• Aucun résultat trouvé

Convective heat transfer in hemispherical cavities with planar inner surfaces

N/A
N/A
Protected

Academic year: 2021

Partager "Convective heat transfer in hemispherical cavities with planar inner surfaces"

Copied!
35
0
0

Texte intégral

(1)

Publisher’s version / Version de l'éditeur:

ASHRAE Transactions, 117, 2, pp. 1-32, 2011-06-01

READ THESE TERMS AND CONDITIONS CAREFULLY BEFORE USING THIS WEBSITE.

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

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.

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 Convective heat transfer in hemispherical cavities with planar inner surfaces

Saber, H. H.; Laouadi, A.

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.

NRC Publications Record / Notice d'Archives des publications de CNRC: https://nrc-publications.canada.ca/eng/view/object/?id=d9a1d96a-bc08-4772-994a-948a4df92f22 https://publications-cnrc.canada.ca/fra/voir/objet/?id=d9a1d96a-bc08-4772-994a-948a4df92f22

(2)

Conve c t ive he at t ra nsfe r in

he m isphe ric a l c avit ie s w it h

pla na r inne r surfa c e s

N R C C - 4 2 6 7 9

S a b e r , H . H . ; L a o u a d i , A .

2 0 1 1 - 0 6 - 0 1

A version of this document is published in / Une version de ce document se trouve dans: Ashrae Transactions, V117, (2) pp.1-14, June, 2011

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

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)

Convective Heat Transfer in Hemispherical

Cavities with Planar Inner Surfaces (1415-RP)

Hamed H. Saber

1

and Abdelaziz Laouadi

Institute for Research in Construction, National Research Council Canada

Bldg. M-24, 1200 Montreal Road, Ottawa, Ontario, Canada K1A 0R6

ABSTRACT

Domed cavities are found in many applications in buildings, particularly in conventional

skylights and tubular daylighting devices. Heat transfer through domed cavities is thus an

important factor in evaluating the energy performance rating of such skylight systems, and in

calculating the heating and cooling loads of buildings. Although there have been many studies

on the convective heat transfer in related geometries, there is very limited information on natural

convective heat transfer in domed cavities with planar inner surfaces. In this paper, a numerical

study is conducted on natural laminar convective heat transfer in horizontal domed cavities with

planar inner surfaces. The bounding surfaces are subject to uniform temperature conditions. The

numerical model is based on the finite element method. The model predictions are successfully

compared with published data in literature for concentric spheres and hemispherical cavities.

The results show that for different boundary temperature conditions, the airflow in the cavities is

mono-cellular and reaches steady state conditions for both cold and hot weather conditions. The

numerical results are used to develop practical correlations for the Nusselt number.

Keywords: Natural laminar convection, hemispherical cavity, domed cavity, skylight, tubular daylighting device.

1

(5)

INTRODUCTION

Domed cavities are found in many applications in buildings, particularly in conventional

skylights and Tubular Daylighting Devices (TDDs). Skylights and TDDs are used mainly to

daylight interior spaces, and therefore reducing lighting energy use. However, these systems

may introduce a significant amount of heat loss and unwanted solar heat gain into indoor that

might off-set their energy saving benefits. Heat transfer through domed cavities is thus an

important factor in evaluating the energy performance rating of such skylight systems, and in

calculating the heating and cooling loads of buildings. Although there have been many studies

on the convective heat transfer in related geometries such as concentric domes, spheres and

cylinders, there is very limited information on natural convective heat transfer in domed cavities

with planar inner surfaces.

A number of studies have addressed the heat transfer in related geometries such spheres and

cylinders. The flow pattern between concentric isothermal spheres with diameter ratio of 2 was

studied by Garg (1991). That study used the vorticity-stream-function formulation and a finite

difference method to obtain the flow and temperature fields, which were in good agreement with

previous experimental work. Chiu and Chen (1996a, 1996b) and Chen (2005) applied the same

numerical method to study the transient natural convection between concentric and vertically

eccentric spheres with diameter ratio of 2 under isothermal and mixed boundary conditions. The

obtained numerical results for the flow pattern were consistent with the previous experimental

work. McGowan et al. (1998) conducted measurements and numerical simulations using a

commercial CFD package to investigate the thermal performance of pyramidal and barrel vault

skylights. They concluded that the convective heat transfer in curved or trapezoidal cavities can

(6)

of the complex cavity. The conclusion drawn from their study is, however, limited to fairly

small Rayleigh numbers (Ra < 7x104), and therefore, their results cannot be generalized to higher

Rayleigh numbers and to other geometry types.

Klems (2000) conducted measurements of the net heat flow through several types of flat

and domed skylights. He compared the measured U-values with calculations using WINDOW 4

and THERM programs. The calculated U-values did not agree with the measurements; in

general, the measured U-factors were higher than the calculated ones. The disagreement

between measurements and calculations was worse if the well air temperature (rather than the

chamber air temperature) was used to calculate the measured U-factor. Klems (2000) pointed up

the importance of having an accurate model of the frame and the environmental conditions.

