HOME JOURNALS CONTACT

Asian Journal of Scientific Research

Year: 2008 | Volume: 1 | Issue: 4 | Page No.: 324-337
DOI: 10.3923/ajsr.2008.324.337
Speed Control for IFOC Induction Machine with Robust Sliding Mode Controller
Noaman M. Noaman

Abstract: In this study, an indirect Field-Oriented Control (IFOC) induction machine drive with a conventional PI and sliding mode controllers is presented. The robustness of ac machine drive speed performance with these controllers is checked in terms of variation of machine parameters. The design includes rotor speed estimation from measured stator terminal voltages and currents. The estimated speed is used as feedback in an indirect vector control system, such that the speed control is performed without the use of shaft mounted transducers. The high performance of the proposed control schemes under load disturbances is studied via simulation cases. The components of the speed controlled indirect field-oriented induction machine with the both controllers are simulated using SIMULINK, while the dynamic of induction machine is simulated using the potential of S-function block and its attached script file.

Fulltext PDF Fulltext HTML

How to cite this article
Noaman M. Noaman , 2008. Speed Control for IFOC Induction Machine with Robust Sliding Mode Controller. Asian Journal of Scientific Research, 1: 324-337.

Keywords: IFOC induction machine drive, PI controller, sliding mode controller, robustness of ac machine drive, speed performance and parameters variation

INTRODUCTION

Field Orientation Control (FOC) or vector control of induction machine achieves decoupled torque and flux dynamics leading to independent control of the torque and flux as for a separately excited DC motor (Denai and Attia, 2002). This is achieved by orthogonal projection of the stator current into a torque-producing component and flux-producing component. This technique is performed by two basic methods: direct and indirect vector control. With direct field orientation, the instantaneous value of the flux is required and obtained by direct measurement using flux sensors or flux estimators, whereas indirect field orientation is based on the inverse flux model dynamics and there are three possible implementation based on the stator, rotor, or air gap flux orientation. The rotor flux indirect vector control technique is the most widely used due to its simplicity (Attia and Denai, 2002). FOC methods are attractive but suffer from one major disadvantage. They are sensitive to parameter variations such as rotor time constant and incorrect flux measurement or estimation at low speeds. Consequently, performance deteriorates and a conventional controller such as a PI is unable to maintain satisfactory performance under these conditions (Denai and Attia, 2002).

A Sliding Mode Control (SMC) is basically an adaptive control scheme that gives robust performance of a drive with parameter variations. The control is nonlinear and applied to linear and nonlinear plant. In SMC, the drive response is forced to slide along a predefined trajectory in a phase plane by a switching control algorithm, irrespective drive`s parameter variation and load disturbance (Edwards and Spurgeon, 2002; Barambones et al., 2002).

Controlled induction motor drives without mechanical speed sensors at the motor shaft have the attractions of low cost and high reliability. To replace the sensor, the information on the rotor speed is extracted from measured stator voltages and currents at the motor terminals. Vector-controlled drives require estimating the magnitude and spatial orientation of the fundamental magnetic flux waves in the stator or in the rotor (Barambones et al., 2002; Bose, 2001).

In this study, two control strategies are considered to adjust the speed of the drive system: PI and sliding mode controller. The robustness of these suggested controllers are checked in terms of motor parameter variations. The speed controller is performed under no mechanical speed sensors and speed observer, based on the software program, is adopted for this purpose.

Dynamic Model of Induction Machine
Induction Machine (IM) equations in arbitrary rotating reference frame can be represented in stator and rotor dq voltage equations (Bose, 2001; Leonhard, 1996):

(1)

where, v is voltage; λ is the flux linkage; i is the current; ω is the arbitrary speed of the reference frame; r is the resistance and p is the time derivative. The subscript r and s denotes the rotor and stator values, respectively referred to the stator and the subscripts d and q denote the dq-axis components in the arbitrary reference frame.

The equations of the machine in the stationary and synchronously rotating reference frame can be obtained from (1) by setting ω to zero and ω = ωe, respectively. To distinguish these two frames from each other, an additional superscript will be used; s for stationary frame variables and e for synchronously rotating frame variables.

