Sunteți pe pagina 1din 12

Construction and Building Materials 43 (2013) 533544

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


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

The variation of thermal conductivity of brous insulation materials


under different levels of moisture content
A. Abdou , I. Budaiwi
Architectural Engineering Department, King Fahd University of Petroleum and Minerals (KFUPM), Dhahran 31261, Saudi Arabia

h i g h l i g h t s
 Higher operating temperature and higher moisture content of brous insulation are always associated with higher k-value.
 The magnitude of change in k-value is generally higher with higher moisture content at a given operating temperature.
 Higher density samples generally exhibit larger changes in k-value at the same moisture content level.
 Different initial moisture contents of brous insulation lead to different k-value and moisture content relationship.
 The rate of change in k-value with moisture content is higher at higher initial moisture content.

a r t i c l e

i n f o

Article history:
Received 8 April 2012
Received in revised form 1 February 2013
Accepted 26 February 2013
Available online 3 April 2013
Keywords:
Fibrous thermal insulation material
Thermal conductivity
Moisture content
Operating temperatures

a b s t r a c t
Thermal insulation plays an important role in determining the thermal and energy performance of a
building. The effectiveness of thermal insulation is dependent on its thermal conductivity (k-value)
and its ability to maintain its thermal characteristics over an extended period of time. However, the kvalue can be greatly reduced by the presence of moisture within the insulation materials. For example,
when circumstances are conducive, under hothumid climatic conditions, condensation may occur
within the insulation material, raising its moisture content well above the hygroscopic level. The objective of this paper is to investigate experimentally the impact of moisture content on the thermal conductivity of commonly used brous insulation materials. Three types of material with different densities are
investigated with the emphasis on berglass. A comparison of the behavior and magnitude of k-value
changes according to moisture content levels indicated appreciable variations. Higher thermal conductivity at a given operating temperature for the investigated densities is always associated with higher moisture content. The relationship between k-value and moisture content is found to be affected by the initial
conditioning moisture content level. Materials having similar density but conditioned at different initial
moisture content levels exhibit different relationships between k-value and moisture content. The rate of
change in thermal conductivity with moisture content is higher at higher initial moisture content. The
results should be of great importance to material manufacturers, building owners and designers when
selecting suitable insulating materials and correctly predicting the thermal and energy performance of
buildings and their energy-efciency.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
In harsh climatic conditions, the use of thermal insulation in
buildings is necessary and is gradually becoming a mandatory
requirement in many countries particularly as energy becomes
more precious and demand increases. The thermal conductivity
of insulation materials is greatly affected by their operating temperature and moisture content, yet limited information is available
on the performance of insulation materials when subjected to ac Corresponding author. Address: KFUPM, P.O. Box 1917, Dhahran 31261, Saudi
Arabia. Tel.: +966 3 860 2762; fax: +966 3 860 3785.
E-mail address: adel@kfupm.edu.sa (A. Abdou).
0950-0618/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.conbuildmat.2013.02.058

tual climatic conditions. Many parameters should be considered


when selecting thermal insulation, including cost, compression
strength, water vapor absorption and transmission and, most
importantly, the k-value of the material when considering thermal
performance of buildings and relevant energy conservation
measures.
Published k-values and those reported by manufacturers are
normally evaluated under standard laboratory conditions of temperature and humidity to allow a comparative evaluation of thermal performance. However, when placed in their locations in the
building envelope, thermal insulation materials are exposed to different temperature and humidity levels depending on the prevailing climatic conditions. Hence, their actual thermal performance

534

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

may substantially differ from that predicted under standard laboratory conditions.
The impact of operating temperature and moisture content on
insulation thermal conductivity varies with the type of insulation
depending on the composition, properties and internal structure
of the materials used, which determine modes of heat transfer
and the moisture storage capacity of the material. The thermal performance of rigid cellular foam insulation was theoretically and
experimentally evaluated under different insulation temperatures
by Aldrich and Bond [1]. The results showed pronounced variations
in the k-value in varying operating conditions. Another set of experiments [2] was conducted on the thermal performance of berglass
using an attic test module in a guarded hotbox facility. Experiments
conducted with one type of loose-ll berglass insulation which
showed that the thermal resistance at large temperature differences
was about 3550% less than that at small temperature differences.
Loose-ll insulation has a somewhat lower thermal resistance at
larger temperature differences across the insulation [3].
Abdou and Budaiwi [4] measured the thermal conductivity of
insulation materials at different operating mean temperatures
using a PC-automated heat ow meter. Their results indicated that
higher temperature leads to higher thermal conductivity values
and that higher insulation density generally results in lower thermal conductivity. The impact of such variations on envelope-induced cooling load was also investigated [5]. Thermal
conductivity increases with increasing temperature. The increase
is relatively large at low temperatures and then levels off with a
further increase in temperature. The relation becomes practically
linear at temperatures above 35 C [6].
In a study conducted by Zhang et al. [7] effective k-values of the
brous insulation were measured over a wide range of temperatures (300973 K to 27700 C). The effective thermal conductivity
was found to increase non-linearly with increasing sample average
temperature due to the fact that radiation heat transfer is related
to the fourth power of temperature, and it is more dominant with
the increase of temperature. In addition to the operating temperature, the material moisture content, which is inuenced by the
ambient humidity level, is another major factor that can diminish
the thermal conductivity of insulation materials [8]. In buildings,
insulation materials used in walls and roofs normally exhibit higher moisture content when compared to test conditions. The ambient air humidity and indoor conditions, as well as the wall or roof
system moisture characteristics, play an important role in determining the moisture status of the insulation material. When conditions are favorable (e.g. hot-humid climates), condensation can
occur within the insulation material, raising its moisture content
well above the hygroscopic level (i.e. un-wetted at RH 98%).
Numerous studies have been carried out to assess the impact of
moisture content on insulation thermal performance. Two parameters of hygroscopic moisture transfer, i.e. the equilibrium moisture content and the thermo-gradient coefcient were studied
[9]. Hedlin [10] reported that the effectiveness of at roof insulation was found to be reduced by the presence of moisture. Results
indicated that a signicant increase in energy exchange (gain and
loss) through the roof occurred for moisture contents less than
1% (by volume). Chyu et al. [11,12] investigated the performance
of berglass insulation and mineral wool used on heating and cooling pipes subjected to underground water attack by measuring the
effective thermal conductivity and the moisture absorption rate.
The effective k-value of the wet berglass insulation was found
to be many times higher than that of the dry insulation [11]. Another study [13] was conducted to simulate the effect of condensation on the performance of berglass slab insulation in a laboratory
environmental chamber. Sandberg [14] showed that for cellulose
ber, the effect of moisture on the thermal transmissivity was in
the order of 0.001 W/m K within the hygroscopic range.