Fomichev and Curcija (2007) used a CFD commercial tool to investigate free convection in

multi-layer eccentric domes tilted from the horizontal. Correlations for Nu were developed for

two types of domes with circular and flat inner surfaces tilted at 20º and 90º. The obtained

correlations are limited to small range of Rayleigh number (Ra ≤ 7x104). Sartipi et al. (2007) conducted numerical simualtions using control volume approach to investigate free convection in

eccentric multi-layer domes with low and high profiles when heated from the exterior surface.

Profiles of Nusselt number as a function of Ra were developed for various dome shapes.

Laouadi and Atif (2001) and Sartipi et al. (2010) used the numerical control volume approach to

investigate the laminar natural convection within domed cavities with uniform gap thicknesses.

The study covered a number of cavity gap spacing to dome radius ratios under a wide range of

Rayleigh numbers (Ra). The numerical results were used to develop practical correlations for

(7)

The present study is a part of an ASHRAE research project to evaluate the thermal

performance of tubular daylighting devices. The study addresses numerically the laminar natural

convection heat transfer in cavities of horizontal domes with planar inner surfaces subject to cold

or hot uniform temperature conditions. The main objectives of this paper are to investigate the

flow pattern within the cavity and to develop heat transfer correlations that account for different

environmental conditions and different dome dimensions.

MATHEMATICAL FORMULATION

A domed cavity is bounded by two finite surfaces: a hemispheric surface at the top and a

planar surface at the bottom. Both surfaces are made of a 6 mm (1/4 inch) acrylic sheets. The

indoor side of the planar surface and outdoor side of the hemispheric surface are maintained at

uniform temperatures Ti and To, respectively. Owing to the temperature differential across the

cavity (Ti – To), a buoyancy-driven flow develops within the air cavity. The flow is considered

to be two-dimensional and axi-symmetric since this configuration is symmetrical with respect to

the vertical revolution axis. Only half of the cavity is, therefore, considered for the numerical

simulation. Figure 1a shows a schematic representation of the domed cavity.

Assumptions

The following assumptions are made for the air inside the cavity:

• The long-wave radiation heat transfer between the cavity surfaces is decoupled from the convective heat transfer, and therefore is not accounted for in this study. This approach has

been successsfully applied to natural flows in enclosed cavities.

• The air is incompressible.

(8)

• The physical properties of the air are constant, except for density in the body force terms in the momentum balance equations. The physical properties of the air inside the cavity are

evaluated at the average temperature of the bounding surfaces.

• The fluid density is given by the Boussinesq approximation.

• The compression work and the viscous dissipation energy in the energy balance equation are neglected.

Using the Boussinesq approximation, the fluid density is expressed as follows:

(

1 ( ref )

)

ref

a =ρ −β TT

ρ (1)

The buoyancy-driven flow in the cavity is governed by the Navier-Stokes equations. The

transient conservation equations for mass, momentum and energy in the spherical coordinate

system read as follows:

Mass balance: 0 ) sin ( sin 1 ) ( 1 2 2 ∂ ⋅ ⋅ = ∂ + ⋅ ⋅ ∂ ∂ ρ θ θ θ ρ u r v r r r a a (2)

Momentum balance in r-direction:

r a v S r p v r u r v v t v + ∇ ⋅ + ∂ ∂ − =       ∂ ∂ + ∂ ∂ + ∂ ∂ µ 2 θ ρ (3)

Momentum balance in θ-direction:

θ µ θ θ ρ p u S r u r u r u v t u a + ⋅∇ + ∂ − =       ∂ ∂ + ∂ ∂ + ∂ ∂ 1 2 (4)

Energy balance (air):

T k T r u r T v t T cpa a a , = ⋅∇2       ∂ ∂ + ∂ ∂ + ∂ ∂ ⋅ θ ρ (5)

(9)

T k t T cps s s , = ⋅∇2 ∂ ⋅ ρ (6) where ∇2

is the Laplace operator, and Sr and Sθ are source terms. These are given by:

) (sin sin 1 ) ( 1 2 2 2 2 θ θ θ θ ∂ ∂ ∂ ∂ + ∂ ∂ ∂ ∂ = ∇ r r r r r (7)       + ∂ ∂ + − + ⋅ − = θ θ µ ρ θ ρ cos 1 2 2 2 22 2 2 cot r u u r r v u r g Sr a a (8)         − ∂ ∂ ⋅ + ⋅ − ⋅ = 2 2 ) sin ( 2 1 sin θ θ µ ρ θ ρ θ r u v r v u r g S a a (9)

Initial and Boundary Conditions

For horozintal hemispherical domed cavity shown in Figure 1, radial symmetry allows

modelling a three-dimensional case in two-dimensions. The initial temperature and velocity

components are assumed equal to 10oC and 0.0, respectively, everywhere in the computational

domain. The momentum equation is subjected to no-slip condition at the air-solid interfaces (at

r = Ri and at θ = 90o, u= v=0), and symmetry condition at the axis of revolution (at θ = 0º,

