼ anew þ Da (20) expressed in more general terms:

On completion of this study it was found that the California Rcolour R DT̄ þ ðaweathered Ḡ=ho Þ
Energy Commission had already proposed a very similar expres- ¼1 ¼1
Rref Rref DT̄ þ ðaave Ḡ=ho Þ
sion (reported by Prado and Ferreira [32]), in which a correspond-
ing solar absorptance of adust = 0.8 had been assumed. Adopting
DT̄solair
¼1 (22)
this value here we note that our expression based on independent DT̄solair;ref
data closely agrees with the one by the California Energy
Rcolour/Rref can be interpreted as the fraction of thermal
Commission. The suggested relationship between the absorptance
resistance that is provided by the roof colour when the average
value for weathered and new material is depicted in Fig. 7. Using
daily heat gain is the same as for a roof of average colour, i.e. for
Eqs. (19) and (20) for the ordinary white steel roof (assumed to
q ¼ qref . While Eq. (22) provides a quantitative measure for the
have a solar absorptance of 0.39 for the new material (see
effect of roof colour, the comparison between the various roof
Appendix A.1)) with adust = 0.8, gives a solar absorptance increase
colours is facilitated with some discrete classes. Table 5 shows
through weathering of 0.11 and a solar absorptance for the
results obtained from Eqs. (22) and (17) for four nominal roof
weathered material of 0.50. The latter value compares to the solar
colours and two outside-to-inside average air temperature
absorptance inferred from reflectance measurement for ordinary
differences. The results obtained for Rcolour/Rref from Eq. (22)
white steel roofing in Table 4 of 0.52.
relate to the case of equal daily downward heat flow, while the
The relationship for the increase in solar absorptance suggested
results obtained for qi =qi;ref from Eq. (17) relate to the change in
in Eq. (19) is only a rough estimate, however, it is noted that the
heat gain for roofs with the same insulation below the roof
19% (=0.15/0.80)9 decrease in solar reflectance for the reflective
surface.
white paint is in line with the 10–30% decrease in solar reflectance
Table 5, for example, shows that a non-air-conditioned
reported by other studies for reflective white roof materials (see
building with a weathered white reflective roof surface (with
Section 2). One would expect the change in reflectance to depend
nominal a = 0.35) can substitute 50% of the thermal resistance
somewhat on the local air-pollution and climate.
required for an average roof with a = 0.7. Alternatively, if the roof
For zinc/aluminium and galvanised bare metal surfaces, the
has the same thermal resistance as the average one, then the heat
changes in solar absorptance are likely to be significantly larger
gain from a weathered white reflective roof surface is only about
8
A roof with having a solar reflectance close to that of the dust and dirt deposits
57% of that from an average roof. It is noted that the percentages
should not change its absorptance significantly (neglecting surface oxidisation). in Table 5 are independent of the reference value Rref that may be
9
Note solar reflectance = 1  a for an opaque surface. prescribed.
H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235 2233

Table 5 applies for high emittance roof surfaces such as paints


and tile surfaces with thermal emittance 0.9. For bare metal
surfaces such as zinc/aluminium the ratios Rcolour/Rref and qi =qi;ref
should be directly calculated from Eqs. (22) and (17), respectively,
with different ho values for the bare and average surface.
Substituting the values of Table A1 for zinc/aluminium coated
steel with the temperature in Kelvin in ho = hc + hr(T  Tsky)/
(T  Ta), where hr ¼ seðT 2 þ Tsky
2
ÞðT þ T sky Þ, gives ho,bare = 20 W/
2 Fig. A1. Schematic diagram of a surface exposed to wind, sky and solar radiation.
m K. Using this reduced ho value and abare = 0.55 for the
weathered state in Eqs. (22) and (17), suggests a bare metal
surface has a similar heat gain than an average surface with
a = 0.7. data. Last but not least we would like to acknowledge the great
encouragement and support for this study by Dr. Jonathan
14. Discussion Harris.

