Вы находитесь на странице: 1из 46

Accepted Manuscript

A life-cycle cost analysis for an optimum combination of cool coating and thermal
insulation of residential building roofs in Tunisia

Khawla Saafi, Naouel Daouas

PII: S0360-5442(18)30600-5

DOI: 10.1016/j.energy.2018.04.010

Reference: EGY 12649

To appear in: Energy

Received Date: 22 November 2017

Revised Date: 07 February 2018

Accepted Date: 03 April 2018

Please cite this article as: Khawla Saafi, Naouel Daouas, A life-cycle cost analysis for an optimum
combination of cool coating and thermal insulation of residential building roofs in Tunisia, Energy
(2018), doi: 10.1016/j.energy.2018.04.010

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

A life-cycle cost analysis for an optimum combination of

cool coating and thermal insulation of residential building

roofs in Tunisia

Khawla Saafi a,*, Naouel Daouas a


a Ecole Nationale d’Ingénieurs de Monastir, Unité de Métrologie et des Systèmes

Energétiques, Département de Génie Energétique, Université de Monastir,

Rue Ibn El Jazzar, 5019 Monastir, Tunisia

* Corresponding author: Tel.: (216) 73 500 244, Fax: (216) 73 500 514

E-mail addresses: k92saafi@gmail.com (Khawla Saafi)

naou.daouas@gnet.tn (Naouel Daouas)

Abstract

The interaction between the roof thermal insulation and the cool roof effects is assessed in

order to determine an optimum combination of the two measures. Dynamic simulations and

estimation of the annual energy requirements are performed using EnergyPlus. Two roof

structures, three insulation materials and three reflectivity scenarios are considered while

taking into account the ageing of the cool material. An energy-based optimization shows that

the summertime benefits induced by the rise of reflectivity outweigh the winter penalties, and

that the optimum value of the roof reflectivity is the highest possible value. Moderate roof

insulation levels with cool roof surfaces of as high as possible reflectivity values are

recommended in the Tunisian climate. A 20-year life-cycle cost analysis proves the cost-

effectiveness of aged and restored cool roof scenarios for uninsulated roofs with a net saving

1
ACCEPTED MANUSCRIPT

up to 44.53 TND/m² and a payback period of 3.4 years. In terms of maximum net savings and

considering the environmental benefits of cool roofs, we recommend an optimum

combination of the restored concrete-based roof and a 5.4 cm-rockwool thickness. A

sensitivity analysis shows that the uncertainties in the economic parameters and changes in

the set-point temperatures have a noticeable effect on the optimal parameters.

Keywords: Cool roof; EnergyPlus; solar reflectivity; thermal insulation; life-cycle cost;

optimization.

Nomenclature

As annual energy savings (TND/m²)

b payback period (years)

Cel cost of electricity (TND/kWh)

Cenr cost of energy consumption (TND/m²)

Cg cost of natural gas (TND/m3)

Ci cost of insulation material in (TND/m²)

Cins cost of insulation material in (TND/m3)

Cr present value of the reflective material cost (TND/m2)

Cref initial cost of reflective material (TND/m2)

Ct total cost (TND/m2)

COP coefficient of performance of air-conditioning system

d discount rate (%)

EPS expanded polystyrene

h heat transfer coefficient (W/m².K)

Hu heating value of natural gas (J/m3)

i inflation rate (%)

k thermal conductivity (W/m.K)

2
ACCEPTED MANUSCRIPT

L layer thickness (m)

LCCA life-cycle cost analysis

Lins insulation thickness (m)

Lopt optimum insulation thickness (m)

n lifetime of building (years)

N number of layers of composite roof

NLCS net life-cycle savings (TND/m²)

PWF present worth factor

q heat flux density (W/m²)

Qc yearly cooling transmission loads (J/m²)

Qh yearly heating transmission loads (J/m²)

Qt total yearly transmission loads (Qt=Qh+Qc) (J/m²)

R uninsulated roof

RI insulated roof

RW rock wool

t time (s)

T temperature (K)

Tc cooling set-point temperature (K)

Th heating set-point temperature (K)

TND Tunisian Dinar (1 TND = 0.42 US$)

U overall heat transfer coefficient (W/m².K)

Uwt overall heat transfer coefficient without insulation layer (W/m².K)

x coordinate direction normal to roof (m)

XPS extruded polystyrene

Greek Symbols

 thermal diffusivity (m²/s)

3
ACCEPTED MANUSCRIPT

 emissivity

s efficiency of the heating system

 solar absorptivity

 solar reflectivity

 constant of Stefan Boltzmann (W/m².K4)

Subscripts and Superscripts

c convective

i inside

j layer number

o outside

1. Introduction

As a result of population and economic growth, global energy demand increased

exponentially in recent decades. In particular, energy consumption in the Mediterranean is

four times higher than in Europe and the OME (Observatoire Méditerranéen de l’Energie) is

forecasting a further 2.6% increase by 2025 [1]. The building sector is the major consumer of

energy in the world, accounting for 40% of the energy consumption and greenhouse gas

emissions. About 60% of energy consumption in the building sector are due to cooling,

heating and ventilation systems. This demand has increased annually by 1.8% over the last

forty years [2]. At the national level, the building sector in Tunisia is the largest consumer of

energy with 37% of the final energy consumption [3]. This sector represents a major

challenge for energy saving issue. In the Tunisian climate, both heating in winter and cooling

in summer are required to reach comfort levels. So a special attention should be paid to

improving the thermal quality of the building envelope. The thermal insulation of the

4
ACCEPTED MANUSCRIPT

envelope is one of the most effective measures in the Tunisian climate context, which can

generate up to 30% energy savings [4].

The optimization of the insulation thickness for the external building envelope has been

performed by numerous investigators where the optimum thickness is the value that provides

the minimum total cost, including the insulation cost and the energy consumption cost, over

the lifetime of the building. In most of these studies, the estimation of the building heating

and cooling loads is based on the Degree-Days concept [5- 8]. This method provides a simple

estimation of the annual loads, but does not consider the impact of the outside surface

radiative properties (solar reflectivity and infrared emissivity) on the energy balance of a

building component. Other studies considered dynamic transient models able to specify,

among others, the radiative properties of the envelope external surface. Al-Sanea et al. [9-10]

used a numerical method for the determination of the optimum insulation thickness in

building walls under a hot-dry climate. All their study has been conducted with values of the

solar absorptivity and the emissivity of the outside surface, fixed respectively to 0.4 and 0.9.

Ozel [11] also used a numerical method to analyze the effect of the orientation of the exterior

walls on optimum insulation thickness. In his study, the solar absorptivity of the outside wall

surface was selected to be equal to 0.9, whereas the effect of surface emissivity indirectly

accounted for in a combined convection and radiation heat transfer coefficient. The same

author investigated the influence of the exterior surface solar absorptivity on optimum

insulation thickness for a south facing wall subject to the climatic conditions of Elazig,

Turkey [12]. According to his economic analysis results, the solar absorptivity has a more

significant effect on the energy savings than on the optimum insulation thickness and the

payback period. Daouas et al. [13-14] are the first who have developed their own analytical

dynamic model, based on the Complex Finite Fourier Transform method, to rigorously

estimate the optimum insulation thickness of exterior walls in Tunisian buildings. The solar

5
ACCEPTED MANUSCRIPT

absorptivity of the outside surface was included in the formulation and taken to be equal to

0.4, whereas a combined convection and radiation heat transfer coefficient was adopted. The

same analytical method has been extended and validated by Daouas [15] to handle the

nonlinear boundary condition on the outside surface of a roof, where incident solar radiation,

convection exchange with ambient air and longwave radiation exchange with the sky are

considered separately. This proposed model allowed to include both the solar absorptivity and

the longwave emissivity of the outside surface and to highlight their important impact on the

transmission loads and optimum insulation thickness.