Field, laboratory and theoretical studies [15] on various types of


insulation including glass ber, cellulose ber and brous biological insulating materials have all been carried out. It was concluded
that the effectiveness of insulating materials at higher moisture
content is reduced in proportion to the moisture content level.
Higher thermal conductivity is obtained as a result of increased energy transfer by conduction and, under certain conditions, by the
evaporationcondensation process, in which moisture moves from
warm to cold regions. Simonson et al. [16] studied simultaneous
heat and mass transfer through a medium-density berglass insulation bay. During the study the insulation was open to ambient air
at a specic humidity on the warm side. The cold side boundary
was an impermeable cold plate at specic temperatures. Thermal
hysteresis was noted at high levels of relative humidity, i.e. when
condensation occurred in the brous insulation material. The study
exemplied how condensation occurs in insulation materials.
Peuhkuri [17] investigated how a number of different porous
insulation materials (mineral wool, cellulose ber, etc.) performed
hygrothermally under conditions similar to those in a typical
building envelope. The liquid distribution, total moisture gain, heat
ux and temperature distributions for moisture absorption in brous insulation materials were studied [18]. Moisture transport
mechanisms along the ber direction and perpendicular to the ber direction were not the same. Capillary rise occurred along
the bers but was not perpendicular to them. Wijesundera et al.
[19] identied four stages of transport processes: (1) A relatively
short initial transient stage in which the temperature and vapor
concentration elds are developing within the insulation slab. (2)
Heat and vapor transfer reach a quasi-steady state, and the temperature and vapor density elds are invariable with time. Liquid
is accumulated in the wet region, but is still at a low level and does
not have a signicant effect on the transport properties. (3) When
liquid accumulation exceeds a critical value the liquid starts to
move due to the generated liquid pressure and ow towards the
wet-dry interface. (4) In the last step the liquid front eventually
reaches the exposed surface, and liquid accumulation in the slab
continues.
Jintu et al. [20] reported on a transient model of coupled heat
and moisture transfer through brous insulation. Based on their
model, numerical simulation was carried out to better understand
the effect of various material and environmental parameters on the
heat and moisture transfer. It was found that the initial water content and thickness of the brous insulation together with the environmental temperature are the three most important factors
inuencing the heat ux. The model was improved and results
were compared to experimental data [21]. Tomas and Kjartan
[22] conducted an experiment in order to study the sensitivity of
brous insulation materials to moisture content. The results were
analyzed and the conclusion was drawn that the moisture properties of brous insulations (i.e. cellulose and stone wool) were not
seriously affected by extreme water condensation, not even by
ice forming in time during the test period. The impact of moisture
content seems to be much more important for thermal properties.
On a longer time scale, the structural stability of the tested insulation materials may deteriorate and the hygienic quality may be
compromised by the growth of mould and fungi in wet structures.
In a more comprehensive study [23] an experimental analysis to
compare the thermal performance of various insulation materials
used in buildings was conducted. The thermal performance of
three typical insulation materials (polyurethane, polystyrene, and
mineral wool) was investigated. For this purpose, four house-like
cubicles were constructed (with a size of 2.4 m  2.4 m  2.4 m)
and their thermal performance throughout the time was measured. The cubicles were built in a conventional Mediterranean
construction system, differing only in the insulation material used.
During 2008 and the rst months of 2009 (summer and winter

535

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

season) the energy consumption of these cubicles was evaluated


and it was found that during the summer season the energy savings from mineral wool (e.g. brous insulation) was around 46%,
and in the winter season it was 34%, with a percentage reduction
of 12%. This was attributed to the high moisture level in the winter
season penetrating through the cubicle and facilitating the ow of
heat by increasing the thermal conductivity of the cubicle.
A transient model was prepared [24] for heat and mass transfer
through brous insulation materials by incorporating the effect of
conduction, water vapor diffusion, evaporation/condensation and
sorption/desorption in the model. A parametric study considering
different amplitudes of temperature change, different moisture
masses and different thicknesses of the insulation matrix was carried out. They found that a relatively small mass of water in the insulation matrix can result in a signicantly increased average heat ux
during a periodic cycle. The numerical code was veried with experiments, which showed good agreement with the numerical analysis.
Bjorn [25] conducted a review of brous materials such as glass
wool, mineral wool, and rock wool. Typical thermal conductivity
values for mineral wool are between 0.030 and 0.040 W/m K. The
thermal conductivity of mineral wool varies with temperature,
moisture content and mass density. An example was given, in
which the thermal conductivity of mineral wool increased from
0.037 to 0.055 W/m K (i.e. 48.6%) with increasing moisture content
from 0% to 10%, by volume respectively. Anastasios et al. [26] studied the thermal performance of stone wool, a widely used inorganic brous material, which is based on the air embodied
between its bers, producing a low thermal conductivity factor.
They evaluated the changes in the k-value under various operational conditions and found that thermal conductivity dramatically
increased due to hydrophilic and water absorption. Diffusion of
water in the brous materials maze caused a signicant increase
of the stone wools k-value.
Research undertaken into the effect of moisture on mineral ber
insulation by Achtziger and Cammerer of FIW (Research Institute
for Thermal Protection, Germany) [27] concluded that 1% moisture
content (by volume) within mineral ber, can increase the k-value
of the material by 36107% [28]. Carey et al. [29] studied the effective k-value of berglass insulation. Under conditions of 97% relative humidity and a temperature of 5 C, the heat ux was found
to be 4.9 times greater than the dry case even though the accumulation of water was quite low (e.g. 0.14% by volume). Under these
boundary conditions the average k-value of the tested insulation
increased by 3.5%. Ochs et al. [30] pointed out that published information about measurements of the k-value of bulk insulation at
higher temperatures and at elevated moisture contents are limited.
Modeling and measurement of the effective k-value of porous bulk
materials at temperatures up to 80 C and moisture contents below
free water saturation were described. A signicant increase of the
effective k-value at temperatures >60 C with water contents <5%
(by volume) was detected. Krishpersad and Gurmohan [31] indicated that accurate prediction of the effective k-value of loose-ll
brous thermal insulation used in buildings remains a challenging
issue. The contributions of the different modes of heat transfer
across a brous material to the total heat transfer depended on
many variables such as effective density, material properties,
arrangement of bers and moisture content. The measurement of
complete sets of heat and moisture transport and storage parameters of ve thermal insulation materials dependent on moisture
content was presented [32]. The k-value of mineral wool was found
to increase very fast with increasing moisture content.
From the above review, it is evident that both operating temperature and humidity factors have a signicant inuence on the
thermal performance of insulation materials. Heat transfer in brous insulation is signicantly affected by the mass transfer even
for low moisture contents. The presence of moisture in insulating

materials can signicantly degrade the effectiveness of their thermal resistance and hence the efciency of the thermal insulation
material. The impact of moisture on the k-value of insulating materials has been the subject of many studies. However, despite the
wide use of brous insulation materials in buildings, limited research has addressed the change of k-value of various types of brous insulation with different densities in the presence of
moisture at different mean operating temperatures and conditioned at different initial moisture contents.
The objective of this paper is to investigate the inuence of
moisture content change at different operating temperatures on
the variation of thermal conductivity and consequently to assess
the relative sensitivity of the k-value of brous insulation materials
commonly used for insulating buildings. Knowledge of thermal
conductivity variation with moisture content will help in correctly
predicting the building thermal loads induced through the exterior
envelope and the subsequent energy performance and potential
energy savings.
2. Methodology
In order to achieve the objective of the study, brous thermal insulation samples with a standard size of 300  300 mm and nominal thickness of 50 mm representing commonly used insulation materials were collected. The samples included
berglass of ve different densities, rock wool of ve different densities and mineral
wool material of one density. Table 1 lists the test specimens of brous materials
acquired for measurements, with their given reference code, density and nominal
thickness.
The samples were conditioned using a moisturizing apparatus and their k-values were measured at different moisture content levels and mean operating temperatures. A state-of-the-art heat ow meter apparatus was utilized to measure
the k-value. Measurement data were analyzed utilizing linear regression to quantify
the impact of moisture content and to establish the relationship between the k-value of different types of brous insulating materials with various mass densities and
the level of moisture content over the hygroscopic and wetting ranges.
In similar studies experiments for testing specic brous insulation specimens
under transient heat and moisture transfer were developed. Larson and Benner [15]
measured transient thermal resistance and moisture gain of spray-applied berglass specimens with a temperature difference of 20 C between the specimen
hot and cold sides and exposed to 10% and 50% relative humidity in an air chamber
at the bottom surface of the specimen. The experimental set-up developed by Wijesundera et al. [19] aimed to simulate mainly the moisture absorption by brous
insulations. The liquid distribution, total moisture gain, heat ux, and the temperature distributions were measured along with the heat ux, which showed a very
gradual increase with time. To achieve the objective of this study a different procedure and experimental setup for conditioning the samples and measurement of the
k-values were devised.
2.1. Measurement apparatus
The heat ow meter (i.e. Holometrix, Type Lambda 2300V, 1997) [33] was utilized. The system is a complete PC-automated system designed to determine the
thermal conductivity as per the ASTM C518-91 [34] and ISO 8301 [35] protocols.
Table 1
List of test samples ordered by type, code, and mass densities.

