11.4.4 Space-Vector-Based Modulating Techniques in CSIs

The objective of the SV-based modulating technique is to generate PWM load line currents that are on average equal to given load line currents. This is done digitally in each sampling period by properly selecting the switch states from the valid ones of the CSI (Table 11.6) and the proper calculation of the period of times they are used. As in VSIs, the selection and time calculations are based upon the space-vector transformation.

11.4.4.1 Space-Vector Transformation in CSIs

Similarly to VSIs, the vector of three-phase line-modulating signals iabcc=[icaicbicc]Tsi490_e can be represented by the complex vector ic=iαβc=[icαicβ]Tsi491_e by means of Eqs. (11.31) and (11.32). For three-phase balanced sinusoidal modulating waveforms, which feature an amplitude îc and an angular frequency ω, the resulting modulating signal complex vector ic=iαβcsi492_e becomes a vector of fixed module îc, which rotates at frequency ω (Fig. 11.37). Similarly, the SV transformation is applied to the line currents of the nine states of the CSI normalized with respect to ii, which generates nine space vectors (iisi493_e, i=1,2,,9si494_e in Fig. 11.37). As expected, i1si233_e to i6si232_e are nonnull line current vectors, and i7si234_e, i8si240_e, and i9si237_e are null line current vectors.

f11-37-9780128114070
Fig. 11.37 The space-vector representation in CSIs.

The SV technique approximates the line-modulating signal space vector icsi500_e by using the nine space vectors (iisi226_e, i=1,2,,9si502_e) available in CSIs. If the modulating signal vector icsi500_e is between the arbitrary vectors iisi226_e and ii+1si227_e, then iisi226_e and ii+1si227_e combined with one zero SV (iz=i7si508_e or i8si240_e or i9si237_e) should be used to generate icsi500_e. To ensure that the generated current in one sampling period Ts (made up of the currents provided by the vectors iisi226_e, ii+1si227_e, and izsi228_e used during times Ti, Ti+1si418_e, and Tz)si516_e is on average equal to the vector icsi500_e, the following expressions should hold:

Ti=Tsˆicsin(π/3θ)

si518_e  (11.52)

Ti+1=Tsˆicsin(θ)

si519_e  (11.53)

Tz=TsTiTi+1

si424_e  (11.54)

where 0ˆic1si521_e. Although the SVM technique selects the vectors to be used and their respective on-times, the sequence in which they are used, the selection of the zero space vector, and the normalized sampled frequency remain undetermined.

11.4.4.2 Space-Vector Sequences and Zero Space-Vector Selection

Although there is no systematic approach to generate a SV sequence, a graphic representation shows that the sequence iisi226_e, izsi228_e, ii+1si227_e (where the chosen izsi228_e depends upon the sector) provides high performance in terms of minimizing unwanted harmonics and reducing the switching frequency. To obtain the zero SV that minimizes the switching frequency, it is assumed that Ic is in sector ②. Then, Fig. 11.38 shows all the possible transitions that could be found in sector ②. It can be seen that the zero vector i9si237_e should be chosen to minimize the switching frequency in all cases. Table 11.7 gives a summary of the zero space vector to be used in each sector in order to minimize the switching frequency. However, it should be noted that Table 11.7 is valid only for the sequence iisi226_e, ii+1si227_e, izsi228_e. Another sequence will require reformulating the zero space-vector selection algorithm.

f11-38-9780128114070
Fig. 11.38 Possible state transitions in sector ② involving a zero SV: (A) transition, i1izi2si25_e or i2izi1si26_e; (B) transition, i1izi1si27_e; and (C) transition, i2izi2si28_e.

Table 11.7

Zero SV for minimum switching frequency in CSI and sequence iisi226_e, ii+1si227_e, izsi228_e

Sector iisi226_e ii+1si227_e izsi228_e
i6si232_e i1si233_e i7si234_e
i1si233_e i2si236_e i9si237_e
i2si236_e i3si239_e i8si240_e
i3si239_e i4si242_e i7si234_e
i4si242_e i5si245_e i9si237_e
i5si245_e i6si232_e i8si240_e

t0040

11.4.4.3 The Normalized Sampling Frequency

As in VSIs modulated by an SV approach, the normalized sampling frequency fsn should be an integer multiple of 6 to minimize uncharacteristic harmonics. As an example, Fig. 11.39 shows the relevant waveforms of a CSI SVM for fsn=18si20_e and ˆic=0.8si29_e. Fig. 11.39 also shows that the first set of relevant harmonics load line current are at fsn.

f11-39-9780128114070
Fig. 11.39 The three-phase CSI. Ideal waveforms for space-vector modulation (ˆic=0.8si29_e, fsn=18si20_e): (A) modulating signals, (B) switch S1 state, (C) switch S3 state, (D) ac output current, (E) ac output current spectrum, (F) ac output voltage, (G) dc voltage, (H) dc voltage spectrum, (I) switch S1 current, and (J) switch S1 voltage.

11.4.5 DC Link Voltage in Three-phase CSIs

An instantaneous power balance indicates that

vi(t)ii(t)=van(t)ioa(t)+vbn(t)iob(t)+vcn(t)ioc(t)

si532_e  (11.55)

where van(t), vbn(t), and vcn(t) are the phase filter voltages as shown in Fig. 11.40. If the filter is large enough and a relatively high switching frequency is used, the phase voltages become nearly sinusoidal balanced waveforms. On the other hand, if the ac output currents are considered sinusoidal and the dc link current is assumed constant ii(t)=Iisi533_e, Eq. (11.55) can be simplified to

f11-40-9780128114070
Fig. 11.40 Phase voltage definition in a wye-connected filter.

vi(t)=1Ii{,2Vonsin(ωt)2Io1sin(ωtϕ)+2Vonsin(ωt120)2Io1sin(ωt120°ϕ)+2Vonsin(ωt240)2Io1sin(ωt240°ϕ),}

si534_e  (11.56)

where Von is the rms ac output phase voltage, Io1 is the rms fundamental line current, and ϕ is an arbitrary filter-load angle. Hence, the dc link voltage expression can be further simplified to the following:

vi(t)=3Io1IiVoncos(ϕ)=3Io1IiVocos(ϕ)

si535_e  (11.57)

where Vo=3Vonsi536_e is the rms load line voltage. The resulting dc link voltage expression indicates that the first line current harmonic Io1 generates a clean dc current. However, as the load line currents contain harmonics around the normalized sampling frequency fsn, the dc link current will contain harmonics but around fsn as shown in Fig. 11.39H. Similarly, in carrier-based PWM techniques, the dc link current will contain harmonics around the carrier frequency mf (Fig. 11.31).

In practical implementations, a CSI requires a dc current source that should behave as a constant (as required by PWM CSIs) or variable (as square-wave CSIs) current source. Such current sources should be implemented as separate units.

11.5 Multilevel Inverters

The traditional inverters can generate an output in two or three different levels. For instance, a single-phase full-bridge inverter can produce vi, visi81_e, and zero; this type of output is unacceptable in the medium- to high-voltage ranges due to the high dv/dt present in the PWM ac line voltages. In order to solve this, the multilevel inverters were initially proposed. However, now, their use has been more extended not only due to the low harmonic content obtained at low switching frequencies but also for the good efficiency obtained.

Different multilevel topologies have been proposed. The goal is to produce an output defined by different levels that can be voltage or current levels. The basic principle is to connect the output to multiple points, one at a time. Fig. 11.41A shows the idea for a voltage-source configuration. Usually, the number M of possible level defines the name of the inverter; for instance, if five points are available as in Fig. 11.41A (M=5), the inverter is usually named a five-level inverter.

f11-41-9780128114070
Fig. 11.41 Basic principle of multilevel output: (A) voltage source and (B) typical output.

For a five-level inverter, the output voltage shown in Fig. 11.41B would be produced. By simple inspection, this output voltage looks more similar to a sinusoidal waveform; then, the harmonic content is expected to be lower than traditional inverters, and the power semiconductors are operating at low switching frequency.

11.5.1 Voltage Source-Based Multilevel Topologies

The three-phase VSI can be called also two-level VSI due to the fact that the inverter phase voltages vaN, vbN, and vcN (Fig. 11.17) are instantaneously either vi/2 or − vi/2. In other words, the phase voltages can take one of the two voltage levels. However, regarding to multilevel inverters, the topologies usually have more than two levels.

