You are on page 1of 23

Transp Porous Med (2018) 122:145–167

https://doi.org/10.1007/s11242-017-0995-9

Numerical Simulation of Turbulent Flow and Heat


Transfer in a Three-Dimensional Channel Coupled with
Flow Through Porous Structures

Guang Yang1,2 · Bernhard Weigand2 ·


Alexandros Terzis2 · Kilian Weishaupt3 ·
Rainer Helmig3

Received: 22 August 2017 / Accepted: 18 December 2017 / Published online: 3 January 2018
© Springer Science+Business Media B.V., part of Springer Nature 2018

Abstract This study investigates numerically the turbulent flow and heat transfer charac-
teristics of a T-junction mixing, where a porous media flow is vertically discharged in a 3D
fully developed channel flow. The fluid equations for the porous medium are solved in a
pore structure level using an Speziale, Sarkar and Gatski turbulence model and validated
with open literature data. Overall, two types of porous structures, consisted of square pores,
are investigated over a wide range of Reynolds numbers: an in-line and a staggered pore
structure arrangement. The flow patterns, including the reattachment length in the channel,
the velocity field inside the porous medium as well as the fluctuation velocity at the interface,
are found to be strongly affected by the velocity ratio between the transversely interacting
flow streams. In addition, the heat transfer examination of the flow domain reveals that the
temperature distribution in the porous structure is more uniform for the staggered array. The
local heat transfer distributions inside the porous structure are also studied, and the general
heat transfer rates are correlated in terms of area-averaged Nusselt number accounting for
the effects of Reynolds number, velocity ratio as well as the geometrical arrangement of the
porous structures.

Keywords Porous structure · Convective heat transfer · Turbulence · Numerical simulation ·


Pore scale

List of Symbols
cp Specific heat capacity of the fluid [J/(kg K)]
D Size of the microcolumn (m)

B Guang Yang
y_g@sjtu.edu.cn; guang.yang@itlr.uni-stuttgart.de
1 Institute of Refrigeration and Cryogenics, Shanghai Jiao Tong University, 200240 Shanghai, China
2 Institut für Thermodynamik der Luft- und Raumfahrt, Universität Stuttgart, 70569 Stuttgart,
Germany
3 Institut für Wasser- und Umweltsystemmodellierung, Universität Stuttgart, 70569 Stuttgart, Germany

123
146 G. Yang et al.

Ek Turbulence kinetic energy (m2 /s2 )


H Distance of neighbouring columns (m)
k Thermal conductivity [W/(m K)]
L Length of the channel cross section (m)
Lr Reattachment length (–)
p Pressure (Pa)
P Pressure (–)
Pr Prandtl number (–)
r Grid refinement ratio
Re Reynolds number (–)
Rec,L See Eq. 4
Rep,D See Eq. 6
t Time (s)
T Temperature (K)
V Velocity (–)
VR Velocity ratio (–)
X, Y, Z Dimensionless coordinate system (–)

Greek Symbols

ρ Density (kg/m3 )
τ Dimensionless time (–)
μ Dynamic viscosity of the fluid (Pa s)
θ Temperature
ϕ Porosity (–)

Subscripts

c Channel
p Porous media
x, y, z Cartesian coordinates

Superscript
 Dimensional variable

1 Introduction

The understanding of coupled free flow and flow through coarse porous structures is of great
significance due to its direct applicability in many industrial processes (transpiration cooling,
fuel cells, oil extraction, etc), environmental flows (water flows over seabed, melt migration)
as well as biological applications (flow at the wall of organic tissues) (Davis 2002; Kollet and
Maxwell 2006; Levenberg et al. 2005; Wu et al. 2017). For example, porous media are used
in transpiration cooling due to their large specific surface area that enhances heat transfer.
The coolant flows through the porous structure removing heat by convection and then forms
a film layer near the surface which blocks the heat transfer from the main flow.

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 147

Due to the complex geometry of the porous structures, a macroscopic model is used
in most of numerical studies, where general transport equations are integrated over a rep-
resentative elementary volume (REV) that accommodates both the fluid and solid phases.
This approach was derived from an empirical point of view by Darcy’s equation, with its
Brinkman’s correction or Forchheimer’s equation in proper cases. Dahmen et al. (2015)
conducted a numerical investigation on the boundary layer of transpiration-cooled turbulent
channel flow. Separate solvers for the hot gas flow in the channel and the porous medium
flow were used, which were coupled to each other by boundary conditions imposed at the
interface. Liu et al. (2013) studied the flow and heat transfer characteristics for a coupled
flow through sintered porous flat plates with hot air flow. In their simulation, the main stream
flow was calculated by solving the Navier–Stokes equations using the standard k − ε, SST
k − ω and RNG k − ε turbulence models, while the Brinkman–Forchheimer-extended Darcy
equation was used to model the coolant flow through the porous media. As for the energy
equation within the porous medium, there are mainly two different models: the local thermal
equilibrium model and the local thermal non-equilibrium model (where two energy equations
are solved). The local thermal equilibrium model assumes that the solid temperature is equal
to the fluid temperature, and hence, the heat exchanges between solid and fluid phases in the
porous medium are not separately considered and only one energy equation is solved. As
the assumption of local thermal equilibrium between the fluid and the solid is inadequate for
many applications, a local thermal non-equilibrium model has also been developed for the
investigation of conduction and convection heat transfer processes in porous media (Nuske
et al. 2014; Vortmeyer and Schaefer 1974). However, both the momentum and two-equation
energy equations of macroscopic models need additional input parameters from experiments,
pore-scale numerical simulations or empirical equations (Mayer 2014; Yang and Nakayama
2010).
Contrary to the macroscopic model, pore-level numerical simulations have been also
performed in porous structures in order to obtain the microscopic aerothermal features of the
flow. By using an upscaling procedure, the microscopic description can be used to improve
macroscopic models. The pore-scale geometry of the porous media can be reconstructed from
scanning electron microscope (SEM) or microcomputed tomography (CT) images (Acosta
and Camacho 2009; Ghanbarzadeh et al. 2015), and then, the transports in solid and fluid
phases are solved separately. The structural units composing the porous structures were also
often simplified as basic shapes such as cylinders, cubes, spheres (Gao and Xu 2015; Torabi
et al. 2016; Uth et al. 2016; Xu and Jiang 2008), as the geometry models are easier to build
with high precision while some macroscopic parameters, such as porosity, can be directly
inferred. By considering the effects of microstructure parameters, such as particle shape,
orientation, pore to throat size ratio, and unit cell distributions (Ozgumus and Mobedi 2016;
Yazdchi et al. 2011; Liu et al. 2017), the features for a particular configuration of porous
medium can be extended to more general cases.
If the velocity in the porous medium is high, e.g. the Reynolds number based on particle
diameter is higher than 100–300 (Pedras and de Lemos 2001; Prausnitz and Wilhelm 1957;
Uth et al. 2016), a turbulence-like structure will occur within the pores. The flow and heat
transfer in the porous structures, the boundary layers in the channel as well as at the interface
of the two domains will become much more complex (Brillant et al. 2008a, b; Suga et al.
2017). This is because the parameters in the two domains strongly interact with each other.
Fetzer et al. (2016, 2017) extended the concept, where compositional single-phase free
flow was coupled to compositional two-phase porous medium flow under non-isothermal
conditions, to turbulent free-flow conditions. Rosti et al. (2015) carried out a number of
direct numerical simulations of turbulent channel flow over two porous walls. Turbulence

