Numerical Analyses of The Flow Past A Short Rotati
Numerical Analyses of The Flow Past A Short Rotati
840
This work studies the three-dimensional flow dynamics around a rotating circular cylinder
of finite length, whose axis is positioned perpendicular to the streamwise direction. Direct
numerical simulations and global stability analyses are performed within a parameter
range of Reynolds number Re = DU∞ /ν < 500 (based on cylinder diameter D, uniform
incoming flow velocity U∞ ), length-to-diameter ratio AR = L/D ≤ 2 and dimensionless
rotation rate α = DΩ/2U∞ ≤ 2 (where Ω is rotation rate). By solving Navier–Stokes
equations, we investigated the wake patterns and explored the phase diagrams of the lift
and drag coefficients. For a cylinder with AR = 1, we found that when the rotation effect is
weak (0 ≤ α 0.3), the wake pattern is similar to the unsteady wake past the non-rotating
finite-length cylinder, but with a new linear unstable mode competing to dominate the
saturation state of the wake. The flow becomes stable for 0.3 α 0.9 when Re < 360.
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
When the rotation effect is strong (α 0.9), new low-frequency wake patterns with
stronger oscillations emerge. Generally, the rotation effect first slightly decreases and then
sharply increases the Re threshold of the flow instability when α is relatively small, but
significantly decreases the threshold at high α (0.9 < α ≤ 2). Furthermore, the stability
analyses based on the time-averaged flows and on the steady solutions demonstrate the
existence of multiple unstable modes undergoing Hopf bifurcation, greatly influenced by
the rotation effect. The shapes of these global eigenmodes are presented and compared, as
well as their structural sensitivity, visualising the flow region important for the disturbance
development with rotation. This research contributes to our understanding of the complex
bluff-body wake dynamics past this critical configuration.
Key words: wakes, vortex instability
of 2-D perturbations by the global stability analysis, which will be extended in the
current study, focusing on a rotating finite cylinder, to account for 3-D perturbations.
Built upon the previous works, El Akoury et al. (2008) extended the neutral stability
curves for these wakes in the Re-α plane by direct numerical simulations (DNS) and the
Landau model. The experiments by Kumar, Cantu & Gonzalez (2011) provided evidence
of the existence of mode II at Re = 200, 300, 400 and 0 < α < 5. The experimental
study of Linh (2011) also reported observations of the low-frequency mode II vortex.
Their experimental Strouhal number and wake patterns agree well with numerical data
of Mittal & Kumar (2003). Later, Pralits, Brandt & Giannetti (2010); Pralits, Giannetti
& Brandt (2013) conducted an extensive study of the linear global dynamics of the 2-D
rotating cylindrical wake flow. The authors explored neutral stability curves on the (Re, α)
plane, providing a comprehensive understanding of this phenomenon. They also observed
multiple steady solutions at high α, explaining the decay of the secondary shedding wake.
More recently, Sierra et al. (2020) fully described the bifurcation, neutral curves and
global instability modes in the parameter space (Re, α) ⊂ [0, 200] × [0, 10], exploring the
relations among Takens–Bogdanov bifurcations, cusps and generalized Hopf bifurcations
975 A15-2
Numerical analyses of flow past a short rotating cylinder
when varying the parameters in the rotating cylinder wake flow. To sum up, for the
infinitely long rotating cylinders, mode I and mode II are fundamentally different flow
phenomena. Mode I undergoes a supercritical Hopf bifurcation and becomes linearly
unstable at 0 ≤ α ≤ 2, which is related to the classical Bénard–von Kármán vortex street,
characterized by alternating vortices with opposite signs of spanwise vorticity. On the
other hand, the physical mechanism of the linear instability mode II (4.5 ≤ α < 6) is
featured by low-frequency vortex shedding with the same vorticity sign.
Further research has shown that rotation can lead to complex 3-D instabilities.
Numerical investigations by Rao et al. (2013a,b) demonstrated several 3-D modes
becoming unstable to spanwise perturbations in the steady and unsteady regimes of
Re = 400 flows. Five 3-D modes were identified to be unstable in the mode I shedding
regime, while four 3-D modes were observed in the steady flow regimes for α ≥ 2. Radi
et al. (2013) proved experimentally the existence of the above numerically predicted 3-D
modes. They additionally showed a highly 3-D wake and the absence of 2-D periodic
shedding at high α (i.e. Re = 200, α = 4.5) previously reported in Mittal & Kumar (2003).
Navrose, Meena & Mittal (2015) conducted 3-D numerical studies and found that the span
length of the rotating cylinder plays an important role in the evolution of the wake with
Re ∈ [200, 350], α ∈ [0, 5]. Specifically, only linear global modes with wavelengths that
are integer multiples of the cylinder span are selected for growth in nonlinear DNS. It is
thus necessary to consider 3-D configurations that take into account the spanwise length
in studies of rotating bluff-body flows.
Researchers have also studied the flow past a rotating bluff body of other forms. It is
instructive to review relevant works on these flows as they will also be discussed in this
study. The flow past a rotating sphere, along either the transverse axis or the streamwise
axis, received considerable attention. Citro et al. (2016) applied the global linear stability
analysis (LSA), adjoint-based structural sensitivity analysis and weakly nonlinear analysis
(WNL) to reveal the mechanism of flow instability around a rotating sphere around the
transverse axis. They characterized the evolution processes of the first (at low α) and
second (at high α) instability modes. Fabre et al. (2017) further improved the WNL from
that in Fabre, Tchoufag & Magnaudet (2012) to achieve a better comparison between the
WNL expansion result and the DNS. √ Namely, the comparison demonstrates that Fabre
et al. (2017)’s expansion ( = Re − Rec /Rec , where Rec is the critical Reynolds
number of a pitchfork bifurcation in their work) provides a better reproduction of the DNS
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
results for both angular velocity and associated lift forces, compared with Fabre et al.
(2012)’s ω expansion (where ω represents the dimensionless rotation rate normalizing
the actual rotation with U0 /D). The ω expansion failed to predict the DNS results for
Re around and beyond Re = 212, whereas the expansion accurately reproduces the
DNS results up to Re ≈ 225. For the flow past a sphere rotating along the streamwise
direction, Lorite-Díez & Jiménez-González (2020) conducted a DNS study on the wake
evolution of a strongly rotating sphere and observed a sequence of continuous bifurcations
from periodic, quasi-periodic and irregular states to chaos, over the parameter range
0 < α < 3 and Re = 250, 500, 1000. Later, Sierra-Ausín et al. (2022) employed global
LSA to determine the neutral curves of three non-zero frequency global modes on the Re-α
plane, and used the normal form expansion to reveal the nonlinear interactions among the
global modes. Their predictions of the normal form analysis were satisfactory and close
to the DNS results, which led to a more detailed phase diagram of the nonlinear patterns.
Besides, Jiménez-González et al. (2014) carried out global LSA of the wake flow past a
streamwise rotating bullet-shaped body and plotted the neutral curves on the Re-α plane.
Their work indicated that the streamwise rotation can also delay the Hopf bifurcation of
975 A15-3
Y. Yang, C. Wang, R. Guo and M. Zhang
a bluff body with a large aspect ratio (AR = 2) when increasing Re, which is different
from a sphere with an aspect ratio of 1. To sum up, in addition to the sphere rotating along
the streamwise direction, the aforementioned rotating bluff bodies of various shapes and
aspect ratios exhibit a moderate rotation regime in the Re-α plane where the neutral curve
of the Hopf bifurcation is significantly shifted to higher Re values. Therefore, the aspect
ratio plays a significant role in the wake transition of a rotating bluff body, which should
be further researched.
Our literature review has identified a research gap in the understanding of the dynamics
of uniform flow past a finite-length rotating cylinder along its cylinder axis. This flow
configuration is common in nature and engineering applications, but its instability
mechanism, bifurcation properties and transition path are still unclear. In a recent study
we conducted a detailed investigation of the wake flow around a non-rotating finite-length
cylinder (Yang, Feng & Zhang 2022). Building on this research, we aim to extend our
investigation to the wake flow around a rotating finite-length cylinder to explore its 3-D
effects, in line with other similar works focusing on the rotation effect in the wake flow
such as Pralits et al. (2010), Citro et al. (2016), Sierra-Ausín et al. (2022) and Zhao &
Zhang (2023), reviewed in Rao et al. (2015). The primary objective of this study is to
clarify the effect of rotation on the unsteadiness of finite-length cylinder wakes using
the (nonlinear) DNS method. Additionally, we aim to determine the instability threshold
at which the unsteadiness occurs using the global stability approach. To identify the
instability region responsible for the unsteadiness, we also probe the structural sensitivity
of the flow. Our work will contribute to a deeper understanding of the wake flow
around rotating finite-length cylinders and provide insights into the dynamics of this flow
configuration, which has practical implications for various engineering applications.
The paper is organised as follows. Section 2 introduces the configuration of a 3-D finite
rotating cylinder flow, the boundary conditions, the governing equations (i.e. nonlinear
Navier–Stokes (NS) equations and their corresponding linearised direct and adjoint
equations) and the numerical methodology. In § 3 we show the results and discuss the base
states (time-averaged flow or steady flow), nonlinear wake patterns, global eigenmodes,
neutral curves in parametric plane Re-α and bifurcations in this flow. Finally, the results are
summarised in § 4 and conclusions are provided. In the appendices we provide additional
results of the Stuart–Landau model, the global modes at different aspect ratios AR and a
verification step of the numerical codes.