Density (kg/m3)

Material type

Sample code

Fiberglass

FG-01
FG-02
FG-03
FG-04
FG-05

27.0
47.0
66.0
70.0
84.0

Nominal thickness (mm)


50
50
50
50
50

FG-02h
FG-03h
FG-05h

47.0
65.0
84.0

50
50
50

Rock wool

RW-01
RW-02
RW-03
RW-04
RW-05

46.0
59.0
64.0
76.0
99.0

30
50
45
30
25

Mineral wool

MW-01

145.0

50

Samples to be conditioned at higher initial moisture contents.

536

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544


vent lateral heat ow. The area of the heat ow transducer is smaller than the crosssectional area of the specimen. The part of the specimen surrounding the area over
which the transducer monitors heat ow acts as an effective guard against lateral
heat ow. Fig. 1 shows the dimensions of the test sample and the self-guarding effect of the specimen.

2.2. Sample preparation

Fig. 1. The self-guarding effect of the sample in the measuring instrument.

A specimen (300  300 mm) with a thickness ranging from 5 to 100 mm can be
placed in the test section of the heat ow meter between two plates, which are
maintained at different temperatures during the test. Upon achieving thermal equilibrium and establishing a uniform temperature gradient throughout the sample,
the k-value is determined. The temperature control system is thermoelectric. An
external computer system with specialized software (i.e. Holometrix Q-LAB software) controls the measuring apparatus, records, and prints results. The heat ow
meter can test materials with thermal conductivities in the range of 0.0050.5 W/
m K. To minimize errors, the apparatus can be easily calibrated with a reference
standard specimen (i.e. supplied by the National Institute of Standards and Technology, NIST) to measure the response of the heat ow transducer at different
temperatures.
The most important features of this automated system are accuracy and reproducibility. Accuracy is the degree of variation for a measurement, taking into account the reproducibility and all physical aspects of the instrument and the test.
The accuracy of the apparatus ranges from 1.0 to 3.0%, while reproducibility is
around 0.3%. This means that the variation of the test results of one specimen is
around 0.3% from one test to another when using the same specimen and testing
conditions. The range of the test sample mean temperature can be varied from 0
to 100 C. Prior to the testing of the selected sample, the measuring instrument
was calibrated using a reference sample as per the manufacturers calibration procedure. In order to verify the accuracy and ensure reproducibility, the reference
sample was tested 10 times over an extended period of time. The accuracy and
repeatability of the ow meter were conrmed within a mean error of around
0.2% with standard deviation of 0.1%.
A temperature difference (DT) of 15 C was maintained between the sample hot
and cold sides. Since over-compression of the brous insulation (particularly loose
ones with relatively low densities) affects the measurements of the k-value, the
samples were not compressed within the heat ow meter. Using a control switch
with caution, the heat ow meter lowers the upper plate until it makes rm contact
with the specimen upper surface. The switch limits further movement of the upper
plate at this point and the thickness transducer reads the sample thickness. In the
test compartment of the heat ow meter, the specimen itself acts as a guard to pre-

In order to account for variations in sample densities and to better assess the
relative moisture content of different samples, the actual density of each test sample at the dry condition was determined using a sensitive digital weighing scale
with 0.01 g accuracy. In order to increase the sample moisture content, the dry test
sample was initially conditioned for 72 h in a bench-top temperature and humidity
conditioning chamber [i.e. Espec, Type SH 641, temperature ( 40 to +150 C, 0.3 C)
and relative humidity (3090%, 3%)] with the temperature set to 24 C and relative
humidity set to 95%. The moisture content of the sample was found to increase from
zero to only 3% by weight. It was not possible to increase the moisture content of
the sample above 3% via the conditioning chamber. The dry samples were kept in
a conditioning oven (i.e. at 22 C) for a number of hours until less than 1% of weight
change was observed [34].
To further increase the moisture content of the test sample, a bench-top stainless steel container with a controllable electric-coil which heats an adequate volume of de-ionized water was also used. The container can accommodate a
sample size of 300  300  100 mm. The apparatus adds moisture to the test sample by heating up the volume of de-ionized water. The sample was mounted over a
stainless steel mesh holder when placed in the container away from the water surface. The holder was high enough to ensure that the central part (10  10 cm) of the
test sample was exposed to the vaporizing water. The sample upper surface was
covered with hardboard wrapped in aluminum foil so that the condensing water
droplets on the containers stainless lid falling on the upper surface of the test sample were drained away. The sample was turned upside down every 15 min to ensure
a uniform exposure of the sample central part to the rising vapor over its thickness.
Executing the turnover required unsealing the testing set-up i.e. opening the moisturizing container lid. This was done efciently in few seconds to minimize any potential effect of relative humidity misbalance on the measurements. It was expected
that shortly after such transient imbalance, relative humidity would regain its balance once the testing setup is re-sealed leading to negligible impact.
The density of the sample was frequently monitored till a desired amount of
moisture content percentage (by weight) was achieved but not necessarily a precise
level of moisture content. The sample was then removed from the container, immediately wrapped with a thin plastic lm in order to maintain its moisture content
during the measurement time period and kept for a few hours at room temperature.
The k-value of the test sample was then measured using the heat ow meter. To test
the same sample at different and reduced levels of moisture content, the conditioned sample was unwrapped and kept at the room temperature with its weight
frequently monitored until its moisture naturally decreased to a lower moisture
content. During this process the sample was ipped over every 15 min to maintain
uniform moisture distribution over its thickness. The sample was then wrapped and
its k-value re-measured. The process was repeated for a second lower level of moisture content. Fig. 2 depicts a cross section of the moisturizing container with the
sample arrangement.
To decide on the moisture content range, the authors have simulated (for a period of 2 years) a 50-mm layer of berglass thermal insulation as actually applied in
a building envelope in the hot-humid climate of Dhahran in Saudi Arabia. This was
modeled and simulated using hygrothermal simulation software. The resulting
accumulation of moisture within the berglass insulation layer in a wallroof system was up to 1215% (by mass) which is representative of actual moisture accumulation at a given time period in hot-humid climates. Moreover, based on the
described experiment an increase of moisture content beyond the range considered

Container Lid
Hardboard wrapped in
aluminum foil

Water-absorbing board

Test Sample
30 x 30 x50 mm

Heating Coil

Holder: Stainless
Steel Frame

Temperature
Control

De-ionized water

Fig. 2. Cross-section of the moisturizing container with the test sample arrangement.

