Tải bản đầy đủ (.pdf) (15 trang)

DSpace at VNU: Eliminated Common-Mode Voltage Pulsewidth Modulation to Reduce Output Current Ripple for Multilevel Inverters

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (3.76 MB, 15 trang )

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

1

Eliminated Common-Mode Voltage Pulsewidth
Modulation to Reduce Output Current Ripple
for Multilevel Inverters
Tam-Khanh Tu Nguyen, Nho-Van Nguyen, Member, IEEE, and Nadipuram (Ram) R. Prasad, Member, IEEE


complete CMV elimination can be found in [11-17].

Abstract— The paper presents an analysis on the output
current-ripple in zero Common Mode Voltage (ZCMV) PWM
control of multilevel inverters. The modulation strategy for
Common Mode Voltage (CMV) elimination in multilevel
inverters is based on the “three zero common mode vectors”
principle. The space vector diagram, which consists of vectors of
zero common mode voltage, is fully explored by properly
depicting the base voltage vectors and their corresponding active
switching vectors. The switching patterns are limited to those of
three switching states each of which is symmetrically distributed.
Based on the PWM process simplified to that of a two-level
inverter with three allowable switching states and the degree of
freedom existing in the switching states arrangement, a novel
carrier-based PWM method with optimized output current
ripple is proposed. Compared to the existing PWM methods
which utilize the same kind of switching patterns, the proposed


PWM method has reduced considerably the RMS current ripple
and total harmonic distortion (THD) of the output-current in a
wide region of the modulation index. The effectiveness of the
proposed method is validated by both simulation and
experimental results.

SW1A

Vdc

SW2A

AC motor
C

B

N

A

O

g

g

Vdc

a)

Vdc

Vdc
SW1A

SW3A

SW2A

SW4A

Vdc

Vdc

A

AC motor
O

g

Vdc

Vdc

B

Multilevel Inverter, Common-mode
voltage, Current-ripple, Harmonic distortion, Pulse Width

Modulation (PWM), PWM inverters.
Index

Terms—

N

g

C

I.

INTRODUCTION

C

ommon-mode voltages are associated with excessive
bearing currents, which may cause premature motor
bearing failure and electromagnetic interference [1-2].
There are a number of approaches to cope with the CMV
issue, including the use of extra hardware with passive and/or
active devices [3-4]. However, the extra hardware causes a
significant increase in the system’s volume or creates much
more complex methods of control.
The multi-level inverters [5-6], as shown in Fig.1, have a
high number of switching states that can either reduce or
eliminate the CMV. Based on this advantage, many studies of
CMV mitigation have been conducted using multi-level
inverters [7-17]. The PWM methods with partial CMV

elimination are presented in [7-10] while PWM methods with
Tam.-K.T. Nguyen and Nho.-V. Nguyen are with the Department of
Electrical Engineering, Ho Chi Minh city University of Technology, Ho Chi
Minh City, Viet Nam (e-mail: ; ).
Nadipuram (Ram) R. Prasad is with the Klipsch Scool of Electrical and
Computer Engineering, New Mexico State University, Las Cruces, NM,
88003-8001, USA (e-mail: )

b)
Fig. 1. Multilevel Inverter system: a) three-level NPC inverter; b) five-level
cascaded inverter

In previously reported works [11-12], the ZCMV PWM is
applied to a three-level Neutral Point Clamped (NPC) Inverter.
In [13], the PWM strategies for CMV reduction/elimination
are developed for a five-level NPC inverter. Work reported in
[14] proposes PWM strategies for partial and complete CMV
elimination in cascaded multilevel inverters. However, the
degree of freedom in the switching states arrangement is not
investigated in the mentioned works. In [16], a PWM strategy
that utilizes one nearest vector of ZCMV with respect to the
desired output voltage vector during one carrier cycle is
presented. The method, however, is suitable for a high
number of inverter levels at which the existing voltage error
can be ignored. A solution for ZCMV PWM in multilevel that
takes into account the degree of freedom in the switching
sequence arrangement is recently reported in [17]. The work
has proposed a simple carrier based PWM method for

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See

for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1
multilevel inverter using the three ZCMV vectors principle.
The resultant switching patterns are limited to those of a
minimum number of switching states, which helps to globally
reduce the switching loss. By using a proposed current-based
mapping technique, the switching loss can be locally reduced
up to 25% as compared to the non-optimized algorithms.
The current ripple reduction is significant in practice.
Reducing the current ripple can result in a lower torque
pulsation of motor drive [24-25], reduction in motor acoustic
noise [27], and also a reduction in motor heating. The PWM
techniques for a reduced current ripple are specifically useful
for high power applications, where the switching frequency is
quite low [25],[39]. For these reasons, the current ripple has
been the subject of extensive research for decades [18-39].
The current ripple analysis of a two-level inverter can be
found in many works [18-33]. The work reported in [18]
analyzed the impact of zero space vector distribution on the
output current ripple. In [19], the switching sequence
arrangement of the space vector PWM (SVPWM) is
considered for the purpose of reducing the output current
ripple. It has been shown in [20-21] that the discontinuous
PWM strategy (DPWM) can be properly applied so that the
output current ripple is reduced compared to PWM methods

with the same average switching frequency. In the work [24],
the geometrical analysis of the current ripple vector
corresponding to a conventional SVPWM pattern is presented
and the CMV is utilized as a degree of freedom to minimize
the RMS output-current. The latter works [25-27] proposed a
more generalized method of output current ripple analysis.
The switching patterns can be extended for division of active
vector time [25],[27] and an effective hybrid PWM technique
which combines multiple switching sequences to reduce the
current ripple is further suggested [26]. Recently, the output
current ripple analysis and optimized control methods have
been developed for the multi-phase two-level systems [28-31].
Also, current ripple analysis in term of the peak-to-peak value
has been reported in the works [32-33].
The current ripple analysis is further extended to high level
inverters as can be found in [33-37]. The works [36-37]
present investigations on the current ripple for the three-level
operation based on dual two-level VSI. The DPWM method is
also employed in [37] to reduce the switching loss as well as
the current ripple. The work [34] introduces novel switching
sequences of the three-level NPC inverter with corresponding
current ripple investigations. Based on [34], a hybrid PWM
technique [35] is proposed that helps reduce the current ripple.
In ZCMV PWM control, greater distances between the three
active vectors lead to increased current ripple and higher THD
compared to the conventional PWM control [12],[22].
Although the ZCMV PWM for multilevel inverters has been
the main subject of many aforementioned works [12-17], no
study has considered the ZCMV PWM technique for current
ripple reduction thus far. The Switching loss optimizing PWM

[17] that considers using the degree of freedom in the
switching states arrangement, however, results in higher
output current ripple than other PWMs with non-optimized
algorithms.

2
In this paper, an output current ripple analysis of multi-level
inverters under condition of ZCMV is presented. The
designed switching patterns satisfy the “three ZCMV vectors”
principle with a minimum number of commutations. The
current ripple is theoretically investigated using the notion of
harmonic flux. Based on the degree of freedom in the
switching states arrangement, a PWM strategy is proposed to
optimize output current ripple. The rules to select the optimum
switching sequence deduced from the current ripple analysis
are simple for an online implementation. Also, the proposed
PWM method can be simply applied to a high level inverter
without losing the generality.
The proposed ZCMV PWM method is developed and
simulated for the three-level NPC inverter and the five-level
cascaded inverter. The harmonic flux characteristics obtained
from simulated data of the proposed ZCMV PWM method and
existing ZCMV PWM methods characterized by the same
kind of patterns in [17] are presented to highlight the
improved performance of the RMS current ripple. Also,
comparisons of the weighted total harmonic distortion
(WTHD) factor characteristics of the line voltage are shown to
demonstrate the improved THD of the proposed PWM
method. In experiment, the proposed PWM method is verified
on a five-level cascaded inverter fed constant volt-per-hertz

