You are on page 1of 20

Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


journal homepage: www.elsevier.com/locate/ymssp

Mechanical circulator for elastic waves by using the


nonreciprocity of flexible rotating rings
Danilo Beli a,⇑, Priscilla Brandão Silva b, José Roberto de França Arruda a
a
Department of Computational Mechanics, Faculty of Mechanical Engineering, University of Campinas, Rua Mendeleyev, 200, Cidade Universitária Zeferino
Vaz, Campinas, São Paulo 13083-860, Brazil
b
Department of Mechanical Engineering, Eindhoven University of Technology, PO Box 513, 5600MB Eindhoven, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: Circulators have a wide range of applications in wave manipulation. They provide a non-
Received 16 February 2017 reciprocal response by breaking the time-reversal symmetry. In the mechanical field,
Received in revised form 23 April 2017 nonlinear isolators and ferromagnetic circulators can be used for this objective.
Accepted 19 May 2017
However, they require high power and high volumes. Herein, a flexible rotating ring is
Available online 12 June 2017
used to break the time-reversal symmetry as a result of the combined effect of Coriolis
acceleration and material damping. Complete asymmetry of oscillating and evanescent
Keywords:
components of wavenumbers is achieved. The elastic ring produces a nonreciprocal
Breaking time-reversal symmetry
Complete asymmetric band diagrams
response that is used to design a three port mechanical circulator. The rotational speed
Coriolis acceleration for maximum transmission in one port and isolation in the other one is determined using
Elastic wave manipulation analytical equations. A spectral element formulation is used to compute the complex dis-
Spectral element formulation persion diagrams and the forced response. Waveguides that support longitudinal and
flexural waves are investigated. In this case, the ring nonreciprocity is modulated by
the waveguide reciprocal response and the transmission coefficients can be affected.
The proposed device is compact, nonferromagnetic, and may open new directions for
elastic wave manipulation.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction

Wave control and manipulation is a topic that raises enthusiasm among researchers of fields such as electromagnetism,
acoustics, mechanics, and material science as it opens new avenues for the development of applications in modern commu-
nication, energy harvesting, and sensing. Recently, the possibility of engineering devices with strongly nonreciprocal
responses has offered new possibilities for wave control [1]. Reciprocity is the fundamental property of most wave phenom-
ena observed in everyday life, which results from the time-reversible property of conventional media. It guarantees that the
wave propagation between source and receiver or vice versa is symmetric [2]. However, in various situations, nonreciprocal
wave phenomena are highly desirable as, for instance, to separate waves propagating in opposite directions, to create one-
way propagation, or to isolate different media [3]. In addition, it has also been shown that breaking time-reversal symmetry
plays an important role in bringing about topological edge states. The main feature of topological states, at first believed as a
quantum Hall effect, is their immunity against wave scattering by disorder [4,5].

⇑ Corresponding author.
E-mail address: dbeli@fem.unicamp.br (D. Beli).

http://dx.doi.org/10.1016/j.ymssp.2017.05.022
0888-3270/Ó 2017 Elsevier Ltd. All rights reserved.
1078 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Nonreciprocal devices, namely isolators and circulators, have initially been proposed in electromagnetism [6,7]. In isola-
tors, nonlinear structures associated with linear periodic structures are employed, whereas in circulators, linear media in the
presence of magnetic or Coriolis forces produce a nonsymmetrical propagation that breaks the time-reversal symmetry [8].
Moreover, an isolator can be constructed using circulators by impedance matching one of the ports. In this work, the objec-
tive is to show that flexible rotating rings have the inherent property of breaking the symmetry of forward and backward
traveling waves producing a nonreciprocal response. Hence, they can be regarded as the mechanical wave counterpart to
the devices recently proposed and demonstrated in electromagnetism [6], acoustics [9], and nonlinear dynamics [10–12].
Indeed, the flexible rotating ring is the mechanical analogue to the acoustic circulator proposed in Fleury et al. [13]. In
magneto-optical media, the magnetic field interacts with light, as a result of the Faraday phenomenon, which splits the spec-
tral lines, resulting in Zeeman’s electronic effect [14]. In acoustic media, angular momentum due to rotational fluid move-
ment induce nonreciprocity by splitting the resonance frequencies of an acoustic cavity. Here, the Coriolis forces acting on
the ring split the elastic waves and induce high interference phenomena.
The first realization of a nonreciprocal elastic wave device was the object of a patent deposed by Comstock [15]. It consists
of a magnetized disk of ferromagnetic material having gyromagnetic properties. Recently, the Coriolis effect was used to
break the time-reversal symmetry in topological active metamaterials. By using gyroscopes attached to a honeycomb lattice
structure, one-way longitudinal and transverse wave propagation immune to backscattering is obtained [16]. For the same
configuration, Nash et al. [4] showed that gyroscopic metamaterials allow one-way wave propagation even in the presence
of defects. Indeed, the unidirectional transport is a result of topological mechanical modes, which are analogues to the quan-
tum Hall effect. Besides, Wang et al. [5] showed that a nonsymmetrical response due to Coriolis forces can also be induced
when a two-dimensional honeycomb lattice of masses and springs is on constant rotation. Later, one-way propagation in
beams and rods was achieved using material modulation in time and space [17].
The present work shows that the Coriolis effect can also induce time-reversal symmetry breaking in a homogeneous
and continuous flexible ring under rotation. The dynamics of a flexible rotating ring was studied by many authors [18–
20] motivated by the need of understanding gear and tire vibrations. However, from the best of the authors knowledge,
the nonreciprocal feature of such structures and their use for the design of mechanical circulators have not been addressed
so far.
This paper is organized as follows. In Section 2, the fundamental basis to demonstrate that time-reversal symmetry is broken
in flexible rotating rings is presented. The equations of motion for an element of flexible rotating ring in an Eulerian reference
frame, corresponding dispersion relations and dynamic stiffness matrix are shown. Using the spectral element (SE) method,
dispersion diagrams and forced response solutions are obtained. In Section 3, the nonreciprocal mechanical circulator is stud-
ied, and analytical equations and numerical simulations verifying the physical phenomenon are shown. More complex exam-
ples with waveguides that can support longitudinal or flexural vibration are discussed in Section 4. Finally, in Section 5, the
conclusions are drawn. In addition, four appendices are included. In Appendix A, the equations of motion for a rotating ring
are provided. In Appendix B, the theoretical background to develop a spectral element from the equations of motion is pre-
sented. Free and forced analytical solution for the rotating ring are shown in Appendices C and D, respectively.

2. Wave propagation in an Eulerian reference frame

2.1. Theoretical background

The use of an Eulerian (spatial) description is convenient to solve coupled mechanical systems involving rotating parts,
which is the case in this paper. In this section, the theoretical formulation for the rotating ring (see Fig. 1) in an Eulerian ref-
erence frame is derived from the Lagrangian equivalent description, which is briefly recalled in Appendix A for the purpose of
clarity. The governing equations of motion for a flexible rotating ring in a Lagrangian reference system can be written in com-
pact form as:

Fig. 1. Schematic representations of a flexible rotating ring supported on an elastic foundation: (a) 3D CAD view, (b) top view and (c) 2D model including
reference axes, model parameters which are defined in Appendix A; the distributed elastic foundation is represented by the hatched area.
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1079

€ þ 2XGu_ þ ðqR2 X2 C þ KÞu ¼ q;


Mu ð1Þ
where M; G; C and K are, respectively, the mass, gyroscopic, centrifugal and stiffness operators, u is the vector of displace-
ments, q is the mass density, R is the mean radius of the centerline, X is the rotational speed, q is the vector of external loads,
and the superscript ðÞ denotes DðÞ=Dt (see Eq. (A.2) in Appendix A). In this description, the gyroscopic operator is skew-self-
adjoint, while the other operators are self-adjoint (see Eq. (A.3) in Appendix A). By employing the material derivative, i.e.
DðÞ=Dt ¼ @ðÞ=@t þ RX@ðÞ=@s — where D and @ denote, respectively, the derivatives with respect to the material (Lagrangian)
reference frame and to the spatial (Eulerian) reference frame —, the governing equations of motion can be rewritten with
respect to an Eulerian reference frame [20,21], as follows:

M@ tt u þ 2XðRM@ st u þ G@ t uÞ þ X2 R2 ½M@ ss u þ ð2=RÞG@ s u þ qCu þ Ku ¼ q; ð2Þ

where the partial derivatives with respect to t and s are expressed in compact form as @ t and @ s , respectively. While the
external loads (q) are attached to the structure in the Lagrangian formulation — i.e., they rotate with the ring —, in the Eule-
rian formulation, the vector of external loads is fixed to a point in space — i.e., it is stationary while the ring rotates. In both
reference systems, the internal efforts are expressed in the same way (see Eq. (A.4) in Appendix A). This is because the time-
dependence is implicit in the displacements.

