• Aucun résultat trouvé

Numerical investigation of thermal response of basement wall systems with low emissivity material and furred airspace

N/A
N/A
Protected

Academic year: 2021

Partager "Numerical investigation of thermal response of basement wall systems with low emissivity material and furred airspace"

Copied!
16
0
0

Texte intégral

(1)

Publisher’s version / Version de l'éditeur:

Vous avez des questions? Nous pouvons vous aider. Pour communiquer directement avec un auteur, consultez la première page de la revue dans laquelle son article a été publié afin de trouver ses coordonnées. Si vous n’arrivez pas à les repérer, communiquez avec nous à PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca.

Questions? Contact the NRC Publications Archive team at

PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca. If you wish to email the authors directly, please see the first page of the publication for their contact information.

https://publications-cnrc.canada.ca/fra/droits

L’accès à ce site Web et l’utilisation de son contenu sont assujettis aux conditions présentées dans le site LISEZ CES CONDITIONS ATTENTIVEMENT AVANT D’UTILISER CE SITE WEB.

Proceedings of the 13th Canadian conference on Building Science and

Technology, Winnipeg, MB, Canada, 10–13 May, 2011-05-10

READ THESE TERMS AND CONDITIONS CAREFULLY BEFORE USING THIS WEBSITE.

https://nrc-publications.canada.ca/eng/copyright

NRC Publications Archive Record / Notice des Archives des publications du CNRC :

https://nrc-publications.canada.ca/eng/view/object/?id=d0606ede-8e51-4bbc-8266-4c0102a61915 https://publications-cnrc.canada.ca/fra/voir/objet/?id=d0606ede-8e51-4bbc-8266-4c0102a61915

NRC Publications Archive

Archives des publications du CNRC

This publication could be one of several versions: author’s original, accepted manuscript or the publisher’s version. / La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur.

Access and use of this website and the material on it are subject to the Terms and Conditions set forth at

Numerical investigation of thermal response of basement wall systems

with low emissivity material and furred airspace

(2)

http://www.nrc-cnrc.gc.ca/irc

N um e ric a l inve st iga t ion of t he rm a l re sponse of ba se m e nt w a ll

syst e m s w it h low e m issivit y m a t e ria l a nd furre d a irspa c e

N R C C - 5 4 4 3 0

S a b e r , H . H . ; M a r e f , W . ; S w i n t o n , M . C .

M a y 2 0 1 1

A version of this document is published in / Une version de ce document se trouve dans:

13th Canadian Conference on Building Science and Technology

Conference,Winnipeg, Manitoba, Canada, May 10-13, 2011, pp. 1-13

The material in this document is covered by the provisions of the Copyright Act, by Canadian laws, policies, regulations and international agreements. Such provisions serve to identify the information source and, in specific instances, to prohibit reproduction of materials without written permission. For more information visit http://laws.justice.gc.ca/en/showtdm/cs/C-42

Les renseignements dans ce document sont protégés par la Loi sur le droit d'auteur, par les lois, les politiques et les règlements du Canada et des accords internationaux. Ces dispositions permettent d'identifier la source de l'information et, dans certains cas, d'interdire la copie de documents sans permission écrite. Pour obtenir de plus amples renseignements : http://lois.justice.gc.ca/fr/showtdm/cs/C-42

(3)
(4)

NUMERICAL INVESTIGATION OF THERMAL RESPONSE OF

BASEMENT WALL SYSTEMS WITH LOW EMISSIVITY MATERIAL

AND FURRED AIRSPACE

Hamed H. Saber*, Wahid Maref, and Michael C. Swinton

Building Envelope and Structure Program

Institute for Research in Construction, National Research Council Canada Bldg. M-24, 1200 Montreal Road, Ottawa, Ontario, Canada K1A 0R6 * Corresponding Author (hamed.saber@nrc-cnrc.gc.ca, phone: 613-993-9772)

ABSTRACT

In basement wall systems, airspaces can contribute in obtaining a higher thermal resistance, if a reflective material such as reflective foil is installed on one side or the other of a furred-airspace. In this paper, the hygIRC-C model was used to investigate the steady-state and transient thermal performance of a basement wall system that incorporate foil bonded to expanded polystyrene (EPS) foam in a furred-assembly having airspace next to the foil. The furring was installed horizontally. The external layer of this wall is a poured-in-place concrete. Walls with and without Furred-Airspace Assembly (FAA) were considered in this study. Also, consideration was given to investigate the effect the above- and below-grade portions of the wall on the thermal performance when these walls were subjected to a Canadian climate. Results showed that the effective thermal resistance of a wall with FAA at steady-state condition depends on the soil, outdoor and indoor temperatures. Additionally, a wall with FAA and low foil emissivity (0.04) bonded to EPS foam resulted in an energy saving of 17.7% compared to a wall without FAA when these walls are subjected to the same climate condition.

