IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 57, NO. 4, APRIL 2010 1217 Real-Time Electrical Load Emulator Using Optimal Feedback Control Technique Y. Srinivasa Rao, Member, IEEE, and Mukul C. Chandorkar, Member, IEEE Abstract—This paper presents a method of emulating electrical loads using power electronic converters. The loads include machines such as induction motors and their associated mechanical load and also more complex machine systems such as wind-driven generators. The load emulator is, effectively, a dynamically controllable source or sink which is capable of bidirectional power exchange with either a grid or another power electronic converter system. Using load emulation, the feasibility of connecting a particular machine to a grid under various load conditions can be examined without the need for any electromechanical machinery. This paper considers the case of a power electronic voltage source inverter (VSI) emulating a three-phase induction motor connected to a three-phase ac grid. The VSI is operated in a mode where the current drawn from the ac grid is controlled by closed-loop control. The consistency of the experimental results with the simulation results proves the ability of the emulator and the proposed testing approach. Index Terms—Digital signal processor (DSP), induction motor, linear quadratic regulator (LQR), load emulator, phase-locked loop (PLL), real-time simulation, voltage source inverter (VSI). I. I NTRODUCTION L OAD EMULATION is the concept of controlling a power electronic converter such as a voltage source inverter (VSI) in such a manner that its behavior resembles that of an electrical load such as an induction machine. Rapid prototyping [1], [2] is a method of verifying the control techniques that are implemented in embedded systems such as digital signal processors (DSP). Reference [1] describes a method of simulating the details of DSPs and their associated analog-to-digital interface. Reference [2] describes a method of combining simulations of the power circuit and the control circuit to closely describe the behavior of the practical system that would be implemented. There is an increasing usage of hardware-in-theloop technologies that design and test control systems [3]– [5]. Currently, there are several kinds of real-time simulation platforms [6]–[9] in academia and industry. However, load emulation, as presented in this paper, provides a platform where the behavior of an electrical load can be emulated in hardware at real power levels using a controllable VSI. The load emulator Manuscript received December 30, 2008; revised November 16, 2009. First published December 4, 2009; current version published March 10, 2010. Y. Srinivasa Rao is with the Department of Electrical Engineering, Indian Institute of Technology Bombay, Mumbai 400076, India, and also with the Sardar Patel Institute of Technology, University of Mumbai, Mumbai 400058, India (e-mail: ysrao@ee.iitb.ac.). M. C. Chandorkar is with the Department of Electrical Engineering, Indian Institute of Technology Bombay, Mumbai 400076, India (e-mail: mukul@ee.iitb.ac.in). Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identifier 10.1109/TIE.2009.2037657 Fig. 1. Schematic diagram of the load emulator. can provide different load characteristics with which the control algorithms and inverter design can be tested. Therefore, load emulation offers a more flexible platform for testing inverters in a laboratory environment. The primary objective of load emulation is to design the power interface for the actual electrical load. Therefore, the behavior of the load emulator must be close to the actual load when connected to a power supply. The load emulator configuration is shown in Fig. 1, where the power supply can be a threephase grid or another three-phase VSI. Therefore, accurate information of the phase angle of the power supply voltages is essential. The phase angle of a three-phase voltage is obtained through the popular method called phase-locked loop (PLL). This takes as its input the three-phase voltages of the power supply and produces a sinusoidal signal of the same frequency as the frequency of the supply voltages. The topic of PLL has been extensively reported in literature [10]–[12]. However, the effectiveness of the PLL has not been investigated in conditions where the input voltages are the pulsewidth modulated (PWM) voltages of a VSI with a switching frequency of approximately 5 kHz. The main component of load emulation is the algorithm