Chapter 11
AC–DC–AC Converters for Distributed Power Generation Systems

Marek Jasinski, Sebastian Stynski, Pawel Mlodzikowski and Mariusz Malinowski

Faculty of Electrical Engineering, Warsaw University of Technology, Warsaw, Poland

11.1 Introduction

Indirect AC–DC–AC converters consist of two connected voltage source converters (VSCs) as presented in Figure 11.1. The AC–DC–AC is a connection through a DC-link of two single-phase, three-phase, or multiphase AC circuits with different voltage amplitude c11-math-0001, frequency c11-math-0002, or phase angle c11-math-0003. The major application of AC–DC–AC converters is in adjustable speed drives. However, recently, these converters have begun to play an increasingly important role in distributed power generation systems (DPGSs) and sustainable AC and DC grids.

c11f001

Figure 11.1 AC–DC–AC indirect converter (a) general block scheme (b) photo of commercially available AC–DC–AC converters manufactured by TWERD type: MFC810ACR, (c) photo of commercially available AC–DC–AC converters manufactured by ABB type: PCS6000 [1 2]

There are several possibilities for an AC–DC–AC converter configuration [1–21]. In this chapter, only the most promising bidirectional topologies are presented and discussed. Recently, bidirectional AC–DC–AC converters are available on the market for different voltage levels [1 2].

Both parts of the converter (i.e., AC–DC and DC–AC) can be controlled independently. However, in some cases, there is a need for improving the control accuracy and dynamics. Therefore, it is useful to use an additional link between both control algorithms, which operates as an active power feedforward (APFF). The APFF gives information about the active power on one side of the AC–DC–AC converter to the other side directly, and consequently, the stability of the DC voltage is improved significantly.

11.1.1 Bidirectional AC–DC–AC Topologies

There are several configurations possible for three-phase to three-phase AC–DC–AC full-bridge converters, which can connect two AC systems. The most popular is a two-level converter, as shown in Figure 11.2(a), which is used mostly in low voltage and low-power or medium-power applications, for example, adjustable speed drives. On the other hand, three-level diode-clamped converters (DCCs) [11–12] and flying capacitor converters (FCCs) [4 6, 7] (Figure 11.2(b and c)) are becoming increasingly popular, but usually in the medium-voltage range for medium- and high-power applications [4], for example, marine propulsion, renewable energy conversion, rolling mills and railway traction. There are several advantages of multilevel converters, such as lower voltage stress of components, low current and voltage, total harmonic distortion factor and reduced volume of input passive filters. The main differences among the mentioned multilevel topologies are as follows [4]:

  • DCC is the most popular topology and needs fewer capacitors. However, for higher voltage levels, it requires serially connected clamping diodes, which increases the losses and switching losses. In addition, for higher voltage levels, the DC capacitor voltage balancing cannot be achieved with classical modulations.
  • FCC is less popular because it needs initialization of the FC voltage, and higher switching frequency is required (greater than 1.2 kHz, whereas for high-power applications the switching frequency is usually between 500 and 800 Hz) in high-power applications because of the FC limits, that is, capacitance versus volume.
c11f002

Figure 11.2 Fully controlled three-phase/three-phase, transistor-based AC–DC–AC converter: (a) two-level (2L-3/3), (b) three-level DCC (3L-DCC-3/3) and (c) three-level FCC (3L-FCC-3/3)

Another group of AC–DC–AC converters are simplified topologies obtained by reducing the number of power electronic switches [8–10]. These attempts were based on the idea of replacing one of the semiconductor legs with a split capacitor bank and connecting a one-phase wire to its middle. In simplified topology, the lower number of switching devices, compared with that of a classical three-phase converter, corresponds to a reduced number of control channels and insulated-gate bipolar transistor (IGBT) driver circuits. Thus, connecting a three-phase AC system to a single-phase AC system is possible by using a single-standard three-leg integrated power module as the AC–DC–AC converter, as shown in Figure 11.3(a).

c11f003

Figure 11.3 Simplified AC–DC–AC converters: (a) two-level three-phase/one-phase (S2L-3/1) and (b) two-level three-phase/three-phase (S2L-3/3)

The same concept can be used to simplify an AC–DC–AC converter by connecting one three-phase system to another using only one additional leg, as is shown in Figure 11.3(b). Despite the advantages of these solutions, there is a necessity to develop new modulation techniques and to keep the DC voltage significantly high in order to maintain all nominal phase-to-phase converter voltages, which places higher voltage stress on the converter semiconductor devices.

This problem can be solved by application of a three-level DCC. When this technology was emerging in industry, not a single integrated power module product for three-level devices was available. Today, manufacturers are selling easy-to-use integrated half-bridge power modules with clamping diodes. New compact devices make it easier to improve the topology with a split capacitor in the DC-link. Thus, the improved topology of a simplified AC–DC–AC converter is shown in Figure 11.4, for applications of three-phase to single-phase, as well as for three-phase to three-phase systems. The first topology is dedicated only to low-power applications, whereas the second is devoted to low- and medium-power applications.

c11f004

Figure 11.4 Simplified DCC AC–DC–AC converters: (a) three-phase/one-phase (S3L-DCC 3/1) and (b) three-phase/three-phase (S3L-DCC 3/3)

The simplified three-level DCC has several advantages compared with that of a simplified two-level topology [13]: reduction of machine torque pulsations (mechanical stress in cases where generator/motor application is decreased), additional zero vectors, and reduction in size of passive filters on the AC side.

11.1.2 Passive Components Design for an AC–DC–AC Converter

This section is devoted to the methods of passive components of an AC–DC–AC converter design. Among them, there are input filters (L or LCL), DC-link capacitors, and flying capacitors (FCs), which have a significant impact on the size, weight and final price of the AC–DC–AC converter.

11.1.3 DC-Link Capacitor Rating

In the literature, many design procedures [17–20] for DC-link capacitors are presented, where the minimum capacitance value is designed to limit the DC-link voltage ripple at a specified level. With the assumption of a balanced three-phase system and of ideal power electronics switches, the DC capacitor current can be expressed as:

11.1 equation

where c11-math-0005 is the capacitance of a DC capacitor, c11-math-0006 is the DC voltage, c11-math-0007 is the DC current, c11-math-0008 is the load DC current, c11-math-0009 are letters of AC circuit phase, c11-math-0010 and c11-math-0011 are the AC circuit phase instantaneous current and switching states at the appropriate phases, respectively, and c11-math-0012 is the load active power in DC. The selected approach of the calculation of the DC capacitors is focused on the following constraints:

  • The voltage ripple, because of the high-frequency components of the modulated DC currents of both AC–DC and DC–AC converters, has to remain within desired limits,
  • The capacitor energy has to sustain the output power demand for a period equal to the time delay of the DC voltage control loop.

The final constraint determines the capacitor value. By assuming that c11-math-0013 is a time delay of the DC voltage control loop, and that c11-math-0014 is a variation of the maximal load power, the energy c11-math-0015 exchanged by the DC capacitor can be estimated as:

11.2 equation

Where c11-math-0017 is the sum of small time constant. From this equation, the maximal DC voltage variation during maximum power change is proportional to energy change:

where c11-math-0019 is the considered maximum DC voltage change within the transient load and c11-math-0020 is the minimum DC capacitance. Taking into account the maximal voltage variation c11-math-0021 and by rearranging Equation (11.3), the minimal capacitance can be calculated as [20]:

The minimum DC capacitance value in an AC–DC–AC converter with respect to different nominal powers is presented in Figure 11.5. Therefore, it can be assumed that for given nominal power, the DC capacitance depends mainly on the switching pattern of the AC–DC converters and on the quality (accuracy and dynamics) of the applied control method for the AC–DC–AC converter, that is, for a shorter sampling time c11-math-0023, the DC capacitance would be lower because the DC voltage regulation accuracy is improved. However, it should be noted that the switching frequency is limited by the switching losses of the power electronics devices applied in the AC–DC–AC converter.

c11f005

Figure 11.5 Value of the DC capacitor versus rated power up to 55 kW (Equation 11.4)

11.1.4 Flying Capacitor Rating

FC capacitance, assuming sinusoidal converter output voltages and currents, can be calculated as [21]:

11.5 equation

where c11-math-0025 is the RMS phase current flowing through the FC, c11-math-0026 is the assumed maximum voltage ripple across the FC and c11-math-0027 is the frequency of changes of the switching states (charging and discharging) applied to the FCC, which are used for FC balancing. It is important to note that c11-math-0028 is not the same as the switching frequency, because the frequency of changes of the switching states applied to FCC depends on the type of modulation and, in the case of FC balancing, is typically a multiple of the switching frequency by a factor 0.5, 1 or 2.

11.1.5 L and LCL Filter Rating

An AC–DC–AC converter is connected to the grid through an AC filter (e.g., L or LCL). The inductance allows current control by the voltage drop on itself. In order to obtain better reduction of current ripples, the LCL T-type filter has been introduced.

For the LCL filter to fulfill the grid codes, or recommended practices for example, IEEE 519-1992 [21 22] a variety of optimization criteria exist, such as minimum costs, losses, volume, weight, or design. However, it is a complex task because many parameters influence this process, for example, DC-link voltage level, phase current, modulation method, modulation index, switching frequency, fundamental frequency, and resonance frequency. Therefore, the design can be done using two different methods:

  • simple trial-and-error method, which has to be supported by simulation to prove that the grid codes are fulfilled (simulation has to take into account all elements of the system),
  • complicated iterative, which calculates directly the filter parameters fulfilling the grid code; however, detailed information is required regarding many parameters, for example, modulation method and modulation index [24 25].