KEYWORDS: Basement wall, above-grade, below-grade, furred-airspace assembly, low emissivity material, thermal modelling, thermal resistance, R-value, heat transfer by natural convection and radiation.

INTRODUCTION

Unlike many parts of the U.S. and other warmer countries, where the basement is considered to be outside the building envelope, in Canada the basement is presumed to be inside the envelope. Over the past decades, a large proportion of basements in newly built Canadian houses have become used as habitable space, not only usable space. This expectation continues to drive builder marketing and Energy Code requirements for new housing in most, if not all, regions of Canada. Although not necessarily lived in, the basement spaces (and heated crawl spaces) are connected to the above-ground spaces through passageways and by air circulation ducts in houses with forced-air systems. Indoor air, including its relative humidity, temperature and its contaminants, is shared with above-ground space. The National Building Code (NBC) of Canada governs minimum requirements for basements spaces and recognizes this feature [1].

The performance requirements for basements and basement guidelines have been detailed in earlier studies [2 – 11]. Because the heat loss from basements accounts for such a significant portion of the energy loss from a home, it is important that basement walls must be insulated. Timusk [4] showed that heat loss from an un-insulated basement can account for up to one third of the heating cost in an average home. This varies depending on many factors, such as the air tightness of the building envelope, the amount of insulation in the house and the height of the above-grade and below-grade portions of the

(5)

basement wall systems [4]. By improving basement thermal performance, the net cost of operating homes can be reduced.

Thermal insulations were applied on basement walls either internally or externally. Some of the reluctance to use external insulation may be simple human nature. Internally insulated basements are often coupled with interior basement finishing and therefore offer a “higher perceived value” to the homebuyer [11]. Many homeowners prefer finished basements and basements with insulation on the outside just don’t look complete. As such, the focus of this paper is on internally insulated basement walls where the thermal resistance of a basement wall was increased by incorporating foil bonded to expanded polystyrene (EPS) foam in a furred-assembly having airspace next to the foil.

In this study, the NRC-IRC’s hygrothermal model “hygIRC-C” is used to investigate both steady-state and transient thermal performance of basement wall systems. This model solves the 2D and 3D Heat, Air and Moisture (HAM) transport equations. This model was benchmarked against the hygIRC-2D model that was previously developed at NRC-IRC, and test results in a number of projects. In the case of accounting for heat, air and moisture transport, the 2D version of the hygIRC-C model was used to predict the drying rate of a number of full-scale wall assemblies subjected to different time dependent exterior and interior boundary conditions [12]. The results showed that the overall agreement between the hygIRC-C model and the hygIRC-2D model [13 and 14] as well as experimental measurements were good. In the case of accounting for heat and air transport (no moisture transport), the 3D version of the hygIRC-C model was used to conduct numerical simulations for different full-scale wall assemblies with and without penetration to represent a window in order to predict the effective thermal resistance (R-value) with and without air leakage [15]. The predicted R-values for these walls were in good agreement (within ±5%) with the measured R-values in Guarded Hot Box (GHB) [16 and 17]. Also, the hygIRC-C model was used to assess the dynamic heat transmission characteristics through two Insulating Concrete Form (ICF) wall specimens installed in the NRC-IRC’s Field Exposure of Walls Facility (FEWF). The results showed that the hygIRC-C model predictions were in good agreement with experimental data [18]. For foundation wall systems, airspace can contribute in obtaining a higher R-value, if a reflective material such as aluminum foil is installed on one side or both sides of a furred airspace. The impact of foil emissivity on the wall R-value was examined in a previous study [19]. In that study, the 2D version of the hygIRC-C model was used to conduct sensitivity analyses in order to investigate the effect of foil emissivity of foil laminated to XPS foam when used within a furred-airspace assembly. In that work, furring strips made of spruce (19 mm x 38 mm) were installed horizontally. Since there were no vertical studs in the wall assembly, the 2D version of the hygIRC-C model was suitable for that study. The results showed that the modelled foundation wall system with foil of emissivity 0.05 increased the effective R-value by ~10% in the case of the indoor and outdoor temperatures of +20oC and -20oC, respectively [19]. Recently, the 3D version of the hygIRC-C model was benchmarked against the experimental data of a full-scale above-grade wall system (2x6 wood frame construction with stud cavities filled with friction-fit glass fibre batt insulation) with foil bonded to wood fibreboard in a furred-assembly having airspace next to the foil [20]. Results showed that the predicted R-value of this wall specimen was in good agreement with the measured value [20]. Thereafter, the hygIRC-C model was used to quantify the contribution of low foil emissivity to the wall R-value.