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
2. Problem formulation
2.1. Flow configuration and governing equations
We study the 3-D stability of the flow around a finite-length rotating cylinder of length
L, diameter D and aspect ratio AR = L/D, subjected to a uniform incoming flow in a
Cartesian coordinate system. As shown in figure 1, the origin of the coordinate system
is located at the centre of the cylinder, the x axis points in the flow direction, the y
axis represents the transverse direction and the z axis extends along the centreline of the
cylinder. The non-dimensional NS equations for the unsteady Newtonian incompressible
flow read
∂U 1
+ (U · ∇)U = −∇P + ∇ 2 U, ∇ · U = 0, (2.1)
∂t Re
where U = (Ux , Uy , Uz ) is the velocity vector and P is the pressure. The Reynolds
number Re = DU∞ /ν is defined based on cylinder diameter D, the velocity of the uniform
incoming flow U∞ at infinity and the kinematic viscosity coefficient ν. Strouhal number
975 A15-4
Numerical analyses of flow past a short rotating cylinder
(a) (b)
In Gmsh In Nek5000
en
eb O-type mesh
eτ
z
Sin y Ω Sxz
O-type mesh
Sc
U∞ x Sout
L
o La
D y
L Sxy
a Cylinder
x
La Lo (AR = 1)
Figure 1. The computational domain and boundary conditions (not to scale) (a) and mesh design (b). The
red unit vectors (en , eτ , eb ) in panel (a) represent the directional vectors of the surface Sxy,t . The finite-length
cylinder is rotating around its axis that is perpendicular to the incoming flow.
where ρ is the fluid density, Fdp = Sc Px dS and Fdv = Sc τwx dS are the pressure drag and
friction drag on the cylinder surface Sc along the streamwise direction, Flp = Sc Py,z dS
and Flv = Sc τwy,wz dS are the pressure lift and wall shear stress lift acting in either y axis
or z axis (defined as Cly and Clz , respectively) and A is the reference area A = LD. Here
(Px , Py , Pz ) are the components of the pressure acting on the cylinder surface along the
x, y and z axes, respectively. And τw is wall (surface) shear stresses. Furthermore, we will
also use the letters C̄d and C̄l to denote the time-averaged values of Cd and Cl , respectively.
As shown in figure 1, Sc represents the surface of the cylinder. Here Sin and Sout
represent the inlet and outlet surfaces of the rectangular computation domain, whose
normal is along the x direction; Sxy,t , Sxy,b , Sxz,f and Sxz,b denote the surfaces of the cuboid
on the top, bottom, front and back side walls, which are parallel to the xy, xy, xz and xz
planes, respectively. The boundary conditions of the system (2.1) are
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
where êz = (0, 0, 1) is a unit vector aligned with the positive z axis. The vector r =
(xc , yc , zc ) is the position vector of a point located on the cylinder surface, i.e. the vector
from the origin of the coordinate system to the point. Here en , eτ , eb are the unit normal,
unit tangent and unit bitangent vectors, respectively. As shown in figure 1(a), the vector
en of surfaces Sxy,t , Sxy,b , Sxz,f , Sxz,b and Sout points out the computational domain. The
directions of the vector eτ of surfaces Sxy,t , Sxy,b , Sxz,f , Sxz,b point along positive x axis,
negative x axis, negative x axis and positive x axis, respectively. The vector eb points in the
direction that is perpendicular to both the normal vector and the tangent vector. Here I is
the identity tensor.
975 A15-5
Y. Yang, C. Wang, R. Guo and M. Zhang
2.2. Linearisation
The global linear stability/instability of the flows past the finite-length rotating cylinder
will be studied. Reynolds decomposition U = U b + u, P = Pb + p will be substituted
into the nonlinear governing equations (2.1). The base-state terms (U b , Pb ) satisfying
the steady NS equations and the nonlinear terms are neglected, yielding the linearised
equations for the infinitesimal perturbations (u, p) residing on these base states, i.e.
∂u 1
+ (U b · ∇)u + (u · ∇)U b = −∇p + ∇ 2 u, ∇ · u = 0, (2.4)
∂t Re
where u is the 3-D perturbation velocity vector u = (ux , uy , uz ) and p is the perturbation
pressure. Homogeneous boundary conditions are applied for the perturbed variables as
follows:
∂q
M = LU b q, (2.6)
∂t
where LU b is the linearised NS operator depending on the base states U b . The elements of
mass matrix M and the Jacobian matrix LU b are
I 0 −U b · ∇ − ∇U b + Re−1 ∇ 2 −∇
M = , LU b = . (2.7a,b)
0 0 ∇· 0
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
As the considered base flow states are steady, we seek the wavelike solution q(x, y, z, t) of
the form
q(x, y, z, t) = q̂(x, y, z)eλt , where λ = σ + i2πω. (2.8)
Substituting this form (2.8) into (2.6), we can get the eigenvalue problem
LU b q̂ = λM q̂, (2.9)
where the stability of the base state U b is dictated by the eigenvalues λ in the
linearised problem with σ being the temporal growth/decay rate of perturbations and
ω the eigenfrequency. The flow is linearly unstable if σ > 0; stable otherwise. The
eigenfrequency ω of the most unstable eigenvalue determines whether the base state U b
experiences a regular bifurcation (ω = 0) or a Hopf bifurcation (ω > 0). Note that the flow
problem considered in this work is not spatially periodic or homogeneous in either x, y, z
directions, and q̂ depends on all the three coordinates, leading to a global stability problem
(Theofilis 2011).
975 A15-6
Numerical analyses of flow past a short rotating cylinder
2.3. Sensitivity analysis
Sensitivity analyses based on the adjoint approach (Luchini & Bottaro 2014) will
be conducted to identify the instability mechanism responsible for the unsteadiness.
Following Giannetti & Luchini (2007), the adjoint equations of the linearised NS equations
read
∂u+ 1
− − U b · (∇u+ ) + (∇U b ) · u+ = −∇p+ + ∇ 2 u+ , ∇ · u+ = 0, (2.10)
∂t Re
where u+ and p+ are the adjoint vector of perturbation field u and p, respectively.
Following Giannetti & Luchini (2007), Marquet, Sipp & Jacquin (2008) and Citro et al.
(2016), the boundary conditions of the adjoint equations are set as
u+ = 0 on Sc , Sin , Sxz and Sxy , (2.11a)
p+ n − Re−1 (∇u+ ) · n = (U b · n)u+ on Sout . (2.11b)
The identification of the core region of the instability can help to understand the
instability mechanism (Giannetti & Luchini 2007; Luchini & Bottaro 2014). According to
Giannetti & Luchini (2007), the sensitivity wavemaker ζ can be identified by overlapping
the direct eigenvector u and adjoint eigenvector u+ ,
|u||u+ |
ζ = . (2.12)
u, u+
the boundary layer elements in the vicinity of the rotating cylinder have been refined by
the O-type mesh (see figure 1b). The only difference lies in the enlarged computational
domain and increased number of elements to accommodate the rotational effects.
We focus on studying two types of base states (both of which are denoted as Qb =
(U b , Pb )T in the text to follow as long as there is no confusion) to analyse their global
dynamics. The flow will become unstable in a certain region of the parametric space. The
unstable base flow Qb in this case cannot be obtained directly by time evolving the NS
equations. For these unstable base flows, the selective frequency damping (SFD) method
proposed by Akervik et al. (2006) was used to obtain an equilibrium solution to the NS
equations (2.1). This type of base state will be called (SFD) base flow. Another base state of
interest is the mean flow, which is obtained by time averaging the periodic flow with vortex
shedding. For the present global stability analysis, at least ten vortex shedding cycles will
be used in the time-average procedure.
For a non-parallel 3-D flow past the finite rotating cylinder, the numerical discretisation
of linearised NS equation (2.4) will result in a large-scale Jacobian matrix LUb in the
generalized eigenvalue problem (2.9). It is impractical to solve a large-scale eigenvalue
problem for its whole eigenspectrum in a 3-D flow. Based on Nek5000 solver and
975 A15-7
Y. Yang, C. Wang, R. Guo and M. Zhang
the ARPACK package (Lehoucq, Sorensen & Yang 1998), the matrix-free time-stepper
method (Theofilis 2011; Doedel & Tuckerman 2012) will be adopted in the present work,
implementing the implicitly restarted Arnoldi method (Radke 1996; Lehoucq et al. 1998).
The validation of the nonlinear DNS code and the linear stability code is provided in
Appendix C together with a convergence study on the size of the computational domain.
that the rotation breaks the symmetry of the wake, leaving the flow only symmetric with
respect to the oxy plane. Compared with the non-rotation case (figure 2a), the positions
of the upper and lower separation points shift significantly along the rotation direction.
Overall, the asymmetric recirculation region generated by the weak rotation is similar
to the asymmetric wake generated by regular bifurcation in the non-rotated case. The
counterclockwise rotation makes the flow velocity near the cylinder arc surface on the
lower side increase, and the velocity near the upper wall surface decrease. It can be
deduced that the pressure near the lower side wall will decrease, and the pressure near
the upper side wall will increase, so a negative lift Cly will be obtained, i.e. the Magnus
effect. It can be seen from figure 2(c–h) that the numerical results conform to the classical
Magnus effect. In the parameter space studied in this paper (especially relatively low Re),
there is no boundary layer transition from a laminar flow to turbulence, that is, there is
no inverse Magnus effect (Kim et al. 2014). As mentioned above, the symmetry breaking
effect caused by the regular bifurcation resembles that due to the rotation (see also the
rotating sphere by Citro et al. 2016). Therefore, the rotation ’strengthens’ the symmetry
breaking caused by the inherent wake mechanism observed in the regular bifurcation
without rotation, resulting in the lift force in the rotating cases in figure 2(c–h) being
975 A15-8
Numerical analyses of flow past a short rotating cylinder
(a) (b) (c) (d )
(i) (i) (i) (i)
Figure 2. Steady SFD base flow past an AR = 1 short cylinder. The colour illustrates pressure contour
and the white lines streamlines. Results are shown for (a) Re = 160, α = 0; (b) Re = 330, α = 0;
(c) Re = 160, α = 0.1; (d) Re = 330, α = 0.1; (e) Re = 160, α = 0.6; ( f ) Re = 330, α = 0.6; (g) Re = 160,
α = 1.2; (h) Re = 330, α = 1.2. For each subgraph, plots (a i,b i,c i,d i,e i, f i,g i,h i) show the flow visualisation
at the plane z = 0 and plots (a ii,b ii,c ii,d ii,e ii, f ii,g ii,h ii) at the plane y = 0. The rotation direction of the
cylinder is counterclockwise. The black translucent thick solid lines denote the recirculating region separatrix,
which is identified by Ux = 0.
larger, to be shown in figure 4(b). On the other hand, the drag coefficient Cd is not very
sensitive to the rotation when at low rotation rate α < 0.3, as shown in figure 4(a).
Figures 2(e) (Re = 160) and 2( f ) (Re = 330) show the base states at moderate rotation
α = 0.6, and both their nonlinear saturation states are steady. So the base flow and mean
flow are the same for these cases. Compared with the previous low rotation case, the flow
topology further changes, i.e. the recirculation zone in the wake almost disappears, in
either the xy or xz plane. But a stagnation point (white point in the panel) similar to that
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
observed in the 2-D rotating cylinder (see figure 3b in Sierra et al. 2020) appears, which
is located in the second quadrant of the cylinder.
