2.7 Grounding

Proper power system grounding is very essential for personnel safety, equipment protection, and power continuity to electrical loads. Special attention should be given to grounding location and means for minimizing circulating ground currents, as well as ground-fault sensing. Electrical power systems require two types of grounding, namely, system grounding and equipment grounding; each has its own functions [55]. The IEEE Green Book, IEEE Recommended Practice for Grounding of Industrial and Commercial Power Systems, ANSI/IEEE C114.1, Std.142, is the authoritative reference on grounding techniques [56].

2.7.1 System Grounding

System and circuit conductor grounding is an intentional connection between the electric system conductors and grounding electrodes that provides an effective connection to ground (earth). The basic objectives being sought are to

 limit overvoltages due to lightning,

 limit line surges, or unintentional contact with high-voltage lines,

 stabilize the voltage to ground during normal operation, and

 facilitate overcurrent device operation in case of ground faults.

2.7.1.1 Factors Influencing Choice of Grounded or Ungrounded System

Whether or not to ground an electrical system is a question that must be faced sometime by most engineers charged with planning electrical distribution. A decision in favor of a grounded system leads then to the question of how to ground.

Definitions

 Solidly grounded: no intentional grounding impedance.

 Effectively grounded: grounded through a grounding connection of sufficiently low impedance (inherent or intentionally added or both) such that ground faults that may occur cannot build up voltages in excess of limits established for apparatus, circuits, or systems so grounded. R0 < X1, X0 ≤ 3X1, where X1 is the leakage reactance, and X0 and R0 are the grounding reactance and resistance, respectively.

 Reactance grounded: X0 ≤ 10X1.

 Resistance grounded: R0 ≥ 2X0.

 High-resistance grounded: ground fault up to 10 A is permitted.

 Grounded for serving line-to-neutral loads: Z0 ≤ Z1.

 Ungrounded: no intentional system grounding connection.

Service Continuity

For many years a great number of industrial plant distribution systems have been operated ungrounded (Fig. 2.45) at various voltage levels. In most cases this has been done with the thought of gaining an additional degree of service continuity. The fact that any one contact occurring between one phase of the system and ground is unlikely to cause an immediate outage to any load may represent an advantage in many plants (e.g., submarines), varying in its importance according to the type of plant.

f02-45-9780128007822
Figure 2.45 Ungrounded system with one fault.
Grounded Systems

In most cases systems are designed so that circuit protective devices will remove a faulty circuit from the system regardless of the type of fault. A phase-to-ground fault generally results in the immediate isolation of the faulted circuit with the attendant outage of the loads on that circuit (Fig. 2.46). However, experience has shown in a number of systems that greater service continuity may be obtained with grounded-neutral than with ungrounded-neutral systems.

f02-46-9780128007822
Figure 2.46 Grounded system with one fault.
Multiple Faults to Ground

Although a ground fault on one phase of an ungrounded system generally does not cause a service interruption, the occurrence of a second ground fault on a different phase before the first fault is cleared does result in an outage (Fig. 2.47). If both faults are on the same feeder, that feeder will be opened. If the second fault is on a different feeder, both feeders may be deenergized.

f02-47-9780128007822
Figure 2.47 Ungrounded system with two faults.

The longer a ground fault is allowed to remain on an ungrounded system, the greater is the likelihood of a second ground fault occurring on another phase resulting in an outage. The advantage of an ungrounded system in not immediately dropping load upon the occurrence of a ground fault may be largely destroyed by the practice of ignoring a ground fault until a second one occurs and repairs are required to restore service. With an ungrounded system it is extremely important that an organized maintenance program be provided so that first ground fault is located and removed as soon as possible after its occurrence.

Safety

Many of the hazards to personnel and property existing in some industrial electrical systems are the result of poor or nonexisting grounding of electrical equipment and metallic structures. Although the subject of equipment grounding is treated in the next section, it is important to note that regardless of whether or not the system is grounded, safety considerations require thorough grounding of equipment and structures.

Proper grounding of a low-voltage (600 V or less) distribution system may result in less likelihood of accidents to personnel than leaving the system supposedly ungrounded. The knowledge that a circuit is grounded generally will result in greater care on the part of the workman.

It is erroneous to believe that in an ungrounded system a person may contact an energized phase conductor without personal hazard. As Fig. 2.48 shows, an ungrounded system with balanced phase-to-ground capacitance has rated line-to-neutral voltages existing between any phase conductor and ground. To accidentally or intentionally contact such a conductor may present a serious, perhaps lethal shock hazard in most instances.

f02-48-9780128007822
Figure 2.48 (a) Ungrounded balanced system (circuits and phasor diagrams) indirectly grounded by the neutral of capacitor bank (neutral is on ground potential). (b, c) Diagrams showing circuit and voltages to ground under steady-state conditions.

During the time a ground fault remains on one phase of an ungrounded system, personnel contacting one of the other phases and ground are subjected to a voltage 1.73 (that is, 3si144_e) times that which would be experienced on a solidly neutral-grounded system. The voltage pattern would be the same as the case shown in Fig. 2.48c.

Power System Overvoltages

Some of the more common sources of overvoltages on a power system are the following:

 lightning (see Application Examples 2.9 and 2.10),

 switching surges,

 capacitor switching,

 static,

 contact with high-voltage system,

 line-to-ground faults,

 resonant conditions, and

 restriking ground faults.

Lightning

Most industrial systems are effectively shielded against direct lightning strikes (Fig. 2.49). Many circuits are either underground in ducts or within grounded metal conduits or raceways. Even open-wire circuits are often shielded by adjacent metallic structures and buildings. Large arresters applied at the incoming service limit the surge voltages within the plant resulting from lightning strikes to the exposed service lines. Other arrester applications may be necessary within the plant to protect low impulse strength devices such as rotating machines and computers.

f02-49-9780128007822
Figure 2.49 Conductor configuration on transmission tower.
Switching Surges

Normal switching operation in the system (e.g., capacitor switching) can cause overvoltages. These are generally not more than three times the normal voltage and of short duration.

2.7.1.2 Application Example 2.9: Propagation of a Surge through a Distribution Feeder with an Insulator Flashover

This example illustrates the performance of a transmission system when the current of a lightning strike is large enough to cause flashover of tower insulator strings [57]. A lightning strike with the rise time τ1 = 1.0 μsec, the fall time τ2 = 5.0 μs, and the peak magnitude of current waveform I0 = 15.4 kA strikes a 138 kV transmission line at node 1 (see Fig. E2.9.1). The span lengths of this transmission line are nonuniform and are given by NL (normalized length). Both ends of the transmission line are terminated by Z0 = 300 Ω.

f02-77-9780128007822
Figure E2.9.1 PSpice model of a transmission line including insulators.

The transmission line has steel towers with 10 porcelain insulators. The tower-surge impedance consists of a 20 Ω resistance in series with 20 μH of inductance. The model used to simulate the first two insulators is shown in Fig. E2.9.1. Note that the transmission line can be simulated by the PSpice (lossless) transmission line model T, where Z0 = 300 Ω, f = 60 Hz, and NL as indicated in Fig. E2.9.1. The current-controlled switch can be simulated by W, where ION = 1 A.

a) Calculate the four span lengths in meters.

b) Calculate (using PSpice) and plot the transmission line voltage (in MV) as a function of time (in μs) at the point of strike, at the first insulator, and at the second insulator.

Solution to Application Example 2.9

a) The normalized length of a distance traveled by a wave (e.g., span) NL is a function of the frequency f, the span length l, and the speed of light c = 2.99792 · 108 m/s [58]: NL=lcfsi145_e or the length of a span is l=cNLfsi146_e. The time to travel the distance NL · λ, where λ is the wavelength, is TD=NLf=lc=NLλsi147_e.
Length of span #1: l1 = 2500 m, length of span #2: l2 = 500 m, length of span #3: l3 = 1000 m, length of span #4: l4 = 1000 m.

b) The PSpice program is listed in Table E2.9.1. One obtains the results of Figs. E2.9.2 and E2.9.3.

Table E2.9.1

PSpice Program Codes for Propagation of a Surge through a Distribution Feeder with an Insulator Flashover (Application Example 2.9)

*Lightning performance of a transmission line
*Insulator flashover
*lightning current source
i1 1 0 exp(0 15.4k 0 1.0u 2.0111u 5u) 
*span#1
t1 1 0 3 0 z0=300 f=60 NL=100u
*insulator #1
d1 13 3 dmod 
L1 13 53 20u 
vs1 23 53 dc 0 
w1 23 33 vs1 wmod1
 rvs1 33 43 20 
vd1 0 43 3700k 
rft1 23 0 20
*span#2
t2 3 0 5 0 z0=300 f=60 NL=200u
*insulator#2
d2 15 5 dmod 
L2 15 55 20u 
vs2 25 55 dc 0 
w2 25 35 vs2 wmod1 
rvs2 35 45 20 
vd2 0 45 3700k
rft2 25 0 20
*span#3
t3 5 0 7 0 z0=300 f=60 NL=200u
*z
RL3 7 0 300
*span#4
t4 1 0 9 0 z0=300 f=60 NL=500u
*z
RL4 9 0 300 
.model dmod d(is=1e-14) 
.model wmod1 iswitch (ron=1 roff=1e6 ion=1  ioff=1.01) 
.tran 0.1u 20u 0 0.05u 
.options chgtol=2u abstol=2u vntol=2u 
.options ITL4=200 ITL5=0 LIMPTS=100000  reltol=0.001 
.probe 
.print tran V(1) V(3) V(5) 
.end

t0040

f02-78-9780128007822
Figure E2.9.2 Transmission line voltage at the point of strike v(1).
f02-79-9780128007822
Figure E2.9.3 Transmission line voltages at the first insulator v(3), and second insulator v(5).