2.2. Spectral element method

The physical phenomena addressed in this paper are demonstrated with the aid of numerical experiments solved by
the SE method. SE method is an analytical and discrete frequency-based approach which can be used to obtain dispersion
relations in the kðxÞ complex form of infinite systems as well as the forced response of finite systems [22–24]. The advan-
tage of the kðxÞ formulation for dispersion analysis compared with xðRefkgÞ formulations such as the plane wave expan-
sion method [25] is that not only propagating wave solutions, but also evanescent and decaying oscillatory ones are
obtained. Herein, real and imaginary parts of wavenumbers correspond, respectively, to propagating and evanescent com-
ponents. In addition, by using the current wave solutions and the same methodology presented in Beli et al. [26] to derive
the spectral element matrix of a rotating ring in a Lagrangian reference frame, it is possible to obtain the spectral dynamic
stiffness matrix of the rotating ring element in an Eulerian reference frame. For the sake of conciseness, the SE theoretical
background is herein omitted, but it is described in Appendix B. By assuming a time-harmonic wave solution (Eq. (B.1) in
Appendix B) to the Eulerian coupled governing equations of motion, Eq. (2), the dispersion relation for the rotating ring
element can be written as:

r6 k6 þ r5 k5 þ r4 k4 þ r3 k3 þ r 2 k2 þ r 1 k þ r 0 ¼ 0; ð3Þ
2 3 2 2 2
where r 6 ¼ EAEIR ; r5 ¼ 2qEIAR Xx; r4 ¼ EIð2EA  R bku þ qAR x2 Þ; r 3 ¼ 2qARXxðEAR  EIÞ; r2 ¼ ðEI þ EAR ÞðqAx2  bkw Þ
2

2
4q2 A2 R4 X2 x2 ; r1 ¼ 2qARXx½R2 bðku þ kw Þ  EA  2qAR2 x2  and r0 ¼ R2 b ku kw þ EAqAx2 þ qAR2 bðku þ kw Þx2
þq2 A2 R2 x2 ð4X2  x2 Þ. Six wavenumbers k are obtained by solving this polynomial equation for discrete angular frequencies
x. Three of them are related to forward waves, herein defined as k1 ; k2 and k3 , and the remaining three, to backward waves,
which are defined as k4 ; k5 and k6 — further information regarding the criterion to determine whether a wavenumber is
related to forward or backward waves is provided in Eq. (B.2) of Appendix B. Then, using the procedure described in Appen-
dix B (Eq. (B.3)) with the current wave solution, the tangential and radial spectral displacements can be written as:
 "e#

^
u R e e

u ¼ K u A; ð4Þ
^
w eI

where
 
e u ¼ diag eik1 s ; eik2 s eik3 s ; eik4 ðss0 Þ ; eik5 ðss0 Þ ; eik6 ðss0 Þ ;
K
e ¼ ½ a1 ; a2 ; a3 ; a4 ; a5 ; a6 T ;
A eI ¼ ½ 1; 1; 1; 1; 1; 1  and
e ¼ ½ a1 ; a2 ; a3 ; a4 ; a5 ; a6 ;
R

with an ¼ i½EIk3n þ EAkn þ 2qAR2 X2 kn  2qARXðx þ RXkn Þ=½ðEIk2n =RÞ  EARk2n  qAR3 X2 k2n  bRku þ qARðx2 þ 2RXkn x þ
R X
2 2 2
kn Þ,
for n ¼ 1; . . . ; 6.
Analogously, the internal efforts can be expressed in terms of the wave solutions, as in Eq. (A.4) of Appendix A. As men-
tioned before, the expressions of the external efforts, and, consequently, those of the internal efforts, remain unaltered under
changes of reference systems. By evaluating the displacements (Eq. (4)) and internal efforts (Eq. (A.4)) at the ends of the
rotating curved beam element, s1 ¼ 0 and s2 ¼ s0 , where s0 is the circumferential unit cell length, yields:

U e
e ¼ GA and e e
F ¼ H A; ð5Þ
1080 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

where
2   
 3
2   3 EA R @ s Ku þ ð1=RÞKu js¼0
R Ku js¼0 6 7
6 7 6   7
6 
7 6 EI ð1=RÞ R 
@ K

 @ K

j 7
6 Ku js¼0 7 6 ss u sss u s¼0 7
6  7 6 7
6    7 6   7
6 @ s Ku  ð1=RÞ R Ku js¼0 7 6    7
6 7 6 EI ð1=RÞ R @ K
s u  @ K j
ss u s¼0 7
6 7 6 7
G¼6   7; H ¼ 6    7;
6 R Ku js¼s0 7 6  7
6 7 6 EA R @ K
s u þ ð1=RÞK j
u s¼s0 7
6 7 6 7
6

Ku js¼s0 7 6   7
6 7 6    7
4   5 6 EI ð1=RÞ R @ ss Ku  @ sss Ku js¼s0 7
  6 7
@ s Ku  ð1=RÞ R Ku js¼s0 4   5
  
EI ð1=RÞ R @ s Ku  @ ss Ku js¼s0
    
U¼ u ^ ð0Þ; wð0Þ;
^ ^
wð0Þ; u ^ 0Þ ¼ u
^ 0 Þ; wðs
^ ðs0 Þ; wðs ^1 ; w^ 1; w^1; u ^2 ; w ^ 2 ; and
^ 2; w

h i h i
F ¼ T^ u ð0Þ; T^ w ð0Þ; T w ð0Þ; þT^ u ðs0 Þ; þT^ w ðs0 Þ; þT w ðs0 Þ ¼ T^ u1 ; T^ w1 ; T^ w1 ; T^ u2 ; T^ w2 ; T^ w2 :

e from the nodal spectral external forces e


Finally, by eliminating the wave amplitude vector A e the
F and displacements U,
elementary dynamic stiffness matrix is obtained, and it is expressed as De ðs0 ; xÞ ¼ HG1 . Using the spectral elements of
curved beams formulated in an Eulerian reference frame, a rotating ring subjected to fixed-point harmonic excitation can
be modeled and solved in the frequency domain.

2.3. Numerical experiments

In the numerical experiments carried out in this paper, the ring (see Fig. 1) is made of aluminum with q ¼ 2800 kg=m3 ,
E ¼ 70 GPa, structural damping g ¼ 0:01, complex elastic modulus E ¼ e
e Eð1 þ igÞ, and it has the following geometric char-
acteristics: R ¼ 50 mm, h ¼ 2 mm and b ¼ 5 mm. In addition, the nondimensional angular frequency (i.e., excitation fre-
qffiffiffiffiffiffiffiffiffi
quency) and rotational speed are, respectively, x ¼ Rx=ð2pcL Þ and X ¼ RX=ð2pcL Þ, where cL ¼ e E=q is the longitudinal

Fig. 2. Real (a, c) and imaginary (b, d) parts of the dispersion diagram in an Eulerian reference frame for X ¼ 0 and g ¼ 0 (a, b) and X ¼ 0:002 and g ¼ 0 (c,
d): forward () and backward () waves.
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1081

wave speed in a straight rod. The nondimensional displacements are given by u ¼ u=R and w ¼ w=R. In this section, the ring
is not supported on an elastic foundation, thus ku ¼ kw ¼ 0. Moreover, the complex nondimensional wavenumber is
lðxÞ ¼ ks0 .

2.3.1. Dispersion diagram


The homogeneous ring considered in this paper is characterized by three wave modes traveling in each direction, i.e.,
clockwise and counterclockwise. Two of them are associated with radial vibration and one, with circumferential motion.
In the undamped stationary condition, i.e., X ¼ 0 and g ¼ 0, Fig. 2(a-b), the oscillating and evanescent components of the
forward (lþ ) and backward (l ) traveling waves are symmetric with respect to the x-axis. In the case of an undamped
and rotating ring, X – 0 and g ¼ 0, Fig. 2(c-d), pairs of forward and backward waves exhibit a nonsymmetrical propagating
part, i.e. Rflþ ðxÞg – Rfl ðxÞg, and a symmetrical evanescent part, i.e. Iflþ ðxÞg ¼ Ifl ðxÞg, which corresponds to the
case of partial asymmetry.
Fully asymmetric band diagrams, i.e., Rflþ ðxÞg – Rfl ðxÞg and Iflþ ðxÞg – Ifl ðxÞg, are only achieved if some
damping is present in the rotating ring (X – 0 and g – 0), which is always the case in real structures. It should be noticed
that any amount of damping is sufficient. This phenomenon is caused by the Coriolis acceleration, which mathematically
implies a skew-self-adjoint gyroscopic operator, and it can be observed in the dispersion diagrams shown in Fig. 3. This fea-
ture leads to unusual dynamic responses that will be discussed in the following sections.