The typical multilevel voltage-source topologies are the neutral-point-clamped (also called diode-clamped) inverters, the flying capacitors, and the cascaded configuration. Since they are voltage-source converters, the compulsory rules described in Section 11.2.1 also apply, but certainly, some remarks are made for each converter.

11.5.1.1 Neutral Point Clamped Inverter

In Fig. 11.42, the neutral-point-clamped inverter is shown (three levels). For this topology, not only an M-level inverter requires (M−1) capacitors and 2(M−1) switching semiconductors per leg, but also 2(M−2) additional diodes must be included per leg; for instance, Da+ and Da− are added for the left leg. M−1 devices per leg must be on to select the desired level.

f11-42-9780128114070
Fig. 11.42 Three-phase three-level VSI topology, neutral-point-clamped inverter.

Table 11.8 gives the switch states of the three-level inverter, just left leg. As it can be observed, not only always two switches are on, but also the signal S1a is the opposite of S4a, and S1b is the opposite of S4b. These conditions must be always satisfied. The switch states for phases b and c are identical to those of phase a.

Table 11.8

Valid switch states for the three-level neutral-point-clamped inverter, the left leg

s1a s1b s4a s4b van Components conducting
1 1 0 0 vi/2 S1a, S1b If ioa>0si250_e
D1a, D1b If ioa<0si251_e
0 1 1 0 0 S1b, Da+si252_e If ioa>0si250_e
S4a, Dasi254_e If ioa<0si251_e
0 0 1 1 vi/2si53_e D4a, D4b If ioa>0si250_e
S4a, S4b If ioa<0si251_e

t0045

11.5.1.2 Flying Capacitors Inverter

In Fig. 11.43, the flying capacitor inverter is shown (three levels). For this topology, an M-level inverter requires M1j=1(Mj)si538_e capacitors per leg (considering that the capacitors have the same voltage rating) and 2(M−1) switching semiconductors per leg, but no additional diodes are required. Also M−1 switches per leg must be on to select the desired level.

f11-43-9780128114070
Fig. 11.43 Three-phase three-level VSI topology, flying capacitors.

Table 11.9 is the switch state of the three-level inverter, just left leg. As it can be observed, not only always two switches are on, but also the signal S1a is the opposite of S4a, and S1b is the opposite of S4b. These conditions must be always satisfied. Table 11.9 shows that there is more than one switching state to produce the zero output level; this redundancy should be employed to distribute the losses or to regulate the capacitor voltages. Moreover, as the number of levels increase, more redundancy exists at different levels.

Table 11.9

Valid switch states for the three-level flying capacitors inverter, left leg

s1a s1b s4a s4b vaN Components conducting
1 1 0 0 vi/2 S1a, S1b If ioa>0si250_e
D1a, D1b If ioa<0si251_e
1 0 0 1 0 S1a, D4b If ioa>0si250_e
D1a, S4b If ioa<0si251_e
0 1 1 0 0 S1b, D4a If ioa>0si250_e
D1b, S4a If ioa<0si251_e
0 0 1 1 visi81_e /2 D4a, D4b If ioa>0si250_e
S4a, S4b If ioa<0si251_e

t0050

11.5.1.3 Casacade Inverter

In Fig. 11.44, the cascade inverter is shown (five levels). This topology is based on several full-bridge single-phase voltage-source inverters with their output connected in series. An M-level inverter requires (M−1)/2 single-phase full-bridge VSI per leg and 2(M−1) switching semiconductors per leg, but no additional diodes are required. Also M−1 switches per leg must be on to select the desired level.

f11-44-9780128114070
Fig. 11.44 Single-phase five-level VSI topology, cascade.

Table 11.10 gives the switch states of the five-level inverter. Notice that only the top switch control signals of each single-phase full-bridge VSI are shown. Their respective bottom switch control signals are complementary, in order to accomplish with the first compulsory rule for VSIs. As it can be observed, always four devices are on; notice that since a single-phase full-bridge inverter is used, the same rules apply for the control signals of the semiconductors. As it can be observed in Table 11.10, there are more than one switching state to produce some output levels; this redundancy should be employed to distribute the losses. Moreover, as the number of levels increases, more redundancy exists at different levels.

Table 11.10

Valid switch states for the five-level cascade inverter, left leg

s21+si268_e s22+si269_e s11+si270_e s12+si271_e van
1 0 1 0 vi
1 0 0 0 vi/2
1 0 1 1 vi/2
0 0 1 0 vi/2
1 1 1 0 vi/2
0 0 0 0 0
0 0 1 1 0
1 0 0 1 0
0 1 1 0 0
1 1 0 0 0
1 1 1 1 0
0 1 0 0 vi/2si53_e
0 1 1 1 vi/2si53_e
0 0 0 1 vi/2si53_e
1 1 0 1 vi/2si53_e
0 1 0 1 visi81_e

t0055

11.5.2 Current Source-Based Multilevel Topologies

Duality is found in many aspects related to voltage- and current-source inverters. Perhaps, the most evident is the duality in terms of modulating techniques. Thus, current-source-based multilevel topologies are available as well. There are different topologies reported: paralleled, two-stage, and embedded configurations.

Fig. 11.45 shows a three-level M=3 current-source topology, which is formed by paralleling two three-phase CSI topologies, the named paralleled CSI (Table 11.11). The main goal is to share evenly the ac current ioabc among the two topologies (iabco/2=iabco1=iabc02si539_e). This should be ensured by having equal dc link currents (ii1=ii2si540_e). Similarly to voltage-source-based multilevel topologies, this could be achieved either by using two independent dc link currents or by properly gating the power semiconductors.

f11-45-9780128114070
Fig. 11.45 Three-phase three-level paralleled CSI topology.

Table 11.11

Valid switch states for the three-level paralleled inverter, left legs

s11 s12 s41 s42 ioa
1 1 0 0 ii2+ii1
1 1 0 1 ii1
1 1 1 0 ii2
1 1 1 1 0
1 0 1 1 ii2
0 1 1 1 ii1
0 0 1 1 −(ii2+ii1)

t0060

11.5.3 Modulation Techniques in Multilevel Inverters

As in traditional inverters, the sinusoidal pulse width and space-vector modulation can be employed in multilevel inverters, for both CSI and VSI.

11.5.3.1 The SPWM Technique

The main objective is to generate the appropriate gating signals so as to obtain fundamental inverter phase voltages or currents equal to a given set of modulating signals. Specifically, the SPWM in three-level VSIs uses a sinusoidal set of modulating signals (vca, vcb, and vcc for phases a, b, and c, respectively) and M−1=2 triangular type of carrier signals (vΔ1 and vΔ2) as illustrated in Fig. 11.46A. The best results are obtained if the carrier signals are in-phase (synchronized with the modulating signal) and feature an odd normalized frequency (e.g., mf=15si24_e). According to Fig. 11.46A and considering a neutral-point-clamped inverter (Fig. 11.42), the switch S1a is turned either on if vca>vΔ1si542_e or off if vca<vΔ1si543_e, and switch S1b is turned either on if vca>vΔ2si544_e or off if vca<vΔ2si545_e. Additionally, the switch S4a status is obtained as the opposite to switch S1a, and the switch S4b status is obtained as the opposite to switch S1b. In order to use the same set of carrier signals to generate the gating signals for phases b and c, the normalized frequency of the carrier signal mf should be a multiple of 3. Thus, the possible values are mf=3,9,15,21,si546_e.

f11-46-9780128114070
Fig. 11.46 Three-level VSI topology. Relevant waveforms using an SPWM (mf=15si24_e, ma=0.8si13_e): (A) modulating and carrier signals, (B) switch S1a status, (C) switch S4b status, (D) inverter phase a voltage, (E) inverter phase a voltage spectrum, (F) load line voltage, (G) load line voltage spectrum, and (H) phase-load a voltage.