The electromagnetic torque equation can be given by (Bose, 2001; Chee-Mun Ong, 1998):

(2)

where, P denotes the number of machine pole pairs. Using Eq. 1 and 2, one can obtain the state-space model for induction motor developed in stationary reference frame as given below (Bose, 2001; Ouhrouche and Volat, 2000; Akin, 2003):

(3)

(4)

where, and . The parameters are rotor, stator and main inductances, respectively. is the rotor time constant; B is the viscous friction coefficient; J is the inertia constant of the motor; TL is the external load; ωr is the rotor electrical speed in angular frequency.

Fig. 1: Phasor diagram explaining indirect vector control

Indirect Field Orientation Control (IFOC)
Indirect vector control is very popular in industrial applications. Figure 1 explains the fundamental principle of indirect vector control with the help of a phasor diagram. The axes are fixed on the stator, but the dr-qr axes, which are fixed on the rotor, are ds-qs moving at speed ωr. Synchronously rotating axes de-qe are rotating ahead of the dr-qr axes by the positive slip angle θsl corresponding to slip frequency ωsl. Since the rotor pole is directed on the de axis and ωe = ωr + ωsl, one can write:

(5)

The phasor diagram suggests that for decoupling control, the stator flux component of current should be aligned on the de axis and the torque component of current should be on the qe axis, as shown. For decoupling control, one can make a derivation of control equations of indirect vector control with the help of de-qe dynamic model of IM, i.e., using Eq. 1 with the addition of superscript e to the variables and setting ω = ωe. If de-axis is aligned with the rotor field, the q-component of the rotor field, , in the chosen reference frame would be zero. One can easily show the following important equations:

(6)

(7)

(8)

To implement the indirect vector control strategy, it is necessary to use the condition in Eq. 6 -8 in order to satisfy the condition for proper orientation. Figure 2 shows an indirect field-oriented control scheme for a current controlled PWM induction machine motor drive.

Fig. 2: Indirect field-Oriented control of a current regulated pwm inverter induction motor drive

The command values for the abc stator currents can then be computed as follows:

(9)

(10)

Motor Speed Calculation
Figure 2 shows the connection of the speed estimator with IFOC IM. The block calculates the synchronous and rotor speed based on the measurements of line voltages and currents. Starting from the flux equations (Denai and Attia, 2002; Barambones et al., 2002):

(11)

the expressions for and can be obtained as:

(12)

Substituting of (12) in the drive voltage equations, Eq. 1, gives:

(13)

Hence,

(14)

(15)

where, and s is the Laplace operator. Equation 14 and 15 represent the rotor flux observers and are termed the voltage model and the current model, respectively. The rotor flux amplitude and phase are:

(16)

Differentiating (16) and substituting (15) leads to the drive speed:

(17)

Therefore, given a complete knowledge of the motor parameters, the instantaneous speed ωr can be calculated from Eq. 17, where the voltage model of Eq. 14 is used to estimate the rotor flux amplitude.

Sliding Mode Control
With the proper field orientation and with rated rotor flux, the torque equation, Eq. 6, can be rewritten as (Denai and Attia, 2002):

(18)

where, KT is the torque constant and is defined as follows:

(19)

where, denotes the command rotor flux. The mechanical equation of an induction motor can be written as (Bose, 2001; Chee-Mun Ong, 1998):

(20)

where, ωm = (2/P)ωr is the rotor mechanical speed. Using Eq. 18, one can obtain:

(21)

Where:

a = B/J, b = KT/J, f = TL/J
(22)

Equation 21 can be written with uncertainties Δa, Δb and Δf in the terms a, b and f, respectively, as follows:

(23)

The tracking speed error can be defined as:

(24)

where, is the rotor speed command. Taking the derivative of Eq. 24 with respect to time yields:

(25)

where, the following terms have been collected in the signal u(t):

(26)

and the uncertainty terms have been collected in the signal d(t):

(27)

The sliding variable S(t) can be defined with an integral component as:

(28)

where, k is a constant gain.

