Sie sind auf Seite 1von 12

Energy and Buildings 40 (2008) 22242235

Contents lists available at ScienceDirect

Energy and Buildings


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

Effect of roof solar reectance 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 School of Architectural, Civil and Mechanical Engineering, Victoria University, Australia c Institute of Sustainability and Innovation Melbourne, Victoria University, Australia d Townsville, Queensland, Australia
b

A R T I C L E I N F O

A B S T R A C T

Article history: Received 20 December 2007 Received in revised form 6 June 2008 Accepted 18 June 2008 Keywords: Roof solar absorptance Solar reectance Roof heat gain Thermal resistance Hot climate Cool roof

The effect of the roof solar reectance on the thermal performance of a building is often ignored. However, there are signicant differences in heat gain from light and dark-coloured roof surfaces. In this paper an equation for the average daily downward heat ow of a sunlit roof is derived. Using building simulation, it is rst shown that the thermal mass of the roof does not signicantly affect the overall daily heat gain (although it causes a time lag and reduction in peak heat ow). As a consequence the daily heat gain from the roof may be estimated by integrating the equation for the steady-state downward heat transfer over the day. For north Australia the derived equation suggests that a light-coloured roof has about 30% lower total (air temperature difference and solar-driven) heat gain than a dark-coloured one. The effect of aging (change in solar reectance with time) is considered in the calculations and a relationship between the solar absorptance of new and aged material is suggested. A classication of roof colours with respect to their solar absorptance (dark, medium, light and reective) is proposed to enable a quick and simple assessment of the effect of roof colour on the heat gain and R-value. 2008 Elsevier B.V. All rights reserved.

1. Introduction The thermal performance of a building is affected by the solar absorptance of the roof. During clear sky conditions up to about 1 kW/m2 of solar radiation can be incident on a roof surface, and between 20% and 95% of this radiation is typically absorbed. The roof colour that is apparent from the reected visible1 part of the solar radiation usually gives an indication of the value of solar absorptance (e.g. a black surface with low visible reectance suggests a high solar absorptance). In this paper we analyse the building heat gain from a roof in a warm/hot climate. In locations close to the equator (tropical areas with latitude angle 23.58 or less) ambient temperatures and solar radiation levels are sufciently high that even during winter buildings do not require active heating. Daytime heat ow from a sun-exposed roof surface is essentially only in downward direction and the

downward heat ow generally is undesired, as it tends to overheat the building or put extra load on an air-conditioning system. Fig. 1 illustrates the difference between downward and upward heat ow through a roof space. For downward heat ow, schematically depicted in Fig. 1(a), the heat transfer between the roof and the ceiling is predominately due to thermal radiation. For upward heat ow, on the other hand, where the air in the roof space is heated from below, heat is transferred by both convection and radiation. The downward heat ow from the roof can be reduced through the use of a light roof colour, reective foil and/or insulation. While the installation of roof insulation and reective foil has now become mandatory in many countries (e.g. Building Code of Australia 2005 [2]), the choice of an appropriate roof colour is not generally accepted as a contribution to the insulation.2 The problem is that no ready and general assessment procedure for the benet of light roof colour appears to be available, although the inuence of roof colour on the building energy efciency 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. E-mail addresses: suehrcke@bigpond.com, harry.suehrcke@jcu.edu.au (H. Suehrcke). 1 Nearly 50% of the solar radiation energy reaching the earths surface is in the visible range (0.380.78 mm) [1]. 0378-7788/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.enbuild.2008.06.015

2 An exception is the Building Code of Australia (BCA) 2006 [3] for commercial buildings occupied during daylight hours. The BCA now enables a reduction in ceiling insulation R-value of 1.0 m2 K/W to be made for a roof solar absorptance of 0.35, or less for particular hot climates regions.

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

2225

Nomenclature area (m2) Factor of increase in heat ow due to solar illumination G solar radiation per unit area (beam + diffuse) (W/m2) G average radiation incident on a (horizontal) roof H=Dt (W/m2) surface, G convective heat transfer coefcient between roof hc and ambient air (W/m2 K) hi heat transfer coefcient between roof upper surface and building inside air, including the sum of thermal resistances through roof construction and inside air lm (W/m2 K) ho outside total heat transfer coefcient between roof and ambient air (W/m2 K). Note that ho is linearized and includes the radiative cooling to the sky, ho = hc + hr(T Tsky)/(T Ta), see Appendix A.1 for details hr radiative heat transfer coefcient between roof 2 T T sky surface and sky hr s eT 2 Tsky (W/m2 K) R H daily horizontal surface irradiation, H G dt (MJ/m2) * * q = Q /A rate of heat ow per unit area (ux) from an illuminated roof to the building inside (W/m2) q = Q/A rate of heat ow per unit area (ux) from an unilluminated (shaded) roof to the building inside (W/m2) R thermal resistance (R = 1/U) (m2 K/W) * effective thermal resistance for an illuminated wall R (m2 K/W) ambient (outside) temperature (8C) Ta inside temperature (average internal space air) (8C) Ti Tsky sky temperature (K) (typically 220 K below ambient temperature) Tsolair solair temperature (Ta + aG/ho) (8C) Dt day length in seconds (=24 3600 s) (s) DT temperature difference between outside and inside air (K) DT* Effective temperature difference for a solar radiation-illuminated surface (DT* = DTsolair) (K) average daily temperature difference between DT outside and inside (K) solair average daily solair temperature difference DT DT solair (K) DT U overall heat transfer coefcient between the ambient and inside (W/m2 K), Note: 1/U = 1/ ho + 1/hi A F = q*/q

aweathered solar absorptance of weathered roong material e surface emissivity for long-wave thermal radiation s StefanBoltzmann constant (5.67 108 W/
m K 4)

Akbari et al. [4] provides extensive experimental information about roof materials. Building simulation (e.g. TRNSYS [5] as used, for example, by [6]) can be used to quantify the relative effect of roof and wall colour on the energy ow of each building element. A quantitative assessment of the effect of roof colour (more correctly roof solar absorptance and thermal emittance) is complicated by the fact that (1) For locations that require both heating and cooling the benets from a light roof colour are not always positive and the downward and upward heat ow cases require separate treatment. (2) The heat ow due to solar absorption on the roof surface combines with that due to air temperature differences between the outside and inside. Both heat ows need to be evaluated in order to judge the relative effect of roof colour. (3) The heat ows due to solar absorption and outside to inside air temperature difference are variable and inuenced by the thermal mass of the roof. (4) The solar absorptance of a roof will change with time due to dust and aging.

