Sie sind auf Seite 1von 13

Mechanism and Machine Theory 48 (2012) 81–93

Contents lists available at SciVerse ScienceDirect

Mechanism and Machine Theory


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

Modal analysis of rotor-shaft system under the influence of rotor-shaft


material damping and fluid film forces
M. Chouksey, J.K. Dutt ⁎, S.V. Modak
Department of Mechanical Engineering, IIT, Delhi, Hauz-Khas, New Delhi, 110016, India

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

Article history: The present work attempts to study the influences of internal rotor material damping and the
Received 29 September 2010 fluid film forces (generated as a result of hydrodynamic action in journal bearings) on the
Received in revised form 26 August 2011 modal behaviour of a flexible rotor-shaft system. This is relevant as both journal bearing and
Accepted 8 September 2011
the internal material damping introduce tangential forces increasing with the rotor spin
Available online 11 October 2011
speed. Such forces considerably influence the dynamic behaviour of a rotor and tend to destabilize
the rotor-shaft system as spin speed increases. Under this system of forces the modal behaviour of
Keywords: the rotor-shaft is studied to get better ideas about the dynamic behaviour of the system, estimated
Rotor modal analysis
in terms of modal damping factors, stability limit speed, the frequency response functions, as well
Journal bearings
as the direction of whirl of the shaft in different modes. It is seen that correct estimation of internal
Internal rotor material damping
Directional frequency response functions friction, in general, and the journal bearing coefficients at the rotor spin-speed are essential to
accurately predict the rotor dynamic behaviour. This serves as a first step to get an idea about
dynamic rotor stress and, as a result, a dynamic design of rotors.
© 2011 Elsevier Ltd. All rights reserved.

1. Introduction

Both anti-friction (ball/roller) and journal bearings are used to support rotor-shaft systems; the former finds good use in aircraft
engines, whereas the latter are mostly used to support heavy rotor-shafts of power plant machinery (Rao [1], Vance [2], Lalanne and
Ferraris [3]). This work is aimed primarily to study the dynamic behaviour of fixed-base heavy rotors used in most of the power plant
industries. Modal analysis is a useful tool to get an idea about the dynamic behaviour of a system and hence is considered to be
important for dynamic design. Irretier [4] has given a detailed mathematical basis for modal analysis of any rotor shaft system
considering it first as a Linear Time Independent (LTI) and later a Linear Time Varying (LTV) system. Modal analysis of rotors differs
from modal analysis of structures due to additional forces induced by rotor spin like gyroscopic, tangential and rotating damping
forces (e.g. by Ewins [5]). These forces change the nature of system matrices and make them asymmetric and speed dependent.
Therefore all modal characteristics of rotors are closely related to rotor spin speed. Rotor-shaft system supported on journal bearings
have been chosen for the study, as, coefficients of these bearings play important role in deciding dynamics of the machine. This is due
to the fact that the coefficients of these bearings are functions of parameters like clearance, oil viscosity, and spin speed of the
rotor-shaft and, as a result, fluid film forces acting on the journal are both conservative and dissipative in nature; such forces
are spin speed dependent as well. Rao [1] and Friswell, et al. [6], among others, have given theoretical derivation to obtain the
bearing model in terms of four stiffness (two direct and two cross coupled) and four damping (two direct and two cross coupled)
coefficients. Though the restoring and dissipative forces due to the hydrodynamic action are actually non linear functions of
displacement and velocity of the journal (Friswell, Penny, Garvey and Lees [6], Cheng, et al. [7]) with respect to the bearing
housing, yet, in this work, linear restoring as well as dissipative forcing functions in terms of displacement and velocity of
the journal with respect to the bearing housing are considered for the sake of simplicity as well as good approximations

⁎ Corresponding author. Tel.: + 91 11 26596334; fax: + 91 11 26582053.


E-mail addresses: manoj_chouksey@yahoo.com (M. Chouksey), jkrdutt@yahoo.co.in (J.K. Dutt), svmodak@mech.iitd.ac.in (S.V. Modak).

0094-114X/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechmachtheory.2011.09.001
82 M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93

(Holmes [8]). Many researchers, e.g. Zhang, et al. [9], Tieu and Qiu [10], Qiu and Tieu [11], Tiwari and Chakravarthy [12]
have suggested methods to determine these co-efficient through experimentally obtained data. Gutierrez-wing [13] in his
doctoral work reported the modal characterisation of rotating machinery using a structural dynamics approach, where the
aspects of modal analysis of a rotor-shaft system were reported; excitations methods and measurement of rotor modes
using few FRFs were also discussed. However the joint influence of material internal damping and the fluid film damping
on mode shapes as well as on FRFs were not reported. This work attempts all these, however, through analytical as well
as numerical simulations.
“Oil whirl” and “Oil whip” are two well known phenomena, which are observed in rotor bearing system supported on journal
bearings. Muszynska [14] has defined and described the two phenomena in detail. Also a good source of information about this
subject is obtained from Stepan, et al. [15]. Oil whirl is a stable sub-synchronous (≈half of the spin speed) whirl motion caused
by cross coupled stiffness and transforms into a sub-synchronous and unstable whirl motion called “Oil whip” at a spin speed
about twice the first critical of the rotor-shaft, when one of the cross coupled stiffnesses is negative (Muszynska [16], Rao
[17]). Presence of negatively cross-coupled terms in the stiffness matrix may be physically interpreted as the presence of a
force in the direction of whirl and tangential to the rotor whirl orbit; strength of the tangential force is proportional to the rotor
spin speed in the case of fluid film bearings. Such forces are always destabilizing in nature and cause instability when the strength
is sufficient to overcome the effect of stationary dissipative forces (Crandall [18]). EI-Shafei, et al. [19] studied, experimentally, the
effect of the parameters that influence the onset of oil whirl and oil whip in journal bearings supporting flexible rotors. The authors
focussed on the effects of unbalance, supply pressure and misalignment on the onset of instability.
In this work the effect of another tangential force, generated by rotor material damping has been included in addition to the
fluid film forces to study the dynamic behaviour of a rotor-shaft system. Viscous form of internal material damping has been
assumed in this work for simplicity after following Tondl [20], Zorzi and Nelson [21] of many others. Hysteretic model to represent
internal friction (Tondl [20], Genta [22]) or very recently viscoelastic model used by Dutt and Roy [23] for representing the internal
damping are in literature and will be considered later. Very few reports on modal analysis of rotors supported on fluid film bearings
are found in literature. Jei and Kim [24] studied the dynamics of rigid rotor supported on oil film bearings, where, obviously, flexible
modes were not reported and the effect of internal damping forces was not considered.
Complex modal analysis of rotating machinery has advantages over classical modal analysis using real coordinates, as, frequency
response functions defined in complex co-ordinates help in finding out modal directivity (direction of whirl of a mode) which is lost if
real co-ordinates are used. Papers by Lee [25], Jei and Kim [24], Joh and Lee [26] show the use of complex modal analysis to study
modal and frequency response behaviour for axi-symmetric rotors put on anisotropic supports as well as non-axi-symmetric rotors
put on isotropic supports. The effect of rotor-shaft material damping on mode shapes and directional frequency response functions for
rotor-shaft systems with constant (spin-independent) bearing coefficients has been investigated by Chouksey, et al. [27].
Thus, this work attempts to present the modal analysis of a flexible internally damped rotor-shaft supported on fluid-film
bearings at the ends to study the stability of different modes, to have a better idea about the frequency response functions of
the rotor-shaft system subjected to external excitations and also to predict the direction of whirl of various modes. For this,
the equations of motion are obtained to a good degree of accuracy by discretizing the rotor-shaft continuum using 2-noded finite
Rayleigh beam elements. These equations are used for analysis of eigenvalues, mode shapes and frequency response.