that generates the desired currents to be drawn so as to mimic the actual load. Thereafter, the control system ensures that the VSI draws or supplies currents that are as close as possible to the desired current references. The load emulator VSI is connected to the power supply through an interface impedance known as the filter. The topology of the filter discussed in this paper has been chosen to be an inductor–capacitor–inductor (L–C–L) filter. Since the load emulator VSI mimics the behavior of an electrical load, it would have to draw or source active power as per load characteristics. In order to enable the emulator VSI to be a sink or source of active power, a back-to-back VSI topology has been proposed. The dc link of the emulator 0278-0046/$26.00 © 2010 IEEE 1218 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 57, NO. 4, APRIL 2010 VSI is common to another VSI known as the reverse power flow converter. This reverse power flow converter is a threephase VSI whose ac output is interfaced to the three-phase ac grid. This converter draws or supplies active power from the three-phase ac grid for the emulator VSI. Two controllers are implemented for the entire system. In order to ensure bidirectional power flow, the reverse power flow converter is controlled in current control mode. The voltage of the dc link common to both the emulator VSI and the reverse power flow converter is maintained constant by this controller. This current controller of the reverse power flow converter is not the focus of this paper, and reported literature contains this in sufficient detail [13]. The load emulator VSI that is interfaced to the three-phase power supply through the L–C–L filter has been controlled through an optimal control strategy. The optimal feedback controller used is a linear quadratic regulator (LQR) that has been discussed in this paper. This paper also investigates the capability of the optimal controller, and therefore, the limits of load emulation are presented. Fig. 2. Inverter interface with an LC filter. (a) Inverter interface with an inductor. (b) Inverter interface with an LC filter. II. P RINCIPLE OF L OAD E MULATION DSPs and microcontrollers have been widely used in implementing digital motor control algorithms [14], [15]. However, the processing speed of DSPs needs to be on the order of microseconds to nanoseconds in order to deal with the hardware-rich tasks. This becomes difficult with the software computation environment. A hardware control system using a field programmable gate array (FPGA) has been reported [16]– [19]. One disadvantage of FPGAs is that they need many resources for complex mathematical operations since they are not arithmetic-oriented like DSPs [20]. Fig. 1 shows the schematic diagram of the load emulator with the associated digital control consisting of a DSP and an FPGA. The diagram shows the L–C–L filter that is used to interface the load emulator VSI to the three-phase power supply. It also shows the three-phase power supply as the three-phase distribution level ac grid. In the later sections, the power supply will also be a three-phase VSI. The supply voltages vs are used in generating references for the currents to be drawn by the emulator VSI. The references are generated by an algorithm in the DSP. The PLL is also an essential component of reference generation and is implemented in DSP. The current is drawn from the three-phase ac grid is the current that must be controlled to be approximately equal to the reference currents. The optimal control algorithm is implemented in DSP and requires the feedback of both the currents is and the emulator VSI currents if . The PWM signals are generated by the FPGA and fed to the gate drivers of the emulator VSI. Before describing the optimal feedback control strategy, a justification for using the L–C–L filter as an interface between the emulator VSI and the ac grid is provided. The well-reported method for current injection requires an inverter with an output inductor filter [Fig. 2(a)]. In Fig. 2(a), the VSI is connected to the ac grid as a node that is called as the point of common coupling (PCC). With this topology, the current injected into the PCC would contain a substantial amount of switching harmonic ripple. Due to the internal impedance of the grid shown in Fig. 2(a), the voltage at the PCC will be Fig. 3. Emulator interface with an LCL filter. distorted. Therefore, by connecting a VSI to the ac grid, a power quality problem has been introduced. As shown in Fig. 