The results in Table 5 demonstrate that the roof colour has a


significant effect on the heat gain, particularly for non-air-
Appendix A
conditioned (or passive) buildings. The assumption that for passive
buildings the inside temperature equals the outside should yield a
conservative estimate for the benefit of a solar reflective roof A.1. Steady-state temperature of a surface
coating.
The reflectance and thermal emittance measurements listed The temperature of a roof surface depends on the environmental
in Table A1 suggest that non-white roof coatings that reflect conditions (solar radiation, ambient temperature, wind velocity
invisible parts of solar radiation still offer appreciable and sky temperature), the optical properties of the roof surface
benefits over their equal coloured counterpart (see also (solar absorptance and thermal emittance) and the underside
[12,35]). insulation.
In hot climates roof insulation may hinder the desired night- Fig. A1 shows a schematic diagram of a surface having incident
time cooling and a light roof colouring in combination with only a
solar radiation and heat loss to the environment. For a surface
reflective roof air space may be preferable for non-air-conditioned
with perfect insulation10 at the underside and steady state the
buildings.
Gaining greater energy efficiency through the choice of absorbed solar radiation must equal the heat loss due to
appropriate roof colours is simple and cost effective. Reflective convection with the air and thermal radiation heat exchange with
(or light-coloured) paints may be used for the retrofit or conversion the sky:
of roof surfaces. aG ¼ hc ðT  T a Þ þ hr ðT  T sky Þ
(A.1)
¼ ho ðT  T a Þ
15. Conclusions
See Nomenclature for the meaning of symbols. Solving Eq. (A.1) for
The above analysis has shown that it is possible to quantity the surface temperature gives:
the effect of colour on the roof heat gain. Consequently, we are
able to assign an R-value credit to the reflectivity of a roof aG þ hc T a þ hr T sky aG
T¼ or T ¼ Ta þ (A.2)
surface. However, the analysis at this stage is restricted to hc þ hr ho
locations where there are no heating requirements (cooling
only). Table A1 shows the results of evaluating Eqs. (A.1) and (A.2) for
The calculations suggest that in hot climates a significant various surfaces for an ambient temperature of 25 8C and solar
reduction in downward heat flow can be achieved by using a light radiation of 1000 W/m2.
or reflective roof colour instead of a dark one. The reduction in
downward roof heat flow means a reduction in air-conditioning A.2. Derivation of the steady-state roof heat flow equation
load or an increase in human comfort.
Measurements of the reduction in reflectance of light roof Consider the thermal network shown in Fig. 5(c). For steady state,
colours due to aging and weathering indicate that much of the
the sum of the energy flowing in and out of the roof surface due to
reduction occurs in the first year and that long-term average values
temperature differences and solar radiation absorption must equal
for the roof reflectance can be estimated.
zero. Moreover, the heat flow to the inside must equal the net heat
A classification of roof colours has been made and the relative
benefits are calculated for the case of downward roof heat flow from between roof and ambient and the solar radiation absorbed
transfer. The concepts presented in this paper have the potential on the roof surface. Hence, using the previously defined symbols, we
to be extended for buildings that require both cooling and can formulate the following energy balances for the roof surface and
heating. heat flows:
ho ðT a  TÞ þ hi ðT i  TÞ þ aG ¼ 0 (A.3)
Acknowledgements
q ¼ hi ðT  T i Þ ¼ ho ðT a  TÞ þ aG (A.4)
The support by Coolshield International Pty Ltd for the
completion of this study and the proof reading of the manuscript 10
Note the back heat losses are normally not zero, but an order of magnitude
by Mr. Graham Aldridge is gratefully acknowledged. Particular smaller than the front heat losses to the environment. Therefore, the temperature in
thanks goes to Nick Holcz from the West Australian Climate Eq. (A.2) approximately gives the temperature of a surface that is well insulated at
Services Centre for providing the long-term air temperature the back.
2234 H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235

Table A1
Steady-state temperatures for surfaces without back loss (G = 1000 W/m2, hc = 17 W/m2 K, Ta = 25 8C, and Tsky = Ta  10 K)

Surface a e T (8C) hr (W/m2 K) ho (W/m2 K)

White highly reflective paint (new Solacoat) 0.20y 0.88y 32 5.2 30


Standard White steel roofing (new) 0.39 0.9* 40 5.6 26
Green steel roofing (new) 0.75 0.9* 55 6.0 25
Grey steel roofing (new) 0.88 0.9* 60 6.2 25
Black (oil paint) 0.90** 0.94** 61 6.4 25
Conventional red (new) 0.70+ 0.9** 53 5.9 25
Coolshield Dynasty Natural red (near-infrared reflective, new) 0.50+ 0.89+ 45 5.6 25
Zinc/aluminium coated steel roofing (new) 0.38 0.3* 44 1.9 20
Galvanised surface (weathered) 0.80*** 0.28*** 66 2.0 19
Light red tilted roof 0.63 0.9* 50 5.8 25

All property values are taken from Suehrcke [33] except: y, Tuquabo and Bell [34]; +, Coolshield International Pty Ltd; *, estimated; **, Siegel and Howell [24]; ***, Incropera
and DeWitt [23].

