You are on page 1of 9

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO.

2, APRIL 2005

867

A Method of a Lightning Surge Analysis Recommended in Japan Using EMTP


A. Ametani, Fellow, IEEE, and T. Kawamura, Fellow, IEEE
AbstractEMTP has been widely used in Japan to analyze switching and lightning overvoltages to design power stations and substations from the viewpoint of insulation coordination. The Japanese standard of high voltage testing, JEC-0102-1994, and the insulation design and the coordination of an 1100-kV line of Tokyo Electric Power Company were mostly based on EMTP simulations. This paper presents a method of a lightning surge analysis using the EMTP recommended in the Japanese standard and related guidelines, and suggests further work to be done based on discussions of applicable limits and problems of the recommended models. Index TermsEMTP, insulation coordination, lightning surge, substation.

that should be done based on discussions of the recommended models. II. RECOMMENDED MODELING METHOD [5][10], [15] Fig. 1 illustrates a basic conguration of a model system recommended for a lightning surge simulation. A. Number of Towers Five towers from a substation gantry are to be considered, because traveling waves reected from the towers affect a lightning surge voltage in the substation. The rst reection, when , from the 6th tower (next lightning strikes the rst tower at to the last = 5th tower in Fig. 1) arrives at the substation entrance after two travel times T between the rst and the sixth towers. with light velocity For example, gives the time . Thus, a simulation result is accurate up to 10 , which is enough in general to observe the maximum overvoltage and the time to the peak. B. AC Sources and Matching Termination The left side of the last tower is represented by a multiphase matching impedance (resistance matrix), which is given as a characteristic impedance matrix including a mutual impedance of the transmission line at a dominant transient frequency dened by [16]:

I. INTRODUCTION LIGHTNING surge overvoltage is a dominant factor to determine the insulation level of a substation. It is very hard to observe the lightning overvoltage experimentally, and thus a numerical simulation has been adopted to investigate it [1][10]. The EMTP [11][13] has been widely used since the 1970s in conjunction with increasing usage of digital computers. In Japan, the EMTP was introduced in 1976, and has become a powerful tool to investigate a transient in a power system since 1980. In 1982, a Working Group (WG) of a Lightning Surge Analysis1 was established in the IEE of Japan to develop and propose a sophisticated but commonly recognized method of a lightning surge analysis using the EMTP [5][10], and was dissolved in 2000. Based on the method developed by the WG, switching, fault and lightning surge analyses had been carried out on an 1100-kV transmission system by the EMTP for its optimized and rationalized insulation design and coordination [14]. A successful completion of the 1100-kV project had led to rationalization of the Japanese standard of high voltage testing, JEC-193-1973 in that time, and a special committee for renewing the standard was established in 1991 in cooperation with the WG. A new standard, JEC-0102-1994, was completed and issued in 1994 [15]. This paper summarizes the method of a lightning surge analysis developed by the WG and applied to the 1100-kV system design and to the JEC-0102-1994, and suggests further work
Manuscript received March 17, 2003; revised January 31, 2004. Paper no. TPWRD-00106-2003. A. Ametani is with Doshisha University, Kyo-Tanabe, Kyoto, Japan (e-mail: aametani@mail.doshisha.ac.jp). T. Kawamura is with the Shibaura Institute of Technology, Minato-ku, Tokyo, Japan. Digital Object Identier 10.1109/TPWRD.2004.839183
1Chairman:

(1) where , : total line length. An ac voltage source is connected to the other side of the matching impedance as illustrated in Fig. 1 to take into account the effect of the ac steady state voltage on a lightning surge. C. Lightning Current and Impedance A lightning current is represented by a ramp waveform with the wavefront duration and the wavetail . Occasionally, a concave current of which the waveform is dened in the following equation is adopted in Japan to represent a real lightning current more accurately [5], [17]:

(2) where is the peak current , i.e., .

T. Kawamura from 1982 to 1994 and A. Ametani from 1994 to

2000.

0885-8977/$20.00 2005 IEEE

868

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 2, APRIL 2005

Fig. 1. Simplied illustration of the recommended model system for a lightning surge simulation.

TABLE I RECOMMENDED VALUES OF LIGHTNING PARAMETERS

The recommended value of the lightning current amplitude is given in Table I for various transmission voltages in Japan. The impedance of a lightning path is represented as a parallel resistance to a current source as illustrated in Fig. 1. The resistance value is taken to be 400 , which was derived by Bewley [18].