(V/f) drive. Comparative results of the RMS current ripple
characteristics as well as the current THD characteristics are
shown for experimental validation.
II. PWM METHOD TO ELIMINATE COMMON MODE
VOLTAGE
The ZCMV PWM in multi-level inverters has been
analyzed in detail in the previous work [17]. For the sake of
clarity of the proposed PWM method development, the
analytical method is briefly summarized in this section.
In the systems of multi-level inverter fed AC motor drive as
described in Fig. 1, the CMV is defined as the voltage
difference between the stator winding neutral N and the midpoint O of the DC-link [13]. The CMV in Fig. 1 can also be
expressed in terms of the pole voltages (each of which
measured from one output terminal to the mid-point O of the
DC-link voltage) as:
V  VBO  VCO
(1)
VCM  VNO  AO
3
The condition of ZCMV defines the space vector diagram
of a five-level inverter with 19 switching combinations of
(VAO, VBO, VCO) of ZCMV as shown in Fig. 2.
Assume that each DC-link voltage defined as Vdc in Fig. 1 is
ideal DC voltage. For a three phase n-level inverter, the
modulation index can be defined as:
v1
m
(2)
n 1
Vdc

3
where v1 is the peak value of the fundamental component of
the actual phase voltage and (n  1)Vdc / 3 is the maximum

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

3

peak value of the fundamental voltage in linear control that the
inverter can produce. In linear modulation of ZCMV PWMs
control, the maximum value of the modulation index defined
by (2) is 0.866 [12-17].
 axis

b axis
040

041

043
044

031


032
033

122

024
014

013

004

213

204

301

302
303

203

103

411

400


integer value of v Xn .
The condition of zero average CMV leads to:

 axis
a axis

F  FL  Fe  3(n  1) / 2

401

403

c axis
Fig. 2. Five-level space vector diagram with zero CMV states (bold letters).

The analytical process for the diode-clamped inverter and
cascaded inverter is unified by a simple voltage modeling. For
an n-level inverter of the two topologies, the pole-voltage
VXO ( X  A, B, C) can be generalized as:
n 1

VXO  (

s

jX

j 1

)Vdc 


n 1
Vdc
2

(3)

s1X  s2 X  ...... sn  2 X s n 1X ; X  A, B,C (for the diode
clamp inverter topology )
(4)
where s1X, s3X,…, sn-1X represent the switching states of
switches SW1X , SW2X ,…, SWn-1X which, for example, are
designated in Fig. 1 for the A-phase of the two topologies; s1X
is 1 if SW1X is on, otherwise its value is 0.
n1

The component

s

jX

(10)

402

404

304


(9)

where, Int (v Xn ) denotes a function that returns a nearest lower

410

311

202

Normalized state
of the base
voltage vector

420
310

312

212

113

104

211

222

 X  v Xn  LX ; X  A, B, C


430

320
321

221

112

114

330

Int (v Xn ) if v Xn  n  1
LX  
;0  LX  n  2; X  A, B, C
n  2 if v Xn  n  1
(8)

Normalized
switching state
of zero CMV

440

220

231
121


123

023

034

131

132
022

340
230

130

141

042

240

140

deduced as follow:

( X  A, B, C ) in (3) is called the

j 1


normalized switching voltage which is denoted as V Xn :

where, FL  LA  LB  LC and Fe   A   B   C
The functions F , FL , Fe are termed the total switching
voltage, total base voltage, and total active voltage,
respectively. The values of FL and Fe which are available for
ZCMV PWM control, are limited to two specific cases as:
-The case of FL  3(n  1) / 2  2 , Fe  2 , which is realized
with three active switching states of (1,1,0), (0,1,1), and
(1,0,1) in the active voltage hexagonal diagram as shown in
Fig. 3(a).
- The case of FL  3(n  1) / 2  1 , Fe  1 , which is realized
with three active switching states of (1,0,0), (0,1,0), and
(0,0,1) in the active voltage hexagonal diagram as shown in
Fig. 3(b).
010

Medium
triangle

110

011

100

101

001


FL  3(n  1) / 2  2

the area meets the
condition of Fe =2

a)

010

Medium
triangle

110

011

100

101

001

FL  3(n  1) / 2  1

the area meets the
condition of Fe =1

b)


Fig. 3. Medium triangle active voltage vector diagrams: a) active switching
states for Fe  2 ( FL  3(n  1) / 2  2 ); b) active switching states
for Fe  1 ( FL  3(n  1) / 2  1 ).

n 1

V Xn 

s

jX

( X  A, B, C )

(5)

j 1

V Xn can be decomposed into two components L X and s X :
V Xn  LX  s X

(6)

where L X is a constant integer value that represents the base
component of V Xn and s X is the active component of V Xn ,
which value can be 0 or 1.
The average value of V Xn can be expressed in terms of  X - the
average active component of s X in a carrier cycle as:
v Xn  LX   X ; (0   X  1,  X  1 if v Xn  n  1)


(7)

The value of the base voltage and the active voltage can be

In the space vector diagram with ZCMV of a five-level
inverter as shown in Fig. 2, considers for example the case
when the tip of the reference voltage vector is enclosed by the
medium triangle defined by tips of three vectors
corresponding to normalized switching states of
(3,2,1),(4,1,1),(4,2,0). The normalized state of the base voltage
vector is determined as ( LA , LB , LC ) = (3,1,0) which yields

FL  4 , and the medium triangle active voltage vector diagram
is of the case in Fig. 3(a). Using similar analysis, total 24
equilateral triangles defined by sets of three zero common
mode vectors are obtained in Fig. 2: twelve triangles meet the
condition FL  FL1  4 (Fig. 3(a)) and confine the light area;
the others satisfy FL  FL1  5 (Fig. 3(b)) and cover the shaded
area.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

4


The PWM switching state sequence of the active voltage
vectors in the ZCMV PWM control can be then grouped into
two PWM patterns related to the total active voltage Fe as
shown in Fig. 4. For three-phase outputs with the use of the
two Patterns in Fig. 4, Table I lists six possible mapping
functions. Different mapping functions result in different
three-phase active switching sequences.
1

v*X 1 (X  A, B, C)

v*X 1 n  1

Vdc
2

v Xn 

L X  int(vXn )

 X  v Xn  LX
Fe   A   B  C

1
Fe  1

 s1

1  s2


 s1

1  s2

y
Select Pattern I

0

0

s1

0 0 1

s2

1 0

d

0 1 0
T1 T2
2 2

a)

s1


0 11

11 0

0 1

s2

1 1 0

0 1 1

1 0 1

1 0 1

0 1 0

d

T2 T1
2 2

T3

Active switching states
and time diagram

1 0 0


Pattern I
Fe   s1   s 2   d  1

T1 T2
2 2

b)

T3

TABLE I
POSSIBLE MAPPING FUNCTIONS AND MODULATING SIGNALS
DETERMINATION
Ad
Ad
A  s2
A  s1 A  s 2 A  s1
Bd

B  s2

B  s1

C  s2

C  s1

C  s2

C  s1


Cd

Cd

s1   B
s2  C

s1  C
s2   B

s1   A

s1  C
s2   A

s1   A

s1   B
s2   A

s2  C

s2   B

Since a commutation of the d-sequence in Fig. 4 happens
simultaneously with one from both sequences s1 and s2 , it is
sufficient to use two modulating voltages  s1,  s 2 to deduce
the switching time diagram of the proposed PWM method.
The modulating voltages 1 ,  2 are determined based on the

mapping function as presented in Table I. The switching time
diagram can be derived accordingly by comparing
1 ,1   2 with a unit carrier as in Fig. 4. The duty cycles can be
obtained as follows:

T1   s 2 .Ts

T3   s1.Ts
T  T  T  T
s
1
3
 2

and

T1  (1   s1 ).Ts

T3  (1   s 2 ).Ts
T  T  T  T
s
1
3
 2

for Pattern I

Pulse
Generator


III. OUTPUT-CURRENT RIPPLE MINIMIZATION

Fig. 4. Two Standardized virtual PWM patterns from the three zero common
mode vectors.

Bd

s A ,s B ,sC

Fig. 5. Block diagram of the proposed PWM method to eliminate
common-mode voltage.

