1785613820
1785613820
1785613820
DC Distribution Systems
and Microgrids
Other volumes in this series:
Volume 1 Power Circuit Breaker Theory and Design C.H. Flurscheim (Editor)
Volume 4 Industrial Microwave Heating A.C. Metaxas and R.J. Meredith
Volume 7 Insulators for High Voltages J.S.T. Looms
Volume 8 Variable Frequency AC Motor Drive Systems D. Finney
Volume 10 SF6 Switchgear H.M. Ryan and G.R. Jones
Volume 11 Conduction and Induction Heating E.J. Davies
Volume 13 Statistical Techniques for High Voltage Engineering W. Hauschild and W. Mosch
Volume 14 Uninterruptible Power Supplies J. Platts and J.D. St Aubyn (Editors)
Volume 15 Digital Protection for Power Systems A.T. Johns and S.K. Salman
Volume 16 Electricity Economics and Planning T.W. Berrie
Volume 18 Vacuum Switchgear A. Greenwood
Volume 19 Electrical Safety: A guide to causes and prevention of hazards J. Maxwell Adams
Volume 21 Electricity Distribution Network Design, 2nd Edition E. Lakervi and E.J. Holmes
Volume 22 Artificial Intelligence Techniques in Power Systems K. Warwick, A.O. Ekwue and
R. Aggarwal (Editors)
Volume 24 Power System Commissioning and Maintenance Practice K. Harker
Volume 25 Engineers’ Handbook of Industrial Microwave Heating R.J. Meredith
Volume 26 Small Electric Motors H. Moczala et al.
Volume 27 AC–DC Power System Analysis J. Arrillaga and B.C. Smith
Volume 29 High Voltage Direct Current Transmission, 2nd Edition J. Arrillaga
Volume 30 Flexible AC Transmission Systems (FACTS) Y.-H. Song (Editor)
Volume 31 Embedded Generation N. Jenkins et al.
Volume 32 High Voltage Engineering and Testing, 2nd Edition H.M. Ryan (Editor)
Volume 33 Overvoltage Protection of Low-Voltage Systems, Revised Edition P. Hasse
Volume 36 Voltage Quality in Electrical Power Systems J. Schlabbach et al.
Volume 37 Electrical Steels for Rotating Machines P. Beckley
Volume 38 The Electric Car: Development and future of battery, hybrid and fuel-cell cars
M. Westbrook
Volume 39 Power Systems Electromagnetic Transients Simulation J. Arrillaga and N. Watson
Volume 40 Advances in High Voltage Engineering M. Haddad and D. Warne
Volume 41 Electrical Operation of Electrostatic Precipitators K. Parker
Volume 43 Thermal Power Plant Simulation and Control D. Flynn
Volume 44 Economic Evaluation of Projects in the Electricity Supply Industry H. Khatib
Volume 45 Propulsion Systems for Hybrid Vehicles J. Miller
Volume 46 Distribution Switchgear S. Stewart
Volume 47 Protection of Electricity Distribution Networks, 2nd Edition J. Gers and
E. Holmes
Volume 48 Wood Pole Overhead Lines B. Wareing
Volume 49 Electric Fuses, 3rd Edition A. Wright and G. Newbery
Volume 50 Wind Power Integration: Connection and system operational aspects B. Fox
et al.
Volume 51 Short Circuit Currents J. Schlabbach
Volume 52 Nuclear Power J. Wood
Volume 53 Condition Assessment of High Voltage Insulation in Power System Equipment
R.E. James and Q. Su
Volume 55 Local Energy: Distributed generation of heat and power J. Wood
Volume 56 Condition Monitoring of Rotating Electrical Machines P. Tavner, L. Ran,
J. Penman and H. Sedding
Volume 57 The Control Techniques Drives and Controls Handbook, 2nd Edition B. Drury
Volume 58 Lightning Protection V. Cooray (Editor)
Volume 59 Ultracapacitor Applications J.M. Miller
Volume 62 Lightning Electromagnetics V. Cooray
Volume 63 Energy Storage for Power Systems, 2nd Edition A. Ter-Gazarian
Volume 65 Protection of Electricity Distribution Networks, 3rd Edition J. Gers
Volume 66 High Voltage Engineering Testing, 3rd Edition H. Ryan (Editor)
Volume 67 Multicore Simulation of Power System Transients F.M. Uriate
Volume 68 Distribution System Analysis and Automation J. Gers
Volume 69 The Lightening Flash, 2nd Edition V. Cooray (Editor)
Volume 70 Economic Evaluation of Projects in the Electricity Supply Industry, 3rd Edition
H. Khatib
Volume 72 Control Circuits in Power Electronics: Practical issues in design and
implementation M. Castilla (Editor)
Volume 73 Wide Area Monitoring, Protection and Control Systems: The enabler for
smarter grids A. Vaccaro and A. Zobaa (Editors)
Volume 74 Power Electronic Converters and Systems: Frontiers and applications
A.M. Trzynadlowski (Editor)
Volume 75 Power Distribution Automation B. Das (Editor)
Volume 76 Power System Stability: Modelling, analysis and control B. Om P. Malik
Volume 78 Numerical Analysis of Power System Transients and Dynamics A. Ametani
(Editor)
Volume 79 Vehicle-to-Grid: Linking electric vehicles to the smart grid J. Lu and J. Hossain
(Editors)
Volume 81 Cyber-Physical-Social Systems and Constructs in Electric Power Engineering
S. Suryanarayanan, R. Roche and T.M. Hansen (Editors)
Volume 82 Periodic Control of Power Electronic Converters F. Blaabjerg, K. Zhou, D. Wang
and Y. Yang
Volume 86 Advances in Power System Modelling, Control and Stability Analysis F. Milano
(Editor)
Volume 87 Cogeneration: Technologies, optimisation and implementation
C.A. Frangopoulos (Editor)
Volume 88 Smarter Energy: From smart metering to the smart grid H. Sun, N. Hatziargyriou,
H.V. Poor, L. Carpanini and M.A. Sánchez Fornié (Editors)
Volume 89 Hydrogen Production, Separation and Purification for Energy A. Basile,
F. Dalena, J. Tong and T.N. Veziroğlu (Editors)
Volume 90 Clean Energy Microgrids S. Obara and J. Morel (Editors)
Volume 91 Fuzzy Logic Control in Energy Systems with Design Applications in MATLAB‡/
Simulink‡ İ.H. Altaş
Volume 92 Power Quality in Future Electrical Power Systems A.F. Zobaa and S.H.E.A. Aleem
(Editors)
Volume 93 Cogeneration and District Energy Systems: Modelling, analysis and
optimization M.A. Rosen and S. Koohi-Fayegh
Volume 94 Introduction to the Smart Grid: Concepts, technologies and evolution
S.K. Salman
Volume 95 Communication, Control and Security Challenges for the Smart Grid
S.M. Muyeen and S. Rahman (Editors)
Volume 97 Synchronized Phasor Measurements for Smart Grids M.J.B. Reddy and
D.K. Mohanta (Editors)
Volume 98 Large Scale Grid Integration of Renewable Energy Sources A. Moreno-Munoz
(Editor)
Volume 100 Modeling and Dynamic Behaviour of Hydropower Plants N. Kishor and
J. Fraile-Ardanuy (Editors)
Volume 101 Methane and Hydrogen for Energy Storage R. Carriveau and D.S.-K. Ting
Volume 104 Power Transformer Condition Monitoring and Diagnosis A. Abu-Siada (Editor)
Volume 108 Fault Diagnosis of Induction Motors J. Faiz, V. Ghorbanian and G. Joksimović
Volume 110 High Voltage Power Network Construction K. Harker
Volume 111 Energy Storage at Different Voltage Levels Technology, Integration, and
Market Aspects Zobaa, Ribeiro, Aleem and Afifi (Editors)
Volume 112 Wireless Power Transfer: Theory, technology and application N. Shinohara
Volume 119 Thermal Power Plant Control and Instrumentation: The control of boilers and
HRSGs, 2nd Edition D. Lindsley, J. Grist and D. Parker
Volume 123 Power Systems Electromagnetic Transients Simulation, 2nd Edition N. Watson
and J. Arrillaga
Volume 124 Power Market Transformation B. Murray
Volume 130 Wind and Solar Based Energy Systems for Communities R. Carriveau and
D.S.-K. Ting (Editors)
Volume 131 Metaheuristic Optimization in Power Engineering J. Radosavljević
Volume 905 Power System Protection, 4 volumes
This page intentionally left blank
DC Distribution Systems
and Microgrids
Edited by
Tomislav Dragičević, Pat Wheeler
and Frede Blaabjerg
This publication is copyright under the Berne Convention and the Universal Copyright
Convention. All rights reserved. Apart from any fair dealing for the purposes of research
or private study, or criticism or review, as permitted under the Copyright, Designs and
Patents Act 1988, this publication may be reproduced, stored or transmitted, in any
form or by any means, only with the prior permission in writing of the publishers, or in
the case of reprographic reproduction in accordance with the terms of licences issued
by the Copyright Licensing Agency. Enquiries concerning reproduction outside those
terms should be sent to the publisher at the undermentioned address:
While the authors and publisher believe that the information and guidance given in this
work are correct, all parties must rely upon their own skill and judgement when making
use of them. Neither the authors nor publisher assumes any liability to anyone for any
loss or damage caused by any error or omission in the work, whether such an error or
omission is the result of negligence or any other cause. Any and all such liability is
disclaimed.