2(b), by connecting a capacitor Cf , a lowpass filter is present at the output of the VSI. This provides a low-impedance path for high switching frequency ripple. Therefore, the PCC voltages will be smooth. However, since is = ic + if , current drawn from the grid will be different from the desired references. Although a larger filter capacitor Cf will provide smoother PCC voltages, the error in the currents drawn from the grid will increase as ic = Cf · dvpcc /dt. Moreover, if the ac grids were to be replaced by another inverter, the current injection strategy would not work as the VSI replacing the grid would not have an internal impedance. Fig. 3 shows an LCL output filter structure which will provide a solution to the two aforementioned drawbacks [21]. The Lf Cf filter removes some of the switching frequency harmonics, resulting in a current through Ls that is relatively smooth. The control strategy employed will attempt to control the current is through the inductor Ls . Therefore, the error in the current is from its desired reference will be negligible. Moreover, in the case of the inverter being interfaced to another inverter, the voltage at the filter capacitor Cf resembles a back-emf, and therefore, the interfacing problems are removed. Two methods of controlling is so that this is approximately equal to the desired references have been examined. The first method is an indirect method where the reference for vf has been generated from the reference for is . The voltage vf can be controlled, making it approximately equal to its reference by using the flux modulator strategy [22]. Current is can be directly controlled by using the optimal feedback controller. This controller is a state feedback controller designed using the LQR technique [23]–[25]. Both currents is and if are measured and fed back to be compared with their respective references. Voltage vf is implicitly controlled. In the case of the VSI SRINIVASA RAO AND CHANDORKAR: ELECTRICAL LOAD EMULATOR USING OPTIMAL FEEDBACK CONTROL Fig. 4. 1219 Closed-loop control scheme of the optimal feedback controller. emulating an induction motor, the voltage across the capacitor Cf is related to the induced rotor voltage of the induction motor. A detailed description of the optimal feedback controller is provided in the next section. III. C ONTROLLER D ESIGN The control system ensures that the emulator draws a current is that is approximately equal to the desired reference. The transient response of the system determines the tracking accuracy between the demanded and actual current drawn from the inverter under test. Here, the control technique used is a state feedback control strategy. Fig. 4 shows the closed-loop control scheme employed for the VSI in Fig. 3. The state vector is x = [ is vf if ]T . (1) The state space equations are obtained by applying Kirchoff’s current law and Kirchoff’s voltage law to Fig. 3 ⎡ ⎤ −rs /Ls −1/Ls 0 ẋ = ⎣ 1/Cf 0 −1/Cf ⎦ x 0 1/Lf −rf /Lf ⎤ ⎡ ⎤ ⎡ 1/Ls 0 ⎦ u + ⎣ 0 ⎦ vs (2) +⎣ 0 0 −Vdc /Lf y = is = [ 1 0 0 ] x. (3) The output variable y is therefore the current is that is injected into the grid. However, as stated in the previous section, both currents if and is are measured. The voltage vf across the filter capacitor Cf also forms the state vector along with if and is . However, its measurement has been excluded (a description has been provided later in the section). The objective of the VSI is to inject a current is that is approximately equal to a desired reference isr . In the aforementioned equation, u is the emulator inverter’s switching function, and ac grid voltage vs is considered to be a disturbance. The state feedback controller is an optimal controller designed using the LQR to achieve the objective [23]. For implementation using DSP, the controller is realized in a discrete-time form. The continuous-time state equations are converted to standard state space discrete-time form x(n + 1) = Gx(n) + Hu u(n) + Hw w(n) y(n) = Cx(n). (4) (5) Here, w(n) = vs (nTs ) is a disturbance, and Ts is the sampling time interval. The LQR seeks to minimize a performance index J= ∞ 1 T e (n)Qe(n) + uT c (n)Ruc (n) 2 n=0 (6) Fig. 5. Closed-loop frequency response of the system. where e = xr − x and T stands for the transpose of a vector. In (6), Q is a diagonal matrix that decides the importance to be given to the state variables, and R is the penalty which is imposed on the controller. The dlqr function of the simulation provides the state feedback matrix K uc (n) = K (xr (n) − x(n)) x(n + 1) = Gx(n) + Hu K (xr (n) − x(n)) + Hw w(n) Y (n) = Cx(n). (7) (8) (9) To obtain the transfer function between the output variable y and the reference state vector xr , disturbance w(n) is neglected. Equation (8) can be written as x(n + 1) = Gx(n) + Hu Kxr (n) − Hu Kx(n). (10) This can be written as x(n + 1) = (G − Hu K)x(n) + Hu Kxr (n) y(n) = Cx(n). (11) (12) Equations (11) and (12) are the closed-loop state equations. The system parameters are chosen as Ls = 3 mH, Rs = 0.01 Ω, Lf = 1 mH, Rf = 0.01 Ω, and Cf = 50 μF. The choices for Q and R are ⎡ ⎤ 50 0 0 Q = ⎣ 0 0.05 0 ⎦ R = 65 (13) 0 0 0.01 where the weightage for is is chosen to be much larger than the other states in the matrix Q. K = [0.6188 0.0404 0.0929], with a sampling time of 20 μs. Since the control coefficient for error in vf is extremely small [0.0404], it is possible to set this constant at zero. This also provides another advantage of eliminating measurement of vf . The closed-loop frequency response characteristics can be represented in a Bode plot as shown in Fig. 5. In Fig. 5, it can be observed that the gain of the closed-loop system is unity for 1220 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 57, NO. 4, APRIL 2010 Fig. 6. Simulation of control system for harmonic references. Fig. 8. Closed-loop scheme of the PLL. (a) Block diagram of the control system. (b) Notch filter design. This system using a notch filter and PID controller which produce accurate information of the phase angle is discussed in this section. The basic control scheme of the PLL can be described as follows. The grid voltages va , vb , and vc are converted to the d–q reference frame through the standard transformation of [10]–[12] 2 cos ω vd t sin ω t 1 √ −1/2 −1/2 √ = vq 3/2 − 3/2 t cos ω t 0 3 − sin ω Fig. 7. Simulation of control system for transient references. frequencies of up to 10 000 rad/s. Moreover, the phase lag is approximately zero for frequencies of up to 350 rad/s. From the simulation results of Fig. 6, current references isr containing fundamental 50-Hz components can be tracked with negligible error. Harmonics in the current reference are tracked with a phase lag. The controller tracks zero current references with negligible error. Fig. 7 shows the results of transient response and stability of the system. To enable the VSI to be able to inject currents containing harmonic components, the bandwidth of the controller can be increased by redesigning the controller. IV. T HREE -P HASE PLL S YSTEM In the proposed system, the reference voltages for the load models in DSP are derived from the PCC. In Fig. 3, voltage vs may be a PWM waveform or may be distorted by various external factors. In order to generate current references of particular harmonic number, the frequency of the grid must be known accurately. Moreover, in order to generate current references that may lead or lag the grid voltages, the phase information of the grid voltages must be known. Therefore, synchronization and frequency estimation using a PLL scheme are employed. × [ va vb vc ]T . (14) The closed-loop control scheme of the PLL is shown in Fig. 8(a). The block “abc to dq transformation” implements the transformation of (14). Voltage vd is forced to zero through a PID controller whose output is the estimated angular frequency ω . In Fig. 8(a), by forcing vd to be equal to zero, the d-axis of the rotating reference frame is forced to be aligned with the voltage vector Vs . Therefore, the estimated angular frequency ω = ω. In the steady state, the cosine wave cos ω t has a frequency equal to the frequency of the grid voltages and is in phase with the voltage va . When the grid voltages va , vb , and vc are balanced, vd can be forced to be very close to zero, and therefore, the estimated angular frequency ω will be constant. However, when the grid voltages are unbalanced, they possess a negative sequence component. On the transformation to the d–q frame, the negative sequence component appears as a component whose frequency is twice that of the fundamental frequency. Therefore, ω will contain oscillations of a frequency twice that of the fundamental frequency. To prevent this, a notch filter [11] is included after the d–q transformation to remove the double harmonic component from vd . Fig. 8(b) shows the design of the notch filter. The “O”s represent zeros, and “X”s represent poles in the complex z plane. The fundamental frequency of the grid is close to the SRINIVASA RAO AND CHANDORKAR: ELECTRICAL LOAD EMULATOR USING