Pattern II
Fe   s1   s 2   d  2

B  s2

Select Pattern II

Selected Mapping
function

T2 T1
2 2

B  s1

n
Fe  2


(11)

A. Simplification of output current ripple analysis in
multilevel inverter with ZCMV PWM control
The approximate current ripples expression in the Nth
carrier cycle can be described as follows [22]:
1
~
iX 
L

( N 1) TS

 (V

XN

 v *X 1 )dt ; X  A, B, C

(13)

NT S

where, V XN ( X  A, B, C ) is the output phase voltage measured
from an output terminal to the load neutral N.
Under
condition
of
zero
CMV,

the
voltage VXN ( X  A, B, C) is identical to the pole voltage V XO and
v *X 1 ( X  A, B, C ) is equal to the average value of V XO which is

denoted as v XO . Equation (13) is further derived as:
1
~
iX 
L

( N 1) TS

 (V

NT S

XO

 v XO )dt 

1
 X ; X  A, B, C
L

(14)

The component  X in (14) is called the X-phase harmonic flux
in the Nth carrier cycle.
Based on (3)-(7) as previously analyzed in Section
II, V XO and v XO both consist of the same constant voltage

component ( LX  (n  1) / 2)Vdc during a carrier cycle. The Xphase harmonic flux can be then expressed in terms of the
instantaneous active switching voltage Vdc .s X and average
active switching voltage Vdc . X as follows:

for Pattern II

(12)

( N 1) TS

 X  Vdc

 (s

X

  X )dt; X  A, B, C

(15)

NT S

If we define v *X 1 ( X  A, B, C ) as the reference load
voltages, the proposed ZCMV PWM method is described as in
Fig. 5.

Eq. (15) shows that the output-currents ripple analysis of a
multilevel inverter under condition of ZCMV can be
simplified to that of a two level inverter.


0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

5

Eq. (15) can also be expressed in another form as:

X 

VdcN
n 1

The normalized harmonic flux of the phase of double
switching is described as follows:

( N 1) TS

 (s

X

  X )dt; X  A, B, C

(16)


NT S

where, VdcN  (n  1)Vdc .
It should be noted that inverters of equal maximum linear
output voltage yield the same value of VdcN . Based on the
normalized (with VdcN ) value of  X in (16), comparative results
of harmonic fluxes given by topologies of different levels are
obtained.
B. Proposed PWM strategy for reduced RMS output current
ripple
Vdc

Vdc

0

Vdc
0

0

dn

( d ).(t  t ( d )0 );

( d ).(t ( d )1  t ( d )0 )  (1   d ).(t  t ( d )1 );
1



( d ).(t  t ( d )3 );
(n  1)TS 
( d ).(t ( d ) 4  t ( d )3 )  (1   d ).(t  t ( d ) 4 );
( ).(t  t );
( d )6
 d

t ( d )0  t  t ( d )1
t ( d )1  t  t ( d ) 2
t ( d ) 2  t  t ( d ) 4 (19)
t ( d ) 4  t  t ( d )5
t ( d )5  t  t ( d ) 6

for pattern I, and:

dn

(1   d ).(t  t ( d )0 );
t ( d )0  t  t ( d )1

(1   d ).(t ( d )1  t ( d )0 )  ( d ).(t  t ( d )1 ); t ( d )1  t  t ( d ) 2
1


t ( d ) 2  t  t ( d ) 4 (20)
(1   d ).(t  t ( d )3 );
(n  1)TS 
(1   d ).(t ( d ) 4  t ( d )3 )  ( d ).(t  t ( d ) 4 ); t ( d ) 4  t  t ( d )5
(1   ).(t  t );
t ( d )5  t  t ( d ) 6

d
( d )6


for pattern II.

t (s1)0 t (s1)1

t (s1)2

t (s1)3

t (s1)4

t (s2)0 t (s2)1

The mean-square values of the harmonic fluxes (over one
carrier cycle) corresponding to the s1 , s2 , d sequences are
determined as:

t (s2)3 t (s2)4

b)

a)
Vdc

Vdc

Vdc

0

t (s2)2

0

0

Vdc

Vdc
0

2s1nrms 

0


t (d)0 t (d)1 t (d)2

t (d)3

c)

t (d)4 t (d)5 t (d)6

t (d)0 t (d)1 t (d)2

t (d)3


2
s 2 n  rms

t (d)4 t (d)5 t (d)6

Considering the generalized PWM patterns described in
Fig. 4, four types of active switching voltage waveform with
their corresponding harmonic fluxes can be classified as in
Fig. 6. For both cases of the generalized switching pattern in
Fig. 4, the active switching voltage waveform corresponding
to the output phase which is mapped to the s1 sequence is
shown in Fig. 6(a).
Since the switching frequency is much higher than the
reference, the average switching voltages during one carrier
cycle can be assumed to be constants. Solving (16), the
normalized (with VdcN .Ts ) harmonic fluxes corresponding to
the s1 , s 2 sequences are described, respectively, as:

s 2 n

1

TS

t( s1) 0 TS



2
s 1n


2dnrms 

1
TS

dt

(21)

dt

(22)

t( s1) 0

t( s 2 ) 0 TS



2
s 2n

t( s 2 ) 0

d)

Fig. 6. Harmonic flux trajectories corresponding to phase of a) s 1- sequence
b) s2- sequence c) d -sequence (Pattern I) and d) d -sequence (Pattern II).


s1n

1
TS

t( d ) 0 TS



2
dn

(23)

dt

t( d ) 0

Using (21)-(22) with s1n , s 2 n expressed in (17)-(18) , and
Fig. 4 taken into account, 2s1nrms , 2s 2 nrms can be obtained as:

2s1nrms 

1
12(n  1) 2

2s 2n  rms 

. s21.(1   s1 ) 2


1
12(n  1)

2

. s22 .(1   s 2 ) 2

(24)

(25)

2
Similarly, the value of  dnrms for pattern I and pattern II can

be determined, respectively, by substituting  d n deduced from
(19) and (20) into (23). The simple form of 2dnrms is finally
obtained as:

( s1 ).(t  t( s1) 0 ); t( s1) 0  t  t( s1)1

1

(1  1 ).(t  t( s1) 2 ); t( s1)1  t  t( s1) 3
(n  1)TS 
 s1.(t  t( s1) 4 ); t( s1) 3  t  t( s1) 4


(17)

(1   s 2 ).(t  t( s 2 ) 0 ); t( s 2 ) 0  t  t( s 2 )1


1

 ( 2 ).(t  t( s1) 2 ); t( s 2 )1  t  t( s 2 ) 3
(n  1)TS 
(1   s 2 ).(t  t( s 2 ) 4 ); t( s 2 ) 3  t  t( s 2 ) 4

(18)

2dnrms 
2dnrms 

1
. 2 .( s21   s22   s1. s 2 )
2 d
12(n  1)
1

12(n  1) 2

for Pattern I (26)

.(1   d ) 2 .((1   s1 ) 2  (1   s 2 ) 2  (1   s1 ).(1   s 2 ))

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.

for Pattern II (27)



This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

6

The mean-squared normalized harmonic flux of X phase
(X=A,B,C) is defined as follows:

2s1nrms if X  s1

FX  2s 2 nrms if X  s2 ; X  A, B, C
2
 dnrms if X  d

(3  2.( A   C )   A . C ) in (34) are always be negative. As a
result, the conditions expressed in (34) are further reduced to
the simplest form as:

 A  Max( A ,  B , C )

(28)

For the pattern II, the condition expressed in (33) leads to:

The total mean-squared normalized harmonic flux of three
phases (over one carrier cycle) is expressed based on (28) as:

F  FA  FB  FC


(29)

 A  Min( A , B ,C )

where the role of sequences s1 , s 2 are altered. Therefore, the
RMS current ripple can be minimized by selecting a phase of
double pulses switching so that the resulting F is at a
minimum.
define F ( Ad ) , F ( Bd ) and

F ( Cd ) the

values

of F

corresponding to three cases of A  d , B  d and C  d ,
respectively. By using (24)-(29) and Table I, the values of
F ( X d ) , X  A, B, C for Pattern I are obtained as:
F ( Ad ) 

1
12(n  1) 2

( B2 .(1   B ) 2   C2 .(1   C ) 2   A2 .( B2   C2   B . C ))

(30)
F ( Bd )


1

( A2 .(1   A )2  C2 .(1  C )2   B2 .( A2  C2   A.C ))
12(n  1)2

(31)
F (C d ) 

1
12(n  1) 2

( A2 .(1   A ) 2   B2 .(1   B ) 2   C2 .( A2   B2   A . B ))