When the sliding mode occurs on the sliding surface of Eq. 28, then and therefore dynamic behavior of the tracking problem Eq. 25 is equivalently governed by the following equation:

(29)

In order to obtain the speed trajectory tracking the following assumption should be formulated (Barambones et al., 2002):

The gain k must be chosen so that the term (k-a) is strictly negative, therefore

(k<0). Then the sliding surface is defined as:

(30)

The variable structure speed controller is designed as:

(31)

where, k is the gain defined previously, β is the switching gain, sgn(.) is signum function.

The gain β must be chosen so that β≥|d(t)|.

The current command , can be obtained directly substituting Eq. 31 in Eq. 26:

(32)

Therefore, the proposed variable structure speed control resolves the speed-tracking problem for the induction motor, with some uncertainties in mechanical parameters and load torque.

Modeling of PI and Sliding Mode Control-Based IFOC IM
Referring to Eq. 6, the stator quadrature-axis current reference is calculated from torque reference as:

(33)

The stator direct-axis current reference is calculated from rotor flux linkage reference using:

(34)

The rotor flux position θe is generated from synchronous speed ωe integration, which estimated from the estimator block.

Equation 33 and 34 represent the main equations responsible for field oriented-control, which is represented by the Field Oriented block diagram shown in Fig. 3.

The conversion of quantities from dqe to abc reference frames are executed by dq_abc block which is shown in Fig. 4 (Humaidi, 2006).

The quantities in abc reference frame are converted to dqs frame using Fig. 5.

The speed controller with the proportional-integral type can be implemented using Simulink blocks shown in Fig. 6 (Humaidi, 2006).

The modeling of the elements of the sliding mode controller is shown if Fig. 7.

Fig. 3: Inside the block diagram of field oriented control

Fig. 4: Conversion block from dqe to abc

Fig. 5: Conversion block from dqe to abc

Fig. 6: Block of the PI speed controller modeling

Fig. 7: Block of sliding mode speed controller modeling

The current regulator indicated in Fig. 8 consists of three hysteresis controllers, is built with Simulink blocks. The motor currents are provided by the multiplexer output of the induction machine block (Humaidi, 2006).

The simulation block diagram for a three-phase, two-level PWM inverter is illustrated in Fig. 9 (Humaidi, 2006).

Each leg of the inverter is represented by a switch which has three input terminals and one output terminal. The output of the switch oscillates between (+0.5Vdc) and (-0.5 Vdc), which is characteristic of a pole of an inverter.

(35)

The SIMULINK modeling of speed observer based on Eq. 17 is illustrated in Fig. 10.

Block diagrams of Fig. 3-10 can be assembled to yield the block diagram of PI or sliding mode controller-based IFOC I.M., which is given in Fig. 11.

Fig. 8: Current regulator modeling

Fig. 9: Modeling of the two-level PWM inverter

Fig. 10: Inside the speed estimator block

Fig. 11: SIMULINK modeling of PI and Slide Mode Control-Based IFOC IM

RESULTS

The speed regulation performance of the proposed sliding mode is compared to that of the PI controller of the field oriented control IM. The performance is checked in terms of load torque variations. The simulation is performed in SIMULINK, while the IM model is simulated using s-function block after dicretizing the IM model defined in Eq. 3 and 4 (Humaidi, 2006). The IFOC IM of Fig. 11 is run at sampling period Ts = 2e-6. The IM used in this case study has the parameters listed in Table 1.

In the following study, the variation of the motor parameters is confined to load change only, while the other parameters, e.g., viscous friction B, magnetizing inductance Lm and rotor inductance Lr are held constant.

In the first test, the PI controller is used as a speed controller and a cyclic change of different load torque levels are subjected to the machine at certain times and as follows:

Time = [0 0.75 0.75 1.0 1.0 1.25 1.25 1.5 1.5 2];

Torque = [0 0 -10 -10 -5 -5 5 5 0 0];

The responses of speed, developed torque and stator current are shown in Fig. (12). It is evident from the figure that the PI controller shows bad speed response at these applied changes of loads and lacks the ability to hold the speed at the required value. Therefore, it is true to say that the PI controller is not robust against the changes of IM parameters. Retuning of the PI parameters may slightly reduce the change of speed due to its corresponding change of load.

