Sei sulla pagina 1di 11

Applied Energy 112 (2013) 235–245

Contents lists available at SciVerse ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Configuring wall layers for improved insulation performance


Danielle E.M. Bond, William W. Clark, Mark Kimber ⇑
Department of Mechanical Engineering and Materials Science, University of Pittsburgh, 206 Benedum Hall, 3700 O’Hara Street, Pittsburgh, PA 15261, United States

h i g h l i g h t s

 Position and distribution of materials are analyzed for insulating performance.


 Insulation for first and last layers improves insulating performance.
 Equal distribution of insulation and thermal mass improves insulating performance.
 Optimum configuration exists for wall with fixed resistance and capacitance.

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

Article history: External walls provide an important barrier between occupied building spaces and variable ambient con-
Received 9 March 2013 ditions. Building insulation is often optimized for static performance based on its R-Value rating, but a
Received in revised form 6 June 2013 growing body of literature has begun to assess the dynamic thermal performance of multi-layer walls.
Accepted 12 June 2013
The focus of this work is to provide fundamental insight into configuring wall layers for improved insu-
Available online 6 July 2013
lating performance. Thirty-three different walls are evaluated based on four primary configurations with
fixed volumes of insulation and thermal mass. Only the layer distribution is varied. Because the total vol-
Keywords:
ume of each material is fixed, the overall thermal resistance and capacitance are equivalent for all con-
Insulation
Thermal mass
figurations studied. An electrical analogy is used to model the one dimensional heat conduction through
Energy savings the wall and determine the magnitude ratio and phase of the frequency response evaluated at the fre-
Optimization quency corresponding to a period of 1 day. The two temperatures used to quantify the wall system trans-
Buildings fer function are the inside and outside surfaces of the wall (i.e., two surfaces exposed to the indoor and
outdoor environments, respectively). The best insulating performance is achieved when insulation layers
are positioned as close as possible to the inside and outside layers of the wall (i.e., near the indoor and
outdoor environments). In addition, optimal results occur when both the insulation and thermal mass
are distributed evenly throughout the wall. Using this optimized configuration, each material is then
divided into an increasing number of thinner layers. Results indicate that an optimum number of layers
exist such that the magnitude ratio is maximized.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction walls for improved thermal insulating performance as quantified


by decrement factor and time lag. The decrement factor is an indi-
Building insulation is a fundamental component in managing cation of steady state heat transfer and defined as the ratio of in-
building energy consumption. Buildings use roughly 40% of the side to outside temperature fluctuations. The time lag indicates
United States total energy consumption, more than industry and how long exterior temperature changes take to affect interior con-
transportation [1]. Within the US residential and commercial ditions. Determining how best to assemble a multi-layer wall to re-
buildings, heating and cooling represent 42% and 29%, respectively, duce decrement factor and increase time lag can significantly
of total energy consumed [1]. Reducing building energy consump- enhance the energy efficiency and reduce the overall energy foot-
tion helps reduce building operating costs as well as emissions and print for multiple types of buildings.
other related impacts. One method towards mitigating building Previous studies optimize the quantity of insulation used in a
energy used for heating and cooling is through the design of exter- given wall based on specific climatic and economic conditions. Sev-
nal walls. The primary objective of this work is to evaluate the eral studies evaluate the effect of insulation position within an
assembly of insulation and thermal mass materials in external external wall. In the current work, in order to remove ambiguity
and consistently define positions within a wall, we let ‘‘inside’’
⇑ Corresponding author. Tel.: +1 (412) 624 8111; fax: +1 (412) 624 4846. and ‘‘outside’’ refer to the indoor and outdoor environments, while
E-mail address: mlk53@pitt.edu (M. Kimber).
‘‘middle’’ refers to those layers sandwiched between the inside and

0306-2619/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.apenergy.2013.06.024
236 D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245

