You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/280032307

Determination of optimum insulation thickness in different wall orientations


and location in Iran

Article  in  Advances in Building Energy Research · July 2015


DOI: 10.1080/17512549.2015.1079239

CITATIONS READS

13 484

3 authors, including:

Hadi Ramin Pedram Hanafizadeh


University of Saskatchewan University of California, Berkeley
20 PUBLICATIONS   45 CITATIONS    129 PUBLICATIONS   608 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Heat transfer and hydrodynamic performance analysis of a miniature tangential heat sink using AL2O3-H2O and TiO2-H2O nanofluids View project

Investigation of the viscous and inertial contributions to the pressure drop of air flow at the end-winding region of generator View project

All content following this page was uploaded by Pedram Hanafizadeh on 17 August 2015.

The user has requested enhancement of the downloaded file.


TAER1079239 Techset Composition India (P) Ltd., Bangalore and Chennai, India 8/14/2015

Advances in Building Energy Research, 2015


http://dx.doi.org/10.1080/17512549.2015.1079239

REVIEW
5
Determination of optimum insulation thickness in different wall orientations
and locations in Iran
Hadi Ramin, Pedram Hanafizadeh* and Mohammad Ali Akhavan-Behabadi

10 Center of Excellence in Design and Optimization of Energy Systems, School of Mechanical Engineering,
College of Engineering, University of Tehran, Tehran, Iran
(Received 15 September 2014; accepted 12 July 2015)

The present study numerically investigated the optimum insulation thickness determination for
conventional walls in Tehran, the capital of Iran. In this study, aerated brick and concrete were
15 considered as the main wall materials, and XPS and EPS as the insulation materials. The one-
dimensional transient heat transfer problem for multi-layer walls has been solved to obtain
temperature distribution within the wall. Different combinations of wall materials and
insulations were examined. Furthermore, the effect of the position of the insulation (inside
and outside of the wall) was studied as well. Finally, in order to determine the optimum
thickness, which minimizes the total cost of insulation and energy dissipation, economic
20
analysis was carried out for a lifetime of 25 years. It is worth mentioning that in the present
study, both cooling and heating seasons were taken into account in the optimization process.
The findings revealed that after using insulation, among different wall configurations, yearly
transmission load can be decreased in the range of 70–82% compared with an uninsulated
wall made from concrete and 31–58% for the aerated brick wall. Moreover, the findings
indicated that two different locations of insulations resulted in an approximately equal
transmission load and optimum insulation thickness, while their time lag and decrement
25 factor were not the same.
Keywords: optimal insulation thickness; multi-layer wall; wall orientations; time lag;
decrement factor
Nomenclature
Alphabetic symbols Rb geometric factor
30
Ao amplitude of the temperature wave at the Se net energy saving per unit area
outer surface of a wall ($/m2 )
Ai amplitude of the temperature wave at the Tb base temperature
inner surface of a wall
As annual energy saving per unit area ($/m2 ) Ti indoor air temperature
B dummy variable (n − 1) 360/365 To outdoor air temperature
35 b payback period (year) Te sol-air temperature
Cenr cost of energy ($/m2 ) Tsurr surrounding temperature
Ce cost of electricity ($/kWh) tTmax
x=0
time when outdoor temperature is
at its maximum (h)
Cg cost of natural gas ($/m3 ) tTmax
x=L
time when indoor temperature is at
its minimum (h)
Ct total cost (energy + insulation) TXmax
=L maximum temperature at the inner
40
surface of a wall (K)

*Corresponding author. Email: hanafizadeh@ut.ac.ir

45 © 2015 Taylor & Francis


2 H. Ramin et al.

Cins cost of insulation material per volume ($/m3 )TXmin


=L minimum temperature at the inner
surface of a wall (K)
c heat capacity (J/kg · K) TXmax
=0 maximum temperature at the outer
surface of a wall
d inflation rate (%) TXmin
=0 minimum temperature at the outer
50 surface of a wall
E equation of time x distance (m)
f decrement factor
Hu lower heating value of the fuel (J/m3 ) Greek symbols
hi indoor surface combined convection heat a thermal diffusivity (m2 /s)
transfer coefficient (J/m2 · K)
55 ho outdoor surface combined convection heat as solar absorptivity (W )
transfer coefficient (J/m2 · K)
i interest rate (%) b slope of the surface (W )
I total solar radiation on a horizontal surface g surface azimuth angle (W )
(W/m2 )
Ib beam solar radiation on a horizontal surface d declination angle (W )
(W/m2 )
60 Id diffuse solar radiation on a horizontal surfaceDT temperature difference
(W/m2 )
IT total solar radiation (W/m2 ) u angle of incidence (W )
k thermal conductivity (W/m · K) uz zenith angle (W )
Lins thickness of the insulation material (m) r density (kg/m3 )
Lloc longitude of the location rg ground reflectance
65 Lst standard meridian F time lag (h)
t time (s or h) f latitude angle (°)
n the day of the year v hour angle (°)
N number of daylight hours eDT /ho correction factor in the sol-air
temperature equation
P period of the temperature wave (24 h) hs efficiency of the heating system
(%)
70 Qc cooling transmission load (J/m2 ) Acronyms
(Qc )un cooling transmission load through an COP coefficient of performance
uninsulated wall (J/m2 )
Qh heating transmission load (J/m2 ) EPS expanded polyurethane
(Qh )un heating transmission load through an PWF present worth factor
uninsulated wall (J/m2 )
75
q heat transfer (W/m2 ) XPS extruded polystyrene

1. Introduction
Increasing demand for energy in the recent years is a result of population growth, urbanization,
migration to larger cities and improvement of living standards. Nowadays, the building sector (resi-
80 dential, commercial and public) is one of the most important parts of energy consumption in each
country. Finding ways for energy conservation in the building sector is an increasingly important
issue (Balaras, Droutsa, Argiriou, & Asimakopoulos, 2000, 2005; Khudhair & Mohammed, 2004).
Energy loss from the building envelope is one of the main sources of energy dissipation in buildings
since it serves as an interface between indoor and outdoor environments (Yu, Tian, Yang, Xu, &
85 Wang, 2011). Implementing insulations in walls is an important option to reduce the energy con-
sumption in buildings. Finding proper materials, designing the building envelope and considering
the location and orientation of its component are efficient means to reduce the annual heating and
cooling load and consequently the need for energy (Kaynakli, 2012).
It is obvious that larger insulation thickness will cause higher energy saving. However,
increasing the insulation thickness will increase the investment cost linearly (Ozel, 2013b). There-
90 fore, it is necessary that thermal and economic analysis be considered simultaneously to obtain the
Advances in Building Energy Research 3

optimum insulation thickness, which is referred to as thermal economic insulation thickness


