Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Accessing the thermal performance of Earth-air heat

exchangers coupled to galvanized materials


Ramalho, J. V. A.a,∗, Fernando, H. J.b , Domingues, A. M. B.a , Navarro
Pastor, N. R.a,c , Burlón Olivera, M. R. B.a,c
a
Programa de Pós-Graduação em Modelagem Matemática, Universidade Federal de
Pelotas (UFPel), Campus Capão do Leão s/n, Pelotas, 96160-000, RS, Brazil.
b
Instituto de Ciências Exatas (ICEx), Departamento de Matemática, Universidade
Federal Fluminense (UFF), Campus Aterrado, Rua Desembargador Ellis Hermydio
Figueira, 783, Aterrado, Volta Redonda, 27213-145, RJ, Brazil.
c
Instituto Tecnológico Regional del Norte (ITRN), Universidad Tecnológica (UTEC) del
Uruguay, Ruta 5 (Guido Machado Brum), km 496, Rivera, 40000, Uruguay.

Abstract
Resumindo... blalblablal
Keywords: Earth-air heat exchangers, keyword two


Corresponding author

Preprint submitted to “Nome da Revista...” March 2, 2022


Nomenclature

b block size (m) Greek letters


cp specific heat (J/kgK) α thermal diffusivity (m2 /s)
D duct diameter (m) ∆y segment size (m)
h convection coefficient (W/m2 K) θ efficiency (%)
L length (m) λ thermal conductivity (W/mK)
ṁ air mass flow (kg/s) µ dynamic viscosity (kg/m s)
Nu Nusselt number ρ density (kg/m3 )
Pr Prandtl number ψ area fraction (%)
Ps soil thermal potential
Pt EAHE thermal potential Subscripts
Q̇ heat transfer rate (W ) 0 initial (reference) value
r radius (m) a air
Re Reynolds number g galvanized
S size (m) h horizontal
t time (s or days) i inlet
T temperature (◦ C) o outlet
UL heat transfer coefficient s soil
per duct length (W/m K) v vertical
U∗ condutance ratio (m)
v speed (m/s) Abreviatures
x horizontal coordinate (m) EAHE Earth-air heat exchanger
z depth (m) RMS Root mean square value

1. Methodology
1.1. Settings
In general, this work aims to evaluate how the EAHE performance is
affected by coupling the ducts to galvanized structures. In particular, our
simulations adapt ideas from the reference [1], regarding the climate and soil
conditions from an experimental EAHE installation in the south Brazilian
city of Viamão [2].
Here, we consider the straight part of the buried duct, as done in other
references like [3, 4, 5]. Moreover, in the duct longitudinal direction, the
soil temperature variations are considered negligible, relative to the ones in
transverse directions. Therefore, we estimate soil temperatures with a 2D
modeling approach, adopting computational domains similar to the sketch

2
in Fig. 1 (a). The overall strategy is to surround the duct by a galvanized
block which can be connected to fins made of the same highly conductive
material.

(a)

(b)

Figure 1: 2D and 3D examples of the computational domain

In the example of Fig. 1 (a), z0 is the duct center depth, initially equal to
1.6 m; D0 = 11 cm is the diameter of the duct; b0 = 18 cm refers to the block
size (which has a square cross section). Moreover, all fins have the same
width Sv = 1 m and thickness Sh = 1 cm. Here, we placed the duct and
block equally centered in the horizontal direction. Hence, their centers have
the coordinate x0 = 5m. Figure 1 (b) also gives a 3D view, where we can
see that the duct, block and fins have the same length L0 = 25.77 m. Such
values for z0 , D0 , and L0 are used to emulate the experimental installation

3
from [2]. As explained in the Section 1.4, the values for the galvanized parts
represent an adaptation of the work [1] aimed to the reality given in [2].
The thermophysical properties of the soil and air in Viamão, as well as
the galvanized materials, are given in Table 1. The values come from the
references [6] and [1]. As done in other works like [7, 4, 8], this study does
not consider the thickness and properties of the ducts; besides, references
like [9] conclude that they have little influence on thermal exchanges because
the thickness of the ducts is relatively small.

