Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Mechanical Systems and Signal Processing 116 (2019) 443–461

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Nonlinear dynamic analysis using the complex nonlinear


modes for a rotor system with an additional constraint
due to rub-impact
Jie Hong a,b, Pingchao Yu a, Dayi Zhang a,c,⇑, Yanhong Ma a,b
a
School of Energy and Power Engineering, Beihang University, Beijing 100191, PR China
b
Collaborative Innovation Center of Advanced Aero-Engine, Beihang University, Beijing 100083, PR China
c
Beijing Key Laboratory of Aero-Engine Structure and Strength, Beijing 100083, PR China

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

Article history: A rub-impact can bring the impact load and friction force onto the rotor, and in the mean-
Received 18 July 2017 time, it can also produce the additional effect of changing the stiffness of the rotor system.
Received in revised form 11 May 2018 Past analysis and numerical simulation work related to rub-impact problems primarily
Accepted 28 June 2018
focus on the complicated nonlinear vibration responses caused by the rubbing force.
However, the influence of the additional effect on the rotor’s modal characteristics is rarely
studied. The present study investigates the modal characteristic of the rotor system with
Keywords:
the additional constraint and how it affects the rotor’s responses. The governing equation
Rub-impact
Additional constraint
for a modified Jeffcott rotor system with a rub-impact is established first. Next, the com-
Complex nonlinear modes plex nonlinear modes concept is introduced, and the corresponding solution method is
Stability derived. Finally, the modal characteristics are analyzed in detail, such as the modal
Dry whip frequency, the modal damping, the stability and the interaction between the nonlinear
mode motion and rotor’s responses. The results show that the rubbing rotor system
possesses both forward whirl mode motion and backward whirl mode motion. The magni-
tudes of the modal frequencies for both the forward whirl and backward whirl increase
with an increase in the amplitude of the mode motion (modal amplitude). Nevertheless,
the magnitudes are limited to an interval range, which can be approximately determined
through the linear rotor without rub-impact and the coupled linear rotor/stator system.
Differing from the forward whirl mode motion, the backward whirl mode motion can be
unstable, since its modal damping may be less than zero in certain cases. Moreover, the
instability of the backward whirl mode motion is only the primary physical mechanism
for the partial rub transmitting into dry whip.
Ó 2018 Elsevier Ltd. All rights reserved.

1. Introduction

The high-performance requirements for rotating machines, such as aero-engines, gas turbines and compressors, always
result in a decreasing clearance between the rotors and stators, which most likely results in a severe rub-impact during
operation. Rub-impact in rotating machines will produce high impact forces and friction that may lead to severe vibration
and even catastrophic failures in the worst-case scenarios [1,2].

⇑ Corresponding author at: School of Energy and Power Engineering, Beihang University, Beijing 100191, PR China.
E-mail address: dayi@buaa.edu.cn (D. Zhang).

https://doi.org/10.1016/j.ymssp.2018.06.061
0888-3270/Ó 2018 Elsevier Ltd. All rights reserved.
444 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

During the past several decades, a significant amount of research has been conducted on the mechanisms and the com-
plicated phenomena of a rub-impact event. Muszynaka [3] presented an excellent review on the rotor-to-stator rubbing con-
tact in rotating machinery. She noted that the rub-impact involves several physical phenomena, including most notably the
impact, friction and stiffness increase. As a result, the rotor becomes highly nonlinear and will exhibit extremely rich
dynamic behaviors, such as a partial rub, backward whirl, quasi-periodic and even chaotic motions [4–6]. Chu and Zhang
[7,8] investigated the non-linear vibration characteristics of a Jeffcott rotor with a rub-impact. These researchers found that
when the rotating speed was increased, the grazing bifurcation, quasi-periodic motion and chaotic motion occurred. Pennac-
chi studied the light and short arc rubs where the fixed part is less stiff than the rotor [9]. Roques investigated the speed
transients during rotor-to-stator rubbing caused by an accidental blade-off imbalance [10]. Jiang explored the existence
boundaries of different responses, e.g., a synchronous full annular rub, partial rub and dry whip for a rubbing rotor, and
an overall picture of the global response in the parameter space was obtained using an analytic method [11]. Ma studied
the rubbing-induced vibration responses based on contact dynamics theory under different rubbing types, such as a
multiple-point rub [12] and a full annular rub [13]. Moreover, Ma built a comprehensive model of a rotational shaft-
disk-blade system, and the effects of the rotational speed, stagger angle of the blade and the casing stiffness on the vibration
responses were discussed in detail [14–17].
Among the different contact responses of a rubbing rotor system, dry whip is the most destructive motion to rotating
machinery. During dry whip, the rotor undergoes a large deformation and is subjected to high-frequency stress, which will
initiate a break or fatigue damage of the shaft and cause a failure of the machine [18,19]. Black built a general model for a
synchronous rub to investigate dry whip and concluded that it was only possible in the frequency band, extending from an
individual rotor/stator natural frequency to the next higher combined system frequency [20]. Applying Black’s model to a
long-cantilevered disk, Zhang [21] accounted for multiple rotor modes in dry whip and dry whirl and identified the same
whirl regions as Black. Jiang studied the physical reason for the onset of dry whip and revealed that the rotor in resonance
at a negative frequency to the rotor-to-stator system results in a dry whip [18]. Childs and Kumar developed analytic dry
whip and dry whirl solutions for a rigid-rotor/rigid-stator model with contact at two rubbing locations [22].
In addition to the complicated dynamic behaviors caused by the impact and friction forces, the transient stiffness of the
rotor will also be increased in the process of the rub-impact. This phenomenon is usually called the stiffening effect [23] or
additional constraint effect [24,25]. Making use of this characteristic, Chu quantitatively analyzed the change of the transient
stiffness of a rotor based on the parameter identification theory and put forward an effective method to detect the rubbing
positions [23,26]. Ma built a constraint mechanical model of a full annular rub and their study showed that the constraint
effect made the rotor’s resonant range expand [24]. After that step, Hong proposed a dynamic modeling for the rotor system
with a non-smooth constraint in the case of an intermittent rub-impact and studied the effects of the non-smooth constraint
on the modal characteristics and vibration response of the rotor system [25,27]. Bently [28] and Child [29] demonstrated that
a partial rub causes a periodic variation of the rotor’s stiffness, which could lead to an instability of the rotor’s vibration.
When a rub-impact occurs, the modal frequency and modal shape of the rotor system change due to the additional con-
straint. However, there are a limited number of papers that attempt to analyze the modal characteristics of a rubbing rotor.
One important reason is that the traditional linear normal modes cannot be applied to this strong nonlinear system. Fortu-
nately, the nonlinear normal modes (NNMs), originated by Rosenberg [30] and developed by Shaw and Pierre [31], provide a
mathematical and practical framework for the analysis of the rubbing rotor system. Since the concept of nonlinear normal
modes were proposed, many approximate methods [32–34], such as asymptotic methods, multiple scale methods and
numerical methods, have been advanced and adopted to construct NNMs, which has strengthened the relevant theoretical
background. More recently, the nonlinear normal modes for the practical engineering systems with non-smooth nonlinear-
ities were investigated by several scholars [35,36]. D. Jiang built a numerical method for calculating the NNMs of piecewise
linear autonomous systems and investigated the existence, stability and bifurcations of the NNMs [37]. Jiang analytically
derived the nonlinear normal modes of a rubbing rotor, but the modal damping was ignored in his study [38]. Considering
the energy dissipation in non-conservative systems, Denis Laxalde defined the complex nonlinear modes that can include the
damping dissipation of the mechanical system and applied the concept to analyze turbo-machinery blades with friction [39].
As the influence of the additional constraint on modal characteristics of the rubbing system is not well-understood to
date, this paper aims at revealing the modal characteristics of the rubbing rotor system with additional constraints, as well
as how the modal characteristics affect the rotor’s rubbing motions. First, the governing equation of a widely used Jeffcott
rotor system with an additional constraint is established. Next, the eigen problem is derived in the frequency-domain by
introducing the complex nonlinear modes concept, and modal frequency and damping characteristics are determined
through numerical solving methods. The modal frequency and modal damping of the rubbing rotor system are obtained,
and the interaction between the nonlinear mode motion and rotor’s responses is analyzed in detail.