0 =

u and ∂ /v ∂θ = 0). The energy equations of both air and solid are subjected to symmetry condition at the axis of revolution (at θ = 0º, ∂ /T ∂θ = 0), and uniform temperatures at the indoor surface of the bottom planar acrylic sheet (T = Ti) and the outdoor surface of the

hemispherical acrylic sheet (T = T0). Due to the energy balance at the air-solid interfaces (at r =

Ri and at θ = 90o), these interfaces are subjected to the continuity condition (i.e. the heat fluxes

in the air and solid normal to these interfaces are equal, and also the temperatures of air and solid

are equal). Finally, in order to cover a wide range of the temperature difference across the

(10)

21oC (70ºF) (indoor temperature), and the temperature at the outdoor surface of the

hemispherical acrylic sheet is taken in the range of 21oC (70ºF) to 70oC (158ºF) for the summer

conditions, and - 40oC (40ºF) to 21oC (70ºF) for the winter conditions.

NUMERICAL PROCEDURE

The National Research Council of Canada’s hygrothermal model (called hygIRC-C) is used

to solve the governing equations (2) to (6). hygIRC-C uses the solver of COMSOL Multiphysics

( COMSOL, 2009) for the numerical simulations. The numerical approach of the tool uses the

finite element method. Recently, the two and three-dimensional versions of hygIRC-C were

benchmarked and used in several tasks to predict the hygrothermal (moisture transport is

accounted for) performance of different building envelope components (Saber et al., 2010a,

2010b; 2010c; 2010d; 2010e). To assure that the results are mesh independent, a non-uniform

mesh was selected with finer sizes near the boundaries. Typically, the numerical mesh was

refined by doubling the number of nodes until the final results do not appreciably change.

Triangular elements were chosen to capture the curved computational domain with less

discritizing error. Furthermore, within the range of temperature difference across the cavity

considered in this study, the cell Reynolds number was calculated to check that the flow is

laminar. Figure 1b shows a typical mesh used through the simulations.

CONVECTIVE HEAT TRANSFER IN THE DOMED CAVITY

The enclosed airspace in the horizontal domed cavity is subjected to a buoyancy-induced

flow. In this study, five of inner radii (Ri) were selected for the simulation: Ri = 0.5 m (1.64 ft),

0.63 m (2.067 ft) , 0.75 m (2.46 ft), 0.88 m (2.89 ft), and 1.0 m (3.28 ft). Based on the set of

(11)

considered so that standard investigation of a convective heat transfer in cavity, subject to

temperature boundary conditions on its bounding surfaces, can be defined. This approach allows

the development of dimensionless setup of the problem where dimensionless convective heat

transfer rates, as expressed by Nusselt number, Nu, can be evaluated as a function of

dimensionless natural convection parameter Raleigh number, Ra.

There are two options to be used in applying temperature boundary conditions on the domed

cavity. The first option is to assume that the temperatures on the indoor surface (facing the

indoor environment) of the bottom planar acrylic sheet (Ti) and the outdoor surface (facing the

outdoor environment) of the top hemispheric acrylic sheet (To) are uniform. The second option

is to assume that the temperatures of the surfaces of the top and bottom sheets in contact with the

cavity air (Toa and Tia, respectively; see Figure 1a) are uniform. The second option was

extensively used in the literature where there is no need to include the walls in the numerical

simulations (Laouadi and Atif, 2001; Sartipi et al., 2010). Due to the buoyancy-induced flow,

the air movement inside the cavity would cause a non-uniform temperature distribution on the

surfaces of the acrylic sheets facing the cavity air (Tia and Toa). Figure 2 shows an example of

the temperature distribution on the side surfaces of the bottom flat acrylic sheet for a domed

cavity with an inner radius of 0.75 m (2.46 ft) under both winter and summer conditions. In this

example, uniform temperature boundary conditions are applied on the outdoor surface of the top

acrylic sheet and indoor surface of the bottom acrylic sheet with temperatures Ti = 21oC (70ºF),

and To = 7oC (44.6ºF) for winter conditions, and 35oC (95ºF) for summer conditions. As will be

shown later, for the same ∆T, the air velocity in the case of cold weather is larger than that in the case of hot weather. As such, the temperature distribution on the surface facing the air cavity of

(12)

approximately uniform except at the edges in the case of hot weather (Figure 2b). Consequently,

the approach used in this work is to include the acrylic walls in the numerical simulations and

apply uniform temperature boundary conditions on the surfaces facing the indoor and outdoor

environments.