(Kaynakli, 2012). Optimum thickness of insulation is one that compensates the investment
cost of insulation with energy-saving benefit in less time. Therefore, insulation (type and thick-
ness) must be selected based on both energy consumption and cost of insulation (Ozel, 2013b).
95
Reliable estimation of the heating and cooling load is an important problem concerning optim-
ization of insulation thickness. A simple and crude model for estimation of heating and cooling
load under a static condition is known as degree-day or degree-hour method which is based on
either ambient air temperature or sol-air temperature. Many studies in the literature used this
model based on ambient temperature to obtain the optimum insulation (Al-Sanea, Zedan, &
Al-Hussain, 2012; Bolattürk, 2008; Kaynakli, 2008; Ucar & Balo, 2009; Wang, Huang, &
100 Heng, 2007). Also, several authors used sol-air temperature instead of ambient air temperature
in this method (Al-Khawaja, 2004; Ozel & Pihtili, 2007; Yu, Yang, Tian, & Liao, 2009). The
dynamic time-dependent model is an accurate model to determine heat loss or gain from the com-
posite walls, which has been used in models proposed by several authors (Al-Sanea, Zedan, & Al-
Ajlan, 2005; Daouas, 2011; Daouas, Hassen, & Ben Aissia, 2010; Ozel, 2012, 2013a; Ozel; Mav- AQ1
105 romatidis, EL Mankibi, Michel, & Santamouris, 2012). The dynamic transient model has been ¶
solved numerically and analytically for multi-layer walls in the literature. It should be noted
that optimal insulation thickness can be obtained based on only cooling loads or only heating
loads and also both heating and cooling loads simultaneously.
In Iran the building sector’s share of final energy consumption is close to 37% of total con-
sumed energy each year (Ministry of Energy, 2013). Emission of carbon dioxide (CO2), sulfur
110
dioxide (SO2) and the other pollutant gases is the most important factor regarding energy con- AQ2
sumption in buildings especially in large cities. During recent years, levels of pollution gases ¶
in Tehran have increased hazardously and become the cause of many diseases and fatalities;
especially in winter during which inversion worsens the situation. Based on the report of the Min-
istry of Energy (Ministry of Energy, 2013), the building sector’s (residential, commercial and
115 public) share of emission of greenhouse gases is 25.6%. Eight million people at night and over
12 million people during the day living in Tehran make it the most populated city in Iran,
which nearly accounts for 10% of the total population in Iran (Statistical Centre of Iran, 2012).
A subsidy reform law, which was passed in Iranian Parliament on 5 January 2010, is planning
for free market energy prices in the next few years. Considering this plan, much attention must
be given to the conservation of energy, and applying thermal insulation material is one of those.
120
The main objective of this study is investigation of optimal insulation thickness in the climate of
Tehran based on the present price of insulation materials and energy. The conventional wall struc-
ture and materials (aerated brick and concrete block) in Tehran are examined here. Expanded poly-
styrene (EPS) and extruded polystyrene (XPS) are used as insulation. The dynamic transient model
is used here and optimal insulation thickness is achieved when insulation is located on both the
125 inside and outside surfaces of the main wall materials. All orientations of a wall, including
south, west, east and north and also horizontal wall (roof), are under consideration. The next
section of the study is devoted to the governing equation, calculation of solar radiation on the
wall, outdoor temperature, the structure of the wall, economic analysis and finally numerical
method to solve the problem. Afterwards, results and discussions are presented.
130

2. Mathematical formulation
2.1. Governing equations and boundary conditions
A schematic of an ordinary multi-layer wall is shown in Figure 1. Thermo-physical properties and
135 length of each layer are shown in this figure. The outside layer of the wall is exposed to periodic
4 H. Ramin et al.

ambient temperature and solar radiation and the inside layer of the wall is exposed to convection
resulting from constant air temperature inside the building.
Assuming no heat generation, constant thermal properties, one-dimensional conduction heat
transfer and negligible interface resistance, the unsteady heat conduction equation for each layer
140
of a multi-layer wall is as follows (Al-Sanea et al., 2005; Ozel, 2013b):

∂2 T 1 ∂T
= , (1)
∂x2 a ∂t

where a = k/rc is the thermal diffusivity of each layer and k, c and r are the thermal conduc-
145
tivity, the heat capacity and the density of each layer, respectively. To solve the problem, two
boundary conditions on both sides of the multi-layer wall and one initial condition are needed.
As the initial condition, the constant inside temperature at t = 0 is considered (Asan, 2000). On
both sides of the wall, convection heat transfer exists. On the inside of the wall, the convection
boundary condition is as follows:
150

∂T 
−k  = hi (Tx=L − Ti ). (2)
∂x x=L

In the above equation, hi is the inside combined (convection and radiation) heat transfer coef-
155 ficient and Ti is the indoor temperature, which is taken to be constant (Ozel, 2013b).

160

165
Colour online, B/W in print

170

175

Figure 1. Multi-layer wall and boundary condition on it; the insulation can be placed either on the inside
surface of the main wall material (current structure) or on the outside of the wall between the cement plaster
180 and the main wall material.
Advances in Building Energy Research 5

And at the outer surface of the wall, the boundary condition can be written as follows:

∂T 
k = ho (Tx=0 − Te (t)), (3)
∂x x=0
185
where ho is the outside convection heat transfer coefficient and Te is the sol-air temperature
including the effect of solar radiation on the outside of the wall. Sol-air temperature is defined
as (Threlkeld, 1998):

aIT eDT
190 Te (t) = To + − . (4)
ho ho

In the above equation, To , IT and a are the outdoor air temperature, total solar radiation and
solar absorptivity of the outdoor wall surface, respectively. The last term of Equation (4), eDT /ho ,
is the correction factor; ranging from 0°C for vertical walls to 4°C for horizontal walls (Mavro-
195
matidis et al., 2012; Ozel, 2013b). The solar absorptivity of daring colour wall is taken as equal to AQ3
0.8 (Ozel, 2013b). ¶

200
2.2. Calculation of solar radiation on the outer wall
Total solar radiation on the wall with different orientations is assessed by the method presented by
Duffie and Beckman (1991), which is modified with given values every minute (Mavromatidis
et al., 2012). Based on this method, the relation between solar time and standard time is given as
205
Solar time = Standard time − 4(Lst − Lloc ) + E, (5)

where Lst is the standard meridian for the local time zone, Lloc is the longitude of the location
and E is the equation of time, which is expressed as follows:
210
E = 229.2(0.000075 + 0.001868 CosB − 0.032077 sinB − 0.014615 cos2B
− 0.04089 sin 2B). (6)

In the above equation, B (in degrees) is defined as follows:


215
B = (n − 1)(360/365) 1 ≤ n ≤ 365 , (7)

where n is the day of the year.