123
148 G. Yang et al.

statistics and instantaneous flow fields were used to study the main changes introduced by
the porous material. Zampogna and Bottaro (2016) studied the interaction between a fluid
flow and a transversely isotropic porous medium, and different interface conditions were
evaluated in their study. Lācis and Bagheri (2017) computed the effective boundary conditions
at the interface between a porous surface and an overlying flow. They derived a tensorial
generalized version of the empirical condition suggested by Beavers &Joseph and compared
the homogenized model results with DNS. However, the coupled effect considering both the
imposed turbulent flows in a channel and in a porous medium, as well as the heat transfer
effect have not been fully understood yet. Therefore, a fully resolved pore-scale numerical
simulation is necessary to obtain the microscopical parameters at the flow domain.
The objective of the present work is to gain a deep insight into the flow and heat transfer
characteristics in coupled turbulent flows consisting of a 3D channel flow and a flow through
a porous medium, and to reveal the interaction effect of the two kinds of flows. The continuity,
energy and Navier–Stokes equations, together with a Reynolds stress SSG turbulence model
for the fluid inside the pore structures, are solved simultaneously with those in the channel
flow. Two types of porous structures are considered: an in-line and a staggered microarray of
square pillars are examined over a wide range of Reynolds numbers. The pore-scale velocity
and temperature profiles inside the porous structure and at the interface, as well as the flow
reversal in the channel are in detail investigated over a wide range of Reynolds numbers. The
results are analysed by various post-processing procedures aiming to quantify and clarify
the flow and heat transfer characteristics of the coupled flow domain at various conditions.
In addition, the heat transfer rates inside the porous medium are successfully correlated in
terms of area-averaged Nusselt numbers.

2 Physical Description and Governing Equations

The geometry and the relevant dimensions are schematically shown in Fig. 1a. Nitrogen gases
with a uniform free-stream velocity (Vc ) and a constant temperature (Tc ) enter a channel with
a square cross section of L × L . Similarly, nitrogen gases with a uniform velocity (Vp ) and
constant temperature (Tp , higher or lower than Vc ), are injected through the porous structure
and mix with the channel flow in a T-junction arrangement. Similar to the channel, the branch
with the porous medium has a cross section of L × L and a length of L/2. The surfaces of the
solid skeleton of the porous structure are kept at constant temperature Tc , equal to the channel
flow. All other walls in the model are considered adiabatic. The channel length extends 10L
upstream of the porous zone in order to ensure fully developed conditions. The channel length
downstream the porous zone is 20L, in order to obtain the full flow patterns resulted from
the transverse mixing of the two flow streams. All solid surfaces are considered by using a
no-slip condition for the fluid velocity.
An in-line and a staggered pattern geometrical arrangements are considered for the porous
medium, both of them consisting of cubic square microarrays, as shown in Fig. 1b, c. All
square columns have a cross section of D × D, and the centrelines are in Z -direction. The
distance between neighbouring columns is H . The porosity of the porous structures can be
determined by ϕ = 1 − (D/H )2 . In this study, D = L/40 and H = 2D resulting a porosity
ϕ value of 0.75.
The flow is considered to be incompressible with constant fluid properties (ρ = 1.2kg/m3 ,
c p = 1006 J/(kg K), k = 2.54 × 10−2 W/(m K), and μ = 1.82 × 10−5 Pa s), and the effect
of buoyancy is neglected. This is reasonable in many applications since the Mach number is

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 149

Fig. 1 a Schematics of flow in a 3D channel coupled with flow through porous structures; b porous structure
with in-line array and c staggered array

less than 0.25 and the Richardson number (Yang et al. 2013) is far less than 0.1, if L = 0.1m
and the temperature difference between Tp and Tc is less than 100 K. As the flow in the
porous structure is modelled on the pore scale, fluids in both the channel and within the
porous structures are governed by the continuity equation, the Navier–Stokes equations and
the energy equation. The non-dimensional form of equations is given by:
∂ Vi
=0 (1)
∂ xi
   
∂ Vi ∂   ∂P ∂ 1 ∂ Vi ∂Vj  
+ Vi V j = − + + − Ui U j (2)
∂τ ∂Xj ∂ Xi ∂ X j Rec,L ∂ X j ∂ Xi
 
∂θ ∂   ∂ 1 ∂θ
+ Vjθ = − V j θ  (3)
∂τ ∂Xj ∂ X j Pr Rec,L ∂ X j
where i = 1, 2 or 3, j = 1, 2 or 3, X 1 = X , X 2 = Y , X 3 = Z , V1 = Vx , V2 = Vy
t·Vc Vi X i
and V3 = Vz . The non-dimensional variables are defined as: τ = L , Vi = Vc , Xi = L ,
T −T
P = ρVp 2 , and θ = T −Tpc . −Ui U j is the Reynolds stress, computed using a Reynolds stress
c
turbulence model (RSM) (Hanjalic and Launder 1972).
The quantity Rec,L in the above equations is the Reynolds number characterizing the
channel flow, based on the inlet velocity into the channel Vc and the characteristic length L,
which is expressed as:

123
150 G. Yang et al.

