You are on page 1of 17

Yuelong Yu

School of Mechanical Engineering,


Gas Turbine Research Institute,
Shanghai Jiao Tong University,
A Review on Fluid-Induced
800 Dongchuan Road,
Shanghai 200240, China
Flag Vibrations
e-mail: yuyuelong@sjtu.edu.cn
Fluid-induced flag vibrations provide unattended, efficient, low-cost, and scalable solu-
Yingzheng Liu1 tions for energy harvesting to power distributed wireless sensor nodes, heat transfer
School of Mechanical Engineering, enhancement in channel flow, and mixing enhancement in process industries. This review
Gas Turbine Research Institute, surveys three generic configurations, the inverted flag, the standard flag, and the forced
Shanghai Jiao Tong University, flag, i.e., an inverted or standard flag located downstream of a bluff body. Their instabil-
800 Dongchuan Road, ity boundaries, vibration dynamics, and vortex dynamics are compared in a unified
Shanghai 200240, China framework to elucidate their common and distinct features and provide insights into
e-mail: yzliu@sjtu.edu.cn the design of vibrating flags for various applications. Some common features are also
identified and analyzed for describing the interaction between multiple flags, three-
Xavier Amandolese dimensional (3D) effects, and Reynolds number effects. The suggestions are intended to
LadHyX, guide future research directions. [DOI: 10.1115/1.4042446]
CNRS-Ecole Polytechnique,
Palaiseau, F-91128, France; Keywords: fluid-induced vibration, flag, vortex dynamics, energy harvesting, mixing
Conservatoire National des Arts et Metiers,
enhancement
Paris F-75141, France
e-mail: xavier.amandolese@ladhyx.polytechnique.fr

1 Introduction configurations share the same governing equations and parameters


but have different boundary conditions. (a) The inverted flag is
Fluid-induced vibrations of flexible flags are of continuing and
clamped at its trailing edge, and its leading edge is free to vibrate.
intense interest in nature and industry applications. The traditional
The flapping dynamics of the inverted flag were first investigated
motivations focused on mitigating fluid-induced vibrations, and flex-
by Kim et al. [19]. The periodic formation and shedding of the
ible flags interacting with axial flow often served as basic models to
leading edge vortex (LEV) and trailing edge vortex (TEV) from
study the flutter of paper during printing presses [1], snoring induced
the deformed inverted flag alter the fluid force on the inverted
by the soft palate in the human upper airway [2], panel flutter in aer-
flag, which leads to large-amplitude periodic vibrations, provided
onautics [3], and nuclear engineering [4]. However, an active area of
the fluid force is well coupled with the bending force of the
research is the utilization of fluid-induced flag vibrations. By explor-
inverted flag. This typical vortex-induced vibration can be
ing their nonlinear motions with large vibration amplitudes and
observed for a light inverted flag, lock-in with the vortex-
deflections, flexible flags of different configurations [5–7] have been
shedding frequency [20] or its higher harmonic fluid response
widely developed to harvest energy from wind or ocean flow in
[21,22], while for a heavy inverted flag, the vortex shedding can
order to power unattended sensor nodes since the early work of
be responsible for the vibration around the flag’s natural fre-
Allen and Smits [8] and Taylor et al. [9], or serve as autonomous
quency [21]. In that context, the fluid-induced vibration of the
flow sensors simultaneously [10]. Abundant vortices that are gener-
inverted flag can be often associated with the vortex shedding
ated by the interaction between vibrating flags with the channel flow
excitation mechanism and thus to an IIE. Under this scenario, the
are favorable for enhancing heat transfer in electronics [11,12] or
nature of vibration of the inverted flag is self-controlled [23]. It
enhancing mixing in massive process industries [13]. The transfer
should be noted that the vortex-induced vibration is the most stud-
enhancement based on fluid-induced flag vibration has been investi-
ied scenario for the inverted flag; however, vortex shedding is not
gated for the past 5 years. Fluid-induced vibrations in flexible flags
significant for small-deflection flapping [21]. (b) The standard flag
serve as renewable, sustainable, and economical solutions for energy
is fixed at its leading edge and is free at its trailing edge. The ear-
harvesting or transfer enhancement. Moreover, knowledge into
liest study on instability and flapping dynamics of the standard
fluid-induced flag vibrations also helps in understanding the flapping
flag goes back to Taneda [24], and extensive literature has been
locomotion of aquatic animals and birds, these problems both
produced to address the instability mechanism since the work of
involve an unsteady balance between inertial forces, body elasticity,
Zhang et al. [25]. The instability in the standard flag occurs due to
and dynamic pressure of fluid flow [14,15], as well as the interaction
the positive feedback of flag’s inertial force with its elastic force
of vortices with a dynamically deformed body [16].
and fluid force. The movement of the standard flag increases the
Fluid-induced vibrations are generally classified into three
aerodynamic load, which in turn drives the standard flag to
categories in terms of the excitation sources, namely, instability-
increase its displacement until an optimal energy exchange with a
induced excitation (IIE), movement-induced excitation (MIE),
right timescale between the two domains is reached. This instabil-
and extraneously induced excitation (EIE) according to Nau-
ity is termed as flutter, and can be associated with the MIE mecha-
dascher and Rockwell [17] and Paidoussis [18]. With regard to the
nism. The nature of the vibration is self-excited vibration. (c) An
main dynamics that can be observed for the three representative
inverted or standard flag can also be forced into vibration by
configurations of flexible flags, namely, the inverted flag, the
upstream fluctuations in the flow velocity and pressure. This EIE
standard flag, and the flag behind a bluff body, the same classifica-
is generally induced by periodic vortices shedding from an
tion will be used here, as shown in Fig. 1. These three
upstream bluff body, such as circular cylinder [26], square cylin-
der [27], D-cylinder [28], and rectangular plate [7,8].
1 Abundant vibration dynamics and vortex dynamics have been
Corresponding author.
Manuscript received July 5, 2018; final manuscript received December 25, 2018; reported separately in the literature for numerous parameters
published online January 31, 2019. Editor: Harry Dankowicz. describing fluid-induced vibrations in flags. The important fluid

Applied Mechanics Reviews Copyright V


C 2019 by ASME JANUARY 2019, Vol. 71 / 010801-1
Fig. 1 Classification of flow-induced flag vibrations

flow parameters are the flow velocity, pressure, kinematic viscos- 2 Mathematical Formulation
ity, and fluid density. And the important structural parameters are
Incompressible viscous fluid flow is described by the
the bending stiffness, length, width, thickness, and flag density.
Navier–Stokes equations and the continuity equation
Coupled vibration dynamics and vortex dynamics are character-
ized by the vibration mode, vibration amplitude, vibration @u 1
frequency, vortex shedding frequency, evolution, and interaction þ ðu  rÞu ¼  ðrpÞ þ r2 u (1)
between vortices, and a dynamically vibrating flag. A unified @t qf
framework for summarizing the features of vibration dynamics
and vortex dynamics in terms of the three distinct mechanisms of ru¼0 (2)
instability and their dependencies on the system parameters is still
lacking. Previous surveys in this series [29,30] have been limited Assuming the flag is sufficiently thin, shear stresses involving the
to the instability in the standard flag. The aim of this review is to normal direction and rotational inertia about tangential directions
comprehensively study the various mechanisms of fluid-induced are ignored [31]. The centerline of the flag is considered to be
vibrations in flags. In particular, a unified framework is provided inextensible. The thickness and the bending rigidity of the flag
to present the instability and vibration modes of various flag are uniform. Accordingly, the relationship between the two-
configurations in the same parameter space, derive the laws dimensional (2D) deflection of a flag and the applied fluid load is
describing vibration dynamics and vortex dynamics and their rela- derived from Newton’s second law [32]
tionships, as well as elucidate their unique features. This frame-    
work sheds light on the design of fluid-induced vibrations in flags @4X @2X @ @X @X @ @4X
B 4 þ qs h 2  T þ þl ¼ Dpn
for applications such as energy harvesting, heat transfer enhance- @s @t @s @s @t @t @s4
ment, and mixing enhancement. (3)
The organization of this review is as follows: The dimension-
less parameters are first derived from the mathematical formula- where s is the curvilinear coordinate and the vector X ¼ (X(s, t),
tions and theoretical methods that address the coupled Y(s, t)) denotes the displacement of the flag. B is the bending
formulations are reviewed in Sec. 2. The instability boundary for rigidity of flag. qs and h are the density and thickness of flag,
different types of fluid-induced vibrations in the space of dimen- respectively. T is the tension force along the flag axis, which
sionless parameters and the dynamic vibrations in terms of fre- enforces inextensibility and is a function of s and t. The terms that
quency and vibration amplitude are analyzed in Secs. 3 and 4, are proportional to the fluid damping coefficient  and the internal
respectively. Section 5 summarizes the vortex dynamics and their damping coefficient l model dissipation due to the fluid viscosity
dependencies on the vibration dynamics. The effects of three- and the Kelvin–Voigt structural damping, respectively. Dp is the
dimensionality and Reynolds number are examined in Sec. 6. The pressure jump across the flag, which is imparted by the surround-
main interaction features between multiple flags are illustrated in ing fluid. n is the unit vector normal to the flag.
Sec. 7. Finally, the applications of fluid-induced flag vibrations The structural influence is imparted to the fluid through the
are addressed in Sec. 8. It is worth noting that the flag is employed velocity boundary condition on the flag boundary
in the review to refer to various terminologies in the list of refer-
ences, including plate, reed, filament, strip, membrane, beam,
sheet, shell, web, ribbon belt, and panel. A plate emphasizes a flag @X
u¼ (4)
with high bending stiffness, membrane refers to a flag with low @t
bending stiffness, strip or ribbon denotes slender flag with low
aspect ratio, filament means the span is zero, and reed refers to a The boundary condition of the free edge of the flag, i.e., the trail-
flag in channel flow. Nonetheless, these terminologies are inter- ing edge of the standard flag at s ¼ L and the leading edge of the
changeable, and flag is used to encompass all the models with one inverted flag at s ¼ 0, is given by
edge fixed and the other edges free to oscillate. In addition, IIE
and MIE henceforth are used to refer to the configurations of the
inverted and standard flags in undisturbed flow, respectively. EIE @2X @3X
¼ 3 ¼0 (5)
denotes either an inverted or standard flag behind a bluff body. @s2 @s

010801-2 / Vol. 71, JANUARY 2019 Transactions of the ASME


The standard flag is fixed at the leading edge (s ¼ 0), while the while a low KB value suggests a more flexible flag or high
inverted flag is fixed at the trailing edge (s ¼ L). The fixed edge flow speed. The tension T  that originates from the viscosity effect
of a flag satisfies clamped or pinned boundary conditions, which of a boundary layer can be modeled empirically [33,34] or
are, respectively, given by expressed as a function of Ms and KB [35]. Meanwhile, the tension
is generally neglected for inviscid flow assumption [36]. Most
@X studies exclude the effects of gravity that would also contribute
X¼ ¼0 (6)
@s to the tension in flag [34,37]. The dissipative properties are
3=2 1=2
characterized by two coefficients   ¼ q2 f ðqs hÞ B and
@2X
X¼ ¼0 (7) l ¼ lq2f ðqs hÞ5=2 B1=2 [38]. The fluid damping and structural
@s2 viscoelastic damping are neglected in the models of inviscid fluid
The inlet fluid velocity is uniform or parabolic in channel flow for and perfectly elastic material [39,40].
IIE and MIE, while the inlet fluid velocity is oscillating for EIE. The reduced flow velocity U and flow-to-mass ratio Mf are
By scaling distance with the length of the flag L, time with L/U, also commonly used in the literature, which are defined as
and pressure difference with qf U 2 , where qf is the density of fluid [34,36,41]
and U is the free-stream velocity of flow, Eq. (3) can be written in rffiffiffiffiffiffiffi
nondimensional form as qs h qf L
U  ¼ LU ; Mf ¼ (10)
  B qs h
@ 4 X @ 2 X @ @X 1=2 @X
KB 4 þ Ms 2   T   þ KB Ms3=2    where U 
is the ratio of the flag’s free vibration characteristic
@s @t @s @s @t pffiffiffiffiffiffiffiffiffiffiffiffi
  2
time L qs h=B to the characteristic time of the fluid flow L/U. In
1=2 5=2  @ @ 4 X
þ KB Ms l  ¼ Dp n (8) order to compare the data in a consistent framework and summa-
@t @s4 rize these laws, the above parameters are converted to KB and Ms
as follows:
The two key nondimensional parameters are defined as

B qs h Ms 1
KB ¼ ; Ms ¼ (9) KB ¼ ; Ms ¼ (11)
qf U 2 L3 qf L U 2 Mf

where KB is the ratio of the flag’s bending force to the fluid’s iner- Theoretical studies that solve the fluid–structure interaction
tial force. A high KB value denotes a stiff flag or low flow speed, system are summarized in Table 1. The immersed boundary (IB)

