You are on page 1of 11

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO.

5, MAY 2012

2277

Measurements and Simulations of DTC Voltage Source Converter and Induction Motor Losses
Lassi Aarniovuori, Lasse I. E. Laurila, Markku Niemel, and Juha J. Pyrhnen, Member, IEEE
AbstractEnergy efcient pulse-width modulation inverters are widely used to control electrical machines accurately for process needs. The pulse-width modulation, however, has also adverse effects and produces additional losses in the motor. These losses increase the motor temperature and result in derating of the machine power in converter use. A reliable and reasonably accurate loss model of an induction motor drive system is important for the performance prediction of a variable-speed drive. A two-level frequency converter main circuit model is coupled to a niteelement method motor model. The drive model is controlled by closed-loop direct torque control. The frequency converter losses are calculated analytically, and the nite-element method motor model provides an analysis of the motor losses. The simulation results are compared with measurement results. Index TermsAC motors, induction motors, magnetic losses, numerical simulation, power semiconductor switches pulse-width modulation converters, variable-speed drives (VSDs).

N OMENCLATURE Esw,fwdiode Esw,IGBT ESRC ESRL f fN fsw idiode IC Ifund iL im Im IN IRated IRMS n nN Nsw,change Pad PC Pchoke Switching loss energy of fw-diode. Switching loss energy of IGBT. Equivalent series resistance of capacitor. Equivalent series resistance of input choke. Frequency. Nominal frequency. Switching frequency. Instantaneous value of diode current. Capacitor current RMS value. Fundamental wave current amplitude. Instantaneous value of line current. Instantaneous value of motor current. Motor current RMS value. Nominal current. Rated current. RMS Current. Rotational speed. Nominal speed. Number of switch state changes. Motor additional losses. Capacitor bank losses. Power loss in frequency converter input choke.

PDClink Pdischarge Pdiode,cond Pdiode,on Pfwdiode,sw PExtra PIGBT,cond PIGBT,sw Pin Pmech PN Pout RCE0 Rdischarge RF RF0 UC UCE0 UDC UF UF0 Ufund UN t T TN

Intermediate circuit losses. Discharge resistor losses. Forward diode conduction losses. Instantaneous conduction losses of diode. Fw-diode switching losses. Frequency converter extra losses. IGBT conduction losses. IGBT conduction losses. Frequency converter input power. Mechanical power. Nominal power. Frequency converter output power. IGBT on-state resistance. Discharge resistor resistance. On state resistance of diode. On state resistance of fw-diode. Capacitor RMS voltage. IGBT threshold voltage. DC-link voltage. Forward voltage drop of diode. Fw-diode threshold voltage. Fundamental wave voltage amplitude. Nominal voltage. Time. Torque. Nominal torque. I. I NTRODUCTION

Manuscript received December 3, 2010; revised February 19, 2011, April 26, 2011, and June 5, 2011; accepted June 13, 2011. Date of publication June 30, 2011; date of current version February 3, 2012. The authors are with the Department of Electrical Engineering, Lappeenranta University of Technology (LUT), 53850 Lappeenranta, Finland (e-mail: lassi.aarniovuori@lut.; lasse.laurila@lut.; markku.niemela@lut.; juha. pyrhonen@lut.). Digital Object Identier 10.1109/TIE.2011.2161061

DDITIONAL losses caused by PWM methods have been studied widely in the literature, e.g., in [1][6]. Several methods have been proposed to calculate the iron losses under nonsinusoidal excitation [7][10]. In [11], a model is proposed for accurately estimating the iron losses in rotating electrical machines. The impact that PWM harmonics, amplitude modulation index, and switching frequency have on induction motor iron losses is investigated in [12] with a special test motor with plastic rotor cage. IEC is developing a new standard: IEC 60034-2-3: Rotating electrical machinesPart 23: Specic test methods for determining losses and efciency of converter-fed ac machines. Converter-fed motors will get their energy efciency classes. However, there is no generally accepted method to evaluate additional losses caused by PWM in electrical machines. The ac motor power derating caused by PWM losses varies from 0% to 20% [13]. In [14], it is stated that the additional losses in the motor caused by the frequency converter can increase the total motor losses up to 15%20% compared with the grid operation.