ρVc L
Rec,L = (4)
μ
The Prandtl number is defined as:
μcp
Pr = (5)
k
The value of the Prandtl number is kept constant, 0.72, in the present study.
The dimensionless physical boundary conditions are given by:
∂θ
Channel inlet (X = −10): Vx = 1, Vy = Vz = 0, θ = 1, and ∂Y = ∂∂θZ = 0.
Inlet of the porous structures (Y = 0): Vy = Vp /Vc , Vx = Vz = 0, θ = 0, and ∂∂θX =
∂θ
∂ Z = 0.
Channel outlet (X = 21): ∂∅∂ X = 0, ∅ is any variable.
∂θ
Channel walls: Vx = Vy = Vz = 0 (no-slip) and ∂n = 0. n is the direction perpendicular
to the wall.
Surface of porous structures: Vx = Vy = Vz = 0, and θ = 1.
The Reynolds number based on the macroscopic velocity in the porous structures and the
size of the microcolumns D is also used in this study, which is defined as

ρVp D
Rep,D = (6)
μ

where Vp is the macroscopic uniform velocity in the porous structure, defined as:

Vp = Vp,D ϕ (7)

where Vp,D is the microscopic average velocity in the pores of the porous structure and is
calculated by the superficial velocity at the inlet of the porous structure Vp :

H
Vp,D = Vp (8)
D
The velocity ratio (VR) of the two inlets is also considered, which is expressed as:

VR = Vp /Vc (9)

The VR also equals to the ratio of the mass flow rate through the porous structures to that
in the channel, since the density is considered to be constant and the areas of the two inlets
are the same in this study. It can be inferred from Eq. (4), (6), (7), (8) and (9) that Rec,L ,
Rep,D , and VR are not independent, but related by the following equation:

Re p,D L
VR = (10)
Rec,L H ϕ
The heat transfer between the solid porous structure and the fluid is calculated by the Nusselt
number:

hL ∂θ 
Nu = =− (11)
k ∂n surface

The average Nusselt number in the porous structure is obtained by averaging the local Nusselt
number over the complete surface of the porous medium.

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 151

Fig. 2 Numerical grid of the flow domain. The porous structure of the in-line array is displayed

3 Numerical Details

3.1 Solution Methodology

The simulations are conducted with the commercial CFD package ANSYS CFX-17.1 on a
cluster of 28 CPU nodes with dual processor Intel(R) Xeon(R) CPU E5-2690 v4 @ 2.60 GHz.
The Reynolds stress model (RSM) of Speziale, Sarkar and Gatski (SSG) (Speziale et al. 1991)
is used for modelling the turbulence, including a scalable wall function (ANSYS Inc 2016).
This model has been proven to perform better than any two-equation turbulence model for
the simulation of the pressure loss in complex metallic porous structures by comparison with
experimental data (Mayer 2014). High-resolution schemes are used for both advection and
turbulence (ANSYS Inc 2016). The calculations were run in stationary mode and they were
considered to convergences when the residuals of all variables were less than 10−5 .
The grid structure is shown in Fig. 2 (only the porous structure of an in-line arrangement is
displayed). The grid cells of the whole computational domain were rectangular cuboids. The
structural grid distribution employed was non-uniform, varied in all directions, providing a
finer mesh near the walls as well as the internal surface of the porous skeleton.

3.2 Grid Independence Test

A grid sensitivity study was performed in order to select the proper number of grid cells to be
used for the calculations. The test was performed for Rec,L = 1 × 105 and Rep,D = 1 × 103
for an in-line array porous structure. Total element numbers of about 4.59 million (N1), 6.4
million (N2), and 9.7 million (N3) were tested, as shown in Table 1. The grid convergence
index (GCI) (Celik et al. 2008), which is based on a Richardson extrapolation and has been
successfully used in complex flow patterns (Caggese et al. 2013; Fechter et al. 2013), is
calculated to quantify the discretization error of the solution between the coarse and fine

123
152 G. Yang et al.

Table 1 Grid independence test N1 N2 N3


results for Rec,L = 1 × 105 , and
Rep,D = 1 × 103 , for the porous 4,594,212 6,441,344 9,017,212
structure of in-line array
r21 r32 Nu1 Nu2 Nu3

1.4 1.4 91.3729 92.3425 92.8134

p GC I 21 GCI32

2.146 1.24% 0.60%

grids. For the present study, the GCI was computed based on the average Nusselt number of
the porous structures. Grids N1 to N3, with global refinement factors of r = 1.4 for each
refinement step, were examined. p = 2.146 is the order of convergence. The discretization
error of N2 is 0.6%, and thus it was used for the calculations. Similar tests were also performed
for the staggered array, and a total grid number of about 6.4 million has been also proven to
be sufficient.

3.3 Validation of the Numerical Method

In order to validate the present numerical method, direct comparisons with experimental
and numerical studies are conducted, and the validation is presented in Fig. 3a, b. Figure 3a
shows the variation of Nusselt number with Reynolds number for the flow through an array
of two-dimensional staggered square columns. All the column surfaces are isothermal and
maintained at a constant temperature different from the temperature of the bulk flowing fluid.
Periodic boundaries are used for the simulations. Over a large range of Reynolds numbers
(characterized by the size of the square column), the here predicted Nusselt numbers agree
well with those reported by Saito and Lemos (2006), Wakao et al. (1979) and Kuwahara et al.
(2001).
In addition, the present numerical scheme and code were used to calculate the flow pattern
in a three-dimensional T-junction flow without porous medium, which had been experimen-
tally studied by Hirota et al. (2006). The same boundary parameters as described in their
experiments are used for the simulation. The contours of Vx /Vin at the middle plane of the
channel for the case of AR (aspect ratio) = 2 and VR (velocity ratio) = 1, are presented in

Fig. 3 Comparison of a area-averaged Nusselt number and b streamwise velocity Vx downstream of the
mixing junction with benchmark results

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 153

Fig. 3(b), indicating excellent agreement with the experimental flow pattern as well as the
length of the reversal zone downstream of the junction. The comparison of the distribution of
the local streamwise velocity at X/B = 1 is presented on the right-hand side of the contours
in Fig. 3b, which also shows a good accordance. Geometrically, the model in the present
study can be regarded as a combination of the above two models. Based on the above two
validation cases, it can be safely concluded that the numerical method used in this study is
able to represent reasonably the complex flow and heat transfer of coupled turbulent flow
with the porous medium.