2.3.2. Dynamic analysis


In this section, the forced response of a rotating ring excited by a harmonic radial force, stationary in the Eulerian refer-
ence frame, is analyzed. Results obtained via the SE method are compared to analytical solutions via modal analysis (the ana-
lytical expressions for eigenfrequencies obtained via modal analysis are provided in Appendix C). The response of the ring
measured at the excited point along the radial direction is shown by means of the Campbell diagram in Fig. 4. As expected,
results obtained via the SE method are in perfect agreement with analytical eigenfrequencies computed with Eq. (C.1c) in
Appendix C. In the Eulerian description, strong bifurcations are observed due to the Coriolis acceleration enhanced by Dop-
pler’s effect. As a consequence, backward resonance modes concentrate at lower frequencies and forward resonance modes
are spread out and occur at higher frequencies. No effect is observed on the rigid body mode due to its translational nature.
Therefore, it always occurs at x ¼ 0. The dynamic behavior presented in the Campbell diagram is in accordance with the

Fig. 3. Real (a, c) and imaginary (b, d) parts of dispersion diagrams in Eulerian reference frame for X ¼ 0 (black), X ¼ 0:001 (red) and X ¼ 0:002 (blue) with
g ¼ 0:01: forward () and backward () waves; (c) zoom in y-axis to ½0; 5  103  and x-axis to ½1:5; 1:5, and (d) zoom in y-axis to ½0; 5  103  and x-axis
to ½3  103 ; 3  103 . (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
1082 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Fig. 4. Campbell diagram of the magnitude of the forced response in an Eulerian reference frame (colormap) and analytical eigenfrequencies ().

degenerate frequency response observed in other circulators and with the electronic Zeeman’s effect [14]. This is sufficient
evidence that the elastic rotating ring can also be viewed as a mechanical meta-atom analogue to the Zeeman’s atom.

3. Nonreciprocal circulator

In this section, the dynamic analysis of the flexible rotating ring, or mechanical meta-atom, is performed with the aim of
demonstrating its nonreciprocal behavior. Nonlinear bifurcation and centrifugal hoop stress are two important effects pre-
sent in the elastic ring which make it different from its optical, magnetic, and acoustic circulator analogues. Circulators usu-
ally present three or four ports that are equally distributed along the circumference — i.e, at each 120 or 90 , respectively.
The former case is considered in this work. Herein the ports located at 0 ; þ120 and 120 are named as port 1, port 2 and
port 3, respectively (see Fig. 5). We look for a solution where the transmission is always unidirectional and between adjacent
ports, i.e., mechanical energy may flow from port 1 to port 2, from port 2 to port 3 and from port 3 to port 1, but not in the
opposite sense, as shown in Fig. 5.

3.1. Analytical analysis

In this section, the aim is to provide a deep physical insight about the circulator dynamics through simple analytical
expressions. Analytical developments for the computation of the forced response of a mechanical circulator in the Eulerian
reference system via mode superposition are provided in Appendix D. In the analysis carried out in this section, we assume
that a radial force — harmonic in time and Dirac delta in space — is applied at port 1. Using Eq. (D.7) from Appendix D.1, for a
spectrum region dominated by a set of degenerate modes þm and m, the transmission (output/input displacement) ratios
for ports 2 and 3 are given by:
^ þ120
w 2p 2p
¼ aei 3 m þ beþi 3 m ;
w^ 0
ð6Þ
^ 120
w 2p 2p
¼ aeþi 3 m þ bei 3 m ;
w^ 0
where a and b are amplitude coefficients. Expanding Eq. (6) in real and imaginary components, yields:
^ þ120 R
w

¼ a þ bR cos hm þ 2aI sin hm þ i aR þ bR sin hm ;


w^ 0
ð7Þ
^ 120 R
w

¼ a þ bR cos hm  2aI sin hm  i aR þ bR sin hm ;


w^ 0

Fig. 5. Circulator with three ports evenly distributed along the ring circunference (120 of angular separation): (a) input at port 1, (b) input at port 2 and (c)
input at port 3. Energy flow consideration is depicted.
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1083

with hm ¼ ð2p=3Þm and aI ¼ bI (see Eq. (D.11) in Appendix D.1).


As shown in Eq. (D.11), the amplitude coefficients a and b can be expressed in terms of the ring rotational speed (X). Then,
from these expressions it follows that a ¼ b when the ring is stationary (X ¼ 0). Considering this relation in Eq. (6), one may
conclude that the responses at output ports are reciprocal as expected. The dynamics of the ring changes considerably under
rotation (X – 0). In the absence of damping (g ¼ 0), a and b must be real, but distinct (this follows again from Eq. (D.11) in
Appendix D.1). In other words, the output/input ratios relative to ports 2 and 3 would have the same amplitude, but a non-
zero phase difference, which characterizes a partial nonreciprocal behavior. A truly nonreciprocal behavior — i.e., output
ports with different amplitude ratios — occurs when some damping is present and the ring is under rotation (X – 0 and
g – 0). In this case, the amplitude coefficients must be complex and distinct. It is important to notice that the nonreciprocal
response of the circulator is directly related to the asymmetric feature of the wavenumbers: both occur under the same cir-
cumstances, i.e., for g – 0 and X – 0.
In the acoustic meta-atom or equivalent systems, the transmission and isolation depend on the amplitude decay rate that
is associated to the quality factor Q [1]. In structural vibration, the Q-factor is governed by damping, which, by introducing a
spatial phase difference between the degenerate modes m, produce a nonsymmetrical interference between them, thus giv-
ing rise to the nonreciprocal response at the output ports. However, when the motion of a damped rotating ring is dominated
by modes 3p; with p ¼ 1; 2; . . ., a reciprocal response is observed. The reciprocity in this case is assured because sin hm in
Eq. (7) becomes zero for these modes, of which the ranks are multiples of 3.
The analytical formulation can also be used to predict the optimal speed, i.e., the ring rotational speed at which total
transmission and complete isolation occur at the output ports. This situation is mathematically expressed as:
^ þ120 =w
jw ^ 0 j ¼ 1 and ^ 120 =w
jw ^ 0 j ¼ 0; or ð8aÞ
^
jwþ120 =w^ 0 j ¼ 0 and ^ ^
jw120 =w0 j ¼ 1: ð8bÞ
From Eq. (7), the conditions for isolation are:
pffiffiffiffiffiffiffi
aR þ bR ¼ 2 ð3ÞaI and ð9aÞ
aR  bR ¼ 0; ð9bÞ

the sign depends on the mode rank m. As discussed before, the second relation for isolation, Eq. (9b), occurs at X ¼ 0.
Indeed, for low flexural modes and at low rotational speed when the behavior is asymptotic around X ¼ 0, one can assume
2
ðaR  bR Þ
0, as it can be verified in Fig. 6. For this frequency range, if one assumes also that ðaR þ bR Þ ¼ 1 and aR and bR
pffiffiffi
satisfy the first condition for isolation, Eq. (9a), the imaginary part of the amplitude coefficients must be aI ¼ 1=ð2 3Þ
p ffiffiffi
and bI ¼ 1=ð2 3Þ. Then, replacing these values in Eq. (7), the following holds:
^ þ120 =w
jw ^ 0 j ¼ 1 and ^ 120 =w
jw ^ 0 j ¼ 0; for m ¼ 3p þ 1; ð10aÞ
^
jwþ120 =w^ 0 j ¼ 0 and ^ ^
jw120 =w0 j ¼ 1; for m ¼ 3p  1; ð10bÞ
with p ¼ 1; 2; . . ..
The optimal rotational speed corresponds to the rotational speed at which the optimal transmission/isolation occurs. This
value can be estimated by solving the polynomial equation that results from Eq. (D.11) (see Appendix D.1) when the optimal
value of aI is considered, i.e.:
I I I
/ðXÞ ¼ f 5 X5  aI g 4 X4 þ f 3 X3  aI g 2 X2 þ f 1 X  aI g 0 ¼ 0 ð11Þ

Fig. 6. (a) Magnitude of ðaR  bR Þ and (b) magnitude of /ðXÞ in Eq. (11) as a function of X for m ¼ 2 (green), m ¼ 5 (red) and m ¼ 8 (blue). (For interpretation
of the references to color in this figure legend, the reader is referred to the web version of this article.)
1084 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

This polynomial equation is solved numerically and the smallest positive real root corresponds to the optimal rotational
speed Xopt for a given degenerate natural angular frequency xopt .