0278-0046/$26.00 2011 IEEE

2278

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 5, MAY 2012

TABLE I TEFC M OTOR PARAMETERS

The aim of this paper is to nd drive loss estimation methods that could be used without extensive conrming measurements. It is important to assure that the motor temperature at full load does not exceed the thermal limits of the insulation and thus have a negative impact on the motor lifetime [15]. The temperature rise of the machine is the main dimensioning boundary condition for the machine. Direct torque control (DTC) is one type PWM control strategy which can be considered as an alternative for vector control technique. DTC was proposed for ac drives by Depenbrock and Takahashi in the 1980s [16], [17]. The DTC has advantages of high torque response, simple design, and robustness against parameter variations. The variable switching frequency and high torque ripple are drawbacks of the classical DTC. DTC has been a topic of numerous scientic works over the past two decades; the switching frequency of the DTC is analyzed in [18][20]. Numerous improvements in the classical DTC have been proposed for instance in [21][26]. The total losses of frequency converters are not studied widely. In [27], a unied loss model of a converter induction machine system is presented that includes steady state as well as dynamic behavior of both machine and converter. The studies are mainly focused on IGBT bridge losses [28] and the effect of modulation method in inverter losses [29], [30]. This paper provides results on how much the induction motor losses and temperature rise of the motor increase when a PWM supply is used. A 37-kW industrial totally enclosed fan-cooled (TEFC) class 130 (B) temperature rise induction motor is used in the tests. The motor parameters with delta connection are given in Table I. A commercial DTC frequency converter is used as a pulsewidth modulation (PWM) supply. The sinusoidal voltages are produced by a synchronous generator in the 25 Hz and 40 Hz points; while in the 50 Hz point, normal utility grid voltage is used. The temperature rise tests were carried out with different average switching frequencies of the inverter and with different rotational speeds of the motor. It should be noted that the cooling conditions with different rotational speed are quite different for a TEFC-motor. A coupled eld-circuit system simulator with a closed-loop control system was used to separate and analyze the drive system losses with the sinusoidal and PWM supply. This paper is organized as follows. Section II concentrates on temperature rise tests, Section III presents the frequency converter loss models used in this study. In Section IV, the simulation method is presented and different loss components are analyzed. The comparison of the simulated and measured

Fig. 1. Measurement setup. The continuous input and output power of 37-kW voltage source converter supplying 37-kW TEFC induction motor was measured with power analyzers. Rotational speed and torque were measured with a torque transducer. The induction motor temperature was measured with Pt-100 sensors.

efciencies is carried out in Section V. Section VI concludes the paper. II. E XPERIMENTAL R ESULTS The measurement setup is given in Fig. 1. The frequency converter was set to the frequency control mode with no slip compensation. Thus, the slip of the induction motor depends on the motor load and losses. The input voltage of the frequency converter was accurately set to 400 V RMS value with a transformer. Because of the nature of the DTC, the switching frequencies used in this paper represent average values of 1-s time intervals. Four different average switching frequencies (1, 2, 3, and 3.75 kHz) were used to nd out the impact of the switching frequency on the motor losses. As a load, another DTC-controlled induction machine was used. The nominal point (n = 100%, T = 100%) of the motor always represents an overload for the motor in a normal frequency converter use. This is a result of the additional harmonic components produced by the PWM and, particularly, converter eld-weakening operation in the 50-Hz operating point if six-step modulation is not used. The nominal load of the 37-kW induction machine in a 50-Hz sinusoidal 400-V supply is 239 Nm. The load value was set to 220 Nm resulting in 92% of the nominal load. With this load, the RMS value of the stator current of the inverter supplied motor is equal to the nominal RMS current in the 50-Hz operating point. To obtain the thermal equilibrium, the motor was kept running for 9 h at a constant load in each test. The temperature rise curves in Figs. 4, 7, and 10 are the average values of three factory installed Pt-100 sensors located in the stator windings. The laboratory temperature was recorded and subtracted from the results. During the temperature tests, the laboratory temperature varied between 25 C and 28 C. The continuous input and output power of the 37-kW voltage source converter supplying the 37-kW TEFC induction motor were measured by two calibrated Yokogawa PZ4000 power analyzers on both sides of the frequency converter. In the following tables, the input power to frequency converter is dened as Pin , and

AARNIOVUORI et al.: MEASUREMENTS AND SIMULATIONS OF DTC VOLTAGE SOURCE CONVERTER

2279

TABLE II M EASURED R ESULTS IN THE 25-Hz O PERATING P OINT

Fig. 2. Waveforms of the 25-Hz three-phase voltages produced with a synchronous generator and their harmonic amplitudes given in percents of the fundamental wave amplitude.

TABLE III M EASURED R ESULTS IN THE 40-Hz O PERATING P OINT

Fig. 3. Measured DTC voltage spectra in 25-Hz operating point. The average switching frequencies are (a) 1 kHz, (b) 2 kHz, (c) 3 kHz, and (d) 3.75 kHz.

the output power from the converter to motor as Pout . The rotational speed, torque, and the shaft power (Pmech in tables) were recorded by a 500-Nm Magtrol torque transducer. Both the THD and TD values provided in Tables II and III are calculated from the measured currents. The THD50 values are calculated from 50 lowest current harmonics. The TD20000 values include all harmonic, interharmonic, and subharmonic components from 1 Hz to 20 kHz with 1-Hz resolution. In both values, the fundamental wave RMS value is used as a scaling factor. In the electric power measurement, the averages of 30 10 s samples with a 10 s sample time were used to minimize the errors. A. 25-Hz Operating Point In the 25-Hz operating point, a synchronous generator was used to produce the nearly sinusoidal supply to the motor. The results of the discrete Fourier analysis of the voltage show that the most signicant harmonic components in the voltage are in the 2nd, 3rd, 5th, and 7th order as shown in Fig. 2. Fig. 3 presents the measured DTC voltage spectra in 25-Hz operating point with different average switching frequencies. Even though the voltage waveform is not purely sinusoidal, the harmonic content of the voltage is minimal compared with the voltage produced with PWM. The THD50 value of the voltage is 0.89%. The motor temperature rises during the last 60 test minutes are shown in Fig. 4. The difference between
Fig. 4. Temperature rises at the end of the temperature test at 25-Hz operating point. The motor has reached its thermal equilibrium.