However, restricting the analysis to locations where there are no heating requirements (cooling only) simplies the problem. Therefore this paper only deals with the downward heat transfer case, as it is typical, for example, in northern parts of Australia (Zone 1 in the Building Code of Australia 2006 [3]). An effective Rvalue for cooling that takes into account the solar absorption (roof colour) is derived. 2. Review of some existing literature The benets of light roof colour have been noted many times, particularly in locations where the sun is almost directly overhead and for single story buildings. For example, Parker et al. [7] reported that the space cooling requirements, after application of reective roof coatings on nine Florida homes, decreased by 19% and the staff of a Florida school found that interior comfort noticeably improved after the grey bitumen roof surface was painted white. Suehrcke [8], on the other hand, using a numerical simulation of a building, suggested that the peak values of heat ow from a roof could reduce by as much as 60% when a white surface replaces a corroded galvanised one. Simpson and McPherson [9] found from measurements of 1/4scale models in Arizona, that ceiling insulation is more effective in reducing daytime heat gain than increased roof albedo. However, Simpson and McPherson also made the interesting observation that on a 24-h basis an increased albedo was about as effective as addition of ceiling insulation in reducing building heat gain. This is explained in the paper due to the enhanced nighttime heat loss through the uninsulated ceiling of the high albedo roof. Griggs et al. [10] provide a comprehensive and very useful study on the effect of roof colour. Their study includes a work sheet to calculate the energy cost savings as a result of a roof reectance change and its use is demonstrated with two examples. However, the report by Griggs et al. unfortunately does not contain explicit

Greek symbols a absorptance for solar radiation (fraction of solar


radiation energy absorbed from all wavelengths and directions) average roof solar absorptance (approx. 0.7) solar absorptance of dust and dirt (assumed here 0.8)

aave adust

2226

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

Fig. 1. Downward and upward heat transfer through a roof space: (a) upward heat transfer; (b) downward heat transfer. In the above gure 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 hidden within the work sheets. There are analytical methods for the calculation of roof heat gain (e.g. see Granja and Labaki [11]). However, many of these methods appear to be limited to simple thermal conduction problems (e.g. a solid concrete roof with insulation) and do not provide general answers on the effect of roof colour. Perhaps the best calculation method for the roof heat gain accompanies a paper on non-white roof coatings that reect the invisible parts of solar radiation (Levinson et al. [12]). The method recognises that changes in solar roof absorption approximately cause proportional changes in ceiling heat ux. However, the issue of the roof thermal mass is not fully resolved and the ceiling heat ux due to outside to inside air temperature differences is not explicitly included in the equations. The Australian/New Zealand standard on thermal insulation AS 4859.1:2002 [13] acknowledges (Appendix G) that external sunlit surfaces with low solar absorptance stay cooler and help reduce the heat ow. Specically, in Appendix G it also says . . ., similar thermal benets may be derived by adding a certain thermal insulation or by changing solar reectance by a certain amount. However, the notion of assigning an insulation value was rejected because of the variability of conditions and because solar reectance offers no benet when the sun is not shining. The CSIRO advisory note (Clarke and Delsante [14]) with the title: Do insulating paints work points out that the thermal resistance of a thin paint lm is effectively zero and that reective paint cannot provide insulation for a building, but keep it cool. This statement is of course correct, though it is noted that for hot and sunny locations bulk insulation or a reective roof coating can both achieve the same (daytime) effectreduce the downward heat ow.3 The recent ANSI/ASHRAE Standard 90.2-2004 [15] now recognises the importance of roof reectance in low-rise residential buildings (see also Akbari et al. [16] for more information). According to the ANSI/ASHRAE 90.2 standard, the resistance (Uvalue) of the ceiling insulation is increased by up to a factor of 1.5 when the roof reectance and thermal emittance values have minimum values of 0.65 and 0.75, respectively. In other words, a reective roof surface increases the effective thickness of the insulation by 50% (e.g. in parts of Florida), but no increase is provided for a climate zone such as Wisconsin. The benet factors for a high roof albedo apparently were obtained from building simulation (Akbari et al. [16]). The cooling benet of a roof surface with high solar reectance decreases with time as the surface accumulates dust and deposits. According to Eilert [17] the reduction in solar reection of white roofs is in the order of 1030% with most of the reduction occurring in the rst year (presumably the degradation usually slows as rain
3 A roof surface with high solar reectance is similar to metal foil reecting thermal radiation, only that it is active for approximately 12 h/day instead of 24 h/ day.

and wind start removing some of the deposits). The paper mentions a study from Sacramento, California that indicates a 20% reduction of the rst year energy savings for all subsequent years (210 years). A detailed recent study on the aging and weathering of cool roong membranes (Akbari et al. [18]), which included 13 white roof material samples that had been exposed for 58 years in eight different locations, found that their solar reectivity had dropped from 0.8 to nearly 0.5. However, in the study it was also found that washing the samples could almost completely restore the original reectivity. See also Levinson et al. [19]. In summary, it is widely recognised that a reective (e.g. white) roof surface in place of a dark one can be of great benet in hot and sunny climates (increase human comfort and/or reduce the cooling load). However, the benets are variable and are not easily accounted for (unlike the thermal resistance of solid material and bulk insulation). 3. Temperature prole in roof space for downward heat ow Downward heat ow from the roof may be caused by (1) A temperature difference between the outside and inside temperature. (2) Solar radiation that is being absorbed on the roof surface. The exposure of the roof surface to solar radiation increases the roof surface temperature and sets up a heat ow in parallel to the one set-up by temperature difference. Fig. 2 illustrates the temperature prole of a roof with and without solar radiation exposure, suggesting the effect of solar radiation absorption usually is the most important factor for the daytime downward heat ow from a roof surface in a hot climate. In tropical regions the daytime heat ow due to the outsideinside air temperature difference is typically in the order of 5 K, while the heat ow due to solar radiation absorption on the roof surface is of the order of 20 K (see also Appendix A.1). It should be noted that the solar absorptance of the roof surface is not the only variable inuencing the roof temperature. The thermal emittance from the roof surface, which determines the radiative heat exchange with the sky, also plays an important role (particularly in low wind conditions). Measurements suggest that roofs in tropical climates cool 16 8C below ambient temperature during nighttime and that the temperature depression may be limited by the formation of dew (Khedari et al. [20]). Typically the sky temperature depression below ambient is 220 K and strongly depends on the level of cloudiness (Cooper et al. [21], Oliveti et al. [22]). For non-metallic surfaces (e.g. painted steel and roof tiles) the thermal radiation emittance to the sky is typically about 0.9 while for bare metal surfaces it varies (for zinc/aluminium and galvanised steel the thermal emittance is approximately 0.3 [23,24]).