2. Governing equation for a rotor system with additional constraints

The rubbing rotor system used in the paper consists of a Jeffcott rotor and a stator, which is modeled with an added
stiffness as shown in Fig. 1. This model is widely used in the mechanism study of the rotordynamics [7,11,24,38] due to
its simplicity and its ability to reveal the basic mechanical characteristics of the rotor system. The rotor has a massless shaft
carrying a mid-span disk with mass m. The mass center of the rotor is located at a distance e from its geometrical center. The
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 445

kc
y
x
m
ω k
e
o z

r0

Fig. 1. Schematic representation of a Jeffcott rotor with a circular rigid stator.

stiffness of the shaft is k. The rotation speed of the rotor is x


 . The stator has a rigid ring with a negligible mass. The rigid ring
is supported elastically, the stiffness of the support is kc, and the damping is ignored. The clearance between the rotor and
the stator is r0.
The governing equations of the above rubbing rotor system can be written as
(  
m€x þ cx_ þ kx þ Hðr  r 0 Þkc 1  rr0 ðx  signðv rel ÞlyÞ ¼ mex
 2 cos x
t
  ð1Þ
my€ þ cy_ þ ky þ Hðr  r0 Þkc 1  r ðsignðv rel Þlx þ yÞ ¼ mex
r0
 2 sin x
t

where x is the horizontal displacement, and y is the vertical displacement, r is the radial displacement of the disk with
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r¼ x2 þ y2 , v rel is the relative speed at the contact point of the rotor with v rel ¼ xw r þ x
 r disk , l is the friction coefficient,
xw is the whirl angular velocity of the rotor and rdisk is radius of the disk. HðÞ is the Heaviside function, and its expression is
shown as Eq. (2). signðÞ is the symbolic function, which is shown as Eq. (3):

0 x60
HðxÞ ¼ ð2Þ
1 x>0
8
< 1 x < 0
>
signðÞ ¼ 0 x¼0 ð3Þ
>
:
1 x>0
As seen from Eq. (1), the rotor goes into contact with the stator only when the rotor’s vibration amplitude is greater than
the clearance. In this case, the additional constraint is applied to the rotor, as shown in Fig. 2.
The governing Eq. (1) can be further rewritten in the following non-dimensional form:
(  
X 00 þ 2nX 0 þ cX þ HðR  R0 Þ 1  RR0 ðX  signðV rel ÞlYÞ ¼ X2 cos Xs
  ð4Þ
Y 00 þ 2nY 0 þ cY þ HðR  R0 Þ 1  RR0 ðY þ signðV rel ÞlXÞ ¼ X2 sin Xs

Fig. 2. Additional constraint caused by a rub-impact.


446 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

where the non-dimensional variables are defined as


rffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
x y r r0 kc c x
 k
X ¼ ; Y ¼ ; R ¼ ; R0 ¼ ; x2 ¼ ; 2n ¼ pffiffiffiffiffiffiffiffiffi ; X ¼ ; c ¼ ; R ¼ X 2 þ Y 2 ; s ¼ x2 t;
e e e e m kc m x2 kc
r disk xw
V rel ¼ XRdisk þ Rx; Rdisk ¼ ;x ¼ ;
e x2
The prime indicates the differentiation with respect to the new time scale s.

3. Solution method based on the harmonic balance method

3.1. Eigenproblem in the frequency domain

In this section, the solution method of complex nonlinear modes for the rubbing rotor system is derived. First, the unbal-
ance forces on the right side of the governing equation are ignored, and the autonomous dynamical system of Eq. (4) is
obtained in the following forms:
(  
X 00 þ 2nX 0 þ cX þ HðR  R0 Þ 1  RR0 ðX  signðV rel ÞlYÞ ¼ 0
  ð5Þ
Y 00 þ 2nY 0 þ cY þ HðR  R0 Þ 1  RR0 ðY þ signðV rel ÞlXÞ ¼ 0

Inspired by the definition of complex modes for linear systems and the definition of nonlinear normal modes given by
Rosenberg [30], Shaw and Pierrre [31], the complex nonlinear modes concept was proposed by Laxalde D [39]. These
complex nonlinear modes are actually an extension of the nonlinear normal modes and the linear complex modes, and
the definition is an oscillation of the autonomous system with a phase difference between its degrees of freedom. Still, by
analogy with the linear complex modes, the eigenvalue of the characteristic equation can be defined in the following form
[39]:
k ¼ b þ ix ð6Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where n ¼ b=x0 is the modal damping ratio, x ¼ x0 1  n2 is the damped natural angular frequency, and x0 is the natural
angular frequency.
According the definition of complex nonlinear modes, the solutions of the free vibration problems for Eq. (5) can be
expanded in the following Fourier series:
8
>
> X
l
>
> X ¼ a0x þ ekbs ðbx cos kxs þ akx sin kxsÞ
k
>
<
k¼1
ð7Þ
>
> X
l
>
> ekbs ðby cos kxs þ aky sin kxsÞ
k
: Y ¼ ay þ
0
>
k¼1

In Eq. (7), the damping b takes into account the energy decay of the nonlinear mode motion caused by the linear damping
and the nonlinear rubbing force. If only the first harmonic k = 1 is retained, the solution degenerates to the well-known expo-
nential assumption for damped linear systems. Moreover, for conservative autonomous systems, the damping term b is zero,
and Eq. (7) is completely equivalent to the conventional harmonic balance method.
The nonlinear rubbing forces in the X direction and Y direction are in the following forms according to Eq. (5).
(
F x ¼ HðR  R0 Þð1  RR0 ÞðX  signðV rel ÞlYÞ
ð8Þ
F y ¼ HðR  R0 Þð1  RR0 ÞðY þ signðV rel ÞlXÞ

Eq. (8) exhibits that the rubbing forces only depend on the rotor’s vibration displacement in the X direction and Y
direction. Therefore, these forces can also be expanded as Fourier series, which are similar to the following displacements:
8
>
> X
l
>
> F x ¼ p0x þ ekbs ðqkx cos kxs þ pkx sin kxsÞ
>
<
k¼1
ð9Þ
>
> X
l
>
> ekbs ðqky cos kxs þ pky sin kxsÞ
: F y ¼ py þ
0
>
k¼1

By substituting Eqs. (7) and (9) into Eq. (5) and letting all of the sine terms and cosine terms be equal to zero, the
following nonlinear algebraic equations are obtained:
KA0 þ P0 ¼ 0 ð10Þ
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 447

" #   
k ðb2  x2 ÞM  kbD þ K 2k bxM  kxD
2 2
Ak Pk
þ ¼ 0 k ¼ 1;    ; l ð11Þ
kxD  2k bxM k ðb2  x2 ÞM  kbD þ K
2 2 Bk Qk

where M, D and K denote the mass matrix, damping matrix and stiffness matrix, respectively. A0 and P0 are the constant
components of displacement and nonlinear rubbing force, respectively. ½Ak Bk T and ½Pk Q k T are the k-th order components
of the displacements and nonlinear rubbing forces. The expressions of the above variables are as follows:
      ! ! ! ! ! !
1 2n c a0x p0x akx
k
bx pkx qkx
M¼ ;D ¼ ;K ¼ ; A0 ¼ ; P0 ¼ ; Ak ¼ ; Bk ¼ ; Pk ¼ ; Qk ¼
1 2n c a0y p0y aky k
by pky qky

