You are on page 1of 11

HYDRODYNAMIC EFFECTS ON MODELING AND CONTROL OF A

HIGH TEMPERATURE ACTIVE MAGNETIC BEARING PUMP WITH A


CANNED ROTOR

Alexander M. Melin, Roger A. Kisner, David L. Fugate, and David E. Holcomb∗


Oak Ridge National Laboratory
1 Bethel Valley Road
Oak Ridge, TN 37831
melina@ornl.gov; kisnerra@ornl.gov; fugatedl@ornl.gov; holcombde@ornl.gov

ABSTRACT

Embedding instrumentation and control (I&C) at the component level in nuclear power plants can im-
prove component performance, lifetime, and resilience by optimizing operation, reducing the constraints
on physical design, and providing on-board prognostics and diagnostics. However, there are difficulties
associated with using embedded I&C in nuclear power plants (NPP). First is that regulatory constraints
make deployment of new technologies difficult. This difficulty is partially alleviated by developing and
refining these technologies for passively safe reactor designs. The second challenge is the extreme envi-
ronments that some components that would benefit from embedded I&C operate in. In this paper, we will
use a high temperature (700 ◦ C) canned rotor pump as an ideal testbed for the development of embedded
I&C that can operate in extreme environments. This pump utilizes active magnetic bearings (AMBs)
to eliminate the need for mechanical bearings and shaft seals. Successfully utilizing embedded I&C in
extreme environments requires developing a deep understanding of the system’s physics and dynamics
and using that knowledge to overcome material and performance limitations on actuators and sensors
imposed by the environment. In this paper, we will develop a coupled dynamic model of a canned rotor
pump that incorporates hydrodynamics, rotordynamics, and active magnetic bearing dynamics. Then
we will compare two control design methods, one that uses a simplified decoupled model of the system
and another that utilizes the full coupled system model. It will be seen that utilizing all the available
knowledge of the system dynamics in the controller design yield an order of magnitude improvement
in the magnitude of the magnetic bearing response to disturbances at the same level of control effort, a
large reduction in the settling time of the system, and a smoother control action. Key Words: Embedded
I&C, Active Magnetic Bearings, Advanced I&C, Extreme Environments, Hydrodynamics

1. INTRODUCTION

Typically magnetic bearings are used for high speed operation for manufacturing and flywheels energy stor-
age, and the rotor operates in air or vacuum. However, the unique characteristics of magnetic bearings
means that they can be used in fluid pumping applications to eliminate the need for pump seals and me-
chanical bearings entirely. This removes the main contributors to maintenance costs and pump failure [1].
For pumps operating in extreme environments pumping harsh chemicals or pumps that need to maintain

This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the U.S. Depart-
ment of Energy. The United States Government retains and the publisher, by accepting the article for publication, acknowledges
that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the
published form of this manuscript, or allow others to do so, for United States Government purposes. The Department of En-
ergy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan
(http://energy.gov/downloads/doe-public-access-plan).
high purity of the pumped fluid, the elimination of seals, bearings, and lubrication is particularly desirable.
Recently, Shanghai blower works built and tested a helium fan for an HTR-PM test platform that utilizes
magnetic bearings to levitate the 4000 kg shaft [2]. Previously, the gas turbine modular helium reactor used
magnetic bearings to suspend the turbine shaft [3]. Pumps operating in harsh environments utilize a protec-
tive layer or ‘can’ around the stator and rotor to prevent fluid contact with the incompatible materials [4, 5].
The pump design described in this paper utilizes both active magnetic bearings and a thin protective can
around the rotor and stator.

While a canned pump design with active magnetic bearings (AMB) can improve performance, the extreme
environments they would be deployed in make designing embedded I&C for the magnetic bearings diffi-
cult [6,7]. In this particular case, implementing embedded I&C reduces some of the mechanical engineering
challenges of the pump and bearing design and increases the challenges in the mechanical and electromag-
netic design of the actuators and sensors. The extreme environment creates limitations on material selection
for sensor and actuators, degrades the performance of sensors and actuators, and can dramatically change the
electrical and mechanical properties of the sensors and actuators over the operating range of the component.
Additionally, proper utilization of these sensors and actuators during degraded performance requires more
advanced control design techniques than are typically employed in less challenging situations.

