Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Journal of Sound and Vibration 442 (2019) 330e349

Contents lists available at ScienceDirect

Journal of Sound and Vibration


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

Sommerfeld effect at forward and backward critical speeds in


a rigid rotor shaft system with anisotropic supports
Saurabh Kumar Bharti, Alfa Bisoi, Anubhab Sinha, A.K. Samantaray*,
R. Bhattacharyya
Systems, Dynamics and Control Laboratory, Department of Mechanical Engineering, Indian Institute of Technology Kharagpur, 721302,
India

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

Article history: A rotor dynamic system requires high power/torque to accelerate through its resonance
Received 12 June 2018 conditions or critical speeds to operate at a super-critical speed. Most practical drive
Received in revised form 29 September 2018 sources are non-ideal in nature, i.e. they can only provide a limited amount of power to the
Accepted 2 November 2018
rotor system. Thus, if there is insufficient power to overcome a resonance then the rotor
Available online 7 November 2018
Handling Editor: D.J Wagg
speed may get arrested at that resonance or take a long time to escape from the resonance;
and thereby damage the system. For the present analysis, a Direct Current (DC) motor is
considered as the non-ideal prime mover for a rigid rotor shaft with unbalanced disk
Keywords:
Non-ideal source
which is supported by flexible anisotropic supports/bearings. Due to the anisotropy in the
Rotor dynamics supports, both forward and backward whirl motions of the rotor get excited. This causes a
Forward and backward whirl multi-Sommerfeld effect where two non-linear jump phenomena are manifested at the
Sommerfeld effect first forward and backward critical speeds during rotor coast up and coast down. When the
Double resonance capture two critical speeds are close to each other, a complex speed capture and jump phenomena
Bond graph occurs. Accordingly, the transitions through both resonance conditions are examined. At
first, the analytical solutions are obtained under steady state assumption and then tran-
sient simulations are performed with the help of a bond graph (BG) model.
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction

Most rotor dynamics studies ignore the possibility resonance at backward whirl critical speeds. When considering the
synchronous critical speeds of rotor-bearing systems, the majority of attention in the technical literature has been focused on
the unbalanced excitation of natural frequencies with forward precession modes (i.e. whirling and spinning in the same
direction) only. However, while not so commonly recognized, a backward precession mode can also be excited using just the
rotor unbalance as the chief driving force. This has been documented before in some defining works in this area [1e5], while
not disregarding others.
Analysis of a Jeffcott rotor-bearing model shows that the occurrence of such a backward whirl response is strongly
dependent on stiffness asymmetry at the support ends [4e6]. A recent analysis of a vertical overhung unbalanced rotor
supported on flexible anisotropic bearings [7] predicted a synchronous critical speed to occur as a direct consequence of

* Corresponding author.
E-mail address: samantaray@mech.iitkgp.ac.in (A.K. Samantaray).

https://doi.org/10.1016/j.jsv.2018.11.002
0022-460X/© 2018 Elsevier Ltd. All rights reserved.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 331

exciting a backward precession mode. As a result of encountering this surprising new critical speed, another such system with
stiffness asymmetry at the supports was also considered [8].
Practical rotor systems, such as a three-stage compressor supported by fluid film bearings (idealised as anisotropic flexible
supports) was recently analysed in Ref. [9]. However, assuming rigid supports for such systems can introduce a huge inac-
curacy in the computed critical speeds; with the error limits for the first and second critical speed reportedly within 21e30%
and 40e88%, respectively [9]. On the other hand, some other notable works have shown that considering flexible supports
(instead of rigid supports) may actually improve the stability characteristics of the rotor systems [10e13].
For a motorised rotor system, the passage through structural resonance of the dynamical system is a key feature that often
goes unnoticed in the design stage. Most of the industrial rotor systems like compressors and turbines are known to operate
in the supercritical speed regions. To operate the systems at such high speeds, it is necessary to smoothly pass the resonance
zone. This is precisely where the consideration of the non-ideal drive dynamics and its mutual interaction with the rotor
system dynamics becomes a matter of concern.
Broadly speaking, researchers mostly assume the drive sources for dynamical systems as ideal. However, such an
assumption has a fundamental failing, since any real drive can output only a limited amount of power which itself is influ-
enced by the load on the drive, i.e., a real drive always behaves as a non-ideal energy source. It has since been documented
that for a non-ideally powered rotor system, the shaft spinning speed may get stuck at various structural resonance or critical
speed regions. This is known as speed entrainment or capture at resonance. Consequently, a significant amount of input
power is dissipated via resonant structural vibrations instead of contributing to increase the motor speed. Continuous op-
erations when the system is captured at resonance may lead to alarmingly high vibration levels and damage to the rotor
bearings; or in extreme case, overall system failure [14e18]. Interestingly though, on increasing the power supply to a certain
limit, the entrained speed jumps up to a much higher value in almost no time. This speed jump is also accompanied by a
corresponding drop in the vibration amplitudes. This ‘jump’ implies a sudden escape from the resonance.
To summarize, if a non-ideal rotor system with eccentric unbalance is operated close to a critical speed then a non-linear
jump phenomenon is observed. The literature acknowledges this jump phenomenon with all its manifestations as the
Sommerfeld effect [19e21]. It is credited to the German theoretical physicist Arnold Sommerfeld who first observed this
‘jump’ during an experiment, back in 1902 [22]. Afterwards, many other researchers started paying attention to various non-
ideal systems and their associated implications. The documented systems mostly feature eccentrically loaded rotor-disk
systems with electric motors (DC motors and induction motors) as the non-ideal drives [14e18,23e30]. Recent evidence
suggests there may be something similar to Sommerfeld effect in vehicle dynamics as well [31].
At its core, the Sommerfeld effect is basically a non-linear resonance phenomenon. The ‘jump’ witnessed at resonance
implies that the resonance peak for a non-ideal system is discontinuous; unlike its ideal counterpart. Due to the discontinuity,
there is a missing range of operating spin speeds [32e35]. The system fails to make a smooth transition through the reso-
nance. More specifically, for a rotor-dynamic system with an undersized actuator, the desired operation above the critical
speed region may not be possible. Therefore, identification of the resonance regions and proper representation of the jump
phenomena inherent within a particular system is deemed essential to avoid unexpected failures.
In view of the foregoing discussions, the authors acknowledge that no study pertaining to the response of rotor systems
with anisotropic supports (for example, fluid film bearings) and driven by non-ideal drive has been documented as yet. To
address that, the present analysis is undertaken to determine the influence of flexible anisotropic supports on the syn-
chronous unbalance response of a rigid rotor; when the vibrating system is electrically powered by a DC motor. The analysis
carefully incorporates the role of gyroscopic effect within the rotor dynamics. It is found that the anisotropic support flexi-
bilities end up exciting both the forward and the backward whirl modes. On account of that, the passages through two
resonance conditions are examined. Subsequently, the multi-Sommerfeld effect as exhibited by the considered rotor system is
then characterised for a detailed study.

2. Sommerfeld effect characterization of non-ideal system

2.1. System description

The schematic representation of the rotor-disc system to be investigated is shown in Fig. 1. It comprises a perfectly
balanced and rigid rotating shaft which is driven by a DC motor. Although the shaft itself is rigid, it is supported at its two ends
on flexible anisotropic bearings. The bearings are ideally represented by two massless spring-dampers in two orthogonal
directions, which permit in-plane small lateral motions.
An unbalanced, heavy disc is then mounted firmly onto the same shaft, with the disc plane perpendicular to the shaft axis.
It is assumed that disc spinning axis coincides with the shaft axis for small deflections and the eccentricity is small. The disc is
asymmetrically placed closer to the left end bearing and as a result, gyroscopic effect and conical modes of vibration come
into play. The shaft mass and rotary inertias are neglected in comparison to the heavy disk. It is assumed that the overall mass
of rotor and disc are concentrated at the eccentric mass centre of the disc itself. Additionally, external aerial damping (for both
translational and rotary modes) is also considered at the disc geometric centre.
Connection between the DC motor and the rotor shaft is established through a dissipative/universal coupling which is
assumed to resist and transmit torsion, but does not provide resistance to bending. The electrical power to the DC motor is
supplied from a constant voltage source; i.e., the idealization boundary is shifted from a constant speed source in the
332 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

Fig. 1. Schematic of the non-ideal rotor dynamic system with anisotropic bearings.

mechanical domain to a constant voltage source in the electrical domain of the motor. In essence, the source needs to be
idealised somewhere to create a model. The model complexity increases as we move the idealization limit to more primitive
causes of the power source, such as chemical reactions at the cell level.