A general algorithm of a simple trial-and-error filter design is presented in Figure 11.6. The starting point for the design procedure is usually assumptions of the converter's rated power, grid, and switching frequencies. Those parameters allow the calculation of base values (base impedance c11-math-0029, capacitance c11-math-0030 and inductance c11-math-0031):

11.6 equation
11.7 equation
c11f006

Figure 11.6 General design algorithm of low-pass LCL filter where c11-math-0034 is the converter-side inductance, c11-math-0035 is the grid-side inductance LCL filter, c11-math-0036 is the converter RMS voltage, and c11-math-0037 is the peak-to-peak ripple current

where c11-math-0038 is the line-to-line grid voltage and c11-math-0039 is the grid voltage pulsation.

The first step is the design of the filter inductance. This should be designed carefully because a large inductance value may, by itself, attenuate higher frequencies and limit current ripples. On the other hand, it brings many disadvantages, such as greater inductor size, a decrease in the converter's dynamics and a smaller range of operation. The operation range is limited through the maximum reachable voltage drop across the inductance. It is dependent indirectly on the DC-link voltage. Thus, in order to achieve a large current through the inductance, a high DC-link voltage level or low input inductance is needed.

The maximum total inductance of the input LCL filter may not be larger than [14]:

11.8 equation

where c11-math-0041 is the grid maximum voltage value and c11-math-0042 is the grid current maximum value.

In addition, the total value of the LCL filter inductance should be lower than 5% of the base impedance value, in order to limit the voltage drop during normal operation and to achieve excellent dynamics and reasonable cost [14 26]. Based on the IEEE 519-1992 limitations and the assumption that peak ripples must not exceed 10% of the grid current, the converter-side inductance c11-math-0043 can be calculated, as shown in Figure 11.6 [14 23]. Typically, the converter-side inductance c11-math-0044 is larger than the grid-side inductance c11-math-0045 in order to attenuate most of the current ripple. In this case, a split factor c11-math-0046, given by Liserre et al. [26]:

11.9 equation

is set by the desirable ripple attenuation described as:

11.10 equation

where c11-math-0049 is the grid higher harmonics current and c11-math-0050 is the converter higher harmonics current.

In order to limit the harmonic currents absorbed by the filter, the capacitor is limited to 5% of the reactive power absorbed at rated conditions (base capacitance) [26]. Therefore, the final capacitance of the LCL filter (Y-connected) cannot be larger than as presented in Figure 11.6. In the case of just an c11-math-0051 filter, the value of the inductance can equal c11-math-0052.

Another constraint in the design process of the LCL filter is an obligation to maintain the resonance frequency c11-math-0053 between 10 times the line frequency c11-math-0054 and one-half of the switching frequency c11-math-0055 [14 26]. These limits guarantee that there will be no resonance problems in both the lower and higher parts of the harmonic spectrum. The actual resonance frequency of the LCL filter may be calculated as in Figure 11.6.

11.1.6 Comparison

Table 11.1 presents a comparison of selected classical AC–DC–AC converter topologies: two-level three-phase/three-phase converter (2L-3/3); three-level three-phase/three-phase DCC converter (3L-DCC-3/3); three-level three-phase/three-phase FCC converter (3L-FCC-3/3), and simplified topologies with reduced number of transistors: simplified AC–DC–AC two-level converter three-phase/one-phase (S2L-3/1), and three-phase/three-phase (S2L-3/3); simplified DCC AC–DC–AC converter three-phase/one-phase (S3L-DCC 3/1), and three-phase/three-phase (S3L-DCC 3/3).

Table 11.1 Comparison of selected AC–DC–AC converter topologies

I II III IV V VI VII
2L-3/3 12 0/1 c11-math-0056 1 1 c11-math-0057 c11-math-0058
S2L-3/3 8 0/2 c11-math-0059 c11-math-0060 c11-math-0061 c11-math-0062 c11-math-0063
S2L-3/1 6 0/2 c11-math-0064 c11-math-0065 c11-math-0066 c11-math-0067 c11-math-0068
3L-DCC-3/3 24 12/2 c11-math-0069 1 c11-math-0070 c11-math-0071 c11-math-0072
S3L-DCC 3/3 16 8/2 c11-math-0073 c11-math-0074 c11-math-0075 c11-math-0076 c11-math-0077
S3L-DCC 3/1 12 6/2 c11-math-0078 c11-math-0079 c11-math-0080 c11-math-0081 c11-math-0082
3L-FCC-3/3 24 0/7b c11-math-0083 c11-math-0084 c11-math-0085 c11-math-0086 c11-math-0087

aSimplified converter only between transistors legs.

bIncluding one standard electrolytic capacitor in DC link and six floating capacitors designed for rapid discharging.

I, Total number of IGBT transistors; II, Number of independent diodes/number of capacitors; III, Total number of converter states; IV, DC voltage level in p.u. in comparison to 2L-3/3; V, Blocking voltage for transistors in p.u. in comparison to 2L-3/3; VI, Minimum voltage level step change; VII, total L filter inductance for all phases.

11.2 Pulse-Width Modulation for AC–DC–AC Topologies

Pulse-width modulation (PWM) for AC–DC–AC converters can be divided into two groups, depending on the single-phase or three-phase topology of the converter. The space vector modulation (SVM) techniques for the different topologies of classical single-phase H-bridge AC–DC VSCs have been presented in Chapter 23. In this section, only the SVM techniques for the following are presented:

  • classical three-phase two-level converter
  • classical three-phase three-level converters: DCC and FCC
  • simplified single-phase and three-phase two-level converter
  • simplified single-phase and three-phase three-level DCC

Modulation techniques for VSCs are responsible for the generation of average output voltage (represented by different widths of short voltage pulses) equal to reference voltage with proper balancing of the additional DC voltage sources for multilevel converters. Depending on the VSC topologies and switching frequency, numerous PWM techniques can be listed [27 28]. Among them, two types of PWM are used most often: carrier-based PWM (CB-PWM) and SVM. However, recent advances and the constant development of digital signal processing systems (allowing high-precision digital implementation of complex control algorithms) has meant that SVM has gained a superior position in research and industry. The digital implementation of SVM is characterized by its simplicity. Furthermore, classical SVM with symmetrical switching placement is equivalent to CB-PWM with an additional c11-math-0088 peak amplitude triangular zero sequence signal (ZSS) of the third harmonic frequency [29].

SVM is based on a single- or three-phase circuit representation in a stationary rectangular coordinate system using space vector (SV) theory. The reference SV c11-math-0089 is described by its length and its angle:

11.11 equation

where c11-math-0091 and c11-math-0092 are the c11-math-0093 and c11-math-0094 components of c11-math-0095 in a stationary rectangular coordinate system. To calculate the duration of the VSC switching states, a value of the modulation index is indispensable, which is proportional to the SV length with respect to the DC-link voltage. Usually, c11-math-0096 is defined as:

11.12 equation

where c11-math-0098 denotes the number of VSC output phase voltage levels. The allowable length of vector c11-math-0099 for each c11-math-0100 angle in the linear operation range is given by c11-math-0101. Therefore, the linear modulation range of an c11-math-0102-level VSC is limited to:

11.13 equation

which is the maximal converter linear range of operation [30].

Table 11.2 Switching states for single leg of the two-level VSC

States c11-math-0104 c11-math-0105
c11-math-0106 OFF c11-math-0107
c11-math-0108 ON c11-math-0109

11.2.1 Space Vector Modulation for Classical Three-Phase Two-Level Converter

Table 11.2 presents all possible switching states for a single leg of a two-level VSC (see Figure 11.2(a), generating output pole voltage c11-math-0110 between the phase terminal c11-math-0111 and the DC-link “c11-math-0112” terminal c11-math-0113, where c11-math-0114 is the leg indicator.

Because the three-phase system is assumed symmetric, the Clarke transformation from natural c11-math-0115 into a stationary c11-math-0116 coordinate system can be used. Figure 11.7(a) shows a graphical representation of the SV c11-math-0117 voltage plane with possible output voltage vectors of the three-phase two-level converter. All voltage vectors are described by three numbers, which correspond to the switching states in leg c11-math-0118, leg c11-math-0119, and leg c11-math-0120, respectively. A three-phase two-level converter provides eight possible switching states, comprising six active (c11-math-0121, c11-math-0122, c11-math-0123, c11-math-0124, c11-math-0125, and c11-math-0126) and two zero switching states (c11-math-0127 and c11-math-0128). The active vectors divide the plane into six sectors, where reference vector c11-math-0129 is obtained by switching on (for proper time) two adjacent vectors. It can be seen that vector c11-math-0130 (see Figure 11.7(b)) can be implemented by the different ON/OFF switch sequences of c11-math-0131 and c11-math-0132, and that zero vectors c11-math-0133 and c11-math-0134 decrease the modulation index.

c11f007

Figure 11.7 Space vector representation of three-phase two-level converter: (a) the c11-math-0135 voltage plane and (b) sector 1 with representation of c11-math-0136

Reference vector c11-math-0137, used to solve equations that describe times c11-math-0138, c11-math-0139, c11-math-0140, and c11-math-0141, is sampled with fixed sampling frequency c11-math-0142. Digital implementation is described with the help of a simple trigonometric relationship for the first sector (11.14):

and is recalculated for the following sectors (from 2 to 6). After the c11-math-0144 and c11-math-0145 calculation, the residual sampling time is reserved for zero vectors c11-math-0146 and c11-math-0147 with condition c11-math-0148. Equation (11.14) is identical for all variants of SVM. The only difference is in the different placement of zero vectors c11-math-0149 and c11-math-0150. This gives different equations defining c11-math-0151 and c11-math-0152 for each method, but the total duration time of the zero vectors must fulfill the following conditions:

11.15 equation

The most popular SVM method is modulation with symmetrical zero states' space vector PWM (SVPWM) where:

11.16 equation

Figure 11.8 shows gate pulses for and the correlation between time c11-math-0155, c11-math-0156, and the duration of vectors c11-math-0157, c11-math-0158, c11-math-0159, and c11-math-0160. For the first sector, commutation delay can be computed as:

11.17 equation
c11f008

Figure 11.8 Vector placement in sampling time for three-phase SVM (SVPWM, c11-math-0162) in sector 1

For classical SVM, times c11-math-0163, c11-math-0164, and c11-math-0165 are computed for one sector only. Commutation delay for the other sectors can be calculated with the help of a matrix:

11.18 equation
11.19 equation

11.2.2 Space Vector Modulation for Classical Three-Phase Three-Level Converter

Tables 11.3 and 11.4 present all possible switching states for the single leg of a three-level DCC and FCC (see Figure 11.2(b and c)), respectively, generating output pole voltage c11-math-0168 between the phase terminal c11-math-0169 and the DC-link “c11-math-0170” terminal c11-math-0171, where c11-math-0172 is the leg indicator.

