Chapter 3

Modeling and Analysis of Induction Machines

Abstract

Investigates the performance of induction machines as a function of the power system’s fundamental and harmonic voltages/currents. It includes sinusoidal model of induction machine, time and space harmonics, forward- and backward-rotating harmonic magnetic fields, and magnetic field- and torque calculations based on the finite-difference method. The steady-state stability of rotating machines is reviewed, and the extension of the variable-speed range is obtained through compensation of flux weakening. Fundamental, integer, inter-, fractional and sub-harmonic torque characteristics of three-phase induction machines are analysed. The interaction of space and time harmonics, voltage-stress winding failures, classification and available harmonic models of induction machines as well as discussions of rotor eccentricity are provided. 14 application examples with solutions and 14 application-oriented problems are listed.

Keywords

Induction Machine Time

Space

Inter

Sub- Harmonics

Magnetic Fields

Torques

Winding Failures

Rotor Eccentricity

Before the development of the polyphase AC concept by Nikola Tesla [1] in 1888, there was a competition between AC and DC systems. The invention of the induction machine in the late 1880s completed the AC system of electrical power generation, transmission, and utilization. The Niagara Falls hydroplant was the first large-scale application of Tesla’s polyphase AC system. The theory of single- and three-phase induction machines was developed during the first half of the twentieth century by Steinmetz [2], Richter [3], Kron [4], Veinott [5], Schuisky [6], Bödefeld [7], Alger [8], Fitzgerald et al. [9], Lyon [10], and Say [11] – just to name a few of the hundreds of engineers and scientists who have published in this area of expertise. Newer textbooks and contributions are by Match [12], Chapman [13], and Fuchs et al. [14,15]. In these works mostly balanced steady-state and transient performances of induction machines are analyzed. Most power quality problems as listed below are neglected in these early publications because power quality was not an urgent issue during the last century.

Today, most industrial motors of one horsepower or larger are three-phase induction machines. This is due to their inherent advantages as compared with synchronous motors. Although synchronous motors have certain advantages – such as constant speed, generation of reactive power (leading power factor based on consumer notation of current) with an overexcited field, and low cost for low-speed motors – they have the constraints of requiring a DC exciter, inflexible speed control, and high cost for high-speed motors. The polyphase induction motors, however, have certain inherent advantages, including these:

 Induction motors require no exciter (no electrical connection to the rotor windings except for some doubly excited machines);

 They are rugged and relatively inexpensive;

 They require very little maintenance;

 They have nearly constant speed (slip of only a few percent from no load to full load);

 They are suitable for explosive environments;

 Starting of motors is relatively easy; and

 Operation of three-phase machines within a single-phase system is possible.

The main disadvantages of induction motors are:

 complicated speed control,

 high starting current if no starters are relied on, and

 low and lagging (based on consumer notation of current) power factor at light loads.

For most applications, however, the advantages far outweigh the disadvantages. In today’s power systems with a large number of nonlinear components and loads, three-phase induction machines are usually subjected to nonsinusoidal operating conditions. Conventional steady-state and transient analyses do not consider the impact of voltage and/or current harmonics on three-phase induction machines. Poor power quality has many detrimental effects on induction machines due to their abnormal operation, including:

 excessive voltage and current harmonics,

 excessive saturation of iron cores,

 static and dynamic rotor eccentricities,

 one-sided magnetic pull due to DC currents,

 shaft fluxes and associated bearing currents,

 mechanical vibrations,

 dynamic instability when connected to weak systems,

 premature aging of insulation material caused by cyclic operating modes as experienced by induction machines for example in hydro- and wind-power plants,

 increasing core (hysteresis and eddy-current) losses and possible machine failure due to unacceptably high losses causing excessive hot spots,

 increasing (fundamental and harmonic) copper losses,

 reduction of overall efficiency,

 generation of inter- and subharmonic torques,

 production of (harmonic) resonance and ferroresonance conditions,

 failure of insulation due to high voltage stress caused by rapid changes in supply current and lightning surges,

 unbalanced operation due to an imbalance of power systems voltage caused by harmonics, and

 excessive neutral current for grounded machines.

These detrimental effects call for the understanding of the impact of harmonics on induction machines and how to protect them against poor power quality conditions. A harmonic model of induction motors is essential for loss calculations, harmonic torque calculations, derating analysis, filter design, and harmonic power flow studies.

This chapter investigates the behavior of induction machines as a function of the harmonics of the power system’s voltages/currents, and introduces induction machine harmonic concepts suitable for loss calculations and harmonic power flow analysis. After a brief review of the conventional (sinusoidal) model of the induction machine, time and space harmonics as well as forward- and backward-rotating magnetic fields at harmonic frequencies are analyzed. Some measurement results of voltage, current, and flux density waveforms and their harmonic components are provided. Various types of torques generated in induction machines, including fundamental and harmonic torques as well as inter- and subharmonic torques, are introduced and analyzed. The interaction of space and time harmonics is addressed. A section is dedicated to voltage-stress winding failures of AC machines fed by voltage-source and current-source pulse-width-modulated (PWM) inverters. Some available harmonic models of induction machines are briefly introduced. The remainder of the chapter includes discussions regarding rotor eccentricity, classification of three-phase induction machines, and their operation within a single-phase system. This chapter contains a number of application examples to further clarify the relatively complicated issues.

3.1 Complete sinusoidal equivalent circuit of a three-phase induction machine

Simulation of induction machines under sinusoidal operating conditions is a well-researched subject and many transient and steady-state models are available. The stator and rotor cores of an induction machine are made of ferromagnetic materials with nonlinear (B–H) or (λi) characteristics. As mentioned in Chapter 2, magnetic coils exhibit three types of nonlinearities that complicate their analysis: saturation effects, hysteresis loops, and eddy currents. These phenomena result in nonsinusoidal flux, voltage and current waveforms in the stator and rotor windings, and additional copper (due to current harmonics) and core (due to hysteresis loops and eddy currents) losses at fundamental and harmonic frequencies. Under nonsinusoidal operating conditions, the stator magnetic field will generate harmonic rotating fields that will produce forward- and backward-rotating magnetomotive forces (mmfs); in addition to the time harmonics (produced by the nonsinusoidal stator voltages), space harmonics must be included in the analyses. Another detrimental effect of harmonic voltages at the input terminals of induction machines is the generation of unwanted harmonic torques that are superimposed with the useful fundamental torque, producing vibrations causing a deterioration of the insulation material and rotor copper/aluminum bars. Linear approaches for induction machine modeling neglect these nonlinearities (by assuming linear (λ i) characteristics and sinusoidal current and voltage waveforms) and use constant values for the magnetizing inductance and the core-loss resistance.

Sinusoidal modeling of induction machines is not the main subject of this chapter. However, Fig. 3.1 illustrates a relatively simple and accurate frequency-based linear model that will be extended to a harmonic model in Sections 3.4 to 3.9. In this figure subscript 1 denotes the fundamental frequency (h = 1); ω1 and s1 are the fundamental angular frequency (or velocity) and the fundamental slip of the rotor, respectively. Rfe is the core-loss resistance, Lm is the (linear) magnetizing inductance, Rs, R′r, Lsℓ, and L′rℓ are the stator and the rotor (reflected to the stator) resistances and leakage inductances, respectively [315]. The value of Rfe is usually very large and can be neglected. Thus, the simplified sinusoidal equivalent circuit of an induction machine results, as shown in Fig. 3.2.

f03-01-9780128007822
Figure 3.1 Complete linear per-phase equivalent circuit of a three-phase induction machine for sinusoidal analysis.
f03-02-9780128007822
Figure 3.2 Simplified per-phase equivalent circuit of a three-phase induction machine for sinusoidal analysis.

