Mechanical Systems and Signal Processing

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

ARTICLE IN PRESS

Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Design and parameter estimation of hybrid magnetic bearings for


blood pump applications
Tau Meng Lim , Dongsheng Zhang, Juanjuan Yang, Shanbao Cheng, Sze Hsien Low,
Leok Poh Chua, Xiaowei Wu
School of Mechanical and Aerospace Engineering, Nanyang Technological University, Singapore 639798, Singapore

a r t i c l e i n f o abstract

Article history: This paper discusses the design and parameter estimation of the dynamics character-
Received 13 August 2008 istics of a high-speed hybrid magnetic bearings (HMBs) system for axial flow blood
Received in revised form pump applications. The rotor/impeller of the pump is driven by a three-phase
21 December 2008
permanent magnet (PM) brushless and sensorless DC motor. It is levitated by two
Accepted 21 March 2009
Available online 8 April 2009
HMBs at both ends in five-degree-of-freedom with proportional-integral-derivative
(PID) controllers; among which four radial directions are actively controlled and one
Keywords: axial direction is passively controlled. Test results show that the rotor can be stably
Magnetic bearings system supported to speeds of 14,000 rpm. The frequency domain parameter estimation
Parameter estimation
technique with statistical analysis is adopted to validate the stiffness and damping
Blood pump
coefficients of the HMBs system. A specially designed test rig facilitated the estimation
Magnetostatic
of the bearing’s coefficients in air—in both the radial and axial directions. The radial
stiffness of the HMBs is compared to the Ansoft’s Maxwell 2D/3D finite element
magnetostatic results. Experimental estimation showed that the dynamics character-
istics of the HMBs system are dominated by the frequency-dependent stiffness
coefficients. The actuator gain was also successfully calibrated and may potentially
extend the parameter estimation technique developed in the study of identification and
monitoring of the pump’s dynamics properties under normal operating conditions with
fluid.
& 2009 Elsevier Ltd. All rights reserved.

1. Introduction

Active magnetic bearings (AMBs) have been widely applied in the field of heart pump research [1–8]. The new-
generation blood pumps, namely the magnetically levitated artificial heart pumps, utilize magnetic bearings to support the
rotating impeller of the pump. There is no mechanical contact between the bearings and the flowing blood, thus
minimizing the risks of red blood cell damage and blood clots during operation. The dynamic characteristics of magnetic
bearings in magnetically levitated blood pumps are critical to the pump performance, and therefore, the experimental
parameter estimation of magnetic bearings is necessary. Baloh et al. [9] propose a transfer function estimation method
based on modal decomposition to obtain system characteristics of their centrifugal blood pump. However, the limitation in
their experimental design is that the actuator gain’s parameter in their plant model cannot be determined satisfactorily. In
addition, the estimates for the actuators’ stiffness and damping are only represented at their operating speed, with no
further detailed experimental results being reported using this method. There is also no statistical information to validate

 Corresponding author. Tel.: +65 6790 5565; fax: +65 6792 4062.
E-mail address: mtmlim@ntu.edu.sg (T.M. Lim).

0888-3270/$ - see front matter & 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ymssp.2009.03.012
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2353

the experimental model. Asama et al. [10] use a sinusoidal excitation method to investigate the Maglev centrifugal pump’s
dynamics characteristics in the axial direction. According to Asama, due to the application of the excitation force in the
axial direction, the parameters in the radial directions cannot be estimated. Although a high-order polynomial estimate is
used to fit the theoretical curve with the experimental frequency response curves, no statistical information is provided to
support the reliability of their results within a certain confidence bounds. Literature review has also indicated that
attempts to investigate the dynamics characteristics of blood pump are only performed on the centrifugal type [11]. Hence,
there is a need to extend such investigation into the axial flow blood pump.
The dynamics characteristics of magnetic bearings are known to be highly dependent on the frequency responses of
the elements in the active magnetic bearing control loop [12]. This is due to the frequency-dependent nature of the
AMB system, i.e. limited bandwidth of the position sensors and amplifiers, time lags in sensors, filters, and the eddy
current and hysteresis losses in the electromagnetic coils. From the design point of view, it is vital to accurately predict
the dynamics response of the magnetically supported system using the AMB data. Hence, the theoretical
frequency-dependent controller’s transfer function has been used in simulation or experimentally to obtain the stiffness
and damping coefficients of magnetic bearings [13–15]. However, this method excludes the system lags due to the
digital signal processing (DSP), amplifiers, feedback sensor, actuator’s cross-axis capabilities and electromagnetic losses
[16,17].
In this paper, a compact hybrid magnetic bearings (HMBs) system consisting of two HMBs for levitating the rotor and
one three-phase permanent magnet (PM) sensorless and brushless DC (BLDC) motor for driving the rotor, has been
designed and developed for the axial flow blood pumps. The Ansoft’s Maxwell 2D/3D finite element software is employed
to check whether the proposed HMB is effective with respect to levitation forces, magnetic properties, electro-mechanical
interactions and calculate its magnetostatic radial stiffness. A special test rig has been designed for parameter estimation of
the HMBs supporting system in both the radial and axial directions. The principles of parameter estimation of a two-
degree-of-freedom (DOF) system and a one DOF system with statistical information are introduced. The specially designed
test rig allowed for the application of a single forcing mechanism onto the pedestal housing the rotor/shaft assembly, at
7451 to the x- and y-axes of the rotor. By injecting a multi-frequency excitation force signal onto the rotor through
the HMBs, it is observed that the maximum displacement linear operating range is 20% of the static eccentricity with
respect to the rotor and stator gap clearance. The experimental estimations were performed on both stationary and rotating
rotor to determine the influence of centrifugal forces due to rotor unbalance mass on the estimated results. The Schroeder
Phased Harmonic Sequences (SPHS) [18], a multi-frequency test signal, was used to persistently excite the pedestal. By
treating the system as a black box, the parameter estimation of the stiffness and damping coefficients of HMBs system in
the frequency domain could be realized experimentally, and statistical information was provided to validate the system
model and verify the estimated coefficients. Using this method, the dynamic properties of magnetically levitated axial flow
blood pumps will be better understood so that they can also be identified and monitored under normal operating
conditions with fluid.

2. Design of axial flow blood pump system

2.1. Magnetic design of HMBs

The magnetic core of the HMBs as shown in Fig. 1 is made of 0.23-mm-thick M19_29G silicon steel laminations that help
to decrease the HMBs’ eddy current and hysteresis losses. Copper wires are also wound in pairs around the inner core to
reduce resistive losses, and the generated flux loop circulates across the two corresponding laminated rings onto the rotor.
Between the two rings of each HMB, there is a permanent magnet ring made of neodymium–iron–boron (NdFeB) material,
having a magnetic coercivity of 9.465 kOe and nominal BH value of 10.43 MGOe, to deliver the bias magnetic flux for the
electromagnet; thereby successfully producing a passive axial control force. This permanent magnet ring is axially
magnetized in the z direction. Due to the composition of the electromagnets and PMs, this kind of magnetic bearing is
known as a hybrid magnetic bearing.
The HMBs have 180 copper turns on each coil with a wire diameter of 0.15 mm, and a DC bias current of 0.2 A is used to
enhance the stiffness of the HMBs in the passive axial direction. The radial stiffness coefficients of the HMBs are calculated
by using magnetostatic analysis to determine the quasi-static force and displacement response in the radial direction. This
is achieved using Ansoft’s MaxwellTM 2D and 3D finite element software. The model drawing is performed using Pro/ETM
software and imported into Ansoft MaxwellTM as universal .sat format.
The magnetic field is produced by the permanent magnets and the DC currents flowing in the coils. The electric field is
restricted to the objects modelled as real (non-ideal, i.e. not infinitely conducting) conductors. The electric field exists
inside the conductors as a consequence of the DC current flow and is totally decoupled from the magnetic field. Therefore,
as far as magnetic material properties are concerned, the distribution of the magnetic field is influenced by the spatial
distribution of the permeability. Firstly, the magnetic solution is assumed to be time invariant, and secondly the rotor
suspended in the HMB is assumed to be stationary. Lastly, the energy transformation in connection with the magnetostatic
solution is only attributable to the ohmic losses associated with the currents flowing in the coils.
ARTICLE IN PRESS
2354 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Aluminum Z
6061 stator N

Silicon steel S
laminations
Axially
Rotor magnetized
assembly with PM ring
lamination Y
rings
sandwiching
PM ring

M4×0.50
Ø2.00

14.70 Z Y Ø21.00 Ø35.60

4 ×Copper
coils
X

Stainless Silicon steel laminations


steel sleeves

2.53 Ø20.30
PM ring
11.62 Ø16.00
5.06
Ø18.60
2.53

Ø20.00
Ø20.10
Ø14.80

Fig. 1. Design of the HMBs: (A) three-dimensional view of the HMB; (B) cross-sectional view of the HMBs at coil level; and (C) sectional view of rotor.