the sinusoidal and frequency converter supply in the motor temperature rise and motor losses is shown in Fig. 5 as a function of the switching frequency. In Fig. 4, the temperature rise behavior at the end of the test with the sinusoidal supply is due to slight unintentional increase in the generator voltage. In Figs. 4 and 5, we can see that the switching frequency of the inverter has a signicant inuence on the motor temperature. The decreasing rates of the losses and temperature rises are almost similar. In the 25-Hz point, the temperature rise of the machine is roughly one degree Celsius higher against 25 watts of losses. Table II shows that in the sinusoidal supply, the motor current is larger than

2280

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 5, MAY 2012

Fig. 5. Temperature rises and changes in losses compared with the 25-Hz sinusoidal supply as a function of switching frequency.

Fig. 7. Temperature rises at the end of the temperature test at 40-Hz operating point. The motor has reached its thermal equilibrium. More efcient fan cooling results in lower temperatures than at 25-Hz supply.

Fig. 6. Measured DTC voltage spectra in 40-Hz operating point. The average switching frequencies are (a) 1 kHz, (b) 2 kHz, (c) 3 kHz, and (d) 3.75 kHz.

in the frequency converter supply, because the DTC converter uses slightly more voltage than the sinusoidal supply. DTC includes ux controller which keeps the ux linkage constant regardless of the load. The lower rotational speed in the sinusoidal supply is a consequence of the slightly lower frequency (24.84 Hz) than what was desired. If the rotational speed is scaled to 25-Hz voltage, the rotational speed will be 734 rpm. The motor losses increase by 4%14% in the frequency converter supply compared with the sinusoidal supply. The drive efciencies are given in Section V, where the simulation results are compared with the measured ones. B. 40-Hz Operating Point Similar as in the 25-Hz operating point, in the 40-Hz operating point, a synchronous generator was used to produce the sinusoidal supply to the motor. The THD50 value of the grid synchronous generator voltage is 0.65% measured at the motor terminals. The voltage spectra of the DTC in the 40 Hz operating point are shown in Fig. 6. The motor temperature rises during the last 60 test minutes in the 40-Hz operating point are shown in Fig. 7, and the difference between the sinusoidal and frequency converter supply

Fig. 8. Temperature rises and changes in losses compared with the 40-Hz sinusoidal supply as a function of switching frequency.

in motor temperature rise and motor losses is shown in Fig. 8 as a function of the switching frequency. When comparing the results to the 25-Hz operating point, it should kept in mind that cooling of the motor in the 40-Hz operating point is much better than in the 25-Hz point. Results in Figs. 7 and 8 do not behave as assumed. The motor temperature rise with 3-kHz switching frequency is smaller than with 3.75-kHz switching frequency, although the direct loss measurement gives smaller losses at higher switching frequency. In the 40-Hz point, the temperature rise of the machine is roughly one degree Celsius higher against 37 watts of losses. The numerical results in Table III show that the slip of the motor remains constant with all switching frequencies. In the 40-Hz operating point, the motor losses increase by 6%14% in the frequency converter supply compared with the sinusoidal supply. When the motor is driven with a frequency converter, the slip is about 1 min1 greater than when using sinusoidal supply. Similar as at the 25-Hz operating point, the DTC is

AARNIOVUORI et al.: MEASUREMENTS AND SIMULATIONS OF DTC VOLTAGE SOURCE CONVERTER

2281

Fig. 9. Measured DTC voltage spectra in 50-Hz operating point. The average switching frequencies are (a) 1 kHz, (b) 2 kHz, (c) 3 kHz, and (d) 3.75 kHz. TABLE IV M EASURED R ESULTS IN THE 50-Hz O PERATING P OINT

Fig. 10. Temperature rise differences and loss changes compared with the 50-Hz sinusoidal supply as a function of average switching frequency in the DTC supply.

driving the motor with a higher voltage than what was used with sinusoidal supply. This leads to a slightly smaller RMS current. C. 50-Hz Operating Point In the 50-Hz operating point, the same load and switching frequencies were used as in the 25-Hz and 40-Hz point. The grid voltage THD50 is 1.24%. Fig. 9 shows the voltage spectra for different average switching frequencies. In frequency converter use, the 50-Hz point is either in the eld weakening or overmodulation range because of the voltage losses in the frequency converter input rectier and lters. In addition to the voltage losses, the converter selects a suitable voltage reserve to be able to control quick load changes. The mechanical power remains almost constant in all measurements when the frequency converter supply is used. The inverter output power decreases as the switching frequency is increased, Table IV. As shown in Fig. 10, in the 50-Hz point, the temperature rise of the machine increases roughly by one degree Celsius against 12 watts of losses. The temperature rises in Fig. 11 at 2, 3, and 3.75 kHz switching frequencies are almost constants and so are the measured losses. The increase in the motor losses is 1820% in the frequency converter supply compared with the sinusoidal supply. The THD values in Tables IIIV show that in the 50-Hz operation point, the currents in the frequency converter supply have a higher harmonic content than in 25-Hz and 40-Hz operating points.