The energy efficiency of roofs raised the interest of many authors who focused their studies

on the improvement and the optimization of its thermal performance. Indeed, this envelope

component receives the most solar radiation in summer, contributes significantly to heat

losses in winter and so, plays an important role in the building energy conservation. Two

widely used types of techniques concern the modification of the roof thermal properties (roof

insulation) and the roof surface treatments (cool roofs, radiant barriers). Yu et al. [16]

determined the optimum insulation thickness for a residential roof with different surface

colors, based on a life-cycle cost analysis and solar-air degree-hours in four typical cities of

China. Results showed that the impact of the roof surface color on the optimum insulation

thickness differs from city to city. Gentle et al. [17] analyzed the combined effect of the

thermal resistance, the solar albedo and the thermal emittance of a roof. A cost benefit

analysis, including only the costs of material and thermal loads, demonstrates the importance

of a high roof albedo, while the sensitivity to the thermal resistance and the thermal emittance

drops away as the albedo rises. Ramamurthy et al. [18] provided guidance on the optimal

combination of roof insulation thickness and roof surface albedo in terms of minimum yearly

heating and cooling costs incurred. They suggested that, for new constructions, a high albedo

exceeding the value of 0.7 associated with a 18 cm-polylso foam insulation thickness is a

6
ACCEPTED MANUSCRIPT

practical solution for significantly reducing energy losses/gains. When including the

insulation cost, maximum savings were obtained for a 4-inch insulation with a too high

recovery period reaching 13 years. The interplay between three energy measures, the roof

insulation, the cool roof and the roof radiant barriers, has been studied by Arumugam et al.

[19] using energy simulations with EnergyPlus. An economic analysis showed that the need

for insulation reduces by using cool roofs and radiant barriers in all the five climatic zones of

India considered in this study.

Other authors limited their study to the optimization of the cool roof properties. Piselli et al.

[20] carried out a parametric analysis in order to estimate the impact of the building end-use,

the air-conditioning system, the internal heat gains and the roof insulation level on the

optimum roof solar reflectance. Results of simulation-based optimizations showed that the

optimum roof solar reflectance, to reduce the annual building energy consumption, is mainly

affected by the climate context. Shi and Zhang [21] proposed a passive solution to reduce the

annual energy loads by changing the solar reflectance and the longwave emissivity of the

exterior surface of the building envelope in 35 different climate conditions around the world.

They found that tropical climates are the most significantly affected by the optical parameters

in term of building energy-savings. Their greatest energy-saving potential is obtained by

considering a high solar reflectance and a high longwave emissivity for the exterior surface of

the building envelope.

All the above-cited studies are limited to the energy-effectiveness of cool roofs. Only a

limited number of investigations dealt with the cost-effectiveness of cool roofs based on the

cool material cost. Hernández-Pérez et al. [22] analyzed the cost-effectiveness of reflective

white and colored roofs compared to conventional gray roofs in six selected cities in Mexico.

Based on the costs of electricity and reflective materials, a 10-year life-cycle cost analysis

showed that, in the absence of insulation, reflective white and colored roofs are more cost-

7
ACCEPTED MANUSCRIPT

effective than gray roofs for all locations. However, for insulated roofs, reflective white

materials are cost-effective only in warmer locations with long payback periods reaching nine

years. Sproul et al. [23], in turn, presented an economic comparison of white, green, and black

flat roofs in the United States. A 50-year life-cycle cost analysis, including installation,

replacement and maintenance costs, showed that white is the best roof color choice. However,

the authors recommended that the choice of white vs. extensive green roof should be based on

the environmental and societal concerns of the decision-maker. Moreover, Zhang et al. [24]

demonstrated the cost-effectiveness of using a cool paint in both unventilated and ventilated

concrete roofs under the tropical climate in Singapore. The life-cycle analysis they conducted

is limited to the initial installation cost of the cool paint. A research, including both on-site

data and a building energy simulation model was presented by Jo et al. [25] in order to

quantify the annual electricity savings of an operational commercial building when replacing

the existing dark aggregate roofing material with a reflective cool roof surfacing system. A

20-year cost benefit analysis, including the additional costs of cool roof retrofit and

maintenance, showed that a 100% cool roof installation resulted in savings of approximately

$22,000 per year in energy costs and a consequent 9-year payback period for the added cost.

Yuan et al. [26] tried to find an optimal combination of the surface reflectivity and the

insulation thickness of exterior walls based on a 10-year life-cycle cost analysis. Besides the

energy cost and the insulation material cost, four steps of the reflective materials prices were

selected depending on their albedo levels. However, the long-term performance of the

reflective material has not been taken into account. The results show a tendency of the

thermal insulation to be thinner and the reflectivity to be higher from high-latitude to low-

latitude regions of Japan.

In the present paper, a more extensive study is proposed in order to determine an optimal

combination between two important measures for energy conservation in residential buildings

8
ACCEPTED MANUSCRIPT

in Tunisia, i.e. the roof thermal insulation and the cool roof application. The main objective is

to conduct a cost optimization analysis where realistic conditions are simulated by taking into

account the long-term performance of the cool material and the impact of ageing on its solar

reflectivity. Two typical roof structures and three types of insulation materials, as commonly

used in Tunisian constructions, are considered. Annual loads, including heating and cooling

requirements and which are the main inputs of the optimization, are estimated using

EnergyPlus software. An energy analysis, based only on minimum energy consumption, is

firstly conducted to determine an optimum value of the roof solar reflectivity, for different

roof insulation levels. After that, a life-cycle cost analysis (LCCA) is performed to evaluate

the cost-effectiveness of cool roof coatings, on the one hand, and to identify the optimum

combination between the surface reflectivity of the roof and its insulation thickness, on the

other hand. Three roof reflectivity scenarios are considered taking into account the

degradation by ageing and weathering of the cool coating. An optimum insulation thickness is

determined for all the roof scenarios, based on the minimum total life-cycle cost including the

energy cost and all other added costs. The resulting net energy savings and payback periods

are estimated and compared. A sensitivity analysis is also investigated to assess the impact of

economic parameters and set-point temperature levels on the obtained results.

2. Methodology

2.1. Modeling of heat transfer in a multilayer roof

In the present study, we consider a multilayer flat roof composed of N parallel layers of

different materials and thicknesses, as shown in Fig. 1. It is assumed that the roof is made of

isotropic and homogeneous layers, whose thermal properties are independent of temperature.

The transient heat transfer through the composite roof is considered one-dimensional since the

thickness of the roof is generally small compared to its other dimensions. Based on these

assumptions and in the absence of heat generation, the heat transfer through each layer of the

9
ACCEPTED MANUSCRIPT

roof is governed by the following heat conduction equation:

 2T j 1 T j
 for 0 < xj < Lj and j = 1, 2,…, N (1)
x 2
j  j t

The initial temperature distribution is taken uniform as:

T j ( x j ,t  0 )  T0 for j = 1, 2,…, N (2)

The roof inside surface (the ceiling) is exposed to the indoor air which is kept at a fixed

design temperature Ti. This surface is subject to combined convection and radiation heat

transfer. The roof outside surface is exposed to a solar radiation flux qs(t) and to ambient air,

with a temperature variation Ta(t). This surface exchanges heat convection with ambient air

and longwave radiation with the sky.