One can rearrange the Eqs. (10) and (11) as following form in order to facilitate numerical solving:
2 32 3 2 3
K A0 P0
6 K1 76 Z 7 6 H 7
6 76 1 7 6 1 7
6 .. 76 . 7 6 . 7
6 76 . 7 6 . 7
6 . 76 . 7 6 . 7
6 76 7 þ 6 7
6 Kk 76 Z 7 6 H 7 ¼ 0 ð12Þ
6 76 k 7 6 k 7
6 .. 76 . 7 6 . 7
6 76 . 7 6 . 7
4 . 54 . 5 4 . 5
Kl Zl Hl
where
" #    
k ðb2  x2 ÞM  kbD þ K 2k bxM  kxD
2 2
Ak Pk
Kk ¼ ; Zk ¼ ; Hk ¼
kxD  2k bxM
2
k ðb2  x2 ÞM  kbD þ K
2 Bk Qk

The nonlinear algebraic Eq. (12) can be solved using a Newton method. As the method is a local convergence algorithm, the
initial value is highly important and chosen as the eigen-solution of the corresponding linear system without a rub-impact in
this paper. In addition, for the complex eigen-problem (12) to be solved, we need to address two additional problems.First, the
number of unknowns in Eq. (12) exceeds the number of equations by two. To overcome this issue, one can define the modes
normalization similar with the normalization of modes in linear systems. In the present study, the modes are normalized with
1 T
respect to a chosen degree of freedom (DOF) X, as shown in Eq. (13), where ½a1x bx  is the first order component of the displace-
  T
ment X, and ½a b  is the specified number given in advance. It should be noted that the DOF Y can also be chosen to perform the
mode normalization, which does not change the results of modal frequency and modal damping. In theory, the mode normal-
ization can be performed with respect to an arbitrary DOF. Moreover, the mode normalization actually includes the amplitude
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 
normalization and phase normalization. The given amplitude is a2 þ b , and the given phase is arctan b =a .
" #  
a1x a
1
¼  ð13Þ
bx b

Second, the nonlinear rubbing forces are the explicit function of the displacement only in the time domain. Therefore, the
frequency-domain components of the nonlinear rubbing forces cannot be derived from the frequency-domain components
of the displacement directly. To address this issue, an alternating frequency-time domain (AFT) scheme is adopted in the
present study. This scheme is widely used in the harmonic balance method, which is described in detail in the Refs. [39–41].
The stability of the nonlinear mode motion is determined by the real parts of the corresponding eigenvalues and its stable
condition can be expressed as
real ðkÞ ¼ b < 0 ð14Þ

3.2. Numerical scheme

According to the deduction in Section 3.1, solving the complex nonlinear modes turns into solving the nonlinear algebraic
equations. The nonlinear algebraic equations are shown in Eq. (15):
8
> HðZ; x; bÞ ¼ Kðx; bÞZ þ bðZÞ ¼ 0
<" #  
a1x a ð15Þ
>
: 1   ¼0
bx b

In Eq. (15), HðZ; x; bÞ ¼ Kðx; bÞZ þ bðZÞ ¼ 0 is the harmonic balance equation shown in Eq. (12), where
T T 1 T   T
K ¼ diagðK; K1 ;    ; Kk ;    ; Kl Þ, Z ¼ ½A0 Z1    Zl  , b ¼ ½P0 H1    Hl  . ½a1x bx 
 ½a b  ¼ 0 is the equation of the modes
normalization. For the convenience of deducing and illustrating, Eq. (15) is rewritten as the following expression:
448 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

FðXÞ ¼ 0 ð16Þ
(
T Kðx; bÞZ þ bðZÞ
where X ¼ ½ZT bx , FðXÞ ¼ 1 T  T .
½a1x bx   ½a b 
Before we start to solve Eq. (16), the values of a⁄ and b⁄ for modes normalization should be given in advance according to
the vibration energy or vibration amplitude that we care about. Taking the rubbing rotor system in this article as an example,
if we want to obtain the modal characteristics of the rubbing rotor system when the vibration amplitude is equal to 1.2, the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 T 2
values of ½a b  should satisfy a2 þ b ¼ 1:2, and they are chosen as ½01:2T in this paper. It should be noted that other
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
values those satisfy a2 þ b ¼ 1:2 will obtain same modal results in the calculation.
Next, a Newton-Raphson solver is used to solve the nonlinear algebraic equations, and the iterative manner used is as
follows:
!1
@FðXðkÞ Þ
Xðkþ1Þ ¼ XðkÞ  FðXðkÞ Þ ð17Þ
@X

The initial value Xð0Þ is formed according to the results of the linear modal analysis. The Jacobi matrix @FðXðkÞ Þ=@X is numer-
ically evaluated with a finite difference scheme. Meanwhile, during the process of iterating, the AFT scheme is used to obtain
the nonlinear rubbing forces in the frequency domain [28]. It should be noted that once the convergence solution is obtained,
the modes normalization regarding the given a⁄ and b⁄ is also achieved.

4. Complex nonlinear modal analysis

In this section, the modal characteristics are calculated and analyzed based on the solution method built in the previous
section. The calculation parameters are given as n ¼ 0:05, c ¼ 0:1, R0 ¼ 1:05, l ¼ 0:15, Rdisk ¼ 20R0 , X ¼ 0:1, and the number
of harmonic terms k is chosen as 5 in following study. The natural angular frequency of the linear rotor system without a rub-
impact for given parameters is 0.3162; therefore, the rotation speed in this instance is a subcritical speed.First, the nonlinear
mode motion characteristics of the rubbing rotor system are analyzed. Since the orbital characteristics of the mode motions
under different modal amplitudes are all similar, only the mode motion orbits with the modal amplitude being 4 are given, as
shown in Fig. 3. It should be noted that the modal damping term ekbs is not considered in Fig. 3 in order to observe the mode
motion characteristics more clearly. Unlike general vibration system, the mode motions reflect that the rotor system pos-
sesses two possible whirling directions, i.e., forward whirl and backward whirl. In addition, we know that for both the for-
ward whirl and backward whirl the mode motion orbits are all circular, meaning that the rotor is in constant contact with
the stator and no partial rub exists for the mode motion of a rubbing rotor system.
Different from the linear system, the modal frequency and modal damping of the rubbing rotor are strongly dependent on
the rotor’s modal amplitude and vibration energy. Therefore, the modal frequency and modal damping for a rubbing rotor
will be discussed in detail in this section.

4.1. Modal frequency

Fig. 4 shows the evolution of the modal angular frequency as a function of the modal amplitude for the forward whirl and
backward whirl mode motions. One can see that the magnitudes of the modal frequency for the forward whirl and backward

Fig. 3. (a) Forward whirl, (b) backward whirl mode motions for the rubbing rotor.
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 449

Fig. 4. Modal angular frequency vs. the modal amplitude for the rubbing rotor.

whirl mode motions at any certain modal amplitude are equal, and only their symbols which represent different whirl direc-
tions are opposite. When the modal amplitude is less than the clearance, no rub-impact occurs, and the modal angular fre-
quency is always equal to that of the linear rotor system. Once the modal amplitude is greater than the clearance, both the
forward whirl and backward whirl modal frequencies increase for an increasing modal amplitude. The reason is that the
stiffness of the rotor system increases due to the additional constraint produced by the rub-impact. However, we should note
that modal frequency increases increasingly slowly as the modal amplitude rises, and it seems that the modal frequency
reaches its maximum value when the modal amplitude reaches infinity.The above analysis indicates that the modal fre-
quency for a rubbing rotor system occurs at a certain interval. The minimum of the interval is the modal frequency value
of the linear rotor system without a rub-impact, and the maximum of the interval, which can be proven in following deriva-
tion, is approximately equal to the modal frequency of the coupled linear rotor/stator system. The coupled linear rotor/stator
system refers to the dynamical system in which the rotor touches the stator through the linear spring kc , and the friction at
the contact point is neglected. To determine the maximum of the interval, a derivations can be conducted with the help of
the above numerical results.First, we can take the forward whirl mode motion as an example since the magnitude of its
modal frequency is the same as that of the backward whirl motion. As the relative velocity V rel ¼ XRdisk þ xR at the contact
point is always greater than zero due to the whirl angular frequency x > 0, Eq.(5) is simplified to the following form with
signðV rel Þ ¼ 1 and HðR  R0 Þ ¼ 1:
(  
X 00 þ 2nX 0 þ cX þ 1  RR0 ðX  lYÞ ¼ 0
  ð18Þ
Y 00 þ 2nY 0 þ cY þ 1  RR0 ðY þ lXÞ ¼ 0