The objective of this paper is to use the hygIRC-C model to investigate the change in the effective R-value at steady-state conditions and the transient thermal response of full-scale foundation wall systems (including the above-grade and below-grade portions of the wall) with low emissivity material and furred-airspace assemblies when these walls are subjected to a Canadian climate. Additionally, the hygIRC-C model is used to determine the energy savings due to having a foundation wall system with a furred-airspace assembly compared to a foundation wall system without a furred-furred-airspace assembly. Since no moisture transport is accounted for in this study, the hygIRC-C model solves simultaneously the energy

(6)

equation in the various material layers, surface-to-surface radiation equation in the furred-airspace assembly, Navier-Stokes equation for the airspace, and Darcy equation (Darcy number < 10-6) and Brinkman equation (Darcy number > 10-6) for the porous material layers. The full description of the hygIRC-C model is available in [19] and [20].

Figure 1. Schematic showing the boundary conditions (B.C.’s) of above- and below-grade basement wall with foil bonded to EPS in furred-airspace assembly (Wall-FAA)

WALL DESCRIPTION, ASSUMPTIONS, AND INITIAL AND BOUNDARY

CONDITIONS

Figure 1 shows a basement wall system with foil bonded to expanded polystyrene (EPS) foam in a furred-assembly having airspace next to the foil referred to as Wall-FAA. The foil has an emissivity of 0.04 [20]. The furred-airspace assembly consists of 19 mm x 64 mm wood furring strips installed horizontally at 16” (406 mm) center-to-center and was closed with a gypsum board (12.7 mm thick). The external layer of this wall is a poured-in-place concrete (200 mm thick). The basement wall was modeled with above-grade and below grade heights of 598 mm and 1,840 mm, respectively. As shown in Figure 1, the soil temperature was measured in a previous experimental study at ~2 m away from the wall at 5 different elevations [21, 22, 23]. Those filed measurements were used as boundary conditions in this modelling study. In order to quantify the contribution of having a furred-airspace assembly with foil bonded to EPS foam on the energy savings, a reference wall was considered. This reference wall is referred to as Wall-R, which is identical to Wall-FAA, but without furred-airspace assembly. It was assumed that all material layers and the soil are in good contact. Since there were no vertical studs in the wall assembly, the 2D version of the hygIRC-C model is suitable for this study.

Soil (2000 mm thick)

C

o

n

cret

e

L

a

y

er

(2

0

0

m

m

t

h

ick

)

EP S i n su la ti o n b o a r d (3 , 7 6 .2 m m th ic k ) w ith l o w fo r e m is si v ity m a te ri a l (0 .0 4 )

y

x

500 m m 1 100 m m 1 570 m m 1690 m m 1 840 m m

Measured soil temperature

Adiabatic (insulation) Adiabatic (insulation) TSo il (t , x= 0, y )

Convective B.C., h and Toutdoor(t)

Co n v ec ti v e B. C., h and T ou td oo r (t ) C o nve ct ive B .C . w it h h an d Tin d oor (t) 5 98 m m G y ps um boar d ( 12. 7 m m t h ic k) E P S ( 76. 2 m m t h ic k) Fur ri n g (1 9 x 6 4 m m ) Spac ing of f u rr in g = 16” o. c. Foi l A irsp a ce (1 9 m m t h ic k )

(7)

Figure 2. Measurements of soil, outdoor and indoor temperatures [21, 22, 23] -30 -25 -20 -15 -10 -5 0 5 10 15 20 25 30 35 0 90 180 270 360 450 540 630 720 810 Indoor Outdoor (b)

Measurements were taken every 6 hrs

Time = 0 corresponding to June 5, 1996 @ 15:19:04

Time (Day)

Air Temperature (

o

C)