Table 11.3 Switching states for single leg of the three-level DCC

States c11-math-0173 c11-math-0174 c11-math-0175
c11-math-0176 OFF OFF c11-math-0177
c11-math-0178 OFF ON c11-math-0179
c11-math-0180 ON ON c11-math-0181

Table 11.4 Switching states for single leg of the three-level FCC

States c11-math-0182 c11-math-0183 c11-math-0184
c11-math-0185 OFF OFF c11-math-0186
c11-math-0187 A ON OFF c11-math-0188
B OFF ON c11-math-0189
c11-math-0190 ON ON c11-math-0191

Figure 11.9 shows a graphical representation of the SV c11-math-0192 voltage plane with possible output voltage vectors of the three-phase three-level converter, which is constructed from the main switching states: c11-math-0193, c11-math-0194, and c11-math-0195.

c11f009

Figure 11.9 The c11-math-0196 voltage plane for three-phase three-level converter

For the three-phase three-level VSC, the 27 voltage vectors can be specified as follows: three zero vectors (c11-math-0197, c11-math-0198, and c11-math-0199); 12 internal small-amplitude vectors (c11-math-0200, c11-math-0201, c11-math-0202, c11-math-0203, c11-math-0204, c11-math-0205, c11-math-0206, c11-math-0207, c11-math-0208, c11-math-0209, c11-math-0210, and c11-math-0211); six middle medium-amplitude vectors (c11-math-0212, c11-math-0213, c11-math-0214, c11-math-0215, c11-math-0216, and c11-math-0217); six external large-amplitude vectors (c11-math-0218, c11-math-0219, c11-math-0220, c11-math-0221, c11-math-0222, and c11-math-0223).

The external vectors divide the vector plane into six sectors (see Figure 11.9), which is similar to the two-level VSC. However, for the three-level VSC, each sector is divided into four triangular regions. Figure 11.10(a and b) shows that the first sector is divided into four regions with all possible switching states for the DCC and the FCC topologies, respectively. As FCC switching state 1 can be divided into two redundant switching states: c11-math-0224 and c11-math-0225 (see Table 11.4, highlighted in Figure 11.10(b)), for this topology, the number of possible switching states increases to 64.

c11f010

Figure 11.10 First sector with all possible switching states for (a) DCC and (b) FCC topologies

Conventional SVM for multilevel VSC is based on the assumption that for the generation of reference vector c11-math-0226 all possible nearest vectors, including their redundant states, are used (c11-math-0227, c11-math-0228, and c11-math-0229 in the first; c11-math-0230, c11-math-0231, and c11-math-0232 in the second; c11-math-0233, c11-math-0234, and c11-math-0235 in the third; and finally, c11-math-0236, c11-math-0237, and c11-math-0238 in the fourth region) with symmetrical placement of the zero and internal vectors.

The calculation of the region number of the location of the reference vector c11-math-0239 and switching times is based on two additional factors called the small modulation indices (c11-math-0240, c11-math-0241). Index c11-math-0242 and c11-math-0243 are the projection of the reference vector on the sector sides, limited by the external vectors (see Figure 11.10). According to the trigonometric dependence, the small modulation indices are calculated as follows:

11.20 equation

In each sector, calculations are carried out to achieve vector switching times that are the same and where the difference is only in the power switch selection for the gating signal. Thus, the reference vector is normalized to the first sector and after evaluation of the vector switching times, a proper transistor switching sequence is created for the reference position. Table 11.5 presents the calculation of the region number and switching times with respect to c11-math-0245 and c11-math-0246.

Table 11.5 Calculation of region number and switching times

Conditions Region Switching times
c11-math-0247 First c11-math-0248, c11-math-0249, c11-math-0250
c11-math-0251, c11-math-0252, c11-math-0253 Second c11-math-0254, c11-math-0255, c11-math-0256, c11-math-0257
c11-math-0258 Third c11-math-0259, c11-math-0260, c11-math-0261, c11-math-0262
c11-math-0263, c11-math-0264, c11-math-0265 Fourth c11-math-0266, c11-math-0267, c11-math-0268, c11-math-0269

11.3 DC-Link Capacitors Voltage Balancing in Diode-Clamped Converter

Proper operation of the DCC requires that the voltage on each DC-link capacitor is stabilized on ½c11-math-0270. Unbalance in the DC-link gives inequality of redundant-internal vectors and distortions in middle vectors. In such cases, the generated voltage is different from the reference. In the DCC, the balancing of the DC-link capacitors depends on the switching states in all phases. When the phase is clamped to the neutral point of the DC-link (e.g., switching state 1 is chosen), it introduces additional current flowing from or to the neutral point. Consequently, inequalities occur in the charging and discharging of the DC-link capacitors. Other sources of imbalance in the capacitor voltages are asymmetries in the system:

  • propagation of nonlinear IGBT gate signals (e.g., different IGBT ON/OFF times),
  • equivalent series and parallel resistance of DC-link capacitors,
  • differences in electrical parameters, such as each leg semiconductor devices and connection points.

It can be observed that selection of redundant-internal small-amplitude vectors always causes clamping of one or two phases to the neutral point of the DC-link. For example, when internal small-amplitude vector c11-math-0271 (redundant switching states c11-math-0272 and c11-math-0273) is chosen, this means that phase a is clamped to the neutral point of the DC-link for c11-math-0274, and for the same time, phases a, b and c are clamped to the neutral point of the DC-link. Therefore, an additional controller (proportional or proportional-integral (PI)) can be introduced, which depending on the imbalance in the capacitors voltages will change the time division between the redundant-internal small-amplitude vectors. In this simple way, the balancing of the capacitor voltages will be ensured without increasing losses.

For proper balancing of DC-link voltages based on the additional controller, the information on the difference between the capacitor voltages is insufficient [31 32]. The second factor needed is the energy flow direction. The energy flow direction decides whether a selected redundant-internal small-amplitude vector will charge or discharge one of the DC-link capacitors. Therefore, the sign of the energy flow direction c11-math-0275 should be used to determine the sign of the additional controller input error between the DC-link capacitor voltages. If we assume that internal small-amplitude vector c11-math-0276 is used to balance the DC-link voltages with a proportional controller, the ratio of c11-math-0277 charging time division of c11-math-0278 (c11-math-0279) and c11-math-0280 (c11-math-0281) is inversely proportional to the ratio of c11-math-0282 and c11-math-0283 voltages with the sign of the energy flow direction c11-math-0284:

11.21 equation

where c11-math-0286. Because of the sampling period, both of the DC-link capacitors during the calculated time c11-math-0287 will be charged or discharged. In the case of a PI controller in the aforementioned example, the ratio of c11-math-0288 charging time division will be the controller output signal.

Determination of the energy flow direction c11-math-0289 in the case of machine-side VSC (MC) can be done by the determination of one of the following quantities [33]: instantaneous active power c11-math-0290, electromagnetic torque c11-math-0291, torque angle c11-math-0292, or current vector to voltage vector angle c11-math-0293. Choosing the appropriate quantity to use in the modulation algorithm depends on the purpose of the converter and the availability in the control. The active power consumed/produced by the MC can be estimated simply from the reference converter voltages c11-math-0294 and c11-math-0295 (reconstructed from switching states) and the actual stator currents c11-math-0296:

11.22 equation

The sign of the power will define the operating mode of the converter: inverter or rectifier, that is, whether the energy is flowing from the capacitors to the motor, or vice versa. A similar procedure of calculation is used for electromagnetic torque c11-math-0298, which is the result of the vector product of stator flux c11-math-0299 and the stator current c11-math-0300 component c11-math-0301:

11.23 equation

where c11-math-0303 is the number of pole pairs and c11-math-0304 is the number of phases. In this case, the sign of the torque multiplied by mechanical speed c11-math-0305 shows the direction of energy flow. In the case of the torque angle c11-math-0306, no additional calculation is needed; only the sign of the c11-math-0307 is important. Therefore, current transformation to the c11-math-0308 stator flux rotating frame and the sign of c11-math-0309 current determines the direction of energy flow. Table 11.6 presents the final determination of the direction of energy flow in MC.

Table 11.6 Determination of direction of energy flow in MC

Parameter Inverter (motoring) mode (c11-math-0310) Rectifier (regenerating) mode (c11-math-0311)
c11-math-0312 c11-math-0313 c11-math-0314
c11-math-0315 c11-math-0316 c11-math-0317
c11-math-0318 c11-math-0319 c11-math-0320 c11-math-0321 c11-math-0322 c11-math-0323
c11-math-0324 c11-math-0325 c11-math-0326

Table 11.7 Determination of direction of energy flow in GC

Parameter Rectifier mode (c11-math-0327) Inverter mode (c11-math-0328)
c11-math-0329 c11-math-0330 c11-math-0331
c11-math-0332 c11-math-0333 c11-math-0334
c11-math-0335 c11-math-0336 c11-math-0337

The determination of energy flow direction c11-math-0338 in grid-side VSC (GC) is similar and can be done by determination of one of the following quantities [33]: instantaneous active power c11-math-0339, instantaneous active current c11-math-0340, or current vector to voltage vector angle c11-math-0341. The active power consumed/produced by the grid-side converter can be estimated simply from the reference converter voltages c11-math-0342 and c11-math-0343 and the actual line currents:

11.24 equation

Table 11.7 presents the final determination of the direction of energy flow in the line-side converter.