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

(a)

Dry_Sample

MC=6.2%

MC=15.7%

MC=43.5%

0.04400

K, W /m. oC

0.04200

Fiberglass, 27 kg/m3
k = 0.000131*t + 0.033761
R2 = 0.981

0.04000

k = 0.000153*t + 0.033965
R2 = 0.976

0.03800
0.03600
k = 0.000187*t + 0.031279
R2 = 0.998

0.03400
k = 0.000188*t + 0.030677
R2 = 0.999

0.03200
0.03000
14.0

19.0

24.0

29.0

34.0

39.0

Temperature, oC

(b)

T= 14 oC

T= 24 oC

T= 34 oC

0.0440

K, W /m. oC

0.0420

Fiberglass, 27 kg/m3

k = 0.000046*mc + 0.037277
R2 = 0.955

0.0400
0.0380
0.0360
0.0340

k = 0.000054*mc + 0.035497
R2 = 0.875

0.0320

k = 0.000062*mc + 0.033716
R2 = 0.791

0.0300
0.0

5.0

10.0 15.0 20.0 25.0 30.0 35.0 40.0 45.0 50.0

Moisture Content, % (by Weight)


Fig. 3. (a) Best-t linear relationship of measurement results of berglass (density
27 kg/m3) at different percentages of moisture content (MC, % by weight), and (b) kmoisture content relationship at 14, 24, and 34 C operating temperatures.

in this investigation (45% by mass) resulted in a major change of the materials


physical characteristics particularly rock and mineral wool test samples. Taking into
account the interest in small range of moisture content as well as the limitation of
the moisture conditioning process and control two to three moisture content levels
were intended for each sample.
In each density category of berglass insulation material (1.0 kg/m3), there
was more than one sample taken from the same batch of test samples. Three samples of similar densities (i.e. FG-02h, FG-03h, and FG-05h) were conditioned to a
higher initial moisture content to study the impact of initial moisture content.
These were stored under the same conditions and tested using the same procedure
under the same mean operating temperatures and conditions.
2.3. Measurement results
Utilizing the test sample preparation and the standard thermal conductivity
measurement procedure discussed above, a total of 11 samples were tested at different levels of moisture content under three operating temperatures. The results of
k-value variation with moisture content and operating temperatures for a berglass
sample having a density of 27 kg/m3 are shown in Fig. 3. Due to limitations in sample preparation and measurement techniques, it was difcult to precisely control
the moisture content at specied constant levels resulting in k-values corresponding to variable moisture content and operating temperatures. This made it difcult
to compare thermal conductivity variations for the different samples at the same
operating conditions. As a rst step toward establishing a relationship between
thermal conductivity and moisture content at a specied operating temperature,
the linear relationship between the measured k-value and the resulting operating
temperature is established. This best-t linear relationship is shown in Fig. 3a
where parameters of this linear relationship are given. Additionally, Fig. 3b illustrates the numerical interpolation of k-values at the operating temperatures of
interest namely 14 C, 24 C and 34 C. Utilizing these relationships, the k-value
of the brous insulation material at specic moisture content and operating temperature can be obtained.
The best-t linear relationships of measurement results of the remaining breglasses samples, rock wool samples and the mineral wool sample are presented in
Table 2. These are graphically presented in parts (a and b) of Fig. 4. Fig. 5 illustrates
the measurement results of rock wool test samples and Fig. 6 illustrates the results
of the mineral wool sample.
Tables 2 and 3 summarize the k-value variations of the tested samples at different moisture contents and operating temperatures. Table 2 illustrates the thermal
conductivity values of the berglass, the rook wool and the mineral ber samples

537

measured at different moisture content levels and at 24 C and 34 C operating temperatures. The percentage of change in thermal conductivity at the different operating temperatures and moisture content levels compared to the dry conditions is
also presented. Additionally, the parameters of the best t linear function that relates temperature with thermal conductivity at different moisture content levels
for all tested samples are also presented. Table 3 presents the values of thermal conductivity at specic moisture content levels (i.e., 0%, 5%, 10% and 15%) under different operating temperatures. Additionally, it illustrates the percentage of change of
thermal conductivity with the specied levels of moisture content at operating
temperatures of 14 C, 24 C and 34 C. These values pertinent to thermal conductivity were obtained from the best t curve of the actual measured operating temperatures versus thermal conductivity measured under the actual measured
moisture content.

3. Analysis and discussion of results


Examination of the above experimental results reveals that
higher operating temperature and higher moisture content are
always associated with higher thermal conductivity for the investigated densities. From Table 2 and Fig. 2a it can be deduced that
increasing the moisture content of a berglass sample with a
density of 27 kg/m3 from 0% to around 16% kg/kg at an operating
temperature of 24 C has resulted in an approximately 5% increase
in k-value. A reduced percentage of increase in k-value of about 3%
is associated with the same level of moisture content at an operating temperature of 34 C. A similar impact of operating temperature can be generally observed for different densities but in
particular at the lower densities of the berglass material. As the
density increases, the impact of operating temperature on
the percentage of change of thermal conductivity at different moisture content levels is generally less. This can be attributed to the
reduced presence of trapped air (the most affected component of
the material matrix by the operating temperature) within the
materials. For example, the percentage of change in k-value for
the 70 kg/m3 berglass sample when the moisture content is increased to 13.8% is 14.4% at an operating temperature of 24 C
and is about 13% at an operating temperature of 34 C. The magnitude of change in k-value is generally higher with a higher moisture content at a given operating temperature. This can be
clearly observed, in particular at higher densities.
Higher density samples generally exhibit larger changes in thermal conductivity at the same moisture content level of berglass as
density increases (as can be seen in Fig. 7, i.e. four out of ve samples). The fact is that higher density samples would hold a higher
quantity of water compared to the lower density samples under
the same moisture content level resulting in a higher k-value due
to waters thermal characteristics. For example, the magnitude of
change in k-value of the berglass sample having a density of
47 kg/m3 is about 5% at a moisture content of around 16% while
it is about 9% and 14% at a similar moisture content and operating
temperature for the densities of 66 kg/m3 and 70 kg/m3 respectively. The magnitude of change of k-value does not necessarily exhibit a consistent and proportional change as density gets higher.
This can be attributed to the unique internal pore structure of
the material sample and the pattern of moisture distribution within the material.
It should be noted from the results of the investigated samples
that the behavior of k-value of berglass insulation is more robust
and generally exhibits a consistent pattern with temperature and
increased moisture contents. On the other hand, the change of
k-value of rock wool with elevated moisture content shows less
consistency. This can be explained by the observed change in the
samples physical characteristics and potentially by the change in
the ber matrix structure of the material due to the effect of moisture and heat on the bers bonding agent. Therefore, further analysis will focus on the behavior of berglass with elevated moisture
and the impact of its initial content. Results for the relatively highdensity sample of the mineral wool which is exposed to higher

Material type

Fiberglass

Rock wool

Mineral wool

Density (kg/m3)

FG-01

27.0

FG-02

47.0

FG-03

66.0

FG-04

70.0

FG-05

84.0

FG-02h

24

FG-03h

65.0

FG-05h

83.0

RW-01

46.0

RW-02

59.0

RW-03

64.0

RW-04

76.0

RW-05

99.0

MW-01

145.0

Samples conditioned with higher initial moisture contents.