(32)
The mapping rule ( A  d ) is selected when the value of

(36)

Similarly, the conditions of using the mapping rules ( B  d )
,( C  d ) are derived, respectively, as:

The problem of output current ripple minimization can be
solved by finding an available three-phase switching sequence
that is corresponding to the minimum value of the function
F in (29) in each carrier cycle. As described in Table I, there
are six possible mapping functions that decide the three-phase
switching sequence. However, it can be concluded from (24)
and (25) that the sum ( 2s1nrms  2s 2 nrms ) is the same in the case

We


(35)

 B  Max ( A , B ,C ) for Pattern I

 B  Min( A , B ,C ) for Pattern II

 C  Max ( A , B , C ) for Pattern I

 C  Min( A , B , C ) for Pattern II

(37)

(38)

The conditions expressed in (35)-(38) divide the two
medium triangles in Fig. 3 into three separate regions each of
which is designed with a different mapping function. As
previously described in (16), the ripple analysis of the PWM
method using three nearest ZCMV vectors can be simplified to
the ripple analysis of a two-level inverter defined by the two
medium triangles in Fig. 3. Thus, the mapping algorithm
designed for the whole vector diagram of the multilevel
inverter can be derived based on Fig. 7. For example, the
space vector diagram with the proposed mapping technique
for minimizing the output current ripple of a five-level inverter
is shown in Fig. 8. It should be noted that each separate region
in the medium triangle defines the phase of double pulse
distribution and the two other phases each of which can be set
arbitrarily to the s1 or s2 sequence. If one mapping rule in

Table I is utilized during active time of one region, there is no
additional commutations between two subsequent carrier
cycles. However, at the transition between two separate
regions when the phase of double pulse is changed, additional
commutations may arise. The s1 or s2 sequence design is then
constrained to the condition of minimized additional
commutations caused by region transitions.

F ( Ad ) satisfies:

F ( Ad )  min(F ( X d ) ); X  A, B, C

(33)

Ad

B
010

By mean of (30)-(32), solving (33) leads to the following
conditions:
( A   B ).(3  2.( A   B )   A . B )  0
(34)

( A   C ).(3  2.( A   C )   A . C )  0

110

the components (3  2.( A   B )   A . B ) and


010

Bd

110

Cd
100

011

101

001

A

Cd

A

011

100

101

001

C


C

Since 0   X ( X  A, B, C)  1 and  A   B   C  1 for Pattern I,

Ad

B

Bd

FL  3(n  1) / 2  2
a)

FL  3(n  1) / 2  1
b)

Fig. 7. Proposed double pulse mapping regions for reducing output current
ripple for a) Pattern I, and b) Pattern II.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

7

 A ,  B , C

b axis
040

042

044

033

132

013

024
014

004

203

103

104

303

204


yes

no

Pattern I
yes

 A  MX

Ad

Ad

 axis

411

400

a axis

401

402

302

yes

 A  MN


410

301

312

Pattern II

420

311

202

 A   B  C  1

310

321

212
213

430
320

211

222


113

114

220
221

112

440
330

230

121

123

340

231

122

022
023

034


131

031

032

043

130

141

MN  Min( A ,  B , C )
MX  Max( A ,  B , C )

240

140

041

 axis

no

no
yes

 B  MN


yes

 B  MX

Bd

Bd

no

no
C  MN

C  MX

Cd

Cd

403

a) determination of phase of double pulses
304

404

c axis
Fig. 8. Illustration of the proposed double pulse mapping algorithm applied
to the whole space vector diagram the five-level inverter.


 new zone( A  d )

sB  0

yes

no

Based on (35)-(38) and Figs.7 and 8, a simple mapping
algorithm for ZCMV PWM with minimized output current
ripple is proposed as shown in Fig. 9. The flow diagram in
Fig. 9(a) is utilized to determine the active regions of the
medium triangles in Fig. 7. Figure 9(b) specifically describes
the control algorithm to minimize the additional commutations
at the instant of transition to the new active region of
( A  d ). The design of the s1 and s2 sequence is based on
information of the current normalized state of the three-phase
active voltages (sA,sB,sC) and phase of double pulses in the old
region. The rule is set as ( B  s2 , C  s1 ) and
( B  s1 , C  s2 ) when the active voltages are (sB=1, sC=0)
and ( sB=0, sC=1), respectively; in this case, no additional
commutations are required. However, when the active
voltages are (sB=0, sC=0) or ( sB=1, sC=1), an additional
commutation is required regardless of the s1 and s2 sequence
design; in this case, the phase of double pulse in the previous
region is selected for non-commutation in order to distribute
the switching stress between the two phases. In cases with the
transitions to the new regions of B  d and C  d , s1,s2
sequences are designed in a similar way as for the case of
transition to the new region A  d .

The proposed ZCMV PWM scheme with reduced output
current ripple is obtained by utilizing mapping results in Fig. 9
as inputs to the block diagram of ZCMV PWM in Fig. 5. The
online algorithm is simple and can be implemented on a realtime microprocessor with small computational burden. The
maximum computations needed for implementation of the
proposed algorithm are 3 multiplications, 4 additions, 3
subtractions, 3 int() operations and 16 comparisons. The
measured execution time of the algorithm implemented on a
DSP28335 processor is less than 10 s .

sC  0

yes

old zone : B  d

sC  1
B  s2 , C  s1

no

yes
B  s1 , C  s2

no

yes

old zone : B  d


no

yes

no
B  s1 , C  s2

B  s2 , C  s1

b) s1 and s2 sequence design in case the transition to new zone
of ( A  d )
Fig. 9. Block diagram of the proposed mapping algorithm for reduced
RMS current ripple.

C. Normalized harmonic flux evaluation
In theoretical analysis, the normalized harmonic flux over
the full fundamental cycle (of A-phase, for example) is
evaluated as follows:

 An  rmsF 

2





 F ( )d
A


(39)

0

where,   2f ot , f o is the output fundamental frequency.
The function FA ( ) in (39) is calculated using (28). It can be
seen that FA ( ) is dependent on the mapping algorithm of each
PWM method. For the proposed ZCMV PWM method, the
designed mapping algorithm is determined as in Fig. 9.
The normalized harmonic flux in (39) is evaluated over the
entire range of the modulation index for the proposed PWM
method and two existing PWM methods that are based on the
same standardized patterns [17]. These two existing ZCMV
PWM methods include the Switching loss optimizing PWM
(SLO PWM) and the Voltage-Based Mapping PWM (VBM
PWM). It has been shown in [17] that characteristic of the
SLO PWM corresponding to   0 is identical to one obtained
by the VBM PWM. In this paper, comparative results of the
evaluated normalized harmonic flux with those PWM methods
would be shown to highlight the improved performance of the
proposed PWM in RMS current ripple reduction.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1


8

a) Three-level NPC inverter
a) Three-level NPC inverter

b) Five-level cascaded inverter
b) Five-level cascaded inverter
Fig. 10. Normalized RMS harmonic flux with the SLO PWM method of the
three-level NPC inverter and the five-level cascaded inverter.

For a given n-level inverter, the normalized harmonic flux
with the proposed PWM method is only dependent on the
modulation index m, while one with the SLO PWM is
dependent on both m and  .The characteristics of the
normalized RMS harmonic flux with SLO PWM method,
corresponding to the three-level inverter and the five-level
inverter, are illustrated in Fig. 10 (a) and Fig. 10(b),
respectively. The normalized RMS harmonic flux
characteristic of the proposed PWM and some slices of the
characteristic with the SLO PWM in Fig. 10(a) are shown in
Fig. 11(a), for the three-level inverter. A similar comparison is
shown in Fig. 11(b) for the five-level inverter. As observed
from each comparison, the normalized harmonic flux
characteristic of the SLO PWM corresponding to   0o is
identical to one obtained by the VBM PWM method.
It can be seen that the proposed PWM yields the optimum
curves in both comparisons. For the three-level inverter, the
normalized harmonic flux characteristic of the proposed PWM
is nearly identical to one obtained by the SLO PWM

at   90o in the region of (0 - 0.5) of the modulation index. In
the region of m higher than 0.5, the proposed PWM method
yields better performance in the RMS harmonic flux, as
compared to the SLO PWM pertaining to all selected slices.
For the five-level inverter, the normalized RMS harmonic flux
characteristic of the proposed PWM method is nearly identical
to the slices of   0o and   90o , respectively, in the regions
of (0.45 - 0.53) and (0 - 0.25, 0.63 -0.69) of the modulation
index, while is significantly lower than all selected slices in
other regions.