Fig. 11. Temperature rises at the end of the 50-Hz temperature rise test.

In frequency converter supply, this is an outcome that results from the eld-weakening operation. The DTC controller leaves a voltage reserve; therefore, the studied converter does not use the full modulation index and hence the eld weakening starts at 45 Hz. Another reason of the higher harmonic content is that the inverter can, naturally, with the same switching frequency use only half of the number of switchings per one fundamental wave when compared with the 25-Hz point, therefore, resulting in a more coarse voltage waveform. In Fig. 12, the grid voltage and the motor current during the temperature test with the 3-kHz average switching frequency are shown. The motor phase current amplitude follows the changes in the grid voltage. In fact, the similar temperature rises in Fig. 10 with different switching frequencies can be explained by different operating points of the motor. As given in Table IV, the fundamental wave voltage amplitude is slightly higher, when the switching frequency is low and the fundamental voltage is decreasing from the switching frequency of 1 kHz to

2282

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 5, MAY 2012

a single series resistance-inductance model. The total losses of the input inductor are Pchoke (t) = 3i2 L (t)ESRL (1)

where iL (t) is the instantaneous inductor line current, and ESRL is the equivalent series resistance. ESRL is calculated by using the measured current and losses of the input inductor. B. Diode Bridge Losses The power dissipation of a diode in forward conduction and reverse blocking state can be modeled as a function of forward and reverse leakage currents and voltages. Losses in the blocking state are negligible. Diode switching losses can be considerable but for a line frequency diode bridge rectier, the switching losses are marginal (fall within error marginal) and only the conduction losses are considered. The parameters needed for calculating the diode bridge losses are the forward voltage drop UF and the on-state resistance RF . The instantaneous conduction losses of the diode are [28] Pdiode,on (t) = UF idiode (t) + RF i2 diode (t). (2)

Fig. 12. Grid phase voltage and motor current during a temperature rise test. The voltage and current curves are from the temperature rise test carried out with a 3-kHz average switching frequency.

The diode bridge losses are six times the single diode losses. C. Intermediate Circuit Losses
Fig. 13. Frequency converter loss model has ve main components.

3 kHz. The higher fundamental wave voltage leads to a lower current, and the dominating resistive losses become smaller. This distorts the impact of the switching frequency on the motor losses. III. F REQUENCY C ONVERTER L OSS M ODELS The losses in the frequency converter are divided into ve groups: input inductor, diode rectier, intermediate circuit, IGBT module, and extra losses. Losses in auxiliary devices such as fans are included, Fig. 13. A. Input Inductor Losses The input inductors of the frequency converter dissipate power in the core and in the windings. The core losses can be divided into hysteresis losses and eddy current losses [31]. Although the exact calculation of these losses is complicated, they can be estimated using data sheet parameters available from magnetic component suppliers or they can be dened by measurements. The input inductor losses and their frequency dependency can be modeled with the lumped parameters model, Cauer, or Foster equivalent models [32]. In the analysis, only the inductance at nominal point, the dc resistance, and total losses at one load point are known. Therefore, lumped parameters cannot be used, and the losses have to be modeled with

The intermediate circuit of the modeled device does not contain dc link reactors. Therefore, the losses consist of only the losses in the capacitor bank and discharge resistors. The total energy loss in a capacitor bank is a function of dielectric losses attributed to the polarizing mechanisms of the electric eld on the molecular structure of the dielectric, and ohmic loss from electrodes and termination metals. Dissipative losses of the capacitor can, again, be represented by equivalent series resistance whose values or a curve of frequency dependency can be found in manufacturers datasheets. The number of the series and parallel capacitors in the bank has to be taken into account to calculate the capacitor bank losses correctly. The dc link current consists of a dc component IDC , harmonics produced by the diode rectier bridge, and the switching harmonics by the IGBT-inverter bridge. The dc voltage produced by a three-phase full-bridge rectier carries large amounts of n times sixth-order harmonics. The average capacitor losses PC can be written PC =
n 2 IC(6 n) ESRC(6n)

(3)

where IC (n) is the RMS value of nth order of the capacitor current and ESRC(n) is the equivalent series resistance of the capacitor for a particular frequency. ESR values for ve different harmonic frequencies have been used in the loss model. At 50-Hz line frequency, these harmonic frequencies are 300, 600, 900, 1200, and 3000 Hz. The lower frequency ESR values describes the losses at harmonic currents produced by the rectier bridge, and the highest frequency ESR value

AARNIOVUORI et al.: MEASUREMENTS AND SIMULATIONS OF DTC VOLTAGE SOURCE CONVERTER

2283

describes the losses in the switching frequency range. The resistive losses in the parallel discharge resistor are Pdischarge =
2 UC

Rdischarge

(4)

where UC is the capacitor RMS voltage. The intermediate circuit conduction losses are neglected in this case. Thus, the total intermediate circuit losses are PDClink = PC + Pdischarge . (5)
Fig. 14. Principle of the simulation software.