Moisture content (%)

k-Value at 24 C

k-Value at 34 C

W/m C

W/m C

% Change

Slope

% Change

Regression data
Y-intercept

Correlation coeff., R2, ratio

0
6.2
15.7
43.5
0
8.8
15.7
22.7
0
8.0
13.6
0
6.6
13.8
0
3.9
13.8

0.0352
0.0358
0.0369
0.0376
0.0335
0.0339
0.0352
0.0382
0.0325
0.0330
0.0353
0.0326
0.0342
0.0373
0.0324
0.0334
0.0342

1.6
4.9
7.0

1.2
5.1
14.2

1.7
8.8

4.8
14.4

2.9
5.5

0.0371
0.0376
0.0382
0.0392
0.0348
0.0355
0.0364
0.0390
0.0339
0.0344
0.0383
0.0338
0.0354
0.0381
0.0336
0.0343
0.0353

1.5
3.1
5.7

2.1
4.7
12.1

1.7
13.2

4.8
12.9

2.1
5.2

0.00019
0.00019
0.00013
0.00015
0.00013
0.00016
0.00012
0.00007
0.00014
0.00014
0.00030
0.00012
0.00012
0.00008
0.00011
0.00009
0.00011

0.03068
0.03128
0.03376
0.03397
0.03047
0.03010
0.03242
0.03649
0.02914
0.02964
0.02811
0.02985
0.03126
0.03536
0.02976
0.03125
0.03159

1.00
1.00
0.98
0.98
0.98
1.00
0.97
0.99
1.00
1.00
1.00
0.99
0.99
0.84
0.99
0.87
0.91

0
5.4
12.5
29.6
0.0
8.1
16.6
29.0
0.0
5.6
12.4
20.2

0.0342
0.0349
0.0356
0.0408
0.0322
0.0362
0.0570
0.0589
0.0318
0.0337
0.0357
0.0443

2.0
4.2
19.4

12.3
76.8
82.5

6.2
12.4
39.4

0.0358
0.0362
0.0371
0.0427
0.0338
0.0388
0.0685
0.0723
0.0329
0.0349
0.0374
0.0507

1.1
3.7
19.4

14.7
102.5
113.7

6.2
13.7
54.0

0.00016
0.00013
0.00015
0.00020
0.00016
0.00026
0.00115
0.00134
0.00011
0.00012
0.00017
0.00064

0.03021
0.03166
0.03195
0.03603
0.02843
0.02997
0.02941
0.02660
0.02904
0.03078
0.03156
0.02889

0.99
1.00
0.93
0.79
0.96
0.76
1.00
1.00
0.95
1.00
0.99
0.99

0
7.4
13.7
23.8
0
7.1
12.7
33.6
0
7.5
10.6
20.2
0
14.6
26.4
0
7.1
12.7

0.0397
0.0400
0.0398
0.0400
0.0366
0.0376
0.0385
0.1456
0.0361
0.0371
0.0480
0.0490
0.0348
0.0360
0.0441
0.0333
0.0357
0.0360

0.6
0.2
0.7

2.6
5.0
297.4

3.0
33.0
35.9

3.5
26.8

7.4
8.2

0.0422
0.0423
0.0422
0.0426
0.0369
0.0391
0.0400
0.2003
0.0378
0.0390
0.0531
0.0545
0.0363
0.0374
0.0490
0.0341
0.0374
0.0377

0.4
0.0
1.1

5.8
8.4
442.1

3.1
40.4
44.0

3.1
35.1

9.7
10.6

0.00024
0.00024
0.00023
0.00026
0.00003
0.00015
0.00016
0.00547
0.00003
0.00015
0.00016
0.00547
0.00015
0.00014
0.00049
0.00008
0.00017
0.00018

0.03394
0.03432
0.03420
0.03387
0.03593
0.03411
0.03477
0.01441
0.03593
0.03411
0.03477
0.01441
0.03128
0.03273
0.03241
0.03125
0.03161
0.03178

1.00
1.00
1.00
1.00
0.99
0.95
0.99
0.99
0.99
0.95
0.99
0.99
1.00
1.00
1.00
0.92
0.99
1.00

0
8.5
22.2
46.6

0.0358
0.0987
0.1273
0.1459

175.8
255.8
307.9

0.0371
0.1228
0.1673
0.1795

230.9
350.9
383.8

0.00013
0.00241
0.00401
0.00336

0.03255
0.04080
0.03114
0.06517

1.00
1.00
0.98
0.86

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

Fiberglassa

Sample code

538

Table 2
Measurement results of k-value of berglass, rock wool and mineral wool, with different densities and moisture content levels (MC, % by weight) under 24 and 34 C mean operating temperatures, and the corresponding percentage
change (%) of k-value with MC increase. Note: regression data and correlation coefcients are indicated.

539

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

(a)

Dry_Sample

MC = 8.8 %

MC =15.7%

(b)

MC = 22.7%

T = 14 oC
Fiberglass, 47 kg/m 3

0.0400

0.0400

0.0380

0.0380

K, W/m. oC

K, W/m. oC

Fiberglass, 47 kg/m 33

0.0360
0.0340

0.0300
14.0

k = 0.0002 * mc + 0.0328
R2 = 0.84

0.0360
0.0340

k = 0.0002 * mc + 0.0313
R2 = 0.80

0.0300
19.0

24.0

29.0

Temperature,
Dry_Sample

34.0

0.0

39.0

oC

MC = 8.0%

5.0

10.0

15.0

20.0

25.0

Moisture Content, % (by weight)


T = 14 oC

MC =13.6%

0.0420

T = 24 o C

T = 34 o C

0.0420
Fiberglass, 66 kg/m 3

Fiberglass, 66 kg/m 3

0.0400

0.0400

0.0380

0.0380

K, W/m. oC

K, W/m. oC

k = 0.0002 * mc + 0.0343
R2 = 0.88

0.0320

0.0320

0.0360
0.0340

k = 0.0003 * mc + 0.0333
R2 = 0.77

k = 0.0002 * mc + 0.0322
R2 = 0.81

0.0360
0.0340

0.0320

k = 9E-05 * mc + 0.031
R2 = 0.97

0.0320

0.0300
14.0

0.0300
19.0

24.0

29.0

34.0

39.0

0.0

Temperature, oC
Dry_Sample

MC = 6.6%

5.0

10.0

15.0

20.0

25.0

Moisture Content, % (by weight)


MC = 13.8%

T = 14 oC

0.0420

T = 24 oC

T = 34 oC

0.0420
3

Fiberglass, 70 kg/m

Fiberglass, 70 kg/m 3

0.0400

0.0400

0.0380

0.0380

K, W/m. oC

K, W/m. oC

T = 34 oC

0.0420

0.0420

0.0360
0.0340

k = 0.0003 * mc + 0.0336
R2 = 0.99

k = 0.0003 * mc + 0.0324
R2 = 0.97

0.0360

k = 0.0004 * mc+ 0.0306


R2 = 0.99

0.0340
0.0320

0.0320
0.0300
14.0

19.0

24.0

29.0

34.0

0.0300
0.0

39.0

Temperature, oC
Dry_Sample

MC = 3.9%

5.0

10.0

15.0

20.0

25.0

Moisture Content, % (by weight)


T = 14 oC

MC = 13.8%

0.0420

T = 24 oC

T = 34 oC

0.0420
Fiberglass, 84 kg/m

Fiberglass, 84 kg/m 3

0.0400

0.0400

0.0380

0.0380

K, W/m. oC

K, W/m. oC

T = 24 oC

0.0360
0.0340
0.0320
0.0300
14.0

k = 0.0001 * mc + 0.0337
R2 = 0.97

0.0360

k = 0.0001 * mc + 0.0327
R2 = 0.99

0.0340

k = 0.0001 * mc + 0.0317
R2 = 0.80

0.0320

19.0

24.0

29.0

34.0

39.0

Temperature, oC

0.0300
0.0

5.0

10.0

15.0

20.0

25.0

Moisture Content, % (by weight)

Fig. 4. (a) Best-t linear relationship of measurement results of berglass (density 47, 66, 70 and 84 kg/m3) at different percentages of moisture content (MC, % by weight),
(regression data of k-values is given in Table 2), and (b) k-moisture content relationship at 14, 24, and 34 C operating temperatures.

moisture content are illustrative of such behavior with as much as


a 300% increase in k-value when exposed to a moisture content level of around 47%, as can be observed from Table 2.