Advanced control design techniques can take advantage of more advanced models of the system dynamics
over a wide range of operating conditions. In this particular case, the large hydrodynamic forces on the rotor
arising from the thin layer of fluid between the rotor and stator can destabilize a controller if they are not
included in the model of the system dynamics used to design the controller.

Control design for magnetic bearings has been extensively studied and advanced control techniques such as
optimal control [8] and robust control techniques have been applied [9]. To the authors knowledge however,
the effect of hydrodynamics on the rotordynamics and control system performance has not been studied.
In the following, a rigid body rotordynamic model will be coupled with hydrodynamics forces on the rotor
caused by the thin fluid layer between the rotor and stator and a model of the active magnetic bearing
dynamics is developed. These hydrodynamic forces cause large coupling forces between the radial axes
of the rotor similar to mass imbalances but with larger magnitude. This makes the design of decoupled
controllers for each bearing axis conservative. We will show that a control designed using knowledge of the
coupling forces provides better performance than the uncoupled controllers.

2. ROTORDYNAMICS MODEL

In this section models for the rotordynamics of the shaft are described. Section 2.1 develops the basic
rotordynamic model treating the rotor as a rigid body. Section 2.2 derives the equations that govern the
hydrodynamics of the fluid layer between the rotor and stator on the rotordynamics. Section 2.3 develops
the linearized model of the active magnetic bearings and finally, section 2.4 presents the rotordynamic model
with the hydrodynamics and magnetic bearing dynamics included.

The pump design used as the basis for the derivation of the dynamic model is a 10 kW pump testbed and
it consists of two radial magnetic bearings A and B, a switched reluctance motor, and two axial magnetic
bearings. Figure 1 shows the rotor geometry for the testbed design.
Figure 1. A cross section of the canned rotor geometry for a 10 kW testbed preliminary design. All
dimensions are in meters.

2.1. Rigid Rotor Dynamics

The section presents the standard rigid body rotordynamics model similar to those presented in [10] and [11].
Normally, models of rotordynamics are concerned with the frequency response of the rotor due to the specific
rotor geometry and material properties [12]. However, when using active magnetic bearings, their effective
stiffness is much lower that conventional mechanical bearings and lower than the rotor stiffness. This allows
the rotor to be treated in some cases as a rigid body [10] as is the case with the testbed design. There are
two coordinate frames that are useful in the development of the rigid body dynamics. The rotor states in the
center of mass (COM) frame are given by z = [x, y, θx , θy ] at the rotor’s COM, where x and y are the two
radial translations and θx and θy are the rotations around the x and y axes, respectively. Figure 2 shows the
axis orientation relative to the rotor COM.

Figure 2. Principal axes corresponding to the principal moments of inertia for the rotor.

While defining the rotor movement and position at the COM simplifies the specification of the equations of
motion, the AMBs operate away from the COM. Defining the bearing reference frame, the rotor states are
the coordinates of the rotor centerline located at the AMBs denoted zb = [xA , yA , xB , yB ]T . Figure 3 shows
the location of the AMBs relative to the COM.

Figure 3. The location of the magnetic bearings in relationship to the COM.

To translate between the two coordinate frames, the small angle theorem can be utilized to derive the linear
transformation matrix B T given by
 
1 0 −L1 0
 0 1 0 −L1 
BT =   1 0 L2
, (1)
0 
0 1 0 L2

where L1 ≥ 0 and L2 ≥ 0 are shown in Figure3 and (·)T denotes a matrix transpose. The transformation
from the COM states z to the AMB states zb is given by zb = B T z and the inverse transformation is given
by z = Bzb .