Figure 11.11 presents examples of switching sequences for the DCC in sector 1, assuming that the ratio of the time division of the internal small-amplitude vector is c11-math-0345.

c11f011

Figure 11.11 Switching sequence for DCC reference vector lying in (a) first, (b) second, (c) third, and (d) fourth regions in sector 1

11.3.1 Flying Capacitors Voltage Balancing in Flying Capacitor Converter

Proper operation of the FCC requires that the FC voltage c11-math-0346 should be half that of the DC-link voltage. With this condition, switching state 1 can be divided into two redundant states: 1A and 1B (highlighted in Table 11.4, Figure 11.10(b)), which generate the same output voltage c11-math-0347. As the output voltage does not depend on the type of selected state, either 1A or 1B can be used for independent control of c11-math-0348. However, to decrease the number of commutations, only one state for each phase is chosen in the sampling period. For proper FC voltage balancing, only information about the sign of the phase current flowing through the FC is needed. Table 11.8 shows the selection between redundant states 1A or 1B based on the sign of phase current c11-math-0349.

Table 11.8 Selection of redundant switching state for FC voltage balancing

Conditions c11-math-0350 c11-math-0351
c11-math-0352 c11-math-0353 c11-math-0354
c11-math-0355 c11-math-0356 c11-math-0357

Figures 11.12 and 11.13 present the switching sequence in all regions in sector 1 for different switching states 1A or 1B for all phases, respectively. When the selected redundant state 1A or 1B is changed, the next modulation period can contain additional switching between sampling periods: two in the second and four in the first, third, and fourth regions. These additional switching can damage the converter; all switches are changing their state, which can generate overvoltage. To eliminate this phenomenon and to provide better switching distribution between particular switches, a modification of the switching pattern should be introduced [6] (see Figure 11.14).

c11f012

Figure 11.12 Switching sequence for FCC reference vector lying in (a) first, (b) second, (c) third, and (d) fourth regions in sector 1 for 1A state selection in each phase

c11f013

Figure 11.13 Switching sequence for FCC reference vector lying in (a) first, (b) second, (c) third, and (d) fourth region in sector 1 for 1B state selection in each phase

c11f014

Figure 11.14 Modification of switching pattern in FCC

11.3.2 Pulse-Width Modulation for Simplified AC–DC–AC Topologies

Modulation for simplified AC–DC–AC converters is also realized separately for the DC–AC and AC–DC stages. It can be considered in this case as a single-phase modulation for the DC–AC part, as shown in Figures 11.3(a) and 11.4(a), or as three-phase modulation for the DC–AC or AC–DC parts, as shown in Figures 11.3(b) and 11.4(b).

Single-phase modulation for two-level DC–AC converters is simple and is described by Equation (11.25). Slightly more complicated is the modulation for a single-phase three-level DC–AC converter [13] shown in Figure 11.4(b), because an additional zero state is included (Figure 11.15).

c11f015

Figure 11.15 Vectors generated by 1DM technique DC/AC converter

It can be utilized using a universal concept of time-domain duty-cycle computation technique (11.26) for single-phase multilevel converters, described as a one-dimensional modulation (1DM) technique and presented in Ref. [34].

Modulation for a simplified three-phase converter requires a different approach than a classical topology, because only two reference signals in the c11-math-0360 coordinate system are needed to obtain a balanced three-phase voltage output, which implies that a higher voltage must be supplied to the DC-link to maintain the reference voltage vector in the linear modulation area (Figure 11.16).

c11f016

Figure 11.16 Generalized αβ voltage plane for three-phase simplified converters: imaginary vectors (dotted line), c11-math-0361, vectors achievable with standard DC-link voltage level, c11-math-0362, vectors achievable with high DC voltage, c11-math-0363, reference vectors

SVM for a simplified converter is based on a modified transformation from c11-math-0364 to c11-math-0365 [35]:

Figure 11.17 shows two cases of a graphical representation of a SV c11-math-0367 voltage plane with four achievable voltage vectors of the three-phase two-level simplified converter shown in Figure 11.3(b). In the first case (Figure 11.17(a)), all active vectors (c11-math-0368, c11-math-0369, c11-math-0370, c11-math-0371) divide the c11-math-0372 plane in four sectors: sector I (c11-math-0373), sector II (c11-math-0374), sector III (c11-math-0375), and finally, sector IV (c11-math-0376). The reference voltage vector c11-math-0377, like in classical converters, could be obtained by switching two adjacent vectors, but this causes a voltage imbalance on DC-link capacitors. For the first sector, switching times can be computed as:

11.28 equation
c11f017

Figure 11.17 Space vector c11-math-0379 (c11-math-0380 are defined in Table 11.2) representation of simplified three-phase two-level AC–DC–AC converter: (a) sector I c11-math-0381 and (b) sector I c11-math-0382

The issue of voltage imbalance on DC-link capacitors can be solved by different divisions of the c11-math-0383 plane (Figure 11.17(b)). The number of sectors does not change, but the sectors are shifted 45° (e.g., sector I c11-math-0384), which gives more vectors forming c11-math-0385 (e.g., c11-math-0386 in sector I). For the first sector, switching state times can be computed as [10]:

11.29 equation
c11f018

Figure 11.18 Switching sequence for simplified two-level AC–DC–AC converter in each of four sectors

Switching signals for a simplified three-phase, 2-level AC–DC or DC–AC converter (Fig. 11.3(b)) are shown in Fig. 11.18.

c11f019

Figure 11.19 Space vector c11-math-0388 representation of three-phase three-level simplified converter

Another simplified three-phase converter is the three-level converter, as shown in Figure 11.4(b). A graphical representation of the nine achievable voltage vectors in the c11-math-0389 voltage plane for such a converter is shown in Figure 11.19.

Eight of them (c11-math-0390, c11-math-0391, c11-math-0392, c11-math-0393, c11-math-0394, c11-math-0395, c11-math-0396, c11-math-0397) are active and there is a zero vector (c11-math-0398). The reference voltage vector c11-math-0399, like in classical converters, could be obtained by switching two adjacent vectors with the use of the zero voltage vector in the middle of the switching period [12]. Switching times for the sector I (c11-math-0400) can be computed as:

11.30 equation

and switching sequences for each of the eight sectors are shown in Figure 11.20.

c11f020

Figure 11.20 Switching sequences for three-level simplified converter in each of the eight sectors

11.3.3 Compensation of Semiconductor Voltage Drop and Dead-Time Effect

The voltage drop on semiconductor devices and the dead-time (between pairs of IGBT's) value distorts the phase current. In both cases, it is the result of increased or decreased amplitude (length) c11-math-0402 of the reference SV, depending on the sign of the phase current, respectively. To avoid that, the voltage drop and the dead-time effect compensation algorithms are used. The impact of voltage drop and dead-time effect of semiconductor devices on phase current distortion have been well investigated for the classical two-level and multilevel converters. These methods can be divided into two groups:

  • modification of amplitude (length) c11-math-0403 of the reference SV [6 36–38],
  • modification of the duty cycles at the output of the modulator [6 36, 38–40].

The solution is slightly more complicated in the case of simplified topologies. Because these converters use transistors in only two of three phases, the impact of voltage drop and dead-time effect of semiconductor devices is more significant, compared with that of classical topologies, owing to the different lengths of individual vectors. Moreover, these issues in reduced topologies cannot be considered as a mature. Therefore, in the following section, only reduced topologies will be considered.

Currents for ideal models of controlled switches and models including voltage drop on semiconductor devices are shown in Figure 11.21. In the other case, the converter currents are distorted significantly.

c11f021

Figure 11.21 Ideal phase currents (c11-math-0404, c11-math-0405, c11-math-0406) and phase currents including voltage drop on semiconductor devices (c11-math-0407, c11-math-0408, c11-math-0409)

The effects of voltage drop on semiconductor devices are compensated by adding a correction to the reference signals on the input of the modulator [41]. In order to estimate the values of the voltage drops, the phase currents should be predicted for the next switching instant. Values of the predicted current in c11-math-0410 and c11-math-0411 phases are given by:

11.31 equation

where c11-math-0413 is the value of the current in the present sample and c11-math-0414 is the value of the current in the earlier sample. Predicted current values are compared with values calculated from transistor and diode data sheets to determine the state of elements in both phases and corresponding voltage drops:

11.32 equation

where c11-math-0416 and c11-math-0417 are the currents for the transistor and diode, respectively. c11-math-0418 denotes the voltage drop on the transistor CE junction, c11-math-0419 denotes the voltage drop on the diode, c11-math-0420 and c11-math-0421 are the conductances of the transistor and diode for conduction operation mode, respectively, c11-math-0422 and c11-math-0423 are conductances of the transistor and diode for blocking operation mode, respectively, and finally, c11-math-0424 and c11-math-0425 are the resulting voltage drop for the transistor and diode, respectively.