Evaluation of Nusselt Number

Under steady state conditions, the heat transfer rates at the boundary surfaces of the bottom

and hemispheric acrylic sheets are equal due to the energy balance. Because of the heat loss/gain

at the edges of the cavity, the heat transfer rates at the surfaces facing the air cavity of the bottom

and top acrylic sheets are not equal. In this work, the Nusselt number, Nu, is defined as follows:

cond conv Q

Q

Nu= / (10)

where Qconv and Qcond are the convective heat transfer rate and the conductive heat rate in the

absence of convection, respectively, normal to the inner surface (facing the air cavity) of the top

hemispherical acrylic sheet.

To evaluate the Nu in equation (10), the conductive heat transfer rate in the absence of

convection, Qcond, should be known a priori. Since there is no analytical solution available for

cond

Q in such hemispherical cavities with two media (fluid and solid), a parametric study was

conducted using the present model in order to calculate Qcond at different temperature difference

across the cavity with different inner radii. The obtained results of Qcond are shown in Figure 3a.

(13)

oa ia i cond R T T Q =0.58345⋅ 1.1637⋅ − (11) dr r r T R T i R i ia =

= ⋅ ⋅ 0 2 ( , /2) 2 θ π and θ θ θ π d R T Toa =

i ⋅ ⋅ 2 / 0 sin ) , ( (12)

where T ,ia Toaare the area-wieghted average temperatures at the surfaces facing the air cavity of

the bottom and top acrylic sheets, respectively. These area-wieghted average temperatures were

obtained by numerical integration of the nodal temperatures along the surfaces (see Eq. (12)).

As shown in Figure 3b, the obtained correlation given by Eq. (11) is in good agreement with the

numerical results. It should be noted that this correlation is developed when the SI units are used

for the heat flux and temperature.

MODEL BENCHMARKING

The present two-dimensional model is benchmarked against published results of domed

skylight devices. Laouadi and Atif (2001) presented correlations for the convective heat transfer

in domed cavities heated from inside. The correlations covered broad ranges of dome shapes and

ratio of cavity gap spacing to radius. Figure 4b compares the values of Nu predicted by the

present model with that obtained by Laouadi and Atif (2001) for concentric full spheres with

ratio of outer radius to inner radius of 2. The model by Laouadi and Atif (2001) was based on

the control voulme approarch using an in-house CFD model. For this sphere configuration, the

airflow reaches the steady state conditions independently of the initial conditions. Figure 4a

shows the airflow field and dimensionless temperature distribution, Φ = (T - To)/(Ti - To) at Ra =

91,000. As shown in Figure 4b, the obtained results using the present model are in good

agreement with the numerical results obtained by Laouadi and Atif (2001), with a maximum

(14)

Another benchmarking exercise was conducted by comparing the predictions of the present

model against the results obtained by Sartipi et al. (2007). The model by Sartipi et al. (2007)

was based on the control voulme approarch using a commercial software tool (Fluent). The

indoor surface of the domed cavity is hotter than the outdoor surface. The cavity was filled by

air and its gap spacing ratio, δ = (Ro-Ri)/Ri = 0.8. Figure 5a shows the airflow field and

dimensionless temperature distribution at Ra = 39,872 or Ra* = 246.1 (see the nomenclatures for

the definition of Ra and Ra*). In this cavity, the airflow was mono-cellular for different values

of Ra and reached the steady state conditions independently of the initial conditions. Figure 5b

shows comparison between the present predictions and those obtained by Sartipi et al. (2007).

As shown in this figure, the prediction of the present model was in good agreement with the

model of Sartipi et al. (2007). Note that in the two benchmarking exercises conducted in this

study, the bounding surfaces facing the air cavity were subjected to uniform temperatures.

RESULTS AND DISCUSSIONS

All results presented in this paper are for the case when air fills the cavity. The physical

properties of the air are evaluated at the bulk temperature of the air in the cavity. During the

transient regime, a multi-cellular airflow with different number of vortex cells are observed for

both the winter (Ti > To) and summer (Ti < To) conditions. The steady state conditions are

reached about three times faster in the case of Ti > To than in the case of Ti < To. At steady state

conditions, however, a mono-cellular airflow (i.e. the number of vortex cells in the cavity = 1)

was observed in both cases of hot and cold weather conditions. Furthermore, a number of

numerical simulations are conducted with different initial temperatures for each cavity

configuration. The results show that both the flow characteristics and calculated Nu at steady

(15)

conditions. Examples of the temperature contours, velocity contours, velocity vectors, and

streamline for both hot and cold weathers are presented in Figure 6 to Figure 9 at steady state

condition for a typical cavity configuration with an inner radius of 0.75 m (2.46 ft) when the

absolute temperature difference across the cavity is equal to 14oC (57.2ºF).