Fig. 2. A model circuit of a tower.

D. Tower and Gantry A transmission tower is represented by four distributed-parameter lines [19], as illustrated in Fig. 2, where tower top to the upper phase arm upper to middle middle to lower; lower to tower bottom. Table I gives typical values of the surge impedance. The propagation velocity of a traveling wave along a tower is taken to be (3)

To represent traveling-wave attenuation and distortion, an parallel circuit is added to each part, as illustrated in Fig. 2. The values of the and are dened in the following equation:

(4) where traveling time along the tower; attenuation along the tower; tower height.

AMETANI AND KAWAMURA: A METHOD OF A LIGHTNING SURGE ANALYSIS RECOMMENDED IN JAPAN USING EMTP

869

A substation gantry is represented by a single distributed line with no loss. E. Tower Footing Impedance A tower footing impedance is suggested in Japan to be mod, although a current-deeled as a simple linear resistance pendent nonlinear resistance is recommended by the IEEE and the CIGRE [1][4] and the inductive and capacitive characteristics of the footing impedance are well-known. A recommended value of the resistance for each voltage class is given in Table I. F. Archorn Flashover An archorn ashover is represented either by a piecewise linear inductance model with time controlled switches as illustrated in Fig. 3(a) or by a nonlinear inductance in Fig. 3(b) based on a leader progression model [20], [21]. The parameters ( to 3) and assuming the initial time in Fig. 3(a) are determined from a measured result of the V-I characteristic of an archorn ashover. Then the rst EMTP simulation with no archorn ashover is carried out in Fig. 1, and the rst ashover phase and the initial time are determined from the simulation results of the voltage waveforms across all the archorns. By adopting the above parameters, the second EMTP simulation only with the rst phase ashover is carried out to determine the second ashover phase. By repeating the above procedure until no ashover occurs, a lightning surge simulation by the piecewise linear model is completed. Thus, a number of pre-calculations are necessary in the case of multiphase ashovers, while the nonlinear inductance model needs no pre-calculation and is easily applied to multiphase ashovers. The detail of the leader progression model is explained in [20], and that of the nonlinear inductance model in [21]. G. Transmission Line Most transmission lines in Japan are of twin-circuit vertical conguration with two ground wires, and thus are composed of eight conductors. It is recommended to represent the line by a frequency-dependent line model of the EMTP. But, a distributed-line model with a xed propagation velocity, attenuation and surge impedance, i.e., xed-parameter distributed-line model explained in [12, Sec. 4.2.2.4], is often used. H. Corona Wave Deformation Japanese guideline neglects corona wave deformation, although it is taken into account in the CIGRE and the IEEE guidelines [1][4]. I. Arrester An inductance of a lead wire and the arrester itself is connected in series to the arrester model, because it affects a transient voltage and current of the arrester. Also, a capacitance of the arrester is considered if necessary. To represent a very fast impulse characteristic of the arrester, an IEEE model [22] or a model of a nonlinear resistance with a nonlinear inductance [23] is occasionally adopted. The latter model takes into account the hysterisis of an arrester characteristic by adding the nonlinear inductance to the nonlinear resistance in series.

Fig. 3. An anchorn ashover model. (a) A piecewise linear mode. (b) A nonlinear inductance model.