Replacing the network to the left of line a–b by Thevenin’s theorem (TH) one gets

V~s1TH=|V~s1|ω1LmθTHRs2+ω1Ls+ω1Lm2,

si4_e  (3-1a)

θTH=π2tan1ω1Ls+ω1LmRs,

si5_e  (3-1b)

RsTH=ω1Lm2RsRs2+ω1Ls+ω1Lm2,

si6_e  (3-1c)

XsℓTH=ω1LsℓTH=ω1LmRs2+ω1Lsω1Lmω1Ls+ω1LmRs2+ω1Ls+ω1Lm2.

si7_e  (3-1d)

Using Eq.3-1, the Thevenin adjusted equivalent circuit is shown in Fig. 3.3.

f03-03-9780128007822
Figure 3.3 Thevenin (TH) adjusted equivalent circuit of a three-phase induction machine for fundamental frequency (h = 1).

The fundamental (h = 1) slip is

s1=ns1nmns1=ωs1ωmωs1,

si8_e  (3-2)

where ns1 is the synchronous speed of the stator field, nm is the (mechanical) speed of the rotor, and ωs1, ωm are the corresponding angular velocities (or frequencies), respectively. Note that subscript 1 denotes fundamental frequency.

For a p-pole machine,

ωs1=ω1/p2=2πns1rps=2πns1rpm60,

si9_e  (3-3a)

ωm=2πnmrps=2πnmrpm60,

si10_e  (3-3b)

where ω1 = 2π f1 = 377 rad/s is the fundamental angular velocity (frequency).

Forp = 2ns1rpm=3600rpmatω1=2πf1=377rad/ssi11_e
p = 4ns1rpm=1800rpmsi12_e
p = 6ns1rpm=1200rpmsi13_e
p = 8ns1rpm=900rpmsi14_e

The torque of a three-phase (q1 = 3) induction machine is (if Vs1TH=|V~s1TH|si15_e)

T=1ωs1q1Vs1TH2Rrs1RsTH+Rrs12+ω1LsℓTH+ω1Lr2

si16_e  (3-4a)

or

T=1ωs1q1Vs1TH2Rrs1RsTHs1+Rr2s12+ω1LsℓTH+ω1Lr2

si17_e  (3-4b)

or

T=1ωs1q1Vs1TH2RrRsTHs1+Rr2s1+ω1LsℓTH+ω1Lr2s1

si18_e  (3-4c)

or

T=1ωs1q1Vs1TH2Rrs1RsTHs1+Rr2+ω1LsℓTH+ω1Lr2s12.

si19_e  (3-4d)

For small values of s1 we have

T1ωs1q1Vs1TH2s1Rr,s1=small.

si20_e  (3-5)

This relation is called the small-slip approximation for the torque.

Based on Eq. 3-4, the fundamental torque-speed characteristic of a three-phase induction machine is shown in Fig. 3.4.

f03-04-9780128007822
Figure 3.4 Fundamental electrical torque-speed characteristics for a three-phase induction machine.

Mechanical output power (neglecting iron-core losses, friction, and windage losses) is

Pout=Tωm.

si21_e  (3-6)

The air-gap power is (Fig. 3.5)

Pgap=q1I~r12Rrs1,

si22_e  (3-7)

where |I~r1|=|I~s1TH|.si23_e

f03-05-9780128007822
Figure 3.5 Definition of air-gap power.

The loss distribution within the machine is given in Fig. 3.6. Note that

Pout=PgapPrloss=PinPslossPrloss

si24_e

and

Pout=Tωm.

si25_e

f03-06-9780128007822
Figure 3.6 Loss breakdown within induction machine neglecting frictional and windage losses.

The complete fundamental torque-speed relation detailing braking, motoring, and generation regions is shown in Fig. 3.7, where ns1 denotes the synchronous (mechanical) speed of the fundamental rotating magnetic field.

f03-07-9780128007822
Figure 3.7 Complete fundamental torque-speed relation detailing braking, motoring, and generation.

3.1.1 Application Example 3.1: Steady-State Operation of Induction Motor at Undervoltage

A Pout = 100 hp, VL-L_rat = 480 V, f = 60 Hz, p = 6 pole, three-phase induction motor runs at full load and rated voltage with a slip of 3%. Under conditions of stress on the power system, the line-to-line voltage drops to VL-L_low = 430 V. If the load is of the constant torque type, compute for the lower voltage:

a) Slip slow (use small-slip approximation).

b) Shaft speed nm_low in rpm.

c) Output power Pout_low.

d) Rotor copper loss Pcur_low = (Ir′)2Rr′ in terms of the rated rotor copper loss at rated voltage.

Solution to Application Example 3.1

a) Based on the small-slip approximation s1V2si26_e, the new slip at the low voltage is

slow=0.03480/4302=0.0374.

si27_e

b) With the synchronous speed ns=120fp=120606=1200rpmsi28_e one calculates the shaft speed at low voltage as

nm_low=1slowns=10.03741200=1155.14rpm.

si29_e

c) The output power at low voltage is Pout = ωm · Tm. For constant mechanical torque one gets Pout_low=10.037410.03100hp=99.23hpsi30_e

d) Due to the relation T=Pgωs1si31_e (where Pg is the air-gap power), the torque is proportional to Pg, and therefore, for constant torque operation the air-gap power is constant. Thus the rotor loss is Pcur_low=slowPg=0.03740.03Pout_rat=1.247Pout_ratsi32_e

One concludes that a decrease of the terminal voltage increases the rotor loss (temperature) of an induction motor.

3.1.2 Application Example 3.2: Steady-State Operation of Induction Motor at Overvoltage

Repeat Application Example 3.1 for the condition when the line-to-line voltage of the power supply system increases to VL-L_high = 510 V.

Solution to Application Example 3.2

a) Based on the small-slip approximation s1V2si33_e the new slip at the high voltage is

shigh=0.03480/5102=0.02657.

si34_e

b) With the synchronous speed ns=120fp=120606=1200rpmsi35_e one calculates the shaft speed at high voltage as nm_high = (1 – shigh)n = (1 – 0.02657)(1200) = 1168.11 rpm.

c) The output power at high voltage is Pout = ωm · Tm. For constant mechanical torque one gets Pout_high=10.0265710.03100hp=100.35hp.si36_e

d) Due to the relation T=Pgωs1si37_e, where Pg is the air-gap power, the torque is proportional to Pg, and therefore, for constant torque operation the air-gap power is constant. Thus the rotor loss is Pcur_high = shighPg = 0.026570.03si38_e Pcur_rat = 0.886 Pcur_rat.

One concludes that an increase of the terminal voltage decreases the rotor loss (temperature) of an induction motor.

3.1.3 Application Example 3.3: Steady-State Operation of Induction Motor at Undervoltage and Under-Frequency

Repeat Application Example 3.1 for the condition when the line-to-line voltage of the power supply system and the frequency drop to VL-L_low = 430 V and flow = 59 Hz, respectively.