Table 1: Thermophysical properties

ρ (kg/m3 ) cp (J/kg K) λ (W/m K) µ (kg/m s)


Soil 1800 1780 2.1 -
Air 1.16 1010 0.0242 1.789×10−5
Galvanized materials 7800 446 52 -

1.2. EAHE Model


We simulate EAHE with the GAEA model (Graphische Auslegung von
Erdwärme Austauschern1 ) introduced in [10]. It is a one-dimensional model
that computes heat transfer coefficients analytically to estimate the heat
exchanges among the air, the duct walls, and the soil. It allows determining
the temperatures at several parts of the duct, including the outlet. Recent
references like [11] and [12] also studied the model; moreover, the latter
validated a version of it considering the conditions of Viamão.
As the temperature in the duct walls depends on the thermal exchange
with the ground and with the air, GAEA estimates the conductance ratio of
heat transfer (from soil to the walls and from airflow to the walls) using the
parameter
2πλs
U∗ =  s 2 , (1)
2z0 2z0
UL ln + −1
D0 D0
Here, λs is the thermal conductivity of the soil around the duct (or the
thermal conductivity of the galvanized material surrounding it) and

UL = πD0 h (2)

1
German for graphical design of EAHE

4
is the overall heat transfer coefficient (per duct length) between the air stream
and the duct walls.
We approximate the convection coefficient at the inner surface of the duct
with
λa Nu
h= , (3)
D0
where the Nusselt number is determined by
Nu = 0.0214 (Re0.8 − 100) Pr0.4 . (4)
The Reynolds and Prandtl numbers are given, respectively, by
ρa va D0
Re = , (5)
µa
µa cp,a
Pr = . (6)
λa
The soil temperature Ts is estimated disregarding the ducts presence. At
this point, we assume that it is already known, but we explain ahead how we
computed it adicionar seção. Hence, GAEA needs to estimate a corrected
temperature on the duct walls, i.e.
U ∗ Ts + Ta
Tc,w = . (7)
U∗ + 1
Here, Tc,w is a weighted average between the air temperature inside the duct,
Ta , and the soil temperature on the duct wall (not influenced by its presence),
Ts . Therefore, U ∗ works as a weighting factor.
However, instead of calculating Tc,w directly by Eq. (7), the duct is divided
into 100 segments, where the thermal exchanges are determined iteratively
[10]. The idea is to take sufficiently small pieces to assume that the air flows
through them at an approximately constant temperature. Therefore, jumps
in temperature values occur between distinct segments of size ∆y = L0 /100.
For each segment k, a correction of the soil temperature in the duct wall
is calculated using the formula

k
U ∗ Ts + Ta,i
k
Tc,w = . (8)
U∗ + 1
k
Here Ta,i is the air temperature at the inlet of the k-th segment, where the
heat transfer rate is given by
k k
Q̇k = UL (Tc,w − Ta,i )∆y. (9)

5
Such rate is also given by
k k
Q̇k = ṁ cp,a (Ta,o − Ta,i ), (10)
k
where Ta,o is the air temperature at the outlet of the k-th segment and ṁ
the air mass flow. As we are considering a circular duct, then
πD02
ṁ = ρa va
. (11)
4
k
Combining equations (9) and (10), we can isolate Ta,o to obtain
k k

k k
∆y UL Tc,w − Ta,i
Ta,o = Ta,i + . (12)
ṁ cp,a
In summary, we calculate with Eq.(12) the air temperature at the outlet
of each segment based on the air temperature at the inlet of it. We note
that the outlet of segment k is the inlet of the k + 1 one. When the iterative
process starts, we know the temperature at the inlet of the first segment
(k = 1) because it is also the duct inlet. The algorithm ends by finding the
temperature at the the last segment outlet (k = 100), which coincides with
the duct outlet.
Juntar em algum lugar informações sobre todas as implementações... To
make the simulations with the GAEA model, we implemented an in-house
code in the Matlab programming language.
1.3. Physical Model
In this work, we compute the temperatures in the soil, and in regions
covered by galvanized materials, disregarding the presence of the ducts. In
particular, we consider cases similar to the one illustrated by Fig. 1 (a),
therefore, we will neglect the temperature variations in the y direction, and
solve the energy conservation equation
 2
∂ 2T

∂T λs ∂ T
= + 2 , (13)
∂t ρs cp,s ∂x2 ∂z
restricted to two-dimensional domains in the plane x − z.
Based on the references (REFS), we can assume the adiabatic boundary
conditions
∂T
= 0◦ C/m at z = 15m. (14)
∂z
∂T
= 0◦ C/m at x = 0 m and x = 10 m. (15)
∂x

6
To keep the continuity of the thermal fluxes on the interfaces Γ between
soil (s) and galvanized parts (g), we impose that

−λs ∇T · n = −λg ∇T · n, (16)

where n is the unit vector normal to Γ.


Following a usual approximation in the references (REFS), we also assume
that the temperature at the soil surface is equal to the air temperature, Ta .
Consequently, the last boundary condition is:

T = Ta at z = 0 m. (17)

For this work, we fitted the annual air temperature data in Viamão by least
squares, as done in other references like [12]. In our model, Ta is given by
the periodic function
 

Ta (t) = 20.49 + 5.66 sin t − 5.30 , (18)
365

where the time t is in days.