The boundary conditions on the inside and the outside surfaces of the roof are written,

respectively, as follows:

T1 
k1   hi [Ti  T1 (x1  0 ,t )] (3)
 x1  x 0
1

TN 
 kN   hc ,o [ TN ( xN  LN ,t )  Ta ( t )]   qs ( t )  qr ,o ( t ) (4)
 xN  xN  LN

hi is the inside-surface combined heat transfer coefficient, hc,o is the outside-surface

convection heat transfer coefficient and qr,o(t) is the longwave radiation exchange flux at the

outside roof surface, given by:

qr ,o (t )   o [TN4 ( LN ,t )  Tsky
4
(t )] (5)

where Tsky is the temperature of the sky generally lower than the outside ambient air

temperature depending on the time of day, clouds and moisture content in the air.

Analysis is based on the loads due to heat transmission from the roof to the indoor space. The

instantaneous transmission load is expressed as follows:

qi (t )  hi [T1 (x1  0 ,t )  Ti ] (6)

10
ACCEPTED MANUSCRIPT

A daily total load is obtained by integrating the above time-variation of the inside-surface heat

flux over a 24 h period.

2.2. Dynamic energy simulation

The thermal behavior of residential roofs is evaluated by means of EnergyPlus software,

which is one of the most known dynamic energy simulation tools. The use of this software in

the energy simulation of buildings by various researchers has proved a good efficiency. It

analyzes both energy consumption (heating, cooling, ventilation, lighting, etc.) and water use

in buildings [27]. EnergyPlus calculates thermal loads based on heat balances on outdoor and

indoor surfaces and transient heat conduction through the building, where both the

Conduction Transfer Function (CTF) method and the finite difference method can be used.

Simulations are conducted based on a weather data file available in the TMY (Typical

Meteorological Year) format and including direct normal, global horizontal and diffuse solar

radiation, dry bulb temperature, relative humidity and wind velocity [28]. Based on the

available solar irradiance components, EnergyPlus uses the Perez model [29] to predict solar

irradiance on tilted surfaces. Radiative sky temperature (see Eq.(5)) is internally calculated by

EnergyPlus using the horizontal infrared radiation from the sky, the cloudiness factors and the

current temperature [30]. In our case, the sky temperature (in °C) is obtained as follows [27]:

0.25
q 
Tsky   IR   273.15 (7)
 

where qIR (in W/m²) is the horizontal infrared radiation (available in the weather file).

If qIR is missing from the weather file, EnergyPlus calculates it from the opaque sky cover

field and the current temperature.

In the present study, we consider a virtual test zone inspired from [17,31-32], whose envelope

is adiabatic except for the roof component being under investigation. The thermal

performance of the roof and the optimization of its conception depend on the operating mode

11
ACCEPTED MANUSCRIPT

of the air-conditioning system. In the literature, different modes have been considered,

including continuous air-conditioning [9-16,18,21-22,24-26] and intermittent air-conditioning

[19] with a fixed indoor set-point temperature, and no air-conditioning with a fluctuating

indoor temperature [33]. In order to analyze the impact of the roof measures on the energy

loads, we consider a continuous air-conditioning operation mode during the heating and

cooling seasons by adopting an "Ideal Loads" HVAC configuration. We assume that the

ambient interior temperature is kept rigorously constant and fix a different indoor set-point

temperature for each season. This temperature is controlled at 24°C in summer and 20°C in

winter. The test zone energy model is developed using EnergyPlus input file. As the

optimization concerns the roof component of the envelope, we will consider only the roof

transmission part of the air-conditioning load, calculated by the above energy model. Among

the two methods proposed by EnergyPlus, the CTF algorithm is adopted to calculate annual

heating and cooling loads needed for the optimization. The CTF method is one of the most

accurate and widely used methods in the building energy calculations [15]. It requires less

computational effort compared to the finite difference method, as it determines the series of

CTF coefficients in a one-time calculation. However, the CTF method is limited to constant

thermal properties. Its accuracy and its computation time are highly affected by the choice of

the simulation time step [30]. Moreover, the stability of the CTF method is mainly influenced

by the number of nodes used to discretize the construction layers. In spite of the limitations of

the CTF approach, EnergyPlus provides stable results for most building constructions [31].

This is due to its local Fourier stability criterion for the choice of the nodal spacing, which is

associated with another procedure for adjusting the number of nodes. Indeed, with an

adequate number of nodes, the CTF method used in EnergyPlus allowed to obtain results in

good agreement with analytical solutions [27]. The CTF method applied to the energy model

considered in the present study has been also validated in a previous work [15] with the

12
ACCEPTED MANUSCRIPT

analytical solution of the Complex Finite Fourier Transform method.

2.3. Roof structures and climatic data

We consider the most widely used variants of roofs in the residential building construction

sector in Tunisia. These roof structures are mentioned in the reference guide of the Tunisian

agency for energy management [34] and were also adopted by Znouda et al. [35] as the most

representative of building construction in Tunisia. As shown in Fig. 2, R1 and R2 are

uninsulated roofs made of common construction materials. R1 is essentially made of

reinforced concrete and R2 contains hollow terracotta units introduced to decrease the roof

total weight and reduce heat transmission [36]. The structures RI1 and RI2 (Fig. 2) represent

insulated concrete-based and terracotta-based roofs, respectively. We analyze three different

insulation materials: expanded polystyrene (EPS), extruded polystyrene (XPS) and rockwool

(RW), which are chosen among the most available in the local markets and the most

commonly used in the Tunisian constructions. In both structures RI1 and RI2, the insulation

layer is placed in the middle whose thickness will be varied during the study. Thermal

properties of building materials, obtained mostly from [37], are provided in Table 1.

Climatic data needed for the simulations are taken from the EnergyPlus weather file of Tunis,

the capital of Tunisia, (latitude: 36°50’N, longitude: 10°14’E, altitude : 4m). The Tunisian

climate is Mediterranean with a high level of solar radiation. As shown in Fig. 3, the

horizontal mean daily solar radiation reaches a maximum of 26.94 MJ/m² (7.48 kWh/m²) in

summer and a minimum of 7.31MJ/m² (2 kWh/m²) in winter. Mean daily outdoor temperature

and relative humidity are also displayed in Fig. 3. Contrarily to mean values, the hourly

outdoor temperature can fall to 4°C in winter and reach 40°C in summer. In the Tunisian

climate, which is rather mild and temperate, both heating in winter and cooling in summer are

required to reach comfort levels [38]. This specificity should be taken into account in the

design of the buildings, where both seasons should be concerned.

13
ACCEPTED MANUSCRIPT

3. Energy optimization analysis

In this first investigation, optimization is based only on energy consumption. The objective is

to determine the optimum value of the roof solar reflectivity, providing the minimum energy

consumption, for different roof insulation levels. In this case study, we consider the concrete-

based roof (structures R1 and RI1) and EPS as the insulation material. The solar reflectivity is

varied from 0.1 to 0.9 and the insulation thickness from 0 to 10 cm. The roof surface

emissivity is kept constant and equal to 0.9

Transmission loads per unit area of the roof surface are evaluated during the cooling season

(June to September) and the heating season (December to February) for different

combinations of solar reflectivity values and insulation thicknesses. As shown in Fig. 4, the

effect of the solar reflectivity on the cooling loads is more pronounced compared to the

heating loads. In both seasons, the building energy need is more sensitive to solar reflectivity

for low roof insulation levels. On the other hand, the increase of the insulation thickness plays

a dominant role in reducing both cooling and heating loads. This reduction is fast at smaller

insulation thicknesses and more gradual for larger values. In summer, the energy loads are

less sensitive to the roof insulation thickness for high solar reflectivity values.

In the case of a typical insulation thickness of 5 cm, the cooling transmission loads obtained

with the highest roof solar reflectivity ( =0.9) are nearly 7 times lower than those obtained

with the lowest roof solar reflectivity ( =0.1). According to the variations of the cooling

loads in Fig. 4, one can also notice that adding a cool paint ( =0.9) to the uninsulated roof R1

is equivalent to adding an insulation thickness of 6 cm to a reference roof ( =0.1). In both

cases the annual cooling loads are equivalent to 45.5 MJ/m².