J. Substation 1) Gas insulated bus and cable: A cable and a gas-insulated bus are represented either as three single-phase distributed lines with its coaxial mode surge impedance and propagation velocity or as a three-phase distributed line system. For a gas-insulated substation involves quite a number of gas-insulated buses/lines, the pipes are, in most cases, eliminated by assuming the voltage being zero. 2) Circuit breaker (CB), disconnector, transformer, bushing: A CB and a disconnector are represented by lumped capacitances between the poles and to the earth. A transformer is also represented by a capacitance to the earth unless a transferred surge to the secondary circuit is needed to be calculated. A bushing is represented by a capacitance. Occasionally, it is represented by a distributed line. 3) Grounding mesh: A grounding mesh is in general not considered in a lightning surge simulation, and is regarded as a zero-potential surface. When dealing with an incoming surge to a low-voltage control circuit, the transient voltage of the grounding mesh should not be assumed zero, and its representation becomes an important but difcult subject. III. DISCUSSIONS OF APPLICABLE LIMITS AND PROBLEMS A. Lightning Current and Impedance 1) Lightning current waveform: Measured results of a lightning current show a concave shape [17], [24]. An archorn ashover characteristic being affected by the current derivative, it has been decided not to adopt a double-exponential waveform. Calculated results of a maximum voltage at a tower top show that the double-exponential waveform produces a far greater voltage than those in the smaller than other waveforms [25]. In the region of 0.7 , no signicant difference is observed. The CIGRE , 5%: by 30/90% waveform (50%: denition) [17] is found to produce a far smaller voltage. It should be noted that the maximum voltage appears at in the ramp and wave cases, the time ( : traveling time while it appears at the time along the tower) in the double-exponential wave case. 2) Lightning-path impedance: The lightning-path in Fig. 1 was derived by Beimpedance of 400 wley [18], but the value seems not correct, because the lightning velocity is assumed equal to the light velocity in free space. On the contrary, Diesendorf suggested the value as 1000 to 2000 [26]. Fig. 4 shows the effect of the lightning-path resistance on a lightning overvoltage

870

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 2, APRIL 2005

Fig. 4. Effect of a lightning-path impedance on a substation entrance overvoltage.

Fig. 6. Simulation results of archorn ashover phases corresponding to Fig. 5 : single-phase FO : two-phase FO. (a) A simple distributed line model. (b) Recommended tower model.

TABLE II MAXIMUM ARCHORN VOLTAGES AND THE TIME OF APPEARANCE Fig. 5. Measured results of archorn ashover phases on a 77-kV transmission line single-phase FO, two-phase FO, three-phase FO.

at the entrance of a 500-kV gas-insulated substation [9]. It is observed that the greater the resistance, the higher , the overvoltage is the overvoltage. With increased by about 7% in comparison with that with . A similar trend is observed at a transformer terminal. The reason for the increase is readily explained as an increase of a current owing into a tower and a ground wire by increasing the lightning-path resistance. When an arrester is installed, the difference is decreased. The impedance value of a real lightning path has not been made clear, and requires further investigation. B. Ac Source Voltage An ac source voltage is often neglected in a lightning surge simulation. It has however been found that the ac source voltage affects a ashover phase of an archorn especially in the case of a rather small lightning current. Fig. 5 is a measured result of archorn ashover phases as a function of the ac source voltage on a 77-kV transmission line in Japan for a summer [24]. The measurements were carried out in two 77-kV substations by surge recorders installed in the substations. From the recorded voltages and currents, Fig. 5 was obtained. The gure clearly shows that the archorn ashover phase is quite dependent on the ac source voltage, i.e., a ashover occurs at a phase of which the ac voltage is in the opposite polarity of a lightning current. Table II shows a simulation result of archorn peak voltages (archorn not operating) on the 77-kV line and on a 500-kV line [27]. The simulation was carried out in a similar circuit to Fig. 1, but another

ve towers were added instead of the gantry and the substation. The parameters are the same as those in Table I for a 77-kV system, except the lightning current of 40 kA based on the eld measurement [24]. The lower phase archorn voltage is relatively smaller than the other phase archorn voltages on the 500-kV line compared with those on the 77-kV line. Thus, an archorn ashover phase on an EHV line is rather independent from the ac source voltage, and the lower phase ashover is less probable than the other phase ashover. On the contrary, ashover probability is rather same on each phase and a ashover is dependent on the ac source voltage on a low-voltage line. C. Tower Model 1) Problem of Recommended Tower Model: Fig. 6 shows simulation results of archorn ashover phases by a simple distributed line tower model i.e., neglecting the RL circuit in Fig. 2 with the parameters in Table I, and by the recommended model illustrated in Fig. 2. The simulation circuit is the same as that described for Table II in the previous section. This gure should be compared with the eld test result shown in Fig. 5. It is clear that the recommended model can not duplicate the eld test result, while the simple distributed line model shows a good agreement with the eld test result. The reason for the poor accuracy of the recommended model is that the model was developed originally for a 500-kV line on which the lower phase