The geometric relationship between a plan of any particular orientation relative to the earth at AQ4
any time and any incoming beam solar radiation can be described in terms of several angles ¶
220
(Duffie & Beckman, 1991). The angular position of the sun at solar noon with respect to the AQ5
plan of the equator is declination angle, d, and varies (−23.45W ≤ d ≤ 23.45W ) throughout the ¶
year and can be calculated by the following relation:

d = 23.45 sin(360(284 + n/365)) n = the day of the year 1 ≤ n ≤ 365 . (8)


225
6 H. Ramin et al.

There is a set of useful relationships between the angles. The angle of incidence (u) between
the sun’s rays and a tilted surface is given by

cos(u) = sin(d) sin(f) cos(b) − sin(d) cos(f) sin(b) cos(g)


230 + cos(d) cos(f) cos(b) cos(v)
(9)
+ cos(d) sin(f) sin(b) cos(g) cos(v)
+cos(d) sin(b) sin(g)sin(v),

where d, f, v and g are declination angle, latitude angle, hour angle and surface azimuth
235 angle, respectively. b is the slope of the surface ranging from 0 for the horizontal
surface to 180 for the horizontally upside down surface. It is notable that b = 0 represents
the vertical wall. The surface azimuth angle varies between −180 ≤ g ≤ 180
(g = −180, −90, 0, 90 and 180 for the surface facing north, east, south, west and again
north, respectively). The hour angle, which is the angular displacement of the sun east or west
240 of the local meridian due to rotation of the earth on its axis at 15W per hour, being negative in
the morning and positive in the afternoon, is calculated in minutes by the following equation:
  
360 (Solar Time − 720)
v= . (10)
24 60
245
The angle between the vertical and the line to the sun is the zenith angle (uz ), which has the
following relationship:

cos(uz ) = sin(d) sin(f) + cos(d) cos(f) cos(v). (11)

250 Setting the cosine of the angle to zero in Equation (11) results in the following relation for the
hour angle (vs ) at sunset:

cos(vs ) = − tan(f) tan(d). (12)

255 The sunrise hour angle is the negative of the sunset hour angle. It follows that the number of
daylight hours is given by

2
N= cos−1 (− tan(f) tan(d)). (13)
15
260 It takes 60 minutes to transverse 15W of longitude; the exact time of sunset (in minutes) can be
found by

60 −1
sunset time = cos (− tan(f) tan(d)). (14)
15
265
And sunrise time can be calculated using equations (13) and (14):

sunrise time = sunset time − N . (15)

The isotropic clear sky method of Hottle (1976) is used to determine total solar radiation.
270 Details of the method of calculation can be found in Duffie and Beckman (1991) and
Advances in Building Energy Research 7

Mavromatidis et al. (2012). Here, we just introduce the equation for calculation of total solar radi-
ation, which is used in the sol-air temperature in Equation (4), as follows:
   
1 + cos(b) 1 − cos(b)
IT = R b Ib + Id + I rg . (16)
275 2 2

where Ib , Id and I are beam, diffuse and total solar radiation on the horizontal surface (three
main components of radiation on a tilted surface). In Equation (16) rg is the ground reflectance
and usually taken as 0.2 and Rb is a geometric factor which can be calculated by
280 cos(u)
Rb = . (17)
cos(uz )

285 2.3. The structures of external walls and the roof


Conventional walls and roofs in Tehran have been considered in this study. The walls and roofs
consist of 2 cm cement plaster at the outer layer, 2 cm gypsum in the inner surface and 20 cm
main wall materials, usually aerated brick or concrete. The insulation layer which can be
placed either on the inside or on the outside surface of the main wall materials is used to save
290 energy. The two most conventional insulation materials, XPS and EPS, are used as insulation
(Figure 1). The thermo-physical properties of materials used in this paper have been tabulated
in Table 1.

2.4. Estimation of heating and cooling transmission load


295
The twenty-first of each month is considered as the representative day for calculation of both
cooling and heating demands (Ozel, 2013b). As mentioned earlier, indoor temperature is
assumed to be constant throughout the day, and hence instantaneous heat transfer through a
wall, after determining the inside wall temperature Tx=L from numerical solution, can be calcu-
lated as follows:
300

qi = hi (Tx=L − Ti ). (18)

Integrating this value over the day, daily heating or cooling transmission load can be obtained.
Yearly heating and cooling transmission loads are also obtained by daily heating or cooling load
305

Table 1. Thermo-physical properties of material used in the paper (most of these data have been extracted
from Najafian and Bahrami (2012), Ozel (2013b) and 19th ntional regulaion for building (2012).
Materials Thermal Conductivity k (W/m · K) Density r (kg/m3 ) Specific Heat c (1/kg · K)
310
Aerated brick 0.3 1000 840
Concrete block 1.37 2076 880
Cement plaster 0.7 2778 840
Gypsum plaster 0.18 800 1090
XPS 0.029 40 1213
315
EPS 0.038 17 1500
8 H. Ramin et al.

over its season (Ozel, 2013b). Constant inside temperature is selected based on thermal comfort
and energy saving for each month (Al-Sanea et al. 2005; Ozel, 2012). These values, which vary
for each month, are summarized in Table 2.