D. IGBT Module Losses The IGBT module losses comprise the conducting and switching losses of a particular device. The same loss models are used for both semiconductor devicesthe IGBT and its antiparallel free-wheeling diode. For the IGBTs and diodes, the instantaneous conducting losses are [33] PIGBT,cond = UCE0 im (t) + RCE0 i2 m (t) and Pdiode,cond = UF0 im (t) + RF0 i2 m (t) (7) (6)
TABLE V F REQUENCY C ONVERTER L OSS S IMULATION PARAMETERS

where im is motor phase current, UCE0 is the IGBTs threshold voltage, RCE0 is the IGBTs on-state resistance, and UF0 and RF0 are the corresponding values of the diode. For the switching losses of the IGBTs and diodes, the same linear loss model is used [33]. The average switching losses for a specic period of time are PIGBT,sw = and Pfwdiode,sw = UDC Im Esw,fwdiode Nsw,change Urated Irated (9) the main features and problems of different techniques are summarized. In this paper, the coupling method referred to as current output approach in [35] is applied. The circuit simulator is used to calculate the frequency converter losses, and the FEM motor model calculates motor losses; additional information of coupling method can be found in [36] and [37] The FEM model is based on a 2-D nite-element method program and a circuit equation of the windings [38]. The copper loss calculation is explained in [38] and the iron loss calculation in [39]. The principle of simulation software is shown in Fig. 14 and the frequency converter simulation parameters are given in Table V. The rst-order mesh of the 37-kW induction motor used in the FE analysis is shown in Fig. 15. The FEM motor model is calculated with 25 s time steps. The FEM motor losses are an average value of a 1-s simulation time containing 40 000 steps. The motor loss results could be improved using second-element mesh or using smaller time steps in the calculation, but in this context, when the total drive system is simulated, the rst-order mesh gives adequate results. The circuit simulator is running with 1 s time steps, and the voltage fed to the FEM model is the average value of 25 steps. The simulation software does not include any cooling or thermal model motor or frequency converter. UDC Im Esw,IGBT Nsw,change Urated Irated (8)

where Nsw,change is the number of the switch changes during the specic time period. Esw is the switching loss energy of a particular device given for the reference commutation voltage and current. UDC and Im are the actual commutation voltage and current. E. Auxiliary Devices Losses The auxiliary devices losses are comprised of the inverter self-usage, for instance microcontroller, internal power supply, display, keyboard, bus-communication, digital and analog inputs and outputs, and the blower and control system power consumption. In this case, these losses are about constant. The fan speed is not load dependent. IV. S IMULATIONS In [34], the coupling of 2-D FEM equations with external circuit equations is exhaustively addressed. [34] also reviews, analyzes, and classies the different coupling methods. Also,

2284

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 5, MAY 2012

TABLE VIII M OTOR L OSS S IMULATION R ESULTS IN THE 40-Hz O PERATING P OINT

TABLE IX F REQUENCY C ONVERTER L OSS S IMULATION R ESULTS IN THE 40-Hz O PERATING P OINT Fig. 15. The rst-order mesh of the 37-kW induction motor used in the FE analysis. The mesh contains 1448 elements and 917 nodes. TABLE VI M OTOR L OSS S IMULATION R ESULTS IN THE 25-Hz O PERATING P OINT

TABLE X M OTOR L OSS S IMULATION R ESULTS IN THE 50-Hz O PERATING P OINT

TABLE VII F REQUENCY C ONVERTER L OSS S IMULATION R ESULTS IN THE 25-Hz O PERATING P OINT

A. Simulation Results at 25 Hz The simulation software does not take friction and windage or all additional (stray) losses into account. These losses are added to the simulated losses. The results are given in Table VI. The additional losses are proportional to the square of the load current and to the power of 1.5 of the frequency [40] Pad I 2 f 1.5 . (10)

by efciency measurements (IEC 60034-2-1) in the nominal point. In Table VI, the simulated and measured motor loss values are close to each other. The rotor losses decrease as the switching frequency is increased, as assumed. The simulation results show that the extra iron losses are the most signicant additional losses that the PWM produces. The frequency converter simulation results are given in Table VII. It is obvious that the input choke, diode bridge, and dc link losses remain constants when the switching frequency is changed. The simulator underestimates the effect of the switching frequency on the IGBT module losses. In the measurements, rising the switching frequency from 1 to 3.75 kHz gives 220 watts of more iron losses, whereas the simulated losses increase only by 120 watts. B. Simulation Results at 40 Hz The simulation results are given in Tables VIII and IX. In Table VIII, the simulated and measured motor loss values with PWM converter supply are close to each other. In the

Friction losses of the motor are directly proportional to the speed, and windage losses are proportional to the third power of speed. Because the ratio of the friction and windage losses is not known, the total friction and windage losses are assumed here to be proportional to the square of the speed. These losses given in Tables IV and VI are scaled loss values obtained

AARNIOVUORI et al.: MEASUREMENTS AND SIMULATIONS OF DTC VOLTAGE SOURCE CONVERTER

2285

TABLE XI F REQUENCY C ONVERTER L OSS S IMULATION R ESULTS IN THE 50-Hz O PERATING P OINT

Fig. 17. Measured and simulated efciencies in the 40-Hz point.