4 Results and Discussion

Numerical simulations of the turbulent flow and heat transfer in a three-dimensional channel
coupled with flow through a porous structure are performed for the parameter range of
1 × 104 ≤ Rec,L ≤ 5 × 105 and 5 × 102 ≤ Rep,D ≤ 1 × 104 , considering two types of porous
structures.

4.1 General Flow Patterns

Figure 4a, b shows the dimensionless streamwise velocity Vx at the middle plane of the
channel and the porous zone (Z = 0.5) for the two types of porous structures at Rec,L =
1 × 105 and Rep,D = 1 × 103 , that result in a velocity ratio of VR = 0.267. First, it can

Vx
26
24
22
20
18
16
14
12
10
8
6
4
L/2 2
0
-2
-4
-6
-8
-10

(a) (b)
Fig. 4 Distribution of dimensionless streamwise velocity Vx in the middle plane of the channel (Z = 0.5) for
Rec,L = 1 × 105 and Rep,D = 1 × 103 for a in-line and b staggered array arrangements of porous structures

123
154 G. Yang et al.

Vy
16
15
14
13
12
11
10
9
8
7
6
5
4
L/2 3
2
1
0
-1
-2
-3

(a) (b)
Fig. 5 Distribution of dimensionless crosswise velocity Vy in the middle plane of the channel (Z = 0.5) for
Rec,L = 1 × 105 and Rep,D = 1 × 103 for a in-line and b staggered arrangements of porous structures

be seen that the velocity distributions in the channel are similar for both pore structures
arrangements. A flow reversal zone in the channel can be found downstream of the porous
zone, which is similar to the those observed in clear T-junction flows (Hirota et al. 2006). In
addition, a close-up look at the interface in the middle of the channel shows that the overall
macroscopic velocity profiles also do not significantly differ. However, on the pore scale of
the porous structures, the inter-pore velocities behave strongly different for the two types of
porous structures. For Rec,L = 1 × 105 and Rep,D = 1 × 103 , a larger region of negative
velocity can be found in the pores of the staggered array and the magnitudes of both negative
and positive velocities are higher than those in the in-line arrangement.
The dimensionless crosswise velocity Vy is presented in Fig. 5. Similarly to Fig. 4, the
velocity contours in the channel are also similar for the two column arrangements, while they
are quite different inside the pores of the porous medium. In the channel, the maximum Vy is
located above the porous zone while the velocity at the trailing edge of the porous structure
and near the interface is significantly larger for the in-line array. This is due to a clearer flow
path where the flow through the porous medium encounters less solid obstacles towards the
free flow channel, and hence, lower pressure drop (The dimensionless pressure differences
(P/(1/2ρVp2 )) between the inlet and outlet of the porous medium region are 21.78 and 51.40,
respectively, for in-line and staggered porous structures at the same boundary conditions in
Fig. 5.) With other words, the pillars of the staggered arrangement somehow eliminate the
development of the crosswise velocity component due to the systematic blockage of the
vertical flow.

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 155

Fig. 6 Distribution of dimensionless velocities Vx and Vy near the interface between the channel and the
porous structure for VR = 0.267. a Vx for in-line array, b Vy for in-line array, c Vx for staggered array, d Vy
for staggered array

4.2 Velocity Near the Interface

A quantitative comparison of the velocity distributions at the interface between the channel
and the porous medium is shown in Fig. 6, which plots the normalized Vx and Vy velocity
components along the X -coordinate of the porous structure. The locations are chosen 0.5D/L
above and below the interface Y = 0.5. As expected, both Vx and Vy fluctuate according
to the locations of the pores of the porous structure. The trends of the velocities at the two
locations are similar for the in-line array but differ for the staggered array. In Fig. 6a, different
inlet velocities at both the channel and porous structure are chosen, but the velocity ratios are
always kept constant at VR = 0.267. It can be seen that the velocities at the same location
are coincident for the different combinations of inlet velocities but the same velocity ratio,
which means that the overall flow pattern is only affected by the velocity ratio. As qualitatively
shown in Fig. 5, a higher Vy velocity component at X = 1 for the in-line array can also be
found when comparing Fig. 6b, d. Although the velocity level is similar at X = 0, it can
be clearly observed that Vy is about 40–45% higher for the in-line array compared to the
staggered arrangement.
The variations of dimensionless streamwise velocity Vx with Y-coordinate at different
locations of the middle plane (Z = 0.5) for in-line porous structure at Rec,L = 1 × 105 and
Rep,D = 1 × 103 are presented in Fig. 7. Though microscale oscillations exist in the porous

123
156 G. Yang et al.

Vx Vx
0 0.8 1.6 0 0.8 1.6
1.5

1.0
Y

0.5

0
0 0.8 1.6 -0.4 0.4 1.2 2.0
Vx Vx

Fig. 7 Variation of Vx with Y-coordinate at different locations

structures, a slip streamwise velocity can be found at the interface between the channel and
the porous media. It has been proved that slip velocity may have a significant physical effect
on the characteristics of the overlying fluid. As the flow develops downstream, the thickness
of the velocity boundary layer in the channel is clearly increased, which is affected by the
upward flow from the porous structures.

4.3 Flow Reversal in the Channel

As observed in Figs. 4 and 7, there is a flow reversal zone in the channel downstream
of the porous structure. The dimensionless lengths of the flow reversal zone for the two
types of porous structures and different velocity ratios are shown in Fig. 8. The length of the
reversal zone (reattachment length, L r ), for both the in-line and the staggered arrays, increases
monotonously with increasing VR. This is expected since the boundary layer in the channel
is affected by the flow through the porous structures. Over the complete range of velocity
ratios, the reattachment length of the in-line array is about 20% longer than the staggered one,
as a direct consequence of the higher Vy velocity components at trailing edge of the porous
medium (X = 1), shown in Fig. 6b. Furthermore, at constant VR, the reattachment length
L r is almost independent from the Reynolds number for both porous structure arrangements,
indicating the incompressible character of the flow. This further proves that for a fixed type
of porous structure, the flow pattern is mainly affected by the velocity ratio.