The linear relationships and the rate of change of k-value with


moisture content for berglass samples with different densities
under operating temperatures of 24 C and 34 C are illustrated

540

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

(a)
0.0460

Dry_Sample

MC = 7.4%

MC = 13.7%

MC = 23.8%

(a)

Rock wool, 46 kg/m3

MC = 22.7%

0.2500

K, W/m. oC

0.0420
0.0400
0.0380
0.0360

0.2000
0.1500
0.1000

0.0340
0.0500

0.0320
0.0300
14.0

0.0000
19.0

24.0

29.0

34.0

39.0

4.0

9.0

14.0

Dry_Sample

(b)
0.0460

MC = 7.1%

19.0

24.0

MC = 12.7%

(b)

Rock wool, 59 kg/m 3

T = 14 C

T = 24 C

K, W/m. oC

0.0420
0.0400
0.0380
0.0360
0.0340

T = 34 C

0.2000
0.1500

k = -0.0003 * mc 2 + 0.0127* mc + 0.0371


R2 = 1
k = -0.0002 * mc 2 + 0.0094 * mc + 0.0358
R2 = 1

0.1000
0.0500

0.0320
19.0

24.0

29.0

Temperature,
Dry_Sample

39.0

Mineral wool,145 kg/m3

0.2500

(c)

34.0

0.3000

0.0440

0.0300
14.0

29.0

Temperature, oC

Temperature, oC

K, W/m. oC

MC = 8.5%

Mineral wool, 145 kg/m3

0.0440

K, W/m. oC

Dry_Sample

0.3000

MC = 7.5%

34.0

39.0

oC

MC = 10.6%

0.0000
0.0

k = -0.0002 * mc2 + 0.0062 * mc + 0.0344


R2 = 1

5.0

10.0

15.0

20.0

25.0

Moisture Content, % (by weight)


MC = 20.2%

0.0600

Fig. 6. (a) Best-t relationship of measurement results of mineral wool (density


145 kg/m3) at different percentages of moisture content (MC, % by weight), and (b)
regression data (non-linear) of k-values at 14, 24, and 34 C operating temperatures.

Rock wool, 64 kg/m3

K, W/m. oC

0.0550
0.0500
0.0450
0.0400
0.0350
0.0300
14.0

19.0

24.0

29.0

34.0

39.0

Temperature, oC

(d)

Dry_Sample

MC = 7.1%

MC = 12.7%

0.0420
Rock wool, 99 kg/m3

K, W/m. oC

0.0400
0.0380
0.0360
0.0340
0.0320
0.0300
14.0

19.0

24.0

29.0

34.0

39.0

Temperature, oC
Fig. 5. Best-t linear relationship of measurement results of rock wool (density (a)
46, (b) 59, (c) 64 and (d) 99 kg/m3) at different percentages of moisture content
(MC, % by weight), (regression data of k-values is given in Table 2).

graphically in Fig. 3. It is evident that higher material density is


generally associated with a higher rate of change in k-value with

the increase in moisture content under different operating temperatures. For example, the rate of change in k-value at 10% moisture
content level under an operating temperature of 24 C is 2.4% for a
27 kg/m3 density berglass sample while it is 9.7% for a 70 kg/m3
density sample, as can be observed from Table 2 and Fig. 6. A similar level of variation is obtained at an operating temperature of
34 C. Discrepancies in measured values or predicted linear behavior of thermal conductivity with moisture content for different
densities may be partially explained by the non-homogeneity of
the material matrix at different densities and/or the inherited
approximation in the linear correlation method utilized.
Comparison of the magnitude and behavior of k-value changes
with moisture content for the different samples as shown in
Fig. 7 indicates considerable variations among the different densities of berglass. Additionally, signicant differences in behavior
and magnitude are attained in samples of the same materials with
similar or different densities when exposed to different initial wetting conditions. This can be attributed to several factors, including
the moisture characteristics of the materials and the internal structure of the insulation sample which determines water absorption/
desorption behavior (i.e. hysteresis), and the amount and distribution of moisture within the material at a given moisture content.
Since the manufacturing processes of thermal insulation materials
produces different material matrix and pore structure, it is expected that different insulation material samples, even from the
same material type, would exhibit different moisture behavior
(i.e. retention and distribution) resulting in different thermal conductivity variations with moisture content.
Referring to Figs. 36 and Tables 2 and 3, the thermal conductivity variation with moisture content generally is noticed to exhibit consistent and stable behavior. However, variation in behavior
is noticed between different materials having the same density and

Table 3
k-Values of berglass, rock wool and mineral wool, with different densities under 14, 24 and 34 C operating temperatures and moisture content levels (MC, % by weight). The corresponding percentage change (%) of k-value with the
increase of MC at 5%, 10%, and 15% are indicated.

Fiberglass

FG-01

27.0

FG-02

47.0

FG-03

66.0

FG-04

70.0

FG-05

84.0

FG-02h

47.0

FG-03h

65.0

FG-05h

83.0

Temperature (C)

k-Value, Dry 0

k-Value at MC %

% Change of k-value at MC %

5%

10%

15%

14
24
34
14
24
34
14
24
34
14
24
34
14
24
34

0.0333
0.0352
0.0371
0.0322
0.0335
0.0348
0.0311
0.0325
0.0339
0.0308
0.0326
0.0338
0.0313
0.0324
0.0336

0.0340
0.0358
0.0375
0.0325
0.0338
0.0352
0.0315
0.0332
0.0349
0.0327
0.0341
0.0352
0.0323
0.0333
0.0344

0.0343
0.0360
0.0377
0.0336
0.0349
0.0361
0.0319
0.0342
0.0364
0.0347
0.0358
0.0368
0.0329
0.0340
0.0350

0.0346
0.0363
0.0380
0.0348
0.0359
0.0370
0.0324
0.0352
0.0380
0.0368
0.0375
0.0384
0.0335
0.0346
0.0357

2.2
1.6
1.2
0.8
1.0
1.3
1.3
2.1
2.9
6.2
4.5
4.2
3.1
2.7
2.4

3.1
2.4
1.8
4.3
4.0
3.8
2.7
5.2
7.6
12.9
9.7
8.9
5.1
4.7
4.4

4.0
3.2
2.4
7.8
7.1
6.4
4.2
8.3
12.2
19.7
15.0
13.6
7.1
6.7
6.3

14
24
34
14
24
34
14
24
34

0.0325
0.0342
0.0358
0.0307
0.0322
0.0338
0.0306
0.0318
0.0329

0.0333
0.0348
0.0363
0.0340
0.0374
0.0409
0.0322
0.0336
0.0351

0.0343
0.0359
0.0375
0.0369
0.0426
0.0483
0.0339
0.0366
0.0394

0.0354
0.0370
0.0387
0.0397
0.0477
0.0557
0.0356
0.0396
0.0436

2.3
1.8
1.4
11.0
16.1
20.8
5.0
5.9
6.7

5.6
5.2
4.8
20.2
32.0
42.7
10.7
15.3
19.7

8.9
8.5
8.2
29.5
47.9
64.6
16.3
24.7
32.6

5%

10%

15%

Rock wool

RW-01

46.0

14
24
34

0.0373
0.0397
0.0422

0.0375
0.0398
0.0422

0.0375
0.0399
0.0423

0.0375
0.0399
0.0424

0.4
0.2
0.1

0.4
0.4
0.3

0.5
0.5
0.5

Mineral wool

MW-01

145.0

14
24
34

0.0344
0.0358
0.0371

0.0548
0.0696
0.0844

0.0682
0.0933
0.1185

0.0798
0.1130
0.2026

59.1
94.6
127.5

98.1
160.9
219.4

131.9
215.9
446.0

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

Sample code

Fiberglassa

Density, (kg/m3)

Material type

Samples conditioned with higher initial moisture contents.