Fig. 11.46 shows the relevant waveforms for a three-level neutral-point-clamped inverter modulated by means of an SPWM technique (mf=15si24_e, ma=0.8si13_e). Specifically, Fig. 11.46D shows the inverter phase voltage, which is clearly a three-level type of voltage, and Fig. 11.46F shows the load line voltage, which shows that the step voltages are at most vi/2. More importantly, the inverter phase voltage (Fig. 11.46E) contains harmonics at lmf±ksi549_e with l=1,3,si550_e and k=0,2,4,si551_e and at lmf±ksi549_e with l=2,4,si358_e and k=1,3,si554_e. For instance, the first set of harmonics (l=1si555_e, mf=15si24_e) are at 15, 15±2si557_e, 15±4,si558_e. The inverter line voltage (Fig. 11.46G) contains harmonics at lmf±ksi549_e with l=1,3,si550_e and k=2,4,si561_e and at lmf±ksi549_e with l=2,4,si358_e and k=1,3,si554_e. For instance, the first set of harmonics in the line voltages (l=1si555_e, mf=15si24_e) are at 15±2si557_e, 15±4,si558_e.

All the other features of carrier-based PWM techniques also apply in multilevel inverters. Firstly, the fundamental component of the inverter phase voltages satisfies

ˆvaN1=ˆvbN1=ˆvcN1=mavi20<ma1

si569_e  (11.58)

and thus, the line voltages satisfy

ˆvab1=ˆvbc1=ˆvca1=ma3vi20<ma1

si570_e  (11.59)

where 0<ma1si571_e is the linear operating region. To further increase the amplitude of the load voltages, the overmodulation operating region can be used by further increasing the modulating signal amplitudes (ma>1si306_e), where the line voltages range in

3vi2<ˆvab1=ˆvbc1=ˆvca1<4π3vi2

si377_e  (11.60)

Second, the modulating signals could be improved by adding a third harmonic (zero sequence), which will increase the linear region up to ma=1.15si574_e. This results in a maximum fundamental line-voltage component equal to vi; as a third point, a nonsinusoidal set of modulating signals could also be used by the modulating technique. This is the case where nonsinusoidal line voltages are required as in active filter applications, and finally, because of the two quadrants operation of VSIs, the multilevel inverter could equally be used in applications where the active power flow goes from the dc to the ac side or from the ac to the dc side.

In general, for an M-level inverter modulated by means of a carrier-based technique, three modulating signals 120° out of phase and M−1 carrier signals are required. Key waveforms are shown in Fig. 11.47.

f11-47-9780128114070
Fig. 11.47 Five-level VSI topology. Relevant waveforms using a SPWM (mf=15si24_e, ma=0.8si13_e): (A) inverter phase a voltage, (B) inverter phase a voltage spectrum, (C) load line voltage, and (D) load line voltage spectrum.

One of the drawbacks of the multilevel inverter is that the dc link capacitor voltages or inductor currents, depending on the type of inverter, should be regulated. Unfortunately, this is not a natural operating condition mainly due to the fact that the storage elements are operated in a not symmetrical manner, and therefore, they will not equally share energy.

11.5.3.2 The SPWM Technique in Three-level CSIs

As in three-level VSIs, the main objective is to generate the appropriate 12 gating signals so as to obtain fundamental inverter line currents equal to a given set of modulating signals. Specifically, the SPWM in three-level inverters uses a sinusoidal set of modulating signals (ica, icb, and icc for phases a, b, and c, respectively) and N1=2si575_e triangular type of carrier signals (iΔ1 and iΔ2) as illustrated in Fig. 11.48A and E. The best results are obtained if the carrier signals are 180° out of phase and feature an odd normalized frequency (e.g., mf=15si24_e). In order to use the same set of carrier signals to generate the gating signals for phases b and c, the normalized frequency of the carrier signal mf should be a multiple of 3. Thus, the possible values are mf=3,9,15,21,si546_e.

f11-48-9780128114070
Fig. 11.48 Three-level CSI topology. Relevant waveforms using an SPWM (mf=15si24_e, ma=0.8si13_e): (A) modulating signals and carrier signal 1, (B) switch S11 status, (C) inverter 1 linear current, (D) inverter 1 linear current spectrum, (E) modulating signals and carrier signal 2, (F) switch S12 status, (G) inverter 2 linear current, and (H) inverter 2 linear current spectrum.

Fig. 11.48 shows the relevant waveforms for a three-level inverter modulated by means of an SPWM technique (mf=15si24_e, ma=0.8si13_e). Specifically, Fig. 11.48B and F show the gating signals obtained as described earlier in this chapter. The inverter line currents shown in Fig. 11.48C and G feature spectra shown in Fig. 11.48D and H, respectively. As expected, the inverter line currents contain harmonics at lmf±ksi549_e with l=1,3,si550_e and k=2,4,si561_e and at lmf±ksi549_e with l=2,4,si358_e and k=1,3,si554_e. For instance, the first set of harmonics in the line currents (l=si586_e 1, mf=15si24_e) are at 15±2si557_e, 15±4,si558_e.

The total inverter line current is shown in Fig. 11.49A and features the first set of unwanted harmonics around 2mf (Fig. 11.49B). This becomes the first advantage of using a multilevel topology as the filtering component requirements become more relaxed. All the other features of carrier-based PWM techniques also apply in current-source multilevel inverters. For instance, (I) the fundamental component of the line currents satisfies

f11-49-9780128114070
Fig. 11.49 Three-level CSI topology. Relevant waveforms using an SPWM (mf=15si24_e, ma=0.8si13_e): (A) total inverter line current and (B) total inverter line current spectrum.

ˆioa1=ˆiob1=ˆioc1=ma32(ii1+ii2)0<ma1

si590_e  (11.61)

where 0<ma1si571_e is the linear operating region. Also, (II) to further increase the amplitude of the load currents, a zero-sequence signal could be injected to the modulating signals, in this case,

ˆioa1=ˆiob1=ˆioc1=ma32(ii1+ii2)0<ma2/3

si592_e  (11.62)

the overmodulation operating region can be used by further increasing the modulating signal amplitudes (ma>2/3si593_e), where the line currents range in

(ii1+ii2)<ˆioa1=ˆiob1=ˆioc1<4π(ii1+ii2)

si594_e  (11.63)

Also, (III) a nonsinusoidal set of modulating signals could also be used by the modulating technique. This is the case where nonsinusoidal line currents are required as in active filter applications, and (IV) because of the two quadrants operation of CSIs, the multilevel inverter could equally be used in applications where the active power flow goes from the dc to the ac side or from the ac to the dc side. In general, for an N-level inverter modulated by means of a carrier-based technique, three modulating signals 120° out of phase and N1si321_e carrier signals are required, and the line currents in the inverters have a peak value of ii/(N1si321_e).

One of the drawbacks of the multilevel inverter is that the dc link capacitors cannot be supplied by a single dc voltage source. This is due to the fact that the currents required by the inverter in the dc bus are not symmetrical, and therefore, the capacitors will not equally share the dc supply voltage vi. To overcome this problem, two alternatives are developed later on.

11.5.3.3 The Space-Vector Modulation

The SV modulating technique can be applied using the same principles used in two-level inverters. However, the higher number of voltage levels increases the complexity of the practical implementation of the technique. For instance, in M=3-level inverters, each leg allows M=3 different switch combinations as indicated in the previous sections. Therefore, in a three-phase system, there are M 3=27 total valid switch combinations, which generate M 3=27 load line voltages that are represented by M 3=27 space vectors (v1si401_e, v2,,v27si598_e) in Fig. 11.50. For instance, v2=0.5+j0.866si599_e is due to the line voltages vab=0.5si600_e, vbc=0.5si601_e, vca=1.0si602_e in per unit. Thus, although the principle of operation is the same, the SV digital algorithm will have to deal with a higher number of states M 3. Moreover, because some space vectors (e.g., v13si603_e and v14si604_e in Fig. 11.50) produce the same load voltage terminals, the algorithm will have to decide between the two based on additional criteria and that of the basic SV approach. Clearly, as the number of level increases, the algorithm becomes more and more elaborate. However, the benefits are not evident as the number of level increases. The maximum number of levels used in practical applications is five. This is based on a compromise between the complexity of the implementation and the benefits of the resulting waveforms.

f11-50-9780128114070
Fig. 11.50 The space-vector representation in a three-level VSI.