320
2.5. Time lag and decrement factor
There is a different temperature profile on the outer surface of the wall. The temperature profile in
the outer surface changes while penetrating the wall. Time lag and decrement factor are two
important characteristics that are defined in problems of propagating heat through a wall
(Asan, 2000). The time taken for maximum or minimum temperature waves’ propagation from AQ6
325 the outer surface of the wall to the inner surface is known as time lag. Also, decrement factor ¶
is defined as the ratio of temperature wave amplitude that enters the wall. Time lag and decrement
factor depend on the thermo-physical properties, thickness and materials of the walls (Asan &
Sancaktar, 1998). The time lag is defined as (Asan, 2000):

330 tTmax . tTmax  tTmax − tTmax ,
f= x=0 x=L x=0 x=L
(19)
tTx=0 , tTx=L  tTx=0 − tTmax
max max max
x=L
+ P,

where tTmax
o
and tTmax
i
are the time when outside and inside temperatures are at their maximum,
respectively, and P is the period of the wave (24 hours here). Also the decrement factor is defined
as (Asan, 2000):
335

=L − TX =L
Ai TXmax min
f = = max , (20)
Ao TX =0 − TXmin
=0

Ao and Ai are the amplitudes of the temperature wave at the outer and inner of the wall. TXmax
=L , AQ7
340
TXmin
=L and TXmax min
=0 , TX =0 are maximum and minimum temperatures at the inner and outer surfaces of

the wall, respectively. Time lag and decrement factor are calculated for the representative days of
each month of the cooling and heating seasons, and finally yearly time lag and decrement factor
are obtained by arithmetic average of these monthly values (Ozel, 2013b).

345 2.6. Economic analyses


As mentioned earlier, optimal insulation thickness is obtained after economic analysis as well as
thermal analysis are carried out. Utilization of insulation in the wall is accompanied by
additional costs of insulation material and installation charge, although it reduces yearly
heating and cooling loads and hence saves money. Economic analysis is performed to obtain
350 the thickness that has the highest economic efficiency. Economic analysis depends on the
cost of energy, yearly heating transmission load of insulated and un-insulated walls, efficiency
of heating and cooling systems, lifetime of the building and inflation and interest rates (Ozel,
2013b). In order to accomplish economic analysis, net energy saving (which has actually
reduced fuel consumption) from using insulation over the building’s lifetime is evaluated in
the present value by using present worth factor (PWF; Daouas, 2011; Kaynakli, 2012). PWF
355
is defined as (Daouas, 2011):
⎧    

⎪ 1+i 1+i n
n 
 u ⎨ 1− , i = d,
1+i
PWF = = d−i 1+d (21)
1+d ⎪
⎪ n
u=1 ⎩ , i = d,
360 1+i
405

400

395

390

385

380

375

370

365
Table 2. Constant Indoor temperature for each month (Al-Sanea et al., 2005; Ozel, 2012).
Month January February March April May June July August September October November December

Advances in Building Energy Research


Temperature(°C) 20 20 20 22 23 23 23 23 23 22 20 20

9
10 H. Ramin et al.

where i, d and n are interest rate, infiltration rate and lifetime period, respectively. Lifetime
period is assumed to be 25 years. The total present cost of heating and cooling per unit area
during the lifetime, Cenr (J/m2), is determined from the following equation (Ozel, 2013b):
410  
Qc Cel Qh
Cenr = PWF + Cg , (22)
COP 3.6∗106 Hu hs

where Qc , Qh , Cel and Cg are the yearly cooling transmission load (J/m2), the yearly heating
415 transmission load (J/m2), the cost of electricity ($/kWh) and the cost of natural gas ($/m3),
respectively. Also, Hu is the lower heating value of the fuel (J/ m3), hs is the efficiency of the
heating system and COP is the coefficient of performance of the air-conditioning system. The
total cost is obtained by adding the present value of the cost of energy consumption over the life-
time of the building and the cost of insulation as below:

420
Ct = Cenr + Ci = Cenr + Cins Lins . (23)

In the above equation Ci is the total cost of the insulated wall, Cins is the cost of the insulation
material per volume ($/m3) and Lins is the thickness of the insulation material (m). The value of
Lins that minimizes Ct is the optimum insulation thickness (Daouas, 2011).
425 At the optimum insulation thickness, the energy saving is calculated by subtracting the cost of
consumed energy of an un-insulated wall from that of an insulated one. Net energy saving per unit
area (over life time of the building, n years) is deduced from the difference between the present
value of energy saving and cost of insulation as (Ozel, 2013a):
 
430 ((Qc )un − Qc ) Cel ((Qh )un − Qh )
Se = PWF + Cg − Cins Lins , (24)
COP 3.6∗106 Hu h s

where (Qc )un and (Qh )un are the total heat gain and loss (J/m2·year) through an un-insulated
wall, respectively. The term (((Qc )un − Qc )/COP)(Cel /3.6∗106 ) + (((Qh )un − Qh )/Hu hs )Cg in
435 the above equation is annual energy saving (As ). The payback period, b, defined as the number
of years that energy saving from optimum insulation thickness compensates the cost of initial
investment of the insulation materials, is defined as (Daouas, 2011):

PWF(b)∗As = Ci, opt , (25)


440 ⎧    

⎪ d − i Ci
⎨ Ln 1 − i=d
1 + i As
b= (26)

⎪ Ci
⎩ (1 + i) i=d
As
445
Figure 2 shows the diagram of the financial flow during the life time of the building. Money
from energy saving is obtained at the end of each year during the lifetime, while the investment
cost of insulation must be paid at the beginning of the life time. Taking the inflation and interest
rates into account, the value of energy saving becomes less than what is obtained by multiplying AQ8
the life time and annul energy saving. In this study, infiltration and interest rates are taken as equal ¶
450 to 20% and 22%, respectively.
Advances in Building Energy Research 11

455

460

465 Figure 2. Diagram of financial flow during the life time of the building.

2.7. Numerical procedure


The one-dimensional transient heat conduction problem within a multi-layer wall, under the
dynamic boundary condition on the outside surface, is considered here. The transient heat con-
470
duction problem has been numerically solved using an implicit finite difference procedure for
a multi-layer wall. General finite difference equations derived by zeal and Pihtili (Optimum
location and distribution, 2007) are used to obtain the set of equations. The Gauss–Seidel iterative
method was used for solving the linear system of equations with the inner wall temperature as the
initial approximation (Burden & Faires, 2011). A FORTRAN code was developed to accelerate
475 the numerical computation. Inputs of the code include thermo-physical and geometrical properties
of each layer of the wall and time-dependent outdoor temperature, while the output is the propa-
gated temperature wave on the inside surface of the wall. The boundary condition was assumed to
be periodic on the outside; therefore until reaching the steady solution, the daily cycle of solutions
of equations with sol-air temperature on the outside was repeated (Ozel, 2012). Time lag, decre-
ment factor, heating and cooling transmission loads were obtained from the output of the Code.
480
The combined heat transfer coefficient on the inside of the envelope, hi , and combined heat trans-
fer coefficient on the outside were taken equal to 8 and 22 w/m2·K, respectively (Ozel, 2013b).