0 5 10 15 20 25 0 90 180 270 360 450 540 630 720 810 y = 1690 mm y = 1570 mm y = 1100 mm y = 500 mm y = 0 mm (a) Time = 0 corresponding to June 5, 1996 @ 15:19:04

Soil measurements were taken at 2 m away from the wall every 6 hrs

Soil Temperature (

(8)

An initial and uniform temperature of 10oC was assumed in the different components of wall systems and soil. The boundary conditions used in the numerical simulations are shown in Figure 1. At the left boundary of the soil (2 m away from the wall), the temperature was measured at five locations (at y = 0, 500, 1100, 1570 and 1690 mm) and shown in Figure 2a. These measured temperatures were taken as temperature boundary condition at the left boundary of the soil. Between these locations, however, the temperature of the soil at the left boundary was determined from the measured temperatures by linear interpolation. Both the top boundary of the soil and the left boundary of the above-grade portion of the wall system are subjected to convective boundary conditions with a heat transfer coefficient of 22.1 W/(m2K) and air temperature equal to the measured outdoor temperature (Figure 2b, [21, 22, 23]). Similarly, the interior surface of the gypsum board is subjected to a convective boundary condition with heat transfer coefficient of 7.11 W/(m2K) and air temperature equal to the measured indoor temperature (Figure 2b, [21, 22, 23]). The top boundary of the wall was assumed adiabatic and sealed (no heat and mass transfer).

The boundary condition on the bottom boundary of the wall and soil are assumed to be adiabatic. During some periods within the year, the bottom boundary may lose/gain heat. As such, the predicted thermal performance of different wall systems in this study is a qualitative performance. To predict the quantitative thermal performance, however, the soil temperature under the wall systems needs to be measured and taken as a temperature boundary condition instead of applying adiabatic condition at the bottom boundary of the wall systems. On the other hand, these temperatures were not measured. Because all boundary conditions described in this section for a wall with a furred-airspace assembly (Wall-FAA) are taken to be the same as for a wall without a furred-airspace assembly (reference wall: Wall-R), it is proposed that the difference in the thermal performance of these walls reflects the contribution due to the furred-airspace assembly.

RESULTS AND DISCUSSIONS

It was shown in previous studies for above-grade wall systems [19 and 20] that a foil with lower emissivity has two interactive and competing effects on the wall R-value, namely: (i) an increase in wall R-value due to lower net radiative heat flux in the furred-airspace assembly, and (ii) a decrease in R-value due to stronger convection currents in the airspace caused by larger temperature gradient across the airspace. The former effect outweighs the latter effect, resulting in a net increase in the effective R-value for a wall system with low foil emissivity. In this study, the effect of both above- and below-grade portions of the wall on the effective R-value as well as the transient thermal response of walls with and without furred-airspace assembly are investigated. The numerical simulations were conducted for a period of two years for all wall systems. The temperature measurements of the soil, outdoor and indoor that were used in the simulations are shown in Figure 2. In these analyses, time = 0 corresponds to June 5, 1996 at 15:19:04.

Steady state calculation of the effective R‐value of the wall as a function of the 

temperature

 gradients across the airspace

The effective R-value of the wall with furred-airspace assembly (Wall-FAA) was calculated at steady-state conditions during the first year. During this year, average temperature measurements of soil, outdoor and indoor were taken at two weeks intervals. Figure 3 shows an example of the average temperatures of soil, outdoor and indoor at t = 7 day (averaging within t = 0 – 14 day: Jun 5, 1996 – Jun 19, 1996), t = 147 day (averaging within t = 140 – 238 day: Oct 23, 1996 – Nov 6, 1996), and t = 231 day (averaging within t = 224 – 238 day: Jan 15, 1997 – Jan 29, 1997). The temperature contours in the furred-airspace assembly (FAA) are shown in Figure 4a, b, and c and the corresponding vertical velocity contours in the six airspaces of the FAA are shown in Figure 5a, b, and c at t = 7, 147 and 231

(9)

day (Jun 12, 1996, Oct 30, 1996 and Jan 22, 1997), respectively. Note that different scales were used for the temperature contours and velocity contours shown in these figures.