11.5.3.4 Balancing Issues

The operation for the multilevel converter was made considering an even distribution of the voltage across the dc link capacitors or current of the inductors, depending of the type of inverter. This even distribution is not naturally achieved and could be overcome by supplying the storage elements from independent supplies or properly gating the power semiconductors of the inverter in order to minimize the unbalance.

Fig. 11.51 shows an ASD based on a three-level VSI, where the dc link capacitors are feed from two different sources. This approach is being commercially used as it ensures a robust balanced dc link voltage distribution and operates with a high-performance type of ac main current. Indeed, for an M-level inverter, M−1-independent dc voltage supplies are required that could be provided by M−1 six-pulse rectifiers feed from an M−1 pulse transformer. Therefore, the ac main currents are an M−1-level type of waveform.

f11-51-9780128114070
Fig. 11.51 ASD based on a three-phase three-level VSI topology.

This approach cannot be used when the inverter does not feature dc link voltage supplies. This is the case of static power reactive power compensators and static power active filters. In this case, the proper gating of the power semiconductors becomes the only choice to keep and balance the dc link voltages. Fig. 11.52A shows this case where the current added by the inverter ioabc provides the reactive power and current harmonics such that the ac main current isabc features a given power factor.

f11-52-9780128114070
Fig. 11.52 Reactive power and current harmonics compensator based on a three-phase three-level VSI topology: (A) power topology, (B) carrier and modulating signals, and (C) δ closed-loop scheme.

The SPWM modulating technique could be used as in Fig. 11.46; however, the zero level of the carrier δ is left as a manipulable variable in Fig. 11.52B. In fact, it is used to control the difference of the upper and lower capacitor voltages Δvi=vi1vi2si605_e. A closed-loop alternative is depicted in Fig. 11.52C to manipulate δ. The modulating signals vcabc are left to control the reactive power and current harmonics injected into the ac mains by regulating the currents ioabc and keep the total dc link voltage vi=vi1+vi2si606_e equal to a reference. Both loops are not included in Fig. 11.52C.

11.6 Closed-Loop Operation of Inverters

In the inverter, as a classic power converter, the output depends on different conditions. For instance, in ASDs, the load usually changes, so the ac waveforms should be adjusted to these new conditions. Also, as the dc power supplies are not ideal and the dc quantities are not fixed, the inverter should compensate for such variations. Such adjustments can be done automatically by means of a closed-loop approach. Inverters also provide an alternative to changing the load operating conditions.

There are two alternatives for closed-loop operation, the feedback and the feedforward approaches. It is known that the feedback approach can compensate for both the perturbations (dc power variations) and the load variations (load torque changes). However, the feedforward strategy is more effective in mitigating perturbations as it prevents its negative effects at the load side. These cause-effect issues are analyzed in three-phase inverters in the following, although similar results are obtained for single-phase VSIs.

11.6.1 Feedforward Techniques in Voltage Source Inverters

The dc link bus voltage in VSIs is usually considered a constant-voltage source vi. Unfortunately and due to the fact that most practical applications generate the dc bus voltage by means of a diode rectifier (Fig. 11.53), the dc bus voltage contains low-order harmonics such as the sixth, twelfth, … (due to six-pulse diode rectifiers) and the second harmonic if the ac voltage supply features an unbalance, which is usually the case. Additionally, if the three-phase load is unbalanced, as in UPS applications, the dc input current in the inverter ii also contains the second harmonic, which in turn contributes to the generation of a second voltage harmonic in the dc bus.

f11-53-9780128114070
Fig. 11.53 Three-phase VSI topology with a diode-based front-end rectifier.

The basic principle of feedforward approaches is to sense the perturbation and then modify the input in order to compensate for its effect. In this case, the dc link voltage should be sensed, and the modulating technique should accordingly be modified. The fundamental ab line voltage in a VSI SPWM can be written as

vab1(t)={vca1(t)ˆvΔvcb1(t)ˆvΔ}32vi(t)ˆvΔ>ˆvca1,ˆvcb1

si607_e  (11.64)

where ˆvΔsi286_e is the carrier signal peak, ˆvca1si609_e and ˆvcb1si610_e are the modulating signal peaks, and vca(t) and vcb(t) are the modulating signals. If the dc bus voltage vi varies around a nominal Vi value, then the fundamental line voltage varies proportionally; however, if the carrier signal peak ˆvΔsi286_e is redefined as

ˆvΔ=ˆvΔmvi(t)Vi

si612_e  (11.65)

where ˆvΔmsi613_e is the carrier signal peak (Fig. 11.54), then the resulting fundamental ab line voltage in a VSI SPWM is

f11-54-9780128114070
Fig. 11.54 The three-phase VSI. Feedforward control technique to reject dc bus voltage variations.

vab1(t)={vca1(t)ˆvΔmvcb1(t)ˆvΔm}32Vi

si614_e  (11.66)

where, clearly, the result does not depend upon the variations of the dc bus voltage.

Fig. 11.55 shows the waveforms generated by the SPWM under a severe dc bus voltage variation (a second harmonic has been added manually to a constant Vi). As a consequence, the ac line voltage generated by the VSI is distorted as it contains low-order harmonics (Fig. 11.55E). These operating conditions may not be acceptable in standard applications such as ASDs because the load will draw distorted three-phase currents as well. The feedforward loop performance is illustrated in Fig. 11.56. As expected, the carrier signal is modified so as to compensate for the dc bus voltage variation (Fig. 11.56B). This is probed by the spectrum of the ac line voltage that does not contain low-order harmonics (Fig. 11.56E). It should be noted that ˆvΔ>ˆvca1,ˆvcb1si615_e; therefore, the compensation capabilities are limited by the required ac line voltage.

f11-55-9780128114070
Fig. 11.55 The three-phase VSI. Waveforms for regular SPWM (ma=0.8si13_e, mf=9si14_e): (A) dc bus voltage, (B) carrier and modulating signals, (C) ac output voltage, and (D) ac output voltage spectrum.
f11-56-9780128114070
Fig. 11.56 The three-phase VSI. Waveforms for SPWM including a feedforward loop (ma=0.8si13_e, mf=9si14_e): (A) carrier and modulating signals, (B) modified carrier and modulating signals, (C) ac output voltage, and (D) ac output voltage spectrum.

The performance of the feedforward approach depends upon the frequency of the harmonics present in the dc bus voltage and the carrier signal frequency. Fortunately, the relevant unwanted harmonics to be found in the dc bus voltage are the second, due to unbalanced supply voltages, and/or the sixth as the dc bus voltage is generated by means of a six-pulse diode rectifier. Therefore, a carrier signal featuring a 15-pu frequency is found to be sufficient to properly compensate for dc bus voltage variations.

Unbalanced loads generate a dc input current ii that contains a second harmonic, which contributes to the dc bus voltage variation. The previous feedforward approach can compensate for such perturbation and maintain balanced ac load voltages.

Digital techniques can also be modified in order to compensate for dc bus voltage variations by means of a feedforward approach.

The duality principle between the voltage- and the current-source inverters indicates that, as described previously, the feedforward approach can be also used for CSIs and for VSIs. Therefore, low-order harmonics present in the dc bus current can be compensated for before they appear at the load side. This can be done for both analog-based (e.g., carrier-based) and digital-based (e.g., space-vector) modulating techniques.

11.6.2 Feedback Techniques in Voltage Source Inverters

Unlike the feedforward advantages, the feedback techniques are more effective to regulate than the system output by gating signals of the inverters properly. Another important difference is that feedback techniques need to sense the controlled variables. In general, the controlled variables (output to the system) are chosen according to the control objectives. For instance, in ASDs, it is usually necessary to keep the motor line currents equal to a given set of sinusoidal references. Therefore, the controlled variables become the ac line currents. There are several alternatives to implement feedback techniques in VSIs, and three of them are discussed next.

11.6.2.1 Hysteresis Current Control

The main purpose here is to force the ac line current to follow a given reference. The status of the power semiconductors are changed whenever the actual ioa current goes beyond a given reference ioa,ref±Δi/2si616_e. Fig. 11.57 shows the hysteresis current controller for one phase. Identical controllers are used in the other phases. The implementation of this controller is simple as it requires a simple comparator operating in the hysteresis mode; thus, the controller and modulator are combined in one unit.

f11-57-9780128114070
Fig. 11.57 Hysteresis current control, one phase.