It is obvious, that summer time benefits induced by cool roofs are associated with wintertime

penalties. The discrepancy between benefits and penalties is the basis for calculating the

14
ACCEPTED MANUSCRIPT

optimum roof solar reflectivity, which corresponds to the minimum total energy consumption,

including heating and cooling loads. Fig. 5 shows the total annual transmission loads versus

the solar reflectivity for roof R1 and roof RI1 with different insulation thicknesses. One can

note that, for all the insulation levels, the optimum solar reflectivity (with respect to minimum

energy consumption) is the highest value ( =0.9) and the worst case is the lowest value

(=0.1). Hence, in the Tunisian climate context, the benefits related to the reduction of

cooling energy consumption (provided by cool roofs) outweigh the winter penalties. The total

energy consumption savings between the optimum and the worst roof solar reflectivity

scenarios varies from 155.486 MJ/m² for the uninsulated roof to 17.582MJ/m² for a 10 cm-

insulation thickness roof.

In order to estimate the impact of the cool roof application on the outside surface temperature,

we consider the period from July 20th to July 25th, which is the hottest of the year with a

horizontal solar radiation up to 981 W/m² and a dry-bulb air temperature reaching 39°C (see

Fig. 6a). Comparison is made between a reference gray roof surface and a cool roof surface

after the application of a white elastomeric coating. Based on the measured values provided

by Synnefa et al. [33], the solar reflectivity changes from =0.2 for the reference roof to

=0.89 for the cool roof. Daily time variations of the outside surface temperature are

calculated using EnergyPlus program for R1 and RI1 with a 10 cm-EPS insulation. One can

notice from Fig.(6b) the important decrease of the surface temperature values after the

application of the cool coating for both, insulated and uninsulated cases. A significant

decrease of daily fluctuations is also observed, which has the advantage of preserving the roof

materials from thermal fatigue and ensuring a longer roof lifetime. On the contrary, the roof

insulation causes a surface temperature increase in summer, which is more pronounced in the

case of the reference roof (Fig. 6b). By analyzing the peak values of the roof outside surface

temperature during the summer season, it is found that the reduction resulting from changing

15
ACCEPTED MANUSCRIPT

the reference roof by the cool roof reaches 17.43°C in the uninsulated case and 18.46°C in the

insulated case. Owing to their reduction effect on the outside surface temperature, cool roofs

could be a good solution to mitigating the heat island effect. The increase in the peak outside

surface temperature due to insulation is equivalent to 1.48°C for the reference roof and

decreases to 0.44 °C for the cool roof.

In the Tunisian Mediterranean climate, summer nights are characterized by some coolness,

where the outdoor temperature could be lower than the required indoor temperature (Fig. 6a).

So, a high level of insulation opposes heat evacuation in nighttime leading to an increase in

the cooling energy loads. Indeed, it has been shown by Ghrab-Morcos [39] that insulation of

the walls is not recommended in the Mediterranean summer climate. However, Fig.4 shows

that the roof insulation is always very efficient, whatever its thickness, which is due to the

very high solar radiation incident on the horizontal roof surface in summer days. The same

conclusions were also given in [39].

Therefore, one can conclude from the above findings that, for Tunisian climate, the two roof

measures are efficient in term of energy benefits. In order to take advantage, in addition, of

the environmental benefit of cool roofs and their positive impact on the roof structure and its

lifetime, it is recommended to use moderate roof insulation levels combined with cool roof

surfaces of as high as possible reflectivity value. In the following section, a more appropriate

choice of the insulation thickness and the cool roof strategy is proposed by taking into account

the economic parameters.

4. Cost optimization analysis

The objective of this study is to assess the cost-effectiveness of cool roof coatings, on the one

hand, and to determine an optimum combination between the solar reflectivity of the roof and

its insulation thickness, on the other hand. In order to simulate realistic conditions, it is

important to take into account the impact of ageing, soiling and weathering of the cool coating

16
ACCEPTED MANUSCRIPT

during the building life span. Indeed, due to dust, air pollution, rain and sun damage, the solar

reflectivity of cool roofs could considerably reduce over time, thus affecting the cooling

benefits of the roof surface. According to Bretz and Akbary [40], most of the decrease in solar

reflectivity occurs in the first year, possibly in the first few months. Then, values tend to

stabilize. Eilert [41] also attested that the reduction in the solar reflectivity of white roofs is in

the order of 10–30%, with most of the reduction occurring in the first year. It has been

concluded from a field survey [42] that the energy calculations should be based on at least

one-year aged reflectivity values, which are typically 75-80% of the initial values. However,

experimental tests [40,42] have shown that the original reflectivity could be restored up to

90-100% by a periodic power washing.

4.1. Methodology

Based on the above recommendations, we considered three scenarios of roof reflectivity: (1)

reference case, (2) restored case and (3) aged case. The reference case is a typical gray surface

roof (=0.2). In the restored case, the original reflectivity of the new white coating (=0.89) is

restored by a periodic washing in the beginning of each summer season. We assume that this

summer value reaches a decrease of 20% when the winter season begins (=0.71). The aged

case represents a weathered roof surface, whose reflectivity undergoes a decrease of 25%

from the original new value and so maintains a stabilized average value =0.67. In all the roof

scenarios, the infrared surface emissivity is equal to 0.9.

Taking into account the average service life of the white reflective coating, we conducted a

20-year LCCA, which includes the costs of energy, insulation material and white reflective

coating. The power washing of the cool surface, planned in the restored case, is also included

in the LCCA as an annual maintenance cost. The P1–P2 method proposed by Duffie and

Beckman [43] is adopted, where P1 is the ratio of the life-cycle energy cost to the first year

energy cost and P2 is the ratio of the life-cycle expenditures incurred to the initial investment.

17
ACCEPTED MANUSCRIPT

The economic indicator P1, also known as the Present Worth Factor (PWF), allows to

evaluate the actual value of the energy consumption cost during the life-cycle period of n

years (amortization period) by accounting for the effects of the inflation rate i (for the energy

cost) and the market discount rate d (for the value of money) over this period. It is defined as

[43]:

 1   1  i n 
 1     id
n
(1  i )u 1  d  i   1  d  
P1 (n ,i ,d )  PWF  
u 1 (1  d )
u
 (8)
 n
 id
1  i

By considering only the miscellaneous costs due to annual maintenance of the reflective roof

coating, the economic indicator P2 is reduced to:

P2  1  P1M s (9)

where Ms is the ratio of the annual maintenance cost to the original initial cost.

Based on the types of energy commonly used in Tunisian buildings (electricity for cooling

and natural gas for heating), the life-cycle total cost per unit area of the roof surface, including

respectively the present value of the energy cost, the insulation cost and the present value of

the reflective roof coating cost is written as:

 Q Cel Qh 
Ct  Cenr  Ci  Cr  P1  c  C   Cins Lins  P2Cref (10)
 COP ( 3. 6  10 6
) H 
u s
g

The optimum insulation thickness is the value of Lins which minimizes the total cost Ct. This

optimum value is mainly dependent on the climatic conditions and the base indoor

temperature, the type and the cost of the insulation material and the reflective roof coating, the

type and the cost of the consumed energy, the efficiency of the heating and cooling systems,

the life-cycle period of the building and the inflation and the discount rates.

Net life-cycle savings (NLCS) are obtained by computing the difference between the saved

energy cost (due to reflective coating and/or insulation) and additional payouts over the

18
ACCEPTED MANUSCRIPT

building life-cycle, compared to the conventional roof (uninsulated roof with a reference solar

reflectivity  =0.2):

 Qc Qh 
C g   Cins Lins  P2Cref 
Cel
NLCS  P1   (11)
 COP ( 3.6  10 ) H u  s 
6

where Qc and Qh are saved cooling and heating energy per unit area of the roof surface,

respectively. A positive NLCS indicates that the measure (insulation and/or cool coating) is

cost-effective compared to the conventional case.