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

2227

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

4. An illustrative case study The effect of a change in roof colour is illustrated in Fig. 3 for a residential building in Townsville, Australia. The high set building is of timber frame and clad construction and has a corrugated steel roof (see Fig. 4). Before the reective white paint (Solacoat, with a = 0.20 in new condition) was applied the roof had been painted with an aluminium type paint, but this was in poor condition with severe aking. From Fig. 3 it is seen that the application of the reective white paint (on the afternoon of the second day) has reduced the roof temperatures by about 20 K. More importantly however, the increased solar roof reectance has reduced the interior temperatures relative to the ambient temperature! More specically, for the 2 days before the application of the reective white paint the average ambient and interior air temperatures were 21.8 and 23.9 8C, respectively, while for the 2 days after the application of the white paint the ambient and house interior temperatures were 23.2 and 24.0 8C. Before painting the interior temperature was 2.1 K above ambient and after the interior temperature was only 0.8 K above ambient. This means the room temperature, relative to ambient, had been lowered by 1.3 K. This reduction in internal room temperature, although small in magnitude, delivered a noticeable improvement in human comfort and reduced the desire for air-conditioning. 5. Steady-state heat transfer rate A roof does not experience a steady state while temperatures and radiation are changing because the thermal mass causes a

signicant time lag in the roof temperature response. However, knowledge of the steady-state heat transfer rate is useful in analysing the daily heat gain from a roof. For steady state the solar radiation absorption on the roof surface in Fig. 5(a) changes the heat ow to the outside and the inside environment as shown in Fig. 5(d). Using the thermal network of Fig. 5(b) and its reduced form in Fig. 5(c), one can nd the steady-state heat ux that ows into the building, q*. The result from Stoecker and Jones [25], which is derived in Appendix A.2 for completeness, is q   Q aG U T a T i ho A (1)

where U = (1/ho + 1/hi)1 is the overall heat transfer coefcient between the ambient and inside environment, Ta Ti the ambient to outside temperature difference, a the roof surface solar absorptance, G the solar irradiance and ho the roof to ambient heat transfer coefcient. Note in particular that ho = hc + hr(T Tsky)/(T Ta) includes both the convective heat transfer with the outside air and radiative heat transfer to the skysee Nomenclature and Appendix A.1 for further explanation. Eq. (1) gives the downward heat transfer rate for steady-state conditions. Collecting the term Ta + (aG/ho) in Eq. (1), which is called the solair temperature [25], the downward steady-state heat ux can also be expressed as q U T solair T i (2)

Fig. 3. Temperature measurements of four successive days (47 September 2005) from a house in Townsville, Australia before and after the application of a white reective paint (Solacoat with a = 0.20 in new condition).

Fig. 4. House in Townsville, Australia used for roof and interior temperature measurements (after application of the reective white paint).

2228

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

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

The solair temperature is a hypothetical outdoor air temperature that would cause the same heat ow for a shaded (roof) surface as that experienced by the illuminated one. Eq. (1) shows the heat ow due to temperature difference and radiation can be obtained from the superposition of the heat ows due to temperature difference, Ta Ti, and radiation, G. It should be stressed that Eq. (1) or (2) cannot be used to calculate the instantaneous heat ow for a roof under non-steady state conditions. 6. Denition of thermal resistance (R-value) For the subsequent development, a clear understanding of the denition of the thermal resistance value (R-value) is important. The steady-state heat ow per unit area (heat ux), Q/A, through a wall or roof across which a xed temperature difference, DT, is applied can be expressed as q Q DT U DT A R (3)

for Ta Ti > 0 the effective thermal resistance, R*, is smaller than R. 7. Effect of thermal mass on the downward roof heat transfer Eq. (1) gives the downward steady-state heat transfer rate per unit area from the roof into the building, q*. However, as already discussed heat ow predicted by Eq. (1) usually is not reached because of the thermal mass of the roof and ongoing changes in environmental conditions. The effect of thermal mass on the heat transfer through the building envelope has been extensively studied (e.g. see [26] for an early example and [27] for a review of methods). However, for this study a specic aspect of the effect of thermal mass is of interestthe roof thermal mass inuence on the integrated daily heat gain for a xed R-value. We have investigated this effect for a 1 mm steel (7.85 kg/m2), clay tile (50.0 kg/m2) and 10 cm thick concrete (240 kg/m2) roof during a clear sky day. All three roof constructions were assumed to have an unventilated air space and a suspended plasterboard ceiling of 10 mm thickness (8 kg/m2) below the roof. Roof insulation was added so that all three roofs had the same overall thermal resistance of 2.2 m2 K/W. The results from a numerical simulation of the three roof constructions types are shown in Table 1 for three sky temperatures and plots of diurnal roof temperatures can be found in Appendix A.3. Table 1 suggests that the integrated daily heat gain from the roof into the building is nearly independent of the roof thermal mass. More specically for the same value of Tsky the daily (24 h) net heat gain differs less than 0.7% between the different roof construction types. This is an important result for this study as it suggests that the integrated daily heat ow from an unventilated roof (not the instantaneous heat ow) to good approximation can be calculated from Eq. (1) without considering heat capacitance effects! The result appears physically reasonable as the amount of solar radiation absorbed by the roof does not depend on
Table 1 Daily heat gain for three roof constructions (steel, tile and concrete roof) with suspended plasterboard ceiling and R = 2.2 m2 K/W. The heat gain was calculated for three values of sky temperature. Ambient and inside building air temperatures were assumed xed at 24 8C Sky temperature Daily heat gain (kJ/m2) Steel roof Tsky = Ta Tsky = Ta 10 K Tsky = Ta 20 K 349.8 266.4 192.6 Clay tiled roof 350.0 266.5 192.7 Concrete roof 351.0 267.6 193.8