The moral rights of the authors to be identified as authors of this work have been
asserted by them in accordance with the Copyright, Designs and Patents Act 1988.
Preface xvii
Index 437
This page intentionally left blank
Preface
that are different and more complex compared to the most conventional single bus
structure were intentionally selected for a detailed overview here, as they can offer
additional degrees of flexibility and redundancy. Moreover, due to the rapid
development of power electronics technology that has resulted in increased relia-
bility of power electronic devices, additional hardware required to establish such
architectures is not considered as a drawback any more. In particular, two archi-
tectures overviewed are relatively generic (i.e., the solid-state transformers in
Chapter 9, and bipolar-type DC microgrids in Chapter 10), while the architecture
detailed in Chapter 8 for electrical vehicle charging infrastructure is more tailored
toward that specific application.
The final part, comprised of seven chapters, presents an analysis of a number of
applications in which the control methods and architectures discussed in the first two
parts can be applied in the specific context of a given application. The selected
applications for this book include aircraft and shipboard systems, electrical vehicles,
data centers, residential buildings, and photovoltaic powered systems, respectively.
The book ends with an overview of some of the DC microgrid demonstration sites
around the world.
The Editors would like to thank all the contributors for their excellent work
and cooperation in the preparation of this book. Furthermore, we would like to
acknowledge the excellent and timely assistance of the editorial and production
staff at the Institution of Engineering and Technology (IET).
The Editors
Tomislav Dragičević, Aalborg University, Denmark
Patrick Wheeler, University of Nottingham, UK and China
Frede Blaabjerg, Aalborg University, Denmark
Chapter 1
DC microgrid control principles – hierarchical
control diagram
Linglin Chen1, Tao Yang1, Fei Gao1, Serhiy Bozhko1, and
Patrick Wheeler1
1.1 Introduction
Different electrical components can be conveniently connected to a DC grid as
there are no issues of reactive power, power factor, different frequencies, etc. as
in AC grids. Control is one of the key disciplines in achieving the objective of
smart grid initiatives. The control of DC microgrids (MGs) covers a wide range of
control objectives and responsive dynamics. Therefore, to better understand the
concept and implement the MG control, division of the MG control into relatively
independent sublevels is essential. The utility AC power system has a history of
over a hundred year of development. Due to its unique physical characteristic, MGs
have more feasibility of controlling individual distributed generation (DG) sub-
systems due to their power electronics interfaces. In order to fully exploit the
potential of power electronic converters in the MG, a hierarchical control for AC
MGs was developed in literature [1].
In the hierarchical structure, MG controls are classified into three groups
according to their different functionalities. Generally, in the primary control level,
controllers are responsible for the active and reactive power sharing among
DGs. In the secondary control level, frequency variation and AC voltage ampli-
tude offset are compensated. In the tertiary control level, power management
algorithm is conducted. The recent development of solid-state circuit breakers for
the DC network has made the DC MG an alternative option to the AC MGs.
Compared with AC networks, the DC network has no reactive power issues and
less power cables, which can potentially reduce weight and cost of the power
grid. A similar hierarchical control structure can also be implemented in DC MGs
as shown in Figure 1.1. The details of each control level will be explained in the
following section.
1
Department of Electrical and Electronics Engineering, The University of Nottingham, United Kingdom
2 DC distribution systems and microgrids
MG central control
Tertiary
Bandwidth
Secondary increases
Primary
Inner loops
Microgrid
In the past few years, different concepts for primary, secondary and tertiary control
have been developed for DC MGs applications [1–4]. The explanation of each level
can be given as follows:
● The primary control is implemented strictly local within the DG subsystems. In
a DC MG, installation of more than one power source is common. Therefore,
how to manage the DGs to share the load proportionally to their rated power is
an issue to consider. Conventionally, droop control [5] is a preferred option in
this control level as it provides autonomous control features to the DGs in the
presence of communication failure. It effectively deals with the load sharing
among DGs connected on the same DC bus and provides references to sub-
control loops (voltage or current loops). Alternatively, the droop control can
also be perceived as virtual resistance, connected in series with converter on
the DC bus side.
● The secondary control, predominately, sets in an MG central controller which
is usually, together with tertiary control level, remotely located to DGs, thus
communications are required. It generates voltage shifting signals to primary
control level in aim to eliminate the DC voltage deviation caused by primary
droop controls. Moreover, it is also possible to implement the secondary con-
trol level locally inside DGs.
DC microgrid control principles – hierarchical control diagram 3
The functionality of this control level is to deal with the power sharing among
different DG units. Generally, the load sharing can be achieved in both centralized
and decentralized ways. In the centralized way [6], usually the DG with the largest
power capacity is chosen to be the dominant power source, and it is responsible for
the regulation of the DC bus voltage (voltage support mode). Other DGs in the DC
MG work in power mode (current mode). They receive power (current) reference
signals from a central controller. In this case, a reliable operation of power sharing
largely depends on the effective communication. As the control bandwidth of the
primary control is the highest among the three control levels (in terms of kHz),
communication within this level between DGs also increases infrastructure
investment. Thus, for large-scale MGs with remotely located DGs, communication
links are not practical in the primary control level, and only local current and
voltage information is available. Thus, as a fully decentralized approach, the droop
control has shown favourable features in large-scale MGs for load sharing. With the
droop control, all measurements take place locally within DGs. The droop control
in DC MGs can be intuitively appreciated as ‘virtual resistance’ [2]. This allows
paralleled DGs to share DC bus currents proportionally in line with their respective
power rating. Due to its modularity and reliability, the droop control has been
widely used in DC MGs in the primary control level [7–9].
is only dependent on frequency ( f), and the reactive power (Q) is solely related to
voltage (V). Therefore, by adjusting the frequency f and the terminal V, the active
power and reactive power from each DG unit can be regulated. These results in the
famous droop regulation in an AC grid as
f f0 ¼ kp ðP P0 Þ (1.1)
V V0 ¼ kq ðQ Q0 Þ (1.2)
with measured terminal voltage V and frequency f, the required P and Q can be
derived from (1.1) and (1.2). The derived P and Q are then sent to the inverter inner
controllers (control level 0), and the DG unit can operate in desired conditions.
A similar concept can also be used to parallel DC converters for power sharing.
Within DC MGs, there is no reactive power, and the active power is only controlled
by terminal DC voltage. Hence, a linear droop characteristic line between V and I
can be implemented. As shown in Figure 1.2, the DC-link voltage vDC and the DC
current iDC is in a linear relation. With the DC current increasing, the DC-link
voltage will be reduced. The relation between DC-link voltage vDC and the DC
current iDC can be expressed as
vDC ¼ V0 RD iDC (1.3)
where iDC is the power converter output DC current; V0 is the voltage at no load
(iDC ¼ 0 A); vDC is the DG terminal voltage; the slope RD is the droop gain and can
also be viewed as a virtual resistor, which will be explained later.
To explain the power sharing using DC V–I droop control concept, a simple
DC MG system is shown in Figure 1.3. The droop curves for the two DGs are
illustrated as in Figure 1.4. The control structure details are depicted in Figure 1.5.
DC currents and voltages of DG1 and DG2 are measured locally and fed into their
own primary controllers. Based on the measured current iDC1 and iDC2, terminal DC
voltage references can be calculated accordingly with the V–I droop curves.
Eventually at steady state, the DC voltage equals to VL. Because the droop gains
RD1 and RD2 are designed based on the rated power of the DG, this will be dis-
cussed later. As shown in Figure 1.4, DG1 and DG2 spend different ‘efforts’ to
share the load current, with iDC1 and iDC2, respectively.
V0
RD
vDC
iDC i
DG 2 DG 1
idc2 idc1
+
VL
RL
vdc
V0 RD1
RD2
vL
idc
idc2 idc1
vdc
vdc Voltage
*
Current PWM
V0 DC source
– loop loop generator
idc
RD
Primary control
Figure 1.5 Practical primary control implementation with V–I droop method [2]
Usually, the inner voltage and current loops have fast dynamic response. The
output impedance of the converter is predominately dependant on the droop gain.
Therefore, the V–I droop gain is termed also as virtual resistance RD [5]. Figure 1.6
illustrates a simplified DC system diagram, where Vo1 and Vo2 are DC voltages at
no load from V–I droop controllers for DG1 and DG2, respectively; RD1 and RD2
are virtual resistance for DG1 and DG2, respectively. If there is no virtual resis-
tance, a small difference between Vo1 and Vo2 will result in a large circulating
6 DC distribution systems and microgrids
DG1 DG2
+ + +
Vo1 VL Vo2
– RL –
–
vdc
εv vo
imax idc
imin
current between DG1 and DG2. This is because without virtual resistance RD1 and
RD2, the circulating current between DG1 and DG2 is only subject to cable resis-
tance which is essentially very small (neglected in Figure 1.6).