AMETANI AND KAWAMURA: A METHOD OF A LIGHTNING SURGE ANALYSIS RECOMMENDED IN JAPAN USING EMTP

871

TABLE III MEASURED AND CALCULATED SURGE IMPEDANCES OF VERTICAL CONDUCTORS

ashover was less probable as explained in Section III-B. Thus, the recommended tower model tends to result in lower ashover probability of the lower phase archorn. An parallel circuit between two distributed lines in Fig. 2 represents traveling-wave attenuation and distortion along a tower. The and values were determined originally based on a eld measurement [ in (4)], and thus those are correct only for the tower on which the measurement was carried out. Sometimes, the R-L circuit generates unreal high frequency oscillation. This indicated a necessity of further investigation of the circuit. 2) Impedance and Admittance Formulas: A number of tower models have been proposed, but most of them are not general, i.e., a tower model shows a good agreement with a specic case explained in the paper where the model is proposed. The following IEEE/CIGRE formula of the tower surge impedance is well known and is widely adopted in a lightning surge simulation [2], [4]: (5) where is the equivalent radius of , and the tower represented by a truncated cone, , , tower top, midsection and base radii [m]; height from midsection to top [m]; height from base to midsection [m]. When the tower is not a cone but a cylinder, then the above equation is rewritten by (6) where is the radius of a cylinder representing a tower The recommended value of a tower surge impedance for each voltage class in Table I was determined by eld measurements in Japan. Although the surge impedance is a representative value, it cannot be applied to every tower (Table III). Wave deformation on tower structures (L- or T-shaped iron conductor) can be included in a lightning surge analysis, if required, based on the approach in [33].

3) Frequency-Dependent Effect of a Tower: The frequencydependent effect of a tower is readily taken into account in a transient simulation by combining the frequency-dependent tower impedance [28] with Semlyens or Martis line model [34], [35] in the EMTP [36]. That is: 1) obtain frequency responses of the propagation constant and the characteristic impedance either from a measured result or from the frequency-dependent impedance and the admittance formulas in references; 2) obtain the parameters of Semlyens or Martis line model for a transient simulation from the above result. 4) Inuence of Surge Impedance and Frequency-Dependent Effect: It should be pointed out that the inuence of the surge impedance and the frequency-dependent effect of a tower is heavily dependent on the modeling of a tower footing impedance, which will be discussed in the following section. When the footing impedance is represented by a resistive model as recommended in Japan or by a capacitance model, then the inuence of the tower surge impedance and the frequency-dependent effect of a traveling wave along the tower becomes rather noticeable. On the contrary, those cause only a minor effect when the footing impedance is represented either by an inductive model or by a nonlinear resistance. Fig. 7 shows an example [25], [36]. The measurement was carried out on a 500-kV tower by applying a current in Fig. 7(a-1) to the top of the tower. The tower top voltage predicted by a distributed-line model with a constant tower surge impedance and no R-L circuit Fig. 7(c-1), differs from that by the frequency-dependent tower model Fig. 7(b), which agrees with the measured result, in the case of the footing impedance being a resistance. On the contrary, in the case of an inductive footing model, the tower top voltage obtained by the distributed-line model shows a rather good agreement with the measured result. It should be also noted that some 10% variation of the tower surge impedance does not affect the result in the inductive footing impedance case. Thus, it is concluded that the frequency-dependent effect of wave propagation along a tower can be neglected and the value of the surge impedance is not signicant unless a tower footing impedance is represented by a resistive or a capacitive model.

872

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 2, APRIL 2005

Fig. 7. Inuence of a tower on the tower top voltage. (a) Measured result. (b) Frequency-dependent tower model with a resistive footing impedance. (c) Distributed-line tower model with various footing impedances.

D. Tower Footing Impedance 1) Linear Footing Impedance: It has been known in general that the footing impedance tends to be capacitive in the case of a high-resistivity earth, and inductive in the low resistivity earth case. A problem of the representation is: the footing impedance can be resistive, inductive and capacitive depending on the season and the weather when a measurement is made, i.e., the impedance is temperature- and soil moisture-dependent. Therefore, it is not easy to select a model of the footing impedance and this is the reason why a resistance model is adopted in Japan. 2) Current-Dependent Nonlinearity: A number of papers have discussed the current-dependence of a tower footing impedance, and have proposed various models of the nonlinear footing impedance. It has been a common understanding that the current-dependence decreases a lightning surge voltage at the tower and thus decreases a lightning overvoltage at a substation. Therefore a simulation neglecting the current dependence gives a severer overvoltage, i.e., a safer side result from the insulation design viewpoint. For this reason, again a pure resistance model is recommended in Japan. E. Archorn Flashover Model There exist a number of archorn ashover models. To investigate the accuracy, phase-wire voltages at the rst tower in Fig. 1 are calculated by various archorn models and are compared [37]. A switch (time-controlled) model and a ashover switch model show not satisfactory agreement especially in the wavefront with the nonlinear inductance model of which the accuracy has been conrmed to be high in comparison with an ex-