Fig. 16. Measured and simulated efciencies in the 25-Hz point.

simulated losses, the most signicant loss change can be seen in rotor copper losses; they are decreasing when the switching frequency is increased, as assumed. Total iron losses decrease from 791 W to 775 W when the average switching frequency is increased from 1 to 3.75 kHz. Also, minor loss changes in the stator copper losses can be seen. The simulated frequency converter losses are slightly smaller than the measured ones. The measured frequency converter losses are increasing linearly as the switching frequency is increased. Similarly, as in the 25-Hz point, the simulator underestimates the effect of switching frequency on frequency converter losses. C. Simulation Results at 50 Hz The 50-Hz simulated losses behave more logically than the measured ones because of the different operating conditions

of the motor as explained above. The differences between measured and simulated losses are very small. The motor losses in the 50-Hz points behave similarly as in the 25-Hz and 40-Hz points, even though the changes in the losses are smaller than in the 25- or 40-Hz points. Corresponding results can be seen in the measured temperature rises and losses. The impacts of the switching frequency and the PWM-caused losses are smaller in the 50-Hz points than in the 25-Hz points. The simulated frequency-converter-caused motor losses are given in Table X and the simulated frequency converter losses in Table XI. The measured losses at the 3.75-kHz switching frequency are less than at 3 kHz; one reason for this is the slightly smaller current at 3.75 kHz and the other one that the switching frequency of the frequency converter has been limited for thermal protection of the IGBTs. Otherwise, the simulated and measured frequency converter losses behave similarly as in the 25-Hz and 40-Hz points. V. C OMPARISON OF THE R ESULTS The simulated motor losses with sinusoidal supply are over 200 watts less in the 25-Hz and 40-Hz operating points than the measured ones. The difference can be explained with pure sinusoidal voltage that was used in the simulations and harmonic components that exist in the generator produced voltage that was used in the measurements. There exists no other explanation than measurement error in the shaft power, why the simulated motor losses at 50-Hz operating point with sinusoidal supply are greater than the measured ones.

2286

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 5, MAY 2012

The additional losses are a function of rotational speed and switching frequency. In this case, the motor efciency drop was 0.2%1.7% units when PWM supply was used compared with the sinusoidal supply depending on the switching frequency and operating point of the motor. Furthermore, the simple, linear loss models used for frequency converter loss calculation are a source of error and may not give accurate results in every operating point of the converter. For the scope of this paper, to obtain the drive system losses without any measurements, the loss models are well suitable. The coupled circuit simulator used in this study gives encouraging results in the VSD simulation. The ability to estimate drive system efciencies with a reasonable accuracy without conrming measurements is a powerful resource in scientic and practical work. R EFERENCES
[1] E. N. Hidebrand and H. Roehrdanz, Losses in three-phase induction machines fed by PWM converter, IEEE Trans. Energy Convers., vol. 16, no. 3, pp. 228233, Sep. 2001. [2] A. Boglietti, A. Cavagino, and A. M. Knight, Factors affecting losses in induction motors with non-sinusoidal supply, in Conf. Rec. Ind. Appl. Conf., New Orleans, LA, 2007, vol. 1, pp. 11931199. [3] J.-J. Lee, Y.-K. Kim, H. Nam, K.-H. Ha, J.-P. Hong, and D.-H. Hwang, Loss distribution of three-phase induction machines fed by pulsewidthmodulated inverter, IEEE Trans. Magn., vol. 40, no. 2, pp. 762765, Mar. 2004. [4] Y. Wu, R. A. McMahon, Y. Zhan, and A. M. Knight, Impact of PWM schemes on induction motor losses, in Conf. Rec IEEE 41st Ind. Appl. Conf., Oct. 2006, vol. 2, pp. 813818. [5] A. Ruderman, Electrical machine PWM loss evaluation basics, in Proc. Energy Efciency Motor Driven Syst., Heidelberg, Germany, Sep. 2005, pp. 5868. [6] Y. Zhan, A. M. Knight, Y. Wu, and R. A. McMahon, Investigation and comparison of inverter-fed induction machine loss, in Proc. IEEE Ind. Appl. Soc. Annu. Meeting, Oct. 2008, pp. 16. [7] Z. Gmyrek, A. Boglietti, and A. Cavagnino, Estimation of iron losses in induction motors: Calculation method, results, and analysis, IEEE Trans. Ind. Electron., vol. 57, no. 1, pp. 161171, Jan. 2010. [8] A. Boglietti, A. Cavagnino, and M. Lazzari, Fast method for the iron loss prediction in inverter-fed induction motors, IEEE Trans. Ind. Appl., vol. 46, no. 2, pp. 806811, Mar./Apr. 2010. [9] A. Boglietti, A. Cavagnino, D. M. Ionel, M. Popescu, D. A. Staton, and S. Vaschetto, A general model to predict the iron losses in PWM inverter-fed induction motors, IEEE Trans. Ind. Appl., vol. 46, no. 5, pp. 18821890, Sep./Oct. 2010. [10] L. Ruifang, C. C. Mi, and D. W. Gao, Modeling of iron losses of electrical machines and transformers fed by PWM inverters, IEEE Trans. Magn., vol. 44, no. 8, pp. 20212028, Aug. 2008. [11] K. Yamazaki and N. Fukushima, Iron loss model for rotating machines using direct eddy current analysis in electrical steel sheets, IEEE Trans. Energy Convers., vol. 25, no. 3, pp. 633641, Sep. 2010. [12] A. Boglietti, A. Cavagnino, and A. M. Knight, Isolating the impact of PWM modulation on motor iron loss, in Conf. Rec. IEEE IAS Annu. Meeting, Edmonton, Canada, Oct. 2008, pp. 17. [13] T. Haring, Design of motors for inverter operation, in Energy Efciency Improvements in Electronic Motors and Drives, P. Bertoldi, A. T. de Almeida, and H. Falkner, Eds. Berlin, Germany: SpringerVerlag, 2000. [14] Rotating Electrical MachinesPart 31: Guide for the Selection and Application of Energy-Efcient Motors Including Variable Speed Drives, IEC/TS 60034-31 Ed 1, 2009. [15] A. Mihalcea, B. Szadados, and J. Hoolboom, Determining total losses and temperature rise in induction motors using equivalent loading methods, IEEE Trans. Energy Convers., vol. 16, no. 3, pp. 214219, Sep. 2001. [16] I. Takahashi and T. Noguchi, A new quick response and high efciency control strategy of an induction motor, IEEE Trans. Ind. Appl., vol. IA22, no. 5, pp. 820827, Sep. 1986.