2.7.1.3 Application Example 2.10: Lightning Arrester Operation

This problem illustrates the performance of a lightning arrester [57]. The same basic parameters as for the lightning strike employed in Application Example 2.9 are used in this example. The only changes are in the value of I0 (in this problem a value of 7.44 kA is assumed) and in the voltage of the second lightning arrester (VD2 = 600 kV). This value is small enough that flashover of the insulator string will not occur; however, it is large enough to cause the lightning arrester to operate due to VD2 = 600 kV. The system being simulated is shown in Fig. E2.10.1.

f02-80-9780128007822
Figure E2.10.1 PSpice model of a transmission line including insulator and lightning arrester.

a) Calculate and plot the transmission line voltage (in MV) as a function of time (in μs) at the point of strike, at the insulator, and at the lightning arrester.

b) Calculate and plot the currents at the point of strike, at the insulator, and at the lightning arrester (in kA) as a function of time (in μs).

Solution to Application Example 2.10

a) Calculations are performed using PSpice, and the plots of the transmission line voltage (in MV) as a function of time (in μs) at the point of strike, at the insulator, and at the lightning arrester are shown in Figs. E2.10.2 and E2.10.3. The PSpice program is listed in Table E2.10.1.

f02-81-9780128007822
Figure E2.10.2 Transmission line voltages at the point of strike v(1), at the first insulator v(3), and at the lightning arrester v(5).
f02-82-9780128007822
Figure E2.10.3 Transmission line voltage at the lighning arrester (magnified) v(5).

Table E2.10.1

PSpice Program Codes for Lightning Arrester Operation (Application Example 2.10)

*Lightning performance of a transmission line
*lightning arrester operation
*lightning current source
i1 1 0 exp(0 7.44k 0 1.0u 2.011u 5u) 
*span#1
t1 1 0 3 0 z0=300 f=60 NL=100u
*insulator #1
d1 13 3 dmod 
L1 13 53 20u 
vs1 23 53 dc 0 
w1 23 33 vs1 wmod1
 rvs1 33 43 20 
vd1 0 43 3700k 
rft1 23 0 20
*span#2
t2 3 0 5 0 z0=300 f=60 NL=200u
*insulator#2
d2 15 5 dmod 
L2 15 55 20u 
vs2 25 55 dc 0 
w2 25 35 vs2 wmod1 
rvs2 35 45 20 
vd2 0 45 600k
rft2 25 0 20
*span#3
t3 5 0 7 0 z0=300 f=60 NL=200u
*z
RL3 7 0 300
*span#4
t4 1 0 9 0 z0=300 f=60 NL=500u
*z
RL4 9 0 300 
.model dmod d(is=1) 
.model wmod1 iswitch (ron=1 roff=1e6 ion=0.8 
+ ioff=0.3) 
.tran 0.1u 30u 0 0.05u 
.options chgtol=2u abstol=2u vntol=2u 
.options ITL4=25 ITL5=0 LIMPTS=100000  reltol=0.001 
.probe 
.print tran V(1) V(3) V(5) -i(i1(t)) i(vs1) i(vs2) 
.end

b) One obtains the results of Figs. E2.10.4 and E2.10.5 using the PSpice program codes of Table E2.10.1. This is a good example where the various versions of PSpice have different convergence properties similar to the Mathematica software (see Application Example 8.5).

f02-83-9780128007822
Figure E2.10.4 Transmission line current at the point of strike -i(i1 (t)).
f02-84-9780128007822
Figure E2.10.5 Transmission line currents at the insulator i(VS1) = 0 and the lightning arrester i(VS2).

2.7.2 Equipment Grounding

All exposed metal parts of electrical equipment (including generator frames, mounting bases, electrical instruments, and enclosures) should be bonded to a grounding electrode.

Basic Objectives

Equipment grounding, in contrast with system grounding, relates to the manner in which nonelectrical conductive material, which encloses energized conductors or is adjacent thereto, is to be interconnected and grounded. The basic objectives being sought are the following:

 to maintain a low potential difference between nearby metal members, and thereby protect people from dangerous electric-shock-voltage exposure in a certain area,

 to provide current-carrying capability, both in magnitude and duration, adequate to accept the ground-fault current permitted by the overcurrent protection system without creating a fire or explosive hazard to buildings or contents,

 to provide an effective electrical path over which ground-fault currents can flow without creating a fire or explosive hazard, and

 to contribute to superior performance of the electrical system (e.g., suppression of differential and common-mode electrical noise, discussed in Chapter 11).

Investigations have pointed out that it is important to have good electric junctions between sections of conduit or metal raceways that are used as equipment grounding paths, and to assure adequate cross-sectional areas and conductivity of these grounding paths. Where systems are solidly grounded, equipment grounding conductors are bonded to the system grounded conductor and the grounding electrode conductor at the service equipment and at the source of a separately derived system in accordance with the National Electrical Code (NEC) Article 250 [55,56].

2.7.3 Static Grounding

The British standard [59] regarding grounding states that the most common source of danger from static electricity is the retention of charge on a conductor, because virtually all the stored energy can be released in a single spark to earth or to another conductor. The accepted method of avoiding the hazard is to connect all conductors to each other and to earth via electrical paths with resistances sufficiently low to permit the relaxation of the charges. This was stated in the original British standard (1980) and is carried out worldwide [60].

Purpose of Static Grounding

The accumulation of static electricity, on materials or equipment being handled or processed by operating personnel, introduces a potentially serious hazard in any area where combustible gases, dusts, or fibers are present. The discharge of the accumulation of static electricity from an object to ground or to another ungrounded object of lower potential is often the cause of a fire or an explosion if it takes place in the presence of readily combustible materials or combustible vapor and air mixtures.

2.7.4 Connection to Earth

Resistance to Earth. The most elaborate grounding system that can be designed may prove to be inadequate unless the connection of the system to the earth is adequate and has a low resistance. It follows, therefore, that the earth connection is one of the most important parts of the whole grounding system. It is also the most difficult part to design and to obtain. The perfect connection to earth should have zero resistance, but this is impossible to obtain. Ground resistance of less than 1 Ω can be obtained, although this low resistance may not be necessary in many cases. Since the desired resistance varies inversely with the fault current to ground, the larger the fault current, the lower must be the resistance [61,62]. Figure 2.50a illustrates the deformation of grounding conductor due to large lightning currents.

f02-50-9780128007822
Figure 2.50 (a) Deformation of grounding conductor due to lightning strike. (b) Force density distribution as it occurs along the x axis for a filamentary 90° bend at a current of I = 100 kA. The force density distribution along the y axis is not shown but is similar to that along the x axis.

2.7.5 Calculation of Magnetic Forces

Considerable work [63,64] has been done concerning the calculation of electromagnetic forces based on closed-form analysis. These methods can be applied to simple geometric configurations only. In many engineering applications, however, three-dimensional geometric arrangements must be analyzed. Based on the theory developed by Lawrenson [65] filamentary conductor configurations can be analyzed. Figures 2.50a,b illustrate the force density distribution for a filamentary 90° bend at a current of I = 100 kA. Three-dimensional arrangements are investigated in [66]. Force calculations based on the so-called Maxwell stress are discussed in Chapter 4.

2.8 Measurement of derating of three-phase transformers

Measuring the real and apparent power derating of three-phase transformers is desirable because additional losses due to power quality problems (e.g., harmonics, DC excitation) can be readily limited before any significant damage due to additional temperature rises occurs [8,9]. Measuring transformer losses from the input power minus the output power in real time is inaccurate because the losses are the difference of two large values; this indirect approach results in maximum errors in the losses of more than 60% for high-efficiency (η > 98%) transformers [7]. The usually employed indirect method consisting of no-load (iron-core loss) and short-circuit (copper loss) tests [67] cannot be performed online while the transformer is partially or fully loaded. IEEE Recommended Practice C57.110 computes the transformer derating based on Rh for various harmonics h, which is derived from the DC winding resistance and the rated load loss [14]. Kelly et al. [68] describe an improved measuring technique of the equivalent effective resistance RT as a function of frequency f of single-phase transformers, which allows the direct calculation of transformer loss at harmonic frequencies from f = 10 Hz up to 100 kHz. This equivalent effective resistance takes into account the total losses of the transformer: the copper losses plus the iron-core losses. Based on the fact that the iron-core losses do not depend on harmonic currents, but depend on harmonic voltages (amplitudes and phase shifts) [69], the total losses determined by [68] are not accurate. In addition, temperature-dependent operating conditions cannot be considered in [68]. Mocci [70] and Arri et al. [71] present an analog measurement circuit to directly measure the total losses for single- and ungrounded three-phase transformers. However, employment of many potential transformers (PTs) and current transformers (CTs) in the three-phase transformer measuring circuits [71] decreases the measurement accuracy.