Continuing to increase the rotation rate, figures 2(g) (Re = 160) and 2(h) (Re = 330)
show the cases of a high rotation rate (α = 1.2), and their wake structure and pressure
distribution are similar to those of figures 2(e) and 2( f ), but the wakes are swung further
upwards. The hyperbolic stagnation point (see the white dot) on the upper side moves
forward along the direction of tangential velocity, which is similar to the results of Citro
et al. (2016) on a rotating sphere at a high rotation rate.
.0
=2
y
1.5
Fixed sphere
1.5
=
xs α = 0.0
α
0
1. 0.0 α = 0.025
Ux = 0 xs α
= α=
0.5 α = 0.1
= α = 0.2
α
x 1.0 α = 0.4
θs θs = π α = 0.6
α = 1.2
0.5 α = 1.6
sphere (Johnson & Patel 1999) and rotating sphere (Sierra-Ausín et al. 2022). The increase
of α decreases the value of xs for 0.025 ≤ α ≤ 1.2, which means that the rotation shortens
the recirculation length. It has been proposed that shortening the recirculation length
stabilises the flow; for example, Sierra-Ausín et al. (2022) discussed such a stabilising
control strategy.
When the rate α is high, xs slightly decreases with the increase of Re, for example, see
α = 0.6, 1.2 in figure 3(b). According to the above discussion on the relation between
rotation, recirculation length and flow stability, we cannot make the present rotating
cylinder flow more stable at high α ≥ 0.6-1.2 by shortening the base flow recirculation
region along the streamwise direction because the value of xs bounces back when α ≥
0.6–1.2. We provide a further mechanistic explanation of this result in the structural
sensitivity analysis § 3.2.3. Besides, the higher the rotation rate, the less sensitive the value
of xs is to the variation of the Reynolds number.
For fixed Re, the effect of rotation rate α on xs is not monotonous, unlike the sphere
rotating along the streamwise axis (Sierra-Ausín et al. 2022). Sierra-Ausín et al. (2022)
also suggested that xs does not have to increase monotonically as α increases after
the bifurcation, which can also be seen in the time-averaged recirculation length of
975 A15-10
Numerical analyses of flow past a short rotating cylinder
the sphere rotating along the streamwise direction (Kim & Choi 2002; Lorite-Díez &
Jiménez-González 2020).
the pressure differential can be regarded as the net area enclosed between the blue solid
line and the Psx = 0 axis, specifically Σ1 − Σ2 + Σ3 . Consequently, it can be observed
that the increase in drag is primarily attributed to a significant reduction (see panel a) in
pressure Ps on the back surface of the cylinder due to rotation, resulting in a substantial
enlargement (see panel b) of the area Σ3 .
As with a fixed infinite cylinder (Schlichting & Gersten 2016), a fixed sphere (Johnson &
Patel 1999) and a fixed finite-length cylinder (Yang et al. 2022), within the parameter range
shown in figure 4(a), an increase in Re generally leads to a decrease in drag coefficient.
On the other hand, unlike the drag coefficient Cd , the values of the lift coefficient Cly
in figure 4(b) present more variation when Re changes. At low rate α ≤ 0.3, Cly first
decreases with the increase of Re and then increases; at high rate α ≥ 0.6, Cly increases
monotonically as Re increases from 50 to 330. Besides, in the interval 0.2 < α < 0.6
of figure 4(b), Cly is relatively insensitive to Re, and the wake is steady (see the neutral
stability curves in the next section). It is worth noting that the lift coefficient Cly |α=c of the
rotating cylinder and the Cly |α=2c of the rotating sphere gradually coincide for Re 270
(with c ≤ 0.1 in present works); see the top-right corner in panel (b). For example, the
values of Cly |α=0.025 of a rotating cylinder and Cly |α=0.05 of a rotating sphere, or Cly |α=0.05
975 A15-11
Y. Yang, C. Wang, R. Guo and M. Zhang
Sphere (Johnson & Patel 1999) α = 0.025
α = 0.0 (Citro et al. 2016) α = 0.05
α = 0.01 (Citro et al. 2016) α = 0.1
α = 0.05 (Citro et al. 2016) α = 0.2
α = 0.1 (Citro et al. 2016) α = 0.4
α = 0.2 (Citro et al. 2016) α = 0.6
α = 0.0 (Yang et al. 2022) α = 1.2
(a) 1.7 (b) 0
1.6
1.5 –0.2
1.4
–0.4
1.3
Cd 1.2 Cly
–0.6
1.1
1.0
–0.8
0.9
0.8 –1.0
0.7
50 100 150 200 250 300 50 100 150 200 250 300
Re Re
Figure 4. Drag (a) and lift (b) coefficients of the steady SFD base flow as a function of Re at AR = 1.
Comparison with a fixed sphere (Johnson & Patel 1999), a fixed finite cylinder (Yang et al. 2022) and a sphere
rotating about the transverse direction (Citro et al. 2016).
(a) (b)
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
0.5 0.8 α
α
Span-averaged pressure Ps
= 0.0 = 0.9
Front surface Back surface α = 0.05 α = 1.2
0.3 α = 0.1 α = 1.4
α α
α decrease 0.6
α
= 0.2
= 0.4 α
= 1.6
= 1.8
0.1 α = 0.6 α = 2.0
–0.1 0.4
Psx
–0.3 y Ps
Psy 0.2
–0.5 θ Psx Σ1 Σ3
o x 0
–0.7 Σ2
–0.9 –0.2 Front surface Back surface
0 30 60 90 120 150 180 0 30 60 90 120 150 180
θ (deg.) θ (deg.)
Figure 5. The distribution of span-averaged pressure Ps (a) acting on the upper surface, as well as its
component Psx = Ps cos θ (b) along the x axis, is characterized with respect to the angle θ . The conditions
considered involve cases of (Re = 160, AR = 1, α < 2). As an example, the areas corresponding to the shaded
regions in panel (b) are denoted as Σ1 (coloured by red), Σ2 (coloured by green) and Σ3 (coloured by grey)
for the case α = 2.
975 A15-12
Numerical analyses of flow past a short rotating cylinder
of a rotating cylinder and Cly |α=0.1 of a rotating sphere, are approximately the same for
Re 270.
mode and the corresponding wake pattern observed in DNS, if their frequencies are close.
Global mode LB, as shown in figure 7(b), with a higher frequency is caused by the vortices
shedding from the curved surface and the associated nonlinear flow pattern wake LB is
shown in figure 13(b). In figure 7 we have additionally shown two additional modes (named
LC and LD) at α = 0.1 and Re = 290. Mode LC in panel (c) has zero frequency and
presents a long smooth streamwise structure. The leading eigenmode in some other cases,
such as LT, may take this form. As shown in figure 8(a), LT represents a low-rotation-rate
transition state where the LA and LB modes are both neutral, to be discussed further below.
Mode LD in panel (d) appears similar to mode B identified in Yang et al. (2022) for the
non-rotating short cylinder. Each global mode in figure 7 maintains the same symmetry as
the base flow, namely, symmetry with respect to the xy plane, consistent with that of the
non-rotating short cylinder (Yang et al. 2022).
Examining closely the solid and dashed lines in figure 6(a), when α is relatively small,
mode LA becomes unstable before mode LB when increasing Re. In the range of larger α,
mode LB starts to become unstable before mode LA. The intersection point of the σ –Re
lines of mode LA and mode LB denotes the condition where σLA = σLB , as represented by
the black stars in figure 6(a). This competitive phenomenon also exists in spheres rotating
975 A15-13
Y. Yang, C. Wang, R. Guo and M. Zhang
(a) 0.10 (b)
α = 0, mode LA 0.21
α = 0, mode LB
0.08 α = 0.025, mode LA 0.20
α = 0.025, mode LB
0.06 α = 0.1, mode LA
α = 0.1, mode LB 0.19
α = 0.15, mode LA
0.04 α = 0.15, mode LB 0.18
σ St
0.02 0.17
0 0.16
0.15
–0.02
0.14
–0.04
270 280 290 300 310 320 330 270 280 290 300 310 320 330
Re Re
(c) 0.10
0.08
LP1
0.06 LP2
LT
0.04 LP3
σ LP4
0.02 LP5
0
–0.02
(a) (b)
1 1
–1 –1
2 2
0 0
15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
0 5 10
(c) (d )
1 1
–1 –1
2 2
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Figure 7. Four representative eigenmodes with weak asymmetry in the flow past a short rotating cylinder at
α = 0.1 and Re = 290. Mode LA (in panel a) and mode LD (in panel d) are similar to mode A and mode
B reported by Yang et al. (2022), respectively. The new mode LB (in panel b) with high frequency is more
asymmetric caused by rotation. The mode LC (in panel c) has zero frequency. The Q-criterion isosurfaces
Q = 0 are coloured by the x component of the vorticity ranging from −2 × 10−3 to 2 × 10−3 . The
corresponding eigenvalues are (a) LA mode λLA = 4.120 × 10−3 + i0.8833, (b) LB mode λLB = −2.202 ×
10−2 + i1.176, (c) LC mode λLC = −6.295 × 10−2 + i0.0, (d) LD mode λLD = −8.683 × 10−2 + i0.8687.
along the streamwise direction (Sierra-Ausín et al. 2022), but differs from the single-mode
instability observed in spheres rotating along the transverse direction (Citro et al. 2016).
Note that our cylinder is rotating along its axis that is in the transverse direction.
975 A15-14
Numerical analyses of flow past a short rotating cylinder
Sphere (Citro et al. 2016)
Mode LA
Mode LB
σLA = σLB
(a) 0.35 (b) 0.35
0.30 0.30
0.25 0.25
0.05 0.05
LP1 LP2
0 LP3
240 260 280 300 320 340 360 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Re St0
Figure 8. (a) Neutral stability curves undergoing the Hopf bifurcation for the flow past a finite rotating cylinder
at AR = 1 and the rotating sphere (the mode I by Citro et al. 2016). The specifications of the points LP1-5
and the codimension-two point LT are shown in table 1. The hollow symbols represent the corresponding
nonlinearly saturated wake that looks similarly to global mode LA, whereas the solid symbols denote the
nonlinearly saturated wake that more closely resembles global mode LB. (b) Plot of critical Strouhal numbers
St0 on the neutral stability curves against rotating ratio α.
The neutral curves in the Re-α plane associated to the two most unstable modes LA
and LB are displayed in figure 8. The present asymmetric steady state is linearly stable
in the blank region and linearly unstable in the shaded region, as shown in panel (a).