2.2. Steady-state equations and modal analysis

With the combination of a rigid shaft and two flexible end supports, the degrees of freedom for the rotor system is reduced
to four - two in the transverse directions and two in angular directions. Apart from these, one additional degree of freedom for
the non-ideal source, representing the rotor spin dynamics, will be introduced later. Note that the due to gyroscopic effect, the
whirl frequencies change with the rotor speed.
The shaft has a uniform circular cross-section. As stated earlier, the rotor disc is asymmetrically placed at one-third of the
shaft length, from its left end support. The disc mass, m, is assumed to be located eccentrically at a distance e from the main
shaft axis. The position of the mass centre M is ðxm ; ym Þ; and the geometric centre G is ðx; yÞ; as depicted in Fig. 2. Accordingly,

xm ¼ x þ e cosðq þ fÞ; ym ¼ y þ e sinðq þ fÞ (1)

where, q is the instantaneous angle of rotation of the rotor disc about the spin axis (i.e. z-axis) and f is the constant phase
corresponding to the initial position of the mass centre of the disc. The angles fx and fy describe the small angular motions of
the rotor disc about the positive x and y directions, respectively. For steady state harmonic motions, the shaft has a constant
angular/spin speed about the z-axis, i.e., q_ ¼ u. The instantaneous angle between the x-axis and the line passing from the
origin to the geometric centre is then expressed as q ¼ ut. With reference to Fig. 1, as per the prescribed co-ordinate axes, the
right hand thumb rule dictates that fx is clockwise in y-z plane (fx is anticlockwise) and fy is anticlockwise in x-z plane.
Here, l1 and l2 are the distances of disc location from the left and the right support ends, respectively. Thus, the displacements

disc

O
x

Fig. 2. Positions of the geometric and mass centre of the rotor disc.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 333

of the left and the right end of the shaft are obtained as ðy þ l1 fx Þ and ðy  l2 fx Þ in y-z plane, respectively. Similarly, for x-z
plane, the displacements of the same points are obtained as ðx  l1 fy Þ and ðx þ l2 fy Þ, respectively. Kx and Rx are the bearing
stiffness and damping in transverse x-direction for both ends of the shaft; while, Ky and Ry represent the same in transverse y-
direction.
For the disc, Re and Ref represent the external translational and rotational (about diametral axes) damping values,
respectively. The external damping forces and moments act at the geometric centre of the disc. The rotary inertia of disc about
the spin axis is Ip and the diametral moment of inertia is Id . For the analysis, the disc is assumed to be thin and laterally
symmetric, which implies Ip ¼ 2Id . Based on the above generalizations, the equations of motion for the rotor disc system
(excluding the non-ideal motor) at a constant rotor speed are obtained as:
       
mx€þRx xþK
_ _
x xþð2L=3Þfy þKx xðL=3Þfy þRx xþð2L=3Þ f_ y þRx xðL=3Þ
_ f_ y ¼meu2 cosðutÞ
   
€ x yþK
myþR _ _
y ðyð2L=3Þfx ÞþKy ðyþðL=3Þfx ÞþRy yð2L=3Þ f_ x þRy yþðL=3Þ
_ f_ x ¼meu2 sinðutÞ
       
€ y þRef f_ y þKx xþð2L=3Þfy ð2L=3ÞKx xðL=3Þfy ðL=3ÞþRx xþð2L=3Þ
Id f _ f_ y ð2L=3ÞþRx xðL=3Þ
_ f_ y ðL=3Þ uIp f_ x ¼0
   
€ x þRef f_ x þKy ðyþðL=3Þfx ÞðL=3ÞKy ðyð2L=3Þfx Þð2L=3ÞþRy yþðL=3Þ
Id f _ f_ x ðL=3ÞRy yð2L=3Þ
_ f_ y ð2L=3Þþ uIp f_ y ¼0
(2)

Excluding the unbalance excitations, Eq. (2) can be written in the state-space form as
h iT h iT
x€ f € x x_ f_ y y_  f_ x
€ y y€  f ¼ ½A88 x_ f_ y y_ f_ x x fy y fx (3)
2 3
a11 a12 0 0 a15 a16 0 0
6 a21 a22 0 a24 a25 a26 0 0 7
6 7
6 0 0 a33 a34 0 0 a37 a38 7
6 7
6 0 a42 a43 a44 0 0 a47 a48 7
where ½A88 ¼6
6 1
7
6 0 0 0 0 0 0 0 77
6 0 1 0 0 0 0 0 0 7
6 7
4 0 0 1 0 0 0 0 0 5
0 0 0 1 0 0 0 0
         
2Rx þ Re L Rx 2Kx L Kx
with a11 ¼  ; a12 ¼  ; a15 ¼  ; a16 ¼  ;
m 3 m m 3 m
           
L Rx Ref 5 uIp L Kx 5
a21 ¼ ; a22 ¼  þ Rx L2 ; a24 ¼  ; a25 ¼  ; a26 ¼  Kx L2 ;
3 Id Id 9Id Id 3 Id 9Id
          (4)
2Ry þ Re L Ry 2Ky L Ky
a33 ¼ ; a34 ¼  ; a34 ¼  ; a35 ¼  ;
m 3 m m 3 m
           
uI p L Ry Ref 5 L Ky 5
a42 ¼ ; a43 ¼  ; a45 ¼  þ Ry L2 ; a47 ¼  and a48 ¼  Ky L2
Id 3 Id Id 9Id 3 Id 9Id

Note that in Eq. (3), negative rotation about x-axis is considered as a state in the state space formulations, as per the right
hand rule convention. To study this rotor dynamic system numerically, the representative parameter values listed in Table 1
are considered. The eigenvalues of ½A88 are calculated by using the parameters values from Table 1. Subsequently, a
Campbell diagram for the rotor dynamic system is generated by plotting the imaginary part of the eigenvalues against the

Table 1
The system parameters, their description and values.

Parameters Description Value


L ¼ l1 þ l2 Shaft length 0.9 m
l1 Distance between left end bearing and disk 0.3 m
l2 Distance between disk and right end bearing 0.6 m
Ip Polar moment of inertia of rotor disc 0.48 kgm2
Id Diametral moment of inertia of rotor disc 0.24 kgm2
m Mass of the rotor disc 15 kg
e eccentricity of the rotor disc 0.001 m
Re Translational aerial damping coefficient on rotor disc 20 N/m
Ref Rotational aerial damping coefficient on rotor disc 1 Nms/rad
Rx Bearing damping coefficient in x-direction 5 N/m
Ry Bearing damping coefficient in y-direction 10 N/m
Kx Bearing stiffness in x-direction 20000 N/m
Ky Bearing stiffness in y-direction 40000 N/m
Rb Spin rotational damping coefficient 0.0005 Nms/rad
334 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

rotor spin speed, as shown in Fig. 3. A similar Campbell diagram (as shown in Fig. 3), for a rigid rotor with flexible support
having asymmetric stiffness, has been previously documented in Ref. [8].
In the Campbell diagram, the intersection of ‘1’ line with the frequency of the rotor at various speeds gives the forward
and backward critical speeds. Evidently, three intersecting points appear and accordingly, three resonance speeds are ob-
tained. These speeds are the first backward, first forward and second backward critical speeds, and are respectively obtained
as ucr1 ¼ 48:76 rad/s, ucr2 ¼ 68:86 rad/s and ucr3 ¼ 132:73 rad/s. Since lines ‘1’ and ‘2FW’ never intersect, no critical speed
is possible for the second forward mode. From the diagram, it is obvious that the first forward and backward whirl modes are
translational. On the other hand, the secondary modes (both forward and backward) emerge conical; the respective branches
show splitting due to gyroscopic effect [37].