perimental result [21]. A - characteristic model and a piecewise linear inductance model show a reasonable accuracy except that the maximum voltage of the former is greater and that of the latter is lower than that calculated by the nonlinear inductance model. It might be noteworthy that a current and energy consumed by an arrester in a substation are dependent on an archorn model. The switch and the ashover switch models result in much higher energy consumed by an arrester. F. Transmission Line, Feeder, Gas-Insulated Bus 1) Frequency-Dependent Transmission Line Impedance: Although a frequency-dependent line model, i.e., Semlyens or Martis model [34], [35], is recommended, the maximum error of Martis model is observed to be about 15% at the wavefront of an impulse voltage on an 1100-kV untransposed vertical twin-circuit line in comparison with a eld test result [38]. The estimation of possible errors incurred by using these models in a lightning overvoltage simulation is not straightforward, because it involves a nonlinearity due to an archorn ashover dependent on a lightning current, an ac source voltage, a ashover phase and so on. This is an important subject to be investigated in future. 2) Finite Length of a Line and a Gas-Insulated Bus: Carsons and Pollaczeks earth return impedance of an overhead line and an underground cable were derived based on the assumption of an innitely long line. A real line is not innitely long at all. The separation distance of a UHV/EHV transmission line between adjacent towers and the length of a gas-insulated bus are in the same order of their height . If the condition that is far greater than and is far greater than the radius , is not satised, Carsons and Pollaczeks impedances

AMETANI AND KAWAMURA: A METHOD OF A LIGHTNING SURGE ANALYSIS RECOMMENDED IN JAPAN USING EMTP

873

are not applicable. Fig. 8 shows an example of the impedance of a nite-length line evaluated by the following equation [39]:

(7) where ; ; horizontal separation between conductors and ; , height of conductors i and j; complex penetration depth [40]. It should be clear in Fig. 8 that Carsons impedance assuming an innitely long line is far greater than that of a real nite line, is not greater enough than 1. The reason for this is when readily understood from the following equation [41]: (8)

(9) As is clear from the above equations, involves mutual incoupling from the innitely long conductor , while volves mutual coupling from the conductor with the nite length . In fact, becomes innite because of the innite length, and thus per unit length impedance is necessarily dened. It should be noted that the per-unit length impedance of (8) has included the mutual coupling from the innitely long conductor . On the contrary, , if we dene the per unit length impedance in (9), includes the mutual cou. Thus pling only for the nite length (10) On the contrary, the per-unit length admittance of an innitely long line is smaller than that of a nite-length line. From the above discussion, it should now be clear that Carsons and Pollaczeks earth return impedances may not be applied to a lightning surge analysis, because the separation distance between adjacent towers is the same order as the line height. The same is true for a gas-insulated bus, because its length, height and radius are in the same order. This requires further work which is interesting and signicant. It is noteworthy that the propagation constants of a nite line is nearly the same as that of an innitely long line, but the characteristic (surge) impedance is smaller, because of a smaller series impedance and a greater shunt admittance of the nite line. Furthermore, the ratio of the surge impedances of two nite lines is nearly the same as that of two innitely long lines. Finally, traveling-wave reection, refraction and deformation on the nite line is not much different from those on the innitely long line.
Fig. 8. Finite line impedance in comparison with Carsons impedance. n = nite line, (7). (a) Resistance. (b) Inductance.