Using c11-math-0426 and c11-math-0427 transformed to the two-phase coordinate system, as in Equation (11.27), a mean value of voltage drop per sampling period is estimated. Assuming that at switching state 1, the phase voltage error is c11-math-0428 (single transistor and a single DCC diode in a branch are conducting) and in the remaining states, it is c11-math-0429 or c11-math-0430 we can write

11.33 equation

the duty cycles for zero states. Therefore, the correcting signals for voltage drop compensation for a leg c11-math-0432 (and similar b leg c11-math-0433) are:

11.34 equation

where c11-math-0435 denotes the direction of phase current flow, c11-math-0436 and c11-math-0437 are the duty cycles for positive and negative voltage output states and c11-math-0438 denotes zero voltage. These voltage-correcting signals are transformed to the c11-math-0439 coordinate system by using the modified transform from Equation (11.27) and adding them to the reference voltages for the modulator. Currents with no compensation and with the applied compensation method for semiconductor voltage drop only are shown in Figure 11.22.

c11f022

Figure 11.22 Phase currents without (c11-math-0440, c11-math-0441, c11-math-0442) and with (c11-math-0443, c11-math-0444, c11-math-0445) voltage drop compensation

Another necessary compensation relates to the influence of the so-called dead time. Because IGBT devices turn off much more slowly than they turn on, there is a need for the introduction of a delay in control signals between complementary switching devices to prevent short-circuiting the DC-link. This effect is more visible for a low modulation index and low-frequency operation [41]. Figure 11.23 shows a comparison of phase currents without and with dead time included between complementary switching devices.

c11f023

Figure 11.23 Phase currents without (c11-math-0446, c11-math-0447, c11-math-0448) and with (c11-math-0449, c11-math-0450, c11-math-0451) dead time

Correcting signals for limiting the effect of dead time c11-math-0452 are calculated based on current sign and saturation level c11-math-0453; usually chosen experimentally to take into consideration the noise near the current zero-crossing. If phase current polarity, which depends on the measurement noise level as well as on the control delays, is not calculated properly, inadequate compensation can worsen the shape of the phase current. Therefore, in the range of c11-math-0454 the dead-time correction value should be limited linearly. An example of a limiting function, modeled by two linear functions marked as 1 and 2, is shown in Figure 11.24.

c11f024

Figure 11.24 Phase current and correction function for dead-time compensation

Because a reduced converter generates voltages that are not equal in length in vector representation (Figure 11.19), after obtaining the above correction, it is necessary to choose which switching combination should not only be increased or decreased, but also multiplied with a scaling constant. For long vectors (c11-math-0455 and c11-math-0456 for the two-level converter and c11-math-0457 and c11-math-0458 for the three-level converter), the scaling constant is equal to c11-math-0459, but for short vectors (c11-math-0460 and c11-math-0461 for two-level converter c11-math-0462, c11-math-0463, c11-math-0464, c11-math-0465, c11-math-0466, and c11-math-0467 for three-level converter), the scaling constant is equal to c11-math-0468. The effect of compensation for converters without and with dead-time compensation between the complementary switching devices is shown in Figure 11.25. As can be seen, the shape of phase currents c11-math-0469 and c11-math-0470 is improved mainly near the zero-crossing.

c11f025

Figure 11.25 Phase currents without (c11-math-0471, c11-math-0472, c11-math-0473) and with (c11-math-0474, c11-math-0475, c11-math-0476) dead-time compensation between complementary switching devices

11.4 Control Algorithms for AC–DC–AC Converters

A typical application of an AC–DC–AC converter is the interconnection of an electrical machine (motor/generator) with the grid. In this case, advanced control methods should be used to obtain high power quality for the grid-side converter (GC) and good energy utilization for the machine-side converter (MC). Control of both converters can be considered as a dual problem (Figure 11.26) [42–44]. We can distinguish among the most popular control methods voltage-oriented control (VOC) and direct power control-space vector modulation (DPC-SVM) for the GC, as well as field-oriented control (FOC) and direct torque control-space vector modulation (DTC-SVM) for the MC. All these methods are very well described in the literature [42–50]. This section presents the theoretical background of selected control methods and gives basic information regarding the design of the controllers.

c11f026

Figure 11.26 Space vector control methods for grid-side converter (GC) and machine-side converter (MC) in an AC–DC–AC indirect converter

c11f027

Figure 11.27 Indirect field-oriented control (IFOC), where c11-math-0477 is the commanded rotor flux

11.4.1 Field-Oriented Control of an AC–DC Machine-Side Converter

The FOC can be divided into direct field-oriented control (DFOC) and indirect field-oriented control (IFOC). A simplified block diagram of the IFOC is presented in Figure 11.27. The IFOC seems to be more convenient than DFOC, especially for permanent magnet synchronous machines (PMSMs) because flux estimation is not necessary. The IFOC needs the coordinate system to rotate synchronously with the rotor flux angular speed c11-math-0478. In this case, the system of coordinates is oriented with the c11-math-0479 rotor flux linkage component, such that:

11.35 equation

All variables are transformed from the natural c11-math-0481 into rotating the c11-math-0482 reference frame. Then, the referenced stator current values c11-math-0483 and c11-math-0484 are compared with the actual values of stator current component c11-math-0485 and c11-math-0486, respectively. It should be stressed that (for steady state) c11-math-0487 is equal to the magnetizing current, whereas the torque state, both dynamic and steady, is proportional to c11-math-0488. The current errors c11-math-0489 and c11-math-0490 are fed to two PI controllers, which generate referenced stator voltage components c11-math-0491 and c11-math-0492, respectively. Furthermore, referenced voltages are transformed from rotating c11-math-0493 coordinates into the stationary c11-math-0494 coordinates system by using the rotor flux vector position angle c11-math-0495. Therefore, the obtained referenced voltage vector c11-math-0496 is delivered to SVM, which generates the appropriate switching signals c11-math-0497.

It should be taken into account that c11-math-0498 and c11-math-0499 are coupled with each other. Any change of c11-math-0500 stator voltage component has an influence not only for d but also for the q current component (the same applies to the q component). Therefore, a decoupling network (DN) in the control loops is necessary. The solution for this phenomenon is presented in Figure 11.28. It should be noted that the decoupling between the d and q axes appears in all control methods, discussed in this chapter, for both GCs and MCs (VOC and FOC). In the case of direct torque control and direct power control, the SV modulated by the coupling phenomenon is reduced significantly and therefore, DN can be omitted.

c11f028

Figure 11.28 Current control with decoupling network (DN) for AC–DC converter, where c11-math-0501 is the angular frequency of the AC system

11.4.2 Stator Current Controller Design

The model is very convenient to use for the synthesis and analysis of current controllers for MC. In the case of FOC, an asynchronous machine can be approximated and treated as a DC machine [48 51]. However, it should be pointed out that presence of coupling requires an application of a DN, as presented in Figure 11.28.

Hence, it can be seen that referenced decoupled machine converter voltages should be generated (after simplification) as follows:

11.36 equation
11.37 equation

where c11-math-0504, and c11-math-0505, c11-math-0506 is the total leakage factor, c11-math-0507 is the stator windings self-inductance, and c11-math-0508 are the stator and rotor windings resistances, respectively.

It simplifies the analysis and enables the derivation of analytical expressions for the parameters of stator current controllers. The control structure will operate in a discontinuous domain (digital implementation); therefore, it is necessary to take into account the sampling period c11-math-0509. This could be done by S&H block. Moreover, the statistical delay of the PWM in VSC c11-math-0510 should be taken into account. In the literature [6 16, 47 50, 51], the delay of the PWM is approximated from 0 to 2 sampling times c11-math-0511. Furthermore, c11-math-0512 is the VSC gain, and c11-math-0513 is the dead time of the VSC (c11-math-0514 for an ideal converter). The block S&H and VSC sum of their time constants is expressed by:

11.38 equation
c11f029

Figure 11.29 Stator current control loop approximation model where c11-math-0516 is the referenced stator current before prefilter and c11-math-0517 is the prefilter time constant

Therefore, the model of the stator current control loop can be approximated as in Figure 11.29.

Please note that c11-math-0518 is the sum of small time constants, c11-math-0519 is a large time constant, and c11-math-0520 is a gain approximation of the stator windings. Hence, c11-math-0521 gives a dominant pole. Among several methods of analysis, there are two simple ways for the design of the controller parameters: modulus optimum (MO) and symmetry optimum (SO) [48]. With the assumption that the internal voltage induced in the machine winding (EMF) is c11-math-0522, for the circuit presented in Figure 11.29, the proportional gain and integral time constant of the PI current controller can be derived as [48 51]

11.39 equation
11.40 equation

After simplifications, the closed-loop transfer function of the VSC can be approximated by the first-order transfer function as:

11.41 equation

11.4.3 Direct Torque Control with Space Vector Modulation

To avoid the drawbacks of the switching table in classical DTC, instead of hysteresis controllers and the switching table, PI controllers with the SVM block have been introduced, as in IFOC. Therefore, DTC with SVM (DTC-SVM) joins DTC and IFOC features in one control structure, as shown in Figure 11.30.

c11f030

Figure 11.30 Direct torque control with space vector modulation (DTC-SVM) control for machine-side converter

DTC-SVM requires that the coordinate system rotates synchronously with the stator flux angular speed c11-math-0526. In this case, the coordinate system is oriented with the c11-math-0527 stator flux linkage component, such that

11.42 equation

Based on actual stator currents c11-math-0529, DC voltage c11-math-0530, mechanical angular speed c11-math-0531, and switching signals c11-math-0532, the commanded electromagnetic torque c11-math-0533 is delivered from the outer PI speed controller (Figure 11.30). Then, c11-math-0534 and the commanded stator flux c11-math-0535 amplitudes are compared with the estimated values of electromagnetic torque c11-math-0536 and stator flux c11-math-0537. The torque c11-math-0538 and stator flux c11-math-0539 errors are fed to the PI controllers. The output signals are the referenced stator voltage components c11-math-0540 and c11-math-0541, respectively.

Furthermore, voltage components in the rotating c11-math-0542 system of coordinates are transformed into c11-math-0543 stationary coordinates by using c11-math-0544 stator flux position angle. The obtained referenced stator voltage SV c11-math-0545 is delivered to the SVM, which generates the desired switching signals c11-math-0546 for the MC.

