You are on page 1of 22

Mechanical Systems and Signal Processing 133 (2019) 106271

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Effects of static and imbalance loads on nonlinear response of


rigid rotor supported on gas foil bearings
Zhiyang Guo a,b, Yuanlong Cao b, Kai Feng b,⇑, Hanqing Guan b, Tao Zhang b
a
College of Petroleum and Engineering, Yangtze University, Wuhan 430072, PR China
b
The State Key Laboratory of Advanced Design and Manufacturing for Vehicle Body, Hunan University, Changsha 410082, PR China

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

Article history: This paper presents an experimental and numerical investigation into the effects of static
Received 27 February 2018 and imbalance loads on the nonlinear rotordynamic performance of a rigid rotor supported
Received in revised form 28 April 2019 on gas foil bearings (GFBs). The interesting phenomenon that static loads can lead to high
Accepted 22 July 2019
sub-synchronous motions in rotor-GFB systems is reported for the first time in this paper.
The effects of static and imbalance loads on coast-down tests, synchronous response
amplitudes, peak response frequency, appearance of sub-synchronous vibrations and time
Keywords:
extent of coast-down responses are measured and discussed. A rotordynamic model based
Gas foil bearings
Nonlinear rotordynamics
on FEM and unsteady Reynolds equation is derived to predict the nonlinear response of the
Static and imbalance loads system. The foil structure and gas film in-series are considered to derive the unsteady
Sub-synchronous vibrations bearing force. The predictions are consistent with the test data. The relationships between
static or imbalance loads and nonlinear response, especially the vibrations in sub-
synchronous frequencies, are discussed. In the numerical simulations, as the static and
imbalance loads increase, the vibration component changes from mainly synchronous to
sub-synchronous and then back to synchronous vibrations. When the static and imbalance
loads change, the natural frequency of the rotor-GFB system could be shifted into or away
from the threshold of the whirl vibration with a typical factor, k = 0.5, and lead to serious
sub-synchronous vibrations. The experimental data and theoretical predictions help with
the engineering design and implementation of GFBs in high-performance micro
turbomachineries.
Ó 2019 Published by Elsevier Ltd.

1. Introduction

Gas foil bearings (GFBs) are widely used in many high-performance compact rotor systems, such as turbojet engines [1],
high-speed oil-free blowers [2,3] and highly efficient micro gas turbines [4], due to their high operational speed, low power
losses and wide range of temperature tolerance [5–7]. However, turbomachineries supported on GFBs may experience extra
static and imbalance loads, especially when they are applied on board transportation equipment, such as air cycling machi-
nes in airplanes [7] and air blowers in fuel cell electrical vehicles [2,8]. Extra static loads, which are due to the stable accel-
eration/deceleration of a foundation, and imbalance loads, which are due to imbalance mass, are quite common and may
cause stability problems and even system failure in their applications [9]. Hence, an analysis of the effects of static and
imbalance loads on rotordynamic response is important for the design of rotor-GFB system.

⇑ Corresponding author.
E-mail address: jkai.feng@gmail.com (K. Feng).

https://doi.org/10.1016/j.ymssp.2019.106271
0888-3270/Ó 2019 Published by Elsevier Ltd.
2 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Nomenclature

GFBs gas foil bearing


Ip polar mass moment of inertia
x; u_ rotor speed
u rotor displacement
C GFBs ; C TBs rotational drag coefficients of the two GFBs/two thrust GFBs
t time
T Turbine drag torque of the turbine
C const constant number
M mass matrix
B damping matrix
G gyroscopic matrix
K stiffness matrix
q system displacement vector
F ub imbalance force vector
F GFB nonlinear force vector
Fg grave force vector
F Static static load vector
m imbalance mass
e radius of imbalance mass
s time step
P dimensionless pressure
Z dimensionless bearing axial width
H dimensionless gas film thickness
K bearing number
m excitation frequency ratio
se dimensionless time
h circumferential coordinate
pa ambient pressure
R bearing radius
C radial nominal clearance
l viscosity of gas
xe the excitation frequency
h gas film thickness
ex ; ey journal eccentricity raito in x and y direction
d deflection of the foil structure
U dimensionless deflection of the foil structure
kb bump stiffness
Kb normalized bump stiffness
Cb equivalent viscous damping
A0 bump effective covering area
g structural loss factor
K sys stiffness of the rotor-GFB system
K GFB stiffness of the GFB
K shaft stiffness of the shaft
K Gas stiffness of the gas film
K Foil stiffness of the foil structure
Xth threshold speed of sub-synchronous vibrations
k fluid circumferential average velocity ratio
M sys mass of the rotor-GFB system

A GFB comprises a flexible top and bump foil structures, as shown in Fig. 1. Given that the pressured gas film is compress-
ible, the nonlinear gas film force is sensitive to the load conditions in GFBs. During operation, the compressed gas film and
foil structure work in series to support the shaft. The supporting force of each bump structure is simplified into a spring and a
damper [10,11], as shown in Fig. 2. Under the top foil, bumps in the foil structure work in parallel to provide supporting
force. Hence, bump deformation heavily depends on the pressure distribution of the gas film. When the bearing load is small
and the region of the generated gas pressure is narrow in the circumferential direction, few bumps work in parallel and
result in low effective foil structural stiffness. The region of gas load widens in the circumferential direction and more bumps
contribute to foil structural stiffness with the increase of the bearing load. Thus, the foil structure shows a typical nonlinear
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 3

Fig. 1. Structure of a typical bump-type GFB.

Fig. 2. Schematic view of light and heavy gas loads on bump foil structure.

load hardening characteristic [12,13]. These factors make the bearing stiffness, which is determined by the gas film and foil
structure in series, nonlinear and sensitive to static and imbalance load conditions. Therefore, understanding the relationship
between the nonlinear performance of the rotor-GFB system and the static and imbalance loads is necessary.
To the best of our knowledge, no experimental study on the influence of static loads on the nonlinear rotordynamic
response of rotor-GFB systems has been reported. Kim and San Andrés [9] presented a simple model to predict the dynamic
force characteristics of GFBs with heavy static loads. As the static load imposed on the foil bearing increases, the predicted
frequency-dependent direct dynamic stiffness coefficients increase steadily as derived. Structural stiffness is an upper limit
factor of the direct dynamic stiffness coefficients, which agree with the model of the gas film and foil structure in series, as
shown in Fig. 2. However, the effects of static loads on the nonlinear response of the rotor-GFB system were not reported.
Heshmat [14] introduced an advanced design of GFBs which has a single top foil and multistage bump foil structures. The
GFBs achieved a breakthrough in the load performance of 673.5 kPa. Under heavy static load conditions, the nonlinear
response of the rotor-GFB system showed a major resonant vibration with high amplitudes. The major vibration had sub-
synchronous frequencies and was related to one of the predicted natural modes under the linear model. Kim et al.
[15,16] built hybrid GFBs utilizing hydrostatic injection of externally pressurized air to overcome system instability. Impulse
external loads were added to the rotor system, and high sub-synchronous motions were recorded in many different frequen-
cies. The fact that heavy static loads could lead to the nonlinear response of the rotor-GFB system was illustrated; however,
the source of the nonlinear vibration was not explored.
Meanwhile, imbalance loads could induce a high nonlinear performance of the rotor-GFB system. San Andrés et al. [17]
conducted a series of coast-down experiments with different imbalance masses and found that increasing imbalance loads
could introduce sub-synchronous motions into the rotor-GFB system in addition to the nonlinear synchronous vibrations.
The instability and forced nonlinearity of the rotor-GFB system because of imbalance loads were discussed by San Andrés
4 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

and Kim [18]. A dynamic model [19,20] was used to investigate the nonlinear response of the rotor system, assuming a rigid
gas film in GFBs. In this model, the results of the static load-deflection test were used to derive the nonlinear stiffness of the
foil bearing with a structural loss factor. However, the nonlinearity of the compressible gas film in GFBs was not considered.
The influence of the foil structure on the nonlinear response of the rotor-GFB system was overestimated because the gas film
and foil structure worked in series (Fig. 2). Balducchi et al. [21] also presented the experimental imbalance response of two
different rigid rotors supported by two identical GFBs. High nonlinear responses due to different unbalance masses of the
rigid rotor supported on GFBs were recorded and discussed. The nonlinear response of rotor-GFBs were estimated with
the rigid gas film model in Ref. [18] and compared with experimental data. However, the prediction results were insufficient
to predict the different nonlinear responses related to imbalance loads as introduced in the paper. Bonello and Hassan [22],
Hoffmann and Liebich [23] also conducted experiments to investigate the effects of imbalance loads on rotor-GFB system.
High sub-synchronous vibrations were recorded and trend to be treated as bifurcation in these researches. Large imbalance
loads bring high synchronised motions of the rotor supported on the GFBs and high-level transient centrifugal force to the
rotor. The gas film in series with the foil structure is loaded and released repeatedly in a large displacement. Thus, the
dynamic stiffness of GFBs varies in a wide range and results in the overall nonlinear performance of GFBs and the rotor
system.
In this study, a test rig with adjustable static and imbalance loads on the rotor is designed to investigate the effects of
static and imbalance loads on the nonlinear response of a rotor-GFB system. Coast-down tests of the rotor-GFB system with
different static loads and imbalance displacements are conducted. Factors such as peak response speeds, synchronous
response amplitudes, onset speeds of sub-synchronous vibrations and time extents of the coast-down process are measured
and discussed to evaluate the influence of static and imbalance loads on the nonlinear performance of the rotor-GFB system.
A rotordynamic model based on FEM and transient Reynolds equation is presented. The predictions are compared with the
corresponding test data. Beyond the experimental limits, further nonlinear phenomena derived from the predictive model
are also discussed in this study. The discussions of test results and predictions help in the engineering design and implemen-
tation of GFBs in high-performance micro turbomachineries.

2. Effects of static load on the nonlinear response of rotor-GFB systems

2.1. Experimental facility

Fig. 3 shows the schematic view of the test rig to investigate the nonlinear response of the rigid rotor system with extra
static and imbalance loads. A steel rotor with a mass of 1.440 kg and a length of 252.0 mm is supported on two GFBs and
driven by a double-impulse turbine. To ensure the loads on the two bearings are the same, that is, half of the rotor weight,

Fig. 3. Schematic view of the rotor-GFB system with the facilities of electromagnet and load cell installed.
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 5

Fig. 4. Data of static load-deflection test conducted with the electromagnetic force.

the mass centre of the rotor is designed to locate at the middle of the two GFBs. A gas source with a maximum air pressure of
0.75 MPa and an air flow of 11.8 m3/min is used to drive the turbine and the rotor. The maximum speed can reach up to
60.0 krpm. A pair of thrust GFBs is also designed to suppress the axial vibration. Four eddy current sensors are fixed on
the housing to record the vibration of the rotor-GFB system at both sides. One custom-built electromagnet is mounted at
the mass centre of the rotor to apply extra static load on the rotor. The static load is measured by a load cell connected
to the electromagnet. The electromagnetic force is controlled by a computer system and can remain stable from 0 to
60.0 N with different power voltages and distances from the rotor surface. All test data are recorded and monitored by a data
acquisition system based on NIÒ. A commercial laser tachometer is used to pick up the speed and keyphaser signals. The
rotor is balanced into imbalance displacements of 2.50 mm in phase by a commercial service before it is installed into the
bearing. The basic physical parameters of the rotor system are listed in the Appendix, Table 1.
Static load-deflection tests are conducted to ensure that the extra electromagnetic force is equally applied on both GFBs.
As shown in Fig. 4, the displacements measured from the corresponding probes are quite near the increase of the static load
on the rotor. The nominal clearances of GFBs can also be estimated as 35.0 mm from the test data. The rotor weight is
included into the electromagnetic load. After the tests, additional measurements of the nominal clearances of the two GFBs
are taken again using a traditional method introduced in Ref. [24,25]. This repeat is to ensure that the nominal clearances are
estimated carefully by the electromagnet. The test rigs and results are listed in the Appendix.

2.2. Experiments with different static loads

Coast-down tests of the rotor-GFB system are conducted for various static loads (21.6 N, 28.8 N and 36.0 N), which are 0.5,
1.0 and 1.5 times the rotor weight along the vertical direction (with gravity subtracted). Other conditions are the same for
the three tests. During the tests, the rotor-GFB system is accelerated rapidly above 40.0 krpm by the pressured air and then
coasts down by cutting off the compressed air. The displacements and tachometer signals are recorded during the tests.
Fig. 5 provides the waterfall plots of the coast-down response of the rotor-GFB system in the horizontal direction at the
turbine side with different static load conditions. The synchronous motions clearly change with the increase of the static load
applied on the rotor. Importantly, the sub-synchronous vibration occurs when the static load increases to 36.0 N. It should be
noticed that the magnitude of imbalance displacement is only 2.5 mm, which is at the same level compared with the man-
ufacturing tolerance of the shaft surface. The runout of the shaft surface would significantly affect the direct probe data in
waterfall plots. Because the directions of vibration vectors of runout and actual response are quite different, the synchronous
amplitudes with runout data is a little lower than 2.5 mm in waterfall plots. Fig. 6 depicts the filtered synchronous ampli-
tudes [0-pk, mm] of the rotor motion with different static loads. The error introduced by the runout response should be
removed from direct probe data in the form of vibration vectors [26]. As the lift-off speed of the rotor-GFB system is about
5.0–6.0 krpm based on the authors’ experience. The responses are compensated with the low speed response at 6.0 krpm to
provide a better view of the synchronous motions. The rotational speed of the synchronous peak response increases with the
static load. The responses show distinctive peak rotor speeds at 8870 rpm, 9650 rpm and 11,080 rpm for the three static load
conditions, and the corresponding amplitudes are 7.52, 9.86 and 9.73 [0-pk, mm], respectively. The synchronous responses
for these three tests are nearly the same in amplitude, about 2.50 [0-pk, mm], when the speed is higher than 25.0 krpm as
the imbalance conditions are the same. Fig. 7 shows the filtered amplitudes of the sub-synchronous motions and the corre-
sponding whirl frequency. The sub-synchronous whirl vibrations with a maximum amplitude of 14.83 [0-pk, mm] onset at a
static load of 36.0 N. The main sub-synchronous motions range from 18.0 krpm to 24.6 krpm. The sub-synchronous motions
maintain a nearly constant whirl frequency ratio (WFR), 0.5.
6 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 5. Waterfall plots of the coast-down response of the rotor-GFB system in the horizontal direction at the turbine side with static load: (a) 21.6 N, (b)
28.8 N and (c) 36.0 N in the vertical direction. Direct probe data.
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 7

Fig. 6. Synchronous coast-down response of the rotor-GFB system in the horizontal direction at the turbine side with static load, 21.6 N, 28.8 N and 36.0 N,
in the vertical direction.

Fig. 7. Sub-synchronous coast-down response of the rotor-GFB system in the horizontal direction at the turbine side with static load 36.0 N in the vertical
direction.

Fig. 8 shows the time extent for coast-down rotor response with different static load conditions. The driving gas is cut off
as the rotor speed reaches x0 ¼ 40:0 krpm. The drag friction force of the rotor has three sources: two GFBs, two thrust GFBs,
and the stirred air by the turbine disc. The speed dynamic function can be derived as
dx
Ip þ ðC GFBs þ C TBs Þx þ T Turbine ¼ 0 ð1Þ
dt
where Ip is the polar mass moment of inertia of the rotor; C GFBs and C TBs are the rotational drag coefficients for the two GFBs
and two thrust GFBs, respectively; T Turbine represents the drag torque from the windage of the turbine; x is the transient rotor
speed; and x0 represents the recorded maximum rotor speed, 40.0 krpm. If the windage of the turbine is ignored, then the
deceleration equation can be expressed as
dx
Ip þ ðC GFBs þ C TBs Þx ¼ 0 ð2Þ
dt

C GFBs þ C TBs
lnx ¼  t þ C const ð3Þ
Ip
where C const is a constant number. Eq. (3) shows that the rotor speed ratio from 40.0 krpm to rest should be one linear equa-
tion. However, the test data are consistent with the linear model when the test rotor speed exceeds 7.0 krpm, or 0.175 in the
rotor speed ratio, as shown in Fig. 8. The gas film still works when the extra static load is added. When the rotor speed is
lower than 7.0 krpm, the dry friction of the thrust GFBs is introduced into the rotor-bearing system because of the low rota-
tional speed. Thus, the line is not straight. The deceleration times from 40.0 krpm to 7.0 krpm are 37.0 sec, 30.5 sec and
8 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 8. Effect of static loads on time extent for coast-down rotor response.

24.5 sec for the static load conditions of 21.6 N, 28.8 N and 36.0 N, respectively. Obviously, the rotational drag coefficients of
the GFBs are strengthened with the static loads.
One more static load test is conducted at a constant rotor speed of 20.0 krpm. The test is designed for comparison with the
nonlinear prediction model in the following section. The extra static loads on the rotor-GFB system in the vertical direction
are controlled by an electromagnet (Fig. 9). Fig. 10 illustrates the corresponding transient response of the rotor-GFB system
in the horizontal direction at the turbine side. When the static load is small, the rotor response is mainly a synchronous
motion (1x) with small amplitudes. When the static load is increased to more than 32.0–34.0 N, as labelled in Fig. 9, it is
interesting to note that the transient system vibration generates one more component, the sub-synchronous motion
(0.5), and shows larger total amplitudes. No significant maximum temperature increases, 3.0 K at the turbine side and
3.5 K at the thrust side, are recorded by the thermocouples attached at the back of the top foils in each GFB. That is to
say, the gas film works well to support the high-speed shaft. No direct physical contact happened between the shaft and
top foil surfaces when the static load is added. The filtered direct component (DC) or the static displacement is also illus-
trated in Fig. 10. It increases with the related electromagnetism force. Two sections are also labelled in Fig. 10 to identify
the selections of data for comparison with the predictions of the nonlinear model. The filtered orbit is shown in Fig. 16. Extra
static loads can clearly improve the dynamic stiffness of rotor-GFB systems, as introduced in Fig. 2. The sub-synchronous or
so-called whirl motion refers to the excitation of one system natural vibration mode by the shearing force of the gas film.
This outcome is discussed in detail in the sections on the nonlinear predicting model.

Fig. 9. Time versus load on rotor-GFB system in the vertical direction at constant speed of 20 krpm. The pop-up region of sub-synchronous motion is
labelled.
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 9

Fig. 10. Transient response of the rotor-GFB system in the horizontal direction at the turbine side with static load changing in the vertical direction,
20.0 krpm.

3. Effects of imbalance loads on nonlinear response of rotor-GFB systems

3.1. Experiments with different imbalance conditions

To investigate the effects of imbalance loads on the nonlinear response of the rotor-GFB system, three more coast-down
tests are conducted. Meanwhile, the static load conditions are set to 36.0 N in these tests because sub-synchronous motions
occurred in the previous test. It is convenient to discuss the relationships between imbalance loads and the nonlinear
response of the rotor-GFB system.
On the turbine and thrust discs, eight holes, 1.2-mm diameter holes are spaced equally for the insertion of small imbal-
ance masses. The radius from the imbalance mass holes to the disc axial centre is 12.0 mm. Pairs of unbalance masses are
inserted into the turbine and thrust discs as listed in Table 2. Imbalance masses are used to balance the rotor-GFB system
because of evident high peak response during the coast-down test. The imbalance masses are separated equally and dis-
tributed in phase at both sides. The coast-down test under the original imbalance condition, 2.5 mm, is the same one in
the last test of Test A. The corresponding imbalance displacements are 1.7, 1.2 and 0.7 mm for Tests B, C and D, respectively.

3.2. Experimental results

In Fig. 11, the waterfall plots of the coast-down response of the rotor-GFB system in the horizontal direction at the turbine
side with different imbalance conditions are depicted. The corresponding waterfall plots for Test A are illustrated in Fig. 5(c).
The synchronous motions are clear in all tests, and the sub-synchronous motions disappear as the imbalance mass decreases.
As the directions of vibration vectors of runout and actual response are quite different, the runout response would dominate
the direct probe data in the waterfall plots. Thus, as the imbalance loads decrease, the measured amplitudes of synchronous
response in waterfall plots grow higher when the rotor speed is above 25.0 krpm. The actual response amplitudes after low
speed compensation should be checked in Fig. 12. The detailed data of filtered synchronous and sub-synchronous motions
are discussed in the following part.

Table 2
Imbalance conditions for rotor-GFB system under static load of 36.0 N.

Imbalance displacements
Test A Original/2.5 lm in phase
Test B 1.7 lm in phase
Test C 1.2 lm in phase
Test D 0.7 lm in phase
10 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 11. Waterfall plots of the coast-down response of the rotor-GFB system in the horizontal direction at the turbine side with different imbalance
conditions: (a) Test B, (b) Test C and (c) Test D. Direct probe data.

Fig. 12 depicts the recorded synchronous amplitudes [0-pk, mm] of rotor motions obtained with different imbalance con-
ditions. The synchronous responses are also compensated by the low speed response at 6.0 krpm. The rotor speeds of the
synchronous peak response are almost unchanged at 11080 rpm. The synchronous responses for these tests are approxi-
mately 2.50, 1.76, 1.28 and 0.89 [0-pk, mm] at the rotor speed region above 25.0 krpm for Tests A, B, C and D, respectively.
Fig. 13 shows the filtered amplitudes of the sub-synchronous motions and the corresponding WFR in Tests A and B. As the
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 11

Fig. 12. Synchronous coast-down response of the rotor-GFB system in the horizontal direction at the turbine side under the static load of 36.0 N in the
vertical direction. The detailed imbalance conditions are listed in Table 2.

Fig. 13. Sub-synchronous coast-down response of the rotor-GFB system in the horizontal direction at the turbine side under the static load of 36.0 N in the
vertical direction. The detailed imbalance conditions are listed in Table 2.

imbalance loads decrease, the sub-synchronous whirl vibrations disappear. The sub-synchronous motions in Tests A and B
maintain a nearly constant WFR, 0.5. Thus, sub-synchronous motions are still the excitation of the system natural vibration
mode by the shearing force of the gas film.
The effect of imbalance conditions on time extent for coast-down rotor response is illustrated in Fig. 14. The coast-down
of the rotor-GFB system from 40.0 krpm to rest is also approximately linear when the rotor speed exceeds 7.0 krpm. Clearly,
the gas film still works under different imbalance conditions. When the rotor speed is lower than 7.0 krpm, the rotor speed
ratio is not linear due to the thrust GFBs, as discussed previously. The total coast-down times for the four tests are approx-
imately 25.0 sec and are not significantly influenced by the different imbalance conditions.

4. Nonlinear prediction of rotor-GFB systems

4.1. Descriptions of nonlinear prediction model and comparison with test data

To predict the nonlinear response of the rigid rotor supported on GFBs, a rotordynamic model based on FEM and transient
Reynolds equation is presented. The continuous shaft is discretised into 18 beam elements with circular sections, as shown
in Fig. 15. The checking of the operation deflection shape (ODS) [27,28] to avoid the existence of flexible rotor is simple in
this model. The Cartesian coordinate system, imbalance mass location, and eddy current sensor location are also labelled in
the schematic figure. A system matrix group is established by employing Timoshenko beam theory, in which shear effect is
considered [29,30]. The gyroscopic effect related to rotational speed is also considered [31]. The two discs are treated as
lumped masses and moments of inertia at the corresponding nodes at the rotor ends, respectively. Nonlinear bearing forces
12 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 14. Effect of imbalance conditions on time extent for coast-down rotor response.

Fig. 15. FEM model of the rotor-GFB system, 76 DOFs in 18 elements. Labelled locations of probes, imbalance mass and GFBs.

are applied at the centres of each bearing location. The detailed physical and geometrical parameters of the rotor-GFB system
under investigation are listed in Tables 1 and 3.
Following the assumptions and settings described above, the governing equation of system dynamic motion for the
investigated GFB-rotor system is derived as
€ðt Þ þ ½B þ u
Mq _ Gq_ ðt Þ þ KqðtÞ ¼ F ub ðt Þ þ F GFB ðtÞ þ F g ðt Þ þ F Static ðtÞ ð4Þ
h iT
where qi ¼ xi yi hxi h yi is the system displacement vector at each node; xi ; yi and hxi ; hyi represent the lateral and
rotational displacements relative to the fixed reference frame shown in Fig. 15; and M; B; G; K are the mass, damping, gyro-
scopic and stiffness matrices, respectively. Damping matrix B is set to zero, and no additional damping is considered when
the bearing damping effect is contained in the nonlinear bearing force. u _ is the rotor axial rotational speed, F g is the gravity
force vector, F Static is the static load vector added by the electromagnet and F GFB is the nonlinear GFB force vector that is
described in the following section, t is time. F ub ðt Þ is the imbalance force vector of the rotor and is denoted as
   
F i;x cosu
F ub ðt Þi ¼ _2
¼ meu ð5Þ
F i;y sinu

where F i;x and F i;y are the imbalance forces in the horizontal and vertical directions, respectively, while the imbalance
moments are not considered; m is the imbalance mass;e is the imbalance radius; u is the rotor axial rotational displacement.
The Wilson-h scheme is used to solve nonlinear dynamic Eq. (4) in the time domain. Acceleration in assumed linear from
time step s to time step s þ Ds as h P 1:0. The optimal value, h ¼ 1:4208, is adopted in the simulation to ensure the stability
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 13

of prediction [31]. The nonlinear GFB force is evaluated with the back-forward time step, i.e. the unsteady Reynolds equation
is solved with journal displacements and velocities in the previous time step. More details should be referred to Ref. [32].
Compared with traditional rigid gas bearings, GFBs have a flexible supporting foil structure. As shown in Fig. 1, GFBs have
two parts, namely, gas film and foil structure, which work in series. The gas film is solved with an unsteady Reynolds equa-
tion that couples the equations of foil structure deflection and rotor motion. Each bump is modelled using a spring with a
constant bump stiffness and a damper with a foil structural loss factor. The top foil is modelled as a thin plate with low radial
stiffness. Whether the bump of the foil affects the model in parallel, the effect heavily depends on the pressure distribution of
the gas film. This simple model describes the nonlinearity of the foil structure and can help determine the source of sub-
synchronous motions in the rotor-GFB system in the following sections. The nonlinear gas film force works in series with
bump foils through the connecting top foil to support the shaft. For simplification, the bump stiffness, kb , is derived on
the basis of the link-spring model with a free segment assumption [33,34] and set to a constant value. By contrast, the damp-
ing factor is modelled with equivalent viscous damping coefficients [35,36], g. This model overcomes the shortcomings of
the inaccurate GFB transient force predicted with the rigid gas film assumption [20]. At the same time, one simple but effec-
tive nonlinear bump foil structure is considered in this model.
To model the pressured gas film between the shaft surface and the top foil inner surface, an unsteady compressible Rey-
nolds equation with isothermal condition is written as
   
@ @P @ @P @ @
PH3 þ PH3 ¼ K ðPHÞ þ 2Km ðPHÞ ð6Þ
@h @h @Z @Z @h @ se
where P is the dimensionless pressure, Z is the dimensionless bearing axial width, H is the dimensionless gas film thickness,
K is the bearing number, m is the excitation frequency ratio and se is the dimensionless time. More details are listed in the
appendix section. The transient GFB hydrodynamic film forces in the x and y directions are calculated by integrating the
pressure field Pðh; zÞ using Simpson’s one-third rule in the following equations:
( R 2p R L=R
F x ¼ pa R2 0 0 ðPðh; Z Þ  1Þsinhdhdz
R 2p R L=R ð7Þ
F y ¼ pa R2 0 0 ðPðh; Z Þ  1Þcoshdhdz
The data in the two labelled regions shown in Fig. 10 are compared with the results from the above rotordynamic model
in Fig. 16 (a.1–3) and (b.1–3). The simulations are conducted at the rotor speed of 20.0 krpm. The static loads are set to (a)
25.0 N and (b) 36.0 N in the vertical direction for the two groups of predictions. The parameters for the two GFBs and the
rotor system are presented in Tables 2 and 3, and the bump stiffness per unit area is calculated k ¼ 8:48  109 N=m3 with
the structural loss factor set to 0.20. The normal clearance is set to 35.0 mm. The imbalance mass is set to 1.44 kg  1.25 lm
in phase on each disc. Only the last 10,000 stable samples of the calculated orbits are pictured in the figure. In practical sit-
uation, the eddy current sensor is not able to locate the accurate center of the bearings under the current testing capacity of
our laboratory. A zero-phase band-pass filter (100–500 Hz) is applied to the test data to eliminate the disturbance from high
orders and extremely low frequency signals. The results of the predictions are only linear trends removed and compared
with the filtered test data in Fig. 16. Obviously, the first labelled region only has one main synchronous component and
has larger amplitude in the vertical direction. In the first region, the horizontal amplitudes of the test and predicted syn-
chronous vibrations are 0.8 and 1.1 [0-pk, lm], respectively, as shown in Fig. 16(a.2–3). However, in the second labelled
region in Fig. 10, the motions have sub-synchronous and synchronous components with larger amplitudes in the horizontal
direction. The horizontal amplitudes of the test and predicted sub-synchronous vibrations (0.5) are 8.9 and 13.9 [0-pk, lm],
respectively, as shown in Fig. 16(b.2–3). Meanwhile, the corresponding amplitudes of the synchronous motions are only 0.4
and 2.3 [0-pk, lm]. The predicted orbits are consistent with the test data but with slightly larger amplitudes, as shown in
Fig. 16(a.1–3) and (b.1–3). Some structural damping is introduced from the thrust GFBs or some other components. Thrust
GFBs are not considered in the rotordynamic model but are considered as the main reason for the difference between the
data from tests and predictions. In this model, the effect of gas film is interactively considered with the foil structure. Thus,
this rotor-GFB system model is more reasonable and can be used to explore additional information on nonlinear response
from the aspect of numerical calculations.
Bently and Charles [26] provided a simple model to estimate the fluid-induced instability in oil journal bearing. The sub-
synchronous vibration in oil journal bearing is related to the transient natural frequency of rotor-bearing system by a fac-
tor,k, which is the fluid circumferential average velocity ratio. Typically, k is slightly less than 0.5 in oil bearing. It is extended
here to explain the onset of sub-synchronous vibrations in the rotor-GFB system. The two GFBs work in parallel to support
the rotor in series to rotate. In each GFB, the gas film and foil structure work in series to provide the total bearing force. These
relationships are depicted in Eq. (8) to show the system’s dynamic stiffness model. As the rotor is rigid within the test speed
region, the model is simplified into Eq. (9). Thus, the system stiffness is approximately determined by the stiffness of the foil
structure and gas film in series. The sub-synchronous vibration is gas fluid-induced with the coupling of the foil structure
instead of the rotor, as derived in Eq. (10). Under high static load condition, more bumps contribute their stiffness to the foil
structure (Fig. 2). The system’s natural frequency is extended and reaches the threshold speed of the whirl vibration. This
outcome can be verified from the experimental synchronous peak response with an increasing static load, as shown in
Fig. 6. Thus, sub-synchronous vibrations pop up until the static load is added to 36.0 N. As the viscous of gas film is much
lower than oil film, k is 0.5 in the test data and demonstrated in the simulating results.
14 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 16. Comparison of filtered data from test and predictions in turbine side probe node with the static loads set (a.1–3) 25.0 N and (b.1–3) 36.0 N in the
horizontal direction. (a.2–3) and (b.2–3) are the corresponding amplitudes spectra of the horizontal data.

( 1
K sys
¼ 2K1GFB þ K 1
ð8Þ
shaft
1
K GFB
¼ K Gas
1
þ K1
Foil
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 15

 
1 1 1 1
 þ ð9Þ
K sys 2 K Foil K Gas

sffiffiffiffiffiffiffiffiffi
1 K sys
Xth ¼ ð10Þ
k M sys

4.2. Effects of static load on nonlinear response of rotor-GFB system

To investigate the effects of static loads on the performance of the rotor-GFB system, different simulating studies and
discussions are presented in this section. The orbits of the rotor centre and the amplitude spectrum are shown in Fig. 17
to illustrate the cases of predictions with different static loads. Dashed line is the nominal clearance circle of the GFBs.
The total imbalance mass displacement, 2.5 lm, is equally set on each disc in the phase as labelled in the FEM model figure.
A nominal clearance of 35.0 lm, as the result of the static load-deflection test, is set to the two GFBs. The bump stiffness per
unit area is calculated from the physical parameters in Table 1, k ¼ 8:48  109 N=m3 , with a structural loss factor of 0.20

Fig. 17. Orbits (a–d.1) of the rotor centre at the turbine side probe node and corresponding amplitude spectra (a–d.2) in the horizontal direction. The static
loads range from 25.0 N to 54.0 N in the vertical direction.
16 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

based on the authors’ experience. The steady-state rational speed of 20.0 krpm is set. The transient simulation time step is
set to 360 per revolution. Only the last 10,000 stable samples of the calculated orbits are pictured in Fig. 17.
As the static loads on the shaft increases from 25.0 N to 54.0 N in the vertical direction, the average centres of the orbits
also increase in the vertical direction, especially when the results shown in Fig. 17(a.1) and (d.1) are compared. As shown in
Fig. 17(a–d.2), sub-synchronous motions occur at the static load of 36.0 N then disappear at 45.0 N in these cases. The orbits
transfer from elliptical into 8-shape then back to elliptical shape again as a result of the different combinations of sub- and
synchronous components. Large vibrations at sub-synchronous frequencies could introduce resonance of other parts in the
rotor-GFB system and even fail the system. As discussed previously, the increasing static loads produce more bump struc-
tures, which contribute their stiffness into the system. Thus, the natural frequency of the rotor-GFB system increases and
encounters the threshold speed of the whirl vibration. The sub-synchronous motions are excited by the gas film with a typ-
ical factor k ¼ 0:5. As the static load is added up to 54.0 N, the bearing stiffness determined by the in-series stiffness of the
gas film and foil structure increases and pushes the natural frequency of the rotor-GFB system out of the region of the whirl
vibration. Thus, the sub-synchronous motions disappear. It is clear that, the stiffness of the bump structure related to the
large sub-synchronous vibrations should be carefully designed and manufactured. This is important to the manufacturers
of rotor-GFB systems.
Extra static load cases, such as 45.0 and 54.0 N, are not conducted. The gap between the electromagnet and the shaft is too
close to keep the loads stable when the static load is large. Meanwhile, the drag torque grows quickly as the static load
increases. The turbine discs driven by pressured gas cannot offer sufficient torque to maintain the constant speed of
20.0 krpm. These two factors limit further experimental studies, though the predictions show more information than do
the tests with convincing results.

4.3. Effects of imbalance loads on nonlinear response of rotor-GFB systems

To investigate the effects of imbalance loads on the performance of the rotor-GFB system, eight groups of predictions are
carried out and discussed in this section. The static load conditions of the rotor-GFB system are set to 36.0 N. The total imbal-
ance mass displacement ranging from 0.5 lm to 7.5 lm is equally set on each disc in the phase as labelled in the FEM model
figure. The other conditions are the same as the simulations of static loads. The orbits of the rotor centre and the amplitude
spectrum are shown in Fig. 18 to illustrate the effects of different imbalance masses on the nonlinear response of the rotor-
GFB system. Only the last 10,000 stable samples of the calculated orbits are plotted in Fig. 18.
As the total imbalance masses of the system increase from 0.5 lm to 7.5 lm, the amplitudes of synchronous motions
increase from 0.4 to 6.7 [0-pk, lm], as labeled in Fig. 18(a–h.2). The sub-synchronous motions pop up at the total imbalance
displacement of 2.5 lm and disappear at 7.5 lm in these cases. The orbits transfer from elliptical into a complex 8-shape and
then back to elliptical shape again as a result of the different combinations of sub-synchronous and synchronous compo-
nents. This finding means that unsuitable imbalance mass displacements will not only introduce large synchronous vibra-
tions of the shaft but also result in the resonance of other parts in the rotor-GFBs at 0.5x WFR. As the imbalance masses
increase, the transient eccentricity ratios of the shaft center can reach much larger than 1.0, and more bumps contribute
their stiffness into the system. Thus, the transient natural frequency of the rotor-GFB system determined by the gas film
and foil structure in series increases and encounters the threshold speed of the whirl vibration. When the imbalance dis-
placements are large enough, they improve the natural frequency of the system beyond the threshold of the whirl vibrations.
The large sub-synchronous vibrations will disappear. The imbalance loads could cause the rotor-GFB system to vibrate at
natural mode alone. The increase of dynamic stiffness due to imbalance loads to the threshold speed of the natural mode
is the source of this kind of sub-synchronous vibration. This holds true for different static load conditions. The specific static
load of 36.0 N in this case is unnecessary, but it could help increase the dynamic stiffness effectively.
The ODS of the rotor-GFB system is pictured in Fig. 19. Some orbits of the nodes are hidden to illustrate the nonlinear
response of the rotor-GFB system better. The locations of the nominal clearances of the two GFBs are represented by two
simple ripples. Obviously, whirl motion is related to the cylindrical modes of the rigid shaft. As discussed above, the whirl
vibrations are related to the in-series stiffness of the gas film and foil structure. However, Chen and Gunter [31,37] discussed
the nonlinear response of the rotor supported on oil bearings. The sub-synchronous motions disappeared as the imbalance
mass increase to high levels. In their research, the stiffness of the oil film was higher than that of the shaft and led to the whip
vibration of a beading shaft mode. This is an important difference between the rotor supported on oil-lubricated bearings
and the rotor supported on GFBs.
Extra imbalance displacement cases are not conducted in the experimental section yet. The rotor-GFB system with large
imbalance displacements cannot easily pass the first peak response at synchronous frequencies. However, a typical advan-
tage of the GFBs is long life. Due to heat distribution, manufacturing tolerance, wear, particles attachment and many other
factors, the balance conditions would get worse in the rotating machinery during operation. In rotor-GFB system, the imbal-
ance loads could cause not only large synchronous motions but also high vibrations in sub-synchronous frequency. The man-
ufacturers of rotor-GFB system should pay attentions to the possible sub-synchronous frequency and avoid related
resonance failure of other parts of the system. The above test and simulating results could help elucidate the nonlinear beha-
viours of the system.
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 17

Fig. 18. Orbits (a–h.1) of the rotor centre at the turbine side probe node and corresponding amplitude spectra (a–d.2) in horizontal direction. The total
imbalance displacements range from 0.5 lm to 7.5 lm.

5. Conclusions

This paper presents an experimental and numerical study on the effects of static and imbalance loads on the nonlinear
response of the rigid rotor-GFB system. A test rig with a rigid rotor supported on two GFBs was designed and manufactured.
Experiments with various static and imbalance load configurations were conducted.
This paper reports an interesting phenomenon in which static loads could lead to high sub-synchronous motions in the
rotor-GFB system. In the coast-down tests starting from 40.0 krpm to rest, the sub-synchronous vibrations onset at 0.5 times
the speed frequencies with relative large amplitudes as the static loads increased. This phenomenon also occurred when the
imbalance loads increased. Extra static loads increased the frequencies of the peak response and reduced the duration of the
coast-down tests in the coast-down tests starting from 40.0 krpm to rest. The relationships between the nonlinear test
results and the rotor-GFB system are explained by one simple in-series model of shaft, gas film and foil structure. Static
and imbalance loads could increase the stiffness of GFBs. Thus, the natural frequency of the rigid rotor system determined
by GFBs increased and encountered the threshold speed of the whirl vibration. Sub-synchronous motions occurred in the
tests.
A rotordynamic model based on the FEM and the transient Reynolds equation was derived. The foil structure and gas film
in-series were considered to derive the unsteady bearing force. The predictions showed good consistency with the corre-
18 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 18 (continued)

Fig. 19. Operation deflection shape of the rotor-GFB system with the total imbalance displacement is set 7.5 lm in phase. The two ripples indicate the
locations of the GFBs. Orbits at some nodes are not pictured for better view.
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 19

sponding data from the tests. As the static and imbalance loads increased, the vibration component changed from mainly
synchronous to sub-synchronous and then back to synchronous vibrations. Static and imbalance loads affected the total
in-series dynamic stiffness of the foil structure and gas film as discussed. Thus, the natural frequency of the rotor-GFB system
was extended to the threshold of the whirl vibration with a typical factor, k ¼ 0:5, and led to serious sub-synchronous
vibrations.

Acknowledgements

The authors acknowledge the financial support of National Key R&D Program of China (No. 2018YFB2000100), the
National Natural Science Foundation of China (Nos. 51575170 and 51875185), and the Foundation of Hunan Province
(2018JJ1006).

Appendix A

Physical parameters of rotor and GFBs

Table 1
Physical parameters of rotor.

Rotor parameters
Young’s modulus, ES 210 GPa
Material density, q 7800 kg/m3
Total mass, M 1.440 kg
Diameter at bearing location, Dj 30 mm
Total length, LT 252.0 mm
Bearings’ axial span, LS 112.0 mm
Distance between rotor CG to drive side end, xG 127.1 mm
Turbine transverse moment of inertia, IT;1 4:04  105 kg  m2
Turbine polar moment of inertia, IP;1 -5
7:18  10 kg  m2
GFTB disk transverse moment of inertia, IT;2 5
5:91  10 kg  m2
GFTB disk polar moment of inertia, IP;2 1:16  104 kg  m2

Table 3
Physical parameters of GFBs.

GFB parameters
Bearing radius, R 15.00 mm
Bearing width, L 30.00 mm
Number of bumps, N b 25
Bump height, Hb 0.55 mm
Bump pitch, Lp 3.77 mm
Radial nominal clearance, C 35.0lm
Foil material Inconel X750
Foil thickness, t 0.115 mm
Bump half length, Lh 1.48 mm
Young’s modulus, E 214 GPa
Poisson’s ratio, P t 0.31

Additional measurements of nominal clearances

Static load-deflection tests are used to estimate the nominal clearances of GFBs in the rotor system. Fig. 20 shows the
schematic view of the test rig for the static load-deflection tests. A fixed dummy shaft with the same diameter as the shaft
at the bearing location was used in the tests. The loads were applied via a head of micrometre to push or pull the housing of
GFBs and recorded by the load cell in series. Static loads were applied on the test bearing at 90° from the welding spot.
20 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

Fig. 20. Schematic view of test rig for nominal clearance of GFBs.

Fig. 21. (a) Data of static load-deflection test and (b) static stiffness coefficients of the foil structure for estimation of GFBs nominal clearances at turbine
and thrust sides.
Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271 21

Three proximity probes were installed in parallel to measure the displacements of the test GFBs. The average displacements
from the three probes were recorded as the final data. Fig. 21(a) illustrates the recorded static load versus the bearing dis-
placement. The behaviours of load-deflection were almost consistent for both test GFBs. The test results showed nonlinear
structural relationships between load and deflection. The maximum load was approximately 40.0 N, and the maximum
deflection was approximately 45.0 lm in both directions. The static stiffness coefficients of the foil structure were derived
by dividing the changes in static load by the corresponding changes in displacements. Fig. 21(b) shows the estimated foil
structural stiffness coefficients versus the displacements of both test GFBs. With low stiffness coefficients around the hori-
zontal axis origin, the nominal clearance was estimated at 35.0 lm; as labeled in the figure.

Derivations of transient bearing force

The dimensionless parameters are expressed as


 2
p z h 6lx R xe
P¼ ;Z ¼ ;H ¼ ;K ¼ ;m ¼ ; s ¼ xe t; ð11Þ
pa R C pa C x e
where pa is the ambient pressure, R is the bearing radius, C is the radial nominal clearance, l is the viscosity of gas, xe is the
excitation frequency. All these are used in the dimensionless aspects of the abovementioned parameters.
The unsteady gas film thickness is expressed as
h
H¼ ¼ 1 þ ex sinh  ey cosh þ dðh; Z Þ; ð12Þ
C
where ex and ey are the journal eccentricity ratios in x and y directions as shown in Fig. 1, respectively; dðh; Z Þ is the deflection
of the foil structure, and U ðh; Z Þ is divided by nominal clearance C, i.e., dðh; Z Þ ¼ U ðh; Z Þ=C. The unsteady deflection of the foil
structure is written as
dd
P ¼ K bd þ Cb ; ð13Þ
dse
where the normalised bump stiffness K b , equivalent viscous damping C b , and bump deflection d are expressed with bump
stiffness kb , bump effective covering area A0 , and structural loss factor g in the following equation:
kb C g

Kb ¼ ; Cb ¼ K : ð14Þ
pa A0 m b
In the orbit simulation, the synchronous excitation frequency ratio (m ¼ 1) is used with equivalent viscous damping coef-
ficients. To solve the foil dynamic equation described in Eq. (13), the time-varying foil deflection equation is expressed in the
following discrete form, which adopts the first-order time derivative term approximation:

g Ds PsþDs
U sþDs ¼ Us þ : ð15Þ
g þ Ds g þ Ds K b
The Reynolds equation is solved using the finite difference method and interacts with the foil structure force modelled
with the FEM method. The Gauss–Seidal iteration method with slow relaxation is applied to obtain a transient solution.
Establishing a balanced equation between the forces from the transient gas film and the dynamic foil structure is simple.
However, when gas film is rarefied with high eccentricity ratios (e.g. e > 1) [38], the iterations are more time consuming
with the slow relaxation method. Sub-ambient pressure situations may occur when the transient gas film pressure is calcu-
lated. Given that the top foil merely rests on the bump foil structure, the sub-ambient pressure region can be suppressed by
the motion of the top foil. The sub-ambient pressure is ignored when obtaining the transient bearing force, i.e. the Gümbel
boundary condition [39] is applied.

References

[1] H. Heshmat, J.F. Walton, M.J. Tomaszewski, Demonstration of a turbojet engine using an air foil bearing, ASME Turbo Expo 2005: Power for Land, Sea,
and Air, American Society of Mechanical Engineers, 2005.
[2] D.-K. Hong et al, Ultra high speed motor supported by air foil bearings for air blower cooling fuel cells, IEEE Trans. Magn. 48 (2) (2012) 871–874.
[3] T.H. Kim et al, Rotordynamic performance of an oil-free turbo blower focusing on load capacity of gas foil thrust bearings, J. Eng. Gas Turbines Power
134 (2) (2011) 022501–022501-7.
[4] H. Heshmat et al., Small gas turbine engine operating with high-temperature foil bearings, in Proceedings of the ASME Turbo Expo 2006, Vol. 5, Pts A
and B 2006, pp. 387–393.
[5] C. DellaCorte, Oil-Free shaft support system rotordynamics: past, present and future challenges and opportunities, Mech. Syst. Sig. Process. 29 (2012)
67–76.
[6] G.L. Agrawal, Foil Air/Gas Bearing Technology — An Overview. in ASME 1997 International Gas Turbine and Aeroengine Congress and Exhibition. 1997.
[7] H. Heshmat, P. Hermel, Compliant foil bearings technology and their application to high speed turbomachinery, Tribol. Interface Eng. Ser. (1993) 559–
575.
[8] P. Costamagna, L. Magistri, A.F. Massardo, Design and part-load performance of a hybrid system based on a solid oxide fuel cell reactor and a micro gas
turbine, J. Power Sources 96 (2) (2001) 352–368.
[9] T.H. Kim, L. San Andrés, Heavily loaded gas foil bearings: a model anchored to test data, J. Eng. Gas Turbines Power 130 (1) (2008) 012504.
22 Z. Guo et al. / Mechanical Systems and Signal Processing 133 (2019) 106271

[10] Y. Chen et al, A condensation method for the dynamic analysis of vertical vehicle-track interaction considering vehicle flexibility, J. Vib. Acoust. 137 (4)
(2015) 041010.
[11] Y. Chen, B. Zhang, S. Chen, Model reduction technique tailored to the dynamic analysis of a beam structure under a moving load, Shock Vib. 2014
(2014).
[12] S. Von Osmanski, J.S. Larsen, I. Santos, A fully coupled air foil bearing model considering friction – Theory & experiment, J. Sound Vib. 400 (2017) 660–
679.
[13] D. Rubio, L.S. Andrés, Bump-type foil bearing structural stiffness: experiments and predictions, J. Eng. Gas Turbines Power 128 (3) (2004) 653–660.
[14] H. Heshmat, Advancements in the performance of aerodynamic foil journal bearings: high speed and load capability, J. Tribol. 116 (2) (1994) 287–294.
[15] D. Kim et al, Rotordynamics performance of hybrid foil bearing under forced vibration input, J. Eng. Gas Turbines Power-Trans. ASME 140 (1) (2017)
012507.
[16] B.Z. Yazdi, D. Kim, Rotordynamic performance of hybrid air foil bearings with regulated hydrostatic injection, J. Eng. Gas Turbines Power-Trans. ASME
140 (1) (2017) 012506.
[17] L. San Andrés, D. Rubio, T.H. Kim, Rotordynamic performance of a rotor supported on bump type foil gas bearings: Experiments and predictions, J. Eng.
Gas Turbines Power-Trans. ASME 129 (3) (2007) 850–857.
[18] L. San Andrés, K. Ryu, On the nonlinear dynamics of rotor-foil bearing systems: effects of shaft acceleration, mass imbalance and bearing mechanical
energy dissipation, ASME 2011 Turbo Expo: Turbine Technical Conference and Exposition, American Society of Mechanical Engineers, 2011.
[19] L. San Andrés, T.-H. Kim, Issues on instability and force nonlinearity in gas foil bearing supported rotors, 43rd AIAA/ASME/SAE/ASEE Joint Propulsion
Conference & Exhibit, American Institute of Aeronautics and Astronautics, 2007.
[20] L. San Andrés, T.H. Kim, Forced nonlinear response of gas foil bearing supported rotors, Tribol. Int. 41 (8) (2008) 704–715.
[21] F. Balducchi, M. Arghir, R. Gauthier, Experimental analysis of the unbalance response of rigid rotors supported on aerodynamic foil bearings, J. Vib.
Acoust. 137 (6) (2015) 061014.
[22] P. Bonello, M.F.B. Hassan, An Experimental and Theoretical Analysis of a Foil-Air Bearing Rotor System, J. Sound Vib. 413 (2018) 395–420.
[23] R. Hoffmann, R. Liebich, Characterisation and calculation of nonlinear vibrations in gas foil bearing systems–An experimental and numerical
investigation, J. Sound Vib. 412 (2018) 389–409.
[24] G. Zhiyang et al, Measurement and prediction of nonlinear dynamics of a gas foil bearing supported rigid rotor system, Measurement (2018).
[25] K. Radil, S. Howard, B. Dykas, The role of radial clearance on the performance of foil air bearings, Tribol. Trans. 45 (4) (2002) 485–490.
[26] D.E. Bently, T. Charles, Fundamentals of rotating machinery diagnostics, Mech. Eng. CIME (2003).
[27] Y. Chen, D. Joffre, P. Avitabile, Underwater dynamic response at limited points expanded to full-field strain response, J. Vib. Acoust. 140 (5) (2018)
051016.
[28] Y. Chen et al, Non-model based expansion from limited points to an augmented set of points using Chebyshev polynomials, Exp. Tech. (2019) 1–23.
[29] H. Nelson, J. McVaugh, The dynamics of rotor-bearing systems using finite elements, J. Manuf. Sci. Eng. 98 (2) (1976) 593–600.
[30] H. Nelson, A finite rotating shaft element using Timoshenko beam theory, J. Mech. Des. 102 (4) (1980) 793–803.
[31] W.J. Chen, E.J. Gunter, Introduction to dynamics of rotor-bearing systems, 2007, Trafford Victoria, BC, Canada.
[32] G. Zhiyang et al, Nonlinear dynamic analysis of rigid rotor supported by gas foil bearings: Effects of gas film and foil structure on subsynchronous
vibrations, Mech. Syst. Sig. Process. 107 (2018) 549–566.
[33] K. Feng, S. Kaneko, Link-Spring Model of Bump-Type Foil Bearings, ASME, 2009.
[34] F. Kai, G. Zhiyang, Prediction of dynamic characteristics of a bump-type foil bearing structure with consideration of dynamic friction, Tribol. Trans. 57
(2) (2014) 230–241.
[35] H. Heshmat, J.A. Walowit, O. Pinkus, Analysis of gas-lubricated foil journal bearings, J. Lubr. Technol. 105 (4) (1983) 647–655.
[36] D. Kim, Parametric studies on static and dynamic performance of air foil bearings with different top foil geometries and bump stiffness distributions, J.
Tribol. Trans. ASME 129 (2) (2007) 354–364.
[37] W.J. Chen, Practical Rotordynamics and Fluid Film Bearing Design, 2015.
[38] S.C. Chapra, R.P. Canale, Numerical methods for engineers, Vol. 2, McGraw-Hill, New York, 1998.
[39] G. Stachowiak, A.W. Batchelor, Engineering Tribology, Butterworth-Heinemann, 2013.

You might also like