2.3. Ideal drive responses

To study the system analytically, the amplitude responses under ideal drive conditions are established first. The ideal drive
assumption dictates that the rotor spin speed (u) is to be treated as a known/controllable parameter. The presence of
anisotropic supports leads to elliptical whirl orbits in both transverse and angular directions. Considering that, the harmonic
solutions for the synchronous whirl amplitudes and phases are assumed as follows:

x ¼ A cosðut þ aÞ; y ¼ B sinðut þ bÞ


(5)
fx ¼ C sinðut þ gÞ; fy ¼ D cosðut þ dÞ

where the pairs A, B and C, D are the amplitudes of the translational and angular motions, respectively. Similarly, the pairs a, b
and g, d are the phases of the translational and angular motions, respectively.
Substituting Eq. (5) into Eq. (2) and thereafter, separating the coefficients of cosðutÞ and sinðutÞ terms, leads to the
following equations:

L L
mAu2 cos a þ 2Kx A cos a þ Kx D cos d  ð2Cx þ Re ÞAu sin a  Du Cx sin d ¼ meu2
3 3
L L
mAu2 sin a  2Kx A sin a  Kx D sin d  ð2Cx þ Re ÞAu cos a  Du Cx cos d ¼ 0
3 3
L   L
mBu2 sin b þ 2Ky B sin b þ Ky C sin g þ 2Cy þ Re Bu cos b þ Cy C u cos g ¼ 0
3 3
L   L
mBu2 cos b þ 2Ky B cos b þ Ky C cos g  2Cy þ Re Bu sin b  Cy C u sin g ¼ meu2
3 3
 2 
L L 5L2 5L (6)
Id Du2 cos d þ Kx A cos a  Cx Au sin a þ Kx D cos d  Cx þ Ref Du sin d þ C u2 Ip cos g ¼ 0
3 3 9 9
2  2 
L L 5L 5L
Id Du2 sin d  Kx A sin a  Cx Au cos a  Kx D sin d  Cx þ Ref Du cos d  C u2 Ip sin g ¼ 0
3 3 9 9
2  2 
L L 5L 5L
Id C u2 sin g  Ky B sin b  Cy Bu cos b  Ky C sin g  Cx þ Ref C u cos g  Du2 Ip sin d ¼ 0
3 3 9 9
2  2 
L L 5L 5L
Id C u2 cos g  Ky Bcosb þ Cy Bu cos b  Ky Ccosg þ Cy þ Ref C u sin g  Du2 Ip cos d ¼ 0
3 3 9 9

In Eq. (6), there are a total of eight equations and eight unknown variables, which are solved numerically by using MATLAB
codes for particular a rotor speed u (i.e. ideal drive assumption) by using the parameter values from Table 1. The whirl
amplitudes (A, B, C, D) and phases (a; b; g; d) are obtained for a specific range of rotor speeds.
Before obtaining the amplitude responses, the parameters are first normalized so as to obtain responses with same order
of magnitude. The normalized whirl amplitudes are defined as x* ¼ 0:33ðucr1 Þ2 AðuÞ=g and y* ¼ 0:4ðucr1 Þ2 BðuÞ=g, where g is
the acceleration due to gravity. Similarly, the normalized rotary whirl amplitudes of the rotor are defined as f*x ¼ 100CðuÞ and
f*y ¼ 120DðuÞ. Lastly, the spin speed is also normalized as u* ¼ u=ucr1 .
Following the normalization procedure, the normalized steady state amplitudes x* and y* are plotted against the
normalized spin speed u* , in Figs. 4 and 5, respectively. Similarly, Figs. 6 and 7 show the responses for normalized rotary
steady state amplitudes f*x and f*y . Clearly, the peak amplitudes for x* and f*y occur at the first backward critical speed, i.e. at
u* ¼ 1, whereas the same for y* and f*x occur at first forward critical speed, i.e. at u* ¼ 1:41. Additionally, for the second
backward critical speed, i.e. at u* ¼ 2:73, two minor rotary amplitude peaks (each for f*x and f*y ) also exist. Overall, the four
figures (Figs. 4e7) describe the frequency response of the rotor system with ideal drive.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 335

450
400
2FW
350
300

Frequency, Hz
250
200
2BW
150 1X

100
1FW
50
1BW
0
0 50 100 150
, rad/s
Fig. 3. Campbell diagram showing critical speed vs. shaft spin speed (u). [FW-Forward Whirl; BW- Backward Whirl; 1-First Mode; 2-Second Mode].

2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3

Fig. 4. Normalized steady state transverse vibration amplitude ðx* Þ vs. normalized spin speed ðu* Þ.

2.4. Non-ideal drive responses and Sommerfeld effect characterization

Here, the non-ideal drive is a permanent magnet DC motor that drives the rotor shaft; the motor modelled with limited
power supply capabilities. Once this non-ideal drive is introduced, one can no longer treat the rotor speed as a known
quantity; instead the spin speed now becomes an output parameter that largely depends on the drive-system energetic
interaction. In fact, the controllable parameters now are the operating voltage and other motor specifications.
DC motors can have various types of construction and/or winding. Here a permanent magnet DC motor is considered
where the motor torque is proportional to the current and the back emf is proportional to the rotor speed [36e38]. In DC
motors, the armature coil has both resistance and inductance. If the motor operates at a constant speed in the steady-state
then the armature current becomes constant. Therefore, for steady-state analysis, the armature inductance can be neglected.
Likewise, for slow angular acceleration/deceleration of motor, the armature resistance can be neglected even in the transient
analysis. However, in systems with large motor speed variation, such as in cam drive, scotch-yoke mechanism, slider-crank
mechanism [39], etc., it is necessary to include the motor inductance effect [40]. In the present study, the rotor disc acts as a
flywheel and limits the angular acceleration/deceleration of the motor speed. Thus, the DC motor inductance is neglected.
The steady-state power developed by the DC motor is

mm uðVs  mm uÞ
Pm ¼ tm q_ ¼ ; (7)
Rm
336 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

2.5

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3

Fig. 5. Normalized steady state transverse vibration amplitude ðy* Þ vs. normalized spin speed ðu* Þ.

Vs Ve ðVs mm uÞ


with tm ¼ mm im and im ¼ Rm ¼ Rm , where, Vs is the voltage supplied across the motor terminal, Ve is the back emf
developed in motor coils, mm is the motor characteristic constant, im is the armature current and Rm is the armature resistance
of the coils. Also, t is the mechanical torque developed by the motor and q_ ¼ u is the rotor angular velocity under steady
m
condition.
For synchronous whirl conditions, the energy supplied by motor for any particular rotor speed is exactly equal to the
energy dissipated by the rotor’s dissipative mediums [34e37,41]. To put it analytically, the average motor power and the
power dissipated by all parts of the system are to be worked out separately and then mathematically equated. However,
instantaneous power balancing is not possible due to highly elliptical nature of vibratory motions. Instead, energy balancing is
carried out by integrating the input-output powers over a cycle of synchronous whirl, which is similar to averaging over a
period of oscillation. By using this energy balance equation, the Sommerfeld effect governing this particular system can be
theoretically estimated.
The energy dissipation primarily takes place through the external damping, i.e. Re , Ref and damping in the bearings, i.e. Rx
and Ry . In steady state, the motor and rotor speeds are equal; so there is no energy dissipation at the rotor-motor coupling.
Overall, the net work done by all the dissipative forces over a fixed cycle can be expressed as