In general, two aspects should be considered when designing the value of
the virtual impedance. One is the maximum allowed voltage deviation (ev in
Figure 1.7). A larger virtual impedance will make the V–I output characteristic
‘softer’ (as the slope is larger). Another aspect to consider is the system stability, as
the droop gain alters the output impedance. A detailed analysis concerning stability
analysis in DC MGs can be found in [5]. Thus, in this chapter, only constraint of the
maximum allowed DC bus voltage deviation is considered. Assuming that ev is the
maximum allowed voltage deviation, the allowed maximum RD is given as
ev
RD ¼ (1.4)
2imax
where imax is the maximum output current. The relation of ev and imax is illustrated
in Figure 1.7.
voltage source). Therefore, voltage deviation will be expected when different loads
are applied to the DC system [7]. Another issue for using the conventional V–I
droop is that some power sharing error would occur during operations. Two main
reasons for this error are discussed below:
1. Error in DC voltage control. In practical situations, due to the measurement
error from voltage and current sensors, the DC bus voltage at each DG unit
might be different even if the references are set the same. If we consider no
load conditions, this control error gives different V0 for each DG unit (Vo1 and
Vo2), as shown in Figure 1.8. The voltage error within DGs will potentially
cause large power sharing discrepancies in MG. As can be seen from
Figure 1.8, if two DGs are with the same power rating and virtual impedances
(i.e. RD), the current error between two DGs is smaller (I1I2) with a larger
droop gain. On the other hand, although a smaller droop gain produces
larger current sharing errors (I1oI2o), it means a stiffer voltage source and a
smaller voltage deviation due to load changes. This results in a dilemma in
designing the droop gain for primary control.
2. Cable resistance: In large-scale DC MGs, DG terminal voltages might be a
variant with load power in presence of cable resistance. A simple DC MG
which takes accounts of cable resistance is shown in Figure 1.9. The cable
resistance between the converter DC terminal and the common DC bus is
considered as Rline1 and Rline2.
idc
I1 I2 I1o I2o
DG1 DG2
RD1 Rline1 Rline2 RD2 idc2
idc1
+ +
+ + +
vdc1 vL vdc2 vo2
vo1 RL
– – –
– –
The DG1 node voltage vDC1 can be derived as in the following equation:
vDC1 ¼ V0 RD1 iDC1 (1.5)
where vDC1 is the voltage reference produced by V–I droop control in MG1; iDC1 is
the output current from DG1; RD1 is the virtual output resistance (droop gain) for
DG1; and V0 is the output voltage at no load. If the line impedance of Rline1 and
Rline2 is considered, (1.5) can be written as
vL ¼ V01 RD1 iDC1 Rline1 iDC1 (1.6)
vL ¼ V02 RD2 iDC2 Rline2 iDC2 (1.7)
Combining (1.6) and (1.7), and assuming Vo1 equals to Vo2, yields
1 1
iDC1 : iDC2 ¼ : (1.8)
RD1 þ Rline1 RD2 þ Rline2
From (1.8), we can conclude that the current sharing between DGs is not only
dependent on the droop gain but also the line impedance. In large-scale DC systems
where the line impedance is not negligible, the current sharing error cannot be
avoided with conventional droop control.
V–P droop
V–I droop idc
idc characteristic
characteristic +
vdc
vdc + P *
vdc
* vdc
idc vdc P
idc vdc
–
–
{
(a) (b)
{ {
(c) (d)
Figure 1.10 Droop characteristic employed in VSCs: (a) V–I droop, (b) V–P
droop, (c) I–V droop and (d) P–V droop
{
(a) (b)
Figure 1.11 I–V droop and its corresponding current-mode control scheme:
(a) I–V droop characteristic and (b) current-mode control scheme
zero in the transfer function. For the voltage-mode droop-controlled system, the DC
voltage dynamics are affected by the DC voltage control bandwidth and the droop
gain. Increasing the droop gain will reduce the voltage loop bandwidth. In contrast
to the current-mode approach, the voltage-mode droop control regulates the term-
inal voltage based on current measurements. The RHP zero causes high gain
instability within the vDC controller, i.e. the system will easily go unstable when the
vDC controller has a high gain value. Furthermore, when utilizing voltage-mode
droop control, the voltage loop bandwidth is mainly determined by the voltage
controller rather than the droop gain. Greater details on stability analysis can be
found in literature [5].
The discussed, droop control techniques so far have assumed linear relations
between the DC voltage and DC current. As opposed to linear droop regulations
mentioned above, a nonlinear droop curve is proposed in [12]. The droop gain can
10 DC distribution systems and microgrids
V
P*
Figure 1.12 Nonlinear droop control scheme considering available headroom for
each DGs
be adapted according to the available headroom available for each individual DGs.
The droop gain is defined as
P jP j l
K ¼ K0 (1.9)
P
where K0 is the rated droop gain, K is the adapted coefficient value, P* is the rated
power and P is the real-time output power. Essentially, the droop gain K gets larger
in heavy load conditions, as shown in Figure 1.12. Therefore, the power sharing
error is reduced in presence of DC voltage control discrepancy.
Figure 1.13 Frequency coordinating virtual impedance with I–V droop [15]
Hybrid ESS
Grid
Lead-acid
AC
DC
DC
DC
Load SC
AC DC
DC DC
FC
Gas
DC H2
AC DC O2
DC
PV
DC
DC
into a current reference i*. Consequently, the high-pass filter alters the virtual
impedance, and the equivalent output impedance becomes
ZðsÞ
Zdroop ðsÞ ¼ (1.10)
Rd
As an example, an MG with both lead acid and super-capacitor energy storages is
shown in Figure 1.14. The lead-acid battery has slower dynamics than the super-
capacitor. Thus, the slower load dynamic spectrum is loaded to battery converter,
while the faster dynamic load is assigned to super-capacitor converter as depicted
in Figure 1.15. A low-pass filter is inserted in series with the droop admittance in
battery control loop to shape the equivalent virtual impedance as illustrated in
Figure 1.16. This allows the output impedance of the battery converter increases
with frequency. On the contrary, a high-pass filter is implemented in the control
loop of super capacitor, and the output impedance of the super capacitor converter
shows low impedance in the medium frequency range. Dynamic range greater than
control bandwidth is mainly managed by passive DC bus filtering capacitors.
12 DC distribution systems and microgrids
idc
Zbat(s)=ωc/(s+ωc)
Vo P* i* Current PWM DC
+ 1/Rd ÷
– loop generator source
ω
vdc
(a)
idc
Zsc(s)=s/(s+ωc)
Vo P* i* Current PWM DC
+ 1/Rd ÷
– loop generator source
ω
vdc
(b)
Figure 1.15 Frequency coordination for hybrid ESS: (a) control loop for battery
and (b) control loop for super-capacitor [15]
Z (dB)
80
Zbat
Zsup
60 Zsup // Zbat
40
20
0
0.1 1 100 1k f (Hz)
Battery Super-cap DCcaps
Z(s) idc +
Vo – P* i*
+ 1/Rd ÷ T(s) Y vdc
ω –
P– P+
P+
P–
Vo Rd
The secondary control aims to restore the DC bus voltage deviation caused by
primary droop control. When MG is connected to an upper grid, the secondary
controller receives command from upper tertiary controllers. This allows voltages
at common coupling points to track the upper grid voltage. The secondary control
can be implemented locally inside DGs. It can also be integrated within the tertiary
control in an MG central controller, as shown in Figure 1.1. When the secondary
control is implemented locally inside DGs, this level of control will be decen-
tralized and is referred to as distributed control. On the other hand, if the secondary
control is integrated into the upper level controller (controller in the tertiary level),
it will be centralized (thus referred as centralized approach).
δVo1
vdc1
*
vdc PWM
vdc Voltage Current
Vo DC source
– loop loop generator
VMG* –
GV
RD1
DC bus
Secondary Control idc1
δVo2
vdc2
vdc* Voltage Current PWM
Vo DC source
– loop loop generator
RD2
idc2
Primary control
Vdc
δVo
V *MG
VMG_p
idc
A droop shifting value dvo is generated from the secondary controller and sent to
each DGs through communication links. To adjust power sharing dynamically,
different shifting values can be sent to different DGs. When embedding the droop
shifting into the DGs, the V–I droop (1.1) becomes
vDC ¼ v0 þ dv0 RD iDC (1.11)
Another figure showing the curve shifting is shown in Figure 1.20. As can be
seen, a droop shifting value dV0 from the secondary controller is used to adjust the
no-load voltage V0 in the primary control level. After this compensation, the droop
curve is lifted up, and DC voltage is increased from VMG_P to VMG .
As can be seen from Figures 1.19 and 1.20, a reliable operation of secondary
control depends on effective communication between the two levels and a reliable
MG central controller. Any fault that occurs in the communication channel or the
DC microgrid control principles – hierarchical control diagram 15
Communication
Secondary
ACS ACS ACS control
DC microgrid
Load-1 Load-2
VNn N
R1 R2 Rn
+ + +
V1 V2 ... Vn
– – –
Figure 1.22 Equivalent circuit for the average current signal bus
R1
Rd idc
*
– vdc Voltage Current PWM
Vo DC source
loop loop generator
Average current signal bus
dVo
∑i vdc
K
n
DC bus
R2
Rd idc
*
– vdc Voltage Current PWM
Vo DC source
loop loop generator
dVo
∑i vdc
K
n
Figure 1.23 ACS control for parallel DC–DC converters using analogue
communication
n vdc
∑i
j =1
pu
j
irated K
j
DC bus
n
pu
idcm
1/imrated
Rd idcm
*
– vdc Voltage Current PWM
Vo DC source
DACS dV loop
o
loop generator
n
∑i
j =1
pu
j
i mrated K
vdc
n
Figure 1.24 ACS-based distributed control using CAN bus as the communication
channel
Figure 1.22 will be created. The voltage on the average current signal bus is, hence,
reflecting the average value of DC currents within different DC/DC converters.