In the above-grade portion of the wall at t = 7 day (Jun 12, 1996), the outdoor temperature (22.6oC) is close to the indoor temperature (20.7oC). As such, the temperature distribution in the upper two airspaces is approximately uniform, resulting in no convection current in these airspaces (see Figure 5a). In the below-grade portion of the wall at this time (t = 7 day), however, the lowest soil temperature occurred in the lower portion of the wall (see Figure 3). As such, the temperature gradient across the furred-airspace assembly decreases with the distance along the wall (Figure 4a), resulting in a stronger convection current in the lower airspaces than that in the upper airspaces (Figure 5a). Conversely, in the above-grade portion of the wall at t = 147 day (Oct 30, 1996), the outdoor temperature (5.4oC) is lower than the indoor temperature (22.2oC) resulting in a high temperature gradient across the upper portion of the furred-airspace assembly (Figure 4b). As such, strong convection current occurred in the upper two airspaces (Figure 5b). In the below-grade portion of the wall at this time (t = 147 day), the highest soil temperature occurred at the bottom of the wall (Figure 3). As such, the temperature gradient across the furred-airspace assembly increases with the distance along the wall (Figure 4b), resulting in a stronger convection current in the upper airspaces than that in the lower airspaces of the below-grade portion of the wall (Figure 5b).

Figure 3. Average temperatures of soil, outdoor and indoor (averaging within 2 week) at different time during the 1st year (Wall-FAA)

0 2 4 6 8 10 12 14 16 18 20 0 0.5 1.0 1.5 2.0

At t = 231 day, Toutdoor,avg = -12.9oC, Tindoor,avg = 21.5oC At t = 147 day, Toutdoor,avg = 5.4oC, Tindoor,avg = 22.2oC At t = 7 day, Toutdoor,avg = 22.6oC, Tindoor,avg = 20.7oC

t = 231 d ay (Jan 2 2, 1997) t = 147 day (Oct 30, 1996) t = 7 da y (Ju n 12, 1996 ) t = 0 corresponding to June 5, 1996 @ 15:19:04 y = 0 at the bottom of the wall

(c) Average soil temperature measurements within t = 224 - 238 day (Jan 15, 1997 - Jan 29, 1997)

(b) Average soil temperature measurements within t = 140 - 154 day (Oct 23, 1996 - Nov 6, 1996)

(a) Average soil temperature measurements within t = 0 - 14 day (Jun 5, 1996 - Jun 19,1996)

y (m) Ave rag e s o il te m per atu re at 2 m away fr om the wall ( o C)

(10)

Figure 4. Temperature contours in the furred-airspace assembly (in oC) at different time for Wall-FAA d ryw al l EPS dryw al l EPS dryw al l EPS 19 mm 243 8 m m

(c) Time = 231 day (Jan 22, 1997)

(11)

Figure 5. Vertical velocity contours in the airspace (in mm/s) at different time for Wall-FAA

(c) Time = 231 day (Jan 22, 1997)

(a) Time = 7 day (Jun 12, 1996)

(b) Time = 147 day (Oct 30, 1996)

19 mm

24

38 m

(12)

Figure 6. Dependence of the effective R-value on the climate conditions for Wall-FAA  2.72 2.74 2.76 2.78 2.80 2.82 0 30 60 90 120 150 180 210 240 270 300 330 360 390-30 -25 -20 -15 -10 -5 0 5 10 15 20 25

Indoor Temperature

Outdoor Temperature

R-Value

J u n 1 2 , 1 9 9 6 O c t 3 0 , 1 9 9 6 J a n 2 2 , 1 9 9 7 M a y 1 , 1 9 9 7

Time = 0 corresponding to June 5, 1996 @ 15:19:04

Time (Day)

Effe

ctive R-Va

lue

(m

2

K/W)

Ai

r T

e

m

p

er

atur

e (

o

C)

(13)

A similar trend in terms of the temperature and velocity distributions in the furred-airspace assembly occurred in both the above- and below-grade portions of the wall at t = 231 day (Jan 22, 1997) as that at t = 147 day. However, because both outdoor and soil temperatures at t = 231 day are lower than that at t = 147 day, the convection current in the airspaces at t = 231 day is stronger than that at t = 147 day (see Figure 5b and Figure 5c). For example, the highest vertical upward velocity at t = 231 day is 38.91 mm/s, which is about 71% higher than that at t = 147 day (22.69 mm/s). As such, the thermal conductance of the furred-airspace assembly at t = 231 day is greater than that at t = 147 day, resulting in a lower R-value at the former than at the latter (see Figure 6).

