Anda di halaman 1dari 8

5th European Thermal-Sciences Conference, The Netherlands, 2008

Conduction / Natural convection analysis of heat transfer


across multi-layer building blocks

H. Baig and M. A. Antar


KFUPM, Dhahran, 31261, Saudi Arabia
King Fahd University of Petroleum and Minerals, Dhahran, Saudi Arabia

Abstract

Heat loss across a multi-layer wall that includes hollow blocks is studied numerically.
Conjugate conduction-natural convection heat transfer is carried out to compute the
heat leak/ R-Value for different numbers of air-filled cavities and for different cavity
layouts. Interaction between conduction and convection at the boundaries is
considered rather than the classical isothermal and adiabatic boundary conditions
reported in open literature. This is believed to provide more accurate estimation of the
heat transfer rate across the walls. Therefore, boundary conditions include a heat
balance at the boundaries between the two heat transfer modes. Results show that
increasing the number of cavities can be effective in reducing the heat flux without
compromising the building blocks strength. Nevertheless, a considerable reduction in
the heat flux or increase in the R-value can be achieved through changing the layout
of the cavities within the wall. The effect of both techniques can compensate for
removing the insulation layer in the cavity used to increase the wall thermal
resistance. This is due to decreasing the cavities width that contributes significantly in
suppressing the convection heat transfer effects (reduced value of Ra).

1. Introduction
Heat leak through building walls and ceilings consume a substantial amount of
energy. Since climate control units require a significant amount of electric energy,
studies of heat leak has received considerable attention in the past decades. An
accurate estimate of the heat leak through the composite, multi-layered walls
accompanied with practical low cost methods for reducing the heat leaks is an
effective way of reducing energy consumption. Hollow blocks are major wall
elements where the modes by which heat transfer occurs are heat conduction in the
solid sections and natural convection within the cavities. Investigations of the heat
transfer through the hollow spaces have been treated both experimentally and
theoretically by several researchers
Lacarrier et al. (2003) analyzed numerically the vertically perforated bricks.
They reported that walls can be constructed without any other materials than clay and
mortar. They reported that heat transfer in these assemblies is not totally understood.
For perforated brick construction, it is indicated that convection heat transfer is
negligible in the perforations. Therefore, the thermal resistance of the brick increases.
In a particular study of the ruptures it is concluded that the convection present in these
regions is a local phenomenon since it breaks the thermal bridges created by the
mortar fill.

Bajorek and Lloyd (1982) carried out an experimental study to investigate the
natural convection heat transfer within a two dimensional partitioned enclosure of unit
aspect ratio using an interferometer. They reported that dividing the cavity along its
5th European Thermal-Sciences Conference, The Netherlands, 2008

vertical axis results in a reduction in the heat transfer by approximately 15%.


Nishimura et al. (1988) reported that Nusselt number is inversely proportional to the
number of partitions which was also confirmed by experiments. They also observed
that effective heat leak reduction is attained using 2-5 partitions. Aviram et al. (1988)
investigated experimentally variable aspect ratio cavity and reported that increasing
aspect ratio decreases flow magnitude, reduces circulation intensity and increases the
cavity thermal resistance. Nusselt number diminished with reduced cavity depth.

Recently, del Coz Diaz et al (2006) carried out an experimental and numerical study
to investigate the thermal transmittance coefficient, U, of a wall made of Arliblock
bricks. They observed that wall insulation decreases with the increase in the mortar
and material conductivities. They also noticed that changing the profiles of the
holes alters the rate of the heat transfer through the hollow blocks. Then, they
(2007) studied major variables influencing the thermal conductivity of masonry
materials and carried out an optimization study for different brick geometries based
on both thermal resistance and weight.

The minimum web thickness for safe construction was reported by Kumar
(2003). Ciofalo and Karayiannis (1991) reported that the mechanism responsible for
the large reduction in heat transfer in partitioned enclosures was because of the
breaking down of the unicellular circulation near the regions. Manz (2003) studied
natural convection heat transfer in rectangular, gas-filled tall cavities in building
elements such as insulating glazing units, double-skin facades and others. He reported
that flow regimes depend on Ra and the aspect ratio. A linear temperature profile
exists as a function of the x-position within the so-called conduction regime.