This section presents a direct method for measuring the derating of three-phase transformers while transformers are operating at any load and any arbitrary conditions. The measuring circuit is based on PTs, CTs, shunts, voltage dividers or Hall sensors [72], A/D converter, and computer (microprocessor). Using a computer-aided testing (CAT) program [7,73], losses, efficiency, harmonics, derating, and wave shapes of all voltages and currents can be monitored within a fraction of a second. The maximum errors in the losses are acceptably small and depend mainly on the accuracy of the sensors used.

2.8.1 Approach

2.8.1.1 Three-Phase Transformers in Δ–Δ or Y–Y Ungrounded Connection

For Δ–Δ or ungrounded Y–Y connected three-phase transformers (see Figs. 2.51 and 2.52 for using shunts, voltage dividers, and optocouplers), one obtains with the application of the two-wattmeter method at the input and output terminals by inspection:

PinPout=1T0TvACiA+vBCiBvacia+vbcibdt,=1T0T[vACiAia+vBCiBib+vACvacia+vBCvbcib]dt.

si148_e  (2-49)

f02-51-9780128007822
Figure 2.51 Eight-channel CAT circuit for accurate Δ–Δ or ungrounded Y–Y connected three-phase transformer loss monitoring using PTs and CTs.
f02-52-9780128007822
Figure 2.52 Eight-channel CAT circuit for accurate Δ–Δ or ungrounded Y-Y connected three-phase transformer loss monitoring using shunts, voltage dividers, and optocouplers.

The term pfet=vACiAib+vBCiBibsi149_e represents the instantaneous iron-core loss, and pcut=vACvacia+vBCvbcibsi150_e is the instantaneous copper loss.

2.8.1.2 Three-Phase Transformers in Δ–Y Connection

Figure 2.53 illustrates the application of the digital data-acquisition method to a Δ–Y connected transformer bank.

f02-53-9780128007822
Figure 2.53 Eight-channel CAT circuit for accurate Δ–Y three-phase transformer loss monitoring using PTs and CTs.

For ungrounded Δ–Y connected three-phase transformers, the currents of the Y side must be referred to the line currents of the Δ side, as shown in Fig. 2.53. The loss of the transformer is given by

PinPout=1T0TvACiAl+vBCiBlvanialvcnibldt,

si151_e  (2-50)

=1T0TvACiAlial+vBCiBlibl+vACvanial+vBC+vcnibldt,

si152_e  (2-51a)

where iAl and iBl are input line currents of phases A and B, respectively; ial=iaibsi153_e and ibl=ibicsi154_e are the two-wattmeter secondary currents; van′ and vcn′ are the output line-to-neutral voltages of phases a and c, respectively.

Because the neutral of the Y-connected secondary winding (N) is not accessible, the secondary phase voltages are measured and referred to the neutral n of the Y-connected PTs (see Fig. 2.53). This does not affect the accuracy of loss measurement, which can be demonstrated below. The output power is

Pout=1T0TvaNia+vbNib+vcNicdt,=1T0Tvan+vnNia+vbn+vnNib+vcn+vnNicdt,=1T0Tvania+vbnib+vcnic+vnNia+ib+icdt,

si155_e  (2-51b)

where N denotes the neutral of the Y-connected secondary winding. Because ia+ib+icsi156_e = 0, the measured output power referred to the neutral n of PTs is the same as that referred to the neutral N of the secondary winding.

2.8.1.3 Accuracy Requirements for Instruments

The efficiencies of high-power electrical apparatus such as single- and three-phase transformers in the kVA and MVA ranges are 97% or higher. For a 15 kVA three-phase transformer with an efficiency of 97.02% at rated operation, the total losses are PLoss = 409.3 W, copper and iron-core losses Pcu = 177.1 W and Pfe = 232.2 W at VL-L = 234.2 V and IL = 32.4 A, respectively, as listed in Table 2.7a.

The determination of the losses from voltage and current differences as described in Fig. 2.51 – where differences are calibrated – greatly reduces the maximum error in the loss measurement. The series voltage drop and exciting current at rated operation referred to the primary of the 15 kVA, 240 V/240 V three-phase transformer are (vACvac) = 3.56 V, (iAia) = 2.93 A, respectively (Table 2.7a). The instruments and their error limits are listed in Table 2.1. In Table 2.1, vAC, vAC – vacʹ, iA – iaʹ, and iaʹ stand for the relevant calibrated voltmeters and ammeters. Because all voltage and current signals are sampled via PTs, CTs (or optocouplers), and current shunts, the error limits for all instruments are equal to the product of the actually measured values and their relative error limits, instead of full-scale errors, as shown in Table 2.1. All error limits are referred to the meter side.

Table 2.1

Instruments and Their Error Limits for Figs. 2.51 and 2.52

Instrument (Fig. 2.51)Instrument (Fig. 2.52)RatingError limits
PT1VDA240 V/240 VentityPT1 = 234 mV
PT2VDa240 V/240 VentityPT2 = 234 mV
CT1SHA30 A/5 AentityCT1 = 5.4 mA
CT2SHa30 A/5 AentityCT2 = 5.4 mA
vACvAC300 VentityV1 = 234 mV
vAC – vacvAC – vac10 VentityV2 = 3.56 mV
iA – iaiA – ia5 AentityA1 = 0.49 mA
iaia5 AentityA2 = 5.4 mA

t0050

In Fig. 2.52, the voltage divider (R1A, R2A) combined with an optocoupler (vAc) emulates the function of a PT without hysteresis. The optocoupler can alter the amplitude of a signal and provides isolation without affecting the phase shift of the signal as it is corrupted by PTs. The current shunt RshAsi157_e and optocouplers iA,ia,iAiasi158_e emulate that of a CT without hysteresis and parasitic phase shift. The prime (ʹ) indicates that iaʹ is about of the same magnitude as iA; this is accomplished by adjustment of the amplifier gain(s) of the optocoupler(s). VD and SH in Table 2.1 stand for voltage divider and shunt, respectively.

The line-to-line voltage is measured with the maximum error of (taking into account the maximum errors of PT1 and voltmeter)

ΔvACvAC=±ɛPT1±ɛV1vAC=±0.234±0.234234=±0.002.

si159_e  (2-52)

The current difference is measured with the maximum error of

ΔiAiaiAia=±ɛCT1±ɛCT2±ɛA1iAia/6=±5.4±5.4±0.49mA0.49A=±0.023.

si160_e  (2-53)

Therefore, the loss component vAC (iA – ia) in Eq. 2-49 is measured with the maximum error of

ΔvACvAC+ΔiAiaiAia=±0.025.

si161_e  (2-54)

The series voltage drop is measured with the maximum error of

ΔvACvacvACvac=±ɛPT1±ɛPT2±ɛv2vACvac=±234±234±3.56mV3.56V=±0.1325.

si162_e  (2-55)

The output current is measured with the maximum error of

Δiaia=±ɛCT2±ɛA2ia/6=±0.0054±0.00545.4=±0.002

si163_e  (2-56)

and the loss component (vAC – vacʹ)iaʹ in Eq. 2.49 is measured with the maximum error of

Δiaia+ΔvACvacvACvac=±0.1345.

si164_e  (2-57)

Thus, the total loss is measured with the maximum error of

ΔPLossPLoss=ΔPcu+ΔPfePLoss=2±0.1345177.1±0.025232.2409.3=±0.145=±14.5%.

si165_e  (2-58)

The above error analysis employs PTs and CTs. If these devices generate too-large errors because of hysteresis, voltage dividers (e.g., VDA) and shunts (e.g., SHA) combined with optocouplers can be used, as indicated in Fig. 2.52 [74]. A similar error analysis using shunts, dividers, and optocouplers leads to the same maximum error in the directly measured losses, provided the same standard maximum errors (0.1%) of Table 2.1 are assumed. The factor 2 in Eq. 2-58 is employed because loss components in Eq. 2-54 and Eq. 2-57 are only half of those in Eq. 2-49.