The payback period b is obtained by setting the NLCS equal to zero [44]. An estimation of the

payback period which accounts for the effects of the inflation and discount rates is deduced

from the following expression:

b
(1  i )u 1 Cins Lins  P2Cref
P1( b ,i , d )   
(1  d )u  Qc Cel Qh 
(12)
u 1
  Cg 
 COP ( 3.6  10 ) H u  s 
6

Therefore, according to Eq.(8), we have:

  C L  P2Cref 
 Ln 1  (d  i ) ins ins 
b   As  id
  1 i 
 Ln   (13)
 1 d 

 C L  P2Cref
b  (1  i ) ins ins id
 As

where As  Qc Cel



Qh
C g ; is the annual energy saving.
COP ( 3.6  10 ) H u  s
6

The inputs needed for the above economic calculations are the yearly values of the heating

and cooling transmission loads, Qh and Qc, obtained from EnergyPlus, as well as nominal

values of economic parameters based on the local market costs and local conditions at the

time of investigation given in Table 2.

19
ACCEPTED MANUSCRIPT

4.2. Results and discussion

The three roof reflectivity scenarios (described in section 4.1) are firstly analyzed, in terms of

energy consumption, for the uninsulated roofs R1 and R2 (see Fig.2). Table 3 presents the

cooling and heating transmission loads and the equivalent annual energy costs for all the

studied cases. One can note the resulting decrease in cooling loads by the application of the

white reflective coating. In the restored case, this decrease reaches 193.81 MJ/m² for R1 and

58.7MJ/m² for R2. On the other hand, a smaller resulting heat penalty is observed (45.3MJ/m²

for R1 and 13.74 MJ/m² for R2). Indeed, for the two roof structures, both restored and aged

cases lead to savings in annual energy cost including cooling and heating. The cool roof effect

is more pronounced for R1 (of higher U-value), where the annual energy cost is lowered by

3.36 TND/m². As expected and due to its higher solar reflectivity, the restored case is better

than the aged case in term of energy cost savings.

In order to assess the cost-effectiveness of the restored and aged cases of cool roofs, the

LCCA, including cool coating installation and maintenance costs, is applied to the uninsulated

roof configurations R1 and R2. According to Table 4, a positive NLCS is obtained for all the

studied cases, which proves the cost-effectiveness of both cool roof scenarios in the Tunisian

climate context. In the absence of insulation, the restored cool roof case R1 is the most cost

effective, providing the highest net saving of 44.53 TND/m² compared to the reference case.

The installation and maintenance of the white coating are paid back after only 3.4 years. On

the other hand, the roof configuration R2 of lower U-value is low cost-effective, notably for

the aged case where the net savings do not exceed 2.62TND/m² and the payback period is up

to 14.98 years. The results of the above analysis concern typical existing buildings where

roofs are not insulated and for which the cool roof measure could be a good alternative to

insulation.

20
ACCEPTED MANUSCRIPT

The second part of the present economic study concerns new constructions where both

insulation and reflective coatings are applied to the roof. The main objective is to determine

an optimal combination between the roof surface reflectivity and the insulation thickness.

Based on annual transmission loads (including heating and cooling energy requirements) and

using the LCCA, the optimum insulation thickness, which minimizes the life-cycle total cost

(Eq. (10)), is evaluated for each of the three roof reflectivity scenarios. The two insulated roof

configurations RI1 and RI2 (see Fig. 2) and the three types of insulation materials are

considered. We present for RI1 with EPS insulation the variations of the energy consumption

cost, the insulation cost, the cool coating cost and the total cost with respect to the insulation

thickness (Fig. 7). As expected, the total cost curve shows a minimum that corresponds to the

optimum insulation thickness. The results for the different analyzed cases are displayed in

Table 5. In terms of minimum life-cycle total cost and maximum NLCS, the reference case of

RI1 with RW as the insulation material is the most economical with an optimum insulation

thickness of 8.8cm, 73.39TND/m² of NLCS and a payback period of 2.66 years. This is

mainly due to the lower cost of RW compared to other insulation materials and to the

additional cost of application and maintenance of the cool coating for the other cases.

However, the NLCS values obtained in the restored case are sufficiently close to those of the

reference case, where a maximum estimated to 73.03TND/m² is also obtained with RW. In

this case, the optimum insulation thickness is estimated to 5.4cm with a payback period of 3.6

years. Although it is not the most cost-effective case, the restored case with RW could be

interesting if we consider its environmental benefits of reducing the urban heat island effects.

The aged case provides slightly lower NLCS values and longer payback periods, where the

RW remains the most cost-effective insulation material with net savings up to 70.95 TND/m².

The same conclusions are obtained for RI2. However, considerably smaller values of NLCS

and longer payback periods are obtained.

21
ACCEPTED MANUSCRIPT

On the other hand, if one chooses to use EPS or XPS as roof insulation material, Table 5

shows that, among the three roof reflectivity scenarios, the restored case is the most cost-

effective for both RI1 and RI2. For RI1, the optimum insulation of EPS is estimated to 3.9 cm

and the NLCS reaches a maximum value of 72.01 TND/m² which represents 71.03% of

savings. The additional costs are paid back after only 3.69 years.

The results of this cost-optimization study could be affected by uncertainties in the economic

parameters. Therefore, a sensitivity analysis is carried out in order to assess the effect of

insulation cost, electricity cost, natural gas cost, reflective material cost and PWF on optimum

insulation thickness and net savings. Changes are applied to the nominal economic parameters

given in Table 2, and resulting changes in optimum insulation thickness and net savings are

displayed in Figs. (8a) and (8b), respectively. It is noted that Lopt decreases with the increase

of Cins, but increases with the rise of Cel, Cg and PWF. Identical trends of variation were

obtained by Yu and al. [16]. Fig. (8a) shows a higher sensitivity of Lopt to Cins, while Cref has

no effect on Lopt since it is independent of the insulation thickness. However, Cins and Cref

have nearly the same effect on the net saving which decreases with the increase of each of the

two costs. On the contrary, the increase of Cel, Cg and PWF results in higher net savings. Base

indoor temperatures are influential parameters which affect the optimum insulation thickness

and associated payback periods and net savings. Figs. (9a) and (9b) show the sensitivity of the

above three parameters to respective changes in cooling and heating set-point temperatures.

These changes are applied to the base temperatures used in the previous investigations

(Tc=24°C and Th=20°C). One can notice a more pronounced sensitivity of the optimal

parameters to the cooling set-point temperature (see Fig. 9a). Opposite trends of sensitivity

are obtained when changing the heating set-point temperature (see Fig. 9b).

5. Conclusions

The thermal performance of residential building roofs under Tunisian climate was simulated

22
ACCEPTED MANUSCRIPT

using EnergyPlus software in order to assess the impact of thermal insulation and cool

coatings. The CTF method proposed by EnergyPlus was adopted to calculate annual heating

and cooling loads. In spite of its limitations and with an adequate choice of its parameters, the

CTF approach is able to provide stable and accurate results in good agreement with analytical

solutions. Two roof structures, as typically used in Tunisian constructions, and three types of

thermal insulation materials were analyzed. Regarding the cool roof surface, realistic

conditions were simulated by taking into account the impact of ageing and weathering on its

solar reflectivity. Indeed, three reflectivity scenarios were considered, including reference,

restored and aged cases. Simulations with different combinations of solar reflectivity values

and insulation thicknesses showed that the cooling energy loads are more sensitive to the

variations of solar reflectivity compared to heating loads. Their sensitivity to the roof

insulation thickness decreases for higher reflectivity values. On the other hand, the effect of

reflectivity on both heating and cooling loads is more pronounced for low roof insulation

levels.