Solving Eqs. (A.3) and (A.4) for the temperature difference and
the sky temperature being below the ambient one, i.e. the roof loses
adding yields:
heat during nighttime to the cold sky.

References

[1] J. Duffie, W. Beckman, Solar Engineering of Thermal processes, 3rd ed., Wiley,
New York, 2006, Chapters 1 and 3.
[2] Building Code of Australia 2005, Canberra, ACT, Australian Building Codes Board,
2005.
[3] Building Code of Australia 2006, Canberra, ACT, Australian Building Codes Board,
2006.
[4] H. Akbari, P. Berdahl, M. Pomerantz, Heat Island Group, Cool Roofs, Lawrence
Berkeley National Laboratory, 2000, Berkeley, California, http://eetd.lbl.gov/Hea-
tIsland/CoolRoofs/, last accessed: 17 September, 2006.
[5] TRNSYS 16 (Transient system simulation program), University of Wisconsin,
where the equation below the line is the sum of the two above the 2005, Madison, USA, http://sel.me.wisc.edu/trnsys, last accessed: 2006.
line and 1/U = 1/ho + 1/hi is the overall heat transfer resistance [6] G.A. Florides, S.A. Kalogirou, S.A. Tassou, L.C. Wrobel, Modeling of the modern
houses of Cyprus and energy consumption analysis, Energy 25 (2000) 915–937.
between the outside and inside. Now solving the above equation [7] D. Parker, J. Sherwin, J. Sonne, S. Barkaszi, FSEC-CR-904-96, Demonstration of
for q* gives Eq. (1). Cooling Savings of Light Colored Roof Surfacing in Florida Commercial Buildings:
Our Savior’s School, Florida Solar Energy Center, University of Central Florida,
Cocoa, Florida, USA, 1996, download: http://www.fsec.ucf.edu/en/publications/
html/fsec-cr-904-96/.
A.3. Roof heat flow for various roof construction types [8] H. Suehrcke, Thermal performance simulation of a building using Engineering
Equation Solver (EES), Energy and Buildings, in preparation.
[9] J. Simpson, E. McPherson, The effects of roof albedo modification on cooling loads
Fig. A2 shows that the reduction in daytime peak heat flow due to on scale model residences in Tucson, Arizona, Energy and Buildings 25 (1997)
127–137.
an increase in thermal mass is accompanied by an approximately
[10] E. Griggs, T. Sharp, J. MacDonald, ORNL-6527, Guide for estimating differences in
proportionate increase in the duration of downward heat flow. This building heating and cooling energy due to changes in solar reflectance of a low-
suggests that the thermal mass of the roof does not significantly affect sloped roof, Oak Ridge National Laboratory-Energy Division, Oak Ridge, USA,
1989, download: http://eber.ed.ornl.gov/commercialproducts/Reflroof.htm.
the integrated daily heat gain through the roof (see also Table 1). Note [11] A. Granja, L. Labaki, Influence of external surface colour on the periodic heat flow
that the slightly negative heat flow during nighttime is the result of through a flat solid roof with variable thermal resistance, International Journal of
Energy Research 27 (2003) 771–779.
[12] R. Levinson, H. Akbari, J.C. Reilly, Cooler tile-roofed buildings with near-infrared-
reflective non-white coatings, Building and Environment 38 (2007) 181–188.
[13] AS 4859.1:2002, Materials for the Thermal Insulation of Buildings—General
Criteria and Technical Provisions, Standards Australia, Sydney.
[14] R. Clarke, A. Delsante, Advisory note NSB 179A, Do ‘insulating’ paints work?
CSIRO, Division of Building and Construction & Engineering, Melbourne, Australia,
1994.
[15] ANSI/ASHRAE Standard 90.2-2004, Energy Efficient Design of Low-Rise Residen-
tial Buildings, Section 5.5, ASHRAE, Atlanta, USA.
[16] H. Akbari, S. Konopacki, D. Parker, Updates on Revision to ASHRAE Standard 90.2:
Including Roof Reflectivity for Residential Buildings, Lawrence Berkeley National
Laboratory, University of California, Berkeley, California, 2000, http://eetd.lbl.-
gov/HeatIsland/PUBS/2000/45369.pdf, accessed: 18 September 2006.
[17] P. Eilert, High Albedo (Cool) Roofs: Codes and Standards Enhancement (CASE)
Study, Pacific Gas and Electric Company, San Francisco, USA, 2000.
[18] H. Akbari, A. Berhe, R. Levinson, S. Graveline, K. Foley, A. Delgado, R. Paroli, Aging
and Weathering of Cool Roofing Membranes, Department of Energy’s (DOE)
Information Bridge, Oak Ridge, USA, 2006, download: www.osti.gov/bridge/serv-
lets/purl/860745-BAdlvk/860745.PDF, accessed: 16 May 2008.
[19] R. Levinson, P. Berdahl, A. Berhe, H. Akbari, Effects of soiling and cleaning on the
reflectance and solar heat gain of a light-colored roofing membrane, Atmospheric
Environment 39 (2005) 7807–7824.
[20] J. Khedari, J. Waewsak, S. Thepa, J. Hirunlabh, Field investigation of night radiation
Fig. A2. Downward heat flow qi for three roof constructions (1 mm steel, clay tiles cooling under tropical climate, Renewable Energy 20 (2) (2000) 183–193.
and 100 mm concrete) into the underlying room during a clear day (R = 2.2 m2 K/W [21] P.I. Cooper, E.A. Christie, R.V. Dunkle, A method of measuring sky temperature,
for all roof constructions and hi = 1/(2.2  0.04)). Solar Energy 26 (1981) 153–159.
H. Suehrcke et al. / Energy and Buildings 40 (2008) 2224–2235 2235