3.2. Spectral element simulation

Using the SE method, the exact forced response — within the extent of the theory under consideration — of the proposed
device is computed. Furthermore, the SE provides a straightforward way to obtain analytical solutions of more complex sys-
tems by assembling the dynamic stiffness matrices of several spectral elements. To begin with, the optimal excitation angu-
lar frequency and optimal rotational speed (xopt and Xopt , respectively) estimated from the analytical expressions and
computed via the SE method for various values of aspect ratio (h=R) and structural damping (g) are shown in Figs. 7 and
8, respectively. The analytical and SE results are mostly in agreement. However, small deviations can be observed as the val-
ues of the aspect ratio and structural damping rise. Notice that Xopt has a linear relation with the mode rank m. Moreover, by
increasing the aspect ratio, higher excitation frequencies and rotational speeds are required to reach the optimal condition.
On the other hand, by increasing the structural damping, the optimal condition is reached at a higher rotational speed, but
roughly constant excitation frequency.
Only one of the three possible configurations for the application of the input load is considered in this section: the one in
which the input is applied at port 1 and outputs are measured at ports 2 and 3, see Fig. 5(a). The other possible configurations
would lead to similar results, but with a phase difference of 120 with respect to the position of measured responses. The
forced response computed by the SE method is shown in Fig. 9. As predicted by analytical equations, a reciprocal result is
obtained at output ports for modes with rank 3p; p ¼ 1; 2 . . .. Moreover, when the excitation is applied at port 1, the vibra-

Fig. 7. Optimal excitation frequency (a) and rotational speed (b) as function of the mode rank for h=R ¼ 0:02 (blue), h=R ¼ 0:04 (red) and h=R ¼ 0:06 (black);
computed using the analytical () and SE ( ) formulation. (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)

Fig. 8. Optimal excitation frequency (a) and rotational speed (b) as function of the mode rank for g ¼ 0:0025 (blue), g ¼ 0:01 (red) and g ¼ 0:025 (black);
computed using the analytical () and SE ( ) formulation. (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1085

Fig. 9. Forced response of the circulator measured at ports 1 (green), 2 (red) and 3 (blue) when the excitation is applied at port 1. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

tion is transmitted to port 2 for modes of rank 3p þ 1 while port 3 remains isolated. The opposite situation — i.e., vibration is
transmitted to port 3 while port 2 remains isolated — occur for modes of rank 3p  1.
The computational efficiency of the SE method compared, for instance, with the finite element method, allows an exten-
sive study of the dynamic response of the circulator for ranges of excitation frequency and rotational speed. Maps of exci-
tation frequency by rotational speed, dynamic response in amplitude and phase, and transmission coefficients, Sio ¼ jwo =wi j,
where wo is the radial displacement at the output port and wi is the radial displacement at the input port, were computed.
For intervals of normalized excitation frequencies x around the natural angular frequencies relative to modes m ¼ 4 and
m ¼ 5, the radial displacement responses for a range of normalized rotational speeds X and transmission coefficients
(Sio ¼ jwo =wi j) are computed, see Figs. 10 and 11 and Figs. 12 and 13, respectively. For the frequency range around the natural
frequency relative to m ¼ 4, the map for port 3 shows a minimum response at x ¼ 0:02674 and X ¼ 2:19  105 . As
expected, at this optimal point, the transmission coefficients are jS12 j ¼ 1 and jS13 j ¼ 0. For small to moderate fluctuations
of the excitation frequency and rotational speed around the optimal values, jS12 j decreases while jS13 j increases, but the cir-
culator behavior is quite robust to such variations. For stronger deviations, jS12 j approximates jS13 j and the response becomes
reciprocal. When the dynamic response of the circulator is analyzed around the natural frequency relative to m ¼ 5, the map
for port 2 shows a minimum value at x ¼ 0:04325 and X ¼ 2:71  105 . At this point, the transmission coefficients are
jS13 j ¼ 1 and jS12 j ¼ 0. As in the previous case, deviations from the optimal point show similar behavior at the three ports.
It should be noted that, for a given damping level, whatever its value is,the tuning may be achieved by varying the rotational
speed.
In both analyses — i.e., responses dominated by modes 4 and 5 —, at the optimal condition, the isolated output port dis-
placement has a phase change passing through 0, while the phase difference between input and output transmission ports is
p. Notice that the radial reference system is pointing outwards, see Fig. 1; therefore, as expect, an input pointing inwards
produces an output displacement pointing outwards. In addition, except for modes of rank 3p with p ¼ 1; 2; 3; . . ., which
exhibit a reciprocal response, and for m ¼ 1, when the degenerate modes do not split, the scattering matrix S at the optimal
condition is nonsymmetric, and given by:
2 3 2 3
S11 S12 S13 0 1 0
6 7 6 7
S ¼ 4 S21 S22 S23 5 ¼ 4 0 0 1 5; ð12Þ
S31 S32 S33 1 0 0

Fig. 10. For a range of excitation frequencies around mode 4, map plot of radial displacement measured at port 1 (a), port 2 (b) and port 3 (c).
1086 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Fig. 11. For a range of excitation frequencies around mode 4 and X ¼ 2:19  105 , radial displacement amplitude (a), phase (b) and transmission
coefficients (c): port 1 (green), port 2 and jS12 j (red), port 3 and jS13 j (blue). (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

Fig. 12. For a range of excitation frequencies around mode 5, map plot of radial displacement measured at port 1 (a), port 2 (b) and port 3 (c).

Fig. 13. For a range of excitation frequencies around mode 5 and X ¼ 2:71  105 , radial displacement amplitude (a), phase (b) and transmission
coefficients (c): port 1 (green), port 2 and jS12 j (red), port 3 and jS13 j (blue). (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

where S11 ; S22 and S33 have the meaning of reflection coefficients, and S12 ; S13 ; S21 ; S23 ; S31 and S32 have the meaning of
transmission coefficients. This S-matrix is consistent with the circulator theory and the nonreciprocal system definition [2].
The optimal transmission/isolation behavior discussed above for the case of radial input and output measurements can-
not be guaranteed on the circumferential direction. For a radial excitation at the optimal rotational speed, Xopt , circumfer-
ential displacements are also computed for responses dominated by modes m ¼ 4 and m ¼ 5 and shown in Fig. 14.
Although the output ports exhibit a nonreciprocal response in this direction, isolation is not observed. Therefore, application
devices that take advantage of the isolation feature must be designed to receive and transmit vibration mainly on the radial
direction, this constraint will be further discussed in Section 4.
A high Q-factor is usually required in circulators, as the ones proposed for electromagnetic and acoustical applications
[6,7]. In mechanics, this condition corresponds to a small value of structural damping g. As shown in previous results, the
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1087

Fig. 14. For a range of excitation frequencies around mode 4 with X ¼ 2:19  105 (a), and around mode 5 with X ¼ 2:71  105 (b), circumferential
displacement response: at port 1 (green), port 2 (red), and port 3 (blue). (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

proposed mechanical circulator works well for g ¼ 0:01. In Fig. 15, the effect of damping in the circulator is analyzed for
g ¼ ½0; 0:05; 0:10; 0:20. Fig. 15 shows that the nonreciprocal behavior is verified for g – 0, and the high transmission/com-
plete isolation condition is also observed for g ¼ 0:05. However, by increasing g to 0:10, although the isolation port condition
continues to be satisfied, the transmission port response is considerably degraded. As a matter of fact, although most engi-
neering materials exhibit internal damping factor below 0:10, one should be aware that damping is critical for the perfor-
mance of the mechanical circulator.

3.3. Elastic foundation

In real applications, rotating rings are not usually free in space, but mounted on a shaft supported in an elastic foundation
and subjected to a controlled torque. For this reason, in this section we study the influence of this elastic support on the non-
reciprocal behavior of rotating rings. Herein, the elastic support is represented by distributed springs in the radial and tan-
gential directions. The additional terms introduced by the consideration of such springs have been added to both analytical
and SE formulations. It is assumed that the elastic support is homogeneous and uniform, and, for the sake of simplicity, the
stiffness in both directions is assumed equal, ku ¼ kw . Its effect in the dynamic response is to shift resonances to higher fre-
quencies, especially for low mode ranks including the rigid body mode (see Fig. 16), which under this effect occurs at
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
xRBM ¼ kef =mring — where mring is the ring mass and kef is the equivalent elastic foundation stiffness. Nevertheless, it does
not hinder the nonreciprocal behavior of the rotating ring, and does not influence the isolation/transmission coefficients, as it
observed in Fig. 16(b-c). Moreover, for a given mode m, while the optimal excitation frequency xopt increases with the elastic
stiffness, the optimal rotational speed Xopt decreases, see Fig. 17. The results obtained by solving Eq. (11) and via SE simu-
lation are also in agreement for the flexible rotating ring supported on an uniform elastic foundation.