An energy-based optimization has been firstly carried out in order to determine the optimum

value of the roof solar reflectivity with respect to minimum energy consumption. It was

noticed that, for all the roof insulation levels, the summer time benefits induced by the

increase of reflectivity outweighed the associated winter penalties. Therefore, the optimum

solar reflectivity is the highest possible value. Daily time variations of the roof outside surface

temperature during the hottest summer period showed a significant decrease after the

application of a cool coating, with a reduction in the peak value that reaches 18.46°C. On the

contrary, thermal insulation of the roof slightly increased the outside surface temperature.

So, in terms of energy conservation and environmental benefits, it is recommended to use

moderate roof insulation levels with cool roof surfaces of as high as possible reflectivity

values.

23
ACCEPTED MANUSCRIPT

A 20-year LCCA including the costs of energy, insulation and white coating, as well as the

maintenance costs, has been conducted in order to determine the optimum combination

between the solar reflectivity of the roof and its insulation thickness. For typical existing

buildings with uninsulated roofs, the results proved the cost-effectiveness of both aged and

restored cool roof scenarios in the Tunisian climate context, with the advantage of the restored

case. The concrete-based roof of higher U-value is more cost-effective than the terracotta-

based roof with a net saving up to 44.53 TND/m² and a payback period of 3.4 years. Thus, the

cool roof measure could be a good alternative to insulation in old constructions. The same

LCCA has been performed in the case of new constructions where both insulation and cool

coating are applied to the roof, and the optimum insulation thickness was evaluated for each

of the roof reflectivity scenarios. In terms of minimum life-cycle total cost and maximum

NLCS, the restored concrete-based roof insulated with RW could be interesting as it is close

to the most cost-effective scenario, i.e. the reference case, with in addition its environmental

benefit of reducing the urban heat island effect. Indeed, NLCS were up to 73.03 TND/m²

(versus a value of 73.39 TND/m² for the reference case) with a moderate optimum insulation

thickness of 5.4cm and a payback period of 3.6 years. However, the terracotta-based roof

provided considerably smaller NLCS values with longer payback periods. By choosing EPS

or XPS as insulation materials, the restored case remains the most cost-effective for both roof

structures.

The results of this cost-optimization study are affected by uncertainties in the economic

parameters. Indeed, a sensitivity analysis showed that the optimum insulation thickness and

the net saving increase with the rise of the energy costs and the present worth factor. The

optimum insulation thickness is highly sensitive to the insulation cost, but not affected by the

cool material cost. On the other hand, the increase of the two above costs has nearly the same

decrease effect on the net saving. Furthermore, the optimum insulation thickness, the payback

24
ACCEPTED MANUSCRIPT

period and the net saving are more sensitive to the indoor cooling set-point temperature with

opposite trends of sensitivity when varying the heating set-point temperature.

The analysis procedure proposed in the present paper is a useful tool to help design building

envelope components. This procedure allows other investigations for various external

envelope compositions, where walls and roof constructions of up to 10 layers can be defined

layer by layer in the EnergyPlus file. The same methodology is reproducible for other

countries and climate conditions by choosing among the weather data for more than 2100

locations, available in EnergyPlus. The proposed approach is also applicable to the study of

the efficiency of cool roofs under different climatic conditions where other reflectivity

scenarios could be considered.

References

[1] Opportunités d'affaires en Méditerrannée-Focus sur l'énergie en Tunisie, ANIMA

investment Network, EuroMed@Change; 2013.

http://www.animaweb.org/sites/default/files/ec_tunisia_energie_final_fr-web.pdf

[2] H. Akeiber, P. Nejat, M.Z.A. Majid, M.A. Wahid, F. Jomehzadeh, I.Z. Famileh, J.K.

Calautit, B.R. Hughes, S.A. Zaki, A review on phase change material (PCM) for sustainable

passive cooling in building envelopes, Renew Sustain Energy Rev 60 (2016) 1470-1497.

[3] R. Missaoui, National Strategy of Energy Conservation in Tunisia, Ministry of Energy and

Industry, Tunisia, 2014.

[4] Agence Nationale pour la Maitrise de l’Energie (ANME). http://www.anme.nat.tn

[5] K. Papakostas, T. Mavromatis, N. Kyriakis, Impact of the ambient temperature rise on the

energy consumption for heating and cooling in residential buildings of Greece, Renew Energy

35 (2010) 1376-1379.

25
ACCEPTED MANUSCRIPT

[6] O. Kaynakli, A review of the economical and optimum thermal insulation thickness for

building applications, Renew Sustain Energy Rev 16 (2012) 415-425.

[7] Ö.A. Dombayci, Ö. Atalay, S.G. Acar, E.Y. Ulu, H.K. Ozturk, Thermoeconomic method

for determination of optimum insulation thickness of external walls for the houses: Case study

for Turkey, Sustain Energy Technol and Assess 22 (2017) 1-8.

[8] L. Sagabanua, F. Balo, Ecological impact & financial feasibility of Energy Recovery

(EIFFER) Model for natural insulation material optimization, Energy Build 148 (2017) 1-14.

[9] S.A. Al-Sanea, M.F. Zedan, S.A. Al-Ajlan, Effect of electricity tariff on the optimum

insulation thickness in building walls as determined by a dynamic heat transfer model, Appl

Energy 82 (2005) 313-330.

[10] S.A. Al-Sanea, M.F. Zedan, Optimum insulation thickness for building walls in a hot-dry

climate, Int. J. Ambient Energy 23 (2002) 115–126.

[11] M. Ozel, Effect of wall orientation on the optimum insulation thickness by using a

dynamic method, Appl Energy 88 (2011) 2429-2435.

[12] M. Ozel, The influence of exterior surface solar absorptivity on thermal characteristics

and optimum insulation thickness, Renew Energy 39 (2012) 347-355.

[13] N. Daouas, Z. Hassen, H.B. Aissia, Analytical periodic solution for the study of thermal

performance and optimum insulation thickness of building walls in Tunisia, Appl Therm Eng

30 (2010) 319-326.

[14] N. Daouas, A study on optimum insulation thickness in walls and energy savings in

Tunisian buildings based on analytical calculation of cooling and heating transmission loads,

Appl Energy 88 (2011) 156-164.

[15] N. Daouas, Impact of external longwave radiation on optimum insulation thickness in

Tunisian building roofs based on a dynamic analytical model, Appl Energy 177 (2016) 136–

148.

26
ACCEPTED MANUSCRIPT

[16] J. Yu, L. Tian, C. Yang, X. Xu, J. Wang, Optimum insulation thickness of residential

roof with respect to solar–air degree-hours in hot summer and cold winter zone of china,

Energy Build 43 (2011) 2304–2313.

[17] A.R. Gentle, J.L.C. Aguilar, G.B.Smith, Optimized cool roofs: Integrating albedo and

thermal emittance with R-value, Sol Energy Mater & Sol Cells 95 (2011) 3207–3215.

[18] P. Ramamurthy, T. Sun, K. Rule, E. Bou-Zeid, The joint influence of albedo and

insulation on roof performance: A modeling study, Energy Build 102 (2015) 317–327.

[19] R.S. Arumugam, V. Garg, V.V. Ram, A. Bhatia, optimizing roof insulation for roofs with

high albedo coating and radiant barriers in India, J Build Eng 2 (2015) 52–58.

[20] C. Piselli, M. Saffari, A. de Gracia, A.L. Pisello, F. Cotana, L.F. Cabeza, Optimization of

roof solar reflectance under different climate conditions, occupancy, building configuration

and energy systems, Energy Build 151 (2017) 81-97.