Figure 6a and Figure 6b show the horizontal and vertical velocity contours, velocity vectors,

and streamline in the case of cold weather at steady state conditions. Also, Figure 7a and Figure

7b show the horizontal and vertical velocity contours, velocity vectors, and streamline in the case

of hot weather. As shown in these figures, only one vortex cell was observed in both hot and

cold weathers. However, both the horizontal and vertical velocities were much higher in the case

of cold weather than that in the case of hot weather. For example, for the same temperature

difference across the cavity, the maximum absolute horizontal and vertical velocities in the case

of cold weather were 207 mm/s (244.5 ft/h) and 308 mm/s (363.8 ft/h), respectively, which were

16 and 11 times higher than that in the case of hot weather. As such, for the same temperature

difference across the cavity, the heat transfer coefficient for cold weather is significantly higher

than warm weather. The heat loss/gain on the other hand will also depend on the temperature

difference between the inside and the outside.

Figure 8a and Figure 8b show the temperature contours without (pure conduction) and with

convection, respectively, in the case of cold weather. Figure 9a and Figure 9b show the

temperature contours without and with convection, respectively, in the case of hot weather. As

shown in these figures, the buoyancy-induced flow results in a significant effect on the

temperature distribution in the cavity in both cold and hot weather conditions.

Figure 10 shows the dependence of Nu at the inner surface (facing air cavity) of the top

(16)

case of cold and hot weather conditions. As shown in this figure, Nu increases significantly with

increasing ∆Ta in the case of cold weather. In the case of hot weather, however, Nu increases with increasing ∆Ta at a much lower rate. Additionally, for the same ∆Ta, the Nu is much higher in the case of cold weather than in the case of hot weather. For example, Nu in the cold

weather is 3.8 and 5.3 times Nu in the hot weather for ∆Ta of 2oC (35.6oF) and 20oC (68oF), respectively. As stated earlier, this is due to the air velocity in the case of cold weather is much

higher than in the case of hot weather for the same ∆Ta across the cavity.

Correlations for the Convective Heat Transfer

Figure 11a and Figure 11b show the effect of the temperature difference across the cavity∆Ta

, and Ra on Nu in the case of cold weather conditions (Ti > To), respectively. For the same

inner radius, increasing ∆Ta results in stronger convection current in the cavity. As such, Nu increases by increasing ∆Ta (see Figure 11a). Additionally, for the same ∆Ta, Nu increases by increasing the inner radius of the cavity. For example, at ∆Ta = 10oC (50oF), Nu is 9.1 and 14.7 for a cavity of radius of Ri = 0.5 m (1.52 ft) and 1.0 m (3.ft), respectively. The numerical results

shown in Figure 11a are used to develop dimensionless heat transfer correlations in the case of

cold weather. A power-law form is used to develop this correlation. The obtained correlation is

given as follows:

[

1

]

, a=1.5205x10-3, b=0.5984, c=0.44845 c b Ra a Nu= + (13)

(17)

α ν β ⋅ ⋅ − ⋅ ⋅ = g Tia Toa Ri3 Ra (14)

In this correlation, Ra was evaluated using the air properties at the bulk temperature of the

air inside the cavity. As shown in Figure 11b, the obtained correlation given by equation (12) is

in good agreement with numerical results for different radii of hemispherical cavities (within an

error of ±3%).

In the case of hot weather (Ti < To), the effect of ∆Ta and Ra on Nu are shown in Figure

12a and Figure 12b, respectively. The results in Figure 12a translate to the following correlation:

[

1

]

, a=9.3949x10-5,b=0.73986, c=0.16495 c b Ra a Nu= + (15)

As shown in Figure 12b, the developed correlation given by equation (15) is in good

agreement with numerical results (within an error of ±3%). Note that in both correlations for the

cold and hot weathers, Nu approaches to 1.0 as the value of Ra goes to 0.0.

The convective heat transfer correlations for the cold weather and hot weather conditions

(Eq. (13) and (15), respectively) were obtianed when the temperature of the indoor side of the

planar surface (Ti) equal to 21oC. In the clod weahter condition, the temperature range of the

outdoor side of the hemispheric surface (To) was 20.95oC to -40oC (|Ti – To|min = 0.05oC and (|Ti

– To|max = 61.0oC). In the hot weahter condition, however, the temperature range of the outdoor

side of the hemispheric surface was 21.05oC to 70oC (|Ti – To|min = 0.05oC and (|Ti – To|max =

49.0oC). The range of the inner radius of the horizontal domed cavity (Ri) was varied from 0.5 m

to 1.0 m. As indicated earlier, a number of numerical simulations were conducted with different

initial temperatures for each cavity configuration. In these simulations, the steady state

condtions were achived where the flow characteristics and calculated Nu were independent on

(18)

Ra number given by Eq. (14), the lowest and highest Ra in the case of the air heated from below

(cold weather condition) were 7.47 x 105 (Ri = 0.5 m, and |Ti – To| = 0.05oC) and 5.51x109 (Ri =

1.0 m, |Ti – To| = 61.0oC), respectivly. For this Ra range, flow instabilities in the cavity are not

known, and were not observed using the present numerical approach. In flat horizontal cavities

heated from below, flow instability due to the Rayleigh-Benard convection may occur at a lower

Ra. For example, Bucchignani and Stella (1999) and Stella and Bucchignani (1999) found out

that the critical Ra values depend on the problem parameters such as the Prandtl number (Pr),

aspect ratio, and flow configurations. For their specific geometries, the critical Ra values

corresponding to flow transition from a steady to unsteady state were found to vary from 3.4x104

to 4.4x104. In the present study, flow instability would occur at higher values of Ra since the

curvature of the horizental domed cavity helps to stabilise the flow (note that natural flows in

vertical cavities are stable for this critcal Ra range). Future studies should therefore consider

flow instability and transition in domed cavities.