As analysed in the primary control section, with conventional droop control, the
voltage variation is subjected to load changes (current changes). If the load
increases, according to the droop curve, the DC voltage will decrease. Thus, the
average current measured from average current signal bus will also increase.
Therefore, the droop shifting value is increased, and the droop curve is shifted up,
as shown in Figure 1.20. The voltage deviation caused by primary droop control is
compensated.
Although ACS method is an effective approach to restore DC bus voltage, this
method is only effective when DGs are configured close to each other. When DGs
are widely spread with long distance in between, since the average current signal
bus has to be deployed along with power lines, this makes average current bus
susceptible to noise. This will degrade the system performance when using ACS
method for the secondary control.
To overcome the above-mentioned shortcomings of ACS, ACS-based dis-
tributed method using controller area network (CAN bus) has been proposed, as
shown Figure 1.24 [7]. The nature of CAN is to provide reliable digital commu-
nications between controllers and devices without a host. This makes CAN bus
rather suitable for the secondary control application. Since all the current infor-
mation shared on the CAN bus are normalized values, this makes the current
sharing algorithm simpler and MG easier to expand. To distinct from analogue-
based ACS, the CAN bus-based implementation is termed as digital ACS (DACS).
Normalized averaged current is calculated in DACS and transformed into a droop
shifting signal by multiplying its current normal base irated
j and shifting gain kj.
18 DC distribution systems and microgrids
idcj Receiving
Secondary BPF
control Broadcasting
–
iref_b
dVo Gr(s)
+ iref_PLS idcj
DC bus
*
vdc PWM DC source
Vo Gv(s) Gi(s)
generator
– vdc –
Rd
DGj
DGm
Afterwards, the droop shifting signal Dv0j (the same purpose as dvo in Figure 1.20)
is sent to the primary controller to restore the bus voltage.
+
VDC VMG
DC microgrid
–
Stiff DC grid sampling
Secondary dVo
Operation control
Tertiary control level
Figure 1.26 Tertiary control to manage power exchange between MG and upper
grid with synchronization loop of a DC MG [2]
VH Mode 1
VM Mode 2
VL Mode 3
MPPT
P P P
Pg–
– +
Pbat Pbat Pg+
Renewables Battery Utility grid
current (or power) control loop in tertiary control level and the current reference is
set by the command iG .
Once the MG is connected to the upper grid, it communicates with the dis-
tribution system operator (DSO) or transmission system operator (TSO) [1] and the
secondary control level. System power management can be implemented based on
DC Bus Signalling (DBS). The concept of DBS technique has been introduced in
literature [15,16]. In this method, the DC bus voltage itself is used as a commu-
nication signal between DGs. Agreements have been made beforehand among
DGs. DGs will take different operation actions under different DC bus voltage
ranges. Sources in the DC MGs can be classified into three main categories:
renewables, storages and utility grid as shown in Figure 1.27. When the bus voltage
is higher than VH (Mode 1), DC bus voltage is regulated by renewables and battery/
utility grid receives power generated from renewables. When bus voltage falls
below VH but stays above VM (Mode 2), DC bus voltage is supported by battery.
In Mode 2, renewables are working in Maximum Power Point Tracking and power
20 DC distribution systems and microgrids
Mode 1
Mode 2
Mode 3
MPPT
P P P
–
Pbat +
Pbat Pg– Pg+
Renewables Depleted battery Utility grid
is feeding into the utility grid. When the bus voltage drops below VM (Mode 3), grid
converters adjust the DC bus voltage. Both renewables and battery work as current
sources. Due to the existence of the secondary control level, the final steady state
bus voltage is always restored to the reference. All the curves shown in Figure 1.27
can be shifted up or down as an entity by the secondary controller.
Each droop curve in Figure 1.27 can be modified in real time online by tertiary
control according to system requirements. For instance, when battery has low State
of Charge (SOC), Pþ bat is forced to zero and the droop curve for utility grid is lifted
up accordingly, as shown in Figure 1.28.
Based on the above mode structure, DSO/TSO gives command to tertiary
control level to determine the power flow between MG and the upper grid by
changing Pþ
g and Pg . Based on information gathered from both DSO/TSO and the
MG, the tertiary controller prepares the source and storage dispatch schedule to
optimize the operation which is communicated to the secondary and primary levels.
From this point, tertiary control level shares similarity for both AC and DC MGs.
Therefore, the existing solutions on energy management for AC MGs can be
adapted for DC MG applications. Greater details concerning energy management
will be discussed in other chapters.
1.6 Summary
load transition is also discussed in this section. Different load dynamic spectrums
are assigned to different DGs according to their dynamic response. Finally, to
receive commands from upper control levels, interfaces are pointed out without
changing the basic primary control loop structure.
The purpose of the secondary control is to restore the DC bus voltage deviation
caused by conventional droop control from primary control level. It can be
implemented remotely in an MG central controller (Centralized approach) or
locally inside each DG (distributed approach). Both dedicated analogue and digital
communication links can be used in a decentralized secondary control to transmit
current signals. To enhance the reliability of the control, a communication link
based on power lines is also possible.
The tertiary control level is responsible for the connecting process of MG to
the upper grid. A basic mode structure based on DBS is introduced to accommodate
energy-management algorithms.
References
[1] T. L. Vandoorn, J. C. Vasquez, J. De Kooning, J. M. Guerrero, and
L. Vandevelde, ‘‘Microgrids: Hierarchical Control and an Overview of the
Control and Reserve Management Strategies,’’ IEEE Ind. Electron. Mag.,
vol. 7, no. 4, pp. 42–55, Dec. 2013.
[2] J. M. Guerrero, J. C. Vasquez, and R. Teodorescu, ‘‘Hierarchical control of
droop-controlled DC and AC microgrids—a general approach towards
standardization,’’ in 2009 35th Annual Conference of IEEE Industrial
Electronics, 2009, pp. 4305–4310.
[3] C. N. Papadimitriou, E. I. Zountouridou, and N. D. Hatziargyriou, ‘‘Review
of Hierarchical Control in DC Microgrids,’’ Electr. Power Syst. Res.,
vol. 122, pp. 159–167, May 2015.
[4] C. Jin, P. Wang, J. Xiao, Y. Tang, and F. H. Choo, ‘‘Implementation of
Hierarchical Control in DC Microgrids,’’ IEEE Trans. Ind. Electron.,
vol. 61, no. 8, pp. 4032–4042, Aug. 2014.
[5] F. Gao, S. Bozhko, A. Costabeber, et al., ‘‘Comparative Stability Analysis of
Droop Control Approaches in Voltage-Source-Converter-Based DC Micro-
grids,’’ IEEE Trans. Power Electron., vol. 32, no. 3, pp. 2395–2415, Mar. 2017.
[6] D. Xu, H. Li, Y. Zhu, K. Shi, and C. Hu, ‘‘High-surety Microgrid: Super
Uninterruptable Power Supply with Multiple Renewable Energy Sources,’’
Electr. Power Compon. Syst., vol. 43, no. 8–10, pp. 839–853, Jun. 2015.
[7] S. Anand, B. G. Fernandes, and J. Guerrero, ‘‘Distributed Control to Ensure
Proportional Load Sharing and Improve Voltage Regulation in Low-Voltage
DC Microgrids,’’ IEEE Trans. Power Electron., vol. 28, no. 4, pp.
1900–1913, Apr. 2013.
[8] Y. A.-R. I. Mohamed, ‘‘Robust Droop and DC-Bus Voltage Control for
Effective Stabilization and Power Sharing in VSC Multiterminal DC Grids,’’
IEEE Trans. Power Electron., vol. 33, no. 5, pp. 4373–4395, May 2018.
22 DC distribution systems and microgrids
[9] P. Karlsson and J. Svensson, ‘‘DC Bus Voltage Control for a Distributed
Power System,’’ IEEE Trans. Power Electron., vol. 18, no. 6, pp. 1405–1412,
Nov. 2003.
[10] J. Jiang, ‘‘Design of an Optimal Robust Governor for Hydraulic
Turbine Generating Units,’’ IEEE Trans. Energy Convers., vol. 10, no. 1,
pp. 188–194, Mar. 1995.
[11] K. De Brabandere, B. Bolsens, J. Van den Keybus, A. Woyte, J. Driesen,
and R. Belmans, ‘‘A Voltage and Frequency Droop Control Method
for Parallel Inverters,’’ IEEE Trans. Power Electron., vol. 22, no. 4,
pp. 1107–1115, Jul. 2007.
[12] N. R. Chaudhuri and B. Chaudhuri, ‘‘Adaptive Droop Control for Effective
Power Sharing in Multi-Terminal DC (MTDC) Grids,’’ IEEE Trans. Power
Syst., vol. 28, no. 1, pp. 21–29, Feb. 2013.
[13] J. Beerten and R. Belmans, ‘‘Analysis of Power Sharing and Voltage
Deviations in Droop-Controlled DC Grids,’’ IEEE Trans. Power Syst.,
vol. 28, no. 4, pp. 4588–4597, Nov. 2013.
[14] F. Gao, Y. Gu, S. Bozhko, G. Asher, and P. Wheeler, ‘‘Analysis of droop
control methods in DC microgrid,’’ in 2014 16th European Conference on
Power Electronics and Applications, 2014, pp. 1–9.