2Zp=u
 
Wd ¼ Pd;left þ Pd;right þ Pd;disc þ Rb u2 dt (8)
0

where,
2 3 2 3 2 3
Rx 0 0 0 Rx 0 0 0 Re 0 0 0
T6
6 0 0 0 07 6
7q_ L; Pd;right ¼ q_T 6 0 0 0 07 6
7q_ ; Pd;disc ¼ q_T 6 0 Ref 0 0 7
7q_
_
Pd;left ¼ qL 4
0 0 Ry 0 5 R4 0 0 Ry 05 R d4 0 0 Re 0 5 d (9)
h 0 0 0 0 i h 0 0 0 0 i 0 h0 0 Ref i
      
q_ TL ¼ x_  l1 f_ y 0 y_ þ l1 f_ x 0 ; q_ TR ¼ x_ þ l2 f_ y 0 _y  l2 f_ x 0 ; and qd ¼ x f_ y y_ f_ x
_ T
_

On the other hand, the energy supplied by the motor in each cycle is given by

2Zp=u

Wm ¼ Pm dt;
0

The steady state energy balance can now be mathematically put forth as

Wm ¼ Wd (10)

or,
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 337

Fig. 6. Normalized steady state rotary vibration amplitude ðf*x Þ vs. normalized spin speed (u* ).

Fig. 7. Normalized steady state rotary vibration amplitude ðf*y Þ vs. normalized spin speed (u* ).

2  2   2  3
2
  2 5L 2 5L 2
ð2R þ R ÞA þ 2R þ R B þ R þ R C þ R þ R D þ
2mm ðVs  mm uÞ 6 7
x e y e x ef y ef
6 9 9 7
¼ 6    7u (11)
Rm 4 2L 2L 5
Rx AD cosða  bÞ þ Ry BC cosðb  gÞ þ 2Rb
3 3

Eq. (11) relates the output rotor spin speed u to the input parameter, i.e. voltage supply Vs. Note that all other parameters in
Eq. (11) are constants and the whirl amplitudes A, B, C and D are functions of u. The motor parameters considered in this study
are given in Table 2. Accordingly, the normalized steady state rotor spin speed versus motor supply voltage is plotted in Fig. 8.
Note that instead of computing u for a given Vs , the computation is easier when Vs is computed for a given u.
Fig. 8 gives a visual representation of the Sommerfeld effect and the associated jump phenomena. With increasing rotor
speed by varying voltage supply, three resonance zones are encountered. However, the resonance at 2BW does not show any
jump phenomena and hence the corresponding speed is kept outside the speed range plotted in Fig. 8. The results show two
jumps in rotor speed e first of which occurs from ucr1 ¼ 48:97 rad/s to u ¼ 57.07 rad/s (i.e. near u* ¼ 1) and the second from
ucr2 ¼ 68:9 rad/s to u ¼ 89.1 rad/s (i.e. near u* ¼ 1:41). The two jumps correspond to 1BW and 1FW critical speed or
resonance zones.
During the rotor coast up operation i.e. when the supply voltage is increased, the rotor speed continues to increase almost
linearly until it reaches point ‘a’. Thereafter, further increase in supply voltage does not result in appreciable change in rotor
speed. For the region ‘a’ to ‘b’, although the supply voltage is being increased, the rotor speed does not increase and is stuck at
the first backward critical speed for the voltages from 26.43 V to 28.81 V. When the voltage is increased above 28.81 V, the
338 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

Table 2
DC motor parameters.

Notation Description Value


Vs DC motor voltage supply 〈Variable〉
Rm DC motor electrical armature resistance 5U
mm DC motor characteristic constant 0.5 Nm/A

rotor speed suddenly jumps from point ‘b’ to a much higher value at point ‘g’. Afterwards, by increasing voltage, the shaft
speed again increases linearly till point ‘c’ (i.e. at 37.7 V). This is the zone for the first forward critical speed. Here, the rotor
speed is stuck again, from ‘c’ to ‘d’ i.e. corresponding to 37.77 Ve44.83 V motor voltage supply. Subsequently, the second jump
takes place from ‘d’ to ‘e’. Thereafter, from point ‘e’ onwards, there is a linear trend without any further discontinuity. Since
both orthogonal bearing stiffnesses have dissimilar values, two distinct ‘jumps’ occur for the two different voltage inputs [42].
The Sommerfeld effect is thus observed for both first backward and first forward whirls of the system (hence, the ascribed tag
- ‘Multi-Sommerfeld effect’). Between the two, the jump size is comparatively higher for the first forward critical speed.
On the other hand, during coast down operation, spin speed reduces as the supply voltage is decreased. The speed trails
the route back from ‘e’ to ‘f’; and as the voltage reduces below 37.77 V (point ‘f’), the speed jumps down from point ‘f’ to ‘c’. The
trajectory then retraces back the line joining ‘c’ to ‘h’. Thereafter, a second coast down jump occurs from point ‘h’ to ‘a’. Clearly,
the jump is much larger during the coast up operations as compared to the coast down operations.
It is to be noted that there are three possible rotor speeds for the voltage ranges varying from points ‘a’ to ‘b’ and also from
‘c’ to ‘d’. The speeds lying within the ranges, 48.79e51.1 rad/s and 68.9e72.7 rad/s, are not physically achievable either in
coast-up or coast-down operations. This is because the branches in between ‘d’ to ‘f’ and ‘b’ to ‘h’ are unstable in nature
[43,44]. As a result, corresponding to two stable branches, only two possible speeds exist near the resonance region; one of
these can be accessed during coast-up operation and the other one under coast-down operation.
It is to be noted that the amplitudes of vibration are also a function of speed. As a direct consequence, there is a concurrent
amplitude jump associated with the speed jump at the respective resonance zones. Near resonance, the energy from the non-
ideal motor is not used to increase the spin speed; instead, this energy is transferred to the vibrating modes which end up
getting excited. Consequently, as speed increment recedes around the resonance region, the system starts vibrating with high
amplitudes. When the speed jumps up to a higher value upon exceeding a critical power input, the amplitudes promptly fall
back. Fig. 9 shows the plot between normalized amplitude x* and voltage supply Vs. An apparent loop in Fig. 9 wrongly
suggests intersecting values. To better visualise it, Fig. 10 depicts the three dimensional plot between voltage Vs , normalized
rotor speed u* and normalized vibration amplitude ðx* Þ.
Here, it is observed that before the first jump takes place around u* ¼ 1, the amplitude values go on increasing. As
mentioned before, around this zone, the energy supplied by motor increases the lateral vibration of system rather than
increasing the shaft speed. Afterwards, the amplitudes suddenly decrease with a simultaneous rise in spin speed (jump from
point ‘b’ to ‘g’ in Fig. 8). This first jump occurs after a peak value of amplitude x* . Later, the second resonance zone is
encountered around u* ¼ 1:41, where the amplitudes again start to increase (albeit, to a much lesser extent). Eventually, all of
these culminates with the second jump (refer jump from ‘d’ to ‘e’ in Fig. 8).

Fig. 8. Normalized spin speed ðu* Þ vs. voltage supply ðVs Þ.


S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 339

Fig. 9. Normalized steady state transverse vibration amplitude ðx* Þ vs. voltage supply ðVs Þ.

1.5

0.5

0
4
3 80
2 60
40
1 20
0 0

Fig. 10. Normalized steady state transverse vibration amplitude ðx* Þ vs. normalized spin speed ðu* Þ vs. voltage supply ðVs Þ.

Figs. 11e13 show the remaining 3D non-ideal amplitude responses (corresponding to y* , f*x and f*y respectively). Alto-
gether, these plots bring out a cumulative picture of the complete non-ideal dynamics, at steady state conditions.