With the rotation speed increasing, the threshold Reynolds number for the instability first
decreases slightly and then increases. The critical Reynolds number approximately reads
Rec = 279 at α = 0.0375. As the rotational speed continues to increase, the threshold
Reynolds number becomes less sensitive to the rotation speed. Thus, the result indicates
that the rotation of a short cylinder can influence and control the stability properties of the
flow.
The present work designates the overlapping area between the unstable regions of modes
LA and LB in figure 8(a) as the bi-unstable region, where points LP2, LP3 and LP4 are
located. In this study the intersection of the neutral curves of mode LA and mode LB is
denoted as the codimension-two point LT, where two different bifurcations (due to mode
975 A15-15
Y. Yang, C. Wang, R. Guo and M. Zhang
(a) (b)
2.0 2.0
HP3 HP4 Sphere (Citro et al. 2016)
1.8 Mode HA 1.8
Mode HB
HT3 Mode HC
1.6 HT1 1.6
HT2 1.4
α 1.4
HP2 HP1
1.2 1.2
1.0 1.0
0.8 0.8
150 200 250 300 350 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Re St0
Figure 9. (a) Neutral stability curves undergoing the Hopf bifurcation for the flow past a finite rotating cylinder
at AR = 1 and the rotating sphere (the mode I by Citro et al. 2016). The shaded area indicates a region of linear
instability. The eigenvalues of points HP1–HP4 are shown in table 2. (b) Plot of critical Strouhal numbers St0
against rotating ratio α.
LA and mode LB) occur simultaneously. Similar to the rotating sphere (Sierra-Ausín et al.
2022), the present point LT, as the organizing centre of the linear system, represents a
turning point of competition between different modes and results in the generation of three
distinct wake patterns around it in the nonlinear system (to be discussed in figure 13). The
dotted line in figure 8(a) passes the codimension-two point LT in the bi-unstable region,
on which the growth rate σLA = σLB . Above this dotted line, the growth rate of mode LB
is greater than that of mode LA; vice versa.
The frequencies St0 corresponding to the neutral conditions in panel (a) are reported
in figure 8(b) as a function of α. As the frequencies are non-zero, the unstable flows will
undergo Hopf bifurcation, that is, the unstable mode will oscillate at a certain frequency.
Similar to the flow past a sphere rotating along the transverse (Citro et al. 2016) and
streamwise directions (Sierra-Ausín et al. 2022), the St0 of present rotating cylinders
increase rapidly as α increases in the regime of low rotation rates α < 0.3.
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
and high frequency for HC. The non-zero frequencies of these modes again indicate
that the unstable wake flow experiences Hopf bifurcation. By comparison to figure 8(b),
the influence of the rotation on the eigenfrequencies of HA, HB and HC is less
significant compared with the mode LB at low rotation rate. This indicates that in the
high-rotation-speed regime, changing the value of α affects less the frequency in the
flow. For the following discussions, when a nonlinear wake flow possesses multiple
characteristic frequencies of these global modes at the same time, we will name the
wake flow by combining the modes; for example, if both the eigenfrequencies of HA and
HB modes are observed in a nonlinear wake, we will call it HAB (see figure 13 to be
discussed).
Besides, the three unstable modes can also interact with each other, resulting in the
three turning points HT1, HT2, HT3 as shown in figure 9(a). In the parameter ranges
of 0.9 < α < 1.64 and 203 < Re < 335, the neutral curves of global modes HA and
HB almost coincide, that is, the low-frequency mode HA and mid-frequency mode HB
simultaneously become linear unstable. For α > 1.64 (see, e.g. the point HT1 in figure 9a),
the Hopf bifurcation in the flow begins to be dominated by the high-frequency mode HC.
In general, the differences of the Hopf bifurcation in our case and that in the rotating
sphere are summarised as follows. (1) When α > 0.7, Citro et al. (2016) reports only one
unstable mode in the flow past a rotating sphere. However, there are three unstable modes
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
in our case, indicating that the Hopf bifurcation process in the present finite cylinder may
be more complex. (2) The eigenfrequency of a sphere is in general higher than that of the
rotating cylinder (see the black line with circles in panel 9b).
In figure 10 we show the flow structure of the global eigenfunctions for the three
eigenmodes HA, HB, HC at the point HP3 in figure 9. It can be seen that global mode HA
(figure 10a) and mode LA (figure 7a) have a similar wake structure, and the difference
is that the greater rotation rate makes the transverse offset of mode HA larger. The
intermediate-frequency mode HB and the high-frequency mode HC have smaller flow
structures than those in mode HA.
(c)
2 Vorticity
–2 2 × 10–3
5 0
–2 × 10–3
0 30 35 40 45 50
0 5 10 15 20 25
Figure 10. The global modes with strong asymmetry for the flow past a rotating finite cylinder at a high
rotation rate at Re = 170, α = 1.8 (point HP3). The Q-criterion isosurfaces Q = 0 are coloured by the x
component of the vorticity ranging from −2 × 10−3 to 2 × 10−3 . The cylinder centroid is located at (0, 0, 0).
Results are shown for the (a) HA mode, (b) HB mode, (c) HC mode.
according to ζ in (2.12), delimit the wavemaker region, which is the superposition of the
leading global mode and its adjoint mode. For both the unstable LA and LB modes, we
compute their adjoint modes in the global LSA based on the SFD base flow and time-mean
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
base flow, respectively. In general, the wavemaker region is located in the near wake region
of the cylinder (in the top-right region of the xy plane). Its structure remains the same
symmetry as the base state, being symmetric with respect to the Oxy plane. The spatial
distributions of wakemakers based on mean flow and base flow are similar. Since the
wavemaker region indicates the most sensitive region in the flow, one can infer from these
observations that (i) the region responsible for the instability is located in the recirculation
region behind the cylinder, and (ii) the instability mainly amplifies the perturbations near
the cylinder surface (Citro et al. 2016).
Figure 12 presents the flow sensitivity for the high rotation case of HP3, showing
the superposition of the global modes HA, HB, HC based on the SFD base flow with
their respective adjoint modes. One can see that the three wavemaker regions differ
significantly, in contrast to the low rotation case. The wavemaker region for the HB mode
is located further downstream of the rotating cylinder compared with the other two modes,
whose sensitivity regions are close to the cylinder. This implies that the control of the
unstable modes in the high rotation case can be treated separately; especially, control of
the HB mode may be achieved more easily as it is not mingled with other modes in the
spatial distribution.
975 A15-18
Numerical analyses of flow past a short rotating cylinder
–0.5
0 0.5 –0.5 0 0.5
–0.5
0 0.5
1 0 0 0
1 1
–0.5 –0.5 –0.5
2 2 2
1 1 1
3 0 3 0 3 0
Figure 12. Flow sensitivity in a high-rotation-rate case. The most unstable wavemaker isosurfaces are plotted
based on the SFD base flow of a rotating cylinder at point HP3. Transparent red is for ζ = 0.2 and opaque blue
is for ζ = 0.4. Panels (a–c) correspond to global modes HA, HB and HC, respectively.
lift coefficient, the drag coefficient seems to be a more robust option for analysing the
time-history data in our case. When the vortices shed alternately, the frequency of the
drag coefficient is twice that of the lift coefficient.
At low rotation, panels 13(a,b) show that the saturated states at points LP2 and LP4
are a limit cycle. Wakes LA and LB represent the wake dominated by vortices shedding
from the cylinder’s flat ends and the circular arc surface, respectively. Both types of wake
structures, LA and LB, are also observed in the fixed cylinder flow (see wake patterns
P3-2 and P3-1 in figure 10 of Yang et al. 2022, respectively). The difference is that, due to
the rotation effect, wake LB undergoes a Hopf bifurcation and becomes a saturated wake
state. In the non-rotating cylinder flow, wake LB is only an intermediate transitional state.
The results of high rotation speeds are shown in panels (c–f ). Three frequencies of
oscillations, including low (HA), medium (HB) and high (HC), are identified in the
present nonlinear wakes; see previous discussions on figure 9. Wake HA pertaining to
the point HP1 (figure 13c) is characterized by a low-frequency oscillation in the spanwise
direction and its higher-order harmonics (see figure 14c). Its phase diagram of the lift-drag
coefficient is also a limit cycle. Wake HAB (figure 13d) is identified with both a low-
975 A15-19
Y. Yang, C. Wang, R. Guo and M. Zhang
(ii) Cd – Cly (iii) Cd – Clz
(a) 0.01
–0.1 0.1 –0.1268
(i)
2 0
0
–2
3 35 40 45 50 –0.1269 –0.01
0 20 25 30 0.74984 0.74992 0.74984 0.74992
–3 0 5 10 15
(ii) Cd – Cly (iii) Cd – Clz
(b) –0.1 0.1
–0.1768 0.001
(i)
2
0 0
–23 40 45 50
0 25 30 35 –0.1785 –0.001
–3 0 5 10 15 20
0.7383 0.73845 0.7383 0.73845
–0.1 0.1 (ii) (iii)
(c) Cd – Cly Cd – Clz
3 0.02
(i) –0.896
–3 0
6
0 50 –0.900 –0.02
20 25 30 35 40 45
0 5 10 15 1.40 1.42 1.40 1.42
4 –0.92 –0.05
0 25 30 35 40 45 50 1.65 1.70 1.65 1.70
0 5 10 15 20
Figure 13. (a i,b i,c i,d i,e i, f i) The nonlinear wakes spatial structure obtained by DNS. The Q = 0 isosurfaces
are coloured by the streamwise vorticity ranging from −0.1 to 0.1. (a ii,iii,b ii,iii,c ii,iii,d ii,iii,e ii,iii, f ii,iii)
The corresponding phase diagrams of Cd − Cl for a rotating cylinder. (a) Wake LA at point LP2, (b) wake
LB at point LP4, (c) wake HA at point HP1, (d) wake HAB at point HP2, (e) wake HC at point HP3 and
( f ) chaotic wake HAC at point HP4. The corresponding Re, α and eigenvalues for each case are shown in
tables 1 and 2.
and a medium-frequency oscillation at point HP2. The phase diagram indicates a limit
torus with two incommensurate frequencies; see also the PSD result in figure 14(d). From
the perspective of the flow structure, the medium-frequency oscillation (HB) is caused by
the vortices shedding from the cylindrical arc surface, while the low frequency (HA) is
associated to the oscillation of the vortices in the z direction. Currently, we are not able
to identify a monochromatic wake HB with only a medium frequency. From the vortex
street structure of the wake HAB at point HP2, it can be seen that the medium-frequency
975 A15-20
Numerical analyses of flow past a short rotating cylinder
(a) (b) (c)
10–1 St = 0.1443 LP2 St1 = 0.1977 LP4 St1 = 0.04055 HP1
1 10–4 10–2
3St1
10–4
5St1
10–6 10–6
PSD
10–7 2St1
3St1
10–10 5St1 10–8 10–10
10–4
10–4
10–10 10–6
10–6
10–14 10–8
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
St St St
Figure 14. The PSD for the cases (a) LP2, (b) LP4, (c) HP1, (d) HP2, (e) HP3 and ( f ) HP4. The PSD is
calculated based on the oscillatory part of the time series of the drag coefficient.
content exists (referring to the grey-coloured structure in the upper part). In § 3.3.2 we
further use the DMD method to decompose the main components in the HAB wake to
understand wake HB. Wake HAC (figure 13e) is identified with a high frequency and a low
frequency at point HP3, and its phase diagram of the lift-drag coefficient is also a limit
torus (see figure 14(e) for the PSD result). Finally, in figure 13( f ) a main low-frequency
oscillation and its higher harmonics are identified at point HP4, along with a broadband
of high-frequency oscillations, leading to a very chaotic signal. Point HP4 is located at
the region where the modes HA and HB are both unstable from the linear analysis. The
corresponding phase diagram is also more chaotic compared with the previous cases.