3. Results and discussions


485 This study focuses on the behaviour of the different wall structures located in Tehran, the capital
of Iran (which has the geographical coordinate of 35W 41′ 39′ ′ N and 51W 25′ 17′ ′ E). All vertical
wall orientations (south, west, east and north) and the horizontal wall were investigated. The
actual outdoor air temperature was obtained by averaging recorded meteorological date, over
the period of 2006–2012 (I. R. Iran’s Meterological Organization, 2013). The data pertain to a
490 representative day of each month during a year. Two-hourly outdoor air temperatures resulting
from averaging over the aforementioned period, one for July 21 and another for January 21,
are shown in Figure 3.
The clear sky model has been used to take the solar radiation on the wall surface into account.
For the representative day of each month during the year, solar radiation has been calculated in the
different wall orientations and the roof. Figure 4 shows hourly incident solar radiation in all wall
495 orientations and the roof in two representative days of the summer and winter seasons, July 21 and
12 H. Ramin et al.

500

505

Figure 3. Average outdoor air temperature on representative days of summer and winter (July 21 for
summer and January 21 for winter).
510

515

520

525

530

535

Figure 4. Hourly incident solar radiation on the horizontal and all vertical wall orientations on representa-
540 tive days of (a) July 21(summer season) and (b) January 21(winter season).
Advances in Building Energy Research 13

January 21, respectively. As it is expected, the maximum values of the incident radiation take
place at different times for each orientation. East- and west-facing walls received the same radi-
ation (Mavromatidis et al., 2012; Ozel, 2013b). It is seen from this figure that unlike July, the
highest value of solar radiation belongs to the south-oriented wall during January (Mavromatidis AQ9
et al., 2012). It is also notable that solar radiation on the north-facing wall in the July has two ¶
545
maximum values at two different times during the day.
Sol-air temperature is obtained by using Equation (4). Sol-air temperature, which is a combi-
nation of radiation and ambient air temperature, is shown in Figure 5 for the representative days of
January and July. As shown in Figure 5, Sol-air temperature is strongly affected by solar radiation.
The inside surface temperatures (and hence the inside surface heat flux), for both concrete and
550 aerated brick on un-insulated walls (in January 21 and July 21), are obtained in the form of a
numerical solution. It was seen that the orientation of the wall has a profound effect on the
inside temperature of the walls and resulting heat flux. Results show that concrete has a higher
maximum inside temperature in summer and lower minimum inside temperature in winter in com-
parison with aerated brick, which implies that aerated brick has a lower decrement factor (Asan,
555 2006). It is also seen that concrete has a higher amplitude of inside temperature compared with
aerated brick. In January, the south-facing wall has the highest inside temperature and therefore
lowest heating is needed, while the horizontal wall in July has the maximum temperature and
hence the maximum cooling load is required. All walls with different orientations and the hori-
zontal wall made from aerated brick reach their maximum temperatures later than the concrete
one, which implies that aerated brick has a higher time lag compared with concrete (Asan, 2006).
560
The daily transmission load is obtained through the integration of the inside heat flux over the
day and the yearly transmission load is the sum of the total daily transmission load over a year.
Figure 6 shows the daily transmission load of an un-insulated wall, for all orientations of the wall
and the roof. Both aerated brick and concrete block walls are shown in Figure 6. From the figure,
aerated brick wall has lower heating and cooling requirements compared with a concrete wall. It is
565 also seen that west- and east-facing walls have the same daily transmission load. The horizontal
wall has the maximum heating and cooling transmission loads during the year for both aerated
brick and concrete block walls. Finally, it is observed that daily heating and cooling loads (for
the horizontal wall) reach their maximum values in January and July, respectively.
Yearly transmission load per square metre of un-insulated walls are also presented in Table 3.
The horizontal wall (roof) has the highest yearly heating and cooling loads in both aerated brick
570
and concrete walls. When insulation is installed in the wall, transmission load decreases. In the
present study, two places where insulation can be placed are chosen, one between the cement
plaster and main wall structure, i.e. outside, and the other between the gypsum plaster and the
main wall structure (concrete or aerated brick in Figure 1), i.e. inside. After utilization of insula-
tion, the time lag increases and the decrement factor decreases.
575

Table 3. Yearly transmission load (MJ/m2) of an un-insulated wall of aerated brick and concrete block.
Yearly transmission load (concrete block) Yearly transmission load (aerated brick)
580 Heating Cooling Heating Cooling
Horizontal 13.18058 −19.2496 Horizontal 28.29076 −41.3215
South 11.05748 −11.682 South 23.7334 −25.0776
West 11.41614 −16.0959 West 24.50328 −34.552
East 11.41614 −16.0959 East 24.50324 −34.552
585
North 8.003887 −18.8242 North 17.17876 −40.4084
14 H. Ramin et al.

590

595

600

605

610

Figure 5. Hourly sol-air temperature on the horizontal and all vertical wall orientations on representative
days of (a) July (summer season) and (b) January (winter season).
615

Figure 7 shows the variation of time lag and decrement factor for all wall orientations and the
horizontal wall versus insulation thickness for aerated brick when EPS insulation was placed on AQ10

the inside; Figure 8 shows the details for the outside. From these two figures it is apparent that the AQ11
orientation of the wall has a strong effect on the time lag and decrement factor. Increasing the ¶
620
insulation thickness decreases the rate of change of time lag and decrement factor. It is also
notable that the east-facing wall has the highest time lag and lowest decrement factor among
all wall orientations and the horizontal wall. A comparison between Figures 7 and 8 reveals
that the place of insulation (inside or outside) does not change the overall trend discussed
above. However, when insulation is placed on the outside of the wall, the time lag will be
625 higher and decrement factor will be lower. The same trend has also been reported in a recent
study by Ozel. AQ12
Other combinations of wall material and insulation show the same trend as above when wall ¶
orientations are examined. In order to determine the effect of different materials and insulations on
time lag and decrement factor, Figures 9 and 10 are presented. Because of many orientations of
630 the wall, in these figures only the south-facing wall is displayed. AB is an abbreviation used for
Advances in Building Energy Research 15