Solution to Application Example 3.3

The small-slip approximation results in the torque equation T=q1Vs2sωs1Rr.si39_e

a) Applying the relation to rated conditions (1) one gets T1=q1Vs21s1ωs11Rr.si40_e
Applying the relation to low-voltage and low-frequency conditions (2) one gets T2=q1Vs22s2ωs12Rr.si41_e
With T(1) = T(2) and the relations ωs11=2πf1p/2andωs12=2πf2p/2si42_e
one obtains s2=s1Vs12Vs22f2f1=0.0367.si43_e

b) The synchronous speed ns(2) at f(2) = 59 Hz is ns2=120596=1180rpmsi44_e, or the corresponding shaft speed nm2=ns21s2=1137rpm.si45_e

c) The torque-output power relations are T1=Pout1ωm1andT2=Pout2ωm2.si46_e
With T(1) = T(2) one gets Pout1ωm1=Pout2ωm2si47_e or Pout1ωs111s1=Pout2ωs121s2si48_e or Pout1120f1p1s1=Pout2120f2p1s2.si49_e For the reduced output power it follows Pout2=1s21s1f2f1Pout1=97.65hp.si50_e

d) The copper losses of the rotor are Pcur=q1Ir2Rrsi51_e and the output power is Pout=q1Ir2Rr1ss.si52_e The application of the relations to the two conditions (1) and (2) yields

Pout1=q1Ir(1)2Rr1s1s1andPout2=q1Ir(2)2Rr1s2s2orIr(2)2Ir(1)2=Pout2Pout1s2s11s11s2=1.20.

si53_e

The rotor loss increase becomes now Pcur2Pcur1=Ir(2)2Ir(1)2=1.20.si54_e

One concludes that the simultaneous reduction of voltage and frequency increases the rotor losses (temperature) in a similar manner as discussed in Application Example 3.1. It is advisable not to lower the power system voltage beyond the levels as specified in standards.

3.2 Magnetic fields of three-phase machines for the calculation of inductive machine parameters

Numerical approaches such as the finite-difference and finite-element methods [16] enable engineers to compute no-load and full-load magnetic fields and those associated with short-circuit and starting conditions, as well as fields for the calculation of stator and rotor inductances/reactances. Figures 3.8 and 3.9 represent the no-load fields of four- and six-pole induction machines [17,18]. Figure 3.10a–e illustrates radial forces generated as a function of the rotor position. Such forces cause audible noise and vibrations. The calculation of radial and tangential magnetic forces is discussed in Chapter 4 (Section 4.2.14), where the concept of the “Maxwell stress” is employed. Figures 3.11 to 3.13 represent unsaturated stator and rotor leakage fields and the associated field during starting of a two-pole induction motor. Figures 3.14 and 3.15 represent saturated stator and rotor leakage fields, respectively, and Fig. 3.16 depicts the associated field during starting of a two-pole induction machine. The starting current and starting torque as a function of the terminal voltage are shown in Fig. 3.17 [19]. This plot illustrates how saturation influences the starting of an induction motor. Note that the linear (hand) calculation results in lower starting current and torque than the numerical solution.

f03-08-9780128007822
Figure 3.8 No-load field of four-pole, 800 W, three-phase induction motor. One flux tube contains a flux per unit length of 0.0005 Wb/m.
f03-09-9780128007822
Figure 3.9 No-load field of six-pole, 75 kW, three-phase induction motor. One flux tube contains a flux per unit length of 0.0006 Wb/m.
f03-10ac-9780128007822f03-10de-9780128007822
Figure 3.10 Flux distribution and radial stator core forces at no load and rated voltage for (a) rotor position #1, (b) rotor position #2, (c) rotor position #3, (d) rotor position #4, and (e) rotor position #5.
f03-11-9780128007822
Figure 3.11 Field for the determination of the unsaturated stator leakage flux of a 3.4 MW, two-pole, three-phase induction motor. One flux tube contains a flux per unit length of 0.005 Wb/m [19].
f03-12-9780128007822
Figure 3.12 Field for the determination of the unsaturated rotor leakage flux of a 3.4 MW, two-pole, three-phase induction motor. One flux tube contains a flux per unit length of 0.015 Wb/m [19].
f03-13-9780128007822
Figure 3.13 Field distribution (first approximation) at starting with rated voltage of a 3.4 MW, two-pole, three-phase induction motor. One flux tube contains a flux per unit length of 0.005 Wb/m [19].
f03-14-9780128007822
Figure 3.14 Field for the determination of the saturated stator leakage flux of a 3.4 MW, two-pole, three-phase induction motor. One flux tube contains a flux per unit length of 0.15 Wb/m [19].
f03-15-9780128007822
Figure 3.15 Field for the determination of the saturated rotor leakage flux of a 3.4 MW, two-pole, three-phase induction motor. One flux tube contains a flux per unit length of 0.005 Wb/m [19].
f03-16-9780128007822
Figure 3.16 Field distribution (second approximation) during starting with rated voltage of a 3.4 MW, two-pole, three-phase induction motor. One flux tube contains a flux per unit length of 0.005 Wb/m [19].
f03-17-9780128007822
Figure 3.17 Starting currents and torques as a function of terminal voltage for a 3.4 MW, three-phase, induction motor [19].

Any rotating machine design is based on iterations. No closed form solution exists because of the nonlinearities (e.g., iron-core saturation) involved. In Fig. 3.13 the field for the first approximation, where saturation is neglected and a linear B–H characteristic is assumed, permits us to calculate stator and rotor currents for which the starting field can be computed under saturated conditions assuming a nonlinear (B–H) characteristic as depicted in Fig. 3.16. For the reluctivity distribution caused by the saturated short-circuit field the stator (Fig. 3.14) and rotor (Fig. 3.15) leakage reactances can be recomputed, leading to the second approximation as indicated in Fig. 3.16. In practice a few iterations are sufficient to achieve a satisfactory solution for the starting torque as a function of the applied voltage as illustrated in Fig. 3.17. It is well known that during starting saturation occurs only in the stator and rotor teeth and this is the reason why Figs. 3.13 and 3.16 are similar.

3.3 Steady-state stability of a three-phase induction machine

The steady-state stability criterion is [20]

TLωm>Tωm,

si55_e  (3-8)

where TL is the load torque and T is the motor torque. Using this equation, the stable steady-state operating point of a machine for any given load-torque characteristic can be determined.

3.3.1 Application Example 3.4: Unstable and Stable Steady-State Operation of Induction Machines

Figure E3.4.1 shows the torque–angular velocity characteristic of an induction motor (Tωm) supplying a mechanical constant torque load (TL). Determine the stability of the operating points Q1 and Q2.

f03-45-9780128007822
Figure E3.4.1 Steady-state stability of induction machine with constant load torque.

Solution to Application Example 3.4

The stability conditions at the two points where the two curves intersect are

Q1:ΔTLΔωm=0ΔTΔωm>positive.0<positive>>>>>>unstable.si56_e
Q2:ΔTLΔωm=0ΔTΔωm=negative.0>negative>>>>>>stable.si57_e

Therefore, operation at Q1 is unstable and will result in rapid speed reduction. This will stop the operation of the induction machine, while operating point Q2 is stable.

3.3.2 Application Example 3.5: Stable Steady-State Operation of Induction Machines

Figure E3.5.1 shows the torque–angular velocity relation (T – ωm) of an induction motor supplying a mechanical load with nonlinear (parabolic) torque (TL). Determine the stability of the operating point Q1.

f03-46-9780128007822
Figure E3.5.1 Steady-state stability of induction machine with nonlinear load torque.

Solution to Application Example 3.5

The stability conditions at the point where the two curves intersect is

Q1:ΔTLΔωm=positiveΔTΔωm=negative.>>>>>>stable.si58_e

Therefore, an induction machine connected to a mechanical load with parabolic torque-speed characteristics as shown in Fig. E3.5.1 will not experience any steady-state stability problems.

3.3.3 Resolving Mismatch of Wind-Turbine and Variable-Speed Generator Torque–Speed Characteristics

It is well known that the torque–speed characteristic of wind turbines (Fig. 3.18) and commonly available variable-speed drives employing field weakening (Fig. 3.19a) do not match because the torque of a wind turbine is proportional to the square of the speed and that of a variable-speed drive is inversely proportional to the speed in the field-weakening region. One possibility to mitigate this mismatch is an electronic change in the number of poles p or a change of the number of series turns of the stator winding [2123]. The change of the number of poles is governed by the relation

Epf=4.44BmaxNrat4RL,

si59_e

where Bmax is the maximum flux density at the radius R of the machine, Nrat is the rated number of series turns of a stator phase winding, and L is the axial iron-core length of the machine.

f03-18-9780128007822
Figure 3.18 Torque–speed characteristic of a wind turbine T si1_e nm2.
f03-19-9780128007822
Figure 3.19 (a) Variable-speed drive torque–speed characteristic without winding switching T si2_e1/nm in field-weakening region. The full line corresponds to the torque T and the dashed-dotted line to the power P. (b) Variable-speed drive torque–speed characteristic by changing the number of series turns of the stator winding from Nrat to Nrat/2. The full line corresponds to the torque T and the dashed-dotted line to the power P. (c) Measured variable-speed drive response with winding switching from pole number p1 to p2 resulting in a speed range from 600 to 4000 rpm [21]. One horizontal division corresponds to 200 ms. One vertical division corresponds to 800 rpm.