Next, noticing that the maximum of modal frequency corresponds to the modal amplitude being infinity, indicating R  R0
 
and 1  RR0  1, Eq. (18) can be further written as

X 00 þ 2nX 0 þ cX þ ðX  lYÞ ¼ 0
ð19Þ
Y 00 þ 2nY 0 þ cY þ ðY þ lXÞ ¼ 0
Let us define W ¼ X þ iY, and the complex equation of Eq. (19) is obtained:
W 00 þ 2nW 0 þ cW þ ð1 þ jlÞW ¼ 0 ð20Þ
According to the numerical results in Fig. 3, we can obtain that only the fundamental component of modal frequency
exists in the frequency domain. Therefore, the complex solution is simplified as the following form with only the first har-
monic k = 1 being retained in Eq. (7):

W ¼ Heks ð21Þ
where H is positive real number. By substituting Eq. (21) in to Eq. (20) and letting the real parts and imaginary parts be zero,
one can obtain
(
b2  x2  2nb þ c þ 1 ¼ 0
ð22Þ
2bx  2nx  l ¼ 0

Based on Eq. (22), the quartic equation with the unknown variable x can be obtained after several manipulations:
450 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

l2
x4 þ ðn2  c  1Þx2  ¼0 ð23Þ
4
The solutions of the Eq. (23) are as follows:
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u 2
tðc þ 1  n2 Þ þ ðc þ 1  n2 Þ þ l2
x1;2 ¼ ð24Þ
2
vqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u 2
t ðc þ 1  n2 Þ þ l2  ðc þ 1  n2 Þ
x3;4 ¼ i ð25Þ
2

It should be noted that the values of x3 and x4 are imaginary numbers, therefore only the solutions x1 and x2 whose values
are real numbers are meaningful. Based on Eq. (24), we can see that the maximum of the modal angular frequency is only
determined by the rotor/stator stiffness ratio c, the friction coefficient l and damping n. Although the precise solution of the
maximum of the modal angular frequency is obtained in Eq. (24), this result is not easy to extend to the more complicated
rubbing rotor system because the physical meaningfulness of the expression is not obvious. Generally, the damping n of the
rotor system is usually on the orders of 103–102, and the friction coefficient l in a rotor/stator contact system is usually
approximately 0.05–0.3 [3,5]. In this case, one can note that the values of n2 and l2 are less than the value of c þ 1 or ðc þ 1Þ2
by approximately 1–2 orders of magnitude. Therefore, the damping n and the friction coefficient l in Eq. (24) can be ignored,
and the approximate solution of the maximum of modal angular frequency for the rubbing system can be simplified as
pffiffiffiffiffiffiffiffiffiffiffiffi
x~ max ¼ c þ 1, which is only the natural angular frequency of the coupled linear rotor/stator system. Using the calculation
parameters in this article as an example, it can be observed that the approximate solution (x ~ max = 1.049) of the maximum of
the modal angular frequency is nearly equal to the precise solution (xmax = 1.05) of the maximum of the modal angular fre-
quency presented in Eq. (24). Based on the above analysis, it is concluded that the modal angular frequency of the coupled
linear rotor/stator can successfully predict the maximum of the modal annular frequency. This conclusion is very meaningful
for the estimation of the maximum modal frequency in the complicated rotor/stator contact system of the practical engi-
neering machine because it is easy to obtain the modal frequency of the coupled linear rotor/stator system through tradi-
tional linear modal analysis. Furthermore, the influence of key parameters, such as friction coefficient l, damping n,
rotation speed X and rotor/stator stiffness ratio c, on modal characteristics of the rubbing rotor system is calculated based
on the solution method in Section 3. In the calculation, the values of the parameters are same with those in above calculation
if they are not mentioned specially. The results are shown in Fig. 5. Since the magnitudes of the modal angular frequency for
forward whirl and backward whirl mode motions are equal, only the modal frequency of the backward whirl is presented in
Fig. 5. The following conclusions can be drawn from Fig. 5:
1) With the increase of the friction coefficient and the decrease of the damping, the magnitude of the modal angular fre-
quency of the rubbing rotor system increases slightly, however the variation is very small. Therefore, the maximum of
the modal frequency is nearly independent of l or n, which also proves above deduction.
2) The modal angular frequencies under different rotation speeds are equal, thus only one modal angular frequency curve
can be seen in Fig. 5(c). It means that the rotation speed does not affect the modal angular frequency of the rubbing
rotor system.
3) The rotor/stator stiffness ratio reflects the strength of the additional constraint caused by a rub-impact. If the rotor/
stator stiffness ratio is higher, then the additional constraint is weaker, because the rotor’s stiffness is greater than
stator’s stiffness in that case. The results shown in Fig. 5(d) show that as the rotor/stator stiffness ratio increases,
the magnitude of the modal angular frequency increases significantly. Meanwhile, the interval of the modal angular
frequency decreases. The reason is that when the rotor/stator stiffness ratio increases, the rotor’s stiffness enhances
relatively, and the additional constraint caused by the rubbing becomes weaker. Moreover, the asymptotic values
of the modal angular frequency under different rotor/stator stiffness ratios in Fig. 5(d) are essentially equal to
pffiffiffiffiffiffiffiffiffiffiffiffi
c þ 1, which validates the effectiveness of the conclusions in the above deduction once again.

Due to the interval feature of the modal frequency, it is very convenient to determine the range of the modal frequency for
a rubbing rotor system through the linear rotor system without a rub-impact and a coupled linear rotor/stator system. This is
very meaningful to the complicated practical engineering system, because if we calculate its modal frequency under every
modal amplitude using the harmonic balance method, the consumed time is remarkable, and converging the solution is
always a problem for the calculation.

4.2. Modal damping and stability

Fig. 6 presents the evolution of the modal damping as a function of the modal amplitude for the forward whirl and back-
ward whirl mode motions. It can be seen that when the modal amplitude is less than the clearance, the modal damping is
constant and equal to n. When the modal amplitude exceeds the clearance, the modal damping for both the forward whirl
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 451

Fig. 5. Influence of (a) friction coefficient, (b) damping, (c) rotation speed, and (d) rotor/stator stiffness ratio on the modal angular frequency for the
backward mode motions.

0.2
Forward whirl
Backward whirl
Modal damping

0.1

Jump up

0.0
<0
Clearance R 0
0 5 10 15
Modal amplitude R
Fig. 6. Modal damping vs. the modal amplitude for rubbing rotor.

and backward whirl mode motions is changed due to the rubbing friction force, but the variation law is different. For the
forward whirl mode motion, the modal damping constantly increases as the modal amplitude increases, with the increasing
tendency becoming slower and slower. While for the backward whirl mode motion, the modal damping first decreases, then
452 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

Fig. 7. Direction of the friction force under (a) forward whirl, (b) backward whirl with XRdisk > jxjR, (c) backward whirl with XRdisk < jxjR.