4. Coupled system analysis

In this section, the effect of the nonreciprocal response of the circulator when it is part of a more complex mechanical
system is investigated. Herein, we assume that a waveguide is connected to each circulator port by elastic bearings, which

Fig. 15. Transmission coefficients for a range of frequencies around mode 5 for g ¼ 0 and X ¼ 0 (o), g ¼ 0:05 and X ¼ 1:34  104 (– –), g ¼ 0:10 and
X ¼ 2:91  104 (—), and g ¼ 0:20 and X ¼ 7:34  104 (j) at port 1 (green), port 2 (red) and port 3 (blue) when the excitation is applied at port 1. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
1088 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Fig. 16. Forced response of the circulator supported on an elastic foundation with ku ¼ kw ¼ 1 GN=m3 and X ¼ 2:44  105 (a) and zoom for a range of
frequencies around mode 5 (b) as well as the corresponding transmission coefficients (c) measured at port 1 (green), port 2 (red) and port 3 (blue) when the
excitation is applied at port 1. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 17. Effect of the elastic foundation on the optimal excitation frequency (a) and rotational speed (b) for: ku ¼ kw ¼ 0 (blue), ku ¼ kw ¼ 0:5 GN=m3 (red),
and ku ¼ kw ¼ 5 GN=m3 (black); computed by solving Eq. (11) () and SE ( ) simulation. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

are modeled by springs. The transmission of vibration from source to receptor through the ring is investigated in terms of
displacements and energy.
If the waveguide inertia is negligible compared to the inertia of ring, the analysis performed in Section 3 would be enough
to describe the dynamic response of the coupled system. However, this assumption is not always valid, and the waveguide
modifies the dynamic response and, consequently, the optimal angular frequency and rotational speed for nonreciprocity.
Two configurations will be analyzed in the present section: one in which the connected waveguides support only longitu-
dinal vibrations (see Fig. 18(a)), and a second one in which the waveguides work as beams and are subjected to flexural
vibrations (see Fig. 18(b)). In both cases, only radial vibration transmission from the ring is allowed, and the waveguides
are supposed to be stationary. The flexible rotating ring equations are solved in an Eulerian reference frame, as this allows

Fig. 18. Coupled systems: (a) configuration 1 - mechanical circulator with waveguides that support longitudinal vibration, (b) configuration 2 - mechanical
circulator with waveguides that support flexural vibration.
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1089

straightforward consideration of the interaction with stationary structures. In addition, coupled systems are easily solved via
the SE formulation, in which coupling conditions are automatically satisfied by the superposition of dynamic stiffness matri-
ces. Moreover, the waveguide-ring contact is assumed to be ideal, and is modeled by linear springs [19], so that vibration is
perfectly transmitted from one structure to another, without dissipation.
The waveguides used in the numerical experiment are also made of aluminum with damping loss factor g ¼ 0:01, and the
geometric characteristics are: length lv ¼ 6R, width bv ¼ 5 mm and thickness hv ¼ 5 mm. The spring connecting each waveg-
uide to the ring port has stiffness kc ¼ 5  108 N/m, and the elastic foundation stiffness coefficients are
ku ¼ kw ¼ 5  108 N=m3 .

4.1. Dynamic analysis

In the first case treated, a longitudinal excitation is applied at the free end of the waveguide connected to port 1 and lon-
gitudinal displacements are measured at the end of the three waveguides that are connected to the ring. The results are pre-
sented in Fig. 19 (a). Notice that, as the response is measured at the waveguides, the nonreciprocal behavior, although
observed, is modulated by the dynamic response of the bars (see Fig. 19 (b)). Moreover, the optimal transmission/isolation
condition is observed within frequency bands dominated by the ring vibration (see Fig. 20 (a) and (b)). For xopt ¼ 0:08464,
which is a region dominated by a bar resonance and the 7th ring mode, and for Xopt ¼ 13:12  105 ; jS12 j
1 at port 2 and
jS13 j ¼ 0 at port 3, the phase difference between the responses at ports 1 and 2 being approximately p.
In the second case treated, a flexural excitation is applied at the free end of the waveguide connected to port 1 and flex-
ural displacements are also measured at the ends of the three waveguides that are connected to the ring, Fig. 21(a). As in the
previous example, the nonreciprocal response is observed within frequency bands in which the ring vibration dominates,
and the response is modulated by the beam dynamics (see Fig. 21(b)). In this case, the transmission coefficients obtained
within a frequency band dominated by the beam resonance and the 5th ring mode — xopt ¼ 0:04287 — are jS13 j
0:95 at

Fig. 19. (a) Displacement responses at the end of waveguides connected to port 1 (green), port 2 (red), and port 3 (blue) for configuration 1 with
Xopt ¼ 13:12  105 , and (b) displacement responses of a free-free bar excited longitudinally at one end: at the non-excited end (black) and at the excited
end (orange). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 20. Dynamic response (a), phase difference (b) and output/input displacement ratio (c) at the end of waveguides connected to port 1 (green), port 2
(red) and port 3 (blue) for configuration 1 with Xopt ¼ 13:12  105 . (For interpretation of the references to color in this figure legend, the reader is referred
to the web version of this article.)
1090 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Fig. 21. (a) Displacement responses at the end of waveguides connected to port 1 (green), port 2 (red), and port 3 (blue) for configuration 2 with
Xopt ¼ 3:93  105 , and (b) displacement responses of a free-free beam excited transversely at one end: non-excited end (black) and excited end (orange).
(For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 22. Dynamic response (a), phase difference (b) and output/input displacement ratio (c) at the end of waveguides connected to port 1 (green), port 2
(red) and port 3 (blue) for configuration 2 with Xopt ¼ 3:93  105 . (For interpretation of the references to color in this figure legend, the reader is referred to
the web version of this article.)

port 2 and jS12 j ¼ 0 at port 3 for X ¼ 3:48  105 . The phase difference between port 1 and port 3 in this case is also approx-
imately p.
In general, for both cases, nonreciprocal responses are also obtained when waveguides are connected to the circulator,
thus, producing jSio j
1 at the transmission port and jSio j
0 at the isolation port. It is important to notice that the nonre-
ciprocal behavior becomes more robust when the optimal excitation frequency of the ring coincides with a resonant fre-
quency of the waveguide as shown in the results of this section. Moreover, the required optimal rotational speed
increases with the waveguide inertia.

4.2. Vibrational power flow and energy density analysis

In the previous section, the forced responses of the coupled system in terms of displacements were analyzed. Similar
analyses can be performed in terms of energy and provides more physical insight. For harmonic motion, the time-
P
averaged flow of vibrational energy in the bar, beam or ring is given by PF j ¼ 12 Refix j T j v j g, where v j is the displace-
ment/rotation degree of freedom and T j is the associated internal efforts on the cross-section [27,28]. The energy density
in the waveguides can also be computed. For a bar which supports only longitudinal motion, as in configuration 1, the kinetic
and potential energies per unit length are given, respectively, by KEbar ¼ 12 qAv ðRf@ v =@tgÞ2 and PEbar ¼ 12 EAv ðRf@ v =@xgÞ2 ,
where Av is the cross-section area. For a beam supporting flexural motion, as in configuration 2, the kinetic and potential
energies per unit length are, respectively, KEbeam ¼ 12 qAv ðRf@ v =@tgÞ2 and PEbeam ¼ 12 EIv ðRf@ 2 v =@x2 gÞ , where Iv is the area
2

moment of inertia. Moreover, the total energy density is given by the sum of kinetic and potential components,
Ev ¼ KEv þ PEv [23].
The results are computed for the same geometries and excitations considered in the previous subsection. The vibrational
power flow in dB — with PF ¼ PF=PF ref and PF ref ¼ 1 W — for configurations 1 and 2 are shown in Fig. 23. As expected, for both
configurations, the energy does not flow at the isolation port where jv j ¼ 0. At the transmission port, jv o j
jv i j with a phase
difference of approximately p (see Figs. 20 and 22). Therefore, the vibrational energy that enters the circulator via port 1 leaves
it through the output transmission port (see Fig. 23(b) and (d)). As a result, along the ring, the energy flow is higher between the
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1091

Fig. 23. Power flow as a function of the excitation frequency (a,c) measured at the end of waveguides connected to port 1 (green), port 2 (red) and port 3
(blue) for a specific Xopt , and power flow distribution throughout the coupled system for specific values of Xopt and xopt (b,d): (a,b) configuration 1 -
Xopt ¼ 13:12  105 and xopt ¼ 0:08464; (c,d) configuration 2 - Xopt ¼ 3:93  105 and xopt ¼ 0:04287. (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)

Fig. 24. Energy density ratio as a function of excitation frequency computed at the end of waveguides connected to port 2 (red) and port 3 (blue): (a)
configuration 1 at Xopt ¼ 13:12  105 , (b) configuration 2 at Xopt ¼ 3:93  105 . (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

input and transmission ports. Between the isolation and the other two ports, the energy flow decreases and reaches its mini-
mum at the isolated port. These results show clearly the direction of energy propagation through the coupled system — i.e.,
mechanical circulator and waveguides — and it is in agreement with the dynamic analysis in the previous section.
Results in terms of energy density are presented in Fig. 24. As in the dynamic analysis, results are expressed in terms of a
ratio of quantities, in this case, the energy density measured at the free end of the waveguides connected to an output and
the input port of the ring. As it can be observed, the transmission and isolation are again associated to the unity and zero,
respectively, as expected.
It is also important to notice that semi-infinite waveguides connected to the mechanical circulator will include an imag-
inary part in the dynamic response of the ports. This is similar to an extra damping effect, which can degrade the transmis-
sion condition depending on the material/geometry of the circulator as well as on the waveguides working on longitudinal or
flexural behavior. Therefore, for the hypothetical semi-infinite waveguide case, the structural damping becomes more
important to design the circulator and a small loss factor g is preferable to guarantee high transmission coefficients.
1092 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