OPTIMAL FEEDBACK CONTROL Fig. 9. Experimental results of PLL for unbalanced utility voltages. nominal expected frequency of ω0 = 100π rad/s. The sampling frequency of the analog-to-digital conversion is fs = 1250 Hz, and ωs = 2.5π ∗ 103 rad/s. Since the objective of the notch filter is to remove the double frequency harmonic term, a pair of complex zeroes is added at z = e±j2ω0 /ωs = e±j0.08 . It is to be noted that conjugate zeroes are added to ensure that the coefficients of the filter are real. To ensure that the notch filter is selective and has a low gain only for frequencies close to the double frequency harmonic, complex poles are added at the same frequency as the zeroes but within the unit circle to ensure stability of the system. By adding the complex poles close to the unit circle at a magnitude of 0.97 (z = 0.97e±j0.08 ), the system is fairly resonant with a sharp notch. To ensure that the dc quantity produced by the d–q transformation corresponding to the fundamental frequency is not attenuated, another real pole is added at z = 0.97. The total transfer function of the notch filter is given by 0.03 z 2 − 1.9974z + 1 Hnf (z) = 4.7692 2 . z − 1.7976z + 0.81z z − 0.97 (15) To prevent quantization errors while implementing in DSP, the notch filter is implemented in cascade form, with the two transfer functions within brackets converted separately into difference equations. The transfer function of the PID regulator is as follows: Ki + Kd s · E(s). (16) Y (s) = Kp + s In the system transfer function (16), the derivative term is added in order to improve system response. Sampling time step dt for the PLL is selected as 80 μs. Therefore, Ki = 12.5, Kp = 1.0, and Kd = 0.00004. 1221 Fig. 10. Performance of the control system and PLL for harmonic currents. TABLE I I NDUCTION M OTOR M ODEL PARAMETERS response slower. This will also make the filter more sensitive to the variations of the center frequency ωo which may occur in some power systems. While designing the notch filter of PLL, it was assumed that the grid frequency will be changed by a maximum of 3%. Therefore, the PLL will continue to function for small changes in grid frequency. A three-phase experimental setup has been built to verify the accuracy of the proposed system. In this setup, the power circuit is rated at 30 A, 415-V input/output line-to-line voltages, and 750-V dc bus. Various linear and nonlinear current references, representing different load characteristics, are emulated experimentally. The average switching frequency of the emulator inverter was 5 kHz. A. Current Harmonics The source of the current harmonics can be considered as nonlinear loads such as rectifiers. The upper trace in Fig. 10 shows the tracking performance of the grid interface emulator for current references containing odd order harmonic components. In the experiment, these magnitudes are taken as one-third, one-fifth, one-seventh, etc., of the fundamental current reference. The experimental data are acquired and stored in real time in the DSP memory and then plotted offline using a suitable plotting software. The lowest trace shows the effectiveness of the PLL for distorted PCC voltages. The PLL is observed to produce smooth waveforms despite notches in the PCC voltages. V. E XPERIMENTAL R ESULTS Fig. 9 shows the performance of the PLL for unbalanced t is seen to grid voltages, where vb = 0. The cosine wave cos ω have the same frequency and phase as that of va . By increasing the filter’s order, it becomes sharper and hence suppresses the undesired components even more effectively, but this makes the B. Induction Motor Model To implement the controller for the three-phase four-pole squirrel cage induction motor emulator, the parameters mentioned in Table I are taken as a reference model in generating reference currents [8]. 1222 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 57, NO. 4, APRIL 2010 Fig. 11. Offline simulation. (a) IM model a-phase stator current. (b) Load emulation of the motor model. Fig. 12. Experimental results of the motor a-phase stator current. The model is simulated to obtain the free acceleration characteristics of the machine when started direct-on-line with no load. While the motor is running under no load steady state, an 80% rated torque is suddenly applied at t = 1 s to the rotor in order to observe the transient behavior of the system. The upper trace in Fig. 11 shows the offline simulation results of the motor model, and the lower trace shows the emulation results. Fig. 12 shows