It is worth remark that we made the simulations consistently, using the
time t in seconds. It is just by convenience that t is given in days during
pre and post-processing stages, where we used least squares to fit all the
temperature results by sine-based functions.
Following [12], we adopted an initial condition based on a one-dimensional
analytical model that considers the temperature variations only in the z
direction, similar to the approach in [13]. Thus, taking the conditions from
Viamão fitted by least squares, we used the following initial condition

T0 = 20.49 − 5.66 sin(5.30 + 0.39z)e−0.39z at t = 0 s. (19)

The tests showed that this initial condition affects only the first two
months of simulation. Hence, we ran simulations covering a total time of
one year and two months, then discarded the first two. This approach is
advantageous because it reduces the computational cost to evaluate a year
of EAHE operation. To make a comparison, in references like [14], where the
authors assume that the initial condition is equal to the annual mean soil
temperature, it affects the solution for almost one year. That means that
they needed to simulate two years and discard entirely the first one.

7
1.4. Numerical approach
The initial-boundary value problem, composed by Eq. (13) to (19), was
solved numerically. For the time discretization, we used finite differences,
more specifically, the first-order implicit Euler method [15, 16]. The finite
element method of Galerkin [17, 18, 19] was employed in the spatial dis-
cretization. We made all the simulations utilizing in-house codes, developed
in the Matlab software language. The codes cover the numerical methods
and the GAEA model.
To explore computational domains like the ones in Fig. 1, the meshes were
composed by the following 6 subdomains: soil (1), block (2), left fin (3), right
fin (4), upper fin (5) and lower fin (6); this structure is illustrated in Fig. 2.
Hence, we could use the same mesh for different cases. For example, in the
case of a simulation involving the four fins, we could define that subdomains
2 to 6 had the thermal diffusivity of galvanized material, while subdomain
1 had thermal diffusivity of the ground. On the other hand, in another
simulation where the left fin is disregarded, then, we could just define the
thermal diffusivity of subdomain 3 to be equal to the soil diffusivity. The
results for such cases are shown ahead in colocar a seção.

Figure 2: Domain parts

As done by [12], we employed discrete time intervals of 1800 seconds. For


the spatial discretization, we conducted various refining tests before adopting
meshes with about 4000 triangular elements, generating them in the GMSH
software [20]. In general, the meshes were more refined in the regions close
to the block and fins (where higher temperature gradients could occur). It
is worth note that the duct is not considered in the mesh, as the solutions of
the energy conservation equation are made without its presence.

8
In all mesh tests, we compared the annual temperature solution vectors
at the center of the block, i.e., at the coordinates x = 5 m and z = 1.6 m.
Using the infinite norm2 , we compared the differences between two successive
solution vectors T s . The conclusive tests and comparisons are presented in
Table 2, which gives also the number of nodes, elements and total simulation
time for each of them. As it can be seen in the same table, the results
converged as expected with the increase in the number of nodes and mesh
elements. Following [12], a maximum difference of the order of 10−4 between
two solutions was considered sufficient to complete the comparisons.
Table 2: Mesh tests

Solution Nodes Elements Time (min) Comparison


s ||T s+1 − T s ||∞
1 1097 2137 2.6
2 1396 2722 3.4 1.80 × 10−3
3 1657 3230 4.2 6.94 × 10−4
4 2017 3935 5.5 2.41 × 10−4

The mesh corresponding to the fourth case of Table 2 can be seen in


Figure 3. The same figure also gives a zoom view in the region of the block
and fins, where the mesh is more refined. As a general criterion for further
simulations in this work, we adopted meshes with a number of nodes greater
than 2000 and with more than 4000 elements. Thus, typical simulations
consumed less than 10 minutes in processing time.
To conclude this subsection, it is worth mention that the galvanized struc-
ture adopted here should be seen as a first adaptation of the geometry pro-
posed in [1] for the case of [6]. A clear difference is that in the work [1], the
fins are connected directly to the duct. This was not done here to maintain a
uniform region around the duct, enabling the use of less complex modeling,
as is the case with GAEA. In fact, the reference [1] does not simulate or esti-
mate the thermal potential of EAHE; it focuses on the problem of improving
heat transfer in the soil. Several other configurations are possible and we
study some of them ahead.

1.5. Additional definitions


We define the soil thermal potential Ps as the difference between the
soil temperature Ts and the air temperature at the EAHE inlet Ti . It is

2
The infinite norm of a vector x = (x1 , · · · , xn ) is defined by: ||x||∞ = maxi=1,...,n |xi |
[21]

9
Figure 3: Selected mesh with 3935 elements

worth note that we are considering Ti = Ta , which is the temperature of the


ambient air near the soil surface. Furthermore, as done in other references
like [14], we define the EAHE thermal potential Pt as the difference between
the temperatures at the outlet To and at the inlet of the ducts. Thus, at any
time t, the instantaneous values of these potentials are given, respectively,
by:

Ps (t) =Ts (t) − Ta (t) (20)


