The Steady Wake of A Wall-Mounted Rectangular Pris
The Steady Wake of A Wall-Mounted Rectangular Pris
The Steady Wake of A Wall-Mounted Rectangular Pris
Article
The Steady Wake of a Wall-Mounted Rectangular Prism with
a Large-Depth-Ratio at Low Reynolds Numbers
Arash Zargar 1,† , Ali Tarokh 1,2,† and Arman Hemmati 1, *,†
Abstract: The wakes of wall-mounted small (square) and large (long) depth-ratio rectangular prisms
are numerically studied at Reynolds numbers of 50–250. The large depth-ratio significantly alters the
dominance of lateral secondary flow (upwash and downwash) in the wake due to the reattachment of
leading-edge separated flow on the surfaces of the prism. This changes the wake topology by varying
the entrained flow in the wake region and changing the distribution of vorticity. Thus, the magnitude
of vorticity significantly decreases by increasing the prism depth-ratio. Furthermore, the length of
the recirculation region and the orientation of near wake flow structures are altered for the larger
depth-ratio prism compared to the square prism. Drag and lift coefficients are also affected due to
the change of pressure distributions on the rear face of the prism and surface friction force. This
behavior is consistently observed for the entire range of Reynolds numbers considered here. The
wake size is scaled with Re1/2 , whereas drag coefficient scaled with Re−0.3 .
Citation: Zargar, A.; Tarokh, A.;
Hemmati, A. The Steady Wake of Keywords: wake dynamics; DNS; wall mounted; prism; CFD; heat exchange; depth ratio
a Wall-Mounted Rectangular Prism
with a Large-Depth-Ratio at Low
Reynolds Numbers. Energies 2021, 14,
3579. https://doi.org/10.3390/ 1. Introduction
en14123579
The flow structures around wall-mounted rectangular cylinders or prisms have been
extensively studied in the literature, partly due to their broad engineering applications
Academic Editor: Francesco
and partly because of their complex dynamics. Particularly at low Reynolds numbers,
Castellani
understanding the wake of a wall-mounted long prism has major implications in improving
the design of electronic chips for better cooling, biomedical devices, vortex generators,
Received: 14 May 2021
Accepted: 9 June 2021
pipe roughness elements, and small heat exchangers [1]. In these applications, the detail
Published: 16 June 2021
understanding of the flow field around the prism is critical in optimizing the design and
performance of various devices, for example fast response accurate measuring equipments
Publisher’s Note: MDPI stays neutral
such as hotwire. Recent developments have revealed that existing experimental and
with regard to jurisdictional claims in
numerical studies do not provide a thorough description of the flow field in the detached
published maps and institutional affil- flow regions of the prism [2], although flow structures around small depth-ratio (tall)
iations. prisms are highly organized [3]. Here, the prism dimensions are normalized by the prism
aspect ratio, defined as AR = h/w, and depth ratio, defined as DR = l/h (see Figure 1).
Wake models have been developed for the flow around wall-mounted circular cylin-
ders, which date back to the work of Taneda (1952) [4]. Since then, there have been several
Copyright: c 2021 by the authors.
modifications and upgrades to these models based on the cylinder characteristics and
Licensee MDPI, Basel, Switzerland.
flow field conditions, a summary of which is provided by Sumner (2013) [5]. Wang et al.
This article is an open access article
(2004) [6] proposed a comprehensive model for the wake of a wall-mounted rectangular
distributed under the terms and cylinder. In this model [6], they classified the wake vortices of a wall-mounted cylinder
conditions of the Creative Commons into four different structures. There are tip and base vortices formed in the wake, which
Attribution (CC BY) license (https:// coincide with the formation of downwash and upwash flow. The spanwise vortices are
creativecommons.org/licenses/by/ Kármán type structures in the middle height of the body, and the horseshoe vortex is
4.0/). formed in front of the body and continue to the wake region. Later on Wang et al. (2009) [3]
modified this model base on their detailed experimental study of the near wake region
immediately behind the cylinder. This study revealed the presence of a single arc-like
structure within the near wake region. They argued that spanwise, tip and base vortices
are inherently connected and form an arch-type structure regardless of the aspect ratio of
the cylinder. They also reported that both asymmetric and symmetric vortex shedding is
observed simultaneously, but the probability of an asymmetrically arranged vortices is
higher at the middle height of the object.
w w l
l
Figure 1. Schematics of the dimensions of short-depth-ratio (left) and large-depth-ratio (right) prisms.
Most recently, da Silva et al. (2020) [7] identified multiple mean wake structures,
instead of a single arc-type structure, formed around a wall-mounted square cylinder
(small depth-ratio prism) with AR = 3 at Re = 500. As shown in Figure 2, it was suggested
that these structures have different origins, contrary to the models discussed by Wang et al.
(2009) [3]. Particularly, da Silva et al. (2020) [7] showed that the structures on the upper
surface of the cylinder appear to fade, while wake tip vortices are formed because of
three-dimensional deflection of the separated flow from the side leading edges of the
cylinder. Moreover, there is a spanwise vortex structure named “Bt ” that is formed by the
folding of the separated shear layer from the free end leading edge of the cylinder. This
newly identified structure has a different origin from the so-called legs of the arc-type
structures. da Silva et al. (2020) [7] identified that the differences observed compared to
the wake model of Wang et al. (2009) [3] may be attributed to the transitional nature of the
latter wake compared to the mean wake considered in the lower Reynolds number study.
These studies, although limited to small depth-ratios, have revealed that, depending on
the cylinder aspect ratio and Reynolds number, there are multiple structures formed in the
wake with distinct characteristics.
Figure 2. The time averaged wake model for a wall-mounted square cylinder from da Silva et al.
(2020) [7].
Energies 2021, 14, 3579 3 of 23
Wei et al. (2001) [8] visualized the effect of increasing Reynolds number, defined
as Re = hU∞ /ν, where h is the prism height, U∞ is freestream velocity, and ν is the
fluid kinematic viscosity, on the horseshoe vortex structure in front of a wall-mounted
cylinder. They showed that increasing Reynolds number can make the horseshoe vortices
oscillatory. Furthermore, these oscillations have a significant impact on the upstream
velocity fluctuations. Zhang et al. (2017) [9] conducted Direct Numerical Simulations
(DNS) at different Reynolds numbers and found that changing the Reynolds number
can change the types of flow structures behind the wall-mounted square cylinder. They
identified a new type of transitional flow structure, namely “Six-Vortices,” in addition to
dipole and quadrupole structures. Using DNS, Rastan et al. (2017) [1] classified the wake
regimes of a short-depth (vertically oriented) cylinder with AR = 7 into five different
regimes: steady flow ( Re < 75), transition to unsteady flow (75 < Re < 85), laminar
flow (85 < Re < 150), transition to turbulent flow (150 < Re < 200) and turbulent flow
( Re > 200). They determined that the wake is characterized by dipole structures at Re < 85,
whereas quadrupole structures dominate the wake at Re > 150. At 85 < Re < 150 the
wake is dominated by other structures that transition the wake into what is called Hexapole
state. The presence of extra vortices in Hexapole wakes is attributed to the bending of
streamlines at the lower part of the cylinder. These studies provided a detailed description
of the wake of small depth ratio cylinders, while the effect of large depth ratios is unknown.
The cross-sectional shape of a wall-mounted geometry is known to influence the flow
separation, and thus, the wake. For example, Uffinger et al. (2013) [10] investigated the
effect of cross-sectional shape of a wall-mounted geometry (i.e., square cylinder, cylinder
with elliptical after body, and cylinder with wedge in front) on its wake using both numeri-
cal and experimental methods at Re = 1.28 × 104 . They determined that cross-sectional
shape of the prism (cylinder) affects the strength of the interaction between the flow over
the top of the body and the Kármán type vortices along its sides. Other similar studies on
circular and rectangular cylinders (e.g., Sattari et al., 2012 [11] and Rastan et al., 2017 [1])
further expanded on this observation at Re = O(104 ). Joubert et al. (2015) [12] investigated
the flow field of a wall-mounted cylinder with the length to width ratio of 2.63 and the
aspect ratio of 5 at Reynolds number of Re = 7.6 × 104 . They reported that the flow
reattachment on the upper surface is one of the main flow changes by increasing the length
of the cylinder. Lim et al. (2015) [13] investigated the effect of depth-ratio (0.5 ≤ DR ≤ 2)
and prism incidence angle (0◦ ≤ i ≤ 45◦ ) at Reynolds number of Re = 4.6 × 104 using
experimental and numerical methods. They reported that changing the depth-ratio and
incidence angle parameters lead to a strong variation of flow field vortex structures and
surface pressure on the prisms. Therefore, analyzing these parameters for designing new
engineering applications is critical. Wang et al. (2019) [14] experimentally investigated
the effect of aspect ratio and depth ratio on the flow field of rectangular cylinders at the
range of Reynolds numbers of 0.78 × 104 –2.33 × 104 . They reported that increasing aspect
ratio changes the flow structures from dipole to quadrupole vortex arrangement. However,
changing the depth ratio does not significantly change the flow structures unless it leads to
flow reattachment on the side surfaces.
The prism aspect ratio is another geometrical parameter that has substantial effect on
the wake dynamics [15,16]. Using DNS, Saha (2013) [15] illustrated that the strength of
tip downwash flow, drag coefficient, Strouhal number, and wake flow unsteadiness of a
short-depth (vertically oriented) cylinder increases with increasing aspect ratio at Re = 250.
Later experimental studies of Sumner et al. (2017) [16] focused on a higher Reynolds
number of 4.2 × 104 , which identified the presence of a critical aspect-ratio, after which the
wake regime is altered. However, these studies focused on high Re and small depth-ratios.
Rastan et al. (2019) [17] explained the effect of changing aspect ratio on the wake flow
structures of a wall-mounted cylinder at Reynolds number of Re = 250. They classified
different streamwise flow structures and mentioned that at AR = 7, the legs of the arc
type structure form with the staggered arrangement, and the vortex shedding is irregular.
However, the arrangement of the arc-type structure legs and vortex shedding is regular for
Energies 2021, 14, 3579 4 of 23
the case of AR = 2. Furthermore, the occurrence of both of these conditions are captured
in the wake of a rectangular cylinder with aspect ratio of AR = 4.
The boundary layer thickness [18,19] and incident angles [20] are also known to effect
the wake dynamics at higher Reynolds numbers. With respect to the former, for exam-
ple, the experiments of Hosseini et al. (2013) [18] and El Hassan et al. (2015) [19] at
Re = 1.2 × 104 identified that the boundary layer has a profound effect on the wake struc-
tures and vortex dynamics, including changing the interaction of horseshoe vortices with
structures in the wake. Although these were limited to higher Reynolds numbers, similar
behavior is intuitively expected for lower Re flows. Behera et al. (2019) [21] investigated the
effect of changing the boundary layer thickness on the flowfield of a square cylinder with
an aspect ratio of 7 at Reynolds number of 250. They reported the simultaneous occurrence
of the two modes of anti-symmetric and symmetric wake shedding processes regardless of
the boundary thickness. Behera et al. (2019) [21] also mentioned that occurrence of the sym-
metric mode of shedding coincides with the strongest values of upwash and downwash
flow. Zhou et al. (2013) [22] investigated the flow features of a two-dimensional rectangular
prism with depth ratio of 5 in non-shear and shear incoming flow at Reynolds number
of Re = 2.2 × 104 . They reported that the peak frequency of the drag and lift coefficients
become identical in shear flow. Furthermore, the unstable separating and reattaching flows
on the side faces of the prism at uniform flow appeared more stable.
The wake features of a wall-mounted prism (cylinder) have been investigated exten-
sively in the literature with respect to the implications of aspect ratio, incidence angle,
boundary layer thickness and Reynolds number. However, these studies almost entirely
focused on short-depth-ratio cylinders at moderate and high Reynolds numbers. Thus,
we aim to address this knowledge gap by extending the wake characterization for larger
depth-ratio (longer) prisms. Therefore, the wake formed at low Reynolds number behind
large depth-ratio wall-mounted prisms will be fully characterized with implications in
design of vortex generators, small-scale heat exchanges, including electronic micro-chips,
and roughness elements in pipes. Particularly related to the fundamentals of wake topology,
studying the wake behind a large depth-ratio prism enables the evaluation of horseshoe
vortex dynamics and its implications on the orthogonal flow behavior in the wake. In this
study, the focus is on examining the wake of a large-depth-ratio (4.17) wall-mounted prism
and compare it with that of a small depth-ratio prism (square cylinder) at low Reynolds
numbers (50 ≤ Re ≤ 250). The main objective of this study is to establish the changes in
the wake due to the larger depth-ratio of the prism. The paper structure follows a detailed
description of the methodology and numerical setup in Section 2. The main results and
discussions are included in Section 3 followed by the summary of conclusions in Section 4.
2. Methodology
This study considers the flow around short and long depth-ratio wall-mounted prisms
at Re = 50 − 250. This range of Reynolds number is selected to best identify the changes
associated with the large depth-ratio effects since the wake of the square prism at Re = 250
should be steady [15]. The prism dominant height (h) is used to normalize all dimensions
and length-scales. For the long rectangular prism, the width is w = 0.83 h and the
length (depth) is l = 4.17 h, which implies a depth-ratio of DR = l/h = 4.17. These
dimensions were motivated by design of electronic chips and a particular flow manipulator
for biomedical applications [23]. Moreover, this depth-ratio is larger than the critical aspect-
ratio of 3–4 previously identified for infinitely span prisms [24]. The square prism has a
width and length of w = l = 0.83 h, which translate to a depth-ratio of DR = 0.83. These
dimensions are selected for the reference case, which satisfies the condition of DR < 1,
and differs from the special case of a cube (DR = 1). The uniform inlet velocity corresponds
to different Reynolds numbers based on the prism height (h). The boundary layer formed
on the ground has a thickness of ≈0.5 h at the prism frontal face. The schematics of the long
prism is shown in Figure 3. The prism cross-sectional area is in the yz-plane and it extends
in the streamwise (x-) direction.
Energies 2021, 14, 3579 5 of 23
∂ui
= 0, (1)
∂xi
∂ui ∂ui u j 1 ∂p ∂2 u j
+ =− +ν , (2)
∂t ∂x j ρ ∂xi ∂xi x j
where ui and p are the velocity and pressure, respectively [25].
In the present study, the below definitions are considered for calculating the coeffi-
cients of drag, lift and pressure:
D
Cd = 2 A
, (3)
0.5ρU∞
L
Cl = 2 A
, (4)
0.5ρU∞
p − p∞
Cp = 2
, (5)
0.5ρU∞
where D and L are the drag and lift forces, respectively, p is the fluid pressure, ρ is the fluid
density, U∞ is the free stream velocity, and A is the prism front face area.
The streamwise (x-), spanwise (y-), and normal (z-) dimensions of the computational
domain, shown in Figure 3, are L = 40 h, H = 6 h, and W = 16 h, respectively. The front
face of the body is located Lup = 10 h from the inlet and Ldo = 25 h from the outlet.
The constant uniform velocity is applied as the inlet boundary condition. Sides and upper
walls of the domain are set to slip boundary condition. The outlet boundary is set as the
Neumann boundary condition (∂φ/∂n = 0, where φ is any flow variable). The no-slip
boundary condition is applied to the ground and body faces.
A non-homogeneous grid made of 4.5 × 106 and 3.9 × 106 hexahedral elements are
used for the long and short rectangular prisms at Re = 250, respectively. To ensure
that the effect of the Reynolds number will be identified thoroughly, the same mesh is
utilized for lower Reynolds numbers. The spatial grid distribution for the long prism is
shown in Figure 4. The grids are designed so that the maximum value of n+ is below
0.55 at the walls at Re = 250, which enables capturing flow fluctuations associated with
separated flow. The long prism domain edges contain 241 elements in the x-direction,
173 elements in the y-direction and 109 elements in the z-direction. Similarly, the short
Energies 2021, 14, 3579 6 of 23
prism contains 206 × 173 × 109 elements along the x × y × z directions, respectively. The
timestep was set as 2.08 × 10−3 h/U0 for the long and short prisms, which ensured that
the maximum CFL number remained below 0.8 in the wake region with the most spatially
refined grid. The spatial and temporal discretization of the governing equations are second-
order accurate. Temporal discretization is based on second order accurate backward Euler
method, and the discretization scheme of the gradient, divergence, and Laplacian terms
are based on the second order Gauss Linear method. The residual momentum root-mean-
square of 10−6 was set as the criteria for convergence for each timestep. The PIMPLE
algorithm, which is a combination of Pressure Implicit with Splitting of Operator (PISO)
and Semi-Implicit Method for Pressure-Linked Equations (SIMPLE) methods, is used for
coupling the pressure and velocity fields. The computational domain for each case is
divided into 16 separate regions for parallel computing. The simulations are completed
using Intel Platinum 8160 F Skylake 2.1 GHz cores at 15,000 core hours.
Figure 4. The spatial grid distribution for the long prism at Re = 250. Top view at z/h = 0.5 (top)
and side view at y/h = 0 (bottom).
present study calculated at x/h = −2.1 and y/h = −7 with the analytically determined
Blasius boundary layer solution. The good agreement between these results indicates that
the simulations are adequate in capturing the main flow features.
Table 1. Numerical setup comparison between current and previous studies. n is the number of elements. The bold case is
used for the simulations.
Figure 5. Grid sensitivity study based on streamwise velocity profile at x/h = 3 and y/h = 0 for the
long prism.
Figure 7. Profiles of time-averaged streamwise velocity profile inside the boundary layer for the long
prism at x/h = −2.1 and y/h = −7, compared to the Blasius Boundary layer solution.
3. Results
This study focuses mainly on characterizing the wake of a large depth-ratio prism at
the highest Re considered here, while comparing the wake topology to well-established
wake of small depth-ratio (square) prisms. The longer surfaces of the long prism is
expected to motivate reattachment of the shear layer formed on the leading edges of the
prism. This hints at potentially major alternations to the wake. To begin, we look at the
integral flow parameters, such as the coefficients of drag and lift, defined as forces acting
in the positive x- and z-directions respectively, at Re = 250. The results showed that
drag and lift coefficients are altered by the larger depth-ratio of the rectangular prims
compared to the square prism. As shown in Table 2, the drag coefficient increases from
Cd = 0.84 for the square prism to 1.05 for the long prism. To find the source of this change,
the distribution of pressure on the front and rear faces of the prisms are depicted in Figure 8.
As expected, the pressure distribution on the front face of both prisms are fairly similar
since the freestream characteristics mostly dictate the pressure of the front stagnation region.
Conversely, the pressure distribution on the rear face of the two prisms strongly differ,
which hints that the wake flow characteristics vary between the two bodies. Magnitude of
the pressure coefficient on the rear face of the long prism is larger than that of the square
prism. Furthermore, Figure 9 presents the pressure distributions across the rear and frontal
face of the prisms on the symmetry plane (y/h = 0). As shown, the pressure coefficient
on the prisms’ front face are identical for both cases, while pressure is larger on the back
face of the large depth-ratio prism. This implies that the increase of drag observed for the
longer prism does not relate to a larger pressure drop in the wake. Thus, the additional
drag must intuitively relate to larger frictional forces on the side and upper faces of the
long prism.
Table 2. Time-average coefficients of lift (Cl ) and drag (Cd ) and recirculation length (Lr ) for both
prisms at Re = 250.
To characterize the wake of a large depth-ratio (long) rectangular prism, the main wake
features are thoroughly examined and compared with those of the square prism. The flow
around a wall-mounted rectangular prism with an aspect ratio of 2.00 and DR = 0.50
at Re = 250 has been characterized as weakly unsteady by Saha (2013) [15]. However,
the current results indicate that the flow around square (DR = 1.00 and AR = 1.20) and
long (DR = 4.17 and AR = 1.20) rectangular prisms are steady at Re = 250. The velocity
Energies 2021, 14, 3579 9 of 23
and pressure are examined at 45 different points within the near and far wake of both prisms
through time, which did not show even weakly unsteady characteristics. Furthermore,
examining the instantaneous isosurface of the Q-criterion at different timesteps did not
exhibit any temporal changes of wake flow structures. The flow steadiness can be attributed
to the suppression of the Kármán type vortices by downwash and upwash flow created
close to the top and bottom of the prisms, respectively. This agrees with observations
reported in the literature (e.g., Sohankar, 2006 [33] and Saha, 2013 [15]) that decreasing
the aspect ratio of a rectangular wall-mounted prism, for the case of the square prism,
and increasing the depth ratio to such large values (>4), for the case of a long prism,
have similar implications on eliminating the wake unsteadiness at low Reynolds numbers.
A higher aspect-ratio square prism case was tested for validation, not shown here for brevity,
which showed unsteadiness at Re = 250, similar to the observations of Saha (2013) [15].
Square Prism
Focusing on the flow field and main characteristics of the wake, the critical points on
different faces of the long rectangular and square prisms are shown in Figures 10 and 11.
The critical points are determined by streamlines of the velocity vector at the first cells from
the walls. The nodal, saddle and focal points are identified as N, S, and F, respectively.
Energies 2021, 14, 3579 10 of 23
(a) Ground
Critical points are identified in the wake using streamlines on 2D planes both on and
around the prisms. The total number of nodal, focal, and saddle points for the long prism
are 14, 4, and 18, respectively. This agrees with the predictions of Tobak et al. (1982) [34],
Chapman et al. (1991) [35] and Delery et al. (2013) [36]. According to them, there is a
critical point theory for a wall mounted object, which also holds:
Σ( N + F ) − Σ(S) = 0. (6)
There is, however, a different wake topology expected for the square prism (short
prism), in which case the result of Equation (6) is two instead of zero, according to
Liakos et al. (2014) [37]. Based on the visualization of the critical points around the square
prism in Figure 11, the total number of nodal, focal, and saddle points around the square
prism are 13, 6 and 17, respectively. Thus, subtracting S from N + F is equal to 2, which
agrees with the findings of Liakos et al. (2014) [37]. It appears that two pairs of saddle
points that are present on the side of the long prism (see Figure 10a) are merged in the
case of the square prism. This leads to a decrease on the number of the saddle points
compared to nodal points for the square prism, which indicates that Equation (6) can be
different for a wall-mounted object based on the conditions of the study. Furthermore,
there are some differences between the distribution of critical points on the wall-mounted
cube in Liakos et al. (2014) [37] and the square wall-mounted prism in the current study.
For example, they reported the presence of some saddle and nodal points on the edge
between the front face and the ground close to the side face of the prism. However, these
critical points are not observed for the square prism in the current study. These differences
between the two cases are expected due to the differences in the geometries and thicknesses
of the boundary layer. Moreover, comparing with Liakos et al. (2014) [37], the distribution
Energies 2021, 14, 3579 11 of 23
of critical points on the long prism at Re = 250 is similar to the distribution of critical
points on a cube at Re ≥ 1000. Looking at the results in Figures 10b and 11b, the nodal
and saddle points on the rear face of the square prism (at y/h = 0) are formed farther
from the ground compared to the long prism. This difference can be attributed to the
more significant upwash flow in the wake of the square prism, which will be discussed
later. It is also notable that there is not any saddle point on the upper rear face of the
square prism. Alternatively, as shown in Figure 11e, there is a saddle point on the ending
part of the upper face of the square prism, which indicates that there is a small region of
reversed flow towards the trailing edge of the square prism upper face. Another significant
difference of the distribution of critical points on faces of the short and long prism is that
there exists an additional attachment nodal point and a saddle point behind the square
prism in Figure 11a. The attachment nodal point is formed because of the interaction of the
separated shear layer with the rear face of the prism. The saddle point is also the result of
the interaction of the former nodal point and the last nodal point of the wake region.
(a) Ground
The iso-surface of the Q-criterion, normalized using the prism height (h) and freestream
velocity (U∞ ) and denoted by Q∗ , are shown in Figures 12 and 13 from two perspectives:
the front and rear views of the prism. The plots in Figure 13 identify that the separated
Energies 2021, 14, 3579 12 of 23
shear layer from the leading edge of the square prism is connected to the wake structures,
which agrees with the existing models [3]. For the long prism, however, the structures are
separated between the leading and trailing edge of the prism in Figure 12. Furthermore,
the legs of the horseshoe vortex are extended into the wake and interact with the structure
formed behind the square prism. This can directly impact the wake dynamics. Conversely,
the wake structure of the long prism is located at a relatively long distance from the legs of
the horseshoe vortex. The shorter depth of the square prism allows for interaction of the
wake structures on the trailing edge of the prism, while the horseshoe vortex legs formed
around the long prism are highly distorted and lack coherence at the trailing edge of the
prism. Thus, there is no interactions between horseshoe structures and the wake behind
the body. This differs from the square prism wake model.
The effect of large depth-ratio on upwash flow is apparent in Figure 14, where there is
evidence of a strong upwash flow for the square prism and not the long prism. The upwash
effect in the wake of the long prism is very narrow and limited to the vicinity of the prism
rear face (see Figure 14a,b). This change in the wake behavior is associated with the change
in interaction between the horseshoe vortex and the tip structure at the trailing of the prism.
Moreover, this provide further evidence in comparison to the larger aspect ratio prisms,
i.e., AR = 2 in Saha (2013) [15], that the height of the boundary layer at the prism rear face
plays an essential role in the wake flow characteristics. Another critical difference between
the two prisms is related to the downwash flow in the middle of the long prism, where
there is a stronger downwash effect in the outer regions. This observation is also supported
by the streamline plots on 2D planes of Figure 15. This implies that the entrainment of
the freestream flow due to the separated shear layer increases the strength of the upwash
and downwash flow. However, its effect on the upwash flow is more apparent near the
symmetry plane, whereas its effect on the downwash is more evident close to the junction
of the side and rear faces.
Energies 2021, 14, 3579 13 of 23
(a) y/h = 0
As shown in Figures 15 and 16, the cross-flow in the wake is symmetric for both prisms.
However, the presence of the boundary layer effects along the z-direction leads to a strongly
antisymmetric wake in the normal (z-) direction. In the normal mid-plane of the prism
(middle plane in the z-direction), there are focal points that correspond to laminar vortical
structures in the wake (see Figure 16b). It can be seen from the contours superimposed
on the isosurface of Q-criterion (Figures 15 and 16) that the rotational motion of the wake
structure at a lower height (close to the ground) and higher part of the long prism are
restricted to the z- and y-directions, respectively. The interconnection of these structures,
which is apparent from the combination of 2D (streamline and contours) and 3D (iso-
surface) visualizations, implies that their behavior is within a single structure that deforms
and diffuses in the wake. The separation bubble close to the leading edge of the body is
highlighted in Figure 15a, on the mid-plane of the prism, which does not extend into the
outer wake. This implies that the reattachment on the prism upper face is stationary and
it does not contribute to the tip structure typically formed on the prism rear, as per the
model of Wang et al. (2009) [3]. There also exists the same separation bubble on sides of the
long prism (Figure 16b) as it was also observed on the top face of the prism in Figure 15a.
The side-bubbles also appear on the mid-plane for the prism, similar to the top-face bubble.
Comparing such behaviors to the wake of the square prism in Figures 17 and 18, there is a
similarity in the formation of bubbles on top and side faces of the prism. However, there
appears to be a larger laminar structure in the outer wake of the square prism, which leads
to a large flow entrainment (downwash flow) compared to the long prism. The center of
Energies 2021, 14, 3579 14 of 23
the wake structure behind the square prism in Figure 17b is 0.5 h farther away from the
rear face of the prism compared to the long prism in Figure 15a.
(a) y/h = 0
The streamline plots of Figure 18b identify a larger streamline curvature angle of 18.4◦
in the shear layer of the square prism compared to the long prism, where the curvature
angle is 10.9◦ in Figure 16. The curvature angles are measured close to the mid-height
plane at y/h = 0.3 using α =tan−1 (v/u), where α is the streamline angle of curvature.
The smaller α for the long prism implies that there is not a significant flow entrainment
into the immediate wake [38]. This wake behavior appears very different for the square
Energies 2021, 14, 3579 15 of 23
prism in Figure 18a. Here, there is a significantly larger streamline curvature, due to
the shorter depth of the prism, which implies a stronger freestream flow entrainment in
the wake region. Streamline of the long and square prisms shows that the length and
height of the wake region of the square prism are greater than the long prism. Particularly,
the vortices formed in the wake of the square prism extend up to 2 h downstream in
the wake, whereas the extension of these structure for the long prism is reduced by a
factor of 2 (see Figures 15a, 16b, 17a and 18b). This trend persists across the flow field in
the spanwise (y-) and normal (z-) directions. The vortex structure in z-direction in the
middle region of the square prism extends to the upper regions, whereas it was constrained
to the mid-plane of the longer prism. This structure for the long prism is completely
different in the middle and upper regions. This indicates that the extension of the flow
structures in the spanwise direction is inherently related to the strength of the upwash
flow. The separation bubble close to the leading edge of the body is highlighted on the
middle planes (see Figures 17a and 18b). This separation bubble does not extend to the
outer regions. Furthermore, the length of this structure is not affected by changing the
depth ratio. Despite this similarity, there exists a tip structure in the wake, which follows
a different dynamics for long prisms compared to shorter prisms at low Re. It appears
that the wake of the square prism is dominated by the downwash flow induced by larger
(in apparent size) and stronger (in vorticity magnitude) trailing structures, which interact
with the horseshoe vortex in the immediate wake. However, the long prism involves a
distinctly smaller (in apparent size) and weaker (in vorticity magnitude) structure formed
at the trailing-edge of the prism that do not interact with the boundary layer. Thus, the flow
appears steady in the wake.
Figure 19 shows the vorticity contours in the near wake region of the prisms. The larger
magnitudes of spanwise and streamwise components of vorticity in the wake region of
the square prism can lead to larger streamline curvatures. This provides further evidence
of the stronger entrainment in the wake of the square prism compared to the long prism.
The separated shear layer from the leading edge exhibits a large vorticity magnitude for
both prisms. In the case of the square prism, the leading edge separated flow impacts the
wake structures through what can be described as flow entrainment. However for the long
prism, the reattachment of the flow on the upper and side faces prevents any entrainment
mechanism. The presence of counter-rotating vortices in two sides of the prism implies that
Kármán type vortices are suppressed by the upwash and downwash flow in both cases,
due to the small aspect ratio of the prisms.
Energies 2021, 14, 3579 16 of 23
For a better understanding of the vortex structures around the body, two-dimensional
streamlines along different planes perpendicular to the streamwise direction (cross-sectional
planes) are shown in Figure 20. Two counter-rotating vortical structures are observed be-
hind the long prism, which expand by moving farther away from the body. Since the
rotational direction of these structures creates a downwash flow in the mid plane region
(y/h = 0), to follow the existing definitions for tip vortex structure [1], these vortices
should be tip vortices. Thus, it seems that the small aspect-ratio of the considered prisms
leads to the elimination of the streamwise base vortices in the wake region, although the
upwash flow dominates the near wake region of the prisms (see Figure 14b,d). These
rotational flow transforms into a pair of streamwise vortical structures within 3.85 h from
the prism, which quickly grow to the same size as the prism height within 8 h downstream
(Figure 20d). These streamwise vortices appear steady and with minimal diffusion far into
the wake, x/h ≥ 20 in Figure 20f.
Figure 23. The profiles of spanwise velocity (uz ) in the symmetry plane of both prisms at different
distances (identified as ‘d’) from the rear face of the prism at Re = 250.
(a) (b)
(c) (d)
Figure 24. Streamlines on a side view (xz-) of the long prism (DR = 4.17) at y/h = 0 highlighted by
the normal component of the velocity vector at Re = (a) 50; (b) 150; (c) 200 and (d) 250.
Moreover, the location of the recirculation structure behind the prism moved away
from the ground with increasing Re. Particularly, the focal point was located ∼0.3 h above
the ground at Re = 50, which moved to 0.65 h at Re = 150 and then settled at ∼0.72 h
for Re ≥ 200. The streamwise location of the structure with respect to the rear face of the
prism increased with Reynolds number in the range of Re = 50–150, but it was almost
unchanged for Re = 150–250. In particular, the focal point was located at x ∼0.75 h for
Re = 50, which then moved to x ∼ 0.8 h for Re = 150–250. It is notable that the shape of
the wake structures remained unchanged in the range of low Reynolds number considered
here (see Figure 26), although the strength of vortical structures changed due to variations
in the strength of upwash and downwash flow. Therefore, establishing a constant flow
skeleton model for this range is reasonable.
Energies 2021, 14, 3579 20 of 23
2.5 1
2
2 0.75
𝑅𝑒 ".$
𝐶!
1.5
∝ 𝑅
Ls/h , hs/h
𝐿! ∝
𝑒
"#
Lr/h
1.5 0.5
.%
Cd
1
1 0.25
0.5 0 0.5
10 100 1000 10 100 1000 10 100 1000
Reynolds number (Re) Reynolds number (Re) Reynolds number (Re)
(a) (b)
(c) (d)
Figure 26. Streamlines on a top view (xy-) of the long prism (DR = 4.17) at z/h = 0.5. Re = (a) 50;
(b) 150; (c) 200 and (d) 250.
The drag coefficient of the prism is another parameter, which changes with respect to
the Reynolds number. Similar to the drag coefficient of a sphere at low Reynolds numbers,
the drag coefficient of the prism decreases from 1.69 at Re = 50 to 1.05 at Re = 250, which
is depicted in Figure 25. The variations in drag appear to scale well with Re−0.3 .
there exists a shear layer formed behind the body that separates from the topper edge of
the prism and folds over its rear face. It generates a small region of upwash in the wake.
However, the presence of two tip vortices induces a downwash flow in the wake. The next
steps in this research involve determining the effects of Reynolds number in the wake
unsteadiness and characterizing the unsteady wake at a higher Reynolds number.
Flow
Separation
Region
Separation
Region Reatt
ac
Regio hment
n
z
y
Horseshoe Vortex
Tip Vortex
Figure 27. The skeleton model of the steady wake formed behind a long rectangular prism.
4. Conclusions
This study focused on fully characterizing the wake of a long prism at a low Reynolds
number in an effort to fill the knowledge gap on the implications of depth ratio on wake
topology. It was hypothesized that potential wake alternations originate from the reattach-
ment of separated shear layer from the leading edges of the prism due to their elongated
shape. The flow around two wall-mounted rectangular prisms with the depth-ratios of 0.83
and 4.17 are numerically studied at low Reynolds numbers (50 ≤ Re ≤ 250). The wakes
are analyzed extensively for both shapes to identify changes in wake topology for the
large depth-ratio (long) rectangular prism compared to a square prism. The wake descrip-
tions identify significant variations in the wake topology between the two prisms, which
are attributed to the larger depth-ratio of the long prism. The wake appears steady for
both prisms due to the small aspect-ratio, which is consistent with existing observations.
A skeleton model of the steady wake is proposed, which depicts the tip vortices and
horseshoe vortex.
It is shown that the distributions of the critical points around the two prisms with
different depth-ratios and the distribution of surface pressure coefficient on the rear faces
are different. The strength of the upwash flow in the wake region of the square prism is
higher than the long prism. This leads to different sizes of the recirculation region due to
the large differences in their depth-ratios. Furthermore, the distribution of downwash flow
from the tip region is different for the two cases.
As a result of flow entrainment from the leading edge separated shear layer, the stream-
wise structures of the near wake region are altered for the short prism. Moreover, there
exist some streamwise structures close to the leading edge of the long prism that do not
appear in the wake of the square prism. This difference is mainly attributed to the large
streamline curvatures in front of the prism with a large-depth-ratio. However, a similar
structure close to the trailing edge of the square prism is observed, where there is evidence
of a large streamline curvature, larger vorticity magnitudes, and thus, more significant
entrainment effects. This is mainly attributed to the smaller depth-ratio of the square prism,
which facilitates the merge of structures in the rear and their interaction with the horseshoe
structure. Horseshoe structures do not appear to interact with the wake for the long prism
at low Reynolds numbers, which the flow behavior is steady.
There is an apparent effect of Reynolds number on the wake size for Re = 50–250 with
a consistent increase in the wake size up to Re = 150 and the movement of the focal point
away from the ground. The distance of the wake structure from the prism rear face scaled
with Re1/2 , while the drag coefficient scaled with Re−0.3 .
Energies 2021, 14, 3579 22 of 23
References
1. Rastan, M.; Sohankar, A.; Alam, M.M. Low-Reynolds-number flow around a wall-mounted square cylinder: Flow structures and
onset of vortex shedding. Phys. Fluids 2017, 29, 103601. [CrossRef]
2. Cimarelli, A.; Leonforte, A.; Angeli, D. Direct numerical simulation of the flow around a rectangular cylinder at a moderately
high Reynolds number. J. Wind Eng. Ind. Aerodyn. 2018, 174, 39–49. [CrossRef]
3. Wang, H.; Zhou, Y. The finite-length square cylinder near wake. J. Fluid Mech. 2009, 638, 453–490. [CrossRef]
4. Taneda, S. An experimental study on the structure of the vortex street behind a circular cylinder of finite length. In Research
Report of Institute of Applied Mechanics; Kyushu University: Fukuoka, Japan, 1952; Volume 1, pp. 131–142.
5. Sumner, D. Flow above the free end of a surface-mounted finite-height circular cylinder: A review. J. Fluids Struct. 2013, 43, 41–63.
[CrossRef]
6. Wang, H.; Zhou, Y.; Chan, C.; Wong, W.; Lam, K. Flow structure around a finite-length square prism. In Proceedings of the 15 th
Australasian Fluid Mechanics Conference, Sydney, Australia, 13–17 December 2004; pp. 13–17.
7. da Silva, B.L.; Chakravarty, R.; Sumner, D.; Bergstrom, D.J. Aerodynamic forces and three-dimensional flow structures in the
mean wake of a surface-mounted finite-height square prism. Int. J. Heat Fluid Flow 2020, 83, 108569. [CrossRef]
8. Wei, Q.D.; Chen, G.; Du, X.D. An experimental study on the structure of juncture flows. J. Vis. 2001, 3, 341–348. [CrossRef]
9. Zhang, D.; Cheng, L.; An, H.; Zhao, M. Direct numerical simulation of flow around a surface-mounted finite square cylinder at
low Reynolds numbers. Phys. Fluids 2017, 29, 045101. [CrossRef]
10. Uffinger, T.; Ali, I.; Becker, S. Experimental and numerical investigations of the flow around three different wall-mounted cylinder
geometries of finite length. J. Wind Eng. Ind. Aerodyn. 2013, 119, 13–27. [CrossRef]
11. Sattari, P.; Bourgeois, J.; Martinuzzi, R. On the vortex dynamics in the wake of a finite surface-mounted square cylinder. Exp. Fluids
2012, 52, 1149–1167. [CrossRef]
12. Joubert, E.; Harms, T.; Venter, G. Computational simulation of the turbulent flow around a surface mounted rectangular prism.
J. Wind Eng. Ind. Aerodyn. 2015, 142, 173–187. [CrossRef]
13. Lim, H.C.; Ohba, M. Detached eddy simulation of flow around rectangular bodies with different aspect ratios. Wind Struct. 2015,
20, 37–58. [CrossRef]
14. Wang, F.; Lam, K.M. Geometry effects on mean wake topology and large-scale coherent structures of wall-mounted prisms.
Phys. Fluids 2019, 31, 125109. [CrossRef]
15. Saha, A.K. Unsteady flow past a finite square cylinder mounted on a wall at low Reynolds number. Comput. Fluids 2013,
88, 599–615. [CrossRef]
16. Sumner, D.; Rostamy, N.; Bergstrom, D.; Bugg, J. Influence of aspect ratio on the mean flow field of a surface-mounted finite-height
square prism. Int. J. Heat Fluid Flow 2017, 65, 1–20. [CrossRef]
17. Rastan, M.; Sohankar, A.; Doolan, C.; Moreau, D.; Shirani, E.; Alam, M.M. Controlled flow over a finite square cylinder using
suction and blowing. Int. J. Mech. Sci. 2019, 156, 410–434. [CrossRef]
18. Hosseini, Z.; Bourgeois, J.; Martinuzzi, R. Large-scale structures in dipole and quadrupole wakes of a wall-mounted finite
rectangular cylinder. Exp. Fluids 2013, 54, 1595. [CrossRef]
19. El Hassan, M.; Bourgeois, J.; Martinuzzi, R. Boundary layer effect on the vortex shedding of wall-mounted rectangular cylinder.
Exp. Fluids 2015, 56, 33. [CrossRef]
20. McClean, J.; Sumner, D. An experimental investigation of aspect ratio and incidence angle effects for the flow around surface-
mounted finite-height square prisms. J. Fluids Eng. 2014, 136, 081206. [CrossRef]
21. Behera, S.; Saha, A.K. Characteristics of the Flow Past a Wall-Mounted Finite-Length Square Cylinder at Low Reynolds Number
With Varying Boundary Layer Thickness. J. Fluids Eng. 2019, 141. [CrossRef]
Energies 2021, 14, 3579 23 of 23
22. Zhou, Q.; Cao, S.; Zhou, Z. Numerical studies on non-shear and shear flows past a 5:1 rectangular cylinder. Wind Struct. 2013,
17, 379–397. [CrossRef]
23. Jia, Y. The Implications of Coarctation and Indentation of the Aorta on Blood Flow Characteristics in Pediatric Patients.
Master’s Thesis, University of Alberta, Edmonton, Alberta, 2020.
24. Okajima, A.; Ueno, H.; Sakai, H. Numerical simulation of laminar and turbulent flows around rectangular cylinders. Int. J.
Numer. Methods Fluids 1992, 15, 999–1012. [CrossRef]
25. Durbin, P.A.; Reif, B.P. Statistical Theory and Modeling for Turbulent Flows; John Wiley & Sons: Hoboken, NJ, USA, 2011.
26. Hemmati, A.; Wood, D.H.; Martinuzzi, R.J. On simulating the flow past a normal thin flat plate. J. Wind Eng. Ind. Aerodyn. 2018,
174, 170–187. [CrossRef]
27. Shah, K.B.; Ferziger, J.H. A fluid mechanicians view of wind engineering: Large eddy simulation of flow past a cubic obstacle.
J. Wind Eng. Ind. Aerodyn. 1997, 67, 211–224. [CrossRef]
28. Krajnovic, S.; Davidson, L. Large-eddy simulation of the flow around a surface-mounted cube using a dynamic one-equation
subgrid model. In First Symposium on Turbulence and Shear Flow Phenomena; Begel House Inc.: New York, NY, USA, 1999.
29. Yakhot, A.; Liu, H.; Nikitin, N. Turbulent flow around a wall-mounted cube: A direct numerical simulation. Int. J. Heat Fluid
Flow 2006, 27, 994–1009. [CrossRef]
30. Krajnovic, S.; Davidson, L. Flow around a simplified car, part 1: Large eddy simulation. J. Fluids Eng. 2005, 127, 907–918.
[CrossRef]
31. Krajnovic, S.; Davidson, L. Flow around a simplified car, part 2: Understanding the flow. J. Fluids Eng. 2005, 127, 919–928.
[CrossRef]
32. Zargar, A.; Gongur, A.; Tarokh, A.; Hemmati, A. Coherent structures in the wake of a long wall-mounted rectangular cylinder at
large incident angles. J. Phys. Rev. Fluids 2020, 6, 034603. [CrossRef]
33. Sohankar, A. Flow over a bluff body from moderate to high Reynolds numbers using large eddy simulation. Comput. Fluids 2006,
35, 1154–1168. [CrossRef]
34. Tobak, M.; Peake, D.J. Topology of three-dimensional separated flows. Annu. Rev. Fluid Mech. 1982, 14, 61–85. [CrossRef]
35. Chapman, G.T.; Yates, L.A. Topology of flow separation on three-dimensional bodies. Appl. Mech. Rev. 1991, 44, 329–345.
[CrossRef]
36. Délery, J. Three-Dimensional Separated Flow Topology: Critical Points, Separation Lines and Vortical Structures; John Wiley & Sons:
Hoboken, NJ, USA, 2013.
37. Liakos, A.; Malamataris, N.A. Direct numerical simulation of steady state, three dimensional, laminar flow around a wall
mounted cube. Phys. Fluids 2014, 26, 053603. [CrossRef]
38. Hemmati, A.; Wood, D.H.; Martinuzzi, R.J. Characteristics of distinct flow regimes in the wake of an infinite span normal thin flat
plate. Int. J. Heat Fluid Flow 2016, 62, 423–436. [CrossRef]