the experimental waveform of the a-phase stator current. The offline simulated load model currents and the experimental emulated currents are nearly identical. Fig. 13 shows the experimental waveforms of rotor mechanical speed, electromagnetic torque, and active power drawn by the emulator. For the given parameter values shown in Table I, the transient and steady-state currents of the emulator are as desired. The current waveforms have a nominal ripple, and the control system is stable. C. PWM Reference Voltages Under the circumstances when both the source and the load are VSIs, the reference voltages are PWM-type signals as the source is also a VSI. The switching frequency of the source VSI is 5 kHz, and the control algorithm is executed every Fig. 13. Waveforms for free acceleration of the IM model. Fig. 14. Emulator performance for PWM reference voltages. 20 μs. The upper trace in Fig. 14 shows the tracking accuracy of the load VSI, where the currents drawn are from the inverter under test. The lower trace shows the effectiveness of the PLL in synchronizing with the PWM output of the source VSI. The PLL is seen to synchronize to PWM voltages with negligible phase error. Fig. 15 shows the performance of the emulator for the variable-frequency source VSI. The frequency and amplitude of the source VSI are modulated from 0 to 50 Hz and 0.1 to 0.9, respectively, over a period of 0–4 s. The upper trace in Fig. 15 shows the tracking accuracy of the emulator for constant v/f reference currents, and the lower trace shows the response of the emulator for the IM load model a-phase startup currents. D. Unbalanced Three-Phase Three-Wire Load Fig. 16 shows the emulator drawing unbalanced currents from the source VSI. It is to be noted that, with the source VSI being a three-phase three-wire VSI, the sum of the currents drawn must be equal to zero. Therefore, the sum of the references for the currents to be drawn by the load VSI is also zero, with a 50% unbalance in the b-phase current. The lower trace SRINIVASA RAO AND CHANDORKAR: ELECTRICAL LOAD EMULATOR USING OPTIMAL FEEDBACK CONTROL 1223 Fig. 17. Single-phase diode bridge rectifier. Fig. 15. Emulator performance for the v/f control source VSI. Fig. 18. Experimental results of the emulator for the delta-connected diode rectifier. Fig. 16. Emulator performance for three-phase unbalanced current references. shows the tracking performance where the reference is tracked with negligible error. These results demonstrate that this control system, along with PLL, can be used to test the inverters with acceptable accuracy. E. Nonlinear Active Load Three single-phase diode bridge rectifiers connected between three phases and a neutral is used as a nonlinear load model. Fig. 17 shows the diagram of a single-phase rectifier with a large filter capacitor. The nonlinear circuit is represented by three cases. Case 1 of the circuit is where diodes D1 and D2 are in the on state and D3 and D4 are in the off state. For this case, when vs > vdc or is > 0, the system can be expressed as d is −Rs /Ls is −1/Ls = 1/Cdc −1/(Rload · Cdc ) vdc dt vdc 1/Ls + vs . (17) 0 Case 2 of the circuit is where diodes D3 and D4 are in the on state and D1 and D2 are in the off state. For this case, when −vs > vdc or is < 0, the system can be expressed as d is −Rs /Ls is 1/Ls = −1/Cdc −1/(Rload · Cdc ) vdc dt vdc 1/Ls + vs . 0 (18) When diodes D1, D2, D3, and D4 are in the off state (i.e., neither case 1 nor 2 are satisfied), therefore, is = 0. To solve the circuit model, the following parameters are taken as a reference, where vs = 230 vrms , Rs = 0.1 Ω, Ls = 100 μH, Cdc = 1000 μf, and Rload = 100 Ω. The model is simulated using Euler’s method, with an integration time step of ΔT = 20 μs. Fig. 18 shows the experimental waveforms of the emulator for the single-phase diode rectifier bank connected between two lines in a delta topology. The figure shows the unbalanced rectifier load model current references isa = 17 A, isb = 16.5 A, and isc = 16 A (peak). The phase currents add up to zero. The lowest trace shows the fourth leg current in as equal to zero. From the results, it can be observed that the controller constantly traces the reference currents of the nonlinear active loads. The tracking performance of the controller remains the same even during the periods when the reference remains zero. VI. E MULATOR T IME A NALYSIS The emulator digital hardware structure is centered on the Texas Instruments TMS320VC33 DSP, which has a clock frequency of 75 