Table 1 Theoretical studies to solve fluid-induced flag vibrations

Flow-induced Method to solve


vibration type fluid–structure interaction Fluid model Structure model Author

IIE Immersed boundary method Incompressible viscous N–S Nonlinear 3D Tang et al. [42], Gilmanov et al. [43]
equations Nonlinear 2D Huang et al. [44], Goza et al. [21], Goza and
Colonius [45], Park et al. [12], Ryu et al. [22]
Theoretical linear stability Incompressible viscous N–S Linear 2D Goza et al. [21]
analysis equations
MIE Immersed boundary method Incompressible viscous N–S Nonlinear 3D Huang and Sung [37]
equations Nonlinear 2D Lee and Choi [46], Akcabay and Young [47],
Goza and Colonius [45]
Linear 2D Huang et al. [48], Zhu and Peskin [49]
Coupled fluid–structure direct Incompressible viscous N–S Nonlinear 3D Banerjee et al. [50]
simulation equations Nonlinear 2D Cisonni et al. [51], Connell and Yue [33]
Linear 1D Balint and Lucey [52]
Combined vortex sheet and Potential flow Linear 2D Michelin et al. [53]
discrete point vortex model Linear 1D Howell et al. [54]
Linear/nonlinear vortex sheet Potential flow Nonlinear 2D Wang et al. [55], Alben [56], Alben and
model Shelley [39]
Linear 1D Shoele and Mittal [57], Yamaguchi et al. [58]
Linear/nonlinear vortex Potential flow Nonlinear 2D Chen et al. [59], Attar et al. [60], Zhao et al. [61],
lattice model Tang and Paidoussis [34]
Nonlinear 1D Dunnmon et al. [62], Tang et al. [3]
Linear 1D Gibbs et al. [63], Eloy et al. [36]
Theoretical linear stability Incompressible viscous N–S Linear 2D Goza and Colonius [64]
analysis equations
Potential flow Linear 1D Eloy et al. [40], Guo and Paidoussis [4],
Kornecki et al. [65], Jia et al. [66]
Potential flow (Theodorsen’s Linear 1D Argentina and Mahadevan [67], Huang [2]
approximation)
Potential flow (slender-body Linear 1D Lemaitre et al. [68], Datta and Gottenberg [69]
approximation)
EIE Immersed boundary methods Incompressible viscous N–S Linear 1D Pan et al. [28]
equations
Arbitrary Incompressible viscous N–S Nonlinear 2D De Nayer et al. [70], H€
ubner et al. [71]
Lagrangian–Eulerian method equations
Theoretical linear stability Potential flow (Theodorsen’s Linear 1D Manela and Howe [72]
analysis approximation)

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-3


method, arbitrary Lagrangian–Eulerian (ALE) method, and solved theoretically by modal analysis; see, for instance, the
coupled fluid–structure direct simulation are three methods that Galerkin method [34,40].
can be used to solve the fluid-structure interaction with the incom-
pressible viscous N–S equations. Those methods facilitate detailed
predictions of the flow field and vortex dynamics. The ALE 3 Instability Boundary
method is “well suited for simulating high Reynolds number flows The foremost problem of fluid-induced vibrations in flags is the
due to its inherent body-fitted mesh structure that conforms to instability boundary. The instability boundaries and vibration
boundaries at all times” [43]. However, the ALE method is lim- modes in IIE, MIE, and EIE are displayed in the parameter space
ited to mild deformations, since frequent remeshing is required to (Ms , KB ) shown in Fig. 2. The instability region of IIE is concen-
“prevent the mesh from becoming severely distorted when large trated between 0:1 ⱗ KB ⱗ 0:4. The inverted flag is first straight at
deformations develop” [43]. As a result, the ALE method has only KB ⲏ 0:4. When KB decreases to 0:3 ⱗ KB ⱗ 0:4, the inverted flag
been employed to model the forced flag with relatively low vibra- vibrates asymmetrically in the biased mode, as shown in Fig. 2(a).
tion amplitude. For coupled fluid–structure direct simulation, fluid Symmetric large amplitude vibration occurs in the flapping mode
dynamics are solved by discretizing the N–S and continuity equa- for 0:1 ⱗ KB ⱗ 0:3, as shown in Fig. 2(b). This flapping mode for
tions onto a grid fitted to the deformable solid body. Two separate the inverted flag resembles the first mode of a vibrating cantilever
flow solver and structural solver require solving the velocity and beam in vacuo [74]. The inverted flag shows large deflections in
pressure fields iteratively in the flow together with iteratively solv- Fig. 2(c) as KB decreases to KB ⱗ 0:1; which results in a limited
ing for the position of the solid, and every update in the solid’s region for the flapping mode. The instability boundary is well cap-
position also requires updating the grid for the fluid [33]. The tured by the global stability analysis [21]. The minor dependence
immersed boundary method is efficient since entire simulation of instability on Ms is shown for IIE, except two minor differences
can be carried out with a fixed Cartesian grid without boundary as Ms increases from O(103 ) to O(1). This results in a narrower
conforming, and the structure motion equation and fluid motion or absent-biased mode with a steeper transition from straight to
equation are solved separately by adding an interaction momen- flapping mode for a higher Ms value and an extension of the flap-
tum force term [49]. It does not require remeshing and can readily ping mode to slightly lower KB values. A further decrease of KB
handle arbitrarily large deformations. However, both the coupled to the instability regime of MIE for the standard flag would flip
fluid–structure direct simulation and IB method are generally lim- the inverted flag by 180deg, and the inverted flag vibrates as the
ited to low Reynolds numbers on the order of a few hundred to standard flag in the flipped-flapping mode, as shown in Fig. 2(d)
thousand. Large Reynolds numbers are prohibitively expensive [42,75].
because very fine grids are required in the flow field. Scare The instability of MIE occurs at lower KB value compared to
attempts have been made to extend the IB method to Re  105 that of IIE, which suggests that the standard flag vibrates with
“using wall models for reconstructing boundary conditions at the larger inertial force of fluid than the inverted flag. In contrast to
immersed boundary nodes” based on large-eddy simulation of tur- the minor dependency of the instability boundary on Ms for IIE,
bulent flows [43]. Recently, a global stability analysis based on Ms plays an important role in determining the instability boundary
the linearization of the fully coupled fluid-structure system of and vibration dynamics of the standard flag because MIE origi-
equations has been performed for the standard flag [64] and the nates from the interaction between the flag’s inertia, the fluid
inverted flag [21], the critical KB is accurately predicted for a dynamic pressure force, and the flexural rigidity of flag. The criti-
wide range of Ms . The flags suffer unstable equilibria that broadly cal KB value for MIE has a general tendency to increase with Ms ,
lead to nonlinear, limit-cycle oscillations, typically through Hopf which means a heavy or short flag corresponds to an extended
bifurcations that can be either subcritical or supercritical depend- unstable region of KB , as indicated in Eq. (9). A close examination
ing on the parameter regime [21,38,64]. of the instability boundary reveals that the critical KB value is
Flow separation occurs along IIE and EIE. Therefore, viscous characterized by a cascade of modal branches, which results from
flow is essential for simulating the postcritical behaviors of these the transitions of vibration modes. The first branch at Ms ⲏ 0:7
types of fluid-induced flag vibrations. However, experimental corresponds to mode 2 (i.e., the single-neck mode) of the
results for the standard flag demonstrate that the flow remains standard flag, as shown in Fig. 2(e). For example, KB ¼
attached to the dynamically deformed flag for high KB and Ms in 2:5  102 when Ms ¼ 2 [53], KB ¼ 1:11  102 when Ms ¼ 1:41
vibration mode 2, which justifies the employment of potential [76], KB ¼ 5:3  103 when Ms ¼ 0.83 [34]. The standard flag
flow theories [1,73]. However, the vortex methods and panel never vibrates in the first mode because the mode always appears
methods in potential flow, including vortex sheet model, vortex to be stable [32,40]. For the second branch (0.3 ⱗ Ms ⱗ 0.7) and
lattice model, combined vortex sheet, and discrete point vortex third branch (0:1 ⱗ Ms ⱗ 0.3), mode 3 with one node in the flag
model, normally fail to deal with the interactions of vortices and profile (see Fig. 2(f)) and mode 4 with two nodes (see Fig. 2(g))
flags. For a theoretical linear stability analysis that aims to predict are observed, respectively, in the middle of those branches
the onset of vibration behavior, the slender-body approximation [24,51]. Those single structure modes are close to the in vacuo
and Theodorsen’s approximation further simplify fluid loading in eigenmode shape. For the transition regions of the branches, the
the model. The slender-body approximation is derived under the flow interacts with two modes competing against each other. As a
assumption that the flag has low aspect ratio. Theodorsen’s result, the flag undergoes mixed modal motion [51] or modal coa-
approximation is derived in the limit of small deformations when lescence [54]. The combination of modes 3 and 4 (or double-neck
the vertical velocity is nearly constant. Since infinitesimal distur- mode) shown in Fig. 2(h) occurs due to the intersections
bances are assumed, vortex shedding from the trailing edge and between the second branch of mode 3 and third branch of mode 4,
the flag boundary conditions is neglected. Thus, the linear stability as reported by Souilliez et al. [76] for Ms ¼ 0:54 and
analysis only provides results on critical KB values that are quali- KB ¼ 4:75  103 , and by Michelin et al. [53] for Ms ¼ 0.33 and
tatively consistent with the experimental results, rather than the KB ¼ 2:3  103 . The interaction between mode 4 with the higher
nonlinear effect in the postcritical region where the disturbance order mode results in a combination of modes 4 and 5 with three
fully develops [66]. Theodorsen’s approach has also been used to necks in Fig. 2(i), as given by Michelin et al. [53] for Ms ¼ 0.1
model fluid in the linearized motion of the flag under EIE, where and KB ¼ 7  104 . More inflections (necks or nodes) are present
vortex shedding from the upstream bluff body was modeled as a along the standard flag for a lower Ms value, i.e., for a longer or
linear vortex sheet [72]. Future developments of theoretical mod- lighter flag. Mode 5 and higher order modes or chaos occur for
els require amending the effects of vortex–body interactions. Ms < 0:1.
The structural model of a flag is numerically solved by The theoretical models [67,77] fit the critical KB of MIE for a
using the finite difference method [28] or finite element method limited range of Ms and fail to capture the modal branches. The
[43,44,51]. For potential flow, the structural equation can be inviscid models [36,39,41,53,56,57,78] predict the instability

010801-4 / Vol. 71, JANUARY 2019 Transactions of the ASME


Fig. 2 Instability boundaries and IIE, MIE, and EIE vibration modes

boundary fairly well for high Ms values. Because flow visualiza- low Ms value favors EIE because the damping of the flag is rela-
tions show that the flow remains attached to the vibrating flag in tively low. Consequently, the flag responds rapidly to the undulat-
mode 2 [73], the potential flow theories, which consider the non- ing wake from the upstream bluff body. For a low Ms value at
linear effect of large geometric deflections are capable of captur- KB < 0:05, the standard flag follows the unsteady pressure field
ing the flow fields around the flag. However, the critical KB value setup by vortex shedding, vibrates with its wavelength and fre-
is overestimated by the inviscid models for Ms < 0.2, the origin of quency close to that of the vortex shedding, and shows a trailing
the mismatch is still an open question. By comparison, simula- wave along the flag [8,27,70,72,81–83]. This is distinctly different
tions that incorporate the viscous effect show an instability bound- from MIE, where the standard flag is more difficult to flutter by
ary that is more consistent with the experimental results for MIE for a low Ms value, as indicated by the low critical KB value.
Ms < 0.2 [33,51]. Both the potential theories and simulations that Copper strips have been adhered along a standard flag to increase
consider the viscous effect qualitatively predict the modal its Ms value in water so that the standard flag is capable of vibrat-
branches of the instability boundary, but obvious quantitative dis- ing by MIE [77]. Although the EIE configuration can be imple-
crepancies exist in the divisions of the modal branches, which mented in the instability region for IIE and MIE, IIE and MIE
require improved numerical methods. Note that the hysteresis dominate while EIE slightly modulates the vibration dynamics of
effect, which means the critical KB value is higher in a Re- the inverted or standard flag, thus causing their motions to be
increasing process compared to that in a Re-decreasing process, is more irregular [7,26]. Therefore, the instability regime for EIE
extensively reported in experimental results rather than in simula- complements the IIE and MIE margins in the parameter space
tions. This is mainly attributed to the unavoidable planeity defects (Ms , KB ), but EIE is not recommended in the instability regime of
of a flag that stiffen the flag in the Re-increasing process during MIE or IIE for any application.
experiments, while the planeity defects are ironed out once the
flag vibrates [38,79]. The critical KB value in the Re-decreasing 4 Vibration Dynamics
process is reproducible and is used when comparing the instability
boundary from experimental and numerical results that neglect The vibration dynamics of the inverted, standard, and forced
the planeity defects. flags obey distant laws due to the various mechanisms of fluid-
The vortices and fluctuating flow due to the upstream bluff induced flag vibrations. This is reflected from the correlation
body force an inverted flag or standard flag to vibrate for higher between the vibration frequency and amplitude, as illustrated in
KB value than the instability boundary of IIE and MIE. The flags Fig. 3. The frequency is normalized by
are forced into the first mode for large KB (see, for example,
KB ¼ 1:17 and Ms ¼ 2.19 [80] for the inverted flag shown in fA
StA ¼ (12)
Fig. 2(j), and the standard flag shown in Fig. 2(k) [7,28,43]). A U