[22] G. Oliveti, N. Arcuri, S. Ruffolo, Experimental investigation on thermal radiation [30] G. Morrison, A. Litvak, Report No. 1/1999, Table 7, 1999, Condensed Solar
exchange of horizontal outdoor surfaces, Building and Environment 38 (2003) Radiation Data Base for Australia, Solar Thermal Energy Laboratory, University
83–89, Cities: Theoretical Applied Climatology 85 (2006) 185–201. of New South Wales, Sydney, Australia, 1999, download: http://solar1.mech.uns-
[23] F.P. Incropera, D.P. DeWitt, Table A.12, in Fundamentals of Heat and Mass w.edu.au/glm/glm-papers.htm.
Transfer, Wiley, New York, 2002. [31] Cool Roof Rating Council website, http://www.coolroofs.org/, last accessed: April
[24] R. Siegel, J. Howell, Appendix D, in: Thermal Radiation Heat Transfer, McGraw Hill, 2008.
New York, 1981. [32] R. Prado, Ferreira, Measurement of albedo and analysis of its influence the
[25] W. Stoecker, J. Jones, Refrigeration and Air Conditioning, Sections 4.12 and 4.13, surface temperature of building roof materials, Energy and Buildings 37 (2005)
McGraw Hill, New York, 1982. 295–300.
[26] E. Danter, Periodic heat flow characteristics of simple walls and roofs, Journal of [33] H. Suehrcke, Solar reflectance of white ‘‘Solacoat insulating paint’’ and other
the Institution of Heating and Ventilating Engineers July (1960) 136–146. common roofing materials, Consultancy Report for COOLshield International,
[27] C.A. Balaras, The role of thermal mass on the cooling load of buildings. An Queensland, 2006.
overview of computational methods, Energy and Buildings 24 (1996) 1–10. [34] T. Tuquabo, J. Bell, Consultancy Report for COOLshield International, Queensland
[28] D.M. Burch, S.A. Malcolm, K.L. Davis, The effects of wall mass on summer space University of Technology, 2006.
cooling of six test buildings, ASHRAE Transactions 90 (1984) 5–21. [35] R. Levinson, P. Berdahl, H. Akbari, Solar spectral optical properties of pigments.
[29] US EnergyPlus Weather Data Site, http://www.eere.energy.gov/buildings/ener- Part II. Survey of common colorants, Solar Energy Materials & Solar Cells 89 (4)
gyplus/cfm/weather_data.cfm, last accessed: June 2007. (2005) 351–389.

Você também pode gostar