CONCLUSIONS

A numerical model based on the finite element method was developed to study natural

convection heat transfer in horizontal domed cavities with planar inner surfaces. The bounding

surfaces of the cavity were subject to uniform temperatures during hot and cold weather

conditions. The model was successfully benchmarked by comparing its predictions against

numerical data available in literature for concentric spheres and domed cavities with a uniform

gap thickness. The results obtained for the domed cavity under consideration showed that at

steady state conditions, a mono-cellular airflow was observed in both cases of hot and cold

weathers. For the same temperature difference across the cavity, the airflow velocity was much

(19)

the convective heat transfer in the cavity were developed, covering broad ranges of the

parameters in both cold and hot weathers.

ACKNOWLEDGEMENT

This work was jointly funded by ASHRAE under the Research Project 1415-RP, and the

Institute for Research in Construction of the National Research Council of Canada. The authors

are very thankful for their financial support.

NOMENCLATURE

Cp : specific heat (J/(kg.K)) g : gravitational acceleration (m2/s) Nu : Nusselt number (-) p : pressure (Pa) Pr : Prandtl number, Pr= Cpµ /λ=ν /α

Qeff : heat transfer rate (W)

Qcond : heat transfer rate by pure conduction (W)

Ri : inner dome radius (m)

Ro : outer dome radius (m)

Ra : Rayleigh number (-)

Ra* : Modified Rayleigh number, Ra*=Ra(δ/2)/

{

(1+δ)−3/5+(1+δ)4/5

}

5

r : radial position (m)

Sr, Sθ : source terms in the radial and circumferential directions.

(20)

Ti : temperature of indoor surface of the bottom planar acrylic sheet of the cavity (K)

Tia : average temperature of the inner surface of the bottom planar acrylic sheet of cavity (K)

To : temperature of the outdoor surface of the top hemispheric acrylic sheet of cavity (K)

Toa : average temperature of the inner surface of top hemispheric acrylic sheet of cavity (K)

Tref : reference temperature (K)

t : time (s)

u : velocity component in θ-direction (m/s) v : velocity components in r-direction (m/s)

Greek Symbols

α : fluid thermal diffusivity (m2/s) λ : thermal conductivity (W/(m.K))

δ : dimensionless cavity gap spacing (δ = (Ro-Ri)/Ri)

β : fluid thermal expansion coefficient (K-1) µ : fluid dynamic viscosity (Pa.s)

ν : fluid kinematic viscosity (m2/s) θ : position angle

ρ : density (kg/m3)

ρref : density evaluated at Tref (kg/m3)

Subscripts

a : air

(21)

REFERENCES

Bucchignani, E. and Stella, F. 1999. Rayleigh – Benard Convection in Limited Domains: Part 2 – Transition to Chaos. Numerical Heat Transfer, Part A, 36: 17-34.

Chiu C.P. and Chen W.R. 1996a. Transient Natural Convection Heat transfer between Concentric and Vertically Eccentric Spheres. International Journal of Heat and Mass Transfer, 39/7: 1439-1452.

Chiu C.P. and Chen W.R. 1996b. Transient Natural Convection Heat Transfer between Concentric and Vertically Eccentric Spheres with Mixed Boundary Conditions. Journal of Heat and Mass Transfer, 31: 137-143.

Chen W.R. 2005. Transient Natural Convection of Micropolar Fluids between Concentric and Vertically Eccentric Spheres. Journal of Heat and Mass Transfer, 48: 1936-1951.

COMSOL. 2009. COMSOL Multiphysics, ver 3.5a. www.comsol.com.

Fomichev A. and Curcija C.D. 2007. Development of a Procedure for U-Factor Rating of Domed Skylights. Final Report. University of Massachussetts. Massachussetts.

Garg V.G. 1991. Natural Convection between Concentric Spheres. International Journal of Heat and Mass Transfer, 35/8: 1935-1945.

Kelms, J.H. 2000. U-Values of Flat and Domed Skylights. ASHRAE 2000 Annual Meeting, June 24-28, 2000, Minneapolis, MN, and published in ASHRAE Transactions, 106 (2) 2000. Laouadi A. and Atif M.R. 2001. Natural Convection Heat Transfer within Multi-layer Domes.

International Journal of Heat and Mass Transfer, 44: 1973-1981.

McGowan A.G., Desjarlais A.O. and Wright J.L. 1998. Simulation and Testing of Pyramid and Barrel Vault Skylights. ASHRAE Transactions, 104/1: 832-844.