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-5


Fig. 3 Correlation between the vibration frequency and amplitude for IIE, MIE, and EIE flags

where f is the vibration frequency and A is the tip-to-tip vibration from mode 2 to higher order modes. Chen et al. [59] have
amplitude of flag. The tip-to-tip vibration amplitude is normalized reported that mode 2 occurs at KB ¼ 0:01, mode 3 occurs
by the length of flag. The increase of A/L is related to a decrease at KB ¼ 3:7  103 , mode 3 and 4 are combined at
in KB , which is generally realized by increasing the flow velocity KB ¼ 1:6  103 , and mode 4 occurs at KB ¼ 1:3  103 for a
U for a certain flag. flag with Ms ¼ 1. Virot et al. [85] reported that mode 2, combined
The maximum flapping amplitude A/L for the inverted flag in mode 3 and 4, combined mode 4 and 5, and chaotic vibration
IIE reaches up to 1.6–1.9, which is approximately two times of arise in sequence with a successive decrease in KB for Ms ¼ 0:72.
that for the standard flag in MIE (A=L  0:9), and larger than two For the light flag in the second branch of the instability boundary
times of that for EIE (A/L < 0.8). The flapping amplitude is (0:3 ⱗ Ms ⱗ 0:7), mode 3 or combined modes 3 and 4 are initi-
proportional to the curvature and strain along the flag, and large ated at instability boundary with large StA  0:26. The vibration
flapping amplitude is favorable for applications such as energy amplitude A/L and StA increase as KB decreases, and the flag tran-
harvesting [22]. For IIE, StA initially increases as A/L increases sits to mode 4 or combined modes 4 and 5 in the postcritical
from around 0.9 to around 1.5. As A/L tends to saturate, a further region, as shown for Ms ¼ 0:35 and 0.4 in Virot et al. [85]. A/L
decrease in KB leads to a slight increase in A/L and a drop in StA reduces to A/L < 0.7 due to augmented inflections along the flag
[19,22,42]. Although the instability boundary shows little depend- in the higher order modes compared to mode 2. The Strouhal
ency on Ms for IIE, as shown in Fig. 2, StA varies from 0.07 to number for various vibration modes collapses to around
0.37 with Ms for a fixed KB value [75]. StA is an important indica- 0.21–0.23 when the length between the trailing edge to its nearest
tor for vortex dynamics, which will be illustrated in Sec. 5. node is taken as the characteristic length [24,86]. Michelin et al.
In the postcritical region of the standard flag in MIE, the [53] reported that the combined modes 3 and 4 retain in the near
evolution of StA with A/L is collapsed onto the line with a slope postcritical region 2:3  103  KB  3:6  103 for Ms ¼ 0:33,
StL  0:23 for heavy flags (Ms > 1) [34,37,60,62,84], where the while chaotic flapping occurs for ultralow KB  1:5  103 . In
slope in the parameter space of (A/L, StA ) denotes the third branch of the instability boundary 0:1 ⱗ Ms ⱗ 0:3, cha-
otic vibrations are extensively reported in the postcritical regime,
fL e.g., at KB ¼ 1  104 [33] and at KB ¼ 3:125  104 (which is
StL ¼ (13)
U only 9.54% of the critical KB ) for Ms ¼ 0.15 [39]. The chaotic
flapping mode is characterized by irregular vibration with no
This means that the dimensionless frequency has no vibration apparent dominant frequency, but rather a wide range of excited
amplitude dependence. The linear aerodynamic model in frequencies in the flag bending energy spectrum. The loss of peri-
Dunnmon et al. [62] underestimates the slope StL . In contrast, the odicity arises from the nonlinear interactions between higher
nonlinear aerodynamic model in Tang and Paidoussis [34] faith- order modes of comparable frequencies [87–89]. The slope StL
fully fits the experimental results. The value StL  0:23 repre- further increases for the flag when Ms < 0:1, and the vibration
sents mode 2 of the standard flag. This vibration mode amplitude reduces to A/L < 0.3 [77].
corresponds to a large range of KB below the instability boundary The criterion for assessing how well the flag couples with the
for large Ms values (see, for example, 8:7  103 < KB < vortices from the upstream bluff body in EIE is to examine the
1:85  102 for Ms ¼ 3 [26]). When Ms ¼ 0:7–1 [59,85], the ratio of the flag’s vibration frequency and the vortex shedding fre-
standard flag flutters in mode 2 when KB is close the critical KB quency. The resonance of the flag’s vibration frequency with the
value and StL ¼ 0:23 for A/L < 0.5. However, a sudden deviation vortex shedding frequency is defined as
of (A/L, StA ) from the line StL ¼ 0:23 is noted for larger A/L due
to a further decrease of KB far from the critical KB value. The
increased slope corresponds to a transition of the vibration mode ðStL  D=LÞ=StD ¼ 1 (14)

010801-6 / Vol. 71, JANUARY 2019 Transactions of the ASME


well established. Every cycle of the inverted flag comprises two
where StD ¼ fv D=U is the Strouhal number for vortex shedding strokes, a LEV followed by a TEV shed in each stroke for large
from the bluff body, fv is the vortex shedding frequency from the StA ¼ 0:22  0:34. Consequently, two pairs of counter-rotating
upstream bluff body, and D is the characteristic length of the bluff LEV and TEV convect downstream toward two sides of the
body. StD is constant for plenty of bluff bodies over a wide Reyn- inverted flag in every cycle as shown in Fig. 4(a); thus, the flow
olds number range, such as StD ¼ 0.2 for a circular cylinder [90], mode is termed the two pairs (2P) mode. The vortex mode was
StD ¼ 0.13 for a rectangular cylinder, and StD ¼ 0.15 for a plate described by Goza et al. [21] at StA ¼ 0:28  0:36, by Ryu et al.
[91]. The EIE results are dispersed in the parameter space (A/L, [22] at StA ¼ 0:26, and by Shoele and Mittal [93] at StA ¼0.22.
StA ) for various bluff bodies and flags, StA reaches up to 0.85 As StA reduces to 0.13–0.19, a secondary LEV (L2) forms behind
when A/L ¼ 0.44 in Allen and Smits [8], while StA is approxi- the TEV during every stroke apart from a pair of LEV and
mately 0.33 when A/L ¼ 0.75 in Shi et al. [82]. However, (A/L, TEV, as shown in Fig. 4(b). Therefore, two pairs plus two single
StA ) in EIE evolves around the corresponding line of ðStL  vortices shed in every cycle, i.e., the 2P þ 2S mode, as reported
D=LÞ=StD ¼ 1 for every experimental set of EIE, as indicated by by Ryu et al. [22] at StA ¼ 0:19, by Shoele and Mittal [93] at
the dotted or dashed lines in Fig. 3. (A/L, StA ) is below ðStL  StA ¼ 0:17 and StA ¼ 0:13, and by Goza et al. [21] at StA ¼ 0:18.
D=LÞ=StD ¼ 1 for low A/L [8,27,82,92], which suggests that the Figure 4(c) shows that a secondary TEV follows the secondary
vibrating frequency is lower than that of the vortex shedding fre- LEV at StA ¼ 0:12, thus forming the 4P mode [22]. More vortices
quency due to the high damping effect at large KB value. The von are generated with a further decrease in StA to 0.1 and 0.08, thus
Karman vortex street becomes strong enough to overcome the generating 4P þ 2S mode and 6P þ 2S mode in Figs. 4(d)
damping effect of the flag as KB decreases. Consequently, the flag and 4(e), respectively [75]. With this framework, the vortex pat-
responds rapidly to the undulating wake from the bluff body, and terns are inferred directly from the vibration dynamics of the
the vibration frequency of the flag locks-in with the shredding fre- inverted flag by StA . StA is the ratio of two velocities as defined by
quency of von Karman vortex street, as indicated by the (A/L, StA ) Eq. (12), where fA is the average speed of the inverted flag’s lead-
values that fall on the lines defined by ðStL  D=LÞ=StD ¼ 1. This ing edge, and U is the flow velocity. For low StA , the flow velocity
kind of resonance between the flag and the undulating wake of the is fast relative to the velocity of the inverted flag and vortices are
bluff body leads to a large vibration amplitude A/L that is close to generated quickly. Therefore, more vortices have time to grow in
the natural wake width of the upstream bluff body [8]. Moreover, every cycle. Although Ryu et al. [22] proposed StL as an indicator
the phase speed of the traveling wave along the standard flag is of the vortex patterns and the 2P, 2P þ 2S, and 4P modes occurred
also close to the convection speed of the vortices [92]. in sequence as StL decreases in their simulations, the 4P þ 2S
mode in Gurugubelli and Jaiman [75] occurred at higher StL ¼
0:1 than that of 4P mode at StL ¼ 0:07 in Ryu et al. [22]. The
5 Vortex Dynamics
disagreement occurs because a fixed length L is selected as the
5.1 Uniform Flow. The inverted and standard flags are used characteristic length in StL as shown in Eq. (13), regardless of
as vortex generators for the applications such as convective heat the flapping dynamics. In contrast, StA ¼ StL  A=L adequately
transfer enhancement or mixing enhancement, which requires considers the interaction between the flapping dynamics and the
examining the vortex dynamics of the flags. Although the inverted vortex dynamics. Moreover, the 2P mode occurs in a large inter-
flag vibrates in mode 1 in the flapping mode, various vortex pat- val of StA around 0.12, the 2P þ 2S mode occurs for a range of
terns have been reported separately in the literature for various StA values around 0.6. The modes with more vortices in 4P,
vibration amplitudes and frequencies of the inverted flag. A 4P þ 2S, and 6P þ 2S modes correspond to a narrower range of
framework for understanding the relationships between the vortex StA  0:02. The 6P mode has not been reported so far, probably
dynamics and flag vibration dynamics is lacking in the literature. due to its corresponding narrow range of StA . However, it is con-
Figure 4 shows that the dependency of vortex patterns on StA is jectured that the 6P mode exists between the 4P þ 2S mode at

Fig. 4 Vortex dynamics interacting with the inverted flag in the flapping mode under uniform
flow, ((a) and (b)) from Goza et al. [21], Shoele and Mittal [93], and Ryu et al. [22], (c) from Ryu
et al. [22], ((d) and (e)) from Gurugubelli and Jaiman [75]

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-7


StA ¼ 0:1 and the 6P þ 2S mode at StA ¼ 0:08. Further work is mode 2, which was reported from the simulation results for
required to validate this conjecture and explore more vortex StA ¼ 0:05–0:23 by Connell and Yue [33], for StA ¼ 0:21 by
modes. Huang et al. [48] and Lee et al. [96], as well as the experiments
Vortex dynamics are essential for understanding the IIE of the conducted in water at StA ¼ 0:06–0:07 by Giacomello and Porfiri
inverted flag since the vibration is induced by the flow instability [97]. Apart from the von Karman vortex street, a series of small
or vortex-induced vibration. LEV is rolled up by the leading edge Kelvin–Helmholtz vortices propagate along the shear layer in
of the inverted flag and is characterized by a region of low pres- flowing soap films [98]. Because the flow velocity at the two sides
sure behind the flag, thus enhancing the force that bends the of flag jumps discontinuously, the velocity shear causes a proces-
inverted flag with large deformation as LEV grows in strength and sion of the same-signed Kelvin–Helmholtz vortices during every
size. As the LEV sheds into the wake, the sudden drop of the fluid stroke of the flag as the flow passes the flag and joins together
force impinging on the inverted flag causes it to bounce due to the in the wake. As a result, the wake structure shows a combination
elastic restoring force. Meanwhile, TEV enhances this process of von Karman vortex streets and Kelvin–Helmholtz vortices,
[94,95]. As the flag passes its mean position, LEV forms again on as schematically shown in Fig. 5(a). The Kelvin–Helmholtz
the other side of the flag for the 2P mode, and symmetric periodic vortices are much smaller than the von Karman vortex street
flapping occurs as this mechanism repeats. Although the second- when the velocity shear is low, as was shown in Jia [98] at
ary or tertiary LEV increases the fluid force that contributes to StA ¼ 0:07  0:33. Those Kelvin–Helmholtz vortices are large
bend the inverted flag for the 2P þ 2S, 4P, 4P þ 2S, and 6P þ 2S and indistinguishable from the von Karman vortex street for high
modes, the resilience of the inverted flag normally dominates over velocity shear [25,99]. In the higher order vibration mode corre-
bounce. However, the biased mode forms when the force of the sponding to a larger StA value, every vortex in the von Karman
secondary or tertiary LEV prevails the resilience, and the inverted vortex street splits into two or three same-signed vortices near
flag bends toward the same side rather than passing the centerline StA ¼ 0:36 and for StA ¼ 0:4–0:5, as shown in Figs. 5(b) and 5(c),
toward the other side. In the deflected mode, the von Karman vor- respectively [48,96]. The vortical wake becomes disorganized in
tex street sheds periodically from the deflected flag [93]. For the the chaotic regime of the standard flag [33,39].
flipped-flapping mode of the inverted flag, the vortex dynamics
are the same as that for the standard flag since the inverted flag in
the flipped-flapping mode is akin to standard flag [93]. 5.2 Channel Flow. The above discussions regarding the
The standard flag with high KB and Ms shows no boundary- instability boundary, vibration dynamics, and vortex dynamics
layer separation along its curved surface in vibration mode 2, concern uniform flow without confinement or with negligible con-
while the flexible flag in higher order vibration mode exhibits finement. However, confinement effect plays an important role in
small-scale separations [1,73]. The trailing edge of the standard some applications, such as heat transfer enhancement in channel
flag rolls up strong vorticity in the wake. These vortices are gener- flow and energy harvesting base on electrostatic conversion using
ated by the agitation of the standard flag’s trailing edge rather channel walls [100,101]. Basically, an increase in the flag’s
than the vortex shedding, which induces the instability as the case length-to-channel-width ratio C destabilizes the flag at higher KB
in the inverted flag. Large vortical structures in the von Karman values compared to the aforementioned instability boundary in
vortex street form in the wake of the standard flag for vibration Fig. 2, both for the inverted flag [79] and the standard flag [56].