The change of the number of turns N is governed by

EfN=4.44Bmax4RL/p.

si60_e

The introduction of an additional degree of freedom – either by a change of p or N – permits an extension of the constant flux region. Adding this degree of freedom will permit wind turbines to operate under stalled conditions at all speeds, generating maximum power at a given speed with no danger of a runaway. This will make the blade-pitch control obsolete – however, a furling of the blades at excessive wind velocities must be provided – and wind turbines become more reliable and less expensive due to the absence of mechanical control. Figure 3.19b illustrates the speed and torque increase due to the change of the number of turns from Nrat to Nrat/2, and Fig. 3.19c demonstrates the excellent dynamic performance of winding switching with solid-state switches from number of poles p1 to p2 [21]. The winding reconfiguration occurs near the horizontal axis (small slip) at 2400 rpm.

3.4 Spatial (space) harmonics of a three-phase induction machine

Due to imperfect (e.g., nonsinusoidal) winding distributions and due to slots and teeth in stator and rotor, the magnetomotive forces (mmfs) of an induction machine are nonsinusoidal.

For sinusoidal distribution one obtains the mmfs

Fa=A1cosθia,Fb=A1cosθ120°ib,Fc=A1cosθ240°ic.

si61_e  (3-9)

Note that cos(θ – 240°) = cos(θ + 120°). For nonsinusoidal mmfs (consisting of fundamental, third, and fifth harmonics) one obtains

Fa=A1cosθ+A3cos3θ+A5cos5θiat,Fb={A1cosθ120°+A3cos3θ120°+A5cos5θ120°}ibt,Fc={A1cosθ240°+A3cos3θ240°+A5cos5θ240°}ict,

si62_e  (3-10)

where

iat=Imcosω1t,ibt=Imcosω1t120°,ict=Imcosω1t240°.

si63_e  (3-11)

The magnetomotive force (mmf) originates in the phase belts a–a’, b–b’, and c–c’ as shown in Fig. 3.20.

f03-20-9780128007822
Figure 3.20 Definition of phase belts and their associated axes in an induction machine.

The total mmf is

Ftotθt=Fa+Fb+Fc,

si64_e  (3-12)

where

Fa={A1cosθ+A3cos3θ+A5cos5θ}Imcosω1t,

si65_e  (3-13a)

Fa=A1Imcosθcosω1t+A3Imcos3θcosω1t+A5Imcos5θcosω1t.

si66_e

Note that cos x cos y = ½{cos(xy) + cos(x + y)}; therefore,

Fa=A1Im2cosθω1t+cosθ+ω1t+A3Im2cos3θω1t+cos3θ+ω1t+A5Im2cos5θω1t+cos5θ+ω1t.

si67_e  (3-13b)

Correspondingly,

Fb=A1Im2cosθω1t+cosθ+ω1t240°+A3Im2{cos3θω1t240°+cos(3θ+ω1t480°)}+A5Im2{cos5θω1t480°+cos5θ+ω1t720°}.

si68_e  (3-13c)

Fc=A1Im2cosθω1t+cosθ+ω1t+240°+A3Im2{cos3θω1t+240°+cos(3θ+ω1t+480°)}+A5Im2{cos5θω1t+480°+cos5θ+ω1t+720°}.

si69_e  (3-13d)

Therefore, with Eq. 3-12 the total mmf is simplified:

Ftot=32A1Imcosω1tθ+3A52Imcosω1t+5θ.

si70_e  (3-14)

The angular velocity of the fundamental mmf dθ1dtsi71_e is

ω1tθ1=0,θ1=ω1t,dθ1dt=ω1,

si72_e

and the angular velocity of the 5th space harmonic mmf dθ5dtsi73_e is

ω1t+5θ5=0,

si74_e

5θ5=ω1t,

si75_e

dθ5dt=ω15.

si76_e

Graphical representation of spatial harmonics (in the presence of fundamental, third, and fifth harmonics mmfs) with phasors is shown in Fig. 3.21. Note that the 3rd harmonic mmf cancels (is equal to zero).

f03-21-9780128007822
Figure 3.21 Fundamental and 5th spatial harmonic of an induction machine rotating in opposite directions (see Eq. 3-14). Note: F1 is rotating in forward (+) or counterclockwise direction and F5 is rotating in backward (–) or clockwise direction.

Similar analysis is performed to determine the rotating directions of each individual space harmonic, and the positive-, negative-, and zero-sequence harmonic orders can be defined, as listed in Table 3.1.

Table 3.1

Positive-, Negative-, and Zero-Sequence Spatial Harmonics

Spatial-harmonic sequence+0
Spatial-harmonic order123
456
789
101112
131415
...

t0025

Note that:

Even and triplen harmonics are normally not present in a balanced three-phase system!

In general one can write for spatial harmonics

dθhdt=+0ω1h,

si77_e  (3-15)

where +, –, and 0 are used for positive-, negative-, and zero-sequence space harmonic orders, respectively. Figure 3.22 illustrates the superposition of fundamental mmf with 5th space harmonic resulting in amplitude modulation where fundamental of 60 Hz is modulated with 12 Hz.

f03-22-9780128007822
Figure 3.22 Superposition of fundamental mmf with 5th space harmonic resulting in amplitude modulation of fundamental of 60 Hz with 12 Hz.

3.5 Time harmonics of a three-phase induction machine

A three-phase induction machine is excited by balanced three-phase f1 = 60 Hz currents containing a fifth time harmonic. The equations of the currents are for ω1 = 2πf1:

iat=Im1cosω1t+Im5cos5ω1t,ibt={Im1cosω1t120°+Im5cos5ω1t120°},ict=Im1cosω1t+120°+Im5cos5ω1t+120°.

si78_e  (3-16)

Note that the three-phase system rotates in a clockwise (cw) manner; that is, in a mathematically negative sense. Assume that the winding has been designed to eliminate all spatial harmonics. Thus for phase a the mmf becomes

Fa=A1cosθiat.

si79_e  (3-17)

Correspondingly,

Fa=A1cosθIm1cosω1t+Im5cos5ω1t,Fb=A1cosθ120°{Im1cosω1t120°+Im5cos5ω1t120°},Fc=A1cosθ+120°{Im1cosω1t+120°+Im5cos5ω1t+120°}.

si80_e  (3-18)

The total mmf is

Ftot=Fa+Fb+Fc,

si81_e  (3-19)

expanded:

Fa=A1Im12cosθω1t+A1Im12cosθ+ω1t+A1Im52cosθ5ω1t+A1Im52cosθ+5ω1t,Fb=A1Im12cosθω1t+A1Im12cosθ+ω1t240°+A1Im52cosθ5ω1t+A1Im52cosθ+5ω1t240°,Fc=A1Im12cosθω1t+A1Im12cosθ+ω1t+240°+A1Im52cosθ5ω1t+A1Im52cosθ+5ω1t+240°,

si82_e  (3-20)

or

Ftot=3A1Im12cosθω1t+3A1Im52cosθ5ω1t.

si83_e  (3-21)