Saber, H.H., Maref, W., Elmahdy, A.H., Swinton, M.C., and Glazer, R. 2010a. 3D Thermal Model for Predicting the Thermal Resistances of Spray Polyurethane Foam Wall Assemblies. Building XI Conference, December 5-9, Clearwater Beach, Florida.

Saber, H.H., Maref, W., Elmahdy, A.H., Swinton, M.C., and Glazer, R. 2010b. 3D Heat and Air Transport Model for Predicting the Thermal Resistances of Spray Polyurethane Foam Wall Assemblies. International Journal of Building Performance Simulation, in print.

Saber, H.H., Swinton, M.C. 2010c. 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: 247-257.

Saber, H.H., Maref, W., Armstrong, M., Swinton, M.C., Rousseau, M.Z., and Gnanamurugan, G. 2010d. 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, Clearwater Beach, Florida.

Saber, H.H., Maref, M., Lacasse, M.A., Swinton, M.C., and Kumaran, M.K. 2010e.

(22)

Assemblies. International Conference on Building Envelope Systems and Technologies, ICBEST 2010, Vancouver, British Colombia Canada, June 27-30: 369-377.

Sartipi, A., Laouadi, A., Naylor, D., and Dhib, R. 2007. Natural convective Heat Transfer within Vertically Eccentric Domed Skylight Cavities. Proceedings of Building Simulation

conference, China.

Sartipi, A., Laouadi, A., Naylor, D., and Dhib, R. 2010. Convective Heat Transfer in Domed Skylight Cavities. Journal of Building Performance Simulation, in press.

Stella, F., and Bucchignani, E. 1999. Rayleigh – Benard Convection in Limited Domains: Part 1 – Oscillatory Flow. Numerical Heat Transfer, Part A, 36: 1-16.

(23)

Figure 1. A schematic representation of a horizontal hemispherical domed cavity with a planar inner surface and corresponding numerical mesh

S

y

m

m

et

ry

(

i.e

. ad

iab

at

ic

)

Temperature (T

i

)

(a) Boundary Conditions

Non-slip

condition

(u = v = 0)

Non-slip

condition

(u = v = 0)

T

ia

θ

v

(b) Numerical Mesh

(24)

Figure 2. Temperature on the surfaces of the bottom planar acrylic sheet at Ri = 0.75 m and ∆T = 14oC 20 22 24 26 28 0 0.15 0.30 0.45 0.60 0.75

Indoor surface of bottom planar sheet

Inner surface of bottom planar sheet

(b) Hot Weather

Ri = 0.75 m

Ti = 21oC

To = 35oC

Distance from center, r (m)

T

e

m

p

e

ra

tu

re

(

o

C

)

12 14 16 18 20 22 0 0.15 0.30 0.45 0.60 0.75

Indoor surface of bottom planar sheet (a)

Inner surface of bottom planar sheet

Cold Weather Ri = 0.75 m Ti = 21oC To = 7oC

T

e

m

p

e

ra

tu

re

(

o

C

)

(25)

Figure 3. Dependence of the conductive heat transfer rate normal to the inner surface of the top hemispherical acrylic sheet

0 5 10 15 20 25 30 0 5 10 15 20 25 30 35 40 45 50

Correlation

R

i

= 0.50 m

R

i

= 0.63 m

R

i

= 0.75 m

R

i

= 0.88 m

R

i

= 1.00 m

(b)

Q

cond

= a R

i1.1637

∆T

a

, a = 0.58345

∆T

a

= |T

ia

- T

oa

| (

o

C)

Q

co n d

/R

i 1 .1 6 3 7 0 5 10 15 20 25 30 0 5 10 15 20 25 30 35 40 45 50

R

i

= 0.50 m

R

i

= 0.63 m

R

i

= 0.75 m

R

i

= 0.88 m

R

i

= 1.00 m

(a)

Q

co n d

(

W

)

(26)

Figure 4. Comparison between the present model and Laouadi and Atif (2001) for concentric spheres heated from inside for (Ro-Ri)/Ri = 1

Flow field and

dimensionless

Temperature at

Ra = 91,000

(a)

(b)

1.0 1.5 2.0 2.5 3.0 3.5 103 104 105

Laouadi and Atif (2001)

Present Model

Ra

(27)

Flow field and

dimensionless

temperature at

Ra = 39,872

(Ra* = 246.1)

for

δ = 0.8

Adiabatic

(a)

(b)

1.0 1.5 2.0 2.5 3.0 3.5 1 2 3 4

Sartipi et al. (2010)

Present Model

δ = 0.8

Ra*

1/4

Nu

(28)

Figure 6. Horizontal and vertical velocity contours, velocity vectors and streamlines in the case of cold weather with temperature difference across the cavity of 14oC

Winter:

T

i

= 21

o

C

T

o

= 7

o

C

R

i

= 0.75 m

(b) Vertical velocity (mm/s)

Winter:

T

i

= 21

o

C

T

o

= 7

o

C

R

i

= 0.75 m

(a) Horizontal velocity (mm/s)

(29)

Figure 7. Horizontal and vertical velocity contours, velocity vectors and streamlines in the case of hot weather with temperature difference across the cavity of 14oC

Summer:

T

i

= 21

o

C

T

o

= 35

o

C

R

i

= 0.75 m

(a) Horizontal velocity (mm/s)

Vertical velocity (mm/s)

Summer:

T

i

= 21

o

C

T

o

= 35

o

C

R

i

= 0.75 m

(b) Vertical velocity (mm/s)

(30)

Figure 8. Temperature contours (oC) in the case of cold weather with temperature difference across the cavity of 14oC

(a) Pure Conduction

(b) Convection

Winter:

T

i

= 21

o

C

T

o

= 7

o

C

R

i

= 0.75 m

Winter:

T

i

= 21

o

C

T

o

= 7

o

C

R

i

= 0.75 m

(31)

Figure 9. Temperature contours (oC) in the case of hot weather with temperature difference across the cavity of 14oC

(a) Pure Conduction

(b) Convection

Summer:

T

i

= 21

o

C

T

o

= 35

o

C

R

i

= 0.75 m

Summer:

T

i

= 21

o

C

T

o

= 35

o

C

R

i

= 0.75 m

(32)
(33)

Figure 10. Dependence of Nusselt number on the temperature difference across the cavity of an inner radius of 0.75 m

1 3 5 7 9 11 13 15 17 0 5 10 15 20 253.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0 5.2 5.4 5.6

Hot weather

Cold weather

Ri = 0.75 m Ti = 21oC

∆T

a

= |T

ia

- T

oa

| (

o

C)

N

u

N

u

co

ld

w

e

a

th

e

r

/N

u

h

o

t

w

e

a

th

e

r

(34)

Figure 11. Dependence of Nu on the temperature difference across the cavity and Ra in the case of a cold weather condition

0

5

10

15

20

25

10

5

10

6

10

7

10

8

10

9

10

10

Correlation

R

i

= 0.50 m

R

i

= 0.63 m

R

i

= 0.75 m

R

i

= 0.88 m

R

i

= 1.00 m

(b)

+3%

-3%

Nu = (1+a*Ra

b

)

c

, r

2

=0.99727

a = 1.5205x10

-3

, b = 0.5984, c = 0.44845

Ra

N

u

0

5

10

15

20

25

0

5

10

15

20

25

30

35

40

45

R

i

= 0.50 m

R

i

= 0.63 m

R

i

= 0.75 m

R

i

= 0.88 m

R

i

= 1.00 m

∆T

a

= |T

ia

- T

oa

| (

o

C)

(a)

N

u

(35)

Figure 12. Dependence of Nu on the temperature difference across the cavity and Ra in the case of a hot weather condition

1.0

1.5

2.0

2.5

3.0

3.5

10

5

10

6

10

7

10

8

10

9

10

10

Correlation

R

i

= 0.50 m

R

i

= 0.63 m

R

i

= 0.75 m

R

i

= 0.88 m

R

i

= 1.00 m

+3%

-3%

(b)

Nu = (1+a*Ra

b

)

c

, r

2

=0.99335

a = 9.3949x10

-5

, b = 0.73986, c = 0.16495

Ra

N

u

1.0

1.5

2.0

2.5

3.0

3.5

0

5

10

15

20

25

30

35

40

45

50

R

i

= 0.50 m

R

i

= 0.63 m

R

i

= 0.75 m

R

i

= 0.88 m

R

i

= 1.00 m

∆T

a

= |T

ia

- T

oa

| (

o

C)

(a)

N

u

Références

Documents relatifs

- Déterminer la motivation des parents et les éléments déclencheurs qui favoriseraient une participation parentale dans un programme de prise en charge du

On the other hand, a slow reaction wave allows the reacted region to cool on the same time scale as the wave propagation, creating a temperature gradient in the opposite

The question is bound to arise with new acuteness: are so many arts and economics graduates really necessary, when Japan, like Western industrial countries, is

Figure 5 displays the profiles of the two-point correlations ρ vv (y) and the length-scale- based resolution indicators φ y (l) in the wall-normal direction computed for the

Consequently, instead of using the formal method of two scale asymptotic expansions in the strong form of problem (8), which is complicated because of the non-local boundary

Of particular interest in the comparison of compression encoding schemes is the relative performance of Run Length encoding and line-to-line Correlation. The

ا هتاه لك و ملعملا نكمت ةيميلعتلا لئاسول ًءاوس دح ىلع ملعتملا ريفوت نم أ ربك دهج ريفوت يلاتلابو نكمم تقو أ ةنرمو ةسلس ةقيرطب تامولعملا لاصياو

In the present study, the spanwise-averaged absolute streamwise vorticity flux is computed only with the longitudinal vortex component, and is used for the discussion on the