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
Imaginary part
Imaginary part
3 3
0.5 0.4 2 0.5 0.5 2
1 1
St
0
St
0 0 0 0 0
–0.5 –0.4 –0.5 –0.5
–1.0 –0.8 –1.0 –1.0
–1 0 1 –15 –10 –5 0 –1 0 1 –6 –4 –2 0
(×10–5) (×10–5)
Real part Growth rate Real part Growth rate
Figure 15. The DMD spectra of point LP2 in panels (a,c) and of point LP4 in panels (c,d). See figure 8(a) and
table 1 for the definitions of points LP2 and LP4. Panels (a,c) are on the unit circle and panels (b,d) are on the
growth-rate-St plane, where the DMD modes 0, 1, 2 and 3 are marked in red, to be discussed in figure 16.
state before capturing the data. In cases where the wake exhibited multiple cycles, we
considered the longest cycle when determining the period of interest.
We first recall the results in figure 8(a) that below the codimension-two point LT for
α < 0.1294, the steady-state flow transitions supercritically to a wave LA as Re exceeds
Rec (represented by the LA curve in this range of α). To demonstrate the supercriticality,
we have calculated the Landau coefficient c1 and c3 in Appendix A; see figure 24. Above
the codimension-two point, i.e. α > 0.1294, the steady-state flow transitions to a wave LB
with high frequencies. That is, the codimension-two point corresponds to a double Hopf
bifurcation, which characterizes the interaction between mode LA and mode LB. In the
overlap shaded area shown on the Re-α plane in figure 8(a), where modes LA and LB both
are linearly unstable in the rotating cylinder flow, we cannot obtain the single-periodic state
corresponding to LA or LB separately for low α < 0.3 by using different initial conditions.
That is, regardless of the initial conditions, our nonlinear flow in the low-rotation-rate
regime always converges to the most unstable mode. This is in contrast to the results of
Sierra-Ausín et al. (2022) on a rotating sphere, where they identified a bi-stable region
where a single-mode state can be obtained separately and different initial conditions may
lead to different flow modes. In figure 15 the DMD spectra of the flows with low rotation
rates corresponding to the points LP2 and LP4 in figure 8 are displayed. The results feature
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
(c) (d)
–1 –1
0 0
1 1
–2 0 5 10 15 20 –2 0 5 10 15 20
(e) (f)
2 2
(g) (h)
2 2
Figure 16. Comparison of the DMD modes with the global modes at the point LP2 (α = 0.05, Re = 305).
Global modes LA (e,g) and LB ( f,h) in the global stability analysis based on the SFD base flow (e, f ) and mean
flow (g,h). All figures coloured by streamwise vorticity. The eigenfrequencies of the DMD modes are in good
agreement with those obtained from the FFT method (figure 14a), with a relative error of 0.48 %. Moreover,
the eigenfrequencies of the DMD modes match well with the eigenfrequencies of the mean flow global modes
(panel g, f ), with a relative error of 0.83 %. However, the difference between the characteristic frequencies of
the DMD modes and the eigenfrequencies of the SFD base flow global mode is slightly larger, with a relative
error of 2.22 %. Additionally, the topological structure of the DMD modes is almost identical to that of the
mean flow global modes. The corresponding eigenvalues are (a) mode 0, σ + iω = 4.5919 × 10−8 + i0.0; (b)
mode 1, σ + iω = −1.436 × 10−6 + i0.1436; (c) mode 2, σ + iω = −3.5685 × 10−5 + i0.2872; (d) mode 3,
σ + iω = −6.4764 × 10−5 + i0.4309; (e) modeLA (σ + iω)BF = 3.3957 × 10−2 + i0.1411; ( f ) modeLB ,
(σ + iω)BF = 1.6972 × 10−2 + i0.1782; (g) modeLA , (σ + iω)MF = 8.4211 × 10−3 + i0.1431 and (h) modeLB ,
(σ + iω)MF = −4.3530 × 10−2 + i0.1915.
Similarly, the DMD and global LSA results for the point LP4 are shown in figure 17.
One can see that the frequencies of the leading DMD mode (panel b) and the global mode
based on the SFD base flow (panel e) are 0.1976 and 0.2077, respectively, with a difference
of 4.9 %. Also, there is no DMD mode with a frequency close to ω = 0.1451, which is a
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
linearly unstable mode (figure 17f ). The leading global mode based on the time-mean flow
in panel g has a closer frequency (0.1917) compared with that based on the SFD base flow
to the DMD mode. The global eigenfunction based on the time-mean flow also looks more
similar to the DMD mode.
The above results pertain to the low-rotation-rate cases. Next, we compare the linear
and nonlinear results for the high-rotation-rate flows. We plot the DMD eigenspectra in
figure 18 for the selected HP2 and HP3 cases. We discuss the DMD modes labelled in red
in the figures. For HP2, the DMD modes are plotted in figure 19(a,b) along with the global
modes based on the SFD base flow (c,d) and the time-mean flow (e, f ). Compared with
the case of a low rotation rate in figures 16 and 17, the similarity of the DMD modes with
the global modes in the high-rotation-rate case is greater; for example, the first column in
figure 19 shows that the three modes look similar, and the global mode based on the mean
flow is again slightly better compared with the DMD mode. In the second column, we can
also find DMD mode 2 that resembles the global modes in the global stability analyses,
where the global mode based on the time-mean base flow looks closer to the DMD mode.
The same conclusion can be drawn for the HP3 point in figure 20. As the discussions are
similar, we will not go into detail about them. To sum up, through the comparison of the
975 A15-23
Y. Yang, C. Wang, R. Guo and M. Zhang
(a) (b)
–1 –1
0 0
1 1
–2 0 5 10 15 20 –2 0 5 10 15 20
(c) (d)
–1 –1
0 0
1 1
–2 0 5 10 15 20 –2 0 5 10 15 20
(e) (f)
2 2
(g) (h)
2 2
Figure 17. Comparison of the DMD modes with the global modes at the point LP4 (α = 0.1, Re = 330).
Global modes LA ( f,h) and LB (e,g) in the global stability analysis based on the SFD base flow (e, f ) and mean
flow (g,h). All figures coloured by streamwise vorticity. The eigenfrequencies of the DMD modes are in good
agreement with those obtained from FFT analysis (figure 14a), with a relative error of 0.05 %. However, the
characteristic frequencies of DMD modes are not in complete agreement with those of mean flow and SFD base
flow, with errors of 3.03 % and 4.81 %, respectively. Additionally, the topological structure of DMD modes is
nearly identical to that of mean flow global modes, as compared with the SFD base flow. The corresponding
eigenvalues are (a) mode 0, σ + iω = 3.7369 × 10−8 + i0.0; (b) mode 1, σ + iω = −5.2874 × 10−6 + i0.1976;
(c) mode 2, σ + iω = −1.9908 × 10−5 + i0.3953; (d) mode 3, σ + iω = −3.8395 × 10−5 + i0.5929;
(e) modeLB , (σ + iω)BF = 9.3466 × 10−2 + i0.2077; ( f ) modeLA , (σ + iω)BF = 4.6550 × 10−2 + i0.1451;
(g) modeLB , (σ + iω)MF = 5.7977 × 10−3 + i0.1917 and (h) modeLA , (σ + iω)MF = 6.8022 × 10−3 + i0.1481.
2
0.5 0.2 2 0.5 0.2
1 1
St
0
St
0 0 0 0
frequencies and the shapes of the linear global modes and DMD modes, we can establish
a connection between the linear and nonlinear systems.
3.4. Effect of AR
In this last section we discuss the effect of AR on the global modes in a short rotating
cylinder wake flow, as this information seems to be scarce in the literature on the short
(rotating) cylinder flows.
Figure 21 shows the neutral stability curves in panel (a) and the shedding frequency
in panel (b) for the case of AR = 0.75. The results of the flow past a sphere (Citro et al.
2016) are also shown for a comparison. Similar to the AR = 1 results in figures 8 and 9,
the low and high rotation flows at AR = 0.75 present dissimilar behaviours. In the case of
975 A15-24
Numerical analyses of flow past a short rotating cylinder
(a) (b)
–1 –1
0 0
1 1
–2 0 5 10 15 20 –2 0 5 10 15 20
(c) (d)
2 2
(e) (f)
2 2
Figure 19. Comparison of the DMD modes (a,b) with the global SFD base flow modes (c,d) and mean
flow modes (e, f ) at point HP2. All figures coloured by streamwise vorticity. The eigenfrequencies of the
DMD modes are in good agreement with those obtained from FFT analysis (figure 14d), with a relative
error of 0.05 %. Because the point HP2 is very close to the neutral curve, the eigenvalues of both mean
flow and SFD base flow can predict the true frequency well, with errors of 0.2 % and 0.8 %, respectively.
The corresponding eigenvalues are (a) mode 1, σ + iω = −9.660 × 10−7 + i0.03994; (b) mode 2, σ + iω =
7.360 × 10−6 + i0.1963; (c) modeHA , (σ + iω)BF = −7.499 × 10−5 + i0.04026; (d) modeHB , (σ + iω)BF =
1.905 × 10−3 + i0.1968; (e) modeHA , (σ + iω)MF = 7.740 × 10−4 + i0.04004 and ( f ) modeHB , (σ + iω)MF =
−3.019 × 10−3 + i0.1964.