(a) (b) 0.20


=0.05
=0.25
=0.5 0.15

Modal damping
Modal damping

"A" "B" "A" "B"


0.10

=0.02
0.05 =0.05
0.0
0 =0.1
0.00

-0.05
0 5 10 0 5 10 15
Modal amplitude R Modal amplitude

(c) 0.20 (d) 0.15

=0.1
=0.3
0.15 =0.5
Modal damping

0.10
Modal damping

0.10

0.05 =0.1
0.05 =0.5
=1

0
0.00 0.00

0 5 10 15 0 5 10 15
Modal amplitude Modal amplitude R

Fig. 8. Influence of (a) friction coefficient, (b) damping, (c) rotation speed, and (d) rotor/stator stiffness ratio on the modal damping for backward whirl
mode motion.

jumps up to a positive value at a certain modal amplitude, after which the modal damping is always identical to that of the
forward whirl mode motion. Moreover, we can note that the modal damping of the forward whirl mode motion is always
greater than zero, indicating that the forward whirl mode motion is stable at any modal amplitude. However, the backward
whirl modal damping can be less than zero during the process of decreasing the modal amplitude, which means that the
backward whirl mode motion may be unstable in some cases. As we can see from Section 5, the instability of the backward
whirl mode motion is just the reason leading to the appearance of the dry whip.
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 453

The change law for modal damping shown in Fig. 6 can be explained from the view of work done by the friction force. For
the forward whirl mode motion, the velocity V rel ¼ XRdisk þ xR is always greater than zero, thus the direction of the friction
force is opposite with the whirl direction (see Fig. 7 (a)), and the friction force does negative work on the rotor’s forward
whirl. That means the friction force herein will dissipate the rotor’s forward whirl energy and increase the modal damping
of the rotor system. While for the backward whirl, the velocity V rel at the contact point is XRdisk  jxjR. If the modal ampli-
tude is small, the velocity V rel is greater than zero, meaning the direction of the friction force is coincident with the whirl
direction (see Fig. 7(b)), and friction force does positive work on the rotor’s backward whirl. Therefore, the friction force will
raise the energy of backward whirl mode motion and the modal damping decreases simultaneously in that case. As the
modal amplitude increases, the work done by the friction force increases, consequently the modal damping constantly
decreases. Furthermore, if the work done by the friction force exceeds the energy dissipation caused by rotor’s original
damping, then b < 0 and the corresponding backward whirl mode motion will be unstable. Once the modal amplitude
reaches the value that makes XRdisk < jxjR (see Fig. 7(c)), the direction of the friction force changes immediately and the fric-
tion force does negative work similar to the case of the forward whirl. Therefore, it is concluded that the value corresponding
to the jumping point in Fig. 6 is the modal amplitude that satisfies XRdisk ¼ jxjR.
Fig. 8 shows the influence of key parameters on the modal damping of the backward whirl mode motion. In the figure,
there exists a region where the modal damping decreases with an increase of the modal amplitude, called region ‘‘A”. The
region where the modal damping increases with an increase of the modal amplitude is called region ‘‘B”. The following con-
clusions are summarized according to the results:

1) With the increase of the friction coefficient and decrease of the damping, the modal damping at a certain modal ampli-
tude in region ‘‘A” decreases significantly, whereas the modal damping at a certain modal amplitude in region ‘‘B”
increases obviously. In addition, the modal amplitude corresponding to the jump point hardly changes with variation
of the friction coefficient and damping, and the ranges of the modal amplitude in region ‘‘A” are also basically invari-
able. The reason for this change law is that the modal amplitude corresponding to the jump point only depends on the
parameters X, Rdisk and jxj according to the above analysis.
2) Both the region where the modal damping decreases with an increase of the modal amplitude and the region where
b < 0 increases for an increase in the rotor’s rotation speed, which indicates that the stability decreases when the rotor
is operating at high rotation speeds. This effect can be explained as the following: if the rotation speed X increases,
then the ranges of the modal amplitude with V rel ¼ XRdisk  jxjR > 0 expands; as a consequence, the modal amplitude
at which the jumping up phenomenon appears increases.
3) The rotor/stator stiffness ratio can also affect the modal damping, because it is closely related to the whirl speed x.
According to Fig. 8(d), both the region where the modal damping decreases with the increase of the modal amplitude
and the region with b < 0 decreases with an increase of the rotor/stator stiffness ratio, which indicates that the sta-
bility of the backward whirl mode motion enhances when the rotor/stator stiffness ratio increases. Meanwhile, one
can also observe that if the modal amplitude is in the region where the modal damping decreases with an increase
of the modal amplitude, the corresponding modal damping will increase with an increase of the rotor/stator stiffness
ratio. This phenomenon can be explained from the view of the work done by the friction force. If the rotor/stator stiff-
ness ratio increases, the radial impact force will be reduced, which consequently leads to the decrease of the friction
force and the work done by the friction force.

(a) (b)

1
10 Dry whip
Displacement Y

Clearance
Partial
Amplitude R

Full Clearance
rub
annular
-1
10 rub
No rub
Ω=0.65
Ω=0.42
Ω=0.24

No rub Full Quasi-periodic Quasi-periodic


annular partial rub dry whip
rub
-3
10
0.0 0.2 0.4 0.6 0.8 1.0

Roation speed Ω Rotation speed Ω


Fig. 9. (a) Vibration amplitude, and (b) bifurcation diagram vs. the rotor’s rotation speed.
454 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

5. Nonlinear vibration response under an unbalance excitation

To date, a large number of papers have studied the nonlinear response of a rubbing rotor, such as the jump phenomenon,
the partial rub, and the chaotic motion, as well as dry whip [1–4]. For the model illustrated in Section 2, Jiang performed a
great deal of theoretical work and obtained many meaningful results [11,18]. Jiang noted that this model is able to exhibit
four classical types of steady-state rotor behavior, i.e., no rub motion, full annular rub, partial rub and dry whip. Meanwhile,
the existence boundary of corresponding typical responses is determined in the parameter space using an analytical tool.
Identical results have also been reported by other authors in their numerical simulation and experimental investigation
[19,42,43]. In this section, the response characteristics are simulated, and similar results are gained. After a detailed analysis
of the responses with an unbalance excitation, the interaction between the nonlinear modes and the vibration response is
studied, and the mechanism for the partial rubbing converting into dry whip is determined.

5.1. Response characteristics with variation of rotation speed

Eq. (4) is solved based on the traditional Newark method, and the vibration amplitudes during the running up of the rotor
system are obtained, as shown in Fig. 9(a). Meanwhile, the boundary for a different typical response is also determined
through observing the rotor’s orbit under each X (see Fig. 10), along with a bifurcation diagram under each X interval
(see Fig. 9(b)). In the simulation, the calculation parameters are same with those in Section 4, and sufficient time for each
rotation speed X is simulated in order to reach the steady-state responses, then results of last fifty revolutions are saved.
The results in Fig. 9 show that the rotor’s response sequence with an increase of the rotation speed inspired by the unbalance
force is no rub motion ? full annular rub ? partial rub ? dry whip. When the rotation speed is less than 0.24, the rotor’s
amplitude is less than clearance and no rub-impact occurs. If the rotation speed is 0:24
0:42, a full annular rub will appear.
For the partial rub, the rotation speed region is 0:42
0:65. Once the rotor’s speed is greater than 0.65, dry whip is
generated.
From the vibration response characteristics presented in Fig. 10, we can see that the orbit of the full annular rub is circular
and there only exists the unbalance excitation component funbalance in the frequency domain, indicating that this rubbing
motion is dominated by the unbalance force. While for the partial rub and dry whip, there is another frequency component
f2 due to the rubbing force. As we can see, the amplitude of frequency component f2 is greater than that of funbalance in the case
of the partial rub with a higher rotation speed and dry whip motion as well, showing that the rubbing force plays an impor-
tant role on the rotor’s response for this case. In particular, for the case of dry whip motion, funbalance can be ignored compared
with f2 and this response is self-excited due to the rubbing force.
Furthermore, the motion orbits of the harmonic component f2 for a partial rub and dry whip are extracted using bandpass
filtering. Since the characteristics of the orbits under different rotation speeds are same, only the result for X ¼ 0:45 is given
as an example, which can be seen in Fig. 11. The result shows that the precession direction is opposite of the rotation direc-
tion, therefore this harmonic component f2 corresponds to a backward whirl motion. As this backward whirl is mainly
excited by a rubbing force, a question arises: are there connections between the backward whirl mode motion and the back-
ward whirl of the harmonic component f2? That question will be discussed in the next subsection.