The angular velocity of the fundamental dθ1dtsi84_e is

θ1ω1t=0,dθ1dt=ω1,

si85_e

and the angular velocity of the fifth time harmonic dθ5dtsi87_e is

θ55ω1t=0,dθ5dt=5ω1.

si88_e

For the current system (consisting of the fundamental and 5th harmonic components where the 5th harmonic system rotates in counterclockwise direction)

iat=Im1cosω1t+Im5cos5ω1t,ibt=Im1cosω1t120°+Im5cos5ω1t+120°,ict=Im1cosω1t+120°+Im5cos5ω1t120°,

si90_e  (3-22)

the fifth time harmonic has the angular velocity

dθ5dt=5ω1.

si91_e

Phasor representation for forward rotating fundamental and forward rotating 5th harmonic current systems is given in Fig. 3.23. Phasor representation for forward rotating fundamental and backward rotating 5th harmonic current systems is shown in Fig. 3.24.

f03-23-9780128007822
Figure 3.23 Forward (+) rotating fundamental mmf of an induction machine superposed with forward (+) rotating 5th time harmonic rotating in the same direction.
f03-24-9780128007822
Figure 3.24 Forward (+) rotating fundamental mmf of an induction machine superposed with backward (–) rotating 5th time harmonic rotating in the opposite direction.

Similar analysis can be performed to determine the rotating directions of each individual time harmonic and to define the positive-, negative-, and zero-sequence harmonic orders, as listed in Table 3.2. As with the space harmonics, even and triplen harmonics are normally not present, provided the system is balanced.

Table 3.2

Positive-, Negative-, and Zero-Sequence Time Harmonics

Time-harmonic sequence+0
Time-harmonic order123
456
789
101112
131415
entityentityentity

t0030

In general one can write for time harmonics

dθhdt=+0hω1,

si92_e  (3-23)

where +, –, and 0 are used for positive-, negative-, and zero-sequence time harmonic orders, respectively.

Therefore, time harmonics voltages have an important impact on induction machines. Forward- and backward-rotating fields are produced by positive-, negative-, and zero-sequence harmonics that produce harmonic shaft torques.

3.6 Fundamental and harmonic torques of an induction machine

Starting with the Thevenin-adjusted circuit as illustrated in Fig. 3.25, one obtains the current of the hth harmonic

I~shTH=V~shTHRsTH+Rrsh+jhω1LsℓTH+Lr.

si93_e  (3-24)

f03-25-9780128007822
Figure 3.25 Thevenin (TH) adjusted equivalent circuit of an induction machine for the hth harmonic. It is assumed that the resistive and inductive circuit parameters are independent of h.

Therefore, the electrical torque for the hth harmonic is

Teh=1ωshq1VshTH2RrshRsTH+Rrsh2+hω12LsℓTH+Lr2.

si94_e  (3-25)

Similarly, the fundamental (h = 1) torque is

Te1=1ωs1q1Vs1TH2Rrs1RsTH+Rrs12+ω12LsℓTH+Lr2.

si95_e  (3-26)

3.6.1 The Fundamental Slip of an Induction Machine

The fundamental slip is

s1=ωs1ωmωs1

si96_e  (3-27)

or

s1=1ωmωs1ωmωs1=1s1

si97_e

where ωs1=ω1p/2si98_e is the (mechanical) synchronous fundamental angular velocity and ωm is the mechanical angular shaft velocity.

The fundamental torque referred to fundamental slip s1 is shown in Fig. 3.26 where Te1 is the machine torque and TL is the load torque. If Rs = 0 then Te1 is symmetric to the point at (Te1 = 0/s1 = 0) or (Te1 = 0/ωs1).

f03-26-9780128007822
Figure 3.26 Fundamental induction motor (Te1) and load (TL) torques as a function of angular velocity and slip s1.

3.6.2 The Harmonic Slip of an Induction Machine

The harmonic slip (without addressing the direction of rotation of the harmonic field) is defined as

sh=hωs1ωmhωs1,

si99_e

where ωs1 = (ω1)/(p/2) and ω1 is the electrical angular velocity, ω1 = 2πf1 and f1 = 60 Hz.

To include the direction of rotation of harmonic mmfs, in the following we assume that the fundamental rotates in forward direction, the 5th in backward direction, and the 7th in forward direction (see Table 3.2 for positive-, negative-, and zero-sequence components).

For motor operation ωm < ωs1; thus for the 5th harmonic component, one obtains

s5=5ωs1ωm5ωs1,

si100_e

where (–5ωs1) means rotation in backward direction, or

s5=5ωs1+ωm5ωs1.

si101_e

Note that in this equation 5ωs1 is the base (reference) angular velocity.

Correspondingly, one obtains for the forward rotating 7th harmonic:

s7=7ωs1ωm7ωs1,

si102_e

where 7ωs1 is the base.

Therefore, the harmonic slip is

sh=hωs1±ωmhωs1,+:forforwardrotatingfield:forbackwardrotatingfield.

si103_e  (3-28)

Harmonic torques Te7 and Te5 (as a function of the angular velocity ωm and the harmonic slips s7 and s5) are depicted in Figs. 3.27 and 3.28, respectively.

f03-27-9780128007822
Figure 3.27 Induction motor torque for 7th time harmonic as a function of angular velocity and slip s7.
f03-28-9780128007822
Figure 3.28 Induction motor torque for 5th time harmonic as a function of angular velocity and slip s5.

3.6.3 The Reflected Harmonic Slip of an Induction Machine

If ωs1 is taken as the reference angular velocity, then for the backward rotating fifth harmonic one obtains

s51=5ωs1+ωmωs1,

si104_e  (3-29)

where s5(1) is the reflected fifth harmonic slip.

Correspondingly, if one takes ωs1 as reference then for the forward rotating seventh harmonic one gets

s71=7ωs1ωmωs1,

si105_e

where s7(1) is the reflected seventh harmonic slip.

In general, the reflected harmonic slip is

sh1=hωs1±ωmωs1,+:forforwardrotatingfield:forbackwardrotatingfield.

si106_e  (3-30)

To understand the combined effects of fundamental and harmonic torques, it is more convenient to plot harmonic torques (Teh) as a function of the reflected harmonic slips (sh(1)):

 The harmonic torque Te7 reflected to harmonic slip s71=7ωs1ωmωs1si107_e is depicted in Fig. 3.29; and

f03-29-9780128007822
Figure 3.29 Induction motor torque for 7th time harmonic as a function of mechanical angular velocity and reflected harmonic slip s7(1).

 The harmonic torque Te5 reflected to harmonic slip s51=5ωs1ωmωs1si108_e is shown in Fig. 3.30.

f03-30-9780128007822
Figure 3.30 Induction motor torque for 5th time harmonic as a function of mechanical angular velocity and reflected harmonic slip s5(1).

3.6.4 Reflected Harmonic Slip of an Induction Machine in Terms of Fundamental Slip

The relation between s1 (fundamental slip) and sh(1) (reflected harmonic slip) is

sh1=hωs1ωmωs1=ωs1ωmωs1+h1ωs1ωs1=1ωmωs1+h1=h1s1.

si109_e

Therefore, the reflected harmonic slip, in terms of fundamental slip, is

sh1=s1+h1.

si110_e  (3-31a)

For the backward rotating 5th harmonic one gets

s51=s1+6thuss1=s51+6,

si111_e

and for the forward rotating 7th harmonic

s71=s1+6thuss1=s716.

si112_e

Note that

s51=5ωs1+ωmωs1.

si113_e

In general, the reflected harmonic slip sh(1) for the forward and the backward rotating harmonics is a linear function of the fundamental slip s1:

sh1=s1+h1forforwardrotatingfieldsh1=s1+h+1forbackwardrotatingfield.

si114_e  (3-31b)

Superposition of fundamental (h = 1), fifth harmonic (h = 5), and seventh harmonic (h = 7) torque Te = Te1 + Te5 + Te7 is illustrated in Fig. 3.31. At ωm = ωmrated the total electrical torque Te is identical to the load torque TL or Te1 + Te5 + Te7 = TL, where Te1 is a motoring torque, Te7 is a motoring torque, and Te5 is a braking torque. Figure 3.32 shows the graphical phasor representation of superposition of torques as given in Fig. 3.31.

f03-31-9780128007822
Figure 3.31 Induction motor torque for fundamental, 5th, and 7th time harmonics as a function of angular velocity and slip s1.
f03-32-9780128007822
Figure 3.32 Graphical phasor representation of superposition of torques as given in Fig. 3.31.