3) Feeding Line From a Transmission Line to a Substation Inclined Conductor: A feeding line from the rst tower to the substation via the gantry in Fig. 1 is inclined, i.e., the height is gradually decreased. As a result, its surge impedance is also decreased gradually and thus no signicant reection of traveling waves occurs along the feeding line until the substation entrance in physical reality. Because the surge impedance (about 70 ) of a gas-insulated bus or a bushing is much smaller than that of an overhead line (300 to 500 ), noticeable reection appears at the substation entrance, if the inclined conguration of the overhead feeding line is not considered. It is better to consider the inclined conguration of a feeding line if an accurate simulation is required. A maximum difference of 7% in a substation entrance voltage is observed when the inclined conguration is considered [42]. G. Corona Wave Deformation The reason for corona wave deformation being not considered in a lightning surge simulation in Japanese guideline is that a simulation result neglecting the corona is expected to be higher than that considering corona and thus the result is on a safer side from the insulation viewpoint. The possible errors incurred by ignoring corona are observed to be less than 10% when lightning strikes the rst tower in Fig. 1 [8]. Although sophisticated corona models have been proposed [43], [44], the reliability and stability in a lightning overvoltage simulation is not conrmed. It is noteworthy that the corona wave deformation can result in a higher overvoltage at a substation entrance under a specic condition. Fig. 9 shows a eld test result of a normalized voltage

874

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 2, APRIL 2005

ratio for the negative polarity case dened in the following equation on a 6.6-kV line [45], [46]: (11) where normalized by applied voltage ; maximum phase-wire (PW) voltage at the receiving end (substation entrance); normalized voltage with no corona discharge. The experiment was carried out on a 6.6-kV line with one phase wire and one ground wire which were terminated by reand at the remote end. An impulse voltage up sistances to 800 kV was applied to the sending end of the ground wire. The back ashover in Fig. 9 was represented by a short circuit of the ground and phase wires. For corona wave deformation decreases a traveling-wave voltage on a line, it is a common understanding that the line voltage is decreased by the corona wave deformation and thus the ratio is less than 1. In the case of no corona discharge, is nearly equal to 1 on a short distance line. on a single-phase It was observed that a measured result of line was less than 1. in the case of no back-ashover beFig. 9(a) shows that comes greater than 1 as the applied voltage is increased, i.e., corona discharge occurs. On the contrary in Fig. 9(b), is less than 1. The reason for the phenomena is readily explained as a result of different attenuation on a phase wire and a ground wire due to corona discharge, and negative reection of a heavily attenuated traveling wave on the ground wire. The phenomena are less noticeable in the positive polarity case. Details are explained in [45] and [46]. The phenomena have been also realized qualitatively by an EMTP simulation. The increase of a phase-wire voltage at a substation entrance is expected to be more pronounced on an EHV/UHV transmission line on which a corona discharge hardly occurs on a phase wire because of a multiple bundled conductor, while a heavy corona discharge is expected on a ground wire. H. Phase-to-Phase Lightning Surge Most of the previous studies on a lightning overvoltage concerned an overvoltage to the earth. A phase-to-phase overvoltage, however, can damage insulation between phases such as core-to-core insulation in a gas-insulated bus in which three-phase cores are enclosed in the pipe. Fig. 10 shows an EMTP simulation result of an inter-phase lightning overvoltage at a substation entrance on a 77-kV vertical twin-circuit line, when phases a and b ashovers. The simulation was carried out on a system composed of a 10-km 245/77-kV quadruple circuit line and a 10-km 77-kV line. The 77-kV line was connected to a substation through a three-phase underground XLPE cable with the length of 500 m. Because of a lower attenuation of aerial propagation modes, the phase-to-phase overvoltage becomes greater than the phase-to-earth overvoltage especially in the case of a lightning strike to a tower far from a substation. The phase-to-phase lightning overvoltage needs further investigation.

Fig. 9. Measured result of normalized voltage ratio Effect of corona wave deformation. (a) Without back-ashover. (b) With back-ashover.

Fig. 10. Phase-to-phase lightning surge on a 77-kV line.

IV. CONCLUSIONS This paper has explained a lightning surge simulation using the EMTP recommended in the Japanese guideline, and discussed problems of the recommended models. Summarizing the discussions, the following remarks, which needs further investigation, have been made clear. 1) Tower modeling should be investigated in connection with modeling of a tower footing impedance. When the footing impedance is inductive, a sophisticated tower model such as a frequency-dependent model is not necessary. The tower footing impedance can be resistive and reactive depending on the earth resistivity and permittivity, i.e., the soil temperature and moisture which show seasonal variation and weather dependence. 2) Carsons and Pollaczecks earth-return impedances assuming an innitely long line can not be applied to an