2.5. Energy requirements for safe passage through resonance

Due to power insufficiency, non-ideal systems show a propensity to get stuck at resonance and hence, must be investi-
gated for quick getaways from there. For the present rotor dynamic system, an attempt is made to decide the energy
requirement conditions to ensure smooth escape from resonance.
To do that, the motor energy Wm curves for various supply voltages are plotted alongside the dissipated energy curve Wd
against normalized rotor speed u* (Fig. 14). The Wm curve for Vs ¼ 28.81 V intersects the dissipated energy curve at u* ¼ 1
and u* ¼ 1:19. For Vs ¼ 44.83 V, Wm curve intersects Wd curve at u* ¼ 1:41 and u* ¼ 1:85. The Wm curves for Vs ¼ 25 V and
Vs ¼ 35 V intersect the Wd curve at subcritical speed points. The corresponding motor energy for these two voltages is
insufficient to cross the resonance region. Accordingly, the rotor speed gets stuck at resonance indefinitely, under both
conditions. On the other hand, the Wm curve for Vs ¼ 40 V intersects the dissipated energy curve at three points, out of which
only two are stable [43]. For the given inputted parameters, Wm for Vs ¼ 28.81 V is the threshold line corresponding to the
first backward critical speed, above which smooth passage through that resonance should be possible. On the other hand, Wm
for Vs ¼ 44.83 V is the threshold line corresponding to the first forward critical speed. It must be noted that 28.81 V and
340 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

0
4
3 80
2 60
40
1 20
0 0

Fig. 11. Normalized steady state transverse vibration amplitude ðy* Þ vs. normalized spin speed ðu* Þ vs. voltage supply ðVs Þ.

44.83 V are in fact the jump voltages obtained earlier by the energy balance formulation (see in Fig. 8). Reasonably enough, to
be on the safe side, voltages above 50 V will ensure smooth passages through all the 1FW and 1BW resonance conditions.
The inset in Fig. 14 shows a zoomed view of the resonance condition at 2BW critical speed. It is seen that the energy
dissipation at 2BW and neighbourhood speeds is too small, and hence, no Sommerfeld effect is observed at 2BW critical
speed. However, with a weakly damped bearing support system and large translational and small rotational damping on the
rotor disc, it is possible to encounter Sommerfeld effect the 2BW speed. High damping often suppresses Sommerfeld effect
[15]. The net external damping in 1FW and 1BW modes comes from the bearing supports and the translational external
damping (Re ) whereas the net external damping on 2BW and 2FW modes some from the bearing supports and the rotational
external damping (Ref ). For a spoked-wheel rotor or a bladed rotor with shrouds, Re overwhelms Ref . On the other hand, Ref
overwhelms Re for a thin solid disc rotor. These relative values of damping depend on the projected surface areas of the rotor.
Note that the 2BW resonance occurs in the conical mode of vibration and this mode is excited here by the asymmetric
position of the rotor disc. When the disc is mounted at the middle of the rotor shaft, the conical mode can be excited due to
asymmetry in left and right bearing parameters or moment unbalance in the disc.
The steady state analysis is valid only if the rotor acceleration is negligible, which happens only when voltage increment is
done gradually or the polar moment of inertia of the rotor is very large. For finite or step input voltages, the steady-state

Fig. 12. Normalized steady state rotary vibration amplitude ðf*x Þ vs. normalized spin speed ðu* Þ vs. voltage supply ðVs Þ.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 341

Fig. 13. Normalized steady state rotary vibration amplitude ðf*y Þ vs. normalized spin speed ðu* Þ vs. voltage supply ðVs Þ.

Fig. 14. Energy vs. Normalized spin speed ðu* Þ.

analysis would not give accurate results. To address that, a separate transient analysis for the rotor-motor system, with non-
ideal source loading, is carried out.

3. Transient analysis of the non-ideal system

The transient analysis can be done by numerical simulations via system modelling with the help of available dynamical
software/tools. One such efficient modelling tool is bond graph. It is well-known mostly due to its ability to model complex
energy interactions between various sub-system boundaries across different domains. At its core, a bond graph is energy
consistent in nature [45,46]; which simplifies the modelling of non-ideal systems. The models are relatively easy to develop,
mostly relying on basic kinematic relations [47,48] and the properties of the power conserving junction (kinematic and
dynamic constraints) structure. Bond graph modelling has been extensively used for rotor dynamics systems [49]. Further-
more, the load reactions for the non-ideal drive are also automatically included without any added concerns [50e52]. Bond
graph modelling has been adopted here for the multi-energy domain (electro-mechanical) nature of the problem and the
ability of the modelling tool to represent the active and reactive torques on the motor in a physically consistent manner. The
computational causalities are applied to the bond graph model as per sequential causality assignment procedure (SCAP) and
thereafter, the state-space equations of the motion are derived from the model for simulation [53].
342 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

3.1. Bond graph model for the system

To develop the bond graph model so as to synthesize the energy domain analogies for the complete motor-rotor system,
few kinematic relations need to be established. First, the velocity relations between the geometric centre and the disc mass
centre are obtained by time-differentiating Eq. (1), and are obtained as
   
x_m ¼ x_  eq_ sin q_ þ f ; y_m ¼ y_ þ eq_ cos q_ þ f (12)

Now, to introduce the non-ideal drive dynamics within the model, one additional degree of freedom must be involved
(mentioned previously in Sec 2.1). As a result, one more equation is added to the four equations available in Eq. (2). This new
equation corresponds to the dynamics of the DC motor and rotor spin, and is given as

Ip €q þ Rb q_ ¼ tm  tl (13)

where, Ip is the rotary inertia of the shaft-disc system about the spinning axis, Rb is the rotational damping to the spinning
shaft, tm is the torque developed by the DC motor and tl is the load torque due to source loading. The load torque need not be
introduced separately; in fact, it can be readily determined from the bond graph model, once the complete model has been
developed.
Accordingly, the complete causalled bond graph model for the non-ideal rotor dynamic system is now developed, as
shown in Fig. 15. The subscripts labelling 1-junctions describe the distinct velocity points. The non-ideal drive source for the
system, i.e. the DC motor, is modelled with the help of Eq. (13). The armature resistance Rm and the voltage supply Vs are
modelled by a R-element and a Source of Effort Se-element, respectively. The motor characteristic constant mm is modelled by
a GY element. Note that inclusion of a small armature inductance in the electrical domain model does not appreciably change
the results.
Motor rotary inertia Ip is modelled by an I-element which is connected to a 1-junction and a flow detector Df is also
attached to detect the instantaneous shaft speed i.e. q_ ¼ u. The net load torque from the rotor-disc system on the motor is
contributed at 1q_ -junction. The spin resistance including the viscous friction in the bearings is modelled by R element as R: Rb
where, Rb is the damping parameter.
The translational velocities x_ and y_ for the shaft are modelled at 1x_ -junction and 1y_ -junction, respectively. Likewise, the
rotational velocities f_ x and f_ y are modelled at 1 _ -junction and 1 _ -junction, respectively.
fx fy
No separate inertial elements are used to model masses for the motor shaft and the rotor shaft; they are assumed to be
negligible in comparison with the mass of the heavy disc or lumped with the mass of the disc (refer Sec 2.1). As for the disc, its
diametrical moment of inertia Id is modelled by an I-element. The gyroscopic couplings are modelled by the gyrator GY-
element with moduli Ip u, where Ip is the polar moment of inertia of the rotor disc. The external aerial damping Re and
damping Ref on the rotor disc are modelled by two R-elements. The disc mass m is modelled as I: m. Thereafter, the disc
eccentricity is incorporated via two transformer-TF elements whose moduli are e sin q and e cos q as per Eq. (12). Note that
the phase f is set to zero without any loss of generality because this phase does not influence the overall steady state results.

