You are on page 1of 9

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 23, NO.

2, APRIL 2008

1063

The Inuence of Induction Motors on Voltage Sag PropagationPart I: Accounting for the Change in Sag Characteristics
Jovica V. Milanovic, Senior Member, IEEE, Myo T. Aung, Member, IEEE, and Sarat C. Vegunta
AbstractThis paper presents an analytical approach to the analysis of the interaction between induction motors (IMs) and voltage sags. The presented methodology enables quick assessment of the inuence of IM on sag characteristics and avoids time-consuming transient simulations. The accuracy of the derived analytical formulae is veried through classical transient simulations using the PSCAD/EMTDC package. This paper further proposes a simple, yet efcient transformation of resulting nonrectangular voltage sags, based on the loss of voltage index, into equivalent rectangular sags suitable for conventional voltage sag performance benchmarking. Index TermsInduction motors (IMs), power quality (PQ), voltage sags.

I. INTRODUCTION OWER-QUALITY (PQ) issues continue to receive significant attention due to their potentially huge economic impact on the industry. Voltage sags, transients, and momentary interruptions are among the most common PQ disturbances and, by large, the most frequent cause of equipment malfunctions and costly disruptions of industrial processes [1]. The IEC 61000-4-30 standard denes the voltage dip (sag) as a temporary reduction of the voltage at a point of the electrical system below a threshold. This threshold is dened by the IEEE 1159-1995 standard as 90% of the nominal voltage. If the magnitude of the retained voltage drops below 10% of the rated voltage, the disturbance is classied as an interruption. Most voltage sags and short interruptions recorded in actual power systems last from a few cycles to 1 min. Voltage sag characteristics (i.e., magnitude, duration, phase-angle jump, sag shape, etc.) change when sags propagate through power system networks. The extent of this change depends on network topology (mesh or radial), line and cable impedances, transformer winding connections, system protection and grounding practices, load connections (star or delta), and load dynamics (e.g., IMs) [2], [3]. The inuence of all of these factors has been
Manuscript received October 18, 2006; revised June 5, 2007. This work was supported in part by the EU Framework V Project Microgrids and in part by the EU Framework VI Project RISE under Contract FP6-INCO-CT-2004-509161. Paper no. TPWRD-00650-2006. J. V. Milanovic and S. C. Vegunta are with the School of Electrical and Electronic Engineering, The University of Manchester, Manchester, M60 1QD, U.K. (e-mail: milanovic@manchester.ac.uk). M. T. Aung is with the University of Bath, Bath, BA2 7AY, U.K. (e-mail: m.t.aung@bath.ac.uk). Color versions of one or more of the gures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identier 10.1109/TPWRD.2007.915846

studied in the past either by considering an individual effect or through more inclusive studies that dealt with several inuences at the same time (typically not all of them). A large number of published and easily accessible reports describe the results of those studies. One of the major deciencies of the past studies remains as the accounting for dynamic effects in large system sag propagation studies. Typically, sag calculation and propagation in large systems are studied by applying static fault calculations based on system impedance (or admittance) matrix and equivalent positive-, negative-, and zero-sequence networks. It is further assumed that voltage sags originate from faults (cleared by primary protection) on transmission and/or distribution systems and that they have a rectangular shape. The assumption about the shape is based on neglecting the change in fault impedance during the fault and the effect of load dynamics. The shape of voltage sags, however, ceases to be rectangular if there is an appreciable number of IMs in the network and the faults last longer (e.g., distribution system faults or faults cleared by secondary protection). In such cases, the shape of the sag changes due to the IMs dynamic responses. When a voltage sag appears at the terminals of an induction motor, the torque and the speed of the motor decrease to levels below their nominal values. Once the voltage sag is removed, the IM attempts to reaccelerate, and starts drawing an excessive amount of current from the power supply. The ow of such current through the supply impedance prevents fast recovery of the voltage and causes a prolonged sag duration and change in the shape of the sag. This new resulting sag may cause sensitive equipment to trip even though it might have been able to ride through the original voltage sag. Effectively, all existing, standard benchmarking approaches (e.g., SARFI, DISDIP, ESKOM, and UNIPED sag tables, 3-D cross tabulation of the rms variation, magnitude-duration barcharts, etc.) register voltage sags as rectangular events [4]. (For example, for a typical voltage sag, only the information about phase-to-neutral or line-to-line voltage magnitude and sag duration is provided.) These tabulated results are then used for comparing the sag performance of different buses/systems and even for the assessment of equipment/process sensitivity to voltage sags. A few past works (e.g., [3] and [5]) dealt with the inuence of IMs on voltage sags. They typically modeled IMs using dynamic models and, therefore, are not suitable for large system applications. The results of that analysis could not be presented using standardized, widely accepted voltage sag tables due to nonrectangularity of the resulting voltage sags.