AMETANI AND KAWAMURA: A METHOD OF A LIGHTNING SURGE ANALYSIS RECOMMENDED IN JAPAN USING EMTP

875

overhead line between towers and a gas-insulated bus, because the line/bus length is in the same order of the height, and the bus height is also in the same order of its radius. Thus, an earth-return impedance of a nite-length conductor has to be taken into account. 3) Similarly, an inclined conguration of an overhead line from the rst tower to a substation entrance is to be considered, if an accurate simulation is required. 4) Corona wave deformation does not always reduce a lightning overvoltage. 5) A phase-to-phase overvoltage can exceed a phase-to-earth overvoltage especially in the case of lightning far from a substation and multiphase ashovers. ACKNOWLEDGMENT The authors are grateful to all the WG members. REFERENCES
[1] A simplied method for estimating lightning performance of transmission lines, IEEE Trans. Power App. Syst., vol. PAS-104, p. 919, 1985. [2] Estimating lightning performance of transmission lines, IIupdate to analytical models, IEEE Trans. Power Delivery, vol. PWRD-8, p. 1254, 1993. [3] IEEE Guide for Improving the Lightning Performance of Transmission Lines, IEEE Standard 1243-1997, 1997. [4] Guide to Procedures for Estimating Lightning Performance of Transmission Lines, CIGRE SC33-WG01, Tech. Brochure, Oct. 1991. [5] A New Method of a Lightning Surge Analysis in a Power System, IEE Japan WG Report, Tech. Rep. 244, Mar. 1987. [6] Various Parameters of a Lightning Surge Analysis in Power Stations and Substations and the Inuences Thereof, IEE Japan WG Report, Tech. Rep. 301, Jun. 1989. [7] A New Method for Estimating Lightning Surge in Substations, IEE Japan WG Report, Tech. Rep. 446, Nov. 1992. [8] An Estimating Method of a Lightning Surge for Statistical Insulation Design of Substations, IEE Japan WG Report, Tech. Rep. 566, Oct. 1995. [9] Modeling for an Accurate Estimation of a Lightning Surge, IEE Japan WG Report, Tech. Rep. 704, Nov. 1998. [10] Power System Transients and EMTP Analyzes, IEE Japan WG Report, Tech. Rep. 872, Mar. 2002. [11] W. S. Meyer, EMTP Rule Book, 1st ed. Portland, OR: B.P.A., 1973. [12] H. W. Dommel, EMTP Theory Book: B.P.A., Aug. 1986. [13] Japanese EMTP Committee, EMTP Test Cases Book, Jun. 1987. [14] N. Yamada et al., UHV AC transmission, J. IEE Japan, vol. 102, pp. 9691051, 1982. [15] High-Voltage Testing Method, IEE Japan, JEC-0102-1994, 1994. [16] A. Ametani, Distributed-Parameter Circuit Theory. Tokyo, Japan: Corona Publishing Co., Feb. 1990. [17] R. B. Anderson and A. J. Eriksson, Lightning parameters for engineering application, Electra, no. 69, p. 65, 1980. [18] L. V. Bewley, Traveling Waves on Transmission Systems. New York: Dover, 1963. [19] M. Ishii et al., Multistory transmission tower model for lightning surge analysis, IEEE Trans. Power Delivery, vol. PWRD-6, no. 3, p. 1372, Jul. 1991. [20] T. Shindo and T. Suzuki, A new calculation method of breakdown voltage-time characteristics of long air gaps, IEEE Trans. Power App. Syst., vol. PAS-104, p. 1556, 1985. [21] N. Nagaoka, An archorn ashover model by means of a nonlinear inductance, Trans. IEE Japan, vol. B-111, no. 5, p. 529, 1991. [22] Modeling of metal oxide surge arresters, IEEE Trans. Power Delivery, vol. PWRD-7, no. 1, p. 302, Jan. 1992. [23] I. Kim et al., Study of ZnO arrester model for steep front wave, IEEE Trans. Power Delivery, vol. PWRD-11, no. 2, p. 834, Apr. 1996. [24] T. Ueda, M. Yoda, and I. Miyachi, Characteristics of lightning surges observed at 77 kV substations, Trans. IEE Japan, vol. 116-B, no. 11, p. 1422, 1996.