3.6.5 Reflected Harmonic Slip of an Induction Machine in Terms of Harmonic Slip

In this section, the relation between the reflected harmonic slip sh(1) and the harmonic slip sh is determined as sh1=fsh.si115_e

The harmonic slip (without addressing the direction of rotation of the harmonic field) is

sh=hωs1ωmhωs1=h1ωs1hωs1+ωs1ωmhωs1

si116_e

with

s1=ωs1ωmωs1.

si117_e

It follows that

sh=h1ωs1hωs1+s1h.

si118_e

With sh1=s1+h1si119_e, one obtains for forward rotating harmonics

s1=sh1h1

si120_e

or

sh=h1ωs1hωs1+sh1h1h=h1h+sh1hh1h=sh1h.

si121_e

Similar analysis can be used for backward-rotating harmonics. Therefore, sh(1) is a linear function of sh:

sh1=hshforforwardandbackwardrotatingfields.

si122_e  (3-32)

3.7 Measurement results for three- and single-phase induction machines

Measurement techniques for the determination of the (copper and iron-core) losses of induction machines are described in [24]. The losses of harmonics of single-phase induction motors are measured in [25]. Typical input current wave shapes for single-and three-phase induction motors fed by thyristor converters are depicted in Figs. 3.33 and 3.34 [26]. In addition to semiconductor converters, harmonics can be generated by the DC magnetization of transformers [27]. A summary of the heating effects due to harmonics on induction machines is given in [28]. The generation of sub- and interharmonic torques by induction machines is presented in [29]. Power quality of machines relates to harmonics of the currents and voltages as well as to those of flux densities. The latter are more difficult to measure because sensors (e.g., search coils, Hall devices) must be implanted in the core. Nevertheless, flux density harmonics are important from an acoustic noise [30] and vibration [3133] point of view. Excessive vibrations may lead to a reduction of lifetime of the machine. Single-phase induction machines are one of the more complicated electrical apparatus due to the existence of forward- and backward-rotating fields. The methods as applied to these machines are valid for three-phase machines as well, where only forward fundamental rotating fields exist.

f03-33-9780128007822
Figure 3.33 Waveforms of input and output voltages and currents of a thyristor/triac controller feeding a 2 hp single-phase induction motor at no load. v1, i1 and v2, i2 are input voltage/current and output voltage/current of controller, respectively [26].
f03-34-9780128007822
Figure 3.34 Waveforms of input voltages vac, vbc and input currents ia , ib of 7.5 hp three-phase induction motor at no load fed by a thyristor/triac controller [26].

3.7.1 Measurement of Nonlinear Circuit Parameters of Single-Phase Induction Motors

The optimization of single- and three-phase induction machines is based on sinusoidal quantities neglecting the influence of spatial and time harmonics in flux densities, voltages, and currents [14,15,34,35]. At nonsinusoidal terminal voltages, the harmonic losses of single-phase induction motors impact their efficiency [25]. The computer-aided testing of [24] will be relied on to measure nonlinear circuit parameters and flux densities in stator teeth and yokes as well as in the air gap of permanent-split-capacitor (PSC) motors. Detailed test results are presented in [36]. One concludes that tooth flux density wave shapes can be nearly sinusoidal although the exciting or magnetizing currents are nonsinusoidal.

A computer-aided testing (CAT) circuit and the computer-aided testing of electrical apparatus program (CATEA) are used in [24] to measure nonlinear parameters. The CAT circuit is shown in Fig. 3.35, where four-channel signals vm, im, va, and ia corresponding to main (m) and auxiliary (a) phase voltages and currents, respectively, are sampled by a personal computer via a 12-bit A/D converter.

f03-35-9780128007822
Figure 3.35 CAT circuit for measuring of nonlinear circuit parameters of single-phase induction machines.

Three (see Table 3.3) fractional horsepower PSC single-phase induction motors serving as prime movers for window air conditioners are tested and their parameters listed.

Table 3.3

Three Fractional Horsepower PSC Single-Phase Induction Motors Serving as Prime Movers for Window Air Conditioners

Single-phase motor R333MCCapacitor: 15 μF, 370 V, output power: 1/2 hp
Single-phase motor Y673MG1Capacitor: 17.5 μF, 366 V, output power: 3/4 hp
Single-phase motor 80664346Capacitor: 25 μF, 370 V, output power: 1/2 hp

All motors to be tested are operated at minimum load, where input and output lines at the compressor are open to air. Measurements are conducted at an ambient temperature of 22°C. The linear stator leakage impedance of a motor is measured by removing the rotor from the stator bore and applying various terminal voltage amplitudes to the stator windings. When the main-phase parameters are measured, the switch leading to the auxiliary phase must be open, and vice versa. From the sampled data one obtains the fundamental voltages and currents of the main- and auxiliary-phase windings through Fourier analysis.

Stator Impedance

The stator input resistance and reactance are derived via linear regression. The stator leakage reactance is obtained by subtracting the reactance corresponding to leakage flux inside the bore of the motor from its input reactance [25].

Rotor Impedance

The rotor leakage impedance is frequency dependent and nonlinear because of the closed-rotor slots. The rotor impedance, referred to the main-phase winding, is measured by passing varying amplitudes of current – from about 5 to 150% of rated current – through the main-phase winding under locked-rotor conditions. The total input resistance is derived from linear regression and the rotor resistance at rated frequency is obtained by subtracting the main-phase stator winding resistance from the total input resistance. The rotor resistance, taking the skin effect into account, can be obtained from [25]:

Rr=φξRr0,

si123_e  (3-33)

φξ=ξsinhξ+sin2ξcosh2ξcos2ξ,

si124_e  (3-34a)

where

ξ=krf.

si125_e  (3-34b)

In Eq. 3-34b, f is the frequency of the rotor current. kr is the skin-effect coefficient derived from the height of the rotor slot hr and the resistivity of the bar ρr:

kr=hrπμ0ρr.

si126_e  (3-34c)

In Eq. 3-33, Rr0 is the DC rotor resistance that can be derived from the rotor AC resistance at rated frequency and its skin-effect coefficient. The nonlinear rotor leakage reactance as a function of the amplitude of the rotor current – which can be described by a (λi) characteristic – is obtained by subtracting the stator main-phase winding leakage reactance from the total input reactance. A linear rotor leakage reactance – which can be used as an approximation of the nonlinear rotor leakage reactance – is also derived from a linear regression.

Magnetizing Impedance

The magnetizing (λi) characteristic referred to the stator main-phase winding is measured by supplying varying amplitudes of voltage to the main-phase winding at no-load operation. The forward induced voltage is determined by subtracting from the input voltage the voltage drop across the stator impedance and the backward induced voltage (the voltage drop across the rotor backward equivalent impedance); note that in [25, Fig. 7a], the main-phase forward circuit, and [25, Fig. 7c], the main-phase backward circuit, are connected in series, whereby h = 1. Thereafter, the magnetizing impedance is derived from the ratio of the forward induced voltage and the input current in the frequency (fundamental) domain. The magnetizing reactance as a function of the amplitude of the forward induced voltage is derived from the imaginary part of the (complex) magnetizing impedance.