2. Modal analysis of rotors

Unlike non rotating structures, rotors have two different kinds of modes, known as forward and backward modes due to the rotor
spin. Modal characteristics associated with forward and backward modes are different from the conventional modal characteristics of
non rotating structures (Lee and Lee [28]). One of the characteristics of rotating machinery is that they generally do not abide by the
principle of reciprocity and are referred to as ‘Non Self Adjoint Systems’ (NSA). This property results in two sets of eigenvectors,
referred to as the ‘left hand’ and ‘right hand’ eigenvectors. The right hand eigenvectors are those generally associated with the
mode shapes, while the left hand eigenvectors are the set, which are obtained by analyzing the transformed equations and indicate
the pattern of forces associated with a single mode (Ewins [5]).

2.1. Finite element model

In this paper, rotor-shaft has been modelled using two noded Rayleigh beam elements with 4 degrees of freedom per node as
shown in Fig. 1, depicting the ith element:
The equations of motion have been formulated by taking into account the effects of rotary inertia, translatory inertia, gyroscopic
moments, support flexibility, stationary and rotational damping forces in the model. Discretizing the rotor shaft into ‘N’ beam
elements, total number of nodes becomes ‘N + 1’. Nodal displacement vector of ith element is given by

n oT
qe ¼ yi ; θzi ; yðiþ1Þ ; θzðiþ1Þ ; zi ; θyi ; zðiþ1Þ ; θyðiþ1Þ ð1Þ

where, the superscript ‘T ’ denotes transpose of a matrix/vector. Matrices and vectors are written in bold letters in this paper and
are mentioned wherever they appear.
M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93 83

Fig. 1. The ith element and nodal degrees of freedom.

2.2. Equations of motion

After assembling the governing equations for all the elements and incorporating the boundary conditions, the equations of
motion of the rotor bearing system are written as:
::ðtÞ þ Cq_ ðtÞ þ KqðtÞ ¼ f ðtÞ
Mq ð2Þ
h i h i
where; M ¼ Mtrs þ Mrot ; C ¼ −ΩG þ ηv Ksh þ Cbrg and K ¼ Kbrg þ ηv ΩHcir þ Ksh ð3Þ

In the above, the ‘M’ is mass matrix and includes mass components due to translation (Mtrs) and rotary inertia (Mrot), C is the
non-symmetric matrix coefficient of the velocity vector and includes skew symmetric gyroscopic matrix (G), dissipation terms
(ηvKsh) relating to shaft internal material damping and the damping matrix due to journal bearing (Cbrg). The symbol ‘K’ is the
matrix coefficient of the displacement vector and is the summation of skew symmetric circulatory matrix (Hcir), stiffness matrices
due to journal bearing (Kbrg) and shaft bending (Ksh) respectively. All the matrices are of size 4(N + 1) × 4(N + 1). The symbol ‘Ω’
stands for rotor spin speed in radian /second and ‘ηv’ is coefficient of viscous form of material damping. Time and time derivative
are represented by ‘t’ and ‘˙’ respectively.
Global displacement and force vectors, q(t) and f(t) respectively, are expressed as below:
8 T 9
  < y1 ðt Þ θz1 ðt Þ y2 ðt Þ θz2 ðt Þ … … yðNþ1Þ ðt Þ θzðNþ1Þ ðt Þ =
yðt Þ2ðNþ1Þ×1
qðt Þ ¼ ¼ h iT ð4Þ
zðt Þ2ðNþ1Þ×1 4ðNþ1Þ×1
: z1 ðt Þ θy1 ðt Þ z2 ðt Þ θy2 ðt Þ … … zðNþ1Þ ðt Þ θyðNþ1Þ ðt Þ ;

8 T 9
  < fy1 ðt Þ Mz1 ðt Þ fy2 ðt Þ Mz2 ðt Þ … … fy ðNþ1Þ ðt Þ MzðNþ1Þ ðt Þ =
f y ðt Þ2ðNþ1Þ×1
f ðt Þ ¼ ¼ h iT ð5Þ
f z ðt Þ2ðNþ1Þ×1 4ðNþ1Þ×1
: fz1 ðt Þ My1 ðt Þ fz2 ðt Þ My2 ðt Þ … … fz ðNþ1Þ ðt Þ MyðNþ1Þ ðt Þ ;