where U is the overall heat transfer coefcient and R = 1/U is the thermal resistance for steady state. The problem with Eq. (3) is that is does not apply for an illuminated surface. To make Eq. (3) applicable to a sun-illuminated surface one needs to use the solair temperature difference (see also Section 5): q Q DT solair U DT solair A R (4)

where DTsolair = Ta + (aG/ho) Ti. Dividing Eq. (4) by (3) gives: q Q DT solair F q Q DT (5)

where F is the factor of increase in heat ow due to the solar illumination (in case of heat loss from the building, the solar heating may decrease the heat loss). The latter factor allows one to express the roof heat ow, which includes the effect of solar illumination as q

DT solair
R

DT F
R

DT
R=F

DT
R

(6)

where R* = R/F may be interpreted as the effective thermal resistance that a (sun-illuminated) roof or wall offers to inward heat ow (note R* ! R as the solar radiation G ! 0). The concept of effective thermal resistance for cooling is useful as it enables a direct comparison between illuminated roof surfaces of different surface colour (solar absorptance), but experiencing the same outdoor to indoor air temperature difference DT. Note

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

2229

the roofs thermal capacitance and for a xed R-value the absorbed solar radiation must be dissipated at about the same ratio to the ambient and to the inside of the building. Moreover, the result is consistent with the observation of Burch et al. [28] that in a hot climate with continuous positive heat gain even the daily space cooling load can be predicted from steady-state theory.4 8. Daily heat gain from the roof In order to judge the benet of light roof colour on the heat gain from a roof it is required to calculate the heat ow due to both temperature difference and solar radiation over a whole day. As was suggested in the previous section, the daily quantity of heat transferred from the roof into the building can be calculated without consideration of the thermal mass by integrating Eq. (1) over a whole day: fdaily heat gain from roof into buildingg   Z aG U T a T i dt ho day Z    Z a a Dt H T a T i dt G d t U DT U ho day ho day   a U DT G solair Dt U Dt DT ho

9.2. Average outside to inside temperature difference for airconditioned buildings In a hot climate, where active heating is not necessary, most5 benet from thermal insulation and solar reective coatings is gained when the ambient temperature rises above 24 8C. We therefore restrict the period of analysis to those times when the ambient temperature rises above 24 8C and evaluate the temperature difference from the number of cooling degrees hours CDH (base 24 8C): P DT
all hours of the year T a

24

24 365

(9)

where the + sign indicates that only temperature differences above 24 8C are included in the summation. For locations where no may be estimated from the hourly temperature data is available DT cooling degree days (CDD) with Ta = (Ta,max Ta,min)/2. We have evaluated Eq. (9) for several north Australian locations using the US EnergyPlus Weather Data Site [25] and the results are presented in Table 2. The population weighted average form the data of Table 2, using the north Australian conditions as an illustrative example, is  2:2 K DT (10)

(7)

is the average daily temperature difference between where DT outside and inside, H the daily horizontal surface irradiation, R H=Dt the average daily irradiance with Dt H day G dt , and G denoting the day length in seconds (=24 3600 s). The above analysis assumes that the heat gain per unit area is based on the projected horizontal roof area or building oor area. In Eq. (7) it has been assumed that U and a/ho are both constant, in particular that a is (approximately) independent of the incidence angle. 9. Average solar absorptance and environmental conditions For the evaluation of Eq. (7) various average quantities, such as the average building outside to inside temperature difference, i , are required. T a T DT 9.1. Average roof solar absorptance The average roof solar absorptance depends on the type of roof 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 green) and 1/3 light colours (e.g. white and cream). Using known solar absorptance values for the above roof surfaces (see also Table A1), suggest the average roof solar absorptance approximately is

9.3. Average outside to inside temperature difference for passive buildings For non-air-conditioned (passive) buildings the inner temperaT a T i tends ture is oating and the temperature difference DT to be negative, i.e. the inside average air temperature tends to be 2:1 K for the house of higher than the ambient one (e.g. DT Section 4 before application of the reective paint). The inside temperature elevation above ambient can be caused by the solar heating of the building envelope and internal heat sources (e.g. electric appliances). means the For a non-air-conditioned building a negative DT roof insulation loses some of it benets (e.g. it may restrict the desired nighttime cooling) while the benets of solar reection remain unchanged. We make the conservative assumption that the outside to inside temperature difference for non-air-conditioned buildings is about zero, i.e.  0K DT (11)

9.4. Average solar radiation The average radiation can be estimated for many locations on earth from available solar radiation data (e.g. see Appendix G of Ref. [1]). Here horizontal surface averages of daily solar radiation data by Morrison and Litvak [30] are used for north Australian conditions (see Table 3). It is noted that close to the equator the summer to winter solar radiation variation is small (less than 10% for Darwin and 20% for Port Headland and Halls Creek). Assuming that the average annual daily radiation on a  21:5 MJ=m2 for tropical regions in horizontal (roof) surface H Australia6 with latitude less than 23.58 (approx. Zone 1 as dened in BCA, 2005 [2]), we can estimate the average radiation during the
5 Due to internal loads some active cooling of a building may be required before the ambient temperature reaches 24 8C. 6 The chosen value is supported by the contour map shown in the Australian Standard AS/NZS 4234:2008 (Fig. A1).

aave  0:9 0:7 0:5 0:7:

1 3

1 3

1 3

(8)

The average solar absorptance of Eq. (8) is only a crude estimate and does not include bare metal coatings (e.g. galvanised and zinc/ aluminium coatings). Note that a roof solar absorptance of 0.7 has also been specied by the Building Code of Australia 2006 [3] for energy simulations as a nominal value upon which to compare variations.
4 Note the daily the space-cooling load is more likely to be affected by the buildings thermal mass than the daily heat gain.