Fig. 11. The Normalized RMS harmonic flux characteristics with the
VBM PWM method (1), the SLO PWM method pertaining to different
phase displacements (2) and the proposed PWM method (3) of the threelevel NPC inverter and the five-level cascaded inverter.

D. Switching loss
In [20], the average value of the local (per carrier cycle)
switching loss over the fundamental (for instance, for phase
A) can be calculated:

Pswave 

1 Vdc (ton  toff )
2
2Ts

2

f


iA

( )d

(40)

0

where, t on and t o ff represent the turn-on and turn-off times of
the switching devices, respectively, and f iA ( ) is the
switching current function, the instantaneous value of which is
defined as a product of the number of commutations on the Aphase in a switching period and the absolute value of its
corresponding current i A ( ) .
2 if A  d
f iA ( )  k. i A ( ) ; k  
1 else

(41)

The switching loss function (SLF) is defined as:
SLF 

Pswave
P0

(42)

where, P0 is the maximum value of the switching loss
attainable for the defined load currents.
The Switching Loss Function (SLF) of the proposed PWM

is dependent on both the modulation index and the phase

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

9

displacement. The SLF surfaces of the proposed PWM method
for the three-level inverter and the five-level inverter are
shown in Fig. 12(a) and Fig. 12(b), respectively.
Comparisons of several slices of the SLF with the proposed
PWM and the two SLF characteristics pertaining to the VBM
PWM and the SLO PWM, are shown in Fig. 13(a) and Fig.
13(b) for the three-level inverter and the five-level inverter,
respectively. In both comparisons, the SLF corresponding to
the SLO PWM is a minimum constant value of 0.756 [17],
while the characteristic corresponding to the VBM PWM and
those pertaining to the proposed PWM with different
displacements are varied in the range (0.756 - 1) depending on
the modulation index.

IV. SIMULATION AND EXPERIMENTAL RESULTS
A. SIMULATION RESULTS
In order to confirm the analytical evaluation obtained for

the normalized RMS harmonic flux of the proposed PWM
method as well as of other existing PWM methods in [17],
numerical evaluation based on the simulated data is
performed. The output frequency is kept at 50 Hz while the
modulation index is increased from 0 with step size of 0.05.
The switching frequency is set to 2 kHz. The DC-link voltages
are set to Vdc =311 V for the three-level NPC inverter, and to
Vdc =155.5 V for the five-level cascaded inverter, so that the
maximum output line voltage of both topologies in the linear
range of the ZCMV PWM control is 380V RMS.
Figs. 14(a) and 14(b) present the normalized RMS harmonic
flux characteristics pertaining to the three ZCMV PWMs of
the three-level NPC inverter and the five-level cascaded
inverter, respectively. The comparisons show good agreement
between the numerical results and the analytical results
evaluated in Fig. 11.

a) Three-level NPC inverter

a) Three-level NPC inverter

b) Five-level cascaded inverter
Fig. 12. SLF characteristic of the proposed PWM method of the threelevel NPC inverter and the five-level cascaded inverter.

a) Three-level NPC inverter

b) Five-level cascaded inverter

Fig. 13. Slices of the SLF function of the proposed PWM method (3) versus
SLF characteristics of the VBM PWM method (1) and SLO PWM method (2)

of the three-level NPC inverter and the five-level cascaded inverter.

THDvL  142.33%
WTHDvL  1.18%

a) VBM PWM

b) Five-level cascaded inverter
Fig. 14. The normalized RMS harmonic flux characteristics pertaining to the
VBM PWM method (1), the SLO PWM method with different phase
displacements (2) and the proposed PWM method (3) of the three-level NPC
inverter and the five-level cascaded inverter.

THDvL  142.03%

THDvL  142.07%

WTHDvL  1.02%

WTHDvL  1.03%

b) SLO PWM (  82 o )

c) Proposed PWM

Fig. 15. Output line voltage spectra of the three-level NPC inverter with the three ZCMV PWM methods for m = 0.346 at 20Hz.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.



This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

THDvL

10

 49.92%

WTHDvL  1.02%

THDvL

 50.18%

THDvL

WTHDvL  1.03%

 50.03%

WTHDvL  0.97%

a) VBM PWM
b) SLO PWM (  82 o )
c) Proposed PWM
Fig. 16. Output line voltage spectra of the three-level NPC inverter with the three ZCMV PWM methods for m = 0.866 at 50Hz.


THDvL

 74.21%

WTHDvL  0.59%

THDvL

 74.11%

THDvL

 74.15%

WTHDvL  0.52%

WTHDvL  0.58%

a) VBM PWM
b) SLO PWM (  82 o )
c) Proposed PWM
Fig. 17. Output line voltage spectra of the five-level cascaded inverter with the three ZCMV PWM methods for m = 0.346 at 20Hz.

THDvL

 26.06%

WTHDvL  0.51%


THDvL

 26.12%

THDvL

 26.08%

WTHDvL  0.48%

WTHDvL  0.53%

a) VBM PWM
c) Proposed PWM
b) SLO PWM (  82 o )
Fig. 18. Output line voltage spectra of the five-level cascaded inverter with the three ZCMV PWM methods for m = 0.866 at 50Hz.

Fig. 15 and Fig. 16 show the line voltage spectra of the
three-level NPC inverter with the three ZCMV PWMs,
corresponding to cases of (m=0.346, fo= 20Hz) and (m=0.866,
fo= 50Hz). The designed phase displacements for the SLO
PWM control are set at 82 o and 84o , respectively, in cases of
fo = 20Hz and fo =50Hz, similar to the measured phase
displacements in the later experiment of V/f induction motor
control.
As seen from the line voltage harmonic spectra in Figs.
15(a)-15(c) of the three-level NPC inverter, the harmonic
component magnitudes in the sideband harmonics around 2
kHz with the proposed PWM are slightly lower than those
with the SLO PWM, while are significantly reduced as

compared to those with the VBM PWM. Figs. 16(a)-16(c)
shows reduced harmonic component magnitudes of around 2
kHz with the proposed PWM over both of the VBM PWM and
SLO PWM. Similar comparisons are performed for a fivelevel cascaded inverter as shown in Figs. 17-18. Similar to the
case of the three-level NPC inverter, the harmonic component
magnitudes around 2 kHz with the proposed PWM method are

substantially lower than those of the two other ZCMV PWM
methods in both comparisons.
The THD and WTHD values of the output line voltage are
defined, respectively, as in [25]:

THDVL 

WTHDVL 

1
VLfund

1
VLfund



V

2
Ln

(43)


n2



(
n2

VLn 2
)
n

(44)

where VLfu n d and VLn are, respectively, the RMS value of the
fundamental component and the RMS value of the n th
harmonic component of the output line voltage.
In order to analyze the THD of the current with no
dependence on the load parameters, the WTHD factor of the
output line voltage in (44) can be utilized. In Figs. 15-18,
THDs and WTHDs of the output line voltage of the three-level
NPC inverter and five-level cascaded inverter with the three

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics


TPEL-Reg-2015-04-0539.R1
ZCMV PWM methods are also provided. All THDs and
WTHDs are analyzed up to the frequency of 20 kHz. It is seen
that, while differences between the THDs pertaining to the
three ZCMV PWM methods are almost negligible in each
comparison, there has been considerable improvement in the
WTHD performance with the proposed PWM over the two
ZCMV PWM methods in [17].