5.2. Influence of complex nonlinear mode on rotor response

To examine the relationship between the backward whirl mode motion and partial rub/dry whip, corresponding complex
nonlinear modes are calculated. As nonlinear modes of a rubbing rotor system are dependent on the system’s energy, the
mechanical energy of the partial rub and dry whip must be obtained in advance. For the rubbing rotor system illustrated
in Section 2, its mechanical energy includes the rotor’s potential energy, kinetic energy and the stator’s potential energy,
as shown in Eq. (26):
1 1 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2
E¼ mðx_ 2 þ y_ 2 Þ þ kðx2 þ y2 Þ þ Hðr  r0 Þ kc ð x2 þ y2  r 0 Þ ð26Þ
2 2 2
where x is the horizontal displacement, and y is the vertical displacement of the rotor. The solutions of x and y can be
obtained by integrating the governing equation of the rubbing rotor system.
The mechanical energy at any moment for dry whip in Fig. 10(d) is gained based on Eq. (26), and the result is shown in
Fig. 12. It can be found that the mechanical energy for the rubbing rotor system varies periodically, indicating that the
mechanical energy is not conserved in the rubbing rotor. The mechanism for the periodic varying of the mechanical energy
is that the energy dissipated by the rotor’s damping reaches a dynamic balance with the energy input by the friction force,
i.e., there exists a minimum of the critical value and a maximum of the critical value. When the vibration energy reaches
the minimum of the critical value, such as 172.87 in Fig. 12, the energy input by the friction force is greater than the energy
dissipated by the rotor’s damping, therefore the vibration energy of the rotor system will increase. Until the vibration energy
reaches the maximum of the critical value, the energy dissipated by the rotor’s damping exceeds the energy input of the fric-
tion force, or even both the rotor’s damping and the friction force dissipate the energy of the rotor system. Consequently, the
vibration energy of the rotor system will decrease until it reaches the minimum of the critical value. Next, the above process is
repeated. The cases for that the friction force dissipates energy and inputs energy are like those presented in Fig. 7(b) and (c).
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 455

(a) 2 2
Axis orbit
Clearance circle

Displacement Y
funbalance

Amplitude
0 1

-2 0
-2 0 2 0.0 0.1 0.2 0.3

Displacement X Frequency

(b) 3 2
Axis orbit
Clearance circle
Displacement Y

funbalance

Amplitude
0 1

f2:0.136

-3 0
-3 0 3 0.0 0.1 0.2 0.3
Displacement X Frequency

(c) 5 2
Axis orbit
Clearance circle

f2:0.123
Displacement Y

funbalance
Amplitude

0 1

-5 0
-5 0 5 0.0 0.1 0.2 0.3

Displacement X Frequency

(d) 22 Axis orbit 15


Clearance circle f2:0.162
Displacement Y

10
Amplitude

funbalance
-22 0
-22 0 22 0.0 0.1 0.2 0.3

Displacement X Frequency
Fig. 10. Orbits and vibration responses in the frequency domain for a rotor at (a) X ¼ 0:35, (b) X ¼ 0:45, (c) X ¼ 0:65, (d) X ¼ 0:7.
456 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

0.5

Precession

Displacement Y
direction
0.0

Rotation
direction

-0.5
-0.5 0.0 0.5

Displacement X
Fig. 11. Motion orbit corresponding to the harmonic component f2 for X ¼ 0:45.

200

190.14
Mechanical energy E

180

172.87

160
1700 1720 1740 1760
Time

Fig. 12. Mechanical energy for dry whip motion at X = 0.7.

Since we only care about the energy level for the certain rubbing motion, the interval range is used to describe the
mechanical energy for the rubbing motion possessing. Consequently, the mechanical energy for this dry whip is denoted
by [172.87, 190.14].
The modal frequency and modal damping of the rubbing rotor are calculated for X = 0.45, 0.55, 0.65 and 0.66, as shown in
Fig. 13. In this figure, the modal frequency curves at different rotation speeds coincide with each other; therefore, we can
only observe one curve in the figure. For the modal damping, there also exists a large region where the modal damping val-
ues at different rotation speeds are equal to each other. It should be noted that the horizontal coordinate here is mechanical
energy. Meanwhile, the mechanical energy intervals of the rubbing motions under these rotation speeds are also marked in
Fig. 13. Moreover, based on the results of the rotor’s responses and modal characteristics, the key characteristic parameters
including the mechanical energy interval, harmonic frequency value f2 for the rubbing motion, modal angular frequency and
modal damping under corresponding rubbing motions are extracted and summarized in Table 1.
The results in Fig. 13 and Table 1 can be concluded as follows:

1) From Table 1, the vibration amplitude increases, and the corresponding mechanical energy of the rubbing motion
increases as well for the increasing of the rotation speed. Meanwhile, for the partial rub with the rotation speed being
at 0.42–0.65, the harmonic frequency value f2 decreases for an increase in the rotation speed. Once the dry whip occurs
when rotation speed exceeds 0.65, this harmonic frequency value suddenly increases.
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 457

(b) 0.25
Energy interval for
Energy interval for
0.20
Energy interval for
Energy interval for

Modal damping
0.15

0.10
Modal damping
0.05 decreasing

0.00

-0.05
0.01 0.1 1 10 100
Mechanical energy E
Fig. 13. (a) Modal angular frequency, (b) modal damping for the rotor’s rubbing motions under different rotation speeds.

Table 1
Key characteristic parameters for the rotor’s response and corresponding nonlinear modes under different rotation speeds.

Rotation speed X Response with an unbalance force Nonlinear modal characteristics


Mechanical energy interval Harmonic frequency f2 Modal frequency x/2p Modal damping b
0.45 [0.3462, 0.4426] 0.1360 [0.0823, 0.0901] [0.022, 0.0225]
0.55 [0.8136, 1.1320] 0.1270 [0.1047, 0.1111] [0.0098, 0.01434]
0.65 [1.8492, 2.6253] 0.1230 [0.1210, 0.1268] [0, 0.0037]
0.66 [171.27, 188.24] 0.1607 [0.1607, 0.1612] [0.0184, 0.1185]

20
Axis orbit
Clearance circle
Ω=0.66
Transient
10 process
Displacement Y

-10
Ω=0.65

-20
-20 -10 0 10 20
Displacement X

Fig. 14. The process of partial rub transmitting to dry whip motion.

2) In Fig. 13, the bars with different colors denote the mechanical energy intervals that rubbing motions possess at
different rotation speeds. Thus, the intersection regions for bars and the modal frequency curve in Fig. 13(a) are
the modal frequency ranges of the corresponding rubbing motions. It can be seen that the magnitudes of the modal
frequency of the rubbing motion are raised significantly as the rotation speed increases. By comparing the harmonic
frequency value of the rubbing motion and the corresponding modal frequency (see Table 1), one can observe that the
harmonic frequency approaches the modal frequency with the increase of the rotation speed. Until the partial rub at
X = 0.65 and dry whip at X = 0.66, their values are basically equal, indicating that this harmonic frequency component
458 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

is just the backward whirl mode motion at that time. Hence, we can infer that the backward whirl mode motion will
be excited for the partial rub with a higher rotation speed or dry whip. Meanwhile, this mode motion will become the
dominant factor of the rotor’s response compared with the unbalance excitation component.
3) In Fig. 13(b), the intersection region for the bar and modal damping curve is the modal damping range for the corre-
sponding rubbing motions. As we can see, the modal damping decreases with an increase of the rotation speed for a
partial rub. Once the rotation speed reaches 0.65, which is the boundary of the partial rub and dry whip, the modal
damping is zero, meaning that mode motion of this partial rub at that time is critically stable. Therefore, the partial
rub herein is also critically stable. It is predictable what occurs if the rotation speed is further raised. The modal damp-
ing will be less than zero, the mode motion will be unstable, and the corresponding partial rub will also diverge. Con-
sequently, the dry whip motion appears. For a stable dry whip, we can see corresponding modal damping is located at
another end of the modal damping curve with a positive number, which is marked by the blue bar.