Fig. 18. Measured and simulated efciencies in the 50-Hz point.

The simulated and measured frequency converter, motor, and drive efciencies are shown in Figs. 1618. In the 25-Hz point, the measured motor efciency is 91.0% with the sinusoidal supply, and the efciency varies from 90% to 90.8% in the frequency converter use. In the 40-Hz point, the measured motor efciency is 93.6% with sinusoidal supply and varies from 91.9% to 92.4% in the frequency converter use. In the 50-Hz point, the measured motor efciency is 93.6% and varies from 92.5% to 92.9% in the frequency converter use. In the 50-Hz point, the measurement shows that the drive system efciency is at its best value at the low switching frequency. In the 40-Hz operating point, the highest switching frequency maximizes the drive efciency. In the 25-Hz point, the 2-kHz switching frequency maximizes the efciency of the drive system. Even though the simulated losses do not exactly match the measured ones, the efciencies of the drive system in different operational point can be simulated with a reasonable accuracy. VI. D ISCUSSION AND C ONCLUSION Frequency converter losses are difcult to measure accurately because of the distorted input and output currents and output voltages of the converter. Extra care should be taken at 50-Hz operation with a frequency converter. An accurate efciency result of the variable-speed drives requires a very constant load and line voltage level as well as good measuring equipment. The differences between the simulated and measured efciencies are so small that they cannot be separated from the error originating from measuring setup.