Fig. 5 Vortex dynamics interacting with the standard flag under uniform flow

010801-8 / Vol. 71, JANUARY 2019 Transactions of the ASME


The effect of confinement on destabilization of the standard flag is studied by Park et al. [12]. The interaction between the inverted
remarkable for heavier flag and is relatively weak for lighter flag. flag and fluid flow generates a pair of TEV and LEV plus a sec-
Moreover, the increased value of C leads to an increased vibra- ondary TEV with every stroke. The TEV first sheds rather than
tion frequency for the standard flag and a transition to a higher the LEV as occurs in the 2P þ 2S mode, since the oncoming fluid
order vibration mode at lower KB value compared to the uniform velocity at the leading edge of the flag close to the channel wall is
flow case [30,56,57]. Although the effect of confinement on flag lower than that at the trailing edge in the channel wall. Moreover,
vibration dynamics has been examined [30], a gap in research the fore TEV induces a pair of vortices near the channel wall.
exists in determining the effect of confinement on vortex dynam- The flow field of the standard flag in channel has been modeled
ics. Some initial studies have addressed the flow field inside a numerically by Alben [56] and Shoele and Mittal [11]. Symmetric
channel for the inverted and standard flags as follows. von Karman vortex streets are formed and the vortices begin to
The confinement effect of channel flow on vortex dynamics of interact with the channel walls further downstream for low con-
the inverted flag has been studied recently by Yu et al. [94] and finement [56]. The von Karman vortex street in the wake of a
Yu et al. [79]. Figure 6 summarizes the evolution of the 2P mode standard flag (see Fig. 5(a)) changes into a reverse von Karman
under uniform flow (see Fig. 4(a)) for various C values. For the vortex street wake when the tip-to-tip vibration amplitude is close
highest value (C ¼ 0.50) at StA ¼ 0.32, the LEV that forms at one to the channel width for C ¼ 0:66–1 [11]. The reverse von
side of the channel develops into bounded circulation around the Karman vortex street is similar to the abovementioned wake of
inverted flag as it convects along the flag, which in turn transfers the inverted flag for C ¼ 0:5, as shown in Fig. 6(a). These vorti-
to TEV at the trailing edge of the flag and sheds at other side of ces are capable of entraining the hot fluid away from the wall,
the channel with every stroke [94]. As a result, only a pair of thus considerably enhancing heat transfer. The vortices are more
reverse von Karman vortices shed in every cycle in highly con- concentrated in the middle of the channel as C increases. The
fined channel flow rather than the 2P mode as occurs in severe collisions of the standard flag with the channel wall
unbounded flow. Figure 6(a) shows that the reverse von Karman for high confinement result in weaker and disordered vortical
vortex street forms with an energetic wake, and the fluid between structures that disappear shortly after the trailing edge in the
the neighboring vortices accelerates and is impinged on the side- wake [11,56].
walls in a staggered fashion. As C decreases to an intermediate Vortex dynamics are vital to fully understand heat transfer and
value of 0.375 at StA ¼0.25, the first LEV in every stroke decays mixing enhancements, as well as other potential applications
quickly close to the wall, and the secondary LEV breaks up into based on the inverted and standard flags. In addition, the vortex
two vortices as the LEV convects to the trailing edge of the flag, dynamics of flag in channel flow are abundant, in particular for
with one vortex shedding directly into the wake as LEV while the the inverted flag. However, the survey shows that vortex dynamics
other vortex of the same sign rotates around the flag and transfers of the flag in channel flow are relatively unexplored. Further
to TEV as occurs for C ¼ 0.50 [79]. Consequently, two counter- investigations into the vortex modes under channel flows for vari-
clockwise and two clockwise vortices alternatively convect into ous StA and C values are required to address the gap.
the wake with every oscillation period, as shown in Fig. 6(b). For
C ¼ 0:25 at StA ¼ 0:25, two pairs of LEV and TEV are gener- 6 Effects of Three-Dimensionality and Reynolds
ated from the interaction between fluid flow and the vibrating
inverted flag as occurs in the 2P mode in unbounded flow. In addi- Number
tion, each LEV induces an opposite sign vortex “W” near the The two-dimensional flag has been the focus of the majority of
channel wall, as shown in Fig. 6(c). The results suggest that the experimental, analytical, and numerical studies. The flag evolves
vortex dynamics of the inverted flag are substantially refereed by into two-dimensional vibration and generates spanwise uniform
the length-to-channel-width ratio in channel flow due to the inter- vortex structures, even when the initial conditions are specified
action between the vortices and the channel wall. using three-dimensional (3D) displacements with spanwise wave-
The evolution of the 2P þ 2S mode under uniform flow at length, as reported by Banerjee et al. [50]. The three-dimensional
StA ¼ 0:18 (see Fig. 4(b)) for channel flow at C ¼ 0:5 has been spanwise wavelength dies out after the flag begins vibrating. How-
ever, this numerical work ignored edge effect and gravity. On the
one hand, real flag has finite span, resulting in the three-
dimensional side edge effect. The side edge effect can be reduced
in experiments by confining the spanwise flow with the presence
of walls, and the flag of finite span converges to the two-
dimensional limit when the gap between the side edge of flag and
the walls tends to zero. However, the side edge effect plays a role
once the gap is larger than O (104 ) multiplied by the chord of the
flag [102,103]. This side edge effect is dependent on the aspect
ratio H  ¼ H=L, where H is the span of the flag. Due to the forma-
tion of vortical structures at the side edges, the pressure difference
across the flag is reduced and vortices that shed from the flag are
attenuated [37]. Consequently, the flag is stabilized by the effect
of side edges. The influence of the side edge effect on the instabil-
ity boundary and flag vibration dynamics can be neglected for
H  2. However, an obvious decrease in the critical KB value
occurs as H  decreases when H  < 1:5, both for the inverted flag
[42] and the standard flag [36,40,84]. A flag with low H  leads to
lower A/L and StA values compared to the two-dimensional case
[37,38,42]. The flag undergoes two-dimensional deformation for
high KB values even with the side edge effect, or slightly three-
dimensional deformations occur only around the side edge of the
flag for intermediate KB , while the three-dimensional deforma-
tions are obvious for low KB values. On the other hand, three-
dimensional deformations occur if gravity prevails overflow
Fig. 6 Vortex dynamics interacting with the inverted flag in the inertia, resulting in sagging, rolling motions of the upper corner,
flapping mode under channel flow and oblique waves in vibrating flags [36,37,104]. In the

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-9


Fig. 7 (a) Destructive and (b) constructive coupling modes for inverted flags, (c) destructive
and (d) constructive coupling modes for standard flags

asymptotic limit of zero aspect ratio, a theoretical model for the with the vortices emanated from the upstream flag. Flags with
stability of slender inverted flag has been developed by Sader equal size have the same vortex shedding and vibration frequen-
et al. [105]. In contrast to the divergence instability of the two- cies, allowing for the synchronization of the vortex streets. The
dimensional inverted flag as KB decreases, the slender inverted vibration of the downstream flag demonstrates a phase delay with
flag is never globally unstable, i.e., in the absence of the flapping respect to the vibration of the upstream flag. The phase delay dic-
mode in Fig. 2(b). Instead, the slender inverted flag exhibits bifur- tates the phase in which the downstream flag encounters vortices
cation of multiple equilibria, including a stable deflected mode as that are shed from the upstream flag. Therefore, the phase differ-
shown in Fig. 2(c) and an unstable biased mode; the latter one is ence between flags depends linearly on the streamwise separation
attracted by the zero-deflection equilibrium. distance Gy for constant convective speed of upstream vortices,
Variation in Reynolds number does not fundamentally modify and the phase difference is periodically mediated by the wave-
the mechanisms of fluid-induced flag vibrations at high Reynolds length or vortex-spacing of the wake, both for the inverted flags
number. The low Reynolds number has a stabilizing effect on [44] and the standard flags [16,106]. Simulations of tandem stand-
flag, and computational simulations demonstrate that the flapping ard flags in potential flow, however, show that the downstream
mode of the inverted flag vanishes when Re ⱗ 50 [22]. For the flag vibrates irregularly for some Gy values and occasionally
standard flag in narrow channels, the high order vibration modes breaks the linear relationship between the phase difference and Gy
are successively stabilized to lower order modes as Reynolds [106]. Those Gy values correspond to the phases where the
number decreases from 1000 to 1. The critical KB decreases with upstream vortices directly impinge on the downstream flag, which
Re for most Ms when the vibration mode remains identical. How- disables the potential flow from providing satisfying results due to
ever, for certain Ms , the critical KB can increase before reaching a flow separation along the downstream flag. In contrast, the
maximum and then decreases, which is due to the “additional “viscosity smoothes out vortices that are encountered by the
instability mechanism related to the wall shear stress” in narrow downstream flag and damps out the irregular flapping motion” in
channels and the “axial stretch” of flag at low Re [51]. As for the viscous flow [16].
vortex dynamics, smaller scale vortex structures form in flows The interaction between the downstream flag and the upstream
with higher Reynolds numbers [33]. The inverted flag transits vortices is analogous to EIE. However, the upstream vortex modes
from the 2P mode to the 2P þ 2S mode as the Reynolds number occurring in the wake of the upstream flag, as illustrated in Fig. 4
increases from 50  Re  130 to Re >130 for KB ¼ 0:35 [22]. for the inverted flag and in Fig. 5 for the standard flag, are much
The instability boundary, vibration dynamics, and vortex dynam- more abundant and complex than the von Karman vortex street
ics become insensitive to Reynolds number for Re ⲏ 200 for the behind the bluff body in EIE, which undoubtedly complicates the
inverted flag [22,75,93] and Re ⲏ 1000 for the standard flag interaction with the downstream flag. The vibration amplitude A/L
[11,33,50,51]. for the upstream flag is smaller than that of an isolated flag for
short Gy values due to the feedback from the downstream flag,
7 Interaction of Multiple Flags both for the inverted [44] and standard configurations [107]. A/L
for the upstream flag increases with Gy and eventually approaches
The interaction between multiple flags through fluid flow the value for an isolated flag and remains stable. Figure 7 shows
involves vortex–vortex and vortex–structure interactions. Insights the mechanisms of the interaction between the upstream vortices
into the interaction between multiple flags facilitate the design of and the downstream flag. The vortices that are generated by
energy harvesting plants and the implementation of transfer upstream flag and downstream flag are denoted by dotted lines
enhancement with a series of flags in a long channel. Tandem and and solid lines, respectively. For the 2P mode wake of the inverted
parallel flags are basic building blocks with more complex config- flag in Fig. 7(a), the pair of LEV and TEV that shed from the
urations, wherein adjacent flags demonstrate the prominent behav- upstream flag moves away from the center position of the flag as
iors of tandem or parallel flags. The features of tandem and they convect downstream. The LEV directly passes around the
parallel arrangements, including the inverted and standard flags, downstream flag, while the TEV merges with the opposite-signed
are examined in the section. LEV of the downstream flag. The destructive merging of vortices
with opposite signs weakens the LEV of downstream flag [44].
7.1 Tandem Flags. In the tandem flags, the upstream flag Because LEV is generally attributed to the force that bends the
sheds vortices into its wake, and the downstream flag interacts inverted flag [94], the A/L value for the downstream flag is

010801-10 / Vol. 71, JANUARY 2019 Transactions of the ASME


Fig. 8 Coupled vibration for two parallel (a) inverted flags and (b) standard flags

reduced compared to that of an isolated flag, which is known as Huertas-Cerdeira et al. [108]. For Gx < 2, parallel inverted flags
destructive coupling. In contrast, both the first LEV and TEV of primarily exhibit irregularly vibration or they collide with each
the upstream flag pass around the downstream inverted flag in the other, and the in-phase coupling and deflected mode occasionally
2P þ 2S mode for the inverted flag, as shown in Fig. 7(b), while occur for Gx  0.3 [109]. The flag vibrations are decoupled for
the secondary LEV impacts the downstream flag. The merging of Gx > 5 [108].
the downstream flag’s LEV with the same-signed secondary LEV Coupling between two parallel standard flags has been investi-
of the upstream flag intensifies the vorticity, which in turn gated by changing Gx or KB while holding the other parameters
increases A/L for the downstream flag and causes constructive constant. The two dashed arrows in Fig. 8(b) indicate the two
coupling [44]. It is demanding to study whether more complex observation directions. For constant KB , Zhang et al. [25] discov-
interactions between the 4P, 4P þ 2S, or 6P þ 2S mode with the ered in-phase, anti-phase, and decoupled vibrations with the suc-
downstream flexible flag lead to constructive coupling or destruc- cessive increase in Gx . The same phenomenon has also been
tive coupling. Similarly, destructive and constructive couplings reported in a series of publications [48,110–112]. A concomitant
occur for the standard flags when the vortices that surround the jump in frequency occurs as the flags transfer from in-phase to
downstream flag merge with upstream vortices with the opposite anti-phase vibrations. For constant Gx , the flags tend to anti-phase
and the same rotational sense, as shown in Figs. 7(c) and 7(d), coupling for high KB values below the instability boundary, and
respectively. Constructive coupling accounts for “anomalous in-phase vibration dominates as KB decreases [113–115]. The
hydrodynamic drafting,” where the downstream flag endures transition region is characterized by the random alternation
larger drag and the upstream flag experiences less drag than that between in-phase and anti-phase vibrations [66] or the co-
of an isolated flag because the drag is correlated with A/L, as existence of the vibration frequencies that correspond to in-phase
reported by Ristroph and Zhang [107]. and anti-phase vibrations [114]. Quantitative coupling division is
not shown in Fig. 8(b) because the coupling distribution also
depends on Ms [116]. The in-phase vibration region expands for
7.2 Parallel Flags. Two parallel flags interact with each other low Ms values, while the anti-phase vibration region enlarges for
through the fluid when their transverse separation distance Gx is high Ms values.
of the order of their lengths. The predominant coupling mecha- The combinations of parallel and tandem flags form more com-
nism between two parallel inverted flags is the anti-phase vibra- plex configurations, such as a triangular configuration [117], three
tion as shown in Fig. 8(a) for 2 ⱗ Gx ⱗ 4. The anti-phase or more parallel flags [115,116,118], diamond configuration
vibration is energetically favorable, where any initial displace- [119], conical configuration [120], and X-shape configuration
ment condition for the two flags leads to inversion-symmetric [121]. Gallegos and Sharma [30] reviewed those configurations
vibration, as reported in the experimental work of Huertas- for the standard flags. In general, the anti-phase and in-phase
Cerdeira et al. [108]. The vibration of each flag is asymmetric vibrations are the two prominent coupling phenomena between
with stronger bending toward the interior of the parallel flags adjacent parallel flags in those combined configurations; and the
since the accelerated flow in the interior forms lower pressure constructive and destructive merging of vortices leads to intensi-
when the flags bend toward each other. A/L and StA increase in the fied and reduced vibrations in the downstream flags, respectively.
anti-phase vibration with respect to an isolated flag, and the gain
in A/L and StA becomes less prominent as Gx increases and satu-
rates to the value for an isolated flag with Gx > 3.2. The anti- 8 Applications
phase vibration corresponds to 0.1 ⱗ KB ⱗ 0.3 for Gxx3. And an The basic idea in mitigating fluid-induced flag vibrations is to
increase in Gx shrinks the regime of KB that corresponds to anti- adjust the flag parameters (L, h, qs , and B) or fluid parameters (U
phase vibration. Instead, coupling between parallel inverted flags and qf ) so that the (Ms ,KB ) values fall outside of the instability
is more sensitive to the initial conditions of the flags. Different regions in Fig. 2. Utilization of fluid-induced vibrations in the
initial conditions give rise to in-phase vibration, staggered vibra- flexible flags involves the selection of IIE, MIE, EIE, and the cor-
tion with constant phase difference between the in-phase and anti- responding vibration modes, design of the flag vibration dynamics
phase, and alternating vibration where the flags switch between in terms of (A/L, StA ) as shown in Fig. 3, optimization of the vor-
two or more of the in-phase, anti-phase, and staggered vibrations. tex modes as shown in Figs. 4–6, and the assembly of multiple
And perturbations cause transitions between those coupled vibra- flags as shown in Figs. 7 and 8. For instance, the development of a
tions. Qualitative consistent anti-phase vibration and relatively self-powered piezoelectric microflow sensor is actually based on
weak coupling between vibrations that are sensitive to initial the linear relationship between A/L and StA for the standard flag in
conditions have been confirmed from numerical results presented the second vibration mode [10]. Vortices behind the vibrating flag
by Ryu et al. [109], but the anti-phase vibration regime at 2 < are used to transform the asymmetric vortices of a slender aircraft
Gx  2:5 shows quantitative discrepancies from 2 ⱗ Gx ⱗ 4 in forebody to symmetric vortices, thus alleviating the lateral force