0885-8977/$25.00 2008 IEEE

1064

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 23, NO. 2, APRIL 2008

Fig. 1. Simplied system impedance model.

To capture the IM behavior during the sag and its inuence on sag characteristics, a comprehensive analytical tool was developed and partially reported in [5]. The method presented there described the general behavior of IMs when subjected to symmetrical and asymmetrical voltage sags using positive- and negative-sequence IM equivalent circuits. It did not discuss the inuence of IM operating condition and parameters on the characteristics of the resulting voltage sag. This paper builds on and is an extension of the work presented in [5]. It considers comprehensively the inuence of IM on individual phase voltages and the conversion of resulting nonrectangular voltage sags into rectangular ones suitable for conventional sag benchmarking purposes. The results of developed analytical method (implemented in Matlab [6] and based on static calculations) are veried using classical transient simulations in PSCAD/EMTDC [7]. The proposed methodology accounts for the inuence of IM on voltage sag characteristics and propagation, without performing lengthy transient simulations, making it applicable for large, practical system studies where hundreds of sags at numerous bases have to be evaluated. It further enables studies of propagation of voltage sags from the point of common coupling (PCC) to the end-user equipment in an industrial plant with a number of IMs or the propagation of voltage sags through the electrical system of an industrial park to be built. The major advantage of the method is that it starts with conventionally recorded/represented voltage sags and produces the output in the same, universally adopted, format. II. MODELLING OF AGGREGATE IM FOR VOLTAGE SAG STUDIES Voltage sagIM interaction is studied based on a group of identical IMs connected at the PCC (as shown in Fig. 1) and subjected to a typical, asymmetrical, three-phase, rectangular voltage sag. IM and sag parameters are given in Appendix A. The model is implemented in Matlab to facilitate static calculations. The integration step of 20 ms (to account for the delay in rms calculation in PSCAD/EMTDC) is used in order to validate the results by transient simulations in PSCAD/EMTDC. The methodology is developed for the most general case of an asymmetrical sag and a group of IMs, but it can be easily simplied and applied either to a single IM or a group of IMs subjected to symmetrical voltage sags. A. Sequence Components Voltages For a given presag, during the sag and after sag voltages (see Appendix C), corresponding sequence component voltages can be calculated as follows: (1)

Fig. 2. Simplied induction motor equivalent circuits. (a) Positive-sequence circuit. (b) Negative-sequence circuit.

With

, where

is a com-

plex operator. It is assumed that at the end of the voltage recovery, the voltages in all three phases reach their respective nominal presag values, i.e., (2)

B. Aggregation of IMs The aggregate model of a group of IMs retains the same form as a model of an individual motor of the group. The aggregated parameters of the equivalent IM representing a group of nidentical (or different) IMs can be calculated as suggested in [8]. However, for simplicity, only identical IMs are chosen in this study and could well extend to a group of nonidentical IMs. For identical IMs, the aggregated parameters are obtained by dividing electrical parameters of an individual motor by (number of aggregated motors) and multiplying mechanical pa, e.g., rameters by

(3) (Note: The aggregated motor parameters are represented in the paper with sufx agg.) C. Electromagnetic Torque Variations Fig. 2 shows the equivalent positive- and negative-sequence circuits of the IM. Since the magnetizing impedance of the IM is much larger than the stator and the rotor impedances, it is neglected. The stator and the rotor currents are assumed equal. The positive- and negative-sequence electromagnetic torques are given in [9] as (4) (5)

MILANOVIC et al.: INFLUENCE OF INDUCTION MOTORS ON VOLTAGE SAG PROPAGATIONPART I

1065

Under balanced or unbalanced voltage sags, (6) can be transformed to a rst-order differential (8) which, when solved, gives the expression (9) for or motor deceleration speed (8)

(9) where
Fig. 3. Speed-torque curves under normal and voltage sag conditions [4].