[21] Z. Shi, X. Zhang, Analyzing the effect of the longwave emissivity and solar reflectance

of building envelopes on energy-saving in buildings in various climates, Sol Energy 85 (2011)

28–37.

[22] I. Hernández-Pérez, J. Xamán, E.V. Macías-Melo, K.M. Aguilar-Castro, Reflective

Materials for Cost-Effective Energy-Efficient Retrofitting of Roofs, Chapter 4 in: Cost-

Effective Energy-Efficient Building Retrofitting, Woodhead publishing, México, 2017, pp.

119-139.

[23] J. Sproul, M.P.Wan, B.H. Mandel, A.H.Rosenfeld, Economic Comparison of White,

Green, and Black Flat Roofs in the United States, Energy Build 71 (2013) 20-27.

[24] Z. Zhang, S. Tong, H. Yu, life-cycle analysis of cool roof in tropical treas, Procedia Eng

169 ( 2016 ) 392-399.

27
ACCEPTED MANUSCRIPT

[25] J.H. Jo, J.D. Carlson, J.S. Golden, H. Bryan, An integrated empirical and modeling

methodology for analyzing solar reflective roof technologies on commercial buildings, Build

Environ 45 (2010) 453-460.

[26] J. Yuan, C. Farnham, K. Emura, M.A. Alam, Proposal for optimum combination of

reflectivity and insulation thickness of building exterior walls for annual thermal load in

Japan, Building Environ 103 (2016) 228-237.

[27] EnergyPlus engineering reference, Version 8.7 documentation, University of Illinois and

Ernest Orlando Lawrence Berkeley national laboratory, USA; 2016.

[28] EnergyPlus energy simulation software, Weather data. https://energyplus.net/weather

[29] R. Perez, R. Stewart R, C. Arbogast, R. Seals, J. Scott, An anisotropic hourly diffuse

radiation model for sloping surfaces: description, performance, validation, site dependency

evaluation. Sol Energy 36 (1986) 481-497.

[30] EnergyPlus input output reference, Version 8.7 documentation, University of Illinois and

Ernest Orlando Lawrence Berkeley national laboratory, USA; 2016.

[31] J. Mun, M. Krarti, Implementation of a new CTF method stability algorithm into

EnergyPlus, Build Simul 8 (2015) 613-620.

[32] M. Rossi, V.M. Rocco, External walls design: The role of periodic thermal transmittance

and internal areal heat capacity, Energy Build 68 (2014) 732-740.

[33] A. Synnefa, M. Saliari, M. Santamouris, Experimental and numerical assessment of the

impact of increased roof reflectance on a school building in Athens, Energy Build 55 (2012)

7–15.

[34] Guide pratique de conception de logements économes en énergie - Réglementation

thermique et énergétique des bâtiments neufs en Tunisie, ANME, Tunisie; 2006.

http://mediplaco.com/pdf/guide Conception Batiments.pdf

28
ACCEPTED MANUSCRIPT

[35] E. Znouda, N. Ghrab-Morcos, A. Hadj-Alouane, Un algorithme génétique pour

l’optimisation énergétique et économique des bâtiments Meditérranéens, 6ème Conférence

Francophone de MOdélisation et SIMulation-MOSIM’06, Rabat-Maroc; 2006.

[36] S.A. Al-Sanea, Finite-volume thermal analysis of building roofs under two-dimensional

periodic conditions, Build Environ 38 (2003) 1039-1049.

[37] S. Trachte, A. De Herde, Elaboration d’un outil d’aide à la conception de maisons à très

basse consommation d’énergie: choix des matériaux - écobilan de parois. Rapport externe,

Wallonie; 2010.

[38] N. Ghrab-Morcos, Energy and financial considerations related to wall design for a

conditioned cell in Tunisian conditions, Renew Energy 1 (1991) 145-159.

[39] N. Ghrab-Morcos, CHEOPS: a simplified tool for thermal assessment of Mediterranean

residential buildings in hot and cold seasons, Energy Build 37 (2005) 651–662

[40] S.E. Bretz, H. Akbari, Long-term performance of high-albedo roof coatings, Energy

Build 25 (1997) 159-167.

[41] P. Eilert, High Albedo (Cool) Roofs: Codes and Standards Enhancement (CASE) Study,

Pacific Gas and Electric Company, San Francisco, USA, 2000.

[42] Durability, Pausing for reflection on aging of the cool roof, Journal of Architectural

Coatings, June/July 2008.

http://www.paintsquare.com/library/articles/Durability_Pausing_for_reflection_on_aging_of_

the_cool_roof.pdf

[43] J.A. Duffie, W. Beckman, Solar engineering of thermal processes, John Wiley and Sons

Inc. (2006).

[44] F.F.C.Vincelas, T. Ghislain, T. Robert, Influence of the types of fuel and building

material on energy savings into building in tropical region of Cameroon, Appl Therm Eng 122

(2017) 806-819.

29
ACCEPTED MANUSCRIPT

Tsky (t)

qs(t) Ta(t)
qr,o(t) hc,o
Outside
xN

Layer N
kN, αN LN

x2
Layer 2
L2
k2, α2
o
x1 Layer 1 L1
k1, α1
o
Inside
hi Ti

Fig. 1
ACCEPTED MANUSCRIPT

R1
Outside
Sealing (0,5 cm)
Cement Plaster (2,5 cm)
Concrete (10 cm)

Reinforced concrete (15 cm)

Cement plaster (2 cm)


Inside

R2
Outside Sealing (0,5 cm)
Cement plaster (2,5 cm)
Concrete (10 cm)

Reinforced concrete (5 cm)

Hollow terracotta (16 cm)

Cement plaster (2 cm)


Inside

RI1
Outside
Sealing (0,5 cm)
Cement Plaster (2,5 cm)
Concrete (10 cm)

Thermal insulation

Reinforced concrete (15 cm)

Cement plaster (2 cm)


Inside

RI2
Outside
Sealing (0,5 cm)
Cement plaster (2,5 cm)
Concrete (10 cm)

Thermal insulation
Reinforced concrete (5 cm)

Hollow terracotta (16 cm)

Cement plaster (2 cm)


Inside

Fig. 2
ACCEPTED MANUSCRIPT

36 100

Horizontal solar radiation (MJ/m².day)


32 Outdoor dry bulb temperature 95
Relative humidity
90
28
26,94
25,53
85
24 23,4 23,41

Relative humidity (%)


Temperature (°C)

80
20
18,86
17,5 75
16 15,84
70
12,41
12 11,15
65
8,92
8,23
8 7,31
60

4 55

0 50

1 2 3 4 5 6 7 8 9 10 11 12
Month number

Fig.3
ACCEPTED MANUSCRIPT

300

 = 0,1
 = 0,3
200  = 0,5
Cooling  = 0,7
 = 0,9
Annual transmission loads (MJ/m²)

100

-100

Heating
-200

-300

0 0,01 0,02 0,03 0,04 0,05 0,06 0,07 0,08 0,09 0,1
Insulation thickness (m)

Fig.4
ACCEPTED MANUSCRIPT

500

No insulation
450 Lins= 2 cm
Lins= 4 cm
400 Lins= 6 cm
Total annual transmission loads, Qt (MJ/m²)

Lins= 8 cm
Lins= 10 cm
350

300

250

200

150

100

50

0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9


Solar reflectivity, 

Fig.5
ACCEPTED MANUSCRIPT

a
60 1100
Horizontal solar radiation Outdoor air temperature
1000
55

900
50
800

45

Solar radiation (W/m²)