The rigid rotor equations of motion will be implemented using a 4 degree of freedom (DOF) system. The
4 DOF model only includes radial translation and non-axial rotation and assumes a constant rotor speed θ̇z .
The rotor position is given by the rotor COM states z. The equation of motion is given by

M z̈ + (G + Cb )ż + (K + Kb )z = u(t), (2)

where M is the symmetric mass/inertia matrix, G is the skew-symmetric gyroscopic matrix, Cb = CbT ≥ 0
is the symmetric positive definite damping matrix, K is the symmetric stiffness matrix, Kb = −KbT is the
nonconservative stiffness matrix, and u(t) are the external forces including the AMB suspension forces. The
damping term Cb and the stiffness term Kb are due to hydrodynamic effects of the fluid film between the
rotor and the stator. These matrices will be calculated in section 2.2. Finally, for AMBs, the matrix K = 0
because there is no physical contact between the rotor and a stationary mechanical bearing. This lack of
physical stiffness is compensated for with the AMBs and their impact on the rotordynamics will be derived
in section 2.3

To develop the mass matrix and gyroscopic matrix for the rotor, a parametric model of rotor imbalances is
used. Imbalances in the rotor will cause the COM to shift slightly, which will cause coupling between the
x and y axes during rotation. This can be modeled as a small mass ∆m attached to the rotor at location
(δx , δy , δz ). This yields a mass matrix given by
 
m + ∆m 0 0 0
 0 m + ∆m 0 0 
M = 2 2
, (3)
 0 0 Ixx + ∆m(δy + δz ) −∆m(δx δy ) 
0 0 −∆m(δx δy ) Iyy + ∆m(δz 2 + δx 2 )
and the gyroscopic matrix given by
0 0 0 0
 
 0 0 0 0 
G= . (4)
 0 0 0 (Izz + ∆m(δx 2 + δy 2 ))θ̇z 
0 0 −(Izz + ∆m(δx 2 + δy 2 ))θ̇z 0
where Ixx , Iyy , and Izz are the principle moments of inertia of the rotor around the x, y, and z axes respec-
tively. It can be seen from the cross-coupling terms −∆m(δx δy ) in the mass matrix that an imbalance in the
rotor will cause a steady state circular motion of the geometric COM. Additionally, note that the gyroscopic
matrix is a function of the rotor rotational speed θ̇z .

2.2. Hydrodynamic Model

The rotor in a canned rotor pump design is fully immersed in the fluid being pumped with a thin fluid layer
between the rotor can and stator can. This causes rotor fluid interactions which are a major source of external
forces on the rotor. The hydrodynamics will generate cross-coupling forces between the radial axes with
significant magnitude. If unmodeled these cross-coupling forces can cause control system instability.

During normal operation, the rotor and stator geometric centers are aligned by the control system and the
fluid forces are tangential to the surface of the rotor and are caused by the velocity differential between
the rotor and stator cans. However, any small movement of the rotor from geometric center caused by
forces on the impeller, motor and mass imbalances, or other disturbances will immediately develop radial
hydrodynamic forces on the rotor. These forces are functions of rotor movement and the rotational speed.

The model of the tangential forces on the rotor due to viscous effects from the angular velocity differential
is equivalent to the standard model of two large plates moving relative to each other. For large identical flat
plates moving relative to each other, the shear force due to fluid viscosity is given by
µAv
F = , (5)
y
where µ is the fluid viscosity, A is the area of one plate, v is the plate relative velocity of the plates, and y
is the distance between the two plates. If the plates are infinite, this is analogous to the cylindrical Couette
flow under the assumption that the edge effects are insignificant compared to the surface effects [13]. The
cylindrical version of Eg. (5) is given by
2πµr3 L
 
T = θ̇z = dz θ̇z , (6)
s0
where r is the radius of the rotor, L is the length of the rotor, s0 is the nominal gap between the rotor and
stator, and θ˙z is the axial rotational velocity in rad/s. [14] The torque given in equation 6 only describes the
resistance to rotation; to determine the effect of transverse rotor motion, a different approach is needed.