5. Conclusions

This work provides the theory for a mechanical circulator by using a flexible rotating ring that breaks the time-reversal
symmetry. The Coriolis acceleration is responsible for splitting the degenerate modes that interfere at the different output
ports to produce the nonreciprocal response. The SE method was used to compute the dispersion diagrams and to perform
the full dynamic analysis of the rotating ring in an Eulerian reference frame. A complete dynamic nonreciprocal response,
with different transmission coefficients at distinct output ports, is related to complete asymmetry for oscillating and evanes-
cent parts of the wavenumbers obtained for rotating cases with material damping, as predicted by the Onsager-Casimir Prin-
ciple [8,29]. The optimal rotational speed for maximum difference in transmission coefficients at different output ports was
computed by analytical equations and compared with SE results for the flexible ring with elastic foundation. The influence of
the waveguide, which can support either longitudinal or flexural vibration, was shown in more complex examples. Waveg-
uides connected to the rotating ring decrease the transmission coefficient values, but the complete isolation in one of the
output ports is always respected. In these complex systems, higher transmission values can be obtained for frequency bands
dominated by waveguide resonances. In addition, the SE method was used to build the complex structures in a straightfor-
ward way by using the superposition of dynamic stiffness matrices. This was possible because the rotating ring equations
were described in an Eulerian (space-fixed) reference frame. The proposed device is largely tunable from lower to higher
frequencies and provide a compact and nonmagnetic solution to construct a switch (energy rectification), an isolator or a
circulator for elastic waves. This mechanical meta-atom may open new possibilities in the research on elastic wave propa-
gation, including vibration control, energy harvesting, and sensing. The experimental realization of such a device is challeng-
ing, as it involves moving parts and connecting these to stationary waveguides. However, the analyses presented in this work
indicate that this should be possible.

Acknowledgments

The authors are grateful to the São Paulo Research Foundation (FAPESP) – Brazil, Project number 2014/19054-6 for the
financial support.

Appendix A. Governing equations of motion in a Lagrangian reference frame

In this work, the one-dimensional ring element is based on the in-plane Euler-Bernoulli beam theory. The assumptions
are that plane sections remain plane after deformation, the centroid of each cross-section is located in the centerline, and
the aspect ratio is small, i.e. h=R < 0:1. These assumptions imply that transverse shear deformation and rotary inertia effects
can be negligible, which is not important for lower-order bending modes. Therefore, a significant strain is only possible in the
circumferential direction. The general extensional-flexural coupled equations of motion for a rotating ring, Fig. 1, described
in Lagrangian reference frame can be obtained upon application of Hamilton’s Principle [18,30], which yields:

     "
# " 00 # "
#    
€ 0 þ1 u_ þ2w0 þ Ru00 ðu =R  w000 Þ 0 00
þqAR
u
þ 2XqAR  qR2 X2 A
 EI  0000
  EA w þ Ru þ bR
ku u
¼ R
qu
;

w 1 0 _
w 2u0 þ Rw00 u000  Rw ðw=R  u0 Þ kw w qw
ðA:1Þ

where u and w are, respectively, the tangential and radial centerline displacements, E ¼ e
Eð1 þ igÞ is the dynamic elastic
modulus, where g is the structural damping and e
3
E is the elastic modulus, q is the mass density, I ¼ bh =12 is the second
moment of area, A ¼ bh is the rectangular cross-sectional area, R is the mean radius of the centerline, h is the beam height,
b is the beam width, X is the rotational speed, ku and kw are the elastic foundation stiffness coefficients, qu and qw are, respec-
tively, the radial and tangential external loads per unit length. In addition, s is the circumferential coordinate along the cen-
terline and t is the time variable; the derivatives with respect to s and t are represented, respectively, by @ðÞ=@s ¼ ð0Þ and
DðÞ=Dt ¼ ðÞ. Besides, w ¼ @w=@s  u=R is the angle of rotation due to bending, and the terms qR2 X2 and 2X½DðÞ=Dt are asso-
ciated to the centrifugal hoop stress and Coriolis acceleration, respectively. In compact form, Eq. (A.1) can be written as
follows:
€ þ 2XGu_ þ ðqR2 X2 C þ KÞu ¼ q;
Mu ðA:2Þ
where

u ¼ ½ u w T ; q ¼ R½ qu qw T ; Mu € ¼ þqAR½ u € w € T ;
   "
#
0 þ1 u_ þ2w0 þ Ru00
Gu_ ¼ qAR ; Cu ¼ A
and
1 0 w_ 2u0 þ Rw00 ðA:3Þ
" 00 # "
#  
ðu =R  w000 Þ w0 þ Ru00 ku u
Ku ¼ EI  0000
  EA þ bR :
000
u  Rw ðw=R  u0 Þ kw w
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1093

Fig. A.1. Schematic of a differential element of flexible rotating ring.

In addition, the internal efforts (see Fig. 1) are also obtained from Hamilton’s principle and can be expressed in terms of
displacements and strains, as follows [23]:
 00 0
w u u
T u ¼ EA u0 þ ; T w ¼ EI  w000 ; T w ¼ EI  w00 : ðA:4Þ
R R R

Here, T u corresponds to the normal force, T w to the transverse shear force, and T w to the bending moment (see Fig. A.1).

Appendix B. Spectral element method: theoretical background

The Spectral Element (SE) formulation, as proposed by Doyle [22], Lee [31] and Mace [24], was applied to the analysis of
rotating rings in a Lagrangian frame of reference by the authors [26]. This appendix provides the general theory that was
used to develop the spectral element of flexible rotating ring in an Eulerian frame of reference. The methodology assumes
a spectral solution for the displacements and external forces, which allows the transformation of governing equations from
time to frequency domain, as follows:
X X
uðs; tÞ ¼ u ^ ðs; xÞ ¼
^ ðs; xÞeixt ; with u an eikn s ; ðB:1Þ
x n

where x is the angular frequency, k is the wavenumber, n is the wave mode count, and i is the imaginary number. By apply-
ing this wave propagation solution in the homogeneous system, Eðk; xÞu ^ ðs; xÞ ¼ 0 is achieved. By setting its determinant to
zero, an analytical polynomial equation of order 2N (detðEðk; xÞÞ ¼ 0), i.e. the dispersion relation, is obtained. Numerically
solving this equation, 2N distinct frequency-dependent complex wavenumbers are found, (1; 2; . . . ; N) correspond to number
of waves that propagate in the forward direction and other (N þ 1; N þ 2; . . . ; 2N) correspond to waves that propagate in the
backward direction. The condition for forward travelling propagation:

Ifkg < 0 or Rf@k=@ xg < 0 if Ifkg ¼ 0: ðB:2Þ


T
The normalized eigenvector associated with each wavenumber is Un ¼ ½an 1 , where an is the amplitude ratio for each
wave mode. Then, for each frequency, displacements and internal forces can be written as a wave mode superposition [24].

X
2N
^ ðs; xÞ ¼
u Un eikn s an ¼ Wðs; xÞAðxÞ; ðB:3Þ
n¼1

where

Wðs; xÞ ¼ ½ U1 ; . . . UN ; UNþ1 ; . . . U2N Ku ðs; xÞ;


 
Ku ðs; xÞ ¼ diag eik1 s ; . . . eikN s ; eikNþ1 ðss0 Þ ; . . . eik2N ðss0 Þ and
AðxÞ ¼ ½ a1 ; . . . aN ; aNþ1 ; . . . a2N T :
For a finite element of length s0 , Eq. (B.1) must satisfy the geometric and natural boundary conditions at end nodes s ¼ 0
and s ¼ s0 . The external nodal displacements are related to the displacement fields by Uðs; ^ xÞ ¼ CGB u^ ðs; xÞ, where CGB is the
linear differential operator for the geometric boundary conditions. And the external nodal forces are related to the displace-
b xÞ ¼ CNB u
ment fields by Fðs; ^ ðs; xÞ, where CNB is the linear differential operator for the natural boundary conditions. Replac-
ing Eq. (B.3) in these equations and evaluating them at the ends, yields:

  h iT
^ T
þUð0Þ ^ 0 ÞT T ¼ þCGB WðsÞjTs¼0 CGB WðsÞjTs¼s A;
Uðs ðB:4aÞ
0

  h iT
^ T
Fð0Þ ^ 0 ÞT T ¼ CNB WðsÞjTs¼0 CNB WðsÞjTs¼s A:
Fðs ðB:4bÞ
0

The dependence upon x of the parameters in the latter expressions was hidden for the purpose of brevity. Eliminating the
wave amplitude vector A from Eq. (B.4) gives the dynamic stiffness matrix, Dðs0 ; xÞ. And by replacing Eq. (B.4a) in Eq. (B.1),
the interpolation functions, Nðs; xÞ, which are used to compute the response at any point of the element, are obtained.
1094 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