11.4.4 Machine Stator Flux Controller Design

Figure 11.31 presents a block diagram of the PI-based stator flux magnitude control loop with neglected voltage drop on the stator resistance [47]. However, the c11-math-0547 delay introduced by the inverter is taken into consideration.

c11f031

Figure 11.31 Block diagram of the stator flux magnitude control loop, where c11-math-0548 is the referenced stator flux before prefilter and c11-math-0549 is the prefilter time constant, which equals the integral time constant of the flux PI controller

According to the SO criterion [48 49], the plant transfer function can be written as:

11.43 equation

Therefore, according to the SO design technique, the stator flux PI controller parameters: proportional gain c11-math-0551 and the integral time constant c11-math-0552 can be calculated as [47 48]

11.44 equation
11.45 equation

11.4.5 Machine Electromagnetic Torque Controller Design

Figure 11.32 presents a block diagram of the simplified PI-based torque control loop with omitted coupling between the torque and stator flux. Consequently, the torque control loop is very simple Thus, for this model, not one criterion can be applied.

c11f032

Figure 11.32 Block diagram of torque control loop, where c11-math-0555 is the referenced electromagnetic torque before prefilter and c11-math-0556 is the prefilter time constant, which equals the integral time constant of the torque PI controller

In this case, according to Ref. [47], the simple (practical) way to design the torque PI controller can be used. Starting from initial values, for example, proportional gain c11-math-0557 and an integral time constant c11-math-0558, the value of c11-math-0559 increases cyclically. From these tests, the best value of c11-math-0560 for fast torque response without oscillation and defined overshoot can be selected.

11.4.6 Machine Angular Speed Controller Design

If the magnitude of the stator flux is constant, c11-math-0561, the dynamics of the IM can be described as follows:

11.46 equation

where c11-math-0563 is the moment of inertia and c11-math-0564 is the load torque.

Accordingly, Figure 11.33 shows a block diagram of the PI-based speed control loop, where c11-math-0565 is the transfer function of the closed torque control loop with prefilter (to compensate the forcing element in the transfer function of the closed torque control loop).

c11f033

Figure 11.33 Block diagram of the speed control loop, where c11-math-0566 is the referenced electromagnetic torque before prefilter and c11-math-0567 is the prefilter time constant, which equals the integral time constant of the speed PI controller

Approximating the torque control loop by the first-order integral part [47], the simplified c11-math-0568 transfer function can be written as:

11.47 equation

where:

11.48 equation

and:

11.49 equation

where c11-math-0572 is the machine pole pairs number, c11-math-0573 is the phase number, c11-math-0574 is the total leakage factor, c11-math-0575 is the stator windings self-inductance, c11-math-0576 is the main magnetizing inductance, c11-math-0577 is the rotor windings self-inductance, and c11-math-0578 are the stator and rotor windings resistances, respectively.

According to the SO criterion [48 49], the plant transfer function can be written as:

11.50 equation

where c11-math-0580 is the gain of the plant and c11-math-0581 is the sum of small time constants and filters:

11.51 equation

The speed PI controller parameters: proportional gain c11-math-0583, and the integral time constant c11-math-0584, can be calculated as [47 48]

11.52 equation
11.53 equation

11.4.7 Voltage-Oriented Control of an AC–DC Grid-Side Converter

The VOC presented in Figure 11.34 guarantees high dynamic and static performance via an internal current control loop. It has become very popular and, consequently, it has been developed and improved [15 53–55]. The goal of the control system is to maintain the DC-link voltage c11-math-0587, at the required level, while currents drawn from the power system should be sinusoidal and in phase with the line voltage in order to satisfy the unity power factor (UPF) condition. The UPF condition is fulfilled when the line current vector c11-math-0588 is aligned with the phase voltage vector c11-math-0589 of the grid.

c11f034

Figure 11.34 Voltage-oriented control (VOC)

For the UPF condition, the commanded value of the reactive component grid current vector c11-math-0590 is set to zero. The reference value of c11-math-0591 is an active component of the grid current vector. After comparing commanded currents with actual values, the errors are delivered to the PI current controllers. Voltages generated by the controllers are transformed to c11-math-0592 coordinates by using the grid voltage vector position angle c11-math-0593. Switching signals c11-math-0594 for the GC are generated by an SVM.

11.4.8 Line Current Controllers of an AC–DC Grid-Side Converter

The model is very convenient to use for the synthesis and analysis of the current controllers for a grid-connected converter. However, the presence of coupling requires an application of a DN (as with FOC in Figure 11.28), as presented in Figure 11.35.

c11f035

Figure 11.35 Current control with decoupling network (DN) for grid-side AC–DC converter

Hence, it can be seen clearly that referenced decoupled GC voltages should be generated as follows:

11.54 equation
11.55 equation

Therefore, the model of the current control loop can be presented as in Figure 11.36.

c11f036

Figure 11.36 Current control loop model where c11-math-0597 is the referenced grid current before prefilter and c11-math-0598 is the prefilter time constant, which equals the integral time constant of the current PI controller

Please note that c11-math-0599 is the sum of small time constants, c11-math-0600 is a large time constant, and c11-math-0601 is a gain of input choke. Hence, c11-math-0602 gives a dominant pole. Between several methods of analysis, there are two simple ways for the design of the controller parameters: MO and SO [48]. With the assumption that disturbance c11-math-0603 for the circuit presented in Figure 11.5, the proportional gain and integral time constant of the PI current controller can be derived as [16]

11.56 equation
11.57 equation

After simplifications, the closed-loop transfer function of the VSC can be approximated by the first-order transfer function as:

For comparison, the parameters of the PI current controller calculated based on MO differ only in integral time constant:

11.59 equation

The VSC with current controllers designed by MO can be approximated as:

11.60 equation

For more information, please refer to Refs. [16 53].

11.4.9 Direct Power Control with Space Vector Modulation of an AC–DC Grid-Side Converter

Direct power control with space vector modulation (DPC-SVM) [52] assures high dynamic and static performance via internal power control loops. Instead of line currents, the active and reactive powers are used as control variables (Figure 11.37). The command active power c11-math-0609 is generated by the outer DC-link voltage controller, whereas command reactive power c11-math-0610 is set to zero for UPF operation. These values are compared with the estimated c11-math-0611 and c11-math-0612 values, respectively [15 16]. Calculated errors c11-math-0613 and c11-math-0614 are delivered to the PI power controllers. Voltages generated by the power controllers are DC quantities, which eliminates steady-state error, as in VOC. Then, after transformation to stationary c11-math-0615 coordinates [16], the voltages are used for the generation of switching signals in the SVM block.

c11f037

Figure 11.37 Direct power control with space vector modulation (DPC-SVM)

11.4.10 Line Power Controllers of an AC–DC Grid-Side Converter

The assumptions are the same as for the design of current controllers in VOC. Because the same block diagram can be applied to both c11-math-0616 and c11-math-0617 power controllers, only the active power c11-math-0618 control loop will be described briefly (Figure 11.38).

c11f038

Figure 11.38 Power control loop with prefilter, where c11-math-0619 is the referenced active power before prefilter, c11-math-0620 is the prefilter time constant, which equals the integral time constant of the power PI controller, and c11-math-0621 is the distortion of the grid voltage, for example, caused by higher harmonics

As in the case of VOC, SO is chosen because of its good response to a disturbance c11-math-0622 step change. With the assumption that c11-math-0623, an open- and closed-loop transfer function can be obtained easily [16]. The proportional gain and integral time constant of the PI current controller in DPC-SVM can be calculated as:

11.61 equation
11.62 equation

The PI controller with such parameters gives power-tracking performance with approximately 40% overshoot. Therefore, for decreasing the overshoot, a first-order prefilter can be used on the reference signal:

11.63 equation

where c11-math-0627 usually equals a few c11-math-0628 [48].

In further investigation, a time delay of the prefilter is set to c11-math-0629. Finally, a closed control loop of the VSC can be approximated by a first-order transfer function as:

11.64 equation

11.4.11 DC-Link Voltage Controller for an AC–DC Converter

For calculation of the parameters of a DC-link voltage controller, the inner current or power control loop can be modeled with the first-order transfer function [16]. Therefore, the power control loop of a grid-connected converter can be approximated by further consideration of the first-order block with an equivalent time constant c11-math-0631 as:

11.65 equation

where c11-math-0633 for inner current or power controllers designed by MO or c11-math-0634 for current/power controllers designed by SO criterion. Therefore, the DC-link voltage control loop can be modeled as in Figure 11.39.

c11f039

Figure 11.39 Block diagram for a simplified DC-link voltage control loop, where c11-math-0635 is the referenced DC voltage before prefilter, c11-math-0636 is the DC voltage prefilter time constant, and c11-math-0637 is the grid-side converter active power

For the sake of simplicity, it can be assumed that

11.66 equation

where c11-math-0639 is the DC voltage filter time constant, c11-math-0640 is a sum of time constants, and c11-math-0641 is an equivalent integration time constant. Hence, the proportional gain c11-math-0642 and integral time constant c11-math-0643 of the DC-link voltage controller can be derived from

11.67 equation
11.68 equation

11.5 AC–DC–AC Converter with Active Power FeedForward

If the accuracy of power matching of an AC–DC GC and DC–AC MC is improved, then a reduction of the DC capacitance is possible, but still providing the same quality of DC voltage stabilization. Because of DC voltage control that is more accurate, the lifetime of the DC capacitor will also be improved. When the DC current c11-math-0646 of a grid converter is equal to the DC current c11-math-0647 of the MC, no current will flow through the DC capacitor C. As a result, the DC voltage will be well stabilized on a desired constant level.

However, despite the very good dynamic behaviors of the presented control methods, the control of the DC voltage could be improved [55–60]. Hence, APFF of the MC side to GC side has to be introduced. The idea of APFF is to give instantaneous power value from the AC MC side directly to the GC active power control loop, faster than the DC voltage feedback control loop. Because control of the power flow between the GC and MC is more precise, the stability of the DC-link voltage would increase significantly. Therefore, the value of the DC capacitance can be reduced significantly, if this is necessary. On the other hand, if the DC capacitance is required to be high for energy storage needs, the APFF can increase the lifetime of the capacitor significantly [57 58].

A simplified block diagram of the AC–DC–AC converter, which consists of a grid converter, machine converter, electrical machine, and APFF, is presented in Figure 11.40. It should be stressed here that in the case of operation with opposite direction power flow, when the MC control stabilizes the DC voltage and GC operates in inverting mode, the opposite direction of APFF is required, especially when standalone operation of GC is considered instead of stiff grid.

c11f040

Figure 11.40 Simplified block diagram of the AC–DC–AC converter, which consists of an AC–DC and DC–AC converter, and APFF. Note: Direction of APFF changes in respect of grid-connected or standalone operation

Hereafter, the AC–DC–AC converter with APFF in control loops will be described. c11-math-0648 is equal to 0 in the case of the UPF operation. Under this condition, the GC input active power is derived by:

11.69 equation

Within a steady-state condition c11-math-0650 and with the assumption that resistance of the input choke c11-math-0651, from the machine side, the power consumed/produced by the MC is

11.70 equation

If virtual flux (VF) [15 16, 53 54] is adopted, it can be noticed that the active power of the GC is proportional to the virtual torque (VT) [16 45, 48]. Thanks to this, a convenient (as in the case of machines) equation for the active grid power calculation is derived:

11.71 equation

where c11-math-0654 are the line current components in a rotating reference frame and c11-math-0655 are the grid VF components in a rotating reference frame.

On the MC, electromagnetic power can be approximated (neglecting power losses) as an active power delivered to the machine c11-math-0656, so

11.72 equation

This is an insufficient assumption because of power losses c11-math-0658 in the real system; therefore, it should be written as

11.73 equation

Let us take into account stand still c11-math-0660 when nominal torque is applied. In such a case, the electromagnetic power will be 0, but the MC power c11-math-0661 will have a nonzero value. Estimation of this power is problematic because the parameters of the machine and power electronics devices are required. Hence, for simplicity of the control structure, a power estimator based on the commanded stator voltage c11-math-0662 and actual stator current c11-math-0663 will be taken into consideration:

11.74 equation

11.5.1 Analysis of the Power Response Time Constant of an AC–DC–AC Converter

Based on Equation (11.58), the time constant of the GC response c11-math-0665 is determined. With the assumption that power losses of the converters can be neglected, the power-tracking performance can be expressed by

Similarly, for the MC, it can be written as

where c11-math-0668 is the equivalent time constant of the MC step response.

11.5.2 Energy of the DC-Link Capacitor

The DC-link voltage can be described as

11.77 equation

So,

Assuming the initial condition is in steady state, the actual DC-link voltage c11-math-0671 is equal to the commanded DC-link voltage c11-math-0672. Therefore, Equation (11.78) can be rewritten as

11.79 equation

where c11-math-0674. Therefore, we can obtain

11.80 equation

If the power losses of the GC and MC are neglected (for simplicity), the energy storage variation of the DC-link capacitor will be the integral of the difference between the GC power c11-math-0676 and the MC power c11-math-0677. Therefore, it can be written as

11.81 equation

From this equation, it can be concluded that for the most accurate control of the GC power c11-math-0679, the commanded power c11-math-0680 is

11.82 equation

where c11-math-0682 describes the power of the DC-link voltage feedback control loop and c11-math-0683 is equal to the instantaneous APFF signal. After the necessary assumptions, we obtain [16]:

11.83 equation

From Equations (11.75) and (11.76), the open-loop transfer function of the grid and machine converter active powers can be expressed as

11.84 equation
11.85 equation

Based on these equations, the analytic model of the AC–DC–AC converter with APFF can be defined, as in Figure 11.41. Such a system can be described by an open-loop transfer function as

11.86 equation
c11f041

Figure 11.41 Simplified transmittances diagram of the AC–DC–AC converter with APFF

Assuming initial steady-state operation, c11-math-0688 and c11-math-0689, the transfer function of the AC–DC–AC converter drive can be derived.

The active power of the machine converter can be calculated from the commanded values of the stator voltages and actual stator current. Assuming the case where c11-math-0690, theoretically, the DC-link voltage should not be affected by the change of the load c11-math-0691 power. However, in the real system, the unbalanced power difference that causes the fluctuations of the DC-link voltage occurs mainly because of the power estimation error, the moment of inertia of the rotor and low-pass filter in the speed control loop, and the saturation of the commanded control variable.

In Figure 11.42, waveforms illustrate the operation with a torque control loop. Torque reversal is applied when the mechanical speed reaches ±71% of the nominal speed. Very good stabilization of DC voltage can be seen with the use of APFF. The selected control method for the AC–DC–AC converter was direct power and torque control with space vector modulations (DPTC-SVM) [16]. The important advantage of this method is that there is no need to rescale the APFF signal.

c11f042

Figure 11.42 Torque reversal from −75% to 75% of nominal torque under DPTC-SVM: (a) only DC-link voltage feedback and (b) with APFF. From the top: DC-link voltage, grid current, active and reactive grid power, electrical machine mechanical speed, and stator current commanded and estimated electromagnetic torque

Moreover, based on Figure 11.42, it can be concluded that for ideal control methods with sufficiently high sampling time, the DC voltage fluctuation during transient for reduced capacitors could be even smaller than for the higher value of capacitance without APFF. Moreover, for the circuit with APFF, it can be seen that the fluctuations of DC voltage during transient for C = 1 p.u. and for C = 0.1 p.u. are almost equal for higher sampling times. Based on this, it can be stated that stabilization of the DC voltage in an AC–DC–AC converter with a controllable GC depends mainly on the quality (control loops accuracy and bandwidths) of the used DC voltage, the grid, and the MC power flow control methods. It can be expected that for high switching frequency, for example, c11-math-069220 kHz, easily available for silicon carbide devices in wider power ranges than in the case of silicon devices, very accurate power flow control will be obtained. A simulation of DC voltage control accuracy in a simplified model (no switching losses) is presented as a function of sampling = switching frequency up to 100 kHz (Figure 11.43) with the DPTC-SVM control method applied [16].

c11f043

Figure 11.43 Module of the DC voltage fluctuation c11-math-0693 as a function of sampling time c11-math-0694, within transient in respect to load change from 0 to 1 p.u. with referenced DC voltage c11-math-0695, and DPTC-SVM control method of the AC–DC–AC converter

11.6 Summary and Conclusions

This chapter has shown the different topologies of AC–DC–AC converters, which becoming increasingly popular in industrial applications, for example, wind turbines, smart grids, drives, and links connecting two electrical networks with different frequency or voltage levels. Among the presented configurations of AC–DC–AC converters, the simplified and classical two-level and multilevel converters are described and compared.

This chapter has also reviewed the modulation and control issues. Among the methods presented in this chapter, the FOC/VOC and DTC-SVM/DPC-SVM are described briefly. Selected issues relating to both the machine-side and grid-side, especially with APFF, have been discussed. Implemented APFFs have no negative impact on system performance in steady states, but do help to reduce DC capacitor and improve dynamic performance of the system. Selected laboratory and simulation tests with DPTC-SVM have been shown.

It is believed that owing to both continuing developments in power semiconductor components and to a reduction of their prices, as well as further elaboration of advanced control and modulation techniques, AC–DC–AC converters will have a significant impact on DPGSs.

References

  1. 1. Power Electronics Division TWERD (Polish: Zaklad Energoelektroniki TWERD) http://www.twerd.pl (accessed 18 December 2013).
  2. 2. ABB www.abb.com (accessed 18 December 2013).
  3. 3. Wu, B., Lang, Y. Zargari., et al. (2011) Power Conversion and control of Wind Energy Systems, John Wiley & Sons, Inc.
  4. 4. Kouro, S., Malinowski, M., Gopakumar, K. et al. (2010) Recent advances and industrial applications of multilevel converters. IEEE Transactions on Industrial Electronics, 57 (8), 2553–2580.
  5. 5. Friedli, T. (2010) Comparative evaluation of three-phase Si and SiC AC-AC converter systems comparative evaluation of three-phase Si and SiC AC-AC converter systems. PhD dissertation. ETH Zurich.
  6. 6. Styński, S. (2011) Analysis and control of multilevel AC–DC–AC flying capacitor converter fed from single–phase. PhD. thesis. Warsaw University of Technology, Warsaw, Poland.
  7. 7. Sędłak, M., Styński, S., Kaźmierkowski, M.P., and Malinowski, M. (2012) Three-level four-leg flying capacitor converter for renewable energy sources. Electrical Review, R.88 (12a), 6–11.
  8. 8. Enjeti, P. and Rahman, A. (1990) A new single phase to three phase converter with active input current shaping for low cost AC motor drives. Industry Applications Society Annual Meeting, 1990, Conference Record of the 1990 IEEE, October 7–12, 1990, Vol. 2, No. 4, pp. 935–942.
  9. 9. Jacobina, C.B., de Rossiter Correa, M.B., da Silva, E.R.C., and Lima, A.M.N. (1999) Induction motor drive system for low-power applications. IEEE Transactions on Industry Applications, 35 (1), 52–61.
  10. 10. Blaabjerg, F., Neacsu, D., and Pedersen, J. (1999) Adaptive SVM to compensate DC-link voltage ripple for four-switch three-phase voltage-source inverters. IEEE Transactions on Power Electronics, 14, 743–752.
  11. 11. Nabae, A., Takahashi, I., and Akagi, H. (1981) A new neutral-point-clamped PWM inverter. IEEE Transactions on Industry Applications, IA-17 (5), 518–523.
  12. 12. Bor-Ren, L. and Ta-Chang, W. (2004) Space vector modulation strategy for an eight-switch three-phase NPC converter. IEEE Transactions on Aerospace and Electronic Systems, 40 (2), 553–566.
  13. 13. Teodorescu, R., Liserre, M., and Rodríguez, P. (2011) Grid Converters for Photovoltaic and Wind Power Systems, John Wiley & Sons, Ltd-IEEE Press, p. 416.
  14. 14. Teichmann, R., Malinowski, M., and Bernet, S. (2005) Evaluation of three-level rectifiers for low-voltage utility applications. IEEE Transactions on Industrial Electronics, 52 (2), 471–481.
  15. 15. Malinowski, M. (2001) Sensorless control strategies for three–phase PWM rectifiers. PhD thesis. Warsaw University of Technology, Warsaw, Poland.
  16. 16. Jasinski, M. (2005) Direct power and torque control of AC–DC–AC converter–fed induction motor drives. PhD thesis. Warsaw University of Technology, Warsaw, Poland.
  17. 17. Carlsson, A. (1998) The back-to-back converter control and design. PhD thesis. Lund Institute of Technology, Lund.
  18. 18. Moran, L., Ziogas, P.D., and Joos, G. (1992) Design aspects of synchronous PWM rectifier-inverter systems under unbalanced input voltage conditions. IEEE Transactions on Industry Applications, 28 (6), 1286–1293.
  19. 19. Winkelnkemper, M. and Bernet, S. (2003) Design and optimalization of the DC-link capacitor of PWM voltage source inverter with active frontend for low-voltage drives. Proceedings of the EPE 2003 Conference.
  20. 20. Malesani, L., Rossetto, L., Tenti, P., and Tomasin, P. (1993) AC/DC/AC PWM converter with minimum energy storage in the dc link. Applied Power Electronics Conference and Exposition, 1993. APEC'93. Conference Proceedings 1993, Eighth Annual, March 7–11, 1993, pp. 306–311.
  21. 21. Fazel, S.S., Bernet, S., Krug, D., and Jalili, K. (2007) Design and comparison of 4-kV neutral-point-clamped, flying-capacitor, and series-connected H-bridge multilevel converters. IEEE Transactions on Industrial Electronics, 43 (4), 1032–1040.
  22. 22. IEEE Std 519–1992. (1993) IEEE Recommended Practices and Requirements for Harmonic Control in Electrical Power Systems.
  23. 23. Bernet, S., Ponnaluri, S., and Teichmann, R. (2002) Design and loss comparison of matrix converters and voltage source converters for modern ac-drives. Transactions of Industrial Electronics Society, Special Edition Matrix Converters.
  24. 24. Jalili, K. (2009) Investigation of control concepts for high-speed induction machine drives and grid side PWM voltage source converters. PhD thesis. Dresden University of Technology.
  25. 25. Jalili, K. and Bernet, S. (2009) Design of LCL filters of active-front-end two-level voltage-source converter. IEEE Transactions on Industrial Electronics, 56 (5), 1674–1689.
  26. 26. Liserre, M., Blaabjerg, F., and Hansen, S. (2001) Design and control of an LCL-filter based three-phase active rectifier. Industry Applications Conference, 2001. Thirty-Sixth IAS Annual Meeting. Conference Record of the 2001 IEEE, September 30–October 4, 2001, Vol. 1, pp. 299–307.
  27. 27. Holmes, D.G. and Lipo, T.A. (2003) Pulse Width Modulation for Power Converters, Principles and Practice, John Wiley & Sons, New York.
  28. 28. Rodriguez, J. (2005) Tutorial on multilevel converters. Proceedings of International Conference PELINCEC 2005, October 2005.
  29. 29. Holmes, D.G. (1996) The significance of zero space vector placement for carrier–based pwm schemes. IEEE Transactions on Industry Applications, 32 (5), 1122–1129.
  30. 30. Jenni, F. and Wueest, D. (1993) The optimization parameters of space vector modulation. Proceedings of EPE 1993, September 1993, pp. 376–381.
  31. 31. Marchesoni, M. and Tenca, P. (2002) Diode-clamped multilevel converters: a practicable way to balance DC-link voltages. IEEE Transactions on Industrial Electronics, 49 (4), 752–765.
  32. 32. Pou, J., Zaragoza, J., Ceballos, S. et al. (2007) Optimal voltage-balancing compensator in the modulation of a neutral-point-clamped converter. Proceedings of IEEE International Symposium on Industrial Electronics, June 2007, pp. 719–724.
  33. 33. Kolomyjski, W. (2009) Modulation strategies for three-level pwm converter-fed induction machine drives. PhD dissertation. Warsaw University of Technology, Warsaw, Poland.
  34. 34. Leon, J., Portillo, R., Vazquez, S. et al. (2008) Simple unified approach to develop a time-domain modulation strategy for single-phase multilevel converters. IEEE Transactions on Industrial Electronics, 55, 3239–3248.
  35. 35. Jacobina, C., da Silva, E., Lima, A., and Ribeiro, R. (1995) Vector and scalar control of a four switch three phase inverter. Industry Applications Conference, 1995. Thirtieth IAS Annual Meeting, IAS'95, Conference Record of the 1995 IEEE, Vol. 3, pp. 2422–2429.
  36. 36. Jeong, S.-G. and Park, M.–.H. (1991) The analysis and compensation of dead-time effects in pwm inverters. IEEE Transactions on Industrial Electronics, 38 (2), 108–114.
  37. 37. Ben-Brahim, L. (2004) On the compensation of dead time and zero-current crossing for a pwm–inverter–controlled ac servo drive. IEEE Transactions on Industrial Electronics, 51 (5), 1113–1118.
  38. 38. Dongsheng, Z. and Rouaud, D.G. (1999) Dead-time effect and compensations of three-level neutral point clamp inverters for high-performance drive applications. IEEE Transactions on Power Electronics, 14 (4), 782–788.
  39. 39. Leggate, D. and Kerkman, R.J. (1997) Pulse-based dead-time compensator for pwm voltage inverters. IEEE Transactions on Industrial Electronics, 44 (2), 191–197.
  40. 40. Oliveira, A.C., Jacobina, C.B., and Lima, A.M.N. (2007) Improved dead-time compensation for sinusoidal pwm inverters operating at high switching frequencies. IEEE Transactions on Industrial Electronics, 54 (4), 2295–2304.
  41. 41. Minshull, S.R., Bingham, C.M., Stone, D.A., and Foster, M.P. (2010) Compensation of nonlinearities in diode-clamped multilevel converters. IEEE Transactions on Industrial Electronics, 57, 2651–2658.
  42. 42. Kazmierkowski, M.P., Krishnan, R., and Blaabjerg, F. (2002) Control in Power Electronics, Academic Press, p. 579.
  43. 43. Wu, R., Dewan, S.B., and Slemon, G.R. (1990) A PWM AC-to-DC converter with fixed switching frequency. IEEE Transactions on Industry Applications, 26 (5), 880–885.
  44. 44. Zhou, D. and Rouaud, D. (1999) Regulation and design issues of a PWM three-phase rectifier. Proceedings of the 25th Annual Conference IECON'99, Vol. 1, pp. 485–489.
  45. 45. Manninen, V. (1995) Application of direct torque control modulation technology to a line converter. Proceedings of the EPE 1995 Conference, pp. 1.292–1.296.
  46. 46. Blaschke, F. (1971) A new method for the structural decoupling of A.C. induction machines. Proceedings of the Conference Record IFAC, Duesseldorf, Germany, October 1971, pp. 1–15.
  47. 47. Zelechowski, M. (2005) Space vector modulated-direct torque controlled (dtc-svm) inverter-fed induction motor drive. PhD dissertation. Warsaw University of Technology, Warsaw, Poland.
  48. 48. Kazmierkowski, M.P. and Tunia, H. (1994) Automatic Control of Converter–Fed Drives, Elsevier, Amsterdam, London, New York, Tokyo, PWN Warszawa.
  49. 49. Levine, W.S. (2000) Control System Fundamentals, CRC Press.
  50. 50. Liserre, M. (2001) Innovative control techniques of power converters for industrial automation. PhD thesis. Politecnico di Bari, Italy.
  51. 51. Blasko, V. and Kaura, V. (1997) A new mathematical model and control of a tree-phase AC-DC voltage source converter. IEEE Transactions on Power Electronics, 12 (1), 116–123.
  52. 52. Malinowski, M., Jasinski, M. and Kaźmierkowski, M.P. (2004) Simple Direct Power Control of Three-Phase PWM Rectifier Using Space-Vector Modulation (DPC-SVM). IEEE Transaction on Industrial Electronics, 51 (2), 447–454.
  53. 53. Wilamowski, B.M. and Irwin, J.D. (2011) The Industrial Electronics Handbook – Power Electronic and Motor Drives, 2nd edn, CRC Press, Boca Raton, FL, London, New York.
  54. 54. Chien, W.S. and Tzou, Y.-Y. (1998) Analysis and design on the reduction of DC-link electrolytic capacitor for AC/DC/AC converter applied to AC motor drives. Power Electronics Specialists Conference, 1998. PESC 98 Record. 29th Annual IEEE Volume 1, May 17–22, 1998, pp. 275–279.
  55. 55. Gu, B.G. and Nam, K. (2002) A DC-link capacitor minimization method through direct capacitor current control. Industry Applications Conference, 2002. 37th IAS Annual Meeting. Conference Record of the Volume 2, October 13–18, 2002, pp. 811–817.
  56. 56. Liao, J.C. and Yen, S.N. (2000) A novel instantaneous power control strategy and analytic model for integrated rectifier/inverter systems. IEEE Transactions on Power Electronics, 15 (6), 996–1006.
  57. 57. EPCOS (0000) Aluminium Electrolic Capacitors – Applications, DataSheet, www.epcos.com (accessed 18 December 2013).
  58. 58. EPCOS (0000) Aluminium Electrolic Capacitors – General Technical Information, DataSheet, www.epcos.com (accessed 18 December 2013).
  59. 59. Sobczuk, D.L. (1999) Application of ANN for control of PWM inverter fed induction motor drives. PhD thesis. Warsaw University of Technology, Warszawa.
  60. 60. Jasinski, M., Rafal, K., Bobrowska-Rafal, M., and Piasecki, S. (2011) Grid interfacing of distributed energy sources by three-level BtB NPC converter under distorted grid voltage. 2011 Workshop on Predictive Control of Electrical Drives and Power Electronics (PRECEDE), October 14–15, 2011, pp. 30–35.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.227.46.229