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 157

Fig. 8 Length of the flow


reversal zone downstream the
porous zone for various velocity
ratios and Reynolds numbers

4.4 Pore-Scale Flow Pattern in the Porous Structures

This subsection focuses on the pore-scale flow pattern inside the porous structures. Figure 8
shows the velocity vectors in the porous structures of the in-line array close to the interface
area with representative regions of 0.4 < X < 0.65 and 0.375 < Y < 0.575. The inter-
pore vortex structure is also marked with red arrows. In Fig. 9a, for VR = 1, the fluid flows
vertically from the bottom to the free flow channel. However, as the velocity ratio decreases (or
the momentum of the porous medium flow decreases), the clockwise flow between vertical
porous elements vanishes, and anticlockwise flow occurs beside the elements. For VR =
0.0267, the momentum of the flow in the porous structure is significantly lower compared to
the free channel flow and the overall flow field becomes more complex since the rear flow
of the porous structure is quickly occupied at the interface by the channel flow which flows
almost parallel to the streamwise direction, see Fig. 9d.
Figure 10 shows the velocity vectors in the porous structure for the porous structure with
a staggered arrangement. For VR = 1, Fig. 10a shows that the effect of the channel flow on
the flow pattern inside the porous medium is small, and symmetrical vortices occur behind
the porous elements. This is similar to the flow past a single rectangular column (Yang and
Wu 2013). As VR decreases, the effect of channel flow increases, which makes the vortex
inclined towards the direction of the free flow. Interestingly, it can be also observed that
when VR decreases to 0.0267, e.g. Fig. 10d, the flow field around the porous elements is
similar to that for an in-line array, as shown in Fig. 9d. This indicates that the influence of the
type of the porous structure on the flow pattern is significantly attenuated for small values of
low velocity ratios, or with other words, that the momentum of the flow through the porous
medium is much lower compared to the free flow channel which dominates the overall flow
field close to the interface.
In order to quantitatively study the effects of velocity ratio and the type of the porous
structure, averaged velocity angles in the porous structures for the various cases are presented
in Fig. 11. The velocity angle is calculated using the volume averaged velocity Vx and the
volume averaged velocity Vy in the porous structure. Thus, it reflects the relative inclination
to the Y -axis. The larger the value of the velocity angle, the stronger the effect of the channel
flow on the porous medium flow. Theoretically, the value should be zero, if VR is infinitely
large, in which case the porous structure flow has no influence on the channel flow. It can be
seen that as VR increases from 0.0267 to 1, the velocity angle significantly decreases from

123
158 G. Yang et al.

Fig. 9 Velocity vectors in the porous structure and at the interface for a VR = 1, b VR = 0.267, c VR = 0.133
and d VR = 0.0267 for in-line array. The representative regions are chosen as: 0.4 < X < 0.65, 0.375 <
Y < 0.575, Z = 0.5

about 29.2◦ to 2.2◦ for the in-line, and from 28.2◦ to 1.1◦ for the staggered arrangement.
The difference in the velocity angle for the two types of the porous structure is as small as
3.36% for VR = 0.0267, which further proves the effect of the type of the porous structure is
negligible for small VR. Note also that the reduction of the velocity angle is more pronounced
for the staggered porous structure for small velocity ratios, e.g. VR < 0.267, which indicates
that flow field generated with a staggered arrangement is more sensitive to free flow channel
velocity variations.

4.5 Turbulence

Figure 12a, b shows the dimensionless turbulence kinetic energy (E k /Vin2 ) at the middle
plane of the channel and the porous zone (Z = 0.5) for the two types of porous structures at
Rec,L = 1 × 105 and Rep,D = 1 × 103 , that result in a velocity ratio of VR = 0.267. In the
channel downstream the porous medium, the turbulence kinetic energy is large in the mixing
layer and around the shear layer formed outside the flow reversal region. After the fluids are
fully mixed, the turbulence kinetic energy is found to decrease along the flow direction.

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 159

Fig. 10 Velocity vectors in the porous structure and at the interface for a VR = 1, b VR = 0.267, c
VR = 0.133 and d VR = 0.0267 for the staggered array. The representative regions are chosen as: 0.4 <
X < 0.65, 0.375 < Y < 0.575, Z = 0.5

Fig. 11 Average velocity angle


in the porous structure for various
velocity ratios

In the pores of the porous structures, the turbulence kinetic energy is larger for staggered
arrangement than in-line arrangement, which indicates the former structure produces more
velocity fluctuations. As the average velocity angle in the porous structures with Y -axis is

123
160 G. Yang et al.

2
Ek/Vin
0.14
0.13
0.12
0.11
0.10
0.09
0.08
0.07
0.06
0.05
0.04
0.03
L/2 0.02
0.01

(a) (b)
2 ) in the middle plane of the channel
Fig. 12 Distribution of dimensionless turbulence kinetic energy (E k /Vin
5 3
(Z = 0.5) for Rec,L = 1 × 10 and Rep,D = 1 × 10 for a in-line and b staggered arrangements of porous
structures

larger for in-line arrangement, the turbulence kinetic energy increases from lower left to
upper right, as the flow develops.

4.6 Temperature Profiles

The dimensionless temperature profiles in the middle plane of the channel (Z = 0.5) and
in the porous zone for Rec,L = 1 × 105 and Rep,D = 1 × 103 are shown in Fig. 13a, b for
in-line and staggered porous elements, respectively. It can be observed that, under the same
boundary conditions, the temperature in the channel is generally higher for the staggered
array, while the temperature distribution is more homogenized in the streamwise direction.
The temperature distributions along the X -coordinate near the interface of the porous
structure (in-line array) and the channel (Y = 0.5 ± 0.5D/L) for various velocity ratios are
presented in Fig. 14. The Reynolds number in the channel is fixed to Rec,L = 1 × 105 and
the corresponding Rep,D is varied to meet the required velocity ratios. For all velocity ratios,
the temperatures at the two lines fluctuate with X , due to the alternately presence of solid
and fluid phases of the porous structure. As VR increases, the average temperature and the
deviation of the temperature between the two lines decrease. This is because the flow rate in
the porous structure is increased with increasing Rep,D , which also increases the heat transfer
efficiency. Note that at VR = 1, the temperature variations between the two lines are almost
identical.
Figure 15 shows temperature distribution along the X -coordinate near the interface for the
porous structure of staggered arrangement. Contrary to the in-line array, the overall trends in

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 161

θ
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1

L/2

(a) (b)
Fig. 13 Dimensionless temperature distribution in the middle plane of the channel (Z = 0.5) for Rec,L =
1 × 105 and Rep,D = 1 × 103 a in-line and b staggered pore structure arrangement

Fig. 14 Temperature distributions near the interface of the in-line array for various velocity ratios of a
VR = 0.133, b VR = 0.2, c VR = 0.267 and d VR = 1, at Rec,L = 1 × 105

the temperatures fluctuations for the two locations are opposite. As VR increases, the average
temperature is found to decrease, due to the increasing velocity in the porous structure. As
previously shown in Fig. 13, the streamwise temperature distribution is more homogeneous
for the staggered array. This can be clearly observed by comparing Fig. 15 with Fig. 14 under
the same VR.