where

and At the end of the sag, the balanced nominal voltage supply is restored, resulting in positive-sequence torque only. This now at time to causes the motor to reaccelerate from at time . During this condition, (6) can be transformed to (10). The solution of (10) gives the expression (11), which represents or motor acceleration speed voltage recovery speed (10)

Under balanced supply conditions, only positive-sequence components exist and they contribute to the development of positive-sequence torque only. During unbalanced supply conditions, however, negative-sequence components appear, giving rise to negative-sequence torque, which opposes the existing positive-sequence torque. The resulting electromagnetic torque, or net torque, in the machine under unbalanced operating conditions, is a vector sum of positive- and negative-sequence torques (6)

(11) where . The reacceleration time from the sag, is given by to and , at the end of

D. Machine Speed Variation During and After Voltage Sag The presence of either balanced or unbalanced voltage sags at the IM terminals will cause electrical and mechanical transients in the IM. The dynamic performance of the IM during the sag can be described by the following electromechanical equation: (12)

where

E. Machine Current Variations During and After Voltage Sag (7) Due to small time constants (fast response of electrical circuits of the IM), the electrical transients in the IM are mostly nished before the onset of the mechanical transients [3], [5]. The electrical torque in the machine therefore will reduce in proportion to the square of the rms value of the voltage sag magnitude, while the mechanical torque will remain largely unchanged. Due to the reduction in electrical torque, the motor operates at a new operating point given by intersection of the motor-speed curve during the voltage sag and load torque characteristic, as shown in Fig. 3. The speed of motor reduces from value at time to that at the end of the the nominal sag at time along the load torque curve. Depending on the duration of the voltage sag, the during voltage sag speed will be between and . Positive- and negative-sequence currents during the sag are given by (13) (14) where ; and

At the onset of the voltage sag, the positive-sequence current falls almost instantly with a drop in terminal positive-sequence voltage. As the voltage sag magnitude stays constant during the sag (for a sag duration ), the positive-sequence current increases gradually followed by a gradual increase in motor slip

1066

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 23, NO. 2, APRIL 2008

during the sag; hence, the motor slows down. However, in the case of unbalanced operation, the negative-sequence current rises instantly with the appearance of voltage unbalance and as negative-sequence voltage stays constant during the sag, the current remains almost unchanged with little or a negligible ). change in slip (as Similarly, positive- and negative-sequence currents after the sag are given by (15) (16) where . At the start of voltage recovery, the voltage rises to nominal balanced value and the slip is large (depending on the speed drop seen at the end of the sag). These contribute to a large positive-sequence inrush current (to build up the ux in the air gap) which gradually decays as the motor accelerates (decreasing motor slip) to nominal speed. As the nominal and balanced voltage is restored, the negative-sequence current drops close to zero. III. INDUCTION MOTORVOLTAGE SAGS INTERACTION A. Basic Change in Sag Shape Due to IM Dynamics As a result of this continuous variation in the motor currents, the motor terminal voltages will not remain constant and a change in terminal voltages occurs in accordance with the variation in currents. To describe this phenomenon, a simple test system consisting of an aggregated IM connected to the in, as shown nite (system) bus through a system impedance in Fig. 1, is considered. During the voltage sag, positive- and negative-sequence voltages at the system bus are calculated using (1) and positive- and negative-sequence currents at the end of voltage sag are obtained and from (13) and (14). Once the initial conditions ( ) are established, the variation in IM terminal voltages (i.e., voltages at the PCC) are calculated from (17) (18) (19) (20) Similarly, positive- and negative-sequence (21), (22) voltages at the system bus ( and ) after the sag can be obtained. Finally, the variation in IM terminal voltages is given by (23) and (24) (21) (22) (23) (24) It can be seen from (23) and (24) that the change in IM terminal voltage depends on the motor current, system voltage, and impedance. The zero-sequence voltage during the sag is not inuenced by the motor dynamic behavior and, hence, remains
Fig. 4. IM dynamic responses following the asymmetrical three-phase voltage sag (sag parameters given in Appendix B) obtained with the Matlab model.