2.8.1.4 Comparison of Directly Measured Losses with Results of No-Load and Short-Circuit Tests

A computer-aided testing electrical apparatus program (CATEA) [73] is used to monitor the iron-core and copper losses of three-phase transformers. The nameplate data of the tested transformers are given in Appendixes 3.1 to 3.4. The results of the on-line measurement of the iron-core and copper losses for a Y–Y connected 9 kVA three-phase transformer (consisting of three 3 kVA single-phase transformers, Appendix 3.1) are given in Table 2.2 for sinusoidal rated line-to-line voltages of 416 V, where the direct (online) measurement data are compared with those of the indirect (separate open-circuit and short-circuit tests) method. The iron-core loss of the indirect method is larger than that of the direct method because the induced voltage of the former is larger than that of the latter. The copper loss of the indirect method is smaller than that of the direct method because the input current of the former (which is nearly the same as the output current) is smaller than that of the latter.

Table 2.2

Measured Iron-Core and Copper Losses of 9 kVA Y–Y Connected Transformer (VLL = 416 V)

LossesDirect method [7,73]Indirect method [67]
Iron-core loss98.4 W114.1 W
Copper loss221.1 W199.3 W
Total losses319.5 W313.4 W

2.8.2 A 4.5 kVA Three-Phase Transformer Bank #1 Feeding Full-Wave Rectifier

A 4.5 kVA, 240 V/240 V, Δ–Δ connected three-phase transformer (Fig. 2.51) consisting of three single-phase transformers (bank #1, Appendix 3.2) is used to feed a full-wave diode rectifier (see Appendix 3.5) with an LC filter connected across the resistive load (see Fig. 6-5 of [75]). In Table 2.3, measured data are compared with those of linear load condition. The measured wave shapes of input voltage (vAC), exciting current (iA – iaʹ), series voltage drop (vAC – vacʹ), and output current (ia) of phase A are shown in Fig. 2.54a,b for linear and nonlinear load conditions. The total harmonic distortion THDi, K-factor, and harmonic components of the output current are listed in Table 2.4.

Table 2.3

Measured Data of 4.5 kVA Transformer Bank #1

Linear load (Fig. 2.54a)Nonlinear load (Fig. 2.54b)
vAC237.7 Vrms237.7 Vrms
iAiasi24_e1.10 Arms1.12 Arms
vACvacsi25_e8.25 Vrms8.19 Vrms
ia11.09 Arms10.98 Arms
Iron-core loss100.7 W103.9 W
Copper loss161.3 W157.6 W
Total loss262.0 W261.5 W
Output power4532 W4142 W
Efficiency94.55%94.09%
f02-54-9780128007822
Figure 2.54 (a) Measured wave shapes of 4.5 kVA three-phase transformer bank #1 feeding linear load (see rms values of Table 2.3). (b) Measured wave shapes of 4.5 kVA three-phase transformer bank #1 feeding full-wave diode rectifier load (see rms values of Tables 2.3 and 2.4).

Table 2.4

Output Current Harmonic Components, THDi, and K-Factor (Fig. 2.54b)

ih (Arms)
h = 1h = 5h = 7h = 11h = 13THDi (%)K
10.43.550.810.770.2436.344.58

t0065

Table 2.3 compares measured data and shows that the transformer is operated at nonlinear load with about the same losses occurring at linear load (261.3 W). With the apparent power derating definition

deratingS=VrmsIrmsVrmsratedIrmsrated,

si166_e  (2-59)

the derating at the nonlinear load of Table 2.3 is 99%. The real power delivered to the nonlinear load is 91.4% of that supplied at linear load.

2.8.3 A 4.5 kVA Three-Phase Transformer Bank #2 Supplying Power to Six-Step Inverter

A 4.5 kVA transformer bank #2 (Appendix 3.3) supplies power to a half-controlled six-step inverter (Appendix 3.6), which in turn powers a three-phase induction motor. The motor is controlled by adjusting the output current and frequency of the inverter. The transformer is operated at rated loss at various motor speeds. Rated loss of bank #2 is determined by a linear (resistive) load at rated operation. The iron-core and copper losses are measured separately and are listed in Table 2.5a. Measured wave shapes of input voltage, exciting current, series voltage drop, and output current of phase A are shown in Fig. 2.55a,b for linear and nonlinear conditions. The output current includes both odd and even harmonics due to the half-controlled input rectifier of the six-step inverter. Dominant harmonics of input voltage and output current are listed for different motor speeds in Table 2.5b. The total harmonic distortion of the input voltage and output current as well as the K-factor are listed in Table 2.5c for the speed conditions of Tables 2.5a and 2.5b.

Table 2.5a

Measured Iron-Core and Copper Losses of 4.5 kVA Δ–Δ Connected Transformer Bank #2 Feeding a Six-Step Inverter at Various Motor Speeds (Fig. 2.55b)

Motor speed (%)aInput voltage (Vrms)Output current (Arms)Core loss (W)Copper loss (W)Total loss (W)Output power (W)
b237.811.465.7154.9220.64542
100237.911.169.6150.5220.23216
90238.111.562.9158.8221.72913
80238.311.265.6153.3218.92398
70238.311.365.7154.8220.42081
60238.211.464.5162.2226.71862
50238.410.869.2151.3220.61530
40238.611.168.5151.2219.71308

t0070

a Refers to rated speed of 1145 rpm.

b Linear resistive load.

f02-55-9780128007822
Figure 2.55 (a) Measured wave shapes of 4.5 kVA three-phase transformer bank #2 feeding linear load (see rms values of Tables 2.5a and 2.5b). (b) Measured wave shapes of 4.5 kVA three-phase transformer bank #2 feeding half-controlled six-step inverter (see rms values of Tables 2.5a and 2.5b).

Table 2.5b

Input Voltage (in rms Volts) and Output Current (in rms Amperes) Harmonics of Phase A (Fig. 2.55a,b)

h12457810
Linear loadvACh236.90.420.482.961.480.200.40
iahsi26_e11.240.040.020.120.080.000.00
100% speedvACh237.21.370.492.881.580.060.30
iahsi27_e9.055.911.080.560.300.320.20
90% speedvACh237.41.580.532.571.600.100.20
iahsi28_e9.066.831.870.540.540.160.30
80% speedvACh237.51.670.432.911.630.020.30
iahsi29_e8.786.822.470.690.670.430.20
70% speedvACh237.61.670.412.781.500.130.20
iahsi30_e8.646.912.820.990.680.600.20
60% speedvACh237.31.710.283.141.210.140.20
iahsi31_e8.596.903.201.290.680.740.20
50% speedvACh237.81.730.353.181.300.210.00
iahsi32_e8.096.573.521.670.540.790.50
40% speedvACh237.71.730.473.081.670.300.30
iahsi33_e7.786.663.702.170.480.820.80

t0075

Table 2.5c

Measured THD-Values and K-Factor (Fig. 2.55a,b)

ConditionTHDvTHDiK-factor
Linear load1.60%1.41%1.01
100% speed1.62%67.19%2.48
90% speed1.62%78.76%2.76
80% speed1.67%83.65%3.25
70% speed1.58%87.99%3.72
60% speed1.71%90.86%4.40
50% speed1.73%96.05%5.53
40% speed1.88%104.05%6.69

t0080

2.8.4 A 15 kVA Three-Phase Transformer Supplying Power to Resonant Rectifier

A 15 kVA, 240 V/240 V, Δ–Δ connected three-phase transformer (Appendix 3.4) is used to supply power to a resonant rectifier [12] (Appendix 3.7). The transformer is operated with the resonant rectifier load at the same total loss generated by a three-phase linear (resistive) load. Measured data are compared in Table 2.6. Measured wave shapes of input voltage, exciting current, series voltage drop, and output current of phase A are depicted in Fig. 2.56a,b. The fundamental phase shift between output transformer line-to-line voltage vacʹ and phase current iaʹ of the resonant rectifier is 67.33°, and the output displacement factor (within transformer phase) is therefore cos(67.33° – 30°) = 0.795. The fundamental phase shift between output line-to-line voltage and phase current for the linear resistive load is 30.95°, and the output displacement factor (within transformer phase) is, therefore, cos(30.95° – 30°) = 1.0. Note that the wave shapes of the output currents of Fig. 2.56a,b are about sinusoidal.

Table 2.6

Measured Data of 15 kVA Three-Phase Transformer with Resonant Rectifier Load (Fig. 2.56a,b)

Linear loadResonant rectifier
vAC215.7 Vrms215.6 Vrms
iAiasi34_e2.39 Arms2.59 Arms
vACvacsi35_e4.42 Vrms4.8 Vrms
iasi36_e27.00 Arms26.94 Arms
Iron-core loss42.0 W49.8 W
Copper loss191.9 W192.2 W
Total loss233.9 W242 W
Output power10015 W7859 W
Efficiency97.72%97.01%
f02-56-9780128007822
Figure 2.56 (a) Measured wave shapes of 15 kVA three-phase transformer feeding linear load (see rms values of Table 2.6). (b) Measured wave shapes of 15 kVA three-phase transformer feeding resonant rectifier (see rms values of Table 2.6).

If the transformer with the resonant rectifier load is operated (see Table 2.6) at about the same loss as linear load (233.9 W), the output current of the transformer with the resonant rectifier load is 26.94 A, which corresponds to the copper loss of 192.2 W = (242 W – 49.8 W). Therefore, the transformer apparent power derating of the transformer for the nonlinear load is 99.7%. The real power supplied to the nonlinear load is 78.5% of that of the linear load.