Journal bearing coefficients to form global stiffness and global damping matrix Kbrg and Cbrg are considered after following Friswell,
Penny, Garvey and Lees [6] and Rao [1], by assuming short bearing assumption (L/Dbb 1, L and D being length and diameter of the
bearing respectively) which enables to give closed form solutions for the bearing coefficients. Other assumptions are:
• Viscosity is constant throughout the oil film.
• The oil is Newtonian, incompressible and the flow is laminar so that Reynolds equation applies.
• The lubricant pressure is zero at the edges of the bearing.
• The radial clearance or the fluid film thickness is so small that the pressured gradient across the clearance is neglected.
• There is perfect adhesion between the lubricant and the bearing surfaces.
The equations of motion (Eq. (2)) may be written in the state space form as:

Awðt Þ ¼ Bwðt Þ þ Fðt Þ ð6Þ

where matrices A, B, the state vector w(t), and the force vector F(t) are given as:
       
0 M M 0 q̇ ðt Þ 0
A¼ ;B¼ ; wðt Þ ¼ ; Fðt Þ ¼ ð7Þ
M C 8ðNþ1Þ × 8ðNþ1Þ
0 −K 8ðNþ1Þ × 8ðNþ 1Þ
qðt Þ 8ðNþ1Þ× 1
f ðt Þ 8ðNþ1Þ × 1

The symbol ‘0’ represents null matrix of size 4(N + 1) × 4(N + 1) and null vector of size 4(N + 1) depending upon where they
appear in the Eq. (7).
84 M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93

2.3. Free vibration analysis

Free vibration analysis of the system of Eq. (2) gives eigenvalues λr, right eigenvector ur and left eigenvector vr which are
connected by equations given in Eq. (8) for the rth mode as
 
λ2r M þ λr C þ K ur ¼ 0 and λ2r MT þ λr CT þ KT vr ¼ 0; where r ¼ 1 to 4ðN þ 1Þ ð8Þ

The rth eigenvalue problem associated with Eq. (6) may be written after following Lee [25] as:
T T
λr Aψr ¼ Bψr and λr A φr ¼ B φr ; where r ¼ 1 to 8ðNþ1Þ ð9Þ

where λr is the rth eigenvalue, and ψr and φr stand for the rth right hand and rth left hand eigenvectors, respectively, in state
space and are given by –
   
λr ur λr vr
ψr ¼ ; φr ¼ ð10Þ
ur vr

ψr and φr may be biorthonormalized so as to satisfy

T T
φi Aψr ¼ δir ; φi Bψr ¼ λr δir ; ð11Þ

The indices ‘i’, ‘r’ vary from 1 to 8(N + 1), and δir is the kronecker delta (i.e.= 1when i = r and = 0 when i ≠ r) For r = 1 to
8(N + 1), ψr,φr and λr may be written in a compact form as Ψ , Φ and Λ respectively and are given by:
h i h i h i
Ψ ¼ ψ1 ψ2 …ψ8ðNþ1Þ ; Φ ¼ φ1 φ2 …φ8ðNþ1Þ ; Λ ¼ diag λ1 λ2 …λ8ðNþ1Þ ð12Þ

Therefore bi-orthogonality condition of Eq. (11) may be written in compact form as

ΦT AΨ ¼ I; ΦT BΨ ¼ Λ ð13Þ

In Eq. (13) the symbol ‘I’ represents identity matrix of size 8(N + 1) × 8(N + 1).

2.4. Frequency response function matrix

Assuming a solution of w(t), the state vector in Eq.(6), as a linear combination of the right hand eigenvectors ψr multiplied by
modal co-ordinates ξr(t), as given in Eq. (14):

8ðNþ1Þ
wðt Þ ¼ ∑ ψr ξr ðt Þ ¼ Ψξðt Þ ð14Þ
r¼1

Substituting it into Eq. (6), premultiplying throughout by ΦT and incorporating the biorthogonality conditions, the Eq. (15) is
obtained.

T T T
Φ AΨξ̇ ðt Þ ¼ Φ BΨξðt Þ þ Φ Fðt Þ ð15Þ

Using orthonormality relations of Eq. (13), the Eq. (15) may be transformed, as shown below, into Eq. (16):

ξ̇ ðt Þ ¼ Λξðt Þ þ nðt Þ ð16Þ

Where nðt Þ ¼ ΦT Fðt Þ, represents the modal excitation vector.


Eq. (16) represents a set of independent modal equations, and may be written as in Eq. (17):

ξ̇ r ðt Þ ¼ λr ξr ðt Þ þ nr ðt Þ; r ¼ 1 to 8ðN þ 1Þ ð17Þ

Where nr ðt Þ ¼ φTr Fðt Þ, is the generalized time varying forcing function exciting the rth mode.
Steady state response, ξr(t) of Eq. (17) under harmonic force Fðt Þof frequency ‘ω’ is given by:

φTr Fðt Þ
ξr ðt Þ ¼ ð18Þ
jω−λr
pffiffiffiffiffiffiffiffi
where ‘j’ is equal to −1.
M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93 85

Substituting Eq. (18) into Eq. (14) gives

8ðNþ1Þ
ψr φTr
wðt Þ ¼ ∑ Fðt Þ ð19Þ
r¼1 jω−λr

Further, Frequency Response Function matrix in state space may be written as:

8ðNþ1Þ T
4ðNþ1Þ T T
ψr φr ψr φr ψ̄ φ̄
H¼ ∑ ¼ ∑ þ r r ð20Þ
r¼1 jω−λr r¼1 jω−λr jω−λ̄r

where overbar ‘ □ ’ represents complex conjugate of a vector/scalar.