123
162 G. Yang et al.

Fig. 15 Temperature distributions near the interface for velocity ratios of a VR = 0.133, b VR = 0.2, c
VR = 0.267 and d VR = 1 of the staggered array at RecL = 1 × 105

4.7 Heat Transfer in the Porous Structure

The local Nusselt number distributions at the solid surface of the porous structure for Rep,D =
5 × 102 and VR = 0.267, Rep,D = 5 × 103 and VR = 0.267, and Rep,D = 5 × 102 and
VR = 0.0267, for both in-line and staggered arrays are presented in Fig. 16. The velocity
ratios are the same for Fig. 16a, b, but the velocities at the inlets are larger for the later case.
Therefore, the Nusselt number is generally increased in Fig. 16b. As the flow patterns, which
are determined by the velocity ratio, are the same for Fig. 16a, b, the distributions of the
Nusselt number on the surfaces are also similar for the two cases.
In Fig. 16c, Rep,D is the same as in Fig. 16a, but Rec,L is as high as that in Fig. 16b. It can
be seen that in the case of Fig. 16c, the Nusselt number distribution is not as uniform as that
for VR = 0.267 in Fig. 16a, b. The distributions of the local Nusselt number for in-line and
staggered arrays are also different, as the maximum values are located at the bottom edges
of the porous elements for the in-line array but at the whole bottom surface for the staggered
array.
The area-averaged Nusselt number in the porous structures for various Rec,L at a fixed
Rep,D = 1 × 103 is presented in Fig. 17a for both porous structures. The Nusselt number
for the in-line array is always smaller than that of the staggered array, as the fluid mixes
better with a staggered arrangement. As the flow rate in the channel increases, the Nusselt
number for in-line array increases drastically, but the increase in the Nusselt number for the

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 163

y
z
x

Nu Nu Nu
120 1000 270

0 0 0

(a) (b) (c)


Fig. 16 Local Nusselt number distributions on the surface of the porous structures for a Rep,D = 5 × 102 ,
VR = 0.267, b Rep,D = 5 × 103 , VR = 0.267, c Rep,D = 5 × 102 , VR = 0.0267 (First row: in-line array;
second row: staggered array)

Fig. 17 Variation of area-averaged Nusselt number in the porous structure a with Rec,L for Rep,D = 1 × 103
and b with Rep,D for Rec,L = 1 × 105 .

staggered array is relatively small. For example, when the Rec,L increases from 2.67 ×104
to 5 × 105 , the Nusselt number for in-line array increases about 44% and only 6.2% for
the staggered configuration. Therefore, as Rec,L increases the Nusselt number for the in-line
array approaches that of the staggered array. This is expected since the velocity profile in the
in-line array changed a lot with the variation of velocity, becoming similar to the one of the
staggered arrays if the velocity ratio is as small as 0.0267, see Fig. 9. When Rec,L is kept
constant at 1 × 105 , the average Nusselt number for various Rep,D is presented in Fig. 17b.
The logarithmically plotted Nusselt numbers increase linearly for both the in-line and the
staggered arrays.

123
164 G. Yang et al.

Fig. 18 Comparison between the


simulated Nusselt number and
the correlation

As the averaged Nusselt number in the porous structure is affected by the Reynolds
number in the porous structure, the velocity ratio, and the type of the porous structure, it can
be correlated according to the semi-empirical expression:

Nuavg = nRem
p,D (12)

where,

m = 0.816 − 0.069δVR2
n = 0.426 − 0.021VR − 0.384VRδ + 0.456δVR2

1, for in-line array
δ=
0, for staggered array
In Eq. (12), m is in the order of 0.8 or more, indicating the turbulent character of the flow
domain. Note that for the in-line array, m slightly decreases with the velocity ratio. For a
staggered array, m is always 0.816. The value of m is close to 0.84, as reported by Zukauskas
(1972) using experimental data for the flow past arrays of tubes and is also very similar to the
values reported at the direct wake of surface mounted obstacles (Terzis et al. 2015). Equation
(12) has an average deviation of 0.16% and a maximum deviation of 4.12%, respectively,
for the 41 sets of numerical data in the range of 1 × 104 ≤ Rec,L ≤ 5 × 105 and 5 × 102 ≤
Rep,D ≤ 1 × 104 . A comparison of the results of the proposed correlation and the numerical
data is depicted in Fig. 18, which shows a good agreement.

5 Conclusion

In this study, the turbulent flow and heat transfer characteristics in a three-dimensional channel
coupled with flow through porous structures is investigated numerically, for the ranges of
1 × 104 ≤ Rec,L ≤ 5 × 105 and 5 × 102 ≤ Rep,D ≤ 1 × 104 . The present calculations
are validated by comparison with benchmark numerical and experimental results from open
literature data.
For Rec,L = 1 × 105 and Rep,D = 1 × 104 (VR = 0.267), the velocity distributions for the
two types of porous structures, in-line and staggered arrays, show significant differences in
the porous structure on the pore scale, while the differences are relative small for the averaged
velocity and for that in the channel. A larger region of negative Vx is found in the pores for