2230

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

Table 2 Average temperature difference between ambient and 24 8C for several North Australian locations from Eq. (8). Temperature raw data from US EnergyPlus Weather Data Site [29] City/town/place State Est. population 165,059 132,765 111,300 73,091 21,000 13,858 12,697 9,409 3,500 1,289 800 4 544,772 DT P CDH=8760 K

Townsville/ Thuringowa Cairns Darwin Mackay Mt Isa Broome Port Hedland Weipa (Cook Shire) Tennant Creek Halls Creek Wyndham Willis Island Sum/average

Queensland Queensland Northern Territory Queensland Queensland Western Australia Western Australia Queensland Northern Territory Western Australia Western Australia Coral Sea Territory

1.7 1.7 3.7 1.1 3.3 3.6 3.4 2.8 3.7 4.0 5.8 2.2 3.1

air DT 2:2 K from Eq. (10) Using the population averaged DT for an air-conditioned building in northern Australia and the 250 W=m2 and approximate average values a 0:7, G 2 ho = 25 W/m K from Eqs. (8), (12) and (13) in Eq. (14), suggests that in DTsolair = 2.2 + 7 K, i.e. the temperature difference due to solar radiation is more than three times as large as that due to air temperature difference (Eq. (5) gives F  4). Note the result for the temperature differences was calculated for a zero thermal mass roof. However, as the daily heat gain shows little dependence on the roof thermal mass of the roof, the factor of increase in heat transfer due to solar radiation should also approximately apply for a roof with non-zero thermal mass. Similarly, from Eq. (7) the average daily rate of heat ow per unit area (ux) for a roof with zero heat capacitance can be expressed as q   a Q G U DT A ho (15)

Table 3 for selected north Australian Annual average values of daily solar radiation H locations from Morrison and Litvak [30] Location Daily average solar (MJ/m2) radiation H 23.1 22.3 21.7 19.0 21.5

Port Headland, WA (20.388S latitude) Halls Creek, WA (18.238S latitude) Darwin, NT (12.428S latitude) Rockhampton (23.388S latitude) Average

Evaluating Eq. (15) for an air-conditioned building under average north Australian conditions for a total R-value (=1/U) of 2.2 m2 K/W, gives an average daily heat ux of approximately 4.2 W/m2 (=1/2.2(2.2 + 7.0) W/m2) for a (horizontal) roof with 0.7 solar absorptance. On a daily average (24 h) basis this results in a roof heat gain of approximately 361 kJ/m2 (=24 3600 s 4.2 W/m2). Note the heat gain is different from that shown in Table 1 due to the air , which was pretemperature difference driven component DT viously set to zero, and due to the fact that the radiation estimate G is based on average rather than clear sky radiation. 11. Effect of colour on roof downward heat transfer

day by dividing the radiation total by the day length (24 h): G H 21:5 106  250 W=m2 Dt 24 3600 (12)

The solar absorptance of a roof surface has been identied as a substantial inuence on the roof heat ow, and we now ask what is the objective benet a light roof colour relative to an average one? Using Eq. (15) for the average rate of heat gain from a lightcoloured roof, and dividing it by the corresponding equation for an average roof gives: a=ho G U DT q aave =ho G q U ref DT ave (16)

9.5. Heat transfer coefcient from roof surface to ambient Eq. (1) and the subsequent equations for the roof heat gain depend on the heat loss coefcient to ambient, ho, which in turn depends on the wind speed and the sky temperature. The higher the heat loss coefcient, ho, and the smaller the effect of roof colour (solar absorptance). For a roof, the heat transfer coefcient, ho, due to wind and sky radiation typically is 1525 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 25 W=m2 K 10. Daily heat gain from roof due to solar radiation and temperature difference The solair 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: (13)

In Eq. (16) aave is the average roof solar absorptance (here assumed to be 0.7) and U and Uref denote the overall heat transfer coefcient between the outside and inside for the coloured and reference roof,

DT solair DT air DT solar aG


T a T i ho

(14)

Fig. 6. Roof average downward heat transfer as a function of solar absorptance relative to an average roof with aave = 0.7 for DT = 2.2 K (north Queensland airconditioned building) and DT = 0 K (no air-conditioning) with ho = 25.0 W/m2 K.

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

2231

Table 4 Solar absorptance values form measurements of solar reectance for weathered roof/coating materials. Note for opaque materials solar absorptance = (1 solar reectance) and that NSW stands for the Australian state New South Wales Roof/coating Re. white NSW home Re. white NSW factory Re. white NSW resort Standard white NSW resort Re. white painted at sheet Weathered galvanised at sheet Age (years) 8 8 4 Unknown Unknown Unknown Solar absorptance 0.36 0.45 0.33 0.52 0.26 0.66 Comments Roof dirty and paint applied too thin Roof very dirty from cement plant Paint professionally applied Front section of building Painted cover Unpainted cover

respectively. The heat transfer coefcients can be expressed in terms of their corresponding thermal resistance as: 1/U = R and 1/Uref = Rref, where Rref is the reference thermal resistance deemed to satisfy building regulations for minimum insulation. 11.1. Equal insulation but different roof colours One way to assess the effect of roof colour is to compare the reduction in downward heat ow when the bulk insulation for the coloured roof is equal to that of the average one (U = Uref). Using Eq. (16) this yields: a=ho G solair q DT DT DT solair;ref aave =ho G q ave DT