Under dynamic loading conditions, the rotor and stator will not remain concentric which leads to changes
in the rotational fluid damping and the addition of damping in the x and y directions. The forces due
to the dynamic motion of the rotor about the origin due to fluid bearing effects can be described by the
dimensionalized stiffness and damping matrices coefficients Kij and Cij respectively. Their relationship to
the axial force on the rotor is given by
       
Fx C11 C12 ẋ K11 K12 x
= + . (7)
Fy C21 C22 ẏ K21 K22 y
To simplify the derivation of the stiffness and damping matrices, the coefficients will be derived as non-
dimensional coefficients. The relationships between the dimensionalized coefficients of the stiffness matrix,
damping matrix, and non-dimensional stiffness and damping matrices coefficients are given by
F0 ∗ F0 ∗
Kij = K , Cij = C , i, j = 1, 2, (8)
s0 ij ωs0 ij
where s0 is the nominal gap between the rotor and stator cans and

∗ µL3 ωri ∗ µL3 ωri ε π 2 − π 2 ε2 + 16ε2
F0 = Fµ S = S = , (9)
2s20 2s20 2(1 − ε2 )2
p
is the static load at a given deflection e where e = εs0 = x2 + y 2 , and 0 ≤ ε ≤ 1 is the eccentricity
ratio [15].

The values of the non-dimensional stiffness matrix are given by [15]



K11 = [2π 2 + (16 − π 2 )ε2 ]Ψ(ε),
π π 2 − 2π 2 ε2 − (16 − π 2 )ε4
 

K12 = Ψ(ε),
4 ε(1 − ε2 )1/2
π π 2 + (32 + π 2 )ε2 + (32 − 2π 2 )ε4
  (10)

K21 = − Ψ(ε), and
4 ε(1 − ε2 )1/2
∗ π 2 + (32 + π 2 )ε2 + (32 − 2π 2 )ε4
K22 = Ψ(ε),
1 − ε2
and the values of the non-dimensional damping matrix are given by [15]

∗ π(1 − ε2 )1/2 2
C11 = [π + (2π 2 − 16)ε2 ]Ψ(ε),

∗ ∗
C12 = C21 = −[2π 2 + (4π 2 − 32)ε2 ]Ψ(ε), and (11)
π π 2 + (48 − 2π 2 )ε2 + π 2 ε4
 

C22 = Ψ(ε),
2 ε(1 − ε2 )1/2
where
4
Ψ(ε) = . (12)
(π 2 + (16 − π 2 )ε2 )3/2
This analysis in standard journal bearings is utilized at nominal rotor positions of ε > 0 due to the constant
loading on the bearing. In the case of the canned rotor with AMBs levitating the shaft, the eccentricity ratio
ε → 0 as t → ∞. For small perturbations around the origin, the damping is given by limε→∞ Cij (ε, ω) and
the stiffness by lime→∞ Kij (ε, ω) for i, j = 1, 2.

Note that the fluid stiffness and damping matrices only assume radial motion of the rotor and do not take
into account rotation of the rotor axis about the COM. The effect of rotor axis rotational movement can
be approximated by applying the fluid stiffness and damping matrices in the AMB coordinate system and
transforming the matrices to the COM coordinate system. The hydrodynamic force equations in the AMB
coordinate system are
       
FxA K11 K12 0 0 xA C11 C12 0 0 ẋA
 FyA 
 =  K21 K22 0 0    yA  +  C21 C22 0 0   ẏA  (13)
     

 FxB   0 0 K11 K12   xB   0 0 C11 C12   ẋB 
FyB 0 0 K21 K22 yB 0 0 C21 C22 ẏB
Fb = Kb zb + Cb żb . (14)
Finally, this is transformed to the COM coordinate system by
F = Fb B T = Kb B T z + Cb B T ż. (15)

2.3. Active Magnetic Bearing Model

The forces generated by the magnetic bearings are integral to the rotordynamics. They enter into the rotor
equations of motion through the input term u of equation 2 as the forces created by each bearing. These
forces are dependent on the rotor position and the coil current for each bearing pole. To be included in the
state-space model of the rotordynamics, the bearing forces are linearized with respect to the rotor radial
movement and the currents of each electromagnet [11]. The linearized force equation for each separate coil
can be written as
f (x, i) = ki i − ks x. (16)
The two constants ki and ks are calculated by linearizing the electromagnetic force equation around the
operating points i0 and s0 which are the nominal coil current and nominal airgap respectively. Assuming
that the bearing has rotational symmetry, the constant ki related to the current is given by