constant. Finally, the phase components of IM terminal voltages during and after the sag are obtained from (25) (26) Responses of the aggregate IM to an asymmetrical voltage sag obtained using (1)(26) are shown in Fig. 4. The top diagram in this gure illustrates the original voltage sag applied, and the subsequent diagrams present speed, voltage, and current responses of the IM. B. Model Validation The developed IM model has been partially validated in [5] by comparing the IM speed response (following a three-phase symmetrical sag) with a previously published speed response obtained with the more detailed model of the IM. A good agreement between the two sets of results was found and discussed in detail in [5]. In order to validate the model further, in particular responses of IM to asymmetrical sags, the test system of Fig. 1 is modeled in PSCAD/EMTDC as shown in Fig. 5. The model consists of an aggregated IM of three identical squirrel-cage induction motors (with the parameters similar to those in Appendix A) connected to the PCC bus (as shown in Fig. 5). A voltage sag generator originally developed in [10] is also connected to a PCC bus via a system impedance dened in Appendix A. Classical transient simulations with full dynamic models of IMs are performed in PSCAD/EMTDC to observe the motor transient response to voltage sags. In order to initiate the PSCAD/EMTDC model smoothly, all motors were began with 0.07914-p.u. torque, and then when they reached full

MILANOVIC et al.: INFLUENCE OF INDUCTION MOTORS ON VOLTAGE SAG PROPAGATIONPART I

1067

Fig. 5. IM connected to a voltage sag generator through system impedance.

Fig. 7. Comparison of the IM speed change with respect to symmetrical and asymmetrical sags (adopted from [4]).

Fig. 8. Comparison of the IM speed change with respect to asymmetrical sags with a small and large amount of negative-sequence contents. Fig. 6. Comparison of Matlab (dashed lines) and PSCAD (solid lines) results.

speed, switched to a full load (0.7914 p.u.). Fig. 6 compares IM terminal voltages following an asymmetrical three-phase sag, obtained with Matlab (dashed lines) and PSCAD (solid lines) models. The top pair of curves represent Phase A voltages, the middle pair is Phase B voltages, while the bottom pair represents Phase C voltages. It is evident from Fig. 6 that the results of Matlab and PSCAD/EMTDC agree well, not only qualitatively but also quantitatively, (both in terms of sag magnitude and duration) with a maximum difference occurring during the sag voltage magnitudes at the beginning of the sag of less than 10%. (Note: The sag magnitude that matters is the minimum value of the rms voltage during the sag, the value that is usually recorded as the equivalent sag magnitude in commonly used in sag representation tables. The difference in this value between two sets of curves, as can be seen from Fig. 6, is far less than 10%.) The observed difference can be attributed not only to simplications made in the development of the Matlab model [based on (1)(26)] but also to the originally different IM model used in PSCAD/EMTDC. Namely, the IM model available in the PSCAD/EMTDC library is based on double-cage IM. A similar or better agreement between responses obtained with Matlab and PSCAD/EMTDC models was observed for other types of voltage sags. Therefore, it can be deduced that the simplied IM model developed in Section II can be used to adequately describe the IM behavior when subjected to any type of single stage (constant sag magnitude for the whole duration of the sag), rectangular voltage sag, and, consequently, the IM inuence on voltage sag shape.

The results of the analysis obtained for different case studies are presented in Sections III-C and D. All results are obtained with the model developed in Section II. C. Effects of Voltage Sags on IM Fig. 7 (reproduced from [5] for the completeness of discussion) illustrates that the voltage sags of different types and severity will result in different variations in motor speed. For this specic case, a symmetrical voltage sag caused by a single line-to-ground fault results in a slight drop in speed compared to that caused by a symmetrical three-phase voltage sag. According to the previous discussion, it can be concluded that this particular asymmetrical voltage sag contributes only to a small amount to the negative-sequence voltage component while its positive-sequence component is much higher than that developed by the symmetrical sag. In the same cases, however, unbalanced voltage sags might cause a signicant drop in speed as shown in Fig. 8. This is particularly the case when they are associated with large negative-sequence voltage components. Fig. 9 illustrates the inuence of a combined inertia of motors and loads on IM behavior during the sag. It can be seen that the IM with a low inertia load decelerates rapidly during the sag though it also restores the nominal speed soon after the voltage sag is over. In contrast, a steady slow speed drop and a slow speed recovery can be expected in the case of the motor with a high inertia load. As far as the inuence of motor parameters on the change in speed is concerned, Fig. 10 shows that different motor parameters will cause different changes in motor speed. The greater

1068

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 23, NO. 2, APRIL 2008

Fig. 9. Comparison of the induction motor speed changes with respect to low and high inertia loads.

Fig. 12. Effects of dynamic responses of 168 aggregated small IMs on sags with different magnitude and duration.

Fig. 10. Comparison of the IM speed changes with respect to low and high rotor resistances.

Fig. 13. Different changes in the shape of voltage sags due to different groups of IMs.

TABLE I GROUPS OF IMS USED IN THE STUDY

Fig. 11. Effects of the dynamic response of 168 aggregated small IMs on sags with different magnitude.

speed drop can be expected in case of motors with high rotor resistance. Speed recovery time, however, is not affected significantly. D. Effects of IM Dynamic Responses on Voltage Sags It can be seen from Figs. 11 and 12 that the shape of voltage sags is not rectangular when the effect of IMs is taken into account. Additionally, Fig. 13 shows that different groups of IMs differently inuence the shape of the resulting voltage sags. (The effect of 168 aggregated small IMs is particularly pronounced). Table I shows that the output power produced by groups 3 and 5 is almost the same though group 5 needs a longer recovery time (i.e., 1.38 s versus 0.48 s). Fig. 14, however, shows an almost identical inuence of individual (small and large) IMs on
Fig. 14. Effects of a single large or small induction motor on the change in the shape of the voltage sag.

voltage sag characteristics. The change in sag characteristics is highly dependent on the number and type of motors involved in the group. Fig. 15 illustrates that different inertia of the group results in a small change in sag shape even dough the difference in inertia is signicant. A motor driving a low inertia load will recover a bit faster than the one driving a high inertia load.

MILANOVIC et al.: INFLUENCE OF INDUCTION MOTORS ON VOLTAGE SAG PROPAGATIONPART I

1069

Fig. 15. Effects of inertia on the change in the shape of voltage sag.

Fig. 18. Original rectangular sags for nonrectangular sags of Fig. 16.

Fig. 19. Original rectangular sags for nonrectangular sags of Fig. 17. Fig. 16. Effects of dynamics responses of IMs on voltage sags with a different magnitude.

Fig. 20. Nonrectangular sag (solid line) originated by a rectangular 0.6.-p.u./0.3-s sag and its respective equivalent rectangular sag (dashed line).

Fig. 17. Effects of dynamics responses of IMs on voltage sags with different duration.

By applying this methodology, voltage sag performance at any given bus, with or without IMs, can be explored. For this purpose, voltage sags can be conveniently arranged in three different categories: severe, with a magnitude of less then 0.5 p.u.; medium with a magnitude in the range of 0.50.7 p.u.; and shallow with a magnitude greater than 0.7 p.u. Figs. 16 and 17 show that IM dynamics do not change the characteristics of every voltage sag. (Note: the original rectangular sags (sags at the bus without the IMs) corresponding to those shown in Figs. 16 and 17 are illustrated in Figs. 18 and 19, respectively). This is particularly true for very shallow voltage sags, with a p.u., or for those lasting less than 0.08 s. magnitude of (Note: The change in duration of sags with magnitudes

p.u. is not relevant anyway since their magnitude is above the sag threshold of 0.9 p.u.) However, when the sag magnitude is p.u., or when the duration is s, the IM dynamics alter sag characteristics as shown by the solid line in Fig. 20. As a result of this, the rectangular voltage sag becomes nonrectangular, and the methodology proposed in Section IV should be applied to convert it back into the equivalent rectangular sag (shown by the dashed line in Fig. 20) for benchmarking purposes. IV. NONRECTANGULAR TO RECTANGULAR SAG CONVERSION A major consequence of the IM dynamics on voltage sag characteristics is the change in sag shape from rectangular to nonrectangular as illustrated in the previous section. The nonrectangular part of the sag has basically a time-varying magnitude [as illustrated in Fig. 21(a)] and, as such, is not suitable for

1070

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 23, NO. 2, APRIL 2008

Fig. 22. Study procedure for sag conversion.

shaded area in Fig. 21(b). A new rectangular sag, represented by the dashed curve abcdghk, is equivalent to the original nonrectangular sag in terms of volt-seconds lost. (Note: The original integration step of 20 ms for determining the nonrectangular voltage sags by procedure described in Section II is reduced in this case to 1 ms to increase the accuracy of the conversion.) The original nonrectangular sag is shown in Fig. 21(a) by a solid line (only the part below 0.9-p.u. voltage threshold is used for conversion). The curve bceh is dened by a set of points and it encompasses the area (28) with a sag duration given by (29) The new rectangular sag will have the same duration as the original sag while its magnitude will be reduced and given by (30) Using (29) and (30), the original nonrectangular sag (bceh) is transformed into an equivalent rectangular sag with magniand duration , as shown in Fig. 21(b). tude The consequence of this conversion is that the resulting rectangular sags have a slightly higher (and constant) magnitude than the original nonrectangular sags and the same duration (only the part below 0.9-p.u. magnitude is considered in accordance with the denition of voltage sags). Since this conversion is applied to all three phases independently, the resulting rectangular sags may have different sag durations as illustrated and discussed in detail in the adjoining paper (Part II). A. Study Procedure The study procedure is divided into two parts (see Fig. 22): IM dynamic response analyzer and nonrectangular to rectangular sag converter. 1) Part 1: Induction Motor Dynamic Response Analyzer: Using the input data (IM equivalent circuit data, original rectangular sag parameters, number of motors connected at the PCC and system impedance) and the model developed in Section II, this part calculates the change in sag parameters at IM terminals using the following algorithm. Parameters of the equivalent aggregate IM connected at PCC are calculated based on parameters of individual IMs. The original (without the effect of IMs) pre, during, and after sag voltages are calculated using (1)(3).

Fig. 21. IMs connected to a voltage sag generator through system impedance. (a) Voltsseconds lost due to IM nonrectangular sag (shaded area). (b) Voltsseconds lost due to new IM rectangular sag (shaded area).

inclusion in existing magnitude-duration tables used for benchmarking purposes. Any conversion of nonrectangular to rectangular shape is associated with a certain level of simplication and the equivalent rectangular sag becomes an approximation of the actual disturbance. Even though it is an approximation, it still represents the actual voltage sags in low-voltage networks (in particular) more accurately then current practices. The existing approaches almost completely ignore the presence and, consequently, the effects of IMs on voltage sags in the network. One possible way for converting nonrectangular to rectangular voltage sags is to use the loss of voltage concept/index [11] and (27). In this way, the original nonrectangular and corresponding rectangular voltage sags have the same loss of voltage (or voltseconds) (27) Fig. 21 illustrates the methodology for voltage sag conversion based on the loss of voltage concept. Consider an IM exposed to rectangular sag represented by the dashed curve abcejm in Fig. 21(a). Due to the effect of IM dynamics, this sag is changed to a nonrectangular sag represented by the solid line abchm. Using (27) and considering only the area below 0.9-p.u. voltage threshold, the nonrectangular sag represented by the shaded area in Fig. 21(a) is converted to a rectangular sag represented by the

MILANOVIC et al.: INFLUENCE OF INDUCTION MOTORS ON VOLTAGE SAG PROPAGATIONPART I

1071

The aggregate IM speed during the sag is calculated every millisecond of the sag duration using (12) and aggregate motor parameters. Motor reacceleration time is computed using (15). Motor speed after the sag is calculated every millisecond during the reacceleration of the motor using (14). The motor currents during and after the sag are determined using (1821). Finally, the change in sag shape due to IM dynamics is determined using (22), (23), and (26)(29) 2) Part 2: Nonrectangular to Rectangular Sag Converter: This block uses the output from the part 1 (i.e., nonrectangular voltage sag parameters) and converts them to parameters of the equivalent rectangular voltage sag using the following algorithm. The voltseconds lost in each phase are calculated using (30). Only the part of the voltage sag curve that falls below the dened voltage sag threshold (90% of nominal voltage) is considered in this calculation. The sag duration in each phase is calculated using (34). The equivalent rectangular sag magnitude in each phase is calculated using (35). V. CONCLUSION This paper presented an analytical tool to describe the inuence of voltage sags on a group of IMs (aggregated model) and the inuence of IM dynamics on voltage sag characteristics. The results show that the presence of IM(s) at a bus may initially provide voltage support, however, during the voltage recovery, the sag duration may increase due to the reacceleration of IM(s). The change in voltage sag shape due to IM dynamics depends on various parameters, such as number of motors, motor electrical and mechanical parameters, number of sagged phases, and severity of voltage sag. To facilitate the assessment of the impact of IMs on voltage sag performance of a bus using conventional benchmarking tables, an analytical nonrectangular to rectangular sag conversion method is introduced. The method accounts for the effect of IMs on sag characteristics and enables quick and convenient representation of resulting voltage sags using conventional benchmarking approaches. Further details about the sag conversion method and practical application issues are discussed in detail in the adjoining paper (Part II). APPENDIX A LARGE INDUCTION MOTOR AND SYSTEM PARAMETERS: V/ph, kW, Eight poles rad/s, rad/s, , , , , , , Nm, , . APPENDIX B SMALL INDUCTION MOTOR PARAMETERS Four poles , , Nm , V/ph, , , kW, ,

APPENDIX C THREE-PHASE TEST ASYMMETRICAL SAG PARAMETERS Phase A presag the sag Phase B presag during the sag: ms. Phase C presag the sag: p.u., p.u., p.u., p.u., p.u., p.u., ACKNOWLEDGMENT The authors would like to thank the Microgrids and RISE partners for their contributions. REFERENCES
[1] W. E. Brumsickle, R. S. Schneider, G. A. Kuckjiff, D. M. Divan, and M. F. Mcgranaghan, Dynamic sag correctors: cost-effective industrial power line conditioning, IEEE Trans. Ind. Appl., vol. 37, no. 1, pp. 212217, Jan./Feb. 2001. [2] R. Gnativ and J. V. Milanovic, Voltage sag propagation in systems with embedded generation and induction motors, in Proc. IEEE Power Eng. Soc. Summer Meeting, Jul. 2001, vol. 1, pp. 474479. [3] M. H. J. Bollen, Understanding Power Quality Problems: Voltage Sags and Interruptions. New York, NY: IEEE, 1995, pp. 153248. [4] D. L. Brooks, E. W. Gunther, and A. Sundaram, Recommendations for Tabulating RMS Variation Disturbances With Specic Reference to Tility Power Contracts, Cigr 36.05/CIRED 2 CC02Voltage Quality Working Group. [Online]. Available: http://www.ccu2.org/count-d2.pdf. [5] M. T. Aung and J. V. Milanovic, Analytical assessment of the effects of voltage sags on induction motor dynamic responses, in Proc. IEEE St Petersburg PowerTech, St. Petersburg, Russia, Jun. 2730, 2005. [6] Matlab Help File. Natick, MA, Mathworks, 2004. [7] PSCAD 4.1.1 On-Line Help File. Winnipeg, MB, Canada, Manitoba HVDC Research Centre, 2004. [8] D. C. Franklin and A. Morelato, Improving dynamic aggregation of induction motor models, IEEE Trans. Power Syst., vol. 9, no. 4, pp. 19341741, Nov. 1994. [9] Y. J. Wang, An analytical study of steady-state performance of an induction motor connected to unbalanced three-phase voltage, in Proc. IEEE Power Eng. Soc. Winter Meeting, Jan. 2000, vol. 1, pp. 159164. [10] S. Z. Djokic, J. V. Milanovic, and K. A. Charalambous, Computer simulation of voltage sag generator, in Proc. 10th IEEE Int. Conf. Harmonics Quality Power, Rio de Janeiro, Brazil, Oct. 69, 2002. [11] IEEE Recommended Practice for Monitoring Electric Power Quality, IEEE Std. 1159-1995, 1995. Jovica V. Milanovic (M95SM98) received the Dipl.Ing. and M.Sc. degrees from the University of Belgrade, Belgrade, Yugoslavia, and the Ph.D. degree from the University of Newcastle, Newcastle, Australia. Currently, he is a Professor of Electrical Power Engineering and Deputy Head of School (Research) of Electrical and Electronic Engineering, The University of Manchester, Manchester, U.K.

; during ms. ; , ; during ms.

Myo T. Aung (M05) received the B.E. degree in electrical engineering from the Yangon Technological University, Rangoon, Burma, the M.E. degree in electric power system management from the Asian Institute of Technology, Bangkok, Thailand, and the Ph.D. degree from the University of Manchester (formerly UMIST), Manchester, U.K. Currently, he is a Lecturer with the Department of Electronic and Electrical Engineering at the University of Bath, Bath, U.K.

Sarat C. Vegunta received the B.E. degree in power electronics from Nagpur University, Nagpur, India, and the M.Sc. degree from the University of Manchester Institute of Science and Technology (UMIST), Manchester, U.K. He is currently a Ph.D. student in the School of Electrical and Electronic Engineering at The University of Manchester, Manchester.

You might also like