The magnetostatic field solution should satisfy the following Maxwell’s equations:
! ! ! 9
Ampere0 s circuital law : rH ¼ J =
! ! (1)
Gauss0 s law for magnetism : r B ¼ 0;
In addition, the following constitutive relationship associated with the material is also applicable:
! ! !
B ¼ m0 mr H þm0 M p (2)
! ! !
where H ðx; y; zÞ is the magnetic field strength
!
in A/m, B ðx; y; zÞ is the magnetic flux density in Tesla, J ðx; y; zÞ is the
conduction current density in A/m2, M p ðx; y; zÞ is the permanent magnetization in A/m, m0 ¼ 4p  107 H/m is the
permeability of free space, and mr is the relative permeability.
The magnetostatic solver handles both 3D linear and nonlinear problems [19]. For nonlinear materials such as
permanent magnets, the dependency between the H and B fields is nonlinear and can be isotropic or orthotropic.
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2355

Additionally, if a demagnetized condition is to be taken into account, Ansoft MaxwellTM allows a solution based on a
previously computed demagnetization operating point. In the case of nonlinear applications, a classic Newton–Raphson
iterative algorithm with user-controlled accuracy is used to define the solution for
! !  ! !
B ¼ B0 þ½m  ðH  H0 Þ (3)

!
 @B 
½m ¼ ! ¼ ½D m þ ½m (4)
@H
! ! 
where B0 and H0 are the previously calculated field solution, ½D m is a general full tensor, and [m] is given by
2 3
mx
6 my 7
½m ¼ 4 5 (5)
mz
with mx, my, mz taking into account the anisotropic effects of any laminations present in the model. The 3D magnetostatic
solver considers the magnetic field H with the following components:
! ! !
H ¼ Hp þrj þ Hc (6)
!
where j is the magnetic scalar potential, Hp is a particular solution constructed by assigning values to ! all the edges in the
mesh in such a way that Ampere’s law holds true for all contours of all the tetrahedral meshes, and Hc accounts for the
permanent magnets. In order to ensure a quadratic approximation inside each finite element for the magnetic scalar
potential, ten DOF per tetrahedron are used at each of the four vertices and all six mid-edge nodes.
Since the mechanical design of the HMB is symmetrical in the orthogonal directions, a significant reduction in
computational time and memory is realised by only analysing a quarter of the HMB in the simulation, as shown in Fig. 2(A).
For the finite element model, a total number of 485,628 elements are generated, as shown in Fig. 2(B). The background is a
vacuum ‘‘Region’’, the surface of the coil has a boundary condition of ‘‘Insulating’’, and the excitation is a stranded-type
current of 36 A (total effect of cross-section of the coil, i.e. 0.2 A  180 Turns ¼ 36 A Turns). The magnetization direction of
the permanent magnet is from the negative Z to the positive Z direction, and the direction of the DC current is clockwise
from the top view of the HMB. The convergence criterion error for the solution is less than 1%. It takes around two hours to
perform each simulation run by using a PC with 2.33 GHz CPU and 2 G RAM.
To illustrate the flow of the magnetic field in the HMB, the cross-section of the HMB in the YZ-plane is analyzed as a
two-dimensional model [20]. The boundary conditions and excitation inputs are similar to those of the three-dimensional
model. As shown in Fig. 3, the magnetic flux lines are mostly confined within the width of the silicon laminations stack and
permanent magnet ring with no risk of saturation.
Since the radial direction of the HMBs is designed to be actively controlled and its effective magnetic air gap at rotor
equilibrium position is 0.45 mm, a small incremental rotor displacement in the radial direction will correspond to an
increase in the control current dI. This change in control current can be expressed as the product of Displace-
ment  EddyCurrentProbeGain  kp (PID’s ProportionalGain)  AmplifierGain. If the sensitivity of the eddy current probe is 8 V/
mm, amplifier gain is 1 A/V and kp ¼ 1, the quasi-static displacement versus force response, as shown in Fig. 4, is calculated
using Eq. (6). The linear relation between magnetic force and displacement over a 45 mm range or 10% of the radial effective
magnetic air gap, gives a radial stiffness of 1694 N/m.

Axial

Radial
Tangent
x

Fig. 2. Sectional view of HMB: (A) 3D view before meshing and (B) finite element model meshed with 485,628 elements.
ARTICLE IN PRESS
2356 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Flux Lines (Weber)


9.7015e-006
8.6599e-006
7.6183e-006
6.5766e-006
5.5350e-006
4.4934e-006 Rotor Stator
3.4518e-006
2.4102e-006
1.3686e-006
3.2697e-007
-7.1464e-007
Silicon steel
Axially magnetized laminations
PM ring

Copper coil

Aluminum
Stainless steel 6061 stator
sleeve
Center

Z (axial)

Y (radial)

Fig. 3. Axial cross-sectional view on one pole of the HMB with flux distributions.

Simulated Data
Best Fit Line
1.45
Radial force, Fy (N)

1.4
dFy
dy
1.35

1.3
dFy
HMB Radial Stiffness, = Ky 1,694N/m
dy
1.25

1.2
0 5 10 15 20 25 30 35 40 45
Rotor displacement in radial direction, dy (um)

Fig. 4. Calculated magnetic force versus rotor displacement in the radial direction.

2.2. Mechanical design of blood pump

The cross-sectional view of the axial flow blood pump is as shown in Fig. 5. The impeller is enclosed in the rotor, which
is driven by the three-phase PM DC motor and supported by two HMBs at both ends without mechanical contact. The blood
flows in the axial direction with respect to the inlet and the outlet cannulae is achieved by passing through the straightener,
the impeller and the diffuser. The straightener and the diffuser that are fixed to the pump housing protrude into the rotor
without coming into physical contact with the rotating rotor, whilst the impeller is shrunk-fit into the bore of the rotor
[21,22].
The left and right HMBs support the two ends of the rotor and the torque from the motor drives the rotor at its mid-
section. The detailed structure and dimensions of the HMBs can be referred to Fig. 1. Rotor displacements relative to the
stator housing are measured by two pairs of eddy current displacement transducers placed at 7451 to the x- and y-axes.
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2357

Rotor Motor
Left HMB Right HMB Pump Casing

Outlet Inlet

Diffuser Straightener
Impeller

Fig. 5. Cross-sectional view of axial flow blood pump.

Motor Stator 4-Pole Magnetic


Bearing Stators

Rotor

Fig. 6. Photograph of the HMB, rotor and motor.

The motor’s stator is also made of 0.23-mm-thick silicon steel laminations and is located at the middle of the rotor, with
an eight-pole permanent magnet glued diametrically opposite the rotor. The motoring coils wound around the stator are
driven by three-phase currents to provide higher average motoring torque on the rotor. The outstanding feature of the
structure of the proposed blood pump is that it includes an enclosed impeller to enhance the motoring electromagnetic
coupling [3].
Fig. 6 shows the photograph of one of the two HMBs, a rotor (weighs 52 gm) and a motor’s stator with wound copper
coils. It is important to emphasize that the mechanical design of the HMBs is symmetrical about the x- and y-axes, and the
axes of symmetry lie in the middle of the pole faces. It is therefore reasonable to assume that the concentration of HMBs’
actuator forces will be in the middle of the pole faces or on the direct-axes.

2.3. Hybrid magnetic bearings controller

The four radial directions of the rotor are actively controlled by the two HMBs driven by digital proportional-integral-
derivative (PID) controllers. A typical system control block diagram for one of the radial directions is as shown in Fig. 7.
The transfer function of the PID controller is as follows:
kd s ki
Gc ðsÞ ¼ kp þ þ (7)
1 þ sT d T i s þ 1
ARTICLE IN PRESS
2358 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Reference Signal
+ PID PWM
HMB Rotor
Controller Amplifier
-

Position
Sensor

Fig. 7. Single control loop of hybrid magnetic bearings.