Pt (t) =To (t) − Ta (t) (21)

To evaluate the performance of EAHE, it is important to measure their


efficiency, which is given by
Pt
θt = , (22)
Ps
at any time t. Consequently, the bigger the efficiency, the closer Pt and Ps
are. On the other hand, such formula has some drawbacks, particularly in
the months near the spring or fall, when Ps is equal or close to zero. Hence,

10
we follow the references [22, 14, 12] and measure the EAHE annual efficiency
by qR
365 2
0
Pt dt
θ = qR . (23)
365 2
0
Ps dt
In this work, all the results for the potentials are fitted by least squares
to periodic functions, like

P (t) = Pm + A sen(ωt + φ), (24)

where, Pm , A, ω, and φ represent, respectively, the mean value, amplitude,


angular frequency, and phase. To do so, we also used an in-house code, made
in Matlab, following the algorithm given in [23].
Finally, we make different comparisons of results using the root mean
square value (RMS) of the annual differences between two functions. Thus,
having obtained two periodic functions f and g, written as in Eq. (24), we
calculate the RMS value of the annual difference between them by the formula
s
R 365
0
[f (t) − g(t)]2 dt
VRM S = , (25)
365
where the time t here is given in days.
é interessante fazer a Descrição geral da metodologia... como abaixo...

1.6. Model validation


We validated the model by comparing its results against experimental
data from [7, 2]. The main parameters needed for modeling are: the duct
length, L0 = 25.77 m, diameter, D0 = 11 cm, and depth, z0 = 1.6 m. Such
values also serve as an initial reference for this study and they have been
presented in Section 1.1. Another important parameter, not mentioned so
far, was the experimental air speed at the ducts inlet, i.e., v0 = 3.3 m/s.
This corresponds to an air mass flow of ṁ0 = 0.0364 kg/s. The other ther-
mophysical properties of air and soil are given in Table 1. Finally, the air
temperature in the Brazilian city of Viamão was modeled with Eq. (18).
We also can find in the literature two works that validate their models
based on the same experimental data. They are: [12], which adopts a GAEA
model where the soil temperature fields are computed by a one-dimensional
methodology based on finite differences; [4], that uses a 3D numerical model,

11
based on conservation equations that are solved by finite volume methods
with the Fluent software.
Hence, Fig. 4 presents a comparison among: (V) the discrete, daily av-
erage temperatures at the outlet of one EAHE installation from [7]; (VF)
a least squares fitted curve of the same data, following a periodic function
similar to that given in Eq. (24); (D) and (H) the numerical results from the
models of [12] and [4], respectively; (CM) our current model results.

28

24
T (oC)

20

V.
16 V.F.
D.
C.M.
H.
12
0 50 100 150 200 250 300 350
t (days)

Figure 4: Comparison among the current model results againt others from literature

As we can see from Fig. 4, the current model gives results approximatelly
equal to the ones from [12]. The Pearson correlation coefficient between such
models results and the experimental data is about 0.89. Furthermore, the
annual RMS difference between the model results and the fitted experimen-
tal values is approximately 1.45◦ C. In [12], the authors estimated that the
correlation between the model results from [4] and the experimental data are
approximately 0.92, while the annual RMS difference between their results
and the fitted experimental values is close to 1.58◦ C.
Therefore, we concluded that our methodology is valid, providing accurate
and highly correlated results to the experimental ones. Furthermore, it can

12
be seen that the methodology is computationally efficient, as the simulations
can be performed in a few minutes (see Table 2), instead of hours, as it
happens with the complete 3D model from [4].
cabe falar um pouco das limitações do GAEA, sugestão Elizaldo.

(a) Summer

(b) Winter

Figure 5: Daily average soil temperature fields in the summer and winter

To conclude this subsection, we present in Fig. 5 the estimates of our


model for the two-dimensional temperature fields in Viamão during the sum-
mer and winter. In summer, the daily average air temperatures are around
25◦ C and the minimum soil temperature is around 20◦ C. In winter, the av-
erage air temperatures drop to almost 16◦ C, while the maximum soil tem-
perature is approximately 21◦ C. In general, the higher magnitude of the soil

13
thermal potential is little more than 5◦ C.

2. Results
2.1. Introducing the block and fins
Using the mesh of Fig. 3, we tested twelve different configurations for the
block and fins coupled to a duct placed at the depth of z0 = 1.6 m. All
cases are illustrated in Fig. 6, and they are numbered from 1 to 12. The case
shown in the validation (where there is no block and no fins) is considered
a reference case, or case number 0. The article [1] considered 7 possible
arrangements, which we adapted here with the cases 0, 2, 3, 4, 5, 6, and 12.
By simetry, the case 4, corresponding to use a fin on the right side of the
block, is equivalent to use a fin on the left.