Al-Hazmy (2006) investigated the heat transfer through a common hollow


building brick. Insulation assessment of the building blocks was examined based upon
the heat transfer rate. Three different configurations for building bricks were studied
including a gas-filled and insulation-filled cavity. Results show that the cellular air
motion inside blocks’ cavities contributes significantly to the heat loads. The insertion
of polystyrene bars reduced the heat rate by a maximum of 36%. Lee and Pessiki
(2006) carried out a study to investigate the performance characteristics of precast
concrete sandwich wall panels with two or three wythes separated by air layers. It was
found that, in general, the thermal performance of three-wythe panels is better than
that of two-wythe panels due to the increased thermal path length.

Ho and Yih (1987) analyzed conjugate natural convection and conduction in a


multi-layer wall. They considered isothermal left and right sides of the wall and
adiabatic boundary condition in both top and bottom surfaces. Tong and Gerner
(1986) analyzed natural convection in partitioned air-filled rectangular enclosures and
reported that placing a partition midway between the vertical walls results in the
greatest reduction in heat transfer. Kangni et al. (1991) investigated natural
convection in partitioned walls for various aspect ratios and for a wide range of Ra
and wall thicknesses. Turkoglu and Yucel (1996) investigated numerically natural
convection heat transfer in enclosures with conducting multiple partitions and side
walls. However, in their analysis the sidewalls were assumed to be isothermal, thus
eliminating the temperature gradient in the y-direction within the solid. They also kept
the top and bottom surfaces perfectly insulated. They reported that Nusselt number
5th European Thermal-Sciences Conference, The Netherlands, 2008

decreases as the number of partitions is increased up to 4. They also reported that the
cavity aspect ratio had an insignificant effect on their calculations

Lorente (2002) published a review article to illustrate the heat flow through
walls with relatively complicated internal structure. He reported the effect of Rayleigh
number and aspect ratio showing that for Re= 3550, no fluctuations were observed
and a unicellular flow was observed. As the Rayleigh number increases, the flow
becomes multi-cellular.

Antar and Thomas (2001) addressed the heat transfer across a hollow building
block and estimated two dimensional effects of the heat transfer across the block. In
another study they (2004) developed a numerical finite-difference analysis for steady-
state heat transfer in a composite wall with a two-dimensional rectangular gray body
radiating cavity with and without natural convection circulation of air. The purpose of
their analysis was to provide a basis for evaluating the accuracy of the first-order two-
dimensional model. Recently, Antar (2006) investigated the significance of multi-
dimensional effects in estimating the rate of heat loss and identified the cases where
simple one-dimensional convection/radiation analysis may be considered a good
approximation for heat transfer rate calculations.

It was reported by Antar and Thomas (2001, 2004) and Antar (2006) that the
approximate simple one dimensional analysis for the problem under investigation has
two alternative thermal circuits, an upper bound thermal circuit and a lower bound
one. Calculations show that the percentage difference in estimating the upper bound
and lower bound heat transfer rate reaches 39 %. This indicates significant two
dimensional effects. Neither the upper bound nor the lower bound solution provides a
reliable value for the heat leak, and a two dimensional model is needed to estimate the
heat transfer rate accurately. Therefore, this work is aimed at developing and using a
two-dimensional model for investigating the effect of cavities layout on the thermal
resistance of the block. A block of 5 cavities in the heat flow direction was considered
with the objective of increasing the thermal resistance. The previous literature search
considered the number of cavities as a major factor for reducing the heat leak .This
study shows that the shape and distribution of the cavity plays an equally important
role in this regard. More practical boundary conditions will be used such as non-
uniform vertical walls temperatures and non-adiabatic horizontal surfaces of the
cavities.