MHz. This gives an instruction cycle time of 13.33 ns. The entire load emulation software is driven by an 1224 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 57, NO. 4, APRIL 2010 TABLE II P ROCESSOR E XECUTION T IME FOR L OAD E MULATION interrupt service routine (ISR). The main code (background loop) consists of simply peripheral initialization (timers, analog-to-digital converter (ADC), interrupt control, and buffer). The remainder of the code is taken up entirely by Timer 0 ISR. This ISR is invoked every 20 μs by the period event flag on Timer 0. Calculation of reference currents from the model, sampling of ADC, and generation of PWM waveforms for power switches are done within 20 μs. Table II shows the tasks performed and time taken by each task. The tasks carried out by DSP–FPGA take 14.8 μs in total for a time step of 20 μs. Three ADC sample time is 3.5 μs, and LQR current controller execution time is 2.06 μs. Data conversion and communication between DSP and FPGA is 640 ns. Motor model using the two-step Adams–Bashforth integration method execution time is 2.04 μs. Buffer save and inverter protection task execution time is 2.8 μs. Data transfer between PC and DSP buffer is offline. The measurements are taken with the help of external flag XF0 of the DSP. VII. C ONCLUSION This paper has discussed load emulation which is capable of emulating an electrical load in real time with power electronic converter. This proposed testing approach eliminates the need of using real loads in the design process and, therefore, greatly reduces the overall design cost. The state feedback controller has been designed and implemented in a DSP–FPGA platform and has been shown to perform satisfactorily in steady-state and transient conditions. The system was operated to emulate various electrical load models. Experimental results of control system and PLL are discussed. The PLL is an additional feature implemented in DSP using which nonlinear and leading/lagging power factor loads can be emulated. The most attractive feature of the emulator is that the emulator is shown to mimic the behavior of induction motors and diode rectifiers. The capacity of the emulator is limited by the power rating and bandwidth of the system. A minimum realtime step of 20 μs has been achieved using this platform. This time resolution is sufficient to simulate the dynamic behavior of most power electronic control systems and electrical load models. R EFERENCES [1] J. Jacobs, D. Detjen, C.-U. Karipidis, and R. W. De Doncker, “Rapid prototyping tools for power electronic systems: Demonstration with shunt active power filters,” IEEE Trans. Power Electron., vol. 19, no. 2, pp. 500– 507, Mar. 2004. [2] A. Monti, E. Santi, R. A. Dougal, and M. Riva, “Rapid prototyping of digital controls for power electronics,” IEEE Trans. Power Electron., vol. 18, no. 3, pp. 915–923, May 2003. [3] B. Lu, X. Wu, H. Figueroa, and A. Monti, “A low-cost realtime hardware-in-the-loop testing approach of power electronics controls,” IEEE Trans. Ind. Electron., vol. 54, no. 2, pp. 919–931, Apr. 2007. [4] H. Li, M. Steurer, K. L. Shi, S. Woodruff, and D. Zhang, “Development of a unified design, test, and research platform for wind energy systems based on hardware-in-the-loop real-time simulation,” IEEE Trans. Ind. Electron., vol. 53, no. 4, pp. 1144–1151, Aug. 2006. [5] W. Ren, M. Steurer, and L. Qi, “Evaluating dynamic performance of modern electric drives via power-hardware-in-the-loop simulation,” in Proc. IEEE Int. Symp. Ind. Electron., Jun. 2008, pp. 2201–2206. [6] V. R. Dinavahi, M. R. Iravani, and R. Bonert, “Real-time digital simulation of power electronic apparatus interfaced with digital controllers,” IEEE Trans. Power Del., vol. 16, no. 4, pp. 775–781, Oct. 2001. [7] S. Abourida, C. Dufour, J. Belanger, G. Murere, N. Lechevin, and Y. Biao, “Real-time PC-based simulator of electric systems and driver,” in Proc. IEEE Appl. Power Electron. Conf., 2002, pp. 433–438. [8] R. Champagne, L. A. Dessaint, H. Fortin-Blanchette, and G. Sybille, “Analysis and validation of a real-time AC drive simulator,” IEEE Trans. Power Electron., vol. 19, no. 2, pp. 336–345, Mar. 2004. [9] G. G. Parma and V. Dinavahi, “Real-time digital hardware simulation of power electronics and drives,” IEEE Trans. Power Del., vol. 22, no. 2, pp. 1235–1246, Apr. 2007. [10] V. Kaura and V. Blasko, “Operation of a phase locked loop system under distorted utility conditions,” IEEE Trans. Ind. Appl., vol. 33, no. 1, pp. 58– 63, Jan./Feb. 1997. [11] S. J. Lee, J. K. Kang, and S. K. Sul, “A new phase detecting method for power conversion systems considering distorted conditions in power system,” in Conf. Rec. IEEE IAS Annu. Meeting, Oct. 1999, vol. 4, pp. 2167– 2172. [12] S. K Chung, “Phase-locked loop for grid-connected three-phase power conversion systems,” Proc. Inst. Elect. Eng.