11.3. Effect of roof thermal emittance Eqs. (17) and (18) consider the effect of solar absorptance on the heat gain, but do not explicitly consider the effect of the roof thermal emittance for long wave radiation exchange with the sky. 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 coloured and average surface. However, for the same environmental conditions a bare metal surface, say with zinc/aluminium coating, will have a lower heat loss coefcient, ho, than a painted surface. Using different values of ho for the coloured and average surface in Eqs. (17) and (18), the effect of roof emittance on the heat gain can be considered (see Appendix A.1 on how to calculate the effect of surface emissivity on ho). A lower roof surface thermal emissivity increases the inuence of the solar absorptance value for bare metal coatings, and explains why, for example, a corroded bare metal surface exposed to sunlight can get nearly as hot as a black one. 12. Effect of dust and aging on the roof solar absorptance Two studies dealing with changes in solar roof absorptance over time have already been discussed in Section 2, and it is clear that white coatings do not retain their original reectivity (e.g. because dust and dirt build up [17]). Cool Roof Rating Council website [31] lists 3-year agedperformance ratings for roong products, where samples are weathered for 3 years at test farms in the U.S. The tested 13 reective products with initial absorptance ranging 0.180.21 exhibited an average 0.12 increase in absorptance after 3 years. Table 4 shows reectance measurements of aged reective white paint, standard white, and galvanised steel, which were carried out by the authors of this paper using ASTM E1918-97. It is apparent from Table 4 that the value and deterioration of solar reectance is dependent on the environment. However, Table 4

(17)

Eq. (17) gives ratio of downward heat gain between a coloured and an average coloured roof surface, both having the same insulation below the roof surface. Eq. (17) is readily evaluated for different values of solar absorptance and air temperature differences, and the results are 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 the roof by about 33% for an air-conditioned building and 43% for a non-air-conditioned building in north Australia. Note for an air T a T i the insulation conditioned building with positive DT offers benets even when the roof solar absorptance a approaches (denoting a passive building with zero. However, for a negative DT internal load) the roof insulation can restrict the cooling of a building (for a < 0.22 in Fig. 6). 11.2. Same downward average heat ow Another way of investigating the effect of roof colour is to nd the thickness of insulation below a light-coloured roof that gives the same downward heat ow as it is achieved for a roof with average solar absorptance and standard insulation thickness. The R-value7 for a roof of a particular colour that gives the same downward heat ow as that for the reference or nominal colour can be calculated by assuming q i qi;ref and using Eq. (16) with 1/ U = R and 1/Uref = Rref: a=ho G R DT aave =ho G Rref DT (18)

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 q =q ref , also can be used for the R/Rref ratio.

Based on the outside to inside air temperature difference DT.

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

2232

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

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 250 W=m2 , ho = 25 W/m2 K (bracketed values are for ho = 15 W/m2 K). Note for DT 0 K ho cancels out of Eqs. (22) and (17) from Eqs. (22) and (18), respectively, using G Roof colour Solar absorptance a value range 0.45 > aweathered 0.6 > aweathered  0.45 0.8 > aweathered  0.6 aweathered  0.8 Nominal a value 0.35 0.55 0.7 0.85 2:2 K Rcolour/Rref DT 0.38 (0.42) 0.16 (0.18) 0.00 0.16 (0.18) 0K Rcolour/Rref DT 0.50 0.21 0.00 0.21
q i =qi;ref DT 2:2 K q i =qi;ref DT 0 K

Reective white (bright white) Light colour (e.g. greyish white, light yellow) Medium colour (e.g. green, red) Dark colour (e.g. grey, black)

0.62 (0.58) 0.84 (0.82) 1.00 1.16 (1.18)

0.50 0.79 1.00 1.21

also suggests that exposed reective paint or white steel roong maintains a signicantly higher reectance than galvanised steel and standard white. The paper by Griggs et al. [10], more generally states that: weathering tends to reduce the reectance of a light roof and increases the reectance of a dark roof. If one assumes that the weathering tends to change the solar reectance towards a particular reectance value,8 the data from Table 4 may be used to estimate the weathered solar absorptance value. The reective white paint used for this study has a solar absorptance of approximately 0.20 (Solacoat) when new and when it is applied with the recommended coverage. The average solar absorptance of the reective white weathered 8 years according to Table 4 is about 0.35 (=(0.36 + 0.45 + 0.33 + 0.26)/4) and therefore the increase in absorptance is 0.15 (=0.35 0.20). If it is now assumed that the absorptance increase (loss in reectance) is the proportional to the difference anew adust, the increase (or decrease) in absorptance may be estimated as

than predicted by Eq. (19) as they assume a visibly duller appearance due to weathering. 13. Equivalent R-value due to roof colour With an estimate for the reduction in roof reectance due to weathering in place, the effective R-value for a roof colour can now be determined. The difference between the reference Rref and R in Eq. (18) may be credited in a hot climate as a heat transfer resistance that is provided by the particular roof colour, i.e. Rcolour Rref R (21)

Da 0:15

adust anew adust 0:20

(19)

so that the weathered value for the solar absorptance may be expressed as

aweathered anew Da

(20)

It is instructive to calculate the Rcolour credit for cooling for a reective white roof surface that has a (weathered) solar absorptance of a = 0.35. Using the previously determined values for northern Australia (Rref = 2.2 m2 K/W, ho = 25 W/m2 K, 250 W=m2 ) in Eq. (18) gives: 0:0 K, aref = 0.7, and G DT R = 1.1 m2 K/W, i.e. a house roof with a solar absorptance of 0.35 only requires a R-value of 1.1 m2 K/W to achieve the same insulation as a nominal roof with 0.7 solar absorptance and Rvalue of 2.2 m2 K/W. Hence, for this particular case one may assign an effective R-value for cooling of 1.1 m2 K/W (=2.2 1.1 m2 K/W) to the reective roof paint. Dividing Eq. (21) by Rref allows the effect of roof colour to be expressed in more general terms: =ho aweathered G Rcolour R DT 1 1 =ho aave G Rref Rref DT 1 solair DT solair;ref DT (22)

On completion of this study it was found that the California Energy Commission had already proposed a very similar expression (reported by Prado and Ferreira [32]), in which a corresponding solar absorptance of adust = 0.8 had been assumed. Adopting this value here we note that our expression based on independent data closely agrees with the one by the California Energy Commission. The suggested relationship between the absorptance value for weathered and new material is depicted in Fig. 7. Using Eqs. (19) and (20) for the ordinary white steel roof (assumed to have a solar absorptance of 0.39 for the new material (see Appendix A.1)) with adust = 0.8, gives a solar absorptance increase through weathering of 0.11 and a solar absorptance for the weathered material of 0.50. The latter value compares to the solar absorptance inferred from reectance measurement for ordinary white steel roong in Table 4 of 0.52. The relationship for the increase in solar absorptance suggested in Eq. (19) is only a rough estimate, however, it is noted that the 19% (=0.15/0.80)9 decrease in solar reectance for the reective white paint is in line with the 1030% decrease in solar reectance reported by other studies for reective white roof materials (see Section 2). One would expect the change in reectance to depend somewhat on the local air-pollution and climate. For zinc/aluminium and galvanised bare metal surfaces, the changes in solar absorptance are likely to be signicantly larger
8 A roof with having a solar reectance close to that of the dust and dirt deposits should not change its absorptance signicantly (neglecting surface oxidisation). 9 Note solar reectance = 1 a for an opaque surface.