2. Formulation
The basic geometry of the problem under investigation is shown in Figure 1. Heat
transfer by conduction occurs in the solid part whereas natural convection heat
transfer occurs within the cavities. Other layouts may have different cavity size or
structure, see Figure 4, but the modes are unchanged.
5th European Thermal-Sciences Conference, The Netherlands, 2008

w1

w2

w1

L1 L2 L1 L2 L1 L2 L1 L2 L1 L2 L1
Figure 1: Hollow brick geometry, case 1

2.1 Conduction heat transfer

Steady two dimensional conduction heat transfer in the block solid material is
governed by the following equation:
∂ 2T ∂ 2 T
+ =0 (1)
∂x 2 ∂y 2
T = Ti at x = 0 T = To at x = L (2)
∂T ∂T
=0 at y = 0 = 0 at y = w (3)
∂y ∂y
∂T
−k = qs,x" at x = L1 and at x = L1 + L2 for w1 ≤ y ≤ w1 + w2 (4)
∂x
∂T
−k = qs,y" at y = w1 and at y = w1 + w2 for L1 ≤ x ≤ L1 + L2 (5)
∂y
where L = (L1 + L2 + L1), w = (w1+ w2 + w1)
qs,x" and qs,y"are the convection heat transfer flux at the interface in both x and y
directions. The same boundary conditions are applied to all cavities.

2.2 Convection heat transfer

The governing differential equations for the free convection within each of the
enclosures are given –for constant properties, no heat dissipation, applying
Boussinesq approximation - as:
∂u ∂v
+ =0 (6)
∂x ∂y
⎛ ∂u ∂u ⎞ ∂p ⎛ ∂ 2u ∂ 2u ⎞
ρ ⎜u +v ⎟ = − +μ⎜ 2 + 2 ⎟ (7)
⎝ ∂x ∂y ⎠ ∂x ⎝ ∂x ∂y ⎠
⎛ ∂v ∂v ⎞ ∂p ⎛ ∂ 2v ∂ 2v ⎞
ρ ⎜u + v ⎟ = − + μ ⎜ 2 + 2 ⎟ − ρ g β (T − T∞ ) (8)
⎝ ∂x ∂y ⎠ ∂y ⎝ ∂x ∂y ⎠
5th European Thermal-Sciences Conference, The Netherlands, 2008

∂T ∂T k ⎛ ∂ 2T ∂ 2T ⎞
u +v = ⎜ + ⎟ (9)
∂x ∂y ρ C ⎜⎝ ∂x 2 ∂y 2 ⎟⎠
At the inner walls of the gaps, the no slip condition applies (U=V=0) (10)
Note that air properties are considered as: ρ = 1.225 kg/m3, cP = 1.006 kJ/kg K and k
= 0.0242 W/mK.
The boundary conditions at the interface of the air gap include continuity of
temperature and heat flux. However, temperatures are unknown, and hence an
iterative solution will be adopted to guess the temperatures at the interface and apply
the heat balance equation to calculate them (refer to equations 4 and 5)
"
q conduction = q convection
"
(11)
Note that the boundary conditions are applied for all the air gaps in the material, i.e.
once for the arrangement given in Fig. 1, and as many as the number of gaps

3. Numerical Solution and Grid independence


The grid used in both solid and fluid phases is designed such that more nodes are
concentrated at all the corners of both solid and gas phases where higher gradients
exist. Other layouts follow the same trend, see figure 2 for a sample of the numerical
grid. The Control volume approach is employed in the numerical scheme. Variables
are computed at ordinary nodal points, except the velocities, which are determined at
staggered grid centered around the faces of the cells. More details of the numerical
scheme are given by Patankar (1980). Grid independent tests are conducted as shown
in Fig. 3 where a grid of 12720 nodes is considered on the basis of less computation
time without compromising the solution accuracy in the case of a 5 cavities. The grid
independence study was repeated for all other configurations. The number of nodes
increases as we use more cavities within the block to take care of the boundary layer
growth within each cavity.

Figure 2: Numerical grid Figure 3: Grid independence

4. Discussion of Results
Since the current layouts are not reported in the open literature, the numerical results
are validated by calculating the Nusselt number using the layouts employed by other
investigators. The percentage difference in calculating the Nusselt number (at Ra =
105 and H/W = 1) compared with Ho and Yih (1987) was 2%. Moreover, Nusselt
5th European Thermal-Sciences Conference, The Netherlands, 2008

number calculations compared with Kangni et al. (1991) for a different case (Ra = 106
and H/W = 5) showed a deviation of 0.2%.

The R-value for the layout shown in Figure 1 (based on one third of the block due to
symmetry) is 0.47 K m2/W (case 1). Changing the layout to the one shown in Figure
4a (case 2) results in an increase in the R-value (decrease in the heat transfer rate) by
13.5 %. This is due to narrowing the path of conduction heat transfer in the solid
material above and below the cavities. A further change in the layout to the case 3,
shown in Figure 4b results in a further increase in the R-value by 17.65 % (R-0.56
K.m2/W) compared to the basic case. This layout is better than the previous one (case
2, Fig 4a) since it does not have small aspect ratio cavities and therefore has less
convection effects. It is interesting to mention that the layout of case 4, Figure 4c did
not show an improvement in the heat leak, R-0.51 K.m2/W. This is due to pronounced
role of conduction heat transfer in the solid surrounding the cavities thus providing
unnecessary thermal bridges. In addition, more cavities in this case with lower aspect
ratio results in a more air circulation that promotes the convection heat transfer rate.

Therefore, changing the layout of the brick can play a substantial role in increasing its
thermal resistance. The layout shown in case 5, Figure 4d is also tested and it results
R-value 0f 0.43 K.m2/W, i.e. an 8.8 % increase in the heat transfer rate/reduction in
R-value. It is fairly obvious that this layout is not a preferred alternative. This is due
to higher convection effects and less conduction resistance (because of more thermal
bridges) layout compared with the other layouts.

(a) case 2 (b) case 3

(c) case 4 (d) case 5


Figure 4: Stream function in the alternative brick layouts
5th European Thermal-Sciences Conference, The Netherlands, 2008

Figure 5 shows the mid-plane (y=w/2) temperature distribution as it drops from To at


the right exterior wall to Ti at the left exterior one for the two layouts “case 2” and
“case 3”. These two layouts were selected since they have the lowest heat transfer
rates (highest R-value). Case 3, shown in Figure 4b has a higher resistance compared
to case 2 shown in Fig. 4a. This can be attributed to more homogeneous and steeper
temperature gradient established across the width of the whole block (20 cm in this
study) leading to better temperature distribution and higher thermal resistance across
the block.

Case 2 Case 3
Figure 5: Centerline temperature profile for cases 2 and 3

5. Conclusion
Conduction / Natural convection conjugate heat transfer within masonry blocks with
hollow cavities is investigated numerically. Changing the layout of the cavities can
play a substantial role in increasing the R-value (decreasing the heat leak) and hence
enhances thermal insulation through increasing the R-value by 13.5% (case 2
compared to case 1) and up to 17.65% (case 3 compared to case 1) without affecting
the structural characteristics of the blocks. Decreasing the thickness of the solid
material used (within the safe limits) decreases the thermal bridges and thus avoids
unnecessary conduction heat leak. In addition, using cavities of high aspect ratio (case
3 instead of case 2) and less width decreases the convection heat transfer effects since
decreasing the cavity width reduces the Rayleigh number and hence reduces the
convection heat transfer coefficient. It is believed that these results will provide
guidelines to masonry block manufacturers for better energy conservation.

Acknowledgement
The support provided by King Fahd University of Petroleum and Minerals, KFUPM
to carry out this investigation through project number IN-0600322 is gratefully
acknowledged.

6. References
[1] Lacarrie`re, B., Lartigue, B. and Monchoux, F., 2003, Numerical study of heat
transfer in a wall of vertically perforated bricks: influence of assembly method,
Energy and Buildings, 35, 229–237
[2] Kumar, S., 2003, Fly ash–lime–phosphogypsum hollow blocks for walls and
partitions, Building and Environment, 38, 291-295.
[3] Bajorek, S. M. and Lloyd, J. R., 1982, Experimental Investigation of Natural
Convection in partitioned Enclosures, Journal of Heat Transfer, 104, 527 – 531.
5th European Thermal-Sciences Conference, The Netherlands, 2008

[4] Nishimura, T., Shiraishi, M., Nagasawa, F. and Kawamura, Y., 1988, Natural
convection heat transfer in enclosures with multiple vertical partitions,
International Journal of Heat and Mass Transfer, 31 (8), 1679-1686.
[5] Aviram, D P., Fried A. N., and Roberts, J. J., 2001, Thermal properties of a
variable cavity wall, Building and Environment , 36, 1057–1072.
[6] del Coz Diaz, J. J., Garcia Nieto, P. J., Martin Rodriguez, A., Lozano Martinez-
L uengas, A., Betegon Biempica, C., 2006, Non-linear thermal analysis of light
concrete hollow brick walls by the finite element method and experimental
validation, Applied Thermal Engineering, 26, 777– 786.
[7] del Coz Dı´az a, J. J., Garcı´a Nieto, P. J., Betego´n Biempica, C., Prendes
Gero, M. P., 2007, Analysis and optimization of the heat-insulating light
concrete hollow brick walls design by the finite element method, Applied
Thermal Engineering, 27, 1445–1456
[8] Ciofalo, M. and Karayiannis, T. G., 1991, Natural convection heat transfer in a
partially—or completely—partitioned vertical rectangular enclosure,
International Journal of Heat and Mass Transfer, 34 (1), 167-179.
[9] Manz, H., 2003, Numerical simulation of heat transfer by natural convection in
cavities of facade elements, Energy and Buildings, 35, 305–311.
[10] Al-Hazmy, M. M., 2006, Analysis of coupled natural convection conduction
effects on the heat transport through hollow building blocks, Energy and
Buildings, 38, 515–521.
[11] Lee, B. J., Pessiki, S., 2006, Thermal performance evaluation of precast
concrete three-Wythe sandwich wall panels”, Energy and Buildings, 38, 1006–
1014.
[12] Ho, C. J. and Yih, Y. L., 1987, Conjugate natural convection heat transfer in
an air-fileld rectangular cavity, International Communication in Heat and Mass
Transfer, 14, 91-100.
[13] Tong, T. W. and Gerner, F. M., 1986, Natural convection in partitioned air-
filled rectangular enclosures, International Communication in Heat and Mass
Transfer, 13, 99-108.
[14] Kangni, A., Yedder B., and Bilgen, E., 1991, Natural convection and
conduction in enclosures with multiple vertical partitions, Int. J. Heat and Mass
Transfer, 34, 2819-2825.
[15] Torkoglu, H. and Yucel, N., 1996, Natural convection heat transfer in
enclosures with conducting multi le partitions and side walls, Heat and Mass
Transfer, 2, 1- 8.
[16] Lorente, S., 2002, Heat losses through building walls with closed, open and
deformable cavities, International Journal of Energy Research, 26, 611-632.
[17] Antar, M. A. and Thomas, L. C., 2001, Heat transfer through a composite wall
with enclosed spaces: A practical two-dimensional analysis approach,
ASHRAE Transactions, 106, 318-324.
[18] Antar, M. A. and Thomas, L. C., 2004, Heat Transfer Through a Composite
Wall with an Evacuated Rectangular Gray body Radiating Space: A Numerical
Solution, ASHRAE Transactions, 110 (2), 36-45.
[19] Antar, M. A., 2006, Multi-Dimensional Effects in Estimating the Heat Loss
Across Building Envelopes” Proceedings of the Second International
Conference on Thermal Engineering Theory and Applications, Al-Ain, UAE 3-
6 Jan.
[20] Patankar, S. V., 1980, Numerical Heat Transfer, McGraw-Hill, New York.

Anda mungkin juga menyukai