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-11


on the aircraft and improving its control characteristics [122,123]. Conversion from vibrational energy in flags into electricity
We review three representative applications of fluid-induced flag occurs via either a deformation-based or displacement-based
vibrations, namely, energy harvesting, heat transfer enhancement, mechanism. Piezoelectric conversion is deformation-based, where
and mixing enhancement. the flags are made of piezoelectric materials. Stretching and com-
pression of the piezoelectric materials due to the local curvatures
8.1 Energy Harvesting. The booming internet of things of the flags induce charge transfer between electrodes attached to
market requires wireless sensor networks, where numerous sen- piezoelectric materials, which is known as the piezoelectric effect.
sor nodes work collaboratively to monitor mechanical stress for The induced electric voltage results in additional internal torque
seismic and infrastructure safety, and humidity and temperature on the piezoelectric flags due to the inverse piezoelectric effect.
for agriculture applications [124]. Those distributive wireless For piezoelectric materials with high electromechanical coupling
sensor nodes require autonomous power ranging from micro- coefficient, such as lead zirconate titanate (PZT), the electro-
watts (lW) to milliwatts (mW) that can be adequately satisfied mechanical coupling model must be integrated with the aforemen-
by energy harvesting. Energy harvesting based on fluid-induced tioned fluid-structure coupling model to account for the inverse
flag vibrations converts kinetic energy of geophysical flows piezoelectric effect, thus forming a full fluid-structure-electricity
(wind, river, and oceanic or tidal currents), or the flow in indus- coupling model [55,128], while the inverse piezoelectric effect is
try ducts/pipes into vibrational energy, which is in turn con- negligible for materials with low electromechanical coupling
verted to electricity via piezoelectric, electromagnetic, or coefficient, such as polyvinylidene fluoride. Electromagnetic con-
electrostatic conversion. The flag is equivalent to an alternating version utilizes relative motions between a conductor and a mag-
current source in parallel with an intrinsic impedance of the con- net, which is displacement-based. The conductor or magnet is
verter, as schematically shown in Fig. 9(a). Power management attached to the vibrating flag [129] or oscillating flagpole [130].
(or conditioner) is used to enhance the efficiency and regulate The conversion efficiency is normally larger than that of piezo-
power in order to satisfy the requirements of wireless sensor electric conversion, but results in stronger dissipation of the vibra-
nodes [125–127]. Regulated power can be used directly by wire- tional energy, which can be modeled as structural damping using
less sensor nodes or stored in an energy reservoir before being the Kelvin-Voigt model [131]. The increase in structural damping
sent to sensor nodes. Energy harvesting based on fluid-induced decreases the critical KB value, reduces A/L and StA , and tends to
flag vibrations is environmentally friendly, incurs low mainte- transform a higher order vibration mode to a low-order vibration
nance costs, and does not require frequent battery replacements mode. Electrostatic conversion has been exploited in special
in sensor nodes, especially in difficult-to-access ocean environ- circumstances, such as energy harvesting in water using ionic
ments or remote areas. polymer metal composites, which are electroactive polymers

Fig. 9 Applications of fluid-induced flag vibrations: (a) energy harvesting, (b) heat transfer
enhancement, and (c) mixing enhancement

010801-12 / Vol. 71, JANUARY 2019 Transactions of the ASME


(EAPs) [97]. For energy harvesting in narrow channel flow, suc- techniques add inserters (twisted tapes, wire coils, etc.) or protru-
cessive contacts between the triboelectric materials on the flags sions (ribs, pins, fins, dimples, etc.) to promote turbulence and dis-
and lateral walls lead to the exchange of electric charges by tribo- rupt the boundary layers at the channel walls [139]. However,
electricity and electrostatic polarization of the channel walls poorer heat transfer performance is found within low speed recir-
[100,101]. More information on smart materials and conversion culation zones around rigid inserters or protrusions. The fluid-
from mechanical energy to electricity are provided in Invernizzi induced flag vibrations generate vortices, as shown in Figs. 4 and
et al. [132]. 5, which significantly promote the turbulent kinetic energy (TKE)
The harvested electrical energy is preliminarily mediated by the of flow. Those vortices interact with the boundary layers and
quality and quantity of the vibrational energy of flags. Vibrational entrain the heated flow near the walls into the cold main stream
energy is optimized through the proper selection of IIE, MIE, and without stable recirculation zones close to the walls, as shown in
EIE, depending on the fluid kinetic energy and conversion Fig. 6. Thus, the vibrating flags improve heat transfer
method. The inverted flag shows significant advantages for energy performance.
harvesting [6,93], including a high critical KB value that is inde- The heat transfer performance evaluation criteria are the Nus-
pendent of Ms (see Fig. 2), and a large A/L value (see Fig. 3) that selt ratio at constant Re and the Nusselt ratio at constant pumping
results in strain energy ranging from Oð102 Þ [22] to Oð103 Þ [75] power. The Nusselt ratio at constant Re quantifies the increased
times that of the standard flag. Furthermore, the symmetric vibra- heat flux due to flag vibration, which is defined as
tion in flapping mode is plausible for piezoelectric conversion due
to the absence of inflection points. As a result, the mechanical Nu
NuRe ¼ (15)
strain along the flag contributes to the total electricity output for a Nu0;Re
continuous piezoelectric layer. However, the flapping mode
occurs in a limited range of KB values, the flag transfers to the where Nu is the Nusselt number in the presence of vibrating flag,
deflected mode at low KB values. Thus, the inverted flag is pre- and the subscripts 0 and Re denote the parameter in plain channel
ferred in both wind and water flow, where the fluid has low yet at constant Re. The pressure expenditure induced by the vibrating
stable kinetic energy, such as occurs in industrial ducts or pipe- flag is taken into account in the Nusselt ratio at constant pumping
lines and ocean currents. Although A/L is lower than that of the power, which is often referred to as the thermal enhancement
inverted flag, the standard flag is robust for a wide range of KB factor [11,140]. This is given by
values beyond the critical value. As positive and negative strains
for mode 2 or higher order vibration modes would generate Nu
opposite charges with continuous piezoelectric layer, piezoelectric Nup ¼ ; (16)
Nu0;p
polymers are segmented in order to prevent cancelation of the
positive and negative charges [78,133]; positions and dimensions where the subscripts 0 and p denote the parameter in plain channel
of the piezoelectric electrodes are optimized [134]. For at constant pumping power. For constant pumping power, Re and
displacement-based electromagnetic conversion, only two embed- the friction factor F satisfy the following relationship:
ded conductor segments in flag are necessary for the vibration in
mode 2, and the higher order modes require several segments due  1=3
Re F
to the presences of more displacement phases along flag. Simi- ¼ (17)
larly, the electrodes on the lateral channel walls must be seg- Re0;p F0;p
mented for high order modes in the standard flag for electrostatic
conversion. Otherwise, the capacitance variations are weak In general, Nu and F in plain channel are related to Re in the fol-
because there is always a contact between the triboelectric flag lowing forms:
and continuous electrodes on the lateral walls [101]. The standard  a  b
flag is favorable in wind rather than in water, since the critical KB Nu0;p Re0;p F0;p Re0;p
¼ ; ¼ (18)
value is unaffordably small for a low Ms value. In contrast, the Nu0;Re Re F0;Re Re
forced flag shows a profound impact for low Ms values due to the
relatively low flag damping with respect to the upstream vortices; where the exponents a and b are fitted using the experimental
hence, the flag generates vibration with higher StA values but results. Eliminating Nu0;p in Eq. (16) using Eqs. (17) and (18)
lower A/L values than that in IIE and MIE, as shown in Fig. 3 [8]. yields
The forced flag also demonstrates better energy-harvesting feasi-
  a
bility for high KB values, such as the stiff PZT flag in Fig. 2(j) Nu F 3þb
[80] and in Fig. 2(k) [10,135]. The low A/L value for the forced Nup ¼ (19)
Nu0;Re F0;Re
flag protects the brittle PZT from damage [7]. Alternatively, a
nonpiezoelectric flag with high KB in EIE is clamped between two
flextensional frames through which the forces are amplified along For fully developed (hydrodynamically and thermally) turbulent
certain axis to compress the PZT stacks and harvest power up to flow in smooth channel, a ¼ 0:8 in the well-known Nusselt num-
50 mW [136,137]. Because the instability originates from the fluc- ber correlation with the Dittus–Boelter equation, and b ¼ 0:25
tuations in fluid, EIE is adaptive for unsteady incoming flow with due to the friction factor correlation in the modified Blasius
strong kinetic energy variations. The power of a single harvester   0:29
equation [141], i.e., Nup ¼ Nu=Nu0;Re F=F0;Re . In prac-
is generally 1 lW10 mW, where multiple flags in tandem or par-   1=3
allel configurations (discussed in Sec. 7) form power plant to meet tice, the approximate form Nu=Nu0;Re F=F0;Re is com-
the requirements for wireless sensor nodes with higher power con- monly used to evaluate Nup for other flow conditions
sumptions [5,138]. [12,30,142,143].
Park et al. [12] and Shoele and Mittal [11] numerically coupled
the thermal energy equation with the aforementioned
8.2 Heat Transfer Enhancement. Convective heat transfer fluid–structure interaction equations in Sec. 2 and modeled heat
enhancement in channel flow is commonly required in cooling transfer in IIE and MIE under laminar flow with parabolic incom-
systems for electronics, heating and refrigeration engineering ing velocity profiles. The inverted and standard flags are installed
fields, and process industries in order to prevent thermal damage in the middle of the channel, with their tip-to-tip vibration ampli-
and ensure proper operation of systems. Convective heat transfer tudes close to the width of the channel. In the abovementioned
is limited by the air-side heat transfer resistance, which can be evolution of the 2P þ 2S mode of the inverted flag in channel
improved through flow manipulation. The traditional passive flow, the opposite-signed LEV and TEV induce heated fluid near