The (λi) characteristics are given in the frequency domain, which means that both λ (flux linkages) and i (current) have fundamental amplitudes only. The fundamental amplitudes of either flux linkages or currents can be derived from each other via such characteristics.

Iron-Core Resistance

The resistance corresponding to the iron-core loss is determined as follows. The iron-core and the mechanical frictional losses are computed by subtracting from the input power at no load

1. the copper loss of the main-phase winding at no load,

2. the backward rotor loss at no load, and

3. the forward air-gap power responsible for balancing the backward rotor loss.

Thereafter, the iron-core and the mechanical frictional losses, the total value of which is a linear function of the square of the forward induced voltage, are separated by linear regression. The equivalent resistance corresponding to the iron-core loss is derived from the slope of the regression line.

Turns Ratio between the Turns of the Main- and Auxiliary-Phase Windings

The ratio between the turns of the main- and auxiliary-phase windings |k~am|si127_e is also an important parameter for single-phase induction motors. The main- and auxiliary-phase windings of some motors are not placed in spatial quadrature; therefore, an additional parameter θ, the spatial angle between the axes of the two stator windings, must be determined. These two parameters |k~am|,θsi128_e are found by measuring the main-phase induced voltage and current as well as the auxiliary-phase induced voltage when the motor is operated at no load. As has been mentioned above, the main-phase forward (f) and backward (b) induced voltages in the frequency domain, f and b, respectively, can be obtained from the input voltage and current. If

k~am=|k~am|ejθ

si129_e  (3-35)

is used to represent this complex turns ratio, then the auxiliary-phase induced voltage is given by

E~a=k~amE~f+k~am*E~b.

si130_e  (3-36)

Therefore, the complex ratio |k~am|si131_e can be iteratively computed from

k~am=E~ak~am*E~b/E~f,

si132_e  (3-37)

where k~am*si133_e is the complex conjugate of k~amsi134_e. Measured parameters of PSC motors are given in Section 3.7.2.

3.7.1.1 Measurement of Current and Voltage Harmonics

Voltage and current harmonics can be measured with state-of-the-art equipment [2426]. Voltage and current waveshapes representing input and output voltages or currents of thyristor or triac controllers for induction motors are presented in Figs. 3.33 and 3.34. Here it is important that DC currents due to the asymmetric gating of the switches do not exceed a few percent of the rated controller current; DC currents injected into induction motors cause braking torques and generate additional losses or heating [26,28,41].

3.7.1.2 Measurement of Flux-Density Harmonics in Stator Teeth and Yoke (Back Iron)

Flux densities in the teeth of rotating machines are of the alternating type, whereas those in the yoke (back iron) are of the rotating type. The loss characteristics for both types are different and these can be requested from the electrical steel manufacturers. The motors to be tested are mounted on a testing frame, and measurements are performed under no load or minimum load with rated run capacitors. Four stator search coils with N = 3 turns each are employed to indirectly sense the flux densities of the teeth and yoke located in the axes of the main- and auxiliary-phase windings (see Fig. 3.36). The induced voltages in the four search coils are measured with two different methods: the computer sampling and the oscilloscope methods.

f03-36-9780128007822
Figure 3.36 Spatial location of four search coils.
Measurement via Computer Sampling

The computer-aided testing circuit and program [24] are relied on to measure the induced voltages of the four search coils of Fig. 3.36. The CAT circuit is shown in Figs. 3.36 and 3.37, where eight channel signals (vin, iin, im, ia, emt, eat, emy, and eay corresponding to input voltage and current, main- and auxiliary-phase currents, main- and auxiliary-phase tooth search-coil induced voltages, and main- and auxiliary-phase yoke (back iron)-search coil induced voltages, respectively), are sampled [36].

f03-37-9780128007822
Figure 3.37 CAT circuit for flux density measurement.

The flux linkages of the search coils are defined as

λt=etdt,

si135_e  (3-38)

and numerically obtained from

λ0=0andλi+1=λi+Δtei+1+ei/2,fori=0,1,2,n1.

si136_e  (3-39)

The average or DC value is

λave=1ni=0n1λi,

si137_e  (3-40)

or the AC values are

λi=λiλave,fori=0,1,2,3,,n,

si138_e  (3-41)

where e(t) of Eq. 3-38 is the induced voltage measured in a search coil and n in Eqs. 3-39 to 3-41 is the number of sampled points. The maximum flux linkage λmax can be obtained from λi. The maximum flux density is given by

Bmax=λmax/Ns,

si139_e  (3-42)

where the number of turns of each search coil is N = 3 and s is given for a stator tooth and yoke by the following equations:

s=kfewtl,

si140_e  (3-43)

s=kfehyl,

si141_e  (3-44)

where wt, hy, l, and kfe are the width of the (parallel) stator teeth (where two search coils reside), the height of the yoke or stator back iron (where two search coils reside), the length of the iron core, and the iron-core stacking factor, respectively. Measured data of PSC motors at rated voltage are presented in Section 3.7.2.

The maximum flux densities in stator teeth and yoke at the axes of the main- and auxiliary-phase windings are measured for various voltage amplitudes by recording the induced voltages in the four search coils. A digital oscilloscope is used to plot the induced voltages of the search coils. These waveforms are then sampled either by hand or by computer based on 83 points per period. Complex Fourier series components C~hsi142_e are computed – taking into account the Nyquist criterion – and the integration is performed in the complex domain as follows:

et=h=127C~hejhωt,

si143_e  (3-45)

λt=h=1271jhωC~hejhωt.

si144_e  (3-46)

The maximum flux linkage λmax is found from λ(t) and the maximum flux densities are determined from Eqs. 3-42 to 3-44.

3.7.2 Application Example 3.6: Measurement of Harmonics within Yoke (Back Iron) and Tooth Flux Densities of Single-Phase Induction Machines

The three permanent-split capacitor (PSC) motors (R333MC, Y673MG1, and 80664346) of Table 3.3 are subjected to the tests described in the prior subsections of Section 3.7, and pertinent results will be presented. These include stator, rotor, and magnetizing parameters, and tables listing no-load input impedances of the main and auxiliary phases (Tables E3.6.1 and E3.6.2). The bold data of the tables correspond to rated main- and auxiliary-phase voltages at no load. The rated auxiliary-phase voltage is obtained by multiplying the rated main-phase voltage with the measured turns ratio |k~am|si145_e. Lm and La are inductances of the main- and auxiliary-phase windings, respectively.

Table E3.6.1

Main-Phase (m) Input Impedance at Single-Phase No-Load Operation as a Function of Input Voltage for R333MC

Vm (V)Rm (Ω)Xm (Ω)Zm (Ω)Lm (mH)
250.39.6546.647.6126
229.911.256.958.0154
210.313.267.869.1183
189.815.177.378.7209
170.617.082.484.1223
149.518.685.487.4232
129.820.586.388.7235
109.222.686.789.5238
89.922.587.890.6240
69.426.387.090.9241

t0040

Table E.3.6.2

Auxiliary-Phase (a) Input Impedance at Single-Phase No-Load Operation as a Function of Input Voltage for R333MC

Va (V)Ra (Ω)Xa (Ω)Za (Ω)La (mH)
298.8830.694.799.5264
279.933.6106.0111.2295
259.336.8117.9123.5328
239.839.1126.2132.1351
219.441.1132.5138.7368
199.043.8135.6142.5378
179.345.6137.4144.7384
160.046.7138.6146.3388
139.149.9139.0147.6392
120.850.3140.2148.9395
100.154.3139.3149.5397

t0045

Solution to Application Example 3.6

3.7.2.1 Data of Motor R333MC: Solution

Stator Parameters

 60 Hz AC main-phase resistance: 4.53 Ω,

 main-phase leakage reactance: 5.02 Ω,

 auxiliary-phase resistance: 17.77 Ω,

 auxiliary-phase leakage reactance: 6.70 Ω,

 turns ratio between auxiliary- and main-phase windings: 1.34, and

 angle enclosed by the axes of auxiliary- and main-phase windings: 90.8°.