Table 1
PID control parameters of hybrid magnetic bearings.

kp 1.00 Td 0.0001
kd 0.00065 Ti 0.0318
ki 0.25 I0 0.20 A

where kp, kd and ki are the proportional, differential and integral coefficients, respectively. Td is the time constant of the
differentiator and Ti is the time constant of the integrator. The PID controller parameters are tuned experimentally, and
their preferred values with the HMBs’ bias current I0 are tabulated in Table 1.
The dSPACE ds1103 in conjunction with the MATLAB 6.1 and Simulink software are used as the rapid prototyping tools
for the development of the digital PID controller of the HMB’s system. The model of the digital controller is firstly designed
and simulated in Simulink, and thereafter this model is built into the digital signal processing module of the dSPACE ds1103
hardware in the loop interface. It is in this manner that the digital PID controller is successfully implemented to operate
real-time via the dSPACE ds1103s hardware. The voltage range of the interface for dSPACE ds1103 is from 10 to +10 V,
sampling frequency of the controller was 20 kHz, 16-bit A/D and 14-bit D/A converters were chosen as the settings for
the experiments. The rotor displacements are measured by four Applied Electronics Corporation’s eddy current probes
(AEC-5503A) which have a resolution of 0.5 mm and gain of 8 V/mm. The coils of the HMBs are driven by the Advanced
Motion Controls (AMC) Pulse Width Modulated servo amplifiers (PWM, 25A series).
The STMicroelectronics’ ST7FMC microcontroller is used to control the PM three-phase brushless DC motor. The ST7FMC
microcontroller is an integrated system designed to provide users with a complete, ready-to-use motor application. Its
Sensorless Control Mode is chosen to drive the motor using a Back Electromotive Force zero-crossing detector as a part of
the ST7 Motor Controller peripheral. As a consequence, the requirement of a flux sensor in the HMBs system is eliminated,
which makes the overall mechanical design of the axial flow pump more compact.

2.4. Measurements of pump’s rotor levitated responses in radial and axial directions

This section discusses the investigation of the stability and measures the maximum levitated responses of the pump
rotor in both the radial and axial directions prior to the parameter estimation procedure to extract the coefficients of the
HMBs support parameters. This is to acquire a better understanding of the rotor’s maximum vibrational amplitudes and
critical speeds, which may violate the linear operating range of the HMBs dynamics support coefficients.
During the experiment, it was observed that the rotor of the pump could be suspended stably by the HMBs system in air
while stationary and with rotation at speeds of up to 14,000 rpm. In general, the radial displacement of the rotor can be
read from the orbit size. However, a direct drive motor drive system, such as ours, can introduce a significant amount of
cogging torque into the rotating system. This cogging torque [3] generally generates asynchronous responses and makes
measuring the orbit size difficult. The solution to this problem is to apply fast Fourier transform (FFT) to the displacement
signals and to record their amplitudes at the rotor speed. The rotor responses were recorded at intervals of 500 rpm in the x,
y (radial) and z (axial) directions, with rotational speeds varying from 0 to 14,000 rpm. Their corresponding frequency
responses during levitation are as shown in Fig. 8. It was observed that the rotor’s suspension during rotation in air was
very stable and no collision between rotor and stator was noted. The hovering displacement of the stationary rotor or 0 rpm
was around 1 mm in the radial and axial directions. The measured maximum rotor displacements occurred at 10,000 rpm in
the radial and axial directions were about 30 and 70 mm, respectively. At the maximum speed of 14,000 rpm, the
displacements in the radial and axial directions were about 8 and 35 mm, respectively.
The difference in responses, shown in Fig. 8, between the left and right supports of the HMBs, is due to the asymmetric
centrifugal forces caused by the distribution of residual unbalance mass along the rotor. This was confirmed by testing the
transfer function of individual HMBs using the calibration test rig as shown in Fig. 18. This was done by exciting one HMB
with SPHS while turning the other off. The obtained transfer function exhibited almost similar characteristics.
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2359

Fig. 8. Levitated response of pump’s rotor in the x, y and z directions. ‘Displacement’ shortened to ‘dp’ in graph.

Fig. 9. Exploded view of the rotor assembly.

On the other hand, Fig. 9 shows that our rotor assembly employing a seamless joint using epoxy to glue all the parts
together. It was after obtaining the experimental rotor responses, we felt that we may not have applied an even distribution
of epoxy on the IDs and ODs of the 13 assorted parts on the rotor; especially the right side. The designed chamfer to trap the
epoxy and enhance the adhesive strength of the spacers, could also aggravate the problem. The authors are confident that
the impact of this problem will be minimized with better design of the rotor.

3. System’s dynamics model and parameter estimation procedures

The experiment design for parameter estimation is usually tackled by the following procedures:

(1) formulation of a mathematical equation for the model based on physical assumptions describing the system’s
dynamics,
(2) selection of a suitable excitation force signal,
(3) selection of a suitable method and location to inject this excitation force signal into the dynamics model,
(4) selection of a data analysis package that provides good sampling rate, and
(5) method of validating the dynamics model and evaluating the estimated results using statistical methods.
ARTICLE IN PRESS
2360 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

These five factors are significant contributions towards achieving the estimated dynamics support’s coefficients of the
system with minimum variance. A multi-frequency signal, such as the SPHS [18], which was successfully employed by Lim
and co-workers [12,17] to persistently excite a simple second-order model test rig, will be adopted as the excitation force
signal in the following parameter estimation procedures.
Assessment of the quality of the estimates is carried out by statistical methods. These methods are advantageous as they
can provide Goodness-of-Fit (G.o.F) calculations to confirm the validity of the system’s dynamics model and the
experimental observations, and serve as an important diagnostic check on the confidence bounds of the estimated
coefficients. The following sub-sections will describe the formulations of the algorithm for the one and two DOF dynamics
models to realise the parameter estimation procedures in a systematic manner.

3.1. One DOF dynamics model

The one DOF dynamics model formulated here will be used to estimate the structural dynamics support parameters of
the experimental test rig, as well as the HMBs in the radial and axial directions.
Fig. 10 indicates that a mass m supported by two linearised spring and damping coefficients, the equation of motion
with a multi-frequency excitation force f applied in the y-direction is:
my€ þ ky þ cy_ ¼ f (8)
Eq. (8) can be expressed in the form of frequency domain:

FðjoÞ ¼ mo2 YðjoÞ þ kYðjoÞ þ cjoYðjoÞ (9)

Defining G(jo) as the Fourier coefficient of the displacement/force transfer function

YðjoÞ
GðjoÞ ¼ (10)
FðjoÞ

where Y(jo) is the displacement and F(jo) is the Fourier transform of the excitation force. The complex transfer function
G(jo) can be separated into real and imaginary parts as
" #
ðGÞr
GðjoÞ ¼ (11)
ðGÞi

Fig. 10. General model of one DOF dynamics model.


ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2361

Therefore, Eq. (9) can be written as


2 3 2 3
ðGÞr oðGÞi mo2 ðGÞr þ 1
6 7  6 7
6 ðGÞi oðGÞr 7 k 6 mo2 ðGÞi 7
6 7 ¼6 7 (12)
6 ;; ;; 7 6 7
4 5 c 4 ;; 5
;; ;; ;;

By replacing o in Eq. (12) with no0 to indicate the use of discrete frequencies where n is an integer in the range of 1 to N,
o0 is the fundamental frequency of the excitation force and N is the total number of harmonics applied, Eq. (12) can be
written as
2 3 2 3
ðGÞrn no0 ðGÞin mn2 o20 ðGÞrn þ 1
6 7  6 7
6 ðGÞin
6 no0 ðGÞrn 7 k
7
6
6 mn2 o20 ðGÞin 7
7
6 ;; 7 c ¼6 7 (13)
4 ;; 5 4 ;; 5
;; ;; ;; n¼1;...;N

By re-arranging coefficients to be estimated on the left and known parameters on the right side of the equation, Eq. (13)
can be generalized as

½W F 2N2 fFF g21 ¼ ½C F 2N1 (14)

where WF is a 2N  2 matrix containing the Fourier coefficients of the transfer function of the displacement and applied
force, FF is a 2N  1 matrix containing the parameters to be estimated and CF is a 2N  1 matrix containing the Fourier
coefficients of the transfer function of the displacement, applied force and mass. The least-squares estimator for the
coefficient matrix, FF, in Eq. (14) is given by Gujarati [23] as

_
FF ¼ ðW TF W F Þ1 W TF C F (15)