AARNIOVUORI et al.: MEASUREMENTS AND SIMULATIONS OF DTC VOLTAGE SOURCE CONVERTER

2287

[17] M. Depenbrock, Direct self-control (DSC) of inverter-fed induction machine, IEEE Trans. Power Electron., vol. 3, no. 4, pp. 420429, Oct. 1998. [18] D. Casadei, G. Serra, and A. Tani, Analytical investigation of torque and ux ripple in DTC schemes for induction motors, in Proc. 23rd Int. Conf. Ind. Electron., Control Instrumen., New Orleans, LA, Nov. 1997, vol. 2, pp. 552556. [19] J.-K. Kang and S. K. Sul, Analysis and prediction of inverter switching frequency in direct torque control of induction machine based on hysteresis bands and machine parameters, IEEE Trans. Ind. Electron., vol. 48, no. 3, pp. 545553, Jun. 2001. [20] V. Ambrozic, M. Bertoluzzo, G. S. Buja, and R. Menis, An assessment of the inverter switching characteristics in DTC induction motor drives, IEEE Trans. Power Electron., vol. 20, no. 2, pp. 457465, Mar. 2005. [21] N. R. N. Idris, C. L. Toh, and M. E. Elbuluk, A new torque and ux controller for direct torque control of induction machines, IEEE Trans. Ind. Appl., vol. 42, no. 6, pp. 13581366, Nov./Dec. 2006. [22] N. R. N. Idris and A. H. M. Yatim, Direct torque control of induction machines with constant switching frequency and reduced torque ripple, IEEE Trans. Ind. Electron., vol. 51, no. 4, pp. 758767, Aug. 2004. [23] M. Hajian, J. Soltani, G. A. Markadeh, and S. Hosseinnia, Adaptive nonlinear direct torque control of sensorless IM drives with efciency optimization, IEEE Trans. Ind. Electron., vol. 57, no. 3, pp. 975985, Mar. 2010. [24] C. Lascu, I. Boldea, and F. Blaabjerg, Direct torque control of sensorless induction motor drives: A sliding-mode approach, IEEE Trans. Ind. Appl., vol. 40, no. 2, pp. 582590, Mar./Apr. 2004. [25] K.-K. Shyu, J. K. Lin, V.-T. Pham, M.-J. Yang, and T.-W. Wang, Global minimum torque ripple design for direct torque control of induction motor drives, IEEE Trans. Ind. Electron., vol. 57, no. 9, pp. 31483156, Sep. 2010. [26] J.-K. Kang and S.-K. Sul, New direct torque control of induction motor for minimum torque ripple and constant switching frequency, IEEE Trans. Ind. Appl., vol. 35, no. 5, pp. 10761082, Sep./Oct. 1999. [27] G. C. D. Sousa, B. K. Bose, J. Cleland, R. J. Spiegel, and P. J. Chappell, Loss modeling of converter induction machine system for variable speed drive, in Proc. IEEE IECON , Nov. 1992, vol. 1, pp. 114120. [28] J. W. Kolar, H. Ertl, and F. C. Zach, Inuence of the modulation method on the conduction and switching losses of a PWM converter system, IEEE Trans. Ind. Appl., vol. 27, no. 6, pp. 10631075, Nov./Dec. 1991. [29] Y. Wu, M. A. Sha, A. M. Knight, and R. A. McMahon, Comparison of the effects of continuous and discontinuous PWM schemes on power losses of voltage-sourced inverters for induction motor drives, IEEE Trans. Power Electron., vol. 26, no. 1, pp. 182191, Jan. 2011. [30] Y. Wu, S. Shao, and R. A. McMahon, Power loss study of inverter-fed machine drives using discontinuous pulse width modulation, in Proc. IEEE Ind. Conf Sustainable Energy Technol., Now. 2008, pp. 11721177. [31] R. Erickson, Fundamentals of Power Electronics. London, U.K.: Chapman & Hall, 1997. [32] F. de Len and A. Semlyen, Time domain modeling of eddy current effects for transformer transients, IEEE Trans. Power Del., vol. 8, no. 1, pp. 271280, Jan. 1993. [33] M. H. Bierhoff and F. W. Fuchs, Semiconductor losses in voltage source and current source IGBT converters based on analytical derivation, in Proc. IEEE 35th PESC, Aachen, Germany, Jun. 2000, vol. 4, pp. 2836 2842. [34] B. Asghari, V. Dinavahi, M. Rioual, J. A. Martinez, and R. Iravani, Interfacing techniques for electromagnetic eld and circuit simulation programs IEEE Task Force on Interfacing Techniques for Simulation Tools, IEEE Trans. Power Del., vol. 24, no. 2, pp. 939950, Apr. 2009. [35] S. Kanerva, Simulation of electrical machines, circuits and control systems using nite element method and system simulator, Ph.D. dissertation, Helsinki Univ. Technol., Espoo, Finland, 2005. [36] S. Kanerva, S. Seman, and A. Arkkio, Simulation of electric drive systems with coupled nite element analysis and system simulator, in Proc. 10th Eur. Conf. Power Electron. Appl. EPE, Toulouse, France, Sep. 24, 2003. [37] S. Kanerva, C. Stulz, B. Gerhard, H. Burzanowska, J. Jrvinen, and S. Seman, Coupled FEM and system simulator in the simulation of asynchronous machine drive with direct torque control, in Proc. ICEM , Cracow, Poland, Sep. 58, 2004.

[38] A. Arkkio, Analysis of induction motors based on the numerical solution of the magnetic eld and circuit equations, Ph.D. Dissertation, Helsinki Univ. Technol., Otaniemi, Finland, 1987. [39] A. Belachen and A. Arkkio, Comprehensive dynamic loss model of electrical steel applied to fe simulation of electrical machines, IEEE Trans. Magn., vol. 44, no. 6, pp. 886889, Jun. 2008. [40] J. Pyrhnen, T. Jokinen, and V. Hrabovcov, Design of Rotating Machines. Hoboken, NJ: Wiley, 2008.

Lassi Aarniovuori was born in 1979, Finland, in Jyvskyl. He received the M.Sc. and D.Sc. degrees in electrical engineering from Lappeenranta University of Technology (LUT), Lappeenranta, Finland, in 2005 and 2010. He is currently a Researcher for the Department of Electrical Engineering at LUT. His current research interests are in the eld of electric drives, particularly simulation of electric drives, efciency measurements, and calorimetric measurement systems.

Lasse I. E. Laurila was born in Helsinki, Finland, in 1971. He received the M.Sc., Lic.Tech. and D.Sc. degrees from Lappeenranta University of Technology (LUT), Lappeenranta, Finland, in 1996, 2000 and 2004, respectively. His research interests include power electronic converters, variable-speed drives, and their control. He is currently interested in electrical energy recovery in mobile work machines. He is currently an Assistant Professor at the Laboratory of Electrical Drives Technology in LUT.

Markku Niemel was born in Mntyharju, Finland, in 1968. He received the B.Sc. degree in electrical engineering from Helsinki Institute of Technology, in 1990. He received the M.Sc. and D.Sc. (Technology) degrees from Lappeenranta University of Technology (LUT), Lappeenranta, Finland, in 1995 and 1999, respectively. He is currently a Senior Researcher with the Carelian Drives and Motor Centre in LUT. His current interests include motion control, control of line converters, and energy efciency of electric drives.

Juha J. Pyrhnen (M06) was born in Kuusankoski, Finland, in 1957. He received the D.Sc. degree from Lappeenranta University of Technology (LUT), Lappeenranta, Finland, in 1991. He has been a Professor of Electrical Machines and Drives since 1997. He is Head of the Department of Electrical Engineering in the Institute of LUT Energy. He is engaged in research and development of electric motors and drives.

You might also like