2.8.5 A 15 kVA Three-Phase Transformer Bank Absorbing Power from a PWM Inverter

The same transformer of Section 2.8.4 absorbs power from a PWM inverter [12] (Appendix 3.8) and supplies power to a resistive load (case #1) and to a utility system (case #2). The transformer losses are also measured when supplying a linear resistive load fed from sinusoidal power supply (linear load). All measured data are compared in Table 2.7a.

Table 2.7a

Measured Data of 15 kVA Three-Phase Transformer Connected to PWM Inverter (Fig. 2.57a,b,c)

Linear load of transformerCase #1Case #2
vAC234.2 Vrms234.4 Vrms248.2 Vrms
iAiasi37_e2.93 Arms3.30 Arms3.82 Arms
vACvacsi38_e3.56 Vrms3.86 Vrms4.28 Vrms
iasi39_e32.4 Arms33.3 Arms28.2 Arms
Iron-core loss232.2 W254.6 W280.5 W
Copper loss177.1 W172.6 W130.4 W
Total loss409.3 W427.2 W410.9 W
Output power13337 W13544 W12027 W
Efficiency97.02%96.94%96.07%

t0090

Measured wave shapes of input voltage, exciting current, series voltage drop, and output current of phase A are shown in Fig. 2.57a,b,c. The total harmonic distortion THDi, K-factor, and harmonic amplitudes of the transformer output current are listed in Table 2.7b.

f02-57-9780128007822
Figure 2.57 (a) Measured wave shapes of 15 kVA three-phase transformer feeding linear load (see rms values of Table 2.7a). (b) Measured wave shapes of 15 kVA three-phase transformer fed by a PWM inverter (case #1) (see rms values of Tables 2.7a and 2.7b). (c) Measured wave shapes of 15 kVA three-phase transformer fed by a PWM inverter (case #2) (see rms values of Tables 2.7a and 2.7b).

Table 2.7b

Output Current Harmonics, THDi, and K-Factor (Fig. 2.57b,c)

Ih(A)THDi
Caseh = 1h = 5h = 7h = 11h = 13(%)K
#132.80.530.870.120.195.841.7
#228.21.151.360.190.3810.02.7

t0095

Figure 2.58 summarizes the total harmonic distortion (THDi), apparent power (kVA) derating, and real power (kW) derating for uncontrolled (a, b), half-controlled (c), and controlled (d, e) converter loads of transformers. In particular the graphs of Fig. 2.58 can be identified [75] as follows:

f02-58-9780128007822
Figure 2.58 Total harmonic distortion of current (THDi), apparent power (kVA) derating, and real power (kW) derating for uncontrolled (a, b), half-controlled (c), and controlled (d, e) converter loads that is, inverter (d), and resonant rectifier (e).

a) 25 kVA single-phase pole transformer [7,11,15,75,76] feeding uncontrolled full wave rectifier load,

b) 4.5 kVA three-phase transformer feeding uncontrolled full wave rectifier load,

c) 4.5 kVA three-phase transformer feeding half-controlled rectifier load,

d) 15 kVA three-phase transformer absorbing power from PWM inverter (14 kW), and

e) 15 kVA three-phase transformer feeding resonant rectifier load (8 kW).

2.8.6 Discussion of Results and Conclusions

2.8.6.1 Discussion of Results

A new approach for the measurement of the derating of three-phase transformers has been described and applied under nonsinusoidal operation. It extends the measurement approach of single-phase transformers [7,73,12,7577] to three-phase transformers [8,75,76,78].

The apparent power (kVA) derating (Eq. 2-59) of three-phase transformers is not greatly affected by the THDi. Even for THDi values of about 70%, derating is about 99%.

The real power (kW) derating is greatly affected (see Fig. 2.58) by the current wave shape generated by solid-state converters, in particular by the phase shift of the fundamental current component. Therefore, inverters and rectifiers should be designed such that they supply and draw power, respectively, at a displacement (power) factor of about 1.

Three-phase transformers have similar derating properties as single-phase transformers [7,12,73,7577]. The maximum error of the directly measured losses is about 15%, which compares favorably with the maximum error of more than 60% [7,73] for loss measurement based on the difference between input and output powers as applied to high-efficiency (η > 98%) transformers.

Transformers of the same type may have significantly different iron-core losses as measured in Table 2.3 (100.7 W) and Table 2.5a (65.7 W). Small transformers (kVA range) have relatively small wire cross sections resulting in small skin-effect losses. Large transformers (MW range) have aluminum secondary windings with relatively large wire cross sections resulting in relatively large skin-effect losses. For this reason substation transformers can be expected to have larger apparent power derating than the ones measured in this section. Unfortunately, transformers in the MW range cannot be operated in a laboratory under real-life conditions. Therefore, it is recommended that utilities sponsor the application of the method of this section and permit on-site measurements.

2.8.6.2 Comparison with Existing Techniques

The maximum error of the directly measured losses is about 15% (using potential and current transformers), which compares favorably with the maximum error of more than 60% [7,73] (employing shunts and voltage dividers) for loss measurement based on the difference between input and output powers as applied to high-efficiency (η > 98%) transformers.

The technique of [68] uses the premeasured transformer effective resistance RT as a function of frequency f to calculate transformer total losses for various harmonic currents. This method can be classified as an indirect method because the transformer losses are obtained by computation, instead of direct measurement. In addition, the approach of [68] neglects the fact that the iron-core losses are a function of the harmonic phase shift [69]; in other words, the RT values are not constant for any given harmonic current amplitude but vary as a function of the harmonic voltage amplitude and phase shift as well. Finally, temperature-dependent operating conditions, for example, cannot be considered in [68]. For the above reasons, the method of [68] must be validated by some direct measurements, such as the method presented in this section.

The method of Mocci [70] and Arri et al. [71] has not been practically applied to three-phase transformers. The presented measurement circuit for three-phase transformers [71] uses too many instrument transformers (e.g., nine CTs and nine PTs), and therefore, the measured results will be not as accurate as those based on the measurement circuits of this section, where only four CTs and five PTs are used as shown in Fig. 2.53.

2.9 Summary

This chapter addresses issues related to modeling and operation of transformers under (non)sinusoidal operating conditions. It briefly reviews the operation and modeling of transformers at sinusoidal conditions. It highlights the impact of harmonics and poor power quality on transformer (fundamental and harmonic) losses and discusses several techniques for derating of single- and three-phase transformers including K-factor, FHL-factor, and loss measurement techniques. Several transformer harmonic models in the time domain and/or frequency domain are introduced. Important power quality problems related to transformers such as ferroresonance and geomagnetically induced currents (GICs) are explored. Different techniques for system and transformer grounding are explained. Many application examples explaining nonsinusoidal flux, voltage and current, harmonic (copper and iron-core) loss measurements, derating, ferroresonance, magnetic field strength, propagation of surge, and operation of lightning arresters are presented for further exploration and understanding of the presented materials.

2.10 Problems

Problem 2.1 Pole Transformer Supplying a Residence

Calculate the currents flowing in all branches as indicated in Fig. P2.1 if the pole transformer is loaded as shown. For the pole transformer you may assume ideal transformer relations. Note that such a single-phase configuration is used in the pole transformer (mounted on a pole behind your house) supplying residences within the United States with electricity. What are the real (P), reactive (Q), and apparent (S) powers supplied to the residence?

f02-85-9780128007822
Figure P2.1 Single-phase transformer loaded by three single-phase loads, where R1 represents a refrigerator, R2 a microwave oven, and R3 an electric stove.

Problem 2.2 Analysis of an Inductor with Three Coupled Windings

a) Draw the magnetic equivalent circuit of Fig. P2.2.

f02-86-9780128007822
Figure P2.2 Analysis of an inductor with three coupled windings.

b) Determine all self- and mutual inductances. Assuming that the winding resistances are known draw the electrical equivalent circuit of Fig. P2.2.

Problem 2.3 Derating of Transformers Based on K-Factor

A substation transformer (S = 10 MVA, VL-L = 5 kV, %PEC-R = 10) supplies a six-pulse rectifier load with the harmonic current spectrum as given by Table X of the paper by Duffey and Stratford [82].

a) Determine the K-factor of the rectifier load.

b) Calculate the derating of the transformer for this nonlinear load.

Problem 2.4 Derating of Transformers Based on FHL-Factor

A substation transformer (S = 10 MVA, VL-L = 5 kV, %PEC-R = 10) supplies a six-pulse rectifier load with the harmonic current spectrum as given by Table X of the paper by Duffey and Stratford [82].

a) Determine the FHL-factor of the rectifier load.

b) Calculate the derating of the transformer for this nonlinear load.