Eq. (20) can be expanded as:
  
λr ur λr vr T
8ðNþ1Þ
ψr φTr 8ðNþ1Þ ur vr
H¼ ∑ ¼ ∑ ð21Þ
r¼1 jω−λr r¼1 jω−λr
" #
λ2r ur vTr λr ur vTr
8ðNþ1Þ λr ur vTr ur vTr
H¼ ∑ ð22Þ
r¼1 jω−λr

Therefore, Frequency Response Function matrix (HÞrelating generalised displacement and forces may be written as:

8ðNþ1Þ T
4ðNþ1Þ T T
ur vr ur vr ū v̄
H¼ ∑ ¼ ∑ þ r r ð23Þ
r¼1 jω−λr r¼1 jω−λr jω−λ̄r

The FRF matrix may be expanded as:


 
Hyy Hyz
H¼ ð24Þ
Hzy Hzz

Examining Eq. (23), as well as following (Lee [25], Mesquita, et al. [29]), it may be seen that both right and left eigenvectors
and the natural frequencies may be found out by estimating any one row and any one column of the FRF matrix H, for a generally
non-self-adjoint system. Thus it is imperative to estimate any complete row and any complete column of the FRF matrix for the
estimation of the complete modal model of a generally non-self-adjoint system. Under some special cases, e.g. for isotropic rotor
bearing systems, there are some special relationship between the right and the left-hand eigenvectors. These relationships simplify
the experimental procedure and thus require fewer measurements. In these cases, even exciting the structure at a few points or
even a single point can yield a complete model (Lee [25], Bucher and Ewins [30]).

2.5. Complex frequency response functions

Due to conjugate even properties of classical frequency response function the information about directivity of modes is lost
and the frequency response function in negative frequency region comes similar to positive frequency region. On the other
hand, complex frequency response functions given by Lee [25] where complex co-ordinates are used for displacement and
force has the advantage of preserving the information about directivity of modes. For a detailed study of complex frequency
response functions the papers by Lee [25], Jei and Kim [24], Kessler and Kim [31], Mesquita, Dias and Miranda [29] can be referred to.
Though the complex frequency response functions have advantages over classical frequency response functions, but applying
rotating force and measuring rotating displacement is not simple. For rotating shafts, Muszynska [16] developed an excitation
technique by mounting an unbalanced disc on a bearing which, in turn, is mounted onto the shaft, such that the unbalanced
disc is able to rotate independently of the shaft.
Considering the difficulty of generating rotating force (g) and measuring the complex displacements (p), an easier way of finding
out the complex FRFs is given by Lee [25], where the classical/conventional FRFs may be used to obtain the complex FRFs as given in
Eq. (25):

g

2Hpg ¼ Hyy þ Hzz −j Hyz −Hzy

2Hpĝ ¼ Hyy −Hzz þ j Hyz þ Hzy
 ð25Þ
2Hp̂g ¼ Hyy −Hzz −j Hyz þ Hzy

2Hp̂ ĝ ¼ Hyy þ Hzz þ j Hyz −Hzy
86 M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93

Here Hpg ; Hp̂ ĝ are directional Frequency Response Function (dFRF) matrix and Hpĝ ; Hp̂g are reverse directional Frequency
Response Function (r-dFRF) matrix (Lee [25], Joh and Lee [26]). Where p and g represent the complex linear displacement
and complex force at any particular node and are given by equation:

pðt Þ ¼ yðt Þ þ jzðt Þ; g ðt Þ ¼ fy ðt Þ þ jfz ðt Þ ð26Þ

And p̂ and ĝ correspond to the complex conjugate of p and g respectively


and represent the complex displacement and force
rotating in the opposite direction to that of shaft spin. Therefore Hpg Hp̂ ĝ is defined as the frequency response of the rotor in the
forward
(backward)
direction to rotating excitation in the forward (backward) direction, and is termed as normal dFRF, whereas
Hpĝ Hp̂g represents the frequency response of the rotor in the forward (backward) direction to rotating excitation in the backward
(forward) direction and is termed as the reverse dFRF (Kessler and Kim [31]).
After following Lee [25] it can be seen that:
— —
Hp̂g ð jωÞ ¼ Hpĝ ð−jωÞ; and Hpg ð jωÞ ¼ Hp̂ ĝ ð−jωÞ ð27Þ

The complex frequency response functions (dFRF and r-dFRF) also help in detecting anisotropy/asymmetry (Joh and Lee [26])
and crack in rotating machinery (Lee [32]). Ginsberg and Wagner [33] developed algorithm for extracting modal parameters from
directional FRFs and named it two sided Algorithm of Mode isolation as it processes data spanning in positive and negative
frequency region.
In the illustrative example of next section, the Eq. (9) giving eigenvalues and eigenvectors and Eq. (25) giving dFRFs are used
to present the results for a flexible rotor-shaft system under influences of rotor-shaft material damping and fluid film forces
simultaneously.

3. Illustrative example

The rotor-shaft system (as shown in Fig. 2), which is a slightly modified version of the one given by Friswell, Penny, Garvey
and Lees [6] as an example, is considered for the purpose of simulation as well as illustration. The two-disc steel-rotor-shaft,
where the density of the shaft and discs is 7810 kg/m 3, is supported on journal bearings at its ends. Viscosity of oil and radial
clearance between journal and the bearing are taken as 0.1 Pa-s and 0.0001 m respectively. Each bearing supports a static load
of 648 N and has length to diameter ratio of 0.3. The viscous damping coefficient of rotor-shaft material is considered as
0.0002 s, after following Zorzi and Nelson [21].
For the sake of comparison, two cases are considered, 1) case-1, where rotor-shaft material damping is not considered, and
2) case-2, where rotor-shaft material damping (ηv = 0.0002 s) has been considered. In the following discussions, rth forward
and rth backward modes have been represented as ‘rF’ and ‘rB’ modes respectively.