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-13


the walls to move into the middle of the channel due to the rota- immiscible liquids, and dispersing a gas phase in a liquid phase.
tional momenta of those vortical structures. The induced vortices The flows can be reactants or catalysts, where mixing is accompa-
can also be attributed to the interruption and redevelopment of nied by chemical reactions or heat transfer in multifunctional
thermal boundary layers. In this vortex mode, NuRe reaches up to mixers/reactors/heat exchangers. The mixing enhancement due to
2.5 at Re ¼ 800, resulting in an optimal value of Nup ¼ 1:2 [12]. the fluid-induced flag vibrations is promising for commercial
Comparable heat transfer performances for NuRe ¼ 2 and Nup ¼ applications in process industries. However, the fluid-induced flag
1:25 occur at Re ¼ 800 when the reverse von Karman vortex street vibrations in those scenarios are unexplored and would be inter-
is well-organized in the wake of the standard flag in vibration esting avenues for future research.
mode 2 [11]. The higher order vibration mode of the standard flag
causes higher energy transfer to the flag; hence, Nup declines.
Heat transfer due to vibrating flags installed in the middle of 9 Conclusions
turbulent channel flow has been experimentally investigated
This paper examined three representative mechanisms of fluid-
recently by two research groups, with the inverted flag at
induced flag vibrations, namely, IIE of the inverted flag due to
Re  Oð104 Þ [79,94] and the standard flag at Re  Oð103 Þ
fluid flow, MIE of the standard flag, and EIE of flag behind a bluff
[144,145]. The value of NuRe increased from 1.08 to 1.31 as C
body. The instability boundary, vibration dynamics, and vortex
increased from 0.125 to 0.5. The evolved 2P mode at C ¼ 0:25,
dynamics of the flags are compared to summarize the characteris-
as shown in Fig. 6, leads to a maximum value of Nup ¼ 0:97
tic features and laws describing the distinct mechanisms of fluid-
among the vortex modes of the inverted flag. Mixing between the
induced vibrations and their constitutive relations. In addition, the
core flow and the near-wall flow is weak for low C ¼ 0:125,
three-dimensional effects and Reynolds number effects are dis-
which does not exhibit apparent benefits in heat removal from the
cussed. After that, the interactions of multiple inverted flags and
walls. A value of C greater than 0.25 induces large pressure loss
standard flags are addressed, where the interactions of tandem
that reduces Nup [79,94]. The reverse von Karman vortex street of
flags and parallel flags received particular focus as building blocks
the standard flag results in NuRe  eT and reduces Nup to approxi-
that facilitate understanding of more complicated arrangements.
mately 0.9 [144,145]. NuRe in turbulent flow is lower than the
Finally, the applications of fluid-induced flag vibrations are
value in laminar flow because the heat flux in plain turbulent flow
reviewed. Energy harvesting concerns flag vibration dynamics
is much higher, which leads to Nup < 1 due to the disproportion-
and has been explored extensively. Heat transfer enhancement
ate energy loss. Nup can be elevated through direct modulation of
and mixing enhancement mainly concern vortex dynamics of
the boundary layers using short flags close to the channel wall, as
flags, which have attracted more attention recently. The following
shown in Fig. 9(b), which effectively reduces energy loss com-
main conclusions are highlighted:
pared to the flags with high C values installed in the middle of
channel. In a long channel, a series of tandem flags disrupt the (1) The inverted flag is unstable, where the instability is
redeveloped thermal boundary layers farther downstream from the determined by KB but is independent of Ms . In contrast, the
flags. critical KB value for the standard flag decreases as Ms
decreases, which is accompanied by the emergences of
higher order modes. For the stable regions of IIE and MIE,
8.3 Mixing Enhancement. Flags with high C values in a flag can still be forced into vibration using an upstream
channel flow have potential applications in mixing enhancement bluff body, i.e., EIE.
in process industries, such as chemical reactions, pharmaceutical, (2) The inverted flag under IIE has the largest vibration ampli-
and water treatment [146]. Pressure loss is not the major concern; tude A/L up to around 1.6–1.9. The dimensionless fre-
instead, energy is pumped into the system in order to reduce quency StA first increases with the vibration amplitude and
inhomogeneities in concentration, density, composition, or tem- subsequently decreases when the vibration amplitude satu-
perature using in-line mixers in pipes or ducts, thereby improving rates. For MIE, the evolution of StA with vibration ampli-
product quality and consistency in continuous processes. Fluid- tude A/L is featured by a constant slope around StL ¼ 0.23
induced flag vibrations serve as passive agitators or passive in vibration mode 2. StL and StA increase for higher order
stirrers in place of mechanically driven agitators. In contrast to modes as KB or Ms decreases, while A/L decreases due to
impellers, rotor–stators, or actuators, fluid-induced vibrating flags the augmented inflections along the flag, and the standard
do not require external power sources and can be easily manufac- flag is finally driven to chaotic vibration for ultralow KB in
tured at low cost. the postcritical region. The vibration amplitude reduces to
Mixing is a multiscale phenomenon. In general, macro, meso, A/L < 0.8 for the forced flag. The (A/L, StA ) values lie
and micromixing are evaluated by the residence time distribution, below ðStL  D=LÞ=StD ¼ 1 at large KB values due to the
TKE, and dissipation of TKE, respectively [147]. The flags per- high damping effect of flag, and the evolution of StA with
turb the fluid path, and they promote secondary transverse and A/L falls on the line ðStL  D=LÞ=StD ¼ 1 as KB decreases,
longitudinal flow in unidirectional channel flow, thereby increas- which is characterized by lock-in of the flag vibration fre-
ing the residence time distribution. Vortices are efficient means of quency at the vortex shedding frequency from the upstream
mixing intensification [148], where the vortices generated by flags bluff body.
considerably enhance the TKE, i.e., meso-mixing [79,94]. Ali (3) The vortex dynamics depend on the dimensionless fre-
et al. [13] numerically studied the mixing enhancement due to quency StA of flags. More LEV and TEV shed from the
instability-induced excitation of flags. The inlet flow is divided inverted flag under IIE in every cycle as StA decreases. For
into two equal parts with two different scalar values. The flags MIE, the tail of the standard flag spins off strong vorticity
resemble the inverted flags but have nonzero initial angle toward at maximum displacements, and the von Karman vortex
the free stream and are fixed on channel walls. They found that street forms in the wake of the standard flag for the mode 2.
the vortices behind vibrating flags improved homogeneity of the Kelvin–Helmholtz vortices may also be generated along
mixture with scalar distribution beyond 97% compared to their the shear layer connected with the von Karman vortex
rigid counterparts. Tandem flags vibrating with phase lags, as dis- street when the velocity differences along the two sides of
cussed in Sec. 6, facilitate refined mixing scales. Figure 9(c) sche- the flag are high, thus causing a procession of the same-
matically shows the mixing of two streams in a Y-shaped channel, signed vortices in every stroke. The higher order vibration
as indicated by the dashed arrows and the solid arrows. As the two mode splits the vortices induced by the vibrating tail of the
streams meander through the tandem vibrating flags, mixing flag into small same-signed vortices, which normally corre-
becomes more homogeneous, eventually producing uniform flow. sponds to a larger StA value. The standard flag behind the
The flags are capable of enhancing mixing of gases, miscible or bluff body of EIE follows the unsteady pressure field

010801-14 / Vol. 71, JANUARY 2019 Transactions of the ASME


created by the upstream vortex shedding and vibrates with concentrations near its mounting region [137]. The fatigue prob-
its wavelength and frequency close to that of vortex shed- lem has been alleviated by optimizing the mounting design. Alter-
ding for a low Ms value at KB < 0:05. natively, the flags can be designed independently of piezoelectric
(4) Confinement by channel walls has destabilizing effects on fatigue limits by introducing novel designs such as the flexten-
flags, which elevates the critical KB and StA values. The sional actuators whereby the flag transfers vibrations to the hous-
vortex dynamics are strongly dependent on the flag’s ing PZT stacks. It is necessary to study the reliability of the
length-to-channel-width ratio C . The vortex mode evolves standard, inverted and forced flags with various KB and Ms for
with C in channel flow. Apart from the vortices generated long-term applications.
from the interaction between flag and fluid flow, vortices
induced close to the channel walls are demonstrated.
(5) Flags with finite spans are stabilized by the effect of side Acknowledgment
edges due to the reduced pressure difference near the edges.
Obvious decreases in the critical KB , A/L, and StA values We would like to thank our colleague Yujia Chen for the fruit-
occur as H  decreases in the range of H  < 1:5. However, ful discussions.
the influence of the side edge effect on the instability
boundary and flag vibration dynamics are negligible for Funding Data
H 2.
(6) Low Reynolds number flow has stabilizing effects on flags,
National Outstanding Youth Foundation of China (11725209).
while fluid-induced flag vibrations become insensitive to
Reynolds number when Re ⲏ 200 for IIE and when Re ⲏ
1000 for MIE. References
(7) In tandem flags, the downstream flag tends to synchronize [1] Watanabe, Y., Suzuki, S., Sugihara, M., and Sueoka, Y., 2002, “An Experi-
with the vibration frequency of the upstream flag, which mental Study of Paper Flutter,” J. Fluids Struct., 16(4), pp. 529–542.
[2] Huang, L., 1995, “Flutter of Cantilevered Plates in Axial Flow,” J. Fluids
causes the phase difference to have a linear dependence on Struct., 9(2), pp. 127–147.
separation distance, both for the standard and inverted [3] Tang, D., Yamamoto, H., and Dowell, E. H., 2003, “Flutter and Limit Cycle
flags. The vibration amplitude of the upstream flag is Oscillations of Two-Dimensional Panels in Three-Dimensional Axial Flow,”
smaller than that of an isolated flag at low separation dis- J. Fluids Struct., 17(2), pp. 225–242.
[4] Guo, C. Q., and Paidoussis, M. P., 2000, “Stability of Rectangular Plates With
tance, and the amplitude increases asymptotically to that of Free Side-Edges in Two-Dimensional Inviscid Channel Flow,” ASME J.
an isolated flag as the separation distance increases. Merg- Appl. Mech., 67(1), pp. 171–176.
ing of the vortices from the downstream flag with the vorti- [5] Yu, Y., and Liu, Y., 2016, “Energy Harvesting With Two Parallel Pinned Pie-
ces from the upstream flag with the same rotational sense zoelectric Membranes in Fluid Flow,” J. Fluids Struct., 65, pp. 381–397.
[6] Orrego, S., Shoele, K., Ruas, A., Doran, K., Caggiano, B., Mittal, R., and
leads to constructive coupling, where the vibration ampli- Kang, S. H., 2017, “Harvesting Ambient Wind Energy With an Inverted Pie-
tude of the downstream flag increases. In comparison, zoelectric Flag,” Appl. Energy, 194, pp. 212–222.
destructive merging of vortices reduces the vibration ampli- [7] Kim, H., Kang, S., and Kim, D., 2017, “Dynamics of a Flag Behind a Bluff
tude of the downstream flag. Body,” J. Fluids Struct., 71, pp. 1–14.
[8] Allen, J., and Smits, A., 2001, “Energy Harvesting Eel,” J. Fluids Struct.,
(8) In-phase and anti-phase vibrations are the predominant 15(3–4), pp. 629–640.
results of coupling between two parallel flags. [9] Taylor, G. W., Burns, J. R., Kammann, S., Powers, W. B., and Welsh, T. R.,
2001, “The Energy Harvesting Eel: A Small Subsurface Ocean/River Power
Future researches should focus in four areas: Generator,” IEEE J. Oceanic Eng., 26(4), pp. 539–547.
First, researchers should explore fluid-induced flag vibrations [10] Liu, H., Zhang, S., Kathiresan, R., Kobayashi, T., and Lee, C., 2012,
“Development of Piezoelectric Microcantilever Flow Sensor With Wind-
under complex incoming flow patterns. Mixing enhancement in Driven Energy Harvesting Capability,” Appl. Phys. Lett., 100(22), p. 223905.
process industries involves incoming fluid with inhomogeneous [11] Shoele, K., and Mittal, R., 2014, “Computational Study of Flow-Induced
concentration and velocity distributions, various species of gases Vibration of a Reed in a Channel and Effect on Convective Heat Transfer,”
or liquids, dispersion of particles or gases in a liquid phase, etc. Phys. Fluids, 26(12), p. 127103.
[12] Park, S. G., Kim, B., Chang, C. B., Ryu, J., and Sung, H. J., 2016,
Those complex incoming flows definitely influence the instability “Enhancement of Heat Transfer by a Self-Oscillating Inverted Flag in a Pois-
boundary, vibration dynamics, and vortex dynamics of flags, euille Channel Flow,” Int. J. Heat Mass Transfer, 96, pp. 362–370.
which in turn determine their performance on mixing enhance- [13] Ali, S., Habchi, C., Menanteau, S., Lemenand, T., and Harion, J.-L., 2015,
ment. Fluid-induced flag vibrations are promising solutions to “Heat Transfer and Mixing Enhancement by Free Elastic Flaps Oscillation,”
Int. J. Heat Mass Transfer, 85, pp. 250–264.
mixing enhancement in process industries, which are of great [14] M€ uller, U. K., 2003, “Fish ‘n Flag,” Science, 302(5650), pp. 1511–1512.
significance but lack investigations to date. The performance of [15] Liao, J. C., Beal, D. N., Lauder, G. V., and Triantafyllou, M. S., 2003, “Fish
the passive stirrer as an in-line mixer requires quantitative Exploiting Vortices Decrease Muscle Activity,” Science, 302(5650), pp.
examinations. 1566–1569.
[16] Kim, S., Huang, W.-X., and Sung, H. J., 2010, “Constructive and Destructive
Second, flag vibration dynamics have been extensively studied, Interaction Modes Between Two Tandem Flexible Flags in Viscous Flow,”
while the resulting vortex dynamics are an active area of research. J. Fluid Mech., 661, pp. 511–521.
In particular, the vortex dynamics of the inverted flag are abun- [17] Naudascher, E., and Rockwell, D., 2012, Flow-Induced Vibrations: An Engi-
dant, and detailed studies into the vortex dynamics would cause neering Guide, Dover Publications, New York.
[18] Paidoussis, M. P., 1998, Fluid-Structure Interactions: Slender Structures and
the vibrating flag to become the standard vortex generator in fluid Axial Flow, Academic Press, San Diego, CA.
engineering. Moreover, the dependence of the vortex convective [19] Kim, D., Cosse, J., Huertas Cerdeira, C., and Gharib, M., 2013, “Flapping
speed, wavelength, or vortex spacing of the wake on KB would Dynamics of an Inverted Flag,” J. Fluid Mech., 736, p. R1.
shed light on the interaction between tandem flags. [20] Sader, J. E., Cosse, J., Kim, D., Fan, B., and Gharib, M., 2016, “Large-
Amplitude Flapping of an Inverted Flag in a Uniform Steady Flow—A
Third, experiments into fluid-induced flag vibrations in laminar Vortex-Induced Vibration,” J. Fluid Mech., 793, pp. 524–555.
flow are highly desirable. The aforementioned numerical studies [21] Goza, A., Colonius, T., and Sader, J. E., 2018, “Global Modes and Nonlinear
suggest significant heat transfer enhancement of vibrating flags in Analysis of Inverted-Flag Flapping,” J. Fluid Mech., 857, pp. 312–344.
laminar flow. While experiments have been performed in the [22] Ryu, J., Park, S. G., Kim, B., and Sung, H. J., 2015, “Flapping Dynamics of an
Inverted Flag in a Uniform Flow,” J. Fluids Struct., 57, pp. 159–169.
range of Re ¼ Oð103 Þ  Oð105 Þ, experimental data are lacking to [23] Weaver, D., 1976, “On Flow Induced Vibrations in Hydraulic Structures and
validate the numerical results in laminar flow. Their Alleviation,” Can. J. Civ. Eng., 3(1), pp. 126–137.
Finally, few efforts have been made to determine the fatigue [24] Taneda, S., 1968, “Waving Motions of Flags,” J. Phys. Soc. Jpn., 24(2), pp.
responses of cyclically vibrating flags under fluid load. The finite 392–401.
[25] Zhang, J., Childress, S., Libchaber, A., and Shelley, M., 2000, “Flexible Fila-
element analysis of brittle PZT bimorph flag showed fatigue ments in a Flowing Soap Film as a Model for One-Dimensional Flags in a
failures after 3.9104 vibration cycles due to the stress Two-Dimensional Wind,” Nature, 408(6814), pp. 835–839.