Table 1: Induction motor parameter

Fig. 12:

PI controller-Based IFOC

Fig.13: Sliding mode control-Based Induction Machine IFOC Induction Machine

Fig.14: Sliding mode control-based IFOC induction machine when higher levels of torque load are exerted

In the second test, the sliding mode controller is proposed and a change of ±20 N.m of load applied for a period of time. For this value of load level change β is found to be 8 and the value of k is set to be 100. The value of β limits the value of exerted load, while the value of k determines the speed of tracking according to Eq. 34. The form of this repeated load change can be clarified as follows:

Time = [0 0.75 0.75 1.0 1.0 1.25 1.25 1.5 1.5 2];

Torque = [0 0 -20 -20 0 0 20 20 0 0];

where, the load is exerted in time ranges [0.75 - 1.0] and [1.25 - 1.5]. The speed response of Fig. 13 shows no detected change in the rotor speed at these ranges, meaning that the sliding mode controller does well in rejecting the applied torque; therefore the rotor speed is by now riding over the sliding trajectory and the controller is robust against IM parameter variations.

To which extent the controller can reject the applied load depends on the value β, which has been determined based on the load level. In the above case, the value of β has been calculated to be greater than 8 for the sliding mode controller to reject the load levels less (greater) than 20 (-20). Otherwise the controller begins to lack the ability to reject the values of load level greater than 20 (or less than -20) unless the boundary of β is increased over the previous one (8). The strong evident of this argument is clearly shown in Fig. 14, where the exerted load level exceeded 20 N.m while the value of β is held constant at the value 8. The speed response deteriorates and a change in the rotor speed will appear at time of exerting ±25 N.m. Therefore the boundary of β should be released and should have a new value other than 8 for the controller to manage the new level of load change.

CONCLUSION

From the simulated results the following points can be concluded:

The PI controller is not robust against IM parameter variations and as has been shown the speed response degrades to different level of load changes
The sliding mode controller can overcome the problem of system degradation, encountered in PI controller, on the condition that the sliding mode controller is well designed for the specified level of load changes

If the load level exceeds the design specification, degradation in the speed response is observed unless another design procedure is adopted.

REFERENCES

  • Akin, B., 2003. State estimation techniques for speed sensorless field oriented control of induction machine. M.Sc. Thesis, The Middle East Technical University.


  • Denai, M.A. and S.A. Attia, 2002. Intelligent control of an induction motor. Electr. Power Compon. Syst., 30: 409-427.
    CrossRef    Direct Link    


  • Barambones, O., A.J. Garrido and F.J. Maseda, 2002. A sensorless robust vector control of induction motor drives. 1st Edn., University of Pais Vasco, UK


  • Bose, B.K., 2001. Modern Power Electronics and AC Drive. 1st Edn., University of Tennessee, Knoxville, Prentice Hall


  • Chee-Mun, O., 1998. Dynamic Simulation of Electric Machinery Using Matlab/Simulink. 1st Edn., Purdue University, Prentice Hall PTR


  • Denai, M.A. and S.A. Attia, 2002. Fuzzy and neural control of an induction motor. Int. J. Applied Math. Comput. Sci., 12: 221-233.
    Direct Link    


  • Edwards, C. and S.K. Spurgeon, 2002. Sliding Mode Control: Theory and Application. Taylor and Francis Ltd.


  • Humaidi, A.J., 2006. Fuzzy learning enhanced speed control of an indirect field-oriented induction machine drive with adaptive hysteresis-band current controller. PhD. Thesis, University of Technology.


  • Leonhard, W., 1996. Control of Electrical Drives. 1st Edn., Springer Press, Berlin.


  • Ouhrouche, A. and C. Volat, 2000. Simulation of a direct field-oriented controller for an induction motor using MATLAB/SIMULINK software package. Proceedings of the IASTED International Conference Modeling and Simulation, Pennsylvania, May 15-17, 2000, USA., pp: 298-302.

  • © Science Alert. All Rights Reserved