—Elect. Power Appl., vol. 147, no. 3, pp. 213–219, May 2000. [13] R. R. Sawant and M. C. Chandorkar, “A multifunctional four-leg grid connected compensator,” IEEE Trans. Ind. Appl., vol. 45, no. 1, pp. 249– 259, Jan./Feb. 2009. [14] M. Boussak and K. Jarray, “A high-performance sensorless indirect stator flux orientation control of induction motor drive,” IEEE Trans. Ind. Electron., vol. 53, no. 1, pp. 41–49, Feb. 2006. [15] M. J. Duran, J. L. Duran, F. Perez, and J. Fernandez, “Induction motor sensorless vector control with online parameter estimation and over current protection,” IEEE Trans. Ind. Electron., vol. 53, no. 1, pp. 154–161, Feb. 2006. [16] E. Monmasson and M. N. Cirstea, “FPGA design methodology for industrial control systems—A review,” IEEE Trans. Ind. Electron., vol. 54, no. 4, pp. 1824–1842, Aug. 2007. [17] Y. F. Chan, M. Moallem, and W. Wang, “Design and implementation of modular FPGA-based PID controllers,” IEEE Trans. Ind. Electron., vol. 54, no. 4, pp. 1898–1906, Aug. 2007. [18] C. Dufour, J. Belanger, S. Abourida, and V. Lapointe, “FPGA-based realtime simulation of finite element analysis permanent magnet synchronous machine drives,” in Proc. IEEE Power Electron. Spec. Conf., Orlando, FL, Jun. 2007, pp. 909–915. [19] D. Zhang and H. Li, “A stochastic-based FPGA controller for an induction motor drive with integrated neural network algorithms,” IEEE Trans. Ind. Electron., vol. 55, no. 2, pp. 551–561, Feb. 2008. [20] S. Jung and S. S. Kim, “Hardware implementation of a real-time neural network controller with a DSP and an FPGA for nonlinear systems,” IEEE Trans. Ind. Electron., vol. 54, no. 1, pp. 265–271, Feb. 2007. [21] M. Liserre, F. Blaabjerg, and S. Hansen, “Design and control of an LCLfilter-based three-phase active rectifier,” IEEE Trans. Ind. Appl., vol. 41, no. 5, pp. 1281–1291, Sep. 2005. [22] M. C. Chandorkar, “New technique for inverter flux control,” IEEE Trans. Ind. Appl., vol. 37, no. 3, pp. 880–887, May/Jun. 2001. [23] H. Komurcugil, O. Kukrer, and A. Doganalp, “Optimal control for single-phase UPS inverters based on linear quadratic regulator approach,” in Proc. Int. Symp. Power Electron., Elect. Drives, Autom. Motion, May 2006, pp. 1137–1142. [24] F. H. F. Leung, P. K. S. Tam, and C. K. Li, “An improved LQR-based controller for switching DC–DC converters,” IEEE Trans. Ind. Electron., vol. 40, no. 5, pp. 521–528, Oct. 1993. [25] T. Murata, T. Tsuchiya, and K. Takeda, “Vector control for induction machine on the application of optimal control theory,” IEEE Trans. Ind. Electron., vol. 37, no. 4, pp. 283–290, Aug. 1990. SRINIVASA RAO AND CHANDORKAR: ELECTRICAL LOAD EMULATOR USING OPTIMAL FEEDBACK CONTROL Y. Srinivasa Rao (M’09) was born in Nandigama, Andhra Pradesh, India, on May 19, 1968. He received the B.E. degree in electronics engineering and the M.E. degree with specialization in computer engineering from Shivaji University, Kolhapur, India, in 1990 and 1996, respectively. He is currently working toward the Ph.D. degree in the Department of Electrical Engineering, Indian Institute of Technology Bombay, Mumbai, India. He is currently an Assistant Professor with the Sardar Patel Institute of Technology, University of Mumbai, Mumbai. His current research interests are in the area of rapid prototyping tools for digital power electronics, embedded systems, DSP applications using VLSI, power converters, and modulation techniques for inverters. 1225 Mukul C. Chandorkar (M’84) received the B.Tech. degree from the Indian Institute of Technology Bombay, Mumbai, India, the M.Tech. degree from the Indian Institute of Technology Madras, Chennai, India, and the Ph.D. degree from the University of Wisconsin, Madison, in 1984, 1987, and 1995, respectively, all in electrical engineering. He has several years of experience in the power electronic industry in India, Europe, and the U.S. In 1996 through 1999, he was with ABB Corporate Research Ltd., Baden–Daettwil, Switzerland. He is currently a Professor with the Department of Electrical Engineering, Indian Institute of Technology Bombay. His technical interests include electric power quality compensation, drives, and real-time simulation of electrical systems.