3.1. Effect of journal bearing on modal frequencies and damping factors

Fig. 3 shows the Campbell diagram (or the whirl speed map) of the rotor-shaft system, for case-1 i.e. when the shaft material
damping is not considered; whirl frequencies (obtained from the imaginary part of the eigenvalues) along with the directions of
whirl corresponding to only the first four modes are plotted in the figure to avoid congestion as well as to illustrate the basic
characteristics. There are 3 forward-whirling modes, marked sequentially in the ascending order of frequency, by ‘1F, 2F, 3F’ in
which the rotor whirls in the direction of the spin and there is one backward whirling mode ‘1B’, in which the rotor whirls
opposite to the direction of spin. The Synchronous Whirl Line is marked as ‘SWL’. The speed corresponding to the first
point of intersection of ‘SWL’ with the Campbell diagram is the critical speed shown as ‘Ωcr’ in Fig. 3; for this example
Ωcr = 1066 rpm. The critical speed ‘Ωcr’ thus divides the Campbell diagram into two regions, the sub-critical region, which falls

Z
Y
Steel shaft,Youngs modulus(E) = 211GPa
kyy kyy
Cyy
Cyy 0.25 m 0.25 m 0.25 m 0.25 m 0.25 m 0.25 m
1 X 2 4 6 7
3 5
kyz
φ 0.1 m φ 0.05 m φ 0.1 m Czy
Czz Czz
Czy
kzz φ 0.31 m kyz kzz

Cyz
kzy φ 0.31 m Cyz
kzy
0.07 m 0.07 m

Fig. 2. Rotor-bearing system supported on journal bearing at the ends.


M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93 87

2000
Before curve veering SWL
1800
1F

Damped Natural freq. (rpm)


1600 2F
1B curve veering
1400 region
3F
1200

1000

800 After curve veering

600 1B
1F
400
2F
200 3F

cr
0
0 500 1000 1500 2000 2500 3000 3500
rotor spin speed in rpm ( )

Fig. 3. Campbell diagram showing whirl-frequency vs. rotor-spin-speed for first four modes.

on the left hand side of ‘Ωcr’ and the post-critical region falling on the right hand side of ‘Ωcr’. On the sub-critical region there is a pair
of eigen-frequencies each above and below the SWL; the pair lying below increases with spin-speed almost proportionally
but at a rate lower than the SWL. These frequencies may be identified to be due to the journal bearings, as, firstly, the film-
stiffness increases with speed and secondly, these frequencies do not appear if the rotor-supports are considered rigid. In both
these modes the rotor-shaft whirls in the forward direction. The other pair of eigen-frequencies lying above the SWL may be
identified to be due to the first bending mode of the rotor-shaft structure. On the post-critical speed region the curve veering
phenomenon of the eigen-frequencies (Jei and Lee [34], is noticed and is marked with a circle. Jei and Kim [24] reported this
phenomenon for rigid rotor-shaft system supported on ball bearings.
The veering phenomenon occurs at a speed of 2000 rpm approximately. During this phenomenon eigen-value plots (in this
plot the eigen-frequencies) do not cross but swap their trends, or veer rapidly, but maintaining continuity, as the speed increases
(Jei and Lee [34]). As a result, the eigen-frequencies corresponding to ‘1B’ and ‘3F’ modes before veering follow the trends of those
corresponding to the modes ‘1F’ and ‘2F’ and vice-versa after veering. So following ascending order of frequencies for a direction
of whirl, the modes 1F, 2F, 1B and 3F before veering become 1B, 1F, 2F and 3F respectively after veering.
Fig. 4 shows the plot of modal damping factors for the four modes as the spin-speed varies. Positive modal damping factor
indicates stability as vibratory energy is dissipated and negative modal damping signifies instability as rotary energy supports
rotor-whirl by adding energy. The veering phenomenon causing an exchange of trends is also seen in the plot. From the diagram, it
is seen that the rotor-shaft system becomes unstable in the 1F mode, after curve veering region, at a spin speed of 2052 rpm, which
is little less than two times the first critical speed. The spin-speed limit, (up to 2052 rpm in this example) till which none of the
modal damping factors is negative is defined by Stability Limit Speed (SLS) of the rotor-shaft system, as shown in Fig. 4.
Fig. 5(a) and (b) shows the details of veering phenomenon corresponding to the eigen-frequencies and the modal damping
factor over a small window of speed over which the same line styles have been maintained corresponding to modes.

0.9

0.8 After curve veering

0.7 1B
Modal damping factor ( )

1F
0.6 Curve veering region 2F
3F
0.5

0.4 Before curve veering

0.3 1F
2F
0.2
1B
0.1 3F

0
SLS
-0.1
0 500 1000 1500 2000 2500 3000 3500
rotor spin speed in rpm ( )

Fig. 4. Stability diagram of rotor-shaft system without rotor-shaft material damping (case 1).
88 M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93

a
1070
3F

Damped Natural freq. (rpm)


2F
1065 3F 2F 1F

1060
2F 1F
1055
1B 1B 1B
1050

2F 1F
1045

1940 1960 1980 2000 2020 2040


rotor spin speed in rpm ( )

b
0.4

0.35
Modal damping factor ( )

0.3

0.25

0.2

0.15

0.1

0.05

1940 1960 1980 2000 2020 2040


rotor spin speed in rpm ( )
Fig. 5. Zoomed view of (a) curve veering region of Fig. 3; (b) curve veering region of Fig. 4.

Figs. 6 and 7 show respectively the first 4 mode shapes at 1800 rpm and 2200 rpm, the spin-speeds below and above the stability
limit speed. In these plots, beginning of the whirl-locus at a plane is marked with a star and the locus is left incomplete at the end
to assess the direction of whirl. It is seen, that before veering, i.e. at a speed of 1800 rpm, the 1F mode is conical having a node at a
distance of 0.7951 m from the left end, the 2F-mode is cylindrical combining rigid-body motion and bending of the rotor-shaft,
whereas the 1B and 3F are both bending modes. At 2200 rpm, i.e. above stability limit speed, the ‘1B’ and ‘1F’ modes are cylindrical
causing bending of the rotor-shaft, whereas the 2F-mode is conical and the 3F-mode is a cylindrical one combining rigid-body motion
as well as bending of the rotor-shaft. By examining, Figs. 4 and 3 at 2200 rpm, it is seen that the first forward mode (i.e. 1F mode) of
the rotor-shaft gets unstable and tends to whirl with large amplitude at a frequency little less than 50% of the spin-frequency.