Problem 2.5 Derating of Transformers Based on K-Factor for a Triangular Current Wave Shape

a) Show that the Fourier series representation for the current waveform in Fig. P2.5 is

it=8Imπ2h=1,3,5,1h2sinhπ2sinhω0t.

si167_e

f02-87-9780128007822
Figure P2.5 Triangular current wave shape.

b) Approximate the wave shape of Fig. P2.5 in a discrete manner, that is, sample the wave shape using 36 points per period and use the program of Appendix 2 to find the Fourier coefficients.

c) Based on the Nyquist criterion, what is the maximum order of harmonic components hmax you can compute in part b? Note that for hmax the approximation error will be a minimum.

d) Calculate the K-factor and the derating of the transformer when supplied with a triangular current wave shape, and %PEC-R = 10. The rated current IR corresponds to the fundamental component of the triangular current.

Problem 2.6 Derating of Transformers Based on K-Factor for a Trapezoidal Current Wave Shape

a) Show that the Fourier series representation for the current waveform in Fig. P2.6 is

it=12Imπ2h=1,3,5,1h2sinhπ3sinhω0t.

si168_e

f02-88-9780128007822
Figure P2.6 Trapezoidal current wave shape.

b) Approximate the wave shape of Fig. P2.6 in a discrete manner, that is, sample the wave shape using 36 points per period and use the program of Appendix 2 to find the Fourier coefficients.

c) Based on the Nyquist criterion, what is the maximum order of harmonic components hmax you can compute in part b? Note that for hmax the approximation error will be a minimum.

d) Calculate the K-factor and the derating of the transformer when supplied with a trapezoidal current wave shape, and % PEC-R = 10. The rated current IR corresponds to the fundamental component of the trapezoidal current.

Problem 2.7 Force Calculation for a Superconducting Coil for Energy Storage

Design a magnetic storage system which can provide for about 6 minutes a power of 100 MW, that is, an energy of 10 MWh. A coil of radius R = 10 m, height h = 6 m, and N = 3000 turns is shown in Fig. P2.7. The magnetic field strength H inside such a coil is axially directed, essentially uniform, and equal to H = Ni/h. The coil is enclosed by a superconducting box of 2 times its volume. The magnetic field outside the superconducting box can be assumed to be negligible and you may assume a permeability of free space for the calculation of the stored energy.

f02-89-9780128007822
Figure P2.7 Superconducting coil for energy storage.

a) Calculate the magnetic field intensity H and the magnetic flux density B of the coil for constant current i = I0 = 8000 A.

b) Calculate the energy Estored stored in the magnetic field (either in terms of H or B).

c) Determine the radial pressure pradial acting on the coil.

d) Devise an electronic circuit that enables the controlled extraction of the energy from this storage device. You may use bidirectional superconducting switches S1 and S2, which are controlled by a magnetic field via the quenching effect. S1 is ON and S2 is OFF for charging the superconducting coil. S1 is OFF and S2 is ON for the storage time period. S1 is ON and S2 is OFF for discharging the superconducting coil.

Problem 2.8 Measurement of Losses of Delta/Zigzag Transformers

Could the method of Section 2.8 also be applied to a Δ/zigzag three-phase transformer as shown in Fig. P2.8? This configuration is of interest for the design of a reactor [67] where the zero-sequence (e.g., triplen) harmonics are removed or mitigated, and where through phase shifting the 5th and 7th harmonics [83] are canceled.

f02-90-9780128007822
Figure P2.8 Delta/zigzag three-phase transformer configuration.

Problem 2.9 Error Calculations

For a 25 kVA single-phase transformer of Fig. P2.9 with an efficiency of 98.46% at rated operation, the total losses are Ploss = 390 W, where the copper and iron-core losses are Pcu = 325 W and Pfe = 65 W, respectively. Find the maximum measurement errors of the copper Pcu and iron-core Pfe losses as well as that of the total loss Ploss whereby the instrument errors are given in Table P2.9.

f02-91-9780128007822
Figure P2.9 Instrumentation for single-phase transformers employing conventional approach for loss measurement.

Table P2.9

Instruments and Their Error Limits of Fig. P2.9

InstrumentFull scaleFull-scale errors
PT1240 V/240 VentityPT1 = 0.24 V
PT27200 V/240 VentityPT2 = 0.24 V
CT1100 A/5 AentityCT1 = 5 mA
CT25 A/5 AentityCT2 = 5 mA
V1300 VentityV1 = 0.3 V
V2300 VentityV2 = 0.3 V
A15 AentityA1 = 5 mA
A25 AentityA2 = 5 mA

Problem 2.10 Delta/Wye Three-Phase Transformer Configuration with Six-Pulse Midpoint Rectifier Employing an Interphase Reactor

a) Perform a PSpice analysis for the circuit of Fig. P2.10, where a three-phase thyristor rectifier with filter (e.g., capacitor Cf) serves a load (Rload). You may assume ideal transformer conditions. For your convenience you may assume (N1/N2) = 1, Rsyst = 0.01 Ω, Xsyst = 0.05 Ω @ f = 60 Hz, vABt=2600Vcosωtsi169_e, vBCt=2600Vcosωt120°si170_e, and vCAt=2600Vcosωt240°si171_e. Each of the six thyristors can be modeled by a self-commutated electronic switch and a diode in series, as is illustrated in Fig. E1.1.2 of Application Example 1.1 of Chapter 1. Furthermore, Cf = 500 μF, Lleft = 5 mH, Lright = 5 mH, and Rload = 10 Ω. You may assume that the two inductors Lleft and Lright are not magnetically coupled, that is, Lleftright = 0. Plot one period of either voltage or current after steady state has been reached for firing angles of α = 0°, 60°, and 120°, as requested in parts b to e. Print the input program.

f02-92-9780128007822
Figure P2.10 Connection of a delta/wye three-phase transformer configuration with six-pulse midpoint rectifier employing an interphase reactor.

b) Plot and subject the line-to-line voltages vAB (t), vab(t), and v′ab(t) to Fourier analysis. Why are they different?

c) Plot and subject the input line current iAL (t) of the delta primary to a Fourier analysis. Note that the input line currents of the primary delta do not contain the 3rd, 6th, 9th, 12th,…, that is, harmonic zero-sequence current components.

d) Plot and subject the phase current iAph(t) of the delta primary to a Fourier analysis. Note that the phase currents of the primary delta contain the 3rd, 6th, 9th, 12th,…, that is, harmonic zero-sequence current components.

e) Plot and subject the output voltage vload(t) to a Fourier analysis.

f) How do your results change if the two inductors Lleft and Lright are magnetically coupled, that is, Lleftright = 4.5 mH?

Problem 2.11 Determination of Measurement Uncertainty

Instead of maximum errors, the U.S. Guide to the Expression of Uncertainty in Measurement [84] recommends the use of uncertainty. As an example the uncertainties and the associated correlation coefficients of the simultaneous measurement of resistance and reactance of an inductor having the inductance L and the resistance R are to be calculated. The five measurement sets of Table P2.11 are available for an inductor consisting of the impedance Z = R + jX, where terminal voltage |t|, current |Ĩt|, and phase-shift angle Φ are measured.

Table P2.11

Measurement Values of the Input Quantities |entityt|, |Ĩt|, and Φ Obtained from Five Sets of Simultaneous Observations

Input quantities
Set number k|V~t|si40_e (V)|I~t|si41_e (mA)Φ (radians)
15.00619.6641.0457
24.99319.6381.0439
35.00419.6411.0468
44.99119.6841.0427
54.99819.6791.0433

t0105

With R = (|V~t|si172_e/ |I~t|si173_e)cosΦ, X = (|V~t|si174_e/ |I~t|si175_e)sinΦ, Z = (|V~t|si176_e/ |I~t|si177_e), and Z2 = R2 + X2 one obtains two independent output quantities – called measurands.

a) Calculate the arithmetic means of |V~t|si178_e, |I~t|si179_e, and Φ.

b) Determine the experimental standard deviation of mean of s(V¯si180_e), s(Ī), and s(Φ¯si181_e)

c) Find the correlation coefficients r(|t|, |Ĩt|), r|V~t|,Φ,si182_e and r|I~t|,Φsi183_e.

d) Find the combined standard uncertainty μ ≡ uc(y) of result of measurement where is the measurand index.

e) Determine the correlation coefficients r(R/X), r(R/Z), and r(X/Z).

References

[1] Fuchs EF, Masoum MAS, Roesler D. Large signal nonlinear model of anisotropic transformers for nonsinusoidal operation, part I: λ–i characteristic. IEEE Trans Power Deliv. 1991;6(1):174–186.

[2] Masoum MAS, Fuchs EF, Roesler D. Large signal nonlinear model of anisotropic transformers for nonsinusoidal operation, part II: magnetizing and core loss currents. IEEE Trans Power Deliv. 1991;6(4):1509–1516.

[3] Masoum MAS, Fuchs EF, Roesler D. Impact of nonlinear loads on anisotropic transformers. IEEE Trans Power Deliv. 1991;6(4):1781–1788.