541

542

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544

(a)

27 kg/m
kg/m33

20

66

70

84

(b)

at 24 C

18

% Change of K-value

16
14
12
10
8
6
4

27 kg/m
kg/m33

20

18

% Change of K-value

47

47

66

70

84

at 34 C

16
14
12
10
8
6
4
2

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Moisture Content, % (by weight)

Moisture Content, % (by weight)

Fig. 7. Comparison of the percentage change of k-value of berglass, with different densities with increase of MC (515%), percentage change of k-value under mean
operating temperature of (a) 24 C and (b) 34 C.

exposed to similar conditioning of moisture content. This can be


noticed when comparing berglass sample FG-02 having a density
of 47 kg/m3 and an initial conditioning moisture content of 22.7%,
with the rock wool sample RW-01 having a density of 46 kg/m3
and a moisture content of 23.8%, as given in Table 2. The k-value
of the berglass tends to be more responsive and shows a relatively greater increase with higher moisture content.
The relationship between thermal conductivity and moisture
content can also be affected by the initial moisture content. Table 3
illustrates the measured k-value of ve density-groups of berglass
samples at operating temperatures of 24 C and 34 C and conditioned at different moisture content levels. The predicted linear relationships between thermal conductivity and operating temperature
as well as moisture content for the tested samples are also presented. Higher initial moisture content is always observed to be
associated with higher thermal conductivity at given moisture contents for all densities. The impact of initial moisture content on the
relative increase in the percentage of change of k-value is evident.
Fig. 8 illustrates the relationship between the k-value and the
moisture content for the different berglass samples. It can be noticed that materials having similar density but conditioned at different conditioning (initial) moisture content levels exhibit
different relationships between the k-value and moisture content.
The rate of change in k-value with moisture content is higher at
higher initial moisture content. This is more apparent when comparing the sample with a density of 65 kg/m3 initially conditioned
at 29.6% by weight and the sample of 66 kg/m3 density conditioned
at 13.6% moisture content. From Fig. 8b it can be deduced that the
rate of change of k-value with a moisture content of 29.6% initial
moisture content is 3.2% compared to 0.62% at 13.6% initial moisture content. A similar trend can be observed for berglass of 83
and 84 kg/m3 for initial moisture contents of 20.2% and 13.8%,
respectively. In general, it can be concluded that the thermal conductivity relationship with moisture content is inuenced by the
wetting/drying history of the material and is not specic to a particular moisture content.

4. Conclusions
The impact of moisture content on thermal conductivity was
investigated by measuring the k-values of 11 different brous insulation samples conditioned at different moisture content levels and
operating temperatures. The relationship between k-value and
moisture content at a specied operating temperature was established and results were discussed. Examination of results revealed
that higher operating temperature and higher moisture content are

always associated with higher thermal conductivity for all densities. A more pronounced impact of operating temperature is observed at the lower densities of the berglass material. As the
density increases, the impact of operating temperature on the percentage of change of k-value at different moisture content levels is
generally less, as particularly applicable to berglass. The magnitude of change in k-value is generally higher with higher moisture
content at a given operating temperature. This can be clearly observed in particular at higher density. Moreover, higher density
samples generally exhibit larger changes in k-value at the same
moisture content level. Variation in behavior is noticed between
different materials having the same density and exposed to similar
initial conditioning of moisture content.
A comparison of the magnitude and behavior of the k-value
changes with moisture content for the different samples indicates
considerable variations among different materials. Additionally,
signicant differences in behavior and magnitude are attained in
samples of the same materials with similar or different densities
when exposed to different initial wetting conditions. This can be
attributed to several factors, including the moisture characteristics
of the materials and the internal structure of the insulation sample
which determines water absorption/desorption behavior (i.e. hysteresis), and the amount and distribution of moisture within the
material at a given moisture content. Since the manufacturing processes of thermal insulation materials produces different material
matrix and pore structure, it is expected that different insulation
material samples, even from the same material, would exhibit different moisture behavior (i.e. retention and distribution) resulting
in different k-value variations with moisture content. Higher initial
moisture content is always associated with higher thermal conductivity at given reduced moisture content for all densities. Materials
having similar density but conditioned at different conditioning
(initial) moisture content levels exhibit different relationships between k-value and moisture content. The rate of change in thermal
conductivity with moisture content is higher at higher initial moisture content. Generally, it can be concluded that the k-values relationship with moisture content is inuenced by the wetting/drying
history of the material and is not specic to a particular moisture
content.
The regression relationships associated with the interpretation
of the results reported in this study, would be benecial and would
provide a good instrument to compute the variation of k-value for
a given brous material and boundary conditions. Given the potentially sufcient effect of moisture content on k-value, it is important to investigate and quantify the impact of the k-value change
of berglass due to moisture content on the thermal and energy
performance of the whole building when subjected to hot-humid

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544


3 - initial MC = 22.7%
47 kg/m
kg/m^3

(a)

3 - initial MC = 29.6%
47 kg/m
kg/m^3

12

543

Arabia. The funding and the supporting facilities provided by


KFUPM are highly appreciated.

% Change of k-value

Rate of change = 0.61, 0.67


10

References

8
6
4
2
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Moisture Content, % (by weight)

(b)

66 kg/m^3
kg/m3 - initial MC = 13.6%

3
65 kg/m
kg/m^3
- initial MC = 29.0%

50

Rate of change = 0.62, 3.2

% Change of k-value

45
40
35
30
25
20
15
10
5
0

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Moisture Content, % (by weight)

(c)

3
84 kg/m
kg/m^3
- initial MC = 13.8%

3
83 kg/m
kg/m^3
- initial MC = 20.2%

30

% Change of k-value

Rate of change = 0.40, 1.9


25
20
15
10
5
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Moisture Content, % (by weight)


Fig. 8. The impact of initial condition of moisture content on the percentage change
of k-value of berglass under 24 C operating temperature. Rates of k-value change
are indicated.

climatic conditions. Ongoing research work [36] is examining and


quantifying the magnitude of thermal load and cooling energy due
to higher effective thermal conductivity as a result of increased
moisture content in typical wallroof systems. Furthermore, the
results of this research work indicated signicant impact of operating temperatures and moisture contents on the k-value. This leads
to the need to consider variable k-values of insulation materials in
building energy simulation programs to address extreme variation
in exterior boundary conditions.

Acknowledgments
This work is part of a research Project Grant IN080390 funded
by King Fahd University of Petroleum and Minerals (KFUPM), Saudi

[1] Aldrich DF, Bond RH. Thermal performance of rigid cellular foam insulation at
subfreezing temperatures. In: Thermal performance of the exterior envelopes
of buildings III, ASHRAE/DOE/BTECC/CIBSE conference, Clean Water Beach,
Florida; 1985. p. 35767.
[2] Wilkes KE, Child PW. Thermal performance of berglass and cellulose attic
insulation. In: Thermal performance of the exterior envelopes of buildings V.
ASHRAE/DOE/BTECC/CIBSE conference, Florida; 1992. p. 35767.
[3] Besan RW, Miller E. Thermal resistance of loose ll berglass insulation spaces
heated from below. In: Thermal performance of the exterior envelopes of
buildings II. ASHRAE/DOE conference, Las-Vegas, 1982; p. 72033.
[4] Abdou AA, Budaiwi IM. Comparison of thermal conductivity measurements of
building insulation materials under various operating temperatures. J Build
Phys 2005;29(2):17184.
[5] Budaiwi I, Abdou A, Al-Homoud M. Variation of thermal conductivity of
insulation materials under different operating temperatures: impact on
envelope-induced cooling load. J Arch Eng 2002;8(4):12532.
[6] Al-Ajlan S. Measurements of thermal properties of insulation materials by
using transient plane source technique. Appl Therm Eng 2006;26(2):218491.
[7] Bo-Ming Zhang, Zhao Shu-Yuan, He Xiao-Dong. Experimental and theoretical
studies of high-temperature thermal properties of brous insulation. J Quant
Spectrosc Radiat Transfer 2008;109:130924.
[8] Kochhar GS, Manohar K. Effect of moisture on thermal conductivity of bers
biological insulating materials. In: Thermal performance of the exterior
envelopes of building VI. ASHRAE/DOE conference, Florida; 1995. p. 3340.
[9] Benner SM, Luu DV. Thermal mass-transfer coefcient and equilibrium
moisture content of insulating materials. In: Thermal performance of the
exterior envelopes of buildings II, ASHRAE/DOE conference, Las Vegas; 1982. p.
105262.
[10] Hedlin CP. Heat ow through a roof insulation having moisture contents
between 0 and 1% by volume, in summer. ASHRAE Trans 1988;94(2):
5791594.
[11] Chyu MC, Zeng X, Ye L. Performance of bers glass pipe insulation subjected to
underground water attack. ASHRAE Trans 1997;103(1):3038.
[12] Chyu MC, Zeng X, Ye L. Effect of underground water attack on the performance
of mineral wool pipe insulation. ASHRAE Trans 1998;104(2):16875.
[13] Wijesundera NE, Hawlader MN, Lian SC. Experimental study of condensation
in berglass insulation. In: ASHRAE far east conference on air conditioning in
hot climates, Kuala lumpur, Malaysia; 1989. p. 13539.
[14] Sandberg PI. Determination of the effects of moisture on the thermal
transmissivity of cellulose ber loose-ll insulation. In: Thermal
performance of the exterior envelopes of building V., ASHRAE/DOE/BTECC/
CIBSE conference, Clearwater Beach, Florida; 1992. p. 51725.
[15] Larson DC, Benner SM. Field and laboratory studies of the thermal resistance of
moist building insulation systems. In: Thermal performance of the exterior
envelopes of building III, ASHRAE/DOE/BTECC conference, Florida; 1985. p.
52738.
[16] Simonson CJ, Tao YX, Besant RW. Thermal performance and hysteresis in
brous insulation exposed to moisture and step changes in boundary
condition. Energy Build 1994;21:2517.
[17] Peuhkuri R. Moisture dynamics in building envelopes. PhD thesis, Report R071, Department of Civil Engineering, Technical University of Denmark,
Denmark; 2003.
[18] Wijesundera NE, Howlander MNA. Effects of condensation and liquid transport
on the thermal performance on brous insulations. Int J Heat Mass Transfer
1992;35:260516.
[19] Wijesundera NE, Cheng BF, Hauptmann EG. Numerical simulation of the
transient moisture transfer through porous insulation. Int J Heat Mass Transfer
1996;39:9951004.
[20] Jintu F, Xinhuo W. Modeling heat and moisture transfer through brous
insulation with phase change and mobile condensates. Int J Heat Mass Transfer
2002;45:404555.
[21] Jintu F, Xiaoyin C, Xinhuo W, Weiwei S. An improved model of heat and
moisture transfer with phase change and mobile condensates in brous
insulation and comparison with experimental results. Int J Heat Mass Transfer
2004;47:234352.
[22] Tomas V, Kjartan G. Comparison of brous insulations cellulose and stone
wool in terms of moisture properties resulting from condensation and ice
formation. Constr Build Mater 2010;24:11517.
[23] Cabeza LF, Castell A, Medrano M, Martorell I, Perez G, Fernandez I.
Experimental study on the performance of insulation materials in
Mediterranean construction. Energy Build 2010;42:6306.
[24] Uros L, Saso M. Heat and moisture transfer in brous thermal insulation with
tight boundaries and a dynamical boundary temperature. Int J Heat Mass
Transfer 2011;54:433340.
[25] Bjorn PJ. Review: traditional, state-of-the-art and future thermal building
insulation materials and solutions properties, requirements and possibilities.
Energy Build 2011;43(10):254963.
[26] Anastasios K, Agis P, Dimitrios A. Heat transfer phenomena in brous
insulating materials. Laboratory of Heat Transfer and Environmental
Engineering, Department of Mechanical Engineering Aristotle University

544

[27]
[28]
[29]

[30]

[31]

A. Abdou, I. Budaiwi / Construction and Building Materials 43 (2013) 533544


Thessaloniki, GR-54124 Thessaloniki, Greece; 2004. URL: <http://
www.aix.meng.auth.gr>;
<http://www.geolan.gr/sappek/docs/publications/
article_6.pdf>.
Forschungsinstitut fr Wrmeschutz (FIW, Research Institute for Thermal
Protection): <http://www.w-muenchen.de/en_thew.php>.
KingspanTarec: High Performance Insulation: <http://www.kingspantarec.
com/tarecUK/performance_kooltherm.htm>.
Carey JS, Yong XT, Robert WB. Thermal performance and hysteresis in brous
insulation exposed to moisture and step changes in the cold temperature
boundary condition. Energy Build 1994;21:2517.
Ochs F, Heidemann W, Muller-Steinhagen H. Effective thermal conductivity of
moistened insulation materials as a function of temperature. Int J Heat Mass
Transfer 2008;51:53952.
Krishpersad M, Gurmohan SK. Experimental investigation of the inuence of
air conduction on heat transfer across brous materials. J Mech Eng Res
2011;3(9):31924.

[32] Milos J, Robert C. Effect of moisture content on heat and moisture transport
and storage properties of thermal insulation materials. Energy Build
2012;53:3946.
[33] Lambda 2000 series. Heat ow meter thermal conductivity instrumentation,
operation manual. USA: Holometrix Inc.; 1997.
[34] ASTM C518-91. Standard test method for steady-state heat ux measurements
and thermal transmission properties by means of ow meter apparatus. PA,
USA: ASTM International; 1991.
[35] ISO 8301. Thermal insulation determination of steady-state thermal
resistance and related properties heat ow meter apparatus, ISO Geneva;
1991.
[36] Budaiwi I, Abdou A. The impact of thermal conductivity change of moist
brous insulation on energy performance of buildings under hot-humid
conditions. Energy Build 2013;60:38899.

S-ar putea să vă placă și