700
Temperature (°C)

40 600

35 500

400
30

300
25
200

20
100

15 0

0:00
16:00

16:00
8:00

8:00
0:00

0:00

0:00

8:00

16:00
16:00

16:00

16:00
8:00

8:00

8:00
0:00

0:00

0:00

July 20 21 22 23 24 25

b 60

Reference roof (uninsulated)


Reference roof (insulated)
55

50
Outside surface temperature (°C)

45

40

35

30

25

20
Cool roof (uninsulated)
Cool roof (insulated)

15
0:00
16:00
8:00
0:00

0:00

8:00

0:00

8:00
16:00

16:00
16:00

16:00
8:00

8:00
0:00

0:00

0:00

8:00

16:00

July 20 21 22 23 24 25
Fig.6
ACCEPTED MANUSCRIPT

60

Energy cost
50 Insulation cost
Cool coating cost
Total cost

40
Life-cycle cost (TND/m²)

30

20 Lopt= 3.9 cm

10

0 0,01 0,02 0,03 0,04 0,05 0,06 0,07 0,08 0,09 0,1
Insulation thickness (m)

Fig.7
ACCEPTED MANUSCRIPT

a
200

Cins
150 Cel
Cg
Cref
PWF
100
Change in Lopt (%)

50

-50

-100

-100 -80 -60 -40 -20 0 20 40 60 80 100


Change in economic parameters (%)

b
120

Cins
Cel
80 Cg
Cref
PWF
40
Change in net saving (%)

-40

-80

-120

-100 -80 -60 -40 -20 0 20 40 60 80 100


Change in economic parameters (%)

Fig.8
ACCEPTED MANUSCRIPT

a
50

40

30
Change in optimum values (%)

20

10

-10

-20
Insulation thickness
Payback period
-30
Net savings

-40

-4 -3 -2 -1 0 1 2 3
Change in cooling set-point temperature, Tc(°C)

b
20

15

10
Change in optimum values (%)

-5

-10
Insulation thickness
Payback period
-15 Net savings

-20

-3 -2 -1 0 1 2 3 4
Change in heating set-point temperature, Th(°C)

Fig.9
ACCEPTED MANUSCRIPT

Figures Captions:

Fig. 1. Schematic representation of N-layered roof and boundary conditions.

Fig. 2. Typical structures of concrete-based roofs, R1 and RI1, and terracotta-based roofs, R2

and RI2, without and with insulation, respectively.

Fig. 3. Climatic data of Tunis [28].

Fig. 4. Effect of the solar reflectivity and EPS insulation thickness of the concrete-based roof

on the annual heating and cooling transmission loads.

Fig. 5. Effect of the solar reflectivity of the concrete-based roof on the total annual loads,

including heating and cooling, for different levels of EPS insulation.

Fig. 6. Impact of the cool roof application on the outside surface temperature for R1 and RI1

(a) Climatic conditions during the hottest period of the year (b) Corresponding daily time

variations of the outside surface temperature of cool roofs compared to reference roofs.

Fig. 7. Total life-cycle cost and its different components versus insulation thickness, for

restored roof RI1 with EPS insulation.

Fig. 8. Sensitivity of the optimum values to economic parameters (a) optimum insulation

thickness changes (b) optimum net saving changes.

Fig. 9. Sensitivity of the optimum values to indoor set-point temperatures (a) cooling set-point

temperature effect (b) heating set-point temperature effect.


ACCEPTED MANUSCRIPT

Highlights

 Optimum combination of surface reflectivity and insulation of roof is proposed.

 Cool roof with moderate insulation is recommended in Mediterranean climate.

 Extensive LCCA considering long-term performance of cool roof is performed.

 Restored concrete-based roof with 5.4 cm-RW thickness is the most cost-

effective.

 The proposed methodology applies to other roof reflectivity scenarios and

climates.
ACCEPTED MANUSCRIPT

Table 1
Thermal properties of materials

k  α×107
(W/m.K) (kg/m3) (m²/s)
Concrete 1.65 2000 8.25
Reinforced concrete 2.5 2400 10.41
Hollow terracotta 0.2 650 3.07
Cement plaster 1.4 2200 6.36
Cement-lime plaster 1 1900 5.26
Sealing 0.23 1150 2
Expanded polystyrene (EPS) 0.037 30 8.5
Extruded polystyrene (XPS) 0.032 36 6.13
Rock wool (RW) 0.047 40 11.4
ACCEPTED MANUSCRIPT

Table 2
Parameters used in the economic calculations

Parameter Value
White reflective material
Initial cost, Cref 9.1 TND/m2
Annual maintenance cost, Ms 1%
Expanded polystyrene (EPS)
Cost, Cins 212 TND/m3
Extruded polystyrene (XPS)
Cost, Cins 330 TND/m3
Rock wool (RW)
Cost, Cins 147 TND/m3
Electricity (for cooling)
Cost, Cel 0.21 TND/kWh
Coefficient of performance, COP 3
Natural gas (for heating)
Cost, Cg 0.2515 TND/m3
Heating value, Hu 34.526  106 J/m3
Efficiency, s 0.8
Inflation rate, i 5.5 %
Discount rate, d 7%
Lifetime, n 20 years
ACCEPTED MANUSCRIPT

Table 3
Annual cooling and heating energy consumption and corresponding energy cost for
uninsulated roofs

R1 R2
(U = 5.55 W/m².K) (U = 1.06 W/m².K)

Roof reflectivity Cooling Heating Energy cost Cooling Heating Energy cost
cases (MJ/m²) (MJ/m²) (TND/m²) (MJ/m²) (MJ/m²) (TND/m²)

Reference case 241.44 163.32 6.18 72.46 49.29 1.85

Restored case 47.63 208.62 2.82 13.76 63.03 0.84

Aged case 102.02 205.03 3.85 30.02 61.95 1.14


ACCEPTED MANUSCRIPT

Table 4
20-year net life-cycle savings (NLCS) and payback periods for two reflectivity scenarios of
uninsulated roofs

R1 R2
(U = 5.55 W/m².K) (U = 1.06 W/m².K)

NLCS Payback NLCS Payback


Roof reflectivity cases (TND/m²) (years) (TND/m²) (years)

Restored case 44.53 3.40 6.16 11.93

Aged case 29.20 4.23 2.62 14.98


ACCEPTED MANUSCRIPT

Table 5
Optimum insulation thickness, minimum total cost, energy savings and payback periods
for the different roof scenarios and insulation materials

RI1 (Uwt = 5.55 W/m².K)a RI2 (Uwt = 1.06 W/m².K)a

Reference Restored Aged Reference Restored Aged

Optimum insulation thickness (cm)


EPS 6.4 3.9 4.8 3.5 1.2 2.0

XPS 4.6 2.8 3.4 2.2 0.4 0.97

RW 8.8 5.4 6.5 5.2 1.9 3.0

Minimum total cost (TND/m²)


EPS 29.66 29.37 31.67 23.66 23.55 25.75

XPS 33.93 31.95 34.84 25.94 24.14 26.87

RW 27.98 28.34 30.42 22.76 23.22 25.21

Payback period (years)


EPS 2.82 3.69 3.86 9.72 12.41 14.18

XPS 3.27 3.96 4.18 11.63 12.36 14.93

RW 2.66 3.60 3.71 9.25 12.28 13.78

Net life-cycle savings, NLCS (TND/m²)


EPS 71.72 72.01 69.70 6.81 6.91 4.71

XPS 67.44 69.42 66.53 4.52 6.32 3.59

RW 73.39 73.03 70.95 7.71 7.24 5.25

Life-cycle savings (%)


EPS 70.74 71.03 68.75 22.35 22.68 15.47

XPS 66.52 68.48 65.63 14.86 20.77 11.78

RW 72.39 72.04 69.98 25.30 23.78 17.24


a Without insulation layer

Вам также может понравиться