Fig. 15. Bond graph model of the shaft-disk system coupled with DC motor.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 343

The rotor disc is situated at a distance l1 from the left end bearing, which is one-third of the total shaft length. Thus, disc
position having two-third distance from the right end bearing is denoted as l2 . Accordingly, the velocities at the support ends
are obtained with the four TF elements (with moduli l2 ,  l2 , l1 and l1 ). The bearing stiffness and damping for x-directions
are modelled by C element as C: Kx and by R element as R: Rx , respectively. The same for y-directions are modelled by C: Ky
and R: Ry element, respectively.
Furthermore, to record the instantaneous displacements, flow detectors Df are placed at all the respective 1-junctions.
These recorded displacements will indicate the vibrational amplitudes.

3.2. Simulation results

Now that the bond graph model is developed, numerical simulations are carried out the help of Symbols-Shakti software
[53]. The objective is to simulate the non-ideal system for various supply voltages and fixed parameters values (as specified in
Tables 1 and 2). The generated time responses will expose both the transient and steady-state natures of the non-ideal system
dynamics.
The simulated responses are obtained and are first normalized. For normalization of displacements and rotor spin speed,
the same normalization scale is used (as used for the steady state analysis in Sec 2.3). Once normalization is done, the
displacement responses are enveloped so as to obtain the corresponding amplitudes of vibration. The steady state analysis
already demonstrated that the system has two resonance zones and accordingly, two corresponding jumps e one for first
backward and first forward critical speeds. Therefore, it is safe to assume that two jumps might also be encountered when the
system is numerically simulated. Keeping that in mind, simulations are carried out to identify the exact voltages at which
these ‘jumps’ take place.
In Fig. 16, the transient response amplitude for a step input voltage of Vs ¼ 28:8 V is shown, for which the shaft speed is
stuck at u* ¼ 1, i.e. at the first backward critical speed. At this point, with even a slight increase of input voltage to Vs ¼ 28:81
V, the first jump is detected, as shown in Fig. 17. The vibration amplitudes have high values during the pre-jump resonating
conditions, and thereby, decrease considerably once the spin speed jumps to a high value. It can be seen that the normalized
amplitudes x* and f*y are comparatively larger than y* and f*x . Similarly, a second jump is also detected when supply voltage is
increased from Vs ¼ 44:73 V (Fig. 18) to Vs ¼ 44:74 V (Fig. 19). This is the resonance zone corresponding to the first forward
critical speed. Once again, the normalized amplitudes y* and f*x are shown to be larger than x* and f*y . A closer comparative
look at the four plots (Figs. 16e19) reveals that the amplitude peaks for the first forward critical speed are larger than those for
the first backward critical speed.
From the steady state analysis, the first jump was predicted exactly at Vs ¼ 28:81 V, and the second jump at Vs ¼ 44:83 V.
However, for the bond graph simulations, although the first jump occurs precisely at Vs ¼ 28:81 V, the second jump occurs
pre-maturely at Vs ¼ 44:74 V. This deviation is mostly an outcome of a basic difference in assumptions. The entire steady state
analysis was based on the idea that there was no rotor acceleration. On the other hand, for the simulation models, the jumps
occur earlier due to sudden rotor acceleration caused by step input voltages. To resolve this incongruity, quasi-static dynamics
simulations are carried out with infinitesimal voltage increments in place of sudden, step voltages; the idea is to noticeably
slow down the rotor acceleration.
To obtain the jump at the first backward critical speed more precisely, an initial voltage supply of 25 V is given for 50 s.
Thereafter, the voltage supply is increased uniformly at a slow rate of 0.0105 V/s for a net duration of 1000 s. The jump occurs
exactly at 412 s, and the corresponding voltage is obtained as 28:81 V , as shown in Fig. 20. Similarly, in Fig. 21, the quasi-static
simulation response for the first forward critical speed is shown. Initial step input voltage input is 40 V, which is supplied for

Fig. 16. Transient response for 1st critical speed during coasting up before the jump for constant Vs ¼ 28:8 V.
344 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

Fig. 17. Transient response for 1st critical speed during coasting up after the jump for constant Vs ¼ 28:81 V.

Fig. 18. Transient response for 2nd critical speed during coasting up before the jump for constant Vs ¼ 44:73 V.

the first 50 s. Thereafter, voltage is incremented at 0.0105 V/s for the next 1000 s. The jump occurs at 509 s, and the voltage at
the instant of jump is then calculated as 44:83 V; which shows clear compliance with the steady state analytical predictions.
Note that with same rotor mass, it is possible to increase the rotor’s polar moment of inertia. As the polar moment of inertia
increases, the rotor acceleration decreases and the results tend to steady state analysis results. For small rotor inertia, the rotor
angular acceleration is high and the results deviate from steady state predictions. Moreover, for small rotary inertia, the
influence of the neglected armature inductance can be significant.

Fig. 19. Transient response for 2nd critical speed during coasting up after the jump for constant Vs ¼ 44:74 V.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 345

Fig. 20. Quasi-static response simulation result for 1st critical speed for constant Vs gradually increased from 25 V.

It must be cautioned that for quasi-static operations, the rotor acceleration is greatly minimised. Hence, supply power for
the system is vastly insufficient to transit through the critical or resonant regime. As a result, the rotor system may be
damaged due to persistent high vibration amplitudes that are encountered as the system is operated around resonance for
long periods. Therefore, quasi-static operations are by no means an advisable option for practical implementations. The
steady state analysis gives the worst case scenario which may be considered during the design phase without any need for any
additional factor of safety in motor sizing.
Overall, the Sommerfeld effect characterization assists careful planning of rotor operations above the critical speed regime
which avoid damage to the costly instruments and set-up. For example, to operate the present rotor at a desired spin speed of
52 rad/s, one needs to first go for coast-up operation and apply full power with Vs  28:81 V, so as to make a quick get away
from the 1st resonance. This is justified because the full power of motor is required to operate the rotor in a stable condition
around the supercritical regime. Effectively, this should stabilize the rotor speed around u  57:07 rad/s (refer point ‘b’ in
Fig. 8). Afterwards, one should opt for a coast-down operation and reduce the voltage supply to 26.58 V so as to stably acquire
the desired operating speed.

4. Additional case e A back to back jump scenario

One additional case will now be investigated for the same system. The idea is to examine a possible scenario, where the
rotor gets stuck at both critical speeds back to back, sequentially. In other words, the rotor speed jumps from the first
backward critical speed and gets caught immediately at the first forward critical speed.
To explore this scenario, the same non-ideal rotor-disc system is re-analysed, with some changes in the system param-
eters. The system is now re-configured with rotor mass m ¼ 25 kg and the bearing stiffness in yedirection Ky ¼ 30000 N/m;
while the other parameter values in Tables 1 and 2 remain intact. The same steady state analysis procedures from Sec 2.3 are
then repeated. Accordingly, the first backward and the first forward critical speeds of the shaft are now obtained as

Fig. 21. Quasi-static response simulation result for 2nd critical speed for constant Vs gradually increased from 40 V.
346 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

Fig. 22. Normalized spin speed ðu*1 Þ Vs. Voltage supply ðVs Þ.

ucr1 ¼ 37:86 rad/s and ucr2 ¼ 46:45 rad/s, respectively. The new spin speed is re-normalized as u* ¼ ðu=ucr1 Þ. The amplitude
parameters are also normalized as per the previous definitions. A new motor speed vs voltage plot is produced in Fig. 22. Once
again, two distinct jumps are observed - one for the first backward whirl mode at u*1 ¼ 1 and the other at the first forward
whirl mode at u*1 ¼ 1:25. Clearly, the two jumps are quite close to each other, which result in back-to-back resonance capture.
The dotted lines in Fig. 22 show the ‘no-unbalance’ condition. For this, power dissipation due to foundation damping and disc
external damping are wholly neglected. Accordingly, the dissipative work done is only due to the bearing resistance and the
no-unbalance speed voltage relation becomes linear, i.e.
 
Vs ¼ Rb Rm þ m2m u=mm (14)

In the coast up operation, the 1st jump occurs at point ‘a’ (i.e., Vs ¼ 24:58 V). This jump corresponds to the passage through
the 1st resonance and should have ideally positioned the spin speed at point ‘c’, which lies on the dotted line. However, in
reality, the speed jumps from ‘a’ and is subsequently arrested at point ‘b’ because the system encounters the 2nd resonance
zone much earlier, which prevents transition to point ‘c’. Once this happens, the rotor speed is stuck at the 2nd critical speed
up to point ‘d’ corresponding to Vs ¼ 31:89 V; the position where the 2nd jump takes place. After the second jump from ‘d’ to
‘e’, the speed-voltage relation nearly follows linear trend given in Eq. (14).
If the resonance zones weren’t positioned one after the other, the 1st jump would have been fairly straightforward, from ‘a’
to directly ‘c’. Instead, due to the back-to-back jump segments, the speed trajectory traces the path from ‘a’ to ‘b’, then ‘b’ to ‘d’,
and finally from ‘e’ and beyond.

Fig. 23. Transient response simulation result during coasting up for constant Vs ¼ 24:1 V.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 347

Fig. 24. Transient response simulation result during coasting up for constant Vs ¼ 31:1 V.

Fig. 23 shows the transient response of the modified system at Vs ¼ 24:1 V. This is where the first jump is numerically
obtained. Note that this is slightly less than the theoretical jump voltage, Vs ¼ 24:58 V (refer point ‘a’ in Fig. 22). At first, the
rotor speed increases steadily when it encounters the 1st or 1BW critical speed (at u*1 ¼ 1) for a small time duration, with
corresponding high amplitudes of x* and f*y . Thereafter, the rotor speed jumps from the 1st resonance with a simultaneous
reduction in the amplitudes of x* and f*y . But instead of making a complete getaway from the resonating conditions, the rotor
speed is immediately stuck again, now at the second critical speed, i.e. at u*1 ¼ 1:25. This is the 2nd resonance zone, where
corresponding amplitudes y* and f*x attain high values.
Likewise, transient response at Vs ¼ 31:1 V is shown in Fig. 24. This is where the 2nd jump is numerically obtained, after
the system escapes from the second (1FW) critical speed.
For such consecutive jump segments, it is pertinent that the system should accelerate through all the resonance condi-
tions, without getting stuck anywhere in between. As such, the power requirement conditions for smooth passage through
resonance need to be carefully worked out. However, the calculated threshold voltage to do this would largely depend on the
particular rotor system parameters.

5. Conclusions

In this paper, multi-Sommerfeld effect is investigated for a rigid shaft with anisotropic flexible supports and carrying an
eccentric disc at an offset position. The considered rotor dynamic system is a five degrees-of-freedom system. At first, the
Campbell diagram is obtained through modal analysis via state space formulations by disregarding the rigid-body rotor spin
mode. The first two modes are cylindrical, having one backward and one forward whirl mode. The third and fourth modes are
conical. Critical speeds exist for the first three modes.
The rotor-disc system is then driven by a non-ideal DC motor. Around resonance, if the supply power from motor is
inadequate then the source power is used to excite the flexural vibration modes rather than to increase the rotor spin speed.
As a consequence, the spin speed may get caught at the various critical speeds for considerable range of input power and
exhibit typical jump phenomena due to Sommerfeld effect.
Due to anisotropic supports, it is shown that the backward whirl mode of the rotor gets excited. Sommerfeld effect is
observed at the first forward and first backward whirl critical speeds. When these two critical speeds are far apart, the
Sommerfeld effect at the backward whirl critical speed is less severe than that at the forward critical speed. Note that severity
is qualitatively defined here as the range of input voltage for which there can be capture at resonance. However, when the
forward and backward critical speeds are close to each other, the severities of Sommerfeld effects at each become comparable.
In addition, escape from resonance at lower critical speed may lead to capture at the other resonance at higher critical speed.
The steady-state analysis shows the existence of stable and unstable operating speeds at certain range of input voltages.
The transition from one stable operating speed to the other, as the input power is varied, is validated through transient
dynamics simulations. The operating speed may as well transit from one stable solution to the other upon external pertur-
bation, such as an impact. When the first forward and first backward critical speeds are far apart, there are two distinct ranges
of input voltages in which two stable operating speeds and one unstable operating speed exist. However, when these two
critical speeds are too close, there can be a range of input voltage for which four stable operating speeds and two unstable
operating speeds co-exist. For example, in the present system, this happens when the parameter values given in Section 4 are
used and the value of horizontal bearing support resistance (Rx ) is reduced further. Such a critical case may lead to chaotic
response near the resonance regime. This can be an interesting area for future research.
348 S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349

Data statement

All the data used in this article are provided with the manuscript (Tables 1 and 2). The following additional information
may be useful to reproduce the results: the solver used during numerical simulations is Runge-Kutta 5th order, with a step
size of 0.0015 and relative error tolerance of 106 .

Acknowledgement

The authors thank Prof. G. Chakraborty, Department of Mechanical Engineering, IIT Kharagpur for his support during this
research. This work is not funded from any source. The authors declare no conflict of interest.