635

640

645

Figure 6. Daily transmission load on the representative day of each month in the horizontal and all other
orientations of a wall made from aerated brick and concrete block without insulation (positive and negative
values represent heating and cooling transmission load, respectively)

650 aerated brick and C is used for concrete in the legend of the figures. As mentioned before, aerated
brick has a higher time lag and a lower decrement factor in comparison with the concrete block in
both positions of insulation. Using XPS instead of EPS in the wall increases the time lag and
decreases the decrement factor. The higher the insulation thickness is, the higher is the perceived
difference between the walls insulated with EPS and XPS. By increasing the thickness of insula-
655 tion, the effect of different wall materials and insulations on the decrement factor will be lessened.
As stated earlier, higher time lag and lower decrement factor are the results of placing the insula-
tion on the outside of the wall instead of on the inside.
Yearly heating and cooling transmission loads for different wall materials and insulations
placed on the inside and the outside of the wall are presented in Figures 11 and 12, respectively.
First of all, it is obvious that heating load is the dominant need for Tehran, which is consistent with
660

665

670

Figure 7. Time lag (left axis) and decrement factor (right axis) for all wall orientations and the horizontal
675 wall of aerated brick and EPS insulation placed on the inside of the wall.
16 H. Ramin et al.

680

685

Figure 8. Time lag (left axis) and decrement factor (right axis) for all wall orientations and the horizontal
wall of aerated brick and EPS insulation placed on the outside of the wall.
690

the values in the national regulation of buildings in Iran. It is seen that, the thinner the insulation
is, the higher the difference between transmission loads of different materials. As the insulation
thickness increases, the yearly transmission load will be relatively independent of materials and
come closer to the same value. The yearly transmission load is higher if the insulation is placed on
695 the outside of the wall. However, this difference is in the order of 1 (MJ/m2); therefore it is neg-
ligible (compared with the value of heating and cooling transmission load of the order of 100 (MJ/
m2)). It is also noteworthy that increasing the insulation thickness will decrease this difference and
yearly transmission load will come closer for two different locations of insulation.
A financial analysis was carried out to find the optimum insulation thickness. The necessary
700 parameters used in the financial analysis are presented in Table 4. Most of the data presented in
this table are taken from available data in a market and official websites of the Ministry of Energy
and Ministry of Economy in 2013.
Insulation cost increases linearly with increasing insulation thickness, while energy cost
decreases. There must be a point at which the total cost, i.e. insulation cost and energy cost,
705

710

715

Figure 9. Effect of different wall materials and insulations on the time lag (left axis) and decrement factor
(right axis) versus insulation thickness placed on the inside of the wall material (AB is an abbreviation for
720 aerated brick and C for concrete).
Advances in Building Energy Research 17

725

730

Figure 10. Effect of different wall materials and insulations on the time lag (left axis) and decrement factor
(right axis) versus insulation thickness placed on the outside of the wall material.
735

740

745

750

755

760

Figure 11. Variation of (a) yearly heating transmission load and (b) yearly cooling transmission load with
insulation thickness for different wall materials and insulations placed on the inside of the wall (AB is an
765 abbreviation for aerated brick and C for concrete).
18 H. Ramin et al.

770

775

780

785

790

Figure 12. Variation of (a) yearly heating transmission load and (b) yearly cooling transmission load versus
insulation thickness for different wall materials and insulations placed on the outside of the wall (AB is an
abbreviation for aerated brick and C for concrete).
795

Table 4. Parameters used in the financial analysis (most of these data have been
extracted from the market and official website of the ministry of energy and ministry
800 of economic affairs).
Parameter Value
Inflation rate, g 20%
Interest rate, i 22%
Life time, n 25
805 Cel 0.027 USD/kWh
Cg 0.04 USD/m3
Heating value, Hu 35.948 × 106 J/m3
Coefficient of performance, COP 3
Efficiency of combustion of natural gas, ηs 0.65
Cins(EPS) 66 USD/m3
810
Cins(XPS) 150 USD/m3
Advances in Building Energy Research 19

reaches its minimum. Figure 13 shows cost diagrams versus insulation thickness for the aerated
brick wall and EPS of the horizontal wall when insulation has been placed on the outside of the
wall. Other orientations of the wall have the same pattern; so to avoid duplication of content their
diagrams are not presented here. However, different optimum insulation thicknesses were
815
obtained for different wall orientations and the horizontal wall. It is seen that approximate
optimum thicknesses of 5, 3.68, 4.31, 4.31 and 4.06 cm were obtained for the horizontal wall
and the south-, west-, east- and north-facing walls, respectively. The south-facing wall has the
lowest optimum insulation thickness and the horizontal wall has the highest one.
The optimum insulation thickness, energy saving during the life time (per square metre), the
payback period for different wall orientations and the roof (for different combinations of wall
820 materials and insulations and also the positions of insulation) are tabulated in Table 5. It is
seen from this table that the aerated brick wall has a lower optimum insulation thickness com-
pared with the concrete one and therefore it has more energy saving and a lower payback
period. Investment cost return for the concrete wall in the reasonable years while aerated brick AQ13
needs more time. The horizontal wall has the highest value of optimal insulation thickness, ¶
825 highest energy saving and lowest payback period while the south-facing wall is positioned oppo-
site to the horizontal wall. It is also notable that the east- and west-facing walls have the same
value of optimum thickness, energy saving and payback period. Because of the higher investment
cost of XPS insulation compared with EPS (approximately three times) in the wall with the same
position and material, it is seen that optimal insulation thickness and energy saving decrease when
XPS is used while payback period increases. Concerning the position of the insulation, as shown
830
in Table 5, it is realized that for EPS, optimal insulation thickness and payback period remain con-
stant or increase when insulation is placed on the outside of the wall, while energy saving
decreases. It is seen that for XPS insulation when it is placed on the outside of the wall,
optimal insulation thickness and payback period decrease while energy saving increases. It
should be noted that the aforementioned difference in the position of insulation is more noticeable
835 for the payback period. Higher payback period and small energy saving are obtained for aerated
brick walls, while reasonable payback period is reported for concrete walls and energy saving as
well.