Appendix C. Analytical modal analysis

Flexural and extensional vibrations are coupled by the ring curvature. However, for low order bending modes and low
rotational speeds, inextensional deformation can be assumed, which considers only flexural behavior and provides decou-
pled equations of motion [18,19]. A stationary inextensional ring resonates at angular frequencies xst , Eq. (C.1a), with back-
ward and forward mode shapes given by e imR , where m is the mode rank [18]. Under rotation, X – 0, the stationary mth
s

mode splits nonlinearly with the rotational speed, which leads to degenerate modes at frequencies xrotL , Eq. (C.1b), in the
Lagrangian description [18], and xrotE , Eq. (C.1c), in the Eulerian one [30].
2
!1=2
EIm2 ðm2  1Þ m2 kw þ ku
xst ¼ þ ; ðC:1aÞ
4
qAR ðm þ 1Þ
2 qhðm2 þ 1Þ
2
!1=2
2m m2 ðm2  1Þ 2
xrotL ¼ X x2st þ X ; ðC:1bÞ
m þ1
2
ðm2 þ 1Þ
2

2 !1=2
ðp2 þ m2 p1 Þ ðp2 p3  m2 p21 Þ p ðm2  1Þ
xrotE ¼ X þ 2 X; ðC:1cÞ
mðp1 þ p2 Þ qAðp1 þ p2 Þ mðp1 þ p2 Þ
 
with p1 ¼ m2 EI=R4 þ EA=R2 þ 2qAX2 ; p2 ¼ m2 EI=R4 þ EA=R2 þ qAX2 þ bku and p3 ¼ m4 EI=R4 þ EA=R2 þ m2 qAX2 þ bkw .
However, the centrifugal hoop stress is not significant at low speeds. Indeed, for low rotational speeds and m – 1, the
rotating ring can be shown to be equivalent to a stationary ring with a traveling load [30], in which vibration modes occur
at angular frequencies xinv ¼ xst mX.

Appendix D. Analytical forced response

The governing equations of motion, Eq. (A.1), describe the complete ring behavior and its analytical solutions can be writ-
ten by making use of modal superposition, as follows:
X
þ1 X
þ1
^ m eiðmRþxtÞ ; uðs; tÞ ¼ ^ m eiðmRþxtÞ
s s
wðs; tÞ ¼ w u ðD:1Þ
m¼1 m¼1

where m is the rank mode. In the case the ring is rotating, forward and backward mode contributions are different; thus, the
sum includes negative and positive terms. Replacing Eq. (D.1) in Eq. (A.1), the following system of equations is obtained:
X
þ1
^ m þ L12 w ^u ;
^ m ÞeiðmRÞ ¼ Rq
s
ðL11 u
m¼1
ðD:2Þ
X
þ1
^ m þ L22 w ^w ;
^ m ÞeiðmRÞ ¼ Rq
s
ðL21 u
m¼1

where, for the Eulerian reference frame

L11 ¼ EIR3 m2 þ EAR1 m2 þ qARX2 m2 þ Rbku  qARðx þ mXÞ2 ;


L22 ¼ EIR3 m4  EAR1  qARX2 m2  Rbku þ qARðx þ mXÞ2 ;
L12 ¼ i½EIR3 m3  EAR1 m  2qARX2 m þ 2qARXðx þ mXÞ;
L21 ¼ L12 :

In order to obtain the analytical solution, Eq. (D.2) is multiplied on both sides by an orthogonal mode, ems=R , and the
resulting equation is integrated with respect to s from 0 to 2pR:
Z 2pR X
þ1 Z 2pR
^ m þ L12 w
^ m ÞeiðmRÞ eiðmRÞ ds ¼ ^u eiðmRÞ ds;
s s s
ðL11 u Rq
0 m¼1 0
Z 2pR X
þ1 Z 2pR ðD:3Þ
^ m þ L22 w
^ m ÞeiðmRÞ eiðmRÞ ds ¼ ^w eiðmRÞ ds:
s s s
ðL21 u Rq
0 m¼1 0

R 2pR
By using the orthogonality relation, 0 eiðmmÞs=R ds ¼ 2pRdmm , where dmm is the Kronecker delta function, the Eq. (D.3) is
reduced to the following system of equations:
   Z 2p R  
L11 L12 ^m
u 1 ^
q u s
¼ eimR ds: ðD:4Þ
L21 L22 ^m
w 2p 0 ^w
q
D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096 1095

Considering only a radial harmonic force as excitation, qu ¼ 0 and qw ¼ Q wm dðs  s Þ, where d is the Dirac delta function
and s is the force position, the amplitude of the spatial solutions can be obtained and replaced in Eq. (D.1), thus yielding
complete analytical solutions:

1 Xþ1
Q wm L22 iðmðss ÞþxtÞ
wðs; tÞ ¼ e R ;
2p m¼1 D
ðD:5Þ
1 Xþ1
Q wm L21 iðmðss ÞþxtÞ
uðs; tÞ ¼ e R ;
2p m¼1 D

where D ¼ L11 L22  L12 L21 is the determinant of the matrix in Eq. (D.4).

D.1. Radial displacement as a function of the rotational speed

The analytical response can be written as a function of the rotational speed, wðsÞ ¼ wðs; XÞ and uðsÞ ¼ uðs; XÞ. By consid-
ering the analytical displacements for a radial harmonic excitation, Eq. (D.5), due to the degenerate modes m, the spatial part
of the radial displacement can be written as:

Q wm ðL22 Dþ ei R m þ Lþ22 D eþi


ss ss
R
m
Þ
^
wðsÞ ¼ : ðD:6Þ
2pD Dþ
where þ and  indicates if the coefficient is evaluated at þm or m, respectively. Then, it is possible to express the relation
between output and input radial displacements as:

L Dþ ei R m þ Lþ22 D eþi


ss
^
ss
m
wðsÞ R
¼ 22 : ðD:7Þ
^ Þ
wðs L22 Dþ þ Lþ22 D


By assuming the degenerate frequency in which the optimal nonreciprocal response occurs as the natural frequency of
the stationary ring, Eq. (C.1a), i.e., the bifurcation center at low rotational speed, and using Eq. (D.2), L-terms can expressed
in terms of X:
Lþ11 ¼ l11;1 X þ l11;0 ; Lþ22 ¼ þl22;1 X þ l22;0 ; Lþ12 ¼ iðl12;1 X þ l12;0 Þ;
L11 ¼ þl11;1 X þ l11;0 ; L22 ¼ l22;1 X þ l22;0 ; L12 ¼ iðl12;1 X  l12;0 Þ; ðD:8Þ
Lþ21 ¼ Lþ12 ; L21 ¼ L12 ;

where l11;1 ¼ 2qARxst m, l11;0 ¼ EIR3 m2 þ EAR1 m2 þ Rbku  qARx2st , l22;1 ¼ l11;1 , l22;0 ¼ EIR3 m4  EAR1
Rbkw þ qARxst , l12;1 ¼ 2qARxst and l12;0 ¼ EIR3 m3  EAR1 m.
The determinants can also be rewritten as a function of X:

Dþ ¼ Lþ11 Lþ22  Lþ12 Lþ21 ¼ d2 X2 þ d1 X þ d0 ;


ðD:9Þ
D ¼ L11 L22  L12 L21 ¼ d2 X2  d1 X þ d0 ;

with d2 ¼ l12;1  l11;1 l22;1 , d1 ¼ 2l12;0 l12;1  l11;1 l22;0 þ l11;0 l22;1 and d0 ¼ l12;0 þ l11;0 l22;0 . Moreover, L þ þ 
2 2
22 D and L22 D can be rewrit-
ten as:

L22 Dþ ¼ þe3 X3 þ e2 X2 þ e1 X þ e0 ;
ðD:10Þ
Lþ22 D ¼ e3 X3 þ e2 X2  e1 X þ e0 ;
2 2 2 2 2 2
where e3 ¼ l12;1 l22;1 þ l11;1 l22;1 , e2 ¼ l12;1 l22;0  2l12;0 l12;1 l22;1  l11;0 l22;1 , e1 ¼ 2l12;0 l12;1 l22;0  l11;1 l22;0  l12;0 l22;1 and
þ 
L Lþ
2 2
e0 ¼ þ
l12;0 l22;0 l11;0 l22;0 .
Thus, þ ¼ 2ðe2 X þ e0 Þ.
22 D 22 D
2

As some structural damping is present, the elastic modulus assumes complex values which leads to complex L and D
terms. Then, it is possible to split the coefficients of the solutions into real and imaginary parts, as follows:
L22 Dþ
a ¼ aR þ iaI ¼ ;
L22 Dþ þ Lþ22 D