Rcolour/Rref can be interpreted as the fraction of thermal resistance that is provided by the roof colour when the average daily heat gain is the same as for a roof of average colour, i.e. for q q ref . While Eq. (22) provides a quantitative measure for the effect of roof colour, the comparison between the various roof colours is facilitated with some discrete classes. Table 5 shows results obtained from Eqs. (22) and (17) for four nominal roof colours and two outside-to-inside average air temperature differences. The results obtained for Rcolour/Rref from Eq. (22) relate to the case of equal daily downward heat ow, while the results obtained for q i =qi;ref from Eq. (17) relate to the change in heat gain for roofs with the same insulation below the roof surface. Table 5, for example, shows that a non-air-conditioned building with a weathered white reective roof surface (with nominal a = 0.35) can substitute 50% of the thermal resistance required for an average roof with a = 0.7. Alternatively, if the roof has the same thermal resistance as the average one, then the heat gain from a weathered white reective roof surface is only about 57% of that from an average roof. It is noted that the percentages in Table 5 are independent of the reference value Rref that may be prescribed.

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

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 q i =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)/ 2 (T Ta), where hr s eT 2 Tsky T T sky , gives ho,bare = 20 W/ 2 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. 14. Discussion The results in Table 5 demonstrate that the roof colour has a signicant effect on the heat gain, particularly for non-airconditioned (or passive) buildings. The assumption that for passive buildings the inside temperature equals the outside should yield a conservative estimate for the benet of a solar reective roof coating. The reectance and thermal emittance measurements listed in Table A1 suggest that non-white roof coatings that reect invisible parts of solar radiation still offer appreciable benets over their equal coloured counterpart (see also [12,35]). In hot climates roof insulation may hinder the desired nighttime cooling and a light roof colouring in combination with only a reective roof air space may be preferable for non-air-conditioned buildings. Gaining greater energy efciency through the choice of appropriate roof colours is simple and cost effective. Reective (or light-coloured) paints may be used for the retrot or conversion of roof surfaces. 15. Conclusions The above analysis has shown that it is possible to quantity the effect of colour on the roof heat gain. Consequently, we are able to assign an R-value credit to the reectivity of a roof surface. However, the analysis at this stage is restricted to locations where there are no heating requirements (cooling only). The calculations suggest that in hot climates a signicant reduction in downward heat ow can be achieved by using a light or reective roof colour instead of a dark one. The reduction in downward roof heat ow means a reduction in air-conditioning load or an increase in human comfort. Measurements of the reduction in reectance of light roof colours due to aging and weathering indicate that much of the reduction occurs in the rst year and that long-term average values for the roof reectance can be estimated. A classication of roof colours has been made and the relative benets are calculated for the case of downward roof heat transfer. The concepts presented in this paper have the potential to be extended for buildings that require both cooling and heating. Acknowledgements The support by Coolshield International Pty Ltd for the completion of this study and the proof reading of the manuscript by Mr. Graham Aldridge is gratefully acknowledged. Particular thanks goes to Nick Holcz from the West Australian Climate Services Centre for providing the long-term air temperature

Fig. A1. Schematic diagram of a surface exposed to wind, sky and solar radiation.

data. Last but not least we would like to acknowledge the great encouragement and support for this study by Dr. Jonathan Harris.

Appendix A A.1. Steady-state temperature of a surface The temperature of a roof surface depends on the environmental conditions (solar radiation, ambient temperature, wind velocity and sky temperature), the optical properties of the roof surface (solar absorptance and thermal emittance) and the underside insulation. Fig. A1 shows a schematic diagram of a surface having incident solar radiation and heat loss to the environment. For a surface with perfect insulation10 at the underside and steady state the absorbed solar radiation must equal the heat loss due to convection with the air and thermal radiation heat exchange with the sky:

aG hc T T a hr T T sky
ho T T a

(A.1)

See Nomenclature for the meaning of symbols. Solving Eq. (A.1) for the surface temperature gives: T

aG hc T a hr T sky
hc hr

or

T Ta

aG
ho

(A.2)

Table A1 shows the results of evaluating Eqs. (A.1) and (A.2) for various surfaces for an ambient temperature of 25 8C and solar radiation of 1000 W/m2. A.2. Derivation of the steady-state roof heat ow equation Consider the thermal network shown in Fig. 5(c). For steady state, the sum of the energy owing in and out of the roof surface due to temperature differences and solar radiation absorption must equal zero. Moreover, the heat ow to the inside must equal the net heat ow from between roof and ambient and the solar radiation absorbed on the roof surface. Hence, using the previously dened symbols, we can formulate the following energy balances for the roof surface and heat ows: ho T a T hi T i T aG 0 q hi T T i ho T a T aG (A.3) (A.4)

10 Note the back heat losses are normally not zero, but an order of magnitude smaller than the front heat losses to the environment. Therefore, the temperature in Eq. (A.2) approximately gives the temperature of a surface that is well insulated at the back.