3.2. Two DOF rotor-bearing model with direct- and cross-axes coefficients

The following two DOF models will be used to estimate the dynamics support parameters of the HMBs in the radial
direction. Like squeeze film dampers [24], the HMBs’ dynamics support parameters can be modelled by eight linearised
direct- and cross-axes coefficients. As shown in Fig. 11, a rotor of mass m is supported by the direct- and cross-axes springs

Fig. 11. Two DOF rotor-bearing model with eight linearised direct- and cross-axes coefficients.
ARTICLE IN PRESS
2362 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

and dampers modelled by eight linearised coefficients of the HMBs:

mx€ þ C xx x_ þ C xy y_ þ K xx x þ K xy y ¼ bx f
my€ þ C yy y_ þ C yx x_ þ K yy y þ K yx x ¼ by f (16)

Eq. (16) can be expressed in the form of frequency domain:

ðmo2 þ C xx jo þ K xx ÞXðjoÞ þ ðC xy jo þ K xy ÞYðjoÞ ¼ bx FðjoÞ


ðmo2 þ C yy jo þ K yy ÞYðjoÞ þ ðC yx jo þ K yx ÞXðjoÞ ¼ by FðjoÞ (17)

where f is the excitation force signal applied to the rotor, bx ¼ b cos a, by ¼ b sin a, a ¼ tan1 (by/bx), is the forcing angle and
b is the input transducer coefficient. Defining Gx(jo) and Gy(jo) as the Fourier coefficients of the transfer function:

XðjoÞ
Gx ðjoÞ ¼
FðjoÞ
YðjoÞ
Gy ðjoÞ ¼ (18)
FðjoÞ

The complex transfer function Gx(jo) and Gy(jo) can be separated into real and imaginary parts as

Gx ðjoÞ ¼ ðGx Þr þ jðGx Þi


Gy ðjoÞ ¼ ðGy Þr þ jðGy Þi (19)

Again, by replacing o in Eq. (17) with no0 to indicate the use of discrete frequencies where n is an integer in the range of
1 to N, and o0 is the fundamental excitation frequency of the applied force, and then expressing Eqs. (17), (18) and (19) in
matrix form, Eq. (20) can be obtained:
2 32 3
ðGx Þrn ðGy Þrn no0 ðGx Þin no0 ðGy Þin 1 K xx K yx
6 76 7
6 76 K xy K yy 7
6 ðGx Þin ðGy Þin no0 ðGx Þrn no0 ðGy Þrn 0 76 7
6 76 7
6 7 C xx C yx 7
6 ;; ;; ;; ;; ; ; 76 7
6 76 7
6 76 C xy C yy 7
6 ;; ;; ;; ;; ; ; 76 5
4 54
;; ;; ;; ;; ;; bx by
2 3
ðGx Þrn ðGy Þrn
6 7
6 ðGx Þi
6 n ðGy Þin 7
7
6 7
26
¼ mðno0 Þ 6 ; ; ;; 7 7 (20)
6 7
6 ;;
4 ;; 7 5
;; ;; n¼1;...;N

Similarly, Eq. (20) can be generalized as Eq. (14) and solved by Eq. (15).

Fig. 12. Two DOF rotor-bearing model with only direct-axis coefficients
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2363

3.3. Two DOF rotor-bearing model with only direct-axis coefficients

As demonstrated by Burrows et al. [24], the cross-axis coefficients in a squeeze film damper can cause the parameters of
the bearing in the direct-axis to be incorrectly estimated. For this reason and the fact that the radial HMBs here is
symmetrically designed about the x- and y-axes, it is only necessary to illustrate the two DOF rotor-bearing model with
only direct-axis coefficients. As shown in Fig. 12, if the excitation forcing angle a is chosen at 451, the applied force in the
vertical direction can be decomposed equally into two components, fx and fy, in the x- and y-directions, respectively. The
rotor of mass m is supported by only the symmetrical direct-axes stiffness and damping coefficients, Kxx ¼ Kyy and Cxx ¼ Cxx,
in the x- and y-directions, respectively.
Since the system is uncoupled, the excitation force and displacements are symmetrical in the x- and y-directions.
Therefore, it is adequate to represent the equation of motion for the rotor-bearing system with just one direction as
my€ þ K yy y þ C yy y_ ¼ f y ¼ f sin a (21)

The parameter estimation algorithm and digital signal processing is simplified, as the two DOF rotor-bearing model is
reduced to one as in Eq. (21). Therefore, based on the same one DOF parameter estimation algorithm described in Eqs.
(9)–(15), we can formulate the following WF, FF and CF matrices:
2 3
ðGÞrn no0 ðGÞin
6 7
6 7
6 ðGÞin no0 ðGÞrn 7
WF ¼ 66 7
7
6 ;; ;; 7
4 5
;; ;;
" #
K yy
FF ¼
C yy
2 3
mn2 o20 ðGÞrn þ 1
6 7
6 mn2 o2 ðGÞi 7
6 0 n 7
CF ¼ 6 7 (22)
6 ;; 7
4 5
;; n¼1;...;N

where
" #
YðjoÞ ðGÞr
GðjoÞ ¼ ¼
F sin aðjoÞ ðGÞi

Similarly, Eq. (22) can be generalized as Eq. (14) and solved by Eq. (15).

3.4. Model validation and statistical information of estimated parameters

The validation of the system model by the simple comparison of theoretical and experimental results is not a
satisfactory measure. A reasonable solution to the problem is by driving the system with the coefficients estimated under
various operating conditions, thereby comparing the magnitude of the difference or errors to that measured during the
course of the experiment. In this aspect, the G.o.F, R2, calculation in the estimation program provides a quantitative
measure of the model validity, when interpreted in conjunction with the confidence bounds associated with each
parameter [23].
R2 can be corrected for the degrees-of-freedom by
2 n1
R ¼ 1  ð1  R2 Þ (23)
nk
2
where R is a measure of G.o.F of the whole experimental data, and k is the number of coefficients to be estimated.
However, in many cases, a well-fitted regression may still contain some insignificant coefficients. The significance test for
the estimated coefficients is established by introducing a scalar, w, representing the sum of squares of the residual error. wi2
can be estimated as

ð1  R2 ÞðSs2i Þ
w2i ¼ (24)
n3
where subscript i represents the corresponding coefficient to be estimated. Ss2i determines the total sum of square error
from the regression estimate. The standard error of the estimated coefficient can be described as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s:e:ðFF Þi ¼ wi jth diagonal element of ðW TF W F Þ1 (25)
ARTICLE IN PRESS
2364 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 13. Parameter estimation procedure for two DOF model.

The standard t-Ratio value was obtained by dividing the estimated coefficients with the corresponding error:
_
ðFF Þi
t-Ratio ¼ _ (26)
s:e:ðFF Þi
Coefficients estimated were considered significant, that is, non-zero, if the absolute t-Ratio is greater than the value
obtained from the standard t-distribution tables for a 95% confidence interval.

3.5. Parameter estimation procedure

By acquiring the appropriate signals, the parameter estimation procedure for the one or two DOF dynamics model could
be successfully performed. Fig. 13 shows that the force signal f(t) and the corresponding displacements, x(t) and y(t), were
sampled and digitized on-line during the experiments. A discrete-time, fast Fourier transform was applied to the signals
f(t), x(t) and y(t), and then based on Eq. (18), Gx(jo) and
_
Gy(jo) could be constructed, to provide the data to form the arrays
of WF and CF in Eq. (14). The least-square estimator FF was then determined using Eq. (15). Lastly, based on Eqs. (24) and
(26), statistical tests using G.o.F were performed to validate the system model and evaluate the significance of the
estimated coefficients, respectively.

4. Preparatory experiments

4.1. Description of test rig

It is clear from the modelling of the one and two DOF dynamics systems earlier that the important measurement to be
undertaken in the radial and axial parameter estimation of the HMBs would be the force and displacement variables. Fig. 14
shows a photograph of the multi-purpose test rig that has been designed to facilitate and achieve the following:

(a) measurement and injection of the excitation force into the dynamics model,
(b) restraint of the angular deflexion of the rotor in the orthogonal directions,
(c) adjustment of air gap between rotor and stator of the HMBs, and
(d) measurement of the x- and y-direction displacements using eddy current probes.