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-15


[26] Yu, Y., and Liu, Y., 2015, “Flapping Dynamics of a Piezoelectric Membrane [61] Zhao, W., Pa€ıdoussis, M. P., Tang, L., Liu, M., and Jiang, J.,
Behind a Circular Cylinder,” J. Fluids Struct., 55, pp. 347–363. 2012, “Theoretical and Experimental Investigations of the Dynamics of
[27] Shi, S., New, T., and Liu, Y., 2013, “Flapping Dynamics of a Low Aspect- Cantilevered Flexible Plates Subjected to Axial Flow,” J. Sound Vib., 331(3),
Ratio Energy-Harvesting Membrane Immersed in a Square Cylinder Wake,” pp. 575–587.
Exp. Therm. Fluid Sci., 46, pp. 151–161. [62] Dunnmon, J., Stanton, S., Mann, B., and Dowell, E., 2011, “Power Extraction
[28] Pan, D., Shao, X., Deng, J., and Yu, Z., 2014, “Simulations of Passive Oscilla- From Aeroelastic Limit Cycle Oscillations,” J. Fluids Struct., 27(8), pp.
tion of a Flexible Plate in the Wake of a Cylinder by Immersed Boundary 1182–1198.
Method,” Eur. J. Mech. B, 46, pp. 17–27. [63] Gibbs, S. C., Wang, I., and Dowell, E., 2012, “Theory and Experiment for
[29] Shelley, M. J., and Zhang, J., 2011, “Flapping and Bending Bodies Interacting Flutter of a Rectangular Plate With a Fixed Leading Edge in Three-
With Fluid Flows,” Annu. Rev. Fluid Mech., 43(1), pp. 449–465. Dimensional Axial Flow,” J. Fluids Struct., 34, pp. 68–83.
[30] Gallegos, R. K. B., and Sharma, R. N., 2017, “Flags as Vortex Generators for [64] Goza, A., and Colonius, T., 2017, “A Global Mode Analysis of Flapping
Heat Transfer Enhancement: Gaps and Challenges,” Renewable Sustainable Flags,” 10th International Symposium on Turbulence and Shear Flow Phe-
Energy Rev., 76, pp. 950–962. nomena (TSFP10), Chicago, IL, July 6–9.
[31] Connell, B. S., 2006, Numerical Investigation of the Flow-Body Interaction of [65] Kornecki, A., Dowell, E., and O’Brien, J., 1976, “On the Aeroelastic Instabil-
Thin Flexible Foils and Ambient Flow, Massachusetts Institute of Technology, ity of Two-Dimensional Panels in Uniform Incompressible Flow,” J. Sound
Cambridge, MA. Vib., 47(2), pp. 163–178.
[32] Paidoussis, M. P., 2004, Fluid-Structure Interactions: Slender Structures and [66] Jia, L.-B., Li, F., Yin, X.-Z., and Yin, X.-Y., 2007, “Coupling Modes Between
Axial Flow, Vol. 2, Elsevier Academic Press, New York. Two Flapping Filaments,” J. Fluid Mech., 581(1), pp. 199–220.
[33] Connell, B. S., and Yue, D. K., 2007, “Flapping Dynamics of a Flag in a Uni- [67] Argentina, M., and Mahadevan, L., 2005, “Fluid-Flow-Induced Flutter of a
form Stream,” J. Fluid Mech., 581, pp. 33–67. Flag,” Proc. Natl. Acad. Sci. U. S. A., 102(6), pp. 1829–1834.
[34] Tang, L., and Paidoussis, M. P., 2007, “On the Instability and the Post-Critical [68] Lemaitre, C., Hemon, P., and de Langre, E., 2005, “Instability of a Long Rib-
Behaviour of Two-Dimensional Cantilevered Flexible Plate’s in Axial Flow,” bon Hanging in Axial Air Flow,” J. Fluids Struct., 20(7), pp. 913–925.
J. Sound Vib., 305(1–2), pp. 97–115. [69] Datta, S. K., and Gottenberg, W. G., 1975, “Instability of an Elastic Strip
[35] Alben, S., 2009, “Simulating the Dynamics of Flexible Bodies and Vortex Hanging in an Airstream,” ASME J. Appl. Mech., 42(1), pp. 195–198.
Sheets,” J. Comput. Phys., 228(7), pp. 2587–2603. [70] De Nayer, G., Kalmbach, A., Breuer, M., Sicklinger, S., and W€ uchner, R.,
[36] Eloy, C., Lagrange, R., Souilliez, C., and Schouveiler, L., 2008, “Aeroelastic 2014, “Flow Past a Cylinder With a Flexible Splitter Plate: A Complementary
Instability of Cantilevered Flexible Plates in Uniform Flow,” J. Fluid Mech., Experimental–Numerical Investigation and a New FSI Test Case (FSI-PfS-
611, pp. 97–106. 1a),” Comput. Fluids, 99, pp. 18–43.
[37] Huang, W.-X., and Sung, H. J., 2010, “Three-Dimensional Simulation of a [71] H€ubner, B., Walhorn, E., and Dinkler, D., 2004, “A Monolithic Approach to
Flapping Flag in a Uniform Flow,” J. Fluid Mech., 653, pp. 301–336. Fluid–Structure Interaction Using Space–Time Finite Elements,” Comput.
[38] Eloy, C., Kofman, N., and Schouveiler, L., 2012, “The Origin of Hysteresis in Methods Appl. Mech. Eng., 193(23–26), pp. 2087–2104.
the Flag Instability,” J. Fluid Mech., 691(1), pp. 583–593. [72] Manela, A., and Howe, M., 2009, “The Forced Motion of a Flag,” J. Fluid
[39] Alben, S., and Shelley, M. J., 2008, “Flapping States of a Flag in an Inviscid Mech., 635, pp. 439–454.
Fluid: Bistability and the Transition to Chaos,” Phys. Rev. Lett., 100(7), [73] Gibbs, S. C., Fichera, S., Zanotti, A., Ricci, S., and Dowell, E. H., 2014, “Flow
p. 074301. Field Around the Flapping Flag,” J. Fluids Struct., 48, pp. 507–513.
[40] Eloy, C., Souilliez, C., and Schouveiler, L., 2007, “Flutter of a Rectangular [74] Shang-Rou, H., Shaw, S. W., and Pierre, C., 1994, “Normal Modes for Large
Plate,” J. Fluids Struct., 23(6), pp. 904–919. Amplitude Vibration of a Cantilever Beam,” Int. J. Solids Struct., 31(14), pp.
[41] Tang, L., 2008, “The Influence of the Wake on the Stability of Cantilevered 1981–2014.
Flexible Plates in Axial Flow,” J. Sound Vib., 310(3), pp. 512–526. [75] Gurugubelli, P. S., and Jaiman, R. K., 2015, “Self-Induced Flapping Dynamics
[42] Tang, C., Liu, N.-S., and Lu, X.-Y., 2015, “Dynamics of an Inverted Flexible of a Flexible Inverted Foil in a Uniform Flow,” J. Fluid Mech., 781, pp.
Plate in a Uniform Flow,” Phys. Fluids, 27(7), p. 073601. 657–694.
[43] Gilmanov, A., Le, T. B., and Sotiropoulos, F., 2015, “A Numerical Approach [76] Souilliez, C., Eloy, C., and Schouveiler, L., 2006, “An Experimental Study of
for Simulating Fluid Structure Interaction of Flexible Thin Shells Undergoing Flag Flutter,” ASME Paper No. PVP2006-ICPVT-11-93864.
Arbitrarily Large Deformations in Complex Domains,” J. Comput. Phys., 300, [77] Shelley, M., Vandenberghe, N., and Zhang, J., 2005, “Heavy Flags Undergo
pp. 814–843. Spontaneous Oscillations in Flowing Water,” Phys. Rev. Lett., 94(9),
[44] Huang, H., Wei, H., and Lu, X.-Y., 2017, “Coupling Performance of Tandem p. 094302.
Flexible Inverted Flags in a Uniform Flow,” J. Fluid Mech., 837, pp. 461–476. [78] Doare, O., and Michelin, S., 2011, “Piezoelectric Coupling in Energy-
[45] Goza, A., and Colonius, T., 2017, “A Strongly-Coupled Immersed-Boundary Harvesting Fluttering Flexible Plates: Linear Stability Analysis and Conver-
Formulation for Thin Elastic Structures,” J. Comput. Phys., 336, pp. 401–411. sion Efficiency,” J. Fluids Struct., 27(8), pp. 1357–1375.
[46] Lee, I., and Choi, H., 2015, “A Discrete-Forcing Immersed Boundary Method [79] Yu, Y., Liu, Y., and Chen, Y., 2018, “Vortex Dynamics and Heat Transfer
for the Fluid–Structure Interaction of an Elastic Slender Body,” J. Comput. Behind Self-Oscillating Inverted Flags of Various Lengths in Channel Flow,”
Phys., 280, pp. 529–546. Phys. Fluids, 30(4), p. 045104.
[47] Akcabay, D. T., and Young, Y. L., 2012, “Hydroelastic Response and Energy [80] Akaydin, H. D., Elvin, N., and Andreopoulos, Y., 2010, “Energy Harvesting
Harvesting Potential of Flexible Piezoelectric Beams in Viscous Flow,” Phys. From Highly Unsteady Fluid Flows Using Piezoelectric Materials,” J. Intell.
Fluids, 24(5), p. 054106. Mater. Syst. Struct., 21(13), pp. 1263–1278.
[48] Huang, W.-X., Shin, S. J., and Sung, H. J., 2007, “Simulation of Flexible Fila- [81] Techet, A. H., Allen, J. J., and Smits, A. J., 2002, “Piezoelectric Eels for
ments in a Uniform Flow by the Immersed Boundary Method,” J. Comput. Energy Harvesting in the Ocean,” 12th International Offshore and Polar Engi-
Phys., 226(2), pp. 2206–2228. neering Conference, Kitakyushu, Japan, May 26–31, pp. 713–718.
[49] Zhu, L., and Peskin, C. S., 2002, “Simulation of a Flapping Flexible Filament [82] Shi, S., New, T. H., and Liu, Y., 2014, “Effects of Aspect-Ratio on the Flap-
in a Flowing Soap Film by the Immersed Boundary Method,” J. Comput. ping Behaviour of Energy-Harvesting Membrane,” Exp. Therm. Fluid Sci.,
Phys., 179(2), pp. 452–468. 52, pp. 339–346.
[50] Banerjee, S., Connell, B. S., and Yue, D. K., 2015, “Three-Dimensional [83] De Nayer, G., and Breuer, M., 2014, “Numerical FSI Investigation
Effects on Flag Flapping Dynamics,” J. Fluid Mech., 783, pp. 103–136. Based on LES: Flow Past a Cylinder With a Flexible Splitter Plate Involv-
[51] Cisonni, J., Lucey, A. D., Elliott, N. S. J., and Heil, M., 2017, “The Stability of a ing Large Deformations (FSI-PfS-2a),” Int. J. Heat Fluid Flow, 50, pp.
Flexible Cantilever in Viscous Channel Flow,” J. Sound Vib., 396, pp. 186–202. 300–315.
[52] Balint, T. S., and Lucey, A. D., 2005, “Instability of a Cantilevered Flexible [84] Bao, C.-Y., Chao, T., Xie-Zhen, Y., and Xi-Yun, L., 2010, “Flutter of
Plate in Viscous Channel Flow,” J. Fluids Struct., 20(7), pp. 893–912. Finite-Span Flexible Plates in Uniform Flow,” Chin. Phys. Lett., 27(6),
[53] Michelin, S., Llewellyn Smith, S. G., and Glover, B. J., 2008, “Vortex Shed- p. 064601.
ding Model of a Flapping Flag,” J. Fluid Mech., 617, pp. 1–10. [85] Virot, E., Amandolese, X., and Hemon, P., 2013, “Fluttering Flags: An Experi-
[54] Howell, R., Lucey, A., Carpenter, P. W., and Pitman, M., 2009, “Interaction mental Study of Fluid Forces,” J. Fluids Struct., 43, pp. 385–401.
Between a Cantilevered-Free Flexible Plate and Ideal Flow,” J. Fluids Struct., [86] Bai, Y., Jia, Y. X., Lee, C., and Zhu, Y., 2016, “Experimental Study of a Peri-
25(3), pp. 544–566. odical Flapping Flag,” Acta Phys. Sin., 65(12), p. 124701.
[55] Wang, X., Alben, S., Li, C., and Young, Y. L., 2016, “Stability and Scalability [87] Abderrahmane, H. A., Paidoussis, M. P., Fayed, M., and Ng, H. D., 2011,
of Piezoelectric Flag,” Phys. Fluids, 28(2), p. 023601. “Flapping Dynamics of a Flexible Filament,” Phys. Rev. E, 84(6), p. 066604.
[56] Alben, S., 2015, “Flag Flutter in Inviscid Channel Flow,” Phys. Fluids, 27(3), [88] Virot, E., Faranda, D., Amandolese, X., and Hemon, P., 2017, “Chaotic
p. 033603. Dynamics of Flags From Recurring Values of Flapping Moment,” Int. J. Bifur-
[57] Shoele, K., and Mittal, R., 2016, “Flutter Instability of a Thin Flexible Plate in cation Chaos, 27(2), p. 1750020.
a Channel,” J. Fluid Mech., 786, pp. 29–46. [89] Abderrahmane, H. A., Pa€ıdoussis, M. P., Fayed, M., and Ng, H. D., 2012,
[58] Yamaguchi, N., Yokota, K., and Tsujimoto, Y., 2000, “Flutter Limits and “Nonlinear Dynamics of Silk and Mylar Flags Flapping in Axial Flow,” J.
Behaviors of a Flexible Thin Sheet in High-Speed Flow—I: Analytical Wind Eng. Ind. Aerodyn., 107, pp. 225–236.
Method for Prediction of the Sheet Behavior,” ASME J. Fluids Eng., 122(1), [90] Roshko, A., 1961, “Experiments on the Flow Past a Circular Cylinder at Very
pp. 65–73. High Reynolds Number,” J. Fluid Mech., 10(3), pp. 345–356.
[59] Chen, M., Jia, L.-B., Wu, Y.-F., Yin, X.-Z., and Ma, Y.-B., 2014, “Bifurcation [91] Knisely, C. W., 1990, “Strouhal Numbers of Rectangular Cylinders at Inci-
and Chaos of a Flag in an Inviscid Flow,” J. Fluids Struct., 45, pp. 124–137. dence: A Review and New Data,” J. Fluids Struct., 4(4), pp. 371–393.
[60] Attar, P., Dowell, E., and Tang, D., 2003, “Modeling Aerodynamic Nonlinear- [92] Shukla, S., Govardhan, R., and Arakeri, J., 2013, “Dynamics of a Flexible
ities for Two Aeroelastic Configurations: Delta Wing and Flapping Flag,” Splitter Plate in the Wake of a Circular Cylinder,” J. Fluids Struct., 41, pp.
AIAA Paper No. 2003-1402. 127–134.