2234

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235

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 White highly reective paint (new Solacoat) Standard White steel roong (new) Green steel roong (new) Grey steel roong (new) Black (oil paint) Conventional red (new) Coolshield Dynasty Natural red (near-infrared reective, new) Zinc/aluminium coated steel roong (new) Galvanised surface (weathered) Light red tilted roof

a
0.20y 0.39 0.75 0.88 0.90** 0.70+ 0.50+ 0.38 0.80*** 0.63

e
0.88y 0.9* 0.9* 0.9* 0.94** 0.9** 0.89+ 0.3* 0.28*** 0.9*

T (8C) 32 40 55 60 61 53 45 44 66 50

hr (W/m2 K) 5.2 5.6 6.0 6.2 6.4 5.9 5.6 1.9 2.0 5.8

ho (W/m2 K) 30 26 25 25 25 25 25 20 19 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 adding yields:

the sky temperature being below the ambient one, i.e. the roof loses heat during nighttime to the cold sky.

References
[1] J. Dufe, 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/HeatIsland/CoolRoofs/, last accessed: 17 September, 2006. [5] TRNSYS 16 (Transient system simulation program), University of Wisconsin, 2005, Madison, USA, http://sel.me.wisc.edu/trnsys, last accessed: 2006. [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) 915937. [7] D. Parker, J. Sherwin, J. Sonne, S. Barkaszi, FSEC-CR-904-96, Demonstration of Cooling Savings of Light Colored Roof Surfacing in Florida Commercial Buildings: Our Saviors 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/. [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 modication on cooling loads on scale model residences in Tucson, Arizona, Energy and Buildings 25 (1997) 127137. [10] E. Griggs, T. Sharp, J. MacDonald, ORNL-6527, Guide for estimating differences in building heating and cooling energy due to changes in solar reectance of a lowsloped roof, Oak Ridge National Laboratory-Energy Division, Oak Ridge, USA, 1989, download: http://eber.ed.ornl.gov/commercialproducts/Reroof.htm. [11] A. Granja, L. Labaki, Inuence of external surface colour on the periodic heat ow through a at solid roof with variable thermal resistance, International Journal of Energy Research 27 (2003) 771779. [12] R. Levinson, H. Akbari, J.C. Reilly, Cooler tile-roofed buildings with near-infraredreective non-white coatings, Building and Environment 38 (2007) 181188. [13] AS 4859.1:2002, Materials for the Thermal Insulation of BuildingsGeneral 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 Efcient Design of Low-Rise Residential Buildings, Section 5.5, ASHRAE, Atlanta, USA. [16] H. Akbari, S. Konopacki, D. Parker, Updates on Revision to ASHRAE Standard 90.2: Including Roof Reectivity 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, Pacic 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 Roong Membranes, Department of Energys (DOE) Information Bridge, Oak Ridge, USA, 2006, download: www.osti.gov/bridge/servlets/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 reectance and solar heat gain of a light-colored roong membrane, Atmospheric Environment 39 (2005) 78077824. [20] J. Khedari, J. Waewsak, S. Thepa, J. Hirunlabh, Field investigation of night radiation cooling under tropical climate, Renewable Energy 20 (2) (2000) 183193. [21] P.I. Cooper, E.A. Christie, R.V. Dunkle, A method of measuring sky temperature, Solar Energy 26 (1981) 153159.

where the equation below the line is the sum of the two above the line and 1/U = 1/ho + 1/hi is the overall heat transfer resistance between the outside and inside. Now solving the above equation for q* gives Eq. (1).

A.3. Roof heat ow for various roof construction types Fig. A2 shows that the reduction in daytime peak heat ow due to an increase in thermal mass is accompanied by an approximately proportionate increase in the duration of downward heat ow. This suggests that the thermal mass of the roof does not signicantly affect the integrated daily heat gain through the roof (see also Table 1). Note that the slightly negative heat ow during nighttime is the result of

Fig. A2. Downward heat ow q i for three roof constructions (1 mm steel, clay tiles and 100 mm concrete) into the underlying room during a clear day (R = 2.2 m2 K/W for all roof constructions and hi = 1/(2.2 0.04)).

H. Suehrcke et al. / Energy and Buildings 40 (2008) 22242235 [22] G. Oliveti, N. Arcuri, S. Ruffolo, Experimental investigation on thermal radiation exchange of horizontal outdoor surfaces, Building and Environment 38 (2003) 8389, Cities: Theoretical Applied Climatology 85 (2006) 185201. [23] F.P. Incropera, D.P. DeWitt, Table A.12, in Fundamentals of Heat and Mass Transfer, Wiley, New York, 2002. [24] R. Siegel, J. Howell, Appendix D, in: Thermal Radiation Heat Transfer, McGraw Hill, New York, 1981. [25] W. Stoecker, J. Jones, Refrigeration and Air Conditioning, Sections 4.12 and 4.13, McGraw Hill, New York, 1982. [26] E. Danter, Periodic heat ow characteristics of simple walls and roofs, Journal of the Institution of Heating and Ventilating Engineers July (1960) 136146. [27] C.A. Balaras, The role of thermal mass on the cooling load of buildings. An overview of computational methods, Energy and Buildings 24 (1996) 110. [28] D.M. Burch, S.A. Malcolm, K.L. Davis, The effects of wall mass on summer space cooling of six test buildings, ASHRAE Transactions 90 (1984) 521. [29] US EnergyPlus Weather Data Site, http://www.eere.energy.gov/buildings/energyplus/cfm/weather_data.cfm, last accessed: June 2007.

2235

[30] G. Morrison, A. Litvak, Report No. 1/1999, Table 7, 1999, Condensed Solar Radiation Data Base for Australia, Solar Thermal Energy Laboratory, University of New South Wales, Sydney, Australia, 1999, download: http://solar1.mech.unsw.edu.au/glm/glm-papers.htm. [31] Cool Roof Rating Council website, http://www.coolroofs.org/, last accessed: April 2008. [32] R. Prado, Ferreira, Measurement of albedo and analysis of its inuence the surface temperature of building roof materials, Energy and Buildings 37 (2005) 295300. [33] H. Suehrcke, Solar reectance of white Solacoat insulating paint and other common roong materials, Consultancy Report for COOLshield International, Queensland, 2006. [34] T. Tuquabo, J. Bell, Consultancy Report for COOLshield International, Queensland University of Technology, 2006. [35] R. Levinson, P. Berdahl, H. Akbari, Solar spectral optical properties of pigments. Part II. Survey of common colorants, Solar Energy Materials & Solar Cells 89 (4) (2005) 351389.

Das könnte Ihnen auch gefallen