Figure 6: Twelve configuration cases for the block and fins

Table 3 presents the annual RMS values for the soil thermal potential
(Ps ) for all considered cases. The highest values occur in cases 3, 8 and 10.
On the other hand, the case 3 requires much less material, as it uses only one
fin (the bottom one). Hence, it is simpler and more economical, therefore,
we considered it the best case of the simulations. As we see ahead, further
parameter tests in this paper use just a block and a bottom fin.
Still analysing the results from Table 3, we note that all cases using a
top fin had a soil thermal potential lower than the reference case with no
galvanized material. Physically, it means that a top fin negatively influences

14
Table 3: Annual RMS values of the soil thermal potential for different cases

Case Description Ps,RM S (◦ C)


0 Reference case: no block nor fins 2.58
1 Only the block 2.58
2 Top fin 2.48
3 Bottom fin 2.64
4 Side fin 2.58
5 Two fins: top and bottom 2.53
6 Two side fins 2.58
7 Two fins: top and side 2.47
8 Two fins: bottom and side 2.64
9 Three fins: top and sides 2.48
10 Three fins: bottom and sides 2.64
11 Three fins: top, bottom and side 2.53
12 Four fins 2.53

the results, as it increases the thermal influence of the soil surface in the duct
region. Besides, using only the block, or side fins (cases 1, 4 and 6) do not
affect the soil thermal potential.
It is worth note that the values in Table 3 are annual RMS values, hence
they cannot achieve the peaks of 5◦ C, which occur only in summer or winter,
as in Fig. 5. Indeed, in spring and fall, there are times when the thermal
potential of the soil is close to zero. Moreover, the installation is at a depth
of only 1.6 m, which is not ideal, as we see ahead in this work.

2.2. Potentials and Efficiency


An annual comparison among the temperatures of the soil (Ts ), duct
outlet (To ), and ambient air (Ta ) is shown in Fig. 7, for the cases 0 and 3.
Naturally, the outlet temperature curve (red) is between the air (black) and
soil (blue) ones. In the graphics, the soil thermal potential is marked by
the distance between Ts and Ta , while the EAHE thermal potential is the
distance between To and Ta . Additionally, the closer the Ts and To are, the
higher is the EAHE efficiency (see Eq. 23).
From the previous discussion, Fig. 7 clearly shows a considerable increase
in the EAHE efficiency with the use of galvanized materials. Numerically
computing the efficiency, we obtained an approximate value of 73% for the
case 0, and 95% for the case 3. Moreover, one can estimate graphically that
the magnitude peaks of the thermal potential, both in summer and in winter,
go closer to 4◦ C in case 3, while it is close to only 3◦ C in the reference case
0.

15
(a) Without galvanized material - case 0

(b) With galvanized material - case 3

Figure 7: Annual comparison of the soil, outlet, and air temperatures.

2.3. Thermal efficiency and duct size


Adopting the block and fin configuration from case 3, we made a para-
metric study, varying the size L of the duct and computing the EAHE annual
efficiency θ. We started using the reference size L = L0 =25.77 m and kept

16
reducing it by 3 m until reaching an efficiency of less than 70%. Such results
are given in Table 4, and they show how θ reduces by reducing L.
Table 4: Comparison between the duct length (L) and the annual thermal efficiency (θ).

L(m) 25.77 22.77 19.77 16.77 13.77 10.77


θ(%) 94.88 92.68 89,58 85,21 79,07 70,47

As observed before, the case 0 has an annual efficiency close to 73%. It


corresponds to the experimental installation from reference [6] using a duct
of 25.77 m. Coupling the galvanized block and the bottom fin to such duct
yields θ ≈ 95%. From Table 4, we conclude that it is also possible to achieve
efficiencies above 70% by employing much smaller ducts, with L ranging
between 10.77 and 13.77 m. Hence, one can reduce the installation size by
almost half.
2.4. Potential and installation depth
Still adopting the block and fin configuration from case 3, we tested how
the installation depth z influences the soil thermal potential Ps . We started
from the reference depth z = z0 =1.6 m and increased it by 1 m until reach
5.6 m. Table 5 shows how the annual RMS values of Ps increase as the z
increases. Down to z =3.6 m there are relative gains of more than 10%.
Then it is harder to improve Ps as its RMS values are not much higher than
4◦ C. For instance, the relative increase between placing the duct at 5.6 m
instead of 4.6 m is less than 2%. Besides, considering that increasing z also
implies increasing digging costs [4], then it is not recommended to adopt z
above 4.6 m.
Table 5: Annual soil thermal potential, and relative gains, with installation depth

Depth(m) Ps,RM S (◦ C) Relative gain (%)


1.6 2.65 -
2.6 3.50 32.1
3.6 3.98 13.7
4.6 4.20 5.8
5.6 4.28 1.7

To complement the results, Fig.8 shows an annual comparison among


the temperatures of the soil (Ts ), duct outlet (To ), and ambient air (Ta )
considering the different values of z. It is worth noting that for all cases, the
efficiency is greater than 90%, thus Ts and To curves are close in all cases.
The same figure also illustrates how the variations in Ts and To reduce as z
increases, as expected from Table 5.