(a) (b)
–1 –1
0 0
1 1
–2 0 5 10 15 20 –2 0 5 10 15 20
(c) (d)
2 2
(e) (f)
2 2
Figure 20. Comparison of the DMD modes (a,b) with the global SFD base flow modes (c,d) and mean
flow modes (e, f ) at point HP3. All figures coloured by streamwise vorticity. The eigenfrequencies of
the DMD modes are in good agreement with those obtained from FFT analysis (figure 14e), with a
relative error of 0.7 %. However, the frequencies of DMD modes are not in complete agreement with
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
those of mean flow and SFD base flow, with errors of 5.8 % and 6.6 %, respectively. Additionally, the
topological structure of DMD modes is nearly identical to that of mean flow global modes, as compared
with the SFD base flow. The corresponding eigenvalues are (a) mode 1, σ + iω = 3.307 × 10−6 + i0.06350;
(b) mode 2, σ + iω = −1.165 × 10−6 + i0.2686; (c) modeHA , (σ + iω)BF = −2.261 × 10−2 + i0.06755;
(d) modeHC , (σ + iω)BF = 3.050 × 10−2 + i0.2646; (e) modeHA , (σ + iω)MF = −5.764 × 10−2 + i0.06698
and ( f ) modeHC , (σ + iω)MF = 1.673 × 10−2 + i0.2707.
low rotation speeds, the two modes LA and LB undergo Hopf bifurcation due to linear
instability around Re = 330, 340, respectively. The linear unstable region of mode LB is
very small as shown in panel 21(a), almost completely inside the unstable region of mode
LA. In the case of high rotation speeds, four global unstable modes are identified up to
α = 2, successively experiencing Hopf bifurcation, resulting in three turning points (TP2
to TP4). All these modes are characterized by different frequencies as shown in panel (b).
The frequencies in the low-rotation-rate flows are close to the frequencies of the slowly
rotating sphere, whereas the frequencies of the high-rotation-rate modes are smaller than
those of the corresponding sphere. The structures of the global modes at low and high
rotation rates are shown in figures 25, 26 and discussed in Appendix B.
975 A15-25
Y. Yang, C. Wang, R. Guo and M. Zhang
Sphere (Citro et al. 2016) Mode HA
Sphere (Citro et al. 2016) Mode HB
Mode LA Mode HC
Mode LB Mode HD
(a) 2.0 (b) 2.0
1.8 1.8
TP4
1.6 1.6
1.4 1.4
TP3
1.2 1.2
α 1.0 1.0
TP2
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 TP1
150 200 250 300 350 400 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Re St0
Figure 21. Neutral stability curves and the corresponding frequencies for the flow past a short rotating
cylinder at AR = 0.75, compared with a vertically rotating sphere (Citro et al. 2016).
0.2
0
50 100 150 200 250 300 350 400 0 0.1 0.2 0.3 0.4 0.5
Re St0
Figure 22. Neutral stability curves and the corresponding frequencies for the flow past a short rotating
cylinder at AR = 2, compared with a vertically rotating sphere (Citro et al. 2016).
The results of AR = 2 are shown in figure 22. In this case, the calculation seems to be
more difficult to converge. We will interpret the results with caution. The shapes of the
corresponding global modes are shown in figure 27 in Appendix B. When the rotation
rate is small in this case, we can identify an unstable mode appearing similarly to the LD
mode in AR = 1; this mode will be similarly called LD. Further increasing α, the shape
of the most unstable global mode changes, see the colour transition from orange to purple
in panels 22(a,b) and the comparison between figures 27(a) and 27(b). In the higher α
regime, the most unstable global mode changes abruptly, denoted by green and blue lines in
figure 22. We tried to converge as many unstable modes as we could, but some calculations
975 A15-26
Numerical analyses of flow past a short rotating cylinder
2.0 Mode I of 2-D cylinder
(Pralits et al. 2010)
1.8 Mode A of 3-D cylinder
(Rao et al. 2015)
1.6 Mode B of 3-D cylinder
(Rao et al. 2015)
Mode I of sphere
1.4 (Citro et al. 2016)
Mode II of sphere
1.2 (Citro et al. 2016)
Mode RW1 of sphere
α 1.0 (Sierra-Ausín et al. 2022)
Mode RW2 of sphere
0.8 (Sierra-Ausín et al. 2022)
Mode RW3 of sphere
(Sierra-Ausín et al. 2022)
0.6 Bullet-shaped body
(Jiménez-González et al. 2014)
0.4 AR = 0.75 at low α
AR = 0.75 at high α
0.2 AR = 1 at low α
AR = 1 at high α
AR = 2
0
50 100 150 200 250 300 350 400
Re
Figure 23. Neutral stability curves of Hopf bifurcation for the flow past a rotating finite cylinder at AR =
0.75, 1, 2. Comparison with streamwise rotating infinite cylinder (Pralits et al. 2010; Rao et al. 2015),
sphere (Citro et al. 2016), and streamwise rotating sphere (Sierra-Ausín et al. 2022) and bullet-like body
(Jiménez-González et al. 2014).
were not converged. Thus, we will not go into details for the high-rotation-rate cases in
AR = 2. Such difficulty in converging the 3-D wake flow is not uncommon, reflecting the
complex nature of these flows and highlighting more research efforts to decipher their
dynamics.
In the end, we consolidate and compare all the significant results in this work spanning
a large parameter space with Re ∈ [100, 500], α ∈ [0, 2] for AR = 0.75, 1, 2 in figure 23.
To place our results in a more general context, we also compare our results with other
rotating bluff-body flows such as 2-D rotating cylinders (Pralits et al. 2010; Rao et al.
2015), spheres (Citro et al. 2016; Sierra-Ausín et al. 2022) and a bullet-like body
(Jiménez-González et al. 2014). It can be seen from figure 23 that in the low-rotation-rate
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
regime, larger AR renders the flow more unstable as the critical Re decreases from
Rec = 340 for AR = 0.75 to Rec = 50 for AR = ∞ (2-D case, Pralits et al. 2010). In
the high-rotation-rate end (with the maximum rotation rate being α = 2 in our work),
it is difficult to summarise a trend of increasing AR from our finite-length cylinders to
the infinitely long cylinder. As discussed above, the computations in this regime are more
difficult, calling for more research efforts to elucidate the difference. The 3-D instability in
the infinitely long cylinder (Rao et al. 2015) presents a smoother transition in the range of
α ∈ [0, 2], different from our 3-D results where distinct behaviours in the low- (α 0.6)
and high-rotation-rate (0.6 α 2) cases can be identified. This is because α = 2 is not
a high rotation rate for infinitely long cylinder flows; in figure 10 of Rao et al. (2015),
dissimilar wake behaviours in this flow are separated by α ≈ 2.5. The dynamics of the
rotating sphere wake flow along the transverse direction (Citro et al. 2016) is similar to that
of the short rotating cylinder with AR slightly larger than 1, whereas the rotating sphere
wake along the streamwise direction (Sierra-Ausín et al. 2022) looks more dissimilar than
ours. This is likely because our cylinder is also rotating along a transverse axis. In the
end, the flow past a spinning bullet-shaped bluff body (Jiménez-González et al. 2014) is
also shown for a comparison and its low-rotation-rate behaviour appears similarly to our
975 A15-27
Y. Yang, C. Wang, R. Guo and M. Zhang
flow. By studying the effect of AR, we can qualitatively connect our results with those for
the sphere, the cylinders of infinite length, etc.and explain the difference between these
benchmark flows in a large parameter space.
4. Conclusions
In this work a 3-D flow stability problem past a short rotating cylinder has been studied.
The motivation for considering this flow configuration is due to its applications in various
engineering settings and its relevance to flow control strategy by rotation. New flow modes
have been identified in our DNS and global stability analyses of this flow. The linear results
have also been compared with the nonlinear results to find their traces in real flows. The
wavemaker region responsible for the instability generation has been delimited. We have
also studied the effect of aspect ratio AR to understand how the 3-D flows change with its
geometry parameters.
Firstly, for a cylinder with AR = 1, when the rotation rate α is slower than 0.3, the
rotation effect only trivially affects the flow past a short cylinder. The two unstable global
modes (see figure 8a) in the low-rotation-rate cases resemble those in the non-rotating
flows, corresponding to the vortices shedding from the flat ends of the cylinder and the
vortices shedding from the curved surface, respectively. The rotating effect swings the
recirculation region towards the rotating direction and, in general, decreases the separation
bubble length, defined in our work. Besides, the rotation also casts its effect on the flow
instability and bifurcation. For example, the rotation strengthens the symmetry breaking
caused by the inherent wake mechanism observed in the regular bifurcation without
rotation, leading to a similar asymmetric recirculation region in the non-rotating flows
generated by the regular bifurcation. We have also investigated the lift and drag coefficients
in the short rotating cylinder flow. Larger rotation rates both increase the absolute value
of the drag coefficient and lift coefficient in the transverse direction. An interesting
correspondence of the lift coefficients between the short rotating cylinder and the rotating
sphere with a doubled rotation rate is observed when Re is relatively large.
The global stability analyses reveal that the parameter space α can be divided into low
and high regimes when Re is relatively low < 500. When the rotation rate is smaller than
approximately 0.3, two unstable global modes exist with non-zero frequencies undergoing
Hopf bifurcation, whose interaction and competition giving rise to a codimension-two
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
transition state where the two Hopf bifurcations can occur simultaneously. The critical Re
at a certain α slightly decreases and then obviously increase with increasing α. When the
rotation rate is large, more unstable modes are observed experiencing Hopf bifurcations.
The critical Re in this case decreases with increasing α, highlighting the different effects
of Re in the two rotation regimes. The eigenvectors as well as their superposition with
the corresponding adjoint modes have also been probed. Especially, we observed that the
sensitivity region of mode HB (with a high rotation rate) is distinguished from other modes
close to the cylinder, indicating that its control may be achieved separately.
The comparison of the linear and nonlinear results aims to attach more physical
significance to the linear analyses. The traces of the global modes are identified in the
nonlinear simulations by comparing their frequencies (i.e. eigenfrequency in the linear
analysis and the shedding frequency in the nonlinear DNS). The DMD method is employed
to conduct the comparison with the global stability analysis based on the steady flow and
the time-mean flow (averaged over several oscillating periods). In general, we can find
better correspondence of the time-mean flow results with the DNS results, simply because
both analyses were applied to the nonlinear saturated oscillation. In term of the phase
diagram, the low-rotation-rate cases characterize limit cycles whereas high-rotation-rate
975 A15-28
Numerical analyses of flow past a short rotating cylinder
cases present an increasing degree of complexity, encompassing limit cycles, limit torus
and chaos, reflecting more complex flow structures and dynamics when the cylinder rotates
faster.