[15] Y. Gu, W. Li, and X. He, ‘‘Frequency-Coordinating Virtual Impedance for
Autonomous Power Management of DC Microgrid,’’ IEEE Trans. Power
Electron., vol. 30, no. 4, pp. 2328–2337, Apr. 2015.
[16] Y. Gu, X. Xiang, W. Li, and X. He, ‘‘Mode-Adaptive Decentralized Control
for Renewable DC Microgrid with Enhanced Reliability and Flexibility,’’
IEEE Trans. Power Electron., vol. 29, no. 9, pp. 5072–5080, Sep. 2014.
[17] K. Sun, L. Zhang, Y. Xing, and J. M. Guerrero, ‘‘A Distributed Control
Strategy Based on DC Bus Signaling for Modular Photovoltaic Generation
Systems With Battery Energy Storage,’’ IEEE Trans. Power Electron., vol.
26, no. 10, pp. 3032–3045, Oct. 2011.
[18] L. Zhang, T. Wu, Y. Xing, K. Sun, and J. M. Gurrero, ‘‘Power control of DC
microgrid using DC bus signaling,’’ in 2011 Twenty-Sixth Annual IEEE
Applied Power Electronics Conference and Exposition (APEC), 2011,
pp. 1926–1932.
[19] T. Dragicevic, J. M. Guerrero, and J. C. Vasquez, ‘‘A Distributed Control
Strategy for Coordination of an Autonomous LVDC Microgrid Based
on Power-Line Signaling,’’ IEEE Trans. Ind. Electron., vol. 61, no. 7,
pp. 3313–3326, Jul. 2014.
Chapter 2
Distributed and decentralized control of
dc microgrids
Saeed Peyghami1, Hossein Mokhtari2, and
Frede Blaabjerg1
2.1 Introduction
Power sharing control of dc power sources in dc microgrids is an important issue in
order to obtain a stable and reliable operation [1–7]. Power sharing approaches in
three levels of the hierarchal control system are described in the last chapter. Two
major objectives of this method of control are proportional power sharing and
appropriate voltage regulation in dc microgrids which can be carried out using
centralized and decentralized/distributed approaches [2,3,6,7]. Centralized
approaches deal with lower reliability and resiliency. In order to enhance the resi-
liency of the power sharing methods, distributed and decentralized controls have
been introduced. Distributed approaches employ sparse communication systems
instead of point-to-point communication control among different converters (and
maybe a central control unit). On the contrary, decentralized approaches use no
communication (or even physical communication link) in the corresponding control
system.
Decentralized approaches can be categorized as
1. Mode adaptive (autonomous) droop control
2. Nonlinear droop control
3. Frequency droop control
The voltage droop method is the simplest decentralized approach employed in dc
microgrids. This method can perfectly control the load sharing among power
sources in the case of short distances where the line resistances can be neglected. In
this case, mode adaptive or autonomous droop approach can be used to control the
output power of different (dispatchable and nondispatchable) sources in the dc
microgrid. Considering the line resistances of long distances among converters, the
conventional droop method cannot properly carry out the power sharing objectives,
1
Department of Energy Technology, Aalborg University, Denmark
2
Department of Electrical Engineering, Sharif University of Technology, Iran
24 DC distribution systems and microgrids
and hence, nonlinear droop approach and frequency droop methods have been
introduced in order to reach the power management objectives.
Furthermore, some distributed approaches have been presented which can be
classified as
1. Fully communicated control
2. Sparse communicated (consensus-based) control
3. Sparse communicated control using current information.
In the distributed approaches, each converter is equipped with a secondary con-
troller cascaded by the primary controller. A communication system is employed to
share the voltage and current information between all of the converters. In a simple
distributed secondary control, it is required that all of the converters have access to
the information each other. This approach may be complex and has a lower resi-
liency; however, consensus strategy with a sparse communication network has been
presented and can use the neighboring converter information to reach an
acceptable operating condition. Furthermore, another sparse communicated control
approach can reach the power management objectives by only employing the cur-
rent information of the converters. In this chapter, decentralized and distributed
control concepts are discussed.
R1
Interlinking
Dispatchable unit converter
AC grid
R4
R2
R3
Rated power
Rated power
Rated power
Rated power
Rated power
MPPT range
SoC range
Pinjected 0 Pabsorbed Pcharge 0 Pdischarge 0 Prated 0 Prated
Figure 2.2 Droop characteristics for different energy units in dc microgrid with
constant droop slopes; (a) IC, (b) storage converter, (c) dispatchable
units and (d) nondispatchable units
SoC 100%
Rated power
Rated power
Rated power
Rated power
Rated power
Rated power
SoC 0%
MPPT range
SoC range
Pinjected 0 Pabsorbed Pcharge 0 Pdischarge 0 Prated 0 Prated
(a) (b) (c) (d)
Figure 2.4 Droop characteristics for different energy units with adjustable
droop slopes; (a) IC, (b) storage converter, (c) dispatchable units
and (d) nondispatchable units
Distributed and decentralized control of dc microgrids 27
Interlinking
R1 converter 1
R4
Dispatchable unit
Interlinking AC grid
R2 converter 2
R5
Nondispatchable unit
R3
IC 1 Voltage IC 2 Voltage
VH Operating point VH
VL VL
Figure 2.5 Typical ac–dc microgrid with two ICs; (a) the single line diagram
of the power system and (b), (c) the droop characteristics of the
interlinking converters (ICs)
28 DC distribution systems and microgrids
Rated power
Mode II
Rated power
VL
Rated power
Rated power
Rated power
MPPT range
SoC range
Converter 1 X1 R1 PCC R2 X2
Converter 2
Io1 Io2
Cdc Cdc
+ V Vo1 Line 1 Line 2 Vo2 Vi2 +
i1
– –
PWM PWM
Vo1 Vo2
Voltage Inner Inner Voltage,
current controller controller current
Io1 Io2
Vref,1 Vref,2
V* Droop Droop
+– Rd,1 Rd,2 +– V*
controller controller
have at least one voltage forming converter in the microgrid. For example, an
autonomous control system based on the droop is shown in Figure 2.6 [10]. As can
be seen in Figure 2.6, in Mode I, the voltage is controlled by the photovoltaic array.
In Mode II, a battery is responsible for forming the voltage, and in Mode III, the IC
controls the dc-link voltage.
40
%
20
0
0 0.5 1 1.5 2 2.5 3
Droop gain (Rd,1 = Rd,2)
Figure 2.8 Performance of the conventional droop control for the simplified dc
microgrid, current sharing error and voltage regulation
Secondary control approaches are explained in the first chapter. The secondary
controller requires communication infrastructures to share the voltage and current
information of converters to regulate the dc-link voltage. In order to avoid utilizing
a communication system, an adaptive nonlinear approach has been used, presented
in [4]. In this approach, the reference voltage Vref,i of the ith converter can be
defined as (2.1), where mi and a are the droop curve constant and coefficients,
respectively [4]. mi can be determined by (2.2), where Vmin is the minimum
allowable voltage and Imax are the maximum current of ith converter.
The nonlinear droop curve is graphically shown in Figure 2.9. As can be seen, the
slope of the droop curve slope is larger in the higher currents, which reflects
accurate current sharing in heavy-loading condition. Furthermore, the dc-link
voltage remains in the acceptable interval. Considering the effective droop curves
(the tangent line of the nonlinear curve) shown in Figure 2.9, the droop parameters
including the slope and voltage references can automatically be adapted by varying
the load current. Actually, with the conventional droop approach, the droop curve
needs to be shifted by a secondary controller to regulate the dc voltage at an
acceptable interval. However, by utilizing the nonlinear droop method, the effec-
tive voltage reference can be determined by the droop curve [4], where the effective
voltage shifting can be calculated as follows:
ða 1ÞðV Vmin Þ a
DVi ¼ a Ii (2.3)
Imax;i
The current sharing error and voltage regulation performance of the conventional
and nonlinear droop methods are illustrated in Figure 2.10. As is shown, employing
30 DC distribution systems and microgrids
∆V2
V* ∆V1
Rd1
Vmin
I
I1 I2 I3
30 Nonlinear approach
20
10
0
2 4 6 8 10 12 14 16 18
(a) Load current (A)
40
Current sharing error (%)
20
10
0
2 4 6 8 10 12 14 16 18
(b) Load current (A)
small droop gains causes acceptable voltage regulation but leads to high current
sharing error. Although selecting higher droop gains causes appropriate current
sharing, it leads to a large voltage drop in the higher load currents. However,
employing the nonlinear curve introduces an acceptable current sharing in a heavy-
load condition and appropriates voltage regulation in different load currents as can
be seen from Figure 2.10.
R2+jX2 Converter N
Converter 1 ViN
Vi1 IoN V
lo1 oN
SN ILN +
+ IL1 S1 P2 –
– Vo1
R1+jX1
R3+jX3
P1
P3
R4+jX4 R6+jX6
P4
Vi2 Converter 2 Converter k Vik
Io2 R5+jX5 P5 lok Vok
+ IL2 S2 Sk ILK +
– Vo2 RN+jXN –
PN
(a)
Power sharing controller for kth unit Inner voltage and
dk current regulators Sk
iok fk
dfk Eq. 1 2π ∫ Sine A
~
Vk PWM
Vo*
Vok Power Qk
dp G(s) –+ +
+– Gv(s) +– Gi(s)
iok calc.