To validate the above analysis for the process of a partial rub converting into dry whip, a numerical simulation is per-
formed based on Newmark method for rotation speed accelerating slowly from 0.65 to 0.66, as shown in Fig. 14. It can
be seen that when the rotation speed is at 0.65, stable partial rubbing occurs. Once the rotation speed increases slowly to
0.66, then the response diverges, and the vibration amplitude is higher and higher until a stable dry whip motion with a pos-
itive modal damping is formed.
According to above results, we can summarize the mechanism for partial rub transmitting into dry whip motion as the
following: first, the partial rub becomes more and more severe with the rotation speed increasing, and the backward whirl
mode motion of the partial rub becomes excited at a certain rotation speed. Meanwhile, the modal damping of the backward
whirl mode motion of the rubbing motion constantly decreases. Until the modal damping is less than zero, the backward
mode motion and the corresponding partial rub is unstable. Consequently, the vibration amplitude increases until the modal
damping becomes positive again. At that time, the stable dry whip with a very high amplitude and vibration energy is
formed.

6. Conclusion

1) The nonlinear modes reveal that the rubbing rotor possesses two potential whirl directions, i.e., the forward whirl with
positive modal frequency and the backward whirl with negative modal frequency. Both the orbits of the forward whirl
and backward whirl mode motions are circular with same modal frequency at a certain modal amplitude. The mag-
nitudes of the modal frequency for these two mode motions increase with an increase of the modal amplitude and
mechanical energy, and the magnitudes are limited to an interval range. This interval can be approximately deter-
mined through the linear rotor system without a rub-impact and through the coupled linear rotor/stator system,
which has been validated by theoretical derivation and numerical simulation.
2) The modal damping of the forward whirl mode motion raises with an increase of the modal amplitude, and this for-
ward whirl mode motion is always stable as the corresponding modal damping is greater than zero for any modal
amplitude. While the modal damping of the backward whirl mode motion first decreases, it subsequently increased
to a positive number at a certain modal amplitude and further increases with an increase of the modal amplitude. The
backward whirl mode motion may be unstable, as its modal damping may be less than zero for a certain modal ampli-
tude range.
3) The backward whirl mode motion can be excited and become the dominant factor of the partial rub for a higher rota-
tion speed and dry whip. By analyzing the rubbing response characteristics and corresponding backward whirl modal
stability, it is found that the mechanism for a partial rub transmitting to dry whip is the instability of the backward
whirl mode motion for the corresponding partial rub.

Acknowledgment

The authors would like to acknowledge the financial support from the National Natural Science Foundation of China
(Grant Nos. 11772022, 51575022 and 51475021).

Appendix A. Calculation of the vibration response and modal characteristics for the rubbing rotor system under
another set of mechanical parameters

In Section 5, by comparing the vibration response and the modal characteristics under a corresponding mechanical
energy, we conclude that in the case of the partial rub, when the rotation speed reaches a certain value, the backward whirl
of harmonic component f2 is just the backward whirl mode motion of the rubbing rotor system. Meanwhile, the backward
whirl mode motion will dominate the vibration response of the rubbing rotor system. When the corresponding backward
whirl mode motion is unstable, the vibration amplitude of the rotor system increases continuously until the backward whirl
mode motion becomes stable. Next, a stable dry whip motion with an extremely high vibration energy is formed.
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 459

(a) 2
3 Axis orbit
Clearance circle

Displacement Y
funbalance

Amplitude
0 1

f2:0.1203

-3
0
-3 0 3 0.0 0.1 0.2 0.3

Displacement X Frequency

(b) 2
3 Axis orbit
Clearance circle
Displacement Y

funbalance

Amplitude
0 1 f2:0.1134

-3
0
-3 0 3 0.0 0.1 0.2 0.3
Displacement X Frequency

2
(c) 3 Axis orbit
Clearance circle f2:0.1108
Displacement Y

Amplitude

0 1 funbalance

-3
0
-3 0 3 0.0 0.1 0.2 0.3

Displacement X Frequency

(d) 20 15
Axis orbit
Clearance circle
f2:0.1557
Displacement Y

10
10
Amplitude

5
-10
funbalance
-20 0
-20 -10 0 10 20 0.0 0.1 0.2 0.3

Displacement X Frequency
Fig. A.1. Orbits and vibration responses in the frequency domain for a rotor at (a) X = 0.45, (b) X = 0.52, (c) X = 0.59, (d) X = 0.60.
460 J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461

Table A.1
Key characteristic parameters for the rotor’s response and corresponding nonlinear modes under different rotation speeds (mechanical parameters in
Appendix A).

Rotation speed X Response with an unbalance force Nonlinear modal characteristics


Mechanical energy interval Harmonic frequency f2 Modal frequency x/2p Modal damping b
0.45 [0.3830, 0.5350] 0.1203 [0.8010, 0.0901] [0.0177, 0.0131]
0.52 [0.6791, 1.0006] 0.1134 [0.0953, 0.1040] [0.0098, 0.0056]
0.59 [1.1655, 1.7918] 0.1108 [0.1069, 0.1153] [0.0045, 0]
0.60 [141.71, 157.09] 0.1557 [0.1554, 0.1560] [0.0201, 0.1205]

(b) 0.20
Energy interval for =0.45
Energy interval for =0.52
0.15 Energy interval for =0.59
Energy interval for =0.60

Modal damping
0.10

Modal damping
0.05 decreasing

0.00

-0.05
0.01 0.1 1 10 100 1000
Mechanical energy E
Fig. A.2. (a) Modal angular frequency, (b) modal damping for the rotor’s rubbing motions under different rotation speeds.

In Section 5, only one set of simulation results are given to demonstrate the above conclusion. To validate the universality
of the results obtained in Section 5, the calculation of the vibration response and modal characteristics is performed again
with another set of mechanical parameters in this appendix. In the calculation, the value of the rotor/stator stiffness ratio is
changed to 0.04, while other parameters remain unchanged.
The vibration responses under different rotation speeds are presented in Fig. A.1. The results show that the super-
harmonic frequency f2 appears in the frequency content in the case of a partial rub and dry whip. In addition, the amplitude
of this harmonic component increases significantly with an increase of the rotation speed. Based on the expression of the
mechanical energy of the rubbing rotor system, which is shown in Eq. (26), the mechanical energies of the rotor/stator con-
tact system under different rotation speeds are obtained, and the results are listed in Table A.1. Meanwhile, the values of the
harmonic frequencies f2 under different rotation speeds are also extracted in Table A.1.
The modal frequency and modal damping with the variation of the mechanical energy under different rotation speeds are
calculated, and the results are shown in Fig. A.2. According to the mechanical energy of the rubbing motions under different
rotation speeds, the modal frequency and modal damping under corresponding rubbing motions are extracted into Table A.1.
One can observe that when X = 0.59, the harmonic frequency f2 is basically equal to the modal frequency of the backward
whirl mode motion as well. Meanwhile, the modal damping of the backward mode motion is nearly equal to zero. When the
rotation speed is disturbed and increases to 0.60, the modal damping of the backward whirl mode motion is less than zero,
and the backward whirl mode motion becomes unstable. In that case, the vibration amplitude of the rotor system increases
continuously until the backward whirl mode motion becomes stable, as shown in Fig. A.2(b). Next, a stable dry whip motion
with an extremely high vibrational energy is formed. The harmonic frequency f2 in the case of X = 0.60 is also equal to the
modal frequency of the backward whirl mode motion as well.
According to the above analysis, it can be seen that the simulation results under the new set of mechanical parameters
also support the conclusions obtained in Section 5, and the similar physical process can be obtained as well.

References

[1] S.K. Sinha, Non-linear dynamic response of a rotating radial Timoshenko beam with periodic pulse loading at the free-end, Int. J. Non Linear Mech. 40
(1) (2005) 113–149.
[2] T.H. Patel, M.J. Zuo, X. Zhao, Nonlinear lateral-torsional coupled motion of a rotor contacting a viscoelastically suspended stator, Nonlinear Dyn. 69 (1)
(2012) 325–339.
[3] A. Muszynska, Rotordynamics, Taylor & Francis, New York, 2005.
[4] F. Chu, Z. Zhang, Periodic, quasi-periodic and chaotic vibrations of a rub-impact rotor system supported on oil film bearings, Int. J. Eng. Sci. 35 (10)
(1997) 963–973.
J. Hong et al. / Mechanical Systems and Signal Processing 116 (2019) 443–461 461

[5] Jiang Jun, Chen Yanhua, Advances in the research on nonlinear phenomena in rotor/stator rubbing systems, Adv. Mech. 1 (2013) 132–148 (In Chinese).
[6] Wang Weimin, Li Qihang, Gao Jinji, Yao Jianfei, Allaie Paul, An identification method for damping ratio in rotor systems, Mech. Syst. Sig. Process. 68–69
(2016) 536–554.
[7] Z. Qin, F. Chu, Z.U. Jean, Free vibrations of cylindrical shells with arbitrary boundary conditions: a comparison study, Int. J. Mech. Sci. 133 (2017) 91–99.
[8] F. Chu, Z. Zhang, Bifurcation and chaos in a rub-impact Jeffcott rotor system, J. Sound Vib. 210 (1) (1998) 1–18.
[9] P. Pennacchi, N. Bachschmid, E. Tanzi, Light and short arc rubs in rotating machines: experimental tests and modelling, Mech. Syst. Sig. Process. 23 (7)
(2009) 2205–2227.
[10] S. Roques, M. Legrand, P. Cartraud, et al, Modeling of a rotor speed transient response with radial rubbing, J. Sound Vib. 329 (5) (2010) 527–546.
[11] J. Jiang, Determination of the global responses characteristics of a piecewise smooth dynamical system with contact, Nonlinear Dyn. 57 (3) (2009) 351–
361.
[12] Hui Ma, Wu Zhiyuan, Xingyu Tai, Bangchun Wen, Dynamic characteristics analysis of a rotor system with two types of limiters, Int. J. Mech. Sci. 88
(2014) 192–201.
[13] Hui Ma, Qianbin Zhao, Xueyan Zhao, Qingkai Han, Bangchun Wen, Dynamic characteristics analysis of a rotor-stator system under different rubbing
forms, Appl. Math. Model. 39 (8) (2015) 2392–2408.
[14] H. Ma, Y. Lu, Z. Wu, et al, Vibration response analysis of a rotational shaft–disk–blade system with blade-tip rubbing, Int. J. Mech. Sci. 107 (2016) 110–
125.
[15] H. Ma, F. Yin, Z. Wu, et al, Nonlinear vibration response analysis of a rotor-blade system with blade-tip rubbing, Nonlinear Dyn. 84 (3) (2016) 1225–
1258.
[16] Qi Sun, Hui Ma, Yunpeng Zhu, Qingkai Han, Bangchun Wen, Comparison of rubbing induced vibration responses using varying-thickness-twisted shell
and solid-element blade models, Mech. Syst. Sig. Process. 108 (2018) 1–20.
[17] Hui Ma, Di Wang, Xingyu Tai, Bangchun Wen, Vibration response analysis of blade-disk dovetail structure under blade tip rubbing condition, J. Vib.
Control 23 (2) (2017) 252–271.
[18] J. Jiang, H. Ulbrich, The physical reason and the analytical condition for the onset of dry whip in rotor-to-stator contact systems, J. Vibr. Acoust. 127 (6)
(2005) 594–603.
[19] P. Yu, D. Zhang, Y. Ma, et al, Dynamic modeling and vibration characteristics analysis of the aero-engine dual-rotor system with Fan blade out, Mech.
Syst. Sig. Process. 106 (2018) 158–175.
[20] H.F. Black, Interaction of a whirling rotor with a vibrating stator across a clearance annulus, J. Mech. Eng. Sci. 10 (1) (1968) 1–12.
[21] W. Zhang, Dynamic instability of multi-degree-of-freedom flexible rotor systems due to full annular rub, IMechE C252 (88) (1988) 305–308.
[22] D.W. Childs, D. Kumar, Dry-friction whip and whirl predictions for a rotor-stator model with rubbing contact at two locations, J. Eng. Gas Turbines
Power 134 (7) (2012) 072502.
[23] F. Chu, W. Lu, Determination of the rubbing location in a multi-disk rotor system by means of dynamic stiffness identification, J. Sound Vibr. 248 (2)
(2001) 235–246.
[24] Y. Ma, C. Cao, D. Zhang, Z. Liang, J. Hong, Constraint mechanical model and investigation for rub-impact in aero-engine system, ASME Turbo Expo 2015:
Turbine Technical Conference and Exposition, American Society of Mechanical Engineers, 2015.
[25] Jie Hong, Yu. Pingchao, Dayi Zhang, et al, Modal characteristics analysis for a flexible rotor with non-smooth constraint due to intermittent rub-impact,
Chin. J. Aeronaut. 31 (3) (2018) 498–513.
[26] F. Chu, W. Lu, Stiffening effect of the rotor during the rotor-to-stator rub in a rotating machine, J. Sound Vib. 308 (3) (2007) 758–766.
[27] Dayi Zhang, Ying Xia, Fabrizio Scarpa, Jie Hong, et al, Interfacial contact stiffness of fractal rough surfaces, Sci. Rep. 7 (2017) 12874.
[28] D. Bently, Forced subrotative speed dynamic action on rotating machines, ASME Paper 74-PET-16; Petroleum Mechanical Engineering Conference,
Dallas, TX, 15–18 September 1974.
[29] D.W. Childs, Rub-induced parametric excitation in rotors, J. Mech. Des. 101 (4) (1979) 640–644.
[30] R.M. Rosenberg, Normal Modes of Nonlinear Dual-Modes Systems, Institute of Engineering Research, University of California, 1959.
[31] S.W. Shaw, C. Pierre, Non-linear normal modes and invariant manifolds, J. Sound Vibr. 150 (1) (1991) 170–173.
[32] H.J. Greenberg, T.L. Yang, Modal subspaces and normal mode vibrations, Int. J. Non Linear Mech. 6 (3) (1971) 311–326.
[33] A.H. Nayfeh, On direct methods for constructing nonlinear normal modes of continuous systems, Modal Anal. 1 (4) (1995) 389–430.
[34] J.C. Slater, A numerical method for determining nonlinear normal modes, Nonlinear Dyn. 10 (1) (1996) 19–30.
[35] Xiao-Dong Yang, Ji-Hou Yang, Ying-Jing Qian, Wei Zhang, Roderick Melnik, Dynamics of a beam with both axial moving and spinning motion: an
example of bi-gyroscopic continua, Eur. J. Mech. A/Solid 69 (2018) 231–237.
[36] Xiao-Dong Yang, Wu Hang, Ying-Jing Qian, Wei Zhang, C.W. Lim, Nonlinear vibration analysis of axially moving string based on gyroscopic modes
decoupling, J. Sound Vib. 393 (2017) 308–320.
[37] D. Jiang, C. Pierre, S.W. Shaw, Large-amplitude non-linear normal modes of piecewise linear systems, J. Sound Vibr. 272 (3) (2004) 869–891.
[38] Y. Chen, J. Jiang, Determination of nonlinear normal modes of a planar nonlinear system with a constraint condition, J. Sound Vib. 332 (20) (2013)
5151–5161.
[39] D. Laxalde, F. Thouverez, Complex non-linear modal analysis for mechanical systems: application to turbomachinery bladings with friction interfaces,
J. Sound Vibr. 322 (4) (2009) 1009–1025.
[40] J.J. Sinou, Non-linear dynamics and contacts of an unbalanced flexible rotor supported on ball bearings, Mech. Mach. Theory 44 (9) (2009) 1713–1732.
[41] Hou Lei, Chen Yushu, Fu Yiqiang, Chen Huizheng, Lu Zhenyong, Liu Zhansheng, Application of the HB-AFT method to the primary resonance analysis of
a dual-rotor system, Nonlinear Dyn. 88 (4) (2017) 2531–2551.
[42] D.E. Bently, G. Paul, J.J. Yu, Full annular RUB in mechanical seals, Part II: analytical study, Int. J. Rotating Mach. 8 (5) (2002) 329–336.
[43] L. Peletan, S. Baguet, M. Torkhani, et al, Quasi-periodic harmonic balance method for rubbing self-induced vibrations in rotor–stator dynamics,
Nonlinear Dyn. 78 (4) (2014) 2501–2515.

You might also like