11
V. EXPERIMENTAL VERIFICATION
The proposed ZCMV PWM method is validated on a 5
KVA five-level cascaded inverter topology fed 2 HP 380 V-50
Hz delta-connected induction motor. Each H-Bridge is made
up of IGBTs using FGL-60N100-BNTD. The DC voltage on
each H-Bridge is held constant at 155.5 V so that the
maximum output line voltage of the ZCMV PWM is 380V
(RMS). The rating of each DC-link capacitor used for the
experimental setup is 6800 µF. In the algorithm for switching
loss optimization, three additional Hall sensors LA55-P are
used to measure the output currents. All three PWM control
methods are implemented on the TMS320F28335 DSP
microcontroller. A three-phase induction motor with the
following parameters: stator resistance RS  8.68 , rotor
resistance Rr  8.3 , stator leakage inductance Lls  17.5mH ,
rotor leakage inductance Llr  17.5mH , and magnetizing

a) Three-level NPC inverter

inductance Lm  0.862H has been used. The switching

frequency is 2 kHz and the induction motor is operated with
constant V/f control. The V/f ratio is fixed so that the rated
line voltage of 380V (RMS) is corresponding to the rated
frequency of 50 Hz.
(V)

CMV

50
25
0
-25
-50

FFT of CMV

b) Five-level cascaded inverter

1.5
1
0.5
0

5

0

0

15


5

20

10

25

30

35

15

40

(kHz)

25
(ms)

20

a) m = 0.346, f o  20Hz

Fig. 19. The WTHD factor characteristics pertaining to the VBM PWM
method (1), the SLO PWM method with different phase displacements (2)
and the proposed PWM method (3) of the three-level NPC inverter and the
five-level cascaded inverter.

CMV

(V)
50
25
0
-25
-50

FFT of CMV

Fig. 19(a) presents the WTHD factor characteristics of the
output line voltage obtained from simulated data of the threelevel NPC inverter due to the VBM PWM method, the SLO
PWM method corresponding to different phase displacements,
and the proposed PWM method. The WTHD factor
characteristics are analyzed to the maximum modulation index
of 0.866, at an output frequency of 50 Hz. Similar
comparisons of the WTHD factor with the three PWM
methods of the five-level cascaded inverter are shown in Fig.
19(b). For better illustration, regions of the remarkable
improvements ((0.5-0.866) for the three-level NPC inverter
and (0.25-0.45) for the five-level cascaded inverter) are
magnified. In both comparisons, it is seen that the proposed
PWM leads to improved line voltage WTHD characteristics
over the VBM PWM and SLO PWM in almost the entire
region of the modulation index.

10

1.5

1
0.5
0

0

0

5

5

10

15

10

20

25

30

35

15

20


40

(kHz)

25
(ms)

b) m = 0.866, f o  50Hz
Fig. 20. Experimental CMV waveforms and CMV harmonic spectra with the
proposed ZCMV PWM method for two cases of a) m = 0.346,
f o  20Hz and b) m = 0.866, f o  50Hz

Fig. 20(a) presents the CMV waveform and its
corresponding harmonic spectrum with the proposed ZCMV
PWM method, at modulation index of 0.346 and output
frequency of 20 Hz. The same quantities are given in Fig.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

12

20(b) corresponding to the rated output frequency of the
induction motor. It is seen that the CMV with the proposed

ZCMV PWM is reduced considerably, compared to those with
conventional PWM methods [5],[6]. The spikes in CMV
waveforms, as observed from Fig. 20, are due to the effect of
double switching which has been referred in [23],[38]. These
high frequency spikes are still the cause of EMI and bearing
current issues as proved in [2-3],[23],[38]. There have been
some feasible approaches to reduce these CMV spikes that
may be taken into consideration [13], [15]. The techniques
include shifting dead-time of switching devices in different
phases [13], and incorporating hysteresis current controllers
into the control algorithm [15].
Fig. 21(a) shows the experimental waveform of the line
voltage with the proposed PWM method when the motor
operates at (m  0.346, f o  20Hz ) . The same quantity is
shown in Fig. 21(b) when the motor operates
at (m  0.866, f o  50Hz ) .

225
200

225
200

175

175

175

150


150

150

125

125

125

100
75
50

Voltage (V)

225
200

Voltage (V)

Voltage (V)

a) m = 0.346 at 20 Hz
b) m = 0.866 at 50 Hz
Fig. 21. Experimental waveforms of the line voltage with proposed PWM
method (X axis: 10 ms/div, Y axis: 200 V/div).

Comparative harmonic spectra of the line voltage with the

VBM PWM method, the SLO PWM method and the proposed
PWM method are shown in Fig. 22 and Fig. 23 for the cases of
(m  0.346, f o  20Hz ) and (m  0.866, f o  50Hz ) ,respectively. It can be observed that the experimental harmonic spectra
in Figs. 22-23 are in close agreement with the corresponding
spectra, obtained by simulation of the five-level cascaded
inverter, shown in Figs. 17-18.
Fig. 24(a) shows the experimental phase current with the
VBM PWM method for modulation index of 0.346 and output
frequency of 20 Hz. The instantaneous current ripple is
obtained by subtracting the fundamental current component
from the stator current. A low pass filter, which cut-off
frequency of 30 kHz, is applied in post-processing of the
current ripple waveform in order to remove the switching
noises. The current ripple waveform is then shown in Fig.
24(b). For comparison, the phase current and phase current
ripple with the SLO PWM method and those with the
proposed PWM method are shown in Figs. 25 and 26,
respectively.
The RMS current ripple values over a fundamental period
can be calculated based on the obtained current ripple
waveforms. Fig. 27 presents the RMS current ripple
characteristics of the measured stator current corresponding to
the VBM PWM, SLO PWM and the proposed PWM .The
motor is operated with constant V/f ratio and step frequency of
5 Hz. It can be seen that the proposed PWM method leads to a
great improvement of the RMS current ripple over the two
other PWM methods in a wide range of the output frequency.

100
75

50

100
75
50

25

25

25

0

0

0

0

1

2

3
4
5
6
Frequency (kHz)


7

8

0

9

1

2

3
4
5
6
Frequency (kHz)

7

8

0

9

1

2


3
4
5
6
Frequency (kHz)

7

8

9

675
600

675
600

525

525

525

450

450

450


375

375

375

300
225
150

Voltage (V)

675
600

Voltage (V)

Voltage (V)

a) VBM PWM
b) SLO PWM (  82 o )
c) Proposed PWM
Fig. 22. Experimental output line voltage harmonic spectra of the five-level cascaded inverter with the three ZCMV PWM methods for m = 0.346 at 20Hz.

300
225
150

300
225

150

75

75

75

0

0

0

0

1

2

3
4
5
Frequency (kHz)

6

7

8


9

0

1

2

3
4
5
Frequency (kHz)

6

7

8

9

0

1

2

3
4

5
Frequency (kHz)

6

7

8

a) VBM PWM
b) SLO PWM (  82 o )
c) Proposed PWM
Fig. 23. Experimental output line voltage harmonic spectra of the five-level cascaded inverter with the three ZCMV PWM methods for m = 0.866 at 50Hz.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.

9


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1

13

a) Phase current ( Y axis: 1A/div, X axis: 10ms/div)

Fig. 27. Experimental RMS current ripple characteristics corresponding to

the VBM PWM (1), the SLO PWM (2) and the proposed PWM method (3)
(the motor operated with constant V/f ratio).

b) Phase current ripple (RMS current ripple

 Irms = 0.127A)

Fig. 24. Experimental waveforms of phase current and phase current ripple
with the VBM PWM for m= 0.346 at 20 Hz.

a) Phase current (Y axis: 1A/div, X axis: 10ms/div)

b) Phase current ripple (RMS current ripple

In Section IV-B, the WTHD has been used as a loadindependent factor to present the THD factor of the current. In
the experiment of a constant V/f induction motor drive, the
THD factor can be obtained directly from the measured stator
current. Fig. 28 presents the experimental current THD
characteristics pertaining to the three ZCMV PWM methods.
It is observed that there is a close agreement between the
experimental THD factor characteristics and RMS current
ripple characteristics obtained in Fig. 27. The current THD
due to the proposed ZCMV PWM is substantially lower than
those due to the VBM PWM and SLO PWM over a wide
frequency range. The best reduction in the THD is from
12.61% with the SLO PWM method to 10.62% with the
proposed PWM method at output frequency of 30 Hz. The
values of the RMS current ripple and the current THD
pertaining to the three ZCMV PWM methods at some specific
frequencies are also given in Table II to highlight the superior

performance in both RMS current ripple and current THD of
the proposed PWM method.