Unfortunately, there are several drawbacks associated with the technique itself. First, the switching frequency cannot be predicted as in carrier-based modulators, and therefore, the harmonic content of the ac line voltages and currents becomes random (Fig. 11.58D). This could be a disadvantage when designing the filtering components. Second, as three-phase loads do not have the neutral connected as in ASDs, the load currents add up to zero. This means that only two ac line currents can be controlled independently at any given instant. Therefore, one of the hysteresis controllers is redundant at a given time. This explains why the load current goes beyond the limits and introduces limit cycles (Fig. 11.58A). Finally, although the ac load currents add up to zero, the controllers cannot ensure that all load line currents feature a zero dc component in one load cycle.

f11-58-9780128114070
Fig. 11.58 The three-phase VSI. Ideal waveforms for hysteresis current control: (A) actual ac load current and reference, (B) switch S1 state, (C) ac output voltage, and (D) ac output voltage spectrum.

11.6.2.2 Linear Control of VSIs

Proportional and proportional-integrative controllers can also be used in VSIs by employing a transformation. The main purpose is to generate the modulating signals vca, vcb, and vcc in a closed-loop fashion as depicted in Fig. 11.59. The modulating signals can be used by a carrier-based technique such as the SPWM (as depicted in Fig. 11.59) or by space-vector modulation. Because the load line currents add up to zero, the load line current references must add up to zero. Thus, the abc/ αβγ transformation can be used to reduce to two controllers the overall implementation scheme as the γ component is always zero. This avoids limit cycles in the ac load currents.

f11-59-9780128114070
Fig. 11.59 The three-phase VSI. Feedback control based on linear controllers.

The transformation of a set of variables in the stationary abc frame xabc into a set of variables in the stationary αβ frame xαβ is given by

eq11-01-9780128114070  (11.67)

The selection of the controller (P, PI, …) is done according to the control procedures such as steady-state error, settling time, and overshoot. Fig. 11.60 shows the relevant waveforms of a VSI SPWM controlled by means of a PI controller as shown in Fig. 11.59.

f11-60-9780128114070
Fig. 11.60 The three-phase VSI. Ideal waveforms for a PI controller in a feedback loop (ma=0.8si13_e, mf=15si24_e): (A) actual ac load current and reference, (B) carrier and modulating signals, (C) ac output voltage, and (D) ac output voltage spectrum.

Although it is difficult to prove that no limit cycles are generated, the ac line current appears very much sinusoidal. Moreover, the ac line voltage generated by the VSI preserves the characteristics of such waveforms generated by SPWM modulators.

However, an error between the actual ioa and the ac line current reference ioa,ref can be observed (Fig. 11.60A). This error is inherent to linear controllers and cannot be totally eliminated, but it can be minimized by increasing the gain of the controller. However, the noise in the circuit is also increased, which could deteriorate the overall performance of the control scheme. The inherent presence of the error in this type of controllers is due to the fact that the controller needs a sinusoidal error to generate sinusoidal modulating signals vca, vcb, and vcc, as required by the modulator. Therefore, an error must exist between the actual and the ac line current references.

Nevertheless, as current-controlled VSIs are actually the inner loops in many control strategies, their inherent errors are compensated by the outer loop. This is the case of ASDs, where the outer speed loop compensates the inner current loops. In general, if the outer loop is implemented with dc quantities (such as speed), it can compensate the ac inner loops (such as ac line currents). If it is mandatory that a zero steady-state error be achieved with the ac quantities, then a stationary (abc frame) to rotating (dq frame) transformation is a valid alternative to use.

11.6.2.3 Linear Control of VSIs in a Rotating Frame

The rotating dq transformation allows ac three-phase circuits to be operated as if they were dc circuits. This is based upon a mathematical operation, that is, the transformation of a set of variables in the stationary abc frame xabc into a set of variables in the rotating dq0 frame xdq0. The transformation is given by

xdq=23[sin(ωt)sin(ωt2π/3)sin(ωt4π/3)cos(ωt)cos(ωt2π/3)cos(ωt4π/3)1/21/21/2]xabc

si617_e  (11.68)

where ω is the angular frequency of the ac quantities. For instance, the current vector given by

iabc=[iaibic]=[Isin(ωtφ)Isin(ωt2π/3φ)Isin(ωt4π/3φ)]

si618_e  (11.69)

becomes the vector

idq0=[idiqi0]=[Icos(φ)Isin(φ)0]

si619_e  (11.70)

where I and φ are the amplitude and phase of the line currents, respectively. It can be observed that (a) the zero-component i0 is always zero as the three-phase quantities add up to zero and (b) the d and q components id, iq are dc quantities. Thus, linear controllers should help to achieve zero steady-state error. The control strategy shown in Fig. 11.61 is an alternative where the zero-component controller has been eliminated due to the fact that the line currents at the load side add up to zero.

f11-61-9780128114070
Fig. 11.61 The three-phase VSI. Feedback control based on dq0 transformation.

The controllers in Fig. 11.61 include an integrator that generates the appropriate dc outputs md and mq even if the actual and the line current references are identical. This ensures that the zero steady-state error is achieved. The decoupling block in Fig. 11.61 is used to eliminate the cross coupling effect generated by the dq0 transformation and to allow an easier design of the parameters of the controllers.

The dq0 transformation requires the intensive use of multiplications and trigonometric functions. These operations can readily be done by means of digital microprocessors. Also, analog implementations would indeed be involved.

11.6.3 Feedback Techniques in Current Source Inverters

Duality indicates that CSIs should be controlled as equally as VSIs except that the voltages become currents and the currents become voltages. Thus, hysteresis, linear, and dq linear-based control strategies are also applicable to CSIs; however, the controlled variables are the load voltages instead of the load line currents.

For instance, the linear control of a CSI based on a dq transformation is depicted in Fig. 11.62. In this case, a passive balanced load is considered. In order to show that zero steady-state error is achieved, the per phase equations of the converter are written as

f11-62-9780128114070
Fig. 11.62 The three-phase CSI. Feedback control based on dq0 transformation.

Cddtvabcp=iabcoiabcl

si620_e  (11.71)

Lddtiabcl=vabcpRiabcl

si621_e  (11.72)

The ac line currents are in fact imposed by the modulator, and they satisfy

iabco=iiiabcc

si622_e  (11.73)

Replacing Eq. (11.73) into the model of the converter Eqs. (11.71) and (11.72), using the dq0 transformation and assuming null zero component, the model of the converter becomes

ddtvdqp=Wvdqp+iiCidqc1Cidql

si623_e  (11.74)

ddtidql=Widql+1LvdqpRLidql

si624_e  (11.75)

where W is given by

W=[0ωω0]

si625_e  (11.76)

A first approximation is to assume that the decoupling block is not there; in other words, idqc=mdqsi626_e. On the other hand, the model of the controllers can be written as

mdq=k{vdqp,refvdqp}+1Tt(vdqp,refvdqp)dt

si627_e  (11.77)

where k and T are the proportional and integrative gains of the PI controller that are chosen to achieve a desired dynamic response. Combining the model of the controllers and the model of the converter in dq coordinates and using the Laplace transform, the following relationship between the reference and actual phase-load voltages is found:

vdqp=iiC{sk+1T}{sI+W+RLI}×[{sI+W+RLI}{s2I+s(W+iiCkI)+iiCTI}+sLCI]1vdqp,ref

si628_e  (11.78)

Finally, in order to prove that the zero steady-state error is achieved for step inputs in either the d or q component of the phase-load voltage reference, the previous expression is evaluated in s=0si629_e. This results in the following:

vdqp=iiC{1T}{W+RLI}[{W+RLI}{iiCTI}]1vdqp,ref=vdqp,ref

si630_e  (11.79)

As expected, the actual and reference values are identical. Finally, the relationship in Eq. (11.78) is a matrix that is not diagonal. This means that both the actual and the reference phase-load voltages are coupled. In order to obtain a decoupled control, the decoupling block in Fig. 11.62 should be properly chosen.