∂f 1 2 i
ki = = µ0 N A 2 cos Θ0 . (17)
∂i 2
i=i0 ,s=s0 s i=i0 ,s=s0
Likewise, the spring constant ks related to the rotor movement is given by
2

∂f 1 2 i

ks = = − µ0 N A 3 cos Θ0 . (18)
∂s
i=i0 ,s=s0 2 s i=i0 ,s=s0
where µ0 is the magnetic permeability of vacuum, N is the number of turns in the coil, A is the cross
sectional area of the flux path in the stator, and 2Θ0 is the angle between the teeth of each pole. Transforming
the bearing forces f (x, i) from the bearing coordinate system to the COM coordinate system yields
u(t) = −BKs B T z(t) + BKi i(t) (19)
where Ks and Ki are the diagonal matrices of the linearized magnetic bearing coefficients ks and ki respec-
tively.

2.4. Combined Model

In the center of mass coordinate system, combining equations 19, 15, 4, 3, and 2 yields the second order
ODE describing the motion of the rotor. After the appropriate transformations are applied, the second-order
linearized ordinary differential equation (ODE) in COM coordinates is given by
M z̈(t) + (G + Cb B T )ż(t) + (Kb B T + BKs B T )z(t) = BKi i(t). (20)
Transforming Eg. (20) into a state-space formulations yields
" #
~04×4 ~4×4 ~04×4
 
~ ~ I
ẋ(t) = Ax(t) + Bu(t) = x(t) + u(t) and
−M −1 (Kb B T + BKs B T ) −M −1 (G + Cb B T ) M −1 BKi
h i
~
y(t) = Cx(t) = B T ~04×4 x(t),
(21)
where x(t) = [x, y, θx , θy , ẋ, ẏ, θ˙x , θ˙y ]T is the system state, y(t) is the vector of position sensor measure-
ments in the bearing coordinate system, and I~4×4 is an identity matrix. The control variable u(t) in this
formulation is the bearing coil current vector i(t).
3. CONTROL SYSTEM DESIGN

In this section two different control design strategies are implemented to demonstrate the advantages of using
the knowledge of the physical system to improve the disturbance response of the feedback controller. Both
a proportional integral derivative (PID) controller and a linear quadratic regulator (LQR) with integrator
were designed using the linearized model of the rotordynamics and simulated using the complete nonlinear
dynamics of the system. The response of each controller was tested with a 106 N static load that occurs one
second after the simulation begins.

A PID controller is a standard control design methodology that is used in a wide variety of devices including
magnetic bearings. When applied to magnetic bearings it is assumed that each axis is decoupled. For
bearings without liquid between the rotor and stator causing large cross-coupling between the rotor axes in
the rotor equations of motion, this is a reasonable assumption. However as will be demonstrated with the
simulations, utilizing a full state LQR controller that does not assume decoupled axes can provide much
better performance in the presence of a system with high coupling forces. For brevity only the translational
response of the rotor center of mass is shown along with the current input for a single axial magnetic
bearing, but the rotational response for the LQR controller show similar improvements in performance to
the translational response.

Figure 4. Rotor COM translational movement to a constant disturbance on the rotor with a PID
controller.

Similarly for the LQR controller,


Figure 5. Coil input currents in amps for the AMB denoted by B using a PID controller. Physically,
the B bearing is on the opposite end of the shaft from the impeller.

Figure 6. Rotor COM translational movement to a constant disturbance on the rotor with an LQR
controller with an integrator.
Figure 7. Coil input currents in amps for the AMB denoted by B using an LQR controller with an
integrator. Physically, the B bearing is on the end of the shaft nearest to the impeller.