17
Figure 8: Annual comparison of the soil, outlet and air temperatures at different depths.

2.5. Thermal potential and the soil thermal conductivity


In [1], the authors measured thermal performance based on the heat ex-
change rate (q) in the soil. Regardless of the soil types explored in that work,
they found that a case with four bridges (fins) gives the maximum thermal
performance in the sense of increasing dissipation flux between the duct and
the ground. Such case is emulated here by the case 12.
Although the addition of fins can increase the general conductivity of
the soil around the duct, as well as the heat flux dissipation capacity, this
work shows that it increases the EAHE efficiency but not the soil thermal
potential. The latter performance aspect was not addressed in the work [1]
and, considering it, the best result here is obtained by adding just one fin.
To further explore such theme, we studied how increasing the soil thermal
conductivity around the duct affects the EAHE potentials. In particular, we
analyse the thermal potential when the shapes of the high conductive area
around the duct changes.

2.5.1. Shape 1: circular high conductive region


We began considering simulations where a circular region around the duct
has the high conductivity of a galvanized material, while the rest of the two

18
dimensional domain has the same thermal conductivity of the soil. Such
configuration is illustrated in Fig. 9, where the galvanized region has the
shape of a disk of radius r, centered on x0 and z0 (i.e. the duct center).

Figure 9: Typical computational domain with a galvanized disc.

Here, the galvanized disk area occupies different proportions of the com-
putational domain. The variations took into account the following consider-
ations:

ˆ The domain area is the constant Ac =150 m2 , while the disk area (Ad )
varies as a function of r.
ˆ The area fraction between disk and domain areas is given by
Ad π r2
ψ= 100% = %.
Ac 1.5

ˆ The radius r is limited to vary between D0 /2=0.055 m and 1.6 m.


Otherwise, the disk area is less than the duct cross sectional area or it
leaves the domain.
ˆ For r=0.055 m, ψ ≈0.006%, while for r= 1.6 m, ψ ≈5.4%. Hence, the
simulations were divided into seven cases, varying ψ from 0.1% to a
maximum of 5%.

As for the rest of the EAHE modeling, we are adopting the same reference
values from the validation, that is, a duct at depth z0 , with length L0 , and

19
diameter D0 . Basically, the same previous methodology from section 1, but
adopting domains similar to those in Fig. 9 instead of using blocks and fins.
From the results in Table 6, it is clear that the annual RMS values of the
soil thermal potential increase with the reduction of the disk area. The last
line of the same Table also includes the reference case without any galvanized
material. In terms of soil thermal potential, there is no gain in adopting a
disk shape for the galvanized area. On the other hand, as the efficiency goes
from 73% to almost 95% using the galvanized material, there is a gain of
almost 30% in the EAHE thermal potential using just ψ=0.1%.

Table 6: Results with the variation of the galvanized disk.

ψ(%) r(m) Ad (m2 ) Ps (◦ C) Pt (◦ C ) θ(%)


5 1.60 8.04 1.34 1.30 94.9
4 1.38 5.98 1.83 1.76 94.8
3 1.20 4.52 2.13 2.04 94.9
2 0.97 2.96 2.33 2.23 94.9
1 0.69 1.50 2.48 2.37 94.9
0.5 0.49 0.75 2.54 2.43 94.9
0.1 0.22 0.15 2.58 2.46 94.9
0 0 0 2.58 1.90 73.0

Another way to see what happens with the introduction of the galvanized
circular area is considering the temperature fields in the computational do-
main. Hence, Fig.10 presents them in the summer (a) and winter (b), taking
ψ=5%. Similarly, Fig.11 shows them for ψ=0.1%.
From Fig.10, one can see that the presence of a large disk (or a large
galvanized region) causes the region around the duct to be strongly influenced
thermally by what happens on the surface. Due to the high conductivity, the
temperatures at the duct position approach those on the surface both, in
summer and winter. Consequently, such unwanted behavior reduces the soil
thermal potential. This situation is well contrasted by observing Fig.11,
where the disk area occupies only 0.1% of the domain. In this case, the
temperatures on the duct region are much different from those on the surface,
thus enabling a higher soil thermal potential.
Figure 12 presents the ambient and outlet air annual temperatures with
the disk occupying 5 and 0.1% of the computational domain. Both cases are
very different throughout the year, as the EAHE thermal potential in the
former is almost half of the latter.
It is also worth taking a look at the result for case 1 in Table 3, where we
used only a square block centered at the duct center. In that case, the block

20
(a) Summer

(b) Winter

Figure 10: Temperature fields for ψ = 5%.

area also occupied a small part of the computational domain (about 0.02%).
The soil thermal potential for that case was also equal to the reference case
without any galvanized materials. Such results suggest that using small and
symmetric galvanized areas around the duct center does not increase the
soil thermal potential. However, they are suitable for increasing the EAHE
efficiency and, consequently, the EAHE thermal potential.
In short, increasing the thermal conductivity of the region around the
duct does not imply increasing the soil thermal potential, as it depends on
the temperature difference between the surface and the duct location.