 Irms = 0.126A)

Fig. 25. Experimental waveforms of phase current and phase current ripple
with the SLO PWM for m= 0.346 at 20 Hz.

Fig. 28. Experimental current THD characteristics corresponding to the VBM
PWM (1), the SLO PWM (2) and the proposed PWM method (3) (the motor
operated with constant V/f ratio).
a) Phase current (Y axis: 1A/div, X axis: 10ms/div)

TABLE II
EXPERIMENTAL CURRENT RIPPLE AND CURRENT THD
COMPARISONS BETWEEN THE VBM PWM, THE SLO PWM AND
THE PROPOSED PWM FOR A V/f CONTROLLED INDUCTION
MOTOR DRIVE FED BY A FIVE-LEVEL CASCADED INVERTER.

fo

b) Phase current ripple (RMS current ripple

 Irms = 0.11A)

Fig. 26. Experimental waveforms of phase current and phase current ripple
with the proposed PWM for m= 0.346 at 20 Hz.

RMS
current

ripple
[A]
THD
[%]

[Hz]

10

20

30

40

50

VBM PWM

0.123

0.127

0.088

0.125

0.108

SLO PWM


0.111

0.126

0.103

0.118

0.109

Proposed PWM

0.11

0.11

0.087

0.117

0.1

VBM PWM

16.19

15.79

10.82


15.11

13.16

SLO PWM

14.41

15.52

12.61

14.3

13.17

Proposed PWM

14.25

13.83

10.62

14.3

12.12

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See

for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1
The proposed PWM method is investigated under condition
of balanced DC voltage sources. It is applicable for cascaded
H-bridge inverters [14-17] or NPC inverters with modified
structures as in [6],[11]. For the cascaded inverters, owing to
the advantage of the modulator structure in which each DC
voltage is supplied by one rectifier through an isolated
transformer, variation of each DC-link voltage is negligible.
For the NPC topologies, for example, the three-level NPC
inverter, the utilization of two isolated DC voltage sources fed
by two diode rectifier front-ends [6] or of an additional
circuitry as in [11] would help solve the problem of neutral
point voltage fluctuations. For other multilevel NPC inverter
structures, if the DC voltage imbalance is problematical, the
ZCMV PWM to reduce current ripple should take into account
the DC voltage balancing as well. The investigation, however,
requires further study and analysis.

VI. CONCLUSION
This paper proposes a novel carrier-based ZCMV PWM
strategy for multi-level inverters using the principle of the
three zero common mode vectors. The resultant PWM patterns
made up of switching states of the three zero common mode
vectors have a minimum number of commutations. It has been

shown in this paper that the analysis of the output current
ripple of an n-level inverter under condition of ZCMV can be
simplified to that of a conventional two level inverter. Based
on the current ripple analysis, an online mapping algorithm is
further developed for the ZCMV PWM method so that a
minimum RMS current-ripple value can be obtained. The
proposed PWM method leads to substantial reduction in both
the RMS current ripple and the current THD over existing
PWM methods that are based on the same generalized
switching patterns. The effectiveness of the proposed PWM
method in CMV elimination as well as its superior
performances in the RMS current ripple and current THD have
been theoretically and experimentally validated.

REFERENCES
[1]

[2]

[3]

[4]

[5]

[6]

Satoshi Ogasawara, and Hirofumi Akagi, “Modeling and damping of
high-frequency leakage currents in PWM inverter-fed AC motor drive
systems”, Transactions on Industry Applications, Vol. 32, No. 5,

September/October 1996, pp.1105-1114.
J. M. Erdman, R. J. Kerkman, D.W. Schlegel, and G. L. Skibinski,
“Effect Of PWM inverters on AC motor bearing currents and shaft
voltages,” IEEE Trans. Ind. Applicat., vol. 32,pp.250-259, Mar./Apr.
1996.
Hirofumi Akagi, and Takafumi Doumoto, “An approach to eliminating
high-frequency shaft voltage and ground leakage current from an
inverter-driven motor”, IEEE Transactions on Industry Applications,
Vol. 40, No. 4, July/August 2004, pp.1162-1169.
Maria Carmela Di Piazza, Giovanni Tinè, and Gianpaolo Vitale, “An
Improved Active Common-Mode Voltage Compensation Device for
Induction Motor Drives, IEEE Trans. On Industrial Electronics, Vol. 55,
No. 4, April 2008, pp.1823-1834.
J.Rodriguez, J. S. Lai and F. Z. Peng, “Multilevel inverters: a survey of
topologies, controls and applications”, IEEE Trans. Industrial
Electronics, Vol. 49, No. 4, pp.724-738, August 2002.
Bin Wu, “High power converter and AC drives”, Wiley-interscience, A
John Wiley & Sons, Inc., Publication

14
[7]

[8]

[9]

[10]

[11]


[12]

[13]

[14]

[15]

[16]

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

[25]

[26]


H. Kim, H. Lee, and S. Sul, “A new PWM strategy for common mode
voltage reduction in neutral-point-clamped inverter-fed AC motor
drives,” IEEE Trans. Ind. Applicat., vol. 37, pp. 1840–1845, Nov./Dec.
2001.
Arnaud Videt, Philippe Le Moigne, Nadir Idir, Philippe Baudesson, and
Xavier Cimetière, “A New Carrier-Based PWM Providing CommonMode-Current Reduction and DC-Bus Balancing for Three-Level
Inverters”, IEEE Trans. On Industrial Electronics, Vol. 54, No. 6,
December 2007, pp. 3001-3011.
Mohan M. Renge and Hiralal M. Suryawanshi, “Three-Dimensional
Space Vector Modulation to Reduce Common-Mode Voltage for
Multilevel Inverter ‘, IEEE Trans. On Industrial Electronics, 2010.
A. K. Gupta and A. M. Khambadkone, “A space vector modulation
scheme to reduce common mode voltage for cascaded multilevel
inverters,” IEEE Trans. Power Electron., vol. 22, no. 5, pp. 1672–1681,
Sep. 2007.
A. V. Jouanne, S. Dai, and H. Zhang, “A multilevel inverter approach
providing DC-link balancing, ride-through enhancement, and Commonmode voltage elimination,” IEEE Trans. Ind. Electron., vol. 49, no. 4,
pp. 739–745, Aug. 2002.
H.Zhang, A.V Jouanne, S. Dai,A.K. Wallace, and Fei Wang, “Multilevel
Inverter Modulation Schemes to Eliminate Common-Mode Voltages”,
IEEE Transactions on Industry Applications, Vol. 36, No. 6,
November/December 2000, pp.1645-1653.
M. M. Renge and H. M. Suryawanshi, “Five-level Diode Clamped
Inverter to Eliminate Common Mode Voltage and reduce dv/dt in
Medium Voltage Rating Induction Motor Drives,” IEEE - Trans. on
Power Electronics, Vol. 23, No.4, July-2008, pp.1598-1607
P.C.Loh, D.G.Holmes, Y.Fukuta, and T. A. Lipo, “Reduced CommonMode Modulation Strategies for Cascaded Multilevel Inverters”, IEEE
Transactions
on
Industry

Applications,
Vol.
39,
No.5,
September/October 2003, pp.1386-1395
P. C. Loh, D. G. Holmes, Y. Fukuta, and T. A. Lipo, “A reduced
common mode hysteresis current regulation strategy for multilevel
inverters,” in Proc. IEEE APEC’03, 2003, pp. 576–582.
J. Rodríguez, J. Pontt, P. Correa, P. Cortés, and C. Silva, “A New
Modulation Method to Reduce Common-Mode Voltages in Multilevel
Inverters”, IEEE Trans. On Industrial Electronics, Vol. 51, No. 4, Aug.
2004, pp.834-839
Nho-Van Nguyen, Tam-Khanh Tu Nguyen and Lee Hong-Hee, “An
Optimized Switching Loss PWM Strategy to Eliminate Common Mode
Voltage in Multilevel Inverters”, IEEE Trans. Power Electron., vol. 30,
no. 10, pp. 5425–5438, Oct. 2015.
D. G. Holmes, “The significance of zero space vector placement for
carrier based PWM schemes,” IEEE Trans. Ind. Applicat., vol. 32,
pp.1122–1129, Sept./Oct. 1996.
V. R. Stefanovi´c and S. N. Vukosavi´c, “Space vector PWM voltage
control with optimized switching sequence,” in Proc. IEEE Ind. Appl.
Soc. (IAS) Annu. Meeting, Houston, TX, 1992, pp. 1025–1033.
A.M.Hava, R.Kerkman, and T.A. Lipo, “A high performance
generalized discontinuous PWM algorithm,” IEEE Trans. Ind. Appl.,
vol. 33, no. 5, pp. 1059–1071, Sep./Oct. 1998.
L.Dalessandro, S.D.Round,U.Drofenik, and J.W.Kolar, “Discontinuous
space-vector modulation for three-level PWM rectifiers,” IEEE Trans.
Power Electron., vol. 23, no. 2, pp. 530–542, Mar. 2008.
A. M. Hava, R. Kerkman, and T. A. Lipo, “Simple analytical and
graphical methods for carrier-based PWM-VSI drives,” IEEE Trans.