References

[1] J.W. Lund, F.K. Orcutt, Calculations and experiments on the unbalance response of a flexible rotor, J. Eng. Ind. Trans. ASME 89 (1967) 785e796.
[2] R.G. Kirk, E.J. Gunter, The effect of support flexibility and damping on the synchronous response of a single-mass flexible rotor, J. Eng. Ind. Trans. ASME
94 (1972) 221e232.
[3] L.E. Barrett, E.J. Gunter, P.E. Allaire, Optimum bearing and support damping for unbalance response and stability of rotating machinery, J. Eng. Power,
Trans. ASME 100 (1978) 89e94.
[4] E.J. Gunter, Dynamic Stability of Rotor-bearing Systems, NASA SP-113, 1966.
[5] J.M. Vance, Rotordynamics of Turbomachinery, John Wiley, 1988.
[6] G. Genta, Dynamics of Rotating Systems, Springer Science & Business Media, 2005.
[7] A. Muszynska, Forward and backward precession of a vertical anisotropically supported rotor, J. Sound Vib. 192 (1) (1996) 207e222.
[8] Lyn M. Greenhill, A. Guillermo, Cornejo.: critical speeds resulting from unbalance excitation of backward whirl modes, in: Design Engineering
Technical Conferences, vol. 3, ASME, 1995 (3).
[9] J.A. Vazquez, L.E. Barrett, R.D. Flak, Flexible bearing supports, using experimental data, J. Eng. Gas Turbines Power, Trans. ASME 124 (2002).
[10] E.J. Gunter, The influence of internal friction on the stability of high speed rotors, J. Eng. Ind., Trans. ASME, Series B 89 (4) (1967) 683e688.
[11] A. Tondl, Some Problems of Rotor Dynamics, Chapman and Hall, London, England, 1965.
[12] M.F. Dimentberg, Flexural Vibrations of Rotating Shafts, Butterworths, London, England, 1961.
[13] J.W. Lund, The stability of an elastic rotor in journal bearings with flexible, damped supports, J. Appl. Mech. 32 (4) (1965) 911e920. Trans. ASME, Series
E, 87.
[14] M.F. Dimentberg, L. McGovern, R.L. Norton, J. Chapdelaine, R. Harrison, Dynamics of an unbalanced shaft interacting with a limited power supply,
Nonlinear Dynam. 13 (1996) 171e187.
[15] A.K. Samantaray, S.S. Dasgupta, R. Bhattacharyya, Sommerfeld effect in rotationally symmetric planar dynamical systems, Int. J. Eng. Sci. 48 (1) (2010)
21e36.
[16] A. Ryzhik, T. Amer, H. Duckstein, L. Sperling, Zum Sommerfeld effect beim selbstt€ atigen Auswuchten in einer Ebene, Tech. Mech. 21 (4) (2001)
297e312.
[17] D.D. Quinn, Resonant dynamics in a rotordynamic system with nonlinear inertial coupling and shaft anisotropy, Nonlinear Dynam. 57 (4) (2009)
623e633.
[18] I.I. Blekhman, Vibrational Mechanics: Nonlinear Dynamic Effects. General Approach, Applications, World Scientific, Singapore, 2000.
[19] I.I. Bleckman, Self-Synchronization of certain vibratory devices, Eng. Trans. 16 (1953).
[20] R.M. Evan-Iwanowski, Resonance Oscillators in Mechanical Systems, Elsevier, 1976.
[21] M.F. Dimentberg, Statistical Dynamics of Nonlinear and Time Varying Systems, John Wiley and Sons, 1988.
[22] A. Sommerfeld, Beitr€ age Zum Dynamischen Ausbau der Festigkeitslehe, Phys. Z. 3 (1902) 266e286.
[23] A. Nayfeh, D. Mook, Nonlinear Oscillations, Wiley-Interscience, NY, 1979.
[24] V.O. Kononenko, Vibrating Systems with Limited Excitation [in Russian], Nauka, Moscow, 1964.
[25] J.M. Balthazar, M.L. Rente, D.T. Mook, H.I. Weber, Some observations on numerical simulations of a non-ideal dynamical system, in: J.M. Balthazar, D.T.
Mook, J.M. Rosario (Eds.), Nonlinear Dynamics, Chaos, Control and Their Applications to Engineering Sciences, vol. 1, 1997, pp. 97e104.
[26] H.I. Weber, A. Fenili, D. Belato, M.C. de Mattos, S. Wieczorek, On vibrating systems with a limited power supply and their applications to engineering
sciences, in: C.S. Honig (Ed.), 49th Brazilian Seminar of Mathematical Analysis, State University of Campinas, Campinas, SP, Brazil, 1999, pp. 137e277.
[27] J.M. Balthazar, D.T. Mook, R.M.L.R.F. Brasil, H.I. Weber, A. Fenili, D. Belato, J.L.P. Felix, Recent results on vibrating problems with limited power supply,
in: J. Awrejcewicz, J. Brabski, J. Nowakowski (Eds.), Sixth Conference on Dynamical Systems Theory and Applications, Lodz, Poland, 2001, pp. 27e50.
[28] J.M. Balthazar, D.T. Mook, R.M.L.R.F. Brasil, A. Fenili, D. Belato, J.L.P. Felix, H.I. Weber, Recent results on vibrating problems with limited power supply,
Meccanica 330 (7) (2002) 1e9.
[29] J.M. Balthazar, R.M.L.R.F. Brasil, F.J. Garzeri, On non-ideal simple Portal Frame structural model: experimental results under a non-ideal excitation,
Appl. Mech. Mater. 1 (2) (2004) 51e58.
[30] P.J.P. Gonçalves, M. Silveira, B.R. Pontes Junior, J.M. Balthazar, The dynamic behavior of a cantilever beam coupled to a non-ideal unbalanced motor
through numerical and experimental analysis, J. Sound Vib. 333 (20) (2014) 5115e5129.
[31] I. Blekhman, E. Kremer, Vibrational resistance to vehicle motion due to road unevenness, J. Sound Vib. 405 (2017) 306e313.
[32] A.A. Alifov, K.V. Frolov, Interaction of Non-linear Oscillatory Systems with Energy Sources, Taylor & Francis, London, 1990.
[33] J.M. Balthazar, D.T. Mook, H.I. Weber, R.M.L.R.F. Brasil, A. Fenili, D. Belato, J.L.P. Felix, An overview on non-ideal vibrations, Meccanica 38 (2003)
613e621.
[34] R. Bhattacharyya, A. Mukherjee, A.K. Samantaray, Harmonic oscillations of non-conservative, asymmetric, two-degree-of-freedom systems, J. Sound
Vib. 264 (2004) 973e980.
[35] A.K. Samantaray, Steady-state dynamics of a non-ideal rotor with internal damping and gyroscopic effects, Nonlinear Dynam. 56 (4) (2009) 443e451.
[36] A. Bisoi, A.K. Samantaray, R. Bhattacharyya, Control strategies for DC motors driving rotor dynamic systems through resonance, J. Sound Vib. 411
(2017) 304e327.
[37] A. Bisoi, A.K. Samantaray, R. Bhattacharyya, Sommerfeld effect in a two-disk rotor dynamic system at various unbalance conditions, Meccanica 53
(4e5) (2018) 681e701.
[38] X. Kong, B. Wen, Composite synchronization of a four eccentric rotors driven vibration system with a mass-spring rigid base, J. Sound Vib. 427 (2018)
63e81.
[39] A. Sinha, S.K. Bharti, A.K. Samantaray, G. Chakraborty, R. Bhattacharyya, Sommerfeld effect in an oscillator with a reciprocating mass, Nonlinear
Dynam. 93 (2018) 1719e1739.
[40] R. Lima, R. Sampaio, Two parametric excited nonlinear systems due to electromechanical coupling, J. Braz. Soc. Mech. Sci. Eng. 38 (2016) 931e943.
[41] A.K. Samantaray, A note on internal damping induced self-excited vibration in a rotor by considering source loading of a DC motor drive, Int. J.
Nonlinear Mech. 43 (9) (2008) 1012e1017.
S.K. Bharti et al. / Journal of Sound and Vibration 442 (2019) 330e349 349

[42] L. Cveticanin, M. Zukovic, D. Cveticanin, Two degree-of-freedom oscillator coupled to a non-ideal source, Int. J. Non Lin. Mech. 94 (2017) 125e133.
[43] M. Karthikeyan, A. Bisoi, A.K. Samantaray, R. Bhattacharyya, Sommerfeld effect characterisation in rotors with non-ideal drive from ideal drive
response and power balance, Mech. Mach. Theor. 91 (2015) 269e288.
[44] A. Bisoi, A.K. Samantaray, R. Bhattacharyya, Sommerfeld Effect in a Gyroscopic Overhung Rotor-disk System, vol.88, 2018, pp. 1565e1585 (3).
[45] D.C. Karnopp, D.L. Margolis, R.C. Rosenberg, System Dynamics: Modelling and Simulation of Mechatronic Systems, John Wiley & Sons Inc., 2006.
[46] W. Borutzky, Bond Graph Methodology Development and Analysis of Multidisciplinary Dynamic System Models, Springer, 2010.
[47] C. Mishra, A.K. Samantaray, G. Chakraborty, Bond graph modeling and experimental verification of a novel scheme for fault diagnosis of rolling
element bearings in special operating conditions, J. Sound Vib. 377 (2016) 302e330.
[48] C. Mishra, A.K. Samantaray, G. Chakraborty, Ball bearing defect models: a study of simulated and experimental fault signatures, J. Sound Vib. 400
(2017) 86e112.
[49] A.K. Jain, V. Rastogi, A.K. Agrawal, A novel approach to study effects of asymmetric stiffness on parametric instabilities of multi-rotor-system, J. Sound
Vib. 413 (2018) 159e172.
[50] R. Merzouki, A.K. Samantaray, P.M. Pathak, B. Ould-Bouamama, Intelligent Mechatronic Systems: Modeling, Control and Diagnosis, Springer, 2012.
[51] A. Mukherjee, R. Karmakar, A.K. Samantaray, Bond Graph in Modeling, Simulation and Fault Identification, CRC Press, FL, 2012.
[52] S.S. Dasgupta, A.K. Samantaray, R. Bhattacharyya, Stability of an internally damped non-ideal flexible spinning shaft, Int. J. Non-linear Mech. 45 (3)
(2010) 286e293.
[53] A. Mukherjee, A.K. Samantaray, SYMBOLS-shakti User’s Manual, HighTech Consultants. STEP Indian Institute of Technology, Kharagpur, 2006.

You might also like