840

845

850

Figure 13. Variation of total cost, energy cost and insulation cost versus the insulation thickness for aerated
brick and EPS insulation (placed on the outside of the wall) for the horizontal wall (south-, west-, east- and
855 north-oriented walls have the same pattern, so they are not presented).
900

895

890

885

880

875

870

865

860

20
H. Ramin et al.
Table 5. Optimum insulation thickness, payback period and energy saving per each square metre of insulation for all wall orientations and a combination of wall
materials and insulations and also position of the insulation (AB is an abbreviation for aerated brick and C for concrete).
Description Inside Outside
Optimal thickness (cm) Wall orientation AB C AB C
EPS XPS EPS XPS EPS XPS EPS XPS
Horizontal 4.97 2.43 6.89 3.78 5.0 2.26 7.0 3.77
South 3.68 1.68 5.62 3.06 3.68 1.42 5.66 3.06
West 4.30 2.08 6.24 3.38 4.31 1.87 6.30 3.37
East 4.30 2.08 6.24 3.38 4.31 1.87 6.30 3.37
North 4.05 1.83 6.0 3.25 4.06 1.68 6.04 3.24
Energy saving (USD/m2) Horizontal 4.43 2.04 18.4 15.40 4.34 2.39 18.27 15.45
South 2.38 0.70 12.1 9.81 2.34 0.95 12.04 9.81
West 3.30 1.34 14.99 12.31 3.23 1.59 14.90 12.35
East 3.30 2.08 14.99 12.31 3.23 1.59 14.90 12.35
North 2.90 1.16 13.72 11.15 2.84 1.27 13.64 11.20
Payback period (years) Horizontal 9.40 14.65 4.2 5.76 9.56 13.18 4.28 5.73
South 10.78 18.33 5.0 6.90 11.4 15.66 5.05 6.90
West 10.28 16.30 4.57 6.27 10.42 14.56 4.72 6.24
East 10.28 16.30 4.57 6.27 10.42 14.56 4.72 6.24
North 10.70 16.23 4.77 6.56 10.86 13.28 4.82 6.51
Advances in Building Energy Research 21

The energy cost in Iran is much lower than the average global price. Followed by the Iran
subsidy reform law mentioned in introduction section, it is predictable that the energy cost in
Iran will be increased in the next few years; therefore, the optimum insulation thickness and
energy saving will increase. For example, if electricity and natural gas cost double simultaneously
905
while the other parameters remain constant, optimal insulation thickness, energy saving and
payback years for the concrete wall with EPS insulation on the inside of the wall will become
10.39 cm, 42.28 USD/m2 and 2.9 years, respectively, which is remarkably better analogously
in comparison with the present values of 6.89 cm, 18.4 USD/m2 and 4.2 years.

910

915

920

925

930

935

940

Figure 14. Effect of (a) inflation rate (b) discount rate (c) electricity price (d) natural gas price and (e) effi-
ciency of combustion of natural gas on optimum insulation thickness and payback period for a wall made of
945 aerated brick and EPS as insulation material on the inside of the wall.
22 H. Ramin et al.

A sensitive analysis has been conducted in order to investigate the effect of different economic
and also thermal parameters on the optimum insulation thickness and payback period. The
obtained results for a south-facing wall made of aerated brick and EPS as insulation material
on the inside of the wall are illustrated in Figure 14 (other combinations show the same trend,
950
so it is not presented). Figure 14(a)–(e) shows the variation of insulation thickness and
payback period as a function of inflation rate, discount rate, electricity price, natural gas price
and efficiency of combustion of natural gas, respectively (all other parameters are set to be con-
stant and equal to the values presented in Table 4). Rising inflation rates increase the optimum
insulation thickness while they decrease the payback period. Both payback period and
optimum insulation thickness decrease as the discount rate increases. The effect of energy
955 price on the payback period and optimum insulation thickness is also investigated. Investigations
reveal that increasing electricity and natural gas price will lessen payback period and increase
optimum insulation thickness. It is also seen from the figure that an increase in the efficiency
of combustion of natural gas tends to decrease optimum insulation thickness while it increases
the payback period; therefore improving the combustion efficiency has a profound effect on
960 the optimum insulation thickness and payback period as well. According to the subsidy law in
Iran, the energy price will increase inevitably; therefore application of insulation material will
be far more reasonable in the future.
Total heating and cooling transmission load are obtained for the optimum insulation thickness
for different combinations of walls and insulations and also for all wall orientations and the hori-
zontal wall. Figure 15 shows the percentage ratio of yearly transmission loads (heating and
965
cooling) for the wall with an optimum insulation thickness and the un-insulated wall. Firstly, per-
centage ratios of cooling and heating transmission load for the same combination of wall and insu-
lation have approximately the same value. It is also realized that the south-facing wall has the
highest percentage ratio of yearly heating and cooling transmission load while the horizontal

970

975

980

985

Figure 15. Yearly cooling and heating transmission load for all orientations of the wall and the horizontal
wall in combination of insulation and material of wall and also position of insulation (AB is an abbreviation
990 for aerated brick and C for concrete).
Advances in Building Energy Research 23

wall has the lowest percentage ratio among all wall orientations. This implies that utilization of
insulation on the south-facing wall has a smaller effect on the yearly transmission load compared
with the other wall orientations. Moreover, concrete walls (when optimum insulation thickness is
applied) reduce the transmission load from 70% to 82% in comparison with the un-insulated wall,
995
while this value lies in the range of 42–58% for the aerated brick wall.

4. Conclusion
In this paper optimal insulation thicknesses, based on heating and cooling seasons, for conven-
tional walls in Tehran are obtained. It was found that the concrete wall has a higher optimum insu-
1000
lation thickness compared with aerated brick. Higher time lag and lower decrement factor have
been achieved when insulation is placed on the outside of the wall. Also the location of insulation
does not change the yearly transmission load considerably. Among all wall orientations, it was
found that the horizontal wall has the highest optimum insulation thickness and the south-
facing wall has the lowest optimum thickness. The results showed that using insulation in
1005 walls made of concrete and aerated brick reduces the yearly transmission load up to 82% and
58%, respectively. Due to the high investment cost of XPS, although obtaining higher payback
period, smaller insulation thickness and energy saving have been achieved. Finally, the results
showed that increase in the energy cost in Iran will make utilization of insulation more reasonable.

1010
Disclosure statement
No potential conflict of interest was reported by the authors.

References
1015 19th national regulaion for building. (2012). The office of national regulations for building. http://inbr.ir/
Al-Khawaja, M. J. (2004). Determination and selecting the optimum thickness of insulation for buildings in
hot countries by accounting for solar radiation. Applied Thermal Engineering, 24, 2601–2610.
Al-Sanea, S. A., Zedan, M. F., & Al-Ajlan, S. (2005). Effect of electricity tariff on the optimum insulation-
thickness in building walls as determined by a dynamic heat-transfer model. Applied Energy, 82, 313–
330.
1020 Al-Sanea, S. A., Zedan, M. F., & Al-Hussain, S. N. (2012). Effect of thermal mass on performance of insu-
lated building walls and the concept of energy savings potential. Applied Energy, 89, 430–442.
Asan, H. (2000). Investigation of wall’s optimum insulation position from maximum time lag and minimum
decrement factor point of view. Energy and Buildings, 32, 197–203.
Asan, H. (2006). Numerical computation of time lags & decrement factors for different building materials.
Building and Environment, 41, 615–620.
Asan, H., & Sancaktar, Y. S. (1998). Effects of wall’s thermophysical properties on time lag and decrement
1025 factor. Energy and Buildings, 28(2), 159–166.
Balaras, C. A., Droutsa, K., Argiriou, A. A., & Asimakopoulos, D. N. (2000). Potential for energy conserva-
tion in apartment buildings. Energy and Buildings, 31, 143–154.
Balaras, C. A., Droutsa, K., Dascalaki, E., Kontoyiannidis, S. (2005). Heating energy consumption and
resulting environmental impact of European apartment buildings. Energy and Buildings, 37, 429–442.
Bolattürk, A. (2008). Optimum insulation thicknesses for building walls with respect to cooling and heating
1030 degree-hours in the warmest zone of Turkey. Building and Environment, 43, 1055–1064.
Burden, R. L., & Faires, J. D. (2011). Numerical analysis. Boston, MA: Brooks/Cole, Cengage Learning.
Daouas, N. (2011). A study on optimum insulation thickness in walls and energy savings in Tunisian build-
ings based on analytical calculation of cooling and heating transmission loads. Applied Energy, 88, 156–
164.
Daouas, N., Hassen, Z., & Ben Aissia, H. (2010). Analytical periodic solution for the study of thermal per-
formance and optimum insulation thickness of building walls in Tunisia. Applied Thermal Engineering,
1035 30, 319–326.
24 H. Ramin et al.

Duffie, J. A., & Beckman, W. A. (1991). Solar engineering of thermal processes (2nd ed.). New York, NY:
John Wiley & Sons.
Hottel, H. C. (1976). A simple model for estimating the transmittance of direct solar radiation through clear
atmospheres. Solar Energy, 18(2), 129–134.
I. R. Iran’s Meterological Organization. (2013). www.weather.ir AQ14
1040 Kaynakli, O. (2008). A study on residential heating energy requirement and optimum insulation thickness. ¶
Renewable Energy, 33, 1164–1172.
Kaynakli, O. (2012). A review of the economical and optimum thermal insulation thickness for building
applications. Renewable and Sustainable Energy Reviews, 16, 415–425.
Khudhair, A. M. & Mohammed, M. F. (2004). A review on energy conservation in building applications with
thermal storage by latent heat using phase change materials. Energy Conversion and Management, 45,
263–275.
1045 Liu, B. Y. H., & Jordan, R. C. (1960). The interrelationship and characteristic distribution of direct, diffuse
and total solar radiation. Solar Energy, 4(3), 1–19. AQ15
Mavromatidis, L. E., EL Mankibi, M., Michel, P., & Santamouris, M. (2012). Numerical estimation of time ¶
lags and decrement factors for wall complexes including multilayer thermal insulation, in two different
climatic zones. Applied Energy, 92, 480–491.
Ministry of Energy. (2013). http://www.moe.gov.ir/ AQ16
1050 Najafian, M., & Bahrami, A. R. (2012). The effect of orientation on optimum insulation position in the wall ¶
of a building with natural ventilation in hot and dry climate. International Journal of Advanced Design
and Manufacturing Technology, 5(3). AQ17
Ozel, M. (2012). Cost analysis for optimum thicknesses and environmental impacts of different insulation ¶
materials. Energy and Buildings, 49, 552–559.
Ozel, M. (2013a). Determination of optimum insulation thickness based on cooling transmission load for
building walls in a hot climate. Energy Conversion and Management, 66, 106–114.
1055 Ozel, M. (2013b). Thermal, economical and environmental analysis of insulated building walls in a cold
climate. Energy Conversion and Management, 76, 674–684.
Ozel, M. Effect of Insulation Location on Dynamic Heat-Transfer Characteristics of Building External Walls
and Optimization of Insulation Thickness. ENERGY BUILDING, http://dx.doi.org/10.1016/j.enbuild.
2013.11.015. AQ18
Ozel, M., & Pihtili, K. (2007). Investigation of the most suitable location of insulation applying on building ¶
1060 roof from maximum load levelling point of view. Building and Environment, 42, 2360–2368.
Ozel, M., & Pihtili, K. (2007). Optimum location and distribution of insulation layers on building walls with
various orientations. Build Environment, 42(8), 3051–3059. AQ19
Statistical Centre of Iran. (2012). http://www.amar.org.ir ¶
AQ20
Threlkeld, J. L. (1998). Thermal environmental engineering. Englewood Cliffs, NJ: Prentice-Hall. ¶
Ucar, A., & Balo, F. (2009). Effect of fuel type on the optimum thickness of selected insulation materials for
the four different climatic regions of Turkey. Applied Energy, 86, 730–736.
1065 Wang, Y. Huang, Z., & Heng, L. (2007). Cost-effectiveness assessment of insulated exterior walls of resi-
dential buildings in cold climate. International Journal of Project Management, 25, 143–149.
Yu, J., Tian, L., Yang, C., Xu, X., & Wang, J. (2011). Optimum insulation thickness of residential roof with
respect to solar-air degree-hours in hot summer and cold winter zone of china. Energy and Buildings, 43,
2304–2313.
Yu, J., Yang, C., Tian, L., & Liao, D. (2009). A study on optimum insulation thicknesses of external walls in
1070 hot summer and cold winter zone of China. Applied Energy, 86, 2520–2529.

1075

1080

View publication stats

You might also like