Vok ILk
(b) (c)
superimposed onto the dc voltage, where the frequency of the ac voltage is pro-
portional to the output dc current of the converter. The rated frequency should be
selected to be smaller than the bandwidth of the inner voltage controller, and hence,
to be regulated by a proportional–integrator (PI)-based voltage regulator. There-
fore, the inner voltage (Gv(s)) and current (Gi(s)) controllers in Figure 2.11(c) can
modulate the reference voltage including dc voltage and superimposed ac voltage.
From the ac voltage point of view, the converters are working like a synchronous
generator; hence, they can be coordinated together with the common frequency
[2,11,12]. From the power system dynamics and control theory, for analyzing the
dynamic behavior of a Synchronous Generator (SG) in an ac power system, it can
be modeled as two SGs; one being the specified SG and the other modeling the
entire power system. Moreover, the two SGs can be simplified as a single-machine-
infinite-bus, where the infinite bus is considered as a stiff ac source [13]. Therefore,
since the proposed approach is based on the SG principles, without losing the
generality, a simplified dc MG, with two converters connected to a load at a point
of common coupling (PCC), is considered, and the block diagram of the system
with the corresponding signals is shown in Figure 2.11.
According to Figure 2.12, if the output dc voltage of the converters (Vo1, Vo2)
is settled at a reference value (Vo* ), their output dc current (Io1, Io2) will be inversely
proportional to the corresponding line resistances (i.e., Io1/Io2 ¼ R2/R1), where R1
and R2 denote the line resistance of the first and second converter, respectively.
Adjusting the output dc voltage of the converters is the only option for controlling
the corresponding output currents at a desired value, for example, proportional to
their rated current, which requires the coordination of the converters. For the
coordination between converters, a small ac voltage, i.e., ~v k ¼ A sin ð2pfk t), is
superimposed onto the dc voltage reference and modulated by each converter.
~
Converter 1 io1 = lo1+io1 X1 R1 Load
+ io1 Line 1
Vi1 S1 vo1 VPCC
–
~
vo1 = Vo1 + vo1 A A
~
vo2 = Vo2 + vo2 d Vo2
Vo1
+ S2 io2 Line 2
Vi2 vo2
–
Figure 2.12 Conceptual illustration of the proposed strategy showing the injected
ac voltages and corresponding currents in a simplified dc MG based
on two dc–dc converters
Distributed and decentralized control of dc microgrids 33
the line impedances. Thereby, the ac power contains the information of the line
impedances. On the other hand, in LV systems with low X/R ratio, the reactive
power can accurately be controlled by the frequency [14]. Therefore, employing the
injected reactive power (Q) of the converters to adjust the dc voltage reference (Vo* )
causes proper current sharing. Applying the proposed control algorithm, the output
dc voltage of the converters can be written as
Vo1 ¼ Vo dp Q1 GðsÞ; Vo2 ¼ Vo dp Q2 GðsÞ (2.10)
where dp is the coupling gain between dc voltage and reactive power, and G(s) ¼
wc/(sþwc) is a low pass filter to eliminate the high-frequency component of the
calculated reactive power. Therefore, the frequency droop can be used to coordinate
the converters, and the small ac power can be employed to adjust the dc voltage and,
consequently, the dc currents. Each converter can be controlled by the local mea-
sured values, and hence, like SGs, there is no need for any communication network
[2,11,12]. Furthermore, the injected ac voltage by the converters must be synchro-
nized with the ac component of the grid voltage at the startup time. The phase of the
connection bus voltage can be extracted using a phase locked loop (PLL) block.
On the other hand, in ac systems, synchronization methods are employed to make
the converter voltage close to the grid voltage in order to limit the inrush current at
the start time, which may damage the converter switches for the high currents.
However, the injected ac voltage and consequently the ac currents are very small in
the proposed approach, and hence, the converters can be connected without utilizing
a PLL. Hence, they can be synchronized based on the droop control functionality as
the grid supporting voltage source converters in ac grids [15].
controllers
V = [V1,...,VN, Vavg V1 1
dV + V1
Inner
VL1,...,VLM]
T Averaging –+ PI +
+ +– R1
I = [I1,...,IN]T Averaging +– PI
Rd1 I1
Current regulator
VL
RN+M
V* M
V* Converter N
controllers
V = [V1,...,VN,
Vavg dV VN
Inner
+ VN
–+ +–+ RN
T
VL1,...,VLM] Averaging PI +
T
I = [I1,...,IN] Averaging +– PI IN
RdN
Current regulator
Vi
ith converter estimator
i
Vavg
+
j + aij 1/s +
–
Vavg
Neighbor average
estimator
controllers
Vavg Dynamic Vavg + dV V1
V1
Inner
+ R1
From Nth N consensus – PI + ++
–
Conv. Ipu
+– PI I1
Rd1
To 2nd Current regulator
Conv.
VN V*
N–1 N Converter N
controllers
Vavg Dynamic Vavg dV VN
Inner
+ VN
From (N–1)th N–1 consensus –+ PI + +–+ RN
Conv. Ipu
+– PI IN
RdN
To 1st Current regulator
Conv.
estimator shown in Figure 2.14, where the coefficient aij is the weight of infor-
mation exchanged between converters i and j. After some iterations, the estimated
average voltage of the converters converges to a value which is equal to the average
voltage of the converters (Vavg). As shown in Figure 2.15, this average voltage is
regulated by the secondary controller and settles at the reference value.
On the other hand, the per-unit current of the neighboring converter can be
used to compensate the mismatch of the currents of all converters. This approach is
analogous to the circular-chain-control (CCCs) in parallel inverters in ac microgrids,
which is used to improve the current mismatches among the inverters [14,21–24].
In this approach, each converter shares the average voltage and per-unit current
with its neighboring converter. Therefore, sparse communication among the con-
verters with low volume of transmitted information improves the reliability and
stability of the system [3]. However, in consensus algorithms, the average voltage
of the grid supporting buses is only regulated. In practice, the loads may not be
connected to the grid supporting converters and might be distributed over the
microgrid. Therefore, regulating the voltage of the grid supporting buses may not
guarantee the voltage regulation at load buses.
V ref
dVi E = Vavg + dVV
controllers
++ Gk
++ – Gv(S) – +–
Inner
N Iavg dVi Rdk
I
1
N Σ a
j
+– Gi(S)
Ik
Current data
j=1 j
1/ak
Current regulator
Secondary control Primary control
l1, l2,...,lN Ik
Figure 2.16 Proposed control approach; voltage regulator Gv, current regulator
Gi, and low-bandwidth communication link with delay function of
Gd ðsÞ ¼ ð1=ð1 þ tsÞÞ [25]
where Vmax and Vmin are the maximum and minimum allowable voltage range,
respectively, and Ink is the rated current of the kth converter. Due to the differences
in line resistances, the output current of converters cannot be proportionally dis-
patched among the converters. The current regulator calculates the weighted
average currents of converters and regulates the corresponding output current
proportional to the rated current of the converter. The average current (Iavg) can be
calculated as [25]
1X N
Ij
Iavg ¼ j ¼ 1 : N; (2.12)
N j¼1 aj
where N is the number of converters, Ij is the measured current and aj is the sharing
coefficient of jth converter, respectively. Furthermore, as with the CCCs approach,
the converters can only employ the neighboring converters information to reach
proportional load sharing.
A simplified single line model of a dc microgrid with two converters is shown
in Figure 2.17. In steady state, droop gain acts as a series resistor (Rd1 and Rd2). The
secondary current regulator behaves as a small positive/negative resistor (rd1 and
rd2) such that the total resistance of each line becomes proportional to the corre-
sponding rated current. Hence, the relationship between rated current (Inj) and
sharing coefficient (aj) and total line resistance between jth converter and PCC can
be written as
Ini aj Rdj þ rdj þ rj
¼ ¼ ; i; j ¼ 1 : N ; i 6¼ j (2.13)
Inj ai Rdi þ rdi þ ri
where rj is the resistance of the line connected to jth converter.
38 DC distribution systems and microgrids
Converter 1 δvi,1
Rd1 I1 r1 Load
rd1 Line 1
Vr E1 V1 VPCC R
Converter 2 δvi,2
Rd2 I2 r2
rd2 Line 2
Vr E2 V2
VPCC + r1I1
Vdc* +δvi
Rd I VPCC + r2I2
–δvi
Rd
VPCC
I
I1= I2
where Vr is the reference value of the voltage loop. If sharing coefficients are
assumed to be one, then the droop gains must be equal (Rd1 ¼ Rd2 ¼ Rd) and I1 ¼ I2
at steady state. The output values of the current regulators can be obtained as
8
>
> I1 þ I2 I2 I1
< dvi1 ¼
>
2
I1 Gd Gi ðsÞ ¼
2
Gd Gi ðsÞ
(2.15)
>
> dvi2 ¼ I1 þ I2 I2 Gd Gi ðsÞ ¼ I1 I2 Gd Gi ðsÞ
>
:
2 2
where Gi(s) is the PI controller and Gd(s) is the delay of communication link. At
steady state, I1 ¼ I2, and hence dvi1 þ dvi2 ¼ 0. Therefore, the average voltage of the
microgrid is
1
V avg ¼ ðV1 þ V2 Þ ¼ Vr Rd I (2.16)
2
Applying the primary droop controller and secondary current regulator causes an
average voltage drop equal to RdI. From the single line model of the microgrid
shown in Figure 2.17, internal voltages (i.e., E1 and E2) are equal to the average
voltage calculated by (2.16). Therefore, the distributed voltage regulator can esti-
mate the internal voltage and regulate it at the reference value. In fact, the cor-
rection term (dvv) shifts up the droop characteristics in Figure 2.18 to restore the
average voltage of the microgrid which can be calculated as (2.17), where V* is the
rated voltage of the microgrid.