21
(a) Summer

(b) Winter

Figure 11: Temperature fields for ψ = 0.1%.

2.5.2. Shape 2: rectangular high conductive region


In this section, we consider simulations adopting a rectangular (block)
region around the duct with the high thermal conductivity of galvanized
materials, while the rest of the two dimensional domain has the same thermal
conductivity of the soil. Such configuration is illustrated in Fig. 13.
Here, x0 and z0 are (again) the coordinates of the duct center; SV and SH
represent, respectively, the vertical and horizontal sizes of the block. Among
the infinite possibilities for the block configuration, we took the following.

ˆ The rectangular block is similar to the rectangular domain.

ˆ The block is centered in the horizontal direction.

22
Figure 12: Comparison among annual air and outlet temperatures.

Figure 13: Typical computational domain with a rectangular galvanized block.

23
ˆ H varies between 0 and 5 m, hence SH can range from 0 to 10 m.

ˆ The vertical variation follows the proportion:


15V1
SV = .
z0
Thus, within the variation limits of V1 , between 0 m and z0 , SV varies
from 0 to 15 m. Note that V2 depends on V1 , ie V2 = SV −V1 . Therefore,
V2 can only vary between 0 and 15 m - z0 .

ˆ The block and the domain have proportional areas. To ensure that, we
assume
2SV
SH = .
3
Thus, there are two limiting cases: (a) SV = 0 m implies SH = 0 m,
and the block converges to a point (the duct center); (b) SV = 15 m
implies SH = 10 m, and the block occupies the entire computational
domain. In short, cases (a) and (b) represent extremes where the entire
domain is composed, respectively, by soil or galvanized material.

ˆ The domain area is again the constant Ac =150 m2 , while the block
area is Ab = SH × SV . Hence, the area fraction between block and
domains areas is given by
Ab SH × SV
ψ= 100% = %.
Ac 1.5

Here, we made several simulations, varying ψ from 1 to 90%, and mea-


suring annual RMS values for the EAHE thermal potential (Pt ), efficiency
(θ) and soil thermal potential (Ps ). For all cases using a block, θ remained
almost constant and close to 95%. The behaviours for Pt and Ps are shown
in Fig. 14.
Using a rectangular block, the maximum for Ts does not happen when
ψ=0% (as it happened for the disk). Both thermal potential decrease when ψ
grows too much, or decreases too much. There is a maximum value for both
potentials when 1%< ψ <10%. To better see such behaviour, Fig. 15 gives a
zoom in the results for ψ varying from 0 to 10%. The thermal potential are
fairly constant when ψ varies between 3 to 10%; in this interval, Ps and Pt
are close to maximum values of 3.2 and 3.0 ◦ C, respectively.

24
Figure 14: Thermal potential variations with ψ.

Figure 15: Zoom - thermal potentials with ψ varying from 0 to 10%.

Figure 17 shows the temperature fields in summer and winter for ψ =90%;
while Fig. ?? shows them for ψ =10%. Hence, for ψ =10%, the block helps
to make the region around the duct colder or warmer than the surface. In
such case, the block region improves the connection between the duct and

25
the deeper soil layers. On the other hand, for ψ =90%, the block connects
the regions close to the surface with the deeper ones. In this situation, the
surface temperature significantly influence the process. It can be noted that
the region around the duct has a temperature close to the surface, which
reduces the soil and EAHE thermal potentials.

(a) Summer (b) Winter

Figure 16: Summer and winter temperature fields for ψ = 90%.

(a) Summer (b) Winter

Figure 17: Summer and winter temperature fields for ψ = 10%.

Such results show once more that only increase the thermal conductivity
on the region around the duct does not imply a better thermal performance.
To improve the soil (and EAHE) thermal potential, it is necessary to think
of suitable shapes for the galvanized area, or galvanized fins coupled to the
ducts.

26
References
[1] R. Hassanzadeh, M. Darvishyadegari, S. Arman, A new idea for improv-
ing the horizontal straight ground source heat exchangers performance,
Sustainable Energy Technologies and Assessments 25 (2018) 138–145.

[2] J. Vaz, M. A. Sattler, R. S. Brum, E. D. Santos, L. A. Isoldi, An ex-


perimental study on the use of earth-air heat exchangers (eahe), Energy
and Buildings 72 (2014) 122–131.

