PHD Thesis PDF
PHD Thesis PDF
PHD Thesis PDF
FACO LTA
PP P AA ' DI
A AA INGE
A PA GNE
P AP RIAA
A DOV
SCUOLA
S CUO LA DI
DI DOTTORATO
DOT TORA TODIDIRICERCA
RICERCAININ
ING EG NER IA INDUSTRIALE
INGEGNERIA INDUS TRIA LE
CICLO: XXII
CICLO: X X II
MODELING
M OD ELIN GAND
AND
ANALYSIS
A N AL YSISOF
OF
MULTIPHASE
M UL TIPH ASE
ELECTRIC
ELEC TR IC MACHINES
M AC H IN ES
FOR
FO R HIGH-POWER
H IGH -PO W ER
APPLICATIONS
A PPLIC A TIO N S
DIRE TT ORE DELLA
DIRETTORE DELLASCUOLA:
SCUO LA:
P rof. P aolo Bar iani
Prof. Paolo Bariani
S UP ERV IS ORE :
SUPERVISORE:
P rof. Roberto
Prof. RobertoMenis
M enis
31 gennaio
31 gennaio 2011
2011
1
Summary
Introduction ............................................................................................................................................................ 8
PART I. Multiphase technology in high-power electric drives and power generation .......... 11
1 Reasons for the multiphase option in the design of high power electric machines ...... 12
2.1 Introduction....................................................................................................................................... 17
3.1.2 Multiphase arrangement selection for the 22500 rpm alternator .................... 41
4 Stator multiphase design and winding construction technology in high power electric
machines................................................................................................................................................................. 49
Part IV. Idealized multiphase machine modeling through Vector-Space Decomposition . 137
11.4 VSD for Symmetrical n-Phase Windings with 360°/n Phase Progression............ 140
11.5 VSD for n-Phase Windings with 180°/n Phase Progression....................................... 142
Appendix......................................................................................................................................................... 144
PART V. Inclusion of space harmonics in multiphase machine modeling through VSD .... 158
13.2 Air-gap field and inductance computation through winding function theory .... 163
13.3.4 Numerical expression of phase inductances for fast computation ................ 175
14.3 VSD of a salient-pole machine with conventional n-phase winding scheme....... 186
PART VI. Analysis of multiphase machines fed by multiple CSIs or multiple VSIs ............... 231
16.3.3 Estimation of internal EMF harmonics due to air-gap flux ............................... 257
16.6 Implications in the Design of High-Power Synchronous Motor Drives ................. 265
Acknowledgements......................................................................................................................................... 271
8
Introduction
Multiphase electric machines are becoming more and more important in today’s high
power applications. To understand this, we can think of an inverter-fed motor, for
instance: increasing its power rating above certain limits, which exceed the maximum
capability of a single inverter unit, necessarily implies using more than one inverter and a
natural consequence of such choice is that the machine stator needs to be equipped with
more than three phases. In addition to power rating enhancement reasons, also reliability
issues play an important role in pushing towards multiphase design options: the fault
tolerance, which is intrinsically guaranteed by a number of phases higher than three, is an
extremely important requirement in many safety-critical applications (such as vehicle
propulsion) and also in many industrial areas where a drive stop following a fault may
cause remarkable economic or material damages.
This work is aimed at giving some possibly useful answers to these (and similar)
questions. Many of the topics addressed have been inspired by the author’s past industrial
experience as an electromagnetic design engineer involved in the development and
analysis of high-power motors and generators. Therefore, also when dealing with
essentially theoretical topics, an effort is spent to highlight how they relate to the
problems that can be encountered in the design and engineering practice. According to
9
The subject matter covered in the thesis is organized into Parts, each including a certain
number of chapters. At the beginning of each Part, a brief summary is proposed which
focus on three elements: the goal (or goals) which that Part is aimed at; the anticipation of
the main results; the comment on how these results relate to the existing literature and on
what they are claimed to bring in terms of new and original contributions.
Part I is mainly introductory and meant to provide a practical overview of the reason why
and on the forms in which multiphase technology is important in today’s high power
electric drives and generation. The topic is covered by addressing some real applications,
most of which falling in the author’s direct engineering experience, chosen as illustrative
examples.
Part III is dedicated to presenting and validating a set of analytical methods that can be
practically used to evaluate the physical parameters in terms of which the machine model
has been expressed. The attention is particularly focused on leakage inductances, whose
number grows as the number of phases increases. The treatment of magnetizing
parameters, in fact, is covered in Part V through the winding function theory.
Part IV introduces the so-called Vector-Space Decomposition (VSD) method for modeling
multiphase machines in their idealized form, i.e. under the hypotheses of uniform air-gap
and sinusoidal winding distribution. The method is revisited with respect to the
references found in the literature giving it a general formulation which applies to all the
number of phases and phase distributions. In particular, the method can be applied to
split-phase machines as an alternative to the approach discussed in Part II.
Part V is probably the most significant and the richest in original results. It deals with the
VSD method extension to the case of non-ideal machines, i.e. to machine topologies with
strongly non-uniform air-gap and non-sinusoidal winding distribution. The purpose
attained is that the lumped-parameter machine model is enriched with some information
about machine geometry which traditionally fall in the finite element analysis domain. The
experimentally proved advantage is obtained that space harmonic effects on machine
operation can be caught through very fast lumped-parameter simulations instead of
resorting to time-consuming transient finite-element analyses.
Part VI, finally, presents some results achieved in the analysis of some particular
phenomena which characterize high-power multiphase machines respectively when
supplied by mutiple Current-Source Inverters and multiple Voltage-Source Inverters. As to
the former case, an analytical-circuit model is presented and experimentally validated
10
PART I. Multiphase
technology in high-power
electric drives and power
generation
This first Part, having an intentionally application-oriented style, is intended to highlight the
importance of multiphase electric machines in the field of high power electric systems,
whether these are variable-frequency drives and or power generation apparatus.
In Chapter 1 the reasons why designers may choose a multiphase configuration when dealing
with high power electric machines are recalled in general terms.
In the rest of the Part, the main advantages brought by adopting a multiphase arrangement,
along with the relevant possible complications and drawbacks, are discussed based on some
industrial application cases falling within the author’s direct experience as a former
electromagnetic designer of this kind of machinery.
Chapter 2, in particular, illustrates some practical features of multiphase machines used as
inverter-fed motors in high-power variable-frequency drives focusing on the quite illustrative
example of a 45 MW synchronous motor fed by four PWM inverters. The main issues involved
in motor development are described, the main design choices justified and the final operating
performance results are presented.
In Chapter 3, multiphase machine application to power generation is covered. The case of a
shipboard power generation system based on a 2 MW high-speed synchronous generator,
implemented both as a 6-phase high-speed and as a 12-phase ultra high-speed prototype , is
taken into account; next, some experimental results on a prototype generator are presented
to practically illustrate in what sense and to what extent the performance of multiphase
alternators may benefit from increasing the number of phases.
Finally, Chapter 4 focuses on a technological aspect (not covered by existing literature)
which is peculiar to large multiphase machinery. It is shown how in a high power multiphase
motor or generator, the number of phases does not only impact on the operating
performance (as it occurs regardless of machine size) but may have remarkable implications
in terms of winding construction technology as well. The thesis is supported and illustrated
by referring to manufactured and tested machines from the author’s design experience.
12 Reasons for the multiphase option in high power electric machines
Quite extensive literature surveys have been recently published on multiphase technology
in electric machines [1], [2] where the advantages of using more than three phases in the
stator winding design have been widely discussed from a general point of view. In this
Section, some introductory considerations will be made on the same topic but focusing on
the case of electric machines for high power applications to empathize the specific reasons
why the multiphase option (in its several design and implementation variants) is
important for such kind of machinery.
Fig .1- 1. (a) Three-phase electric drive; (b) multi-star electric drive; (c) symmetrical polyphase electric drive.
Reasons for the multiphase option in high power electric machines 13
Fig. 1- 2. (a) Multi-phase machine supplied by N n-phase inverters; (b, c) example of 19 MW15-phase induction
motor with N=3 and n=5.
carrying P/n megawatts. In the former case, the electric machine will have a so called
“split-phase” or asymmetrical multiphase winding structure [1], [2], characterized by N
independent three-phase sets suitably displaced in space (the most common solution
entails N sets displaced by 60/N electrical degrees apart). In the latter case, the electric
machine will have a symmetrical polyphase winding configuration, characterized by n
phases symmetrically distributed over the stator by a typical shift angle (“phase
progression”) of 360/n electrical degrees [1], [2].
In high power applications, the solution illustrated by Fig. 1-1b (based on a “multiple star”
machine) is the most widespread because it enables the designer to use existing and
proven three-phase inverter units combined together instead of developing new
polyphase topology with the relevant control algorithms. In fact, in high power
applications, where project risk menagement issues play an important role throughout
system development because of the high investment or project costs, the possibility to rely
on individually proven and assessed subsystem is often regarded as a preferable option.
This does not exclude that other multiphase topologies can be implemented where, for
example, multiple (N) symmetrical n-phase inverters feed a motor equipped with N n-
phase stator winidngs (Fig. 1-2a). An example with n=5 and N=3 (Fig. 1-2b, c) is reported
in [4], dealing with a 19 MW 15-phase ship propulsion induction motor fed by three 5-
phase inverters.
14 Reasons for the multiphase option in high power electric machines
Fig. 1-3. Four shunt-connected PWM inverters supplying a 35 MW motor for LNG applications.
Fig. 1-4. Test facility set-up in Massa Carrara, Italy, during the full load system testing of a 34 MW 3000 rpm
motor supplied by four shunt-connected inverters.
Reasons for the multiphase option in high power electric machines 15
This solution, although viable and possibly successful [5], suffers from the potential risk
that important circulation current harmonics may arise due to the inverters being parallel-
connected. In order to contain such harmonics and make drive performance better than
achievable with an ordinary three-phase design, large thread-decoupling inductors had to
be used along with a suitable PWM control strategy (referred to as “interleaving”, [5]).
It is also intuitive and experimentally proven that the higher the number of stator phases
the less the degradation and the power de-rating that is to be expected following a fault on
a machine phase [6], [7]. Therefore, increasing the number of phases is a provision which
generally increments system fault tolerance, in the sense that it reduces the effect of the
fault in terms of machine performance.
The topic of fault tolerance in multiphase machine design will be further addressed in the
next Chapters from a practical and implementation perspective.
1.3 Performance
It is well known from multiphase machine classical theory [8], [9] that increasing the
number n of stator phases enhances the harmonic content of the air-gap flux density field,
making its waveform closer and closer to the sinusoidal profile as n grows. This can be
easily explained because the harmonic rotating fields sustained by different phase sets in a
multiple star machine undergo a beneficial mutual cancellation effect. A description of this
phenomenon in analytical and quantitative terms in proposed in [10].
The benefits which originate for the better air-gap flux waveform due to a high number of
stator phases are mainly the following:
• Reduction of rotor losses due to flux pulsations and consequent induced eddy
currents in rotor circuits (field, dampers if present) and permanent magnets (if
present).
The former benefit is especially important for high-speed multiphase electric machines
equipped with permanent magnet rotors (permanent magnet eddy current losses tend to
increase as the speed grows) as in the practical case mentioned in 3.1.2. The latter benefit
is crucial in those applications where the multiphase machine is subject to strongly
distorted phase currents [10] and limits are imposed on the maximum allowed torque
ripple [11]. This is the typical case of synchronous machines supplied by Load
16 Reasons for the multiphase option in high power electric machines
1.4 References
[1] E. Levi, R. Bojoi, F. Profumo, H.A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Application, IET, 2007, July
2007, vol. 1, pp. 489-516.
[2] E. Levi, “Multiphase electric machines for variable-speed applications”, IEEE Trans.
on Industrial Electronics, vol. 55, May 2008, pp. 1893-1909.
[3] B. Bose, “Power Electronics and Motor Drives: Recent Progress and Perspective”,
IEEE Trans. on Industrial Electronics, vol. 56, Feb. 2009, pp. 581-588.
[4] T. J. McCoy, “Trends in ship electric propulsion”, IEEE PES Summer Meeting, 2002, 25
July 2002, Chicago, IL, USA, pp. 343-346.
[5] S. Schroder, P. Tenca, T. Geyer, P. Soldi, J. L. Garces, R. Zhang, T. Toma, P. Bordignon,
“Modular high-power shunt-interleaved drive system: a realization up to 35 MW for
oil and gas applications”, IEEE Trans. on Industry Applications, vol. 46, no. 2, Mar.-
Apr. 2010, pp. 821-830.
[6] A. Tessarolo, G. Zocco, C. Tonello, "Design and Testing of a 45-MW 100-Hz
Quadruple-Star Synchronous Motor for a Liquefied Natural Gas Turbo-Compressor
Drive", IEEE Trans. on Industry Applications, in press.
[7] A. Tessarolo, “Experimental performance assessment of multiphase alternators
supplying multiple AC/DC power converters”, Journal of Energy and Power
Engineering, vol. 4, no. 12, Dec. 2010, pp. 43-50.
[8] E. A. Klingshirn, “High phase order induction motors, Part I”, IEEE Trans. on Power
Apparatus and Systems, Vol. PAS-102, no. 1, Jan. 1883, pp. 47-53.
[9] E. A. Klingshirn, “High phase order induction motors, Part II”, IEEE Trans. on Power
Apparatus and Systems, Vol. PAS-102, no. 1, Jan. 1883, pp. 54-59.
[10] A. Tessarolo, “Analysis of split-phase electric machines with unequally-loaded stator
windings and distorted phase currents”, International Conference on Electric
Machines, ICEM 2010, 6-9 Sept. 2010, Rome, Italy, CD-ROM paper RF-013331.
[11] A. Tessarolo, S. Castellan, R. Menis, “Feasibility and performance analysis of a high-
power drive based on four synchro-converters supplying a twelve-phase
synchronous motor”, IEEE Power Electronics Specialists Conference, IEEE PESC 2008,
15-19 June 2008, Rhodes, Greece, pp. 2352-2357.
[12] A. Tessarolo, S. Castellan, R. Menis, G. Ferrari, “On the Modeling of Commutation
Transients in Split-Phase Synchronous Motors Supplied by Multiple Load-
Commutated Inverters”, IEEE Transactions on Industrial Electronics, vol. 57, issue 1,
Jan. 2010, pp. 35-43.
Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry 17
2.1 Introduction
In today’s Oil & Gas industry, Liquefied Natural Gas (LNG) plays a key role due to the
growing demand for this energy source compared to other more polluting fossil fuels [1].
The high-power turbo-compressors employed for gas refrigeration and liquefaction
processes in LNG production plants are traditionally driven by large Gas Turbines (GTs),
whose power may exceed 100 MW, but the use of electrical Adjustable Speed Drives
(ASDs), as GT starters, helpers or substitutes, is a steadily increasing trend as well as a
noticeable technological challenge [2], [3]. Together with wind galleries and feed pumps
for power station steam boilers, turbo-compressors in LNG production plants represent
the applications where electric motor drives of highest power are required and presently
installed. According to a recent survey paper on this topic [2], electric motor compressor
drives with up to 65 MW power rating are presently installed in Norwegian LNG plants.
Due the very large overall output power, a multiphase arrangement is mandatory for such
drives, with a tendency to move towards higher and higher number of phases and towards
increasing levels of expected performance. In order to practically illustrate the design
issues and the technological choices involved in the realization of very high power and
high performance drives based on multiphase motors, the case is examined in this Chapter
of a 45 MW 12-phase motor for LNG industry, whose development (in which the author
was directly involved as electromagnetic designer) inspired many of the investigations
presented in the following of the thesis. The case is deemed illustrative not only for the
motor being one of the largest installed in the world, but also for the wide set of
technological choices made at drive system level which place it on the cutting edge of
multiphase technology for high power and high performance applications.
Fig. 2-1. Natural gas liquefaction process and relevant refrigeration cycle.
motor conception is outlined starting with its genesis from system-level requirements,
throughout the design and development stage, up to the full-load factory testing under
converter supply and the final site commissioning. Presently, the motor is installed and
successfully operating in the word largest LNG refrigeration production plant, located in
Qatar [7], [8]. The experience reported in this Chapter validates the proposed design as an
effective state-of-the-art answer to the most demanding requirements that come from
today’s LNG industry and, possibly, from other industry application fields.
In the specific application considered in this Chapter — which refers to the largest LNG
production plant in the world [7], [8] — the TC power demand exceeds 100 MW and
cannot be delivered by a GT alone, particularly during overloads. Therefore, an ASD is
Fig. 2-3. Concept map representing the flow from requirements to final system and motor design choices.
needed as a “helper”. The resulting arrangement for the complete “refrigeration string” is
depicted in Fig. 2-2, featuring a double-ended shaft inverter-fed electric motor installed
between the GT and the TC. The fundamental role of the electric motor is to integrate the
GT torque so as to fulfill the overall compressor power demand during normal operation;
furthermore, during possible GT faults or depressurizations, the electric motor serves the
purpose to compensate for the transient power loss guaranteeing the operation continuity
of the refrigeration process.
• Maximum power capability. The rated power output of the ASD is 45 MW at 3000
rpm within thermal class B for the electric motor. Additionally, a 130% overload
capability, corresponding to the torque of 60 MW at 3000 rpm, is required for 90 s
to accelerate the string during pressurized starting transients.
• Fault tolerance. The ASD is demanded to continue operating at full load (45 MW at
3000 rpm) even in case of fault on one of the converters. This is crucial for
avoiding expensive string stops due to possible power electronics faults or
malfunctioning. During full-power operation with one out-of-service converter, the
motor is allowed to exceed the limits of thermal class F.
damage TC blades [5], [7], [8]. A torque ripple amplitude lower than 1% peak-to-
peak is established as a target.
• Line-side harmonic distortion. Due to the weakness of the on-site electric grid, the
Total Harmonic Distortion (THD) of the line-side voltage is required not to exceed
3% so as to minimize disturbances to the other loads connected to the same
network.
• Total transferred torque. Because of its particular location between the TC and the
GT, the motor shaft shall be capable of transferring the total torque, equal to the
sum of motor’s (143 kNm) and GT’s (347 kNm) torques.
• Lateral dynamics performance and reduced bearing span. Due to the overall shaft
line length (Fig. 2-2), lateral vibrations constitute a very delicate issue. Therefore,
motor-generated lateral exciting forces (vibrations) are to be minimized to avoid
lateral resonance issues [5] and the bearing span (double-ended shaft) it to be kept
as small as possible.
Fig. 2-4. (a) Three-phase converter topology; (b) single cell structure.
Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry 21
Fig. 2-5. Individual converter power rating Pmax per requirements 1 and 2 as a function of the number of
converters (N) supplying the motor.
makes it possible to attain some important goals like: a voltage output that approaches the
sinusoidal waveform as the number of cells is increased [10]; possibility to operate the
motor at unity power factor [10] (with consequent reduction in both stator and excitation
currents compared to the case of LCI supply); fault tolerance with respect to single cell
faults thanks to the implementation of a faulty-cell by-pass function [9]; grid-side
performance with high power factor, low harmonic injection and possibility of
regenerative operation through AFE’s [10].
Actually, alternative VSI converter technologies could be chosen to obtain similar ratings
and performance. For example, [11] and [12] report on a VSI drive solution, conceived for
moving LNG compressors up to 35 MW, where four shunt-connected three-level Neutral
Point Clamped (NPC) converters with IGCT switches are used to supply a three-phase
(single-star) motor. In this solution, a PWM technique (referred to as “interleaving”, [11])
is employed to enhance motor voltage and current quality by a suitable switching
coordination among the four shunt-connected inverters. Experimental results reported in
[12] on a 32 MW prototype show how nearly sinusoidal motor voltage and current
waveforms can be practically attained in this realization along with a torque ripple
amplitude lower than 4%. Such amplitude exceeds the 1% target pursued in the project
being covered in this Chapter (Section 2.3), but can be still considered satisfactory due to
the very high frequency of torque harmonic components compared to the expected shaft
line resonance frequencies [12].
Returning to the application addressed in this Chapter, the final decision on what
converter solution to adopt does not imply only performance considerations, but is also
affected by the supplier’s know-how and industrial experience gained through previous
similar realizations. This accounts for the cascaded-cell topology to be finally selected.
As concerns the selection of the rated motor-side voltage, a key reason for limiting it to 7.2
kV regards stator winding insulation. In fact, higher voltage ratings would imply larger
dielectric stresses in terms of voltage spikes and dv/dt values; additionally, they would
require a thicker ground insulation wall, thus leading to a lower slot fill factor and lower
heat transfer coefficient by conduction from coil conductors to surrounding materials.
N of converters (that is the power segmentation level) is mainly dictated by the maximum
power and the fault tolerance targets. In fact, calling N the number of converter units to be
employed, each unit is demanded (see Section 2.3) to deliver a maximum power of:
(1)
Pmax = 60 / N (MW) (2-1)
On the other side, as discussed in Section 2.3 the rated power of 45 MW shall be delivered
even in case of an out-of-service converter. Therefore, the rated power per converter unit
shall be at least:
(2)
Pmax = 45 /(N − 1) (MW) (2-2)
Finally, N is to be chosen so that the power capability of a single inverter can be realized
with the chosen power electronics technology (2.4.1)
If we plot the converter power capabilities given by (2-1) and (2-2) as functions of N (Fig.
2-5), we can see that for N=4 the two equations lead to the same converter sizing of 15
MVA (15 MW at unity power factor). This rating can be practically realized with the
Fig. 2-6. Three-phase cascaded converters with n cells (C1..Cn) per phase supplying a four-star motor through
output reactors L.
Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry 23
Fig. 2-7. Phase belt arrangement over a pole span: (a) single three-phase winding; (b) quadruple-star winding.
Angles are in electrical degrees.
Choosing N=4 converter units (or more if necessary) is possible thanks to the VSI
technology chosen for motor supply. In fact, with LCI-based drives, increasing the number
of supplying converters above two would be theoretically feasible but would imply the
risk of commutation overlapping issues [13].
The structure of the chosen quadruple-star winding is illustrated in Fig. 2-7. Practically, it
can be obtained from a usual three-phase winding (Fig. 2-7a) by splitting phase belt A into
four sub-belts called A1..A4 (each spanning 15 electrical degrees), phase belt B into B1..B4
and phase belt C into C1..C4. At this point, belts A1, B1, C1 are star-connected to form the first
star, A2, B2, C2 to form the second, and so forth. The four stars thus obtained (Fig. 2-7b) are
naturally displaced by 15 electrical degrees apart in space and suitable for being supplied
by voltage systems mutually shift by 15 electrical degrees in time.
As outlined in Fig. 2-3, the use of a quadruple-star winding design is also beneficial for
meeting the low torque ripple requirement. In fact, it is known that with such a winding
scheme, should phase currents contain time harmonics of orders 5, 7, 11, 13, 17, 19, the
resulting space harmonic fields in the air gap would be globally null due to mutual
cancellation effects [14]. This means that possible phase current distortions, due to
harmonic components with orders below the 23rd (see Section 2.5.7), do affect the air-gap
field and torque.
24 Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry
Fig. 2-8. Sketches of a 4-pole (left) and 2-pole (right) rotor designs.
Fig. 2-9. Pre-formed coils and their assembly in the stator core.
Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry 25
Fig. 2-10. Picture of the motor during installation. The four terminal boxes (one per star) can be seen on motor
side.
frequency can be easily obtained with the selected PWM VSI technology [6], [9]. The main
concern, in the specific case under discussion, relates to possible thermal issues due to the
increased magnetic losses inside the stator core. In each point of the stator core, in fact,
the specific power dissipation due to magnetic losses can be approximated by Steinmetz
formula as:
turns. As the rated phase current of the machine increases, the current flowing though
each turn also grows, calling an increasingly large turn section area. Above certain power
levels, the turn cross section is so large that only one turn per coil should be used. This
implies abandoning the formed coil winding technology for Roebel bars [18].
Manufacturing and assembly of Roebel bars require a dedicated expensive technology (for
strand transposition and end-turn connections above all) which is typical of large
turboalternators [18]. Whenever possible economic and technological reasons thus
encourage to avoid Roebel bars and retain multi-turn coil windings. For this purpose, the
adoption of a high phase order winding design is beneficial Errore. L'origine riferimento
non è stata trovata.. In fact, the higher the number of phases, the smaller the current per
phase, assuming a fixed motor power rating.
In the application under study, the total power of 45 MW at 7200 V would give a current
higher than 3 kA if a single three-phase design were used, resulting in a Roebel bar
construction Errore. L'origine riferimento non è stata trovata.. Conversely, the decision
to split the winding into four stator sets reduces the phase current to one forth, which is
compatible with a coil winding technology (Fig. 2-9). The adoption of the coil winding
technology is also facilitated by the rotor four-pole design (2.4.4), which helps shorten the
stator core length and then reduce the coil axial dimension with benefits in assembly and
impregnation processes.
A wide testing campaign is performed both on the motor alone and under inverter supply.
The main testing and site commissioning achievements are reported next.
The no-load test, conducted at rated speed, gives the open-circuit curve reproduced in Fig.
2-11. It can be seen that the rated voltage (7200 V) is located in a slightly saturated region
of the no-load characteristic. This is due to the relatively low tooth and yoke flux density
values decided in the design stage to avoid excessive magnetic losses at 100-Hz supply
frequency (see 2.4.5).
Based on the no-load and short-circuit tests, the various loss contributions can be
segregated in accordance with [19]. This leads to the loss diagram reported in Fig, 12. It
can be seen that the majority of the losses comes from friction and windage, as often
happens in high-power high-speed machinery. The large power dissipation due to
mechanical losses is partly due to the strong cooling air flow produced by the shaft-
mounted fans and is therefore associated with a good internal cooling effectiveness, as
confirmed by the motor thermal behavior at full load (see 2.5.3). The amount of core
losses, on the other side, proves to be well under control despite of the relatively high
rated frequency (100 Hz).
The computation of the total losses according to [19] brings to a conventional efficiency of
98.1%.
The power flow during the back-to-back test is illustrated in Fig. 2-13. The four converters
supplying the motor machine draw the power from the testing facility grid; the power is
next transferred from the motor to the generator through the mechanical coupling; finally
the generator machine feeds its four converters which return the power to the grid by
Fig. 2-12. Loss segregation from tests on the individual machine (loss values are in kilowatts).
28 Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry
Fig. 2-13. Back-to-back arrangement for full-load motor testing. Arrows indicate the power flow direction.
means of their AFE’s. From an overall power balance, it is clear that most of the power
involved in the test flows in a closed loop, while the testing facility grid is demanded to
supply only a power equal to the total losses of the system under test.
The most significant results of the back-to-back testing campaign are summarized in
Tables II-IV and discussed in the next subsections.
TABLE II
THERMAL MEASUREMENTS DURING BACK-TO-BACK TESTING
Number of converters 4 3
Temperature rise 59 K 78 K
TABLE III
VIBRATION MEASUREMENTS DURING BACK-TO-BACK TESTING
TABLE IV
TORQUE RIPPLE MEASUREMENTS DURING BACK-TO-BACK TESTING
Fig. 2-14. Back-to-back test arrangement. (a) Motor and generator machines mechanically coupled. (b) Coupling
with torque transducers.
30 Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry
Fig. 2-15. Speed profile and vibration measurements during the maximum-torque acceleration test.
mechanically damped while being transferred from the air-gap to the shaft through the
solid steel rotor body, so that the torque ripple measured on the shaft is lower than 1% of
the rated torque, which satisfies the initial design target (see Section 2.3).
As well as the start-up and validation activities performed on turbines and compressors,
the commissioning of electric drives involves several technical challenges. In particular,
electric power oscillations between the ASD and the network have to be thoroughly
monitored on site in any operating conditions and for different network configurations;
this is necessary to make sure that such oscillations do not excite the natural vibration
modes of the string resulting in mechanical resonance issues [8].
The ASD performance is successfully assessed on site in all the required operating modes
[8], i.e. as a helper (allowing for turbine unloading without string shut down), as a starter
32 Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry
Fig. 2-17. Picture of the motor in its final position between the turbine and the compressor.
(accelerating the string in case of sudden unpredicted turbine trip) and as a generator
(feeding the local network when turbine power exceeds compressor demand).
2.7 Conclusions
LNG production plants are one of the industrial areas where electric drives with the
highest power ratings in the world, up to several tens of megawatts, are demanded today.
A consolidated technology for such drives, based on dual-star two-pole LCI-fed
synchronous motors with supply frequencies between 50 and 80 Hz, has been steadily
employed for decades. A new electric drive solution, based on a VSI-fed quadruple-star
100-Hz 4-pole synchronous motor, rated 45 MW at 3000 rpm, is presented in this Chapter
as a high-performance and high-reliability alternative. The entire process from
specifications through the design and development stage up to the final system testing and
commissioning is outlined. The design solution adopted is proved to successfully meet the
demanding requirements for which it has been conceived. Particular advantages over
traditional LCI-based solutions are highlighted in terms of torque ripple (lower than 1%
peak-to-peak), very low vibrations thanks to the four-pole design, high fault-tolerance due
to the 4-star 4-converter topology, high motor efficiency (98.1%).
Thanks to the large number of phases (12) and the 4-pole design, the stator winding can
be implemented with coil (instead of Roebel bar) technology with noticeable
manufacturing and cost benefits. The 100 Hz supply frequency is shown not to give
excessive core losses. The 5th and 7th current harmonic distortion, observed in stator phase
currents as a consequence of motor internal EMF’s, are proved not to negatively impact on
Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry 33
Fig. 2-18. Field distributions produced by phase current harmonics of orders: (a) 1st (fundamental); (b) 5th ; (c)
7th; (d) 11th.
torque performance thanks to the mutual cancellation effects of the relevant air-gap fields
assured by the quadruple-star winding design.
2.8 Appendix
To investigate the distortion phenomena observed in stator currents and its effects on the
motor torque performance, a 2D motor model is built in the Ansoft/Maxwell environment
(Fig. 2-18a). External circuits are added to account for phase resistances (R) and external
reactors (L) and perfectly sinusoidal supply voltages are applied to the 12 machine
terminals. A time-stepping FE simulation is then run. The current waveforms obtained in
steady-state conditions well match measured currents (Fig. 2-16). This proves that the
current distortion does not depend on converter non-idealities or control (which are not
34 Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry
included in the model), but originates from the motor internal back EMF’s. In a
synchronous motor (even with a round rotor) small back-EMF 5th and 7th order harmonics
are in fact inevitable due to the non-perfectly-uniform air-gap permeance [21].
On the other side, time-stepping calculations also show that the current distortion under
discussion does not have detrimental effects on the motor torque ripple (Fig. 2-17b),
whose computed amplitude is lower than 0.25% p-p according to simulation results.
2.9 References
[1] Zuyi Li, “Natural gas for generation”, IEEE Power and Energy Magazine, July/Aug.
2005, pp. 16-21.
[2] A. Sannino, T. Nestli, M.Kjall-Ohlsson, P.E. Holsten, “All-electric LNG liquefaction
plants”, IEEE IAS Annual Meeting 2007, 23-27 Sept. 2007, New Orleans, Louisiana,
USA, pp. 2407-2413.
[3] B. Martinez, C.B. Meher-Homji, J. Paschal, A. Eaton, “All-electric motor drives for
LNG plants”, Gastech 2005, 14-17 March 2005, Bilbao, Spain.
[4] G.J. Neidhofer, A.G. Troedson, “Large converter-fed synchronous motors for high
speeds and adjustable speed operation: design features and experience”, IEEE
IEMDC ’97, 18-21 May 1997, Milwaukee, WI, USA, pp. MA2/6.1- MA2/6.3.
[5] H. E. Albright, “Application of large high-speed synchronous motors”, IEEE Trans. on
Industry Applications, vol. IA-16, Jan./Feb. ’80, pp. 134-143.
[6] B. Bose, “Power electronics and motor drives: Recent progress and perspective,” IEEE
Trans. Ind. Electron., vol. 56, no. 2, pp. 581–588, Feb. 2009.
[7] R. Salisbury, P. Rasmussen, T. Griffith, A. Fibbi, “Design, manufacture and test
campaign of the world’s largest LNG refrigeration compressor strings”, LNG Journal,
July-August 2007, PS2.2.1-22.
[8] S. Judd, R. Salisbury, P. Rasmussen, P. Battagli, D. Mosier, A. Smith, “Successful Start-
up and Operation of GE FRAME 9 Gas Refrigerant Strings”, 16th International
Conference and Exhibition on Liquefied Natural Gas, 18-21 Apr. 2010, Oran, Algeria,
paper PS4-1.
[9] J. Rodriguez, P.W. Hammond, J. Pontt, R. Musalem, P. Lezana, M.J. Escobar,
“Operation of a medium-voltage drive under faulty conditions”, IEEE Trans. on
Industry Applications, vol. 52, Aug. 2005, pp. 1080-1085.
[10] M. Rastogi, P.W. Hammond; R.H. Osman; “High performance, high reliability medium
voltage drives”, IEEE PEDS 2001, 22-25 Oct. 2001, Indonesia, pp. 259-264.
[11] S. Schroder, P. Tenca, T. Geyer, P. Soldi, J. L. Garces, R. Zhang, T. Toma, P. Bordignon,
“Modular high-power shunt-interleaved drive system: a realization up to 35 MW for
oil and gas applications”, IEEE Trans. on Industry Applications, vol. 46, no. 2, Mar.-
Apr. 2010, pp. 821-830.
[12] R. Baccani, R. Zhang, T. Toma, A. Iuretig, M. Perna, “Electric systems for high power
compressor trains in oil and gas applications—System design, validation approach
and performance”, 36th Turbomachinery Symposium, 10-13 Sept. 2007, Huston,
Texas, pp. 61-68.
[13] A. Tessarolo, S. Castellan, R. Menis, G. Ferrari, “On the modeling of commutation
transients in split-phase synchronous motors supplied by multiple Load-
Commutated Inverters”, IEEE Trans. on Industrial Electronics, vol. 57, no. 1, Jan.
2010, pp. 35-43.
[14] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, and S. Williamson, “Multiphase induction
motor drives—A technology status review,” IET Elect. Power Appl., vol. 1, no. 4, Jul.
2007, pp. 489–516..
[15] V. B. Honsinger, “Sizing equations for electrical machinery”, IEEE Trans. on Energy
Conversion, vol. EC-2, no. 1, March 1987, pp. 116-121.
Multiphase motor drives: the case of a 45 MW 12-phase motor for LNG industry 35
High power multiphase machines have important applications not only in electric drives
but also in electric power generation systems. Unlike three-phase alternators, which can
be directly connected to the grid (possibly through a transformer), multiphase generators
are typically connected to a suitable power electronic equipment, which can be a set of
rectifiers supplying DC grids or a set of AC/AC converters supplying AC networks. The
former case is definitely the most frequent. In many isolated power systems, like in marine
and aeronautical applications, in fact, a DC distribution network is employed. One of its
advantages over AC distributions systems, which are characterized by a fixed-frequency
operation, is the possibility for electric generators to rotate at different speeds depending
on the load. This enables to minimize the fuel consumption and polluting emissions of
prime movers, which can always operate in the neighborhood of their maximum-efficiency
working point [1].
Next, the topic is addressed in more general terms presenting a set of tests performed on a
special 2-pole alternator, whose stator winding is designed so as it can be reconfigured
according to various multiphase schemes and consequently connected to a different
number of output rectifiers. The performance of the generator in different multiphase
configurations under the same operating conditions in terms of overall generated DC
power are compared to empathize the advantages that can be practically expected from
increasing the number of stator phases of the generator.
Fig. 3-1. Generation system schematic for: (a) NP1; (b) NP2. T: gas turbine; G: gear-box; L: DC load; E: rotor
excitation system.
The NP project involved the realization of two 2-MW generation system prototypes
suitable for supplying a 3000 V shipboard DC network using a 22500-rpm gas turbine
(GT) as the prime mover (Fig. 3-1).
The first and more conventional system prototype, referred to as “Naval Package 1” (NP1)
is a 6300 rpm alternator coupled to the GT through a gear-box. The second, more
innovative and technologically challenging prototype, referred to as “Naval Package 2”,
(NP2) is a ultra-high speed alternator suitable for direct coupling to the GT and thereby
with a rated speed of 22500 rpm. The main characteristic data of the two prototypes are
summarized in Table I.
TABLE I
COMPARISON BETWEEN NP1 AND NP2 GENERATION SYSTEMS
NP1 NP2
Alternator rotor type Wound rotor Permanent-magnet rotor
Number of alternator phases 2⨯3 4⨯3
GT / alternator coupling Epicycloidal gearbox Direct coupling
Number of alternator poles 4 4
Alternator speed 6300 rpm 22500 rpm
Alternator frequency 210 Hz 750 Hz
Rectifying power electronics Diode bridges Diode bridges and choppers
A schematic of the two generation system layout and components is provided in Fig. 3-1. It
can be seen that the stator winding of both generators have a multiphase arrangement
consisting of N three-phase sets (N=2 for NP1 and N=4 for NP2), each connected to a
power-electronics rectifier (a simple diode bridge for NP1 and a cascade of a diode bridge
and a chopper for NP2). The N rectifiers are connected in cascade and their series closed
on the DC load.
A picture of the two manufactured prototypes is reported in Fig. 3-2, which clearly
empathizes how increasing the rated speed in NP2 (and thereby reducing its rated torque)
leads to a significant reduction in the overall machine dimensions.
Fig. 3-2. Generation system schematic for: (a) NP1; (b) NP2. T: gas turbine; G: gear-box; L: DC load; E: rotor
excitation system.
milled to accommodate concentric excitation coils. The main challenge in such generator
development is scaling down a turbo-alternator design (tailored on very large machine
sizes, with rotor lengths of up to several meters) to the relatively compact dimensions of a
generator whose external frame does not exceed 1.5 meters of length (Fig. 3-2a). The use
of a solid steel rotor technology with end-winding retaining ring was dictated by the high
rotor peripheral speed (machine rated speed is 6300 rpm), which discouraged the use of a
salient-pole design, although this would be much more usual for the machine dimensions.
The advantages offered by the dual star arrangement compared to the ordinary three-
phase design in terms of performance and fault tolerance can be understood by comparing
the voltage and current waveforms obtained for the system configurations depicted in Fig.
3-3, i.e.: (a) the dual star healthy configuration; (b) the dual star configuration with a faulty
diode; (c) the single star (three-phase) healthy configuration; (d) the single star
configuration with a faulty diode.
In Fig. 3-4 the stator phase currents are shown, while in Fig. 3-5 and Fig. 3-6 we can see
the output DC current (whose waveform is practically the same as for the output voltage)
and the resistant electromagnetic torque developed by the generator. It is clear how, in
health operation, the dual star configuration allows for a smaller output current ripple and
40 Multiphase technology in electric power generation
a smaller torque ripple thanks to the 12-pulse reaction. In presence of a faulty diode, the
system performance undergoes a significant degradation in both single and dual star
Multiphase technology in electric power generation 41
generator arrangement, but it can be seen that in the latter case the consequences of the
fault are much less severe both in terms of rectified current quality and in terms of torque
quality. Similar considerations can be made with regard to the rotor field current, whose
ripple reflects the magnitude of the harmonic rotating fields which originate in machine
air-gap. The smaller field ripple amplitude observed in the case of the dual star generator
is the consequence of the better air-gap field harmonic content which characterizes this
configuration, where no 5th and 7th order rotating fields are present.
Fig. 3-8. System schematic of a split-phase alternator G supplying a DC load L though multiple cascaded rectifiers
Fig. 3-9. Prototype generator with prime mover (left-hand side) and with halogen lamps used as DC load (right-
hand side).
Multiphase technology in electric power generation 43
Fig. 3-10. Schematic illustrating stator coil arrangement in one winding layer: each circle represent a coil side.
The same 36 coils are used to implement the different phase arrangements illustrated in
Fig. 3-11 for N=1, N=2, N=3 and N=4, where Uk, Vk, Wk denote the phases of the k-th
winding (k = 1.. N) and each solid-line arrow represents the spatial orientation of a stator
phase axis.
Fig. 3-11. Phase vectors for generator winding configurations with different N.
44 Multiphase technology in electric power generation
Fig. 3-12. Phase belt arrangement and extension for generator split-phase winding configurations with different
N.
For this purpose, the 36 stator coil of the alternator are connected as per Fig. 3-12, where
the stator circumference is unrolled and the coils arranged in a linear array.
For example, it can be seen from Fig. 3-12 that the usual three-phase configuration (N=1)
is obtained by series connecting coils 1 ÷ 6 to form phase belt U1, and series connecting
coils −1 ÷ −6 to form the phase belt –U (the same pertains to the other two phases).
By splitting each phase belt into 2, 3 and 4 sections, then, the split-phase configurations
with N=2, N=3, N=4 are obtained respectively. In particular, to implement the quadruple
three-phase arrangement (Fig. 3-12) a fractional-slot winding configuration needs to be
adopted.
Regardless of the number of phases (3N) and of the number of rectifiers (N), the
aforementioned values of DC voltage and current are obtained on the load by operating
the generator at nearly the same flux and with nearly the same excitation current If. In fact,
from Fig. 3-8 the following relation is immediately derived:
V dc ≅ 1.35 × N × V (3-1)
where V is the line-to-line rms voltage on a stator winding. The voltage V can be in turn
written as:
V = 3 × 4.44 × N s × f × φ × k w (3-2)
where f is the stator frequency, φ the flux per pole, kw the stator winding factor and Ns the
number of series-connected turns per phase. Ns is given by (3-3), where q indicates the
number of coils per pole per phase and Nt the number of turns per coil.
N s = 2× q × Nt (3-3)
36 (3-4)
q=
2× 3× N
Substitution of (3-2)-(3-4) into (3-1) causes N to cancel out, which proves that regardless
of N the output voltage Vdc can be obtained with the same flux per pole (i.e. with the same
generator magnetic loading), apart from slight differences due to the weak dependency of
kw on N. This fact is confirmed by the measured field current If which is approximately the
same in all test configurations.
5 8
4
6
0
4
2
−5
N=1 2 N=1 N=1
− 10 0 0
0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08
10 10 6
5 8
4
6
0
4
2
−5
N=2 2 N=2 N=2
− 10 0 0
0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08
10 10 6
5 8
4
6
0
4
2
−5
N=3 2 N=3 N=3
− 10 0 0
0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08
10 10 6
5 8
4
6
0
4
2
−5
N=4 2 N=4 N=4
− 10 0 0
0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08
s s s
10 10 5
4 4 2
2 2 1
0 0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
10 5
2 2 1
0 0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
10 5
2 2 1
0 0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 2 4 6 8 10 12 14 16 18 20 22 24 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
10 5
4 4 2
2 2 1
0 0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Harmonic order Harmonic order Harmonic order
Fig. 3-13. Recorded current waveforms (top diagrams) and relevant harmonic spectra (bottom diagrams) in tests
with different N.
The importance of minimizing air-gap field harmonics is connected with two main design
goals:
1) The improvement of the torque quality: in fact, the harmonic spectrum of the
electromagnetic torque is qualitatively the same as that of the field current [10] and
high values of torque ripple, especially at relatively small frequency, can produce
severe damages on the mechanical equipment coupled to the generator (prime
Multiphase technology in electric power generation 47
9
8
8 8 7
6
Healthy bridge
6 6 5
Faulty bridge
4
4 4 3
2
2 2
1
0 0 0
0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08 1 2 3 4 5 6 7 8 9 10 11 12 13 14
8
7
6 6 6
5 Healthy bridge
4 Faulty bridge
4 4
3
2 2 2
1
0 0 0
0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Fig. 3-14. Output DC current waveforms and relevant harmonic spectra during normal operation and with an
open diode.
mover, couplings, gear box, shaft, etc.) due to fatigue and torsional resonance issues
[11].
2) The reduction of stray-load losses in rotor circuits, i.e. in the field winding and in the
damper cage. The losses in the field circuit taking the DC resistance of the field (i.e.
neglecting skin effects) can be easily computed from the measured excitation
current (Fig. 3-13), leading to definitely negligible values compared to the losses
produced by the DC field current component. Conversely, the losses in the damper
cage, which cannot be directly estimated, are expected to be non-negligible in
presence of important air-gap field space harmonics, as discussed in [12]. Moreover,
while the field circuit is usually composed of thin round wire or thin flat turns,
damper bars may have an important solid section, leading to much higher eddy-
current and skin effect additional losses.
3.3 References
[1] G. Sulligoi, A. Tessarolo, V. Benucci, M. Baret, A. Rebora, A. Taffone, “Modeling,
simulation and experimental validation of a generation system for Medium-Voltage
48 Multiphase technology in electric power generation
DC Integrated Power Systems”, IEEE Transactions on Industry Applications, vol. 46, no.
4, July/Aug. 2010, pp. 1304-1310.
[2] A. Tessarolo, “Analysis of split-phase electric machines with unequally-loaded stator
windings and distorted phase currents”, International Conference on Electric Machines,
ICEM 2010, 6-9 Sept. 2010, Rome, Italy, CD-ROM paper RF-013331.
[3] A. Tessarolo, S. Castellan, R. Menis, G. Ferrari, “On the modeling of commutation
transients in split-phase synchronous motors supplied by multiple load-commutated
inverters”, IEEE Trans. on Industrial Electronics, vol. 57, no. 1, Jan. 2010, pp. 35-43.
[4] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul. 2007,
pp. 489-516.
[5] H. E. Jordan, R. C. Zowarka, S. B. Pratap, “Nine-phase armature windings design, test
and harmonic analysis”, IEEE Trans. on Magnetics, vol. 41, issue 1, part 2, Jan. 2005,
pp. 299-302.
[6] G. Buja, S. Castellan, “A flicker compensation strategy”, IEEE Trans. on Power
Electronics, vol. 24, no. 5, May 2009, pp. 1243-1247.
[7] B.K. Bose, Modern Power Electronics and AC Drives, 2001, Pearson Education.
[8] E.A. Klingshirn, “High phase order induction motors-Part I-Description and
theoretical considerations”, IEEE Trans. on Power Apparatus and Systems, Jan. 1983,
vol. PAS-102, pp. 47-53.
[9] E.A. Klingshirn, “High phase order induction motors-Part I-Experimental results”,
IEEE Trans. on Power Apparatus and Systems, Jan. 1983, vol. PAS-102, pp. 54-59.
[10] Tessarolo A., Castellan S., Menis R., “Feasibility and Performance Analysis of a High-
Power Drive Based on Four Synchro-Converters Supplying a Twelve-Phase
Synchronous Motor”, IEEE PESC 2008, pp. 2352-2357.
[11] K.S. Smith, Li Ran, “Torsional resonance risk management in islanded industrial
power systems supplying large VFDs”, IEEE Trans. on Industry Applications, vol. 44,
no. 6, Nov./Dec. 2008, pp. 1841.
[12] G. Traxler-Samek, T. Lugand, A. Schwery, “Additional losses in the damper winding of
large hydrogenerators at open-circuit and load conditions”, IEEE Trans. on Industrial
Electronics, vol. 57, no. 1, Jan. 2010.
[13] M.G. McArdle, D.J. Morrow, “Noninvasive detection of brushless exciter rotating diode
failure”, IEEE Trans. on Energy Conversion, vol. 19, no. 2, June 2004.
Multiphase design and winidng construction technology in high-power machines 49
The benefits that can be obtained from splitting the stator winding of an electric machine
into more than three phases have been largely investigated in the field of small and
medium power applications [1], [2]: they mainly have a “functional” nature (concerning
inverter power segmentation, redundancy and reliability improvement, torque ripple and
efficiency enhancement, power density increase, etc.), whereas machine structure and
construction technology substantially remain unchanged compared to conventional three-
phase design. When it comes to very high-power electric machines (in the multi-MW
range), however, adopting an n-phase stator configuration (with n>3) may have
considerable “structural” or “constructive” impacts as well. In fact, this enables to retain a
coil winding technology, which is the same as for small-sized machines, instead of
resorting to the much more complicated and costly Roebel-bar constructions, which is
typical or large turboalternators [3], [4]. Firstly, the point is discussed in this Section in
general terms, by means of appropriate sizing equations which relate the number of
phases to the design variables involved in the selection of the machine winding
technology. Secondly, some industrial application examples will be reported referring to
built high-power multi-phase machines.
Fig. 4-1. Slot cross section for increasing power and fixed voltage in a coil winding electric machine.
50 Multiphase design and winidng construction technology in high-power machines
Fig. 4-2. Stator coil winding (a) made from preformed closed coils (b).
necessarily grows too, which results in an increasingly high individual coil cross-section
and low number of turns per coil (Fig. 4-1). Above a certain power level, a design with a
single turn per coil becomes eventually mandatory, which leads to move from coil winding
to Roebel bar technology (Fig. 4-3). A typical example of the latter can be found in large
turboalternators [3]. Compared to coil windings, Roebel bar technology is much more
costly and complicated, in general, especially due to strand transpositions and special
manufacturing techniques required for end-bar connections.
When an n-phase stator configuration with n>3 is adopted instead of a traditional three-
Fig. 4-3. Roebel winding technology: (a) strand transposition; (b) bar section; (c) insulated and formed bars
ready for assembly.
Multiphase design and winidng construction technology in high-power machines 51
Fig. 4-4 Phase belt distribution over a pole-span in a double-layer shortened-pitch split-phase winding composed
of N three-phase sets (a, b, c).
phase one, the power is split into n phases instead of three. Consequently, if the voltage is
maintained the same, the phase current diminishes by approximately a factor n/3. This
may enable the designer to retain a coil design for the stator winding avoiding the use of
Roebel bars, with significant savings in cost and production times.
As an alternative, the n stator phases can be distributed uniformly over each pole span
instead of being grouped into three-phase sets. This results in a “symmetrical” n-phase
configuration, where n is not necessarily required to be a multiple of three. This winding
topology requires a non-conventional n-phase inverter (and control strategy) to be used
for motor supply [2].
Fig. 4-5. Phase arrangement schematic for a split-phase machine with N stator windings displaced by τ=60/N
electrical degrees apart.
52 Multiphase design and winidng construction technology in high-power machines
machine windings with a “hybrid” winding composition, which results from combining
split-phase and symmetrical schemes. This is the case of the 15-phase 21 MW induction
machine reported in [8], where three 5-phase symmetrical phase sets are employed, each
supplied by a 5-phase inverter.
Calling n the number of stator phases (however arranged in space), the rated phase
voltage is given by:
V = 1 / 2 Φ p N s k w (2π f ) (4-1)
where Φp is the flux per pole, Ns the number of series-connected turns per phase Ns, kw the
winding factor [5] and f the rated frequency. Quantities Φp, Ns and f can in turn be
expressed as follows:
Φ p = Bg L π D /( 2 p ) (4-2)
N s = q ( 2 p) N t / b (4-3)
f = N p / 60 (4-4)
in terms of: the average flux density Bg in the air-gap, the useful core length L, the machine
average diameter D at the air-gap, the number of pole pairs p, the number of slots per pole
per phase q (possibly fractional), the number of series-connected turns per coil Nt, the
number of parallel paths per phase b and the speed N in revolutions per minute.
A further design figure, called the output coefficient C and defined as per [11], can be also
introduced to describe the degree of utilization of the machine volume (roughly
proportional to D2L) in terms of useful machine torque (proportional to P/N, where P is
the rated active power):
C = P /( N D 2 L ) (4-5)
π 2 2 (2p)P B m k w q N t
V= (4-6)
120 N pp D C
πD
τs = (4-7)
q n (2 p )
π 3 2 (2p) q n P B m k w N t π 3 2 P B m k w N t
V= = (4-8)
120 π D N pp C n 120 N pp C n τ s
2N t At σ s
τs = (4-9)
λs
in terms of the current density σs flowing though stator conductors, the copper cross
section area of a turn At and the stator electric loading λs [12]. In fact, the following
definitions apply for σs and λs:
I / b 2N t (I / b)
σs = = (4-10)
At As k f
2 Nt I / b
λs = (4-11)
τs
where I is the phase current, As and kf are the cross-section area and the fill factor of a
stator slot.
Substitution of (4-9) into (4-8), after elementary algebraic manipulation, finally yields:
P C σ s k f As
= k × b × n
V B
142
m λ s k w N t (4-12)
43
R
where k is a non-dimensional constant whose value only depends on the units used to
express the other quantities. Coefficient in brackets (R) does not depend on the winding
structure, but only on the magnetic, thermal and electrical loading of the machine;
therefore, for machines of homogeneous design in terms of thermal class, insulation
technology, cooling system effectiveness, etc., R can be regarded as a constant to a good
approximation, as confirmed by the application examples reported in the following
Section.
Hence Equation (4-12) expresses the explicit relationship between the following design
quantities:
• machine power (P) and voltage (V) ratings;
• winding structure in terms of slot cross-section area (As), number of turns per coil
(Nt), number of phases (n) and number of parallel ways per phase (b).
Equation (4-12) shows that if the power rating P increases while the voltage V below a
certain level, this naturally leads to decrease the number of turns per coil Nt, which may
result in the need for Roebel bars (Nt=1) above a given power level. Equation (4-12) also
demonstrates that there are three design “levers” available to counteract the decrease of
Nt, namely:
• increasing the slot cross-section area As;
• increasing of the number b of parallel ways per phase;
• the increase of the number of phases n.
54 Multiphase design and winidng construction technology in high-power machines
The first strategy is of limited help, since it generally implies a growth of the overall
machine size: in fact the slot opening Ws (Fig. 4-1) needs to be lower than the tooth width
at the air-gap to contain slot harmonics [5] and the increase of Hs brings to a growth of the
stator outer diameter so as to keep the yoke flux density within acceptable limits.
The second strategy can be actually pursued until the number of parallel ways b equals 2p
since the number of parallel ways cannot exceed the number of machine poles in any case
[5].
Hence, it is easily understood that, after the limit b=2p has been reached, the only way left
to avoid the use of Roebel bar construction without incrementing the machine size
consists of increasing the number of its stator phases n.
TABLE I
RATINGS OF THREE MACHINES TAKEN AS EXAMPLES
A B C
Machines A and B (Fig. 4-6) are high-speed inverter-fed motors used in turbo-compressor
applications. In particular, machine A is a dual three-phase (split-phase) synchronous
motor fed by two Load-Commutated inverters (Fig. 4-6, A.1), while machine B is a 12-
phase synchronous motor equipped with four three phase windings, shifted by 15
electrical degrees apart and supplied by four Voltage-Source PWM inverters (Fig. 4-6, B.1).
Machine C is a high-speed naval generator which feeds two diode rectifier bridges (one
per stator section) connected to a DC on-board power grid.
The three machines are all required to operate within thermal class B and are designed
with homogeneous thermal, electrical and magnetic loading values.
With regard to the sizing Equation (4-12) derived in the previous Section, coefficients R
are computed using the design quantities of the three machines and their values are
reported in Table II.
Multiphase design and winidng construction technology in high-power machines 55
Fig. 4-6. Actual system configurations for machines A, B, C (left column); possible alternative arrangements with
a three-phase design for the electric machine (right column)
TABLE II
VALUES OF COEFFICIENT R FOR MACHINES A, B, C
A B C
Coefficients R 0.53 0.45 0.52
It can be seen that, despite of the difference in machine size, number of phases and ratings
(Table II), the values of R are relatively close. This confirms that R, at least for preliminary
sizing purposes, can be actually regarded as a design constant depending on thermal,
magnetic and electrical loading only.
As a consequence of the high P/V ratios, and for a given slot size As, a low number of turns
per coil Nt may result according to Equation (4-12), possibly leading to a Roebel bar design
(Nt=1) if the number of phases n is kept at its minimum value (4-3).
Conversely, raising the number of phases enables to keep the number of turns per coil
higher than one (Table I) and allows for a coil winding construction to be used in all the
three cases, with significant savings in terms of manufacturing cost, lead times and tooling.
The relationship between the winding construction technology and the number of phases
can be better highlighted by considering some possible design alternatives for the systems
under consideration. Such alternatives are illustrated in the left-hand column of Fig. 4-6
for comparison with the actual design configurations, represented in the left-hand side
column.
Case of machine B — If machine B were conceived as a single three-phase one under the
same power and voltage requirements (Fig. 4-6, B.2), for instance according to the design
reported in [14], Equation (4-12) would apply with:
Case of machine C — If machine C were conceived as a single three one with the same
power and voltage requirements (Fig. 4-6, C.2), Equation (4-12) would apply with:
4.2 References
[1] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul. 2007,
pp. 489-516.
[2] S. Williamson, S. Smith, “Pulsating torque and losses in multiphase induction
machines”, IEEE Trans. on Industry Applications, vol. 39, July/Aug. 2003, pp. 986-993.
[3] J. Haldemann, “Transpositions in stator bars of large turboalernators”, IEEE Trans. on
Energy Conversion, vol. 19, Sept. 2004, pp. 553-560.
[4] R.V. Musil, W.Schmatloch, “Dimensioning and design of high-power variable speed
synchronous motors”, Energy and Automation X (1988), Special − Large Electric Motor
A.C. Variable Speed Drives, pp. 32-41.
[5] A. Still, C.S. Siskind, Elements of Electrical Machine Design, McGraw-Hill, 1954.
[6] J. J. Simond, A. Sapin, M. T. Xuan, R. Wetter, P. Burmeister, “12-pulse LCI synchronous
drive for a 20 MW compressor modeling simulation and measurements”, Industry
Application Society Annual Meeting, IAS 2005, pp. 2302-2308.
[7] H. E. Jordan, R. C. Zowarka, S. B. Pratap, “Nine-phase armature windings design, test
and harmonic analysis”, IEEE Trans. on Magnetics, vol. 41, issue 1, part 2, Jan. 2005, pp.
299-302.
[8] F. Terrein, S. Siala, P. Noy, “Multiphase induction motor sensorless control for electric
ship propulsion”, IEE Power Electronics, Machines and Drives Conference, PEMD 2004,
pp. 556-561.
[9] E. A. Klingshirn, “High phase order induction motors, Part I”, IEEE Trans. on Power
Apparatus and Systems, Vol. PAS-102, no. 1, Jan. 1883, pp. 47-53.
[10] E. A. Klingshirn, “High phase order induction motors, Part II”, IEEE Trans. on Power
Apparatus and Systems, Vol. PAS-102, no. 1, Jan. 1883, pp. 54-59.
[11] M. Yamamoto, Y. Takeda, T. Oishi, “Output coefficient of synchronous motors”, IEEE
Trans. on Power Apparatus and Systems, vol. PAS-104, July ’85, pp. 1849-1855.
[12] S. Huang, J. Luo, F. Leonardi, T. A. Lipo, “A general approach to sizing and power
density equations for comparison of electrical machines”, IEEE Trans. on Industry
Applications, vol. 34, Jan./Feb. ’98, pp. 92-97.
[13] G. Sulligoi, A. Tessarolo, V. Benucci, M. Baret, A. Rebora, A. Taffone, “Modeling,
simulation and experimental validation of a generation system for medium-voltage DC
integrated power systems”, IEEE Trans. on Industry Applications, IEEE Transactions on
Industry Applications, vol. 46, no. 4, July/Aug. 2010, pp. 1304-1310.
[14] S. Schroeder, P. Tenca, T. Geyer, P. Soldi, L. Garces, R. Zhang, T. Toma, P. Bordignon,
“Modular high-power shunt-interleaved drive system: a realization up to 35 MW for oil
and gas applications”, IEEE IAS Annual Meeting, 2008.
58
In Part I, it has been pointed out that multiphase stator windings can be designed according
to various topologies. An important one is constituted by the so called “split-phase” or
“asymmetrical” or “multiple star” arrangement. It consists of splitting the stator winding into
multiple (N) identical three-phase sections, displaced by 60/N electrical degrees apart. Split-
phase configurations are important as the N winding sections can be supplied by the same
well proven three-phase inverter modules as otherwise used to feed three-phase machines of
smaller power ratings.
In this Part, the modeling of multiple star machines is treated in a quite traditional way,
following an approach which was first proposed by Nelson and Krause in the 1970’s and
which practically extends the well known Park’s dq theory (valid for three-phase machines)
to the case when multiple three-phase sets are present.
The original contribution given here with respect to Nelson and Krause’s work consists of the
following points:
Some measurements on real multiple star machines are proposed to experimentally validate
the results presented and in particular to confirm and quantify the cross coupling
phenomena theoretically predicted.
The final Chapter of this Part is devoted to the numerical implementation of the multi-star
machine model in the Matlab/Simulink environment and to the discussion on how it has been
used to successfully simulate the performance of various multi-star machines together with
the power electronics and control systems interfaced to them. A practical example of the
accuracy that can be achieved in such simulations is illustrated referring to a dual-star
synchronous machine supplied by two Load Commutated Inverters.
General dq equivalent circuit representation of multi-star machines 59
Split-phase electric machines are of increasing importance in modern electric drives as far
as very high power and high reliability applications are concerned. They are characterized
by a stator winding composed of N three-phase sets displaced by 60/N electrical degrees
apart. This winding configuration exhibits various advantages over the traditional three-
phase one [1], [2], such as: a performance enhancement in terms of torque ripple and
efficiency; an intrinsically redundant structure (the drive can keep in operation even with
a faulty converter); the possibility to use three-phase converter modules (which are more
technologically-proven than poly-phase ones); the possibility to reach an overall drive
power rating (up to several tens of MWs) that could not be attained using a single three-
phase supply unit due to the current capability limits of power electronic devices.
Different approaches have been proposed in the literature to model split-phase motors,
mainly depending on the strategy chosen for their control [1]. Some control techniques
require stator phase variables to be mapped into an orthonormal 3N-dimensional
coordinate system, by means of the so-called Vector-Space Decomposition (VSD) method
[3], [4], [5]. Other control strategies [6], [7], [8] are based on an extension of the two-axis
Park’s theory [9], as described by Nelson and Krause [10]. According to the latter
approach (referred to as “dq0” approach in the following), the triplet of currents or
voltages pertaining to each three-phase set is represented through a rotating space vector
plus a homopolar component; the resulting N space vectors are then projected onto a
couple of orthogonal d, q axes.
Application examples of the dq0 methodology for split-phase motors can be found in the
control and analysis of induction [6], [7] and of synchronous machines as well [9].
In principle, the dq0 modelling of a split-phase electric machine enables to represent its
dynamics in terms of three equivalent circuits (the “d-axis”, the “q-axis” and the
homopolar ones) that generalize the well-known dq0 equivalent circuits widely used for
ordinary three-phase machines [11], [12]. Such generalization, though, has been explicitly
carried out only with regard to the dual three-phase configuration [1], [13] which is in fact
the most common among split-phase arrangements.
The purpose of this Chapter is to investigate the detailed dq0 equivalent circuit topology
for split-phase synchronous machines equipped with more than two stator three-phase
sets (N>2), given their growing importance in the field of today’s drive technology and
research [8], [14]-[18].
It will be shown that, while the magnetizing and rotor portion of the equivalent circuit
remains unchanged with respect to the three-phase case [11], the branch including stator
leakage inductances results in different topologies depending on N and can be fully
described through exactly 3N/2 or (3N+1)/2 independent parameters, respectively in the
case of even and odd N. Furthermore, it will be shown that when N>2, some magnetic
coupling arises, in general, between the d-axis and q-axis equivalent circuits, even in the
case of unsaturated round-rotor machines [19], due to stator leakage flux.
60 General dq equivalent circuit representation of multi-star machines
Fig. 5-1. Phase arrangement in a quadruple three-phase winding (N=4): phasor diagram for one layer (top);
phase-belt sequence over a pole-span in a double-layer shortened-pitch winding configuration (bottom).
The content of the Chapter is organized into two steps. Firstly, the general analytical dq0
model of a split-phase machine equipped with a generic number N of stator three-phase
sets is set forth including the detailed effects of phase couplings through leakage
inductances. Secondly, such analytical model is turned into an equivalent circuit
representation.
The hypothesis is made in the following that the split-phase machine under consideration
is a synchronous salient-pole one, provided with a field circuit and a couple of damper
windings, located along the polar and inter-polar axes respectively. In fact, as discussed
further on, the main complications arising in the modelling process concern stator leakage
inductances, the rotor portion of the dq0 equivalent circuit remaining the same as in three-
phase machines. Therefore, any model enhancement intended to fit other rotor types and
to improve the accuracy of their representation could be applied in the same ways as
proposed for three-phase machines [20], [21].
General dq equivalent circuit representation of multi-star machines 61
• space harmonics in the air-gap field due to winding distribution are neglected;
• space harmonics in the air-gap field due to rotor saliency are considered up to the
second-order harmonic;
• the leakage flux magnitude and distribution are supposed independent of the rotor
position;
The impact of assumptions 1, 2, 3 is again limited by the fact that, as anticipated, the focus
of the Chapter is on stator magnetic couplings through leakage inductances, which can be
reasonably assumed unsaturated and independent of rotor position [12].
v1 R L 03 03 i1 ϕ1
M M O M M M d M
v = 0 L R
+
03 i N dt ϕ N
(5-1)
N 3
vˆ 0 L 03 R r iˆr ϕˆ
r 3 r
ϕ 1 L 1 ,1 L L 1 ,N L 1 ,r i 1
M M O M M M
ϕ = L L L N ,N L N ,r i N
(5-2)
N N ,1
ϕˆ L L L r ,N Lˆ r ,r iˆ r
r r ,1
va, j ia, j ϕ a, j
v j = v b , j , i j = i b , j , ϕ j = ϕ b , j (5-3)
vc , j ic, j ϕc , j
0 ikd
v̂ r = 0 , î r = ikq (5-4)
v f
if
Rkd 0 0
R = R I 3N , Rr = 0 Rkq 0 (5-5)
0 0 R f
t
L j +1 ,r = L r , j +1
while the structure of each matrix block Lj,r, Lr,j is detailed in (5-7) at the top of the next
page. In (5-5)-( 5-7) the following nomenclature has been used:
Rkq, Rkq, Lkq, Lkq: stator-referred resistances and leakage inductances of damper
circuits;
Rf, Lf: stator-referred resistance and leakage inductance of the field circuit.
Coefficients 2 / 3 and 2/3, which appear in (5-6) and (5-7), depend on the transformation
ratio chosen for referring rotor quantities to the stator [12]. Superscript “t” indicates the
transposition operator.
As concerns stator inductance matrix blocks Li,j, their structure is detailed in (5-8), where
λφ,ψ denotes the mutual inductance between two stator phases located at θ=φ and
θ=ψ respectively (i, j = 0, 1, …, N−1).
Each inductance λφ,ψ can be expressed as the sum of a leakage and a magnetizing term,
respectively denoted by superscripts “l” and “m”:
General dq equivalent circuit representation of multi-star machines 63
Fig. 5-2. A pair of phases (grey and white phase belts) and their magnetic axes, at θ=φ and θ=ψ (a); the former
phase is rotated by π radians (b).
Under hypotheses 2 and 3 of 5.1, the magnetizing term takes the well-known form below:
The leakage inductance term, instead, is more difficult to formulate explicitly without
knowing the detailed geometry of the machine to be modelled [22]-[24]. Nevertheless, for
the purposes of this Chapter, a very general expression of λφ(l,ψ) is next proposed in terms of
Fourier series as per (5-11).
∞
λ(φl,ψ) = ∑ Λ2n+1 cos[(2n + 1)(φ −ψ )] (5-11)
n =0
the function is thus proved to be an even one; finally, the periodicity properties (13)-(14)
are easily established, assuring that λφ(l,ψ) can be expanded in Fourier series with only odd
cosine terms, as per (5-11).
Fig. 5-2 provides a schematic illustration of property (5-14), showing how the rotation of
one phase by π electrical radians is the same as reversing the direction of its current.
64 General dq equivalent circuit representation of multi-star machines
[
vˆ j = v d , j vq, j v0, j ] t
[
= T j va, j v b, j vc, j ]t
(5-15)
[
iˆ j = i d , j iq, j i0, j ]
t
[
= T j i a, j i b, j ic, j ]t
(5-16)
[
ϕˆ j = ϕ d , j ϕ q, j ϕ 0 , j ]t = T j [ϕ a , j ϕ b, j ϕ c , j ]t (5-17)
where j=1..N, subscript “0” indicates the homopolar component of the j-th triplet and
matrices Tj are defined by (5-18).
The whole set of machine variables can be thus transformed into the dq0 reference frame
as per (5-19), with matrix T defined by (5-20).
vˆ 1 v1 iˆ1 i1 ϕˆ1 ϕ1
M
ˆv = = T , iˆ = = T , ϕˆ = = T
M M M M M
ˆ (5-19)
ˆv N v i ϕˆ ϕ
N iN N N N
vˆ v iˆ iˆ ϕˆ ϕˆ
r r r r r r
T1 L 03 03
M O M M
T= (5-20)
0 L TN 03
3
0 L 03 I N
3
In terms of the new variables, machine voltage equations (5-1)-(5-2) take the time-
invariant form reported below:
vˆ 1 R L 03 03 iˆ1 J L 03 03 ϕˆ 1
M M O M M M dθ r M O M M M
vˆ = 0 L R
+
03 iˆ N dt 0 L J 03 ϕˆ N
N 3 3
vˆ 0 L 03 R r iˆr 0 L 03 03 ϕˆ r
r 3 3
Lˆ 1 ,1 L Lˆ Lˆ iˆ 1 (5-21)
1,N 1 ,r
M O M M d M
+
ˆL ˆ Lˆ N ,r dt iˆ N
N ,1 L L N ,N
Lˆ ˆ
r ,1 L L r , N Lˆ r ,r iˆ r
General dq equivalent circuit representation of multi-star machines 65
ϕˆ 1 Lˆ 1 ,1 L Lˆ 1 ,N Lˆ 1 ,r iˆ1
M M O M M M
ϕˆ = Lˆ L Lˆ N ,N
Lˆ N ,r iˆ N
(5-22)
N N ,1
ϕˆ Lˆ L Lˆ r ,N Lˆ r ,r iˆ r
r r ,1
The 3×3 constant sub-matrices which appear in (5-21)-(5-22) can be found, through
algebraic manipulation, to take the following form, with i and j ranging from 1 to N:
0 − 1 0
J = Tj
d
dθ r
t
( )
T j = 1 0 0 (5-23)
0 0 0
Lmd 0 Lmd
3
Lˆ j ,r = Lˆ r , j = T j L j ,r = 0
t
Lmq 0 (5-24)
2
0 0 0
Lmd 0 0 Mi − j − X i− j 0
ˆL = Lˆ t = T L T t = 3 0 Lmq 0 + X i − j Mi− j 0 (5-25)
i, j j ,i i i, j j
0 0 H i − j
2
0 0 0
3 N N N N
= p ∑
2 j =1 j =1∑ j =1 ∑
i d , j (Lmd − Lmq ) i q , j + i q , j Lmd (i kd + i f ) − i d , j Lmq i kq
j =1 ∑
where p is the number of pole pairs. The first term in curly brackets constitutes the
reluctance torque, accounting for the interaction between d-axis and q-axis stator current
components in the presence of rotor saliency; the second term is the main torque resulting
from stator and rotor current interaction. It can be seen that the torque expression is
exactly the same as for a three-phase synchronous machine [12] provided that id and iq are
respectively replaced by the sums ( id , j and ∑j ∑j
iq , j ) of the d and q currents pertaining
to all stator sets. In particular, (5-26) formally confirms that, as expected, the leakage
inductance parameters M k , X k , H k do not have any effect as far as the torque generation
is concerned. Conversely, it is well known how stator leakage parameters play a crucial
role in determining phase-split machine harmonic impedances as defined in [23], [26] and
strongly affect their performance under VSI supply [22].
66 General dq equivalent circuit representation of multi-star machines
As concerns the practical evaluation of (5-11) from the machine design data, this task can
be accomplished though either numeric or analytical approaches, as discussed in [23],
[24].
The focus at this point is, instead, on finding the general split-phase machine equivalent-
circuit form. To this end, a valuable help comes from looking at the explicit expression
of M k , X k , H k in terms of Fourier coefficients Λ2n+1 used in (5-11). Such expressions are
detailed below for any integer k.
3 3 ∞ 2πnk
Mk =
2
Λ1 + ∑ ( Λ6n−1 + Λ6n+1 ) cos
2 n =1 N
(5-27)
3 ∞ 2πnk
Xk = ∑ ( Λ6n−1 − Λ6n+1 ) sin
2 n =1 N
(5-28)
∞
π(2n + 1)k
Hk = 3 ∑ Λ
n =1
6 n +3 cos
N
(5-29)
Identities (5-27)-( 5-29) have been derived through a symbolic math expansion of (5-25)
by taking (5-8), (5-9), (5-11) and (5-18) into account.
Fig. 5-4. Self and mutual inductances of stator dq0 circuits relevant to the i-th and j-th stator three-phase set.
General dq equivalent circuit representation of multi-star machines 67
homopolar circuit. The latter takes into account a generic couple of stator sets (namely the
i-th and the j-th) and illustrates how the self and mutual inductances of the corresponding
dq0 circuits can be expressed in terms of Lmd, Lmq and the leakage parameters (5-27)-(5-
29).
In particular, it can be seen that the mutual leakage inductance Xi−j couples the d-axis
circuit corresponding to the i-th set with the q-axis circuit corresponding to the j-th set.
Such d-q cross-coupling does not depend on magnetic saturation like analogous
phenomena observed in three-phase salient-pole machines [19]; conversely, it depends on
leakage fluxes alone and may occur only between d and q circuits representing different
stator sets, i.e. only if i≠j. In fact, (5-28) proves that X0 is always zero.
From the inspection of (5-27)-(5-29) some more properties of the dq0 leakage inductance
set can be inferred, as summarized by the following identities holding for any integer k:
Equations (5-30)-(5-31) enable us to simplify the 3N×3N stator leakage inductance matrix,
dramatically reducing the number of independent inductive parameters it depends on.
This is made clear by taking into account some particular cases (namely those with N=2,
N=3, N=4).
For example, in the dual three-phase configuration (N=2), the stator leakage inductance
matrix in the dq0 reference frame takes the form (5-32) and can be fully characterized by
three independent parameters (M0, M1, H0); in fact (5-30)-(5-31) assure that X0=X1=H1=0.
Fig. 5-3. Physical stator phase circuits (a) and stator dq0 circuits after Park’s transform (b), in the case of N=3.
68 General dq equivalent circuit representation of multi-star machines
M0 0 0 M1 0 0
0 M0 0 0 M1 0
0 0 H0 0 0 0
(5-32)
M1 0 0 M0 0 0
0 M1 0 0 M0 0
0 0 0 0 0 H0
In the cases of N=3, the matrix structure (5-33) is obtained, exhibiting five independent
parameters (M0, M1, H0, H1, X1), since H1=−H2, M1=M2, X1= −X2 in virtue of (5-30)-(5-31).
M0 0 0 M1 − X1 0 M1 X1 0
0 M0 0 X1 M0 0 − X1 M1 0
0 0 H0 0 0 H1 0 0 − H1
M1 X1 0 M0 0 0 M1 − X1 0
− X1 M1 0 0 M0 0 X1 M1 0 (5-33)
0 0 H1 0 0 H0 0 0 H1
M1 − X1 0 M1 X1 0 M0 0 0
X M1 0 − X1 M1 0 0 M0 0
1
0 0 − H1 0 0 H1 0 0 H0
Finally, in the case of N=4, the leakage inductance matrix takes the general form (5-34),
showing six independent parameters (M0, M1, M2, H0, H1, X1), since H1=−H3, M1=M3, X1= −X3,
X2=H2=0 in accordance with (5-30)-(5-31).
M0 0 0 M1 − X1 0 M2 0 0 M1 X1 0
0 M0 0 X1 M0 0 0 M2 0 − X1 M1 0
0 0 H0 0 0 H1 0 0 0 0 0 − H1
M1 X1 0 M0 0 0 M1 −X1 0 M2 0 0
− X1 M1 0 0 M0 0 X1 M1 0 0 M2 0
0 0 H1 0 0 H0 0 0 H1 0 0 0
(5-34)
M2 0 0 M1 X1 0 M0 0 0 M1 − X1 0
0 M2 0 − X1 M1 0 0 M0 0 X1 M1 0
0 0 0 0 0 H1 0 0 H0 0 0 H1
M −X1 0 M2 0 0 M1 X1 0 M0 0 0
1
X1 M1 0 0 M2 0 − X1 M1 0 0 M0 0
0 0 − H1 0 0 0 0 0 H1 0 0 H0
The procedure leading to build (5-32)-(5-34) can be easily repeated for an arbitrary
number N of stator three phase sets. In such general case, the number ν of independent
parameters required to fully define the stator leakage inductance matrix is found to be:
3N / 2 if N is even
ν = (5-35)
(3N + 1)/ 2 if N is odd
General dq equivalent circuit representation of multi-star machines 69
Fig. 5-5. Expanded forms of the magnetizing and rotor portion of the equivalent circuits.
The prove of (5-35) is obtained as per Table I, where the independent dq0 leakage
inductance parameters are explicitly listed and “counted” in the two cases of odd and even
N.
TABLE I
STATOR LEAKAGE PARAMETERS LIST IN D-Q-O REFERENCE FRAME
Number of
Symbols Notes
parameters
M 0 , M 1 , …, M N / 2 N/2+1
Even X 1 , X 2 , …, X N / 2−1 N/2−1 X 0 = X N /2 = 0
N H 0 , H 1 , …, H N / 2−1 N/2 HN / 2 = 0
Total: ν =3N/2
M 0 , M 1 , …, M ( N −1 ) / 2 (N−1)/2+1
X 1 , X 2 , …, X ( N −1 ) / 2 (N−1)/2
Odd N
H 0 , H 1 , …, H ( N −1 ) / 2 (N−1)/2+1
Total: ν = (3N+1)/2
Fig. 5-6. D-q equivalent circuits for a dual three-phase machine (N=2).
70 General dq equivalent circuit representation of multi-star machines
Fig. 5-7. D-q equivalent circuits for a triple three-phase machine (N=3).
Fig. 5-8. D-q equivalent circuits for a quadruple three-phase machine (N=4).
In the study of higher phase order configurations, the dq0 equivalent circuit approach has
never been used, instead [1], while other methods, like vector-space decomposition (VSD),
have been introduces as an alternative [3], [15], [22]. The resulting machine modeling,
characterized by a diagonal inductance matrix [22], is certainly useful for many purposes,
such as for control applications, but has the disadvantage that no distinction among the
stator three-phase sets is preserved in transformed coordinates.
General dq equivalent circuit representation of multi-star machines 71
Fig. 5-9. Homopolar equivalent circuit for a dual three-phase machine (N=2).
Fig. 5-10. Homopolar equivalent circuit for a triple three-phase machine (N=3).
Fig. 5-11. Homopolar equivalent circuit for a quadruple three-phase machine (N=4).
On the other hand, a dq0 circuital representation of phase-split machine might be helpful
in some ways, like:
The topology of the dq equivalent circuit for N≥2 can be directly inferred from the
structure of the leakage inductance matrices like (5-32)-(5-34), as exemplified in Figures
6-8 (where ω=dθr/dt). Regarding the magnetizing and rotor portions, indicated by Rd and
Rq, they keep the same form as for three-phase machines (Fig. 5-5), regardless of N. In fact,
72 General dq equivalent circuit representation of multi-star machines
(5-24)-(5-25) confirm that, after Park’s transformation, all stator sets share the same d, q
self magnetizing inductances and mutual inductances with rotor circuits ( 3 2 Lmd , 3 2 Lmq ).
Similarly, from leakage inductance matrices expressed as per (5-32)-(5-34), the dynamics
of homopolar components of all stators sets can be represented in circuit form as
exemplified in Figures 9-11.
The procedure followed to derive the dq0 equivalent circuits in the N=2, N=3, N=4 cases
apply for any higher number of stator sets. This may be of practical interest given that
some 15-phase stator designs have already been proposed and implemented for high
power ship propulsion [27].
5.6 Conclusions
In this Chapter phase-split machine modeling through dq0 equivalent circuit
representations has been investigated. It has been shown that, after applying Park’s
transform to all the N stator sets, the machine dynamic equations take a time-invariant
form that lends itself to an effective circuit representation. Equivalent circuit topologies
for the cases of two, three and four stator three-phase sets have been explicitly derived
based on a methodology that applies for any N. It has been shown that, while the stator
and rotor portion of the equivalent circuit is the same regardless of N, increasingly
complicated topologies are needed to represent the magnetic coupling among stator
phases through leakage fluxes. The number of independent parameters required for this
purpose has been determined as a function of N, under the only hypothesis that stator
leakage inductances are unsaturated and independent of rotor position. Finally, it has
been proven that, when N is greater than two (i.e. for more than two stator sets) the usual
decoupling between d and q equivalent circuits is lost. Such cross-coupling phenomenon
has been explained as a consequence of stator leakage flux distribution. In the next
Chapters, some analytical and numeric methods will be described to practically compute
the equivalent circuit parameters of split-phase machines and some experimental results
will be presented for validation.
5.7 References
[1] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul.
2007, pp. 489-516.
[2] D.G. Dorrell, C.Y. Leong, R.A., R.A. McMahon, “Analysis of performance assessment of
six-pulse inverter-fed three-phase and six-phase induction machines”, IEEE Trans on
Industry Applications, Nov./Dic. 2006, vol. 6, pp. 1487-1495.
[3] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction machine
using vector space decomposition”, IEEE Trans. on Industry Application, Sept.-Oct.
1995, vol. 31, pp. 1100-1109.
[4] A. Tessarolo, “On the modeling of poly-phase electric machines through vector-space
decomposition: theoretical considerations”, International Conference on Power
Engineering, Energy and Electrical Drives, POWERENG 2009, Lisboa, Portugal, 18-20
March 2009, pp. 519-523.
[5] A. Tessarolo, “On the modeling of poly-phase electric machines through vector-space
decomposition: numeric application cases”, International Conference on Power
Engineering, Energy and Electrical Drives, POWERENG 2009, Lisboa, Portugal, 18-20
March 2009 pp. 524-528.
General dq equivalent circuit representation of multi-star machines 73
Society Annual Meeting, IAS 2008, 5-9 Oct. 2008, Edmonton, Canada, CD-rom paper
n. 79.
[24] A. Tessarolo, D. Giulivo, “Analytical methods for the accurate computation of stator
leakage inductances in multi-phase synchronous machines”, International
Symposium on Power Electronics, Electrical Drives, Automation and Motion, SPEEDAM
2010, 14-16 June 2010, Pisa, Italy, pp. 845-852.
[25] E.A. Klingshirn, “High phase order induction motors−Part I−Description and
theoretical considerations”, IEEE Trans. on Power Apparatus and Systems, Jan. 1983,
vol. PAS-102, pp. 47-53.
[26] E.A. Klingshirn, “High phase order induction motors−Part II−Experimental results”,
IEEE Trans. on Power Apparatus and Systems, Jan. 1983, vol. PAS-102, pp. 54-59.
[27] F. Terrein, S. Siala, P. Noy, “Multiphase induction motor sensorless control for
electric ship propulsion”, IEE Power Electronics, Machines and Drives Conference,
PEMD 2004, pp. 556-561.
Closed-form expression for dq0 model parameters 75
In the previous Chapter, the structure of multi-star machine model and of the
corresponding dq equivalent circuits has been determined based on analytical Fourier
expansion of machine inductances. This approach is useful for the presented theoretical
treatment but does not yield final formulas suitable for numerical evaluation. In fact,
model parameters in transformed coordinates are expressed as infinite Fourier series. To
overcome the problem, in this Chapter the same results will be obtained with a different
methodology, which makes use of “conventional winding arrangement”, a concept which is
anticipated here and which will be widely employed in Part IV and V, dealing with Vector
Space Decomposition modeling. The advantage of the approach proposed in this Chapter is
that Fourier decompositions are not used and the model is always formulated (both before
and after transformation) as a function of physical parameters, that can be either
measured or computed from machine design data with the techniques described in Part III
The model formulation proposed in this Chapter allows for a quite easy comparison with
measurement results on some real multi-star machine to experimentally validate the
theoretical results presented. Such comparison will be presented in the next Chapter.
v 1 R L 03 i 1 ϕ e
d 1 1
M = M O M M + M + M (6.1)
v 0 L R i dt ϕ e
N 3 N N N
ϕ 1 L1 ,1 L L1 ,N i1
M = M O M M (6.2)
ϕ L
N N ,1 L L N ,N i N
va , j ia , j ϕ a , j ea , j
v j = vb , j , i j = ib , j , ϕ j = ϕ b , j , e j = eb, j (6.3)
vc , j ic , j ϕc , j ec , j
where, differently from what has been done in the previous Chapter, rotor effects are
accounted for through the e.m.f.’s ej induced by the rotor in the phases of the jth star. This
allows for the rotor equations not to be explicitly included in the model; in fact, it has been
shown in the previous Chapter that transformation of rotor equation in the multiple dq
reference frames gives identical results as for three-phase machines.
76 Closed-form expression for dq0 model parameters
T1 L 03
T= M O M (6.4)
0 L T
3 N
vˆ 1 v1 iˆ i1 ϕˆ 1 ϕ1 eˆ 1 e1
ˆ 1
vˆ = M = T M , i = M = T M , ϕˆ = M = T M , eˆ = M = T M (6.6)
vˆ v ˆ i ϕˆ ϕ eˆ e
N N iN N N N N N
Lˆ 1,1 Lˆ 1 ,2 L Lˆ 1,N Λˆ 1 Λˆ 2 L Λˆ N L1 ,1 L 1 ,2 L L 1 ,N
Lˆ 2,1 Lˆ 2,2 L Lˆ 2,N Λˆ 2t Λˆ 1 L Λˆ N −1 L2,1 L 2 ,2 L L 2 ,N t
= = T M T (6.7)
M M O M M M O M M O M
Lˆ Lˆ N ,2 L Lˆ N ,N Λˆ N t Λˆ N −1
t
L Λ 1
ˆ L L N ,2 L L N ,N
N ,1 N ,1
Λˆ i − j = Lˆ i , j (6.8)
since in the previous Chapter it has been demonstrated that L̂i , j depends only on the
difference i−j.
The purpose of this Chapter is to derive explicit expressions for the transformed
y 2× ABC = ( y A1 yC 2 )
t
y A2 y B1 yB2 yC 1 (6.9)
where y stands for a generic phase quantity (current, voltage, etc.), and that the
conventional winding variables are collected as:
y c = ( y0 y5 ) ,
t
y1 y2 y3 y4 (6.10)
it is clear that for the schemes in Fig. 6-2a and Fig. 6-2b to be equivalent we must impose:
1 0 0 0 0 0
0 0 0 1 0 0
0 0 −1 0 0 0
y c = W2×3 y 2× ABC , W2×3 =
0 0 0 0 0 − 1 (6.11)
0 1 0 0 0 0
0 0 0 0 1 0
The geometrical transformation matrix W2×3 can be written in a general form, which
enables an N-star winding to be mapped into a conventional 3N-phase scheme, as follows:
78 Closed-form expression for dq0 model parameters
where mod(x, y) indicates the remainder on dividing x by y . The inverse transformation is:
1 if j − trunc(i/ 3) − 2N mod(i, 3 ) = 0
{W } = {W }
N×3
−1
i, j N×3
t
i, j
= − 1 if j − trunc(i/ 3) − 2N mod(i, 3 ) = 3N i , j = 0, ..., 3N − 1
(6.13)
0 otherwise
The advantage of working with the conventional 3N-phase winding scheme instead of the
original split-phase one is that, due to the higher degree of symmetry, it is much easier to
write model matrices (in particular the inductance matrix) as functions of physical
parameters. This will be better discussed in the next Section.
d
v c = R c ic + ϕ c + ec (6.14)
dt
where Rc is the resistance matrix, i c , ϕ c and e c are respectively the phase current, flux
linkage and rotor induced e.m.f. vectors defined below:
v c = (v 0 v n−1 )
t
v 1 L v n −2 (6.15)
i c = (i0 in−1 )
t
i 1 L i n −2 (6.16)
ϕ c = (ϕ 0 ϕ1 L ϕ n−2 ϕ n−1 )
t
(6.17)
R c = rI n (6.18)
In the above equations, n is equal to 3N, In is the n⨯n identity matrix, r is the stator
resistance matrix.
As concerns the inductance matrix Lc, as done in the previous Chapter we shall suppose
that it consists of a constant term including leakage inductances (superscript l), and of a
magnetization term (superscript m) which can depend on the rotor position θr, measured
in electrical radians from the symmetry axis of phase 0 in the conventional axis scheme:
Thanks to the phases being arranged in a conventional structure, it is easy to prove (see
14.3.2.1 for details) that L(lc ) can be written as:
Closed-form expression for dq0 model parameters 79
l0 l1 l2 − l3 − l2 − l1
l1 l0 l1 L − l4 − l3 − l2
l l1 l0 − l5 − l4 − l3
2
Lc =
(l )
M M (6.20)
− l3 − l4 − l5 l0 l1 l2
− l2 − l3 − l4 L l1 l0 l1
− l1 − l2 − l3 l2 l1 l 0
For an alternative expression of (6.20), we can introduce the auxiliary constant n⨯n
matrix
0 1 0 0 0
0 0 1 0 0
M O 0 I n−1 ,n−1
B= = n−1 ,1 (6.21)
0 0 0 1 0 − 1 01 ,n−1
0 0 0 0 1
−1 0 0 L 0 0
where 0h,k indicates the h⨯k null matrix. It is easy to prove that:
644k4 74448
columns 64n−4 7448
k columns
0 0 ⋅ 0 0 1 0 ⋅ 0 0
0 0 ⋅ 0 0 0 1 ⋅ 0 0
⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅
0 0 ⋅ 0 0 0 0 ⋅ 1 0
0 k ×k I k×( n−k )
B k = B B KB = 0 0 ⋅ 0 0 0 0 ⋅ 0 1 = (6.22)
1
424 3
− 1 0 ⋅ 0 0 0 0 ⋅ 0 0 − I( n−k )×k 0( n−k )×( n−k )
k times
0 −1 ⋅ 0 0 0 0 ⋅ 0 0
⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅
0 0 ⋅ −1 0 0 0 ⋅ 0 0
0 0 ⋅ 0 −1 0 0 ⋅ 0 0
n−1
trunc
2
L(cl ) = l 0I n + ∑ l (B
k =1
k
k
− B n−k ) (6.23)
As to the magnetizing term, it will be assumed, as in the previous Chapter, that the mutual
magnetization inductance between two phases whose symmetry axes are placed at
positions φ and ψ respectively can be written as:
80 Closed-form expression for dq0 model parameters
where Lmd and Lmq indicate d-axis and q-axis magnetizing inductances. In a conventionally-
arranged winding scheme, two generic phases of indices i, j (0 ≤ i, j ≤ n−1) have their
symmetry axes placed at iα and jα electrical radians (Fig. 6-1); so their mutual inductance
according to (6.24) is:
Equation (6.25) is the element of indices i, j of the magnetizing inductance matrix L(m )
c ,
where
and
Matrix D is non-zero for both round-rotor and salient-pole machines, while matrix S(θ r ) ,
accounting for saliency, is null in case of round-rotor machines.
Using the geometrical transformation WN⨯3 defined in 6.3, the inductance matrix of the N-
star machine, defined as per (6.2), can be written as:
Closed-form expression for dq0 model parameters 81
L 1 ,1 L L 1 ,N
t
L c = WN×3 M O M WN×3 (6.29)
L
N ,1 L L N ,N
L 1 ,1 L L 1 , N
t
M O M = WN×3 L c WN×3 . (6.30)
L
N ,1 L L N ,N
Λˆ 1 L Λˆ N
M O M = T WN×3 L c WN×3 T t .
t
(6.31)
ˆ t
Λ L Λˆ 1
N
Λˆ 1 L Λˆ N n−1
trunc
t
2
M
ˆ t
O M = T WN×3 l 0 I n +
∑ (
l k B k − B n− k ) WN×3 T
t
Λ L Λˆ 1 k = 1
N
t Lmd + Lmq Lmd − Lmq
+ T WN×3 D+ S(θ r ) WN×3 T t
2 2 (6.32)
n−1
trunc
2
t
= l 0 T WN×3 WN×3 T + ∑ [ t
( )
l k T WN×3 B k − B n−k WN×3 T ]
k =1
By means of a symbolic math tool, it is now easy to prove the following identities:
t
T WN×3 WN×3 T = I n (6.33)
32 0 0 32 0 0
0 3
2
0 0 2 0 L
3
0 0 0 0 0 0
t
T WN×3 DWN×3 T t = 32 0 0 32 0 0 (6.34)
3 3
0 2
0 0 2 0 L
0 0 0 0 0 0
M M
82 Closed-form expression for dq0 model parameters
32 0 0 32 0 0
0 − 32 0 0 − 32 0 L
0 0 0 0 0 0
T WN×3 S(θ r )WN×3 T t = 32 0
t
0 32 0 0 ∀θ r (6.35)
3
0 − 2 0 0 − 32 0 L
0 0 0 0 0 0
M M
t
( )
From the symbolic expansion of T WN×3 B k − B n−k WN ×3T in (6.32) and also based on
(6.33)-(6.35), we finally obtain the closed form expression for the generic bloc Λ̂ N of the
inductance matrix, that is:
3N −1
32 Lmd 0 0 1 0 0 trunc 2
Λˆ k = 0
0
3
L
2 mq
0 + δ k ,0 l 0 0 1 0 +
h=1
∑
l hU h ,k (6.36)
0 0 0 0 1
cos(hα ) sin(hα ) 0
U h ,k = δ mod( h−k ,N ),0 − sin(hα ) cos(hα ) 0
h−k
0 0 (− 1)trunc N
cos(hα ) − sin(hα ) (6.37)
0
+ δ mod( h+k ,N ),0 sin (hα ) cos(hα ) 0 .
h+ k
0 0 (− 1)trunc N
The symbol δ x , y is the Kronecker symbol, which is equal to 0 or 1 respectivly whether x≠y
or x=y. The function mod(m, n) returns the reminder on dividing integer m by n. Therefore
the first term of (6.37) is non-zero if n−k is an integer multiple of N, the second term is
non-zero if n−k is an integer multiple of N.
Using expression (5.25) and comparing it with (6.36), the leakage portion of Λ̂ k can be
written as:
3 N −1
Mk − Xk 0 1 0 0 trunc 2
Xk
0
Mk 0 = δ k ,0 l 0 0 1 0 +
h=1
∑
l hU h ,k (6.38)
0 Hk 0 0 1
where Mk, Xk, Hk are the inductances that appear in the equivalent-circuit representation
and, in particular, Xk represent the d-q cross-coupling coefficients. By substitution of (6.37)
into (6.38) one can derive the explicit closed-form expression of parameters Mk, Xk, Hk as
functions of the physical (measurable or calculable) inductances l k . These expressions
are:
Closed-form expression for dq0 model parameters 83
3N −1
trunc
2
M k = δ k ,0 l 0 + ∑ {l [δ
h=1
h mod( h−k , N ),0 ]
+ δ mod( h+k ,N ),0 cos(hα )} (6.39)
3N −1
trunc
2
Xk = ∑ {l [− δ
h=1
h mod( h−k , N ),0 ]
+ δ mod( h+k ,N ),0 sin(hα )} (6.40)
3N −1
trunc
2 h+ k
h−k
H k = δ k ,0 l 0 + ∑
h=1
trunc trunc
l h δ mod( h−k ,N ),0 (− 1) N + δ mod( h+k ,N ),0 (− 1) N
(6.41)
Equation (6.36), (6.37) allow for the entire transformed inductance matrix (6.7) to be
computed once the magnetizing inductances Lmd, Lmq and the leakage inductances l k are
known, with 0 ≤ k ≤ trunc[ (3N − 1) / 2 ]. The computation of Lmd and Lmq can be conducted
as for three-phase machines (for a deeper insight considering also air-gap space
harmonics, reference can be made to Chapter V). The computation of the leakage
parameters l k is usually more critical and will be addressed in Part III.
84 Model experimental evaluation on sample machines
In this Chapter, some measurements performed on real multi-star machines are used to
compute their transformed models and to verify the results found in the previous two
Chapters.
The focus will be on the leakage inductance portion of the model since the magnetizing
and rotor-related portions have been shown to be quite trivial. The leakage inductance
matrix, transformed into the multiple dq reference frame, will be experimentally evaluated
from measurements and shown to have the form theoretically predicted in the previous
Chapters; furthermore, the relevant equivalent circuits will be drawn assigning a
numerical value to the self and mutual inductances appearing in them.
TABLE I
RATINGS OF THE MACHINES USED FOR TESTING
A2 A3 A4
Type Synchronous generator Induction motor Synchronous motor
Number of phases 2⨯3 3⨯3 4⨯3
Rated voltage 720 V 6600 V 7200 V
Rated current 481 A 520 A 917 A
Rated frequency 5.33 Hz 50 Hz 100 Hz
Number of poles 8 2 4
The prototype data in its three-phase configurations are: 21 kVA, 760 V, 3000 rpm, 2
poles, 0.8 power factor. For a detailed description of the prototype with reconfigurable
winding reference can be made to Section 3.2.
Fig. 7-1. Solved FE analysis models reproducing the test with the rotor removed and one stator phase energized
on machines A2, A3, A4.
Fig. 7-2. Solved FE analysis models reproducing the test with the rotor removed and one stator phase energized
on the reconfigurable prototype machines in multi-star configurations B2, B3, B4.
Energizing one stator phase with an AC current I at frequency f after rotor removal and
measuring the e.m.f. Ek induced in the phase displaced by an electrical angle kα with
respect to the energized phase, the inductance parameters l(rr
k
)
is computed as:
Ek
l(krr ) = . (7.1)
2π f I
Fig. 7-3. Leakage inductances ℓk experimentally obtained on the machines used for testing.
The values obtained for l k with the above procedure are shown in Fig. 7-3.
l0 l1 l2 0 − l2 − l1
l1 l0 l1 l2 0 − l2
l l1 l0 l1 l2 0
L(cl ,)6 = 2 (7.3)
0 l2 l1 l0 l1 l2
− l2 0 l2 l1 l0 l1
−l − l2 0 l2 l1 l 0
1
l0 l1 l2 l3 l4 − l4 − l3 − l2 − l1
l1 l0 l1 l2 l3 l4 − l4 − l3 − l2
l l1 l0 l1 l2 l3 l4 − l4 − l3
2
l3 l2 l1 l0 l1 l2 l3 l4 − l4
L(cl ,)9 = l4 l3 l2 l1 l0 l1 l2 l3 l4 (7.4)
− l4 l4 l3 l2 l1 l0 l1 l2 l3
− l3 − l4 l4 l3 l2 l1 l0 l1 l2
−l − l3 − l4 l4 l3 l2 l1 l0 l1
2
− l1 − l2 − l3 − l4 l4 l3 l2 l1 l 0
l0 l1 l2 l3 l4 l5 0 − l5 − l4 − l3 − l2 − l1
l1 l0 l1 l2 l3 l4 l5 0 − l5 − l4 − l3 − l2
l l1 l0 l1 l2 l3 l4 l5 0 − l5 − l4 − l3
2
l3 l2 l1 l0 l1 l2 l3 l4 l5 0 − l5 −l4
l4 l3 l2 l1 l0 l1 l2 l3 l4 l5 0 − l5
l l4 l3 l2 l1 l0 l1 l2 l3 l4 l5 0
L(cl ,)12 = 5 . (7.5)
0 l5 l4 l3 l2 l1 l0 l1 l2 l3 l4 l5
−l 0 l5 l4 l3 l2 l1 l0 l1 l2 l3 l4
5
− l4 − l5 0 l5 l4 l3 l2 l1 l0 l1 l2 l3
−l − l4 − l5 0 l5 l4 l3 l2 l1 l0 l1 l2
3
− l2 − l3 − l4 − l5 0 l5 l4 l3 l2 l1 l0 l1
− l1 − l2 − l3 −l4 − l5 0 l5 l4 l3 l2 l1 l0
where (7.4) is used for machines A2, B2; (7.5) for machines A3, B3; (7.6) for machines A4,
B4.
The geometrical transformation matrices W2×3 , W3×3 , W4×3 which map the multi-star
winding scheme into the conventional one have been the computed as per Chapter 6
obtaining:
1 0 0 0 0 0
0 0 0 1 0 0
0 0 −1 0 0 0
W2×3 = . (7.6)
0 0 0 0 0 − 1
0 1 0 0 0 0
0 0 0 0 1 0
88 Model experimental evaluation on sample machines
1 0 0 0 0 0 0 0
0
0 0 0 1 0 0 0 0
0
0 0 0 0 0 0 1 0
0
0 0 −1 0 0 0 0 0
0
W3×3 = 0 0 0 0 0 −1 0 0 0 (7.7)
0 0 0 0 0 0 0 0 − 1
0 1 0 0 0 0 0 0 0
0 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 1 0
1 0 0 0 0 0 0 0 0 0 0 0
0 0 0 1 0 0 0 0 0 0 0 0
0 0 0 0 0 0 1 0 0 0 0 0
0 0 0 0 0 0 0 0 0 1 0 0
0 0 −1 0 0 0 0 0 0 0 0 0
0 0 0 0 0 −1 0 0 0 0 0 0
W4×3 = (7.8)
0 1 0 0 0 0 0 0 −1 0 0 0
0 0 0 0 0 0 0 0 0 0 0 − 1
0 1 0 0 0 0 0 0 0 0 0 0
0 0 0 0 1 0 0 0 0 0 0 0
0 0 0 0 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 0 0 0 1 0
In accordance with Chapters 5 and 6, the transformation matrices TN⨯3 are then defined
as:
T1 03 03 03
T1 03 03
T 03 0 T2 03 03
T2×3 = 1 , T3×3 = 03 T2 03 , T4×3 = 3 (7.9)
03 T2 0 0 T 0 03 T3 03
3 3
3
3 0 0 0 T4
3 3 3
The transformed inductance matrix for N=2, N=3 and N=4, at this points, are respectively
computed as:
Lˆ(dq
l) t (l ) t
,2×3 = T2×3 W2×3 L c ,6 W2×3 T2×3 (7.11)
Lˆ(dq
l) t (l ) t
,3×3 = T3×3 W3×3 L c , 9 W3×3 T3×3 (7.12)
Model experimental evaluation on sample machines 89
Lˆ(dq
l) t (l ) t
, 4×3 = T4×3 W2×3 L c ,12 W4×3 T4×3 (7.13)
The numerical results of the evaluations for machines A2, A3, A4 are given below:
1.377 0 0 0.812 0 0
0 1.377 0 0 0.812 0
0 0 0.75 0 0 0
Lˆ(dq
l)
,2×3 = mH (7.14)
0.812 0 0 1.377 0 0
0 0.812 0 0 1.377 0
0 0 0 0 0 0.75
The numerical evaluation results for the configurations B2, B3 and B4 of the prototype
machine are reported below:
5.8 0 0 4.2 0 0
0 5.8 0 0 4.2 0
0 0 2.6 0 0 0
Lˆ(dq
l)
,2×3 = mH (7.17)
4.2 0 0 5.8 0 0
0 4.2 0 0 5.8 0
0 0 0 0 0 2.6
90 Model experimental evaluation on sample machines
The numerical forms found for all the matrices match theoretical expectations according
to Chapter 5 and Chapter 6. In particular, the presence of non-null (even small) d-q cross
coupling elements is confirmed, although their magnitude appears small compared to the
other matrix entries.
2
M k = δ k ,0 l 0 + ∑ {l [δ
h=1
h mod( h−k ,2),0 ]
+ δ mod( h+k ,2),0 cos(hα )} (7.20)
2
Xk = ∑ {l [− δ
h=1
h mod( h−k ,2 ),0 ]
+ δ mod( h+k ,2),0 sin (hα )} (7.21)
2
h−k h+ k
H k = δ k ,0 l 0 + ∑ l
h=1
δ
h mod( h− k ,2),0 (− 1)trunc 2
trunc
+ δ mod( h+k ,2),0 (− 1)
2
(7.22)
4
M k = δ k ,0 l 0 + ∑ {l [δ
h=1
h mod( h−k ,3 ),0 ]
+ δ mod( h+k ,3),0 cos(hα )} (7.23)
4
Xk = ∑ {l [− δ
h=1
h mod( h−k ,3),0 ]
+ δ mod( h+k ,3),0 sin(hα )} (7.24)
4
h−k h+ k
H k = δ k ,0 l 0 + ∑ l
h=1
δ
h mod( h− k ,3),0 (− 1)trunc 3
+ δ mod( h+ k ,3),0 (− 1)
trunc
3
(7.25)
5
M k = δ k ,0 l 0 + ∑ {l [δ
h=1
h mod( h−k , 4 ),0 ]
+ δ mod( h+k , 4 ),0 cos(hα )} (7.26)
5
Xk = ∑ {l [− δ
h=1
h mod( h−k , 4 ),0 ]
+ δ mod( h+k , 4 ),0 sin (hα )} (7.27)
5
h−k h+ k
H k = δ k ,0 l 0 + ∑ l
h=1
δ
h mod( h−k , 4 ),0 (− 1)trunc 4
trunc
+ δ mod( h+k , 4 ),0 (− 1)
4
(7.28)
where δi,j indicates the Kronecker symbol, equal to 0 or 1 depending on whether i≠j or i=j,
and mod(m,n) is the reminder on dividing integer m by integer n.
The numerical computation results for the six machine typologies considered are
summarized in Table I.
TABLE II
DQ0 LEAKAGE INDUCTANCE PARAMETERS COMPUTED FROM MEASUREMENTS (VALUES IN MILLI-HENRIES)
Machine
A2 A3 A4 B2 B3 B4
Parameter ID
M2 — — 0.270 — — 1.84
The fact that some parameters do not appear for some multi-star configuration is due to
the fact that the number of independent parameters that appear in the transformed model
depend on the number N of stator stars, as discussed in Chapter 5.
If we compare the parameters shown in Table II with the numerical matrices reported in
the previous Section, by bearing in mind the general matrix structure derived in Chapter 5,
we can conclude that the two methods lead exactly to the same results.
92 Model experimental evaluation on sample machines
Fig. 7-5. Homopolar equivalent circuits of the machines tested with relevant parameter numerical values in mH.
Fig. 7-4. d-q equivalent circuits of the machines tested with relevant parameter numerical values in mH.
Fig. 7-6. Homopolar equivalent circuits of the machines tested with relevant parameter numerical values in mH.
dq0 coordinates have been determined. As a result, the theoretical inductance matrix and
equivalent circuit topologies expected from the theory presented in the two previous
Chapters have been fully confirmed.
In particular, measurements have confirmed that while for N=2 no d-q cross-coupling
phenomenon occurs, for higher values of N some mutual inductances appear which link
the d-axis and q-axis equivalent circuit as a result of the stator winding leakage flux. The
magnitude of such cross-coupling inductances, however, has been numerically evaluated
and shown to be much smaller than the other leakage inductance parameters involved in
all the cases under examination.
94 Model Simulink implementation and experimental validations
d
v dq0 = R dq0 i dq0 + ω J L dq0 + L dq0 i dq0 (8.1)
dt
where R dq0 , L dq0 and J are constant matrices, ω is the electrical speed in radians per
second and vector variables v dq0 , i dq0 are linked to the natural machine variables in
multiphase coordinates ( v abc , i abc ) through the transformation T(θ) depending on rotor
position θ:
The machine instantaneous electromagnetic torque Tem has been derived as:
t
Tem = pi dq0 J L dq0 i dq0 (8.3)
d ω ω
Tem − Text = J +B (8.4)
dt p p
where Text is the torque applied to the machine shaft, J is the inertia coefficient and B the
viscous friction coefficient.
d
i dq0 = L dq0 (v dq0 − R dq0 i dq0 − ω J L dq0 i dq0 ) .
−1
(8.5)
dt
Model Simulink implementation and experimental validations 95
Fig. 8-1. Block-scheme used to integrate multi-star electric machine model equations. Symbol”⨯” denotes
multiplication of a vector by a scalar, symbol “•” denotes inner product.
d ω 1 B
= (Tem − Text ) − ω . (8.6)
dt p J J
At this point, it is natural that the system of differential equations (8.5)-(8.6) can be
integrated in ω and i dq0 with Text and v dq0 as forcing (input) functions, making use of (8.3)
for Tem. The block scheme used to integrate the model is depicted in Fig. 8-1.
The input and output variables vabc and iabc are vector signals which, according to the
conventions used in Chapter 5, have the following structure:
v abc ,1 i abc ,1
v abc ,2 i abc ,2
v abc = ,i =
M abc M
(8.7)
v i
abc ,N abc ,N
where the generic sub-vectors v abc ,k and iabc ,k respectively contain the triplets of phase
voltages (vak, vbk, vck) and currents (iak, ibk, ick) pertaining to the kth star of the machine
winding (1 ≤ k ≤ N):
va ,k ia ,k
v abc ,k = vb ,k , iabc ,k = ib ,k . (8.8)
v i
c ,k c ,k
Therefore, in order to access the single phase quantities, some multiplexing and de-
multiplexing operations are needed on vector variables vabc and iabc, as illustrated in Fig.
8-2.
When the model is implemented in the Matlab/Simulink environment, the phase current
and voltage variables are scalar signals that can be used as they are or be somehow
interfaced with “power circuit” blocks, representing the power electronics equipment or
96 Model Simulink implementation and experimental validations
Fig. 8-2. Multiplexing and de-multiplexing operations on variables iabc and vabc exchanged with the machine
block scheme model (Fig. 8-1).
Fig. 8-3. Interfacing phase voltage and current Simulink signals with SymPowerSystems blocks.
the grid elements to which the machine is physically connected. These power circuit
blocks are often conveniently implemented using a dedicated Simulink library called
Model Simulink implementation and experimental validations 97
Fig. 8-4. (a) Dual-star machine used for validation; (b) and (c): test set-up including the LCI converters which
supply the machine and the DC generator used as a load.
98 Model Simulink implementation and experimental validations
Fig. 8-5. Equivalent circuit of the dual-star machine used for validation. Parameters are in per unit of the base
impedance Z = 0.667 Ω.
Fig. 8-6. Simulink block scheme for the simulation of a dual-star synchronous motor supplied by two Load
Commutated Inverters.
The ratings of the dual star machine are given in Table I. Its parameter characterization in
terms of equivalent circuit is given in Fig. 8-5 (the equivalent circuit parameters have been
derived from previous testing).
Model Simulink implementation and experimental validations 99
TABLE I
RATINGS AND CHARACTERISTICS OF THE DUAL-STAR MACHINE USED FOR VALIDATION
A2
Type Brushless synchronous machine
Rotor type Round solid-steel wound rotor
Rated voltage 2 ⨯ 1200 V
Rated current 2 ⨯ 517 A
Rated frequency 210 Hz
Number of poles 4
Number of phases 2⨯3
The Simulink block scheme used to simulate the drive operation is shown in Fig. 8-6: the
dual-star motor model is interfaced with the blocks, built using the SimPowerSystem
library, which represent the Load Commutated Inverters and the field excitation system.
Stator phase flux linkage and the actual machine speed are picked up as machine model
output and brought to the LCI blocks as feedback signals used for synchronizing SCR firing
pulses with stator voltages [6].
The comparison between simulation and measurement results is shown in Fig. 8-7, where
current and voltage waveforms are considered in two steady-state operating modes
summarized in Table II.
TABLE II
OPERATING MODES OF THE LCI DRIVE USED FOR VALIDATION
First operating mode Second operating mode
Frequency 100 Hz 100 Hz
DC link current 720 A 720 A
Number of active inverters 2 1
SCR firing angle 150 deg 155 deg
In both operating modes the drive operates at 100 Hz and with a 720 A DC link current. In
the first operating mode, however, both LCIs are supplying the motor, while in the second
operating mode one of the two LCIs is disconnected and the corresponding machine star is
at no load.
The comparison reported in Fig. 8-7 shows a good accordance between the simulated and
measured waveforms in both operating modes. The same accuracy is found in other drive
working conditions differing by load, LCI firing angle and field excitation current.
8.4 References
[1] S. Castellan, R. Menis, M. Pigani, G. Sulligoi, A. Tessarolo, “Modeling and simulation of
electric propulsion systems for all-electric cruise liners”, IEEE Electric Ship
Technologies Symposium, IEEE ESTS 2007, 21-23 May 2007, Arlington, VA, USA, pp. 60-
64.
[2] S. Castellan, G. Sulligoi, A. Tessarolo, “Comparative performance analysis of VSI-fed and
CSI-fed supply solutions for high power multi-phase synchronous motor drives”,
100 Model Simulink implementation and experimental validations
Fig. 8-7. Comparison between voltage and current waveforms resulting from numerical simulation and
measurements for two different operating modes of the dual-star synchronous motor.
In Part II, the modeling has been presented of multiple-star machines with an extended
Clark’s and Park’s approach. In the final expression of the model, a certain number of leakage
inductive parameters have appeared along with the usual magnetizing and rotor-related
parameters common to three-phase machines. Such leakage inductance parameters must be
determined in some way for the machine model to be fully identified and usable.
The Finite Element Method (FEM) is certainly a possible solution for approaching the
problem. However, it suffers from well known disadvantages due to the high computational
resources and times it often requires and its poor flexibility with respect to geometric model
variations.
The approach investigated in this Part is analytical and employes the design data which
characterize machine winding and dimensions as input information. Following most of the
literature on the subject, leakage flux is classified into end-coil, slot and air-gap components.
Chapter 9 deals with end-coil leakage inductance computation; Chapter 10 is devoted to slot
and air-gap leakage inductances and it also presents the experimental validations which
have been conducted to assess the accuracy of the proposed calculation techniques.
It is important to observe that the same leakage inductance parameters whose computation
is covered in this Part will be also used in Part V for expressing multiphase machine model
through the Vector-Space Decomposition (VSD) technique.
End-winding leakage inductance computation 103
9.1 Introduction
Multi-phase machines are of increasing importance in today’s electric drives [1]. An issue
in their modeling and design is the computation of stator leakage inductances, which may
be crucial in limiting the extra harmonic currents that originate in case of supply from
multiple three-phase voltage source inverters [22]. The leakage inductance portion due to
the slot leakage flux is relatively simple to predict [22], [4] based on the approximate
magnetic field distribution inside the motor slots. A more challenging task is to predict the
leakage inductance portion due to stator end-coils. This is generally done through
simplifying assumptions: in some authors’ approach only self leakage inductances are
considered [3] for example; in other works the assumption is made that the end-coil
leakage flux distribution around the stator periphery is the same as for slot leakage flux
[22]. A more accurate evaluation could be performed resorting to 3D finite element (FE)
techniques [11], [12]. The drawback of this approach is that it requires the construction of
complicated models for any design configuration to be analyzed. An alternative method for
end-coil leakage inductance calculation of three-phase turboalternators was proposed in
[5] resorting to Neumann integrals and to the principle of images as the main theoretical
background. In this Chapter the analytical-numeric approach based on Neumann integrals
and on the method of images is extended to the case of multi-phase machines with a
generic phase number and arrangement. Some original contributions are also proposed
with regard to the computation of end-coil self inductance, for which the Neumann
integral method cannot be directly used. Two algorithms are presented for this purposes.
Finally, the experimental validation of the method through measurements on three multi-
phase machines (6-phase, 9-phase and 12-phase) with different number of poles are
presented, showing a good accordance with the calculation results.
(a) (b)
Fig. 9-1. End-coils in a stator imbricated double-layer winding, during assembly (a) and after assembly (b).
104 End-winding leakage inductance computation
Fig. 9-2. Curves γ and γint used for coil geometric modeling.
For the purpose of mutual inductance calculation, the current is supposed to flow along an
infinitely thin curve γ passing through the center of the end-coil cross-section (Fig. 9-2).
For the self inductance calculation, a slightly more accurate model will be adopted,
including the auxiliary end-coil profile γaux and the end-coil cross-section, approximated
with a rectangle of sides h, w.
The 3D geometry and dimensions that characterize the curve γ are represented in Fig. 9-3.
In particular, Rsup and Rinf (such that Rsup−Rinf=h) are the distances of the two active coil
sides from the machine axis and the coil span β equals 2πr/np being np the number of
Fig. 9-4. Approximation of the end-coil curve γ with straight segments Si(γ ) , with i=1..2N+5 (case of N=3).
machine poles and r the winding coil pitch. Finally, the xyz orthogonal reference frame is
set so that z lies along the machine axis and x, y lie on the stator core end plane.
The slant end-coil portions C2C3 and C6C7 (Fig. 9-3) are approximated as helical arcs, lying
on the cylindrical surfaces of radius Rinf and Rsup respectively. In order to facilitate
computations, the helical curves are replaced by sequences of N straight filaments of equal
length whose end points belong to the curves themselves. Each end-coil is then
approximated with a set of 2N+5 straight filaments (Fig. 9-4).
Fig. 9-6. Flux lines produced by an end-coil element P in presence of stator and rotor materials (a) and from the
method of images with µr=∞ (b).
Fig. 9-7. End-coils connecting two active coil side paris ab and cd displaced by 180 mechanical degrees.
Fig. 9-8. Flux lines produced on PQz plane by two end-coil elements P, Q displaced by 180°, neglecting the shaft
effect (a), and considering shafts of different diamenters (b), (c).
infinitely long straight lines δ’ and ε’, parallel to z axis and respectively beginning from
points C1 and C8 [5].
The end-coil mirror image carries an auxiliary current I’ and the two infinite straight lines
a current I+I’, with the directions shown in Fig. 9-4b and with I’ given by [5]:
To assess the appropriateness of this approach and for a realistic choice of parameter µr, a
FE analysis can be of help. In fact, let P be a generic infinitesimal end-coil element and P’
its mirror image (Fig. 9-5b). The magnetic flux lines produced by P in the plane containing
P and z axis, taking stator and rotor magnetic materials into account, are represented in
Fig. 9-6a. For the FE simulation, a solid iron rotor and a laminated stator core are assumed
with laminations parallel to xy and a filling factor of 0.94; the relative permeability of both
stator and rotor magnetic materials is set µr=7000 but any further increase in µr does not
produce any appreciable changes. It can be seen from Fig. 9-6 that the average angle of
incidence of the magnetic field on the z=0 plane is very close to 90° in the stator region,
hence it is reasonable to set µr=∞, which implies I’=I according to (9-2). Fig. 9-5b shows
the flux lines obtained from the method of images in this hypothesis, i.e. with the image
element P’ carrying the same current as P. A comparison between Fig. 9-5a and Fig. 9-5b
confirms the validity of the method of image as far as the stator core effect is concerned.
Regarding the effect of the rotor shaft, the method of images would more difficult to apply
because of the non-planar geometry of the reflecting surface. Fortunately, this problem is
alleviated by the fact that for each end-coil element P, there is an element Q, symmetrical
to P with respect to z axis, displaced by 180 mechanical degrees and carrying the same
current as P (Fig. 9-7). Though Q does not coincide with its mirror image P” with respect to
the shaft surface element Σ, we observe that the smaller the shaft diameter, the closer Q to
P”. This concept is illustrated by FE analysis as shown in Fig. 9-8: supposing to neglect the
shaft effect (Fig. 9-8a), the flux distribution in the end-coil region does not differ
significantly from the case when the shaft effect is considered (Fig. 9-8b, 8c). This is
particularly true for relatively small shaft diameters (Fig. 9-8b).
To summarize, for the leakage inductance calculation the stator core effect is taken into
account through the method of images, setting µr=∞, whereas the rotor shaft effect is
ignored.
Fig. 9-9. Two end-coils η, γ, displaced by ∆θ mechanical radians, with the fictitious conductors γ’, ε’, δ’ required by
the method of images.
108 End-winding leakage inductance computation
µ0 ds × dt ds × dt ds × dt ds × dt
M( Δθ ) = ∫∫
4 π s∈η s − t
+
s∈η
∫∫
s−t
+2
s∈η
s−t
+2 ∫∫
s∈η
s−t
∫∫ (9-2)
t∈γ t ∈γ ' t ∈ε ' t∈δ '
where “×” indicates the inner product and ||.|| indicates the Euclidean norm. If curves η, γ
and γ' are approximated respectively with the sequences of straight oriented
filaments S i(η ) , Si(γ ) , S i(γ ') (1≤i≤2N+5) as described above, (9-2) becomes:
2N +5
µ0
M( Δθ ) = ∑ [N (S η
4 π i , j =1
( )
i , S (jγ ) ) + N ]
( S i(η ) , S (jγ ') ) +2 N ( S i(η ) ,δ ' ) + 2 N ( S i(η ) , ε ' ) (9-3)
The function N( AB , ab ), where AB and ab are two finite oriented straight segments,
indicates the Neumann integral:
ds × ds ′
N ( AB, ab) = ∫ s − s′ (9-4)
s∈ab , s′∈ AB
The closed form analytical solution of (9-4) is available from [6] for all possible reciprocal
positions of AB and ab , on condition that the two segments do not overlap (except for the
end points at most). This enables to compute all the terms of (9-3) in square brackets. As
concerns the last two terms, they cannot be directly computed with the help of [6] because
they involve infinitely long filaments δ’, ε’. Therefore the formula given by [5] for Neumann
integrals involving infinitely long anti-parallel straight filaments is used to compute each
of the sums N ( S i(η ) , δ ' ) + N ( S i(η ) , ε ' ) .
The computation of (9-4) is to be repeated for any possible displacement ∆θ between the
two end coils. It is easy to realize that the number of possible mutual displacements ∆θ
between two end-coils equals the number of stator slots Z. In particular, being the slot
pitch αs defined as
α s = 2π/Z (9-5)
1
The conceptual difficulty of η not being a closed circuit is easily overcome observing that the same identical situation occurs at
the opposite side of the machine, so that what is being computed is actually half the flux linkage of the closed circuit formed by the
two end-coils located at opposite machine sides. A theoretical discussion on the use of self and mutual inductance concepts for non-
closed circuits can be found in [7].
End-winding leakage inductance computation 109
we can compute the sequence of inductances Mk, for k=0..Z, such that M k = M (α s k ) . In
words, Mk for 1≤k≤Z−1 is the mutual inductance between two end-coils displaced by k
slots apart, whereas M 0 = M (0) = M Z = M (α s Z ) is the self inductance of one end-coil.
The formulas given by [5], [6] for the closed-form solution of (9-3) have been
implemented in the Mathcad environment and used to compute the sequences of mutual
inductances Mk (1≤k≤Z−1) in the practical case of the three multi-phase machines M1, M2,
M3 referred to later in the Chapter. The resulting values of Mk are represented with the
dotted diagrams of Fig. 9-10 (one dot per value).
110 End-winding leakage inductance computation
Fig. 9-11. Substitution of a straight infinitely thin filament with the corresponding end-coil volume.
Unfortunately, (9-3) cannot be used to find the end-coil self inductance. In fact, when the
curve η coincides with γ (Fig. 9-9), all Neumann integrals N ( S i(η ) , S (jγ ) ) for i=j become
divergent because Si(η ) ≡ S (jγ ) . Two possible ways to overcome this difficulty are presented
hereinafter; they will be conventionally referred to as method “a” and method “b”.
I dV
I ds = I ds uˆ Z = j dV = dV uˆ Z ⇒ ds = (9-7)
hw hw
Substituting ds with dV/(hw) in (9-4), the divergent term N ( S i(γ ) , S i(γ ) ) is replaced by the
following double volume integral:
l i h w l i h w
1 dV dV ′ 1 dX ′dY ′dZ ′
L(iγ ) = ∫ =
(h w )2 s , s′∈V ( γ ) s − s′ (h w )2 ∫ ∫∫ ∫ ∫∫ dXdYdZ
( X − X ′)2 + (Y − Y ′)2 + ( Z − Z ′)2
(9-8)
i
00 0 0 0 0
End-winding leakage inductance computation 111
which is proved to be convergent [8]. The physical reason why (9-8) converges and
N ( S i(γ ) , S i(γ ) ) does not is that modeling the end coil as an infinitely thin filament is an
approximation that holds only when the magnetic field outside the coil is of interest,
whereas the computation of self inductance also involves the magnetic field inside the
conductor and in its close neighborhood.
In the hypotheses that the cross-section sides are much smaller than the length l i , i.e.
w2 h2 h2 w2 4w h
L(iγ ) ≈ l i 2 ln(2l i ) − 2 ln(w ) − 2
ln( h) −
1 − 2
− ln w 2 + h2 −
2
(
arctan )
3h 3w 6w 6h 3 h w
(9-10)
4 h w 11
− arctan + .
3w h 6
2l i 1
L(iγ ) ≈ l i 2 ln −K + (9-11)
h+ w 2
According to method “b”, the end-coil self inductance M0 can be then computed by
replacing N ( S i(γ ) , S i(γ ) ) with L(iγ ) , given by either (9-10) or (9-11), into (9-3) written for
∆θ=0. In symbols:
µ 0
M 0 = M (0) = ∑[ N ( S i(γ ) , S (jγ ) ) + N ( S i(γ ) , S (jγ ') )
4 π i , j =1..2 N + 5
i ≠ j (9-11)
] ∑L
+ 2 N ( S i(γ ) , δ ' ) + 2 N ( S i(γ ) , ε ' ) + (γ)
i
i =1..2 N + 5
M 0 = Lext + ΔL (9-12)
Regarding the self inductance Lext, its computation is reduced back to the computation of
the mutual inductance between two distinct circuits, namely the central line γ and the
peripheral line γaux. Moreover, in order to account for the stator core effect through the
method of images, the fictitious conductors γ’, δ’, ε’ need to be introduced too (Fig. 9-12) as
discussed earlier.
The external self-inductance Lext can be then expressed by (9-3) with the position η =γaux,
which yields:
2N + 5
µ0
Lext = ∑ [N (S γ
4 π i , j =1
(
i
aux )
]
, S (jγ ) ) + N ( S i(γ aux ) , S (jγ ') ) +2 N ( S i(γ aux ) , δ ' ) + 2 N ( S i(γ aux ) , ε ' ) (9-13)
Regarding the correction ∆L, one can observe that Lext computed above is the flux
produced by a unitary end-coil current across a semi-surface Σ having γ as its contour (Fig.
9-13), the complementary semi-surface pertaining to the end-coil at the opposite machine
side. If the end-coil could be modeled as a circular cross-section conductor (Fig. 9-14a), ∆L
would simply equal the end-coil internal inductance, namely µ 0 l /(8π ) [10], denoting the
conductor length with l . Taking the rectangular cross-section into account, the situation
is slightly more complicated. In fact, let us refer to Figure Fig. 9-14b, which represents a
portion of end-coil of length ∆l and the corresponding portion of surface Σ. Let us denote
the flux through ∆Σ with Φ and consider the three points P, Q, R respectively outside,
upon and inside the flux line κ tangent to the end-coil vertical sides. The end-coil
elementary conductor passing through R has a flux linkage greater than Φ (due to the
contribution of the flux lines between R and κ); the elementary conductor passing through
Q has a flux linkage equal to Φ; finally, the elementary conductor passing through P has a
flux linkage lower than Φ (as the flux lines between P and κ do not link it). Based on this
qualitative reasoning on can observe that both positive and negative contributions are to
be added to Lext for the total self-inductance calculation; moreover, ∆L is expected to
strongly depend on cross-section aspect ratio.
The quantitative approach to find ∆L is illustrated in Fig. 9-14c and 14d. The magnetic flux
due to a current I uniformly distributed over a rectangular cross section of sides h, w is
first determined with a FE analysis, setting a model depth equal to ∆l . Next the flux φVT
through the segment VT (trace of the surface ∆Σ) is computed, with V located at a very
large distance from the cross section ( VT ≈ 200 w ). The total flux linkage φtot of the cross
section is determined as φtot=2Wtot/I, where Wtot is the total magnetic energy over the
entire domain. Hence, the correction ∆L is found as:
Repeating the procedure for several values of h/w ratio, the points of Fig. 9-15 are found.
The adimensional quantity ∆L 8π /( µ 0 ∆l) is diagrammed in order to empathize that, if the
cross-section were approximated with a circle, the constant value of 1 would be obtained
regardless of the circle radius. In case of rectangular cross section, instead, one can see
that the correction term ∆L decreases for increasing h/w and becomes negative for
h/w>2.5.
The values of ∆L 8π /( µ 0l) for 1≤h/w≤ 10 are well interpolated (Fig. 9-15) by the following
hyperbolic equation:
8π k1
ΔL = − k3 (9-15)
µ 0 Δ l h / w + k2
Writing (9-15) for all the 2N+5 straight parallelepipeds that approximate the end-coil (Fig.
9-2), each of length l i , and summing over i yields the total correction
µ0l k1
− k3 (9-16)
8π h / w + k2
where l = ∑ l i is the total end-coil length. Summing (9-16) to (9-13) finally gives the
i
µ 0
2 N +5
M 0 = Lext + ∆L = ∑[
4 π i , j =1
N ( Si (γ aux )
, S (jγ ) ) + N ( S i(γ aux ) , S (jγ ') ) +
(9-17)
l
+2N ( S i(γ aux ) , δ ' ) +2N ( S i(γ aux ) , ε ' ) ] +
k1
− k3
2 h / w + k2
Actually, one should not expect a perfect agreement because both methods rely on some
approximations that may fit the physics of the studied configuration to a different extent.
In particular, the analytical solution of (9-8), expressed by either (9-9) or (9-10), is
derived under the hypothesis (9-9), which could be not well fulfilled if a high number N of
short segments are used to approximate end-coil shape.
End-winding leakage inductance computation 115
Between the two proposed methods, “b” is deemed the most reliable and is chosen for
experimental validation (see Section 9.6). In fact, zooming on the diagrams of Fig. 9-10,
one should reasonably expect that M0 lies on or above the curve interpolating the mutual
inductances Mk, which is not the case if method “a” is employed (Fig. 9-16).
With the above conventions it is straightforward to realize that the coils in the pole p
(0≤p≤np−1) belonging to the phase h (0≤h≤n−1) are defined by the following set of indices:
U h , p = {z : hq + pQ ≤ z ≤ (h + 1) q − 1 + pQ} (9-18)
Given two phases h and h’, the computation of their mutual inductance mh,h’ due to end-coil
Fig. 9-17. Conventional phase names and directions for a n-phase double-layer shortened-pitch winding.
116 End-winding leakage inductance computation
U1
P1 −V 2 U2
− P6 P2
− P5 P2 − V1 −W1
− P4 P3 W2 −W 2
− P3 P5 W1 V1
− P2 P6 −U 2 V2
− P1 − U1
Fig. 9-18. Example of phasor diagrams for an 6-phase winding: conventional phase arrangement (a); dual three-phase configuration
(b).
leakage flux requires summing the mutual inductances between all their end-coils, i.e.:
n p −1 n p −1
nt 2
mh, h ' = 2
nw 2
∑ ∑ ∑ ∑ (−1)
p = 0 p '= 0 i∈U h , p i '∈U h ', p '
p+ p'
M i −i ' =
where nt and nw respectively denote the number of turns in series per coil and the number
of parallel ways per phase. The initial coefficient 2 is needed to account for the end-coils
on both machine sides; the factor (−1)p+p’ takes into consideration that the conventional
sign of the coils in pole p can be expressed as (−1)p, so that the mutual inductance between
an end-coil in pole p and an end-coil in pole p’ has sign (−1)p(−1)p’=(−1)p+p’.
mh,h’=mh’,h (9-20)
in accordance with the reciprocity theorem for mutual inductances. Moreover, with simple
algebraic manipulations (9-19) can be written in the equivalent form:
n p −1 q −1
nt 2
mh, h ' = 2
nw 2
∑ ∑ (−1)
p , p ′ = 0 i ,i ′ = 0
p+ p'
M Q ( p − p′) + q ( h − h′) + (i −i′) (9-21)
which highlights that mh,h’ only depends on the difference h−h’ and in particular:
with k=0..n−1. From a physical point of view, µk represents the mutual inductance between
two phases displaced by qk slots (i.e. k phase belts) apart.
Properties (9-20) and (9-22) say that the matrix [mh,h’] of the end-coil phase leakage
inductances is a symmetrical Toeplitz matrix, i.e. has the form:
End-winding leakage inductance computation 117
µ0 µ1 µ2 L µ n −1
µ µ0 µ1 L µ n − 2
1
[mh,h' ] = µ2 µ1 µ0 L µ n −3 (9-24)
M M M O M
µ n −1 µ n−2 µ n −3 L µ 0
This is an important result because it assures that the entire system of the self and mutual
end-coil inductances of the machine is completely determined if the n quantities µk are
determined through (9-19).
It is straightforward to write the transformation matrix W that links the electrical phasors
in the cases of Fig. 9-18a and 18b as follows (superscript “t” indicates matrix
transposition):
[P1 P 2 P 3 P 4 P 5 P 6] = W [U1 V1 W1 U2 V2 W2 ]
t t
(9-25)
1 0 0 0 0 0
0 0 0 1 0 0
0 0 −1 0 0 0
W = (9-26)
0 0 0 0 0 − 1
0 1 0 0 0 0
0 0 0 0 1 0
Therefore, assuming the phase variables of the dual three-phase system arranged as in the
right-hand side of (9-25), the end-coil mutual inductance (9-24) becomes:
µ0 µ4 − µ2 µ1 µ5 − µ3
µ µ0 − µ2 µ3 µ1 − µ1
4
− µ − µ2 µ0 − µ1 − µ3 µ1
[
W t mh , h ' ]
W= 2 (9-27)
µ1 µ3 − µ1 µ0 µ4 − µ2
µ5 µ1 − µ3 µ4 µ0 − µ2
− µ 3 − µ1 µ1 − µ2 − µ2 µ 0
The same procedure can be naturally applied to any phase number and arrangement.
unit M1 M2 M3
np - 8 2 4
n - 2×3 3×3 4×3
D, E, Rinf mm 195 , 237, 931 32, 368, 341 94, 301, 566
h, w mm 45, 14 41, 17 63, 15
Z, nw, nt. r - 192, 2, 3, 0.79 54, 1, 15, 0.63 96, 2, 3, 0.83
The experimental set-up is sketched in Fig. 9-19. The tests were performed on the
machine wound stator with the rotor removed. The machine phases are marked according
to the scheme of Fig. 9-19. The phase 0 was supplied with a voltage V0 and current I0 at
fixed frequency f0, with all the other phases left open. The induced voltage Vk on the the k-
th open phase was measured, for any k=1..n−1, along with the supply voltage and current
V0, I0. The mutual inductance µ k(test ) between the phase 0 and the phase k was computed as
µ k(test ) = Vk /(2πf 0 I 0 ) . The sign of µ k(test ) was determined based on the sign of the reactive
power, measured by a wattmeter, associated with the pair Vk, I0. In fact, neglecting
resistance effect, the induced voltage Vk is at 90° leading or lagging with respect to I0
depending on whether µ k(test ) is positive or negative.
The mutual inductance µ k(test ) is due to the end-coil leakage flux and to the flux produced in
the stator core region. As proposed in [5], the latter contribution can be calculated with
good accuracy from a FE analysis. By means of FE analysis (Fig. 9-22) it was possible to
determine the mutual inductance term µ k(core ) due to the flux in the core region with the
same technique as described in [5].
The end-coil inductance term was then computed as µ k(test ) − µ k( core) . This value was
compared to the inductance µ k obtained from calculation by means of (9-19), for all k
values between 0 and n−1 (in particular, for k=0 the self inductance is obtained). The
results are displayed in Fig. 9-21.
9.8 Conclusions
In this Chapter a method to compute end-coil leakage inductances in multi-phase electric
machines has been investigated, extending an approach, based on Neumann integral
solution, previously proposed for three-phase machines. Two dedicated algorithms have
been presented to compute the end-coil self inductance, which cannot be determined
through Neumann integrals directly. Compared to 3D FE approaches, the method does not
imply complicated geometric models to be prepared and requires small computational
resources. It can be easily implemented in a software program performing a fast end-
End-winding leakage inductance computation 119
Fig. 9-21. Comparison between test and computation results for the three machines M1, M2, M3.
leakage inductance calculation for any phase number and arrangement, needing coil and
machine main dimensions are the only input data. The method has been experimentally
120 End-winding leakage inductance computation
tested on three machines with different numbers of phases and poles, finding a
satisfactory agreement between measurements and calculation results.
9.9 References
[1] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul. 2007,
pp. 489-516.
[2] D. Hadiouche, H. Razik, A. Rezzoug, “On the modeling and design of dual-stator
windings to minimize circulating harmonic currents for VSI-fed AC machines”, IEEE
Transactions on Industry Applications, vol. 40, Mar./Apr. 2004, pp. 506-515.
[3] E.A. Klingshirn, “High phase order induction motors−Part I−Description and
theoretical considerations”, IEEE Transactions on Power Apparatus and Systems, Jan.
1983, vol. PAS-102, pp. 47-53.
[4] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction machine
using vector space decomposition”, IEEE Transactions on Industry Application, Sept.-
Oct. 1995, vol. 31, pp. 1100-1109.
[5] D. Ban, D. Zarko, I. Mandic, “Turbogenerator end-winding leakage inductance
calculation using a 3-D analytical approach based on the solution of Newmann
integrals”, IEEE Transactions on Energy Conversion, vol. 20, Mar. 2005, pp. 90-105.
[6] G.A. Campbell, “Mutual inductances of circuits composed of straight wires”, Phys. Rev.,
1915, vol. 5, pp. 452-458.
[7] E.B. Rosa, “The self and mutual inductances of Linear Conductors”, Bulletin of the
Bureau of Standards, vol. 4, No. 2, 1908, pp. 301-344.
[8] M.A. Bueno, A.K.T. Assis, “A New Method for Inductance Calculations”, Journal of
Physics D: Applied Physics, 1995, vol. 28, pp. 1802-1806.
[9] Gover, F.W. Inductance Calculations, Working Formulas and Tables, Dover Publications,
2004, pp. 22, 35.
[10] R. Bansal, Handbook of Engineering Electromagnetics, CRC Press, 2004, pp. 154-155.
[11] A.T. Brahimi, A. Foggia, G. Meunier, “End winding reactance computation using a 3D
finite element program”, IEEE Transactions on Magnetics, Mar. 1993, vol. 29, pp. 1411-
1414.
[12] Y. Kawase, Y. Hayashi, T. Yamaguchi, “3-D finite element analysis of motors excited
from voltage source taking into account end-coil effects”, IEEE Transactions on
Magnetics, vol. 33, Mar. 1997, pp. 1686-1689.
121
10.1 Introduction
In three-phase machines, stator leakage flux phenomena are exhaustively accounted for
through a single parameter in the dq0 equivalent circuit [1], [2], [3]. As the number of
stator phases increases, however, the number of stator leakage inductances to be
considered grows accordingly due to the mutual magnetic couplings among phases [4].
Recent and past studies have shown how stator leakage inductances may strongly impact
on multiphase machine performance, especially under inverter supply. For example,
circulation current harmonics which appear in multi-phase machines (like split-phase
motors) when supplied by voltage-source inverters have an amplitude which mainly
depends on stator leakage inductances [5], [6], [7]. Also in multi-phase synchronous
motors, supplied by Load-Commutated Inverters, these parameters play a remarkable role
in determining the current dynamics during normal and abnormal commutation
transients as discussed in [8].
The aforementioned approximations are partly due to the objective difficulty of accurately
computing the parameters in issue analytically, i.e. without resorting to complete
geometric models of the machine, to be processed though detailed and time-consuming
Finite Element (FE) analyses [13]. Moreover, test procedures for measuring stator leakage
inductances in synchronous machines are quite critical, even in case of ordinary three-
phase windings [10].
This Chapter proposes a set of analytical techniques (partly original and partly derived
from past cited works) that can be employed for a fast but sufficiently accurate
computation of stator leakage inductances of multi-phase machines without using FE
techniques.
The proposed methods apply to any number of stator phases and hold for both
symmetrical and asymmetrical winding configurations [1]. The assumptions are made that
leakage inductances are not affected by magnetic saturation (which is reasonable except in
fault and overload conditions [12]). Furthermore, the n-phase winding is supposed of
double-layer imbricated-coil type with an integer number of slots/pole/phase and with n
phase belts per pole.
Fig. 10-1. Conventional phase arrangement scheme for an n-phase machine. (a) Position of phase axes, with
angles in electrical radians; (b) double-layer short-pitch phase-belt arrangement.
More precisely: experimental results are used to validate slot and end-coil leakage
inductances by means of measurements on two high-power synchronous machines (a 12-
phase round-rotor motor and a 6-phase salient 8-pole generator), following the testing
guidelines provided in [3], [8], [11]; FE simulations are used to assess air-gap leakage
inductances and for assessing the dependency of slot leakage inductances on the coil pitch.
In this Chapter we try to make abstraction from the particular multi-phase scheme in
establishing the algorithms for self and mutual inductance determination. For this
purpose, we propose to map a generic n-phase winding (with n phase belts per pole) into
an equivalent scheme with sequentially distributed phases arranged as per Fig. 10-1,
similarly to what already done in [11], [14].
For illustration purpose, the mapping procedure is exemplified in Fig. 10-2 and Fig. 10-3
respectively in the case of a dual-star and a symmetrical 5-phase windings. The former is
mapped into a 6-phase and the latter into a 5-phase scheme, both having phase belts
arranged sequentially as per Fig. 10-1. The phases of the original winding are mapped
univocally into those of the equivalent scheme as per Table I and II.
TABLE I
MAPPING TABLE FOR DUAL-STAR WINDING
Original phase U,1 U,2 V,1 V,2 W,1 W,2
Corresponding phase 1 2 5 6 −3 −4
TABLE II
MAPPING TABLE FOR FIVE-PHASE WINDING
Original phase 1 2 3 4 5
Corresponding phase 1 3 5 −2 −4
Once the winding is mapped into its equivalent scheme arranged as per Fig. 10-1, the
problem reduces to determining n leakage inductance values λ0, λ1, ..., λn−1: the generic λk
is the self inductance if k=0, while for k=1, ..., n−1 it indicates the mutual inductance
between two phases displaced by k phase belts (i.e. by kπ/n electrical radians) apart.
124 Computation of leakage inductances and experimental validations
Each λk can be expressed as the sum of three components λsk, λek, λak which respectively
represent the slot, end-coil and air-gap leakage components.
In the following Sections, the focus is on separately determining analytical expression for
such these leakage components λsk, λek, λak.
In three-phase machines, self and mutual slot leakage inductances can be analytically
computed by means of suitable slot permeance coefficients combined with factors to
account for the coil pitch [15]. The method has been extended to five-phase [16] and six-
phase windings [6], [9]. In the following, a easy-to-handle analytical formula is derived to
extend the computation for any number n of phases.
2pq
λks = {(Lt + Lb )δ k ,0 + (1 + δ k ,0 )⋅ M tb ⋅ [R(1 − k − n(1 − r ) )− R(1 − k − nr )]} (10.1)
b2
with the following set of definitions:
Lt: self-inductance (due to slot leakage flux) of a coil side lying in the top layer
(air-gap slot side);
Lb: self-inductance (due to slot leakage flux) of a coil side lying on the bottom
layer (opposite the air-gap);
Mtb: mutual inductance (to to slot leakage flux) of two coil sides lying in the same
slot;
δi,j: Kronecker symbol, such that δi,j=1 if i=j and δi,j=0 otherwise;
R(x) ramp function, such that R(x)=0 if x≤0 and R(x)=x otherwise.
The derivation of (10-1) is reported in Appendix A. As concerns parameters Lt, Lb and Mtb,
they can be difficult to find analytically for a generic slot shape. Nevertheless, for
rectangular slot cross sections (typical of high-power machines wound with flat turns) a
simple expression can be given for them as follows:
h s: slot height;
The approach proposed for the analytical air-gap leakage inductance computation is
based on the winding function theory extended to machines with possibly non-uniform
air-gap [19]. According to this theory, the total mutual inductance (including both leakage
and useful fluxes) between the i-th and j-th phases due to air-gap flux can be computed as:
2π
∫
M i , j = R L µ 0 P( x − x r )Wi ( x )W j ( x )dx
0
(10-2)
where:
xr: rotor position;
Analytical expressions for P(x) and Wi(x) are proposed in the following subsections.
equal to the reciprocal of the air-gap width. In general, P(x) should account for possible
rotor saliencies and stator slot openings as well. For this purpose P(x) is written as:
Psal ( x )Pslot ( x )
P( x ) = (10-3)
g
where
g minimum air-gap width;
The factor Psal(x) can be estimated based on the approximation that the air-gap width, in
salient pole machines, results from superimposing a constant and a sine wave with a
period equal to a pole pitch (i.e. π in electrical radians), that is:
Lmq − Lmd
D= (10-5)
Lmd
Concerning slotting effects, quite accurate results have been found using Weber
approximation for the air-gap field dips caused by slot openings [20]. This approximation
leads to write:
2α
Z mod(x , 2π / Z )
Pslot ( x ) = 1 − 2β cos (10-6)
2
where Z is the number of slots per pole pairs and non-dimensional coefficients α, β depend
on slot opening ws, tooth width wt and air-gap width θ as follows:
Fig. 10-4. Phase self inductance of a salient-pole machine as a function of the rotor position neglecting higher-
order harmonics.
127
α = wt / w s , β =
(1 − u)2 , u=
ws w
+ 1 + s . (10-7)
(
2 1 + u2 ) 2θ 2θ
If also the rotor surface were slotted, an additional factor should be included in (10-3) to
account for that.
i − 1
Wi ( x) = ∑ Ah cosh x −
h =1,3,5, 7 ,...
π
n
(10-8)
Nn 1 πh πhr
Ah = 8q sin sin (10-9)
b h2 2n 2
where:
N: number of series-connected turns per coil;
Fig. 10-5. FE-analysis of a salient-pole machine (dual-star 8-pole) for the determination of its air-gap field profile.
128 Computation of leakage inductances and experimental validations
Fig. 10-6. Comparison of air-gap profiles obtained analytically and from FE simulations (Fig. 5).
i −1 i − 1
Wi ( x ) = A1 cos x −
π + ∑ Ah cosh x −
n h=3,5 ,7 ,...
π
n
(10-11)
2π 2α
i −1 j −1 Z mod(x − xr , 2π / Z )
λai − j = −2Rβ L µ0 ∫ A12 cos x − π cos x − π [1 + D sin(2x − 2xr )]cos dx
0 n n 2
2π
i − 1 i − 1
+ R L µ0 ∫ ∑ Ah cos h x −
0 h=3 ,5 ,7 ,...
π Am cos m x −
n
π ×
n
m=3,5,7 ,...
Z mod(x − x r , 2π / Z )
2α
Fig. 10-7. Models of (a) quadruple-star round rotor machine; (b) dual-star salient-pole machines. Phase names
are indicated for a winding layer and over a pole span.
In the above expansion two terms are isolated: the first considers only the contribution of
MMF fundamental without slotting effects, the second (equal to the sought air-gap leakage
inductance λak ) collects all other contributions.
For the FE assessment of slot leakage inductances, auxiliary points (Pb, Pt) are included in
the model at each slot opening (Fig. 10-8). Then, for any stator coil with sides Cb and Ct
130 Computation of leakage inductances and experimental validations
Fig. 10-8. Sides (Cb, Ct) of a stator coil and auxiliary points (Pb, Pt) introduced for leakage inductance
computation..
(Fig. 10-8), its flux linkage due to slot leakage field is computed from FE analysis as
follows:
1 1
N L
C
∫∫ A ( x , y )dxdy − C ∫∫ A ( x , y )dxdy − [A (P ) − A (P )]}
z z z b z t (10-14)
Cb Ct
where: C coil cross-section area, N number of turns per coil, L useful core length, Az vector
potential (z component) resulting from FE analysis. In fact, the first term in square
brackets is proportional to the total coil flux linkage, while the second term is proportional
Fig. 10-9. Slot leakage phase inductances (self and mutual) as functions of the coil pitch for the 12-phase machine
(a, b) and for the 6-phase machine (c). Solid line diagrams are from analytical computation [eq. (1)], circled
points are from FE analysis.
131
Fig. 10-10. Air-gap leakage inductances λak computed analytically and from FEA: (a) for the 6-phase machine; (b)
for the 12-phase machine.
to the total coil flux linkage minus the portion due to slot leakage field.
The comparison between analytical and FE simulation results is shown in Fig. 10-9 (each
circle corresponds to a FE simulation), showing a very good matching.
The results of the comparison, reported in Fig. 10-10, show a good accordance between
the numerical and analytical evaluations both in the case of the salient-pole machine and
in the case of the round-rotor one. It is however to be considered that, in a salient-pole
machine, the air-gap leakage inductance is slightly dependent on the rotor position as can
be seen from (10-13), so diagrams of Fig. 10-10a refer to a particular rotor position.
132 Computation of leakage inductances and experimental validations
Fig. 10-11. Search coil placed in the stator bore in a test with rotor removed.
Fig. 10-12. Slot (λsk), end-coil (λek) and air-gap (λak) leakage inductances for the 6-phase machine, from the
proposed analytical method (top graph) and from measurements and calibrated FE analysis (bottom graph).
133
described in [8], [11] is followed: basically, while a stator phase is supplied with an AC
current I, the open-circuit induced voltage is measured on the other ones. This enables to
determine phase self and mutual inductances (λrrk) with the rotor removed. As discussed
in [3], each inductance λrrk with the rotor removed can be written as:
The values obtained with the described procedure are compared in Fig. 10-12 with those
obtained with the proposed analytical computations [air-gap leakage components (from
Fig. 10-10) are also added to visualize the entity of the various leakage flux contributions].
Fig. 10-12 shows a satisfactory agreement.
10.8 Conclusions
In this Chapter the analytical calculation of stator leakage inductances in multi-phase
machine with n phases (either symmetrically or asymmetrically distributed) is addressed.
Fig. 10-13. Slot (λsk), end-coil (λek) and air-gap (λak) leakage inductances for the 12-phase machine, from the
proposed analytical method (top graph) and from measurements and calibrated FE analysis (bottom graph).
134 Computation of leakage inductances and experimental validations
Analytical expressions are proposed to compute slot and air-gap leakage inductances
based on machine design data, while for the end-coil contribution reference is made to
previous works. The formulas for slot and air-gap leakage inductances are assessed by
means of FE analysis taking into account different numbers of phases and coil pitch values.
Measurements on a 6-phase (dual star) and 12-phase (quadruple star) machines with the
rotor removed are also presented and compared with analytical results. All the validations
presented (both through measurements and FE simulations) show a satisfactory matching
with analytical predictions.
Appendix A
In this Appendix the derivation of (10-1) is illustrated.
With the mapping procedure discussed in Section 10.2, any double-layer n-phase winding
with coil to pole pitch r can be represented as sketched in Fig. 10-14a. For the sake of
simplicity, Fig. 10-14a illustrates a 2-pole machine (the extension to a generic number of
pole pairs is straightforward).
Given two generic and distinct phases i and j, such that k=|i−j|, their respective phase belts
may or may not overlap depending on the coil pitch (changing the coil pitch is equivalent
to “sliding” one winding layer keeping the other fixed). Let us call Z'i,j the overlapping
fraction of phase belts with equal sign belonging to phases i, j and Z''i,j the overlapping
fraction of phase belts with opposite signs belonging to phases i, j. For example, it is
evident from Fig. 10-14a that if r=1 (full pitch winding) there is no overlapping between
the phase belts of the two phases, so Z'i,j=Z''i,j=0. Conversely, a complete overlapping of the
phase belts with equal sign (i.e. “+i”, “+j” and “−i”, “−j”; Z'i,j=1) occurs for coil pitch values r
Fig. 10-14. (a) Conventional n-phase winding scheme; (b) Z'i,j and Z''i,j diagrams.
135
such that:
π (1 − r ) = (π / n ) i − j (A1)
while a complete overlapping between phase belts of opposite signs (i.e. “+i”, “−j” and “−i”,
“+j”; Z''i,j=1) occurs for those coil pitch ratios r such that:
π (1 − r ) = π − (π / n ) i − j (A2)
Based on similar geometric considerations, the diagrams reported in Fig. 10-13b are
obtained for the functions Z'i,j and Z''i,j. Because Z'i,j and Z''i,j do not depend on i, j
individually but only on k=|i−j|, it is convenient to write them as Z'k and Z''k. Analytically,
these functions can be written in the form:
What proved above enables us to say that the two phases i and j have 2pqZ'k shared slots
where they have equal directions and 2pqZ''k shared slots where they have opposite
directions. Calling Mtb the mutual inductance between two coil sides lying in the same slot,
the mutual slot leakage inductance will then be:
2 pq (Z k′ − Z k′′)M tb (A5)
which is to divided by b2 in the case when there are b parallel ways per phase. Considering
(A3) and (A4), the expression (10-1) is then proved for the case i≠j i.e. k≠0 (mutual
inductance).
For the self inductance case (k=0), the contribution is to be added of the self slot leakage
inductance of each coil side. Calling Lt and Lb such self inductances respectively for a coil
side placed in the top or bottom layer, this contribution will be:
2 pq(Lt + Lb ) (A6)
since each phase has 2pq coil sides lying in the top layer and 2pq lying in the bottom layer.
Furthermore we have to consider possible slots occupied by coil sides of the same phase:
each of such slots gives a contribution of 2Mtb. This ends the proof of (10-1) also for the
self leakage inductance case (k=0).
10.9 References
[1] E. Levi, “Multiphase electric machines for variable-speed applications”, IEEE Trans. on
Industrial Electronics, vol. 55, May 2008, pp. 1893-1909.
[2] E. Levi, R. Bojoi, F. Profumo, H.A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Application, IET, 2007, July 2007,
vol. 1, pp. 489-516.
[3] IEEE Std. 115-1995.IEEE Guide: Test Procedures for Synchronous Machines, p.159.
[4] J. Figueroa, J. Cros, P. Viarouge, « Generalized Transformations for Polyphase Phase-
Modulation Motors”, IEEE Trans. on Energy Conversion, vol. 21, no. 2, June 2006, pp.
332-341.
136 Computation of leakage inductances and experimental validations
[5] Y. Zhao, T. Lipo, “Space Vector PWM Control of Dual Three-Phase Induction Machine
Using Vector Space Decomposition”, IEEE Trans. on Industry Applications, vol. 31, no.
5, Sept.-Oct. 1995.
[6] D. Hadiouche, H. Razik, A. Rezzoug, “On the modeling and design of dual-stator
windings to minimize circulating harmonic currents for VSI-fed AC machines”, IEEE
Trans. on Industry Applications, vol. 40, Mar./Apr. 2004, pp. 506-515.
[7] A. Tessarolo, C. Bassi, “Stator harmonic currents in VSI-fed synchronous motors with
multiple three-phase armature windings”, IEEE Transactions on Energy Conversion,
vol. 25, no. 4, Dec. 2010, pp. 974-982.
[8] A. Tessarolo, S. Castellan, R. Menis, G. Ferrari, “On the Modeling of Commutation
Transients in Split-Phase Synchronous Motors Supplied by Multiple Load-
Commutated Inverters”, IEEE Transactions on Industrial Electronics, vol. 57, issue 1,
Jan. 2010, pp. 35-43.
[9] T. A. Lipo, “A d-q Model for Six-Phase Induction Machines”, Proceedings of the
International Conference on Elecric Machines, Athens, pp. 860-867, Athens, Sept.
1980.
[10] IEC 34-4 Std., 1995, Rotating Electrical Machines, Part 4: Methods for Determining
Synchronous Machine Quantities from Tests.
[11] A. Tessarolo, F. Luise, “An analytical-numeric method for stator end-coil leakage
inductance computation in multi-phase electric machines”, IEEE Industry Application
Society Annual Meeting, IAS 2008, 5-9 Oct. 2008, Edmonton, Canada, CD-rom paper n.
79.
[12] J. C. Flores, G. W. Buckley, G. McPherson, “The Effects of Saturation on the Armature
Leakage Reactance of Large Synchronous Machines”, IEEE Trans. on Power Apparatus
and Systems, vol. PAS-103, no. 3, March 1984.
[13] K. Shima,K. Ide, M. Takahashi, “Analysis of Leakage Flux Distributions in a Salient-Pole
Synchronous Machine Using Finite Elements”, IEEE Trans. on Energy Conversion, vol.
18, no. 1, March 2003.
[14] S. Williamson, S. Smith, “Pulsating torque and losses in multiphase induction
machines”, IEEE Transactions on Industry Applications, vol. 39, July/Aug. 2003, pp.
986-993.
[15] Liwschitz-Garik, M.M., Whipple, C.C., Alternating Current Machines, D. Van Nostrand
Company, Princeton, NJ, 1961.
[16] L.A. Pereira, C. C. Scharlau, L.F.A. Pereira, J.F. Haffner, “General model of a five-phase
induction machine allowing for harmonics in the air-gap”, IEEE Trans. on Energy
Conversion, vol. 21, issue 4, Dec. 2006, pp. 891-899.
[17] D. Ban, D. Zarko, I. Mandic, “Tubogenerator end-winding leakage inductance
calculation using a 3-D analytical approach based on the solution of Neumann
integrals”, IEEE Trans. on Energy Conversion, vol. 20, Mar. 2005, pp. 98-105.
[18] A. Tessarolo, “On the modeling of poly-phase electric machines through vector-space
decomposition: numeric application cases”, International Conference on Power
Engineering, Energy and Electrical Drives, POWERENG 2009, pp. 524-528.
[19] J. Faiz, I.Tabatabaei, “Extension of winding function theory for nonuniform air-gap in
electric machinery”, IEEE Trans. on Magnetics, vol. 36, no. 6, Nov. 2002, pp. 3654-
3657.
[20] I. A. Viorel, K. Hamemeyer, L. Strete, “On the Carter’s factor calculation for slotted
electric machines”, Advances in Electrical and Computer Engineering, vol. 7, no. 2,
2007.
137
The modeling approach developed in PART II suffers from some limitations, the main of
which is that it applies only to multiple-star machines. A more general approach to the
modeling of multiphase machinery is offered by the so called Vector-Space Decomposition
(VSD) technique. This Part is dedicated to show how VSD is a suitable methodology for
treating all kinds of multiphase machines, whether with symmetrical or asymmetrical (split-
phase) windings. The attention, in any case, is to idealized machines, where a uniform air-
gap and a sinusoidal wininding distribution are assumed. In the next Part, the extension of
the VSD method to real machines (with stator and rotor saliencies and with non-ideal
winding distribution) will be covered, instead.
The application of VSD is not new: it was used by Lipo in the last 1990’s to dual-star
induction machines and by Gataric to symmetrical polyphase machines in 2000. Even before
(during the 1950’s), the basis of VSD methodology had been laid by applying Fortescue’s
symmetrical component transform to polyphase machines and observing how it could turn
the inductance matrix into a diagonal form.
The original contribution brought by the Chapters of this part, therefore, is not the method in
itself but rather its generalization. What will be essentially proved is that VSD can be applied
to all multiphase topologies, but the transformations which implement it may take two
distinct forms. The criterion for choosing either of such forms is the phase progression which
characterizes the multiphase winding. A distinction is in fact made between full phase
progression and half phase progression systems and two different general formulations of
VSD transforms are proposed to treat them.
More precisely, Chapter 11 covers the subject at a theoretical level, while Chapter 12 presents
some numerical application examples which serve both illustrative and validation purposes.
138 VSD modeling of idealized multiphase machines: theory
11.1 Introduction
In the electromagnetic conception of modern electric machines, the possibility to select
the number of phases without restricting the choice to the traditional three-phase scheme
provides the designer with an additional degree of freedom. This might be used for
various purposes, such as [1]: increasing the motor power rating without changing the
power electronic device size in the supplying inverter; matching higher reliability
requirements since drive operation continuity may be guaranteed also in presence of a
faulty inverter phase; increasing motor performance, power density and efficiency [6], [2];
using innovative multi-motor drive configurations [3].
As discussed in Section 11.2, a variety of possibility exists when it comes to select the
physical phase distribution within the stator winding. Regardless of the particular
arrangement selected, the Vector Space Decomposition (VSD) is an effective technique
currently used to model electric machines equipped with poly-phase windings [4], [3].
Nevertheless, we can observe that in the case when a symmetrical poly-phase
configuration is chosen, with a phase progression of 360/n electrical degrees, the
implementation of the VSD relies on consolidated and well defined methods, based on
Fortescue transform [6], recently enhanced by adding a rotational Park’s transform [7]
and including some space harmonic effects [7], [8]. Conversely, this Chapter shows that
when it comes to model other kinds of poly-phase windings, like split-phase ones [1], this
theory is not always applicable and other alternative strategies, hereinafter discussed in
detail, need to be adopted so as to achieve the VSD successfully.
Fig. 11-1. Examples of symmetrical n-phase winding schemes (n=5) with full phase progression, equal to 360/n
electrical degrees.
VSD modeling of idealized multiphase machines: theory 139
Fig. 11-2. Example of n-phase winding schemes (n=4) with half phase progression, equal to 180/n electrical
degrees (a). Example of physical phase arrangements in a double-layer shortened-pitch winding implementation
(b).
2n-phase winding (where phase belts displaced by 180 electrical degrees are series-
connected and assigned to the same phase).
It is important to notice that the latter kind of winding arrangement is not always feasible.
For example, we can consider the case of a double-layer semi-eight-phase winding (Fig.
11-2), composed of n=4 phases, each comprising both goes and returns in each winding
layer (Fig. 11-2b). It is intuitive that such a winding cannot be arranged in such a way that
the phase progression equals 360/n electrical degrees (Fig. 11-1b). A feasible
arrangement, instead, is that shown in Fig. 11-2, where the phase progression is of 180/n
electrical degrees (“half phase progression”).
The same situation as in the mentioned example occurs when considering a split-phase
winding [10] composed of N three phase sets with N equal to an even number. For
example, setting N=2 (Fig. 11-3a), we have the well known case of a dual three-phase
winding, also denoted as semi-12-phase [23] or quasi-6-phase [2]. Regardless of how the
n=6 stator phases are rearranged or redefined, there is no possibility to reduce such
winding scheme to a symmetrical 6-phase one with a 360/n phase progression. Instead, it
is possible (and also convenient for some purposes) to map such a winding into a 6-phase
one having the structure shown in Fig. 11-3b, i.e. characterized by a 180/n phase
progression. For instance, such an approach has been usefully adopted in [6] to study
multi-phase machine losses and torque ripple making abstraction of the specific phase
arrangement adopted in the stator winding design.
Hypothesis 1 is reasonable when the machine operation under study is not subjected to
significant flux changes. Assumption 2 is usually made regardless of the kind of machine
under investigation [7]. Conversely, assumption 3 is well matched only in case of round-
140 VSD modeling of idealized multiphase machines: theory
Fig. 11-3. A dual three-phase winding (a) and its correspondent six-phase scheme with half phase progression (30
electrical degrees).
rotor machines [12], otherwise it may constitute a significant restriction. The latter
hypothesis, though, will be removed in Part V, where the effect of all space harmonics, due
to possible rotor magnetic anisotropy and non-sinusoidal winding distribution, will be
described in detail from the VSD viewpoint.
In this case the stator inductance matrix L, under the hypotheses of Section 11.3, exhibits a
symmetrical “circulant” matrix [7], with the following n×n structure:
L0 L1 L2 L L2 L1
L1 L0 L1 L L3 L2
L L1 L0 L L4 L3
L( f ) = 2 (11-1)
M M M O M M
L2 L3 L4 L L0 L1
L L2 L3 L L1 L0
1
where subscript f means that a “full” phase progression angle is used in the stator poly-
phase winding.
The generic element Lk of (11-1) represents the mutual inductance between two stator
phases physically displaced by kπ/n electrical radians, degenerating into the phase self
inductance, L0, when k=0.
It is a well known fact that any n×n circulant matrix (regardless of whether it is
symmetrical or not) can be diagonalized as per (11-2) by means of the complex Fortescue
transformation F [6], [7] defined by (11-3), with “i” denoting the imaginary unit.
λ1 0 L 0
0 λ2 L 0
F L ( f ) F −1 = (11-2)
M M O M
0 0 L λn
VSD modeling of idealized multiphase machines: theory 141
The transformation matrix (11-2) suffers from the disadvantages of being complex and
non-orthonormal, i.e. F F t ≠ I n , where superscript t denotes the transposition operator
and In the n×n identity matrix.
C(1)
S( 1)
C(2)
S(2)
Fr = M (11-4)
C(ν )
S(ν )
C(0)/ 2
C(n / 2)/ 2
Vectors C(j), S(j) are n-sized row-vectors defined as per (11-6)-(11-7) and the constant ν
results from truncating (n−1)/2, i.e.:
For both even and odd n, the real matrix Fr is easily proved to be orthonormal and capable
of transforming the stator inductance matrix (11-1) into the diagonal structure (11-8),
where the last row is to be omitted in case of odd n.
d1 0 0 0 L 0 0 0 0
0 d1 0 0 L 0 0 0 0
0 0 d2 0 L 0 0 0 0
0 0 0 d2 L 0 0 0 0
t
Fr L ( f ) Fr = M M M M O M M M M (11-8)
0 0 0 0 L dν 0 0 0
0 0 0 0 L 0 dν 0 0
0 0 0 0 L 0 0 d0 0
0 0 0 0 L 0 0 0 d n / 2
Obviously, the eigenvalues that appear in (11-8) are the same as in (11-2), simply
arranged in another order. Their physical significance is well known [7], as each diagonal
142 VSD modeling of idealized multiphase machines: theory
elements dj is the “harmonic inductance” (or the imaginary part of the “harmonic
impedance”) of the stator poly-phase system with respect to the j-th order harmonic [23].
The key point on which this Chapter intends to draw the attention is that the poly-phase
windings under consideration cannot be modeled, in terms of VSD, with the same
approach as explained in the previous Section for full phase-progression systems.
In fact, under the hypotheses of Section 11.3, the stator inductance matrix L(h) (subscript h
denoting “half” phase progression) has the structure shown below, where Lk is the mutual
inductance between two stator phases displaced by kπ/n electrical radians.
L L1 L2 L − L2 − L1
L0 L1 L Ln−2 Ln−1 0
L L0 L1 L − L3 − L2
L1 L0 L Ln−3 Ln−2 1
L L1 L0 L − L4 − L3
L ( h) = M M L M M = 2 (11-9)
M M M O M M
Ln−2 Ln−3 L L0 L1
− L2 − L3 − L4 L L0 L1
Ln−1 Ln−2 L L1 L0
− L1 − L2 − L3 L L1 L0
For the sake of clarity, (11-9) is next expanded in the particular cases of n=5 and n=6,
when it takes the forms (11-10) and (11-11) respectively [8], [12]. For numerical
examples and validations, the next Chapter [12] should be referred to.
L0 L1 L2 − L2 − L1
L1 L0 L1 L2 − L2
L ( h) = L2 L1 L0 L1 L2 (11-10)
− L2 L2 L1 L0 L1
− L1 − L2 L2 L1 L0
L0 L1 L2 0 − L2 − L1
L1 L0 L1 L2 0 − L2
L L1 L0 L1 L2 0
L ( h) = 2 (11-11)
0 L2 L1 L0 L1 L2
− L2 0 L2 L1 L0 L1
−L − L2 0 L2 L1 L0
1
The form of L(h) can be easily explained by inspection of Fig. 11-2a or Fig. 11-3b. For
example, under the hypotheses of Section 11.3, it is clear that, for n=6 (Fig. 11-3b), the
mutual inductance between phases 1 and 2 (i.e. L1) is equal in amplitude and opposite in
sign with respect to the mutual inductance between phases 1 and 6 (i.e. L5), so that L1=−L5;
the same pertains to the phase pairs 1-3 and 1-5, implying L2=−L4. In this way the first row
VSD modeling of idealized multiphase machines: theory 143
of (11-11) is justified. With similar reasoning, the rest of the matrix can be easily found
out.
C(1)
S(1)
C(3)
S(3)
Gr = M (11-12)
C(2ν − 1)
S(2ν − 1)
C(2ν + 1)/ 2
S(2ν + 1)/ 2
In (11-12), ν is again defined by (11-5) and the row-vectors C, S are given by (11-13)-( 11-
14), with m ranging from 0 to ν and k ranging form 1 to n.
It can be easily seen that Gr has a very similar structure to Fr, defined by (11-4), the main
difference being the fact that only odd values of j are chosen for row-vectors C(j), S(j) in
(11-12). The application of Gr to the stator inductance matrix (11-9) always leads to a
diagonal form as required for VSD purposes. This is shown in (11-15) where, as usual, the
last row is to be omitted in case of odd n.
144 VSD modeling of idealized multiphase machines: theory
d1 0 0 0 L 0 0 0 0
0 d1 0 0 L 0 0 0 0
0 0 d3 0 L 0 0 0 0
0 0 0 d3 L 0 0 0 0
t
G r L ( h) G r = M M M M O M M M M (11-15)
0 0 0 0 L d 2ν −1 0 0 0
0 0 0 0 L 0 d 2ν −1 0 0
0 0 0 0 L 0 0 d 2ν +1 0
0 0 0 0 L 0 0 0 d 2ν +1
As recalled in the previous Section, each diagonal elements dj represents the harmonic
inductance of the n-phase system with respect to the j-th order harmonic. It is interesting
to notice that only odd order harmonic inductance appear in (11-15) as a consequence of
how the decoupling matrix Gr has been built: some physical interpretation of this fact can
be found in [23], [12].
The diagonal values dj can be determined through a simple formula from the elements of
the inductance matrix (11-9), i.e.:
n
d2m+1 = ∑L
r =1
r −1 cos[(π / n)(2m + 1)(r − 1)] ∀m = 0..ν (11-16)
where we recall that Lk is the mutual inductance between two stator phases displaced by
kπ/n electrical radians.
11.6 Conclusions
In this Chapter, poly-phase machine modeling through the Vector-Space Decomposition
(VSD) technique has been investigated. It has been pointed out how the well-known VSD
method based on Fortescue transform, usually applied to symmetrical n-phase machines
with a phase progression of 360/n electrical degrees, needs to be properly modified in
order for it to work when a half phase progression (of 180/n electrical degrees) is
adopted, as in the case of split-phase machines. A transformation that serves this purpose
has been explicitly presented and its validity demonstrated analytically. In the next
Chapter, the theory herein presented is applied to some numeric case studies and
validated through finite-element analysis.
Appendix
In this Appendix a formal proof is provided for (11-15)-(11-16), assuming that the
transformation matrix Gr is built as per (11-12)-(11-14). For this purpose, it convenient to
define the mutual inductance µ between two stator phases whose magnetic axes are
VSD modeling of idealized multiphase machines: theory 145
located respectively at φ1 and φ2 with respect to a common reference frame. In virtue of the
hypotheses listed in Section 11.3, we can certainly state that µ depends only on the
relative displacement ∆φ =φ2− φ1. Moreover, the reciprocity theorem for linear systems
[13] assures that
µ( Δφ ) = µ( − Δφ ) (11-17)
Finally, provided that the winding structure has a half phase progression, with a winding
scheme like those depicted in Fig. 11-2a and Fig 11-3b, if straightforward to prove that
µ( Δφ ) = − µ( Δφ + π ) (11-18)
since increasing the displacement angle ∆φ between the two phases by 180 electrical
degrees is the same as reversing the current direction in one of them.
Properties (11-17) and (11-18) assure that µ(∆φ) can be expanded in Fourier series with
only odd cosine terms, i.e. in the form:
µ( Δφ ) = ∑ Mk
k =1 ,3,5,7 ,...
cos(k ⋅ Δφ ) (11-19)
Using (11-19) the mutual inductance Lj between two stator phases displaced by jπ/n
electrical radians apart becomes:
L j = µ( jπ / n) = ∑ Mk
k =1 ,3,5 ,7 ,...
cos(k jπ / n) (11-20)
L( h)C( H )t = d H C( H )t (11-21)
∑L
j =1
i− j 2/ n cos[H (π / n)( j − 1)] = d H 2/ n cos[H (π / n)(i − 1)] (11-22)
Substituting (11-20) into (11-16) we can write dH as per (11-23) and parameter Li−j as per
(11-24).
n n
dH = ∑
j =1
L j −1 cos[(π / n)H ( j − 1)] = ∑ ∑M
j =1 k =1 ,3 ,5 ,...
k cos[k (π / n)( j − 1)]cos[H (π / n)( j − 1)] (11-23)
1 ,3 ,... n
1 ,3 ,... n
(11-25)
= ∑ M ∑ cos[k (π / n)( j − 1)]cos[H (π / n)( j − 1)]cos[H (π / n)(i − 1)]
k
k
j =1
At this point, by easy algebraic manipulation and based on the fact that both H and k are
odd integers, one can prove that:
n n
∑cos[k (π / n)(i − j )]cos[H (π /n)( j − 1)] = ∑cos[k (π / n)( j − 1)]cos[H (π / n)( j − 1)]cos[H (π / n)(i − 1)]
j =1 j =1
which finally shows that equality (11-21) holds for any odd integer H.
With identical procedure, one can demonstrate that S(H) is an eigenvector of L(h) for any
odd H and the associated eigenvalue is dH. The demonstration is omitted for the sake of
brevity. As a consequence, we have proved that the matrix Gr, defined according to (11-
12), is the sought transformation suitable for diagonalizing L(h) as per (11-15) for any
number of phases n and regardless of the particular values taken by the elements of L(h).
11.7 References
[1] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul.
2007, pp. 489-516.
[1] S. Williamson, S. Smith, “Pulsating torque and losses in multiphase induction
machines”, IEEE Transactions on Industry Applications, vol. 39, July/Aug. 2003, pp.
986-993.
[2] D.G. Dorrell, C.Y. Leong, R.A., McMahon R.A., “Analysis of performance assessment of
six-pulse inverter-fed three-phase and six-phase induction machines”, IEEE
Transactions on Industry Applications, Nov./Dic. 2006, vol. 6, pp. 1487-1495.
[3] E. Levi, M. Jones, S.N. Vukosavic, H.A. Toliyat, “A novel concept of a multi-phase,
multimotor vector controlled drive system supplied from a single voltage source
inverter”, IEEE Transactions on Power Electronics, March 2004, vol. 19, pp. 320-334.
[4] S. Gataric, “A Polyphase Cartesian Vector Approach to Control of Polyphase AC
Machines”, IEEE IAS Annual Meeting, 2000.
[5] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction machine
using vector space decomposition”, IEEE Transactions on Industry Application, Sept.-
Oct. 1995, vol. 31, pp. 1100-1109.
[6] D.C. White, H.H. Woodson, Electromechanical Energy Conversion, John Wiley and
Sons, New York, 1959.
[7] J. Figueroa, J. Cros, P. Viarouge, “Generalized Transformations for Polyphase Phase-
Modulated Motors”, IEEE Transactions On Energy Conversion, vol. 21, June 2006, pp.
332-341.
[8] L.A. Pereira, C.C. Scharlau, L.F.A. Pereira, J.F. Haffner, “General model of a five-phase
induction machine allowing for harmonics in the air-gap”, IEEE Transactions on
Energy Conversion, Dec. 2006, vol. 21, pp. 891-899.
VSD modeling of idealized multiphase machines: theory 147
[9] E.A. Klingshirn, “High phase order induction motors−Part I−Description and
theoretical considerations”, IEEE Transactions on Power Apparatus and Systems, Jan.
1983, vol. PAS-102, pp. 47-53.
[10] E.A. Klingshirn, “High phase order induction motors−Part I−Experimental results”,
IEEE Transactions on Power Apparatus and Systems, Jan. 1983, vol. PAS-102, pp. 54-
59.
[11] R.H. Nelson, P.C. Krause, “Induction machine analysis for arbitrary displacement
between multiple winding sets”, IEEE Transactions on Power Apparatus and Systems,
May./June 1974 vol. PAS-94, pp. 841-848.
[12] A. Tessarolo, “On the modeling of poly-phase electric machines trhough vector-space
decomposition: numeric application cases”, submitted paper for POWERENG 2009.
[13] Y. Li, Z.Q. Zhu, D. Howe, C.M. Bingham, “Modeling of cross-coupling magnetic
saturation in signal-injection-based sensorless control of permanent magnet
brushless motor”, IEEE Transactions on Magnetics, June 2007, vol. 43, pp. 2552-
2554.
148 VSD modeling of idealized multiphase machines: application examples
12.1 Introduction
In the design of poly-phase electric machines, the possibility to select the number of
phases, instead of restricting the choice to the conventional three-phase solution,
introduces a valuable degree-of-freedom, that can be exploited in different ways, like
performance, reliability and power rating enhancement. The number n of stator phases,
though, is not the only variable of which the electromagnetic designer may take advantage.
Some further freedom can result from the way in which the n phases are physically
arranged to build up the winding. Apart from such possible variants as fractional slot
designs, differences may originate, for example, depending on whether opposite (i.e.
displaced by 180 electrical degrees) phase belts are series-connected or not [23], [26];
furthermore, if n is multiple of three, two or more three-phase groups may be formed,
each suitable for being supplied by a different three-phase inverter independently, as
happens in the so called split-phase configurations [10]; examples can also be found in the
literature, along with practical applications, of split-phase solutions where a fifteen-phase
winding is split into three five-phase sections [4]; finally, successive phases may be
displaced by a different electrical angle (called “phase progression” [6]), the usual choice
being between 360/n (“full” phase progression) and 180/n (“half” phase progression).
In order to restrict the variety of poly-phase solutions to be studied and classified, some
authors have pointed out that phase-split arrangements can be easily reduced to other
phase arrangements where the n stator phases are sequentially distributed, [1]. This is the
reason why the focus is hereinafter put on the modeling of such reduced winding schemes,
which can be classified according to whether they use a full or half phase progression.
Fig. 12-1. Variants of a five-phase winding: (A) properly called five-phase scheme with full phase progression; (B)
semi-ten-phase scheme with full phase progression; (C) semi-ten-phase scheme with half phase progression..
VSD modeling of idealized multiphase machines: application examples 149
Fig. 12-2. Physical phase arrangement in shortened-pitch, double-layer implementations of winding schemes A, B,
C of Fig. 1.
In the former case, Fortescue’s theory on symmetrical poly-phase systems [6], [7], may be
directly employed to achieve the Vector-Space Decomposition of the machine model [3],
i.e., in algebraic terms, to diagonalize its stator inductance matrix. When a half phase
progression is adopted, instead, the use of the same transformation matrices does not lead
to the desired result, even in the case when space harmonics are neglected. This fact has
been theoretically proven and discusses in the previous Chapter [4], where a modified
decoupling matrix has been proposed for the VSD of half-phase-progression systems. This
approach will be hereinafter illustrated numerically through the application to a five-
phase machine, with possible alternative stator winding designs. The importance of
extending the VSD method to machines with half-phase progression will be also
150 VSD modeling of idealized multiphase machines: application examples
Three possible stator phase arrangements have been taken into account, as schematically
shown by the phasor diagrams of Fig. 12-1. In the scheme A, a full phase progression,
equal to 360/5 electrical degrees, is employed, so that each phase is composed of one
single phase belt spanning 72 electrical degrees. In the scheme B, the number of phase
belts is doubled up to 10 and opposite phase belts are series-connected, as usually done in
ordinary three-phase windings; the obtained configurations is referred to in the literature
as “semi-10-phase” in [23], [26], or quasi “quasi-5-phase”, [2]; in scheme B, the phase
progression is equal to 360/5 electrical degrees as well. Finally, scheme C originates from
simply rearranging the phases so that a half phase progression (equal to 180/5 electrical
degrees) is employed.
The physical implementation of the three winding solutions is depicted in Fig. 12-2, where
the same color is used to indicate any given phase and the sign “−” refers to the
conventional current direction in a specific coil side.
L0 L1 L2 L3 L4
L1 L0 L1 L2 L3
L = L2 L1 L0 L1 L2 (12-1)
L3 L2 L1 L0 L1
L4 L3 L2 L1 L0
By analyzing the motor cross-section models, as depicted in Fig. 12-2, through a Finite
Element (FE) Analysis tool, the stator inductance matrix L can be easily determined
numerically. In the cases of winding schemes A, B, C, the computed inductances L0, …, L4
are reported in Table I. The corresponding matrices LA, LB, LC are reported in (12-2)-( 12-
4).
VSD modeling of idealized multiphase machines: application examples 151
TABLE I
SELF AND MUTUAL INDUCTANCE VALUES RESULTING FROM FEA
Winding Inductances (mH)
scheme L0 L1 L2 L3 L4
A 52.79 13.14 −39.41 −39.40 13.13
B 58.97 14.41 −42.87 −42.91 14.40
C 58.96 42.87 14.40 −14.40 −42.90
As expected [7], in the case of full phase progression, L turns out to be not only a
symmetrical Toeplitz matrix, but also circulant [13]. This does not occur, instead, for the
scheme C, where a half phase progression is used.
When the stator inductance matrix is circulant, like LA and LB, a real-valued transformation
serving the purpose is that recalled in [1]. For the five-phase machine under consideration,
it takes the form (12-5), through which the diagonalization of both (12-2) and (12-3) can
be attained as per (12-6)-( 12-7).
152 VSD modeling of idealized multiphase machines: application examples
(
cos 25π 1 ⋅ 0
) ( )
cos 25π 1 ⋅ 1 ( )
cos 25π 1 ⋅ 2 ( )
cos 25π 1 ⋅ 3 (
cos 25π 1 ⋅ 4
)
(
sin 25π 1 ⋅ 0) ( )
sin 25π 1 ⋅ 1 ( )
sin 25π 1 ⋅ 2 ( )
sin 25π 1 ⋅ 3 (
sin 25π 1 ⋅ 4 )
F = 2n cos 25π 2 ⋅ 0
( ) cos( 2 ⋅ 1)
π 2
5
cos( 2 ⋅ 2)
π 2
5
cos( 2 ⋅ 3)
π2
5
(
cos 25π 2 ⋅ 4 ) (12-5)
(
sin 2π 2 ⋅ 0 ) sin( π 2 ⋅ 1)
2
sin( π 2 ⋅ 2)
2
sin( π 2 ⋅ 3)
2
(
sin 25π 2 ⋅ 4 )
5 5 5 5
1 / 2 1/ 2 1/ 2 1/ 2 1 / 2
d 1A 0 0 0 0
0 d 1A 0 0 0
FL F = 0
A t
0 d 2A 0 0
(12-6)
0 0 0 d 2A 0
0 0 0 0 d 0A
d 1B 0 0 0 0
0 d 1B 0 0 0
FL F = 0
B t
0 d 2B 0 0
(12-7)
0 0 0 d 2B 0
0 0 0 0 d 0B
n
dk = ∑L
j =1
j −1 [ π k ( j − 1)]
cos 2
5
∀k ∈ {0,1,2} (12-8)
As concerns matrix LC, owing to its non-circulant form, it cannot be diagonalized through
(12-5), even if the factor 2π/5 is replaced by π/5 in order to take the half phase
progression into account. This circumstance can be easily checked numerically. As
proposed in [4], the decoupling matrix (12-9) can be adopted, instead. This assures
diagonalization as per (12-10).
(
cos π5 1 ⋅ 0
) ( )
cos π5 1 ⋅ 1 ( )
cos π5 1 ⋅ 2 ( )
cos π5 1 ⋅ 3 ( )
cos π5 1 ⋅ 4
(
sin π5 1 ⋅ 0 ) sin(π 1 ⋅ 1)
5
sin(π 1 ⋅ 2)
5
sin(π 1 ⋅ 3)
5
( )
sin π5 1 ⋅ 4
(
G = 2n cos π5 3 ⋅ 0 ) cos(π 3 ⋅ 1)
5
cos(π 3 ⋅ 2)
5
cos(π 3 ⋅ 3)
5
( )
cos π5 3 ⋅ 4 (12-9)
(
sin π5 3 ⋅ 0 ) sin(π 3 ⋅ 1) sin(π 3 ⋅ 2) sin(π 3 ⋅ 3) ( )
sin π5 3 ⋅ 4
5 5 5
π
(
cos 5 5 ⋅ 0 ) cos(π 5 ⋅ 1)
5
cos(π 5 ⋅ 2)
5
cos(π 5 ⋅ 3)
5
( )
cos π5 5 ⋅ 4
d1C 0 0 0 0
0 d1C 0 0 0
GLC G t = 0 0 d 3C 0 0
(12-10)
0 0 0 d3C 0
0 0 0 0 d5C
The diagonal elements can be also computed through the compact form proposed below:
VSD modeling of idealized multiphase machines: application examples 153
n
dk = ∑L
j =1
j −1 [ ]
cos π5 k ( j − 1) ∀k ∈ {1,3,5} (12-11)
The latter fact can be easily checked by means of FE harmonic simulations (Section 12.5)
in which a k-sequence symmetrical system of currents, characterized by phasors (12-12),
is applied to the stator winding, where I indicates the current amplitude and “i” represents
the imaginary unit. The FE analysis yields the flux linkage phasors as per (12-13), where Φ
denotes the common flux linkage module.
Hence, the sought machine inductance with respect to the k-th sequence component is
computed as:
d k = Φ/Ι (12-14)
Fig. 12-3 shows some examples of FE analysis outputs that result from supplying the
stator phases with current systems of different sequences. It is important to notice that, in
the case of winding schemes A and B, the a homopolar current component gives rise to a
homopolar flux linkage, as per (12-12), (12-13) with k=0; conversely, if the winding
scheme C is adopted, a homopolar current component gives rise to a magnetic flux
distribution which highly resembles that produced by the fundamental (Fig. 12-3a). This is
a physical interpretation for the fact that zero-sequence, as well as even-sequence
components, cannot be used in the decoupling matrix G, in accordance with (12-9).
The slight slight discrepancies found in some values can be accounted for based on two
reasons: the intrinsic approximations introduced by the FE methods; the fact that space
harmonics, including slotting effects, are not taken into account in the proposed VSD
modeling approach.
154 VSD modeling of idealized multiphase machines: application examples
Fig. 12-3. Example of magnetic flux lines, computed through FE analysis, which result from applying a
symmetrical phase current system of sequence 1 (a), 3 (b), 5 (c) to the winding scheme C of Fig. 1. Magnetic flux
lines due to a zero-sequence current system in the winding scheme A of Fig. 1 (d).
TABLE II
STATOR MATRIX EIGENVALUES AND SEQUENCE INDUCTANCES
Fig. 12-4. Variants of a five-phase winding: (A) properly called five-phase scheme with full phase progression; (B)
semi-ten-phase scheme with full phase progression; (C) semi-ten-phase scheme with half phase progression..
Actually, if only the five-phase case were to be considered, it might be objected that the
winding scheme C (Fig. 12-1) does not need to be studied, since it can be easily reduced to
scheme A, with full phase progression, based on the obvious identity:
(i B
1 i 2B i 3B i 4B i 5B )
t
(
= W i 1C i 2C i 3C i 4C i 5C )
t
(12-15)
1 0 0 0 0
0 0 0 − 1 0
W = 0 − 1 0 0 0 (12-16)
0 0 0 0 1
0 0 1 0 0
where superscript t denotes transposition and symbols ikB , ikC indicate the k-th phase
current in winding schemes B and C respectively.
Equation (12-15) implies the following relationship between stator inductance matrices:
L B = W LC W t (12-17)
as one can easily check numerically by substitution of (12-3), (12-4), (12-16) into (12-17).
In reality, the reduction of an n-phase system with half phase progression (like that of Fig.
12-1C) to another one having full phase progression (like that of Fig. 1B) is feasible only
when n is odd.
As an example, one can consider the 12-phase schemes depicted in Fig. 12-4: the split-
phase scheme (Fig. 12-4a) cannot be reduced to any equivalent 12-phase arrangement
having full phase progression. Conversely, it is easily possible to map it into the equivalent
twelve-phase scheme with half phase progression shown in Fig. 12-4b (as suggested in [6],
156 VSD modeling of idealized multiphase machines: application examples
for example). The latter can be in turn diagonalized by means of the transformation
matrices reported in [4].
In more general terms, the study of poly-phase winding schemes is important for the study
of those split-phase configurations composed of an even number N of three-phase sets.
Such configurations, in fact, can be usefully mapped into equivalent 3N-phase schemes [6],
similar to those of Fig. 12-1C and Fig. 12-3b, characterized by a half phase progression.
12.7 Conclusions
In this Chapter, the method of Vector-Space Decomposition has been applied to
diagonalize the stator inductance matrix of a five-phase induction machine, taking into
account various possible phase arrangements in its winding design. It has been shown that
the winding arrangements where a full phase progression (equal to 360/5 electrical
degrees) is adopted can be effectively treated through a well known approach, derived
from Fortescue’s theory of symmetrical poly-phase systems. It has been also shown how
the same methodology fails when a half phase progression is used, instead. To obtain the
stator inductance matrix diagonalization in this case too, an alternative transformation has
been applied and numerically validated based on Finite Element calculations. Finally, the
practical importance of studying poly-phase systems with half phase progression has been
empathized as a useful way to approach the VSD of split-phase schemes composed of an
even number of three-phase sets.
12.8 References
[1] E.A. Klingshirn, “High phase order induction motors−Part I−Description and
theoretical considerations”, IEEE Transactions on Power Apparatus and Systems, Jan.
1983, vol. PAS-102, pp. 47-53.
[2] E.A. Klingshirn, “High phase order induction motors−Part I−Experimental results”,
IEEE Transactions on Power Apparatus and Systems, Jan. 1983, vol. PAS-102, pp. 54-
59.
[3] R.H. Nelson, P.C. Krause, “Induction machine analysis for arbitrary displacement
between multiple winding sets”, IEEE Transactions on Power Apparatus and Systems,
May./June 1974 vol. PAS-94, pp. 841-848.
[4] F. Terrein, S. Siala, P. Noy, “Multiphase induction motor sensorless control for
electric ship propulsion”, IEE Power Electronics, Machines and Drives Conference,
PEMD 2004, pp. 556-561.
[5] S. Williamson, S. Smith, “Pulsating torque and losses in multiphase induction
machines”, IEEE Transactions on Industry Applications, vol. 39, July/Aug. 2003, pp.
986-993.
[6] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul.
2007, pp. 489-516.
[7] D.C. White, H.H. Woodson, Electromechanical Energy Conversion, John Wiley and
Sons, New York, 1959.
[8] J. Figueroa, J. Cros, P. Viarouge, “Generalized Transformations for Polyphase Phase-
Modulated Motors”, IEEE Transactions On Energy Conversion, vol. 21, June 2006, pp.
332-341.
[9] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction machine
using vector space decomposition”, IEEE Transactions on Industry Application, Sept.-
Oct. 1995, vol. 31, pp. 1100-1109.
[10] A. Tessarolo, “On the modeling of poly-phase electric machines through Vector-
Space Decomposition: theoretical considerations”, International Conference on
VSD modeling of idealized multiphase machines: application examples 157
Power Engineering, Energy and Electrical Drives, POWERENG 2009, 18-20 March
2009, Lisbon, Portugal, 18-20 March 2009, pp. 519-523.
[11] R.H. Nelson, P.C. Krause, “Induction machine analysis for arbitrary displacement
between multiple winding sets”, IEEE Transactions on Power Apparatus and Systems,
May./June 1974 vol. PAS-94, pp. 841-848.
[12] D.G. Dorrell, C.Y. Leong, R.A., R.A. McMahon, “Analysis of performance assessment of
six-pulse inverter-fed three-phase and six-phase induction machines”, IEEE
Transactions on Industry Applications, Nov./Dic. 2006, vol. 6, pp. 1487-1495.
[13] F.R. Gantmacher, The Theory of Matrices, Chelsea Publishing Co., NY 1960.
[14] D. Hadiouche, H. Razik, A. Rezzoug, “On the modeling and design of dual-stator
windings to minimize circulating harmonic currents for VSI-fed AC machines”, IEEE
Transactions on Industry Applications, vol. 40, Mar./Apr. 2004, pp. 506-515.
[15] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction machine
using vector space decomposition”, IEEE Transactions on Industry Application, Sept.-
Oct. 1995, vol. 31, pp. 1100-1109.
158
In Part IV, Vector-Space Decomposition (VSD) has been investigated as a method for
modeling multiphase machines with whatever winding topology (whether symmetrical or
asymmetrical) under the assumption of uniform air-gap and sinusoidal winding distribution.
Under such hypothesis, it has been also demonstrated that a suitable decoupling
transformation matrix can be always defined (an explicit expression has been given for it)
such that the transformed inductance matrix takes a diagonal form. However, it is well
known from both literature and experience that in multiphase machine operation (especially
in case of voltage-source supply) space harmonics play an important role (examples of their
effects will be extensively investigated in Chapter 16). The typical way to capture space
harmonic effects through simulations is to use time-stepping Finite Element analysis, where
the detailed machine geometry is intrinsically considered. The disadvantage of such
approach is terms of computational resources required, however, are known. The attempt is
thereby worthwhile to include space harmonic effects in the lumped-parameter modeling of
the machine through VSD. This would enable to run extremely fast but sufficiently accurate
simulations of those phenomena in which space harmonics are involved.
Based on the aforementioned premises, this Part is aimed at extending the VSD method to the
case of a non-idealized machine. In other words, the problem is faced as to if and how space
harmonics (due to non-uniform air-gap and non-sinusoidal winding distribution) can be
included in the multiphase machine model through VSD. It will be shown that the inclusion
of space harmonics causes, in general, some out-of-diagonal elements to appear in the
transformed inductance matrix of the multiphase machine. Actually, this fact had already
been reported in the literature in qualitative terms. The original contributions brought
hereinafter are listed next:
1) A practical method is defined for numerically identifying the permeance and winding
functions that characterize the real machine topology through a minimal set of
Finite Element magnetostatic simulations; this is equivalent to quantifying air-gap
flux space harmonics.
159
2) A new general VSD approach is formalized which applies to all kinds of multiphase
winidng topologies, which are all reduced to a single “conventional” phase scheme
through explicitly-formulated geometrical transformations.
Most of the subject matter of this Part is included in Chapter 14 where the points from 2)
through 4) listed above are covered. Chapter 13 is mainly propedeutical in the sense that it
lays the basis for the theory developed in Chapter 14 by covering point 1).
At the end of Chapter 14 two experimental validation examples of the modeling technique
proposed in this Part are presented referring to a six-phase and nine-phase machine with
strongly non-uniform air-gap and consequent large air-gap flux space harmonics. Comparing
measurements and simulations for a permanent short-circuit test, it will be proved that the
inclusion of space harmonics in the machine lumped-parameter model leads to very accurate
simulation results. The level of accuracy achieved in this way through a Simulink simulation
taking few seconds is practically the same as (or even better than) that obtained from a
time-stepping Finite-Element simulation lasting more than two hours on the same hardware
platform.
160 Inductance expression through winding function theory
The winding function theory has been used for a long time in multiphase machine theory
as a way to analytically express self and mutual phase inductances for a generic rotor
position in terms of the so-called phase winding function and air-gap permeance function
[1][2]. The practical usefulness of the theory for the numerical computation of machine
inductances is limited by two problems: no rigorous permeance function definition exists
for machine topologies with strongly non-uniform air-gap; the computation formula
requires a definite integral to be solved. In this Chapter, such issues are addressed. Firstly,
a method for numerically identifying the permeance function of salient-pole machines
through a minimal set of magnetostatic finite-element analyses is proposed. Secondly, an
inductance expression through winding and permeance functions, suitable for fast and
simple numerical implementation, is derived. The proposed techniques are assessed
against direct mutual inductance computation by FE analysis and measurements on a real
salient-pole machine prototype.
The inductance expression proposed in this Chapter will be also used in the next Chapter
to include space harmonic effects in multiphase machine modeling through Vector Space
Decomposition.
Fig. 13-1. Example of a two-pole 5-phase electric machine where each phase has coil groups shifted by
π electrical radians.
Inductance expression through winding function theory 161
Fig. 13-2. Example of a two-pole 10-phase electric machine where each phase is not composed of coil groups
shifted by π electrical radians. (a) Physical winding topology; (b) conventional winding representation.
Fig. 13-3. Examples of different 9-phase winding arrangements: (a) symmetrical scheme; (b) split-phase scheme;
(c) conventional scheme. All schemes can be reduced to a conventional one by a geometrical variable
transformation W.
162 Inductance expression through winding function theory
Fig. 13-4. Transformation of a 5-phase (a) and a dual three-phase (c) winding schemes into conventional
arrangements (b, d).
Using the aforementioned conventions, we can notice that in an n-phase machine phases
can be differently arranged over the stator circumference. For example, a 9-phase winding
may have either the symmetrical structure shown in Fig. 13-3a or the asymmetrical (split-
phase) structure shown in Fig. 13-3b. In order to approach n-phase windings in a unified
manner, however, it is convenient to make abstraction of whether phases are
symmetrically or asymmetrically distributed. This is made possible by simply renaming
the stator phases over a pole span in a sequential manner, for example using indices from
0 to n−1, as exemplified by Fig. 13-3c for the case of a 9-phase winding. Fig. 13-4 illustrates
the same procedure for the case of a 5-phase and of a dual three-phase winding
arrangement. It is easily understood that the procedure can be applied to any n-phase
winding scheme respecting conditions 4 and 5 of 13.1.1 (either symmetrical or
asymmetrical) through a suitable geometrical variable transformation W (explicit
expressions for W will be given in the next Chapter).
Fig. 13-5. Polar system of coordinates (ρ, ξ) and air-gap mean circumference Γ.
δ0
ρ = R = Rinn − (13.1)
2
where Rinn is the stator bore inner radius and δ0 is the minimum air-gap width. Unless
differently specified, air-gap points lying on circumference Γ (ρ=R) will be taken into
account and thereby identified by the sole angular coordinate ξ. In the polar coordinate
system defined above, the rotor position is indicated by the angular coordinate x of its
polar axis d (Fig. 13-5).
The winding function theory offers a more synthetic alternative based on an a single
analytical expression of the mutual inductance between two whatever phases i, j as a
function of the rotor position. This consists of a definite integral involving the so-called
phase winding and air-gap permeance function: the former describes phase geometrical
position and distribution, the latter describes air-gap shape accounting for slotting and
saliency effects.
In this Section, the winding function theory as a method for computing the air-gap
magnetic field and inductances of an electric machine will be reviewed and particularized
to the conventional n-phase winding topology described in 13.1.2. The expression of the
164 Inductance expression through winding function theory
mutual inductance that will be derived as a result will be used in the following Sections for
permeance function numerical identification.
π
α= . (13.2)
n
Let us suppose to energize a generic stator phase i with a unity current with all other
machine circuits open and the rotor at stand-still in position x. According to the winding
function theory, the radial component of the resulting magnetic field and flux density at
position ξ along the air-gap circumference Γ can be expressed as follows:
H (ξ , x ) = ps (ξ )pr (ξ − x )w i (ξ ) (13.3)
B (ξ , x ) = µ0 H i (ξ , x ) (13.4)
where wi(ξ) is the winding function of phase i, ps(ξ) is the stator permeance function and
pr(ξ) is the rotor permeance function. The two permeance functions account for the
slotting and/or saliency effects due to stator and rotor surfaces respectively.
In the assumed conventional phase arrangement (13.1.2), the winding function of the
generic phase i is equal to the winding function of phase “0” shifted by iα electrical
radians. In symbols:
w i (ξ ) = w 0 (ξ − iα ) (13.5)
In case of distributed winding with q slots per pole per phase, coil pitch equal to γ
electrical radians, N series-connected turns per coil and b parallel paths per phase, the
winiding function w0(x) can be expressed analytically through the following Fourier
expansion:
Inductance expression through winding function theory 165
Fig. 13-7. Examples of winding function profiles obtained from Fourier series expansion.
(a) Distributed winding; (b) concentrated winding.
r γ α qr
sin sin s
4N 2 2 cos(r ξ )
w0 (ξ ) = ∑
π b r =1 ,3,5,7... α r . (13.6)
r sin s
2
π α
αs = = (13.7)
qn q.
In case of concentrated-coil winding (Fig. 13-6) with N series connected turns per coil,
tooth angular width τt and coil side angular width τc, the winding function expanded in
Fourier series is:
τ +τ τ
sin r c t sin r c
8N 2 2 cos(r ξ ) (13.8)
w0 (ξ ) = ∑
π b r =1 ,3,5,7... π r 2τ c
Examples of winding function profiles obtained from (13.6) and (13.8) with r between 1
and 151 are given in Fig. 13-7.
2π 2π
∫ ∫
mi(,agj ) (x ) = R L B (ξ , x )w j (ξ )dξ = R Lµ0 ps (ξ )pr (ξ − x )w i (ξ )w j (ξ )dξ
0 0
2π
(13.9)
∫
= R Lµ0 ps (ξ )pr (ξ − x )w0 (ξ − iα )w0 (ξ − jα )dξ
0
At this point, we notice that the stator permeance function ps(ξ) (which accounts for stator
surface slotting) is a periodic function, the period of which coincides with the slot pitch αs
given by (13.7); the phase progression α, which equals qαs according to (13.7), is also a
period for ps(ξ) then. So we can write:
ps (ξ ) = ps (ξ − iα ) (13.10)
∫
mi(,agj ) (x ) = R L µ 0 pr (ξ − x )w 0 (ξ − jα )p s (ξ − iα )w 0 (ξ − iα )dξ
0
2π
(13.11)
∫
= R L µ0 pr (ξ − x ) f (ξ − iα )w0 (ξ − jα )dξ
0
f (ξ ) = p s (ξ )w 0 (ξ ) (13.12)
For illustration purposes, the case of the 36-slot salient-pole machine shown in Fig. 13-8a
is considered. Its stator winding is supposed to be arranged according to either of the
Inductance expression through winding function theory 167
Fig. 13-8. (a) Cross section of the actual machine considered and (b-d) of the auxiliary models needed for
permeance function identification.
schemes illustrated in Fig. 13-9, i.e. in either a 6-phase either 9-phase arrangement. The
former is characterized by q=2 slots per pole per phase, the latter by q=3 slots per pole per
phase. In both cases the coil pitch ratio is 16/18.
The proposed FE identification technique for the permeance function makes use of the
three auxiliary machine models (indicated as Model I, II and III) shown in Fig. 13-8b-d and
described next.
Model II is obtained from the actual machine model replacing its stator with a hollow
cylinder of inner radius Rinn, while the rotor is placed at position x=0; in Model II the MMF
source is constituted by an auxiliary winding which, in principle, could have any
168 Inductance expression through winding function theory
Fig. 13-9. Stator winding multiphase arrangements of the 36-slot example machine: (a) 6-phase; (b) 9-phase.
Model III combines model I stator and model II rotor and is therefore characterized by
smooth stator and rotor air-gap surfaces and by a uniform air-gap width δ0. The MMF
source in model III is the same as that chosen for model II.
1
H I (ξ ) = p s (ξ ) w 0 (ξ ) (13.13)
δ0
In fact, the rotor permeance function is identically equal to δ0 for Model I. From (13.13),
the stator permeance function can be identified as follows (Fig. 13-13):
Fig. 13-10. (a) Auxiliary turn sides (A, B) used to energize model I.
Inductance expression through winding function theory 169
Fig. 13-11. Model I FE analysis results: (a) 6-phase winding; (b) 9-phase winding.
δ0
p s (ξ ) = H I (ξ ) (13.14)
w 0 (ξ )
f (ξ ) = p s (ξ )w 0 (ξ ) = H I (ξ )δ 0 . (13.15)
Fig. 13-12 and Fig. 13-14 show the diagrams of f (ξ ) = H I (ξ )δ 0 found from the FE analysis
on Model I respectively assuming a 6-phase and a 9-phase winding arrangements. The
Fig. 13-12. Winding function of phase “0” and air-gap field of Model I. Case of 6-phase winding.
170 Inductance expression through winding function theory
Fig. 13-14. Winding function of phase “0” and air-gap field of Model I. Case of 9-phase winding.
Fig. 13-13. Stator permeance functions. (a) 6-phase winding; (b) 9-phase winding.
diagrams are compared to the winding function w0 (ξ ) profiles obtained from (13.6). It can
be seen that H I (ξ )δ 0 practically coincide with w0 (ξ ) apart from the slotting effect.
Fig. 13-13 shows the stator permeance function obtained from (13.14) for the two
assumed winding configurations. Since stator permeance function depends only the air-
gap profile and not on the winding structure, it is confirmed (as expected) that the same
p s (ξ ) profiles are obtained regardless of the winding arrangement.
1
H III (ξ ) = w aux (ξ )I aux (13.17)
δ0
Inductance expression through winding function theory 171
Fig. 13-15. FE analysis results of (a) model II; (b) model III.
where w aux (ξ ) indicates the winding function of the auxiliary circuit. The magnetic fields
H II (ξ ) and H III (ξ ) obtained from FE analysis are given in Fig. 13-16.
From (13.16) and (13.17) the following expression is obtained for the rotor permeance
function:
H II (ξ )
pr (ξ ) = (13.18)
H III (ξ )δ 0
Looking at H II (ξ ) and H III (ξ ) diagrams in Fig. 13-11, one can see that a numerical
problem in evaluating (13.18) occurs for those points (A, B) where H III (ξ ) is equal to zero
(the points where this happens coincide with the positions of the auxiliary turn sides).
Fig. 13-16. FE analysis results for: (a) Model II and (b) Model III.
172 Inductance expression through winding function theory
Fig. 13-17. FE analysis results for: (a) Model II and (b) Model III.
Anyway, numerical issues can be avoided by observing that, thanks to its being a π-
periodic even function, pr (ξ ) is fully identified once only its values on the interval ∆=[0,
π/2] are known (Fig. 13-16); in fact, for any ξ between π/2 and π we can write:
p r (ξ ) = p r (π − ξ ) (13.19)
Hence, to avoid convergence issues, it is required that H III (ξ ) be non-zero only in the
interval [0, π/2], as happens in Fig. 13-16. This imposes restrictions on selecting the
auxiliary circuit used to energize Models II and III and justifies the choice of punctual
conductors A and B, placed at ξ = −π/4 and ξ = −3π/4, for this purpose.
H II (ξ ) π
H (ξ )δ if 0 ≤ mod(ξ , π ) ≤
2
pr (ξ ) = III 0
(13.20)
H II (π − ξ ) π
if < mod(ξ , π ) < π
H III (π − ξ )δ 0 2
This definition assures that the permeance function pr (ξ ) can be defined for any real ξ
without numerical convergence problems. The obtained diagram is shown in Fig. 13-17.
13.3.3.1 Sensitivity to the auxiliary winding selections for models II and III
The question arises at this point as to whether the rotor permeance function
determination discussed above may depend on the choice of the auxiliary winding used to
energize models II and III. To answer this question, the determination will be next
performed with the same methodology but using a completely different auxiliary winding.
Fig. 13-18. (a) Linear current sheet on stator bore of Models II and III; (b) its approximation for FE analysis.
Inductance expression through winding function theory 173
Fig. 13-19. FE analysis results of (a) model II; (b) model III. Models energized with sinusoidal current sheet
distributions.
The latter is constituted by a sinusoidal linear current distribution λ(ξ) spread on the
smooth stator bore surfaces of Models II and III (Fig. 13-18a):
λ (ξ ) = Λ0 cos(ξ ) (13.21)
where amplitude Λ0 is an arbitrary constant, which can be set equal to 1 A/mm, for
instance.
The FE analysis of models II and III energized with the linear current distribution (13.21)
gives the results shown in Fig. 13-19 and the air-gap magnetic field profiles shown in Fig.
13-20.
By means of (13.20), the rotor permeance function shown in Fig. 13-21 is obtained. The
same figure also compares the diagrams of pr(ξ) obtained with the two auxiliary windings
used to energize models II and III (concentrated conductors A, B and sinusoidal current
sheet distribution), showing that the result is practically independent of the auxiliary
winding selection.
Fig. 13-20. Air-gap magnetic field computed by FE analysis on models II and III.
174 Inductance expression through winding function theory
Fig. 13-21. (a) Rotor permeance function diagrams obtained with different auxiliary circuits to energize Model II
and III. (b) Rotor air-gap function diagram.
The rotor air-gap function gr(ξ) represents the magnetic air-gap width at position ξ. This
coincides with the physical air-gap in case round-rotor machines, but may differ from it in
case of salient-pole machines. The discrepancies between magnetic and physical air-gap
are larger and larger as the air-gap non-uniformity increases. An effective way to visualize
the difference between magnetic and physical air-gaps is to plot the function Rinn−gr(ξ)
(where Rinn is the inner stator bore radius) in polar coordinates (Fig. 13-22a) and
superimpose the resulting diagram to the actual rotor profile (Fig. 13-22b). The
superimposition (Fig. 13-22c) shows that the magnetic air-gap width precisely follows the
physical air-gap in the pole shoe region, while in the inter-pole region (where the air-gap
remarkably increases), the two profiles differ to a significant extent.
Fig. 13-22. (a) Polar plot of the function Rinn−gr(ξ); (b) real rotor; (c) superimposition of (a) and (b).
Inductance expression through winding function theory 175
(k )
pr = pr (k ∆ξ ) , f (k ) = f (k ∆ξ ) (13.24)
with k indicating the sample number and ∆ξ the discretization step of the interval
[0, 2π]. For numerically evaluating the integral in (13.11), an integration step
different from ∆ξ is to be used, in general, which implies the need for some
interpolation among samples.
This Section is dedicated to proposing a solution to the above issues. The starting point is
to express functions pr (ξ ) and f (ξ ) through Fourier series expansions, i.e.:
f (ξ ) = ∑ Fh
h=1 ,3 ,5 ,...
cos(hx ) (13.25)
pr (ξ ) = ∑ Ph cos
h=2 , 4 ,6 ,...
(hx ) (13.26)
(k )
Fourier series coefficients Fr and Ph are computed using function samples pr and f (k )
(obtained by FE analysis) as follows:
1
Fh =
π ∑f
k
(k )
cos(hk ∆ξ )∆ξ (h = 1, 3, 5, 7, …) (13.27)
1
P0 =
2π
∑pk
r
(k )
∆ξ (13.28)
1
Ph =
π ∑p
k
r
(k )
cos(hk ∆ξ )∆ξ (h = 2, 4, 6, 8, …) (13.29)
where the integer index k (sample number) ranges in the interval [0, 2π/∆ξ].
2π
mi(,agj) (x ) = R L µ0 ∫ ∑Fs cos[s(ξ − i α )] ∑ Ph cos[h(ξ − x )] ∑Wr cos[r (ξ − j α )]dξ (13.30)
0 s =1,3,5,...
h=2,4,6 ,... r =1,3,5,...
176 Inductance expression through winding function theory
2π
= R L µ0 ∫ ∑ ∑ ∑F
0 r =1 ,3 ,5 ,... s =1 ,3 ,5 ,... h=2 , 4 ,6 ,...
s Ph Wr cos[s (ξ − i α )]cos[h(ξ − x )]cos[r (ξ − j α )]dξ
Fourier coefficient Fk, Pk computed as per (13.27)-(13.29), along with Fourier coefficients
Wk given by (13.31), are shown in Fig. 13-23 for the example machine considered, in its 6-
phase and 9-phase winding arrangements.
r γ α qr
sin sin s
Wr =
4N 2 2
(13.31)
π b α r
r sin s
2
At this point, the integral in (13.32) can be symbolically solved. In doing this, one can
Fig. 13-23. Space harmonic spectra of functions w0(ξ) and f(ξ) for (a) the 6-phase winding; (b) the 9-phase
winding arrangement. (c) Rotor permeance function harmonic spectrum.
Inductance expression through winding function theory 177
observe that, out of all the terms in the right-hand side member of (13.32), only those
where h = s + r or h = s − r are different from zero, i.e.:
2π
h= s −r ∧ s ≠ r
2π
π (13.35)
⇒ ∫ cos[s(ξ − i α )]cos[r (ξ − j α )]cos[h(ξ − x )]dξ = 2 cos[(s i − r j )α − (s − r )x ]
0
2π
h = s −r ∧ s = r ⇒ ∫ cos[s(ξ − i α )]cos[r (ξ − j α )]cos[h(ξ − x )]dξ = π cos[s (i − j )α ]
0
(13.36)
π
mi(,agj ) (x ) = R L µ0 ∑
s ,r =1 ,3,5,... 2
Fs Wr Ps + r cos[(s i + r j )α − (s + r )x ]
(13.37)
π
+ ∑ 2
Fs Wr P s −r cos[(s i − r j )α − (s − r )x ] + ∑ π Fs Ws P0 cos[s (i − j )α ]
s ,r =1 ,3 ,5 ,... s =r =1 ,3 ,5 ,...
s ≠r
For further simplification, one can observe that the following equality holds:
π
∑
s ,r =1 ,3,5 ,... 2
Fs Wr P s −r cos[(s i − r j )α − (s − r )x ]
s ≠r
(13.38)
π π
= ∑
s ,r =1 ,3 ,5 ,... 2
Fs Wr P s −r cos[(s i − r j )α − (s − r )x ] − ∑
s =1 ,3 ,5 ,... 2
Fs Ws P0 cos[s (i − j )α ]
π
mi(,agj ) (x ) = R L µ 0 ∑
s ,r =1 ,3,5,... 2
Fs Wr Ps +r cos[(s i + r j )α − (s + r )x ]
(13.39)
π π
+ ∑
s ,r =1 ,3 ,5 ,... 2
Fs Wr P s −r cos[(s i − r j )α − (s − r )x ] + ∑
s =1 ,3 ,5 ,... 2
Fs Ws P0 cos[s (i − j )α ]
π R L µ0
mi(,agj ) (x ) =
2
∑ {F W P
s ,r =1 ,3,5,...
s r s +r cos[(s i + r j )α − (s + r )x ]
(13.40)
+ Fs Wr P s −r cos[(s i − r j )α − (s − r )x ] + } π R L µ0
2 ∑ Fs Ws P0 cos[s (i − j )α ]
s =1 ,3 ,5 ,...
1) Phase i is energized with unity current, while the rotor is placed at position x and
all the other phases are at no load.
3) Inductance m(i ,agj ) (x ) is computed from the solved model as the flux linkage of phase
j due to air-gap flux only. Such flux linkage is computed as follows:
Z −1
mi(,agj ) (x ) = N ∑ A(P )[χ
k =0
k
( top )
k,j + χ k( bot
,j
)
] (13.41)
where: N is the number of series-connected turns per coil; Z is the number of slots;
Pk is the intersection between the air-gap circumference Γ and the symmetry axis
of the kth stator slot (Fig. 13-24); A(Pk) is the vector potential at point Pk and
Fig. 13-24. Air-gap points for the FE computation of phase air-gap inductances.
Inductance expression through winding function theory 179
coefficients χ k(top
,i
)
and χ k( bot
,i
)
are defined as follows:
The described procedure assures that the FE computation of inductances takes into
account only the air-gap field and is then suitable for comparison with the same quantities
computed analytically through winding function theory.
13.4 Conclusions
In the literature, the use of the winding function theory is substantially confined in the
field of analysis and modeling. In this Chapter, it has been assessed as an effective and
numerically efficient method for the practical computation of phase inductances in a
Fig. 13-25. 6-phase machine mutual inductance between phase “0” and phase j as a function of the rotor position
x, found by direct FE computation and by the proposed winding function theory method.
180 Inductance expression through winding function theory
Fig. 13-26. 9-phase machine mutual inductance between phase “0” and phase j as a function of the rotor position
x, found by direct FE computation and by the proposed winding function theory method.
13.5 References
[1] J. Faiz, I. Tabatabaei, “Extension of winding function theory for non uniform air-gap
in electric machinery”, IEEE Trans. on Magnetics, vol. 38, no. 6, 2002, pp. 3654-3657.
[2] J. Figueroa, J. Cros, P. Viarouge, “Generalized Transformations for Polyphase Phase-
Modulated Motors”, IEEE Transactions On Energy Conversion, vol. 21, June 2006, pp.
332-341.
[3] I. Tabatabaei, J. Faiz, H. Lesani, M. T. Nabavi-Razavi, “Modeling and simulation of a
salient-pole synchronous generator with dynamic eccentricity using modified
winding function theory”, IEEE Trans. on Magnetics, vol. 40, no. 3, 2004, pp. 1550-
1555.
Inclusion of space harmonics in multiphase machine modeling through VSD 181
Vector Space Decomposition (VSD) is a modeling techniques which has been widely
applied to multiphase machines, with both split-phase [1]-[2] and symmetrical [3]-[4]
stator winding configurations. Its theoretical foundation can be traced back to the
Fortescue symmetrical component (SC) theory for polyphase systems [5]. In fact, it is well
known that, when applied to an ideal round-rotor symmetrical n-phase machine, the SC
transformation is capable of decomposing its model into n−1 fully-decoupled component
models [5]. From an algebraic viewpoint, this means that the machine phase inductance
matrix after transformation assumes a diagonal time-invariant structure, which is suitable
for simulation, analysis and control synthesis purposes. It is also a known fact, however,
that such full decoupling effect is achieved only under some restrictive hypotheses
regarding the sinusoidal winding distribution and the uniform air-gap width [4]. For a real
machine with non-sinusoidal winding distribution and non-uniform air-gap, a Park’s
rotational transformation can be introduced in combination with the SC one [4] in order
for the transformed machine model to keep time-invariant. Nevertheless, the transformed
inductance matrix, although time-independent, loses its diagonal structure in this case. In
other words, the presence of space harmonics causes some out-of-diagonal inductance
elements to appear after transformation.
This chapter attempts to solve the aforementioned issues by setting forth a general VSD
method featuring the following peculiarities:
1) The first is a merely geometrical transformation (W) capable of mapping the actual
winding structure into a conventional one (Fig. 14-1); the precise meaning of this
“mapping” operation will be clarified in next.
Fig. 14-1. Two-step transformation for the VSD of a generic multiphase model
Inclusion of space harmonics in multiphase machine modeling through VSD 183
Fig. 14-2. Mapping of a triple-star winding (a) into a symmetrical 9-phase scheme with 2π/9 phase progression
(b); a dual-star winding (c) cannot be mapped into any symmetrical 6-phase scheme with 2π/6 phase
progression.
The overall VSD transformation V(x)=T(x)W will then result from combining the two
transformations. The advantage of this approach is that the properly called VSD theory can
be developed only for the conventional multiphase model (thereby making abstraction of
the particular phase arrangement of the actual machine), instead of tailoring VSD
procedures on any particular multiphase winding topology that may occur in practice.
α = π /n (14.1)
With such a choice, any n-phase winding (whether symmetrical or asymmetrical, with
even or odd phase count) can be mapped into a conventional n-phase arrangement such as
that in Fig. 14-3 by means of a geometrical transformation W, built as detailed in the next
Section.
y A..E = ( y A yE )
t
yB yC yD (14.2)
y 2× ABC = ( y A1 yC 2 )
t
y A2 y B1 yB2 yC 1 (14.3)
y n = ( y0 y2 L yn−1 )
t
y1 (14.4)
where y indicates a generic phase variable, such as a current, voltage or flux linkage and
superscript t indicates transposition. It can be easily seen that the following relationships
must hold for the windings (a), (c) to be respectively equivalent to windings (b), (d) in Fig.
14-4:
1 0 0 0 0
0 0 0 −1 0
y 5 = W5 y A..E , W5 = 0 1 0 0 0 (14.5)
0 0 0 0 − 1
0 0 1 0 0
Inclusion of space harmonics in multiphase machine modeling through VSD 185
1 0 0 0 0 0
0 0 0 1 0 0
0 0 −1 0 0 0
y 6 = W2×3 y 2× ABC , W2×3 =
0 0 0 0 0 − 1 (14.6)
0 1 0 0 0 0
0 0 0 0 1 0
1 if 2 j − i = 0
{Wn }i , j = − 1 if 2 j − i = n i , j = 0, ..., n − 1 ;
(14.7)
0 otherwise
The formal proof of (14.7) is omitted; the formula can be easily checked on a case-by-case
basis.
In this Section the decoupling transformation matrix T(x) is sought which fits an n-phase
machine with conventional multiphase winding structure (Fig. 14-3).
The aim will be attained considering machine steady-state symmetrical and balanced
operation in presence of time harmonics. The full DT definition will encompass three
steps:
2) Secondly, the DT form will be found which projects machine variables into
stationary (αβ) mutually-decoupled reference frames;
mutually-decoupled reference frames. This will be the final DT form will be applied
in the following for setting forth the VSD in case of salient-pole machine topology.
∑ Yh cos[hωt − φh ]
h=h1 ,h2 ,h3 ,...
y0(t )
Y cos[h(ωt − α ) − φh ]
∑
y1(t ) h=h ,h ,h h,...
1 2 3
y sc (t ) = y2(t ) = (14.9)
M ∑ Yh cos[h(ωt − 2α ) − φh ]
h=h1 ,h2 ,h3 ,...
y (t )
n−1 M
∑ Yh cos{h[ωt − (n − 1)α ] − φh }
h=h1 ,h2 ,h3 ,...
where Yh and φh are the hth time harmonic amplitude and phase and h1, h2, h3, …. are
harmonic orders which appear in the phase quantities. As usually done in the literature for
symmetry reasons, we shall suppose in the following that only odd time harmonics are
present, hence h1, h2, h3, … are supposed to be positive odd integers.
yαh1 (t )
yβh1 (t ) y0(t )
y (t ) y (t )
αh2 1
yαβ (t ) = yβh2 (t ) = C y2(t ) (14.10)
yαh3 (t ) M
yβh (t ) yn−1(t )
3
M
For transformation C to perform a VSD, it is required that the couple of variables yαh, yβh
represent the hth harmonic space vector while not depending on any other harmonic
component. In other words, yαh, yβh shall be the components of a space vector of amplitude
proportional to Yh which rotates at hω electrical radians per second in the αh−βh plane.
In order to find the appropriate form of C which leads to VSD, let us expand (14.9) as
follows:
188 Inclusion of space harmonics in multiphase machine modeling through VSD
cos(hωt − φh )
cos(hωt − φh )cos(hα ) + sin(hωt − φh )sin(hα )
y sc (t ) = ∑
h=h1 ,h2 ,h3 ,...
Yh
cos(hωt − φh )cos(2hα ) + sin(hωt − φh )sin(2hα )
M
cos(hωt − φh )cos[(n − 1)hα ] + sin(hωt − φh )sin[(n − 1)hα ]
(14.11)
1 0
cos(hα ) sin(hα )
= ∑ Yh cos(hωt − φh ) cos(2hα ) + sin(hωt − φh ) sin(2hα )
h=h1 ,h2 ,h3 ,... M M
cos[(n − 1)hα ] sin[(n − 1)hα ]
1 0
cos(hα ) sin(hα )
c h = cos(2hα ) , s h = sin(2hα ) (14.12)
M M
cos[(n − 1)hα ] sin[(n − 1)hα ]
y sc (t ) = ∑Y {cos(hωt − φ )c
V
h h h + sin(hωt − φh )sh } (14.13)
LEMMA
Vectors ch, sh ck, sk for odd values of h and k have the following properties:
PROOF
t
Based on (14.12), the product c h c k can be expanded as:
n−1
1 n−1
t
ch ck = ∑ cos( jhα )cos( jkα ) = 2 ∑{cos[ j(h + k )α ] + cos[ j(h − k )α ]}
j =0 j =0
(14.16)
1 n−1 (h + k )π 1 (h − k )π
n−1
= ∑
2 j =0
cos j
+
n 2 j =0 ∑
cos j
n
Since both h and k are positive odd integers, the sum h+k in (14.16) is certainly a positive
even integer and (h + k )π is thereby an integer multiple of 2π: this implies that the first
Inclusion of space harmonics in multiphase machine modeling through VSD 189
sum in the right-side member of (14.16) is always null. As concerns the second sum, if h ≠
k then the difference h − k is a positive even integer too and the sum is zero as well.
Conversely, if h = k, all the terms in the second sum are equal to 1 and the sum thereby
t
equals n. Equation (14.14) is thus proved as far as the product c h c k is concerned.
t
Concerning the product s h s k , this can be expanded as follows:
n−1
1 n−1
t
sh sk = ∑
j =0
sin( jhα )sin( jkα ) = − ∑
2 j =0
{cos[ j (h + k )α ] − cos[ j (h − k )α ]}
(14.17)
1 n−1 (h + k )π 1 n−1 (h − k )π
=− ∑
2 j =0
cos j
+
n 2 j =0
∑cos j
n
Since both h and k are positive odd integers, the sum h+k in (14.17) is certainly a positive
even integer and (h + k )π is an integer multiple of 2π: this implies that the first sum in the
right-side member of (14.17) is always null. As concerns the second sum, if h ≠ k then the
difference h − k is a positive even integer too and the sum is zero as well. Conversely, if h =
k, all the terms in the second sum are equal to 1 and the sum thereby equals n. Equation
t
(14.14) is thus proved also as far as the product s h s k is concerned.
t
As regards (14.15), the products s h c k can be expanded as:
n−1
1 n−1
t
sh c k = ∑
j =0
sin( jhα )cos( jkα ) = ∑
2 j =0
{sin[ j (h + k )α ] + sin[ j (h − k )α ]}
Since h and k are odd integers, their sum and their difference are certainly null or positive
even integers, hence both sums in the right-side members of (14.18) are always equal to
t
zero. The same kind of proof applies to product c k s h , so (14.15) is fully demonstrated.
⃞-
n
υ = trunc (14.19)
2
As a candidate for the VSD transformation through (14.10), let us consider matrix C
defined as follows in terms of vectors (14.12):
190 Inclusion of space harmonics in multiphase machine modeling through VSD
ch t
1
sh t
1t
ch
2 2t
Ch1 ..hυ = s (14.20)
n h2 t
ch
3t
sh3
M
The subscript “h1..hυ” indicates that the transformation C h1 ..hυ is not univocal but depends
on the υ harmonic orders h1, …, hυ which are supposed to exist in phase variables.
For the reasons which will become clear later, we suggest that the structure of C should be
defined in a slightly different way depending on whether n is odd or even. In particular, if
n is even (hence υ=n/2), the definition is (14.21), while if n is odd [hence υ = (n−1)/2 ] the
definition is (14.22).
ch t
1
sh t
1t
ch
2 2t
Ch1 ..hυ = s (14.21)
n h2
M
t
c hυ
s t
hυ
c h1
t
sh1
t
t
c h2
t
2 sh2
Ch1 ..hυ = M (14.22)
n
t
c hυ
shυ
t
1 t
cn
2
It can be seen from (14.12) that, in case of odd n, the last row of C h1 ..hυ is set as a constant
row, equal to:
1 t
c n = (1 − 1 1 − 1 1 L) (14.23)
2
Since the definition of C h1 ..hυ is not univocal, but it depends on the set of υ odd integers
h1..hυ, the question arises as to whether the choice of these υ integers is free or subject to
any restrictions. To answer this question, we need to define the properties which we want
Inclusion of space harmonics in multiphase machine modeling through VSD 191
matrix transformation C h1 ..hυ to have. For the reasons which will be better explained in
14.3.3, we require that matrix C h1 ..hυ be orthonormal, i.e. invertible and such that its
inverse coincides with its transpose. In symbols the following condition must hold:
t t
Ch1 ..hυ Ch1 ..hυ = Ch1 ..hυ Ch1 ..hυ = I (14.24)
THEOREM
A necessary and sufficient condition for (14.24) to hold is that indices h1 .. hn satisfy the
following:
ch t ch t
c h1 sh1
t
c h1 c h2
t
c h1 sh2 L
1 1
sh t c h t
sh1 sh1
t
sh1 c h2
t
sh1 sh2 L
t 2 1 1
Ch1 ..hυ Ch1 ..hυ = ch t ch t
c h2 sh1
t
c h2 c h2
t
c h2 sh2 L (14.26)
n 2t 1 t t t
sh2 c h1 sh2 sh1 sh2 c h2 sh2 sh2 L
M M M M
In virtue of properties (14.14)-(14.15), it is clear that condition (14.25) assures that all
out-of-diagonal elements are zero in (14.26) and all the elements on the main diagonal are
equal to 1. It is worth noticing that this property holds also for odd values of n thanks to
1 t 1
the definition (14.22). In fact it is easy to prove from (14.23) that cn cn = 2n .
2 2
⃞-
192 Inclusion of space harmonics in multiphase machine modeling through VSD
The task is still left to check that C h1 ..hυ accomplishes the VSD. For this purpose, let us
suppose that the phase variable y contains the set of time harmonics of orders h1, h2, …, hυ
and substitute (14.11) into (14.10), obtaining:
0 0
2(i − 1) 2(i − 1)
M M
rows 0 rows 0
1 0
Ch1 ..hυ c hi = , Ch1 ..hυ shi = (14.28)
0 1
0 0
M M
0 0
Therefore (14.27) can be written as:
0
yαh1 (t )
2(i − 1)
(
cos h1ωt − φh1
)
M
yβh1 (t ) rows 0 (
sin h1ωt − φh1 )
y (t ) cos h ωt − φ
( )
αh2 2 cos hiωt − φh ( ) 2
2
(
h2
)
yαβ (t ) = yβh2 (t ) =
∑
Yh i
yαh3 (t ) 0
(
cos h3ωt − φh3 )
yβh (t )
(
sin h3ωt − φh )
M
3 3
M M
0
from which we obtain:
yαh (t ) 2 cos(hωt − φh )
yβh (t ) = nYh sin(hωt − φ ) (14.30)
h
for any h ∈ {h1 ,..hυ }. This means that, at steady state, the couple of transformed variables
yαh and y βh actually describe only the hth time harmonic of phase variables. Such
In conclusion, in this Section it has been shown that C h1 ..hυ is an orthonormal matrix under
condition (14.25) and is suitable for decoupling machine time harmonics of orders h1,..,hu
by projecting them onto independent subspaces. The transformed reference frame is
composed by a set of stationary αh−βh orthogonal axis pairs, such that the hth order
harmonic is represented by a space vector rotating at hω electrical speed in the αh−βh
plane.
Inclusion of space harmonics in multiphase machine modeling through VSD 193
cos(ωht ) − sin(ωht )
ûdh = , ûqh = (14.31)
sin(ωht ) cos(ωht )
The coordinates of space vector (14.30), representing the hth order harmonic, written in
the new rotating dh−qh reference frame defined by (14.31) are:
t yαh (t )
uˆ
ydh (t ) dh yβh (t ) uˆ dht yαh (t ) cos(ωht ) sin(ωht ) 2 cos(ωht − φh )
yqh (t ) = ( ) = t y (t ) = Yh
uˆ t yαh t
uˆ qh βh − sin(ωht ) cos(ωht ) n sin(ωht − φh )
qh yβh (t ) (14.32)
2 cos(φh )
= Yh
n − sin(φh )
Writing (14.32) for all h ∈ {h1 ,..hυ } and resorting to matrix notation, the entire variable
vector y dq (t ) in the rotating reference frames dh−qh can be obtained from the vector
yαβ (t ) in the stationary reference frames αh−βh as follows:
ydh1
Yh1 cos φh1
( )
yqh1 − Yh1 sin φh1 ( )
y Y cos φ ( )
dh2 2 2
h h2
y dq = yqh2 = − Yh2 sin φh2 ( )
n
ydh3 Yh3 cos φh3 ( )
yqh
3
− Yh sin φh
3 3
( )
M M
(14.33)
cos(ωh1t ) sin(ωh1t ) 0 0 0 0 L yαh1 (t )
− sin(ωh1t ) cos(ωh1t ) 0 0 0 L yβh1 (t )
0 0 cos(ωh2t ) sin(ωh2t ) 0 0 L yαh2 (t )
= 0 0 − sin(ωh2t ) cos(ωh2t ) 0 0 L yβh2 (t )
0 0 0 0 cos(ωh3t ) sin(ωh3t ) L yαh3 (t )
0 0 0 0 − sin(ωh3t ) cos(ωh3t ) L yβh3 (t )
M M M M M M O M
A Park’s real transformation matrix Ph1 ..hυ (x ) can be usefully introduced. Its definition is
given by either (14.34) or (14.35) depending on whether the number of phases n is even
or odd respectively.
194 Inclusion of space harmonics in multiphase machine modeling through VSD
In (14.34)-(14.35), the symbol 0p,q indicates the p⨯q null matrix and
cos(hx ) sin(hx )
Ph (x ) = . (14.36)
− sin(hx ) cos(hx )
In conclusion, the decoupling transformation which performs the VSD into rotating
reference frames in presence of υ time harmonics of orders h1, h2, …, hυ is:
matrix) and thereby observing that Ph1 ..hυ (x )Ph1 ..hυ (x ) = I . This guarantees that Th1 ..hυ (x ) ,
t
defined as per (14.38), is orthonormal, too, being the product of orthonormal matrices.
THEOREM
The generic set H of feasible harmonic orders h1, h2, …, hυ which can be used to build DT
matrix C h1 ..hυ has the form:
Inclusion of space harmonics in multiphase machine modeling through VSD 195
Fig. 14-5. Time harmonics and their expression through the formula 2nk±q, with k=0, 1, 2, … and q=1, 3, …, 2υ−1.
(a) Case of even n (n=6): all odd integers are covered; (b) Case of odd n (n=9): all odd integers are covered except
for multiples of n.
h1 2k1n ± 1
h2 2k2n ± 3
H = h3 = 2k3n ± 5 , k1 , k2 ,..., kυ ∈ {0,1,2,3,....} (14.39)
M M
hυ 2kυ n ± (2υ − 1)
PROOF
The aim of the proof is to demonstrate that the set H of indices is the most general set of
integers satisfying condition (14.25).
1) If the number of phases n is even, for any given odd harmonic order h, there exist a
positive or null integer k and a positive integer q ∈ {1, 3, 5, …, 2υ−1} such that:
h = 2kn ± q (14.40)
2) If the number of phases n is odd, proposition 1) is true except for the harmonic
order n and its integer multiples; harmonic order n is in any case included in the
VSD process through the last row of (14.22), which is independent of how indices
h1… hυ are chosen.
3) For any choice of integers k1, …, kυ the set of harmonic orders (14.39) satisfies
(14.25).
196 Inclusion of space harmonics in multiphase machine modeling through VSD
The first two points are quite intuitive and are illustrated by Fig. 14-5 in the cases of n=6
and n=9. To prove point 3), let us take two generic pair of orders hi = 2ki n ± qi and
h j = 2k j n ± q j according to (14.39), i.e. with ki, kj ∈ {0, 1, 2, 3, …}; qi, qj ∈ {1, 3, 5, …, 2υ−1}; qi
≠ qj. Let us consider their sum and difference:
Because qi and qj are distinct and belong to the set {1, 3, 5, …, 2υ−1}, we can certainly write
the following inequalities:
⃞-
Equation (14.39) is of practical use because it assures that choosing a feasible set of
harmonic orders h1, …, hυ is equivalent to choosing υ whatever integers or null numbers
k1, .., kυ together with the relevant signs “±” in (14.39).
d
v s = R sis + ϕs + es (14.44)
dt
v s = (v 0 v n−1 )
t
v1 L v n−2 (14.45)
i s = (i0 i n−1 )
t
i1 L i n−2 (14.46)
ϕ s = (ϕ 0 ϕ1 L ϕ n−2 ϕ n−1 )
t
(14.47)
e s = (e0 en−1 )
t
e1 L en−2 (14.48)
The symbol xk , with x ∊ {v, i, φ, e} and k ∊ {0, 1,…, n−1} represents the kth phase voltage (v),
current (i), flux linkage (φ) or e.m.f. due to the rotor (e). The resistance matrix R is the n⨯n
diagonal matrix having all its diagonal elements equal to phase resistance r:
Inclusion of space harmonics in multiphase machine modeling through VSD 197
r 0 0
0 r 0
Rs = (14.49)
O
0 0 r
Phase flux linkage and current vectors are linked by the stator inductance matrix L which,
for salient-pole machines, is a function of the rotor position x.
ϕ s = L s (x )i s (14.50)
The stator inductance matrix is assumed to be composed of a leakage inductance term L(sl ) ,
not dependent on rotor position, and of an air-gap inductance term L(sag ) (x ) :
v s = R sis +
d
(L s i s ) + e s = R s i s + d L s i s + L s d i s + e s (14.52)
dt dt dt
THEOREM
The leakage inductance matrix of an n-phase machine with phases arranged according to
the conventional scheme has the peculiar Toeplitz structure shown below:
l0 l1 l2 − l3 − l2 − l1
l1 l0 l1 L −l4 − l3 − l2
l l1 l0 − l5 −l4 − l3
2
L(sl ) = M M (14.53)
− l3 − l4 − l5 l0 l1 l2
− l2 − l3 − l4 L l1 l0 l1
− l1 − l2 − l3 l2 l1 l 0
where ℓk indicates the leakage inductance between two phases displaced by kα electrical
radians.
PROOF
The structure of the matrix, besides being confirmed by measurements on real machines
(see 9.7, [6]), can be easily justified also from a theoretical viewpoint based on the two
following considerations.
1. The assumption that all phases are geometrically identical and that their leakage
inductance does not depend on the rotor position necessarily implies that the
mutual leakage inductance depends only on their mutual displacements. This
198 Inclusion of space harmonics in multiphase machine modeling through VSD
accounts for the elements on each diagonal (representing the mutual inductances
between equally-distanced phases) to be equal.
2. Furthermore, let us take a generic index j with 1 ≤ j ≤ n−1 and consider phase j and
phase n−j. Looking at the conventional phase arrangement (Fig. 14-3), it is
immediately seen that phases j and n−j have the same displacement from phase “0”
but opposite conventional directions. This implies that their mutual inductances
with respect to phase “0” have equal magnitude and opposite sign, that is:
l j = − l n− j (14.54)
The combination of the two points above fully justifies the leakage inductance structure
(14.53) for machine with conventionally-arranged phases.
⃞-
0 1 0 0 0
0 0 1 0 0
M O
B= (14.56)
0 0 0 1 0
0 0 0 0 1
−1 0 0 L 0 0
644k4 74448
columns 64n−4 7448
k columns
0 0 ⋅ 0 0 1 0 ⋅ 0 0
0 0 ⋅ 0 0 0 1 ⋅ 0 0
⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅
0 0 ⋅ 0 0 0 0 ⋅ 1 0
k
B = B BKB = 0 0 ⋅ 0 0 0 0 ⋅ 0 1 (14.57)
1
424 3
k times − 1 0 ⋅ 0 0 0 0 ⋅ 0 0
0 −1 ⋅ 0 0 0 0 ⋅ 0 0
⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅
0 0 ⋅ −1 0 0 0 ⋅ 0 0
0 0 ⋅ 0 −1 0 0 ⋅ 0 0
Property (14.57) allows for the matrix structure (14.53) to be expressed as follows:
Inclusion of space harmonics in multiphase machine modeling through VSD 199
n−1
trunc
2
L(sl ) = l 0 I + ∑ l (B
k =1
k
k
− B n− k ) (14.58)
where I represents the n⨯n identity matrix. Expression (14.58) will be useful in the following
to compute the inductance matrix form after VSD transformation.
π R L µ0
mi(,agj ) (x ) =
2
∑ {F W P
s ,r =1 ,3,5,...
s r s +r cos[(s i + r j )α − (s + r )x ]
(14.59)
+ Fs Wr P s −r cos[(s i − r j )α − (s − r )x ] + } π R L µ0
2
∑F W s s P0 cos[s (i − j )α ]
s =1 ,3 ,5 ,...
where α=π/n is the phase progression, R is the radius of the mean air-gap circumference
Γ (see 13.1.3 for exact definition), L is the core length, Fk, Wk and Pk are the Fourier
coefficients computed as per Chapter 13. Here it is important to observe that Fourier
coefficients Fk, Wk and Pk include all the information about the detailed geometry of the
machine, concerning winding distribution and air-gap topology.
Considering that (14.59) is the i, j element of matrix L(scag ) (x ) , the latter can be expressed
as:
π R L µ0
L(scag )( x ) =
2
∑ [
s ,r =1,3,5,...
Fs Wr Ps+r C(r+,s) (x ) + Fs Wr P s−r C(r−,s) (x ) + ] ∑
Fs Ws P0C(s0)
(14.60)
s =1 ,3,5,...
[C (+)
r ,s (x )]i , j = cos[(s i + r j )α − (s + r )x ] (14.61)
[C ( −)
r ,s (x )]i , j = cos[(s i − r j )α − (s − r )x ] (14.62)
[C ]( 0)
s i, j = cos[s(i − j )α ] (14.63)
[C (+)
r ,s (x )]i , j = cos(s iα − s x )cos(r jα − r x ) − sin(s iα − s x )sin(r jα − r x ) (14.64)
[C ( −)
r ,s (x )]i , j = cos(s iα − s x )cos(r jα − r x ) + sin(s iα − s x )sin(r jα − r x ) (14.65)
[C ]
(0)
s i,j = cos(s iα )cos(s jα ) + sin(s iα )sin(s jα ) (14.66)
Equations (14.64)-(14.66) suggest the following factorized expression for C(r+,s) (x ) , C(r−,s) (x )
and C(s0) :
200 Inclusion of space harmonics in multiphase machine modeling through VSD
(14.67)
, s (x ) = c s (x )c r (x ) − s s (x )sr (x )
t t
C(r+)
(14.68)
, s (x ) = c s (x )c r (x ) + s s (x )s r (x )
t t
C(r−)
cos(− kx ) sin(− kx )
cos (α k − kx ) sin (αk − kx )
cos(2αk − kx ) sin(2αk − kx )
c k (x ) = , s k (x ) = (14.70)
cos(3αk − kx ) sin(3αk − kx )
M M
cos[(n − 1)αk − kx ] sin[(n − 1)αk − kx ]
1 0
cos(αk ) sin(αk )
cos(2αk ) sin(2αk )
wk = , (14.71)
cos(3αk ) sin(3αk )
M M
cos[(n − 1)αk ] sin[(n − 1)αk ]
cos(kx ) (14.72)
uk (x ) = ,
sin(kx )
0 − 1
J2 = , (14.73)
1 0
cos(kx )
cos(αk )cos(kx ) + sin(αk )sin(kx )
cos(2αk )cos(kx ) + sin(2αk )sin(kx )
c k (x ) =
cos(3αk )cos(kx ) + sin(3αk )sin(kx )
M
cos[(n − 1)αk ]cos(kx ) + sin[(n − 1)αk ]sin(kx )
1 0 (14.74)
cos(αk ) sin(αk )
cos(2αk ) sin(2αk )
= cos(kx ) + sin(kx ) = w k u k (x )
cos(3αk ) sin(3αk )
M M
cos[(n − 1)αk ] sin[(n − 1)αk ]
Inclusion of space harmonics in multiphase machine modeling through VSD 201
− sin(kx )
sin(αk )cos(kx ) − cos(αk )sin(kx )
sin(2αk )cos(kx ) − cos(2αk )sin(kx )
s k (x ) =
sin(3αk )cos(kx ) − cos(3αk )sin(kx )
M
sin[(n − 1)αk ]cos(kx ) − cos[(n − 1)αk ]sin(kx )
(14.75)
1 0
cos (αk ) sin (α k )
cos(2αk ) sin(2αk ) − sin(kx )
= − sin(kx ) + cos(kx ) = w k = w k J2 u k (x )
cos(3αk ) sin(3αk )
cos(kx )
M M
cos[(n − 1)αk ] sin[(n − 1)αk ]
(14.76)
[
= w s u s (x )ur (x ) + J2 u s (x )ur (x ) J2 w r
t t
] t
(14.77)
[
= w s u s (x )ur (x ) − J2 u s (x )ur (x ) J2 w r
t t
] t
(14.78)
[
= w s u s (0)u s (0) − J2 u s (0)u s (0) J2 w s
t t
] t
sin[(s + r )x ] − cos[(s + r )x ]
cos[(r − s )x ] sin[(r − s )x ]
u s (x )u r (x ) − J2 u s (x )ur (x ) J2 = = Pr − s (x ) ,
t t
(14.80)
− sin[(r − s )x ] cos[(r − s )x ]
where
C(r+,s) (x ) = w s Q s +r (x )w r
t (14.83)
202 Inclusion of space harmonics in multiphase machine modeling through VSD
C(r−,s) (x ) = w s Pr −s (x )w r
t (14.84)
t
C(s0) = w s w s (14.85)
π R L µ0
L(scag )( x ) =
2
∑
s ,r =1 ,3,5,...
[ ]
w s Fs Wr Ps +r Q s +r (x ) + Fs Wr P s −r Pr −s (x ) w r
t
(14.86)
t
+ ∑ Fs Ws
s =1 ,3 ,5 ,...
P0 w s w s
Equation (14.86) is the final expression which will be used for VSD in the next Section.
v dq = T( x ) v s , i dq = T( x )i s , ϕ dq = T( x )ϕ s , e dq = T( x )e s (14.87)
In (14.89) we have made the assumption that the transformation T(x), applied to L s (x ) ,
reduces it to a time-invariant form, i.e. to the constant matrix Ldq which does not depend
on rotor position. This assumption, which is essential for the forthcoming passages [see
(14.93)], is not necessarily verified. The precise conditions for it to hold will be
investigated in Section 14.3.3.1.2 when detailing the explicit form assumed by matrix Ldq.
Using (14.87), the relationship between the flux linkage and current vectors (14.50)
becomes:
ϕ dq = L dq i dq (14.91)
Using the above relationships, the stator voltage equation (14.53) becomes:
Inclusion of space harmonics in multiphase machine modeling through VSD 203
v s = T( x )t v dq = R s T( x )t i dq +
d
dt
[
T( x )t ϕ dq + e s ]
d
= Rs T( x )t idq +
d
[
T( x )t Ldqidq ] [
+ e s = R s T( x )t idq + T( x )t Ldqidq + T( x )t
d
] [ ]
Ldqidq + es (14.92)
dt dt dt
d d
[ ]
= R s T( x )t i dq + T( x )t L dq i dq + T( x )t L dq i dq + e s
dt dt
It is important to remark that in the last passage of (14.92), the assumption has been used
of Ldq being time-invariant, i.e. it has been assumed that
d
dt L dq = 0 (14.93)
d
dt
[L dq i dq ]= [d
dt
]
L dq i dq + L dq d
dt i dq = L dq d
dt i dq (14.94)
d d
T( x )v s = v dq = T( x )R s T( x )t i dq + T( x ) T( x )t L dq i dq + L dq i dq + T( x )e dq
dt dt
dx d d
= R dq i dq + T( x ) T( x )t L dq i dq + L dq i dq + e dq (14.95)
dt dx dt
d d
= R dq i dq + ω T( x ) T( x )t L dq i dq + L dq i dq + e dq
dx dt
where the rotor speed in electrical radians per second has been introduced:
dx
ω= (14.96)
dt
[ ]
The product T( x ) dxd T( x )t in (14.95) can be expanded using (14.38) as follows:
d d d d
T( x ) T( x )t = P( x )C Ct P( x )t = P( x )C Ct P( x )t + Ct P( x )t
dx dx dx dx
t
(14.97)
d d d
= P( x )C C P( x )t = P( x ) P( x )t = P( x ) P( x )
t
dx dx dx
expanded as per (14.98) or (14.99) depending on whether the phase count n is even or
odd [see (14.34)-(14.35)]:
204 Inclusion of space harmonics in multiphase machine modeling through VSD
P( x )[dtd P( x )]
t
t
d cos(hx ) sin(hx ) − hsin(hx ) hcos(hx )
Ph (x ) Ph (x ) =
t
dx − sin(hx ) cos(hx ) − h cos(hx ) − hsin(hx )
(14.100)
cos(hx ) sin(hx ) − hsin(hx ) − hcos(hx ) 0 − 1
= = h = h J2
− sin(hx ) cos(hx ) hcos(hx ) − h sin(hx ) 1 0
having used definition (14.73) for J2. In conclusion, the product P( x )[dxd P( x )] becomes a
t
P( x )[dxd P( x )] = J
t
(14.101)
where J is defined by either (14.102) or (14.103) depending on whether the phase count n
is even or odd.
h1 J 2 02×2 ⋅ 02×2
0 hJ ⋅ 02×2
J = 2×2 2 2 (14.102)
⋅ ⋅ ⋅ ⋅
0 ⋅ hυ J2
2×2 02×2
Inclusion of space harmonics in multiphase machine modeling through VSD 205
It is important to notice that the definition of J [as well as the definitions of C and P(x)] is
not univocal and depends on the choice of harmonic orders h1..hυ; therefore, according to
the extensive notation employed in Section 14.3.1, it would be correct to indicate J as Jh1..hυ
for the same reason why T(x), P(x) and C were extensively designated as Th1..hυ (x), Ph1..hυ
(x) and Ch1..hυ.
Based on (14.101)-(14.103), the final expression for the machine voltage equation (14.95)
in orthonormal coordinates becomes:
d
v dq = R dq i dq + ω JL dq i dq + L dq i dq + e dq (14.104)
dt
From (14.104) a simple expression for the machine electromagnetic torque can be also
derived. In fact, if we left-multiply both sides of (14.104) by idqt we obtain:
t t t t d t
i dq v dq = i dq R dq i dq + ω i dq JL dqi dq + i dq L dq i dq + i dq e dq . (14.105)
dt
n−1
pe = i dq v dq = [T( x )i s ] T( x )v s = i s T( x )t T( x )v s = i s v s =
t t t t
∑v i
k =0
k k (14.106)
is the instantaneous electrical power entering machine terminals; using (14.88) we have
that
n−1
t t t t
p j = i dq R dq i dq = r i dq i dq = ∑ ri
k =0
k
2
(14.107)
t d d d 1 t 1 d t 1 t d
pmag = i dq L dq i dq = w mag = i dq L dq i dq = i dq L dq i dq + i dq L dq i dq (14.108)
dt dt dt 2 2 dt 2 dt
t
is the amount of power used to change the magnetic energy w mag = 12 i dq L dqi dq stored in
machine magnetic circuits. Therefore, (14.105) can be written as:
pe = p j + pmag + pm . (14.109)
where
206 Inclusion of space harmonics in multiphase machine modeling through VSD
t t
pm = ω i dq JL dq i dq + i dq e dq . (14.110)
is the part of the power converted into mechanical power, thereby such that it can be also
written in terms of electromagnetic machine torque Tem and mechanical rotor speed ωm:
ω
pm = Tem ωm = Tem . (14.111)
p
where p is the number of pole pairs. By equaling (14.110) and (14.111) one obtains for the
electromagnetic torque:
t p t
Tem = p i dq JL dq i dq + i dq e dq (14.112)
ω
where the first term represents the reluctance torque component (due to rotor saliency
and acting even in absence of rotor MMF) and the second term represents the torque
component due to the interaction between stator and rotor MMF fields.
The conditions under which Ldq is actually time invariant will be investigated in this
Section by detailing the explicit expression assumed by Ldq. For this purpose, let us
substitute (14.51) into (14.89) and obtain:
[ ]
Ldq = T( x ) L(sl ) + L(sag ) (x ) T( x )t = T( x )L(sl )T( x )t + T( x )L(sag )T( x )t = L(dq
l) ag )
+ L(dq (14.113)
where
l)
L(dq = T( x ) L(sl ) T( x )t (14.114)
ag )
L(dq = T( x ) L(sag ) (x ) T( x )t (14.115)
are respectively the transformed leakage inductance matrix and air-gap inductance
matrix. We shall examine the two terms (14.114) and (14.115) separately.
n −1
trunc
2 t
l)
L(dq = P( x ) C l 0 I +
k = 1
∑ (
l k B k − B n− k )
C P( x )
t
(14.116)
Inclusion of space harmonics in multiphase machine modeling through VSD 207
n−1
trunc
2
= l 0 P( x ) C I C t P( x )t + ∑ l [P( x )C (B
k =1
k
k
)
− B n−k C t P( x )t ]
As concerns the term proportional to l 0 , the property of matrix T(x)=P(x)C of being
orthonormal (see Section 14.3.1) assures that:
P( x )C I Ct P( x )t = P( x )C Ct P( x )t = I (14.117)
As concerns the terms included in the indexed sum, their expansion is slightly more
involved and requires the following preliminary lemma to be proved.
LEMMA
For any odd integer k with 1 ≤ k ≤ n, the following identities hold respectively for even and
odd phase count n:
PROOF
The proof is quite cumbersome and will then be reported by omitting details for the sake
of brevity.
Through a symbolic math tool it is possible to check that matrix B can be always
decomposed as follows:
B = G H DG (14.120)
exp(− ih1α ) 0 0 0 ⋅
0 exp(− ih2α ) 0 0 ⋅
H
G DG = 0 0 exp(− ih3α ) 0 ⋅ (14.121)
0 0 0 exp − ih4α )
( ⋅
⋅ ⋅ ⋅ ⋅ ⋅
1 1 1 1 L
exp(− h1iα ) exp(− h2iα ) exp(− h3iα ) exp(− h4iα ) L
exp(− 2h iα ) exp(− 2h2iα ) exp(− 2h3iα ) exp(− 2h4iα ) L
1 1 (14.122)
G=
n exp(− 3h1iα ) exp(− 3h2iα ) exp(− 3h3iα ) exp(− 2h4iα ) L
M M M M
exp[− (n − 1)h iα ] exp[− (n − 1)h iα ] exp[− (n − 1)h iα ] exp[− (n − 1)h iα ] L
1 2 3 4
1
[G]m,q = exp(− iαhq+1m) ∀m, q = 0, 1,.., n − 1 (14.123)
n
In the above equations, i indicates the imaginary unit and superscript H the complex
conjugate.
Identity (14.120) can be verified by checking that the rows of matrix G are eigenvectors
for B and the diagonal elements of D the relevant eigenvalues.
Bk = G H Dk G . (14.124)
In fact
424
(
3 14444244443
)( ) (
B k = B BKB = G H DG G H DG K G H DG = G H D DKD G
1 1
424 3
) (14.125)
k times k times k times
GGH = I (14.126)
Ct B k C = C t G H D k GC . (14.127)
At this point, using definition of C and G, one can immediately prove that:
1 0 ⋅ 0 0 ⋅ 0 1
−i 0 ⋅ 0 0 ⋅ 0 i
0 1 ⋅ 0 0 ⋅ 1 0
1
(14.128)
GC = 0 −i ⋅ 0 0 ⋅ i 0 ,
2
⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅
0 0 ⋅ 1 1 ⋅ 0 0
0 0 ⋅ −i i ⋅ 0 0
Inclusion of space harmonics in multiphase machine modeling through VSD 209
H
1 0 ⋅ 0 0 ⋅ 0 1
−i 0 ⋅ 0 0 ⋅ 0 i
0 1 ⋅ 0 0 ⋅ 1 0
1
Ct G H = 0 −i ⋅ 0 0 ⋅ i 0
2
⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ⋅
0 0 ⋅ 1 1 ⋅ 0 0
0 0 ⋅ −i i ⋅ 0 0
Substitution of (14.128) and of (14.121) into (14.127), written for k and n−k, leads, after a
few algebraic manipulations, to prove identities (14.118)-(14.119).
⃞-
Equations (14.117) and (14.118) are ready for being substituted into (14.116) together
the expression (14.34)-(14.35) for P(x), giving for even n:
(14.129)
where index k in all the sums ranges from 1 to trunc[(n−1)/2].
l 0 + 2
∑lk
k cos(h1kα )I2 ⋅
02×2 02×1
⋅ ⋅ ⋅ ⋅
l) (14.130)
L(dq =
02×2 ⋅ l 0 + 2
∑l
k
k cos(hυ kα )I2
02×1
01×2 ⋅ 01×2 ∑
l 0 + 2 l k cos(nkα )
k
As a conclusion, we can say that transformation T(x), for any choice of orders h1..hu, leads
to a transformed leakage inductance matrix always having a time-invariant and diagonal
form.
π R L µ0
L(dq
ag )
(x) =
2
∑
s ,r =1 ,3 ,5,...
[
T( x )w s Fs Wr Ps +r Q s +r (x ) + Fs Wr P s −r Pr −s (x ) w r T( x )t ] t
+ ∑ F W P T( x )w w
s
s =1 ,3 ,5 ,...
s 0 s s
t
T( x )t
(14.131)
π R L µ0
= ∑
2 s ,r =1 ,3 ,5,...
[
P( x )Cw s Fs Wr Ps +r Q s +r (x ) + Fs Wr P s −r Pr − s (x ) w r T( x )t
t
]
+ ∑ F W P P( x )Cw w
s
s =1 ,3 ,5 ,...
s 0 s s
t
Ct P( x )t
If we consider the definition (14.21)-(14.22) of C and the definition (14.71) of wk, we can
rewrite C as follows:
wh t
1
2 w h2
t
C= (14.132)
n M
w t
hυ
w h1 t
w h2
t
2
C= (14.133)
n w hυ
t
1 t
cn
2
At this point, we need to make a restrictive hypothesis about the number of space
harmonics (of indices s, r) included in the sum in (14.131). The hypothesis consists of
Inclusion of space harmonics in multiphase machine modeling through VSD 211
supposing that the υ harmonic orders h1…hυ used to build matrices C and P(x) (see Section
14.3.1) are considered in such sum; in symbols, we shall suppose that:
Later, we shall prove that, if this hypothesis is not satisfied, the transformation (14.38)
ag )
does not lead to a time-invariant result L(dq . Conversely, we shall also prove below that, if
ag )
the condition is satisfied, L(dq is time invariant. Therefore, we shall be able to conclude
ag )
that the (14.134) is both a necessary and sufficient condition for L(dq to be time invariant.
π R L µ0 2
ag )
L(dq = ×
2 n
wh t wh wh t wh
t
1 i 1 j
w tw w tw
υ
× P( x ) h2 hi Fhi Whj Phi +hj Q hi +hj (x ) + Fhi Whj P h −h Phj −hi (x ) h2 hj P( x )t
∑
i , j=1 M M
i j
w tw w tw
hυ hi hυ hj (14.135)
wh t wh wh t wh
t
1 i 1 i
υ wh t wh wh t wh
t
∑ P( x ) 2
M
i
Fhi Whi P0 2
M
i
P( x )
i =1
w tw w tw
hυ hi hυ hi
π R L µ0 2
L(dq
ag )
= ×
2 n
wh t wh wh t wh
t
1 i 1 j
wh t wh w tw
υ 2 i
h2 hj
∑
× P( x ) M Fhi Whj Phi +hj Q hi +hj (x ) + Fhi Whj Phi −hj Phj −hi (x ) M
w tw
P( x )
t
i , j=1 w tw
hυ t hi hυ hj
1 cn wh 1 c nt w h
2 i 2 j
(14.136)
wh t wh wh t wh
t
1 i 1 i
wh t wh wh t wh
υ 2 i
2 i
∑ P( x ) M Fhi Whi P0 M t
P( x )
i =1 w tw w tw
hυ t hi hυ t hi
1 c n w hi 1 c n w hi
2 2
where the former applies for even n and the latter for odd n. Concerning the products
whwk, (14.14)-(14.15) together with (14.71) immediately yield:
212 Inclusion of space harmonics in multiphase machine modeling through VSD
n I if h = k
∀h, k ∈ {h1 ,.., hυ }
t
wh wk = 2 2 (14.137)
02×2 if h ≠ k
where I2 and 02×2 are the 2⨯2 identity and null matrices. Therefore, for any k=1..υ:
02×2
P ( x ) 0 0 ⋅ 0
wh t wh h1 2×2 2×2 2×2 M
k − 1 blocks
1 k
0 Ph2 (x ) 02×2 ⋅ 02×2 0
w h t w h n 2×2 2×2
P(x ) 2 k
= 02×2 02×2 Ph3 (x ) ⋅ 02×2 I2
M 2 ⋅ ⋅ ⋅ O ⋅ 02×2
w tw
hυ hk 02×2 02×2 02×2 ⋅ Phυ (x ) M
0
2×2
02×2 (14.138)
M i − 1 blocks
0
n 2×2
= Phk (x )
2
0
2×2
M
0
2×2
02×2
Ph1 (x ) 02×2 02×2 ⋅ 02×2 02×1 M k − 1 blocks
w h1 w hk
t
02×2 Ph2 (x ) 02×2 ⋅ 02×2 02×1 02×2
w h2 w hk
t
n 02×2 02×2 Ph3 (x ) ⋅ 02×2 02×1 I2
P(x ) M =
2 ⋅ ⋅ ⋅ O ⋅ ⋅ 02×2
w hυ w hk
t
02×2 02×2 02×2 ⋅ Phυ (x ) 02×1 M
1
c w hk
t
0
2 n 1×2 01×2 01×2 ⋅ 01×2 1 02×2
0
1×2
02×2 (14.139)
M k − 1 blocks
0
2×2
n P (x )
= hk
2 02×2
M
02×2
0
1×2
where (14.138) and (14.139) respectively hold for even and odd phase count n.
The above results allows us to expand the matrix products involved in (14.135) and
(14.136) as follows:
Inclusion of space harmonics in multiphase machine modeling through VSD 213
t
wh t wh w h t w h
1 i 1 j
wh t wh w t w
P(x ) 2 i
Q hi +hj (x )P(x ) 2
h h j
M M
w tw w t w
hυ hi hυ hj
t
02×2 02×2
i − 1 blocks M M j − 1 blocks
n2 02×2 02×2
= Phi (x )Q hi +h j (x ) Ph j (x )
4 02×2 0
2×2
M M
0
2×2 02×2 (14.140)
The only non-null 2⨯2 block element in (14.140)-(14.141), expanded using (14.81)-
(14.82), is shown to be equal to a constant (x-independent) diagonal matrix as shown
below for any of the index pairs h, k involved:
Pk (x )Q h+k (x )Pk (x )
t
t
wh t wh w h t w h
1 i 1 j
wh t wh w t w
P(x ) 2 i
Q hi + h j (x )P(x )
h2 hj
M M
w tw w t w
hυ hi hυ h j
(14.143)
02(i −1 )× 2( j −1 ) 02(i −1 )×2 02(i −1 )×[2(n− j )+1]
2
n 1 0
= 02×2( j −1 ) 02× [2(n− j )+1]
4 0 −1
0 0 0
[2(n−i )+1 ]×2( j −1 ) [2(n−i )+1 ]× 2 [2(n−i )+1]×[2(n− j )+1]
or
t
wh t wh w h t w h
1 i
1 i
wh t wh w h t w h
2 i
2 i
P(x ) M Q hi +hj (x )P(x ) M (14.144)
w tw w t w
hυ t hi hυ t hi
1 cn wh 1 c n w h
2 i 2 i
Inclusion of space harmonics in multiphase machine modeling through VSD 215
The same analytical procedure followed for the product P(x)CwhQh+kwktCtP(x)t, which led
to equations (14.143)-(14.144), can be repeated for the product P(x)CwhPk−hwktCtP(x)t
with the only difference that we have:
Pk (x )Pk −h (x )Pk (x )
t
t
wh t wh w h t w h
1 i 1 j
wh t wh w t w
P(x ) 2 i
Phj −hi (x )P(x ) 2
h h j
M M
w tw w t w
hυ hi hυ hj
(14.146)
02(i −1 )× 2( j −1 ) 02(i −1 )×2 02(i −1 )×[2(n− j )+1]
2
n 1 0
= 02×2( j −1 ) 02× [2(n− j )+1]
4 0 1
0
[2(n−i )+1 ]×2( j −1 ) 0[2(n−i )+1 ]× 2 0[2(n−i )+1]×[2(n− j )+1]
or
t
wh t wh w h t w h
1 i
1 i
wh t wh w h
t
w h
2 i
2 i
P(x ) M Phj −hi (x )P(x ) M
w tw w t w
hυ t hi hυ t hi
1 cn wh 1 c n w h
2 i 2 i (14.147)
02(i −1 )× 2( j −1 ) 02(i −1 )×2 02(i −1 )×[2(n− j )+1]
2
n 1 0
= 02×2( j −1 ) 02× [2(n− j )+1]
4 0 1
0 0 0
[2(n−i )+1 ]×2( j −1 ) [2(n−i )+1 ]× 2 [2(n−i )+1]×[2(n− j )+1]
The last row of matrix (14.139), which applies for odd n, is always null because none of the
indices h1..hυ can be equal to n (see Section 14.3.1).
t
w h t w h w h t w h
1 i 1 i
02(i −1 )×2(i −1 ) 02(i −1 )×2 02(n−i )×2(n−i )
w h t w h w h t w h t n2
P( x ) 2 i
2 i
P( x ) = 02×2(i −1 ) I2 02×2(n−i ) (14.148)
M M 4
w t w w t w 02(n−i )×2(n−i ) 02(n−i )×2 02(n−i )×2(n−i )
hυ hi hυ hi
or
t
wh t wh w h t w h
1 i 1 i 02(i −1 )×2(i −1 ) 02(i −1 )×2 02(i −1 )×2(n−i ) 02(i −1 )×1
wh t wh w h t w h
2 i
2 i
t n 02×2(i −1 )
2 I2 02×2(n−i ) 02×1
P( x ) M M P( x ) =
w tw w t w 4 02(n−i )×2(n−i ) 02(n−i )×2 02(n−i )×2(n−i ) 02(n−i )×1 (14.149)
hυ t hi hυ t hi 01×2(i −1 ) 01×2 01×2(n−i ) 0
1 c n w hi 1 c n w hi
2 2
Λ (1ag )
Λ (1ag )
⋅ Λ (1ag )
,1 ,2 ,υ
π R L µ 0n Λ (2ag )
Λ 2 ,2 ⋅ Λ 2,υ
( ag ) ( ag )
L(dq
ag )
=
,1
(14.150)
4 ⋅ ⋅ ⋅ ⋅
Λ ( ag ) Λυ( ag,2 ) ⋅ Λυ( ag,υ )
υ ,1
or
Λ (1ag )
Λ (1ag )
⋅ Λ (1ag )
02×1
,1 ,2 ,υ
Λ (2ag
,1
)
Λ 1 ,1 ⋅ Λ 2 ,υ 02×1
( ag ) ( ag )
π R L µ 0
n
L(dq
ag )
= ⋅ ⋅ ⋅ ⋅ ⋅ (14.151)
4
Λ υ( ag,1 ) Λυ( ag,2 ) ⋅ Λ υ( ag,υ ) 02×1
01×2 01×2 ⋅ 01×2 0
depending on whether n is even or odd, respectively, where the generic 2⨯2 block Λ (i ag )
,j ,
F W P
hi h j hi −h j + Phi +h j 0 Fh Wh P0 0
Λ(i ag )
= + δ i , j i i (14.152)
,j
0 Fhi Wh j P h −h − Phi +h j 0 Fhi Whi P0
i j
where δi,j is the Kronecker symbol, equal to 0 or 1 respectively whether i≠j or i=j.
t
We can notice from (14.150) that Λ(i ag ) ( ag ) ( ag )
, j = Λ j ,i = Λ j ,i for any i, j = , which guarantees that
the transformed inductance matrix (14.149) is symmetrical. Because it is also real-valued,
the well known spectral theorem for real symmetric matrices guarantees that it can be
certainly diagonalized through an orthonormal matrix Θ as follows:
Inclusion of space harmonics in multiphase machine modeling through VSD 217
Θ = ΘT (x )L(sag )T (x ) Θt
t
Δ (dqag ) = ΘL(dq
ag ) t
(14.153)
where Δ(ag
dq
)
is diagonal and the orthonormal matrix Θ can be numerically determined by
well-established algorithms known from Numerical Analysis.
In conclusion, we have proved that if the transformation matrix T(x) is built using the
same odd harmonic orders which are considered in expanding machine self and mutual
air-gap inductances inductances as function of the rotor position by means of the winding
function theory through (14.59), then the transformed air-gap inductance matrix is time-
invariant and thereby suitable for use in the VSD model of the machine illustrated in
Section 14.3.3.1.
What is still left to prove is that the condition above is not only sufficient but also
necessary. To prove that, let us call H = {h1 ,..., hυ } the set of harmonic orders used to build
the transformation matrix T(x) and call K = {k1 ,..., kυ } the set of harmonic orders involved
in the air-gap inductance Fourier decomposition according to (14.59). Let us then suppose
that the two sets do not coincide. In this case, all the procedure leading to (14.142) would
hold, and the generic form of (14.142) would be:
Phi (x )Q ki +k j (x )Ph j (x )
t
(14.155)
where, for some pair of indices i and j, we would have that hi≠ hj and/or ki≠ kj, so that
hi + h j ≠ k i + k j .
(14.156)
k i + k j = hi + h j + δ
(14.157)
Phi (x )Q ki +k j (x )Ph j (x )
t
=
[
cos(hi x ) sin(hi x ) cos (hi + h j + δ )x ] [ ] cos(h x )
sin (hi + h j + δ )x j − sin(hj x )
[
− sin(hi x ) cos(hi x ) sin (hi + h j + δ )x
] [ ] sin(h x )
− cos (hi + h j + δ )x j cos(h j x ) (14.158)
cos(δ x ) sin (δ x )
=
sin(δ x ) − cos(δ x )
If δ=0, this means that the product depends on x and, following the procedure seen above,
it is easy to see that the entire transformed inductance matrix would not be constant, as a
consequence. This completes the proof about the condition being necessary.
218 Inclusion of space harmonics in multiphase machine modeling through VSD
Fig. 14-7. Prototype machine: (a) frame; (b) rotor; (c) cross-section. model
Fig. 14-6. (a) Connections from stator coils to the terminal box; (b) terminal box.
Fig. 14-8. Prototype stator winding configurations used for testing: (a) dual star; (b) triple star.
Inclusion of space harmonics in multiphase machine modeling through VSD 219
Fig. 14-9. Connections inside the terminal box to implement: (a) dual-star winding configuration; (b) triple-star
winding configuration.
14-7b) was used in this case in order not to introduce model complications due to rotor
ammortisseur circuits (whose inclusion in machine model is widely covered by existing
literature); it is clear that such a wound-field rotor, if supplied by a constant field current,
behaves like a salient-pole or reluctance-assisted permanent magnet machine.
Implementing different winding schemes on the same machine was possible thanks to the
particular design of the prototype, whose stator coils are individually connected to the
terminal box (Fig. 14-6). Therefore, stator coils can differently connected by changing the
connections in the terminal box (Fig. 14-9).
As a three-phase machine with two parallel paths per phase, the prototype is rated 22
kVA, 760 V, 50 Hz, 3000 rpm.
The prototype testing condition reproduced to validate the modeling strategy proposed in
this Chapter is the sustained short-circuit operation at 1500 rpm and constant field
excitation current. This is achieved by turning the machine with a suitable drive motor
(Fig. 14-7a) while stator phases are short circuited and the field winding is supplied with a
fixed excitation current. Sustained short circuit operation has been chosen as testing
condition because during such operation mode the machine space harmonics, which are
responsible for air-gap field distortion, cause very large circulation currents to circulate in
in shorted stator phases and therefore produce large effects, which should be also caught
by the proposed model where space harmonics are incorporated.
220 Inclusion of space harmonics in multiphase machine modeling through VSD
d
v dq = R dq i dq + ω JL dq i dq + L dq i dq + e dq (14.159)
dt
Written in space-state form with idq chosen as the state variable vector, it becomes:
d
i dq = −L dq (R dq + ω JL dq )i dq + L dq (v dq − e dq )
−1 −1
dt (14.160)
x& = A x + Bu
(14.161)
y =Cx
where:
A = −L dq
−1
(R dq + ω JL dq ) (14.162)
−1
B = −L dq (14.163)
C=I (14.164)
x = i dq (14.165)
u = v dq − e dq (14.166)
By natural phase variables we mean the quantities of machine phases arranged as per Fig.
14-8, that is written in vector form as follows (in the instance of phase currents):
i s ,6 = (i A1 iC 2 )
t
i B1 iC 1 i A2 i B2 (14.167)
i s ,9 = (i A1 iC 3 )
t
i B1 iC 1 i A2 i B2 iC 2 i A3 i B3 (14.168)
where subscripts 6 and 9 respectively indicate the number of phases (n=6, n=9) of the
stator winding configuration.
The geometrical transformation of the actual winding scheme into the conventional one
(see Section 14.2.2.2) is accomplished through matrix (14.8) written for m=3 and N equal
to either 2 (dual-star configuration) or 3 (triple-star configuration), i.e.:
Inclusion of space harmonics in multiphase machine modeling through VSD 221
Fig. 14-10. Magnitude of coefficients FsWsPs+r and FsWsP|s+r| for s and r between 1 and 11 for the 6-phase machine.
1 0 0 0 0 0
0 0 0 1 0 0
0 0 −1 0 0 0
W2×3 = ;
0 0 0 0 0 − 1 (14.169)
0 1 0 0 0 0
0 0 0 0 1 0
1 0 0 0 0 0 0 0 0
0 0 0 1 0 0 0 0 0
0 0 0 0 0 0 1 0 0
0 0 −1 0 0 0 0 0 0
W3×3 = 0 0 0 0 0 −1 0 0 0 .
(14.170)
0 0 0 0 0 0 0 0 − 1
0 1 0 0 0 0 0 0 0
0 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 1 0
As concerns the decoupling matrix T(x), its definition requires choosing the set of υ
harmonics to be included in the VSD model, where υ=trunc(n/2) equals respectively 3 and
4 in the case of the 6-phase and 9-phase winding. The criterion to be followed is obviously
to include those space harmonic orders r, s which are associated to the highest values of
Fourier coefficients Fs Wr Ps +r and Fs Wr P s −r in the expansion (14.59) of self and mutual
air-gap inductances. For example, the magnitude of coefficients Fs Wr Ps +r and Fs Wr P s −r
are reported in Fig. 14-10 with respect to the 6-phase winding configuration. Based on this
criterion, the harmonic orders 1, 3, 5 are chosen for the 6-phase configuration and the
harmonic orders 1, 3, 5, 7 are chosen for the 9-phase configuration.
t
2 c3 2 1 cos(3α6 ) cos(6α6 ) cos(9α6 ) cos(12α6 ) cos(15α6 )
C6 = = (14.171)
6 s3t 6 0 sin(3α6 ) sin(6α6 ) sin(9α6 ) sin(12α6 ) sin(15α6 )
t
c5 1 cos(5α6 ) cos(10α6 ) cos(15α6 ) cos(20α6 ) cos(25α6 )
s t 0 sin(5α ) sin(10α ) sin(15α ) sin(20α6 ) sin(25α6 )
5 6 6 6
where
π π
α6 = , α9 = . (14.173)
6 9
As concerns Park’s matrix P(x), application of (14.34)-(14.35) with the selected harmonic
orders gives:
cos(x ) sin(x ) 0 0 0 0
− sin(x ) cos(x ) 0 0 0 0
0 0 cos(3x) sin(3x) 0 0
P6 (x ) = (14.174)
0 0 − sin(3x) cos(3x) 0 0
0 0 0 0 cos(5x) sin(5x)
0 0 0 0 − sin(5x) cos(5x)
cos(x) sin(x) 0 0 0 0 0 0 0
− sin(x) cos(x) 0 0 0 0 0 0 0
0 0 cos(3x) sin(3x) 0 0 0 0 0
0 0 − sin(3x) cos(3x) 0 0 0 0 0
P9 (x ) = 0 0 0 0 cos(5x) sin(5x) 0 0 0
(14.175)
0 0 0 0 − sin(5x) cos(5x) 0 0 0
0 0 0 0 0 0 cos(7x) sin(7x) 0
0 0 0 0 0 0 − sin(7x) cos(7x) 0
0 0 0 0 0 0 0 0 1
R dq ,6 = r6 I6 (14.178)
R dq ,9 = r9 I 9 (14.179)
where r6 and r9 are the phase resistances in the two configurations and I6, I9 are the 6⨯6
and 9⨯9 identity matrices.
After rotor removal (Fig. 14-11a), one of the stator phases (which we call h) is supplied
with a measured 50 Hz current I. The e.m.f. induced in the other phases at no load is
measured. Calling E k(meas ) the e.m.f. induced in the kth phase, let us call
E k( meas )
mk( meas
,h
)
= (14.180)
I
the mutual inductance between the two phases with the rotor removed. Now we can
repeat the same test with a FE analysis simulation (Fig. 14-11b, Fig. 14-11c), imposing the
same current as in the test and making the FE software compute the induced e.m.f. E k(FEM )
Fig. 14-11. (a) Test on the machine with the rotor removed and one phase supplied; (b) simulation of the test by
FE analysis in the 6-phase configuration; (c) simulation of the test in the 9-phase configuration.
224 Inclusion of space harmonics in multiphase machine modeling through VSD
Fig. 14-12. Self and mutual inductances with rotor removed measured and computed by FE analysis; the
difference gives the leakage inductances. (a) 6-phase configuration; (b) 9-phase configuration.
in phase k, but considering only contribution due to the flux in the stator bore region. This
can be done in the same way as described in Section 16.3.3 where air-gap inductances are
to be computed with the rotor in place, i.e. by adding suitable auxiliary air-gap point for
vector potential computation.
E k( FEM )
mk( FEM
,h
)
= . (14.181)
I
The mutual inductance computed by (14.180), i.e. from measurement, is larger (in
magnitude) that that computed by (14.181). In fact, (14.181) accounts only for the flux in
the stator bore region, while (14.180) includes the contribution of the same stator bore
flux plus the leakage flux (end-coil and slot components). Therefore, the leakage
inductance contribution can be segregated as:
mk( leak
,h
)
= m(kmeas
,h
)
− m(kFEM
,h
)
. (14.182)
Inclusion of space harmonics in multiphase machine modeling through VSD 225
Supposing that phases are numbered according to the conventional multiphase scheme
(Section 14.2.1) and that phase 0 is supplied, the results shown in Fig. 14-12 are obtained
and differences (14.182) directly give leakage inductances l k , with
k=0,1,..,trunc[(n−1)/2]), which appear in (14.53).
π R L µ0 6
ag )
L(dq ,6 = ×
4
226 Inclusion of space harmonics in multiphase machine modeling through VSD
π R L µ0 9
L(dq
ag )
,9 = ×
4
F1 W1 (2P0 + P2 ) 0 L F1 W7 (P6 + P8 ) 0 0
0 F1 W1 (2P0 − P2 ) L 0 F1 W7 (P6 − P8 ) 0
M M L M M 0
×
F7 W1 (P6 + P8 ) 0 L F7 W7 (2P0 + P14 ) 0 0
0 F7 W1 (P6 − P8 ) L 0 F7 W7 (2P0 − P14 ) 0
0 0 L 0 0 0
418 0 − 42.8 0 1.61 0 − 0.994 0 0
0 138 0 − 37.9 0 0.266 0 0.477 0
− 42.8 0 23.7 0 − 5.43 0 0 0 0
0 − 37.9 0 22.6 0 − 4.39 0 0.241 0
= 1.61 0 − 5.43 0 4.05 0 − 0.791 0 0 mH
0 0.266 0 − 4.39 0 4.39 0 − 0.812 0
− 0.994 0 0 0 − 0.791 0 0.615 0 0
0 0.477 0 0.241 0 − 0.812 0 0.626 0
0 0 0 0 0 0 0 0 0
(14.186)
To determine edq the machine no-load voltage at the desired field current (1.58 A) is
measured and computed by FE analysis (as can be seen from Fig. 14-13, the two
waveforms are very close to each other). The vectors es,6(t) and es,9(t) of the n phase
e.m.f.’s arranged in the same order as per (14.167)-(14.168), are then transformed into
Inclusion of space harmonics in multiphase machine modeling through VSD 227
Fig. 14-13. No-load phase voltage waveforms obtained from FE analysis and measurements, with a rotor
excitation current of 1.58 A, (a) in the 6-phase machine configuration; (b) in the 9-phase configuration.
e dq ,6 (t ) = V(ωt )e s ,6 (t ) . (14.187)
e dq ,9 (t ) = V(ωt )e s ,9 (t ) . (14.188)
where the rotor position is set equal to x = ωt with ω indicating the electrical rotation
speed.
The question may arise at this point as what role is played by space harmonics in the
simulated and measured phenomenon. The answer can be given by running the same
simulation assuming a merely diagonal inductance matrix, i.e. neglecting the out-of-
diagonal elements in (14.185)-(14.186). This equivalent to building the machine model
without taking space harmonics into account, i.e. considering only P0 and P2 permeance
function Fourier coefficients and only W1 winding function Fourier coefficient. The result
is given in Fig. 14-14 which clearly highlights that neglecting space harmonic in the VSD
machine model causes simulation results to become definitely inaccurate.
228 Inclusion of space harmonics in multiphase machine modeling through VSD
Fig. 14-15. Current waveforms recorded during the test and obtained by integrating the VSD model including
space harmonics: (a) six-phase configuration; (b) nine-phase configuration.
Therefore we can conclude that the inclusion of space harmonics in the lumped-parameter
modeling of the machine is crucial for obtaining accurate simulation results in multiphase
machines with non-uniform air-gap and non-ideal winding distribution.
Fig. 14-14. Current waveforms resulting from: measurement; simulation with machine VSD model including space
harmonics; simulation with machine VSD neglecting space harmonics.
Inclusion of space harmonics in multiphase machine modeling through VSD 229
Fig. 14-16.Dual-star machine FE model used for time-stepping simulation. Comparison between current
waveforms resulting from measurement and time-stepping FE analysis.
Fig. 14-17.Triple-star machine FE model used for time-stepping simulation. Comparison between current
waveforms resulting from measurement and time-stepping FE analysis.
The possibility to enrich the multiphase machine lumped-parameter model with space
harmonic information is practically important because it provides the designer and the
analyst with a highly fast and convenient alternative to time-stepping FE simulations for
the simulation of all those phenomena where important space harmonic effects are
expected.
14.5 References
[1] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction machine
using vector space decomposition”, IEEE Trans. on Industry Application, Sept.-Oct.
1995, vol. 31, pp. 1100-1109.
[2] D. Hadiouche, H. Razik, A. Rezzoug, “On the modeling and design of dual-stator
windings to minimize circulating harmonic currents for VSI-fed AC machines”, IEEE
Trans. on Industry Applications, vol. 40, Mar./Apr. 2004, pp. 506-515.
[3] S. Gataric, “A polyphase Cartesian vector approach to control of polyphase AC
machines”, 2000 IEEE IAS Annual Meeting, Rome, Italy, 2000, pp. 1648-1654.
[4] J. Figueroa, J. Cros, P. Viarouge, “Generalized Transformations for Polyphase Phase-
Modulated Motors”, IEEE Transactions On Energy Conversion, vol. 21, June 2006, pp.
332-341.
[5] D.C. White, H.H. Woodson, Electromechanical Energy Conversion, John Wiley and Sons,
New York, 1959.
230 Inclusion of space harmonics in multiphase machine modeling through VSD
[6] A. Tessarolo, C. Bassi, “Stator harmonic currents in VSI-fed synchronous motors with
multiple three-phase armature windings”, IEEE Trans. on Energy Conversion, vol. 25,
no. 4, Dec. 2010, pp. 974-982.
[7] A. Tessarolo, “On the modeling of poly-phase electric machines through vector-space
decomposition: theoretical considerations”, International Conference on Power
Engineering, Energy and Electrical Drives, POWERENG 2009, Lisboa, Portugal, 18-20
March 2009, pp. 519-523.
[8] E. Levi, R. Bojoi, F. Profumo, H. A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Applications, IET, vol. 1, Jul. 2007,
pp. 489-516.
[9] M.J. Duran, E. Levi, M. Jones, “Indipendent vector control of asymmetrical nine-phase
machines by means of series connection”, IEEE International Electric Machines and
Drives Conference, IEMDC, 2005, 15-18 May 2005, San Antonio, Texas, USA, pp. 167-
173.
231
The modeling and simulation work described in the Previous Parts has provided with the
analysis tools required to describe and explain in detial some aspects of multiphase machine
operation. In particular, this part collects two examples of multiphase machine analysis
concerning the operation of multiple-star motor when supplied by the two main types of
inverters presently used in high power applications: the Load Commtated Inverter (LCI) as
which is an example of Current Source Inverter (CSI) and the multilevel Medium Voltage
PWM inverter, which is an example of Voltage-Souruce Inverter.
Converning LCI-fed motors, the typical case is investigated in Chapter 15 of a machine with N
three-phase sets, each supplied by an LCI independently. It has been theoretically predicted
and experimentally observed that, in these machine, the phase current commutation
phenomena may exhibit a strongly different dynamics compared to the case of three-phase
LCI-fed cases. The difference is due to the fact that the commutations occurring in one
winding section affect the other winding sections through stator mutual inductances and this
may significantly slow down the commutation transient if two commutations occur
simultaneously in different motor stars. These considerations have been formalized in
analytical terms and led to set forth a relatively simple circuit-analytical model capable of
describiling the commutation transient including the mutual interaction effects among
simultaneously commutating stator stator phases. The model has been successfully validated
against experimental measurements on a dual-star synchronous motor fed by two LCIs.
As concerns VSI-fed motors, the case is investigated in Chapter 16 of a machine with N three-
phase sets, each supplied by a PWM inverter independently. Some experimental observations
have shown that, during the operation of such drive systems, quite large low-freqeuncy
current harmonics arise in addition to the expected PWM-excited high-freqeuncy harmonics.
The phenomenon could be accurately reproduced by time-stepping Finite Element analysis
but some uncertainty remained about its origin and on the possibility to predict it
analytically. Chapter 16 takes the example of a quadruple-star synchronouys motor fed by
four PWM inverters and, based on the Vector-Space Decomposition (VSD) technique
described in Parts IV and V, it is shown that low-frequency circulation harmonics can be
predicted also analytically with a definitely good accuracy with respect to experimental
measurements. The analysis has also led to physically interpreting the phenomoneon as a
consequence of the 5th and 7th harmonic distortion of the air-gap flux due to the non-uniform
air gap and to the non-sinusoidal winding distribution. The phenomenon thereby constitutes
a typical example of how space hamrmonic effects can strongly affect multiphase machine
232
Nomenclature
N Number of stator windings of the split-phase machine.
τ Displacement angle between stator windings.
µ Commutation angle.
α1, α2 Firing angles of the LCIs that supply windings 1, 2.
vab1, vab2 Line-to-line (a-b) voltages of stator windings 1, 2.
eab1, eab2 Motor e.m.f. between phases a, b in windings 1, 2.
ia1, ib1 Currents in phases a, b of stator winding 1.
ia2, ib2 Currents in phases a, b of stator winding 2.
Idc DC-link current (average value).
Lc Commutation inductance of a three-phase machine.
i1, i2 Commutation circulating current in windings 1, 2.
Ld′′ , Lq′′ Sub-transient inductances of an individual winding.
V, E, I Fundamental of measured supply voltage (V), e.m.f. (E) and current (I).
φV,I Phase shift between V and I.
φE,I Phase shift between E and I.
15.1 Introduction
Split-phase synchronous motors are special electric machines whose stator phases are
grouped into two or more (N) three-phase windings, normally displaced by 60/N electrical
degrees apart [2]-[6] each suitable for being supplied by a three-phase inverter
independently. The advantages of electrical drives based on split-phase motors have been
widely investigated [2], [1], [1]; they mainly relate to the overall drive power rating
234 Modeling of commutation transients in split-phase LCI-fed motors
Fig. 15-1. Stator phase arrangement in a phase-split motor with N windings, each composed of three symmetrical
phase set (a, b, c).
As concerns split-phase motors used in Current-Source Inverter (CSI) drives, a dual three-
phase wound-rotor synchronous machine (N=2), fed by a couple of LCIs, is normally
employed [10], [11]. This solution is consolidated for those applications where a very high
power rating (in the tens of megawatts range), along with some degree of fault-tolerance,
is to be guaranteed while a relatively poor dynamic performance may be accepted [11].
Under some conditions, the use of LCI-fed synchronous motors with more than two three-
phase sections (N>2) would magnify the merits of the dual three-phase configuration
(N=2) compared to the single three-phase one [12]. In particular, [12] investigates the
expected benefits of adopting a quadruple three-phase synchronous machine (N=4)
supplied by four LCIs in terms of power factor, efficiency and output torque quality.
This electromagnetic transient is an issue that has been investigated in [14] in detail. As a
result, it has been shown how the split-phase motor behaviour during overlapping
commutations can be described through a simple model, which may be expressed in either
circuit or analytical form. In this Chapter, such model will be briefly discussed (Section
15.3) and validated through measurements collected on a real dual LCI synchronous
motor drive (Section 15.4). Finally, the results of some dedicated tests on a prototype
Modeling of commutation transients in split-phase LCI-fed motors 235
Fig. 15-2. Current waveforms in two motor phases (e.g. a1 and a2) displaced by an electrical angle τ, at electrical
pulsation ω.
synchronous machine will be presented (Section 15.5) to assess the model validity in
various split-phase winding configurations and for different rotor positions.
τ = π /(3N ) . (15.1)
Fig. 15-2 schematically highlights how the current waveforms of a generic couple of
phases, physically displaced by τ radians apart (like phases a1 and a2, Fig. 15-1), are
shifted by the same electrical angle τ. In virtue of (15-1), it is evident that for increasing N
such phase shift τ diminishes progressively so that the following condition may occur:
τ ≤µ, (15.2)
where µ indicates the commutation angle [13]. Condition (15-2) implies that
commutations, which take place in different stator windings, overlap (as can be inferred
from Fig. 15-2).
In practice, for N=2, N=3, N=4, N=5 we have that τ respectively equals 30, 20, 15, 12
electrical degrees, thus possibly drooping lower than µ. Furthermore, as discussed in [14],
should the LCIs which supply different motor windings be controlled with different firing
angles (e.g. in case of absent or faulty synchronization), a commutation overlap could
occur even for smaller values of µ than given by (15-2).
236 Modeling of commutation transients in split-phase LCI-fed motors
Fig. 15-3. Simultaneous commutation of a1-b1 and a2-b2 phase pairs. Thin lines and empty SCRs denote paths
where no current flows.
In the event of simultaneous commutation, phases a2, b2 start commutating (Fig. 15-2,
point B) before ia1 has extinguished. The resulting scenario is depicted in Fig. 15-3, where
i2 denotes the circulation current that the e.m.f. eab2 drives in the loop formed by phases
a2, b2 (only windings 1 and 2 are included in Fig. 15-3 for the sake of simplicity).
During simultaneous commutations (15-3) does not hold any more. In fact, the circulation
current i2 affects the rate of change of i1 due to the magnetic coupling among phases
through stator mutual inductances, and vice versa. Furthermore, the presence of the rotor
and its electric circuits ― a ϐield winding (f) and a damper (kd) on the d axis; a damper (kq)
on the q-axis ― needs to be taken into account.
The analytical formalization of the phenomenon is reported in [14], where the differential
equations of a split-phase machine during single and overlapping commutations are
Modeling of commutation transients in split-phase LCI-fed motors 237
manipulated leading to a simple model that describes how i1, i2 evolve during the
transient. The model representation in terms of equivalent circuit is shown in Fig. 15-4.
Explicit expressions for parameters L11, L22, L12, that appear in the circuit, are given by (15-
4)-(15-6) in terms of the sub-transient inductances Ld′′ , Lq′′ (of an individual three-phase
stator winding) and of rotor position θ, measured in electrical radians with respect to the
magnetic axis of phase a1 (Fig. 15-3).
( ) ( )
L11 = Ld′′ + Lq′′ + Ld′′ − Lq′′ cos[2(θ + π / 6 )] , (15.4)
( ) ( )
L22 = Ld′′ + Lq′′ + Ld′′ − Lq′′ cos[2(θ − τ + π / 6 )] , (15.5)
( ) ( )
L12 = Ld′′ + Lq′′ cos(τ ) + Ld′′ − Lq′′ cos[2(θ − τ / 2 + π / 6)] + ∆Lσ , (15.6)
Furthermore, (15-6) includes a leakage inductance term ∆Lσ, whose expression has been
derived as follows [14]:
(
∆Lσ = 2 Lσa1, a 2 − Lσa1, b 2 − Lσb1, a 2 − 2 cos(τ ) Lσa1, a1 − Lσa1, b1 , ) (15.7)
where Lσa1, a1 denotes the self leakage inductance of a stator phase and LσX ,Y denotes the
mutual leakage inductance between two generic phases X and Y. Because the self leakage
inductance Lσa 1, a 1 is usually much larger than the mutual ones, ∆Lσ is expected to be
negative, as confirmed by measurements (Section 15.5).
The equivalent circuit of Fig. 15-4 directly yields the time derivatives of the circulation
currents i1, i2 during simultaneous commutation as per (15-8).
In virtue of the method used in its derivation [14], the model also holds in the hypothesis
that one commutation happens at a time (which is the case of normal drive operation):
setting i2=0 and i1=0, Fig. 15-4 respectively yields:
Equations (15-9) are consistent with (15-3), provided that the commutation inductance Lc
is approximated as a θ-independent value equal to the average of Ld′′ and Lq′′ as proposed
in [13]. The approximation is equivalent to neglecting the θ-dependent terms proportional
to Ld′′ − Lq′′ in (15-4)-(15-5), which seems reasonable because the difference Ld′′ − Lq′′ is
much less than Ld′′ + Lq′′ in high-power synchronous machines with normal damper circuit
construction (see Appendix A for an example).
To summarize, the proposed model (expressed by Fig. 15-4 in circuit form, by (15-8)-(15-
9) in analytical form) enables to extend and refine the concept of commutation inductance
Lc, introduced for three-phase machines [13], in such a way that: (a) the influence of rotor
position θ is considered; (b) the case of simultaneous commutation events, that may occur
in split-phase machines, can be accounted for, as well.
Fig. 15-5. Functional block-scheme of the dual LCI synchronous motor drive.
Modeling of commutation transients in split-phase LCI-fed motors 239
Fig. 15-6. Main drive components: (a) dual 3-phase synchronous motor; (b) dual converter power electronics; (c)
control boards.
To further assess the model validity and to illustrate its practical applications,
measurements collected on a real drive are reported hereinafter for comparison with
model outputs.
The equipment is mainly composed of two SCR inverters (LCI#1 and LCI#2) which supply
the stator windings of a dual three-phase synchronous motor (N=2, τ=π/6). The two LCIs
are synchronized to the respective stator windings with firing angles (α1, α2) that can be
manually and independently adjusted. The two grid-side SCR rectifiers, instead, are
synchronized to the grid voltages with firing angles (β1, β2) which are the outputs of DC-
link current controllers. The reference DC-link current (Idc*), in turn, results from a speed
controller which regulates the motor speed to the desired value (ω*). Finally, the motor
voltage is maintained at its commanded value (V*) acting on the field excitation supply.
A digital oscilloscope is used to record phase currents ia1, ia2 and line-to-line voltages vab1,
vab2 with a sampling interval of 10−7 s.
Fig. 15-7a highlights that both voltage and current waveforms of the two windings are
shifted by τ=π/6 electrical radians apart; the commutation of phase a1 starts (at point C)
with a delay of α1 electrical radians from the zero-crossing instant (A) of voltage vab1;
similarly, the commutation of phase a2 starts (D) with a delay of α2 electrical radians from
the zero-crossing instant (B) of voltage vab2.
240 Modeling of commutation transients in split-phase LCI-fed motors
Fig. 15-7. Normal drive operation with equal firing angles α1=α2 =125°. (a) Recorded phase currents ia1, ia2 and
recorded line-to-line voltages vab1, vab2; (b) zoomed view around point C, where phase a1 starts commutating.
By zooming on point C, the normal commutation transient can be seen in more detail (Fig.
15-7b). This enables to appreciate how the rate of increase of ia1 is well predicted through
(15-9), where i1′=−ia1′ (Section 15.3), so that:
In fact, since the machine under testing has Ld′′ ≅ Lq′′ , parameters L11 and L22 are almost
independent of θ and, for any rotor position, their measured value is (Appendix A):
Concerning eab1, it follows vab1 except for commutation voltage dips (Section 15.3); hence
form Fig. 15-7a one can write:
Fig. 15-8. Drive operation with different firing angles (α1=155°, α1=128°). (a) Recorded phase currents ia1, ia2 and
recorded line-to-line voltages vab1, vab2; (b) zoomed view around point A, showing a simultaneous commutation
transient that involves phases a1-b1 and a2-b2.
At instant t0 phase a1 starts commutating first. While ia1 is still growing, namely at instant
t1, phase a2 starts commutating, too. Therefore, between t1 and t2 both a1-b1 and a2-b2
phase pairs are commutating simultaneously. During this interval it is evident from Fig.
15-8b how a2-b2 commutation current strongly affects the dynamics of a1-b1
commutation due to the mutual magnetic couplings between phases [14]. The influence is
such that the slope of ia1 is even reversed (time derivative becomes negative) until a2-b2
commutation ends, i.e. until t2. After this instant, current ia1 continues increasing with the
same time derivative it had for t0<t<t1, up to the end of the commutation (t3).
As shown in Fig. 15-8b, the dynamics of the overall commutation transient can be well
predicted by the proposed model. In fact, during intervals t0-t1 and t2-t3, when only one
winding (i.e. winding 1) commutates, the first of (15-9) holds, with
ia1′ = 1.33×105 A/s (t0< t< t1, t2< t< t3). (15.14)
During t1-t2 interval, when both windings are commutating together, equations (15-8)
become applicable instead, with:
ia1′ = − i1′, ia2′ = − i2′, eab1 ≅ −VM sin(α1), eab2 ≅ −VM sin(α2),
(15.15)
L11 ≅ L22 ≅ 1.72×10−3 H, L12 ≅ 1.24×10−3 H.
The numeric evaluation of (15-8) with the above data yields to:
ia1′ = −9.51×104 A/s, ia2′ = 3.16×105 A/s (t1< t< t2). (15.16)
The slopes given by (15-14) and (15-16) are those used to plot the diagrams “from
calculation” in Fig. 15-8b.
The simple model proposed in this Chapter is able to predict the split-phase machine
behavior during both normal and overlapping commutation transients with adequate
accuracy. The use of such a simple analytical or circuit model may represent a useful
alternative to the complete modeling and numeric simulation of the entire drive system
[12], [14], especially when a very large amount of commutation transient scenarios needs
to be investigated in the drive design stage.
Modeling of commutation transients in split-phase LCI-fed motors 243
Fig. 15-9. Prototype machine used for testing: (a) view of internal structure with connections from each single
stator coil to the terminal box; (b) terminal box.
For the above purposes, a laboratory test-bench has been prepared based on a prototype
salient-pole wound-rotor synchronous machine (nameplate ratings are reported in
Appendix B). The double-layer short-pitch stator winding of the prototype has been
designed so that both leads of each individual coil are accessible from the terminal box
(Fig. 15-9a). Hence, a wide variety of winding arrangements can be implemented by
simply changing the connections inside the terminal box (Fig. 15-9b). In particular, for the
purposes of this Chapter, the phase-split configurations characterized by N=2, N=3, N=4
three-phase windings have been implemented, according to the schemes of Fig. 15-10.
For each of the three configurations, the following tests were carried out:
Fig. 15-10. Physical phase-belt extension and arrangement over a single double-layer pole span, for N=2 (A), N=3
(B) and N=4 (C).
244 Modeling of commutation transients in split-phase LCI-fed motors
Fig. 15-11. Test circuit and measured quantities for determination of L11, L12.
2. Detection of the magnetic axis of phase a1. For this purpose, the rotor field winding
is supplied with a 50 Hz fixed current and the open-circuit e.m.f. induced in phase
a1 was recorded for several equally-distanced rotor positions. The position
corresponding to the maximum of the e.m.f. curve is considered as the magnetic
axis of phase a1, i.e. as the zero-reference for rotor position θ (Fig. 15-3).
3. Measurements of parameters L11, L22, L12. Parameters L11 and L22 are determined as
the self inductances of the circuit loops respectively formed by phases a1, b1 and
phases a2, b2 (Fig. 15-3, Fig. 15-4). Parameter L12 was measured as the mutual
inductance between the mentioned loops (Fig. 15-4). Therefore, to measure L11
and L12 the test circuit of Fig. 15-11 is used: while supplying the series of phases
a1, b1 with a voltage V and current I of frequency f=50 Hz, the open-circuit e.m.f
E induced across the series of phases a2, b2 was recorded for several equally-
distanced rotor positions θ, measured from the magnetic axis of phase a1 as per 2).
The values of L11 and L12 are then computed as per (15-17), where φV,I and φE,I are
the measured phase shifts between phasors V , I and E , I respectively.
For the measurement of L22, the same test circuit is employed, but with the series
of phases a2, b2 supplied and the series of phases a1, b1 at open circuit.
Fig. 15-12. Inductive parameters L11, L22, L12 as functions of rotor position for different split-phase configurations
(N=2, N=3, N=4). Circles represent measured values, solid lines result from plotting equations (4)-(6).
Fig. 15-12 illustrates the comparison between the values of L11, L22, L12 measured as per 3)
and the values obtained from the analytical model (15-4)-(15-6) for each of the split-phase
configurations taken into account; the abscissa θ=0 in all diagrams corresponds to the
position of the phase a1 magnetic axis detected as per 2). In Table I the parameters used
for the analytical model (15-4)-(15-6) are reported, being Ld′′ and Lq′′ measured as per 1)
and ∆Lσ computed through (15-7) from stator leakage inductances (Appendix C).
TABLE I
VALUES OF ANALYTICAL MODEL PARAMETERS
τ (deg) 30 20 15
246 Modeling of commutation transients in split-phase LCI-fed motors
15.6 Conclusion
High-power electrical drives using split-phase synchronous machines equipped with N
stator windings, each supplied by a Load Commutated Inverter (LCI), are investigated in
this Chapter. The need for growing power levels, reliability and performance in such drive
systems would lead designers to increase the number N of stator windings. One of the
main criticalities thus encountered is the possible overlap between commutations which
take place in different machine windings. As experimentally proved in this Chapter, the
occurrence of such event, although not necessarily resulting from a fault or a trip, gives
rise to abnormal phenomena like significant changes in commutation current derivatives.
The use of the proposed model in the drive design stage enables to analytically predict the
split-phase machine behavior during various possible commutation scenarios (different
voltage and load conditions, different firing angles, different machine parameters, etc.)
with no need for the entire drive system (motor, inverter, control) to be modeled,
implemented and studied on a case-by-case basis through time-consuming numerical
simulations.
Appendix A
The technical data of the dual three-phase LCI drive used for the testing activity reported
in Section 15.4 are provided next.
Ratings of the dual three-phase motor (Fig. 15-6a): 250 kW, 380 V, 50 Hz, 0.9 power factor,
1500 rpm, 4 poles, two stator 3-phase windings shifted by 30 electrical degrees apart.
Motor sub-transient inductances: Ld′′ =0.85 mH, Lq′′ =0.87 mH. These inductances have
been determined as per Section 15.5-1. Model parameters L11, L22, L12 have been identified
with the procedure described in Section 15.5-3 obtaining: L11=1.72×10−3 H, L12=1.24×10−3
H regardless of rotor position.
Each SCR converter used to supply the motor (Fig. 15-5, Fig. 15-6b-d) is equipped with a
DC-link inductance Ldc=3.8 mH.
Appendix B
The nameplate ratings of the prototype synchronous machine used for the testing
described in Section 15.5 (referred to the single three-phase stator winding configuration)
are as follows: 22 kVA, 760 V, 50 Hz, 0.8 power factor, 3000 rpm.
Modeling of commutation transients in split-phase LCI-fed motors 247
Appendix C
In this Appendix the procedure followed to measure stator phase leakage inductances of
the split-phase machine prototype is presented. The sought parameters, which appear in
the expression (15-7) for ∆Lσ, are in fact quite cumbersome to estimate by calculation,
especially as concerns their component due to end-winding leakage flux (some analytical-
numeric methods for this purpose are discussed in [23]). On the other hand, even for
ordinary three-phase machines, stator leakage inductances are known to be difficult both
to compute and to measure [17], [10]. The strategy used for this Chapter basically follows
the guidelines provided by [23], [17], [10] extending them to the multi-phase winding
configurations in a quite natural way.
As in three-phase machines, a good basis for stator leakage inductance measurement is the
test with the rotor removed. Let us suppose that the mutual leakage inductance LσX ,Y
between two generic stator phases X, Y is to be determined (obviously becoming a self
inductance if X=Y). For this purpose, the total mutual inductance LrrX ,Y between X and Y,
with the rotor removed (rr), is first measured. This value is actually due not only to the
slot and end-coil leakage flux, but also to the flux passing through the region normally
occupied by the rotor [17], i.e. in the stator bore region. In other words, Lrr
X ,Y can be
expressed as:
where Lbore
X ,Y denotes the inductance component due to the flux in the stator bore. The
latter term, however, is relatively easy to determine, e.g. through a Finite Element (FE)
analysis of the machine with the rotor removed, as proposed in [17]. The machine stator,
in each split-phase configuration under test, was then modeled and analyzed through a 2D
FE tool, energizing the supplied phase with the same current as measured during the test
on the actual machine. From the solved FE model, it was straightforward to determine the
σ
value of Lbore
X ,Y [17] and thus to find L X ,Y by subtracting it from the measured
Fig. 15-13. Test-set up (a) used for checking the FE model correct tuning (b) through the use of a search coil
inside the stator bore after rotor removal.
248 Modeling of commutation transients in split-phase LCI-fed motors
inductance Lrr
X ,Y .
This procedure is repeated for all the phase pairs involved in (15-7) and its results are
reported in Table II and Table III. In particular, the values of Table III are used to compute
∆Lσ according to (15-7).
Each time a FE program has to be used to compute some quantities that cannot be directly
measured (as occurs in this case for the flux in the stator bore region), it is fundamental
that the FE model be previously checked to be correctly tuned, so that the computed
values can be deemed as reliable. This check is successfully performed configuring the
prototype as an ordinary three-phase machine through the appropriate connections in its
terminal box and then testing it as prescribed by [10] for the test with the rotor removed.
Namely, all the three stator phases are supplied with a symmetrical system of 50 Hz
currents, while a search coil, shaped and placed according to [10], is located inside the
stator bore (Fig. 15-13). The test is then simulated on the FE model, with the same current
and frequency, setting the axial depth of the 2D model equal to the actual stator core
length (150 mm). The e.m.f. measured at the search coil terminals is finally compared with
the value obtained from the simulation. As summarized in Table IV, a discrepancy around
5% is found, which represents a sufficient level of accuracy for the FE model employed.
TABLE II
PHASE INDUCTANCES WITH THE ROTOR REMOVED (mH)
Lrr bore
a1, a1 , La1, a1 11.347, 4.85 5.354, 2.280 3.710, 1.387
Lrr bore
a1,b1 , La1,b1 −2.257, −1.126 −0.971, −0.490 −0.510, −0.278
Lrr bore
a1, a 2 , La1, a 2 5.878, 2.959 3.452, 1.728 1.962, 1.034
Lrr bore
a1,b 2 , La1,b 2 −5.681, −2.985 −1.808, −0.933 −0.850, −0.426
Lrr bore
b1,a 2 , Lb1,a 2 −0.004, 0.004 −0.314, −0.152 −0.263, −0.116
TABLE III
MEASURED PHASE LEAKAGE INDUCTANCES (mH)
TABLE IV
CHECK FOR FE MODEL ACCURACY
FE simulation Test
15.7 References
[1] E. Levi, “Multiphase electric machines for variable-speed applications”, IEEE Trans.
on Industrial Electronics, vol. 55, May 2008, pp. 1893-1909.
[2] K. Marouani, L. Baghli, D. Hadiouche, A. Kheloui, A. Rezzoug, “A new PWM strategy
based on 24-sector vector space decomposition for a six-phase VSI-fed dual stator
induction motor”, IEEE Trans. on Industrial Electronics, vol. 55, May 2008, pp. 1910-
1920.
[3] D. Yazdani, S.A. Khajehoddin, A. Bakhshai, G. Joòs, “Full utilization of the inverter in
split-phase drives by means of a dual three-phase space vector classification
algorithm”, IEEE Trans. on Industrial Electronics, vol. 56, Jan. 2009, pp. 120-129.
[4] M.A. Shamsi-Nmejad, B. Nahid-Mobarakeh, S. Pierfederici, F. Meibody-Tabar, “Fault
tolerant and minimum loss control of double-star synchronous machines under
open phase conditions”, IEEE Trans. on Industrial Electronics, Vol. 55, May 2008, pp.
1956- 1965.
[5] M.J. Duran, E. Levi, M. Jones, “Indipendent vector control of asymmetrical nine-phase
machines by means of series connection”, IEEE International Electric Machines and
Drives Conference, IEMDC, 2005, 15-18 May 2005, San Antonio, Texas, USA, pp. 167-
173.
[6] L. Hua, Z. Yunping, H. Bi, “The vector control strategies for multiphase synchronous
motor drive systems”, IEEE Industrial Electronics International Symposium, ISIE
2006, 9-13 July 2006, Montréal, Canada, pp. 2205-2210.
[7] E. Levi, R. Bojoi, F. Profumo, H.A. Tolyat, S. Williamson, “Multiphase induction motor
drives – a technology status review”, Electric Power Application, IET, 2007, July
2007, vol. 1, pp. 489-516.
[8] B. Bose, “Power Electronics and Motor Drives: Recent Progress and Perspective”,
IEEE Trans. on Industrial Electronics, vol. 56, Feb. 2009, pp. 581-588.
[9] K.K. Mohopatra, R.S. Kanchan, M.R. Baiju, P.N. Tekwani, K. Gopakumar, “Independent
field-oriented control of two independent induction motors from a single six-phase
inverter”, IEEE Trans. on Industrial Electronics, vol. 52, Oct. 2005, pp. 1372-1382.
[10] A. N. Alcaso, A. J. M. Cardoso, “Remedial operating strat[10]egies for a 12-pulse LCI
drive system”, IEEE Trans. on Industrial Electronics, vol. 55, May 2008, pp. 2133-
2139.
[11] B. Wu, J. Pontt, J. Rodriguez, S. Bernet, S. Kouro, “Current-source converter and
cycloconverter topologies for industrial medium-voltage drives”, IEEE Trans. on
Industrial Electronics, vol. 55, July 2008, pp. 2786-2797.
[12] A. Tessarolo, S. Castellan, R. Menis, “Feasibility and performance analysis of a high
power drive based on four synchro-converters supplying a twelve-phase
synchronous motor”, IEEE Power Electronics Specialists Conference, PESC 2008, pp.
2352-2357.
250 Modeling of commutation transients in split-phase LCI-fed motors
[13] B. K. Bose, Modern Power Electronics and AC Drives, Prentice-Hall, 2001, pp. 283-284.
[14] A. Tessarolo, S. Castellan, “Analytical and circuital modeling of commutation
transients in phase-split synchronous motors supplied by multiple Load-
Commutated Inverters”, International Conference on Electric Machines, ICEM 2008, 6-
9 Sept. 2008, Vilamoura, Portugal, CD-rom paper n. 110.
[15] IEEE Std 115-1995. IEEE Guide: Test Procedures for Synchronous Machines, Part II,
Sections 11.13: Stationary or unbalanced test for determining X d′′ , X2, or X q′′ .
[16] A. Tessarolo, F. Luise, “An analytical-numeric method for stator end-coil leakage
inductance computation in multi-phase electric machines”, IEEE Industry Application
Society Annual Meeting, IAS 2008, 5-9 Oct. 2008, Edmonton, Canada, CD-rom paper
n. 79.
[17] D. Ban, D. Zarko, I. Mandic, “Tubogenerator end-winding leakage inductance
calculation using a 3-D analytical approach based on the solution of Neumann
integrals”, IEEE Trans. on Energy Conversion, vol. 20, Mar. 2005, pp. 98-105.
[18] IEC 34-4 Std., 1995, Rotating Electrical Machines, Part 4: Methods for Determining
Synchronous Machine Quantities from Tests.
Stator harmonic currents in VSI-fed split-phase synchronous motors 251
16.1 Introduction
Multi-phase machines are of increasing importance in today’s electric motor drives as they
allow for higher performance, reliability and power rating levels than achievable with
traditional three-phase configurations [1].
Out of the various multi-phase arrangements that can be adopted in stator winding design,
an important role is played by the so called asymmetrical multi-phase or split-phase
configurations. They result from grouping stator coils into multiple (N) three-phase
windings, displaced by 60/N electrical degrees apart, each suitable for being separately
supplied by a three-phase inverter [1], [11].
When Voltage-Source Inverters (VSIs) are employed to supply the stator windings, all
inverters should ideally output three-phase voltage systems of identical amplitude shifted
by 60/N electrical degrees. Each time this ideal condition is violated, circulation currents
appear in stator phases, mainly limited by phase leakage inductances [23], [26], [22].
This phenomenon has been widely studied in low-voltage Dual Stator Induction Machines
(DSIMs) where the main source of voltage imbalance is given by PWM inverter switching,
causing high-frequency circulation current harmonics [22], [3].
On the other hand, in modern medium-voltage motor drives, the use of VSIs with multi-
level topology in combination with synchronous machines is becoming important to
achieve high power and performance levels [1]. The Chapter illustrates how, in this case,
other possible sources of circulation currents can gain importance, due to the internal
motor structure. In fact, in wound-rotor synchronous machines, rotor conductors are not
uniformly distributed as in squirrel-cage induction motors and the air-gap permeance is
not uniform due to the different rotor reluctance along d and q axes. Both effects,
combined with possible magnetic saturation, may cause the air-gap field to significantly
deviate from its ideal sinusoidal waveform and some space harmonics (of 3rd, 5th and 7th
order above all) to appear in it. As a consequence, internally-generated low-frequency
phase circulation currents appear, which cumulate with possible PWM-related high-
frequency harmonics, already known from split-phase induction motor experience [22],
[3].
Fig. 1. Overall drive layout (a); motor phase arrangement (b) with displacement angles represented in electrical
degrees.
Each of the four VSI’s (rated 15 MVA, 7200 V, 100 Hz) has a cascaded topology with
several series-connected IGBT H-bridge cells per phase, [1]. Thanks to the PWM strategy
adopted and due to the high number of cascaded cells, the output phase voltage
approximates a sinusoidal waveform very closely, as confirmed by measurements (see
Section 16.5).
Finally, in order to limit the effects of stator circulation currents which will be investigated
next, external reactors (of 0.215 mH each) are added on all phases between motor and
inverter terminals.
Fig. 2. Original phase arrangement (a) and equivalent scheme with sequentially-distributed phases (b).
phase [7] and nine-phase [10] winding arrangements, as well as to asymmetrical or split-
phase induction motors with N stator three-phase sets displaced by 60/N electrical
degrees apart (N=2 in [22], [3]; N=3 in [11]). The main common purpose of these VFD
applications consists of controlling the high-frequency current harmonics induced in
stator phases as an effect of the inverter PWM switching.
In this Section, the VSD approach is adopted to model a VSI-fed synchronous motor with
an asymmetrical 12-phase winding arrangement (composed of N=4 three-phase sets). The
aim of the study is to predict the current harmonics not injected by the inverters (which
output almost sinusoidal voltages), but generated inside the machine due to its air-gap
field distortion. For this purpose, it is straightforward that a lumped parameter approach
is not adequate to fully catch the details of the motor electromagnetic structure (including
saturation, permeance harmonics and winding distribution effects). These details will be
properly taken into account, instead, by resorting to a sequence of FE magneto-static
analyses, as discussed in Section 16.3.3.
where: v is the vector of inverter output phase voltages; rs is the resistance of a motor
phase; lext is the external reactor inductance; ψtot is the vector of total flux linkages of
motor phases.
ψ tot = Lσ i + ψ (16.2)
the former accounts for stator leakage flux by means of the leakage inductance matrix Lσ,
which can be assumed constant (i.e. not depending on rotor position and magnetic
saturation, [7]); the latter accounts for the machine air-gap flux, caused by both stator and
254 Stator harmonic currents in VSI-fed split-phase synchronous motors
rotor MMFs, and includes the effects of magnetic saturation and air-gap permeance
harmonics.
d d
v = rs i + lext dt
i + Lσ dt
i + dtd ψ (16.3)
The structure of leakage inductance matrix Lσ obviously depends on the order in which
phase quantities are arranged to build up vector variables v, i, ψ. A convenient structure
for VSD purposes, in particular, is obtained by thinking of stator phases not grouped into
four three-phase sets (Fig. 16-2a) but sequentially arranged over a pole span and
numbered with sequential indices (1, 2, …, 12), as illustrated in Fig. 16-2b. The current
vector i will then be:
1 0 0 0 0 0 0 0 0 0 0 0
i1 0 0 0 1 0 0 0 0 0 0 0 0 i a1
i2 0 0 0 0 0 0 1 0 0 0 0 0 i b1
0 0 0 0 0 0 0 0 0 1 0 0 i c 1
0 0 −1 0 0 0 0 0 0 0 0 0 i a2
M 0
0 0 0 0 −1 0 0 0 0 0 0 i b2
i= = (16.4)
M 0 0 0 0 0 0 0 0 −1 0 0 0 i c 2
0
0 0 0 0 0 0 0 0 0 0 − 1 M
0 1 0 0 0 0 0 0 0 0 0 0 i a 4
i 11
0 0 0 0 1 0 0 0 0 0 0 0 i b4
i
12 0 0 0 0 0 0 0 1 0 0 0 0 i c 4
0 0 0 0 0 0 0 0 0 0 1 0
where ia1, ib1, ic1, ia2, ib2, ic2, etc. are the physical currents flowing in phases a1, b1, c1, a2, b2,
c2, etc. The same obviously apply to vector variables v and ψ.
l0 l1 L l10 l11
l1 l0 L l9 l10
Lσ = M M O M M (16.5)
l10 l9 L l0 l1
l11 l10 L l1 l0
where l0 is the self leakage inductance of a stator phase, while lk indicates the mutual
leakage inductance between two phases displaced by k × 15 electrical degrees, for
k=1,..,11. The values of lk resulting from measurements on the actual machine are reported
in Appendix A. As can be expected from winding geometry, measurements confirm that
l1=−l11, l2=−l10, l3=−l9, etc, i.e. in general
ln = −l12 − n (16.6)
for any n=1,..11. In virtue of (16-6), the stator inductance matrix (16-5) can be reduced to
a diagonal form L̂σ as follows:
Stator harmonic currents in VSI-fed split-phase synchronous motors 255
lˆ1 0 0 0 L 0 0
0 lˆ1 0 0 L 0 0
ˆ
0 0 l3 0 L 0 0
Lˆ σ = 0 0 0 ˆl
3 L 0 0 = T Lσ T
t
(16.7)
M M M M O M M
0 0 0 0 L lˆ11 0
0 0 0 0 L 0 ˆl
11
where superscript “t” denotes transposition and T is the constant decoupling matrix given
below, [4], [5]:
If model variables and parameters transformed through T are marked with a circumflex
accent “∧”, the stator voltage equation (16-3) becomes:
vˆ = rs iˆ + lext d
dt
iˆ + Lˆ σ d
dt
iˆ + dtd ψ
ˆ (16.9)
where
(
iˆ = iˆd 1 iˆq1 iˆd 3 iˆq3 L iˆd 11
t
)
iˆd 11 = T (i1 L i12 )
t
vˆ dn = rs iˆdn + lext d ˆ
dt i dn + lˆn dtd iˆdn + dtd ψˆ dn (16.11)
vˆ qn = rs iˆq n + lext d ˆ
dt i qn + lˆn dtd iˆqn + dtd ψˆ qn (16.12)
with n=1,3,…,11.
256 Stator harmonic currents in VSI-fed split-phase synchronous motors
0
[
I24q±n cos (24q ± n)ωt + φ24q±n ]
M [
I24q±n cos (24q ± n)(ωt − α ) + φ24q±n ]
0 I [
cos (24q ± n)(ωt − 2α ) + φ24q±n ]
24q±n
ˆ
idn
ˆ = T M
(16.13)
iqn
0
M [
I 24q±n cos (24q ± n)(ωt − 10α ) + φ24q±n ]
0
I [
24q±n cos (24q ± n)(ωt − 11α ) + φ24q±n ]
with
[
iˆdn = 6 I24q±n cos (24q ± n)ωt + φ24q±n ] (16.14)
[
iˆqn = 6I24q±n sin (24q ± n)ωt + φ24q±n ] (16.15)
With complex notation, the (iˆdn iˆqn ) space vector can be represented as:
t
where j denotes the imaginary unit and I 24q ± n = I 24q ± n exp( jφ24q ± n ) . Definition (16-16) can
be naturally extended to voltage and flux linkage variables as well, leading to the below
complex expression of (16-11)-(16-12):
[ ( )]
V24q±n = rs + jω (24q ± n) lext + lˆn I24q±n + jω (24q ± n)Ψ 24q±n (16.17)
for any q=0,1,2,… and any n=1,3,…,9,11; finally, (16-17) can be put in the form:
Vh = Z hIh + E h , (16.18)
with h=24q±n, having introduced motor harmonic impedances Z h and internal EMF
harmonics Eh due to internal air-gap flux as follows:
( )
Z h = Z 24q ± n = rs + jω (24q ± n ) lext + lˆn , (16.19)
Equation (16-18) relates voltage, current and internal EMF phasors for any odd harmonic
order h. It thus enables to compute current harmonic amplitudes Ih based on:
Obviously this holds under the usually-verified hypothesis that the above spectra include
only odd-order harmonics, [7].
In the case of the machine under study, parameters rs, lext and lˆn are reported in Appendix
A. As concerns the supply voltage spectrum, it will be assumed that the inverter output
voltage is perfectly sinusoidal (i.e. Vh=0 for any h different from 1) since the focus of the
Chapter is on the effects of internally-generated low-order harmonics: the assumption is
anyway consistent with measurements (Fig. 16-11a) apart from high-frequency voltage
harmonics due to inverter switching, which are shown to cause negligible current
circulation effects.
The most delicate point is determining internal EMF harmonics (Εh) due to air-gap field
distribution in full-load conditions. For this purpose, some 2D magneto-static FE
simulations can be used as a source of information, as explained in the next Section.
Fig. 3. Comparison of open-circuit saturation curves from measurement and form FE analysis. Abscissa is in per unit of the
field current that produces the rated flux at no-load.
258 Stator harmonic currents in VSI-fed split-phase synchronous motors
where V1 = 7200 × 2 / 3 V , I1= 917 × 2 A , ω = 2π 100 rad/s and rs, lˆ1 as per Appendix A.
The solved FE model yields the flux density distribution in the air gap at time t. If B
indicates the normal flux density component across the mean air-gap circumference γ (Fig.
16-5), then the air-gap flux linkage of a stator coil is computed as:
P2
ψ coil (t ) = N s L∫ B(t )dx (16.22)
P1
where Ns is the number of turns per coil, L is the core length, x is the spatial coordinate
along γ and P1, P2 are the intersections between γ and the segments joining the coil sides
with the center of the machine. The total air-gap flux linkage of a stator phase is simply
obtained by summing the contributions, computed through (16-22), of all its series-
connected coils.
By Fourier analysis of flux linkage function versus time, the sought EMF harmonics
Eh = jωhΨ h are finally obtained with the amplitudes shown in Fig. 16-7. It can be seen that
Eh are very small in percentage of the fundamental E1; nevertheless, they are capable of
driving non-negligible currents due to the very low harmonic impedances Zh which appear
in (16-18).
It is important to notice that such phase circulation currents cannot be included in the FE
simulations described above as they are the unknowns of the problem. Nevertheless, this
leads to theoretically null (and practically negligible) errors as far as the air-gap flux
Fig. 4. Motor FE analysis at no-load (a) and full load (b). Magnetic field produced by a 5th harmonic (c) and a 7th
harmonic in stator phase currents.
Stator harmonic currents in VSI-fed split-phase synchronous motors 259
The computed Eh harmonics originate from both magnetic saturation and rotor anisotropy
(i.e. the slightly non-uniform air-gap permeance and non-sinusoidal current distribution
which characterize the rotor of a synchronous machine). The two contributions can be
Fig. 6. Air-gap flux linked by the three phases (a, b, c) of stator winding 1. Each point is obtained from a 2D
magneto-static FE analysis.
260 Stator harmonic currents in VSI-fed split-phase synchronous motors
Fig. 7. Internal EMF harmonic En due to air-gap flux in percentage of the fundamental E1,
The evaluation of (16-23), together with the application of the phasor diagram of Fig 16-
8a for the fundamental components (h=1) at unity power factor, leads to the numerical
results reported in Table I.
Phasor harmonic currents I h can be finally brought back to the time domain as follows:
ik (t ) = Re
∑I h
h exp[ jhωt ]
(16.24)
Stator harmonic currents in VSI-fed split-phase synchronous motors 261
Fig. 8. Phasor diagrams for the fundamental (a) and higher order harmonic quantities (b) at steady-state.
for one stator phase; the other phase current waveforms are simply shifted by integer
multiples of the phase progression α=15 electrical degrees. The resulting diagrams for one
of the four motor windings are shown in Fig. 16-11c; they are in good accordance with
current waveforms recorded during motor testing in rated steady-state conditions (Fig.
16-11b).
TABLE II
NUMERIC COMPUTED VALUES FOR MOTOR HARMONIC QUANTITIES IN STEADY-STATE RATED CONDITIONS
Fig. 9. External circuits for time-stepping FE simulation. Each “phase” circuit element is linked to the
corresponding coils of motor FE model.
stepping FE analysis. For the above purpose, the motor 2D geometrical model (Fig. 16-4) is
imported in the Ansoft Maxwell environment and interfaced with the external circuit
elements that represent phase resistances, reactors and inverters (Fig. 16-9). The latter
are intentionally modeled as perfectly sinusoidal voltage sources since the aim of the
simulation is to prove that harmonic circulation currents are not excited by external
voltages but by the EMFs that arise inside the machine due to rotor geometry.
The difference between magneto-static FE simulations used in Section 16.3 and the time-
stepping one being described now is that in the former a fixed current is imposed in each
phase, while in the latter inverter output voltages vak, vbk, vck, for k=1, 2, 3, 4, are prescribed
as simulation inputs according to the functions below:
Furthermore, the rotor of the model is assigned a total inertia including that of coupled
rotating equipment and a ramp of resistant torque is applied to it up to the rated value of
143 kNm; a DC component of the field current such that the rated flux is obtained in the
air-gap is also imposed.
The FE analysis output is the evolution of all machine quantities (air-gap torque, speed,
phase currents and total field currents including possible induced ripples) as functions of
time. After the synchronization transient, the steady-state is reached and the phase
current waveforms given in Fig. 16-11d are obtained (for a motor winding), which well
Stator harmonic currents in VSI-fed split-phase synchronous motors 263
match experimental results (Fig. 16-11b). This accordance confirms that the observed
circulating harmonics are generated inside the machine due to its geometric structure,
since no inverter switching nor control features are included in the time-stepping FE
simulation.
The system test is carried out in the back-to-back arrangement sketched in Fig. 16-10, by
mechanically coupling two identical machines, of which one is operated as a motor and the
other as a generator. The four converters connected to the generator machine work in
Fig. 11. Supply voltage (a) and current (b) waveforms recorded on a motor winding during the system testing,
over one period interval (10 ms); voltage and current amplitudes are 7200 V rms and 900 A rms respectively.
264 Stator harmonic currents in VSI-fed split-phase synchronous motors
Fig. 12. Current waveforms resulting from motor VSD analysis (a) and time-stepping FE simulation (b); relevant
current harmonic amplitudes in percent of the fundamental (c).
regenerative mode (through their Active Front Ends) and feed the four converters
connected to the machine acting as a motor. This allows for a system test to be performed
with rated motor power P flowing through the shaft, while only the power corresponding
to total tested system losses Ptest is drawn from the mains.
Phase currents and line-to-line voltages are measured at all the twelve stator terminals of
the motoring machine during steady-state operation in rated balanced conditions by
digital oscilloscopes, showing identical amplitudes and waveforms. The waveforms
recorded on one of the four motor windings are shown in Fig. 16-11a-b, to be compared
with the current diagrams obtained by analysis and FE simulation of the machine in the
same operating conditions (Fig. 16-12a, 16-12b). It can be seen that the circulation
harmonics which affect phase currents can be predicted by both analysis and simulation
with satisfactory accuracy.
As a general remark resulting from the system test campaign, the circulation currents in
issue, although apparent from oscilloscope recordings, are shown not to harmfully impact
on the system performance. In fact, both temperature rise and torque ripple requirements
are met with adequate margins. In particular, the torque pulsation are assessed by means
of torque meters based on shaft-mounted strain gauges and reveal a ripple amplitude
lower than 1% peak-to-peak. This confirms the theoretical prediction that all the
measured current harmonics (i.e. up to the 11th order) do not take part in the energy
conversion process nor pass the air-gap, their magnetic energy being all stored in stator
leakage flux (Fig. 16-4c-d).
Stator harmonic currents in VSI-fed split-phase synchronous motors 265
1) Load-Commutated Inverters (LCIs) are more and more frequently being replaced by
voltage-source PWM inverters for synchronous motor supply in order to achieve
better performance in terms of efficiency, torque ripple and power factor [1], [15].
2) The need for high power rating and reliability levels often forces the designer to
segment the overall power into multiple inverters, which in turn implies a split-phase
motor winding structure [1], [15].
The resulting drive configuration is of the type investigated in this Chapter, i.e.
characterized by multiple VSIs supplying a split-phase synchronous motor. The results
discussed in the Chapter show that, even though the VSI output voltage has a nearly ideal
multi-level waveform and the motor has a uniform air-gap, non-negligible issues are still
to be expected due to internally-generated low-frequency circulation currents excited by
air-gap space harmonics.
These issues need to be properly taken into account in the drive system design stage so as
to adopt the adequate countermeasures (such as the inclusion of properly-sized input
reactors between motor and inverter terminals) and avoid system performance
degradation. In any case, the performance degradation to be expected mainly concerns
possible overheating caused by additional copper losses in motor and inverter phases.
Conversely, no detrimental effects in terms of torque pulsations and rotor circuit extra-
losses are to be expected, as proved by the reported system tests (Section 16.5), thanks to
the mutual cancellation effects among the air-gap fields produced by stator circulation
currents.
16.7 Conclusion
In this Chapter the performance of synchronous machines with multiple three-phase
stator windings, independently supplied by voltage source inverters (VSIs), is investigated
with respect to the occurrence of circulation currents in stator phases during normal
steady-state operation. The problem is well known in case of VSI-fed dual stator induction
motors, where it depends on high-frequency switching harmonics in inverter output
voltages. This Chapter illustrates how, in case of synchronous machines used as electric
motors, a further and possibly major source for low-frequency circulation currents arises
due to space harmonics in the air-gap field. These are explained in the Chapter as a
consequence of the synchronous motor non-uniform air-gap permeance and rotor field
circuit distribution. The phenomenon is investigated focusing on a quadruple three-phase
45-MW round-rotor synchronous motor supplied by four multilevel VSIs. To investigate
the origin of its phase circulation currents (mainly consisting of 5th and 7th order
harmonics), these are computed in two independent ways (analytically and with time-
stepping FE analysis) by intentionally disregarding inverter-related effects. The
accordance of the results obtained in both ways with experimental measurements
confirms that the observed circulation currents originate inside the machine.
266 Stator harmonic currents in VSI-fed split-phase synchronous motors
The results presented in the Chapter are finally contextualized in the field of modern high-
power split-phase synchronous motor drives, where the choice of replacing traditional
Load-Commutated Inverters with multi-level PWM ones (even with almost sinusoidal
voltage waveform) should take into account the investigated non-negligible low-frequency
circulation current issues.
Appendix A
The ratings of the quadruple three-phase synchronous motor investigated in this Chapter
are listed next:
Power: 45 MW
Number of poles: 4
Concerning motor parameters, the model (16-3) adopted in Section 16.3 for the quadruple
three-phase synchronous machine requires the knowledge of stator phase resistance rs
and stator phase inductances l0..l11 which constitute matrix Lσ as per (16-5). In fact, once
l0..l11 are known, parameters lˆn for h=1,3,..,11 can be computed through (16-7) and, finally,
leakage harmonic impedances Z h , for any odd harmonic order h, can be obtained from
(16-19).
The measured values of stator phase inductances l0..l11, along with the parameters lˆn that
result from (16-7), are reported in Table II.
TABLE II
Fig. 13. Solved FE model with rotor removed and one phase supplied, used to compute lhbore .
Self and mutual phase leakage inductances lh were determined based on impedance
measurements with the rotor removed. The test method adopted, which is detailed in [23],
[15], mainly follows the guidelines given in [10], [5] for leakage inductance determination
in three-phase machines, by extending it to the case of more than three stator phases. The
procedure is shortly recalled next for the sake of commodity.
In brief, to determine the parameter lh a stator phase is supplied with current I, voltage V
and frequency f=50 Hz after rotor removal; the voltage E thus induced across the stator
phase displaced by hα electrical degrees is measured; for h=0, E coincides with the voltage
on the supplied phase once the resistive drop has been removed as follows:
E = V 2 − rs I 2 (A1)
A total inductance l hrr with the rotor removed (self inductance if h=0, mutual otherwise)
is then determined as:
of which the former coincides with the sought leakage inductance, the latter is due to the
flux which passes through the stator bore, i.e. in the region usually occupied by the rotor.
As proposed in [5], the term lhbore can be accurately estimated with a 2D FE analysis of the
motor cross-section after rotor removal (Fig. 16-13).
268 Stator harmonic currents in VSI-fed split-phase synchronous motors
The axial depth L of the model to be used for this purpose (including the effects of radial
ventilation ducts which may slightly reduce the useful core length compared to its
geometric value) can be thoroughly calibrated by means of a search coil placed inside the
stator bore as recommended in [10].
Table III provides the values of l hrr obtained directly from measurements through (A2) and
the values of lhbore obtained with the calibrated FE analysis. The values of leakage
inductances lh given in Table II are finally determined as:
TABLE II
16.8 References
[1] E. Levi, R. Bojoi, F. Profumo, H.A. Tolyat, S. Williamson, “Multiphase induction
motor drives – a technology status review”, Electric Power Application, IET,
2007, July 2007, vol. 1, pp. 489-516.
[2] R.H. Nelson, P.C. Krause, “Induction machine analysis for arbitrary
displacement between multiple winding sets”, IEEE Trans. on Power
Apparatus and Systems, May./June 1974 vol. PAS-94, pp. 841-848.
[3] E.A. Klingshirn, “High phase order induction motors−Part I−Description and
theoretical considerations”, IEEE Trans. on Power Apparatus and Systems, Jan.
1983, vol. PAS-102, pp. 47-53.
[4] E.A. Klingshirn, “High phase order induction motors−Part I−Experimental
results”, IEEE Trans. on Power Apparatus and Systems, Jan. 1983, vol. PAS-
102, pp. 54-59.
[5] D. Hadiouche, H. Razik, A. Rezzoug, “On the modeling and design of dual-
stator windings to minimize circulating harmonic currents for VSI-fed AC
machines”, IEEE Trans. on Industry Applications, vol. 40, Mar./Apr. 2004, pp.
506-515.
[6] Y. Zhao, T.A. Lipo, “Space vector PWM control of dual three-phase induction
machine using vector space decomposition”, IEEE Trans. on Industry
Application, Sept.-Oct. 1995, vol. 31, pp. 1100-1109.
[7] J. Rodriguez, S. Bernet, B. Wu, J.O. Pontt, S. Kouro, “Multilevel voltage-source-
converter topologies for industrial medium-voltage drives”, IEEE Trans. on
Industrial Electronics, vol. 54, no. 6, Dec. 2007, pp. 2930-2944.
[8] J. Figueroa, J. Cros, P. Viarouge, “Generalized Transformations for Polyphase
Phase-Modulated Motors”, IEEE Transactions On Energy Conversion, vol. 21,
June 2006, pp. 332-341.
Stator harmonic currents in VSI-fed split-phase synchronous motors 269
[9] L.A. Pereira, C. C. Scharlau, L.F.A. Pereira, J.F. Haffner, “General model of a
five-phase induction machine allowing for harmonics in the air-gap”, IEEE
Trans. on Energy Conversion, vol. 21, issue 4, Dec. 2006, pp. 891-899.
[10] D. Dujic, M. Jones, E. Levi, “Space-Vector PWM for nine-phase VSI with
sinusoidal output voltage generation: analysis and implementation”, 33rd
Annual Conference of the IEEE Industrial Electronic Society, 2007, IECON 2007,
Taipei, Taiwan, Japan, 5-8 Nov. 2007, pp. 1524-1529.
[11] M.J. Duran, E. Levi, M. Jones, “Independent vector control of asymmetrical
nine-phase machines by means of series connection”, 2005 IEEE International
Conference on Electric Machines and Drives, S. Antonio, Texas, USA, 15-18
May, 2005, pp. 167-173.
[12] A. Tessarolo, “On the modeling of poly-phase electric machines through
vector-space decomposition: theoretical considerations”, International
Conference on Power Engineering, Energy and Electrical Drives, POWERENG
2009, pp. 519-523.
[13] A. Tessarolo, “On the modeling of poly-phase electric machines through
vector-space decomposition: numeric application cases”, International
Conference on Power Engineering, Energy and Electrical Drives, POWERENG
2009, pp. 524-528.
[14] A. Tessarolo, F. Luise, “An analytical-numeric method for stator end-coil
leakage inductance computation in multi-phase electric machines”, IEEE
Industry Application Society Annual Meeting, IAS 2008, 5-9 Oct. 2008,
Edmonton, Canada, CD-rom paper n. 79.
[15] A. Tessarolo, S. Castellan, R. Menis, G. Ferrari, “On the modeling of
commutation transients in split-phase synchronous motors supplied by
multiple load-commutated inverters”, IEEE Transactions on Industrial
Electronics, accepted paper, in press. Online preview available at:
http://ieeexplore.ieee.org/stamp/stamp.jsp?tp=&arnumber=5229192&isnu
mber=4387790.
[16] IEC 34-4 Std., 1995, Rotating Electrical Machines, Part 4: Methods for
Determining Synchronous Machine Quantities from Tests.
[17] D. Ban, D. Zarko, I. Mandic, “Turbogenerator end-winding leakage inductance
calculation using a 3-D analytical approach based on the solution of
Newmann integrals”, IEEE Transactions on Energy Conversion, vol. 20, Mar.
2005, pp. 90-105.
270
271
Acknowledgements
Roberto Menis, Giorgio Sulligoi, Simone Castellan, Alfredo Contin, Mario Mezzarobba,
Michele Degano, Luca Spangaro, Mauro Favot
Fabio Luise, Cristina Bassi, Davide Giulivo, Piero Raffin, Riccardo Macuglia, Gianfranco
Zocco, Antonio Calonico, Antonio Odorico, Carlo Tonello, Giancarlo Ferrari, Antonella
Scaglia
A special thank is due to Prof. Roberto Menis, who has always made it possible for
me to work in freedom and serenity.