ðD:11Þ
Lþ22 D
b ¼ bR þ ibI ¼ ;
L22 Dþ þ Lþ22 D


where
R 5 R 4 R 3 R 2 R R I 5 I 3 I
aR ¼ þf 5 X þf 4gX Xþf4 þg3 X Xþf2 þg
2 X þf 1 Xþf 0
; aI ¼ þ fg5 XX4þfþg3 XX2þfþg1 X ;
4 2 0 4 2 0

R f R5 X5 þf R4 X4 f R3 X3 þf R2 X2 f R1 Xþf R0 I f I5 X5 þf I3 X3 þf I1 X
b ¼ g 4 X4 þg 2 X2 þg 0
; b ¼ g 4 X4 þg 2 X2 þg 0
;
1096 D. Beli et al. / Mechanical Systems and Signal Processing 98 (2018) 1077–1096

R R
2
2 R R R
with f 5 ¼ eR2 eR3 þ eI2 eI3 ; f 4 ¼ eR2 þ eI2 ; f 3 ¼ eR1 eR2 þ eI1 eI2 þ eR0 eR3 þ eI0 eI3 , f 2 ¼ 2ðeR0 eR2 þ eI0 eI2 Þ, f 1 ¼ eR0 eR1 þ eI0 eI1 ,
R 2 2 I I I 2 2
f0 ¼ ðeR0 Þ þ ðeI0 Þ , f5 ¼ eR2 eI3 þ eI2 eR3 , f3 ¼ eR1 eI2 þ eI1 eR2 þ eR0 eI3  eI0 eR3 , f1 ¼ eR0 eI1  eI0 eR1 , g 4 ¼ 2ððeR2 Þ þ ðeI2 Þ Þ,
2 2
g 2 ¼ 4ðeR0 eR2 þ eI0 eI2 Þ and g 0 ¼ 2ððeR0 Þ þ ðeI0 Þ Þ.

References

[1] D.L. Sounas, A. Alú, Angular-momentum biasing: a new paradigm for linear, magnetic-free, non-reciprocal devices, in: Antennas and Propagation
Society International Symposium (APSURSI), 2014, IEEE, 2014, pp. 1232–1233, http://dx.doi.org/10.1109/APS.2014.6904943.
[2] A. Maznev, A. Every, O. Wright, Reciprocity in reflection and transmission: what is a ’phonon diode’?, Wave Mot 50 (4) (2013) 776–784, http://dx.doi.
org/10.1016/j.wavemoti.2013.02.006.
[3] S.A. Cummer, J. Christensen, A. Alú, Controlling sound with acoustic metamaterials, Nat. Rev. Mater. 1 (16001) (2016) 13, http://dx.doi.org/
10.1038/natrevmats.2016.1.
[4] L.M. Nash, D. Kleckner, A. Read, V. Vitelli, A.M. Turner, W.T.M. Irvine, Topological mechanics of gyroscopic metamaterials, Proc. Natl. Acad. Sci. 112 (47)
(2015) 14495–14500, http://dx.doi.org/10.1073/pnas.1507413112.
[5] Y.-T. Wang, P.-G. Luan, S. Zhang, Coriolis force induced topological order for classical mechanical vibrations, New J. Phys. 17 (7) (2015) 073031.
[6] D.L. Sounas, C. Caloz, A. Alú, Giant non-reciprocity at the subwavelength scale using angular momentum-biased metamaterials, Nat. Commun. 4 (2407)
(2013) 1–7, http://dx.doi.org/10.1038/ncomms3407.
[7] Z. Yu, S. Fan, Complete optical isolation created by indirect interband photonic transitions, Nat. Photon. 3 (2009) 91–94, http://dx.doi.org/10.1038/
nphoton.2008.273.
[8] L. Onsager, Reciprocal relations in irreversible processes. i., Phys. Rev. 37 (1931) 405–426, http://dx.doi.org/10.1103/PhysRev.37.405.
[9] R. Fleury, D.L. Sounas, A. Alù, Subwavelength ultrasonic circulator based on spatiotemporal modulation, Phys. Rev. B 91 (2015) 174306, http://dx.doi.
org/10.1103/PhysRevB.91.174306.
[10] B. Liang, B. Yuan, J.-c. Cheng, Acoustic diode: Rectification of acoustic energy flux in one-dimensional systems, Phys. Rev. Lett. 103 (2009) 104301,
http://dx.doi.org/10.1103/PhysRevLett.103.104301.
[11] N. Boechler, G. Theocharis, C. Daraio, Bifurcation-based acoustic switching and rectification, Nat. Mater. 10 (9) (2011) 665–668.
[12] Q. Wang, Y. Yang, X. Ni, Y.-L. Xu, X.-C. Sun, Z.-G. Chen, L. Feng, X.-p. Liu, M.-H. Lu, Y.-F. Chen, Acoustic asymmetric transmission based on time-
dependent dynamical scattering, Sci. Rep. 5 (10880). doi:http://dx.doi.org/10.1038/srep10880.
[13] R. Fleury, D.L. Sounas, C.F. Sieck, M.R. Haberman, A. Alù, Sound isolation and giant linear nonreciprocity in a compact acoustic circulator, Science 343
(6170) (2014) 516–519, http://dx.doi.org/10.1126/science.1246957.
[14] C. Cohen-Tannoudji, B. Diu, F. Laloe, Quantum Mechanics, vol. 1, Wiley, New York, 1991.
[15] R.L. Comstock, Nonreciprocal Elastic Wave Coupling Network, US Patent 3,292,109, December 1966.
[16] P. Wang, L. Lu, K. Bertoldi, Topological phononic crystals with one-way elastic edge waves, Phys. Rev. Lett. 115 (2015) 104302, http://dx.doi.org/
10.1103/PhysRevLett.115.104302.
[17] G. Trainiti, M. Ruzzene, Non-reciprocal elastic wave propagation in spatiotemporal periodic structures, New J. Phys. 18 (8) (2016) 083047.
[18] S. Huang, W. Soedel, Effects of coriolis acceleration on the free and forced in-plane vibrations of rotating rings on elastic foundation, J. Sound Vib. 115
(2) (1987) 253–274, http://dx.doi.org/10.1016/0022-460X(87)90471-8.
[19] C.G. Cooley, R.G. Parker, Limitations of an inextensible model for the vibration of high-speed rotating elastic rings with attached space-fixed discrete
stiffnesses, Euro. J. Mech. - A/Solids 54 (2015) 187–197, http://dx.doi.org/10.1016/j.euromechsol.2015.06.012.
[20] J. Lee, S. Wang, P. Kindt, B. Pluymers, W. Desmet, Identification of the direction and value of the wave length of each mode for a rotating tire using the
phase difference method, Mech. Syst. Signal Process. 68–69 (2016) 292–301, http://dx.doi.org/10.1016/j.ymssp.2015.07.003.
[21] W.M. Lai, D. Rubin, E. Krempl, Introduction to Continuum Mechanics, Butterworth-Heineman, Burlington, 1999, pp. 556.
[22] J. Doyle, Wave Propagation in Structures: Spectral Analysis Using Fast Discrete Fourier Transforms, Mechanical Engineering Series, Springer, New York,
1997.
[23] S.-K. Lee, B. Mace, M. Brennan, Wave propagation, reflection and transmission in curved beams, J. Sound Vib. 306 (3–5) (2007) 636–656, http://dx.doi.
org/10.1016/j.jsv.2007.06.001.
[24] B.R. Mace, D. Duhamel, M.J. Brennan, L. Hinke, Finite element prediction of wave motion in structural waveguides, J. Acoust. Soc. Am. 117 (5) (2005)
2835–2843, http://dx.doi.org/10.1121/1.1887126.
[25] C. Kittel, P. McEuen, Introduction to Solid State Physics, John Wiley, New York, 2005.
[26] D. Beli, P.B. Silva, J.R.F. Arruda, Vibration analysis of flexible rotating rings using a spectral element formulation, J. Vib. Acoust. 137 (2015) 041003,
http://dx.doi.org/10.1115/1.4029828.
[27] K.M. Ahmida, J.R.F. Arruda, Spectral element-based prediction of active power flow in timoshenko beams, Int. J. Solids Struct. 38 (10–13) (2001) 1669–
1679, http://dx.doi.org/10.1016/S0020-7683(00)00128-1.
[28] R. Langley, Analysis of power flow in beams and frameworks using the direct-dynamic stiffness method, J. Sound Vib. 136 (3) (1990) 439–452, http://
dx.doi.org/10.1016/0022-460X(90)90455-9.
[29] H.B.G. Casimir, On onsager’s principle of microscopic reversibility, Rev. Mod. Phys. 17 (1945) 343–350, http://dx.doi.org/10.1103/RevModPhys.17.343.
[30] S. Huang, W. Soedel, Response of rotating rings to harmonic and periodic loading and comparison with the inverted problem, J. Sound Vib. 118 (2)
(1987) 253–270, http://dx.doi.org/10.1016/0022-460X(87)90524-4.
[31] U. Lee, Spectral Element Method in Structural Dynamics, Wiley, 2009.

You might also like