Power Electron., vol. 14, no. 1, pp. 49–61, Jan. 1999.
Hava, A.M.; Un, E., "Performance Analysis of Reduced Common-Mode
Voltage PWM Methods and Comparison With Standard PWM Methods
for Three-Phase Voltage-Source Inverters," Power Electronics, IEEE
Transactions on , vol.24, no.1, pp.241,252, Jan. 2009.
D. Casadei, G. Serra, A. Tani, and L. Zarri, “Theoretical and experiment
analysis for the RMS current ripple minimization in induction motor
drives controlled by SVM technique,” IEEE Trans. Ind. Electron., vol.
51 no. 5, pp. 1056–1065, Oct. 2004.
G. Narayanan and V. T. Ranganathan, “Analytical evaluation of
harmonic distortion in PWM AC drives using the notion of stator flux
ripple,” IEEE Trans. Power Electron., vol. 20, no. 2, pp. 466–474, Mar.
2005.
G. Narayanan, D. Zhao, H. K. Krishnamurthy, R. Ayyanar, and V. T.
Ranganathan, “Space vector based hybrid PWM techniques for reduced
current ripple,” IEEE Trans. Ind. Electron., vol. 55, no. 4, pp. 1614–
1627, Apr. 2008.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.


This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI
10.1109/TPEL.2015.2489560, IEEE Transactions on Power Electronics

TPEL-Reg-2015-04-0539.R1
[27] A. C. Binojkumar, J. Prasad, and G. Narayanan, “Experimental
investigation on the effect of advanced bus-clamping pulsewidth
modulation on motor acoustic noise,” IEEE Trans. Ind. Electron., vol.
60, no. 2, pp.433–439, Feb 2013.

[28] P. A. Dahono, Deni, and E.G. Supriatna, “Output-current-ripple analysis
of five-phase PWM inverters,” IEEE Trans. Ind. Appl., vol. 45, no. 6,
pp. 2022–2029, Nov./Dec. 2009.
[29] D. Dujic, M. Jones, and E. Levi, “Analysis of output-current ripple rms
in multiphase drives using space vector approach,” IEEE Trans. Power
Electron., vol. 24, no. 8, pp. 1926–1938, Aug. 2009.
[30] D. Dujic, M. Jones, E. Levi, J. Prieto, and F. Barrero, “Switching ripple
characteristics of space vector PWM schemes for five-phase two-level
voltage source inverters—Part 1: Flux harmonic distortion factors,”
IEEE Trans. Ind. Electron., vol. 58, no. 7, pp. 2789–2798, Jul. 2011.
[31] M. Jones, D. Dujic, E. Levi, J. Prieto, and F. Barrero, “Switching ripple
characteristics of space vector PWM schemes for five-phase two-level
voltage source inverters—Part 2: Current ripple,” IEEE Trans. Ind.
Electron., vol. 58, no. 7, pp. 2799–2808, Jul. 2011.
[32] D. Jiang, F. Wang, “Current-ripple prediction for three-phase PWM
converters,” IEEE Trans. Ind. Appl., vol. 50, no. 1, pp. 531–538,
Jan/Feb 2014.
[33] Grandi, G.; Loncarski, J. and Dordevic, O. “Analysis and Comparison of
Peak-to-Peak Current Ripple in Two-Level and Multilevel PWM
Inverters” Industrial Electronics, IEEE Transactions on, vol.62, no.5,
pp.2721-2730, May 2015.
[34] Das, S.; Narayanan, G. "Novel Switching Sequences for a Space-VectorModulated Three-Level Inverter", Industrial Electronics, IEEE
Transactions on, pp.1477 - 1487 Volume: 59, Issue: 3, March 2012.
[35] Das, S.; Narayanan, G.; Pandey, M. "Space-Vector-Based Hybrid Pulse
width Modulation Techniques for a Three-Level Inverter", Power
Electronics, IEEE Transactions on, pp.4580 - 4591 Volume: 29, Issue: 9,
Sept. 2014.
[36] K. R. Sekhar and S. Srinivas, “Discontinuous decoupled PWMs for
reduced current ripple in a dual two-level inverter fed open-end winding
induction motor drive,” IEEE Trans. Power Electron., vol. 28, no. 5, pp.

2493–2502, May 2013.
[37] S. Srinivas, K. Ramachandra Sekhar, “Theoretical and experimental
analysis for current in a dual-inverter-fed open-end winding induction
motor drive with reduced switching PWM,” IEEE Trans. Ind. Electron.,
vol. 60, no. 10, pp. 4318–4328, Oct. 2013.
[38] Kalaiselvi, J.; Srinivas, S., "Bearing Currents and Shaft Voltage
Reduction in Dual-Inverter-Fed Open-End Winding Induction Motor
With Reduced CMV PWM Methods," Industrial Electronics, IEEE
Transactions on , vol.62, no.1, pp.144,152, Jan. 2015.
[39] A. R. Beig, S. Kanukollu, K. Al Hosani, and A. Dekka, “Spacevectorbased synchronized three-level discontinuous PWM for mediumvoltage high-power VSI,” IEEE Trans. Ind. Electron., vol. 61, no. 8, pp.
3891–3901, Aug. 2014.

Tam-Khanh Tu Nguyen received his
B.S. and M.S. in Electrical
Engineering from Ho Chi Minh City
University of Technology, Viet Nam,
in 2010 and 2012, respectively. He is
currently a researcher at Power
Engineering
Research
Lab,
Department
of
Electrical
and
Electronics Engineering, Ho Chi Minh
City University of Technology,
Vietnam. His current research interests include active power
filters, PWM techniques for Matrix converter, Multi-level
inverters and advanced control of AC motor drives.


15
Nho-Van Nguyen was born in
Vietnam, in 1964. He received
the M.S. and PhD. degrees in
electrical engineering from
University of West Bohemia,
the Czech Republic in 1988
and 1991, respectively. Since
1992, he has been with
Department of Electrical and
Electronics Engineering, Ho
Chi Minh City University of
Technology, Vietnam, where he is currently an associate
professor. He was with KAIST as a post-doc fellow for six
months in 2001 and a visiting professor for a year in 20032004. His research interests include modeling and control of
switching power supplies, ac motor drives, active power
filters, and PWM techniques for power converters. He is a
member of the Institute of Electrical and Electronics
Engineers (IEEE).

Nadipuram (Ram) R. Prasad
is
a
tenured,
Associate
Professor in the Klipsch School
of Electrical and Computer
Engineering Department at
New Mexico State University.

He is also the Director of the
Rio Grande Institute for Soft
Computing (RioSoft) at New
Mexico State University and
RioRoboLab, a NASA Ames
funded Advanced Robotics
Laboratory. He is a Senior Research Fellow for SPAWAR
Systems Center of the Office of Naval Research, and is a
recipient of several NASA fellowships including the
prestigious NASA Administrator’s Fellowship in 2001-2002.
More recently, his contributions are in energy harvesting,
specifically in the area of low-head hydropower generation.
Dr. Prasad’s research interests are in neural networks, fuzzy
logic based systems, and evolutionary computation platforms.
Dr. Prasad has authored and co-authored over 150 research
publications in journals and conference proceedings and is the
co-author of three books on fuzzy and neural control.

0885-8993 (c) 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See
for more information.



×