outside surfaces. Many studies compare performance when insula- insulation towards the outside. The highest energy demand results
tion is placed near the inside or outside wall surface. Others from a light weight wall configuration with no significant thermal
experiment with various multi-layer configurations. Some studies mass layers. The lack of thermal mass in this poorly performing
consider both insulation optimization and investigation of the ac- assembly illustrates its importance.
tual configuration (i.e., order and number of layers). Unlike other studies that consider different combinations and
For example, Al-Sanea and Zedan [2] compare two wall config- configurations of insulation and thermal mass layers in a wall,
urations using single layers of insulation and thermal mass. The Asan and Sancaktar [8] evaluate the thermal performance of vari-
decrement factor is analyzed for the two wall configurations and ous building materials independently. They develop a numerical
for different response time periods, including an initial transient solution to the one dimensional heat equation and investigate
or instantaneous response and a steady periodic response. Between the effects of wall thickness and material properties on time lag
the two configurations, insulation towards the inside yields a low- and decrement factor. Results demonstrate common conclusions,
er instantaneous load, but insulation towards the outside has a such as the time lag increases and decrement factor decreases with
much longer time lag (almost 10 times). The results are somewhat increasing heat capacity; time lag increases and decrement de-
conflicting, since insulation on the inside is better for reducing the creases with increasing thickness; and, low thermal conductivity
load while insulation at the outside is better for increasing time and high heat capacity yield longer time lags. In a subsequent
lag, both features of effective insulating walls. In a follow up study study, Asan [9] evaluates heat conduction in 26 building materials.
[3], the two configurations are studied while varying the amount The time lag and decrement factor for each building material are
and position of thermal mass under steady periodic conditions. calculated for 8 variations of a single layer wall thickness. For all
The total wall resistance is maintained by varying the mass and the materials, the decrement factor decreases with increased wall
insulation layer thicknesses. Results show that increasing the thickness, and for small thicknesses (roughly less than 0.050 m)
amount of thermal mass in the wall leads to lower transmission the decrement factor is nearly constant. The decrement factor
loads during swing seasons when temperatures are not extreme. tends to zero for wall thickness greater than 1 m, and for wall
They find that for any set amount of thermal mass, positioning thickness less than 1 cm, all materials exhibit zero time lag. In
the mass towards the inside surface and positioning the insulation the second part of the study, two multi-layer walls are analyzed:
near the outside surface yields better overall thermal performance. one with three layers and no insulation and one with five layers
Kossecka and Kosny [4] evaluate six wall configurations using a including insulation. For the same thickness wall, the wall with
simplified building model and a one story residential building insulation has a longer time lag and lower decrement factor. In
model simulated with six different US climates. The thermal per- these two works, Asan analyzes a large set of walls and provides
formance of the walls is evaluated in terms of annual building generalized conclusions relating material properties to thermal
heating and cooling demands and total annual energy demands performance, but suggests that future work will aim to optimize
as calculated by the building modeling program. Each wall has insulation thickness and position to achieve larger time lags and
internal layers of concrete and insulation with either one or two smaller decrement factors.
layers of each such that all of the walls have the same total resis- In a more recent study, Asan [10] evaluates six wall configura-
tance and capacitance values. The study agrees with previous find- tions for time lag and decrement factor based on a Crank Nicolson
ings that walls with concrete layers at the inside and insulation numerical solution to the one-dimensional heat equation with
layers at the outside show the best building energy savings. In periodic convection boundary conditions. Fewer than six configu-
terms of dynamic performance criteria, part of the study indicates rations are actually studied due to overlap between the different
that a concrete/insulation/concrete configuration has a lower dec- configuration ‘families’. Two different thermal mass and insulation
rement factor than a configuration of insulation/concrete/insula- materials are selected each, so that a total of four thermal mass-
tion, which has a longer time lag. The authors describe an even insulation combinations are studied. The minimum decrement fac-
distribution of insulation on either side of the mass, but they do tor is achieved with half the insulation at the inner and outer walls,
not consider other wall configurations with unevenly distributed while the maximum time lag is achieved with two layers of insu-
layers. Furthermore, because these results seem to conflict, the lation located within the wall, equidistant from the inner and outer
need for clear and unambiguous guidelines for assembling multi- surfaces. While practical guidelines for renovating building walls
layer walls remains. for improved insulating performance are provided from the results,
Two other studies consider different mass/insulation configura- the number of layers studied was limited and different weighting
tions. Bojić and Loveday [5] find an alternating insulation/mass/ configurations were not considered.
insulation configuration is better under intermittent heating con- Another such study that addresses insulation thickness and po-
ditions and an alternating mass/insulation/mass configuration is sition is Al-Sanea and Zedan [11]. They evaluate the number and
better under intermittent cooling conditions. In a separate study distribution of insulation layers in a wall with fixed amounts of
from Bojic et al. [6], six wall configurations are analyzed within a insulation and thermal mass. The wall width is fixed and the total
full building model of two high-rise residences comprised of vari- amount of insulation is optimized for cost and energy performance.
ous flats facing different directions on each floor. Results show that Different wall configurations are arranged and thermal perfor-
thicker insulation reduces the cooling load and insulation at the in- mance evaluated. The study concludes that the best configurations
side position yields the lowest annual cooling load. Cooling de- for insulating performance have either two layers of insulation di-
mand results are varied, however. It is lower for insulation at the vided between the outside and middle layers or three layers of insu-
outside position for some flats and at the inside positions for other lation split evenly among the outside, middle, and inside layers. The
flats. The authors suggest this is due to the different orientation of study only considers evenly distributing insulation layers, and the
the flats within the building. The mixed results suggest that whole analysis is limited to using one, two, and three layers of insulation.
building simulation may not be the best tool for analyzing thermal The effects of increasing the number of insulation layers further is
dynamic wall performance. not explored nor is the concept of using unevenly distributed layers.
In a slightly different study, Aste et al. [7] analyzes six represen- In a supporting study, Mavromatidis et al. [12] investigate the ef-
tative wall configurations as a Southern facing wall within a whole fects of wall composition and orientation on time lag and decrement
building analysis. All of the walls have the same R-value but differ- factor. The authors conclude that their multi-layer insulation (MTI)
ent values of thermal capacity and mass per unit area. The wall has the best overall thermal performance and increases time lag
system with the lowest energy demand is a two layer wall with compared to insulation alone, citing similar results from Al-Sanea.
D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245 237