The schematic arrangement in Fig. 15 shows an 8 mm diameter horizontal steel shaft assembled to the rotor by means of
two rolling element bearings on both sides of the rotor. This design allows the rotor to rotate freely on the two bearings.
However, the rotor axial movement relative to the shaft is fixed by two end-collars. The design also ensures that only rotor
translational movement is allowed in the HMBs stator bore and angular deflexion movement is restrained. The horizontal
shaft is able to slide freely on the two horizontal linear bearings (HLBs), which are mounted onto the pedestal. The pedestal
assembly consisting of the horizontal shaft, rotor and pedestal, is able to slide vertically along the two 8 mm diameter
vertical steel shafts located diagonally on the pedestal. The two ends of the vertical shafts are fixed to the top and base
plates. The stator housing of the pump is held by two stator-stands rigidly secured onto the base plate. At the start of the
experiment, the concentric adjustment of the rotor and stator bore is facilitated by the two vertical air gap adjusting screws
that are inserted through the four supporting spiral springs. This is due to the adjustment of the pedestal height, which is
connected to the rotor, directly affecting the air gap thickness between the rotor and HBMs’ stator. Besides realizing the
application of excitation force onto the rotor, the spring-supported pedestal also assists in the partial levitation of the rotor
assembly; which otherwise will be slightly too heavy for the HMBs’ actuator forces to manage.
A force transducer is connected to a shaker mounted on top of the pedestal via a stinger rod to measure the excitation
force. The SPHS test signal is fed to the shaker, which is mounted on a stand to perturb the pedestal. The eddy current
probes are used to measure the rotor displacement in x- and y-directions. During the experiment, the SPHS was digitally
generated in the dSPACE ds1103 with MATLAB Simulink and then pre-recorded onto the TEAC RD-145 T DAT DATA Recorder
for playback purposes. It was in this manner, that the pedestal was persistently excited by the shaker (B&K 4810) with
the SPHS test signal via the tape recorder. The DSPACE ds1103 acquired the force and the resulting displacements signals
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2365

Shaker

Eddy
Current
Probe
Sliding
Vertical
Shaft

Adjusting
Screw With
Supporting
Spring

Pump Casing
With HMBs
and motor
inside

Horizontal
Linear
Bearings
Vertical (HLBs)
Linear
Bearings Horizontal
(VLBs) Shaft

Fig. 14. Test rig used for parameter estimation of the HMBs system.

real-time for the successful implementation of the on-line parameter estimation procedure. In order to minimize the
variance of the estimated coefficients, five periods of force and displacement signals were acquired and averaged
throughout the experimental estimation procedure.

4.2. Parameter estimation of pedestal structural support using one DOF model

Successful experimental determination of the estimated coefficients for the HMBs requires a good prior knowledge of
the supporting structural parameters. Otherwise, the estimated coefficients will be a composite of both HMBs and
supporting structure. Awareness of this outcome led to the performance of structural identification test on the pedestal
support spring stiffness.
For this test, the horizontal shaft with rotor assembly was removed, but the pedestal was first adjusted to the desired
height where the rotor would be concentric with the HMBs stator bore. As the pedestal supported by the four spiral springs
could only move in the vertical direction, the one DOF parameter estimation algorithm formulated in Section 3.1 was used
to estimate the effective stiffness coefficient of the spiral springs. The mass of the pedestal is 0.79 kg. The SPHS signal
bandwidth was set as 1–100 Hz with fundamental frequency of 0.5 Hz. Based on the procedure of parameter estimation of
one DOF dynamics model, the acquired excitation force and displacement signal data was read into the MATLAB program
of the one DOF parameter estimation algorithm, from which the estimated results of spring stiffness were then obtained.
Fig. 16 shows the time domain plots of the excitation force and the corresponding pedestal displacement.
In Table 2, the estimated pedestal support stiffness coefficient of 134.5 kN/m indicated that the best G.o.F of 0.9952 was
obtained by using the transfer functions data in the frequency band of 60–70 Hz, that is, in the neighborhood of the test rig
natural frequency. The G.o.F, calculated by Eq. (24) of almost unity indicated that a perfect mathematical model structure
was adopted in the system dynamics’ model. On the other hand, the damping coefficient of 3.682 N s/m was statistically
interpreted by Eq. (26) to be insignificant; that means it could be treated as zero value. The author had recently published a
paper [25] addressing this issue. It examines the reliability of the statistical information on significance testing using the
least-square estimator. In this paper, the sensitivity analysis of the synchronous responses of a multi-mode rotor-bearing
system using both significant and insignificant estimated coefficients of the bearing pedestals was studied. It demonstrates
the benefits of using statistical analysis to determine and eliminate insignificant or redundant estimated coefficients from a
ARTICLE IN PRESS
2366 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 15. Schematic arrangement of parameter estimation test rig.

Fig. 16. Excitation force and displacement plots with 5 periods of signal averaging.
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2367

Table 2
Parameter estimation of pedestal dynamics characteristics.

Frequency band used 60–70 Hz


Goodness-of-Fit 0.9952
Stiffness (kN/m)1 134.5
Damping (N s/m)1 3.682*

Note: * denotes insignificant coefficient(s).

Fig. 17. Amplitude and phase transfer functions of pedestal structure.

practical rotor-bearing model. It was observed that the computer simulation and experimental results of a multi-mode
rotor-bearing system responses were insensitive to insignificant estimated stiffness and damping coefficients.
The measured amplitude and phase transfer functions of the pedestal structural system are as shown in Fig. 17.
The natural frequency of the pedestal system was determined at 66 Hz as shown in Fig. 17. With the stiffness coefficient
estimated at 134.5 kN/m, and based on the equation of system’s undamped natural frequency response (27):
rffiffiffiffiffi
k
on ¼
m
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
134:5  103
¼ ¼ 412:6 rad=s or 65:6 Hz (27)
0:79
we could obtain the theoretical fundamental response frequency of 65.6 Hz, which was in excellent agreement with the
experimentally obtained natural frequency of 66 Hz. Hence, in the following radial parameter estimation procedures, the
HMBs’ stiffness coefficient can be determined by subtracting the pedestal support stiffness of 134.5 kN/m from the total
estimated stiffness coefficient.

4.3. Determination of HMBs’ actuator characteristics

The HMBs’ actuator gain is one of the important properties of the magnetic bearing system. With the knowledge of
actuator gain, the parameter estimation technique introduced here can be potentially extended to in-situ identification of
the pump dynamics properties during operating conditions. Furthermore, before the embarkation of the HMBs’ parameter
ARTICLE IN PRESS
2368 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 18. Experimental setup for calibrating actuator gain.

SPHS (V)

Force (N)
PWM Amplifiers HMBs Rotor
I0

Fig. 19. Open loop calibration of HMBs’ actuator gain.

estimation procedure, this section presents the investigation of the maximum displacement linear operating range of the
rotor perturbation limits; or the rotor movement range where the HMBs’ actuator gain does not change significantly. This is
to ensure that the linearised stiffness and damping coefficients adopted in the dynamics model are not violated throughout
the experimental investigations. Any violation, however, will be detected and reflected in the G.o.F calculation in the
parameter estimation algorithm, which is the distinct benefit of employing statistical method.
In order to obtain the HMBs’ actuator gain, some minor setup changes have been made to the test rig shown in Fig. 14.
The photograph of the setup shows that in Fig. 18 has the shaker removed, and the bottom surface of the force transducer is
connected directly onto the pedestal, while its top surface is connected to a height adjusting screw. There are two lock-nuts
going through the height adjusting screw. By fastening the two lock-nuts on both sides of the shaker stand, the pedestal
with the rotor/shaft assembly is simply connected to the force transducer at one point only. Therefore, in this manner, any
HMBs’ actuator force applied onto the pedestal–rotor assembly will be transmitted to the force transducer.

4.3.1. Frequency response of test rig for calibrating actuator gain


The block diagram of the open loop transfer function test of HMBs’ actuator gain is as shown in Fig. 19. The SPHS test
signal is injected into HMBs’ open loop controller summing block, while the force transducer in Fig. 18 measures the
corresponding force (N), from which the transfer function of HMBs’ actuator gain force/voltage (N/V) can be measured.
In order to ensure that the dynamics magnification factor of the test rig does not affect the test results of the actuator
gain, the test rig is excited with a SPHS with bandwidth of 1–500 Hz and a fundamental frequency of 1 Hz to determine its
frequency response. The measured amplitude and phase transfer functions of the HMBs’ actuator gain (force/voltage) are as
shown in Fig. 20.
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2369

Fig. 20. Amplitude and phase transfer functions of actuator gain.

Fig. 21. Flat amplitude transfer function of actuator gain from 1 to 50 Hz.