Then, the effect of the aspect ratio AR has also been investigated. The aim of this
investigation is to compare our flow configuration with other bluff-body wake dynamics in
a large parameter space. Even though more data points are needed, we can find the trend
of how increasing AR renders the short rotating flow more unstable in the low-rotation-rate
cases α < 0.6. When the rotation rate is large but less than α = 2, the critical Re decreases
with increasing α. Compared with other bluff-body dynamics, we found that the dynamics
of the rotating sphere wake flow along the transverse direction is similar to that of the
short rotating cylinder with AR slightly larger than 1, and the rotating sphere wake along
the streamwise direction differs more significantly than our results due to the different flow
configuration.
The current work focuses on the first instability in the short rotating cylinder wake
flow. The flow dynamics already presents a high degree of complexity. In order to
further understand the underlying mechanism, as a future direction, the subsequent flow
bifurcations in this flow can be studied in detail by employing the global linear stability
and weakly nonlinear stability analyses.
Acknowledgements. The simulations were performed at the National Supercomputing Centre, Singapore
(NSCC). The authors would like to thank Ms X. He and Mr D. Wan for insightful discussions.
Funding. We acknowledge the financial support of a Tier 1 grant from the Ministry of Education, Singapore
(WBS No. A-8001172-00-00). Y.L. is supported by the National Natural Science Foundation of China (No.
12202200) and China Postdoctoral Science Foundation (No. 2022M711641).
Author ORCIDs.
Yongliang Yang https://orcid.org/0000-0002-4965-8167;
Chenglei Wang https://orcid.org/0000-0003-0771-8667;
Mengqi Zhang https://orcid.org/0000-0002-8354-7129.
The calculation of the Landau coefficient is presented in this appendix. The growth rate
(c1 = 1.4473 × 10−2 ) obtained by nonlinear DNS is in good agreement with the growth
rate (1.4642 × 10−2 in table 1) of SFD base flow LSA. The coefficient c3 = −3.673
indicates that the Hopf bifurcation caused by mode LA is supercritical. This is the same
as the bifurcation property of the non-rotating cylinder (Yang et al. 2021), also caused
by mode LA. Now, based on the Stuart–Landau equation dAm /dt = c1 Am + c3 Am |Am |2
(where Am can be viewed as the amplitude of Clz ) or, equivalently, d(ln |Am |)/dt =
c1 + c3 |Am |2 , the Landau coefficients c1 and c3 can be calculated by plotting d(ln |Am |)/dt
vs |Am |2 (Thompson, Leweke & Provansal 2001; Sheard, Thompson & Hourigan 2004),
as shown in panel (c). Therefore, the transverse intercept point gives an estimation of
c1 = 1.4473 × 10−2 and the gradient near this point is an approximation of c3 = −3.673.
dln|A|/dt
2
ln(A)
Slope ≈ –3.673
Clz
0 –8
–2 –9 5
–4 –10
–6
–11
0 200 400 600 800 0 200 400 600 800 0 1 2 3 4
Time Time |A|2 (×10–5)
Figure 24. Landau coefficients c1 = 1.4473 × 10−2 and c3 = −3.673 computed by the nonlinear DNS at
point LP1. The blue circular markers in panel (a) represent the amplitude A of Clz .
(a) (b)
1 1
–1 –1
1 1
–1 –1 30 35 40
5 10 15 20 25 30 35 40 0 5 10 15 20 25
0
Figure 25. The global mode LA (a) and LD (b) for the flow past a rotating finite cylinder at (AR = 0.75,
Re = 365, α = 0.05). The Q = 1 × 10−6 isosurfaces are coloured by the streamwise vorticity ranging from
−2 × 10−3 to 2 × 10−3 . The corresponding eigenvalues are (a) LA mode λLA = 4.062 × 10−2 + i0.1749 and
(b) LD mode λLD = −5.222 × 10−3 + i0.1431.
(a) (b)
1 1
–1 –1
4 4
0 40 45 50 0 35 40 45 50
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30
(c) (d )
1 1
–1 –1
4 4
0 30 35 40 45 50 0 45 50
0 5 10 15 20 25 0 5 10 15 20 25 30 35 40
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
Figure 26. The global modes HA1 (a), HA2 (b), HB (c) and HC (d) for the flow past a rotating finite cylinder
at (AR = 0.75, Re = 330, α = 1.5). The Q = 1 × 10−6 isosurfaces are coloured by the streamwise vorticity
ranging from −2 × 10−3 to 2 × 10−3 . The corresponding eigenvalues are (a) HA1 mode λHA1 = 6.177 ×
10−2 + i0.07840; (b) HA2 mode λHA2 = 9.240 × 10−3 + i0.08931; (c) HB mode λHB = 0.1078 + i0.1950;
(d) HC mode λHC = 0.1146 + i0.3012.
in figure 7(a,d). Figure 26 features the global modes HA1, HA2, HB and HC. Again,
comparison can be made to the global modes for the AR = 1 flow in figure 10 at a similar
rotation rate.
The structures of some selected global linear modes for AR = 2 are displayed in
figure 27. Panel (a) shows an unstable mode at a low rotation rate α = 0.1. It is called
the LD mode because the wake structure looks very similar to the LD mode in figure 7
for AR = 1, Re = 290, α = 0.1. Under the action of enhanced speed ratio, mode LB in
panel (b) can be regarded as the result of mode LD losing the vortex shedding from
the cylinder’s arc surface that rotates along the streamwise direction. It closely resembles
mode HB in figure 10 for AR = 1, Re = 170, α = 1.8 at a high rotation rate. Mode HB1 in
975 A15-30
Numerical analyses of flow past a short rotating cylinder
(a) (b)
0 0
3
3
–1 0
25 30 35 40 45 50 20 25 30 35 40 45 50
0 5 10 15 20 0 5 10 15
(c) (d )
2
1 0
–1 –2
5 8
4
0 40 45 50 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35 40 45 50
Figure 27. Selected global modes (SFD base flow) for AR = 2. (a) Mode LD (Re = 150, α = 0.1, which is
similar to mode LD in figure 7d). (b) Mode LB (panel (b) at Re = 130 and α = 0.4), mode HB1 (panel (c) at
Re = 310 and α = 0.6) and mode HB2 (panel d at Re = 190 and α = 1.5). The Q = 1 × 10−7 isosurfaces are
coloured by streamwise vorticity ranging from −0.02 to 0.02. The corresponding eigenvalues are (a) LD mode
λLD = 1.105 × 10−2 + i0.1255, (b) LB mode λLB = 9.314 × 10−3 + i0.1470, (c) HB1 mode λHB1 = 2.899 ×
10−2 + i0.2720 and (d) HB2 mode λHB2 = 2.028 × 10−2 + i0.1562.
(a) (b)
2.4 8
AR = 1, Re = 80
2.2 AR = 1, Re = 160 Potential theory
AR = 1, Re = 200 7
2.0 AR = 1, Re = 300
1.8 6
1.6 5
1.4
l|
d
4
|C
C
0.6
Cdv(AR = 1, Re = 160)
1
0.4 |(AR = 1, Re = 160)
|Clv
0.2
0 0.5 1.0 1.5 2.0 2.5 0 0.5 1.0 1.5 2.0 2.5
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
(c) (d ) (e)
0.4 0.20
0.12
0.3 0.10 0.15
0.08
Re = 60, Kang and Choi (1999)
d
l
AC
St
AC
0 0.5 1.0 1.5 2.0 2.5 0 0.5 1.0 1.5 2.0 2.5 0 0.5 1.0 1.5 2.0 2.5
α α α
Figure 28. Validations by comparing the lift and drag coefficients of a 2-D rotating cylinder flow between the
present DNS code results with those in Kang et al. (1999) and Stojković et al. (2002). (a) Time-averaged drag
coefficient; (b) time-averaged lift coefficient; (c) amplitude of Cl ; (d) amplitude of Cd ; (e) Strouhal number
calculated using the Cl signal. Besides, the drag and lift coefficients of steady SFD base flow are added in
panels (a,b) for comparative analysis, and decompose them into C̄l = C̄lp + C̄lv , C̄d = C̄dp + C̄dv for the case
(AR = 1, Re = 160).
975 A15-31
Y. Yang, C. Wang, R. Guo and M. Zhang
(a) (c)
0.14 0.14
0.12 0.12
0.10 0.10
0.08 0.08
(b) 0.06 (d ) 0.06
0.04 0.04
0.02 0.02
0 0
0 2 4 6 –2 0 2 4 6
Figure 29. Comparison of the wavemaker region ζ between the results generated by the present code (panels
a,c) and those in Marquet et al. (2008) (panel b), Giannetti & Luchini (2007) (panel d) for 2-D non-rotating
cylinder flows at Re = 46.8 (panels a,b) and Re = 50 (panels c,d).
Table 3. A grid sensitivity test for the nonlinear DNS case Re = 290, α = 1.2, AR = 1, Δt = 10−3 . As shown
in figure 1, La is the length from surfaces Sin , Sxz and Sxy to the cylinder centre; Lo is the length from surface
Sout to the cylinder centre. Here Ntot. is the total number of hexahedral elements inside the computational
domain; Nord. is the polynomial order of each hexahedral element.
panel c exhibits similar vortex structures to mode LB, but with smaller vortices, resulting
in higher oscillation frequencies. Mode HB2 in panel d can be regarded as the result of a
stronger lateral deflection of mode LB due to the stronger rotation.
text as qualitative proof of the accuracy of our codes. To further test the accuracy in a
quantitative manner, we compare our simulated results of the lift and drag coefficients
in the 2-D rotating cylinder flow with those in Kang et al. (1999) and Stojković et al.
(2002). Figure 28 displays a good comparison of our results with theirs at Re = 60 and
100, indicating the accuracy of the used nonlinear numerical code in the current work. For
the verification of the linear code, figure 29 presents the wavemaker region in the classical
2-D cylindrical wake flow in Giannetti & Luchini (2007) and Marquet et al. (2008). We
can see a very good comparison is achieved between our results (top) and theirs (bottom).
This good comparison entails the linear code solving correctly both the global modes and
the adjoint modes.