Rotor Parameters

 DC resistance: 4.687 Ω,

 60 Hz AC resistance: 4.703 Ω,

 120 Hz AC resistance: 4.751 Ω,

 linear leakage reactance: 5.635 Ω,

 skin-effect coefficient: kr = 0.0572, and

 nonlinear rotor leakage (λi) characteristic (referred to stator) is shown in Fig. E3.6.1.

f03-47-9780128007822
Figure E3.6.1 Nonlinear rotor leakage (λ – i) characteristic of R333MC.
Magnetizing Parameters

 unsaturated magnetizing reactance: 166.1 Ω,

 resistance corresponding to iron-core loss: 580 Ω, and

 nonlinear magnetizing (λi) characteristic is depicted in Fig. E3.6.2.

f03-48-9780128007822
Figure E3.6.2 Nonlinear magnetizing (λ – i) characteristic of R333MC.
No-Load Input Impedance

The main- and auxiliary-phase input impedances are measured and recorded in Tables E3.6.1 and E3.6.2, respectively, at single-phase no-load operation as a function of the rms value of the input voltage. Note that in Table E3.6.1 the parameters Zm and Lm are referred to the main-phase (m) winding, whereas in Table E3.6.2 Za and La are referred to the auxiliary-phase (a) winding.

Geometric Data

 length of iron core: 80 mm,

 width of stator tooth: 3.5 mm,

 height of main-phase yoke: 14.85 mm,

 height of auxiliary-phase yoke: 11.5 mm, and

 iron-core stacking factor: 0.95.

Measured Waveforms of Induced Voltages and Flux Densities

Figures E3.6.3a and E3.6.3b show the induced voltage waveforms within search coils wound on the teeth located at the axes of the main-phase and auxiliary-phase windings, respectively.

f03-49-9780128007822
Figure E3.6.3a Waveforms of input (vin = Vin, left-hand side ordinate) and main-phase winding tooth induced voltages (emt ≡ Emt, right-hand side ordinate, w = ω) of R333MC.
f03-50-9780128007822
Figure E3.6.3b Waveforms of input (vin = Vin, left-hand side ordinate) and auxiliary-phase winding tooth induced voltages (eat ≡ Eat, right-hand side ordinate, w = ω) of R333MC.

Figures E3.6.4a and E3.6.4b depict the flux density waveforms associated with the axis of the main-phase winding of all three machines as listed in Table 3.3. Figures E3.6.4c and E3.6.4d depict the flux density waveforms associated with the axis of the auxiliary-phase winding for all three machines.

f03-51-9780128007822
Figure E3.6.4a Flux density waveforms of tooth (Bmt) located in axis of main-phase winding at rated voltage of R333MC, 80664346, and Y673MG1.
f03-52-9780128007822
Figure E3.6.4b Flux density waveforms of yoke (Bmy) associated with main-phase winding at rated voltage of R333MC, 80664346, and Y673MG1.
f03-53-9780128007822
Figure E3.6.4c Flux density waveforms of tooth (Bat) located in axis of auxiliary-phase winding at rated voltage of R333MC, 80664346, and Y673MG1.
f03-54-9780128007822
Figure E3.6.4d Flux density waveforms of yoke (Bay) associated with auxiliary-phase winding at rated voltage of R333MC, 80664346, and Y673MG1.

Figures E3.6.5a, E3.6.5b, E3.6.5c, and E3.6.5d show that the flux densities are approximately a linear function of the input voltage. In Figs. E3.6.6a and E3.6.6b comparison of the flux densities in the teeth and yoke (back iron) are presented for the methods discussed in the prior subsections of Section 3.7. The no-load currents of R333MC and 80664346 are depicted in Figs. E3.6.7a and E3.6.7b, respectively.

f03-55-9780128007822
Figure E3.6.5a Maximum flux density in tooth located in axis of main-phase winding as a function of input voltage of R333MC, 80664346, and Y673MG1.
f03-56-9780128007822
Figure E3.6.5b Maximum flux density in tooth located in axis of auxiliary-phase winding as a function of input voltage of R333MC, 80664346, and Y673MG1.
f03-57-9780128007822
Figure E3.6.5c Maximum flux density in yoke located in axis of main-phase winding as a function of input voltage of R333MC, 80664346, and Y673MG1.
f03-58-9780128007822
Figure E3.6.5d Maximum flux density in yoke located in axis of auxiliary-phase winding as a function of input voltage of R333MC, 80664346, and Y673MG1.
f03-59-9780128007822
Figure E3.6.6a Comparison of maximum flux densities in teeth located in axis of main- and auxiliary-phase windings, and yokes located in axis of main- and auxiliary-phase windings for Y673MG1, 80664346, and R333MC (computer-sampling method).
f03-60-9780128007822
Figure E3.6.6b Comparison of maximum flux densities in teeth located in axis of main- and auxiliary-phase windings, and yokes located in axis of main- and auxiliary-phase windings for Y673MG1, 80664346, and R333MC (oscilloscope method).
f03-61-9780128007822
Figure E3.6.7a Oscillogram of total input current waveform of R333MC, 2 A/div for current iin(t) and input AC voltage vin(t) with 100 V/div at minimum load at f = 60 Hz.
f03-62-9780128007822
Figure E3.6.7b Oscillogram of total input current waveform of 80664346, 1 A/div for current iin(t) and input AC voltage vin(t) of 230.5 Vrms at minimum load at f = 60 Hz.

3.7.2.2 Discussion of Results and Conclusions

The permanent-split capacitor (PSC) motors of the air conditioners 80664346 and R333MC have the same output power and about the same overall dimensions (length and outer diameter). However, their stator and rotor windings are different. Measurements indicate that the 80664346 motor [36] generates a nearly sinusoidal flux density in the stator teeth (Fig. E3.6.4a), whereas that of the R333MC motor exhibits nonsinusoidal waveform although its (Bmt)max is smaller. In addition, the no-load current of the 80664346 (Fig. E3.6.7b) is about half that of R333MC (Fig. E3.6.7a), although the latter’s maximum flux density in stator teeth and back iron (yoke) is less (Figs. E3.6.4a and E3.6.4c). This indicates that the proper designs of the stator and rotor windings are important for the waveform of the stator tooth flux density and its associated iron-core loss, and for the efficiency optimization of PSC motors.

The yoke flux densities (Figs. E3.6.4b and E3.6.4d) are mostly sinusoidal for all machines tested [36]. Although 80664346 has a larger flux density than R333MC, the no-load losses of 80664346 are smaller as can be seen from their stator, rotor, and magnetizing parameters. References [30], [37], and Appendix of [38] address how time-dependent waveforms of the magnetizing current depend on the spatial distribution/pitch of a winding. Figures E3.6.5a through E3.6.5d show that in these three PSC motors saturation in stator teeth and yokes is small as compared to transformers, and therefore the principle of superposition with respect to flux densities and losses can be applied. Figures E3.6.6a and E3.6.6b demonstrate that both the computer sampling and the oscilloscope methods lead to similar results with respect to maximum flux densities. One concludes that although a single-phase motor (e.g., 80664346) has relatively large flux densities in stator yoke and teeth the no-load current can be relatively small, although it is quite nonsinusoidal. One of the reasons for this is that the power factor at no load is relatively large. The best design of a PSC motor is thus not necessarily based on sinusoidal current wave shapes. Further work should investigate the following:

 the losses generated by alternating flux densities as they occur in the stator teeth, and rotating flux densities as they occur in the stator yoke, and

 the interrelation between time and space harmonics and their effects on loss generation.

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

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