As can be seen from Fig. 20, the system is sufficiently stiff as its first natural frequency is at 270 Hz. Except for the small
fluctuation in amplitude response at the low-frequency range (1–20 Hz) caused by the larger frequency sampling intervals,
the actuator gain is observed not to be affected by the test rig’s dynamics magnification in the frequency range of 1–100 Hz;
which can be approximated as almost flat and is suitable for calibrating the actuator gain in the following section.
In order to overcome the amplitude response fluctuation in the initial frequency range (1–20 Hz), the bandwidth of
SPHS is decreased to the range of 1–50 Hz with a fundamental frequency of 1 Hz. The corresponding amplitude transfer
function of actuator gain in the range of 1–50 Hz is as shown in Fig. 21.
ARTICLE IN PRESS
2370 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 22. Measured HMBs’ actuator gain at e0 ¼ 0%.

It is clearly shown in Fig. 21 that the actuator gain (force/voltage) is almost flat, being about 5.7 (N/V), in the low-
frequency range 1–50 Hz. Therefore, it can be concluded that the amplitude transfer function of the actuator gain can be
accurately calibrated within this frequency range.

4.3.2. Calibration of HMBs’ actuator gain


In order to measure the HMBs’ actuator gain (force/voltage), the rotor was fixed at the centre position of the stator bore,
which means at a rotor’s static eccentricity e0 ¼ 0%. At e0 ¼ 0%, with the effective magnetic air gap between rotor and stator
being 0.45 mm. SPHS voltage signal amplitude was then gradually increased, and its corresponding excitation force measured
by the force transducer. The seven sets of SPHS and force data are plotted in Fig. 22, from which it can be clearly seen that the
HMBs’ actuator gain at e0 ¼ 0%, is almost linear, with a sensitivity of around 5.8 N/V. Determination of the HMBs saturation
force was not possible during the experiment as the HMBs were found to heat up after 3.5 N of force output.
In order to determine the maximum displacement linear operating range, or insensitivity of the HMBs’ actuator forces
with respect to rotor’s perturbations at different orbital positions, the rotor position was adjusted such that e0 was
gradually increased from 10%, 15%, 20%, 30%, 40% to 50%, respectively. Fig. 23 indicated that maximum displacement linear
operating range of the HMBs is up to 20% (or 90 mm) of the rotor’s static eccentricity ratio. However, the linear operating
range deteriorated from 30% to 50% of e0.

5. Radial experimental parameter estimation of HMBs

5.1. Radial estimation of HMBs modelled with direct- and cross-axes coefficients

This sub-section discusses the design of radial experimental parameter estimation of the HMBs’ eight linearised direct-
and cross-axes coefficients as represented by Eq. (16) and Fig. 11. By rotating the x–y axes orientation of the HMBs system
by 451 as shown in Fig. 24, a single excitation force f is applied in the vertical direction to excite both its direct- and cross-
axes coefficients.
The SPHS bandwidth for the excitation force was set at 1–250 Hz with a fundamental frequency of 1 Hz. The sampling
frequency was 20 kHz for acquiring the force signal f and the corresponding displacement signals in x- and y-directions. The
mass of the pedestal with rotor assembly was weighed at 0.92 kg. During the experiment, several estimation frequency
bands were attempted and a bandwidth of 15 Hz was found to achieve the highest G.o.F in the parameter estimation
procedure, of almost unity in the x- and y-directions as tabulated in Table 3. Since the estimation bandwidth was 15 Hz
with the fundamental frequency of the SPHS at 1 Hz, 15 discrete frequency points were used by the estimation algorithm.
For example, firstly, if the estimation frequency band used were from 65 to 80 Hz, then the stiffness and damping
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2371

Fig. 23. Determination of HMBs’ maximum displacement linear operating range.

Fig. 24. Experimental arrangement of x–y axes, excitation force and displacement probes for estimating eight direct- and cross-axes coefficients.

coefficients would be estimated at the mean frequency of 72.5 Hz. Secondly, the parameter estimation in the frequency
domain was achieved by using a frequency step interval of 3 Hz to obtain 79 sets of the estimation results. Lastly, in order to
ensure that the parameter estimation procedure was performed within the maximum displacement linear operating range
of the HMBs, the maximum rotor displacement was kept within 40 mm in the x- and y-directions.
The measured transfer functions in x- and y-directions with G.o.F in the frequency domain are plotted as shown in Fig. 25.
As can be seen from Fig. 25, the almost similar magnitudes of the transfer functions in the x- and y-directions confirm that
the mechanical design of the HMBs is indeed symmetrical in the x- and y-directions. The accompanying G.o.Fs in the x- and
y-directions are observed to be almost perfect from 40 to 250 Hz. However, the lower G.o.F of 0.65–0.95 in the 0–40 Hz range
is attributed to the integral feedback constant used in Table 1 of the HMBs controller. This is because integral feedback is well
known to eliminate low-frequency position offset error, thus acting as a low pass filter in the control loop [17].
By comparing the structure’s transfer function of Fig. 25 with that in Fig. 17, as shown in Fig. 26, one can conclude that
the introduction of HMBs’ stiffness active control force naturally increases the natural frequency of the system from 66 to
70 Hz.
ARTICLE IN PRESS
2372 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Table 3
Radial parameter estimation of HMBs with 8 direct- and cross-axes coefficients.

Frequency band used 65–80 Hz


Parameter estimated at 72.5 Hz
Goodness-of-Fit in x-direction 0.9999
Goodness-of-Fit in y-direction 0.9999
Kxx (kN/m)1 111.7
Kyy (kN/m)1 46.65
Kyx (kN/m)1 43.87
Kyy (kN/m)1 22.37
Cxx (N s/m)1 66.50
Cxy (N s/m)1 28.00
Cyx (N s/m)1 26.39
Cyy (N s/m)1 13.19*
Forcing angle, a 44.71

Note: * denotes insignificant coefficient(s).

Fig. 25. The measured transfer function in the x and y directions of HMBs system with G.o.F. ‘Transfer function’ is shortened to ‘TF’ in graph.

Fig. 27 shows the four estimated stiffness coefficients of the HMBs system throughout the whole frequency range. The
stiffness coefficients are statistically tested by Eq. (26) to be significant. The direct-axis stiffness coefficients Kxx and Kyy
ranging from 90 to 1400 kN/m are observed to be frequency dependent, that is, their magnitudes increase with frequency.
This observation has been proven theoretically by Aoyama and Okazaki [13], Matsumura et al. [14], and experimentally by
Williams et al. [15].
On the other hand, the four estimated damping coefficients as shown in Fig. 28 are observed to be small and irregular
with respect to frequency, and range from 50 to 200 N s/m. They are statistically tested by Eq. (26) to be mostly
insignificant or can be considered as zero values.
In spite of the perfect G.o.F in the 40–250 Hz range, it can be seen that the direct-axis stiffness coefficients in the initial
frequency range of 0–80 Hz are sometimes smaller in magnitude as compared to their cross-axis terms. This result
contradicts the original intended design of the HMBs mentioned in Section 2.1, that is, the concentration of the actuator
forces should be in the middle of the pole faces or on the direct-axes. Further investigation found that this is due to the
experimental approach being unable to distinguish between the effects of the complex quantities involving the direct- and
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2373

Fig. 26. Transfer function comparison of Fig. 25 to Fig. 17.

Fig. 27. Radial estimation of HMBs with direct- and cross-axes stiffness coefficients.

cross-axes stiffness and damping forces in the x- and y-directions of Eq. (16). These phenomena and limitations of the
method are consistent to observations made by Burrows et al. [24]. The explanation for these phenomena is that, in general,
the stiffness and damping coefficients are direct and coupled, and assuming that the displacement orbit is circular, the
ARTICLE IN PRESS
2374 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 28. Radial estimation of HMBs with direct- and cross-axes damping coefficients.

Fig. 29. Almost symmetrical resultant force vectors applied on mass subjected to circular displacement orbit at 72.5 Hz in the x- and y-directions: (a)
stiffness coefficients and (b) damping coefficients.

force vectors attributed to the stiffness and damping coefficients plotted in Fig. 29, indicated almost symmetry in the x- and
y-directions.
In conclusion, we observed that the parameter estimation procedure employing the least-square method, while being
able to provide a very high G.o.F, is limited in that it can yield erroneous estimates. To overcome this limitation, the direct-
axis support coefficients of the HMBs system will be estimated in the following sections. However, despite the limitation in
the method used, it is clear from the estimated results that the stiffness coefficient is very dominant and frequency
dependent in the dynamics characteristics of the HMBs system.