Figure 6 shows the dependence of the effective R-value of Wall-FAA on the environmental conditions (soil, outdoor and indoor temperatures) during the first year of simulation at steady-state conditions. As indicated earlier, the temperature measurements of the soil, outdoor and indoor were averaged over a 2 week periods. As shown in Figure 6, the lower temperature difference across the wall results in a higher effective R-value. The highest effective R-value of this wall was 2.81 m2K/W which occurred at t = 91 day. Additionally, the lowest effective R-value (2.73 m2K/W) occurred at t = 231 day, at which Wall-FAA was subjected to the highest temperature difference across it. During the first year, the maximum change in the effective R-value was ~3% (see Figure 6). The R-value for the reference Wall-R (same as Wall-FAA but without FAA) is 2.28 m2K/W. The lowest and highest R-values of Wall-FAA are about 20% and 23% higher than that of Wall-R. As such, it is expected that the heat loss from Wall-FAA will be smaller than that from Wall-R as explained below.

Transient simulation of the wall thermal response with contribution of the 

furred

 airspace and low emissivity foil

Two transient numerical simulations were conducted for a period of two years in order quantify the contribution of having furred-airspace assembly with low foil emissivity (0.04) bonded to EPS foam on the energy saving. These simulations were conducted for Wall-FAA and Wall-R (same as Wall-FAA but without FAA) using the same soil, outdoor and indoor temperatures shown in Figure 2. Figure 7a shows the rate of heat loss/gain per unit width of the wall (in KJ/(d.m)) from the interior surface of the drywall of Wall-FAA and Wall-R. In this figure, positive and negative heat rate represent heat gain and heat loss, respectively. As shown in this figure, the rate of heat loss from wall with FAA (Wall-FAA) is lower than that for wall without FAA (Wall-R). For example, the heat loss from Wall-FAA at t = 270 day is 1510 KJ/(d.m), which is about ~16% lower than that for Wall-R (1745 KJ/(d.m)). Furthermore, for both walls, the heat gain during two years is approximately zero (i.e. no cooling load is needed).

The rate of heat loss/gain shown in Figure 7a was used to derive the accumulative energy loss/gain per unit width of the wall (in MJ/m) over a two years period. This result is shown in Figure 7b. During the two years of simulation, this figure clearly shows that the energy loss from the wall with FAA is always lower than that for the wall without FAA. At the end of the second year, the energy loss from Wall-R was 771 MJ/m, which is 17.7% higher than that of Wall-FAA (655 MJ/m, see Figure 7a). As such, wall with FAA and low foil emissivity (0.04) bonded to EPS foam results in an energy savings of 17.7% compared to the wall without FAA when these walls were subjected to the same climate conditions.

(14)

Figure 7. Comparison between heat loss in wall with furred-airspace assembly (Wall-FAA) and wall without furred-airspace assembly (Wall-R).

-800 -700 -600 -500 -400 -300 -200 -100 0 0 90 180 270 360 450 540 630 720 Wall-FAA Wall-R (b)

Time = 0 corresponding to June 5, 1996 @ 15:19:04

Heal loss in wall without FAA is 17.7% higher than wall with FAA

Time (day)

Energy Loss/G

a

in (M

J/

m)

-3000 -2500 -2000 -1500 -1000 -500 0 500 0 90 180 270 360 450 540 630 720 Wall-FAA Wall-R Heat Gain Heat Loss (a) Time = 0 corresponding to June 5, 1996 @ 15:19:04

H

e

at Loss/Gai

n

(15)

SUMMARY AND CONCLUSIONS

The hygIRC-C model was used to conduct numerical simulations in order to investigate the steady-state and transient thermal performance of a foundation wall system (including the above-grade and below-grade portions of the wall) having a Furred-Airspace Assembly (FAA) and incorporating low emissivity foil material. This model was benchmarked in a previous study by comparing its predictions against experimental data generated by a commercial laboratory for an above-grade wall assembly. The external layer of the foundation wall is poured-in-place concrete and the internal layer is gypsum board. In order to quantify the contribution of a FAA with foil bonded to EPS foam on energy savings in foundation wall systems (Wall-FAA), a reference wall (Wall-R) was considered in this study. This wall is identical to Wall-FAA but without a FAA. Walls with and without FAA were subjected to the same climate load where the measurements of soil temperature (2 m away from the wall), outdoor temperature and indoor temperature were used. Results showed that at steady-state condition, the effective R-value of wall with FAA can vary by as much as ~3%, depending on the soil, outdoor and indoor temperatures. Moreover, a wall with FAA and incorporating low foil emissivity (0.04) bonded to EPS foam resulted in an energy savings of ~17.7% compared to a wall without FAA. This is on-going research. The hygIRC-C model is being used to investigate the transient thermal response of foundation wall systems with furring installed horizontally and vertically, and subjected to different Canadian climate conditions. The results of this effort will be reported at a later date.

REFERENCES

1. CMHC-Research Highlight, Technical Series 06-109 “Occupancy-based Classification for Design and Construction of Residential Basements”, August 2006.

2. Kesik, T., Basement System Performance, Proceedings, CSCE 7th Conference on Building Science and Technology, March 20 & 21, 1997, Toronto, pp.227-241.

3. Swinton, M.C., Kesik, T.J., "Performance Guidelines for Basement Envelope Systems and Materials : Final Research Report”, Research Report, NRC Institute for Research in Construction, (199), pp. 193. 2005-10-01.

4. Timusk, J., 1981. Insulation Retrofit of Masonry Basements. Department of Civil Engineering, University of Toronto, Toronto, Canada.

5. Swinton, M.C., Bomberg, M.T., Maref, W., Normandin, N., Marchand, R.G. In-Situ

Performance Evaluation of Exterior Insulation Basement System (EIBS) - EPS Specimens. Institute for Research in Construction, NRCC, Ottawa, 1999 (A-3132.1).

6. Swinton, M.C., Bomberg, M.T., Maref, W., Normandin, N., Marchand, R.G. In-Situ

Performance Evaluation of Exterior Insulation Basement System (EIBS) . Spray Polyurethane Foam. Institute for Research in Construction, NRCC, Ottawa, 2000 (A-3132.3).

7. Swinton, M.C., Bomberg, M.T., Maref, W., Normandin, N., Marchand, R.G. In-Situ

Performance Evaluation of Exterior Insulation Basement System (EIBS) - Glass Fibre Specimens. Institute for Research in Construction, NRCC, Ottawa, 2000 (A-3129.1).

8. Swinton, M.C., Bomberg, M.T., Maref, W., Normandin, N., Marchand, R.G. In-Situ

Performance Evaluation of Exterior Insulation Basement System (EIBS) - Mineral Fibre Specimens. Institute for Research in Construction, NRCC, Ottawa, 2000 (A-3129.2).

9. Swinton, M.C., Bomberg, M.T., Kumaran, M.K., Normandin, N. and Maref, W. Performance of Thermal Insulation on the Exterior of Basement Walls. Construction Technology Update No. 36, Institute for Research in Construction, NRCC, Ottawa, 1999.

10. Kesik, T., Economic Assessment of Basement Systems, Contract Report No. 338326, Institute for Research in Construction, National Research Council of Canada, Ottawa, March 2000.

11. Yost, N. and Lstiburek, J., “Basement Insulation Systems”, Research Report – 0202 by Building Science Corporation, PP. 21, 2002.

(16)

12. Saber, H.H., Maref, W., Lacasse, M.A., Swinton, M.C., and Kumaran, M.K., “Benchmarking of hygrothermal model against measurements of drying of full-scale wall assemblies”, 2010 International Conference on Building Envelope Systems and Technologies, ICBEST 2010, Vancouver, British Colombia Canada, June 27-30, 2010, pp. 369-377.

13. Maref, W., Kumaran, M.K., Lacasse, M.A., Swinton, M.C., and van Reenen, D., “Laboratory measurements and benchmarking of an advanced hygrothermal model”, Proceedings of the 12th International Heat Transfer Conference (Grenoble, France, August 18, 2002), pp. 117-122, October 01, 2002 (NRCC-43054).

14. Maref, W., Lacasse, M.A., Kumaran, M.K., and Swinton, M.C., “Benchmarking of the advanced hygrothermal model-hygIRC with mid-scale experiments”, eSim 2002 Proceedings (University of Concordia, Montreal, September 12, 2002), pp. 171-176, October 01, 2002 (NRCC-43970). 15. Elmahdy, A.H., Maref, W., Swinton, M.C., Saber, H.H., Glazer, R., “Development of Energy

Ratings for Insulated Wall Assemblies”, 2009 Building Envelope Symposium (San Diego, CA. 2009-10-26) pp. 21-30, 2009.

16. Saber, H.H., Maref, W., Elmahdy, A.H., Swinton, M.C., and Glazer, R., “3D thermal model for predicting the thermal resistances of spray polyurethane foam wall assemblies”, Building XI Conference, December 5-9, 2010, Clearwater Beach, Florida, USA.

17. Saber, H.H., Maref, W., Elmahdy, H., Swinton, M.C., and Glazer, R. “3D Heat and Air Transport Model for Predicting the Thermal Resistances of Insulated Wall Assemblies”, Journal of Building Performance Simulation, (http://dx.doi.org/10.1080/19401493.2010.532568), First published on: 24 January 2011 (iFirst), pp. 1-17.

18. Saber, H.H., Maref, W., Armstrong, M., Swinton, M.C., Rousseau, M.Z., and Ganapathy, G., “Benchmarking 3D thermal model against field measurement on the thermal response of an insulating concrete form (ICF) wall in cold climate”, Building XI Conference, December 5-9, 2010, Clearwater Beach, Florida, USA.

19. Saber, H.H., and Swinton, M.C., “Determining through numerical modeling the effective thermal resistance of a foundation wall system with low emissivity material and furred – airspace”, 2010 International Conference on Building Envelope Systems and Technologies, ICBEST 2010, Vancouver, British Colombia, Canada, June 27-30, 2010, pp. 247-257.

20. Saber, H.H., Maref, W., Swinton, M.C. and St-Onge, C., “Thermal Analysis of above-Grade Wall Assembly with Low Emissivity Materials and Furred-Airspace” Journal of Building and Environment, volume 46, issue 7, pp. 1403-1414, 2011 (doi:10.1016/j.buildenv.2011.01.009). 21. Swinton, M.C., Bomberg, M.T., Kumaran, M.K., and Maref, W. “In-situ performance of

expanded molded polystyrene in the exterior insulation systems (EIBS)”, Journal of Thermal Envelope & Building Science, 23, (2), pp. 173-198, 1999.

22. Maref, W., Swinton, M.C., Bomberg, M.T., Kumaran, M.K., Normandin, N., and Marchand, R.G. “Three-Dimensional Numerical Analysis Technique for Interpreting Monitored Result of the Exterior Insulation Basement Systems (EIBS) Experiment” Internal Report, Institute for Research in Construction, National Research Council Canada, 814, pp. 76, 2000.

23. Swinton, M.C., Maref, W., Bomberg, M.T., Kumaran, M.K., and Normandin, N. “Assessing heat flow patterns in basement walls with exterior insulation” Performance of Exterior Envelopes of Whole Building VIII: Integration of Building Envelopes (Clearwater Beach, Florida), pp. 1-9, 2001.

Figure

Figure 1. Schematic showing the boundary conditions (B.C.’s) of above- and below-grade basement wall  with foil bonded to EPS in furred-airspace assembly (Wall-FAA)
Figure 2. Measurements of soil, outdoor and indoor temperatures [21, 22, 23] -30-25-20-15-10-505101520253035090180270360450540630 720 810IndoorOutdoor(b)
Figure 3. Average temperatures of soil, outdoor and indoor (averaging within 2 week) at different time  during the 1 st  year (Wall-FAA)
Figure 4. Temperature contours in the furred-airspace assembly (in  o C) at different time for Wall-FAA drywallEPSdrywallEPS drywallEPS19 mm2438 mm
+4

Références

Documents relatifs

If fitting the temperature variation of the bulk modulus requires a value of K’ 0 that is inconsistent with that required to fit either the volume variation with pressure or

In order to separately quantify the effects of climate and biogeochemical changes, we made two additional control simulations: (1) CTRL_R with climatic and Atlantic condi-

menting an object from the input interactively, we incor- porate the shape and color features into a template match- ing framework to identify possible similar object locations

This paper uses data from Vietnam Household Living Standard Survey (VHLSS) 2010-2014 to investigate whether the current increasing block tariffs have negative

Phylogenetic reconstruction performed with single-copy genes (n=8,030) showed high levels of discordance among the different individual gene trees obtained, both at the nucleic

To confirm the importance of Bid to doxorubicin-induced cell death in vivo, we injected three cohorts of syngeneic mice with Eu-Myc p19 Arf −/− lymphoma cells expressing one of

- 66 -  :تايضرفلاب جئاتنلا ةنراقمو ةشقانم 1 - :ةيئزجلا تايضرفلاب جئاتنلا ةنراقمو ةشقانم تاباصلإا جلاع يف يضايرلا بطلا مايسإ ىدم ام ةفرعم دصق ايب انمق

Since the behavior of water in soils and many related porous materials is similar, the suction concept is not limited to soil.. Both the apparatus and technique used