11.7 Regeneration in Inverters

Industrial applications are usually characterized by a power flow that goes from the ac distribution system to the load. This is, for example, the case of an ASD operating in the motoring mode. In this instance, the active power flows from the dc side to the ac side of the inverter. However, there are an important number of applications in which the load may supply power to the system. Moreover, this could be an occasional condition and a normal operating condition. This is known as the regenerative operating mode. For example, when an ASD reduces the speed of an electric machine, this can be considered a transient condition. Downhill belt conveyors in mining applications can be considered as a normal operating condition. In order to simplify the notation, it could be said that an inverter operates in the motoring mode when the power flows from the dc to the ac side and in the regenerative mode when the power flows from the ac to the dc side.

11.7.1 Motoring Operating Mode in Three-phase VSIs

This is the case where the power flows from the dc side to the ac side of the inverter. Fig. 11.63 shows a simplified scheme of an ASD where the motor has been modeled by three RLe branches, where the sources eabc are the back emf. Because the ac line voltages applied by the inverter are imposed by the pulse width modulation technique being used, they can be adjusted according to specific requirements. In particular, Fig. 11.64 shows the relevant waveforms in a steady state for the motoring operating mode of the ASD. To simplify the analysis, a constant dc bus voltage vi=Visi631_e has been considered.

f11-63-9780128114070
Fig. 11.63 Three-phase VSI topology with a diode-based front-end rectifier.
f11-64-9780128114070
Fig. 11.64 The ASD based on a VSI. Motoring mode: (A) dc bus voltage, (B) dc bus current, (C) ac line-load voltage, (D) ac phase-load voltage, (E) motor line current and back emf, and (F) shaft power.

It can be observed that (i) the dc bus current ii features a dc value Ii that is positive and (ii) the motor line current is in phase with the back emf. Both features confirm that the active power flows from the dc source to the motor. This is also confirmed by the shaft power plot (Fig. 11.64F), which is obtained as

pl(t)=ea(t)ila(t)+eb(t)ilb(t)+ec(t)ilc(t)

si632_e  (11.80)

11.7.2 Regenerative Operating Mode in Three-phase VSIs

The back emf sources eabc are functions of the machine speed, and as such, they ideally change just as the speed changes. The regeneration operating mode can be achieved by properly modifying the ac line voltages applied to the machine. This is done by the speed outer loop that could be based on a scalar (e.g., V/f) or vectorial (e.g., field-oriented) control strategy. As indicated earlier, there are two cases of regenerative operating modes.

11.7.2.1 Occasional Regenerative Operating Mode

This mode is required during transient conditions such as in occasional braking of electric machines (ASDs). Specifically, the speed needs to be reduced, and the kinetic energy is taken into the dc bus. Because the motor line voltage is imposed by the VSI, the speed reduction should be done in such a way that the motor line currents do not exceed the maximum values. This boundary condition will limit the ramp-down speed to a minimum, but shorter braking times will require a mechanical braking system.

Fig. 11.65 shows a transition from the motoring to regenerative operating mode for an ASD as shown in Fig. 11.63. Here, a stiff dc bus voltage has been used. Zone I in Fig. 11.65 is the motoring mode, zone II is a transition condition, and zone III is the regeneration mode. The line voltage is adjusted dynamically to obtain nominal motor line currents during regeneration (Fig. 11.65D). Zone III clearly shows that the shaft power gets reversed.

f11-65-9780128114070
Fig. 11.65 The ASD based on a VSI. Motoring to regenerative operating mode transition: (A) dc bus current, (B) ac line motor voltage, (C) ac phase motor voltage, (D) motor line current and back emf, and (E) shaft power.

Occasional regeneration means that the drive rarely goes into this operating mode. Therefore, such energy can be (a) left uncontrolled or (b) burned in resistors that are paralleled to the dc bus. The first option is used in low- to medium-power applications that use diode-based front-end rectifiers. Therefore, the dc bus current flows into the dc bus capacitor, and the dc bus voltage rises accordingly to

Δvi=1CIiΔt

si633_e  (11.81)

where Δvi is the dc bus voltage variation, C is the dc bus voltage capacitor, Ii is the average dc bus current during regeneration, and Δt is the duration of the regeneration operating mode. Usually, the drives have the capacitor C designed to allow a 10% overvoltage in the dc bus.

The second option uses burning resistors RR that are paralleled in the dc bus as shown in Fig. 11.66 by means of the switch SR. A closed-loop strategy based on the actual dc bus voltage modifies the duty cycle of the turn-on/turnoff of the switch SR in order to keep such voltage under a given reference. This alternative is used when the energy recovered by the VSI would result in an acceptable dc bus voltage variation if an uncontrolled alternative is used.

f11-66-9780128114070
Fig. 11.66 The ASD based on a VSI. Burning resistor strategy.

There are some special cases where the regeneration operating mode is frequently used. For instance, electric shovels in mining companies have repetitive working cycles, and 15%si634_e of the energy is sent back into the dc bus. In this case, a valid alternative is to send back the energy into the ac distribution system.

The schematic shown in Fig. 11.67 is capable of taking the kinetic energy and sending it into the ac grid. As reviewed earlier, the regeneration operating mode reverses the polarity of the dc current ii, and because the diode-based front-end converter cannot take negative currents, a thyristor-based front-end converter is added. Similarly to the burning resistor approach, a closed-loop strategy based on the actual dc bus voltage vi modifies the commutation angle α of the thyristor rectifier in order to keep such voltage under a given reference.

f11-67-9780128114070
Fig. 11.67 The ASD based on a VSI. Diode-thyristor-based front-end rectifier with regeneration capabilities.

11.7.2.2 Regenerative Operating Mode as Normal operating Mode

Fewer industrial applications are capable of returning energy into the ac distribution system on a continuous basis. For instance, mining companies usually transport their product downhill for a few kilometers before processing it. In such cases, the drive maintains the transportation belt conveyor at constant speed and takes the kinetic energy. Due to the large amount of energy and the continuous operating mode, the drive should be capable of taking the kinetic energy, transforming it into electric energy, and sending it into the ac distribution system. This would make the drive a generator that would compensate for the active power required by other loads connected to the electric grid.

The schematic shown in Fig. 11.68 is a modern alternative for adding regeneration capabilities to the VSI-based drive on a continuous basis. In contrast to the previous alternatives, this scheme uses a VSI topology as an active front-end converter, which is generally called voltage-source rectifier (VSR). The VSR operates in two quadrants, that is, positive dc voltages and positive/negative dc currents as reviewed earlier. This feature makes it a perfect match for ASDs based on a VSI. Some of the advantages of using a VSR topology are the following: (i) The ac supply current can be as sinusoidal as required (by increasing the switching frequency of the VSR or the ac line inductance), (ii) the operation can be done at a unity displacement power factor in both motoring and regenerative operating modes, and (iii) the control of the VSR is done in both motoring and regenerative operating modes by a single dc bus voltage loop.

f11-68-9780128114070
Fig. 11.68 The ASD based on a VSI. Active front-end rectifier with regeneration capabilities.

11.7.3 Regenerative Operating Mode in Three-phase CSIs

There are drives where the motor side converter is a CSI. This is usually the case where near sinusoidal motor voltages are needed instead of the PWM type of waveform generated by VSIs. This is normally the case for medium-voltage applications. Such inverters require a dc current source that is constructed by means of a controlled rectifier.

Fig. 11.69 shows a CSI-based ASD where the dc current source is generated by means of a thyristor-based rectifier in combination with a dc link inductor Ldc. In order to maintain a constant dc link current ii=Iisi456_e, the thyristor-based rectifier adjusts the commutation angle α by means of a closed-loop control strategy. Assuming a constant dc link current, the regenerating operating mode is achieved when the dc link voltage vi reverses its polarity. This can be done by modifying the PWM pattern applied to the CSI as in the VSI-based drive. To maintain the dc link current constant, the thyristor-based rectifier also reverses its dc link voltage vr. Fortunately, the thyristor rectifier operates in two quadrants, that is, positive dc link currents and positive/negative dc link voltages. Thus, no additional equipment is required to include regeneration capabilities in CSI-based drives.

f11-69-9780128114070
Fig. 11.69 The ASD based on a CSI. Thyristor-based rectifier.