[4] Masoum MAS, Fuchs EF. Transformer magnetizing current in harmonic power flow. IEEE Trans Power Deliv. 1994;9(1):1020.

[5] Fuchs EF, Fei R. A new computer-aided method for the efficiency measurement of low-loss transformers and inductors under nonsinusoidal operation. IEEE Trans Power Deliv. 1996;PWRD-11(1):292–304.

[6] Stensland T, Fuchs EF, Grady WM, Doyle M. Modeling of magnetizing and core-loss currents in single-phase transformers with voltage harmonics for use in power flow. IEEE Trans Power Deliv. 1997;12(2):768–774.

[7] Fuchs EF, Yildirim D, Batan T. Innovative procedure for measurement of losses of transformers supplying nonsinusoidal loads. IEE Proc Generat Transm Distrib. 1999;146(6) IEE Proceedings online no. 19990684.

[8] Fuchs EF, Lin D, Martynaitis J. Measurement of three-phase transformer derating and reactive power demand under nonlinear loading conditions. IEEE Trans Power Deliv. 2006;21(2):665–672.

[9] Lin D, Fuchs EF. Real-time monitoring of iron-core and copper losses of three-phase transformers under (non)sinusoidal operation. IEEE Trans Power Deliv. 2006;21(3):1333–1341.

[10] Montsinger VM. Loading transformer by temperature. Trans AIEE. 1930;49:776.

[11] Fuchs EF, Yildirim D, Grady WM. Measurement of eddy-current loss coefficient PEC-R, derating of single-phase transformers, and comparison with K-factor approach. IEEE Trans Power Deliv. 2000;15(1):148–154 and 2000;15(4):1331–1333 (discussions/closure) and p. 1357.

[12] Yildirim D. Commissioning of 30 kVA variablespeed direct-drive wind power plant. Ph.D. Thesis University of Colorado, Boulder; May 1999.

[13] Standard UL 1561. Dry-type general purpose and power transformers. April 22, 1994.

[14] Standard ANSI/IEEE C57.110. Recommended practice for establishing transformer capability when supplying nonsinusoidal load currents. February 1998.

[15] Yildirim D, Fuchs EF. Measured transformer derating and comparison with harmonic loss factor FHL approach. IEEE Trans Power Deliv. 2000;15(1):186–191 and 2000;15(4):1328–1329 (discussions/closure) and p. 1357.

[16] Fuchs EF, Masoum MAS, Ladjevardi M. Effects on distribution feeders from electronic loads based on future peak-load growth, part I: measurements. In: Proceedings of the IASTED international conference on power and energy systems (EuroPES 2005); June 15–17, 2005 Paper #468-019, Benalmadena, Spain.

[17] Dommel H. Digital computer solution of electromagnetic transients in single and multiphase network. IEEE Trans Power App Syst. 1974;PAS-88:487–493.

[18] Dommel H. Transformer models in the simulation of electromagnetic transient. In: 5th power systems computation conference; September 1–5, 1975:1–16 Paper 3.1/4, Cambridge, England.

[19] Hatziantoniu C, Galanos GD, Milias-Argitis J. An incremental transformer model for the study of harmonic overvoltages in weak AC/DC systems. IEEE Trans Power Deliv. 1988;3(3):1111–1121.

[20] Dolinar D, Pihler J, Grcar B. Dynamic model of a three-phase power transformer. IEEE Trans Power Deliv. 1993;8(4):1811–1819.

[21] Leon F, Semlyen A. Complete transformer model for electromagnetic transients. IEEE Trans Power Deliv. 1994;9(1):231–239.

[22] Garcia S, Medina A, Perez C. A state space single-phase transformer model incorporating nonlinear phenomena of magnetic saturation and hysteresis for transient and periodic steady-state analysis. 2417–2421. IEEE Power Engineering Society summer meeting, 2000. 16–20 July 2000;vol. 4.

[23] Garcia S, Medina A. A state space three-phase multilimb transformer model in time domain: fast periodic steady-state analysis. 1859–1864. IEEE Power Engineering Society summer meeting, 2001. 15–19 July 2001;vol. 3.

[24] Reference manual for electromagnetic transient program (EMTP), Bonneville Power Administration, Portland, Oregon.

[25] Chan JH, Vladimirescu A, Gao XC, Liebmann P, Valainis J. Nonlinear transformer model for circuit simulation. IEEE Trans Comput Aided Des. 1991;10(4):476–482.

[26] Huang SR, Chung SC, Chen BN, Chen YH. A harmonic model for the nonlinearities of single-phase transformer with describing functions. IEEE Trans Power Deliv. 2003;18(3):815–820.

[27] Medina A, Arrillaga J. Generalized modeling of power transformers in the harmonic domain. IEEE Trans Power Deliv. 1992;7(3):1458–1465.

[28] Semlyen A, Acha E, Arrillaga J. Newton-type algorithms for the harmonic phasor analysis of nonlinear power circuits in periodical steady state with special reference to magnetic nonlinearities. IEEE Trans Power Deliv. 1988;3(3):1090–1099.

[29] Pedra J, Corcoles F, Sainz L, Lopez R. Harmonic nonlinear transformer modeling. IEEE Trans Power Deliv. 2004;19(2):884–890.

[30] Dugui W, Zheng X. Harmonic model of transformer. 1045–1049. Proceedings of 1998 international conference on power system technology, POWERCON ‘98. 18–21 Aug. 1998;vol. 2.

[31] Greene JD, Gross CA. Nonlinear modeling of transformers. IEEE Trans Ind Appl. 1988;24(3):434–438.

[32] Yamada S, Bessho K, Lu J. Harmonic balance finite element method applied to nonlinear AC magnetic analysis. IEEE Trans Magn. 1989;25(4):2971–2973.

[33] Silvester P, Chari MVK. Finite element solution of saturable magnetic field problem. IEEE Trans Power App Syst. 1970;PAS-89(7):1642–1651.

[34] Enokizono M, Soda N. Finite element analysis of transformer model core with measured reluctivity tensor. IEEE Trans Magn. 1997;33(5):4110–4112.

[35] Masoum MAS. Generation and propagation of harmonics in power system feeders containing nonlinear transformers and loads. Ph.D. Thesis University of Colorado at Boulder; April 1991.

[36] Stensland TD. Effects of voltage harmonics on single-phase transformers and induction machines including pre-processing for power flow. MS Thesis Colorado: University of Colorado at Boulder; 1995.

[37] Marti JR, Soudack AC. Ferroresonance in power systems: fundamental solutions. IEEE Proc C Generat Transm Distrib [see also IEE Proc Generat Transm Distrib]. 1991;138(4):321–329.

[38] Li Y, Shi W, Qin R, Yang J. A systematical method for suppressing ferroresonance at neutral-grounded substations. IEEE Trans Power Deliv. 2003;18(3):1009–1014.

[39] Fuchs EF, You Y, Roesler DJ. Modeling, simulation and their validation of three-phase transformers with three legs under DC bias. IEEE Trans Power Deliv. 1999;14(2):443–449.

[40] Fuchs EF, You Y. Measurement of (λ–i) characteristics of asymmetric three-phase transformers and their applications. IEEE Trans Power Deliv. 2002;17(4):983–990.

[41] Young FS, Schmid RL, Fergestad PI. A laboratory investigation of ferroresonance in cable-connected transformers. IEEE Trans Power App Syst. 1968;PAS-87(5):1240–1249.

[42] Koch HJ. Gas-insulated transmission line (GIL). 0-7803-7989-6/03/$17.00 copyright 2003 IEEE, Panel Discussion: Gas-Insulated Transmission Lines. In: Proceedings of the IEEE PES general meeting, Toronto. p. 2480–2483.

[43] Koch H. Gas-insulated line (GIL) of the 2nd generation. In: AC-DC power transmission. 28–30 November 2001:39–43 Conference Publication No. 485 copyright IEEE 2001.

[44] Koch H, Hillers T. Second-generation gas-insulated line. Power Eng J. June 2002;111–116.

[45] Kieny C, Le Roy G, Sbai A. Ferroresonance study using Galerkin method with pseudo-arclength continuation method. IEEE Trans Power Deliv. 1991;6(4):1841–1847.

[46] Lahtinen M, Elovaara J. GIC occurrences and GIC test for 400 kV system transformer. IEEE Trans Power Deliv. 2002;17(2):555–561.

[47] Boteler DH, Shier RM, Watanabe T, Horita RE. Effects of geomagnetically induced currents in the BC Hydro 500 kV system. IEEE Trans Power Deliv. 1989;4(1):818–823.

[48] Kappenman JG, Norr SR, Sweezy GA, Carlson DL, Albertson VD, Harder JE, et al. GIC mitigation: a neutral blocking/bypass device to prevent the flow of GIC in power systems. IEEE Trans Power Deliv. 1991;6(3):1271–1281.

[49] Gish WB, Feero WE, Rockefeller GD. Rotor heating effects from geomagnetic induced currents. IEEE Trans Power Deliv. 1994;9(2):712–719.