In a recent review of building insulation work, Kaynakli [13] re-


ports that most studies determine optimum insulation thickness
from heating and cooling loads, material and energy costs, and
other economic factors for a given climate. Most studies also eval-
uate insulation thickness using static conditions. A few, however,
consider the dynamic performance of multi-layer walls. The stud-
ies described here indicate that multi-layer walls yield better over-
all thermal performance than single or double layer walls [4,11,12].
Two separate studies [4,11] find insulation towards the inside and
outside to be best. Several find that insulation towards the outside
(with thermal mass towards the inside) provides better thermal
performance than insulation towards the inside [2–4,7,14,15],
and most agree that, generally, more thermal mass yields lower
peak loads, a lower decrement factor, and a larger time lag
[3,7,8]. Other studies cited in Aste et al. [7] that consider insulation
position and use different analyses and performance criteria, such
as cooling loads, time lag, and decrement factor, also find that insu-
lation at the outside layer yields better thermal performance.
While the studies mentioned above provide some meaningful
results about assembling insulation and thermal mass layers in a
wall, a limited number of configurations are evaluated and the
study of multi-layered insulation walls is limited. Since the dy-
namic effect of increasing layer divisions and re-positioning insula-
tion layers within an assembly is not fully characterized, the study
presented in this paper focuses on the dynamic thermal response
of multi-layered walls. The present work uses frequency analysis
to evaluate various wall assemblies. In this analysis, the walls are
decoupled from any specific climatic data or building orientation,
and results provide insight into the general dynamic performance
of each wall system as an assembly of multiple insulation and ther-
mal mass layers.
The primary objective of this work is to evaluate the assembly
of insulation material and thermal mass in external walls for im-
proved thermal insulating performance. Given a fixed amount of
insulating material and thermal mass, determining how best to di-
vide these materials and assemble multiple layers to reduce or
minimize the decrement factor and increase or maximize the time
Fig. 1. Assembling wall layers for improved thermal performance.
lag may help reduce material use and energy consumption. In this
study, the primary focus is decreasing the decrement factor with a
secondary emphasis on increasing the time lag. A total of 33 wall thermal mass will be used for each wall design, the primary vari-
configurations are evaluated to assess (1) the effect of configura- ables left to define are the number of layers, the thickness of the
tion (2) the effect of weighting insulation and/or mass towards layers, and the order in which the layers are arranged. Two exam-
each side of the wall compared to distributing layers evenly, and ples of potential wall designs are also shown in Fig. 1. Any partic-
(3) the effect of increasing the overall number of insulation and ular layer exhibits thermal resistance and capacitance behavior,
thermal mass layers. which is influenced by the material properties, namely thermal
From the frequency analysis of the multi-layered wall, the re- conductivity (k), density (q), and heat capacity (Cp), and the thick-
sponse is evaluated at a frequency whose period is 1 day ness of each layer. However, since the same amount of each mate-
(xday = 7.27  105 rad/s). Although many different methods (e.g., rial is used for any wall design, the total resistance and capacitance
finite volume or finite element modeling) could be employed to remain constant, regardless of how it is divided or arranged.
compute the metrics of interest, analyzing the wall from its fre- It is important to note that the dynamic thermal response may
quency response provides certain benefits that are worth noting. vary significantly based on the configuration of layers. While
First is the speed at which calculations can be performed. Multiple improving the thermal insulating performance of external walls
wall configurations can be easily assessed in a matter of a few sec- is desirable for many climates, it is recognized that increasing insu-
onds even with underwhelming computational resources. Another lation is not always the single most effective strategy towards
major benefit is the fact that data can be extracted for additional reducing building energy consumption. In some instances, in-
driving frequencies without the need to run more simulations. Ide- creased insulation actually results in greater energy consumption
ally, the weather patterns fall in line with a frequency whose per- [14]. Both climatic and economic factors will help determine the
iod is 1 day, but for many scenarios (e.g., approaching storms), best overall amount of insulation and thermal mass to use in an
additional frequencies might also be of interest. external wall, and knowing how to arrange the layers may help
The problem of concern is illustrated in Fig. 1. Suppose a wall is achieve the same insulating performance while using less material.
to be comprised of fixed amounts of insulation (e.g., fiberglass) and Reducing material consumption typically has positive economic
thermal mass (e.g., concrete). For a particular wall design, the left returns and sustainable impacts. In the remainder of this paper,
hand side is subject to an oscillatory ambient temperature (i.e., we first present the development of the dynamic models for mul-
outdoor environment). The corresponding internal surface temper- ti-layered walls. This accounts for the thermal resistance and
ature (filtered by the wall) is seen on the right hand side (i.e., inside capacitance of each individual layer, and includes validating key
surface of a room). Since a fixed amount of insulating material and configurations with finite element models. Next, we analyze simu-
238 D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245

lation results from different fundamental wall configurations A block diagram in Fig. 2c shows the system frequency response
which ultimately provides insight into optimizing designs for mul- function G(jx) with an ambient sinusoidal input and the output
ti-layered walls. as a modified amplitude and peak time shifted by /.

2. RC model development and frequency analysis 2.1. RC model development

The studies referenced in this work use either simplified analyt- Based on the RC wall system model described above, the system
ical calculations, numerical analysis, or whole building simulation transfer function is determined for frequency analysis. Fig. 3a shows
to evaluate the performance of different wall configurations. In this the RC circuit model for a wall with N layers, where Tamb is the input
study, each wall layer is modeled as a T-shaped RC circuit with the and Ts,rm is the system output. The outside and inside surface heat
resistance divided evenly on each side of the capacitance, as shown transfer coefficients (hamb and hrm) are approximated using the ASH-
in Fig. 2a. This analysis is in accordance with multiple studies [16– RAE [19] procedure for estimating surface conductance. The inner
18] that determine the time lag and decrement factor for heat con- and outer layer material is cement plaster with a summer time wind
duction in external walls using an electrical analogy. The RC sys- speed of 7.5 mp h and building material emissivity of 0.90 used in
tem is formed by connecting each layer in series with resistance the calculation to obtain hamb = 16.74 W/m2 K and hrm = 8.23 W/
added at each end to capture the convective heat transfer at both m2 K. However, only the inside convection coefficient is applicable
interior (Rh,rm) and exterior (Rh,amb) sides of the wall. The wall tem- to this model, because the decrement factor and time lag relate in-
perature at the outside surface is denoted as Ts,amb, while the wall side and outside surface temperatures. The outside convection coef-
temperature at the inside surface of the wall is denoted Ts,rm. The ficient affects outside surface temperature, but the outside surface
inside room and surface temperature will oscillate as a result of temperature is considered the input to the wall system modeled
the outside temperature fluctuation. The outside temperature is here. In other words, the transfer functions compared for various
modeled as a sinusoidal input with a 24 h period. The wall acts wall configurations are from Ts,amb to Ts,rm, and do not include Rh,amb,
as a filter to reduce the amplitude of the transmitted signal and which is there simply to complete the circuit analogy that includes
to shift the signal in time, as in Fig. 2b. The amplitude ratio of an ideal ambient temperature source. As described previously, this
the outside surface temperature Ts,amb to the inside surface temper- model is decoupled from any specific weather signal.
ature Ts,rm is known as the decrement factor (df), and defined The resistance (R) and capacitance (C) of each layer are based on
equivalently to the magnitude ratio (M): three material properties: thermal conductivity (k), density (q),
and heat capacity (Cp). They also depend on the layer thickness
Amplitude of T s;rm Arm (L) according to the following equations:
df ¼ ¼ ¼j GðjxÞ j¼ M ð1Þ
Amplitude of T s;amb Aamb
L
The magnitude ratio (magnitude of the transfer function G(jx), a R¼ ð3Þ
k
term more familiar with traditional frequency response), is evalu-
ated at x = xday. The phase angle (/) can likewise be evaluated at C ¼ Cp  q  L ð4Þ
x = xday in order to find the time lag (tlag) in hours according to
the following equation: The wall system transfer function is determined using an imped-
ance method and voltage divider calculation. The impedance value
24 h for the resistance of layer ‘‘i’’ is:
t lag ¼ /ðxday Þ ¼ ð2Þ
2p
1
Z Ri ¼ Ri ð5Þ
2
where the division by two is included to account for the fact that
the thickness of each layer is divided into two as previously dis-
cussed. The corresponding impedance for the capacitance of layer
‘‘i’’ is:
1
Z Ci ¼ ð6Þ
sC i
The RC circuit can be simplified by recognizing that the transfer
function of primary interest, G(jx) = Ts,rm/Ts,amb, can be expressed
as the product of intermediate transfer functions, shown in the fol-
lowing equation:
T s;rm T n T n1 T 1 T s;rm
¼    ð7Þ
T s;amb T s;amb T n T2 T1
To determine the intermediate transfer functions, the circuit is col-
lapsed from right to left by combining individual impedances in ser-
ies or in parallel. The transfer function relating each layer is
calculated as the circuit is reduced. The transfer function for the
innermost layer, including convective resistance (Rh,rm), is calcu-
lated first. In Eq. (8), T1 represents the temperature in the first layer,
measured as the voltage across the first capacitor C1.
T s;rm Rh;rm
¼ ð8Þ
T1 Rh;rm þ Z R1
Fig. 2. RC system analysis: (a) RC network diagram, (b) ambient temperature (left)
and inside wall temperature after filtering (right), and (c) frequency response The circuit can then be reduced as shown in Fig. 3b using an equiv-
representation of wall. alent impedance Z eq1 defined in Eq. (9) as:
D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245 239

Fig. 3. N-layer RC network diagram: (a) full circuit, (b) collapsed circuit to first equivalent impedance, and (c) collapsed circuit to second equivalent impedance.

 1
1 1 Then taking the product of each intermediate transfer function, as
Z eq1 ¼ þ ð9Þ shown in Eq. (7), gives the final relationship of interest G(jx) = Ts,-
Z R1 þ Rh;rm Z C 1
rm/Ts,amb, which is easily evaluated for any value of x. The magni-
The transfer function relating T1 to T2 is also calculated as a voltage tude ratio and phase at a period of 24 h (xday = 7.27e5 rad/s) are
divider: extracted from the frequency analysis for each wall assembly sys-
T1 Z eq1 tem. This procedure is used to develop transfer functions for each
¼ ð10Þ wall assembly considered, thereby enabling a quick and simple fre-
T 2 Z eq1 þ Z R1 þ Z R2
quency analysis procedure.
Defining a new nested impedance, Z eq2 , the circuit can be further
collapsed as shown in Fig. 3c, where equivalent impedance Z eq2 is
calculated in a similar manner. The general expression for equiva- 2.2. RC model validation
lent impedance for the ith layer Z eqi is simply:
To validate the model used in this study, the seven wall config-
 1
1 1 urations from Al-Sanea and Zedan [11] are modeled using this RC
Z eqi ¼ þ ð11Þ
Z eqi1 þ Z Ri1 þ Z Ri Z C i circuit methodology and results are compared with those pub-
lished. The seven wall configurations are shown in Fig. 11 in the
Subsequent transfer functions relating T2 to T3, T3 to T4, and eventu- Appendix. Each wall contains 200 mm concrete and 78 mm insula-
ally Ti1 to Ti, are calculated as: tion distributed in varying configurations between two 15 mm lay-
T i1 Z eqi1 ers of cement plaster on both inner and outer surfaces. The same
¼ ð12Þ overall wall thickness (308 mm) and total volume of each material
Ti Z eqi1 þ ðZ Ri1 þ Z Ri Þ
is fixed, so the total resistance and capacitance values do not

Table 1
Model validation results.

Wall system Decrement factora (%) Magnitude ratio (%) Time laga (h) Phase (h)
Wall 1a 1.35 1.51 6.13 5.92
Wall 1b 1.34 1.55 7.33 7.53
Wall 1c 0.74 0.90 6.71 6.10
Wall 2a 0.42 0.51 9.33 8.97
Wall 2b 0.24 0.33 8.19 6.76
Wall 2c 0.26 0.31 10.44 10.07
Wall 3 0.13 0.16 12.13 11.27
a
Decrement factor and time lag results from Al-Sanea et al. [15] compared with magnitude ratio and phase results from frequency analysis. Al-Sanea wall configurations
are reproduced for visual reference in the Appendix Fig. 11.
240 D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245

side, middle, and outside, with concrete divided evenly between


the three insulation layers. Because the decrement factor and time
lag reported by Al-Sanea and Zedan [11] are equivalent to the mag-
nitude ratio and phase lag produced in the frequency analysis, re-
sults may be compared directly. This data is provided in Table 1 for
the decrement factor and magnitude ratio as a percentage, along
with the time lag and phase lag, which is converted from radians
to hours (considering a 24-h period).
Comparing the two sets of results, the numerical solution to the
one dimensional heat equation is expected to provide a more accu-
rate solution than the lumped parameter RC model. Any lumped
parameter model necessarily sacrifices some accuracy to enable
or expedite calculation. Here, the magnitude ratio from the RC
model differs from the decrement factor results by 21% on average,
while the time lag results differ by 7% on average. A lower percent
error would be desirable, but further analysis reveals that the RC
model consistently predicts performance trends. In addition, the
ease and quickness of implementing numerous wall configurations
using the RC model makes it ideal for analyzing general trends and
investigating many different configurations.
The results show consistent trends with those from Al Sanea
and Zedan [11]. Both models predict that Wall 1c has better insu-
lating performance, with insulation positioned towards the out-
side, than Walls 1a and 1b with insulation positioned towards
the inside and middle, respectively. The models also agree that
Wall 3 has the best performance among all of the walls studied,
and Walls 2b and 2c have similar performance, which improves
over Wall 2a. In terms of time lag, similar trends are observed. Wall
3 in both models has the longest time lag among all seven studied.
Comparing results demonstrates the ability of the RC model to
accurately capture thermal performance trends produced by a
numerical solution to the one dimensional heat conduction equa-
tion. The RC model will be used in this paper to investigate trends
that have not been well covered in the past and suggest improve-
ments over previous conclusions about wall assembly.

2.3. Multi-layer wall study

Specific objectives of this multi-layer wall study are to distin-


guish between the performance of an insulation/mass/insulation
configuration compared to mass/insulation/mass and insulation
towards the inside compared to insulation towards the outside,
to investigate the effect of distributing layers un-evenly, and to
evaluate the effect of increasing the overall number of insulation
and thermal mass layers. In order to understand the thermal
behavior trends for all possible multi-layer wall arrangements, four
basic patterns are studied, with each pattern being studied as a
family of configurations.
The four fundamental wall configurations to be studied are
shown in Fig. 4. Each wall has two cement plaster layers at the out-
er most position exposed to ambient and room conditions. The to-
tal volume of insulation and concrete in all configurations is the
same, but the layers are divided into percentages of the total mate-
Fig. 4. Four wall configurations: (a) Wall 0, (b) Wall 1, (c) Wall 2, and (d) Wall 3.
Each wall has thin layers of cement plaster at the outermost layers. Wall 0 has 7 rial, so the total wall resistance and capacitance values are fixed.
total layers with insulation at the outer layers. Wall 1 has 7 total layers with The four basic patterns are defined by how the insulation and con-
concrete at the outer layers. Wall 2 has 6 total layers with insulation at the outside crete are positioned. For instance, the Wall 0 family has insulation
layer and concrete at the inside layer. Wall 3 reverses the insulation and concrete at the outside, middle, and inside of the wall with concrete divided
layers from Wall 2.
evenly among the evenly-distributed insulation layers. The three
insulation layers are each one third (33%) of the total amount
of insulation in the wall configuration, while the concrete is split
change regardless of how the layers are arranged. The first family evenly into two layers, each 50% of the total amount of concrete.
considers a single 78 mm layer of insulation placed towards the in- The Wall 1 family reverses the position of concrete and insulation
side, middle, and outside of the wall. The second family splits the from Wall 0 such that concrete comprises the outside, middle, and
insulation into two 39 mm layers and distributes them at the in- inside layers of the assembly with insulation layers divided evenly
side/middle, inside/outside, and middle/outside positions. The final among the concrete. The Wall 2 family has layers of insulation at
configuration positions a third of the insulation (26 mm) at the in- the middle and inside, interspersed with concrete. Finally, the Wall
D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245 241

Table 2
Weighting variations for each of three wall configurations and extended Wall 0 configurations (8, 9, and 10). Layer lengths are displayed as percentages of total wall insulation or
thermal mass.

3 family is reversed with insulation at the middle and outside. thickness, or distribution, and material position. Wall 0.0, with
Variations of these four primary or base configurations are ex- evenly distributed layers, is extended to investigate the effect of
tended to investigate different weighting scenarios for the insula- increasing the total number of layer divisions. As with the other
tion and concrete layers. wall variations, Walls 0.8, 0.9, and 0.10 have the same total volume
Any weighting distribution might be selected for study with of insulation and concrete, thus maintaining the same overall resis-
infinitely many scenarios possible. For this paper, scenarios are tance and capacitance. However, each wall configuration divides
studied in which each base configuration, 0.0, 1.0, 2.0, and 3.0, is the two materials into a greater number of thinner layers. One
extended using different volumetric weightings. Specifically, con- might expect further divisions to continually improve perfor-
figurations are constructed to investigate alternatively weighting mance, but results will be discussed that suggest otherwise.
insulation, concrete, or both insulation and concrete by 80% of each
material volume towards the outside, middle, or inside of the wall.
Table 2 shows the configurations for all walls studied. For example, 3. Results and discussion
Wall 0.1 weights 80% of the insulation towards the outside with
the remaining 20% divided evenly between the middle and inside. Figs. 5–8 illustrate the (a) magnitude ratio and (b) phase results
Similarly, Wall 1.1 weights 80% of the concrete towards the out- for each material weighting configuration of the four wall families.
side. Wall 1.2 weights 80% of the concrete towards the middle, Fig. 5a, for example, shows the magnitude ratio results for Wall 0.0
and Wall 1.3 shifts it to the inside. Walls 1.4 and 1.5 alternate through 0.7 from the RC model frequency analysis, and Fig. 5b
weighting the two insulation layers towards the outside or inside, shows the respective phase results. Each bar is shaded and labeled
while Walls 1.6 and 1.7 alternate weighting concrete and insula- to indicate which material is weighted (no weighting or evenly dis-
tion layers towards the outside or inside together. tributed, insulation weighting, concrete weighting, or both in a
The four base configurations and the various weighting scenar- combined weighting) and in what direction the weighting is ap-
ios are designed to clarify wall assembly guidelines regarding layer plied (outside, middle, or inside). Refer to Fig. 4 and Table 2 for
242 D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245

Fig. 5. (a) Magnitude ratio and (b) phase results for Wall 0 weighting Fig. 6. (a) Magnitude ratio and (b) phase results for Wall 1 weighting
configurations. configurations.

more detailed configuration descriptions of each wall


configuration. with concrete or both concrete and insulation weighted towards
Wall 0.0 with evenly distributed layers and insulation at the the inside and outside show the highest magnitude ratio, with
outside, middle, and inside, has a lower magnitude ratio than the exception of Wall 1.3. These wall configurations – 1.1, 1.3,
any of the other Wall 0 weighting configurations (0.1 through 1.6, and 1.7 – weight 80% of the concrete towards either the inside
0.7) and outperforms all other configurations when considering or outside. Walls 1.4 and 1.5, that weight insulation only, show
magnitude and phase. For the Wall 0 family, weighting insulation some improvement, although 1.0 and 1.2 are still better with the
towards the inside performs better than weighting towards the concrete distributed evenly or weighted towards the middle.
middle or outside, but not as well as distributing the three layers Among the four layer Wall 2 and 3 configurations (Figs. 7 and
evenly. Weighting insulation and concrete together (either on the 8), 2.2 and 3.2 have the highest phase lag, and 50% of the concrete
inside or outside) has the highest magnitude ratio or worst insulat- is positioned toward one of the outer most layers. Walls 2.3 and 2.5
ing performance. Results also show that among the Wall 0 varia- weight 80% of the concrete towards the middle of the assembly,
tions, the two configurations, 0.4 and 0.5, that maintain and show the lowest phase lag among Wall 2 configurations. This
insulation evenly at the inside and outside and simply shift the suggests that increasing concrete at an outer layer helps increase
concrete between the two middle layers performs better than the phase lag performance of the wall assembly. Among Wall 1
any configuration that deviates more dramatically from the 0.0 configurations, Wall 1.2 is lowest and has just 10% of concrete at
base configuration. Wall 1 results also favor a configuration that both outermost layers. This lack of concrete at the outer layers
resembles the 0.0 assembly, namely that concrete is weighted to- resulting in decreased phase lag results supports the indication
wards the middle of the wall with insulation on both sides. The that increasing concrete at an outer layer increases phase lag. How-
best performing configuration for Wall 1 (Wall 1.2) occurs when ever, all of the Wall 3 configurations have relatively low phase lag
80% of the concrete is weighted towards the middle layer, leaving results (around 8 h) among the four families of wall configurations,
only 10% of the concrete at the inside and outside positions. If the and other results show some inconsistency in this trend. Such dis-
outer concrete layers were considered negligible, this configura- crepancy suggests that the model accuracy may need improve-
tion resembles a three layer version of the top performing five ment in order to capture phase lag trends among different wall
layer Wall 0.0, indicating a configuration favoring insulation at families as well as different weighting configurations within a
the outside and inside and concrete in the middle. The assemblies family.
D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245 243

Fig. 7. (a) Magnitude ratio and (b) phase results for Wall 2 weighting Fig. 8. (a) Magnitude ratio and (b) phase results for Wall 3 weighting
configurations. configurations.

When magnitude ratio is considered, Walls 2.3 and 3.4 show


the best performance among Wall 2 and 3 variations. This indicates
that weighting concrete towards the middle of the assembly is
best. This is consistent with Wall 1 results, where Wall 1.2 has
the lowest magnitude ratio and concrete weighted towards the
middle. Wall 0 results agree as well, where Walls 0.0, 0.4, and
0.5 show concrete weighted towards the middle layers, with at
least one third of the insulation on either side, and relatively low
magnitude ratios. Walls 2.5 and 3.6 also weight 80% of the concrete
towards the middle and have comparatively low magnitude ratios.
The evenly distributed configurations 2.0 and 3.0 also perform
comparatively well, while the worst performing Wall 2 and 3 con-
figurations – 2.6 and 3.5 – weight 80% of the concrete towards one
of the outer most layers, leaving only 20% towards the middle.
Since Wall 0.0 shows the best overall thermal performance
among the thirty variations, it is chosen as the configuration to ex-
tend for further analysis. The configuration is extended by dividing
the insulation and concrete into a great number of thinner layers.
Wall 0.0 has five layers of insulation and concrete, increasing to Fig. 9. Magnitude ratio (%) for Wall 0 extended configurations shown versus total
7, 9, and 11 in Walls 0.8, 0.9, and 0.10, respectively. Magnitude number of layers.
and phase results comparing these four evenly distributed Wall 0
configurations with an increasing number of layers are plotted in
Figs. 9 and 10. Fig. 9 shows that rather than decreasing continu- figurations studied. Insulation and concrete are distributed evenly
ously as the number of layers increases, the magnitude ratio ap- throughout the seven layers, and insulation is located at the inner
pears to reach a minimum with a smaller number of divisions. and outer most positions. Phase lag shows a continuously increas-
Wall 0.8 has the lowest magnitude ratio (M = 0.14) of all the con- ing trend as the number of layers increases. In terms of the RC
244 D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245

frequency. Previous work has also shown, generally, that phase


lag increases with an increasing number of layers. Even though
some results suggest that concrete at an outer layer helps increase
the phase lag, complete analysis shows that phase lag is not com-
promised when design focuses on reducing the magnitude ratio.
The assembly that minimizes the magnitude ratio among the pri-
mary thirty configurations considered also maximizes the phase
lag.
The degree to which maximizing the phase lag is desirable may
be debated in the same way that maximizing the decrement factor
is not always beneficial, as explained in the Masoso study [14],
where it is shown that increased insulation may actually lead to
higher annual energy consumption based on annual heating and
cooling loads. Similarly, maximizing the decrement factor may
not always be beneficial from a practical standpoint, unless meet-
ing specific thermal performance criteria while reducing material
usage is considered.
Fig. 10. Phase (h) for Wall 0 extended configurations shown versus total number of It is important to note that the analysis described herein need
layers. not include region specific climate data. Although the absolute
thermal performance might be dependent on this factor as shown
in [4], the qualitative relationships between the number of layers
and optimal configuration is expected to be maintained. In other
words, the same wall might not perform exactly the same in differ-
ent climates, but a wall that insulates better than its alternative in
one climate will still insulate better than its alternative in a differ-
ent climate. Also expected to have an impact is the solar insolation,
which would suggest another current source is needed in the RC
circuit or a sol–air method may be used to incorporate the effect
within the outside air temperature.