dvv1 ¼ ðV ðV1 þ dvi1 ÞÞGv ðsÞ
(2.17)
dvv2 ¼ ðV ðV2 þ dvi2 ÞÞGv ðsÞ
The relationship between converter voltage and PCC voltage can be shown as
V1 VPCC ¼ r1 I1
(2.18)
V2 VPCC ¼ r2 I2
where
VPCC ¼ RL ðI1 þ I2 Þ (2.19)
The set point value for the primary controller is
V1 ¼ Vref 1 Rd I1
(2.20)
V2 ¼ Vref 2 Rd I2
and the set point value for the inner voltage loop can be determined by the primary
controller as
V1 ¼ Vref 1 Rd I1
(2.21)
V2 ¼ Vref 2 Rd I2
40 DC distribution systems and microgrids
This equation shows that the term of RdI which is related to the primary controller
can be eliminated in low frequencies, i.e., in the secondary controller frequency
bandwidth. Therefore, the primary controller shares the load current between
converters based on droop gain, and the secondary controller reduces the mismatch
in current sharing as well as decreasing the voltage drop of the droop gain. This
approach uses the only current information to appropriately control the load sharing
and voltage regulation in the dc microgrid.
References
[1] Peyghami, S., Mokhtari, H., Davari, P., Loh, P.C., Blaabjerg, F.: ‘On Sec-
ondary Control Approaches for Voltage Regulation in DC Microgrids’ IEEE
Trans. Ind. Appl., 2017, 53, (5), pp. 4855–4862.
[2] Peyghami, S., Davari, P., Mokhtari, H., Loh, P.C., Blaabjerg, F.: ‘Syn-
chronverter-Enabled DC Power Sharing Approach for LVDC Microgrids’
IEEE Trans. Power Electron., 2017, 32, (10), pp. 8089–8099.
[3] Nasirian, V., Davoudi, A., Lewis, F.L., Guerrero, J.M.: ‘Distributed Adap-
tive Droop Control for dc Distribution Systems’ IEEE Trans. Energy Con-
vers., 2014, 29, (4), pp. 944–956.
[4] Khorsandi, A., Ashourloo, M., Mokhtari, H., Iravani, R.: ‘Automatic Droop
Control for a Low Voltage DC Microgrid’ IET Gener. Transm. Distrib.,
2016, 10, (1), pp. 41–47.
[5] Khorsandi, A., Ashourloo, M., Mokhtari, H.: ‘A Decentralized Control
Method for a Low-Voltage DC Microgrid’ IEEE Trans. Energy Convers.,
2014, 29, (4), pp. 793–801.
[6] Guerrero, J.M., Vasquez, J.C., Matas, J., De Vicuña, L.G., Castilla, M.:
‘Hierarchical Control of Droop-Controlled AC and DC Microgrids—
A General Approach Toward Standardization’ IEEE Trans. Ind. Electron.,
2011, 58, (1), pp. 158–172.
[7] Peyghami, S., Mokhtari, H., Blaabjerg, F.: ‘Hierarchical Power Sharing
Control in DC Microgrids’, in Magdi S Mahmoud (Ed.): ‘Microgrid:
advanced control methods and renewable energy system integration’
(Oxford: Elsevier Science & Technology, 2017, first), pp. 63–100.
[8] Boroyevich, D., Cvetković, I., Dong, D., Burgos, R., Wang, F., Lee, F.:
‘Future electronic power distribution systems—A contemplative view’ Proc.
Int. Conf. Optim. Electr. Electron. Equipment, OPTIM, 2010, pp. 1369–1380.
[9] Loh, P.C., Li, D., Chai, Y.K., Blaabjerg, F.: ‘Autonomous Operation of ac–
dc Microgrids with Minimised Interlinking Energy Flow’ IET Power Elec-
tron., 2013, 6, (8), pp. 1650–1657.
[10] Gu, Y., Xiang, X., Li, W., He, X.: ‘Mode-Adaptive Decentralized Control
for Renewable DC Microgrid with Enhanced Reliability and Flexibility’
IEEE Trans. Power Electron., 2014, 29, (9), pp. 5072–5080.
[11] Peyghami, S., Mokhtari, H., Loh, P.C., Davari, P., Blaabjerg, F.: ‘Distributed
Primary and Secondary Power Sharing in a Droop-Controlled LVDC
Microgrid with Merged AC and DC Characteristics’ IEEE Trans. Smart
Grid, 2018, 9, (3), pp. 2284–2294.
[12] Peyghami, S., Mokhtari, H., Blaabjerg, F.: ‘Decentralized Load Sharing in a
Low-Voltage Direct Current Microgrid with an Adaptive Droop Approach
Based on a Superimposed Frequency’ IEEE J. Emerg. Sel. Top. Power
Electron., 2017, 5, (3), pp. 1205–1215.
[13] Kundur, P., Balu, N., Lauby, M.: ‘Power system stability and control’
(New York: McGraw-Hill, 1994).
42 DC distribution systems and microgrids
[14] Guerrero, J.M., Hang, L., Uceda, J.: ‘Control of Distributed Uninterruptible
Power Supply Systems’ IEEE Trans. Ind. Electron., 2008, 55, (8), pp. 2845–
2859.
[15] Rocabert, J., Luna, A., Blaabjerg, F., Rodriguez, P.: ‘Control of Power
Converters in AC Microgrids’ IEEE Trans. Power Electron., 2012, 27, (11),
pp. 4734–4749.
[16] Lu, X., Guerrero, J.M., Sun, K., Vasquez, J.C.: ‘An Improved Droop Control
Method for DC Microgrids Based on Low Bandwidth Communication with
DC Bus Voltage Restoration and Enhanced Current Sharing Accuracy’ IEEE
Trans. Power Electron., 2014, 29, (4), pp. 1800–1812.
[17] Meng, L., Dragicevic, T., Vasquez, J.C., Guerrero, J.M.: ‘Tertiary and Sec-
ondary Control Levels for Efficiency Optimization and System Damping
in Droop Controlled DC–DC Converters’ IEEE Trans. Smart Grid, 2015,
6, (6), pp. 2615–2626.
[18] Meng, L., Dragicevic, T., Guerrero, J.M., Vasquez, J.C.: ‘Dynamic con-
sensus algorithm based distributed global efficiency optimization of a droop
controlled DC microgrid’ ENERGYCON 2014—IEEE Int. Energy Conf.,
2014, pp. 1276–1283.
[19] Shafiee, Q., Dragicevic, T., Vasquez, J.C., Guerrero, J.M.: ‘Hierarchical
Control for Multiple DC-Microgrids Clusters’ IEEE Trans. Energy Convers.,
2014, 29, (4), pp. 922–933.
[20] Moayedi, S., Davoudi, A.: ‘Distributed Tertiary Control of DC Microgrid
Clusters’ IEEE Trans. Power Electron., 2015, 31, (2), pp. 1717–1733.
[21] Wu, T.-F., Chen, Y.-K., Huang, Y.-H.: ‘3C Strategy for Inverters in Parallel
Operation Achieving an Equal Current Distribution’ IEEE Trans. Ind.
Electron., 2000, 47, (2), pp. 273–281.
[22] Ding, G., Gao, F., Zhang, S., Loh, P.C., Blaabjerg, F.: ‘Control of Hybrid
AC/DC Microgrid under Islanding Operational Conditions’ J. Mod. Power
Syst. Clean Energy, 2014, 2, (3), pp. 223–232.
[23] Hatziargyriou, N., Asano, H., Iravani, R., Marnay, C.: ‘Microgrids’ IEEE
Power Energy Mag., 2007, 5, (4), pp. 78–94.
[24] Lu, X., Guerrero, J.M., Sun, K., Vasquez, J.C., Teodorescu, R., Huang, L.:
‘Hierarchical Control of Parallel AC–DC Converter Interfaces for Hybrid
Microgrids’ IEEE Trans. Smart Grid, 2014, 5, (2), pp. 683–692.
[25] Peyghami-Akhuleh, S., Mokhtari, H., Loh, P.C., Blaabjerg, F.: ‘Distributed
secondary control in DC microgrids with low-bandwidth communication
link’ 2016 7th Power Electronics and Drive Systems Technologies Con-
ference (PEDSTC), IEEE, 2016, pp. 641–645.
Chapter 3
Stability analysis and stabilization
of DC microgrids
Alexis Kwasinski1
1
Department of Electrical and Computer Engineering, University of Pittsburgh, USA
44 DC distribution systems and microgrids
(POL) converter. Hence, within the context of this book, it is more relevant to focus
the discussion of DC microgrids stability on the effect of CPLs and
suitable approaches for mitigating the stability issues that these loads introduce.
As Figure 3.1(a) exemplifies, instantaneous CPLs are commonly realized by
very efficient POL converters with fast controllers that regulate their output voltage
tightly. Since the output voltage of the POL converter remains sufficiently constant,
its output power with a resistive load will remain approximately constant, too.
Hence, its input power will also remain constant despite input voltage changes
because the POL converter is assumed to have a sufficiently high efficiency so
that the input power equals to the output power. Figure 3.1(b) shows a common
practical instantaneous CPL: a data center server. In the particular case of this
figure, the servers are powered in part by photovoltaic arrays through a 380 V DC
power distribution network.