Similarly, an active front-end rectifier could be used to improve the overall performance of the thyristor-based rectifier. A PWM current-source rectifier (CSR) could replace the thyristor-based rectifier with the following added advantages: (i) The ac supply current can be as sinusoidal as required (e.g., by increasing the switching frequency of the CSR), (ii) the operation can be done at a unity displacement power factor in both motoring and regenerative operating modes, and (iii) the control of the CSR is done in both motoring and regenerative operating modes by a single dc bus current loop.

Further Reading

Inverters Applications

[1] Huang Chih-Yi, Wei Chao-Peng, Yu Jung-Tai, Hu Yeu-Jent. Torque and current control of induction motor drives for inverter switching frequency reduction. IEEE Trans. Ind. Electron. 2005;52(5):1364–1371.

[2] Huang Chih-Yi, Wei Chao-Peng, Yu Jung-Tai, Hu Yeu-Jent. Torque and current control of induction motor drives for inverter switching frequency reduction. IEEE Trans. Ind. Electron. 2005;52(5):1364–1371.

[3] J. Espinoza, L. Morán, J. Guzmán, Multi-level three-phase current source inverter based AC drive for high performance applications, Conf. Rec. PESC'05, Recife, Brazil, June 2005.

[4] Joós G., Espinoza J. Three phase series var compensation based on a voltage controlled current source inverter with supplemental modulation index control. IEEE Trans. Power Electron. 1999;15(3):587–598.

[5] Jain P., Espinoza J., Jin H. Performance of a single-stage UPS system for single-phase trapezoidal-shaped ac voltage supplies. IEEE Trans. Power Electron. 1998;13(5):912–923.

[6] Akagi H. The state-of-the-art of power electronics in Japan. IEEE Trans. Power Electron. 1998;13(2):345–356.

[7] Wu T., Yu T. Off-line applications with single-stage converters. IEEE Trans. Ind. Appl. 1997;44(5):638–647.

[8] Wu T., Yu T. Off-line applications with single-stage converters. IEEE Trans. Ind. Appl. 1997;44(5):638–647.

[9] Espinoza J., Joós G. A current source inverter induction motor drive system with reduced losses. IEEE Trans. Ind. Appl. 1998;34(4):796–805.

[10] Ryan M., Brumsickle W., Lorenz R. Control topology options for single-phase UPS inverters. IEEE Trans. Ind. Appl. 1997;33(2):493–501.

[11] Jungreis A., Kelly A. Adjustable speed drive for residential applications. IEEE Trans. Ind. Appl. 1995;31(6):1315–1322.

[12] Rajashekara K. History of electrical vehicles in General Motors. IEEE Trans. Ind. Appl. 1994;30(4):897–904.

[13] Bose B. Power electronics and motion control – Technology status and recent trends. IEEE Trans. Ind. Appl. 1993;29(5):902–909.

[14] Bhowmik S., Spée R. A guide to the application-oriented selection of ac/ac converter topologies. IEEE Trans. Power Electron. 1993;8(2):156–163.

Current Source Inverters

[15] J. Espinoza, L. Morán, N. Zargari, Multi-level three-phase current source inverter based series voltage compensator, Conf. Rec. PESC'05, Recife, Brazil, June 2005.

[16] Pande M., Jin H., Joós G. Modulated integral control technique for compensating switch delays and nonideal dc buses in voltage-source inverters. IEEE Trans. Ind. Electron. 1997;44(2):182–190.

[17] Espinoza J., Joós G. Current-source converter on-line pattern generator switching frequency minimization. IEEE Trans. Ind. Appl. 1997;44(2):198–206.

[18] Joós G., Moschopoulos G., Ziogas P. A high performance current source inverter. IEEE Trans. Power Electron. 1993;8(4):571–579.

[19] Loh Poh Chiang, Holmes D.G. Analysis of multiloop control strategies for LC/CL/LCL-filtered voltage-source and current-source inverters. IEEE Trans. Ind. Appl. 2005;41(2):644–654.

[20] Salo M., Tuusa H. Vector-controlled PWM current-source-inverter-fed induction motor drive with a new stator current control method. IEEE Trans. Ind. Electron. 2005;52(2):523–531.

[21] Dong Shen, Lehn P.W. Modeling, analysis, and control of a current source inverter-based STATCOM. IEEE Trans. Power Delivery. 2002;17(1):248–253.

[22] Bendre A., Wallace I., Nord J., Venkataramanan G. A current source PWM inverter with actively commutated SCRs. IEEE Trans. Power Electron. 2002;17(4):461–468.

[23] Han B.M., Moon S.I. Static reactive-power compensator using soft-switching current-source inverter. IEEE Trans. Ind. Electron. 2001;48(6):1158–1165.

[24] Zmood D.N., Holmes D.G. Improved voltage regulation for current-source inverters. IEEE Trans. Ind. Appl. 2001;37(4):1028–1036.

Impedance Source Inverters

[25] Peng Fang Zheng. Z-Source Inverter. IEEE Trans. Ind. Appl. 2003;39(2):504–510.

[26] Peng F.Z., Yuan X., Fang X., Qian Z. Z-source inverter for adjustable speed drives. IEEE Power Electron. Lett. 2003;1(2):33–35.

[27] Vázquez N., Baeza E., Perea A., Hernández C., Vázquez E., López H. ‘Z’ and ‘qZ’ Source Inverters as Electronic Ballast. IEEE Trans. Power Electron. 2016;31(11):7651–7660.

[28] Siwakoti Y.P., Loh P.C., Blaabjerg F., Town G.E. Y-Source Impedance Network. IEEE Trans. Power Electron. 2014;29(7):3250–3254.

[29] Xia C., Li X. Z-source inverter-based approach to the zero-crossing point detection of back EMF for sensorless brushless DC motor. IEEE Trans. Power Electron. 2015;30(3):1488–1498.

[30] Fernao Pires V., Cordeiro A., Foito D., Martins J.F. Quasi-Z-source inverter with a T-type converter in normal and failure moden. IEEE Trans. Power Electron. 2016;31(11):7462–7470.

[31] Beer K., Piepenbreier B. Properties and Advantages of the Quasi-Z-Source Inverter for DC-AC Conversion for Electric Vehicle Applications. In: Anti-Counterfeiting, Security and Identification (ASID), 2012 International Conference on; 2010:1–6.

[32] Yuan L., Peng Z. AC Small Signal Modeling, Analysis and Control of Quasi-Z-Source Converter. In: Power Electronics and Motion Control Conference (IPEMC), 2012 7th International; 2012:1848–1854.

Modulating Techniques and Control Strategies

[33] Espinoza J., Joós G. DSP implementation of output voltage reconstruction in CSI based converters. IEEE Trans. Ind. Electron. 1998;45(6):895–904.

[34] Kazmierkowski M., Malesani L. Current control techniques for three-phase voltage-source PWM converters: A survey. IEEE Trans. Ind. Electron. 1998;45(5):691–703.

[35] Tilli A., Tonielli A. Sequential design of hysteresis current controller for three-phase inverter. IEEE Trans. Ind. Electron. 1998;45(5):771–781.

[36] Chung D., Kim J., Sul S. Unified voltage modulation technique for real-time three-phase power conversion. IEEE Trans. Ind. Appl. 1998;34(2):374–380.

[37] Malesani L., Mattavelli P., Tomasin P. Improved constant-frequency hysteresis current control of VSI inverters with simple feed-forward bandwidth prediction. IEEE Trans. Ind. Appl. 1997;33(5):1194–1202.

[38] Rahman M., Radwin T., Osheiba A., Lashine A. Analysis of current controllers for voltage-source inverter. IEEE Trans. Ind. Electron. 1997;44(4):477–485.

[39] Trzynadlowski A., Kirlin R., Legowski S. Space vector PWM technique with minimum switching losses and a variable pulse rate. IEEE Trans. Ind. Electron. 1997;44(2):173–181.

[40] Tadakuma S., Tanaka S., Naitoh H., Shimane K. Improvement of robustness of vector-controlled induction motors using feedforward and feedback control. IEEE Trans. Power Electron. 1997;12(2):221–227.

[41] Holtz J., Beyer B. Fast current trajectory tracking control based on synchronous optimal pulse width modulation. IEEE Trans. Ind. Appl. 1995;31(5):1110–1120.

[42] Espinoza J., Joós G., Ziogas P. Voltage controlled current source inverters. In: Conf. Rec. IECON'92; USA: San Diego CA; 1992:512–517 November.