4. Conclusions

A lumped parameter RC model is used to simplify and expedite


solution of the one dimensional heat equation for a multi-layer
wall. Results indicate the model is sufficiently accurate, having rea-
sonable agreement with benchmarks found in the literature. The
first part of the analysis explores various weighting arrangements
developed from four primary configurations; a total of 33 configu-
rations are studied. Frequency response analysis results suggest
placing insulation at both inside and outside layers and positioning
thermal mass towards the middle of the assembly will result in the
best insulating performance. In the second part of the analysis, the
wall with the best performance from the first set of results, Wall
0.0, is extended by splitting insulation and mass layers into an
increasing number of divisions. These results suggest that there ex-
ists an optimum number of layers where the thermal performance
is maximized. In other words, continually increasing the number of
layers will at some critical point begin to result in a decrease in
performance. Additional studies are underway to clarify whether
distributing the thermal mass evenly among the inner layers is
best or if weighting those layers towards the middle yields better
thermal performance for the same overall resistance and capaci-
tance. For a more complete optimization procedure, the fabrication
and operating costs must also be considered. The current study
represents the first step in achieving this comprehensive analysis.

Appendix A
Fig. 11. Al-Sanea [11] Wall configurations: Wall 3 has insulation at inside, middle,
and outside.
See Fig. 11.

model, dividing each thermal mass layer into a greater number of References
layers increases the number of poles in the transfer function of the
model, since the impedance of each capacitance layer adds an inte- [1] Chen A. Closing in on zero-energy buildings. Environ Energy Technol Division
News 2008;8(2).
grator. The increased number of poles increases the negative slope
[2] Al-Sanea SA, Zedan MF. Effect of insulation location on initial transient thermal
of the phase response, resulting in a longer phase lag at the same response of building walls. J Therm Envelope Build Sci 2001;24:275–300.
D.E.M. Bond et al. / Applied Energy 112 (2013) 235–245 245

[3] Al-Sanea SA, Zedan MF, Al-Ajlan SA, Hadi ASA. Heat transfer characteristics and [16] Fraisse G, Viardot C, Lafabrie O, Achard G. Development of a simplified and
optimum insulation thickness for cavity walls. J Build Phys accurate building model based on electrical analogy. Energy Build
2003;26(3):285–307. 2002;34:1017–31.
[4] Kossecka E, Kosny J. Influence of insulation configuration on heating and [17] Peng C, Wu Z. Thermoelectricity analogy method for computing the periodic
cooling loads in a continuously used building. Energy Build 2002;34:321–31. heat transfer in external building envelopes. Appl Energy 2008;85:735–54.
[5] Bojić ML, Loveday DL. The influence on building thermal behavior of the [18] Davies MG. Optimum design of resistance and capacitance elements in
insulation/masonry distribution in a three-layered construction. Energy Build modelling a sinusoidally excited building wall. Build Environ 1983;18:19–37.
1997;26(2):153–7. [19] ASHRAE Handbook. Fundamentals. American society of heating. Refrigerating
[6] Bojic M, Yik F, Sat P. Influence of thermal insulation position in building and Air-Conditioning Engineers, Inc., Atlanta, GA; 2009.
envelope on the space cooling of high-rise residential buildings in Hong Kong.
Energy Build 2001;33:569–81.
[7] Aste N, Angelotti A, Buzzetti M. The influence of the external walls thermal
Glossary
inertia on the energy performance of well insulated buildings. Energy Build
2009;41:1181–7. xday: frequency whose period is one day
[8] Asan H, Sancaktar YS. Effects of wall’s thermophysical properties on time lag k: thermal conductivity
and decrement factor. Energy Build 1998;28:159–66. q: density
[9] Asan H. Numerical computation of time lags and decrement factors for Cp: heat capacity
different building materials. Build Environ 2006;41:615–20. df: decrement factor
[10] Asan H. Investigation of wall’s optimum insulation position from maximum M: magnitude ratio
time lag and minimum decrement factor point of view. Energy Build G(jx): system transfer function
2000;32(2):197–203. /: phase angle
[11] Al-Sanea SA, Zedan MF. Improving thermal performance of building walls by
tlag: time lag
optimizing insulation layer distribution and thickness for same thermal mass.
hout: outside surface heat transfer coefficients
Appl Energy 2011;88:3113–24.
[12] Mavromatidis LE, Mankibi MEL, Michel P, Santamouris M. Numerical hin: inside surface heat transfer coefficients
estimation of time lags and decrement factors for wall complexes including R: resistance
multilayer thermal Insulation, in two different climatic zones. Appl Energy C: capacitance
2012;92:480–91. L: length
[13] Kaynakli O. A review of the economical and optimum thermal insulation Z Ri : impedance value for the resistance of layer ‘‘i’’
thickness for building applications. Renew Sustain Energy Rev 2012;16:415–25. Z C i : impedance for the capacitance of layer ‘‘i’’
[14] Masoso OT, Grobler LJ. A new and innovative look at anti-insulation behaviour Zeq: equivalent impedance
in building energy consumption. Energy Build 2008;40:1889–94.
[15] Al-Sanea SA, Zedan MF, Al-Hussain SN. Effect of thermal mass on performance
of insulated building walls and the concept of energy savings potential. Appl
Energy 2012;89:430–42.

Potrebbero piacerti anche