3.2. Effect of shaft material damping on stability diagram

Influence of shaft material damping on the stability of the rotor-shaft system has been investigated in this section. Internal ma-
terial damping has very marginal effect on the eigen-frequencies. Following Zorzi and Nelson [21], it may be seen that the internal
material damping introduces a symmetric damping matrix and a spin-speed dependent skew symmetric circulatory matrix when

a b c d

Fig. 6. Mode shapes at 1800 rpm, (a) Mode 1F, (b) Mode 2F, (c) Mode 1B, (d) Mode 3F.
M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93 89

a b c d

Fig. 7. Mode shapes at 2200 rpm, (a) Mode 1B, (b) Mode 1F, (c) Mode 2F, (d) Mode 3F.

the equations of motion are written with respect to the inertial frame of reference. Therefore, the eigen-frequencies for whirl-motion,
and, as a result, the Campbell diagram do not change much. In this case the new critical speed of the rotor-shaft system is obtained as
1072 rpm compared to 1066 rpm when the internal material damping is not considered. However the internal shaft material
damping has a pronounced effect on the modal damping factors. For this purpose, the modal damping factors for the first 4
modes are generated for case-2, i.e. by including in the model, viscous form of shaft material damping and plotted with the
spin speed in Fig. 8. For a comparison, modal damping factors in first four modes for case-1 and case-2 are tabulated in
Table 2. From Fig. 8 it is noticed that as the spin speed is changed, the 1F mode (forward bending mode, same as the case
1) becomes unstable but at a spin speed close to 1800 rpm, in comparison to 2052 rpm approximately, when the internal
material damping is not considered. Hence shaft material damping pulls down the SLS of the rotor-shaft-system to
1800 rpm approximately. At the same time comparison of Figs. 4 and 8 show that the modal damping factor decreases
more rapidly in case-2 in comparison with case-1.
Next the values of SLS, obtained for the system for a few increasing values of ηv, are given in Table 1 to show the extent by
which the SLS is reduced. From the table it is apparent that SLS is fairly sensitive to the variation of ηv, so, correct estimation of
internal material damping factor or, in general, correct modeling of internal friction forces acting in a rotor-shaft system is essential
for accurate prediction of the stability limit speed.

4. The dFRF plots for the rotor-shaft system

In this section, initially, the dFRF at node 2 (Hp2g2) for the rotor-shaft system of Case 1 (i.e. without considering rotor-shaft
material damping) has been plotted at two spin speeds: at 200 rpm, which is below the stability limit speed and at 2052 rpm,
the stability limit speed (SLS). The damped natural frequencies and modal damping factors at these speeds in the first four
modes are shown in Tables 2 and 3 respectively. Since, below the stability limit speed (200 rpm) the modal damping in the initial
two modes is very high, (as also may be seen from the plot of modal damping factor in Fig. 4), the response is highly over-damped
and is not visible in the dFRF plot, shown by using black line, in Fig. 9. However, to prove this point, the actual dFRF plot has been
compared with a hypothetical dFRF plot, using a red (dashed) line, in which the modal damping factor in the initial four modes is
assumed equal to zero. The comparison shows significant effect of journal bearing damping in the initial four modes. It should be
noted here that the initial two modes, which are not visible in the dFRF plot, are corresponding to modes of “Oil Whirl”. Similarly,
Fig. 10(a) shows the dFRF (Hp2g2) plot of the rotor shaft system for case 1 at a spin speed of 2052 rpm. Since the peaks are very
close and are not visible in the Fig. 10(a), a zoomed views of the negative and positive frequency regions are plotted in Fig. 10(b)
and (c) respectively. It is seen from the Fig. 10 that only initial two modes get excited. Also comparing Figs. 10(b) and (c) in the
light of the theory of complex modal testing (Lee [25]), the directivity of first mode is identified as backward (1B) whereas that of
the second mode is identified as forward (1F) since the magnitude of 1st (2nd) mode is more in negative (positive) frequency
region and less in positive (negative) frequency region. The 3rd (2F) and 4th (3F) modes do not appear in the dFRF plot due to

0.9
After curve veering
0.8
1B
0.7
Modal damping factor ( )

1F
curve 2F
0.6 veering
region 3F
0.5
Before curve veering
0.4 1F
0.3 2F
1B
0.2 3F

0.1

0
SLS
-0.1
0 500 1000 1500 2000 2500 3000 3500
rotor spin speed in rpm ( )

Fig. 8. The effect of shaft material damping on stability diagram (case 2).
90 M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93

Table 1
Effect of amount of shaft material damping on stability limit speed.

No. ηv (s) SLS (rpm)

1 0.00 2052
2 0.0002 1803
3 0.0003 1705
4 0.0004 1625

Table 2
Modal parameters at spin speeds of 200 rpm and 2052 rpm for case 1.

Speed 1st mode 2nd mode 3rd mode 4th mode


(rpm)
ωd (cpm) ζ ωd (cpm) ζ ωd (cpm) ζ ωd (cpm) ζ

200 147.2(F) 0.7742 147.7(F) 0.7735 1072.9(B) .0099 1074.4(F) 0.0016


2052 1051.1(B) 0.0058 1063.8(F) − 0.000007 1073.8(F) 0.4158 1101.1(F) 0.4254

Table 3
Modal parameters at spin speed of 1803 rpm for case 2.

Speed 1st mode 2nd mode 3rd mode 4th mode


(rpm)
ωd (cpm) ζ ωd (cpm) ζ ωd (cpm) ζ ωd (cpm) ζ

1803 951.1(F) 0.4506 977.1(F) 0.4534 1053.9(B) 0.0362 1062.4(F) − 3.31e−5

high damping in these modes. The 1F mode in the dFRF plot is corresponding to the mode of “Oil Whip”. Fig. 11 shows the dFRF
(Hp2g2) plot of the rotor-shaft system for case 2 (i.e. considering the rotor-shaft material damping) at the stability limit speed of
1803 rpm. Here it is seen that the 3rd (1B) and 4th (3F) mode appears in the plot, whereas the 1st (1F) and 2nd (2F) mode which
are high damping modes do not appear in the plot. The directivity of 3rd mode is identified as backward (1B) whereas that of the
4th mode is identified as forward (3F) since the magnitude of 3rd (4th) mode is more in negative (positive) frequency region and
less in positive (negative) frequency region.

5. Conclusions

This work investigates the modal analysis of a rotor-shaft system supported on journal bearing by including the effect of rotor-
shaft material damping, which has not yet been addressed in the literature. Damping in the mode corresponding to oil whirl is
found to be very high which gradually decreases with increase in spin speed of rotor-shaft and becomes negative at a speed of
about two times the first critical speed the rotor-shaft-system ushering in the phenomenon of “oil whip”, however, the mode
shapes taken by the rotor-shaft during whirl and whip are different. It has been found that the shaft material damping

-25

-30
Abs. dFRF,dB scale, ref 1 m/N

-35

-40

-45

-50

-55

-60

-65

-70
-1500 -1000 -500 0 500 1000 1500
Excitation frequency ( ) in cpm

Fig. 9. dFRF at node 2 (Hp2g2) for the rotor shaft system supported on journal bearing at 200 rpm for case 1.
M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93 91

a
-20
Negative Frequency Positive Frequency
Region Region 1F

-30 1F

Abs. dFRF,dB scale, ref 1 m/N


-40

1B 1B
-50

-60

-70
-1000 -500 0 500 1000
Excitation frequency ( ) in cpm
b c
Abs. dFRF,dB scale, ref 1 m/N

Abs. dFRF,dB scale, ref 1 m/N

-20 -20
1F
-30 1F -30

-40 -40
1B
1B
-50 -50

-60 -60

-70 -70
-1150 -1100 -1050 -1000 1000 1050 1100 1150
Excitation frequency ( ) in cpm Excitation frequency ( ) in cpm

Fig. 10. (a) dFRF at node 2 (Hp2g2) for the rotor shaft system supported on journal bearing at 2052 rpm (case 1), (b) zoomed view of negative frequency region,
(c) zoomed view of positive frequency region.

considerably decreases the stability limit speed which is very sensitive to the coefficient of internal material damping. Strength of
instability (the rate at which the modal damping falls) in the mode corresponding to “oil whip” is also considerably increased. So,
correct estimation of internal material damping factor or, in general, correct modeling of internal friction forces acting in a rotor-

-20
3F
Abs. dFRF,dB scale, ref 1 m/N

Negative Positive
-30 Frequency Frequency
Region Region

-40 3F

1B
-50
1B

-60

-70
-1000 -500 0 500 1000
Excitation frequency ( ) in cpm

Fig. 11. dFRF at node 2 (Hp2g2) for the rotor shaft system supported on journal bearing at 1803 rpm (case 2).
92 M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93

shaft system is essential for accurate prediction of the stability limit speed. Finally the directional frequency response analysis
shows that, at low speeds, the modes corresponding to oil whirl does not appear in the plot as these are highly damped, but as
the spin speed approaches the instability limit speed the mode corresponding to “oil whip” becomes visible.

Acknowledgements

The first author, hereby, acknowledges the authorities of SGSITS, Indore, for granting him leave of absence to conduct his doctoral
research work in the department of Mechanical Engineering, IIT Delhi under the Quality Improvement Program.