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 165

the staggered array, and magnitudes of both negative and positive Vx are higher than for the
in-line array. Vy at the interface is more uniform for the staggered array as for the in-line
array, where locally large values are found in the vicinity of X = 1. For a fixed geometry,
the flow patterns in both the channel and the porous structures are determined by the velocity
ratio between the transversely interacting flow streams.
A flow reversal zone (negative Vx ) is found in the channel downstream of the porous
structure. The length of this flow reversal zone increases with increasing velocity ratios, and
it is always larger for the in-line arranged porous structure. The pore-scale flow field in the
porous structures is also strongly affected by the velocity ratio. However, when the velocity
ratio decreases to values as low as 0.0267, the velocities behave similar for the in-line and
staggered arrays.
Under the same inlet boundary conditions, the temperature in the channel is generally
higher for the staggered array, and the streamwise distribution of the local temperature is also
more uniform for this case. The temperature at the interface fluctuates along the flow axial
coordinate due to the alternately presence of fluid and solid phases of the porous structure,
but the average value decreases with increasing Rep,D for a fixed Rec,L . In contrast to the
flow pattern, the heat transfer in the porous structure is not only affected by the velocity ratio,
but also directly by the absolute velocity magnitude at the inlets.
The average Nusselt number in the porous structure increases drastically with increasing
Rec,L (or decreasing VR) for a constant Rep,D for the porous structure of an in-line array,
while for the staggered array it is insensitive. A correlation for Nu is also proposed considering
the effects of Reynolds number, velocity ratio and the type of porous structures, which has
an average deviation of 0.16% and a maximum deviation of 4.12%, respectively, compared
to the numerical data.
The problem of turbulent flow and heat transfer in a three-dimensional channel coupled
with flow through porous structures is of particular importance. This study, for the first
time, presents fully resolved numerical results for such a model, without any filtering errors.
Even though idealized porous structures are used for the porous medium, our future studies
will further investigate the effect of different microscale structural parameters to extend the
present results to general porous structures. Furthermore, the present pore-scale results can
be used to validate and improve REV and macroscopic models for flow and heat transports
in porous media.

Acknowledgements Guang Yang kindly acknowledges the support of the Sino-German (CSC-DAAD) Post-
doc Scholarship, the National Postdoctoral Program for Innovative Talents (No. BX201600102) and the China
Postdoctoral Science Foundation (No. 2016M601591). In addition, Alexandros Terzis is grateful to Alexander
von Humboldt Foundation for the funding support.

References
Acosta, J.L., Camacho, A.F.: Porous Media: Heat and Mass Transfer, Transport and Mechanics. Nova Science
Publishers Inc, New York (2009)
ANSYS Inc.: ANSYS CFX-Solver Theory Guide-Release 17.1 (2016)
Brillant, G., Husson, S., Bataille, F.: Experimental study of the blowing impact on a hot turbulent boundary
layer. Int. J. Heat Mass Transf. 51, 1996–2005 (2008). https://doi.org/10.1016/j.ijheatmasstransfer.2007.
06.022
Brillant, G., Husson, S., Bataille, F., Ducros, F.: Study of the blowing impact on a hot turbulent boundary layer
using thermal large eddy simulation. Int. J. Heat Fluid Flow 29, 1670–1678 (2008). https://doi.org/10.
1016/j.ijheatfluidflow.2008.06.011

123
166 G. Yang et al.

Caggese, O., Gnaegi, G., Hannema, G., Terzis, A., Ott, P.: Experimental and numerical investigation of a fully
confined impingement round jet. Int. J. Heat Mass Transf. 65, 873–882 (2013). https://doi.org/10.1016/
j.ijheatmasstransfer.2013.06.043
Celik, I.B., Ghia, U., Roache, P.J., Freitas, C.J., Coleman, H., Raad, P.E.: Procedure for estimation and reporting
of uncertainty due to discretization in CFD applications. J. Fluids Eng. 130, 078001 (2008). https://doi.
org/10.1115/1.2960953
Dahmen, W., Müller, S., Rom, M., Schweikert, S., Selzer, M., Von Wolfersdorf, J.: Numerical boundary layer
investigations of transpiration-cooled turbulent channel flow. Int. J. Heat Mass Transf. 86, 90–100 (2015).
https://doi.org/10.1016/j.ijheatmasstransfer.2015.02.075
Davis, M.E.: Ordered porous materials for emerging applications. Nature 417, 813–821 (2002). https://doi.
org/10.1038/nature00785
Fechter, S., Terzis, A., Ott, P., Weigand, B., von Wolfersdorf, J., Cochet, M.: Experimental and numerical
investigation of narrow impingement cooling channels. Int. J. Heat Mass Transf. 67, 1208–1219 (2013).
https://doi.org/10.1016/j.ijheatmasstransfer.2013.09.003
Fetzer, T., Smits, K.M., Helmig, R.: Effect of turbulence and roughness on coupled porous-medium/free-flow
exchange processes. Transp. Porous Media 114(2), 395–424 (2016)
Fetzer, T., Grüninger, C., Flemisch, B., Helmig, R.: On the conditions for coupling free flow and porous-
medium flow in a finite volume framework. In: International Conference on Finite Volumes for Complex
Applications, pp. 347–356. Springer, Cham (2017)
Gao, C., Xu, R.-N.: Pore-scale numerical investigations of fluid flow in porous media using lattice Boltzmann
method. Int. J. Numer. Methods Heat Fluid Flow 25, 1957–1977 (2015). https://doi.org/10.1108/HFF-
07-2014-0202
Ghanbarzadeh, S., Hesse, M.A., Prodanovi, M., Gardner, J.E.: Deformation-assisted fluid percolation in rock
salt. Science 350, 1069–1072 (2015). https://doi.org/10.1126/science.aac8747
Hanjalic, K., Launder, B.E.: A Reynolds stress model of turbulence and its application to thin shear flows. J.
Fluid Mech. 52, 609 (1972). https://doi.org/10.1017/S002211207200268X
Hirota, M., Asano, H., Nakayama, H.: Three-dimensional structure of turbulent flow in mixing T-junction.
JSME Int. J. 49, 1070–1077 (2006). https://doi.org/10.1299/jsmeb.49.1070
Kollet, S.J., Maxwell, R.M.: Integrated surface–groundwater flow modeling: a free-surface overland flow
boundary condition in a parallel groundwater flow model. Adv. Water Resour. 29, 945–958 (2006).
https://doi.org/10.1016/j.advwatres.2005.08.006
Kuwahara, F., Shirota, M., Nakayama, A.: A numerical study of interfacial convective heat transfer coefficient
in two-energy equation model for convection in porous media. Int. J. Heat Mass Transf. 44, 1153–1159
(2001). https://doi.org/10.1016/S0017-9310(00)00166-6
Lācis, U., Bagheri, S.: A framework for computing effective boundary conditions at the interface between free
fluid and a porous medium. J. Fluid Mech. 812, 866–889 (2017). https://doi.org/10.1017/jfm.2016.838
Levenberg, S., Rouwkema, J., Macdonald, M., Garfein, E.S., Kohane, D.S., Darland, D.C., Marini, R., van
Blitterswijk, C.A., Mulligan, R.C., D’Amore, P.A., Langer, R.: Engineering vascularized skeletal muscle
tissue. Nat. Biotechnol. 23, 879–884 (2005). https://doi.org/10.1038/nbt1109
Liu, Y.Q., Jiang, P.X., Xiong, Y.-B., Wang, Y.P.: Experimental and numerical investigation of transpiration
cooling for sintered porous flat plates. Appl. Therm. Eng. 50, 997–1007 (2013). https://doi.org/10.1016/
j.applthermaleng.2012.08.028
Liu, X., Wang, J., Ge, L., et al.: Pore-scale characterization of tight sandstone in Yanchang Formation Ordos
Basin China using micro-CT and SEM imaging from nm-to cm-scale. Fuel 209, 254–264 (2017)
Mayer, B.: Investigations of Pressure Loss and Heat Transfer in a regular metallic Porous Structure. http://
www.dr.hut-verlag.de/9783843918176.html (2014)
Nuske, P., Joekar-Niasar, V., Helmig, R.: Non-equilibrium in multiphase multicomponent flow in porous
media: an evaporation example. Int. J. Heat Mass Transf. 74, 128–142 (2014). https://doi.org/10.1016/
j.ijheatmasstransfer.2014.03.011
Ozgumus, T., Mobedi, M.: Effect of pore to throat size ratio on thermal dispersion in porous media. Int. J.
Therm. Sci. 104, 135–145 (2016). https://doi.org/10.1016/j.ijthermalsci.2016.01.003
Pedras, M.H.J., de Lemos, M.J.S.: Simulation of turbulent flow in porous media using a spatially periodic
array and a low Re two-equation closure. Numer. Heat Transf. Part A Appl. 39, 35–59 (2001). https://
doi.org/10.1080/10407780121486
Prausnitz, J.M., Wilhelm, R.H.: Turbulent concentration fluctuations in a packed bed. Ind. Eng. Chem. 49,
978–984 (1957)
Rosti, M., Cortelezzi, L., Quadrio, M.: Direct numerical simulation of turbulent channel flow over porous
walls. J. Fluid Mech. 784, 396–442 (2015). https://doi.org/10.1017/jfm.2015.566
Saito, M.B., de Lemos, M.J.S.: A correlation for interfacial heat transfer coefficient for turbulent flow over an
array of square rods. J. Heat Transf. 128, 444 (2006). https://doi.org/10.1115/1.2175150

123
Numerical Simulation of Turbulent Flow and Heat Transfer in a… 167

Speziale, C.G., Sarkar, S., Gatski, T.B.: Modelling the pressure–strain correlation of turbulence: an
invariant dynamical systems approach. J. Fluid Mech. 227, 245 (1991). https://doi.org/10.1017/
S0022112091000101
Suga, K., Nakagawa, Y., Kaneda, M.: Spanwise turbulence structure over permeable walls. J. Fluid Mech.
822, 186–201 (2017). https://doi.org/10.1017/jfm.2017.278
Terzis, A., Wolfersdorf, J.Von, Weigand, B., Ott, P.: A method to visualise near wall fluid flow patterns using
locally resolved heat transfer experiments. Exp. Therm. Fluid Sci. 60, 223–230 (2015). https://doi.org/
10.1016/j.expthermflusci.2014.09.009
Torabi, M., Peterson, G.P., Torabi, M., Karimi, N.: A thermodynamic analysis of forced convection through
porous media using pore scale modeling. Int. J. Heat Mass Transf. 99, 303–316 (2016). https://doi.org/
10.1016/j.ijheatmasstransfer.2016.03.127
Uth, M.F., Jin, Y., Kuznetsov, A.V.: A direct numerical simulation study on the possibility of macroscopic
turbulence in porous media: effects of different solid matrix geometries, solid boundaries, and two
porosity scales. Phys. Fluids 28, 065101 (2016). https://doi.org/10.1063/1.4949549
Vortmeyer, D., Schaefer, R.J.: Equivalence of one- and two-phase models for heat transfer processes in
packed beds: one dimensional theory. Chem. Eng. Sci. 29, 485–491 (1974). https://doi.org/10.1016/
0009-2509(74)80059-X
Wakao, N., Kaguei, S., Funazkri, T.: Effect of fluid dispersion coefficients on particle-to-fluid heat transfer
coefficients in packed beds. Correlation of Nusselt numbers. Chem. Eng. Sci. 34, 325–336 (1979). https://
doi.org/10.1016/0009-2509(79)85064-2
Wu, Y., Zhu, G., Gao, B., Sun, Y.: Phase-changed transpiration cooling: material selection, permeability anal-
ysis, and experimental tests in high heat flux. In: 21st AIAA International Space Planes and Hypersonics
Technologies Conference. p. AIAA 2017–2265. American Institute of Aeronautics and Astronautics,
Xiamen, China (2017)
Xu, R.-N., Jiang, P.-X.: Numerical simulation of fluid flow in microporous media. Int. J. Heat Fluid Flow 29,
1447–1455 (2008). https://doi.org/10.1016/j.ijheatfluidflow.2008.05.005
Yang, C., Nakayama, A.: A synthesis of tortuosity and dispersion in effective thermal conductivity of porous
media. Int. J. Heat Mass Transf. 53, 3222–3230 (2010). https://doi.org/10.1016/j.ijheatmasstransfer.
2010.03.004
Yang, G., Wu, J.: Effect of side ratio and aiding/opposing buoyancy on the aerodynamic and heat transfer
characteristics around a rectangular cylinder at low Reynolds numbers. Numer. Heat Transf. Part A
Appl. 64, 1016–1037 (2013). https://doi.org/10.1080/10407782.2013.811057
Yang, G., Wu, J.Y., Yan, L.: Flow reversal and entropy generation due to buoyancy assisted mixed convection
in the entrance region of a three dimensional vertical rectangular duct. Int. J. Heat Mass Transf. 67,
741–751 (2013)
Yazdchi, K., Srivastava, S., Luding, S.: Microstructural effects on the permeability of periodic fibrous porous
media. Int. J. Multiph. Flow 37, 956–966 (2011). https://doi.org/10.1016/j.ijmultiphaseflow.2011.05.003
Zampogna, G., Bottaro, A.: Fluid flow over and through a regular bundle of rigid fibres. J. Fluid Mech. 792,
5–35 (2016). https://doi.org/10.1017/jfm.2016.66
Zukauskas, A.: Heat transfer from tubes in crossflow. Adv. Heat Transf. 8, 93–160 (1972). https://doi.org/10.
1080/07373938908916581

123

You might also like