[43] Wang Fei. Sine-triangle versus space-vector modulation for three-level PWM voltage-source inverters. IEEE Trans. Ind. Appl. 2002;38(2):500–506.

[44] Tse K.K., Chung Henry Shu-Hung, Ron Hui S.Y., So H.C. A comparative study of carrier-frequency modulation techniques for conducted EMI suppression in PWM converters. IEEE Trans. Ind. Electron. 2002;49(3):618–627.

[45] Shi K.L., Li H. Optimized PWM strategy based on genetic algorithms. IEEE Trans. Ind. Electron. 2005;52(5):1558–1561.

Overmodulation

[46] Hava A., Sul S., Kerkman R., Lipo T. Dynamic overmodulation characteristics of triangle intersection PWM methods. IEEE Trans. Ind. Appl. 1999;35(4):896–907.

[47] Hava A., Kerkman R., Lipo T. Carrier-based PWM-VSI overmodulation strategies: Analysis, comparison, and design. IEEE Trans. Power Electron. 1998;13(4):674–689.

[48] Bae Bon-Ho, Sul Seung-Ki. A novel dynamic overmodulation strategy for fast torque control of high-saliency-ratio AC motor. IEEE Trans. Ind. Appl. 2005;41(4):1013–1019.

[49] Park Hee-Jhung, Youn Myung-Joong. A new time-domain discontinuous space-vector PWM technique in overmodulation region. IEEE Trans. Ind. Electron. 2003;50(2):349–355.

[50] Mondal S.K., Bose B.K., Oleschuk V., Pinto J.O.P. Space vector pulse width modulation of three-level inverter extending operation into overmodulation region. IEEE Trans. Power Electron. 2003;18(2):604–611.

[51] Khambadkone A.M., Holtz J. Compensated synchronous PI current controller in overmodulation range and six-step operation of space-vector-modulation-based vector-controlled drives. IEEE Trans. Ind. Electron. 2002;49(3):574–580.

[52] Narayanan G., Ranganathan V.T. Extension of operation of space vector PWM strategies with low switching frequencies using different overmodulation algorithms. IEEE Trans. Power Electron. 2002;17(5):788–798.

[53] Bakhshai A.R., Joos G., Jain P.K., Jin Hua. Incorporating the overmodulation range in space vector pattern generators using a classification algorithm. IEEE Trans. Power Electron. 2000;15(1):83–91.

Selective Harmonic Elimination

[54] Bowe S., Grewal S. Novel space-vector-based harmonic elimination inverter control. IEEE Trans. Ind. Appl. 2000;36(2):549–557.

[55] Li L., Czarkowski D., Liu Y., Pillay P. Multilevel selective harmonic elimination PWM technique in series-connected voltage inverters. IEEE Trans. Ind. Appl. 2000;36(1):160–170.

[56] Karshenas H., Kojori H., Dewan S. Generalized techniques of selective harmonic elimination and current control in current source inverters/converters. IEEE Trans. Power Electron. 1995;10(5):566–573.

[57] Patel H., Hoft R. Generalized techniques of harmonic elimination and voltage control in thyristor inverters, Part I-Harmonic elimination. IEEE Trans. Ind. Appl. 1973;9(3):310–317.

[58] Wells J.R., Nee B.M., Chapman P.L., Krein P.T. Selective harmonic control: a general problem formulation and selected solutions. IEEE Trans. Power Electron. 2005;20(6):1337–1345.

[59] Newman M.J., Holmes D.G., Nielsen J.G., Blaabjerg F. A dynamic voltage restorer (DVR) with selective harmonic compensation at medium voltage level. IEEE Trans. Ind. Appl. 2005;41(6):1744–1753.

[60] Espinoza J.R., Joos G., Guzman J.I., Moran L.A., Burgos R.P. Selective harmonic elimination and current/voltage control in current/voltage-source topologies: a unified approach. IEEE Trans. Ind. Electron. 2001;48(1):71–81.

Effects of PWM-Type of Voltage Waveforms

[61] Aoki N., Satoh K., Nabae A. Damping circuit to suppress motor terminal overvoltage and ringing in PWM inverter-fed ac motor drive systems with long motor leads. IEEE Trans. Ind. Appl. 1999;35(5):1015–1020.

[62] Rendusara D., Enjeti P. An improved inverter output filter configuration reduces common and differential modes at the motor terminals in PWM drive systems. IEEE Trans. Power Electron. 1998;13(6):1135–1153.

[63] Chen S., Lipo T. Bearing currents and shaft voltages of an induction motor under hard- and soft-switching inverter excitation. IEEE Trans. Ind. Appl. 1998;34(5):1042–1048.

[64] von Jouanne A., Zhang H., Wallace A. An evaluation of mitigation techniques for bearing currents, EMI and overvoltages in ASD applications. IEEE Trans. Ind. Appl. 1998;34(5):1113–1122.

[65] Akagi H., Doumoto T. A passive EMI filter for preventing high-frequency leakage current from flowing through the grounded inverter heat sink of an adjustable-speed motor drive system. IEEE Trans. Ind. Appl. 2005;41(5):1215–1223.

Multilevel Inverters

[66] Tolbert L., Habetler T. Novel multilevel inverter carrier-based PWM method. IEEE Trans. Ind. Appl. 1999;35(5):1098–1107.

[67] Walker G., Ledwich G. Bandwidth considerations for multilevel converters. IEEE Trans. Power Electron. 1999;15(1):74–81.

[68] Liang Y., Nwankpa C. A new type of STATCOM based on cascading voltage-source inverters with phase-shifted unipolar SPWM. IEEE Trans. Ind. Appl. 1999;35(5):1118–1123.

[69] Schibli N., Nguyen T., Rufer A. A three-phase multilevel converter for high-power induction motors. IEEE Trans. Power Electron. 1998;13(5):978–986.

[70] Lai J., Peng F. Multilevel converters – A new breed of power converters. IEEE Trans. Ind. Appl. 1997;32(3):509–517.

[71] Tallam R.M., Naik R., Nondahl T.A. A carrier-based PWM scheme for neutral-point voltage balancing in three-level inverters. IEEE Trans. Ind. Appl. 2005;41(6):1734–1743.

[72] Perez M.A., Espinoza J.R., Rodriguez J.R., Lezana P. Regenerative medium-voltage AC drive based on a multicell arrangement with reduced energy storage requirements. IEEE Trans. Ind. Electron. 2005;52(1):171–180.

[73] Antunes F.L.M., Braga H.A.C., Barbi I. Application of a Generalized Current Multilevel Cell to Current-Source Inverters. IEEE Trans. Ind. Electron. 1999;46(1):31–38.

[74] Barbosa P.G., Carvalho H.A., Barbosa M. do C., Coelho E. Boost current multilevel inverter and its application on single-phase grid-connected photovoltaic systems. IEEE Trans. Power Electron. 2006;21(4):1116–1124.

[75] Vázquez N., López H., Hernández C., Vázquez E., Osorio R., Arau J. A Different Multilevel Current Source Inverter. IEEE Trans. Ind. Electron. 2010;57(8):2623–2632.

Regeneration

[76] Verdelho P., Marques G. DC voltage control and stability analysis of PWM-voltage-type reversible rectifiers. IEEE Trans. Ind. Electron. 1998;45(2):263–273.

[77] Espinoza J., Joós G., Bakhshai A. Non-linear control and stabilization of PWM current source rectifiers in the regeneration mode. In: Conf. Rec. APEC'97; USA: Atlanta GA; 1997:902–908 February.

[78] Hinkkanen M., Luomi J. Stabilization of regenerating-mode operation in sensorless induction motor drives by full-order flux observer design. IEEE Trans. Ind. Electron. 2004;51(6):1318–1328.

[79] Tanaka T., Fujikawa S., Funabiki S. A new method of damping harmonic resonance at the DC link in large-capacity rectifier-inverter systems using a novel regenerating scheme. IEEE Trans. Ind. Appl. 2002;38(4):1131–1138.

[80] Rodriguez J., Pontt J., Silva E., Espinoza J., Perez M. Topologies for regenerative cascaded multilevel inverters”. In: Conf. Rec. PESC'03, Acapulco, Mexico; 2003:519–524.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.139.97.157