4. CONCLUSIONS

Embedded I&C has the potential to enable previously unattainable NPP component performance. Embedded
I&C is being implemented in other industrial applications and in the non-nuclear elements of nuclear power
plants (i.e. turbine control systems). NPPs with significant passive safety elements enable application
of more advanced techniques to control components without significant impact on plant operational risk.
Developing these techniques require developing more advanced models of the dynamics of the components
so that these models can be used to design controllers with improved performance and stability margins. In
this paper, a dynamic model of a rigid rotor was developed that includes both the dynamics of the active
magnetic bearings and the hydrodynamic effects of the this fluid layer between the rotor and stator on
the rotordynamics. This model was then used to develop two controllers, one of which assumes the axes
are decoupled, and one that doesn’t use this simplifying assumption. Through simulations, the full rank
LQR controller with integrator that incorporates the cross-coupling effects in the controller design showed
significant performance improvements over the decoupled PID controllers.

ACKNOWLEDGMENTS

This work was sponsored by the US Department of Energy (DOE) Nuclear Energy Enabling Technologies
(NEET) program on Advanced Sensors and Instrumentation.
REFERENCES

[1] H. P. BLOCK, “Root Cause Analysis of Five Costly Centrifugal Pump Failures,” Proc. 7th Interna-
tional Pump User Symposium, 1990.
[2] “Helium Fan Produced for Chinese HTR-PM,” World Nuclear News, 2014, http://www.world-nuclear-
news.org/NN-Helium-fan-produced-for-Chinese-HTR-PM-1908144.html.
[3] “Preliminary Assessment of the GT-MHR Power Conversion System,” Rolls Royce, 2007,
DNS133534 v1.0.
[4] B. J. BLAIR, R. R. HUMPHRIS, P. E. ALLAIRE, D. W. LEWIS, and L. E. BARRETT, “A Canned
Pump with Magnetic Bearings for Industrial Use - Laboratory Testing,” Proc. Energy Conversion
Engineering Conference, 1990.
[5] D. M. KITCH, J. M. KUJAWSKI, D. R. FARRUGGIA, J. L. MATOS, and C. T. FARR, “Nuclear
Reactor High Temperature Spool Pump,” Patent, US 6,813,328 B2, 2004.
[6] D. L. FUGATE, A. M. MELIN, R. A. KISNER, and J. WILGEN, “Advanced Instrumentation for Ex-
treme Environments: Nuclear Reactor Environments,” Proc. Future of Instrumentation International
Workshop, 2012.
[7] A. M. MELIN, R. A. KISNER, and D. L. FUGATE, “Advanced Instrumentation for Extreme Environ-
ments,” IEEE Instrumentation & Measurement Magazine (2013).
[8] T. SCHUHMANN, W. HOFMANN, and R. WERNER, “Improving Operational Performance of Ac-
tive Magnetic Bearings Using Kalman Filter and State Feedback Control,” IEEE Transactions on
Industrial Electronics, 59 (2012).
[9] S.-L. CHEN and C.-C. WENG, “Robust Control of a Voltage-Controlled Three-Pole Active Magnetic
Bearing System,” IEEE Transactions on Mechatronics, 15 (2010).
[10] G. SCHWEITZER, “Dynamics of the Rigid Rotor,” in Magnetic Bearings: Theory, Design, and
Application to Rotating Machinery, edited by G. SCHWEITZER and E. H. MASLEN, Springer, 2010.
[11] A. CHIBA et al., Magnetic Bearings and Bearingless Drives, Elsevier (2005).
[12] A. MUSZYNSKA, Rotordynamics, Taylor and Francis Group, LLC, Boca Raton, FL (2005).
[13] B. R. MUNSON, D. F. YOUNG, and T. H. OKIISHI, Fundamentals of Fluid Mechanics, John Wiley
& Sons (1998).
[14] M. D. HERSEY, Theory and Research in Lubrication, John Wiley & Sons (1966).
[15] U. YÜCEL, “Calculation of the Dynamic Coefficients for Fluid Film Journal Bearings,” Journal of
Engineering Sciences (2005).

You might also like