[50] Bozoki B, Chano SR, Dvorak LL, Feero WE, Fenner G, Guro EA, et al. The effects of GIC on protective relaying. IEEE Trans Power Deliv. 1996;11(2):725–739.

[51] Dickmander DL, Lee SY, Desilets GL, Granger M. AC/DC harmonic interactions in the presence of GIC for the Quebec–New England Phase II HVDC transmission. IEEE Trans Power Deliv. 1994;9(1):68–78.

[52] Walling RA, Khan AN. Characteristics of transformer exciting-current during geomagnetic disturbances. IEEE Trans Power Deliv. 1991;6(4):1707–1714.

[53] You Y, Fuchs EF, Lin D, Barnes PR. Reactive power demand of transformers with DC bias. IEEE Ind Appl Soc Mag. 1996;2(4):45–52.

[54] Fardoun AA, Fuchs EF, Masoum MAS. Experimental analysis of a DC bucking motor blocking geomagnetically induced currents. IEEE Trans Power Deliv. January, 1994;TPWRD-9:88–99.

[55] Castenschiold R, Johnson GS. Proper grounding of on-site electrical power systems. IEEE Ind Appl Soc Mag. 2001;7(2):54–62.

[56] American National Standard ANSI/IEEE C114.1. IEEE recommended practice for grounding of industrial and commercial power systems, the IEEE green book. 1973.

[57] Kraft LA. Modelling lightning performance of transmission systems using PSpice. IEEE Trans Power Syst. 1991;6(2):543–549.

[58] Johnk CTA. Engineering electromagnetic fields and waves. New York, NY: John Wiley & Sons; 1975.

[59] British standard code of practice for control of undesirable static electricity, pt. 1, sec. 4, 1991.

[60] Luttgens G. Static electricity hazards in the use of flexible intermediate bulk containers (FIBCs). IEEE Trans Ind Appl. 1997;33(2):444–446.

[61] Pillai P, Bailey BG, Bowen J, Dalke G, Douglas BG, Fischer J, et al. Grounding and ground fault protection of multiple generator installations on mediumvoltage industrial and commercial systems-Part 2: grounding methods working group report. IEEE Trans Ind Appl. 2004;40(1):17–23.

[62] Meliopoulos APS, Dunlap J. Investigation of grounding related problems in AC/DC converter stations. IEEE Trans Power Deliv. 1988;3(2):558–567.

[63] Frick CW. Electromagnetic forces on conductors with bends, short lengths and cross-overs. Gen Electr Rev 36(5).

[64] Dwight HB. Calculation of magnetic forces in disconnecting switches. AIEE Trans. 1920;39:1337.

[65] Lawrenson PJ. Forces on turbogenerator end windings. Proc IEE. 1965;112(6).

[66] Al-Barrak AS. Electromagnetic forces and bending moments on conductors of arbitrary polygonal filament and finite shape. Independent Study University of Colorado, Boulder; 1979.

[67] Electrical transmission and distribution reference book. Westinghouse Electric Corporation; 1964.

[68] Kelly AW, Edwards SW, Rhode JP, Baran M. Transformer derating for harmonic currents: a wideband measurement approach for energized transformers. In: Proceedings of the industry applications conference, 1995; 840–847. Thirtieth IAS annual meeting, IAS 1995. Oct. 8–12, 1995;1.

[69] Masoum MAS, Fuchs EF. Derating of anisotropic transformers under nonsinusoidal operating conditions. Electr Power Energ Syst. 2003;25:1–12.

[70] Mocci F. Un nuovo methodo per la determinazione della potenza assorbita dai doppi bipoli. L‘Energia Elettrica. (7–8):1989;323–329.

[71] Arri E, Locci N, Mocci F. Measurement of transformer power losses and efficiency in real working conditions. IEEE Trans Instrum Meas. 1991;40(2):384–387.

[72] Current and voltage transducer catalog. 3rd ed. Milwaukee, WI: LEM U.S.A., Inc.

[73] Lin D, Fuchs EF, Doyle M. Computer-aided testing of electrical apparatus supplying nonlinear loads. IEEE Trans Power Syst. 1997;12(1):11–21.

[74] Rieman S. An optically isolated data acquisition system. Independent Study University of Colorado, Boulder; December 1997.

[75] Grady WM, Fuchs EF, Lin D, Stensland TD. The potential effects of single-phase power electronic-based loads on power system distortion and losses, Vol. 2: Single-phase transformers and induction motors. Technical Report published by the Electric Power Research Institute (EPRI), Palo Alto, CA, 2003, 1000655; September 2003.

[76] Fuchs EF. Transformers, liquid filled. In: Webster JG, editor. Encyclopedia of electrical and electronics engineering 2000. New York, NY: John Wiley & Sons; 10158, paper No. 934C, 40 pages.

[77] Batan T. Discussion of “Measurement of eddy-current loss coefficient PEC-R, derating of single phase transformers, and comparison with K-factor approach”. IEEE Trans Power Deliv. 2000;15(4):1331–1333.

[78] Fuchs EF, You Y, Lin, D. Development and validation of GIC transformer models. Final Report for Contract 19X-SK205V, Prepared for Martin Marietta Energy Systems, Inc., Oak Ridge, TN, June 1996.

[79] High speed bipolar monolithic sample/hold amplifier. Burr-Brown, SHC5320.

[80] DAS-50 User’s guide. Taunton, MA: Keithley MetraByte Corporation, 440 Myles Standish Boulevard.

[81] Fuchs EF, Fei R, Hong J. Harmonic losses in isotropic and anisotropic transformer cores at low frequencies (60–3,000 Hz). Final Report for Contract No. 19X-SD079C. Prepared for Martin Marietta Energy Systems, Inc., by University of Colorado, Boulder; January 1991.

[82] Duffey CK, Stratford RP. Update of harmonic standard IEEE-519: IEEE recommended practices and requirements for harmonic control in electric power systems. IEEE Trans Ind Appl. 1989;25(6):1025–1034.

[83] Neutral Current Eliminator (NCE) and Combined Neutral Current Eliminator (CNCE). MIRUS International Inc., private communication, http://mirusinternational.com.

[84] American national standard for calibration—U.S. guide to the expression of uncertainty in measurement. NCSL International, ANCI/NCSL Z540–2-1997 (r2002).

Additional bibliography

[85] Hesterman B. Time-domain K-factor computation methods. In: Proc. PCIM 1994, 29th Int. power conference; Sept. 1994:406–417.

[86] Fuchs EF, Roesler DJ, Kovacs KP. Aging of electrical appliances due to harmonics of the power system’s voltage. IEEE Trans Power Deliv. 1986;TPWRD-1(3):301–307.

[87] Fuchs EF, Roesler DJ, Alashhab FS. Sensitivity of electrical appliances to harmonics and fractional harmonics of the power system’s voltage, part I. IEEE Trans Power Deliv. 1987;TPWRD-2(2):437–444.

[88] Fuchs EF, Roesler DJ, Kovacs KP. Sensitivity of electrical appliances to harmonics and fractional harmonics of the power system’s voltage, part II. IEEE Trans Power Deliv. 1987;TPWRD-2(2):445–453.

[89] Masoum MAS, Fuchs EF. Transformer magnetizing current and core losses in harmonic power flow. IEEE Trans Power Deliv. 1994;PWRD-9:10–20.

[90] Chowdhury AH, Grady WM, Fuchs EF. An investigation of harmonic characteristics of transformer excitation current under nonsinusoidal supply voltage. IEEE Trans Power Deliv. 1999;14(2):450–458.

[91] Fuchs EF, Roesler DJ, Masoum MAS. Are harmonic recommendations according to IEEE and IEC too restrictive? IEEE Trans Power Deliv. 2004;19(4):11775–11786.

[92] Fuchs EF, Masoum MAS. Suppression of harmonic distortion in power systems due to geomagnetically induced currents (GICs) through enforcing gic balance in all phases of a system. Invention Disclosure University of Colorado, Boulder; June 1, 1992.

[93] Fuchs EF. A new method for measuring the efficiency of low-loss electrical apparatus. Invention Disclosure University of Colorado, Boulder; February, 1994.

[94] You Y, Fuchs EF. Means for suppressing harmonic distortion due to DC bias currents/fluxes in electric circuits containing any three-phase transformers through balancing and canceling of DC MMF/flux of every phase. Invention Disclosure University of Colorado, Boulder; June, 1994.

[95] You Y, Fuchs EF. High AC impedance, low DC resistance inductor. Invention Disclosure University of Colorado, Boulder; June, 1994.

[96] Fuchs EF, Lin D. On-line monitoring of iron-core and copper losses of single and three-phase transformers under (non)sinusoidal operation at any load. Invention Disclosure University of Colorado, Boulder; January 1995.

[97] Fuchs EF, Yildirim D. On-line technique for separate measurement of copper loss, iron-core loss and (rated) eddy-current loss PEC-R of transformers supplying nonlinear loads. Invention Disclosure University of Colorado, Boulder; April 1998.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset