(Power Electronics and Applications Series) Bianchi, Nicola - Electrical Machine Analysis Using Finite Elements-CRC Press (2005)
(Power Electronics and Applications Series) Bianchi, Nicola - Electrical Machine Analysis Using Finite Elements-CRC Press (2005)
MACHINE ANALYSIS
USING FINITE ELEMENTS
POWER ELECTRONICS AND
A P P L I C AT I O N S S E R I E S
Muhammad H. Rashid, Series Editor
University of West Florida
PUBLISHED TITLES
Complex Behavior of Switching Power Converters
Chi Kong Tse
NICOLA BIANCHI
A CRC title, part of the Taylor & Francis imprint, a member of the
Taylor & Francis Group, the academic division of T&F Informa plc.
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2005 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business
This book contains information obtained from authentic and highly regarded sources. Reasonable
efforts have been made to publish reliable data and information, but the author and publisher cannot
assume responsibility for the validity of all materials or the consequences of their use. The authors and
publishers have attempted to trace the copyright holders of all material reproduced in this publication
and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any
future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information stor-
age or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copy-
right.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222
Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that pro-
vides licenses and registration for a variety of users. For organizations that have been granted a pho-
tocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Preface
For a long time the course Design Methodology of Electrical Machines has been
a teaching course for electrical engineering aimed at electrical applications.
This is due both to the subject that is dealt with and to the organization of
the course by Prof. Luciano Merigliano, based on many years’ experience
with electromechanical design.
The laboratory activities of the course started in 1996, directed by myself
along with Prof. Silverio Bolognani, who was the teacher in charge in those
years, and to whom the credit for course innovation has to be given.
This book reports some of the laboratory exercitations of the course Design
Methodology of Electrical Machines. The aim is to examine the key concepts of
the finite element method, as well as to focus attention on the applications
of such a method to electrical machine analysis.
Not long after the course began, the students expressed the need to have
a text illustrating the analysis of the electric and magnetic devices by means
of the finite element method. In addition, the electromechanical industries,
which require accurate analyses, also had need for such a text. This was a
result of the popularity of software programs for the finite element analysis
of electromagnetic problems, some of which are even available on the Web.
Among them, some well-known packages are ANSYS, CADEMA, FEMM,
FLUX, MAXWELL, MEGA, QFIELD, and VECTOR-FIELD.
A first draft of this text was published in Italian in 2001, for the electrical
engineering students at the University of Padova. It was also adopted at
other Italian universities and by some industries. The present text is essen-
tially a translation of the Italian version of the same text, with a few revisions.
The book is organized to meet the goal described above. Several examples
of finite element analysis of the electrical machines are pointed out. Some
repetitions in the various chapters are intentional, for the sake of a more
fluent treatment. Some theoretical concepts, numerical techniques, and sim-
ulation methods are reported gradually in the various chapters, in an effort
to make for an easier read and to avoid many boring introductory chapters.
To the aim of facilitating understanding, some electrotechnical principles
and some electrical machines concepts are sometimes reviewed. Moreover,
simple numerical algorithms, written using MATLAB or BASIC software,
are also reported. They are examples of processing the field analysis results
and of the automatic computation procedures.
Let me thank my two generous teachers, Prof. L. Merigliano and Prof. S.
Bolognani, for the unbroken respect and the constant teaching, and not only
in the technical field. Let me also thank my friends and colleagues Prof. F.
Dughiero and Dr. A. Tortella, whose help was essential for the fulfillment of
this text. Last, but not least, I would like to express my thanks to my students,
whose lively curiosity and diligent work contributed significantly to the
improvement of the text.
Nicola Bianchi
The Author
“…a man traveling to a far country called his own servants and delivered
his goods to them. And to one he gave five talents, to another two, and
to another one, to each according to his own ability…”
Greek symbols
αc rad electric slot angle
βr rad electric shortened angle
βr βs rad rotor and stator width angle
δ rad load angle
∆i A current variation
∆ϑ red mechanical angle variation
ε F/m electric permittivity
φ, ϕ, ψ — unknown potential to be determined
φ* — function that approaches the unknown
potential
Φ Wb magnetic flux
Φo Wb magnetic flux at no-load
ϕ rad power angle
ϕ rad initial voltage phase
Φj — j-th coefficient to approach the potential φ
[φ] — matrix vector of the unknown coefficients Φi
Γ — boundary of the domain (line or surface)
γFe kg/m3 specific iron weight
Λ, λ Wb, Vs flux linkage
Λd Λq Vs d-axis and q-axis flux linkages
Λrms Vs RMS flux linkage
Λj Vs flux linkage of the j-th phase (j = a, b, c)
λ1 λ2 Vs primary and secondary flux linkages
λ1σ λ2σ Vs primary and secondary leakage flux linkages
λpm Vs flux linkage due to the permanent magnet
λr λs rad rotor and stator pole pitch angle
µ H/m magnetic permeability
µr — relative magnetic permeability
h — efficiency
νj — j-th interpolanting function
ρ C/m3 volume density of electric charge
ρs C/m2 surface density of electric charge
ρ Ωm electric resistivity, ρ = σ −1
ϑ rad angular coordinate (electric angle)
ϑm rad angular coordinate (mechanic angle)
ϑoff rad switch-off angle
ϑon rad switch-on angle
ϑr rad angular coordinate referred to the rotor polar
axis
ϑs rad angular coordinate referred to the stator
a-phase axis
ξ — saliency ratio, ξ = Lq/Ld
ω rad/s electric frequency
ωm rad/s mechanic angular speed
ωr rad/s rotor electric angular speed ωr = 2πfr
σ S/m electric conductivity
τ m3 volume
τCu m3 conductive material volume
τm Nm instantaneous electromagnetic torque
τL Nm load torque
Ψ A magnetic voltage
Further Symbols
Im imaginary part
Re real part
~ conjugate complex
1
Outline of Electromagnetic Fields
where nAB is the unity vector normal to the plane defined by A and B. The
vector (1.2) results normal to both vector A and vector B, and its direction
1
2 Electrical Machine Analysis Using Finite Elements
A
A⫻B A P2
A n
B t t
S
A P l P
l
γ
P1
FIGURE 1.1
Operations among the vectors.
A ⋅ B = A x B x + A y B y + A zB z (1.3)
ux uy uz
A × B = Ax Ay Az
Bx By Bz (1.4)
( ) ( )
= A y B z − A zB y u x + A zB x − A xB z u y + A xB y − A y B x u z( )
The following properties exist:
2
A·A = A A·(B + C) = A· B + A· C
A· B = B ·A A × (B + C) = A × B + A × C (1.5)
A × B = −B × A A × (B × C) = (A· C) B − (A· B) C
The loop integral of the vector field corresponds to the linear integral along
a closed curve, shown in Figure 1.1(c), which is
Outline of Electromagnetic Fields 3
Γ=
∫ A ⋅ t dl
l
(1.7)
The flux of the vector field A through the surface S, oriented by the normal
unity vector n in each point, as shown in Figure 1.1(d), is given by
Φ=
∫
S
A ⋅ n dS (1.8)
∂V
gradV = max u max (1.9)
∂l
Along any direction defined by the unity vector u, the scalar product of
gradV by u gives the derivative of V along the direction of u, which is
∂V
gradV·u = (1.10)
∂l u
(
grad Σ V = V2 − V1 n Σ ) (1.11)
divA = lim
∫ Sc
A· n dS
(1.12)
τ→0 τ
4 Electrical Machine Analysis Using Finite Elements
( )
div Σ A = A 2 − A1 ⋅ n Σ (1.13)
curlA = lim
∫ A ⋅ t dl
l
(1.14)
S→0 S
(
curl Σ A = − A 2 − A1 × n Σ) (1.15)
P2
Γ 12 =
∫ P1
gradV ⋅ t dl = V2 − V1 (1.18)
∫ gradV dτ = ∫ Vn dS
τ S
(1.19)
The divergence theorem (or Gauss’s theorem) states that the volume integral
of a divergence divA is equal to the integral of the vector A through the
closed surface that contains the volume, which is the flux of A going out
through the closed surface:
Φ=
∫ Sc
A ⋅ n dS =
∫ (divA) dτ
τ
(1.20)
The curl theorem (or the Stokes theorem) states that the integral through
the surface S of a vector curlA is equal to the linear integral of the vector A
along a closed line l bordering the surface S itself. The unity vector t tangen-
tial to the line l has to be chosen in accordance with the unity vector n normal
to the surface S (with the rule of the right screw). Finally, it is
∫ A ⋅ t dl = ∫ curlA ⋅ n dS
l S
(1.21)
∂V
∫ U div(h gradV) + h gradU ⋅ gradV dτ = ∫
τ Sc
hU
∂n
dS (1.22)
∂V ∂U
∫ U div(a gradV) − Vdiv(a gradU) dτ = ∫
τ Sc
AU
δn
−V
δn
dS (1.23)
6 Electrical Machine Analysis Using Finite Elements
( )
grad VU = V(gradU) + U(gradV) (1.26)
( )
grad kU + hV = k gradU + h gradV (1.27)
∂V
gradV U =( ) ∂U
gradU (1.28)
Divergence:
( )
div UA = U(divA) + gradU ⋅ A (1.29)
( ) ( )
div kA + hB = k div A + h div B ( ) (1.30)
( )
div A × B = − A ⋅ (curlB) + (curlA) ⋅ B
(1.31)
div ( curlA ) = 0
Curl:
( )
curl UA = U ⋅ (curlA) + (gradU) × A (1.32)
( ) ( )
curl kA + hB = k curl A + h curl B ( ) (1.33)
Laplacian:
( )
∇ 2 UV = U ⋅ ∇ 2 V + V∇ 2 U + 2 gradU ⋅ gradV (1.35)
( )
∇ 2 kU + hV = k ∇ 2 U + h ∇ 2 V (1.36)
∂V ∂V ∂V
gradV = ux + uy + uz (1.37)
∂x ∂y ∂z
∂A x ∂A y ∂A z
divA = + + (1.38)
∂x ∂y ∂z
ux uy uz
∂ ∂ ∂
curlA =
∂x ∂y ∂z
Ax Ay Az (1.39)
∂A z ∂A y ∂A x ∂A z ∂A y ∂A x
= − ux + − uy + − uz
∂y ∂z ∂z ∂x
∂x ∂y
y y y
uy
uϑ uϑ uϕ
P P P
ux ϑ ux ux
ϑ
uz uz
ϕ
O x O r O r
z z z
FIGURE 1.2
Systems of orthogonal coordinates: Cartesian (a), cylindrical (b), and spherical (c) coordinates.
8 Electrical Machine Analysis Using Finite Elements
∂2 U ∂2 U ∂2 U
∇ 2 U = div gradU = + + (1.40)
∂x 2 ∂y 2 ∂z 2
( )
∇ 2 A = grad divA − curl curlA ( )
= ∇ 2 A x u x + ∇ 2 A y u y + ∇ 2 A zu z
∂2 A x ∂2 A x ∂2 A x ∂2 A y ∂2 A y ∂2 A y (1.41)
= + + 2 x
u + + + uy +
∂x ∂y ∂z ∂x ∂y 2 ∂z 2
2 2 2
∂2 A z ∂2 A z ∂2 A z
∂x 2 + ∂y 2 + ∂z 2 u z
∂V 1 ∂V ∂V
gradV = ur + uθ + uz (1.42)
∂r r ∂θ ∂z
divA =
(
1 ∂ rA r
+
)1 ∂A θ ∂A z
+ (1.43)
r ∂r r ∂θ ∂z
1 ∂A z ∂A θ
curlA = −
∂A r ∂A z
ur + −
1 ∂ rA θ
uθ + −
(∂A r )
u z (1.44)
r ∂θ ∂z ∂z ∂r r ∂r ∂θ
1 ∂ ∂V 1 ∂2 V ∂2 V
∇2V = r + +
r ∂r ∂r r 2 ∂θ2 ∂z 2
(1.45)
1 ∂V ∂2 V 1 ∂2 V ∂2 V
= + + +
r ∂rr ∂r 2 r 2 ∂θ2 ∂z 2
∂V 1 ∂V 1 ∂V
gradV = ur + uθ + uϕ (1.46)
∂r r ∂θ r sin θ ∂ϕ
Outline of Electromagnetic Fields 9
divA =
(
1 ∂ r Ar
2
+
) (
1 ∂ sin θ A θ
+
)
1 ∂A ϕ
(1.47)
r 2
∂r r sin θ ∂θ n θ ∂ϕ
r sin
curlA =
(
1 ∂ sin θ A ϕ
−
)
∂A θ
ur +
r sin θ ∂θ ∂ϕ
(1.48)
−
(
1 ∂A r ∂ sin θ rA ϕ ) (
1 ∂ rA θ
uθ +
)−
∂A r
uϕ
r sin θ ∂ϕ ∂r r ∂r ∂θ
1 ∂ 2 ∂V 1 ∂ ∂V 1 ∂2 V
∇2V = r + sin θ + (1.49)
r 2 ∂r ∂r r 2 sin θ ∂θ ∂θ r sin θ
( )
2
∂ϕ 2
is unique if the field is normal at the infinite (a field whose magnitude tends
to 1/r2 when the distance r from the origin tends to infinite). Thus, it is
divB = divBirr
B = Birr + Bsol with (1.50)
curlB = curlBsol
( )
∇ 2φ = div gradφ = 0 (1.51)
∇2A = 0 (1.52)
Then, the vector field B that is obtained as curl of a harmonic vector field
A, i.e., B = curlA, is solenoidal because div(curlA) = 0, and irrotational.
∂V ∂ ∂A ∂
grad = gradV div = divA
∂t ∂t ∂t ∂t
(1.53)
∂A ∂ ∂A ∂ 2
curl = curlA ∇2 = ∇A
∂t ∂t ∂t ∂t
= A e jα x
A x = A xM cos(ωt + α x ) ↔ A x xM
= A xr + jA xi (1.54)
= A xM cos α x + jA xM sin α x
Outline of Electromagnetic Fields 11
and analogously for the other components. The vector field A, whose com-
ponents are described by the symbolic function (1.54), is indicated as
=A
A u +A u +A
u (1.55)
x x y y z z
∆q
ρ = lim (1.56)
∆τ→0 ∆τ
that are measured in (C/m3) and generally variable with the time.
If the element on which the charge is distributed is characterized by a
dimension negligible with respect to the other two dimensions, it is possible
to define a surface density of charge ρs, measured in (C/m2), as
∆q
ρs = lim (1.57)
∆S→0 ∆S
∫ Sc
D ⋅ n dS = q (1.59)
The second equation is the well-known Gauss’s law. It indicates that the
flux of the vector D going out through the closed surface Sc, oriented by the
unity vector n normal to the surface and with external direction, is equal to
the free charge q, which is contained by the surface Sc itself.
J = ρ vρ (1.60)
whose reference positive direction is that of the positive charges. Its magni-
tude is measured in (A/m2). The vector J defines a vector field, called the
current field. Let S be an open surface, with normal unity vector n, then the
current intensity i measured in (A) is given by
i=
∫ J ⋅ n dS
S
(1.61)
∂ρ
divJ = − (1.62)
∂t
∆q
i out =
∫ Sc
J ⋅ n dS = −
∆t
(1.63)
∂D
Jtot = J + (1.64)
∂t
which is the sum of the current density vector J and the displacement current
density vector ∂D/∂t. By property (1.58) about D and the continuity Equation
(1.62), the current field Jtot is solenoidal.
Outline of Electromagnetic Fields 13
divB = 0 (1.65)
∫
Sc
B ⋅ n dS = 0 (1.66)
B = curlA (1.67)
This relationship defines the field A apart from a generic irrotational field.
The divergence of A can be defined in an arbitrary way; the positions that
are commonly adopted are as follows.
divA = 0 (1.68)
∂V
divA = −µε (1.69)
∂t
B = µH (1.70)
∂D
curlH = J + (1.71)
∂t
curlH = J (1.72)
In integral form, the Ampere’s law is expressed by equating the line inte-
gral of H along an oriented closed line l to the current intensity i flowing
through the surface enclosed by the line l itself, which is
∫ H ⋅ t dl = i
l
(1.73)
Q
Ψ PQ =
∫P
H ⋅ t dl (1.74)
The magnetic field strength H and the magnetic flux density B are wholly
defined by Equation (1.64) and Equation (1.71) together with the constitutive
equation [Equation (1.70)]. From them, the divergence of H is
H
divH = − gradµ (1.75)
µ
H = −gradΨ (1.77)
δFk
E k = lim (1.78)
δq→0 δq
E c = − gradV (1.79)
Also the specific induced electric force Ei , produced by the rate of change
of the magnetic flux density field B with the time, is a function of the points
of the space, so that it defines a vector field as well. The fundamental
property of the induced electric field is the definition of its curl, which is
∂B
curlE i = − (1.80)
∂t
TABLE 1.1
Classification of the Specific Electric Forces
conservative
Ec Coulomb
E i induced
} E electric field
electromagn.
Ek non-conservative E L Lorentz
E motional
m
non-electromagn. E n.e.
16 Electrical Machine Analysis Using Finite Elements
∂A
Ei = − (1.81)
∂t
∂A
E = E c + E i = − gradV − (1.82)
∂t
The electric field E is defined by its divergence and its curl, that are divE =
divEc and curlE = curlEi , respectively. In a dielectric medium with electric
permittivity ε (in general of tensorial nature), the electric field E is linked to
the electric displacement field D by means of the constitutive relationship
D = εE (1.83)
The Lorentz’s specific electric force EL acts on the electric charges moving
in a magnetic flux density field B at a velocity vq with respect to the adopted
reference system. It is
EL = vq × B (1.84)
The motional specific electric force Em acts on the electric charges posed
on a conductor moving at a velocity vm in a magnetic flux density field B. It is
Em = v m × B (1.85)
Both these two specific forces are not property of the points of the space,
but they depend on the velocity of the charges and of the conductor, with
respect to the reference system. Thus, they do not define a vector field.
However, by means of a change of the reference system it is possible to
consider the effects of the motional specific forces as induced specific forces.
An example will be shown in the study of the induction motor, in Chapter 13.
The specific electric force of nonelectromagnetic nature, Ene, can be of
chemical, piezoelectric, photovoltaic nature, and so on. They are all noncon-
servative. They do not define vector fields, but they must be considered as
specific forces external to the electromagnetic system.
Outline of Electromagnetic Fields 17
The total specific electric force Et is the sum of the field E and the Lorentz’s,
motional, and external specific forces, which is
Et = E + EL + Em + Ene (1.86)
J = σE t (1.87)
The nonconservative specific electric forces Eemf are rotational specific elec-
tric forces. They are called specific electromotive forces. Among them, Ei , EL,
and Em are of electromagnetic nature.
B
v AB =
∫ A
E ⋅ t dl (1.88)
∫ (E + v × B) ⋅ t dl
B
v AB = l (1.89)
A
Analogously, the electromotive force (EMF) eBA, between two points B (+)
e A (–), along a line l is given by the line integral along l from point A to
point B of the specific electromotive force Eemf. It is
B
eBA =
∫A
Eemf ⋅ t dl (1.90)
In analyzing electrical machines, the induced EMF and the motional EMF
are of particular interest. The induced EMF is caused by the time-variation
of the flux density B. The motional EMF is caused by the movement of the
line in a constant field B. The sum of the two EMFs, computed along a closed
18 Electrical Machine Analysis Using Finite Elements
dλ
∫ (E + E ) ⋅ t dl = − dt ∫ B ⋅ n dS = − dt
d
elc = i m (1.91)
lc S
where S is the surface enclosed by lc and λ is the flux of the vector B linked
by the line lc. The subdivision of the EMF elc in its two components, induced
and motional EMF, depends by the adopted reference system only.
divP = div(E × H)
= H ⋅ (curlE) − E ⋅ (curlH)
∂B ∂D (1.92)
= H⋅ − − E ⋅ J +
∂t ∂t
∂B ∂D
= −H ⋅ − E⋅ − E⋅J
∂t ∂t
p=
∫ divP dτ
τ
(1.93)
∂B ∂D
τ ∫
= − H⋅
∂t
+ E⋅
∂t
dτ −
∫ (E ⋅ J)ddτ
τ
∂B 1 ∂H ⋅ B
H⋅ =
∂t 2 ∂t
(1.94)
∂D 1 ∂E ⋅ D
E⋅ =
∂t 2 ∂t
Outline of Electromagnetic Fields 19
H
n
H
S n
dS
θ Hn
τ Ht
dS t
FIGURE 1.3
Computation of the magnetic force by means of the Maxwell’s stress tensor.
p=
∫ τ
divP dτ
=
∫ P ⋅ n dS
S
(1.95)
∂ H⋅B E ⋅D
=−
∂t ∫ τ
2 + 2 dτ − ∫τ
E ⋅ J dτ
µo 2
dF = − H ndS + µ o (H ⋅ ndS)H (1.96)
2
where n is the unity vector normal to the surface dS. By expressing the vector
H as the sum of its components tangential and normal to the surface dS, as
shown in Figure 1.3(b), as
20 Electrical Machine Analysis Using Finite Elements
H = Ht t + Hnn (1.97)
dF = −
µo
2
( ) (
H t 2 + H n 2 dS t + µ o H t t + H n n ⋅ n H t t + H n n dS ) ( )
(1.98)
( ) µ
(
= µ oH t H n dSt + o H t 2 − H n 2 dSn
2
)
It is possible to identify the two components of the force along the two
preferential directions, as
(
dFt = µ oH t H n dS )
(1.99)
µ
dFn = o H n 2 − H t 2 dS
2
( )
The amplitude of the force is computed as
dF = (µ H H dS )
o t n
2 µ
(
+ o H n 2 − H t 2 dS
2
)
≈
1
2
(
µ o H t 2 + H n 2 dS ) (1.100)
1
≈ µ oH 2dS
2
The angle α between the force vector and the unity vector n normal to the
infinitesimal surface dS is computed as
Ht Hn
tan α =
1
2
(
Hn 2 − Ht 2 )
Ht / Hn
=2 (1.101)
( )
2
1 − Ht / Hn
tan θ
=2 = tan(2θ)
1 − tan 2 θ
Outline of Electromagnetic Fields 21
TABLE 1.2
Particular Cases of the Computation of the Magnetic
Force by Means of the Maxwell’s Stress Tensor
F
H Ht = 0 dFt = 0
Hn ≠ 0 dFn ≠ 0 (>0)
Hn = Ht
dFt ≠ 0
Ht = H n ≠ 0
F dFn = 0
Ht t
H
Ht ≠ 0 dFt = 0
F
Hn = 0 dFn ≠ 0 (<0)
where tanθ = Ht/Hn, that expresses the angle between the vector H and the
unity vector n. One can notice that the angle α results two times the angle
θ, i.e., α = 2θ.
Particular cases are shown in Table 1.2.
where P is the point where the vector is considered, while t is the time when
the vector is considered. In other words, P indicates the space dependence,
while t indicates the time dependence.
22 Electrical Machine Analysis Using Finite Elements
∂D(P,t)
curlH(P,t) = J(P,t) + (1.102)
∂t
∂B(P,t)
curlE(P,t) = − (1.103)
∂t
divB(P,t) = 0 (1.104)
∂ρ(P,t)
divJ(P,t) = − (1.107)
∂t
In general the media are not isotropic; then the parameters depending on
the material are of tensorial nature:
magnetic permeability electric permittivity electric conductivity
µ x 0 0 εx 0 0 σx 0 0
µ= 0 µy 0 ε=0 εy 0 σ= 0 σy 0
0 µ z 0 ε z 0 σ z
0 0 0
(1.108)
In general µi, εi, and σi are not constant along the generic i-th direction. In
fact they can be a function of the position (nonhomogeneous media) and/or
a function of the magnetic field H and the electric field E (nonlinear media).
Outline of Electromagnetic Fields 23
1
curl curlA = J (1.109)
µ
∇ 2 A = −µJ (1.110)
If the current density field J is null in the considered domain, the problem
is described by the Laplace equation [Equation (1.52)], which is
∇2A = 0 (1.111)
When the magnetic field and the current density field are varying with
the time, they are mutually coupled. Let us suppose that such a variation is
sinusoidal with the time; then the field quantities can be expressed using
the symbolic notation, pointed out in Equation (1.54). A uniform medium is
considered, so that a constant electric conductivity σ and a constant magnetic
permeability µ are obtained. Finally, the time-derivative of the displacement
field D could be neglected in comparison with the other fields. Thus the
electric field E is the unique specific electric force. Hence the current density
J is obtained as the sum of the source current density Js and the induced
current density Ji, given by
∂A
J i = σE i = − σ
= − jωσA (1.112)
∂t
− jωσA
∇2A = −µJ (1.113)
s
24 Electrical Machine Analysis Using Finite Elements
References
1. J.A. Stratton, Electromagnetic Theory, McGraw-Hill, New York, 1941.
2. M. Abramowitz and I.A. Stegun, Handbook of Mathematical Functions, Dover
Publications, Inc., New York, 1965, pp. 374–376.
3. D.R. Corson and P. Lorrain, Introduction to Electromagnetic Field and Waves, W.H.
Freeman & Co., San Francisco–London, 1970.
4. J.D. Kraus and K.R. Carver, Electromagnetics, McGraw-Hill, New York, 1973.
5. G. Someda, Elementi di Elettrotecnica Generale, Patron Editore, Bologna, 1977 (in
Italian).
6. P. Lorrain and D.R. Corson, Electromagnetism. Principles and Applications, San
Francisco, 1978.
7. M.A. Plonus, Applied Electromagnetics, McGraw-Hill, New York, 1978.
8. R.E. Collin, Field Theory of Guided Waves, IEEE Press, New York, 1991.
9. D.K. Cheng, Fundamentals of Engineering Electromagnetics, Addison-Wesley,
Reading, MA, 1993.
10. M. Guarnieri and G. Malesani, Elettromagnetismo Stazionario e Quasi-Stazionario,
Edizioni Progetto, Padova, 1999 (in Italian).
11. J.M. Jin, Electromagnetic Analysis and Design in Magnetic Resonance Imaging, CRC
Press, Boca Raton, FL, 1998.
2
Basic Principles of Finite Element Methods
Some basic concepts of finite element method are dealt with in this chapter
and in the next. In this chapter, the differences of the finite element method
compared to the classical methods of field problem analysis are highlighted.
The mathematical notions of the method are presented. In the next chapter
the construction of the system of equations for the solution is investigated,
and the finite element method is applied to some two-dimensional field
problems. The concepts here reported are mainly quoted from two excellent
books (see References 16 and 17).
These two chapters are useful for acquiring a complete understanding of
how the method works. However, they might be passed over by the reader
primarily interested in the applications of the finite element method.
2.1 Introduction
The requirement of more and more accuracy during the process of design
and analysis of the electrical machines fostered the spreading of numerical
models appropriate for computing electric and magnetic fields. These
numerical methods are essentially based on the determination of the distri-
bution of the electric and magnetic fields in the structures under study, based
on the solution of Maxwell’s equations. An analytical solution is barely
achieved, because of the complex geometrical machine structures and the
nonlinear characteristic of the materials. Then, in most cases only a numerical
solution is possible.
The finite element method is a numerical technique that is suitable for this
purpose. It allows a field solution to be obtained, even with time-variable
fields and with materials that are nonhomogeneous, anisotropic, or non-
linear. Using the finite element method, the whole analysis domain is divided
into elementary subdomains, which are called finite elements, and the field
equations are applied to each of them.
This method was proposed in the 1940s, but it was firstly applied almost
ten years later in aeronautical design and in structural analysis. As years
25
26 Electrical Machine Analysis Using Finite Elements
went by, the finite element method was largely adopted in almost all physical
and mathematical problems. Today it is the most diffused method for the
solution of vector field problems.
The study of the field distributions, and in particular of electromagnetic
field problems, exhibits the following advantages. It allows a meticulous
local analysis to be carried out, highlighting dangerous field gradient, mag-
netic field strength, saturation, and so on. It allows a good estimation of the
performance of the electromagnetic devices under analysis (especially when
the classical methods of analysis give unsatisfactory results). Finally, it per-
mits one to reduce substantially the number of prototypes.
However, the method has some drawbacks, too. Because of its numerical
nature, the solution is necessarily approximate. Then, if the method is not
correctly applied, it might generate inaccurate results. Finally, since the com-
puted quantities are distributed in the space, the required computation time
is generally long.
In order to reduce the computation time, and to improve the analysis at
the same time, each periodicity and symmetry (both geometric and electro-
magnetic symmetry) of the structure is used. The resulting accuracy is influ-
enced by the dimension of the finite elements and by the uniformity of the
subdivision. To increase the accuracy, a fine subdivision of the structure is
carried out, adopting finite elements of smaller dimension. Nevertheless, an
excessive subdivision of the analysis domain causes an aggravation of the
computation time.
L φ(P , t ) = f (P , t ) (2.1)
together with the boundary conditions. The latter constrain the fields along
the boundary Γ of the domain under analysis. In Equation (2.1) L is a
differential operator, φ is the unknown function to be determined, and f is
the forcing function. Equation (2.1) highlights that both φ and f are functions
of the position in the space, P(x,y,z), and of the time, t.
( )
−div ε grad V = ρ (2.2)
(
L = −div ε grad ) (2.3)
φ=0 on Γ 1 (2.4)
2. Nonhomogeneous condition:
φ = φf on Γ 1 (2.5)
∂φ
=0 on Γ 2 (2.6)
∂n
∂φ
+ kφ = 0 on Γ 2 (2.7)
∂n
3. Nonhomogeneous condition:
∂φ
+ kφ = φg on Γ 2 (2.8)
∂n
φ * (P , t ) = ∑ Φ ν (P , t )
j=1
j j (2.9)
where νj are interpolating functions (that are also called expansion functions
or base functions), while Φj are unknown coefficients that have to be deter-
mined during the computation process. Such a combination has to approx-
imate appropriately the exact solution, satisfying the differential operator
[Equation (2.1)] and the boundary conditions at the same time.
The first two methods, the classical residual method and the classical
variational method, take into account the whole analysis domain. The func-
tions νj are defined on the whole domain. Conversely, in the finite element
method the whole domain is divided in subdomains; then the function φ* is
a combination of functions νj that are defined in the subdomains. Conse-
quently, since the subdomains are of reduced dimensions, the interpolating
functions νj can be very simple.
Basic Principles of Finite Element Methods 29
Before illustrating the different procedures, let us introduce the inner prod-
uct between two functions φ and ϕ. Let us refer to the volume τ; then the
inner product is defined as
φ, ϕ =
∫ φϕ dτ
τ
(2.10)
where the symbol ~ indicates the complex conjugate. This inner product is
a linear operation, since the properties of additivity and the product by a
constant are satisfied:
φ1 + φ2 , ϕ = φ1 , ϕ + φ2 , ϕ (2.11)
αφ, ϕ = α φ, ϕ (2.12)
> 0 φ≠0
Lφ, ϕ = (2.13)
= 0 φ=0
Lφ, ϕ = φ, Lϕ (2.14)
r = Lφ * − f (2.15)
equal to zero (or at least very low) in the whole analysis domain. Fixing
some weight functions wi, the residual method forces the integral of the
residuals, weighed by wi to be zero over the domain volume τD. This is to
force the following condition:
30 Electrical Machine Analysis Using Finite Elements
Ri =
∫τD
(
w i L φ * − f dτ = 0 ) (2.16)
There are different weighed residual methods. The best known and most
used method is Galerkin’s method, where the weight functions wi are chosen
equal to the interpolating function νi, i.e.,
w i = νi i = 1, 2 , 3, … , N (2.17)
N
Ri =
∫
τD
νi L
∑j=1
Φ j ν j − νi f dτ
i = 1, 2 , 3, … , N (2.18)
SS φ = T (2.19)
where [φ] is the column vector of the unknown coefficients Φi. [SS] is a matrix
vector that depends on the interpolating functions whose elements are given
by
∫ ( ν Lν + ν Lν ) dτ
1
s ij = i j j i (2.20)
2 τD
If the operator L satisfies the property (2.14), the matrix vector [SS] is
symmetrical, and its elements are
s ij =
∫τD
νi Lν j dτ (2.21)
ti =
∫ τD
νi f dτ (2.22)
Basic Principles of Finite Element Methods 31
1 1 1
F(φ) = Lφ, φ − φ, f − f , φ (2.23)
2 2 2
δF
=0 i = 1, 2 , 3, … , N (2.24)
δΦ i
It is
F + δF = F(φ + δφ)
(2.25)
=
1
2
( ) 1
L φ + δφ , φ + δφ − φ + δφ, f −
2
1
2
f , φ + δφ
1 1 1 1
δF = Lφ, δφ + Lδφ, φ − δφ, f − f , δφ (2.26)
2 2 2 2
1 1
δF = δφ, Lφ − f + Lφ − f , δφ (2.27)
2 2
Thus, from the definition of the inner product (2.10), the following results:
(
δF = Re δφ, Lφ − f ) (2.28)
(
Re δφ, Lφ − f = 0 ) (2.29)
( ) ( ) ()
δ δF = δF φ + δφ − δF φ = Re δφ, Lδφ > 0 ( ) (2.30)
1 1 1
F(φ′) = Lφ′ , φ′ − φ′ , f ′ − f ′ , φ′ (2.31)
2 2 2
1 1 1 1 1
F(φ) = Lφ, φ − Lφ, ψ + φ ,Lψ − φ, f − f , φ (2.32)
2 2 2 2 2
In this equation, the second and the third term in the second member,
containing the function ψ, can be changed in integrals along the domain
boundary, by means of the Stokes theorem and the divergence theorem. Thus,
the function ψ disappears when the fixed boundary conditions are applied.
Further modifications can be applied to both the definition of the func-
tional and the definition of the inner product, which allow the method to be
generalized. In particular, it is possible to adapt the method to complex
differential operators (used in computation of media with losses) or to oper-
ators that do not satisfy the property (2.13) and (2.14). The reader who wants
to know more about these techniques may refer to the tests in the References.
FIGURE 2.1
Elements for the partition of the domain.
Basic Principles of Finite Element Methods 35
φ m * (x , y , z , t ) = ∑Φ
j=1
mj νmj (x , y , z , t ) (2.33)
where n is the number of the nodes of the element, Φmj is the value of φ in
the j-th node of the m-th element. Finally, νmj is the interpolating function
referred to the j-th node of the m-th element. The highest order of the function
defines also the order of the element.
R im =
∫ ν ( Lφ
τ
i m )
* − fm dτ
n (2.34)
τ ∫
= νi L
∑Φ
j=1
mj νmj dτ −
∫ν f
τ
i m dτ i = 1, 2 , … , n
are posed equal to zero. A system of n equations with the n unknown Φmj
is obtained. Applying Equation (2.34) to all the Nm elements that form the
domain, and considering the relationships that link the adjacent elements, a
system of this kind is obtained:
SS φ − T = 0 (2.35)
F(φ*) = ∑ F(φ* )
m=1
m
(2.36)
M
1
= ∑ ∫
m=1
2
τ
φm
* Lφm
* dτ −
1
2 ∫ τ
* dτ
fm φm
1 t t
F(φ*) = φ SS φ − φ T (2.37)
2
∂F
=0 j = 1, 2 , 3, … , N (2.38)
∂Φ j
References
1. O.C. Zienkiewicz, The Finite Element Method in Engineering Science, McGraw
Hill, London, 1971.
2. H.C. Martin and G.F. Carey, Introduction to Finite Element Analysis: Theory and
Application, McGraw-Hill, New York, 1973.
3. D.H. Norrie and G. de Vries, The Finite Element Method: Fundamentals and Appli-
cations, Academic Press, New York, 1973.
Basic Principles of Finite Element Methods 37
4. M.V.K. Chari and P.P. Silvester, Finite Elements in Electrical and Magnetic Field
Problem, New York, John Wiley & Sons, 1980.
5. R.D. Cook, Concepts and Applications of Finite Element Analysis, Wiley, New York,
1981.
6. S.S. Rao, The Finite Element Method in Engineering, Pergamon Press, Oxford,
1982.
7. R.K. Livesley, Finite Element: an Introduction for Engineers, Cambridge University
Press, Cambridge, 1983.
8. P.P. Silvester and R.L. Ferrari, Finite Element Analysis and Design of Electromag-
netic Devices, Cambridge University Press, Cambridge, England, 1983.
9. J.N. Reddy, An Introduction to Finite Element Method, McGraw Hill, New York,
1984.
10. R. Wait and A.R. Mitchell, Finite Element Analysis and Applications, Wiley, New
York, 1985.
11. D.A. Lowther and P.P. Silvester, Computer Aided Design in Magnetics, Springer
Verlag, New York, 1986.
12. H. Grandin, Fundamentals of the Finite Element Method, Macmillan, New York,
1986.
13. D.S. Burnett, Finite Element Analysis: from Concepts to Applications, Addison-
Wesley Publishing Company, Reading, MA, 1987.
14. S.R.H. Hoole, Computer-Aided Analysis and Design of Electromagnetic Device,
Elsevier, New York, Amsterdam, London, 1989.
15. W.B. Bickford, A First Course in the Finite Element Method, Richard D. Irwin,
Homewood, 1990.
16. J.M. Jin, The Finite Element Method in Electromagnetics, John Wiley & Sons, New
York, 1992.
17. N. Ida and J.P.A. Bastos, Electromagnetics and Calculation of Fields, Springer-
Verlag, New York, 1992.
18. S.J. Salon, Finite Element Analysis of Electrical Machine, Kluwer Academic Pub-
lishers, Boston, MA, 1995.
19. A. Reece and T. Preston, Finite Element Method in Electric Power Engineering,
Oxford University Press, UK, 2000.
3
Applications of the Finite Element Method
to Two-Dimensional Fields
3.1 Introduction
3.1.1 Statement of the Two-Dimensional Field Problem
In 2D field problems, the considered domain is a surface S, and its boundary
is a curve. Let φ be the unknown function that is to be determined. It is a
scalar function of the space coordinates x and y, i.e., φ = φ(x,y). The time
dependence is omitted. Let f be the forcing function, which is also function
of x and y, and independent of the time. The 2D field problem is defined by
the differential equation of second order
∂ ∂φ ∂ ∂φ
( )
−div α ⋅ gradφ + βφ = − αx − αy
∂x ∂x ∂y ∂y
+βφ = f (3.1)
together with the boundary conditions that are imposed on the boundary Γ
of the domain. They are Dirichlet’s boundary conditions on the portion Γ1 of
the boundary:
φ = φf on Γ 1 (3.2)
39
40 Electrical Machine Analysis Using Finite Elements
∂φ ∂φ
α x ∂x u x + α y ∂y u y ⋅ n + kφ = φg on Γ 2 (3.3)
In Equation (3.1) and Equation (3.3), αx , αy, and β are known parameters
that are related to the physical property of the materials in the domain. In
Equation (3.2) and Equation (3.3), k, φf, and φg are known parameters that
are related to the physical property of the boundary. In particular, φf and φg
are functions describing the sources along the boundary curves.
The differential notation of Equation (3.1) is of a general nature, whose
Laplace’s, Poisson’s, and Helmholtz’s equations are the particular forms. It
corresponds to that given in Equation (2.1), where the differential operator
corresponds to
∂ ∂ ∂ ∂
L=− αx − αy +β (3.4)
∂x ∂x ∂y ∂y
Lφ, ϕ =
∫ −div (α ⋅ gradφ) + βφ ϕ dS
S
(3.5)
=
∫ −div (α ⋅ gradφ) ϕ dS + β ∫ φϕ dS
S S
∂φ ∂ϕ
Lφ, ϕ =
∫
S
( )
φ −div α ⋅ gradϕ dS −
∫
Γ
α ⋅ ϕ
∂n
− φ dl + β
∂n ∫ φϕ dS
S
(3.6)
Applications of the Finite Element Method to Two-Dimensional Fields 41
The second addendum in the second member is null, since it has been
assumed that both the functions φ and ϕ satisfy homogeneous conditions on
the boundary Γ. Then, it is
Lφ, ϕ =
∫ φ −div (α ⋅ gradϕ ) + βϕ dS
S
(3.7)
= φ, Lϕ
Lφ, φ =
∫ −div (α ⋅ gradφ) + βφ φ dS
S
(3.8)
=
∫S
(
−div α ⋅ gradφ φ dS + β
) ∫ S
φ φ dS
∂φ
Lφ, φ =
∫S
α ⋅ gradφ ⋅ gradφ dS −
∫Γ
α ⋅ φ
∂n
dl + β
∫
S
φ φ dS (3.9)
∫ α gradφ ∫ ∫φ
2 2 2
Lφ, φ = dS + k φ dl + β dS
S Γ2 S
(3.10)
∫ α gradφ ∫
2 2 2
= +β φ dS + k φ dl
S Γ2
1 1 1
F(φ) = Lφ, φ − φ, f − f , φ
2 2 2
(3.11)
∫ ∫ ∫( )
1 k 1
φf + fφ dS
2 2 2
= α gradφ +β φ dS + φ dl −
2 S 2 Γ2 2 S
Given that φ and f are real functions, it is possible to obtain the following
formulation:
2 2
∂φ ∂φ
∫ ∫ ∫ f φ dS
1 k 2
F(φ) = α x + α y + βφ2 dS + φ − φφg dl − (3.13)
2 S ∂x ∂y Γ2 2 S
∂φ ∂δφ ∂φ ∂δφ
∫α
1
δF(φ) = x + αy + βφδφ dS +
2 S ∂x ∂x ∂y ∂y
(3.14)
∫ ( kφ − φ ) δφ dl − ∫
Γ2
g
S
f δφ dS
The two addenda of the first integral in the second member can be rewrit-
ten as
∂φ ∂δφ ∂ ∂φ ∂ ∂φ
αx = αx δφ − αx δφ
∂x ∂x ∂x ∂x ∂x ∂x
(3.15)
∂φ ∂δφ ∂ ∂φ ∂ ∂φ
αy = αy δφ − αy δφ
∂y ∂y ∂y ∂y ∂y ∂y
∂ ∂φ ∂ ∂φ
δF(φ) =
∫ − α x −
S ∂x
αy
∂x ∂y ∂y + βφ − f δφ dS +
(3.16)
∂φ ∂φ
∫ α x
Γ ∂x
ux + α y u y ⋅ n δφ dl +
∂y ∫Γ2
kφ − φg δφ dl
∂ ∂φ ∂ ∂φ
δF(φ) =
∫ − ∂x α
S
x − αy
∂x ∂y ∂y
+ βφ − f δφ dS +
(3.17)
∂φ ∂φ
∫ α x
Γ2 ∂x
ux + α y u y ⋅ n + kφ − φg δφ dl
∂y
∂ ∂φ ∂ ∂φ
− αx − αy + βφ − f = 0
∂x ∂x ∂y ∂y
(3.18)
∂φ ∂φ
α x ∂x u x + α y ∂y u y ⋅ n + kφ − φg = 0
φm (x , y) = a + bx + cy (3.19)
In particular, in the three nodes of the triangle, the three i-th values are
given by
44 Electrical Machine Analysis Using Finite Elements
Φ3 (x3, y3)
Φ2 (x2, y2)
Φ1 (x1, y1)
3(x3, y3)
Am 2 (x2, y2)
1 (x1, y1)
FIGURE 3.1
Linear interpolation of the potential function φm in the m-th triangular element.
Φ1 = a + bx1 + cy1
Φ 2 = a + bx 2 + cy 2 (3.20)
Φ = a + bx + cy
3 3 3
From the knowledge of the value of the function in the nodes of each finite
element, i.e., Φ1, Φ2, and Φ3, by means of Equation (3.19), it is possible to
compute the potential function in any other point of the element. This is
represented in Figure 3.1.
If the three values of the potential are given in the three nodes of the
element, it is possible to solve the system (3.20) in the three unknowns a, b,
and c. At first, it is posed
1 x1 y1
1
Am = 1 x2 y2
2
1 x3 y3 (3.21)
=
1
2
( ) ( ) (
x 2 y 3 − x 3 y 2 + x 3 y1 − x1y 3 + x1y 2 − x 2 y1 )
that represents the area of the m-th triangular element, as shown in Figure
3.1. By Cramer’s rule, it results that
a=
1
2A m
( ) ( )
Φ1 x 2 y 3 − x 3 y 2 + Φ 2 x 3 y1 − x1y 3 + Φ 3 x1y 2 − x 2 y1 ( )
b=
1
2A m
( ) ( )
Φ1 y 2 − y 3 + Φ 2 y 3 − y1 + Φ 3 y1 − y 2 ( ) (3.22)
c=
1
2A m
( ) ( )
Φ1 x 3 − x 2 + Φ 2 x1 − x 3 + Φ 3 x 2 − x1 ( )
Applications of the Finite Element Method to Two-Dimensional Fields 45
It is
1
a= Φ1p1 + Φ 2 p2 + Φ 3 p3
2A m
1
b= Φ1q1 + Φ 2q2 + Φ 3q3 (3.23)
2A m
1
c= Φ1r1 + Φ 2 r2 + Φ 3r3
2A m
where
p1 = x 2 y 3 − x 3 y 2 q1 = y 2 − y 3 r1 = x 3 − x 2
p2 = x 3 y 1 − x 1 y 3 q2 = y 3 − y 1 r2 = x1 − x 3 (3.24)
p3 = x 1 y 2 − x 2 y 1 q3 = y 1 − y 2 r3 = x 2 − x1
Equation (3.19) expresses the unknown function φm(x,y), inside the m-th
triangular finite element. It can be rewritten as
1 3 1 3 1 3
φ m (x , y ) =
2 A m ∑
i =1
pi Φ i +
2 A m ∑ i =1
qi Φ i ⋅ x +
2 A m ∑ r Φ ⋅ y
i =1
i i
3
(p + q x + r y) ⋅ Φ
= ∑
i =1
i
2A m
i i
i (3.25)
= ∑ ν (x , y ) ⋅ Φ
i =1
i i
Φ3 = φm(x3, y3) Φ3 Φ3
Φ2 = φm(x2, y2) Φ2 Φ2
Φ1 = φm(x1, y1) Φ1 Φ1
3 (x3, y3) 3 (x3, y3) 3 (x3, y3)
Am Am Am
2 (x2, y2) 2 (x2, y2) 2 (x2, y2)
FIGURE 3.2
Representation of the linear interpolating functions of the function φm in the m-th triangular
element drawn in Figure 3.1.
46 Electrical Machine Analysis Using Finite Elements
the nodes of the triangle itself. It is then evident that, to get the field, it is
sufficient to compute the values of the unknown function φm(x,y) in the nodes
of each element that form the whole domain. In Equation (3.25), the inter-
polating functions νi(x,y), with i = 1, 2, 3, that interpolate the function φm(x,y)
have been highlighted. Each of them can be represented graphically, as
illustrated in Figure 3.2.
The values (gradφm · gradφm) and (curlφm · curlφm), inside each triangular
element, are constant, since they do not depend on the coordinates x and y.
In fact, the results are
∂F
=0 with i = 1, 2 , 3, … , N n (3.30)
∂Φ i
Nm
F= ∑F
m =1
m (3.31)
i
=0 with i = 1, 2 , 3, … , N n (3.32)
2 2
∂φ ∂φ
∫ ∫
1
Fm = α x m + α y m dS + fm φm dS (3.33)
2 Am ∂x ∂y Am
48 Electrical Machine Analysis Using Finite Elements
3
∂φm
∂x
=
1
2A m ∑q Φ
i =1
i i
(3.34)
3
∂φm
∂y
=
1
2A m ∑r Φ
i =1
i i
so that
2 3 3
∂φm 1
∂x = 4A 2
m
∑∑q q Φ Φ
i =1 j=1
i j i j
(3.35)
2 3 3
∂φm
∂y
=
1
4A m 2 ∑∑r rΦ Φ
i =1 j=1
i j i j
Both the terms in Equation (3.35) are constant inside the elements. They
can be put outside the integral of Equation (3.33). After some manipulations,
the functional Fm of the m-th element results in
3 3 3
Fm =
1
8A m ∑ ∑(
i =1 j=1
)
α xqiqj + α y ri rj Φ i Φ j + fm
Am
3 ∑Φ
i=1
i (3.36)
where the terms of the matrix [Sm] and the vector [Tm] are
s ij =
1
4A m
(
α xqiqj + α y ri rj )
(3.38)
A
t i = fm m
3
Applications of the Finite Element Method to Two-Dimensional Fields 49
It is observed that the terms sij depend only on the geometry and the
material characteristic (expressed by the coefficients αx and αy). If the mate-
rial is linear, the terms sij do not depend on the value of the potentials in the
nodes (since αx and αy are constant with the field). By analogy, the terms ti
are only functions of the geometry of the element and of the forcing quantity
fm.
Since the research of the stationary point of the functional Fm is carried
out by posing to zero the derivative of Fm with respect to the values Φ1, Φ2,
and Φ3 in the three nodes of the m-th element. Deriving Fm in Equation (3.37)
with respect Φ1, Φ2, and Φ3 and equating the result to zero, the results are
t
∂Fm ∂Fm ∂Fm
∂Φ , ∂Φ , ∂Φ = S m Φ123 − Tm = 0 (3.39)
1 2 3
Equation (3.39) shows that, for the generic m-th triangular element, a
system of three equations is obtained, with the three unknown potential
values Φ1, Φ2, and Φ3 in the nodes of the element. When this system of
equations is satisfied, the field problem solution is obtained in the m-th
element.
3 (x3, y3)
4 (x4, y4)
Fb
Fa
2 (x2, y2)
1 (x1, y1)
FIGURE 3.3
Adjacent elements with shared nodes 2 and 3.
50 Electrical Machine Analysis Using Finite Elements
The two triangles are characterized by the local functionals Fa and Fb, given
by
1 t t
Fa = Φ123 Sa Φ123 − Φ123 Ta
2
(3.40)
1 t t
Fb = Φ 234 S b Φ 234 − Φ 234 Tb
2
1 t t
Fa + Fb = Φ1234 Sab Φ1234 − Φ1234 Tab = (3.41)
2
Φ1
t
sa(11) sa(12 ) sa(13) 0 Φ Φ t t a(1)
1 1
1 Φ2 sa( 21) sa( 22 ) + s b(11) sa( 23) + s b(12 ) s b(13) Φ 2 Φ 2 t a( 2 ) + t b(1)
s −
2 Φ3
a( 31) sa( 32 ) + s b( 21) sa( 33) + s b( 22 ) s b( 23) Φ 3 Φ 3 t a( 3) + t b( 2 )
Φ 4 0 s b( 31) s b( 32 ) s b( 33) Φ 4 Φ 4 t b( 3)
Proceeding in the same way for each triangular element with shared nodes,
the complete functional of the whole domain is obtained as
1 t t
F= Φ SS Φ − Φ T (3.42)
2
where [Φ] is the column vector of all the Nn values of the function φ in all
the nodes of the domain. The dimension of the matrix [SS] is Nn × Nn. It is
called the “stiffness” matrix, due to the analogy with the structural analysis
to which the finite element method has been originally applied. Finally, [T]
is the column vector of the known terms, and its dimension is Nn.
The stiffness matrix is a sparse matrix, i.e., with several zeros. As observed
in Equation (3.38), in linear conditions, this matrix is built and is not modified
during the process of search of the solution. The stationary point of the
functional F (3.42) corresponds to
SS Φ − T = 0 (3.43)
All the time derivatives are null and the current density J is null as well.
The unique specific electric force is Coulomb’s electric field, E = Ec. The
second equation highlights that the electric field is irrotational in the whole
domain, which is conservative. Thus, it is possible to define a scalar electric
potential V such as Ec = −gradV. From these equations, the field problem is
described by the following almost harmonic scalar equation:
∂ ∂V ∂ ∂V
εx + εy = −ρ (3.45)
∂x ∂x ∂y ∂y
∂2 V ∂2 V ρ
+ =− (3.46)
∂x 2 ∂y 2 ε
1
F=
∫S
2 εE c − ρV dS
2
(3.47)
∫ ∫
1
Fm = ε (gradV · gradV)dS − ρ V dS
2 Am Am
∫ ∫
1
= ε ( b2 + c 2 )dS − ρ (a + bx + cy) dS (3.48)
2 Am Am
1 x + x2 + x3 y + y2 + y3
= εA m ( b2 + c 2 ) − ρA m a + b 1 +c 1
2 3 3
The coefficients a, b, and c are functions of the potential V1, V2, and V3 of
the nodes of the m-th triangular element, where the functional is computed.
Substituting the expressions (3.22), that have been obtained for a, b, and c
as a function of the potentials, into Equation (3.48), Fm can be written as
Applications of the Finite Element Method to Two-Dimensional Fields 53
1 t t
Fm = V123 S m V123 − V123 Tm (3.49)
2
where [V123] is the column vector of the potentials V1, V2, V3 of the three
nodes of the triangular element, and [Sm] is the stiffness matrix, with dimen-
sion 3 × 3. The generic component corresponding to the i-th row and j-th
column of [Sm] is given by
ε
s ij = (qiqj + ri rj ) (3.50)
4A m
Finally [Tm] is a column vector of known terms, which is [Tm] = [t1, t2, t3]t,
where
Am
ti = − ρ (3.51)
3
Of course, Equation (3.50) and Equation (3.51) coincide with those given
in Equation (3.38), once equating fm = −ρ and αx = αy = ε.
∂ ∂V ∂ ∂V
σx + σy =0 (3.53)
∂x ∂x ∂y ∂y
∂2 V ∂2 V
+ =0 (3.54)
∂x 2 ∂y 2
54 Electrical Machine Analysis Using Finite Elements
1 2
Fm =
∫Am
2 σE dS
∫
1
= σ (gradV ⋅ gradV)dS (3.55)
2 Am
1
= σA m ( b2 + c 2 )
2
Since the coefficients a, b, and c are functions of the potentials V1, V2, and
V3 of the nodes of the triangular element, the functional Fm can be expressed
as
1 t
Fm = V123 S m V123 (3.56)
2
where [V123] = [V1, V2, V3]t, and the generic component of the matrix [Sm] is
given by
σ
s ij = (qiqj + ri rj ) (3.57)
4A m
J 1
curlE = curl = curl curlN = 0
σ σ
( ) (3.58)
The coordinate system is chosen so that the electric vector potential N has
only a z-axis component, which is N = [0, 0, Nz]. In this way, there are only
x-axis and y-axis components of the current density vector J:
∂N ∂Nz
J = curl 0, 0, Nz = z , − , 0 (3.59)
∂y ∂x
Applications of the Finite Element Method to Two-Dimensional Fields 55
∂2 Nz ∂2 Nz
+ =0 (3.60)
∂x 2 ∂y 2
1 2
Fm =
∫
Am 2σ
J dS
∫
1
= (curlN ⋅ curlN)dS (3.61)
2σ Am
1
= A m ( b2 + c 2 )
2σ
Since a, b, and c are functions of the potential values N1, N2, and N3 of the
three nodes of the triangular element, the functional Fm becomes
1 t
Fm = N123 S m N123 (3.62)
2
where [N123] = [N1, N2, N3]t, and [Sm] is the stiffness matrix. Its generic
component is
1
s ij = (qiqj + ri rj ) (3.63)
4σA m
B 1
curlH = curl = curl curlA = J (3.64)
µ µ
In the 2D field, using Cartesian coordinates, the current density vector J has
only a component normal to the plane (x, y), which is only a z-axis compo-
nent. Consequently, the magnetic vector potential A has only a z-axis com-
ponent, i.e., the vector A is parallel to the vector J. Then these vectors can
be expressed as J = [0, 0, Jz] and as A = [0, 0, Az]. Equation (3.64) becomes
56 Electrical Machine Analysis Using Finite Elements
∂ 1 ∂A z ∂ 1 ∂A z
+ = − Jz (3.65)
∂x µ x ∂x ∂y µ y ∂y
∂A ∂A z
B = curl 0, 0, A z = z , − , 0 (3.66)
∂y ∂x
∂2 A z ∂2 A z
+ = − µJ z (3.67)
∂x 2 ∂y 2
1
Fm =
∫Am
2
B ⋅ H − J ⋅ A dS
∫ ∫
1
= (curlA ⋅ curlA)dS − J z A zdS (3.68)
2µ Am Am
1 x + x2 + x3 y + y2 + y3
= A m ( b2 + c 2 ) − J z A m a + b 1 +c 1
2µ 3 3
Since a, b, and c are functions of the potentials A1, A2, and A3 of the three
nodes of the triangular element, Fm becomes
1 t t
Fm = A 123 S m A 123 − A 123 Tm (3.69)
2
where [A123] = [A1, A2, A3]t, and [Sm] is the stiffness matrix, whose generic
component is
1
s ij = (qiqj + ri rj ) (3.70)
4µA m
Am
ti = − Jz (3.71)
3
Applications of the Finite Element Method to Two-Dimensional Fields 57
∂2 Ψ ∂2 Ψ
+ =0 (3.73)
∂x 2 ∂y 2
1
Fm =
∫Am
2 B ⋅ H dS
∫
1
= µ (gradΨ ⋅ gradΨ )dS (3.74)
2 Am
1
= µA m ( b2 + c 2 )
2
Since a, b, and c depend on the potentials Ψ1, Ψ2, and Ψ3 of the three nodes
of the triangular element, Equation (3.74) becomes
1 t
Fm = Ψ 123 S m Ψ 123 (3.75)
2
where [Ψ123]=[Ψ1, Ψ2, Ψ3]t, and [Sm] is the stiffness matrix whose generic
component corresponds to
µ
s ij = (qiqj + ri rj ) (3.76)
4A m
B − Bres B B
curlH = curl = curl − curl res
µ µ µ
(3.77)
B
1
( )
= curl curlA − curl res = 0
µ µ
B
curl
1
µ
( µ
)
curlA = curl res (3.78)
The functional that is associated with the field, referred to as the area Am
of the m-th triangular element, is given by
∫
1
( )
2
Fm = B − B res dS
A m 2µ
∫
1
= (B 2 − 2BB res + B res 2 )dS (3.79)
2µ Am
∫
1
= (curlA ⋅ curlA − 2B res curlA + B res 2 )dS
2µ Am
1 t t
Fm = A 123 S m A 123 − A 123 Tm (3.80)
2
where [A123] = [A1, A2, A3]t, and the generic component of the stiffness matrix
[Sm] is
1
s ij = (qiqj + ri rj ) (3.81)
4µA m
ti =
1
µ
(
B res , yqi − B res ,x ri ) (3.82)
Applications of the Finite Element Method to Two-Dimensional Fields 59
x1 + x 2 + x 3 y1 + y 2 + y 3
= =0
3 3
∫Am
x dx dy = 0
∫Am
y dx dy = 0
1 x1 y1
∫
1
dx dy = 1 x2 y2 = Am
Am 2
1 x3 y3
∫Am
x 2 dx dy =
Am
12
(
x 12 + x 2 2 + x 3 2 )
∫Am
y 2 dx dy =
Am
12
(
y 12 + y 2 2 + y 3 2 )
∫Am
xy dx dy =
Am
12
(
x1y1 + x 2 y 2 + x 3 y 3 )
References
See references in Chapter 2.
4
The Analysis Procedure Using the Finite
Element Method
4.1 Introduction
This chapter deals with the analysis of the electrical machines by means of the
finite element method. Instead of a generic problem, the magnetic field prob-
lem is investigated. Such a problem is the most common in the analysis of
electrical machines, including transformers, rotating machines, and actuators.
In any case, the study of the magnetic field alone is not sufficient for a
complete analysis of the machine. For instance, the evaluation of the electric
field between the turns of the coils requires an electrostatic field analysis;
the computation of the Joule losses in conductors with restricted sections,
especially in the terminals, requires a current field analysis. However, the
extension of the finite element analysis to vector fields of a nature different
from the magnetic field is left to the reader.
61
62 Electrical Machine Analysis Using Finite Elements
z
y
z ϑ
r
(a) (b)
FIGURE 4.1
Planar symmetry (a) and axial symmetry (b).
∂2 A z ∂2 A z
+ = − µJ z (4.1)
∂x 2 ∂y 2
The Analysis Procedure Using the Finite Element Method 63
Homogeneous
Dirichlet
condition
FIGURE 4.2
Homogeneous Dirichlet’s condition along the external circumference of a synchronous generator.
64 Electrical Machine Analysis Using Finite Elements
stator back iron, the condition Az = 0 is assigned along all the external
circumference of the synchronous machine. Figure 4.2. highlights this homo-
geneous Dirichlet’s condition, by means of a bold line.
A A' A A'
SCu+ SCu−
Neumann
condition
(a) (b)
FIGURE 4.3
Reduction of the analysis domain by means of Neumann’s conditions.
The Analysis Procedure Using the Finite Element Method 65
ϑ ϑ
FIGURE 4.4
Reduction of the analysis domain by means of the periodic condition.
π
( )
A z r , ϑ = + A z r , ϑ + (2 k )
p
k = 1, 2 , 3, … (4.2)
π
( )
A z r , ϑ = − A z r , ϑ + (2 k − 1)
p
k = 1, 2 , 3, … (4.3)
y
x
FIGURE 4.5
Flux lines, i.e., lines of the flux density vector B.
∂A z
Bx =
∂y
(4.4)
∂A z
By = − =0
∂x
Equation (4.4) shows that the flux density vector has a component only in
the direction of the equipotential line of Az.
y B n y
B n
S
1 2 1 IS 2
L
z x x
(a) (b)
FIGURE 4.6
Surface integral in the 2D problem.
Φ=
∫ B ⋅ n dS = L ∫ B ⋅ n dl
S lS
(4.5)
Φ=
∫ B ⋅ n dS = ∫ ( curlA) ⋅ n dS = ∫ A ⋅ t d l
S S l
(4.6)
Φ = (A z1 − A z 2 )L (4.7)
L
Φ=
S Cu ∫ S Cu +
A z dS −
∫ S Cu −
A z dS
(4.8)
68 Electrical Machine Analysis Using Finite Elements
where SCu+ and SCu– are the conductor surfaces, positively oriented and
negatively oriented, respectively.
If the current is supposed to be uniformly distributed on the conductor
surface, the flux linkage Λ is obtained by multiplying the magnetic flux Φ
by the number of turns Nt that link the flux.
pJ (t ) =
∫ ρJ
τ
2
dτ (4.9)
pJ (t ) = L
∫ ρJ
S
2
z dS (4.10)
J = J zu z = J z e jϕu z (4.11)
The average power that is lost in a period due to Joule effect is given by
2
Jz
∫ ∫
1
pJ = ρ dτ = ρ J ⋅ J dτ (4.12)
τ 2 2 τ
where the factor 1/2 appears, because Jz denotes the maximum of the
sinusoidal variation.
B A
Wm =
∫∫
τ 0
H ⋅ dB dτ =
∫∫
τ 0
J ⋅ dA dτ (4.13)
The Analysis Procedure Using the Finite Element Method 69
∫ H ⋅ B dτ = 2 ∫ µH
1 1
Wm = 2
dτ
2 τ τ
(4.14)
∫ ∫JA
1 1
= J ⋅ A dτ = z z dτ
2 τ 2 τ
In the case of stationary fields, the energy is constant, but in the case of
variable fields, this energy has to be considered as instantaneous, at the
time t. In particular, if the variable fields have a sinusoidal variation, and
they are described by using the symbolic notation, it is possible to compute
the average energy stored in a period as
B⋅H Jz ⋅ Az
∫ ∫
1 1
Wm = dτ = dτ (4.15)
2 τ 2 2 τ 2
where 1/2 appears because B and H represent the maximum value of the
flux density and the magnetic field strength, respectively, and also Jz and
Az represent the maximum value of the current density and magnetic vector
potential, respectively.
Wm′ =
∫ H ⋅ B dτ − W
τ
m
(4.16)
=
∫τ
J ⋅ A dτ − Wm
Ft = Lµ o
∫ H H
l
t n dl
(4.17)
∫ ( H ) dl
µ
Fn = L o n
2
− Ht 2
2 l
The magnitude of the force and the direction of the force (which is
expressed with respect the direction normal to the surface) are given by
Equation (1.100) and Equation (1.101), respectively.
As far as the computation time is concerned, it is worth noticing that
Maxwell’s strength tensor requires only one field solution.
As far as the computation accuracy is concerned, it is observed that with
a correct magnetic field distribution, the method of Maxwell’s strength tensor
provides an exact value of the force, disregarding the choice of the surface
(i.e., the line) of integration. However, due to the intrinsic approximation of
the adopted numerical method, the continuity of the field components
among the adjacent elements is not guaranteed. As a result, the computed
forces depend on both the path of integration and the adopted mesh of the
domain.
Some preliminary tests are essential, in order to verify the accuracy obtain-
able with the adopted mesh of the structure and the selected path for the
integration. For instance, a possible check may be the evaluation of the force
in a circumstance where this force should be null. An example to check the
accuracy of the results by using the method of Maxwell’s strength tensor is
illustrated in Chapter 12.
pJ
R= (4.18)
I2
2 Wm Λ
L= = (4.19)
I2 I
References
See references in Chapter 2.
5
Cylindrical Magnetic Devices
5.1 Introduction
The cylindrical magnetic device is used in applications requiring high thrust
but small stroke, e.g., to move power switches, to drive valves, and so on.
A section of the cylindrical magnetic device is shown in Figure 5.1. It consists
of a fixed magnetic part, i.e., the core, and of a cylindrical magnetic plunger
moving axially within the core, guided by a nonmagnetic guide. A coil
composed by Nt turns is wound inside the core.
When no currents feed the coil, the plunger is in its rest position, at the
lower position, at which the air-gap thickness is maximum, i.e., g = gmax.
When the coil is fed by a current with a sufficiently high amplitude, the
plunger pops up, reaching the upper position, at which the air-gap is min-
imum, i.e., g = gmin.
The radial gap between the core and the plunger, corresponding to the
driver thickness, is constant and equal to t.
73
74 Electrical Machine Analysis Using Finite Elements
fixed magnetic
part (core)
coil
g variable
airgap
ht non-magnetic
driver
r
cylindrical
plunger t d
FIGURE 5.1
Sketch of the cylindrical magnetic device.
Nt i = H g g + H t t (5.1)
where Hg and Ht are the magnetic field strength in the air-gap g and t,
respectively. Because of the small air-gap thickness t, a constant value of Ht
is considered.
Gauss’s law yields
πd 2
Bg = B t h t π(d + t ) (5.2)
4
where Bg and Bt are the flux density in the air-gap g and t, respectively. In
Equation (5.2), π(d + t) is the average circumference in the middle of the air-
gap t. From the two equations (5.1) and (5.2), the flux density Bg is obtained as
Nt i Nt i
Bg = µo = µo (5.3)
d2t g + t'
g+
4 h t π(d + t )
Cylindrical Magnetic Devices 75
Nsp i
g Bg Hg
Rg ϕg
Ht Bt
Rt
t d
FIGURE 5.2
Definition of the magnetic strength and the flux density in the two air-gaps of the device (a)
and equivalent magnetic circuit (b).
g
Rg = (5.4)
πd 2
µo
4
Rt =
(
ln 1 + 2 t / d ) (5.5)
µ o 2 πh t
With a small ratio t/d, so that ln(1 + 2t/d) is approximated by t/(d + t),
Equation (5.5) can be expressed as
t
Rt = (5.6)
µ oπh t (d + t )
The magnetic flux ϕg through the base surface of the plunger is computed
as
Nt i
ϕg = (5.7)
Rg + Rt
The flux density Bg is achieved as the ratio between ϕg and the air-gap
surface πd2/4, as
76 Electrical Machine Analysis Using Finite Elements
ϕg
Bg = (5.8)
πd 2
4
Nt 2 i
λ = Nt ϕ g = µ oπd 2 (5.9)
4(g + t ')
λo
Wm =
∫ 0
i(λ)dλ (5.10)
λ magnetic g = constant
energy magnetization
λo curve
Wm
W'm magnetic
coenergy
io i
FIGURE 5.3
Magnetization curve, magnetic energy, and coenergy.
λo
Wm′ =
∫0
λ(i)di
(5.11)
= λ oi o − Wm
1
Wm = Wm′ = λ oio (5.12)
2
1 πd 2 1
Wm = Bg 2 g+ B t 2 h t π(d + t )tt (5.13)
2µ o 4 2µ o
dWm g = go
λo
di
io i
FIGURE 5.4
Current-flux linkage characteristic and surface corresponding to the apparent inductance com-
putation.
(the core and the plunger), they are also functions of the current i flowing
through the coil.
Let us consider the operating point (io, λo), defined by the air-gap go, the
current io, and the flux linkage λo = λ(io, go). The apparent inductance of the
coil is given by the ratio between the flux linkage λo and the current io, which
is
λo
(
Lapp i o , g o = ) io
(5.14)
The ratio between an infinitesimal variation of flux linkage, dλ, and the
corresponding variation of current, di, around the operating point (io ,λo),
defines the differential inductance of the coil, which is
dλ
(
Ldif i o , g o = ) di ( i
(5.15)
o ,λo )
Wm + Wm′ λ i
( )
Lapp i o , g o =
io2
= o 2o
io
(5.16)
The quantity (½λoio) has the same dimension of an energy, but it does not
have a physical meaning. It corresponds to the surface shown in Figure 5.4.
The same figure highlights the magnetic energy variation dWm corresponding
Cylindrical Magnetic Devices 79
O io i O i O io i
FIGURE 5.5
Energy variations corresponding to the air-gap variation –dg.
to the current variation di, around the working point. From this variation,
since dWm ≈ dλio, the differential inductance can be roughly estimated as
( )
Ldif i o , g o ≈
1 dWm
i o di ( λ ,i )
(5.17)
o o
2 Wm
Lapp = Ldif = (5.18)
io2
λ λ λ λ
B B B B
A A A A
i i i i
dWmech Wm(A) Wm(B) dWel
FIGURE 5.6
Energy balance during the air-gap variation -dg.
where dWm is the variation of the magnetic energy stored in the magnetic
field, during the movement of the plunger from A to B, which is
and dWel is the electrical energy that is furnished by the electrical source via
the coil terminals, which is
λ o + dλ
dWel =
∫λo
i(λ)dλ (5.21)
These quantities can be drawn in the (i, λ) plane, as shown in Figure 5.6.
Neglecting the infinitesimal term of second order, the area OAB can be
approximated by the area OAB', shown in Figure 5.5(b), which corresponds
to the difference between the stored magnetic energy before and after the
movement –dg, with a constant flux linkage λ = λo. Then the force component
is computed as
(
Fz λ o , g = − ) dWm
=−
dWm λ , g ( ) (5.22)
dz λ = λo
−dg
λ = λo
λ λ λ
B B B
A A A
i i i
FIGURE 5.7
Mechanical work with different finite movements: although the starting and final points are
the same, the dependence on the way of variation of λ and i is evident.
( )
Fz i o , g = +
dWm′
=+
( )
dWm′ i , g
(5.23)
dz i = io
−dg
i = io
( )
2
1 Nt i πd 2
Wm′ = µo g (5.24)
2 g + t′ 4
πd 2
1
( ) 1
2
Fz = Nt i (5.25)
2µ o ( )
2
4 g + t′
82 Electrical Machine Analysis Using Finite Elements
pn =
µo
2
( µ
) B 2
Hn 2 − Ht 2 = o Hn 2 = g
2 2µ o
(5.26)
where Hn and Ht are the normal and tangential components of the field
strength at the surface of the plunger in front of the air-gap. Then the total
force on the plunger is
1 πd 2
Fz = Bg 2 (5.27)
2µ o 4
∂A ϑ
Br = −
∂z
(5.28)
Bz =
1 ∂ rA ϑ( A
= ϑ+
)
∂A ϑ
r ∂r r ∂r
Cylindrical Magnetic Devices 83
z
Aϑ = 0
Aϑ = 0 Aϑ = 0
r
Aϑ = 0
FIGURE 5.8
Sketch of the section of the cylindrical magnetic device, which is used in the finite element
analysis and boundary conditions.
In addition, all the field quantities depend on the variables r and z only,
i.e., Jϑ = Jϑ(r,z), Aϑ = Aϑ(r,z), Br = Br(r,z), Bz = Bz(r,z). The components of the
magnetic field strength vector H = (Hr , 0, Hz) are connected to those of the
flux density vector by the constitutive law.
The field problem is magnetostatic and is expressed by the Poisson differ-
ential equation
1 ∂ ∂A ϑ ∂2 A ϑ
∇2A ϑ = r + = −µJ ϑ (5.29)
r ∂r ∂r ∂z 2
B = Br 2 + Bz 2 (5.30)
ro 2π
λ ro =
∫ ∫
0 0
B z (r , z) r dϑ dz (5.31)
z
Bz uz r
ro
dϑ
dr = r dϑ
(a)
z
z
r r
ro
Aϑ uϑ
dϑ A = (0, Aϑ, 0)
SCu
dr = r dϑ
(b) (c)
FIGURE 5.9
Integral lines for the computation of the flux linkage.
Cylindrical Magnetic Devices 85
λ ro =
∫ A (r, z) dr
ro
ϑ
2π
=
∫0
roA ϑ (r , z) dϑ (5.32)
= 2 πroA ϑ (ro , z)
as shown in Figure 5.9(b). If the turn is not dimensionless but has a section
SCu, as shown in Figure 5.9(c), the average flux linkage becomes
∫
1
λ= 2πrA ϑdS (5.33)
S Cu S Cu
At last, if Nt turns are considered that cover the whole surface SCu, the flux
linkage is
∫
Nt
λ= 2πrA ϑdS (5.34)
S Cu S Cu
B B
Wm =
∫∫
τ 0
HdB dτ = 2 π
∫ ∫
S
r
0
HdB dS (5.35)
where r is the dummy radius. Referring to the current density and the
magnetic vector potential, Wm is given by
Aϑ Aϑ
Wm =
∫∫
τ 0
J ϑdA ϑ dτ = 2 π
∫ ∫
S
r
0
J ϑdA ϑ dS (5.36)
Wm′ =
∫ BH dτ − W
τ
m (5.37)
86 Electrical Machine Analysis Using Finite Elements
∫ r 2 BH dS
1
Wm = 2 π (5.38)
S
∫ r2J A
1
Wm = 2 π ϑ ϑ dS (5.39)
S
The core and the plunger of the magnetic device show a nonlinear mag-
netic characteristic. However, when the magnitude of the flux density is low,
they work in the linear part of the B-H curve. Hence, a constant permeability
can be assigned even in these objects.
5.3.2.4 Inductances
The apparent inductance is obtained by dividing the flux linkage (5.34) by
the current:
∫
Nt
2πrA ϑdS
S Cu
Lapp =
S Cu
(5.40)
i
If the magnetic circuit is linear, the inductance Lapp can be computed from
the energy quantities as
1
Lapp =
2
2 ∫ BH dτ = 2π ∫ rBH dS
τ S
(5.41)
i2 i2
The result depends on the value of the current variation ∆i. A more accurate
result is obtained if the field problem is solved for two current values around
i, which are i + ∆i and i – ∆i. Adopting the Taylor series expansion of the
Cylindrical Magnetic Devices 87
flux linkage around the current value i, and neglecting the terms of order
higher than the second order, it is
dλ(i) 1 d 2 λ( i ) 2
λ(i + ∆i) = λ(i) + ∆i + ∆i +…
di 2 di 2
(5.43)
dλ(i) 1 d 2 λ( i ) 2
λ(i − ∆i) = λ(i) − ∆i + ∆i −…
di 2 di 2
By subtracting the second equation from the first one, the second-order
terms disappear and the differential inductance results in
Fn =
∫Sz
µo
2
(
H n 2 − H t 2 dS ) (5.45)
On the surface Sr , whose normal unity vector has an r-axis direction, the
force contribution to Fz corresponds to the component of the force tangential
to Sr , which is
Ft =
∫Sr
µ oH n H t dS (5.46)
Alternatively, the force in the z-axis direction can be computed from the
magnetic coenergy variation, see Equation (5.23), as
Fz = −
( )
dWm′ i , g
dg
i = const
(5.47)
=−
( ) (
Wm′ i , g + ∆g − Wm′ i , g − ∆g )
2 ∆g
i = const
88 Electrical Machine Analysis Using Finite Elements
FIGURE 5.10
Definition of the regions that are interesting to the movement of the plunger; they are adopted
to avoid errors due to variation of the mesh.
where W′m (i, g + ∆g) and W′m (i, g – ∆g) are the magnetic coenergy values that
are computed from the field solutions, and correspond to the air-gap length
(g + ∆g) and (g – ∆g), respectively, and to the given current i.
The result of Equation (5.47) depends on the variation of the air-gap length
∆g. This has to be neither too small, so that the difference in the numerical
solution is not appreciable, nor too large, so that an excessive variation of
magnetic coenergy occurs.
In addition, the result may be influenced by the variation of the mesh of
the domain, since the geometry changes when the plunger is in the two
different positions. In order to avoid this numerical error, it is useful to define
the regions involved in the plunger movement, as shown in Figure 5.10.
When the plunger occupies these regions, they are considered as iron, assign-
ing the corresponding magnetic property. Conversely, when the plunger
does not occupy these regions, they are considered as air.
5.4 Example
This section illustrates the analysis of the cylindrical magnetic device shown
in Figure 5.11, together with its main dimensions. The coil is formed by Nt =
1000 turns, whose diameter is dc = 0.9 mm. The coil is represented by the
rectangle with sides 90 mm and 17.5 mm. The nonmagnetic driver, which
determines the lateral gap between the core and the plunger, is t = 0.5 mm.
When the plunger is in its lower position, the air-gap length is gmax = 15 mm,
while when the plunger is in its higher position, the air-gap length is gmin =
1 mm.
Since the problem is axial-symmetric, the section of the magnetic device
of Figure 5.8 is analyzed. The current density Jϑ is fixed as source, and the
magnetic vector potential Aϑ represents the field solution. The boundary
condition Aϑ = 0 is fixed on the boundary of the domain, as illustrated in
Figure 5.8. The mild steel B-H curve is used for the core and the plunger.
Cylindrical Magnetic Devices 89
15
150 100 90 1 – 15 mm
22
0.5 45
60
100
125
FIGURE 5.11
Main dimension of the cylindrical magnetic device.
FIGURE 5.12
Mesh of the domain: initial mesh (a) and refined mesh (b).
Figure 5.12 shows the mesh of the device: Figure 5.12(a) shows the initial
mesh, while Figure 5.12(b) shows the refined mesh. The mesh was mainly
refined in the air-gap region, where the highest field gradients take place.
The field solution consists of the knowledge of the magnetic vector poten-
tial Aϑ. The equipotential lines are shown in Figure 5.13(a). The flux density
90 Electrical Machine Analysis Using Finite Elements
FIGURE 5.13
Equipotential lines of Aϑ (a) and flux density vectors (b).
vector components are computed from the magnetic vector potential Aϑ, as
reported by Equation (5.28). The vector field B is shown in Figure 5.13(b).
As an example, three configurations of the magnetic device are analyzed,
corresponding to three air-gap lengths: g = 1 mm, g = 5 mm, and g = 15 mm.
The behavior of the z-axis component of the flux density in the middle of
the air-gap is shown in Figure 5.14, with a source current equal to i = 5 A.
The dependence of the flux density value on the air-gap length is evident.
When g = 1 mm, the effect of the plunger and core edges is also manifest.
FIGURE 5.14
Behavior of the flux density component Bz in the middle of the air-gap g (with i = 5 A).
Cylindrical Magnetic Devices 91
TABLE 5.1
Computation on the Solved Structure (with a
current i = 5 A)
g = 15 mm g = 5 mm g = 1 mm
∫τ Cu
A ϑ dτ 2.023⋅10–6 4.123⋅10–6 5.204⋅10–6
∫ J A dτ
τ
ϑ ϑ 6.422 13.09 16.52
∫ BHdτ
τ
6.415 13.16 16.79
∫∫
B
TABLE 5.2
Comparison between Analytical and Finite Element (FE) Results
g = 15 mm g = 5 mm g = 1 mm
Analytical FE Analytical FE Analytical FE
Bg (T) 0.415 0.407 1.219 1.097 5.435 (!) 1.87
λ (Vs) 0.659 1.284 1.939 2.618 — 3.304
Wm (J) 1.648 3.217 4.847 6.119 — 3.978
Fz (N) 217.5 169.4 1881 881.5 — 2086
92 Electrical Machine Analysis Using Finite Elements
FIGURE 5.15
Flux linkage as a function of the current, with three air-gap lengths.
TABLE 5.3
Computation of the Force by Means of Maxwell’s Stress Tensor
Force (N)
g = 5 mm i=5A Line Points 100 Points 200 Points 400
surface higher than the plunger surface. Finally, the curve C completely
contains the plunger, and the effects of the lateral surface of the plunger are
also considered in the force computation. The numerical integration is car-
ried out with a different number of points. It is always a good rule to check
the influence of the number of points, to avoid errors due to numerical
integration.
By comparing the results of Table 5.3, obtained on the three different lines
of integration, the following conclusions arise. The attractive force takes
place essentially on the upper surface of the plunger: in fact the force com-
puted on the line A is almost 90% of the total force, considered equal to that
computed on the line C. A good approximation of the force, almost 94% of
the total force, is obtained computing the force on the line B.
The force acting on the plunger for different air-gap lengths is reported in
Table 5.3. These results refer to the integrals along the line C, with 400 points.
The force is also computed by means of the virtual work principle, in the
same conditions, i.e., i = 5 A and g = 5 mm. Two further field solutions are
carried out, considering a constant current i = 5 A, and two variations of the
air-gap length of ∆g = 0.5 mm, the first positive and the second negative.
The values obtained from the field solutions are reported in Table 5.4. Using
Equation (5.46) yields
Fz = −
( ) (
Wm′ i , g + ∆g − Wm′ i , g − ∆g ) =−
6.5799 − 7.477
= 898 N
2 ∆g 2 · 0.5 · 10−3
i = const
which coincides with the force computed by means of Maxwell’s stress tensor.
94 Electrical Machine Analysis Using Finite Elements
TABLE 5.4
Computation of Some Quantities Around g = 5 mm
(current i = 5 A)
g = 4.5 mm g = 5 mm g = 5.5 mm
∫τ Cu
A ϑ dτ 4.278 · 10–6 4.123 · 10–6 3.961 · 10–6
∫ J A dτ
τ
ϑ ϑ 13.58 13.09 12.58
∫ BHdτ
τ
13.63 13.16 12.59
∫∫
B
References
1. D.C. White and H.H. Woodson, Electromechanical Energy Conversion, John Wiley &
Sons, New York, 1959.
2. G. Xiong and S.A. Nasar, Analysis of Field and Forces in a Permanent Magnet
Linear Synchronous Machine Based on the Concept of Magnetic Charge, IEEE
Trans. on MAG, vol.25, no.3, pp. 2713–2719, 1989.
3. R. Akmese and J.F. Eastham, Dynamic performance of a brushless dc tubular
drive system, IEEE Trans, MAG, vol.25, no.5, pp. 3269–3271, 1989.
4. J. Hur, S.B. Yoon, D.Y. Hwang, and D.S. Hyun, Analysis of PMLSM Using Three
Dimensional Equivalent Magnetic Circuit Network Method, IEEE Trans, MAG,
vol.33, no.5, pp. 4143–4145, 1997.
5. I. Boldea and S.A. Nasar, Linear Electric Actuators and Generators, Cambridge
University Press, Cambridge, UK, 1997.
6. N. Bianchi, S. Bolognani, F. Tonel, Design Considerations for a Tubular Linear
PM Motor for Servo Drives, in Proc. EPE-PEMC 2000 International Conference,
Kosice, Slovak Republic, 5–7 September 2000.
7. N. Bianchi, Analytical Computation of Magnetic Fields and Forces of a Tubular
PM Linear Servo Motor, in Proc. di IEEE IAS Annual Meeting, Roma, Italy,
8–12 October 2000.
8. J.F. Gieras and Z.J. Piech, Linear Synchronous Motors. Transportation and Auto-
mation Systems, CRC Press, London–New York, 2000.
6
The Single-Phase Transformer
The aim of this chapter is to describe the finite element analysis of a low-
power, single-phase transformer. We start by defining the equivalent circuit
of the transformer; then we go on to describe the procedure to obtain its
parameters.
95
96 Electrical Machine Analysis Using Finite Elements
(a) (b)
FIGURE 6.1
Single-phase core-type (a) and shell-type (b) transformers.
B B B
B' B'
(a) (b) (c)
FIGURE 6.2
Core-type transformer: simplification of the structure.
B B B
B' B'
(a) (b) (c)
FIGURE 6.3
Shell-type transformer: simplification of the structure.
geometric structure and the flux lines. The section drawn in Figure 6.3(a)
can be reduced to the section part drawn in Figure 6.3(b) and then to that
drawn in Figure 6.3(c).
dλ 1 di1 di 2
v1 = dt = L1 dt + M dt
(6.1)
dλ 2 di di
v2 = = M 1 + L2 2
dt dt dt
where v1, v2, i1, i2, λ1, and λ2 are the voltages, currents, and flux linkages at
the two terminal pairs. The sign notation is such as the positive direction of
the currents is against the voltage in both terminal pairs. Finally, L1, L2, and
M are the self- and mutual inductances of the transformer. With the assump-
tion that there is no saturation, these inductances are constant.
The flux linkages can be expressed as
λ 1 = L1i1 + Mi 2
(6.2)
λ 2 = Mi1 + L 2 i 2
For the sake of convenience, the self-inductances L1 and L2 are split into
two addenda, which are
L1 = L1σ + L1m
(6.3)
L2 = L2 σ + L2 m
L1m M
a= = (6.5)
M L2 m
v′1 = a v′′2
i2 (6.6)
i m = i1 + a
98 Electrical Machine Analysis Using Finite Elements
+ + im + +
i1 i2
L1m
v1 v'1 v"2 v2
_ _ _ _
FIGURE 6.4
Equivalent electric circuit of the real transformer, T representation.
L'2σ
a' : 1
+ + +
i1 i'm i2
_ _ _
FIGURE 6.5
Equivalent electric circuit, Γ′ representation .
The choice of how the inductances L1 and L2 are separated in the two
addenda is arbitrary. It is sufficient that Equation (6.2) is satisfied and that
the inductances of the model are positive or, at least, null. The latter condition
limits the possible values of the transformation ratio:
M L
≤a≤ 1 (6.7)
L2 M
L′1m = L1
L 1L 2 − M 2 (6.8)
L′2 σ =
L1
L"1σ
a" : 1
+ +
i1 i"m i2 +
_ _ _
FIGURE 6.6
Equivalent electric circuit, Γ″ representation.
M2
L′′1m =
L2
(6.9)
L′′ = L1L 2 − M
2
1σ
L2
N1
a= (6.10)
N2
2
L 1 m N1
= (6.11)
L 2 m N2
TABLE 6.1
Transformations of the Inductive Parameters of the Transformer
From T to Γ′ Representation From T to Γ″ Representation
L 1 m2
L′1 m = L 1 σ + L 1 m L′′1 m =
L1m + a2 L 2 σ
primary winding is fed at the nominal frequency, with voltage equal to the
nominal voltage, measuring the primary current and the induced secondary
voltage. Similarly, in the simulation, the primary winding current is imposed
as the field problem source and the corresponding voltages are computed
in the two windings. It is also possible to compute the variation of the
induced voltage at no-load, corresponding to different primary current val-
ues. Of course, this makes sense when iron saturation occurs.
A current source I1o is forced in the primary winding. For the sake of
simplicity, the latter is simplified as a unique conductive bar, as shown in
Figure 6.2(c) and Figure 6.3(c), fed by the total current:
We must not forget that the current flows in the conductors and then a
free distribution of the current in the equivalent bar is not realistic. It has to
be imposed that the current is uniformly distributed in the bar. Of course,
if the current assigned in the simulation is constant, a magnetostatic problem
arises; then the problem of a nonuniform distribution of the current does
not exist. This kind of simulation should be preferred. In this case, the
constant current I1o denotes the maximum value of the sinusoidal current
waveform.
The Single-Phase Transformer 101
1 1
wm = B ⋅ H = µH 2 (6.12)
2 2
∫
1
Wm = 4L Fe µH 2 dS (6.13)
S 2
102 Electrical Machine Analysis Using Finite Elements
∫ ∫
1 1
Wm = 4L Fe J ⋅ A dS = 4L Fe J z A z dS (6.14)
S 2 S 2
∫
Φo = L Fe B ⋅ n dl ,
l
in case of core-type transformer; (6.15)
∫
Φo = 2 L Fe B ⋅ n dl ,
l
in case of shell-type transformer.
The ends of the lines are generally chosen within the windings.
Λ1o = N1 Φo (6.16)
1
Λ1o = N1L Fe
S Cu1 ∫ S Cu 1+
A z dS −
∫S Cu 1−
A z dS in case of core-type transformer;
(6.17)
1
Λ1o = 2 N1L Fe
S Cu1 ∫ S Cu 1+
A z dS
in case of shell-type transformer.
The Single-Phase Transformer 103
The same computation is carried out for the secondary winding. The
relationships are the same, only subscript 2 will be used in place of 1.
E1o = ω Λ1o
(6.18)
E2o = ω Λ 2o
We wish to point out that the simulation is carried out considering the
maximum value of the sinusoidal current, say I1o. Consequently, the quan-
tities Φo, Λ1o, and E1o, as well as the secondary winding quantities represent
the maximum value that is reached. Since the problem has been assumed to
be linear, the RMS values of these quantities are obtained by dividing the
maximum value by 2. This is not possible when the iron saturation occurs.
Λ1o
L1 =
I1o
(6.19)
Λ
M = 2o
I1o
2 Wm
L1 = (6.20)
I1o 2
For obtaining the self-inductance L2, the secondary winding is fed by the
current I2o, and the primary winding is open-circuited (that is, it assigns a
null conductivity of the primary winding equivalent bars). Then L2 is given
by the ratio between the secondary flux linkage Λ2o and the current I2o.
104 Electrical Machine Analysis Using Finite Elements
ω Λ1o
dc magnetizing
I characteristic
I1o I1o(rms)
Corresponding
current
(a) (b)
FIGURE 6.7
Drawing of the current waveform (a) and the magnetizing characteristic of the transformer (b).
The Single-Phase Transformer 105
where γFe is the specific weight of the iron(more or less γFe = 7700 kg/m3).
The value BM must be the maximum flux density amplitude, reached in the
volume dτ. This is the reason for using the maximum value of the current
I1o in the (magnetostatic) simulation. The power α of BM in Equation (6.21),
which ranges from 1.6 to 2.2 in Steimnetz’s formula, can be approximated
by 2.
Then the iron losses are given by
pFe =
∫
τ Fe
ps ,FeB M 2 γ Fedτ = 4L Fe
∫
S Fe
ps ,FeB M 2 γ FedS (6.22)
6.3.5 Example
Let us consider a shell-type transformer, characterized by the following
nominal data:
106 Electrical Machine Analysis Using Finite Elements
25
125 71 75
25
25 25 50 25 25
150
20
50
FIGURE 6.8
Main dimensions of the transformer.
The aspect and the dimensions of the iron core are reported in Figure 6.8.
The rated currents of the secondary winding are
Sn
I2 n = = 8.33 A , at V2n = 24 V
V2 n
Sn
I2 n = = 11.1 A , at V2n = 18 V
V2 n
Assuming an efficiency η = 90% and a unity power factor, the rated currents
of the primary winding are
Sn
I1 n = = 1 A , at V1n = 220 V
ηV1n
The Single-Phase Transformer 107
Sn
I1 n = = 1.74 A , at V1n = 127 V
ηV1n
Such values have been used for the choice of the diameters of the coils,
fixing a current density about 2 A/mm2. Each coil is characterized by three
terminals, as shown in Figure 6.9. The number of conductors have been
chosen according to the iron flux density BM = 1.1 T in the iron core. In
addition, in the choice of the number of secondary winding turns, a 4%
voltage drop under full load has been assumed. With reference to Figure 6.9,
the following number of turns and diameters have been chosen for the
secondary winding:
The following number of turns and diameters have been chosen for the
primary winding:
+ +
N1' + +
V1'
N2' V2'
_ _
V1 V2
N1'' N2''
_
_
FIGURE 6.9
Sketch of the transformer windings.
108 Electrical Machine Analysis Using Finite Elements
75
62.5
35 37.5
8.7 9.3
1.5 0.5
FIGURE 6.10
Section of the transformer that is used in the following simulations.
Figure 6.11 shows the flux plot, obtained by drawing the equipotential
lines of the magnetic vector potential.
Figure 6.12 shows the flux density along the line AA″ (see Figure 6.11)
inside the magnetic iron. The average value of the flux density is approxi-
mately equal to 1.125 T. The flux density is not exactly constant, due to the
different length of the magnetic paths within the iron core. The total magnetic
energy stored in the transformer, using Equation (6.13) or Equation (6.14),
results in
Wm = 58.46 mJ
The Single-Phase Transformer 109
SCu2 SCu1
A A" A'
N1 I1o
2
FIGURE 6.11
Flux plot.
1.15
1.14
Flux density (T)
1.13
1.12
1.11
1.1
0 5 10 15 20 25
(A) Length (mm) (A")
FIGURE 6.12
Flux density within the iron core.
The integral of the magnetic energy density only over the iron core results
in
Wm(Fe) = 58.42 mJ
which represents almost the totality of the stored magnetic energy. Con-
versely, by applying Equation (6.14) to the iron core, the integral results null.
In fact this equation represents the energy of the transformer as work fur-
nished by the external sources. It is different from zero only in the coils of
the primary winding, where the current density is not null.
From the second part of Equation (6.15), the magnetic flux is
Φo = 2.7952 mWb
Then, from Equation (6.16), with N1 = 416 and N2 = 47, the flux linkages are
Λ1o = 1.1628 Vs
110 Electrical Machine Analysis Using Finite Elements
Λ2o = 0.1314 Vs
Λ1o = 1.1696 Vs
Λ2o = 0.1321 Vs
which differ from the previous values just of 0.5%. Finally the voltage
induced in the windings are
E1o = 367.4 V
E2o = 41.5 V
Since the system has been supposed to be linear, the RMS values of these
voltages result in
E1o(rms) = 220 V
E2o(rms) = 24.85 V
L1 = 11.69 H
M = 1.321 H
TABLE 6.2
Values Corresponding to the Field Analysis
I1o N1I1o/2 BFe Λ1o E1o
(mA) (A) (T) (Vs) (V)
25.0 5.2 0.36 0.3774 118.5
50.0 10.4 0.73 0.7542 237.0
75.0 15.6 0.94 0.9780 307.2
100.0 20.8 1.05 1.0977 344.9
112.5 23.4 1.09 1.1361 356.9
125.0 26.0 1.12 1.1684 367.1
137.5 28.6 1.15 1.1976 376.2
150.0 31.2 1.17 1.2202 383.3
FIGURE 6.13
Current-voltage characteristics.
described in Section 6.3.3. Figure 6.14 shows the sinusoidal voltage wave-
form, corresponding to the rated value of E1o(rms) = 220 V, and the correspond-
ing current waveform. The distortion of the current from the sinusoidal
waveform is noticeable.
250 62.5
Current
200 50.0
150 37.5
100 25.0
50 12.5
0 0
0 2 4 6 8 10 12 14 16 18 20
Time (ms)
FIGURE 6.14
Voltage and current waveforms corresponding to the rated voltage E1o(rms) = 220 V.
C D F
FIGURE 6.15
Standardized lamination for a low-power shell-type transformer (a) and flux plot in half a
section of the transformer (b).
TABLE 6.3
Effect of the Air-Gap on the Magnetic Quantities
g BFe
∫ S Cu 1+
A z dS
∫
S Cu 2 +
A z dS
∫ J A dS
S
z z
i m = i1 +
i2
= i1 +
(
− N1 / N 2 i 1
=0
) (6.23)
a N1 / N 2
Thus the two flux linkages (6.2) correspond to the leakage flux linkages
λ 1σ = L1σ i 1
(6.24)
λ 2 σ = L 2 σ i 2
∫
1
Wm = 4l average µH 2 dS (6.25)
S 2
where S is the total surface of the domain and laverage indicates the average
length of a conductor (half a length of one turn). Let us highlight that the
effect of the end windings has been considered the same of the part of the
winding in the section under study. Such an approximation is reasonable,
since the fields are mainly localized between the two coils.
Alternatively, the magnetic energy is computed as the energy that is fur-
nished from the source to the system, as
∫
1 1
Wm = 4 l average J1z A z + J 2 z A z dS (6.26)
S 2 2
1
Λ1sc = N1l average
S Cu1 ∫ S Cu 1+
A z dS −
∫
S Cu 1−
A z dS , with core-type transformer
1
Λ1sc = 2 N1l average
S Cu1 ∫ S Cu 1+
A z dS , with shell-type transformer
(6.27)
E1sc = ω Λ1sc
(6.28)
E 2 sc = ω Λ 2 sc
Since the field problem is linear, the quantities I1sc, I2sc, Λ1sc, Λ2sc, and E1sc,
E2sc are in proportion. The maximum values are achieved by multiplying the
RMS values by the square root of two.
Λ1sc
L1σ =
I1sc
(6.29)
Λ
L2 σ = 2 sc
I 2 sc
2
∫ ∫
1 1
L1σ = 2 average
4l J1z A z dS − J1z A z dS
I1sc Scu 1+ 2 Scu 1− 2
(6.30)
2
∫ ∫
1 1
L2 σ = 2 average
4l J 2 z A z dS − J 2 z A z dS
I 2 sc Scu 2 + 2 Scu 2 − 2
The Single-Phase Transformer 117
2
∫
1
L1σ = 4l average J1z A z dS
I1sc 2 Scu 1 2
(6.31)
2
∫
1
L2 σ = 4l average J 2 z A z dS
I 2 sc 2 Scu 2 2
that correspond to those given in Equation (6.29), observing that in the core-
type transformer
N1 I 1 N2 I 2
J 1z = J2z =
4S Cu1 4S Cu 2
N1 I 1 N2 I 2
J 1z = J2z =
2 S Cu1 2 S Cu 2
The total leakage inductance L′2σ (or alternatively L″1σ) is obtained, apply-
ing the equivalences of Table 6.1.
Ft = l average
∫ (µ H H ) dl
l
o t n
(6.32)
∫ ( )
µo
Fn = l average H n 2 − H t 2 dl
l 2
The x-axis force is the sum of the forces Ft of the vertical lines with the
forces Fn of the horizontal lines. Similarly, the y-axis force is the sum of the
forces Ft of the horizontal lines with the forces Fn of the vertical lines.
Due to the symmetry with respect to the axis BB′ that has been used in
the simulations, the forces refer to half a coil, and then the overall y-axis
force component is null while the overall x-axis force component is obtained
by doubling the force obtained in half a winding.
Alternatively, the force on the coils can be computed as Lorentz’s force.
According to Lorentz’s specific electric force (1.84) and the definition of the
118 Electrical Machine Analysis Using Finite Elements
current density vector (1.60), the x-axis and the y-axis that force components
in the half coil are
Fx = − l average
∫
S Cu 1
J zB y dS
(6.33)
Fy = l average
∫ S Cu1
J zB x dS
the infinite. This is because of the 2D problem. Then, the bar Scu2+
links the flux produced by the primary currents, while Scu2– does not
link any flux, so that only Scu2+ is interested by eddy currents. It is
thus compulsory to impose that the sum of the currents induced in
Scu2+ and Scu2– be null, which is Icu2– = –Icu2+. A method to impose this
condition is to declare that the bars Scu2+ and Scu2– are interested by
a total current in any time. This is equivalent to imposing that Icu2+ +
Icu2– = 0.
6.4.4 Example
Let us consider the transformer described in Section 3.6, with the aim of
computing the leakage inductances.
The iron relative permeability is fixed to the constant value µFe = 5500.
The current I1sc = 1 A is forced in the primary winding, and then the total
current in the equivalent primary bar of the analyzed section part is
N1I1sc
= 208 A
2
N2 I 2 sc NI
= − 1 1sc = −208 A
2 2
which is a current I2sc = –8.851 A. The flux plot is shown in Figure 6.16.
The flux density within the windings assumes the behavior of Figure 6.17(a)
with a maximum value B = 7.44 mT. This is in accord with the theoretical
behavior, as reported in Figure 6.17(b). For the sake of comparison, the
maximum flux density computed analytically results in
FIGURE 6.16
Flux plot during short-circuit operation.
120 Electrical Machine Analysis Using Finite Elements
FIGURE 6.17
Flux density behavior within the two windings.
N1I1sc
B = µo = 7.47 mT
h coil
Wm = 2.547 mJ
and results are stored almost completely in the primary winding (45.59%),
secondary winding (42.87%), and the surrounding space (11.53%). In the
iron, it is lower than 0.01%.
From Equation (6.26) it results always that
Wm = 2.547 mJ
L1σ = 1.9533 mH
2 Wm
Lσ = = 5.094 mH
I12
N1 2 s s
Lσ = µo 2 l average 1 + s + 2 = 5.365 mH
h coil 3 3
Fx = 0.391 N
1 2 l average
( )
2
Fx = µo N1I1sc = 0.413 N
2 h coil
which is slightly higher than that obtained from the field solution.
Note: the computation of the x-axis force component on the secondary coil
is illustrated as follows. A line surrounding the coil is fixed as shown in
Figure 6.18. The x-axis force Fx, is computed in the different paths as
122 Electrical Machine Analysis Using Finite Elements
Q R
P S
x
Secondary coil
FIGURE 6.18
Line surrounding the secondary coil to compute the x-axis force.
BxHx − B yH y
FxPQ = 4l average
∫
l PQ 2
dl = 1.22 ⋅ 10−3 N
FxQR = 4l average
∫
l QR
B x H ydl = 6.64 ⋅ 10−3 N
BxHx − B yH y
FxRS = 4l average
∫
l RS 2
dl = 383.1 ⋅ 10−3 N
CONST pi = 3.14156
a$ = CHR$(34)
DIM f(205), i(205), ang(205), v(205)
Imax = 0
FOR k = 1 TO 201
theta = 2 * pi * (k - 1)/ 200
ang(k) = theta
v(k) = Vm * SIN(theta)
f(k) = Fm * COS(theta)
i(k) = current(f(k))
IF i(k) > Imax THEN Imax = i(k)
NEXT k
isum = 0
vsum = 0
FOR k = 1 TO 200
isum = isum + i(k) ^ 2
vsum = vsum + v(k) ^ 2
NEXT k
Vrms = SQR(vsum/ 200)
Irms = SQR(isum/ 200)
FUNCTION current(FlxLnk)
‘sign evaluation
sign = 1
IF FlxLnk < 0 THEN
sign = -1
FlxLnk = - FlxLnk
END IF
Nb = 9
DIM FluxVect(Nb)
DIM CurrVect(Nb)
FluxVect(1) = 0: CurrVect(1) = 0
FluxVect(2) = 0.3774: CurrVect(2) = .025
FluxVect(3) = 0.7543: CurrVect(3) = .050
FluxVect(4) = 0.9780: CurrVect(4) = .075
FluxVect(5) = 1.0977: CurrVect(5) = .100
FluxVect(6) = 1.1361: CurrVect(6) = .1125
FluxVect(7) = 1.1684: CurrVect(7) = .125
FluxVect(8) = 1.1976: CurrVect(8) = .1375
FluxVect(9) = 1.2202: CurrVect(9) = .150
124 Electrical Machine Analysis Using Finite Elements
‘Interpolation (linear)
FOR i = 0 TO (Nb - 1)
IF (FlxLnk >= FluxVect(i) AND FlxLnk < FluxVect(i + 1)) THEN
Cf =(CurrVect(i+1)-CurrVect(i))/ (FluxVect(i+1)-
FluxVect(i))
CurrAux = Cf * (FlxLnk - FluxVect(i)) + CurrVect(i)
END IF
NEXT i
‘Sign correction
END FUNCTION
References
1. E. Arnold, Die Wechselstromtechnick. Die Transformatoren, Vol. 2, Julius Springer,
Berlin, 1904.
2. R. Richter, Elektrische Maschinen. Die Transformatoren, Vol. 3, Julius Springer,
Berlin, 1932.
3. A.S. Longsdorf, Theory of Alternating Current Machinery, McGraw-Hill, New
York, 1937.
4. T. Bödefeld and H. Segueuz, Elektrische Maschinen, Springer-Verlag, Wren, 1942.
5. A. Carrer, Macchine Elettriche. Parte prima: Trasformatori, Editrice Universitaria
Levrotto & Bella, Torino, 1954.
6. S. Ramo, J.R. Whinnery, and T. Van Duzer, Fields and Waves in Communication
Electronics, John Wiley & Sons, New York, 1965.
7. G.R. Slemon, Magnetoelectric Devices: Transducers, Transformers and Machines,
John Wiley & Sons, New York, London, Sydney, 1966.
8. A.E. Fitzgerald, C. Kingsley, Jr., and S.D. Umans, Electric Machinery, McGraw-
Hill, New York, 1983.
9. I.J. Nagrath and D.P. Kothari, Electric Machines, Tata McGraw-Hill Publishing
Company Limited, New Delhi, 1985.
7
Single-Phase Variable Reactance
This chapter deals with the finite element analysis of a single-phase variable
reactance. In particular the dependence of the reactance on the air-gap length
is analyzed. The results obtained by the finite element method are compared
with those computed analytically.
125
126 Electrical Machine Analysis Using Finite Elements
FIGURE 7.1
Magnetic structures of single-phase reactance.
B B B
A A'
B'
B' B'
(a) (b) (c)
FIGURE 7.2
Reduction of the domain of the structures of Figure 7.1 using their symmetry.
1 1
wm = B ⋅ H = µH 2 (7.1)
2 2
∫
1
Wm = L Fe µH 2 dS (7.2)
S 2
128 Electrical Machine Analysis Using Finite Elements
∫
1
Wm = L Fe J z A z dS (7.3)
S 2
which results different from zero only on the surfaces of the coils.
1
Λ = Nt L Fe
S Cu ∫
S Cu +
A z dS −
∫
S Cu −
A z dS
(7.4)
The voltage at the terminal pairs of the reactance coils is obtained as the
geometric sum of the voltage given in Equation (7.5) with the voltage drop
on the resistance, computed analytically.
7.2.2.5 Self-Inductance
The value of the self-inductance is achieved by dividing the flux linkage by
the corresponding current, which is
Λ
L= (7.6)
I
2 Wm
L= (7.7)
I2
Single-Phase Variable Reactance 129
P O
M N M N
x
(a) (b)
FIGURE 7.3
Lines along which the attractive force is computed.
In Equation (7.8) there are the normal components of the force along the
segments MN and OP:
FyMN = − L fe
∫
MN
µo
2
(
H y 2 − H x 2 dl )
(7.9)
FyOP = L fe
∫
µo
OP 2
(
H y 2 − H x 2 dl )
and the tangential components of the force along the segments NO and PM:
FyNO = L fe
∫ (µ H H ) dl
NO
o y x
(7.10)
FyPM = − L fe
∫ (µ H H ) dl
PM
o y x
Since the magnetic field assumes a negligible value outside the structure,
the main addendum is the normal force on the segment MN. The attractive
force can be estimated from the integration of the magnetic pressure on the
line MN only, as illustrated in Figure 7.3(b).
For computing the force, an alternative of Maxwell’s stress tensor is the
method of the virtual works. Let Wm(y) be the magnetic energy computed
with the movable part in the generic position y. The movable part is shifted
130 Electrical Machine Analysis Using Finite Elements
∆Wm ( y) W ( y + ∆y) − Wm ( y)
Fy = = m (7.11)
∆y i= const ∆y i = const
Of course, this result depends on the chosen value of the movement ∆y.
An improvement to the force computation is obtained considering the
changes of position +∆y and –∆y. The force yields
Wm ( y + ∆y) − Wm ( y + ∆y)
Fy = (7.12)
2 ∆y i = const
In this way, all the terms of the force due to the second derivative of the
energy are eliminated. This is verified by expanding the magnetic energy
around the position y via the Taylor series. Neglecting the upper-order
addenda, it is
dWm ( y) 1 d 2 Wm ( y) 2
Wm ( y + ∆y) = Wm ( y) + ∆y + ∆y +… (7.13)
dy 2 dy 2
dWm ( y) 1 d 2 Wm ( y) 2
Wm ( y − ∆y) = Wm ( y) − ∆y + ∆y +… (7.14)
dy 2 dy 2
By subtracting the second equation from the first one, it results that
E Magnetizing
Erms characteristic
ωΛ dc magnetizing
characteristic
I Irms
FIGURE 7.4
Magnetizing characteristics.
7.3 Example
A single-phase variable reactance, designed for laboratory tests, is analyzed.
It has been designed for giving rise to resonance phenomena, when con-
nected with capacitor loads. The reactance is changed so that a very high
voltage is applied to the capacitor load under test, even though the voltage
source is lower. The air-gap length is adapted from time to time depending
on the capacitor load, in order to obtain the resonance.
According to a capacitor load in the range C = 10 ÷ 200 nF, the variation
of the inductance is in the range L = 50 ÷ 1000 H.
The rated data of the single-phase reactance are as follows:
1. f = 50Hz — frequency
2. Vn = 50 kV — RMS voltage in resonance conditions
3. In = 3.15 A — maximum RMS current
4. Sn = 160 kVA — apparent power
5. magnetic material — grain-oriented TERNI M5T30, with thickness
0.3 mm and specific iron losses ps,Fe = 0.5 W/kg (at BM = 1 T, f = 50 Hz)
6. maximum flux density between 1.4 and 1.5 T (to avoid saturation)
7. the reactance is in oil, so that a class A insulation is chosen
The leg is almost circular, with SFe = 4253 mm2 and external diameter D =
249.6 mm (the stacking factor is kstk = 0.93, and the utilization factor, due to
the steps of the leg, is 0.935). The distance between the iron leg and the
winding is 25 mm with two insulating layers of 5 mm thickness each and
three oil channels. The distance between the coil and the iron yoke is 100 mm,
and between the two external diameters of the coils it is 110 mm. The main
dimensions of the reactance are reported in Figure 7.5. The movable part is
chosen with a rectangular section with a width equal to the diameter D of
the leg and a height equal to 170.4 mm.
132 Electrical Machine Analysis Using Finite Elements
875.9
170.4
t
100
39.4
30
627.6
100
41.7 25
243.3
742.5
FIGURE 7.5
Main dimensions of the reactance.
2
Nt
2 µ S
L= ≈ Nt 2 o Fe (7.16)
2t l Fe ,average 8t
+
µ oS Fe µ Feµ oS Fe
L L L M L
M
(a) (b)
FIGURE 7.6
Connection between the two coils.
µ oS Fe
Ls ≈ 2(L + M) ≈ 4L ≈ Nt 2 (7.17)
2t
When the two coils are connected in parallel, as in Figure 7.6(b), the total
inductance is reduced to a quarter, that is
1 µ S
Lp ≈ (L + M) ≈ L ≈ Nt 2 o Fe (7.18)
2 8t
The ratio between the maximum inductance, obtained with the minimum
air-gap length and with series-connected coils, and the minimum inductance,
obtained with the higher air-gap length and with parallel-connected coils,
becomes equal to 20. In conclusion, with the parallel connection, the resonance
is obtained with capacitor loads in the range C = 40 ÷ 200 nF, while with the
series connection, the resonance is obtained with C = 10 ÷ 50 nF.
is halved with the series connection. The number of turns has been chosen
equal to 3780 for each coil, which is a total number of turns Nt = 7560.
The maximum value of the current occurs when the reactance is in reso-
nance with the capacitive load C = 200 nF with a current I = ωCVn = 3.15 A,
to which corresponds the apparent power Sn = 160 kVA. Such resonance is
obtained with the lowest value of inductance, which is with the parallel
connection. The current in the conductors is then halved, resulting in 1.575 A.
The diameter of the conductor corresponds to dCu = 1.093 mm with a
double insulating thickness (UNEL 01722-3-4).
Each coil is divided in three elementary coils of 1260 turns each. The turns
are composed of 35 layers of 36 turns each. Each layer is insulated with a
paper sheet of thickness 0.1 mm. The dimensions of each elementary coil
result: a 41.7 mm width and a 39.4 mm height, as shown in Figure 7.5.
7.3.3 Analysis
The analysis of the reactance is reduced to a 2D analysis on the plane (x,y).
To have a planar symmetry, the circular shape of the leg is modeled as it
would be rectangular, but preserving the same cross area. This is shown in
Figure 7.7. The circular leg has the diameter D = 249.6 mm, then keeping
the width of the rectangular leg equal to D, the z-axis length becomes LFe =
170.4 mm.
The height of the movable part has to be adapted as well, in order to refer
the magnetic quantities at the same machine z-axis length LFe. Such a height
becomes equal to the diameter D, as shown in Figure 7.7. Finally, with the
aim of considering the effective iron volume together with the effect of the
squared edges of the yoke, the horizontal length of the movable part is
increased, as shown by the dotted line in Figure 7.7.
With the required simplifications, the 2D analysis allows a good estimation
of the reactance behavior, with the advantage of a faster analysis.
D D
FIGURE 7.7
Equivalent structure of the reactance to apply the planar symmetry.
Single-Phase Variable Reactance 135
FIGURE 7.8
Mesh of the section of the reactance.
We will carry out the analysis on the whole section of the machine of Figure
7.5. However, the structure presents a symmetry axis, corresponding to the
BB′ axis of Figure 7.2(c). Then, only half a structure could be analyzed,
assigning Neumann’s boundary condition on the BB′ axis.
Since the magnetic field strength is higher in the air-gap, the mesh is
increased in this zone, as shown in Figure 7.8. The subdivision of the struc-
ture is carried out using some 2000 triangular elements. This choice proved
to be appropriate to obtain accurate results: a 20% increase of the number
of the finite elements generates a variation of the magnetic energy in the
field solution lower than 0.3%.
The magnetic B–H curve Terni M5T30 is used for the iron core.
Various simulations are carried out with different values of air-gap length t
and current I. Since the elementary coils are schematized by means of conduc-
tive bars, the equivalent current in each coil is obtained as the product of the
current I by the number of turns of each elementary coil, which is Nt/6 = 1260.
Figure 7.9 shows the flux plot. The detail of Figure 7.9(b) highlights that
the flux lines are not confined within the air-gap, but they expand at the
extremity of the legs. The existence of such flux lines points out that, at the
same current (and then at the same MMF), the flux linkage is higher than
that analytically predicted with the assumption of flux lines that are only
normal to the air-gap. Consequently, the inductance is higher. It is then
necessary to verify that the inductance assumes the required values, espe-
cially when the air-gap length is maximum.
Figure 7.10 shows the flux density magnitude in the middle of the air-gap,
corresponding to different values of air-gap length and current. The flux
density is drawn as a function of the distance from the symmetry axis of the
reactance. Figure 7.10(a) corresponds to an air-gap t = 1.5 mm and a current
I = 0.154 A. Figure 7.10(b) corresponds to an air-gap t = 9 mm and a current
I = 3.19 A.
136 Electrical Machine Analysis Using Finite Elements
(a) (b)
FIGURE 7.9
Flux plots.
B (T) B (T)
0.8 1.4
0.7 1.2
0.6 1
0.5
0.8
0.4
0.6
0.3
0.4
0.2
0.1 0.2
0 0
0 50 100 150 200 250 300 350 400 450 0 50 100 150 200 250 300 350 400 450
l (mm) l (mm)
FIGURE 7.10
Flux density in the middle of the air-gap as a function of the distance of the symmetry axis;
(a): t = 1.5 mm and I = 0.154 A; (b): t = 9 mm and I = 3.19 A.
In both figures, the edge effect is evident: in the central zone the value of
the flux density magnitude is essentially constant; near the edges it increases
about 10%.
Table 7.1 contains the values of the inductance obtained varying the air-
gap t. Low current is adopted to avoid saturation effect. The inductances are
computed by Equation (7.7) using the magnetic energy, given by Equation
(7.2) or Equation (7.3), since the problem is essentially linear. The same result
is obtained, however, if the inductances are computed by Equation (7.6)
using the flux linkages, given by Equation (7.4). The inductances with par-
allel-connected coil are obtained simply dividing by 4 the inductances with
series-connected coils.
For an easy comparison, in the same table the inductances computed
analytically are reported as well, as given by Equation (7.17) and Equation
Single-Phase Variable Reactance 137
TABLE 7.1
Values of the Self-Inductance, Computed at Low Current
t Ls Ls(analytic) Lp Lp(analytic)
(mm) (H) (H) (H) (H)
1.5 1047.0 1002.3 261.7 250.6
1.6 972.4 940.6 243.4 235.1
1.7 917.6 886.0 229.4 221.5
1.8 870.2 837.4 217.6 209.4
1.9 826.8 793.9 206.7 198.5
2.0 789.8 754.7 197.5 188.7
2.1 752.2 719.2 188.1 179.8
2.2 720.0 686.8 180.0 171.7
2.3 690.0 657.3 172.6 164.3
2.4 662.8 630.1 165.7 157.5
2.5 638.0 605.2 159.5 151.3
3.0 538.4 505.1 134.6 126.3
4.0 412.8 379.6 103.2 94.9
5.0 336.4 304.0 84.1 76.0
6.0 285.8 253.6 71.5 63.4
7.0 249.0 217.5 62.3 54.4
8.0 221.0 190.4 55.3 47.6
9.0 199.5 169.3 49.9 42.3
10.0 182.1 152.4 45.5 38.1
(7.18). As expected, the values obtained from the finite element method are
higher than those analytically computed. This is essentially due to the flux
lines external to the air-gap, as highlighted in Figure 7.9. The relative differ-
ence between the two inductances is as high as the air-gap increases. In fact,
with the higher values of t, the flux outer the air-gap increases.
Figure 7.11 illustrates graphically the results reported in Table 7.1.
1100
1000
900
800
Inductance (H)
700
600
500
400 Ls
Ls (analytical)
300
200
Lp
100
Lp (analytical)
0
2 3 4 5 6 7 8 9 10
Airgap length (mm)
FIGURE 7.11
Behavior of the inductance as a function of the air-gap, with constant current.
138 Electrical Machine Analysis Using Finite Elements
TABLE 7.2
Effect of the Saturation (t = 1.5 mm)
I Wm Λ Ls
(mA) (J) (Vs) (H)
58 1.76 60.71 1046.6
104 5.66 108.91 1047.2
150 11.78 157.12 1047.5
196 20.12 205.35 1047.7
242 30.68 253.58 1047.9
288 43.40 301.36 1046.4
334 58.19 348.46 1043.3
373 71.91 387.99 1040.2
419 89.41 432.30 1031.7
458 105.30 468.97 1024.0
504 123.10 506.53 1005.0
550 136.00 531.26 965.9
596 144.40 545.89 915.9
642 149.20 553.40 862.0
688 153.60 560.23 814.3
734 158.50 567.33 772.9
780 162.8 573.31 735.0
826 166.50 578.24 700.0
872 169.90 582.62 668.1
918 173.10 586.60 639.0
964 176.10 590.17 612.2
500 1000
Inductance
400 800
300 600
200 400
Flux linkage
100 200
0 0
0 200 400 600 800 1000
Current (mA)
FIGURE 7.12
Flux linkage and inductance as functions of the current (at t = 1.5 mm).
ΛM 600
Λ rms = = = 424.3 Vs
2 2
Nj
∑i
2 π/ω
ω
∫
1
I rms = i(t ) dt =
2 2
= 0.556 A
2π
j
0 Nj j=1
Λ rms
L* = = 763.7 H
I rms
140 Electrical Machine Analysis Using Finite Elements
FIGURE 7.13
Current waveforms, corresponding to a sinusoidal flux linkage with peak value (a): 400 Vs; (b):
500 Vs; (c): 550 Vs; (d): 600 Vs.
References
1. G. Kron, “Equivalent Circuits to Represent the Electromagnetic Field Quanti-
ties,” The Physical Rev., vol. 64, Aug. 1–15, 1943, pp. 126–128.
2. L.F. Blume, A. Boyajian, Transformer Engineering, John Wiley, New York, 1951.
3. G. Rago, Costruzioni Elettromeccaniche e Disegno, Sansoni, Firenze, 1968.
4. N. Bianchi, F. Dughiero, “Optimal Design Techniques Applied to Transverse-
Flux Induction Heating Systems,” in IEEE Trans. on Magnetics, vol. 31, n. 3,
Maggio 1995, pp. 1992–1995.
5. N. Bianchi, F. Dughiero, S. Lupi, “Design of Induction Heating Systems by
Optimisation of Field Shape,” in Proceedings of the International Induction Heating
Seminar, IHS ’98, Padova, 13–15 Maggio, 1998, pp. 413–423.
6. W.T. McLyman, Transformer and Inductor Design Handbook, CRC Press, Boca
Raton, FL, 2004.
8
Synchronous Generators
8.1 Introduction
The synchronous generator is composed of a fixed part, the stator, and a
rotating part, the rotor. In both of them, magnetic materials are used to guide
the magnetic flux, and windings are employed to carry the current. The
structure of a 2p = 2 pole synchronous generator is represented in Figure 8.1.
The rotor winding, called field winding or magnetizing winding, is fed by
a direct current, by means of dragging contacts, i.e., brushes and conductive
rings. This is the primary source of the main flux of the machine. The polar
axis in the middle of the rotor pole, sketched in Figure 8.1, is called the direct
axis, or d-axis. The interpolar axis is 90 electrical degrees in advance with
respect to the rotating direction and is called the quadrature axis, or simply
q-axis. The rotor rotates at constant mechanical angular speed ωm (an elec-
trical angular speed ω = pωm), giving rise to a rotating magnetic field.
The stator winding, also called armature winding, links a variable mag-
netic flux so that a variable EMF is induced. The three phases of the winding
are named a, b, and c. Each phase winding is spaced out of 120 electrical
degrees. They are sketched in Figure 8.1 by means of a single turn and are
identified by the axis normal to the turn itself. Observe that the orientation
of the turns agrees with their axis directions, according to the right-handed
screw advance (a rotation in direction of the positive current would cause
the right-handed screw to advance in the direction of the turn axis). Finally,
the rotor position with respect to the stator is pointed out by the electrical
angle ϑ, between the polar axis and the a-phase axis. The angle ϑ is ϑ = pϑm
where ϑm is the mechanical angle between rotor and stator.
When the armature windings are open-circuited, no current flows and the
induced EMF can be measured at the winding terminals. This is the no-load
141
142 Electrical Machine Analysis Using Finite Elements
Rotor coordinate
b-phase axis ϑr
Interpolar axis Polar axis
(or quadrature (or direct axis)
axis)
+a
ϑm Angular position
between
−c −b rotor and stator
Field winding ω a-phase axis
(or magnetizing Rotation
+b
winding)
+c
−a
Armature
winding
c-phase axis
FIGURE 8.1
Structure and references of the three-phase synchronous generator.
voltage of the generator. On the contrary, when the armature windings are
connected to the load, variable currents flow in the windings, with the same
frequency of the EMF. Such currents cause a magnetic field that rotates at
the same speed (called the synchronous speed) of the rotor speed. The
electromechanical conversion is obtained by the interaction between the
stator and rotor magnetic fields.
Neglecting the edge effects and assuming an identical behavior along the
generator, a planar symmetry is supposed: only the plane (x, y) is considered,
the magnetic vector potential and the current density have a z-axis compo-
nent only, which is A = (0, 0, Az) and J = (0, 0, Jz). Then the flux density and
magnetic field strength vectors have a component on the plane (x, y) only,
which is B = (Bx, By , 0) and H = (Hx, Hy , 0).
The machine length is chosen equal to the effective length of the iron LFe,
considering the stacking factor kstk, which considers the insulation among
the laminations (kstk = 0.95 – 0.98) and the possible ventilation channels (e.g.,
Ncv channels, whose wideness is lcv). Then the net length is
( )
L Fe = L − Ncv l cv k stk (8.1)
q d q d q d q d
O O O O
(a) (b) (c) (d)
FIGURE 8.2
Sections of the machine and periodicity.
assigned no matter what axis is chosen. However, the choice of the d- or the
q-axis is particularly suitable in the assignment of the boundary conditions.
In fact, special operating conditions allow Dirichlet’s or Neumann’s bound-
ary conditions to be assigned, as will be hereafter described.
Finally, some cases exist in which the analysis is reduced to half a pole, as
shown in Figure 8.2(d).
Dirichlet Az = 0 Az = 0
Az = 0
dAz
dAz =0
=0 dn
dn
dAz
=0
dn Az = 0 Az = 0 Az = 0
Neumann
FIGURE 8.3
Boundary conditions in the no-load simulation.
−b −b +a +a −c −c +a −c −c +b +b −a
FIGURE 8.4
Distribution in the slots of the three-phase windings: single-layer, nonchorded winding.
−b +a −c +a −c +b −a
FIGURE 8.5
Distribution in the slots of the three-phase windings: double-layer, one-slot chorded winding.
∫
1
A zdS (8.2)
Sq Sq
Let nq be the number of conductors per slot and npp be the number of
parallel paths of the machine. It follows that the total number of conductors
of the machine is npp times the number of series conductors. Considering
that the analysis domain is one pole only — see Figure 8.3(a) or Figure
8.3(b) — the j-th phase flux linkage is given by
Q/2 p
∑k ∫
nq 1
Λ j = 2 pL Fe jq A zdS j = a , b, c (8.3)
n pp q= 1
Sq Sq
where 2p is the number of poles and LFe is given by Equation 8.1; Q is the
number of the slots of the generator, so that Q/2p is the number of slots per
pole. Finally, kjq is the coefficient taking into account whether the conductors
in the q-th slot are of the j-th phase or not, as well as the conductor orientation.
Such coefficient assumes the following values.
kjq = 0 — if the coil side in the q-th slot does not belong to the j-th phase;
kjq = +1 — if the coil side in the q-th slot belongs to the j-th phase and
its orientation is positive with respect to the z-axis direction (leaving
the sheet with the references of Figure 8.1);
kjq = –1 — if the coil side in the q-th slot belongs to the j-th phase and
its orientation is negative with respect to the z-axis direction (going
into the sheet with the references of Figure 8.1).
kjq = 0 — if the coil sides in the q-th slot do not belong to the j-th phase;
kjq = +0.5 — if only one coil side in the q-th slot belongs to the j-th phase
and its orientation is positive with respect to the z-axis direction;
kjq = –0.5 — if only one coil side in the q-th slot belongs to the j-th phase
and its orientation is negative with respect to the z-axis direction;
kjq = +1 — if both two coil sides in the q-th slot belong to the j-th phase
and their orientation is positive with respect to the z-axis direction;
kjq = –1 — if both two coil sides in the q-th slot belong to the j-th phase
and their orientation is negative with respect to the z-axis direction.
146 Electrical Machine Analysis Using Finite Elements
TABLE 8.1
Values of kjq Referring to the
Distribution of Figure 8.4(a)
phase (j)
slot (q) a b c
0 –1 0
0 –1 0
+1 0 0
+1 0 0
0 0 –1
0 0 –1
TABLE 8.2
Values of kjq Referring to the
Distribution of Figure 8.5(a)
phase (j)
slot (q) a b c
1 0 –0.25 +0.25
2 0 –1 0
3 +0.5 –0.5 0
4 +1 0 0
5 +0.5 0 –0.5
6 0 0 –1
7 0 +0.25 –0.25
In other words, the magnitude of kjq specifies the relative filling of the q-th
slot by the j-th phase conductors. The sign of kjq considers the sign of the
scalar product A·t of the loop integral: the vector A has only the component
Az while the tangential unity vector t (which defines the turn’s orientation)
coincides with uz when a positive current flows in the z-axis direction, or
with –uz when the same positive current flows opposite to the z-axis direc-
tion. Thus, A·t = kjqAz is equal to +Az or to –Az.
For instance, with the single-layer winding of Figure 8.4(a), the values
reported in Table 8.1 are obtained, while with the double-layer winding of
Figure 8.5(a), the values reported in Table 8.2 are obtained. In the latter case,
the value 0.25 has been used in the slot q = 1 and in the slot q = 7, because
only half a slot is considered. Finally, it is worth noticing that the sum of the
magnitudes of the coefficients of each row (with whole slot) is always equal
to one.
−b −b +a +a −c −c
Uo(n)
+c −b −b +a +a −c
+c +c −b −b +a +a
0 π 2π
ϑ = ωt
FIGURE 8.6
Construction of the voltage waveform by dots.
the d-axis. Then, indicating Λma the maximum flux linkage, with constant
rotor speed, the RMS value of the induced EMF is
1
Ea = ωΛ ma (8.4)
2
which is the no-load line voltage with a wye winding connection. Con-
versely, with a star winding connection, the no-load line voltage is
U 0 = 3 Ea (8.5)
From the same field solution, supposing a step rotation of the coils of one
slot pitch, it is possible to achieve some points of the no-load voltage wave-
form. Assuming this rotation, different coil sides are associated to each slot,
as shown in Figure 8.6, the coefficients kjq are suitably modified. This is
possible because only the rotor is fed, which means the flux lines do not
depend on the distribution of the conductors into the slots.
The waveform Uo(n) is built by Nc points (e.g., it is common that Nc = Q/p).
The harmonic content can be approximately estimated. From the discrete
Fourier series expansion, the fundamental value is obtained as
Nc
U1sin =
2
Nc ∑ U ( n ) ⋅ sin N2π n
n =1
o
c
Nc
U1 cos =
2
Nc ∑ U ( n ) ⋅ cos N2π n
n =1
o
c
(8.6)
U1 = U1sin 2 + U1 cos 2
148 Electrical Machine Analysis Using Finite Elements
Nc
U ν sin =
2
Nc ∑ U ( n ) ⋅ sin N2π νn
n =1
o
c
(8.7)
Nc
U ν cos =
2
Nc ∑ U ( n ) ⋅ cos N2π νn
n =1
o
c
i a (t ) = I M sin ωt ( )
2π
i b (t ) = I M sin ωt − (8.8)
3
4π
i c (t ) = I M sin ωt −
3
Ic
Ia
ωt
Ib
FIGURE 8.7
Reference phasor diagram of the stator winding currents.
Synchronous Generators 149
a-phase axis
Ia
−b −b +a +a −c −c
ω t = π /2
Polar axis
(d-axis)
Ic Ib
(a)
a-phase axis
Ic
+a −c −c +b +b −a
Polar axis
(d-axis) Ia
ωt=0
Ib
(b)
FIGURE 8.8
Combinations of the winding distribution and time instant, for computing the direct axis
inductance.
Once the distribution of the winding is fixed, the three-phase currents are
chosen so that the maximum of the MMF distribution coincides with the
polar axis (d-axis). This corresponds to set Id = In and Iq = 0. Referring to one
pole of the generator, Figure 8.8 shows two suitable combinations of winding
distribution and time instant. For both of them, the boundary conditions of
Figure 8.3(b) can be assigned.
With the configuration of Figure 8.8(a) the time instant with ωt = π/2 has
been chosen, so that
ia = I M
IM
ib = − (8.9)
2
IM
ic = −
2
With the configuration of Figure 8.8(b) the time instant with ωt = 0 has
been chosen, so that
ia = 0
3
ib = − IM (8.10)
2
3
ic = + IM
2
150 Electrical Machine Analysis Using Finite Elements
Once the three-phase currents ia, ib, and ic, and the distribution of the
windings have been fixed, i.e., the coefficients kaq, kbq, and kcq for each slot,
the current flowing in the q-th slot can be expressed as
∑k i
nq
iq = jq q
n pp j= a , b , c
(8.11)
( )
n
= q k aqi a + k bqi b + k cqi c
n pp
31
Wmd = Ld Id2 (8.12)
22
With d-axis current only, it is Id = IM, and the synchronous d-axis inductance
results in
2 2 Wmd
Ld = (8.13)
3 I 2M
Slightly different results are obtained with the distributions of Figure 8.8(a)
and Figure 8.8(b) and the corresponding currents indicated in Equation (8.9)
and Equation (8.10). This is due to the different harmonic contents of the
two configurations. At last, we must bear in mind that this computation is
correct only in case of a linear magnetic circuit.
The flux linked with the a-phase winding, Λa = Λd, is computed from
Equation (8.3). The inductance is given by
Λa Λd
Ld = = (8.14)
Ia IM
Λb 2Λ b
Ld = = (8.15)
Ib 3IM
The difference between Equation (8.14) and Equation (8.15) indicates the
different harmonic contributions of the two solutions.
∫B
2
B1dM = (ϑ r )cos(ϑ r )dϑ r (8.16)
π
gn
ο
Np
B1dM =
2
Np ∑B
n =1
gn ( n )cos( n∆ϑ)∆ϑ (8.17)
k w N DL Fe
Λ d = B1dM (8.18)
2 p
152 Electrical Machine Analysis Using Finite Elements
Λd
Ldm = (8.19)
IM
The last computation adopts the fundamental harmonic of the flux density
distribution, then is almost independent of the harmonic content.
In addition, while the leakage flux in the air-gap and in the slots is con-
sidered in Equation (8.13) and Equation (8.14), it is not considered when
Equation (8.19) is used. Thus, letting Lσ be the leakage inductance, the direct
axis inductance is obtained as
Ld = L σ + Ldm (8.20)
It is worth noticing that Equation (8.14) and Equation (8.19) can be applied
even without setting the magnetizing current equal to zero. In such a case,
the flux linkage due to the magnetizing current must be previously com-
puted and considered constant. It has to be subtracted from the total flux
linkage. With reference to the a-phase, Equation (8.14) and Equation (8.19)
become
Λ d − Λ dm Λ a − Λ ma
Ld = = (8.21)
IM IM
where Λdm = Λma indicates the a-phase flux linkage due to the magnetizing
current.
Dirichlet Az = 0 Az = 0
Az = 0
Az = 0 dAz
Az = 0 =0
dAz dAz dn
=0 =0
Az = 0 dn dn
Neumann
FIGURE 8.9
Boundary conditions for computing the quadrature axis inductance.
2 2 Wmq
Lq = (8.22)
3 I 2M
Λq
Lq = (8.23)
IM
∫B
2
B1qM = (ϑ r )sin(ϑ r )dϑ r (8.24)
π
gn
ο
k w N DL Fe
Λ q = B1qM (8.25)
2 p
Λq
Lqm = (8.26)
IM
154 Electrical Machine Analysis Using Finite Elements
Λ a (ϑ )
L a (ϑ ) =
Ia
Λ b (ϑ )
M ba (ϑ) = (8.27)
Ia
Λ c (ϑ )
M ca (ϑ) =
Ia
2 2π 4π
Λd = Λ a cos ϑ + Λ b cos ϑ − + Λ c cos ϑ −
3 3 3
(8.29)
2 2π 4π
Λ q = − Λ a sin ϑ + Λ b sin ϑ − + Λ c sin ϑ −
3 3 3
Az = 0 Az = 0 Az = 0
+Az (r)
+Az (r) −Az (r)
+Az (r) +Az (r) −Az (r)
(a) (b) (c )
FIGURE 8.10
Periodic boundary conditions.
156 Electrical Machine Analysis Using Finite Elements
Alternatively, starting from the radial component of the air-gap flux density
distribution Bgn(ϑ), the two-axis fundamental harmonic components are
obtained from Equation (8.16) and Equation (8.24). The d- and q-axis flux
linkages Λ1d and Λ1q are then obtained from Equation (8.18) and Equation
(8.25). At this point, the d- and q-axis currents are computed from the three-
phase currents Ia, Ib, Ic, always by means of the transformation Tabc/dq (see
the Appendix to this chapter), as
2 2π 4π
Id = Ia cos ϑ + I b cos ϑ − + I c cos ϑ −
3 3 3
(8.30)
2 2π 4π
Iq = − Ia sin ϑ + I b sin ϑ − + I c sin ϑ −
3 3 3
Finally, neglecting the cross coupling between the d- and q-axis (that will
be considered in Chapter 10), the d- and q-axis inductances result in
Λd
Ld =
Id
(8.31)
Λ
Lq = q
Iq
XdId
Uo
RId
XqIq
RIq
Uq U
δ
I
ϕ
Iq γ
Id Ud
Polar axis
(d-axis) Interpolar axis
(q-axis)
FIGURE 8.11
Phasor diagram of the synchronous generator.
Computed quantities:
• From the field solution, the three-phase flux linkages Λa, Λb, and Λc
are computed.
• By means of the transformation Tabc/dq, the d- and q-axis components
of current, i.e., Id Iq, and flux linkages, i.e., Λd Λq, are computed.
• The d- and q-axis voltages are
Ud = − RId + ωΛ q
(8.32)
Uq = − RIq − ωΛ d
• The sign refers to the generating mode of the electrical machine, that
is, the positive current direction is in phase with the voltage. The
phasor Uo in Figure 8.11 is π/2 radiants lagging the polar axis (d-axis).
• The electrical angles are
Ud
δ = tan load angle
Uq
Iq
γ = tan current angle (8.33)
Id
π
ϕ= −γ −δ angle between voltage and current
2
ϑr d-axis
Stator
Infinitesimal 60°
conducting
sheet Rotor ϑs
y
Field
winding
200
400
FIGURE 8.12
Ideal synchronous machine.
8.9 Example
The aim of the following example is to show the flux plots and the air-gap
flux density distribution of the synchronous generator, during no-load and
full-load operations. The effects of the armature reaction are highlighted,
feeding the machine with d-axis currents only, q-axis currents only, and both
of them together.
Rather than analyzing an effective figure of the synchronous generator, let
us refer to the ideal structure drawn in Figure 8.12. This is a two-pole
machine, formed by a rotor with salient poles through which the field wind-
ings are wound, and by a slotless stator. In the inner surface of the stator
there is an infinitesimal conducting sheet carrying a linear current density
distribution. Such a distribution is considered to be sinusoidal, correspond-
ing to an ideally sinusoidal distribution of the winding.
In Figure 8.12, the dimensions of the generator are reported in mm. A
linear magnetic material is used, with a constant relative permeability µr =
5000.
The first simulation deals with the no-load operation, without stator cur-
rent and with a magnetizing current feeding the rotor winding. In the con-
ducting bar, equivalent to the magnetizing winding, the total current is set
NeIe = 13500 A. Figure 8.13 shows the corresponding flux plot and the air-
gap flux density distribution. Since the pole shoe is not shaped — which is
a constant air-gap under the pole — the flux density distribution looks like
a quasi-square distribution. The maximum flux density corresponds to the
Synchronous Generators 159
0.8
0.6
0.4
0.2
−0.2
−0.4
−0.6
−0.8
−1
0 90 180 240 360
Angle (electr. degrees)
FIGURE 8.13
No-load flux plot (a) and air-gap flux density distribution (b).
predictable value by Ampere’s law Bo = µ0NeIe/(2 g), with the air-gap length
g = 10 mm.
In the two following simulations, the machine is fed by a d-axis stator
current only and by a q-axis stator current only, respectively. In the stator
conducting sheet, a linear current density is imposed with the suitable dis-
tribution. In order to have a sinusoidal MMF distribution with the maximum
value right on the d-axis, the following linear current density must be
imposed:
( ) ( )
J sd ϑ r = J dM sin ϑ r (8.34)
where ϑr is the rotor coordinate and the positive sign means a current
direction equal to the z-axis direction.
Similarly, to have a sinusoidal MMF distribution with the maximum value
centered right on the q-axis, the following linear current density must be
imposed:
( ) ( )
J sq ϑ r = − J qM cos ϑ r (8.35)
Figure 8.14 shows the simulation with d-axis current density distribution.
It is useful to consider the stator coordinate ϑs = ϑm + ϑr (see Figure 8.12).
Since the rotor position is ϑm = π/2 (see Figure 8.1), a linear current density
Jsd(ϑs) = JdMcos(ϑs) has to be imposed. A negative current value has been
fixed, which is a demagnetizing current, so that JdM = −20000 A/m. Figure
8.14 shows the corresponding flux plot and the air-gap flux density distri-
bution. The flux density distribution looks like a sinusoidal distribution in
correspondence to the pole shoes, while it decreases elsewhere, due to the
decrease of the permeance of the magnetic paths.
160 Electrical Machine Analysis Using Finite Elements
FIGURE 8.14
Flux plot (a) and the air-gap flux density distribution (b) with d-axis current only.
FIGURE 8.15
Flux plot (a) and the air-gap flux density distribution (b) with q-axis current only.
Similarly, Figure 8.15 shows the flux plot and the air-gap flux density
distribution when the machine is fed by a q-axis current. A linear current
density Jsq(ϑs) = −JqMsin(ϑs) is assigned with JqM = 20000 A/m (positive). Once
again, the flux density distribution looks like a sinusoidal distribution in
correspondence with the pole shoes, while it decreases elsewhere. Compar-
ing the flux density distributions of Figure 8.14 and Figure 8.15, the reduction
of the flux density corresponding to the symmetry axis is almost 10 times.
Figure 8.16 shows the flux plots and the air-gap flux density distributions,
with magnetizing current in the rotor and d-axis current in the stator. Because
of the negative d-axis current, the demagnetizing effect of the stator current
is evident.
Figure 8.17 shows the flux plots and the air-gap flux density distributions,
with magnetizing current in the rotor and q-axis current in the stator. In this
Synchronous Generators 161
FIGURE 8.16
Flux plot (a) and air-gap flux density distribution (b) with magnetizing and d-axis currents.
0.5
−0.5
−1
−1.5
0 90 180 240 360
Angle (electr. degrees)
FIGURE 8.17
Flux plot (a) and air-gap flux density distribution (b) with magnetizing and q-axis currents.
case, the distorting effect of the stator current is evident. Since the material
has a constant µr , the saturation does not occur.
Finally, Figure 8.18 shows the flux plot and the air-gap flux density distri-
bution, when the rotor winding is fed by the magnetizing current, and the
stator winding is fed by both d- and q-axis currents. Comparing Figure 8.18
and Figure 8.13, the effect of the armature reaction is evident.
FIGURE 8.18
Flux plot (a) and air-gap flux density distribution (b) with magnetizing, d- and q-axis currents.
2 2π 4π
fd = fa cos ϑ + fb cos ϑ − + fc cos ϑ −
3 3 3
(8.36)
2 2π 4π
fq = − fa sin ϑ + fb sin ϑ − + fc sin ϑ −
3 3 3
Synchronous Generators 163
c-phase axis
FIGURE 8.19
Synchronous rotating d- and q-axis windings.
2π 4π
cos ϑ cos ϑ − cos ϑ − f
fd 2 3 3 a
= ⋅ fb (8.37)
fq 3 2π 4π
− sin ϑ − sin ϑ − − sin ϑ − fc
3 3
dλ d
vd = Ri d + − ωλ q
dt
(8.38)
dλ
vq = Ri q + q + ωλ d
dt
λ d = Ld i d + λ dm
(8.39)
λ q = Lqi q
164 Electrical Machine Analysis Using Finite Elements
T=
3
2
(
p λ d i q − λ qi d ) (8.40)
1. Since the d- and q-axis are rotating with the rotor, and the stator
reluctance variation is negligible (apart from almost negligible reluc-
tance variation due to the stator slotting), the self-inductance of the
two d- and q-axis windings is constant, unlike for the three-phase
a, b, c windings in which it is a function of the rotor position ϑm.
2. Since the d- and q-axis are at π/2 electrical radiants and posed on
two symmetry machine axes, they are not mutually coupled, as is
clear by Equation (8.39). Conversely, the a-, b-, and c-phase windings
are mutually coupled and this coupling is a function of the rotor
position ϑm as well. In reality, with high saturation, the d- and q-
axis windings may show a mutual coupling (the so-called cross cou-
pling). This phenomenon will be analyzed in Chapter 10.
3. During steady-state operations, the electrical quantities in the sta-
tionary reference frame a, b, c are sinusoidal in time, while the
electrical quantities in the rotating reference frame d, q are constant.
For instance, observe the voltage components along the d- and
q-axis, reported in Equation (8.32).
4. As a consequence of the points (1) and (2), the dynamic analysis of
the machine in the synchronous reference frame is very simplified.
Finally, let us point out that the factor in Equation (8.36) and Equation
(8.37) could be not only 2/3, even though such a choice is the most used in
the literature. With this choice, it is
References
1. R.H. Park, “Two Reaction Theory of Synchronous Machines, Part I,” in Trans.
of American Institute of Electrical Engineers, vol. 48, p. 716, 1929.
2. R.H. Park, “Two Reaction Theory of Synchronous Machines, Part II,” in Trans.
of American Institute of Electrical Engineers, vol. 52, p. 352, 1933.
3. M. Liwschitz-Garik and C.C. Whipple, Electric Machinery, Vol. I–II, D: Van
Nostrand Co., New York, 1947.
4. C. Concordia, Synchronous Machines, John Wiley & Sons, New York, 1951.
5. G. Kron, Equivalent Circuits of Electric Machinery, John Wiley & Sons, New York,
1951.
6. W.V. Lyon, Transient Analysis of Alternating Current Machinery, John Wiley &
Sons, New York, 1954.
7. A.S. Langsdorf, Theory of Alternating Current Machinery, McGraw-Hill, New
York, 1955.
8. B.J. Chalmers, Electric Motor Handbook, Butterworth, London, 1988.
9. A.E. Fidzgerald, C. Kingsley, Jr., and S.D. Umans, Electric Machinery, McGraw
Hill, New York, 1983.
10. S.A. Nasar (editor), Handbook of Electric Machines, McGraw-Hill, New York,
1987.
9
Surface-Mounted Permanent Magnet Motors
9.1 Introduction
The surface-mounted permanent magnet motor can be considered as a syn-
chronous machine where the magnetizing winding is replaced with a per-
manent magnet. Thanks to the actual performance of the recent permanent
magnet materials, such motors exhibit high efficiency and high torque-to-
volume ratio.
In addition, the permanent magnet motor may be designed in different
shapes. They may be built with high length-to-diameter ratio, when high
speed and low inertia are required, e.g., for machine tools, or with low
length-to-diameter ratio, when low speed and high torque are required, e.g.,
for direct drives and motor-wheel-in-traction applications. In recent years,
permanent magnet generators and motors were developed; however, the
following analysis deals with the permanent magnet motors.
The magnetic structure of two configurations of a four-pole surface-
mounted permanent magnet motor is shown in Figure 9.1. The first config-
uration represents the classical solution with an inner rotor and an outer
stator; the second configuration is with an outer rotor and an inner stator.
In both configurations, the permanent magnets are radially or parallel mag-
netized, and produce an almost rectangular waveform of air-gap flux density.
The d-axis corresponds to the polar axis in the middle of the rotor pole. The
q-axis is leading the d-axis of π/2 electrical radians.
In order to highlight the motor versatility to be designed in various shapes,
Figure 9.2 shows a radial-flux surface-mounted permanent magnet motor
configuration. This is formed by two flat cylinders over which the permanent
magnets, which are axially magnetized, are placed. Within the two permanent
167
168 Electrical Machine Analysis Using Finite Elements
d d
q q
FIGURE 9.1
Surface-mounted permanent magnet motor configurations: (a) with inner rotor and (b) outer
rotor.
(a) (b)
FIGURE 9.2
Axial flux surface-mounted permanent magnet motor configuration.
magnet cylinders, a stator cylinder has been inserted, holding the three-
phase armature windings. Figure 9.2(b) shows a detail of the rotor cylinder.
The surface-mounted permanent magnet motors fed by an electrical drive
are classified as: (1) squarewave current-fed motors (also called trapezoidal
brushless or dc brushless motors), and (2) sinewave current-fed motors (also
called sinusoidal brushless or ac brushless motors).
They are characterized by different winding distributions in such a way
as to obtain a different induced EMF waveform. In the squarewave current-
fed motors, the EMF waveform should be ideally trapezoidal and the windings
are fed by squarewave currents, which are synchronized with the EMFs. The
ideal waveforms of the three-phase EMFs and the corresponding forced
currents are shown in Figure 9.3. In the sinewave current-fed motors, the
induced EMF should be ideally sinusoidal and the windings are fed by
sinewave currents, synchronized with the EMFs. In both cases, the instan-
taneous motor torque is ideally constant.
Surface-Mounted Permanent Magnet Motors 169
ea t
eb t
ec t
(a)
ia t
ib t
ic t
(b)
FIGURE 9.3
Ideal EMF (a) and current (b) waveforms in a trapezoidal brushless motor.
H
Hc Hknee Ho
FIGURE 9.4
B-H curve of the permanent magnet and straight-line approximation.
170 Electrical Machine Analysis Using Finite Elements
B = Bres + µ rµ 0H (9.1)
where Bres = (Bres,x, Bres,y , 0). Then the two-dimensional magnetostatic problem
is described by the differential equation
∂ 1 ∂A z ∂ 1 ∂A z ∂ B res , y ∂ B res ,x
+ = −µ 0 J z − + (9.2)
∂x µ r ∂x ∂y µ r ∂y ∂x µ r ∂y µ r
The symmetry between the pole pairs and the corresponding periodic
boundary conditions are the same of the synchronous generator; see Chapter 8.
Dirichlet Az = 0 Az = 0
Az = 0
dAz dAz
=0 =0
dn dAz dn
=0 Az = 0
dn Az = 0
Az = 0
Neumann
FIGURE 9.5
Boundary conditions in the no-load simulation.
Surface-Mounted Permanent Magnet Motors 171
The three-phase winding is placed within the stator slots, and may be
single-layer, full-pitched winding or a double-layer, chorded winding. The
first solution is mainly adopted with squarewave current-fed motor, in order
to obtain EMFs with a trapezoidal waveform. The second solution is adopted
in sinewave current-fed motor, in order to obtain EMFs with a sinusoidal
waveform. Some examples of such windings are shown in Figure 8.4 and
Figure 8.5 of Chapter 8.
Q/2 p
∑k ∫
nq 1
Λ j, pm = 2 pL Fe jq A zdS j = a , b, c (9.3)
n pp q= 1
Sq Sq
where nq is the number of the conductors in the slot, npp is the number of
parallel paths of the winding, Q/2p is the number of slots per pole, and kjq
assumes the values 0, ±1, or ±0.5, depending on the orientation and whether
the coils in the q-th slot belong to the j-th phase (see Chapter 8, Section 8.2.1.1).
1
Ea = ωΛ a , pm (9.4)
2
one winding linking the maximum magnetic flux, as shown in the previous
section.
Conversely, a more accurate computation should make provision for an
analysis at various rotor positions. In this way, it is possible to built the flux
linkage waveform for each phase winding, as a function of the mechanical
angle ϑm, i.e., λpm(ϑm). Therefore, the fundamental and the higher-order
harmonics of the flux linkage can be obtained by means of the Fourier series
expansion. It yields
Nk
λ pm (ϑ m ) = ∑Λ
k =1
pm , k cos(kϑ m ) (9.5)
where the higher order of the series is Nk. In the series expansion, only cosi-
nusoidal waveforms have been adopted, thanks to a suitable choice of the
reference angle ϑm = 0. Really, this assumption is not valid in the case of
fractional-slot winding without symmetry between the North and South poles.
The no-load EMF is obtained by deriving the flux linkage with respect to
the time. Assuming a constant rotor speed, it is practical to derive the flux
linkage with respect to the mechanical angle ϑm, and to multiply by the
mechanical speed ωm, which is
dλ pm (ϑ m ) dλ pm (ϑ m ) dϑ m dλ pm (ϑ m )
e(ϑ m ) = = = ωm (9.6)
dt dϑ m dt dϑ m
Since the flux linkage in Equation (9.6) is derived numerically, this oper-
ation may give rise to errors. It is convenient to express the flux linkage
using the Fourier series expansion, as in Equation (9.5), and then to compute
the EMF as the series of the derivatives of each flux linkage harmonic, as in
Nk
e(ϑ m ) = ω m ∑ −k Λ
k =1
pm , k sin(kϑ m ) (9.7)
lg in the middle of the air-gap, and then multiplying the result by the active
length of the rotor, LFe. Assuming a 2p-pole machine, it yields
D − g L Fe
Tcog =
2 µ0
p
∫ B B dl
lg
r ϑ (9.8)
where Br is the radial component of the flux density (normal to the line lg),
Bϑ is the azimuthal component of the flux density (tangential to the line lg),
D is the inner stator diameter, and g is the air-gap length. The number of
pole pairs p is used assuming that the simulation has been carried out on
two pole-pieces of the machine only. A different factor has to be used with
different simulation (e.g., 2p should be used instead of p when only one
pole-piece is simulated).
Because of the numerical nature of the finite element method, the result
may depend on both the position of the integration line and the number of
points chosen for the numerical integration. Instead of Equation (9.8), it is
better to compute the average value of the torque over the entire air-gap
surface Sg. Then the torque is
∫
L Fe
Tcog = p rB rB ϑdS (9.9)
gµ 0 Sg
B
Wm( pm ) =
∫ ∫
τ pm 0
H ⋅ dB dτ
(9.10)
B
= L Fe
∫ ∫
S pm 0
H ⋅ dB dS
174 Electrical Machine Analysis Using Finite Elements
B B B
Bo
∆B ∆B
B
H H H
Ho ∆H ∆H
(a) (b) (c)
FIGURE 9.6
Magnetic energy and coenergy density in permanent magnets.
∫ (B )
1
Wm( pm ) = L Fe 2
− B resB dS
2µ r µ 0 S pm
(9.11)
µµ
= r 0 L Fe
2 ∫ S pm
2
H dS
H
Wm′ ( pm ) =
∫ ∫
τ pm 0
B ⋅ dH dτ
∫
1
= L Fe B 2 dS (9.12)
2µ r µ 0 S pm
2
µ rµ 0 B
=
2
L Fe
∫
S pm
H + res dS
µ rµ 0
dWm
Tcog = − (9.13)
dϑ m
(a)
(b) (c)
FIGURE 9.7
Tricks to rotate the structure without mesh change, so as to avoid numerical errors.
so high that they significantly alter the solution, especially when low torques
are investigated (such as the cogging torque).
To avoid or to reduce such errors, two tricks are suggested.
The first is to divide the permanent magnet in various pieces, as shown
in Figure 9.7(a). The length of each piece should correspond to the minimum
rotation that has been fixed. According to the rotor position, the magnetic
B-H curve of the air or of the permanent magnet is assigned to each piece.
In this way, a fictitious rotation is obtained, except that it operates on the
material definition and not on the drawing of the motor. As a consequence,
the mesh of the domain remains always the same.
The second trick is to split the air-gap into two parts, as illustrated in
Figure 9.7(b). Then the rotor and the corresponding air-gap part is rotated,
keeping fixed the other air-gap part together with the stator, as shown in
Figure 9.7(c). The mesh remains the same, fixed to the structure, both in the
movable and in the fixed part. Of course, for each position, the boundary
conditions have to be adapted to the exact domain contour. It is worth
noticing that the boundary line between the two air-gap parts must be
divided properly, in such a way as the stator mesh correctly matches the
rotor mesh in each rotor position.
The three ways for computing the motor torque are described next.
∫
L Fe
T= p rB rB ϑdS (9.14)
gµ o Sg
The pole pairs number p is used in Equation (9.14), assuming that the
simulation is carried out again on two pole-pieces of the machine.
Surface-Mounted Permanent Magnet Motors 177
dWm′
T=+ (9.15)
dϑ m i = const
T=
3
2
(
p Λ d I q − Λ qI d ) (9.16)
where the d- and q-axis currents are given by Equation (8.30), and the d-
and q-axis flux linkages are given by Equation (8.29).
The angle ϑ is the electrical angle between the reference a-phase axis and
the d-axis. The flux linked with the j-th phase (j = a, b, c) is obtained from
the average value of the magnetic vector potential Az over the slot surfaces SCu.
It is worth bearing in mind that the torque computed by Equation (9.16)
is obtained as the energy balance at the terminals of the motor windings.
Such a balance is based on the assumption of sinusoidal distributed wind-
ings. As a consequence, the average torque is correct, but the torque ripple
is lower. It could be verified that MMF harmonics are only partially included
in Equation (9.16). In addition, such a computation method does not consider
the cogging torque due to the interaction of the permanent magnet and the
stator slotting. In fact, the torque computed from Equation (9.16) is always
null when the stator currents are null (i.e., Id = Iq = 0).
3 2
2 1.5
Torque (Nm)
Current (A)
1
1
0 Reference (A)
0.5 ia(t)
2
−1 ia(t)
−2 0 0
ia(t) ib(t) ic(t)
−3 −0.5 −2
0 2 4 6 8 10 0 2 4 6 8 10
Time (ms) Time (ms)
FIGURE 9.8
Currents and torque waveforms.
inductances and the induced EMFs, together with the limited value of the
source voltage, the electrical drive is not always able to feed the motor with
ideal currents. On the contrary, the currents often follow their references
with a delay.
This delay is mainly appreciable in squarewave current-fed motors, since
the currents cannot exhibit rapid changes as those required by the ideal
currents shown in Figure 9.3(b). The current is as distorted as the rotor speed
is high. Consequently, a constant torque is not obtained, but a torque ripple
takes place, which is called the commutation effect. Figure 9.8(a) shows the
typical three-phase current waveforms, obtained by a dynamic simulation
at a supply frequency f = 100 Hz. In the figure, the rise and fall times of the
current are clear. The delay of the phase current with respect to its reference,
highlighted in Figure 9.8(b), causes the torque ripple, shown in Figure 9.8(b)
as well.
For the computation of the instantaneous torque, a simple procedure is
adopted. All the motor parameters are known, i.e., the resistance R, self- and
mutual inductances L and M, the induced EMF waveform e(ϑm). With fixed
motor speed, the latter may be expressed as a function of the time, as ϑm =
ωmt. Using a two-level inverter, the voltage forced at the motor terminals
can assume the two values v = −Vdc or v = +Vdc (v = 0 using a three-level
inverter), where Vdc is the dc bus voltage. The voltage level is chosen by the
control comparing the current reference (ideal current) with the effective
current.
The dynamic analysis is carried out, from the step integration of the dif-
ferential equation
di j (t )
v j (t ) = Ri j (t ) + (L − M) + e j (t ) + v N (t ) (9.17)
dt
for each j-th phase winding (j = a, b, c). In Equation (9.17), vN(t) indicates
the star-center potential, which is computed at each step from the forced
terminal voltages.
Surface-Mounted Permanent Magnet Motors 179
9.5 Example
This section discusses some results of finite element analysis of a surface-
mounted permanent magnet motor. The motor data are reported in Table 9.1;
the magnetic structure and the winding are sketched in Figure 9.9. It is a
squarewave current-fed brushless motor with a low number of slots per pole
and a chorded winding. Such a solution is adopted in low-power applica-
tions, with the aim of having minimum end-winding length and higher
motor efficiency, in spite of a higher torque ripple.
With such a motor structure, a two-pole section of the machine is simu-
lated, using the periodic conditions on the two boundaries. Different rotor
positions are simulated splitting the permanent magnet in elementary pieces,
as shown in Figure 9.7(a).
A rotation of 40 mechanical degrees (120 electrical degrees) is considered
although a rotation of 30 mechanical degrees would be enough.
TABLE 9.1
Brushless Motor Data
Stator Data
Rotor Data
a1
b1
c3 a1 b1 c1
c1
b3 a2 b2 c2
a2
a3 b3 c3
a3
b2
c2
FIGURE 9.9
Brushless motor magnetic structure.
FIGURE 9.10
Flux lines at no-load operation; a rotation of ϑm = 10 mechanical degrees is considered.
The three-phase flux linkages are shown in Figure 9.11(a). They are approx-
imately flat for a portion of 40 mechanical degrees (120 electrical degrees).
It follows that, with constant speed, the back EMFs exhibit a constant value
for a corresponding period of time (one third of the electrical period), as
required in a trapezoidal brushless drive; see Figure 9.3(a).
The EMF waveform is not exactly trapezoidal because of the chorded
winding. Thus, a constant torque is not achieved even when ideal square-
wave currents of Figure 9.3(b) feed the stator windings.
The magnetic energy is shown in Figure 9.11(b). The energy variation
indicates that, during the rotation, there is a cogging torque due to the
interaction of the permanent magnet with the stator slotting.
Surface-Mounted Permanent Magnet Motors 181
0.2 0.229
0.15 0.228
0.1 0.227
0.05 0.226
0.225
0
0.224
−0.05 0.223
−0.1 0.222
−0.15 0.221
−0.2 0.22
0° 5° 10° 15° 20° 25° 30° 35° 40° 0° 5° 10° 15° 20° 25° 30° 35° 40°
Mechanical angle (degrees) Mechanical angle (degrees)
FIGURE 9.11
Flux linkages and magnetic energy during the no-load operation.
Maxwell
0.1 Energy 0.1
0.08 Coenergy
Cogging torque (Nm)
0.08
0.06 0.06
0.04 0.04
0.02 0.02
0 0
−0.02 −0.02
−0.04 −0.04
−0.06 −0.06
−0.08 −0.08
−0.1 −0.1
0° 5° 10° 15° 20° 25° 30° 35° 40° 0 1 2 3 4 5 6
Mechanical angle (degrees) Time (s)
FIGURE 9.12
Cogging torque: simulated (a) and measured (b) values.
The cogging torque behavior is shown in Figure 9.12(a). It has been com-
puted in different ways by means of Maxwell’s stress tensor, using Equation
(9.9), drawn in a thin line; by means of the derivative of the magnetic energy,
using Equation (9.13), drawn in a bold line (almost indistinguishable from
the previous one); and by means of the derivative of the magnetic coenergy,
using Equation (9.15) with null currents, drawn using dots. It is worth
noticing the good match among the three methods.
At last, Figure 9.12(b) shows the comparison between the simulated and
the measured cogging torque. The measure system has been regulated to
complete a whole rotation in 6 seconds. The figure shows a satisfactory
agreement between simulated results and measurements. A small difference
can be attributed to a nonperfect symmetry of the rotor poles and to a
residual flux density slightly lower than that used in the simulations.
2.5
1.5
Maxwell stress tensor
1 Balance in d-q ref. frame
Magnetic coenergy variation
0.5
Cogging torque
0
0° 5° 10° 15° 20° 25° 30° 35° 40°
Mechanical angle (degrees)
FIGURE 9.13
Electromechanic torque with ideal squarewave currents, I = 2 A.
References
1. N. Demerdash et al., “Comparison between features and performance charac-
teristics of fifteen-hp samarium-cobalt and ferrite-based brushless d.c. motors,”
IEEE Trans. on PAS, vol. 102, pp. 104–112, 1983.
2. H. Weh, H. Mosebach, H. May, “Design concepts and force generation in
inverter-fed synchronous machines with permanent magnet excitation,” in
IEEE Trans on Magnetics, vol. MAG-20, no. 5, pp. 1756–1761, 1984.
3. B.J. Chalmers, S.A. Hamed, G.D. Baines, “Parameters and performances of a
high-field permanent-magnet synchronous motor for variable-frequency oper-
ation,” IEEE Proc. Pt. B, vol. 132, no. 3, 1985.
4. N. Boules, “Prediction of No-Load Flux Density Distribution in Permanent
Magnet Machines,” IEEE Trans. on IA, vol. 21, no. 4, pp. 633–643, 1985.
5. K. Kobayashi and M. Goto, “A brushless DC motor of a new structure with
reduced torque fluctuations,” Elec. Eng. Jpn, vol. 105, no. 3, pp. 104–112, 1985.
Surface-Mounted Permanent Magnet Motors 183
This chapter deals with the finite element analysis of an interior permanent
magnet motor. The interior permanent magnet motor is used in traction
applications or when high dynamic performance is required. It is a synchro-
nous motor, fed by sinewave currents via an electrical drive; therefore the
analysis of such a motor is carried out by means of the direct and the
quadrature axis theory. Two magnetic models of the motor are presented,
and the procedures for obtaining them are illustrated. An automatic algo-
rithm is also described to process the data obtained from the finite element
analysis. The synchronous reluctance motor is discussed at the end of the
chapter, considering such a motor as a particular case of the interior perma-
nent magnet motor.
10.1 Introduction
The key characteristic of the interior permanent magnet motor is that the
permanent magnets are not on the surface of the rotor but are buried in the
rotor laminations. Apposite holes are stamped in the rotor laminations to
hold the permanent magnets and to obstruct the d-axis flux due to the stator
current. For this reason, they are also called flux barriers.
Since the differential relative permeability of the permanent magnet is only
slightly higher than the air permeability, the rotor exhibits magnetic paths
with different magnetic permeance. Let the d-axis correspond to the polar
axis (i.e., the permanent magnet axis). Thus the d-axis permeance is lower
than the q-axis permeance, and the inductances are Ld < Lq. This is opposite
of the usual situation of the salient pole synchronous machines with wound
rotor (it is useful to compare with Chapter 8).
As a consequence of the rotor anisotropy, the motor takes advantage of
the reluctance torque and the permanent magnet torque, showing a favorable
torque-to-volume ratio. In addition, the interior permanent magnet motor is
185
186 Electrical Machine Analysis Using Finite Elements
d d
q q
N
N N S
S S S S
N N
S S
N N
S S
N N N
S N
S N
S
FIGURE 10.1
Different rotor structures of the interior permanent magnet motors.
FIGURE 10.2
Air-gap flux density distributions due to the permanent magnet (a), d-axis current (b), q-axis
current (c), and flux density distribution along the permanent magnet with negative d-axis
current (d).
The effect of the slot opening is well manifested. Figure 10.2(b) and Figure
10.2(c) illustrate the air-gap flux density distribution due to the stator cur-
rents only (the permanent magnet has been demagnetized, i.e., Bres = 0), with
only a d-axis current and only a q-axis current, respectively (the PM is
unmagnetized). The fundamental harmonic of the distribution is highlighted
by dashed lines. The slotting effect is again manifested, together with the
effect of the winding distribution. Because of the flux barriers, when a d-
axis current flows in the stator windings, some parts of the rotor assume a
magnetic potential different from zero. These rotor parts are the iron
“islands,” bordered by the flux barriers and the air-gap. As a consequence,
188 Electrical Machine Analysis Using Finite Elements
the maximum value of the flux density due to the d-axis current of Figure
10.2(c) is lower than that due to the q-axis current of Figure 10.2(b), although
the stator is fed by the same current.
Finally, Figure 10.2(d) shows the flux density distribution on the surface
of the permanent magnet, caused by a negative d-axis current only. The flux
density is negative (demagnetizing) and is uniformly distributed on the
whole surface. This check should be carried out at the maximum stator
current and is necessary to ensure that the permanent magnets are not
demagnetized irreversibly.
T=
3
2
(
p Λ d I q − Λ qI d ) (10.1)
Vd = RId − ωΛ q
(10.2)
Vq = RIq + ωΛ d
(
V 2 = ω 2 Λd 2 + Λq2 ) (10.3)
The flux linkages Λd and Λq are always expressed by using the flux linkage
due to the permanent magnet Λm and the two axis inductances Ld and Lq.
In general, they are dependent on the stator current, due to the saturation.
These parameters represent the magnetic model of the motor and are easily
obtained from the field solution.
0.02 Lqu
0
−200 −100 0 100 200
d- and q-axis current (A)
FIGURE 10.3
Simplified magnetic model of the interior permanent magnet motor.
Λ d (I d ) = Λ m + L d I d (10.4)
The magnetic characteristics that describe the dependence of the flux link-
age on the currents are shown in Figure 10.3. An advantage of this magnetic
model is that simple equations are employed. These are particularly advan-
tageous in the design step of the interior permanent magnet motor or of the
current control. Moreover, it is rapidly obtained from only four field solutions:
190 Electrical Machine Analysis Using Finite Elements
1. A first simulation with null stator currents allows the flux linkage
due to the permanent magnet, Λm, to be obtained.
2. A second simulation, with low d- and q-axis current values (an
indicative value could be Id = Iq equal to the 10% of the rated stator
current), allows the two unsaturated inductances to be computed.
Letting Λd and Λq be the flux linkages corresponding to the currents
Id and Iq, the two-axis inductances are
Λd − Λ m Λq
Ld = Lqu = (10.6)
Id Iq
Λ′q − Λ′′q
Lqs = (10.7)
I′q − I′′q
Λ qs
Iqs =
Lqu − Lqs
TABLE 10.1
Computed Values for the Interior Permanent Magnet
Motor Model
Id (A) Iq (A) Λd (mVs) Λq (mVs) Motor Parameter
#1 0 0 67.5 — Λm = 67.5 (mVs)
#2 –85 85 46.8 46.2 Ld = 0.243 (mH)
Lqu = 0.543 (mH)
#3 0 192 — 95.05
#4 0 240 — 108.9 Lqs = 0.288 (mH)
Interior Permanent Magnet and Reluctance Synchronous Motors 191
Λ d = Λ d (I d , I q )
(10.9)
Λ q = Λ q (I d , I q )
FIGURE 10.4
Flux linkages Λd and Λq vs. currents Id and Iq.
192 Electrical Machine Analysis Using Finite Elements
on the currents is particularly useful for the design of electrical drive with
very precise positioning and very high dynamic performance.
Constant torque region — At low speed, the motor operates at the rated
current, say In, with Id and Iq components chosen in such a way to
obtain the maximum torque. It is the base torque Tb. This operating
point can be maintained from null speed up to the base speed, say
nb, at which the terminal voltage reaches the rated value Vn.
Constant power region — The motor operates at speed higher than nb
with a demagnetizing (i.e., negative) d-axis current. Such a current
weakens the flux linked by the stator winding, so that it is possible
to increase the motor speed without exceeding the nominal voltage,
as indicated by Equation (10.3) where ω = (2π/60)np. On the other
hand, a decrease of the torque is achieved, since the Id and Iq currents
move from the operating point of maximum torque. The operating
region presents a torque decreasing with the speed and is so called
the “constant power” region, even though it only approximates this
behavior. Referring to the operating limit, the current phasor Id + jIq
maintains a constant magnitude and describes a circular trajectory,
always increasing the negative value of the d-axis current. A wide
operating region is obtained if the interior permanent magnet motor
exhibits Λm ≈ LdIn.
2. Once the maximum torque operating point has been found, impos-
ing the nominal voltage V = Vn in Equation (10.3), the base speed is
computed as nb = (60/2π)ω/p, where
Vn
ω= (10.11)
Λd 2 + Λq2
10.3.1 Example
An interior permanent magnet motor is designed for an application requiring
a constant torque region with base torque Tb = 10 Nm and base speed nb =
2000 rpm, and a constant power region up to the maximum speed nfw =
6000 rpm, at which the torque should be Tfw = 3.33 Nm.
The motor structure and the main data are reported in Figure 10.5. A single-
layer, full-pitched winding is used. The rotor exhibits an anisotropy ratio
(also called the saliency ratio) ξ = Lq/Ld almost equal to 4, confirming the
high anisotropy of this kind of rotor configuration.
Figure 10.6 shows the flux lines in one section of the motor pole. Figure
10.6(a) refers to the no-load operation. The flux leakage through the rotor
wt
ht
De g S
2 lm
N
hm
2 S N
D/2
FIGURE 10.5
Motor structure and main data.
194 Electrical Machine Analysis Using Finite Elements
FIGURE 10.6
Flux lines with only permanent magnet (a) and only q-axis currents (b).
0.5 0
0 −0.5
−0.5 −1 Iq = 30 (A)
no PM
only PM
−1 −1.5
0° 60° 120° 180° 0° 60° 120° 180°
Electrical angle (deg) Electrical angle (deg)
(a1) (b1)
−0.5 −1
−1 −1.5
0° 60° 120° 180° 0° 60° 120° 180°
Electrical angle (deg) Electrical angle (deg)
(a2) (b2)
−0.5 −1
−1 −1.5
0° 60° 120° 180° 0° 60° 120° 180°
Electrical angle (deg) Electrical angle (deg)
(a3) (b3)
FIGURE 10.7
Effect of different currents on the air-gap flux density distribution.
At first, the code is reported for the automatic computation of the d- and
q-axis flux linkages, corresponding to a generic current value.
In the example, the motor structure of Figure 10.10 is considered. A double-
layer, one-slot chorded winding is adopted. The portion used in the simula-
tion is that shown in Figure 8.5(b) in Chapter 8.
196 Electrical Machine Analysis Using Finite Elements
12
70 (A) peak Simulated
10 Measured
Torque (Nm)
50 (A) peak
6
30 (A) peak
4
0
90° 100° 110° 120° 130° 140° 150° 160° 170° 180°
Current vector angle (elect. degrees)
FIGURE 10.8
Motor torque versus the angle of the current vector (Id, Iq).
FIGURE 10.9
Torque and power versus speed, at different stator current values.
7 d
6
5
4
3
2
1
g
D
De
FIGURE 10.10
Structure of the interior permanent magnet motor and references.
Interior Permanent Magnet and Reluctance Synchronous Motors 197
Some FUNCTIONs are employed to assign the currents, the material char-
acteristics, and the boundary conditions. Then the field problem is solved.
From the field solution the magnetic quantities are processed. Let us assume
that the air-gap flux density distribution is stored in the file FileInduz$,
and the average magnetic vector potential within the stator slots are stored
in files FilePotential$, one for each slot.
' Name of the file for flux density distribution
FileInduz$ = “bgload”
' Name of the file for average magnetic vector potential
FilePotential$ = “MagnVectPot”
' Assignment of the currents
CALL AssignCurrent (Id, Iq)
' Assignment of the iron permeability
CALL AssignMu
' Assignment of the permanent magnet parameters
CALL AssignPM (mu, Br, ang)
' The field problem is solved
CALL Solution
' the flux density distribution is stored in the file
CALL Induction (FileInduz$)
' the flux linkage is computed by integrating Az
CALL GlobalFlux
The FUNCTION that is used to assign the current in the stator slots is
reported next. This FUNCTION is based on the distribution of the coils as
shown in Figure 10.10(b). In order to impose the current components Id and
Iq, since the polar axis has been fixed corresponding to the a-phase axis, so
that ϑ = 0, the three-phase currents are given by
Ia = Id Ib = −
1
2(Id − 3 Iq ) Ic = −
1
2(Id + 3 Iq ) (10.12)
( ) ( ) ( )
n qs n qs n qs
I1 = − Ia − Ia I2 = − Ia + I c I3 = Ic + Ic
4 2 2
( ) ( ) ( )
n n n
I 4 = qs I c − I b I 5 = qs − I b − I b I6 = qs − I b + Ia (10.13)
2 2 2
( )
n
I7 = qs Ia + Ia
4
198 Electrical Machine Analysis Using Finite Elements
where nqs is the number of series conductors per slot, given by nq/npp, with
npp the number of parallel paths (see Chapter 8). The current I1 and I7 are
divided by 4, since only half a slot is simulated. Let N be the number of
series conductors per phase, 2p the number of poles, and qs the number of
slots per pole per phase. Then, it results in
N 3N
n qs = = (10.14)
2 pqs Q
After the field solution has been obtained, the results stored in the files
are processed. At first the air-gap flux density distribution is managed. The
distribution of the air-gap flux density is given by 180 values; it is read from
the file FileInduz$. These 180 values are stored in the variable Bn1(nnn).
The two fundamental harmonics are computed, corresponding to the direct
axis (cosϑr) and the quadrature axis (sinϑr) as depicted by Equation (8.16)
and Equation (8.24). The d- and q-axis flux linkages Λd and Λq are obtained
as depicted by Equation (8.18) and Equation (8.25).
∫
1
A slot (q) = A z dS (10.15)
Sq Sq
and it is stored in the files FilePotential$, one for each slot. Then these
values are read from the files, for each slots of the domain (7 in the present
example) and assigned to the variables fslot(nnn).
From the distribution of the coil sections in the slots and from the sym-
metry conditions, the three-phase flux linkages Λa, Λb, and Λc are computed.
In the specific example they are given by
( ) ( )
n qs
Λa = 2 p − Λ1 + Λ 2 + Λ 6 + Λ 7
2
( )
n qs
Λ b = 2p − Λ 4 + 2 Λ 5 + Λ6 (10.16)
2
( )
n qs
Λc = 2p Λ2 + 2Λ3 + Λ4
2
The two-axis flux linkages Λd and Λq are then computed, using the trans-
formation Tabc/dq (see the Appendix in Chapter 8). Since ϑ = 0, the d- and q-axis
flux linkage components are given by
2 Λ + Λc
Λd =
3
Λa − b
2
Λq =
1
3
(Λ b − Λc ) (10.17)
Then it results:
The flux lines corresponding to one pole of the machine are shown in
Figure 10.11, referring to the no-load and the full-load operation, respectively.
Some results obtained by applying the two proposed methods are reported
in Figure 10.12. Figure 10.12 shows a good agreement between the two
computation methods. In addition, the results are compared with experi-
mental results obtained by feeding the motor at industry frequency 50 Hz
and different loads.
FIGURE 10.11
Flux lines during no-load (a) and the full-load (b) operations.
γ = 23.6°
* Method 1 60 27.6° * Method 1
50 x Method 2 x Method 2
40
γ = 23.6°
40 20
50 150 250 350 450 50 150 250 350 450
Current: magnitude (A), phase (deg) Current: magnitude (A), phase (deg)
FIGURE 10.12
Comparison between simulated and measured results.
Interior Permanent Magnet and Reluctance Synchronous Motors 201
d d
q q
(a) (b)
FIGURE 10.13
Synchronous reluctance motor configurations.
202 Electrical Machine Analysis Using Finite Elements
are bended and placed along the axial direction. The structure is supported
by nonmagnetic spacings. The shaft is nonmagnetic as well. This rotor con-
figuration shows a saliency ratio more than 20, in linear conditions. However,
the rotor assembly is difficult and iron rotor losses occur, caused by the flux
fluctuation in the rotor laminations due to the stator slotting.
Sometimes, for increasing the power factor or the performance in the con-
stant power region, permanent magnets are inset within the flux barriers. They
are plastic bonded low-energy permanent magnets, which are shaped easily.
FIGURE 10.14
Flux lines in a synchronous reluctance motor fed by only d-axis current (a), only q-axis current
(b), and two-axis current (c).
Interior Permanent Magnet and Reluctance Synchronous Motors 203
FIGURE 10.15
Flux plots with d-axis current only (a) and q-axis current only (b).
mNIˆ
( )
J sd ϑ r =
πD
sin(ϑ r ) (10.18)
mNIˆ
( )
J sq ϑ r = −
πD
cos(ϑ r ) (10.19)
204 Electrical Machine Analysis Using Finite Elements
FIGURE 10.16
Flux lines with (a) d-axis and (b) q-axis surface current density distribution.
0.25 0.02
0.01
0.2
0
0.15
−0.01
0.1
−0.02
0.05 −0.03 Current Iq = 1 A
Current Id = 1 A
0 −0.04
−90° −80° −70° −60° −50° −40° −30° −20° −10° 0° −90° −80° −70° −60° −50° −40° −30° −20° −10° 0°
Electric angle (deg) Electrical angle (deg)
FIGURE 10.17
Air-gap flux density distributions in simulations of rotor only.
References
1. V.B. Honsinger, “The field and parameters of interior type AC permanent
magnet machines,” in IEEE Trans. on PAS, 101, 1981, pp. 867–876.
2. B. Sneyers, D.W. Novotny, T.A. Lipo, “Field weakening in buried permanent
magnet AC motor drives,” in IEEE Trans. on Ind. Appl., vol. 21, no. 2, pp.
398–407, March/April 1985.
3. T.M. Jahns, G.B. Kliman, T.W. Neumann, “Interior permanent-magnet synchro-
nous motors for adjustable-speed drives,” in IEEE Trans. on Ind. Appl., vol. 22,
no. 4, pp. 738–747, July/August 1986.
4. R.F. Schiferl and T.A. Lipo, “Power capability of salient pole permanent magnet
synchronous motors in variable speed drive applications,” in IEEE Trans. on
Ind. Appl., vol. 26, no. 1, pp. 115–123, January/February 1990.
Interior Permanent Magnet and Reluctance Synchronous Motors 205
20. A. Fratta and A. Vagati, “A Reluctance Motor Drive for High Dynamic Perfor-
mance Applications,” in IEEE IAS Annual Meeting, Atlanta, pp. 295–302, 1987.
21. R.E. Betz, “Control of Synchronous Reluctance Machines,” in Ind. Appl. Society
Annual Meeting Records, pp. 456–462, 1991.
22. R.E. Betz, M. Javanovic, R. Lagerquist, T.J.E. Miller, “Aspects of the control of
synchronous reluctance machines including saturation and iron losses,” in Ind.
Appl. Society Annual Meeting Records, pp. 456–463, 1992.
23. A. Vagati, G. Franceschini, I. Marongiu, G.P. Troglia, “Design criteria of high
performance synchronous reluctance motors,” in Ind. Appl. Society Annu. Meet.,
1992, vol. I, pp. 66–73.
24. N. Bianchi, S. Bolognani, “Synchronous Reluctance Motor Drives: Overload
and Flux-Weakening Performance Taking into Account Iron Saturation,” in
Proc. of International Conference on Electrical Machines, ICEM ’96, Vigo (Spagna),
10–12 September 1996, vol. I, pp. 360–365.
25. N. Bianchi, S. Bolognani, “Parameters and Volt-Ampere Ratings of a Synchro-
nous Reluctance Motor Drives for Flux-Weakening Applications Taking into
Account Iron Saturation,” in Proc. of Conference on Power Electronics and Appli-
cations, EPE ’97, Trondheim (Norvey), 8–10 September 1997, vol. 3, pp. 631–636.
26. B.J. Chalmers, L. Musaba, “Design and Field-Weakening Performance of a
Synchronous Reluctance Motor with Axially-Laminated Rotor,” IEEE Ind. Appl.
Society Annual Meeting Conference Records, 1997, pp. 271–278.
27. N. Bianchi, B.J. Chalmers, “Effect of the Distribution of the Laminations in an
Axially Laminated Relactance Motor,” in Proc. of IEEE 9th International Confer-
ence of Electrical Machines and Drives, EMD ’99, Canterbury, UK, 1–3 September
1999, pp. 376–380.
28. N. Bianchi and T.M. Jahns, Eds., “Design, Analysis and Control of Interior PM
Synchronous Machines,” Tutorial Course Notes, IEEE IAS Meeting, Seattle,
WA, Oct. 3, 2004, CLEUP, Padova.
11
Self-Starting Single-Phase
Synchronous Motors
11.1 Introduction
Electrical motors of very small dimensions are used in many devices, espe-
cially in civil applications. Since an alternative voltage is commonly avail-
able, the motors are often ac motors. Among them, the most utilized is the
single-phase induction motor, with an auxiliary winding or a shaded pole
to allow the rotor starting. In some applications, where a synchronous speed
is required, hysteresis motors, reluctance motors, or permanent magnet
motors are adopted. Additionally, when reduced dimensions are explicitly
required, the use of permanent magnets allows miniaturized motors to be
obtained.
Apart from the hysteresis motor, the main problem of the synchronous
motors is the rotor starting. In order to obviate such a problem, a conducting
cage may be inset in the rotor, so that an asynchronous starting is achieved.
Also, an anisotropic stator may be designed, so as the rest positions of the
rotor are different according to the stator current. The analysis and the design
of such motors are laborious and require remarkable work.
207
208 Electrical Machine Analysis Using Finite Elements
N N
S S
N
y S S
N
z x
FIGURE 11.1
Structure of the single-phase permanent magnet motor.
1
( )
λ ϑ m , i = Nt L Fe
∫ A zdS −
1
∫ A zdS (11.2)
S Cu S Cu + S Cu S Cu −
where LFe is the motor net length, SCu+ is the equivalent conductive bar
carrying a positive current (which is a current with direction corresponding
to the z-axis), and SCu– is the equivalent conductive bar carrying a negative
current (which is a current with direction opposite to the z-axis). The ratio
between the integral of Az and the bar surface gives the average value of the
magnetic vector potential.
210 Electrical Machine Analysis Using Finite Elements
La =
( )
λ ϑ m , i − λ pm ϑ m ( ) (11.3)
i
and can be computed from two field solutions at the same rotor angle ϑm,
the former with null current, to obtain the flux linkage λpm(ϑm), and the latter
with a stator current i, to obtain the flux linkage λ(ϑm,i).
Alternatively, the permanent magnet is demagnetized, so that λpm(ϑm) = 0.
This is achieved by assigning a null residual flux density, i.e., Bres = 0 to the
permanent magnet elements, but without varying the relative permeability
µr . It results in
La =
(
λ ϑm , i ) (11.4)
i
λ pm = 0
intervals ∆ϑm for a complete angle 2π/p, corresponding to two motor poles.
Similarly, a number of Ni currents with fixed increment is chosen. Positive
currents only may be studied. It can be assumed that the motor torque for
positive current i and angle ϑm is the same torque developed with a negative
current and a rotation of π/p, i.e., τm(ϑm + π/p, –i) = τm(ϑm,i). Such an
observation allows the number of magnetostatic field analysis to be reduced.
Computing the motor torque in all the Nϑ rotor positions and with all the
Ni stator currents, a matrix with Ni column and Nϑ rows is built, containing
the numerical evaluation of the state function τm = τm(ϑm,i). It is called the
torque matrix.
v(t ) = Ri(t ) + La
di(t ) dλ pm t
+
() (11.5)
dt dt
dω m (t )
τ m (t ) = τ L + k Bω m (t ) + J (11.6)
dt
where v(t) is the forcing voltage source and R is the coil resistance. In
Equation (11.6), τm(t) is the electromechanical torque developed by the motor,
τL is the load torque, kB is the friction coefficient, and ωm(t) is the mechanical
speed.
The dynamic simulation requires that the motor performance be known
with any rotor position and with any stator current. Since in the previous
field analyses, the flux linkage and the torque have been obtained in a
discrete number of points, these values have been interpolated.
The flux linkage vector has been computed at Nϑ different rotor positions.
A one-dimensional interpolation is required to obtain the value correspond-
ing at each angular position.
Similarly, the torque matrix contains the torque computed at Nϑ rotor
position with Ni current values. A 2D interpolation is needed to obtain the
motor torque at any position and current.
212 Electrical Machine Analysis Using Finite Elements
dλ pm dλ pm dϑ m dλ pm
= = ωm (11.7)
dt dϑ m dt dϑ m
since the variation of λpm with ϑm has been computed by the magnetostatic
field analysis.
Assuming that the voltage v, and the derivative dλpm/dt are given in the
n-th instant, the time variation of the current is computed from Equation
(11.6) as
di 1 dλ pm
dt = L v − Ri( n − 1) − dt (11.8)
a
In Equation (11.8), i(n–1) is the current value at the (n–1)-th instant, and
the current variation refers to the n-th instant. Then the current at the n-th
instant is given by
di
i( n ) = i( n − 1) + dt (11.9)
dt
i( n ) + i( n − 1)
i m ( n − 1) = (11.10)
2
ϑ m ( n ) = ϑ m ( n − 1) + ω mdt (11.11)
dω m 1
dt = J τ m ( n ) − τ L − k Bω m ( n − 1) (11.12)
dω m
ω m ( n ) = ω m ( n − 1) + dt (11.13)
dt
dω m dt 2
ϑ m ( n ) = ϑ m ( n − 1) + ω m ( n − 1)dt + (11.14)
dt 2
3.4
0.4 2.6 0.4
0.25
0.4
0.95
1.1 1.9
0.05
N
1.0 S S
N
2.65
FIGURE 11.2
Sketch and main dimensions of the motor (measured in mm).
11.6 Example
The motor used in this example is a micro-motor, whose main dimensions
are shown in Figure 11.2. The rotor is formed by an isotropic permanent
magnet, which is a ceramic Barium-Ferrite magnet, characterized by a resid-
ual flux density Bres = 80 mT and a coercive field strength Hc = –60 kA/m.
The external diameter is 1 mm and the inner diameter is 0.15 mm. Then the
permanent magnet thickness is 0.425 mm.
The stator is formed by magnetic steel laminations with a thickness of
0.4 mm, and a stack length LFe = 5 mm. The coil is placed on one side of the
lamination and is formed by Nt = 1000 turns with dc = 25 µm copper wire
diameter.
The four poles of the rotor are modeled by four different objects; the shape
of each object is a circular sector. Each is characterized by a suitable magne-
tization direction. The flux lines due to the permanent magnet only are
shown in Figure 11.3.
S S
FIGURE 11.3
Flux lines due to the permanent magnet only.
Self-Starting Single-Phase Synchronous Motors 215
0 deg 45 deg
30 deg 60 deg
FIGURE 11.4
Flux lines at different rotor positions ϑm (at no-load).
0 deg 30 deg
45 deg 60 deg
FIGURE 11.5
Flux lines at different rotor positions ϑm (under load).
216 Electrical Machine Analysis Using Finite Elements
⫻ 10−5
1.5
−0.5
−1
−1.5
0 0.5 1 1.5 2 2.5 3
Angular position (rad)
FIGURE 11.6
Flux linkage due to the permanent magnet only, simulated values, and interpolating function.
π
λ pm (ϑ m ) = −1.431 ⋅ cos 2 ϑ m − (11.15)
36
In Figure 11.6, stars indicate the simulated values while the solid line
represents the interpolating function.
TABLE 11.1
Torque Matrix (Torque as a Function of the Rotor Position and Current)
FIGURE 11.7
Motor torque as a function of the rotor position and the current.
FIGURE 11.8
Current and the torque waveforms, assuming a constant rotor speed.
Self-Starting Single-Phase Synchronous Motors 219
FIGURE 11.9
Rotor starting at ϕ = 0, ϑ(0) = 0, f = 100 Hz.
waveform, with an average value different from zero, and an almost sinu-
soidal variation at a frequency double of the source frequency.
FIGURE 11.10
Rotor starting at ϕ = π/4, ϑ(0) = 0, f = 100 Hz.
FIGURE 11.11
Rotor starting at ϕ = π/2, ϑ(0)= 0, f = 100 Hz.
600
400
Speed (rad/s)
200 157
0
−200
−400
−600
0 0.01 0.02 0.03 0.04 0.05
Time (s)
FIGURE 11.12
Rotor starting at ϕ = π/2, ϑ(0) = 0, f = 50 Hz.
y(x1, x2) Y4
Y2
Y3
Y1
x2
x1
x2(k + 1)
x1(h + 1)
x2(k)
x1(h)
0 0
FIGURE 11.13
Interpolating plane.
M × N values. From the knowledge of such values, the value of the function
y(x1, x2) corresponding to a couple of variables (x1, x2) can be interpolated,
with the constraint that x1(1) < x1 < x1(M) and x2(1) < x2 < x2(N).
The first step of the interpolation is to individuate the interval of the vector
[x1(1), …, x1(M)] in which x1 is contained. This involves individuating the
number h so that x1 > x1(h) and x1 < x1(h + 1). Analogously, with the second
variable, the number k is individuate, so that x2 > x2(k) and x2 < x2(k + 1).
The second step is to compute the four values of the function y(x1, x2)
corresponding to the combinations two by two of the variables, which is
where
x1 ( h + 1) − x1 x 2 (k + 1) − x 2
y1 (x1 , x 2 ) = Y1 (11.17)
x1 ( h + 1) − x1 ( h ) x 2 (k + 1) − x 2 (k)
x1 ( h + 1) − x1 x 2 − x 2 (k )
y 2 (x1 , x 2 ) = Y2 (11.18)
x1 ( h + 1) − x1 ( h ) x 2 (k + 1) − x 2 (k)
222 Electrical Machine Analysis Using Finite Elements
Y1
x2 x2
x1 x1
x2(k + 1) x2(k + 1) x1(h + 1)
x2(k) x1(h + 1) x2(k)
x1(h) x1(h)
0 0 0 0
Y3
x2 x2
x1 x1
x2(k + 1) x2(k + 1)
x1(h + 1) x1(h + 1)
x2(k) x1(h) x2(k) x1(h)
0 0 0 0
FIGURE 11.14
Shape of the interpolating functions.
x1 − x1 ( h ) x 2 (k + 1) − x 2
y 3 (x1 , x 2 ) = Y3 (11.19)
x1 ( h + 1) − x1 ( h ) x 2 (k + 1) − x 2 (k)
x1 − x1 ( h ) x 2 − x 2 (k )
y 4 (x1 , x 2 ) = Y4 (11.20)
x1 ( h + 1) − x1 ( h ) x 2 (k + 1) − x 2 (k)
Each of these functions assumes a value different from zero in only one
couple of variables, while it assumes a null value in the other three couples
of variables (x1, x2). For instance, the first function, say y1(x1, x2), assumes
the value Y1 in the point of coordinates x1(h), x2(k); while a null value is
assumed in the points of coordinates x1(h), x2(k + 1); x1(h + 1), x2(k); x1(h + 1),
x2(k + 1).
The shapes of the interpolating functions are reported in Figure 11.14. Once
the value that the four interpolating functions assume corresponds to the
point (x1, x2), the final value is given by their sum.
Torq =[
-0.0199 -0.1181 -0.1991 -0.2824 -0.3540 -0.4353 -0.5209 -0.6018 -0.6887
-0.3703 -0.3401 -0.3130 -0.2826 -0.2529 -0.2176 -0.1886 -0.1830 -0.1638
-0.5952 -0.4534 -0.3149 -0.1718 -0.0353 0.1128 0.2469 0.3013 0.4250
-0.6588 -0.4114 -0.1657 0.0852 0.3348 0.5853 0.7502 0.9783 1.2065
-0.5578 -0.2168 0.1297 0.4777 0.8259 1.1717 1.5294 1.7640 2.0893
-0.3540 0.0876 0.5192 0.9523 1.3854 1.8184 2.2353 2.6842 3.1178
0.0272 0.5236 1.0197 1.5073 2.0051 2.4983 2.9818 3.4268 3.9067
0.3777 0.8839 1.3895 1.8838 2.3901 2.8951 3.3917 3.8985 4.4046
0.5891 1.0601 1.5304 1.9973 2.4642 2.9299 3.4005 3.8920 4.3682
0.6945 1.0918 1.4894 1.8905 2.2862 2.6760 3.0759 3.4718 3.8642
0.5901 0.8897 1.1887 1.4882 1.7891 2.0906 2.3927 2.7005 2.9935
0.3231 0.5133 0.6991 0.8913 1.0878 1.2881 1.4801 1.6721 1.8619
-0.0203 0.0604 0.1451 0.2260 0.3109 0.3902 0.4690 0.5427 0.6223
-0.3645 -0.3963 -0.4255 -0.4644 -0.4906 -0.5186 -0.5459 -0.5756 -0.6033
-0.6089 -0.7573 -0.8969 -1.0267 -1.1709 -1.3098 -1.4490 -1.5531 -1.7047
-0.6686 -0.9161 -1.1638 -1.4114 -1.6627 -1.9120 -2.1584 -2.4063 -2.6453
-0.5572 -0.9090 -1.2660 -1.6137 -1.9622 -2.3111 -2.6610 -3.0055 -3.3552
-0.3561 -0.7805 -1.1635 -1.5888 -2.0822 -2.4954 -2.9406 -3.3722 -3.8163
0.0212 -0.4722 -0.9652 -1.4525 -1.9518 -2.4460 -2.9427 -3.4392 -3.9367
0.3763 -0.1310 -0.6395 -1.1479 -1.6571 -2.1612 -2.6639 -3.1714 -3.6780
0.5953 0.1270 -0.3495 -0.8196 -1.2911 -1.7654 -2.1984 -2.6678 -3.1371
0.6792 0.2940 -0.1014 -0.4999 -0.8977 -1.2966 -1.6961 -2.0675 -2.4615
0.5894 0.2921 -0.0111 -0.3093 -0.6091 -0.9091 -1.2146 -1.5130 -1.8462
0.3177 0.1208 -0.0708 -0.2639 -0.4556 -0.6505 -0.8508 -1.0405 -1.2357
-0.0199 -0.1178 -0.1988 -0.2821 -0.3540 -0.4353 -0.5210 -0.6018 -0.6887
]*(1e-7);
224 Electrical Machine Analysis Using Finite Elements
k=1;
z=1;
t1= Torq(z-1,k-1);
t2= Torq(z-1,k);
t3= Torq(z,k-1);
t4= Torq(z,k);
Self-Starting Single-Phase Synchronous Motors 225
com=(VectI(k)-VectI(k-1))*(VectA(z)-VectA(z-1));
val1=t1*(VectI(k)- iact)*(VectA(z)-tetamec);
val2=t2*( iact -VectI(k-1))*(VectA(z)-tetamec);
val3=t3*(VectI(k)- iact)*(tetamec-VectA(z-1));
val4=t4*( iact -VectI(k-1))*(tetamec-VectA(z-1));
value=(val1+val2+val3+val4)/com;
end;
clear;
close all;
% Constant declaration
for n=2:5000
v=vm*sin(we*t+fi); % voltage in n-th time
didt=(v-res*ip(n-1)-dfdz*wm)/la; % current increment
ip(n)=ip(n-1)+didt*dt; % current in n-th time
im(n-1)=(ip(n)+ip(n-1))*0.5; % c. average in interval
tmot(n-1)=coppia(im(n-1),tetam(n-1)); % torque in the interval
tetam(n)=tetam(n-1)+wm*dt; % position in n-th time
tavrg=tavrg+tmot(n-1); % average torque
flux(n)=(-1.431e-5)*cos(2*(tetam(n)-0.0277778*pi)); % PM flux
dfdz=flux(n)-flux(n-1); % flux variation
t=t+dt; % time increment
end;
end
clear;
close all;
% Constant definition
for n=2:5000
v=vm*sin(we*t+fi); % voltage in n-th time
Self-Starting Single-Phase Synchronous Motors 227
end
References
1. P.C. Krause, Analysis of Electrical Machinery, McGraw-Hill, New York, 1986.
2. K. Deng, M. Mehregany, and A.S. Dewa, “A Simple Fabrication Process for
Polysilicon Side-Drive Micromotors,” in IEEE Journal of Micromechanical System,
vol. 3, n. 4, December 1994, pp. 126–133.
3. A. Teshigahara, M. Watanabe, N. Kawahara, Y. Ohtsuka, and T. Hattori, “Per-
formance of a 7 mm Microfabbricated Car,” in IEEE Journal of Micromechanical
System, vol. 4, n. 2, June 1995, pp. 76–80.
12
Switched Reluctance Motors
This chapter deals with the analysis of the switched reluctance motor. Since
such a motor works in highly saturated, the chapter gives full details of how
to choose the number of finite elements and how to compute the torque
developed by the motor. We underline the effect of the number of finite
elements on the local components of the magnetic fields, from which the
torque is computed.
After we compute the electromagnetic quantities by the finite element
method, they are used for the prediction of the dynamic performance of the
motor, in a similar way as done in the previous chapter.
12.1 Introduction
The switched reluctance motor originated in 1842. However, it was not
employed until several decades later, and then thanks to the development
of the power electronic devices. The first variable-speed applications were
proposed in the 1970s by Harris and Lawrenson.
The motor has a double saliency, i.e., salient poles in both stator and rotor.
The phase currents are switched depending on the rotor position, by means
of a very simple control strategy. The motor differs from the stepper motor
for the following four reasons:
229
230 Electrical Machine Analysis Using Finite Elements
∂Wm′ (ϑ m , i)
T( ϑ m , i ) = (12.1)
∂ϑ m
where ϑm is the mechanical angle between the rotor pole and the stator pole,
as shown in Figure 12.1(a), and i is the current flowing through the stator
coil. The analysis has to be carried out considering the dependence on both
the variables ϑm and i.
Rotor ϑm Stator
axis axis
+a −a a-phase b-phase c-phase d-phase
−b −d
Voltage
+b +d source
+
−c −c Vdc
βr βs −
+c +c
λr
−d λs −b
+d +b
+a −a
(a) (b)
FIGURE 12.1
Structure of a switched reluctance motor with Qs = 8 and Qr = 6 (a); sketch of the static power
electronic converter (b).
Switched Reluctance Motors 231
1
Wm′ = Wm = L(ϑ m )i 2 (12.2)
2
1 dL(ϑ m ) 2
T( ϑ m , i ) = i (12.3)
2 dϑ m
L(ϑm)
Lmax
dL > 0 dL < 0
dϑ dϑ
Lmin
ϑm
ϑ0 ϑ1 ϑ2 ϑ3 ϑ4
βs βr − βs
λr
FIGURE 12.2
Phase inductance versus the rotor position.
dλ(t )
v(t ) = Ri(t ) + (12.4)
dt
Since the flux linkage λ is univocally determined by the current i and the
rotor position ϑm, that is, λ = λ(i,ϑm), as explained in Chapter 5, Equation
(12.4) becomes
dL(ϑ m ) di(t )
v(t ) = Ri(t ) + ω m (t )i(t ) + L(ϑ m ) (12.6)
dϑ m dt
In the motoring operations, the phase coils are fed during the inductance
increasing, fixing a specific conducting period. Let ϑon be the turn-on angle
and ϑoff be the turn-off angle. According to Figure 12.1(b), the switches are
turned on when the rotor reaches the angular position ϑon. The voltage v =
+Vdc is applied to the coil terminal forcing the current to increase. Then the
switches are turned off when the rotor reaches the angular position ϑoff, the
current flows through the two diodes, so that the voltage v = –Vdc is auto-
matically applied to the coil terminal, forcing the current down to zero.
Az = 0 Az = 0
+a −a +a −a
dAz
=0
dn
+a −a
FIGURE 12.3
Structure of the switched reluctance motor for the finite element analysis.
variables. They are necessary for the following prediction of the steady-state
and dynamic performance of the motor.
A dynamic analysis of the switched reluctance motor is practically com-
pulsory, due to the phase commutations and the nonlinearity of the magnetic
circuit. The aim of the dynamic analysis is to evaluate the waveform of the
motor torque for given turn-on ϑon and turn-off ϑoff angles. It is also possible
to determine these two angles so as to maximize the average torque and
minimize the torque ripple.
A finite element step-to-step analysis is possible, but it is not convenient.
In fact, for any change in the control strategy, a new finite element simulation
is required, with a consequent long-time consumption. Conversely, if the
magnetic model of the motor is built, so that the state functions are known,
the integration of the differential equations is carried out with no further
finite element analysis. An example is reported at the end of this chapter.
The rotor speed ωm is assumed to be fixed, neglecting the variation of the
mechanical quantities. Conversely, we focus on the electrical quantity dynam-
ics, which generally is faster.
Starting from Equation (12.5), it results that
di(t ) 1 ∂λ(i , ϑ m )
= v(t ) − Ri(t ) − ω m
∂λ(i , ϑ m ) ∂ϑ m
(12.7)
dt
∂i
where
∂λ(i , ϑ m )
Lapp = (12.8)
∂i
∂λ(i , ϑ m )
Em = ω m (12.9)
∂ϑ m
represents the motional EMF. Both of them are functions of the state variables
current i and angular position ϑm. Equation (12.7) has to be integrated numer-
ically, considering the angular position ϑm = ωmt as the integration variable.
The values of Lapp and Em are updated at each integration step ∆ϑm. At last,
the current is computed by integrating Equation (12.7), and the flux linkage
λ(ϑm,i) and the motor torque T(ϑm,i) are obtained from the model built from
the magnetostatic analysis.
Interpolation methods are used to limit the number of field analyses. Some
values of the state variables are chosen; then the tables of the corresponding
flux linkage and torque are built, and the other values are obtained by means
236 Electrical Machine Analysis Using Finite Elements
12.5 Example
As an example, a switched reluctance motor with Qs/Qr = 8/6 is analyzed.
Its main dimensions are reported in Table 12.1.
Figure 12.4 shows the flux lines when the a-phase coil is supplied. Figure
12.4(a) shows the complete section of the motor, Figure 12.4(b) shows only
half a section, and Figure 12.4(c) displays a detail of the flux lines in the
region of the stator and rotor poles when they are partially aligned. It is
important to check this region, since the two poles operate in a very saturated
TABLE 12.1
Main Dimension of the Switched Reluctance Motor
Quantity Value Quantity Value
Qs Stator pole number 8 g Air-gap 0.625 mm
Qr Rotor pole number 6 Nt Number of turns 93
βs Stator polar arc 20.1 deg R Resistance 3Ω
βr Rotor polar arc 21.6 deg Pn Rated power 4 kW
LFe Axial length 152 mm Imax Maximum current 18 A
(b)
(a)
(c)
FIGURE 12.4
Flux lines in a switched reluctance motor.
Switched Reluctance Motors 237
FIGURE 12.5
Flux linkage as a function of the rotor position (a) and of the current (b).
FIGURE 12.6
Magnetic coenergy (a) and static torque (b) versus rotor position and current.
manner. At the higher currents, they reach such a high saturation level that
the differential magnetic permeability approaches unity.
Figure 12.5 shows the flux linkage as a function of the angular position
and of the current, obtained from magnetostatic analysis. It is observed that
when the pole pairs are not overlapped or when the current is low, the flux
linkage is low. The motor works in linear conditions and the flux linkage is
proportional to the current. On the contrary, when the current is high and
when the pole pairs are overlapped, the effect of the magnetic saturation is
manifest. The flux linkage is not more proportional to the current, and the
slope of the curves decreases.
Figure 12.6(a) shows the magnetic coenergy as a function of the rotor
angular position and the phase current. At given ϑm, the magnetic coenergy
is proportional to the squared current only when the current is low and the
poles are not aligned. When the current increases, due to the saturation, it
238 Electrical Machine Analysis Using Finite Elements
Airgap flux density, normal component (T) Airgap flux density, tangential component (T)
1.8
1.6 0.4
1.4 0.3
1.2
0.2
1
0.8 0.1
0.6 0
0.4
−0.1
0.2
0 −0.2
60° 65° 70° 75° 80° 85° 90° 95° 100° 105° 110° 60° 65° 70° 75° 80° 85° 90° 95° 100° 105° 110°
Angle position (mechanical degrees) Angle position (mechanical degrees)
(a) (b)
FIGURE 12.7
Normal component (a) and tangential component (b) of the air-gap flux density with different
numbers of elements of the mesh.
Switched Reluctance Motors 239
28
(Nm)
26
18
12 (A)
16
14
12
500 1000 1500 2000 2500 3000 3500
Number of finite elements
FIGURE 12.8
Dependence of the motor torque on the number of finite elements of the mesh.
−15 0 −15 0
−30° −25° −20° −15° −10° −5° 0° −30° −25° −20° −15° −10° −5° 0°
Rotor position (mech. degrees) Rotor position (mech. degrees)
(a) (b)
−15 0 −15 0
−30° −25° −20° −15° −10° −5° 0° −30° −25° −20° −15° −10° −5° 0°
Rotor position (mech. degrees) Rotor position (mech. degrees)
(c) (d)
FIGURE 12.9
Simulations of the switched reluctance motor at constant rotor speed.
Switched Reluctance Motors 241
TABLE 12.2
Speed and Control Angles for the Dynamic Simulations
n(rpm) ϑon (deg) ϑoff (deg) Current Limitation
(a) 1000 –22.5 –10.0 No
(b) 1000 –22.5 –10.0 Yes
(c) 2000 –30.0 –15.0 No
(d) 2000 –30.0 –15.0 Yes
(e) 3000 –30.0 –15.0 No
(f) 6000 –30.0 –15.0 No
References
1. P.J. Lawrenson et al., “Variable speed switched reluctance motors,” in IEE Proc.,
Electr. Power Appl., pt. B, vol. 127, no. 4, pp. 253–65, 1980.
2. J.F. Lindsay, R. Arumugam, and R. Krishnam, “Finite-Element Analysis Char-
acterization of a Switched Reluctance Motor with Multitooth per Stator Pole,”
IEE Proc., Electr. Power Appl., vol. 133, pp. 347–353, Nov. 1986.
3. P. Materu, R. Krishnan, “Estimation of switched reluctance motor losses,” in
Conf. Rec. IEEE Ind. Applicat. Soc. Annu. Meeting, Pittsburgh, pp. 79–90, October
1988.
4. I.D. Mayergoyz, “Dynamic Preisach Models of Hysteresis,” in IEEE Transactions
on Magnetics, vol. 24, pp. 2925–2977, 1988.
5. T.J.E. Miller, Switched Reluctance Motor Drive, Claredon Press, Oxford, 1989.
6. T.J.E. Miller, Brushless Permanent-Magnet and Reluctance Motor Drive, Claredon
Press, Oxford, 1989.
7. I. Husain and M. Ehsani, “Torque Ripple Minimization in Switched Reluctance
Motor Drives by PWM Current Control,” Proc. Appl. Power Electr. Conf., Orlando,
FL, 1996, pp. 72–77.
8. P.C. Kjaer, P. Nielson, L. Anderson, and F. Blaabjerg, “A New Energy Optimiz-
ing Control Strategy for Switched Reluctance Motor,” IEEE Trans. Industry
Appl., vol. 31, no. 5, pp. 1088–1095, 1995.
9. N. Bianchi, S. Bolognani, M. Zigliotto, “Prediction of Iron Losses in Switched
Reluctance Motors,” in Proc. of 7th International Power Electronics and Motion
Control Conference, PEMC ’96, Budapest, Hungary, 2–4 September 1996, vol. 3,
pp. 223–228.
10. R. Krishnan, P. Vijayraghavan, “State of the Art: Acustic Noise in Switched
Reluctance Motor Drives,” Proc. of IEEE-IECON Conf., Aachen (D), pp. 929–934,
1998.
11. P.G. Barrass and B.C. Mecrow, “Flux and Torque Control of Switched Reluc-
tance Machines,” IEE Proc., Electr. Power Appl., vol. 145, no. 6, pp. 519–527, Nov.
1998.
13
Three-Phase Induction Motors
13.1 Introduction
Each component of the field quantities is assumed to vary sinusoidally with
the time. The symbolic notation is adopted. A dot is placed over each field
quantity to represent the complex phasor. The value at the instant t assumed
by the components Gx, Gy , and Gz of the generic field quantity G(P,t), that
depends on the point P = (x,y,z) and the time t, is given by
(P , t ) = Im G
G x (P , t ) = Im G ˆ (P)e j(ωt + ϑ x ( P ))
x x
(P , t ) = Im G
G y (P , t ) = Im G ˆ (P)e j(ωt + ϑ y ( P )) (13.1)
y y
(P , t ) = Im G
G z (P , t ) = Im G ˆ (P)e j(ωt + ϑ z ( P ))
z z
243
244 Electrical Machine Analysis Using Finite Elements
y
ϑ
x
FIGURE 13.1
Frontal section of the three-phase induction motor.
With reference to Figure 13.1, the current density vector J and the magnetic
vector potential A are normal to the considered (x, y) plane, i.e., they have
only the component parallel to the z-axis. They are
J = (0, 0, J z )
(13.2)
)
A = (0, 0, A z
This implies that the magnetic field strength vector H and the flux density
vector B have components on the (x, y) plane only, normal to the z-axis, i.e.,
,H
H = (H , 0)
x y
(13.3)
B = (B x , B y , 0)
Is1 Rs L σs, 3D
Vs1
Is2 Rs L σs, 3D
Vs2 R r, 3D
Is3 Rs L σs, 3D
Vs3 L σr, 3D
Rotor rings
effects
_
FE - 2D simulation
FIGURE 13.2
Sketch of the analysis of the three-phase induction motor.
Figure 13.2 highlights the resistive and leakage inductance of the rotor rings,
i.e., Rr,3D and Lσr,3D. In addition, the parameters can be modified to take into
account the possible rotor slot skewing.
Thanks to the geometrical and magnetic symmetry of the induction motor,
the analysis domain is always reduced to a portion of the motor section.
Such a portion is essentially established by the stator and rotor slot numbers
and by the stator winding. Often the analysis domain is equal to one pole
piece. In the case of the four-pole induction motor of Figure 13.1, the analysis
domain is as shown in Figure 13.3, to which we will refer in the following
sections.
As far as the geometry symmetry is concerned, usually there are no diffi-
culties, especially if the slot number is a multiple of the pole number. On
the contrary, particular care has to be paid to the magnetic symmetry, essen-
tially due to the stator winding.
The analysis domain can be reduced, imposing suitable periodic boundary
conditions. Moreover, the stator and the rotor can be analyzed separately.
This kind of field analysis is based on the studies by Prof. S. Williamson,
reported in the References, but is omitted hereafter.
FIGURE 13.3
A quarter of the induction motor section of Figure 13.1.
246 Electrical Machine Analysis Using Finite Elements
R r,2D
Lσ,2D
Rs Lσs,3D S
+ +
Is Ir R r,3D
S
LM
Vs jωΛs
Lσr,3D
− −
2D field analysis
FIGURE 13.4
Equivalent circuit of the three-phase induction motor, highlighting the parameters obtained
from the 2D finite element analysis.
Three-Phase Induction Motors 247
placed inside the dashed box, while outside there are the parameters due to
the 3D effects.
Figure 13.4 highlights all the parameters inside the box that are not con-
stant. The magnetizing inductance LM is a function of the stator flux linkage
Λs, considering the saturation effects, even if attributed to the stator current
only. The total leakage inductance Lσ,2D and the rotor resistance Rr,2D are
functions of the rotor frequency fr , which is of the rotor slip s, considering
the effects of the current distribution in the rotor bars.
Conversely, the parameters corresponding to the 3D effects are considered
constant. The stator winding resistance Rs and the end-winding leakage
inductance Lσs,3D are assumed to be independent of the nonuniform distri-
bution of the current. Similarly, the rotor parameters Rr,3D and Lσr,3D are
considered constant. They take into account the rotor rings and the bar
skewing. They are computed analytically and are thus included in the equiv-
alent circuit. The iron losses are considered separately; therefore the resis-
tance Ro, that represents such losses is not considered.
At last, let us observe that a Γ-type representation is adopted for the part
of circuit associated with the 2D parameter (inside the dashed box). A trans-
formation from the classic T-type representation to the Γ-type representation
is operated (see Chapter 6).
∂ 1 ∂A z ∂ 1 ∂A z
+ = J osz (13.4)
∂x µ ∂x ∂y µ ∂y
Rs Lσs,3D
+ +
Iso
LM
Vso jωΛso
− −
FIGURE 13.5
Reduction of the equivalent circuit during the no-load simulation.
∑n
nq
Iq = k jqi oj (t *) q = 1, 2 , … , Q s /2 p (13.5)
j= a , b , c pp
Qs /2 p
∑k ∫
nq 1
Λ oj = 2 pL Fe jq A zdS j = a , b, c (13.6)
n pp q= 1
Sq Sq
Three-Phase Induction Motors 249
where LFe is the net axial length of the lamination stack. From the three-
phase flux linkages, the RMS value of the no-load flux linkage Λso is obtained.
The same stator flux linkage may be also obtained starting from the air-
gap flux density distribution. It is necessary to evaluate the radial component
of the flux density Br(ϑ,t*) = B(ϑ,t*)·ur along a circumference in the middle
of the air-gap. Using the Fourier series expansion, from the distribution
Br(ϑ,t*), the amplitude of the fundamental harmonic Br1 is computed. Then,
the average value of the fundamental harmonic distribution in a polar pitch
is (2/π)Br1. The corresponding average magnetic flux per pole is
2 πDL Fe DL Fe
Φ o1 = B r 1 = B r1 (13.7)
π 2p p
and then the corresponding stator flux linkage (the peak value) becomes
k wN
Λ̂ so = Φ o1 (13.8)
p
ˆ
Λ π
Eo = ωΛ so = ω so
= f k w NΦo1 (13.9)
2 2
Λ so
( )
L M Λ so =
Iso
(13.10)
In the locked rotor test, the rated currents correspond to relatively low
source voltages, then to low magnetic fluxes. As a consequence, the magnetic
materials are assumed to be linear. The field problem is then linear and
characterized by field quantities varying sinusoidally, at the electrical fre-
quency ω = ωr = 2πfr . The magnetic vector potential has only the z-axis
component, as shown in Equation (13.2), that may be expressed by means
of the phasor notation of Equation (13.1), which is
=A
A ˆ e j(ωt + ϑ ) (13.11)
z z
with ω = ωr .
Thus, with constant magnetic permeability µ, the field problem is
described by the differential equation
1 ∂2 A
∂2 A
z
+ z
= Jsz − jω r σA (13.12)
µ ∂x 2
∂y 2
z
Is Ir Lσ,2D
Rs Lσs,3D A
+ +
Vs jωΛs LM R r,2D
S
− −
A′
2D Simulation
FIGURE 13.6
Equivalent circuit corresponding to the locked rotor test.
Three-Phase Induction Motors 251
Also in this case, the field problem is solved referring to the equivalent coil
sides. The shaft may be disregarded in the field analysis. In fact, the currents
induced in the rotor bars shield the inner parts of the rotor significantly.
From the field solution, the Joule losses in the rotor bars are
∫
1
PJr = 2 pL J z ⋅ J zdS (13.13)
2 σ Al S Al
where L is the total length of the rotor bars, σAl is the Aluminium conduc-
tivity, and SAl is the overall cross-section of the rotor bars in the simulated
pole piece. The factor 2p considers that one pole piece only is simulated.
The average value of the magnetic energy in the magnetic material is
Bˆ H
ˆ
∫
1
Wm = 2 pL Fe dS (13.14)
2 D 2 2
PJr 2 Wm
R eq = Leq = (13.15)
3I s 2 3I s 2
Is Leq
A
jωΛs Req
A′ 2D Simulation
FIGURE 13.7
Electrical parameters corresponding to the locked rotor test.
252 Electrical Machine Analysis Using Finite Elements
N2
L σs ,ew = µ o Lew λ ew (13.16)
2p
where Lew is the effective end-winding length, and λew is the specific per-
meance coefficient. Its value ranges between 0.35 and 0.55, according to the
winding type.
The leakage inductance caused by skewing is estimated by
(
L σs ,sk = L M 1 − k sk
2
) (13.17)
k sk =
(
sin ε sk 2 ) (13.18)
ε sk 2
( )
2
3 k wN
R r ,3D = ρAl (13.19)
2 πp2 S ring
Once the model of the induction motor is obtained, i.e., all the parameters
of the equivalent circuit of Figure 13.4 are found, it is possible to predict the
motor performance at different operating conditions. The motor character-
istics are computed, for instance the mechanical characteristic, the current-
speed curve, and so on. As explained earlier, the iron and mechanical losses,
needed for the computation of the efficiency, have to be considered separately.
Three-Phase Induction Motors 253
TABLE 13.1
Main Data of the Three-Phase
Induction Motor
Rated power 22000W
Rated voltage 280/485V
Line rated current 32.7 A
Rated frequency 60 Hz
Rated speed 1750 rpm
Number of poles 4
Stack length 300 mm
Air-gap length 0.5 mm
Rotor bar skewing 1 slot
Service S1
Protection IP55
Insulation Class F
Cooling IC 41
13.2.5 Example
The finite element method is applied to a commercial three-phase induction
motor with a double bar rotor. The geometry is shown in Figure 13.1. The
main data of the induction motor are reported in Table 13.1.
The number of stator slots is Qs = 36 so that the number of slots per pole
per phase is qs = 3 and the electrical slot angle is αc = 20 degrees. The stator
winding is a double-layer winding, with two coil sides per slot, each of them
formed by 14 copper conductors, thus nq = 28. The coil pitch is yq = 9,
corresponding to one slot chording. At last there are npp = 2 parallel paths.
A delta connection is adopted. The number of rotor aluminum bars is Qr = 28.
Dirichlet Dirichlet
+B −C −C +B −C −C
+B −C +B −C
−A +B +B +B −C −C +A −A +B +B +B −C −C +A
−A +B −C +A −A +B −C +A
−A +A −A +A
Ic Ib
Ib
FIGURE 13.8
Two solutions of assignment of the stator currents and the boundary conditions for the no-load
test: solution (a) and solution (b).
TABLE 13.2
Three-Phase Currents at the Time t*
for the No-Load Simulation
No-load simulation
Phase Solution (a) Solution (b)
A + 2 Ic 0
B 2 3
− Ic + 2 Ic
2 2
C 2 3
− Ic − 2 Ic
2 2
–A − 2 Ic 0
–B 2 3
+ Ic − 2 Ic
2 2
–C 2 3
+ Ic + 2 Ic
2 2
FIGURE 13.9
Flux plots in the two no-load simulations: referring to Table 13.2, solution (a) and solution (b).
FIGURE 13.10
Air-gap flux density distributions in the two no-load simulations: referring to Table 13.2,
solution (a) and solution (b).
used to halve the analysis domain. Finally, the effect of the slot openings is
clear on the distribution of the air-gap flux density.
The magnetization curve is illustrated in Figure 13.11, showing the mag-
netizing current as a function of the source voltage. In order to stress the
effect of the air-gap length, the no-load simulation is carried out on induction
motor with different air-gaps.
At last, referring to the equivalent electric circuit of Figure 13.5, the mag-
netizing inductance LM is computed as a function of the stator flux linkage
Λs = Λso. It is shown in Figure 13.12.
It is worth noticing that using a current source allows the obtained result
to be applied to motors of different stack length, but with the same lamina-
tions and the same winding distribution. In fact, the result is extended to
motors with different stack length maintaining the same slot current. Of
course, the values of line current and voltage have to be arranged according
to the number of the actual conductors.
256 Electrical Machine Analysis Using Finite Elements
14
(c) (b) (a)
(a) g = 0.4 mm
12
(b) g = 0.5 mm
(c) g = 0.6 mm
10
Current (A)
8
2
100 150 200 250 300 350 400 450 500
Voltage (V)
FIGURE 13.11
Magnetizing current versus source voltage with different air-gap lengths.
0.16
0.15
Inductance (H)
0.14
0.13
0.12
0.11
0.1
0.09
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Flux linkage (Vs)
FIGURE 13.12
Magnetizing inductance LM versus stator flux linkage Λs.
TABLE 13.3
Three-Phase Currents for the
Locked Rotor Simulation
Phase Locked Rotor Simulation
A + 2 Ic + j0
B 2 6
− Ic − j Ic
2 2
C 2 6
− Ic + j Ic
2 2
–a − 2 Ic − j0
–b 2 6
+ Ic + j Ic
2 2
–c 2 6
+ Ic − j Ic
2 2
FIGURE 13.13
Flux plots of the locked rotor simulations, with source frequency (a) f = fr = 60 Hz, and (b) f =
fr = 5 Hz.
0.9 1.6
0.8 1.4
Resistance (Ohm)
Inductance (H)
0.7 1.2
0.6 1
0.5 0.8
0.4 0.6
0.3 0.4
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Frequency (Hz) Frequency (Hz)
FIGURE 13.14
Variation of the resistance Rr,2D and of the total leakage inductance Lσ,2D with the rotor frequency fr .
In the same Figure 13.15, some experimental results are reported, by circles.
Comparing simulated and measured data, the agreement is satisfactory,
during operations both at standstill and under load.
Analogously, Figure 13.16 shows the stator current as a function of the
motor speed, while Figure 13.17 shows the power factor as a function of the
phase current.
400
350
250
200
FIGURE 13.15
Torque-speed characteristic.
Current (A)
160
140
120
100
80
FIGURE 13.16
Stator current versus speed.
Power factor
0.9
0.8
0.7
0.6
0.5
0.4
FIGURE 13.17
Power factor versus stator current.
∂A
J = Js − σ + σv × B (13.20)
∂t
Because of the 2D problem, with only the z-axis components of the current
density vector J and the magnetic vector potential A, as indicated in Equation
(13.2), the field problem is described by the following differential equation:
∂ 1 ∂A
∂ 1 ∂A ∂A
z
+ z
= Jsz − σ z )
+ σ v × curl(0, 0, A (13.21)
∂x µ ∂x ∂y µ ∂y ∂t
z
∂A
J i = −σ z
∂t
)
J m = σ v × curl(0, 0, A z
In the same way, the z-axis component of the magnetic vector potential is
=A
A ˆ e j(ωt − ϑ ) (13.23)
z z
∂A
z ˆ e j(ωt − ϑ ) = jωA
= jωA (13.24)
∂t
z z
From the first Maxwell’s equation (1.102), in which the displacement cur-
rent density is negligible because of the low frequency, it is
curlH = J (13.25)
∂A
curlH + σ − σv × B = Js (13.26)
∂t
v = rω ru ϑ (13.28)
z
P
uz uϑ
ϑ
ur
r
FIGURE 13.18
Cylindrical coordinate system.
(
v × B = rω ru ϑ × curlA ⋅ u r ⋅ u r ) (13.29)
1 ∂A ∂A
curlA = z
ur − z
uϑ (13.30)
r ∂ϑ ∂r
1 ∂A ∂A
v × B = rω ru ϑ × z
ur − z
uϑ ⋅ ur ⋅ ur (13.31)
r ∂ϑ ∂r
∂A
v × B = −ω r z
uz (13.32)
∂ϑ
ω − ωr
Finally, introducing the equivalent conductivity σ eq = σ , it is possi-
ble to write ω
∂ 1 ∂A
∂ 1 ∂A
z
+ z
∂x µ ∂x ∂y µ ∂y
( )
= Jsz − j ωσ
eq
A z (13.35)
B
Beq (Hm)
Bm (Hm )
FIGURE 13.19
Equivalent B-H curve assuming a sinusoidal waveform of the magnetic field strength.
Beq T B(t)
∫ ∫∫
1 1
H m dBeq = H dB dt (13.36)
2 0 T 0 0
∂ 1 ∂A
∂ 1 ∂A
∂V
z
+
z
∂x µ ∂x ∂y µ ∂y =σ
∂z
− j ωσ ( ) eq
A z (13.37)
Three-Phase Induction Motors 265
13.3.6 Example
The second analysis approach is applied on the same induction motor of the
previous example. The simulation is carried out forcing three symmetric
sinusoidal voltages with the RMS value Vs = 385 V at the motor terminals,
which is in Figure 13.2. The various electrical lumped-parameters are added,
to consider the 3D effects, and the magnetization characteristic is set up.
Each simulation is carried out at a fixed rotor speed.
To have an easy comparison with the previous approach, the results are
reported in Figure 13.14 to Figure 13.16, using stars. It is observed that the
results obtained by means of this approach agree with the measured one.
The highest error is observed on the mechanical characteristic comparing
the maximum torque. The value of the maximum torque is the same with
the two analysis approaches, while the speed at which this value is reached
is different.
References
1. T. Trnhuvud, K. Reichert, “Accuracy problems of force and torque calculation
in FE-systems,” IEEE Trans. Industry Appl., vol. 24, n. 1, pp. 443–446, 1988.
2. G.R. Slemon, “Modelling of induction machines for electric drives,” IEEE Trans.
on Industry Applications, vol. 25, n. 6, pp. 1126–1131, 1989.
3. E. Vassent, J.C. Sabonnadiere, “Simulation of induction machine operating
using complex magnetodynamic finite elements,” IEEE Trans. on Magnetics,
vol. 25, n. 4, pp. 3064–3066, 1989.
4. R. Belmans, D. Verdyck, T.B. Johansson, W. Geysen, R.D. Findlay, “Calculation
of the no-load and torque speed characteristic of induction motors using finite
elements,” Int. Conf. on Electrical Machines, Boston, 1990, pp. 2724–2729.
5. S. Williamson, M.J. Robinson, “Calculation of induction motor equivalent cir-
cuit parameters using finite elements,” IEE Proc., Pt. B, vol. 138, n. 5, pp.
264–276, 1991.
6. P.L. Alger, Induction Machines. Their Behavior and Uses, 2nd ed., Gordon and
Breach, Basel, 1995.
7. S. Williamson, D.R. Gersh, “Finite element calculation of double-cage rotor
equivalent circuit parameters,” IEEE Trans. on Energy Conversion, vol. 11, n. 1,
pp. 41–48, 1996.
266 Electrical Machine Analysis Using Finite Elements
267
268 Electrical Machine Analysis Using Finite Elements
2.0
1.9
1.8
1.7
1.6
Induzione tesla
1.5
1.4
1.3
1.2 TERNI-ARMCO M5T30
1.1 Curva di magnetizzazione inc.c.
D-C magnetization curve
1.0 Prova dopo ricottura di distensione
0.9 Test: SRA; parallel; A 341
0.8
0.7
0.6
0.5
1 2 3 5 7 10 2 3 5 7 102 2 3 5 7 103 2 3 5 7 104
Forza di magnetizzazione- Asp/m
FIGURE A.1
Grain-oriented lamination.
2.0 8 10−3
1.8 7 10−3
1.6 6 10−3
Permeabilita’ µ-H/m
1.4 5 10−3
Induzione tesla
0.8 2 10−3
0.6 1 10−3
0.4 0
10 2 3 5 7 102 2 3 5 7 103 2 3 5 7 104 2 3 5 7 105
Forza magnetica-Asp/m
FIGURE A.2
Silicon iron laminations for rotating machines.
Material Data 271
B (T)
1.3
3
1.1
1 Ferrite di Bario
2 AINiCo 5 2 0.9
3 AINiCo 8
0.7
4 MnAIC 7
5 SmCo5 4 0.5
5
6 Sm2Co17
1 0.3
7 NdFeB
6
0.1
H (kA/m)
900 700 500 300 100
FIGURE A.3
B-H curves of some permanent magnet materials.
272 Electrical Machine Analysis Using Finite Elements
1.5 2 2.5
500
B·H
B
kJ/m330 mT
−µ0 · H 20 10 5
1 400
KOEROX 300
300
KOEROX 330
KOEROX 360
0.5 200 B
KOEROX 300 K
KOEROX 380
KOEROX 330 K
100
KOEROX 150
KOEROX 100
0
−300 −250 −200 −150 −100 −50 kA/m 0
H
FIGURE A.4
Ferrite permanent magnet.
6 600
0.2
4 400
0.1
2 200
0
KOe −26 −24 −22 −20 −18 −16 −14 −12 −10 −8 −6 −4 −2 −0
2000 1800 1600 1400 1200 1000 800 600 400 200 0
H (kA/m)
FIGURE A.5
Neodimium–iron–boron permanent magnet.