010801-16 / Vol. 71, JANUARY 2019 Transactions of the ASME


[93] Shoele, K., and Mittal, R., 2016, “Energy Harvesting by Flow-Induced Flutter [123] Zhai, J., Zhang, W., Gao, C., Zhang, Y., Ye, Z., and Wang, H., 2016, “Side
in a Simple Model of an Inverted Piezoelectric Flag,” J. Fluid Mech., 790, pp. Force Control on Slender Body by Self-Excited Oscillation Flag,” Theor.
582–606. Appl. Mech. Lett., 6(5), pp. 230–232.
[94] Yu, Y., Liu, Y., and Chen, Y., 2017, “Vortex Dynamics Behind a Self- [124] Sudevalayam, S., and Kulkarni, P., 2011, “Energy Harvesting Sensor Nodes:
Oscillating Inverted Flag Placed in a Channel Flow: Time-Resolved Particle Survey and Implications,” IEEE Commun. Surv. Tutorials, 13(3), pp. 443–461.
Image Velocimetry Measurements,” Phys. Fluids, 29(12), p. 125104. [125] Xia, Y., Michelin, S., and Doare, O., 2015, “Fluid-Solid-Electric Lock-In of
[95] Gurugubelli, P., and Jaiman, R., 2017, “On the Mechanism of Large Energy-Harvesting Piezoelectric Flags,” Phys. Rev. Appl., 3(1), p. 014009.
Amplitude Flapping of Inverted Foil in a Uniform Flow,” epub, [126] Ottman, G. K., Hofmann, H. F., Bhatt, A. C., and Lesieutre, G. A., 2002,
arXiv:1711.01065. “Adaptive Piezoelectric Energy Harvesting Circuit for Wireless Remote
[96] Lee, J. H., Huang, W.-X., and Sung, H. J., 2014, “Flapping Dynamics of a Power Supply,” IEEE Trans. Power Electron., 17(5), pp. 669–676.
Flexible Flag in a Uniform Flow,” Fluid Dyn. Res., 46(5), p. 055517. [127] Robbins, W. P., Morris, D., Marusic, I., and Novak, T. O., 2006, “Wind-
[97] Giacomello, A., and Porfiri, M., 2011, “Underwater Energy Harvesting From a Generated Electrical Energy Using Flexible Piezoelectric Materials,” ASME
Heavy Flag Hosting Ionic Polymer Metal Composites,” J. Appl. Phys., 109(8), Paper No. IMECE2006-14050.
p. 084903. [128] Michelin, S., and Doare, O., 2013, “Energy Harvesting Efficiency of Piezo-
[98] Jia, L., 2014, The Interaction Between Flexible Plates and Fluid in Two- electric Flags in Axial Flows,” J. Fluid Mech., 714, pp. 489–504.
Dimensional Flow, Springer, Berlin, Germany. [129] Tang, L., Pa€ıdoussis, M. P., and Jiang, J., 2009, “Cantilevered Flexible Plates
[99] Portaro, R., Fayed, M., Gunter, A.-L., Abderrahmane, H. A., and Ng, H. D., in Axial Flow: Energy Transfer and the Concept of Flutter-Mill,” J. Sound
2011, “Fractal Geometry of the Wake Shed by a Flapping Filament in Flowing Vib., 326(1–2), pp. 263–276.
Soap-Film,” Fractals, 19(3), pp. 311–316. [130] Virot, E., Amandolese, X., and Hemon, P., 2016, “Coupling Between a Flag
[100] Perez, M., Boisseau, S., Gasnier, P., Willemin, J., and Reboud, J. L., 2015, and a Spring-Mass Oscillator,” J. Fluids Struct., 65, pp. 447–454.
“An Electret-Based Aeroelastic Flutter Energy Harvester,” Smart Mater. [131] Chen, M., Jia, L. B., Wang, S. Y., and Yin, X. Z., 2014, “Effects of Material
Struct., 24(3), p. 035004. Damping on Flag Flutter,” Sci. China Technol. Sci., 57(1), pp. 117–127.
[101] Perez, M., Boisseau, S., Geisler, M., Gasnier, P., Willemin, J., Despesse, G., [132] Invernizzi, F., Dulio, S., Patrini, M., Guizzetti, G., and Mustarelli, P., 2016,
and Reboud, J. L., 2018, “Aeroelastic Flutter Energy Harvesters Self- “Energy Harvesting From Human Motion: Materials and Techniques,” Chem.
Polarized by Triboelectric Effects,” Smart Mater. Struct., 27(1), p. 014003. Soc. Rev., 45(20), pp. 5455–5473.
[102] Doare, O., Mano, D., and Carlos Bilbao Ludena, J., 2011, “Effect of Spanwise [133] Shan, X., Song, R., Fan, M., Deng, J., and Xie, T., 2016, “A Novel Method for
Confinement on Flag Flutter: Experimental Measurements,” Phys. Fluids, Improving the Energy Harvesting Performance of Piezoelectric Flag in a Uni-
23(11), p. 111704. form Flow,” Ferroelectrics, 500(1), pp. 283–290.
[103] Doare, O., Sauzade, M., and Eloy, C., 2011, “Flutter of an Elastic Plate in a [134] Pi~neirua, M., Doare, O., and Michelin, S., 2015, “Influence and Optimization
Channel Flow: Confinement and Finite-Size Effects,” J. Fluids Struct., 27(1), of the Electrodes Position in a Piezoelectric Energy Harvesting Flag,” J. Sound
pp. 76–88. Vib., 346, pp. 200–215.
[104] Hoepffner, J., and Naka, Y., 2011, “Oblique Waves Lift the Flapping Flag,” [135] Pobering, S., Ebermeyer, S., and Schwesinger, N., 2009, “Generation of Elec-
Phys. Rev. Lett., 107(19), p. 194502. trical Energy Using Short Piezoelectric Cantilevers in Flowing Media,” Proc.
[105] Sader, J. E., Huertas-Cerdeira, C., and Gharib, M., 2016, “Stability of Slender SPIE 7288, p. 728807.
Inverted Flags and Rods in Uniform Steady Flow,” J. Fluid Mech., 809, pp. [136] Lee, H. J., Sherrit, S., Tosi, L. P., and Colonius, T., 2016, “Design and Experi-
873–894. mental Evaluation of Flextensional-Cantilever Based Piezoelectric Trans-
[106] Alben, S., 2009, “Wake-Mediated Synchronization and Drafting in Coupled ducers for Flow Energy Harvesting,” Proc. SPIE 9806, p. 980610.
Flags,” J. Fluid Mech., 641, pp. 489–496. [137] Lee, H. J., Sherrit, S., Tosi, L. P., Walkemeyer, P., and Colonius, T., 2015,
[107] Ristroph, L., and Zhang, J., 2008, “Anomalous Hydrodynamic Drafting of “Piezoelectric Energy Harvesting in Internal Fluid Flow,” Sensors, 15(10), pp.
Interacting Flapping Flags,” Phys. Rev. Lett., 101(19), p. 194502. 26039–26062.
[108] Huertas-Cerdeira, C., Fan, B., and Gharib, M., 2018, “Coupled Motion of Two [138] Pobering, S., and Schwesinger, N., 2004, “A Novel Hydropower Harvesting
Side-by-Side Inverted Flags,” J. Fluids Struct., 76, pp. 527–535. Device,” International Conference on MEMS, NANO and Smart Systems,
[109] Ryu, J., Park, S. G., and Sung, H. J., 2018, “Flapping Dynamics of Inverted (ICMENS’04), Banff, AB, Canada, Aug. 25–27, pp. 480–485.
Flags in a Side-by-Side Arrangement,” Int. J. Heat Fluid Flow, 70, pp. [139] Dewan, A., Mahanta, P., Raju, K. S., and Kumar, P. S., 2004, “Review of Pas-
131–140. sive Heat Transfer Augmentation Techniques,” Proc. Inst. Mech. Eng., Part A,
[110] Farnell, D. J. J., David, T., and Barton, D. C., 2004, “Coupled States of Flap- 218(7), pp. 509–527.
ping Flags,” J. Fluids Struct., 19(1), pp. 29–36. [140] Promvonge, P., and Eiamsa-ard, S., 2007, “Heat Transfer Augmentation in a
[111] Zhu, L., and Peskin, C. S., 2003, “Interaction of Two Flapping Filaments in a Circular Tube Using V-Nozzle Turbulator Inserts and Snail Entry,” Exp.
Flowing Soap Film,” Phys. Fluids, 15(7), pp. 1954–1960. Therm. Fluid Sci., 32(1), pp. 332–340.
[112] Tang, L., and Pa€ıdoussis, M. P., 2009, “The Coupled Dynamics of Two Canti- [141] Bhagoria, J., Saini, J., and Solanki, S., 2002, “Heat Transfer Coefficient and
levered Flexible Plates in Axial Flow,” J. Sound Vib., 323(3–5), pp. 790–801. Friction Factor Correlations for Rectangular Solar Air Heater Duct Having
[113] Sun, C. B., Wang, S. Y., Jia, L. B., and Yin, X. Z., 2016, “Force Measurement Transverse Wedge Shaped Rib Roughness on the Absorber Plate,” Renewable
on Coupled Flapping Flags in Uniform Flow,” J. Fluids Struct., 61, pp. Energy, 25(3), pp. 341–369.
339–346. [142] Promvonge, P., and Thianpong, C., 2008, “Thermal Performance Assessment
[114] Wang, S.-Y., Duan, W.-G., and Yin, X.-Z., 2013, “Transition Mode of Two of Turbulent Channel Flows Over Different Shaped Ribs,” Int. Commun. Heat
Parallel Flags in Uniform Flow,” Chin. Phys. Lett., 30(11), p. 110502. Mass Transfer, 35(10), pp. 1327–1334.
[115] Schouveiler, L., and Eloy, C., 2009, “Coupled Flutter of Parallel Plates,” Phys. [143] Wang, L., and Sunden, B., 2002, “Performance Comparison of Some Tube
Fluids, 21(8), p. 081703. Inserts,” Int. Commun. Heat Mass Transfer, 29(1), pp. 45–56.
[116] Michelin, S., and Llewellyn Smith, S. G., 2009, “Linear Stability Analysis of [144] Jha, S., Crittenden, T., and Glezer, A., 2017, “Enhancement of Forced Con-
Coupled Parallel Flexible Plates in an Axial Flow,” J. Fluids Struct., 25(7), pp. vection Heat Transfer Using Aero-Elastically Fluttering Reeds,” 23rd IEEE
1136–1157. International Workshop on Thermal Investigations of ICs and Systems
[117] Jeong, Y. D., and Lee, J. H., 2017, “Passive Control of a Single Flexible (THERMINIC), Amsterdam, The Netherlands, Sept. 27–29.
Flag Using Two Side-by-Side Flags,” Int. J. Heat Fluid Flow, 65, pp. 90–104. [145] Crittenden, T., Jha, S., and Glezer, A., 2017, “Forced Convection Heat Transfer
[118] Tian, F.-B., Luo, H., Zhu, L., and Lu, X.-Y., 2011, “Coupling Modes of Three Enhancement in Heat Sink Channels Using Aeroelastically Fluttering Reeds,”
Filaments in Side-by-Side Arrangement,” Phys. Fluids, 23(11), p. 111903. 16th IEEE Intersociety Conference on Thermal and Thermomechanical Phenom-
[119] Tian, F. B., Luo, H., Zhu, L., Liao, J. C., and Lu, X. Y., 2011, “An Efficient ena in Electronic Systems (ITherm), Orlando, FL, May 30–June 2, pp. 114–121.
Immersed Boundary-Lattice Boltzmann Method for the Hydrodynamic Inter- [146] Anxionnaz, Z., Cabassud, M., Gourdon, C., and Tochon, P., 2008, “Heat
action of Elastic Filaments,” J. Comput. Phys., 230(19), pp. 7266–7283. Exchanger/Reactors (HEX Reactors): Concepts, Technologies: State-of-the-
[120] Uddin, E., Huang, W.-X., and Sung, H. J., 2013, “Interaction Modes of Multi- Art,” Chem. Eng. Process.: Process Intensif., 47(12), pp. 2029–2050.
ple Flexible Flags in a Uniform Flow,” J. Fluid Mech., 729, pp. 563–583. [147] Ghanem, A., Lemenand, T., Della Valle, D., and Peerhossaini, H., 2014,
[121] Son, Y., and Lee, J. H., 2017, “Flapping Dynamics of Coupled Flexible Flags “Static Mixers: Mechanisms, Applications, and Characterization Methods—A
in a Uniform Viscous Flow,” J. Fluids Struct., 68, pp. 339–355. Review,” Chem. Eng. Res. Des., 92(2), pp. 205–228.
[122] Zhang, W., Liu, X., Zhai, J., and Ye, Z., 2012, “Experimental Study on Side [148] Habchi, C., Lemenand, T., Della Valle, D., and Peerhossaini, H., 2016,
Force Alleviation of Conical Forebody With a Fluttering Flag,” Phys. Fluids, “Turbulence Statistics Downstream of a Vorticity Generator at Low Reynolds
24(12), p. 124105. Numbers,” Phys. Fluids, 28(10), p. 105106.

Applied Mechanics Reviews JANUARY 2019, Vol. 71 / 010801-17

You might also like