References

[1] J.S. Rao, Rotor Dynamics, 3rd ed New age international (P) Ltd, 1996.
[2] J.M. Vance, Rotordynamics of turbomachinery, A Wiley-Interscience Publication, New York, 1988.
[3] M. Lalanne, G. Ferraris, Rotordynamics Prediction in Engineering, Second ed John Wiley and Sons, 1998.
[4] H. Irretier, Mathematical foundations of experimental modal analysis in rotor dynamics, Mech. Syst. Sig. Process. 13 (1999) 183–191.
[5] D.J. Ewins, Modal testing: Theory, practice and application, Research Studies Press, Baldock, Hertfordshire, England, 2000.
[6] M.I. Friswell, J.E.T. Penny, S.D. Garvey, A.W. Lees, Dynamics of Rotating Machines, Cambridge University Press, 2010.
[7] M. Cheng, G. Meng, J. Jing, Numerical and experimental study of a rotor-bearing-seal system, Mech. Mach. Theory 42 (2007) 1043–1057.
[8] R. Holmes, Oil‐whirl characteristics of a rigid rotor in 360° journal bearings, Proceedings of the I. Mech. E. 177 (1963) 291–307.
[9] Y.Y. Zhang, Y.B. Xie, D.M. Qiu, Identification of linearized oil-film coefficients in a flexible rotor-bearing system, part I: Model and simulation, J. Sound Vibr.
152 (1992) 531–547.
[10] A.K. Tieu, Z.L. Qiu, Identification of sixteen dynamic coefficients of two journal bearings from experimental unbalance responses, Wear 177 (1994) 63–69.
[11] Z.L. Qiu, A.K. Tieu, Identification of sixteen force coefficients of two journal bearings from impulse responses, Wear 212 (1997) 206–212.
[12] R. Tiwari, V. Chakravarthy, Simultaneous identification of residual unbalances and bearing dynamic parameters from impulse responses of rotor-bearing
systems, Mech. Syst. Sig. Process. 20 (2006) 1590–1614.
[13] E.S. Gutierrez-wing, Modal analysis of rotating machinery structures, Deptt. of Mech. Engg, Imperial College London, London, 2003.
[14] A. Muszynska, Whirl and whip–rotor/bearing stability problems, J. Sound Vibr. 110 (1986) 443–462.
[15] G. Stepan, P. Frank, T. Muller, F. Szlivka, Vibration problems on steam turbine-generators at low power, Proceedings of IFToMM 9th World Congress on Theory of
Machines and Mechanisms, Milan, 1995, pp. 1344–1388.
[16] A. Muszynska, Rotor dynamics, CRC Press, Taylor and Francis, 2005.
[17] J.S. Rao, Instability of rotors mounted in fluid film bearings with a negative cross-coupled stiffness coefficient, Mech. Mach. Theory 20 (1985) 181–187.
[18] S.H. Crandall, Physical Nature of Rotor Instability Mechanisms, Proceedings of the Rotor Dynamical Instability, ASME special publication, AMD, 55, 1983,
pp. 1–18.
[19] A. EI-Shafei, S.H. Tawfick, M.S. Raafat, G.M. Aziz, Some Experiments on Oil Whirl and Oil Whip, J. Eng. Gas Turbines Power 129 (2007) 144–153.
[20] A. Tondl, The effect of internal damping on the stability of rotor motion and the rise of self excited vibrations, Some Problems of Rotor Dynamics, Publishing
House of Czechoslovak Academy of Sciences, Prague, 1965, pp. 1–6.
[21] E.S. Zorzi, H.D. Nelson, Finite element simulation of rotor- bearing systems with internal damping, J. Eng. Power, Trans. ASME 99 (1977) 71–76.
[22] G. Genta, Time domain simulation of rotors with internal damping, Proceedings of ImechE, 2008, pp. 799–810.
[23] J.K. Dutt, H. Roy, Viscoelastic modelling of rotor–shaft systems using an operator-based approach, Proc. IMechE . Mech. Eng. Sci. 224 Part C: J.
[24] Y.G. Jei, Y.J. Kim, Modal Testing Theory of Rotor-Bearing Systems, J. Vibr. Acoust. Trans. ASME 115 (1993) 165–176.
[25] C.W. Lee, A complex modal testing theory for rotating machinery, Mech. Syst. Sig. Process. 5 (1991) 119–137.
[26] C.-Y. Joh, C.-W. Lee, Use of dFRFs for diagnosis of asymmetric/anisotropic properties in rotor-bearing system, J. Vibr. Acoust. Trans. ASME 118 (1996) 64–69.
[27] M. Chouksey, S.V. Modak, J.K. Dutt, Influence of rotor-shaft material damping on modal and directional frequency response characteristics, Proceedings of
ISMA-2010, Katholieke Universiteit, Leuven, Belgium, 2010, pp. 1543–1557.
[28] C.-W. Lee, S.-K. Lee, An efficient complex modal testing theory for asymmetric rotor systems: use of undirectional excitation method, J. Sound Vibr. 206
(1997) 327–338.
[29] A. Mesquita, M. Dias, U. Miranda, A comparison between the traditioal Frequency Response Function(FRF) and the directional Frequency Response Function
(dFRF) in rotor dynamic analysis, First South American congress on computational mechanics, MECOM, Santa fe-parana, Argentina, 2002, pp. 2227–2246.
[30] I. Bucher, D.J. Ewins, Modal analysis and testing of rotating structures, Phil. Trans. R. Soc. London 359 (2001) 61–96.
[31] C. Kessler, J. Kim, Vibration analysis of rotors utilizing implicit directional information of complex variable descriptions, J. Vibr. Acoust. Trans. ASME 124
(2002) 340–349.
[32] C.-W. Lee, Crack detection in rotating machinery by modal testing, 7th International Conference on Vibrations in Rotating Machinery, Nottingham, 2000,
pp. 535–543.
[33] J.H. Ginsberg, B.B. Wagner, Modification of the algorithm of mode isolation to process directional frequency response functions, Proceedings of 23rd International
Modal Analysis Conference Orlando, Florida, 2004.
[34] Y.G. Jei, C.-W. Lee, Does curve veering occur in the eigenvalue problem of rotors? J. Vibr. Acoust. Trans. ASME 114 (1992) 32–36.

Notations
qe : Nodal displacement vector
M: Mass matrix
G: Gyroscopic matrix
ηv: Coefficient of viscous form of shaft material damping
Ω: Spin speed of rotor-shaft
Ksh : Global stiffness matrix due to shaft bending
Kbrg : Global stiffness matrix due to journal bearing
Kyy, Kyz, Kzy, Kzz: Journal bearing stiffness coefficients
Cbrg : Global damping matrix due to journal bearing
Cyy, Cyz, Czy, Czz: Journal bearing damping coefficients
Hcir : Circulatory matrix
N: Number of finite beam elements
qðt Þ: Global displacement vector
yðt Þ: Displacement vector for motion in XY plane
zðt Þ: Displacement vector for motion in ZX plane
M. Chouksey et al. / Mechanism and Machine Theory 48 (2012) 81–93 93

f ðt Þ: Global force vector


A and B: System matrices of state space equation
wðt Þ: State vector
λr: rth eigenvalue
ur ; vr : rth right hand and left hand eigenvector relating to generalised displacement
ψr ; φr : rth right hand and left hand eigenvector in state space
Ψ; Φ: Right hand and left hand eigenvector matrix in compact form
Λ: Eigenvalue matrix
ξðt Þ: Modal co-ordinate vector
nðp t Þ:ffiffiffiffiffiffiffi
Modal
ffi excitation vector
j: −1
ω: Frequency of excitation
ωd: Damped natural frequency (imaginary part of eigenvalue)
Ωcr: Critical speed
ζ: Modal damping factor
H: FRF matrix in state space
H: FRF matrix relating to generalised displacement and forces
Hyy ; Hyz ; Hzy ; Hzz : Element FRF matrices of the H matrix
Hpg : dFRF matrix
Hpĝ : r-dFRF matrix
y, z: Translations along Y and Z axis
fy(t), fz(t): Forces along Y and Z axis
My(t), Mz(t): Moments about Y and Z axis

Subscripts

y, z: Directions along y and z axis


p, g: Represent complex displacement and force rotating in same direction as shaft spin.
p̂ ; ĝ : Represent complex displacement and force rotating in opposite direction of shaft spin.

Abbreviations

FRF: Frequency response function


dFRF: Directional FRF
r-dFRF: Reverse directional FRF
SLS: Stability limit speed
rF, rB: rth forward and rth backward mode
SWL: Synchronous whirl line

Das könnte Ihnen auch gefallen