5.2. Radial estimation of HMBs modelled with only direct-axis coefficients

Earlier in Section 5.1, we identified the distinguishing difficulties associated with the HMBs cross-axis capabilities. This
can be overcome by adopting in the experiment a reduced order dynamics model as shown in Fig. 30, which is the
equivalent of Fig. 12, in the following parameter estimation procedure. Moreover, since the mechanical design of the HMBs
is symmetrical about the x- and y-axes with a concentric rotor, it is reasonable to assume that its direct-axis stiffness and
damping coefficients will be identical in the orthogonal directions, that is, Kxx ¼ Kyy and Cxx ¼ Cyy.
It is in this manner, that is, by using the system equation of motion (21) in the estimation algorithm, that the stiffness
and damping coefficients are successfully estimated as shown in Fig. 31. The estimated coefficients in Fig. 31 must be
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2375

Fig. 30. The decomposition of excitation force in x- and y-directions and HMBs modelled with four direct-axis coefficients.

Fig. 31. Radial estimation of HMBs with only four direct-axis coefficients.

halved for each set of HMB. However, it has topbe ffiffiffi emphasized that the estimated stiffness coefficients are accurately
represented as the pedestal spring stiffness K s = 2 (equivalent of vertical pedestal stiffness in the x- and y-directions)
subtracted from the total estimated stiffness coefficient. Likewise, the stiffness coefficients here are statistically tested to be
significant with a maximum value of 1400 kN/m. This maximum value is similar to those obtained with HMBs modelled
with eight possible coefficients, and the damping coefficients are small and tested to be insignificant. The slightly low G.o.F
from 160 to 250 Hz range is attributed to hysteresis and eddy current losses of the HMBs.
A graph is plotted in Fig. 32 to study the differences in the estimated stiffness coefficients obtained between the direct-
axis and direct- with cross-axes dynamics models. It can be observed that the direct-axis stiffness coefficients estimated
with both models are in good agreement throughout the frequency range. In addition, it is also noted that the direct-axis
model displayed a smoother curve at the lower-frequency range of 0 to 80 Hz, which was quite indistinguishable earlier in
ARTICLE IN PRESS
2376 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 32. Comparison of estimated stiffness coefficients obtained with the direct-axis, and direct- with cross-axes dynamics models.

7000

6000
Stiffness (N/m)

5000

4000 Estimated Value


Calculated Value
3000
2000 Kr = 1694N/m
1650
1000
2 3 4 5 6 7 8 9 10 11 12
Hz

Fig. 33. Comparison of calculated magnetostatic and direct-axis estimated radial stiffness.

Section 5.1 with the direct- and cross-axes model. In conjunction with the G.o.F calculations and intended HMBs
mechanical design, we can therefore conclude that there is no deficiency in HMBs modelled with only the direct-axis
coefficients.

5.2.1. Comparison between magnetostatic and estimated radial stiffness


Comparison of the calculated magnetostatic and estimated radial stiffness using the direct-axis coefficients model is as
shown in Fig. 33. Excellent agreement can be observed at the minimum frequency of 2 Hz used in the parameter estimation
experiment. The radial stiffness found by the parameter estimation method is about 2.6% less than the finite element
prediction. However, their differences increase significantly with frequency above 5 Hz. This observation support that the
stiffness of the HMBs is frequency dependent [17], and that the current approach using the magnetostatic method [26] to
calculate for the stiffness of magnetic bearing is highly inaccurate in the determination of performance for rotating
machinery supported by active magnetic bearings.

5.3. Radial estimation of HMBs modelled with only direct-axis coefficients and with varying kp

This section discusses the investigation of the influences of the PID proportional controller kp on the HMBs stiffness
coefficients. Fig. 34 shows the estimated HMBs stiffness coefficients with different values of kp. Although it can be seen that
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2377

Fig. 34. Radial estimation of HMBs stiffness coefficients with different kp.

the increase in stiffness coefficients’ is marginal with respect to kp ¼ 1.2, the system response and natural frequency is
increased by 10 Hz in the amplitude transfer function plot. In the frequency range of 10–130 Hz, the G.o.F is favorable and
the increase of stiffness is quite distinct. However, at the higher frequency range 130–250 Hz, these properties are not very
indistinguishable, and are likely attributed to the eddy current and hysteresis losses.

5.4. Radial estimation of HMBs modelled with only direct-axis coefficients with rotating rotor

By adopting the same setup as in Fig. 15, this section presents the comparison of the estimated coefficients between
stationary and rotating rotor at a speed of 8000 rpm (133 Hz), by using HMBs modelled with only direct-axis coefficients.
The two estimation results are compared in Fig. 35 where it can be seen that their transfer function, stiffness, damping, and
even the G.o.F are almost identical in magnitude. This implies that there is negligible skin effect in the thickness of the
rotor. A conclusion can therefore be drawn here that parameter estimation procedure on the rotor-bearing system at
stationary position may replace the estimation procedure in rotation. This can greatly facilitate the estimation procedure
ARTICLE IN PRESS
2378 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 35. Comparisons of radial HMBs estimated results modelled with only direct-axis coefficients at 0 and 8000 rpm.

because the rotor does not need to be rotated, and therefore the possibility of coinciding excitation test signal with rotor
rotational speed can be avoided.

6. Axial experimental estimation of HMBs using one DOF model

Due to reaction forces caused by flowing fluid during operations in an axial flow blood pump, the stiffness and damping
coefficients in the axial direction of the HMBs system are of fundamental importance to the performance and stability of
the axial blood pump. This section therefore discusses the investigation of the dynamics characteristics of the pump in the
axial direction. Referring to Fig. 36, the experimental arrangement for parameter estimation in the axial direction is similar
to that of radial parameter estimation, except that the shaker and the eddy current probe are placed horizontally. This
enables the shaker to excite the shaft and rotor assembly axially at its centre axis. The pedestal is fixed at the desired
vertical position by four mechanical clamping rings secured onto the vertical shafts.
In this manner, the horizontal shaft and rotor assembly can slide freely in the horizontal linear bearings in the axial
direction. The photograph of the test rig for axial parameter estimation of the HMBs system is shown in Fig. 36 and its
schematic arrangement is illustrated in Fig. 37.
The schematic experimental arrangement of axial parameter estimation of the HMBs system is as shown in Fig. 38. The
mass of the rotor and shaft assembly is weighed at 0.13 kg and its equation of motion in the z direction can be expressed as

mz€ þ K z z þ C z z_ ¼ f z (28)
As this is a one DOF system, the earlier parameter estimation algorithm formulated with Eq. (8) and estimation
procedure can be utilized to obtain the stiffness Kz and damping Cz coefficients. The force and the displacement signals
were measured on-line and the computed transfer functions data was filled in the matrices of equations (14) and estimated
with (15), respectively. The bandwidth of the SPHS excitation force signal was chosen as 1–250 Hz, with a fundamental
frequency of 1 Hz. The estimation frequency band was set at 10 Hz with an overlapping range of 8 Hz, or with a step interval
of 2 Hz. The estimation results for a discrete frequency point around the natural frequency, given as an example here, are
illustrated in Table 4, and the whole range of estimation results in the frequency domain are shown in Fig. 39. It can be
observed that the stiffness coefficients are statistically tested to be significant and frequency dependent too. It has a
maximum stiffness of 270 kN/m corresponding to 245 Hz. Similar to the radial direction, the estimated damping
coefficients are observed to be very small and range from 1 to 6 N s/m. However, some of the values were statistically tested
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2379

Fig. 36. Experimental arrangement of axial experimental estimation of HMBs.

Fig. 37. Schematic arrangement of axial estimation of HMBs using one DOF model.

Fig. 38. One DOF model of HMBs axial parameter estimation.

Table 4
Sample of parameter estimation results in the axial direction.