[3] R. S. Brum, J. Vaz, L. A. O. Rocha, E. D. D. Santos, L. A. Isoldi, A


new computational modeling to predict the behavior of earth-airheat
exchangers, Energy and Buildings 64 (2013) 395–402.

[4] V. F. Hermes, J. V. A. Ramalho, L. A. O. Rocha, E. D. dos Santos,


W. C. Marques, J. Costi, M. K. Rodrigues, L. A. Isoldi, Further realistic
annual simulations of earth-air heat exchangers installations in a coastal
city, ELSEVIER - Sustainable Energy Technologies and Assessments 37
(2020).

[5] M. K. Rodrigues, F. S. Coswig, K. R. Camargo, L. A. Isoldi, R. S.


Brum, J. V. A. Ramalho, J. Vaz, L. A. O. Rocha, E. D. Santos, Thermal
performance simulations of earth-air heat exchangers for different soils
of a coastal city using in-situ data, Sustainable Energy Technologies and
Assessments 30 (2018) 224–229.

[6] J. Vaz, Estudo experimental e numérico sobre o uso do solo como reser-
vatório de energia para o aquecimento e resfriamento de ambientes
edificados, Ph.D. thesis, Universidade Federal do Rio Grande do Sul
(UFRGS), Porto Alegre, Brazil (2011).

[7] J. Vaz, M. A. Sattler, E. D. Santos, L. A. Isoldi, Experimental and


numerical analysis of an earth-air heat exchancher, Energy and Buildings
43 (2011) 2476–2482.

[8] A. M. Dominguez, E. S. Nóbrega, J. V. Ramalho, R. S. Brum, R. S.


Quadros, Towards owards earth-air heat exchangers in pelotas, in: As-
sociação Brasileira de Engenharia e Ciências Mecânicas - ENCIT2020,
Ribeirão Preto, Brazil, 2020.

27
[9] F. Ascione, L. Bellia, F. Minichiello, Earth-to-air heat exchangers for
italian climates, Renewable Energy 36 (2011) 2177–2188.

[10] S. Benkert, F. D. Heidt, D. SchÖler, Calculation tool for earth heat ex-
changers gaea, in: Proceedings Building Simulation, Fifth International
IBPSA Conference 2, -, 1997.

[11] T. Papakostas, A. Tsamitros, G. Martinopoulos, Validation of modified


one-dimensional models simulating the thermal behavior of earth-to-
air heat exchangers-comparative analysis of modelling and experimental
results, Geothermics 82 (2019) 1–6.

[12] A. M. B. Domingues, E. S. B. NÓbrega, J. V. A. Ramalho, R. S. Brum,


R. S. Quadros, Parameter analysis of earth-air heat exchangers over
multi-layered soils in south brazil, Geothermics 93 (2021) 1–14.

[13] O. Ozgener, L. Ozgener, J. W. Tester, A practical approach to predict


soil temperature variations for geothermal (ground) heat exchangers ap-
plications, International Journal of Heat and Mass Transfer 62 (2013)
473–480.

[14] R. S. Brum, J. V. A. Ramalho, M. K. Rodrigues, L. A. O. Rocha, L. A.


Isoldi, E. D. Santos, Design evaluation of earth-air heat exchangers with
multiple ducts, Renewable Energy 135 (2019) 1371–1385.

[15] M. N. Özisik, Heat Conduction, John Wiley & Sons, New York, 1993.

[16] C. F. GERALD, P. O. WHEATLEY, Applied Numerical Analysis,


Addison-Wesley, New York, 1990.

[17] T. J. R. Hughes, The finite element method (Linear Static and Dynamic
Finite Element Analysis), Prentice Hall, INC., New Jersey, 1987.

[18] E. B. Becker, G. F. Carey, J. T. Oden, Finite Elements An Introduction,


Prentice Hall, New Jersey, 1981.

[19] A. Logg, K. A. Mardal, G. N. Wells, Automated Solution of Differential


Equations by the Finite Element Method, Springer, Berlin, 2012.

[20] Geuzaine, Christophe and Remacle, Jean-Francois, Gmsh (2020).


URL http://http://gmsh.info/

28
[21] L. N. Trefethen, D. Bau, Numerical Linear Algebra, SIAM, 1997.

[22] R. S. Brum, Teoria construtal e desempenho térmico de trocadores de


calor solo-ar, Ph.D. thesis, Universidade Federal do Rio Grande do Sul
(UFRGS), Porto Alegre, Brasil (2016).

[23] R. S. Brum, J. V. A. Ramalho, L. A. O. Rocha, L. A. Isoldi, E. D. Santos,


A matlab code to fit periodic data, Revista Brasileira de Computação
Aplicada 7 (2015) 16–25.

29

You might also like