In the end, we furnish a test study on the size of the computational domain. In our
previous work on the 3-D non-rotating short cylinder (Yang et al. 2021), we converged a
suitable computational domain balancing the computational efficiency and accuracy. With
the rotation effect, we found that the size of the computational domain should increase to
accommodate the swinging effect brought by the rotation. The result is shown in table 3
for different values of La and Lo ; see their definitions in figure 1. The symmetry boundary
975 A15-32
Numerical analyses of flow past a short rotating cylinder
condition is imposed on the surfaces Sxy,t , Sxy,b , Sxz,f and Sxz,b , where the position is set to
a range of ±10D to ±15D, that is, La = 10 − 15.
As shown in table 3, unlike the non-rotating cylinder (Cadieux, Sun & Domaradzki
2017), the drag coefficient is more sensitive to the domain size than St number. The relative
errors of the drag, lift coefficients and the frequencies from mesh M2 to M4 are less
than 1 % compared with the reference results of mesh M5. Thus, in order to balance the
efficiency and accuracy, mesh M2 and the order Nord. = 7 are adopted in the present work
to compute all computational instances.
R EFERENCES
A FROZ , F., L ANG , A. & J ONES , E. 2017 Use of a rotating cylinder to induce laminar and turbulent separation
over a flat plate. Fluid Dyn. Res. 49 (3), 035509.
A KERVIK , E., B RANDT, L., H ENNINGSON , D.S., H ŒPFFNER , J., M ARXEN , O. & S CHLATTER , P. 2006
Steady solutions of the Navier–Stokes equations by selective frequency damping. Phys. Fluids 18 (6),
068102.
BAGHERI , S. 2013 Koopman-mode decomposition of the cylinder wake. J. Fluid Mech. 726, 596–623.
BAGHERI , S. 2014 Effects of weak noise on oscillating flows: linking quality factor, Floquet modes, and
Koopman spectrum. Phys. Fluids 26 (9), 094104.
C ADIEUX , F., S UN , G. & D OMARADZKI , J.A. 2017 Effects of numerical dissipation on the interpretation of
simulation results in computational fluid dynamics. Comput. Fluids 154, 256–272, iCCFD8.
DE C ELIS , R., C ADARSO , L. & S ÁNCHEZ , J. 2017 Guidance and control for high dynamic rotating artillery
rockets. Aerosp. Sci. Technol. 64, 204–212.
C ITRO , V., T CHOUFAG , J., FABRE , D., G IANNETTI , F. & LUCHINI , P. 2016 Linear stability and weakly
nonlinear analysis of the flow past rotating spheres. J. Fluid Mech. 807, 62–86.
D OEDEL , E. & T UCKERMAN , L.S. 2012 Numerical Methods for Bifurcation Problems And Large-Scale
Dynamical Systems, vol. 119. Springer Science & Business Media.
E L A KOURY, R., B RAZA , M., P ERRIN , R., H ARRAN , G. & H OARAU, Y. 2008 The three-dimensional
transition in the flow around a rotating cylinder. J. Fluid Mech. 607, 1–11.
FABRE , D., T CHOUFAG , J., C ITRO , V., G IANNETTI , F. & LUCHINI , P. 2017 The flow past a freely rotating
sphere. Theor. Comput. Fluid Dyn. 31, 475–482.
FABRE , D., T CHOUFAG , J. & M AGNAUDET, J. 2012 The steady oblique path of buoyancy-driven disks and
spheres. J. Fluid Mech. 707, 24–36.
F ISCHER , P., K ERKEMEIER , S. & P EPLINSKI , A. 2020 Nek5000 home page. Website, https://nek5000.mcs.
anl.gov/.
GAD - EL H AK , M. & B USHNELL , D.M. 1991 Separation control: review. Trans. ASME J. Fluids Engng
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
975 A15-33
Y. Yang, C. Wang, R. Guo and M. Zhang
L EHOUCQ , R.B., S ORENSEN , D.C. & YANG , C. 1998 ARPACK Users’ Guide: Solution of Large-Scale
Eigenvalue Problems with Implicitly Restarted Arnoldi Methods. SIAM.
L INH , D.T.T. 2011 Flow past a rotating circular cylinder. PhD thesis, Department of Mechanical Engineering,
National University of Singapore.
L ORITE -D ÍEZ , M. & J IMÉNEZ -G ONZÁLEZ , J.I. 2020 Description of the transitional wake behind a strongly
streamwise rotating sphere. J. Fluid Mech. 896, A18.
LUCHINI , P. & B OTTARO , A. 2014 Adjoint equations in stability analysis. Annu. Rev. Fluid Mech. 46, 493–517.
M ARQUET, O., S IPP, D. & JACQUIN , L. 2008 Sensitivity analysis and passive control of cylinder flow. J. Fluid
Mech. 615, 221–252.
M ITTAL , S. 2004 Three-dimensional instabilities in flow past a rotating cylinder. J. Appl. Mech. 71 (1), 89–95.
M ITTAL , S. & K UMAR , B. 2003 Flow past a rotating cylinder. J. Fluid Mech. 476, 303–334.
M ODI , V.J. 1997 Moving surface boundary-layer control: a review. J. Fluids Struct. 11 (6), 627–663.
NAVROSE , M EENA , J. & M ITTAL , S. 2015 Three-dimensional flow past a rotating cylinder. J. Fluid Mech.
766, 28–53.
PATERA , A.T. 1984 A spectral element method for fluid dynamics: laminar flow in a channel expansion.
J. Comput. Phys. 54 (3), 468–488.
P RALITS , J.O., B RANDT, L. & G IANNETTI , F. 2010 Instability and sensitivity of the flow around a rotating
circular cylinder. J. Fluid Mech. 650, 513–536.
P RALITS , J.O., G IANNETTI , F. & B RANDT, L. 2013 Three-dimensional instability of the flow around a
rotating circular cylinder. J. Fluid Mech. 730, 5–18.
R ADI , A., T HOMPSON , M.C., R AO , A, H OURIGAN , K. & S HERIDAN , J. 2013 Experimental evidence of
new three-dimensional modes in the wake of a rotating cylinder. J. Fluid Mech. 734, 567–594.
R ADKE , R.J. 1996 A Matlab implementation of the implicitly restarted arnoldi method for solving large-scale
eigenvalue problems. PhD thesis, Rice University.
R AO , A., L EONTINI , J., T HOMPSON , M.C. & H OURIGAN , K. 2013a Three-dimensionality in the wake of a
rotating cylinder in a uniform flow. J. Fluid Mech. 717, 1–29.
R AO , A., L EONTINI , J.S., T HOMPSON , M.C. & H OURIGAN , K. 2013b Three-dimensionality in the wake of
a rapidly rotating cylinder in uniform flow. J. Fluid Mech. 730, 379–391.
R AO , A., R ADI , A., L EONTINI , J.S., T HOMPSON , M.C., S HERIDAN , J. & H OURIGAN , K. 2015 A review
of rotating cylinder wake transitions. J. Fluids Struct. 53, 2–14.
ROSLAN , R., SALEH , H. & H ASHIM , I. 2012 Effect of rotating cylinder on heat transfer in a square enclosure
filled with nanofluids. Intl J. Heat Mass Transfer 55 (23), 7247–7256.
ROWLEY, C.W., M EZI Ć , I., BAGHERI , S., S CHLATTER , P. & H ENNINGSON , D.S. 2009 Spectral analysis of
nonlinear flows. J. Fluid Mech. 641, 115–127.
S CHLICHTING , H. & G ERSTEN , K. 2016 Boundary-Layer Theory. Springer.
S CHMID , P.J. 2010 Dynamic mode decomposition of numerical and experimental data. J. Fluid Mech.
656, 5–28.
S CHMID , P.J. 2022 Dynamic mode decomposition and its variants. Annu. Rev. Fluid Mech. 54, 225–254.
S EIFERT, J. 2012 A review of the Magnus effect in aeronautics. Prog. Aerosp. Sci. 55, 17–45.
S HEARD , G.J., T HOMPSON , M.C. & H OURIGAN , K. 2004 From spheres to circular cylinders:
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
non-axisymmetric transitions in the flow past rings. J. Fluid Mech. 506, 45–78.
S IERRA , J., FABRE , D., C ITRO , V. & G IANNETTI , F. 2020 Bifurcation scenario in the two-dimensional
laminar flow past a rotating cylinder. J. Fluid Mech. 905, A2.
S IERRA-AUSÍN , J., L ORITE -D ÍEZ , M., J IMÉNEZ -G ONZÁLEZ , J.I., C ITRO , V. & FABRE , D. 2022 Unveiling
the competitive role of global modes in the pattern formation of rotating sphere flows. J. Fluid Mech.
942, A54.
S TOJKOVI Ć , D., B REUER , M. & D URST, F. 2002 Effect of high rotation rates on the laminar flow around a
circular cylinder. Phys. Fluids 14 (9), 3160–3178.
S TOJKOVI Ć , D., S CHÖN , P., B REUER , M. & D URST, F. 2003 On the new vortex shedding mode past a
rotating circular cylinder. Phys. Fluids 15 (5), 1257–1260.
T ENNANT, J.S., J OHNSON , W.S. & K ROTHAPALLI , A. 1976 Rotating cylinder for circulation control on an
airfoil. J. Hydronaut. 10 (3), 102–105.
T HEOFILIS , V. 2011 Global linear instability. Annu. Rev. Fluid Mech. 43 (1), 319–352.
T HOMPSON , M.C., L EWEKE , T. & P ROVANSAL , M. 2001 Kinematics and dynamics of sphere wake
transition. J. Fluids Struct. 15 (3-4), 575–585.
W ILLIAMSON , C.H.K. 1996a Three-dimensional wake transition. J. Fluid Mech. 328, 345–407.
W ILLIAMSON , C.H.K. 1996b Vortex dynamics in the cylinder wake. Annu. Rev. Fluid Mech. 28 (1), 477–539.
YANG , Y., F ENG , Z. & Z HANG , M. 2022 Onset of vortex shedding around a short cylinder. J. Fluid Mech.
933, A7.
975 A15-34
Numerical analyses of flow past a short rotating cylinder
YANG , Y., G UO , R., L IU, R., C HEN , L., X ING , B. & Z HAO , B. 2021 Quasi-steady aerodynamic
characteristics of terminal sensitive bullets with short cylindrical portion. Def. Technol. 17 (2), 633–649.
Z HAO , M. & Z HANG , Q. 2023 Three-dimensional numerical simulation of flow past a rotating step cylinder.
J. Fluid Mech. 962, A45.
https://doi.org/10.1017/jfm.2023.840 Published online by Cambridge University Press
975 A15-35