Frequency band used 18–28 Hz


Parameter estimated at 23 Hz
Goodness-of-Fit 0.9842
Stiffness (kN/m)1 2.471
Damping (N s/m)1 3.108
ARTICLE IN PRESS
2380 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

Fig. 39. Parameter estimation results of HMBs system in the axial direction.

x 105
15
Radial direction
Stiffness (N/m)

Axial direction
10

0
0 50 100 150 200 250
350
Damping (N.s/m)

300
250 Radial direction
200 Axial direction
150
100
50
0
-50
0 50 100 150 200 250
Hz

Fig. 40. Comparisons of estimation results of the HMBs system between axial and radial directions.

to be significant. The structural natural frequency of the test rig was tested to be about 38 Hz, which explains for the low
G.o.F in the neighborhood frequencies of 30–44 Hz as shown in Fig. 39.
A comparison study is also made between the stiffness coefficients in the radial and axial directions as shown in Fig. 40.
It is observed that the maximum stiffness coefficient attained in the radial direction is five times larger than its axial
counterpart. This is because the axial direction is only passively controlled, while the radial direction is actively controlled.
In addition, the damping coefficients in the axial direction are very much smaller than those in the radial directions.
However, it was observed during the experiment that even when a pen is thrust at the rotor in the axial direction, this
passive control force could still assist in maintaining the rotor position in the pump housing.
ARTICLE IN PRESS
T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382 2381

Z
Y
Y

X X

Fig. 41. An improved model of the hybrid magnetic bearing.

Lastly, in order to improve the efficiency and magnetic force of the HMBs, an improved design shown in Fig. 41 is
currently under investigation in the laboratory. In this design, the two holes used for alignment purposes on the lamination
core are made redundant by epoxy gluing. This is to increase the cross-sectional area of the middle lamination core, and to
decrease the reluctance of the lamination loop. Hence, the magnetic flux can easily flow through the lamination loop.

7. Conclusions

A compact HMBs system has been designed and developed for axial flow blood pumps in this research process. With a
specially designed test rig for parameter estimation, the stiffness and damping coefficients of the HMBs system in the radial
and axial directions have been systematically obtained in the frequency domain accompanied with statistical information.
It is shown that the direct-axis model is adequate and accurate in representing the system dynamics. The Ansoft’s Maxwell
2D/3D finite element software is very accurate in calculating the levitation forces of the proposed HMB. However, we wish
to highlight and illustrate in our work the severe limitations of the current practice of using magnetostatic analysis to
determine the supporting properties of the active magnetic bearing systems. As the name implies, the magnetostatic
analysis is only accurate in calculating the radial stiffness of the HMBs under static conditions. The maximum estimated
stiffness coefficients of the HMBs are 1400 and 270 kN/m in the radial and axial directions, respectively. Both the stiffness
coefficients are found to be frequency-dependent and increase with frequency, dominating its dynamics characteristics.
These behaviours would increase the rotor-bearing system’s natural frequencies with rotor speed, thus explaining why
rotor systems supported by active magnetic bearings could operate at very high speed and are only limited by the amplifier
and actuator outputs. By injecting a multi-frequency excitation signal through the HMBs, the actuator gain is determined to
be linear and has a sensitivity of 5.8 N/V, and a maximum displacement linear operating range of e0 ¼ 20%. This
understanding may potentially extend the parameter estimation technique developed here to the identification and
monitoring of the pump’s dynamics properties under normal operating conditions with fluid.

References

[1] T.M. Lim, S.B. Cheng, D.S. Zhang, A compact magnetic bearing system for axial flow blood pump, IMECS 2007: 21–23 March, 2007, Hong Kong, 2007.
[2] T.M. Lim, S.B. Cheng, The development of a hybrid magnetic bearings system for axial flow blood pump, in: Advances in Industrial Engineering and
Operations Research. Series: Lecture Notes Electrical Engineering. vol. 5, Springer, Berlin, 2008, pp. 391–399.
[3] T.M. Lim, D.S. Zhang, Development of Lorentz force type self-bearing motor for an alternative axial flow blood pump design, Artificial Organs 30
(2006) 347–353.
[4] H. Hoshi, J. Asama, T. Shinshi, S. Takatani, K. Ohuchi, M. Nakamura, T. Mizuno, H. Arai, A. Shimokohbe, S. Takatani, Disposable magnetically levitated
centrifugal blood pump: design and in vitro performance, Artificial Organs 29 (2005) 520–526.
[5] H.M. Loree, K. Bourque, D.B. Gernes, J.S. Richardson, V.L. Poirier, N. Barletta, A. Fleischli, G. Foiera, T.M. Gempp, R. Schoeb, K.N. Litwak, T. Akimoto, M.
Kameneva, M.J. Watach, P. Litwak, The HeartMate W: design and in vivo studies of a maglev centrifugal left ventricular assist device, Artificial Organs
25 (2001) 386–391.
[6] J. Asama, T. Shinshi, H. Hoshi, S. Takatani, A. Shimokohbe, A new design for a compact centrifugal blood pump with a magnetically levitated rotor,
Journal of ASAIO 50 (2004) 550–556.
[7] P.E. Allaire, E. Hilton, M. Baloh, et al., Performance of a continuous flow ventricular assist device: magnetic bearings design, construction, and testing,
Artificial Organs 22 (1998) 475–480.
[8] H. Hoshi, T. Shinshi, S. Takatani, Third-generation blood pumps with mechanical noncontact magnetic bearings, Artificial Organs 30 (2006) 324–338.
[9] M.J. Baloh, P.E. Allaire, E.F. Hilton, N. Wei, O.B. Olsen, G.B. Bearnson, P.S. Khanwikar, Characterization of a magnetic bearing system and fluid
properties for a continuous flow ventricular assist device, Artificial Organs 23 (1999) 792–796.
[10] J. Asama, T. Shinshi, H. Hoshi, S. Takatani, A. Shimokohbe, Dynamic characteristics of a magnetically levitated impeller in a centrifugal blood pump,
Artificial Organs 31 (2007) 301–311.
ARTICLE IN PRESS
2382 T.M. Lim et al. / Mechanical Systems and Signal Processing 23 (2009) 2352–2382

[11] M.K.H. Chung, N. Zhang, G.D. Tansley, Y. Qian, Experimental determination of dynamic characteristics of the VentrAssist implantable rotary blood
pump, Artificial Organs 28 (2004) 1089–1094.
[12] T.M. Lim, S.B. Cheng, Parameter estimation of one-axis magnetically suspended system with a digital PID controller, in: Proceedings of First
International Conference on Sensing Technology, New Zealand, 2005, pp. 419–424.
[13] Y. Aoyama, Y. Okazaki, Active magnetic bearing for spindle, in: Proceedings of Fourth International Workshop on Rare Earth Cobalt Permanent
Magnets and Their Applications, Japan, 1979, pp. 169–176.
[14] F. Matsumura, Y. Tanaka, M. Kido, Y. Akiyama, Composition of a magnetic-bearing system for horizontal shaft and its experimental results, Electrical
Engineering in Japan 103 (1983) 121–128.
[15] R.D. Williams, F.J. Keith, P.E. Allaire, Digital control of active magnetic bearings, IEEE Transactions on Industrial Electronics 37 (1990) 19–27.
[16] T.M. Lim, D.S. Zhang, Control of Lorentz force-type self-bearing motors with hybrid PID and robust model reference adaptive control scheme,
Mechatronics 18 (2008).
[17] T.M. Lim, S.B. Cheng, Parameter estimation and statistical analysis on frequency-dependent active control forces, Mechanical Systems and Signal
Processing 21 (2007) 2112–2124.
[18] M.R. Schroeder, Synthesis of low-peak-factor signals and binary sequences with low auto correlation, IEEE Transactions on Information and Theory
(1970) 85–89.
[19] Ansoft Maxwell 3D Field Simulator v11 User’s guide: ANSOFT Corporation.
[20] Getting Started: A 2D Magnetostatic Problem: ANSOFT Corporation.
[21] T.M. Lim, D.S. Zhang, Numerical analysis of blood trauma in an enclosed-impeller axial flow pump, 13th Congress of the ISRBP, September 2005,
Tokyo, Japan, 2005.
[22] L.P. Chua, B.Y. Su, T.M. Lim, T.M. Zhou, Numerical simulation of an axial blood pump, Artificial Organs 31 (2007) 560–570.
[23] D. Gujarati, Essential of Econometrics, McGraw-Hill International Editions, 1992.
[24] C.R. Burrows, N.C. Kucuk, M.N. Sahinkaya, R. Stanway, Linearised squeeze film dynamics: model structure and the interpretation of experimentally
derived parameters, Proceedings of the Institution of Mechanical Engineers Part C 204 (1990) 263–272.
[25] T.M. Lim, G.B. Chai, Validating the dynamic coefficients of bearing pedestals in a multi-mode rotor-bearing system, Proceedings of the Institution of
Mechanical Engineers Part C 223 (2009), doi:10.1243/09544062JMES1161.
[26] G. Schweitzer, A. Traxler, H. Bleuler, Active Magnetic Bearings, Verlag der Fachvereine, 1994.

You might also like