[25] H. Motoyama, Y. Okumura, and A. Ametani, The effects of a tower footing resistance and a lightning current waveform on a lightning surge, in IEE Japan, Research Meeting, Jul. 1987, Paper PE-87-19. [26] W. Diesendorf, Insulation Co-Ordination in High Voltage Electric Power Systems. London, U.K.: Butterworths, 1974. [27] A. Ametani et al., Investigation of ashover phases in a lightning surge by new archorn and tower models, in Proc. IEEE PES T&D Conf. 2002, Yokohama, Japan, 2002, pp. 12411246. [28] , A frequency-dependent impedance of vertical conductors on a multiconductor tower model, IEE Proc.-GTD, vol. 141, no. 4, pp. 339345, 1994. [29] C. A. Jordan, Lightning computation for transmission line with overhead ground wires, G. E. Rev., vol. 37, no. 4, pp. 180186, 1931. [30] C. F. Wagner, A new approach to calculation of lightning performance of transmission lines, AIEE Trans., vol. 75, p. 1233, 1956. [31] M. A. Sargent and M. Darveniza, Tower surge impedance, IEEE Trans. Power App. Syst., vol. PAS-88, no. 5, p. 680, 1969. [32] T. Hara et al., Empirical formulas of surge impedance for single and multiple vertical cylinder, Trans. IEE Japan, vol. 110-B, p. 129, 1990. [33] A. Ametani et al., Wave propagation characteristics of iron conductors in an intelligent building, Trans. IEE Japan, vol. 120-B, no. 1, p. 31, 2000. [34] A. Semlyen and A. Dabuleanu, Fast and accurate switching transient calculations on transmission lines with ground return using recursive convolutions, IEEE Trans. Power App. Syst., vol. PAS-95, no. 5, p. 561, 1975. [35] J. R. Marti, Accurate modeling of frequency-dependent transmission lines in electromagnetic transient simulations, IEEE Trans. Power App. Syst., vol. PAS-101, no. 1, p. 147, 1982. [36] N. Nagaoka, Development of frequency-dependent tower model, Trans. IEE Japan, vol. 111-B, p. 51, 1991. [37] N. Nagaoka and A. Ametani, A lightning surge analysis considering multiphase ashovers, in IEE Japan, Research Meeting, 1995, Paper HV-95-50. [38] A. Ametani, K. Adachi, and T. Narita, An investigation of surge propagation characteristics on an 1100 kV transmission line, IEEJ Trans. PE, vol. 123, no. 4, p. 513, 2003. [39] A. Ametani and A. Ishihara, Investigation of impedance and line parameters of a nite-length multiconductor system, Trans. IEE Japan, vol. 113-B, no. 8, p. 905, 1993. [40] A. Deri et al., The complex ground return plane: a simplied model for homogeneous and multi-layer earth return, IEEE Trans. Power App. Syst., vol. PAS-100, no. 8, p. 3686, 1981. [41] A. Ametani, Wave propagation on a nonuniform line and its impedance and admittance, Sci. Eng. Review, Doshisha Univ., vol. 43, no. 3, p. 136, 2002. [42] A. Ametani and A. Ishihara, Impedance of a nonparallel conductor system and its circuit analysis, in IEE Japan, Research Meeting, 1992, Paper PE-92-173. [43] C. de Jusus and M. T. Correia de Barros, Modeling of corona dynamics for surge propagation studies, IEEE Trans. Power Delivery, vol. PWRD-9, no. 3, p. 1564, Jul. 1994. [44] J. F. Guiller, M. Poloujadoff, and M. Rioul, Damping model of traveling waves by corona effect along extra high voltage three phase line, IEEE Trans. Power Delivery, vol. PWRD-10, no. 4, p. 1851, Oct. 1995. [45] A. Ametani et al., A study of phase-wire voltage variations due to corona wave-deformation, in Proc. IPST 99, Budapest, Hungary, June 1999, pp. 433438. , A basic investigation of substation entrance voltage variation due [46] to corona wave deformation, Trans. IEE Japan, vol. 120-B, no. 3, p. 403, 2000.

A. Ametani, (M71SM84F92) received the Ph.D. degree from the University of Manchester Institute of Science and Technology, Manchester, U.K., in 1973. Currently, he is a Professor at Doshisha University, Kyoto, Japan.

T. Kawamura, (M59SM89F91) received the D.Eng. degree from Tokyo University, Tokyo, Japan, in 1959. Currently, he is a Professor with the Shibaura Institute of Technology, Tokyo, Japan, and is Professor Emeritus at Tokyo University.

You might also like