Mathematically, ideal instantaneous CPLs are modeled by
PL
i ðt Þ ¼ (3.1)
v ðt Þ
where v(t) is the voltage at the CPL, i(t) is the current through the CPL and PL is the
power of the CPL. In practical applications, CPLs do not maintain indefinitely a
same value for PL, but although PL may change over time, changes in PL occur at
time scales much longer than the time constants associated to the dynamics of the
system under study. Additionally, real CPLs only show the behavior modeled by
(3.1) above a given threshold voltage. Below such voltage, protections that are part
of the CPL controller act in order to disconnect such load to prevent the current to
reach excessively high values. Nevertheless, in order to provide a systematic ana-
lysis of DC microgrids stability characteristics and control that is not dependent on
Point-of-load DC–DC
Constant DC
converter
output
voltage
DC Point-of-load converter +
input power stage VL R
voltage –
Constant power
(a) (b)
specific design choices, such as selection of CPLs low voltage threshold, for the
rest of this chapter, a voltage threshold is omitted and, thus, it is assumed that CPLs
behave as described by (3.1) irrespective of the value of v(t).
The next section of this chapter discusses the effect of CPLs on DC microgrids
stability characteristics followed by two sections describing approaches for miti-
gating the destabilizing effect of CPLs. These mitigating approaches are described
first considering passive methods followed by actively controlled strategies.
Stability challenges appear with CPLs when connecting at least two converters in a
cascade configuration, as shown in Figure 3.2. In a DC microgrid, one of these
converters would be a POL converter, whereas the other would be a converter
interfacing an electric power source with the power distribution grid powering the
POL converter. In Figure 3.2, the converter interfacing the power source is repre-
sented by a buck converter. However, the same general stability characteristics are
observed when using other type of converters to interface a power source with the
microgrid power distribution grid [5,6]. In order to initiate the discussion of DC
microgrids stability characteristics with CPLs, consider that the buck converter
interfacing a power source is controlled with a fixed duty cycle D and that it
operates in continuous conduction mode. Hence, this buck converter can be
mathematically represented by its average model given by
8
>
> di L
>L
< ¼ DE v C
dt
>
> dv C PL (3.2)
>
:C ¼ iL
dt v C
v C > e; i L 0
where E is the power source voltage; v C is the power source converter interface
(PSCI) average capacitor voltage, which equals the DC bus average voltage; iL is
the PSCI average inductor current; PL is the CPL power; L is the PSCI inductance;
and C is the PSCI output capacitor capacitance. The equilibrium point for this
converter is then
0 1 0 1
DE VO
VC
¼ @ PL A ¼ @ P L A (3.3)
IL
DE VO
It is possible to observe that this equilibrium point is not stable. In order to prove
this characteristic, it is relatively simple to find that the characteristic equation for
the small signal model of the buck PSCI with a CPL is given by
PL 1
l2 lþ ¼0 (3.4)
CVC2 LC
where l are the unknown eigenvalues. Since the coefficient for the first-order term
is negative, the characteristic polynomial associated to the buck PSCI with a CPL
does not satisfy the Routh–Hurwitz criterion and, thus, the equilibrium point for a
system represented by (3.1) is not stable.
Although small signal models allows to determine the stability characteristics
of the equilibrium point in a relatively simple way, linearization analytic methods
omit important behavior characteristics of converters with CPLs. A nonlinear
analysis approach, such as the one in [5–7], shows that depending on the PSCI
capacitor voltage and inductor current initial conditions, it is possible to observe
that the PSCI shows two behaviors: for sufficiently large initial capacitor voltages,
the voltage and current waveforms eventually settle into an oscillatory behavior,
which in a state space plot represents a limit cycle behavior. Otherwise, the capa-
citor voltage drops and the inductor current increases to very high values. As [5–7]
also show, it is possible to find that the curve—a separatrix—in the state space that
acts as a boundary to both behaviors can be approximated to
PL Cv2C
iL ¼ ðE v C Þ (3.5)
vC LPL
A simulation based on the model represented by (3.1) can be used to exemplify the
two possible behaviors of a PSCI with a CPL. For example, consider a buck PSCI
with the following parameters: E ¼ 400 V, C ¼ 3 mF, L ¼ 0.15 mH, D ¼ 0.5 and
PL ¼ 50 kW. As Figure 3.3 shows, different choices for initial conditions result in
either the oscillatory behavior or the voltage collapse-high current conditions.
Evidently, both of these behaviors are undesirable in a DC microgrid in which the
PSCI output voltage is intended to be constant. Although the assumption for the
analysis is that the converter operates in continuous conduction mode, in [7], it was
shown that in most practical conditions, a more exact model, than (3.1) that con-
siders operation in discontinuous conduction mode, does not deviate significantly
from the already discussed observations. Additionally, as it was indicated above,
Stability analysis and stabilization of DC microgrids 47
–
iL (A)
1,200
Separatrix Oscillatory limit
cycle behavior
1,000
Limit cycle
800
600
400 Equilibrium
point
Voltage
200 collapse –
behavior vC (V)
0
0 100 200 300 400 500
–200
–
iL [A]
1,200 1,200
1,000 1,000
800 – 800
iL [A]
600 600
400 400
v–C [V]
200 200
t (s) v–C [V]
0 0
0 0.02 0.04 0.06 0.08 0.1 0 100 200 300 400 500
–200 –200
(a) (b)
Figure 3.4 Time domain (a) and state-space response (b) of a buck PSCI
attempting to regulate its output voltage that is supplied to a CPL
although the analysis uses a buck converter as an example for the discussion of the
observations, as [5,6] shows, the same general behavior described for the buck
converter is also observed in other DC–DC converter topologies.
One common initial thought to address the stability issues introduced by CPLs
and to achieve a constant DC voltage on the microgrid main bus is to attempt to
regulate the PSCI output voltage with a PI controller. However, as Figure 3.4 shows
with the same PSCI used in Figure 3.3 but now with a PI controller with iL(t ¼ 0) ¼
1 A, vC (t ¼ 0) ¼ 150 V, ki ¼ 10, kp ¼ 0.1 that attempts to regulate the output vol-
tage at 200 V, the oscillatory behavior still persists exemplifying the fact that a
common PI controller, that could achieve a stable equilibrium point with a PSCI
48 DC distribution systems and microgrids
powering a resistive load, may not achieve a stable equilibrium point when the load
is an equivalent CPL. Hence, the next sections will discuss various strategies to
mitigate the stabilizing effect of CPLs and to achieve a constant regulated DC
voltage in a microgrid power distribution grid.
PL Cv2C
iL > ðE v C Þ (3.6)
vC LPL
Hence, the stability characteristics of a DC microgrid with CPLs improves when
C increases, L decreases or PL decreases. Still, (3.6) only provides an initial
approximation of suitable passive approaches for stabilizing DC microgrids.
In order to further determine passive stabilization approaches for DC micro-
grids, consider the expanded model of a buck PSCI
8
>
> diL
<L ¼ qðtÞðE RSW iL Þ ð1 qðtÞÞðVD þ iL RD Þ iL RL vC
dt
>
> C dvC ¼ iL PL vC (3.7)
:
dt v C RO
with iL 0; vC > e
in which q(t) is the main switch switching function, RSW is the conduction resis-
tance of the main switch, RD is the conduction resistance of the diode, VD is the
diode’s forward voltage drop, RL is the inductor’s series resistance and RO is an
output resistance acting as an additional load in parallel with the CPL. The new
characteristic equation of the linearized average model is
Ri PL 1 Ri 1 PL 1
l2 þ þ l þ þ ¼0 (3.8)
L CVO2 RO C L RO C CVO2 LC
where Ri is the sum of RSWD, RL and RD(1D). Based on the Routh–Hurwitz
criterion, the two necessary and sufficient conditions for a stable equilibrium
point are
2 C 1
P L < VO Ri þ (3.9)
L RO
and
1 1
PL < VO2 þ (3.10)
Ri R O
Stability analysis and stabilization of DC microgrids 49
–
iL [A]
1,400 1,400
1,200 1,200
1,000 1,000
800 – 800
iL [A]
600 600
400 400
v–C [V]
200 200
t [s] v–C [V]
0 0
0 0.02 0.04 0.06 0.08 0.1 0 50 100 150 200 250 300 350
(a) (b)
Figure 3.5 Time domain (a) and state-space behavior (b) of a buck PSCI
powering a CPL and a resistive load
50 DC distribution systems and microgrids
–
iL [A]
1,000 1,200
800 1,000
800
600
600
400 –
iL [A] 400
200
v–C [V] 200
v–C [V]
t (s)
0 0
0 0.02 0.04 0.06 0.08 0.1 0 100 200 300 400 500
–200 –200
(a) (b)
Figure 3.6 Time domain (a) and state-space behavior (b) of a buck PSCI with
increased output capacitance and powering a CPL
–
iL [A]
500 1,200
400 1,000
– 800
300 iL [A]
600
200
v–C [V] 400
100 200
t [s] v–C [V]
0 0
0 0.02 0.04 0.06 0.08 0.1 0 100 200 300 400 500
–100 –200
(a) (b)
Figure 3.7 Time domain (a) and state-space behavior (b) of a buck PSCI
powering a CPL and controlled with a PD controller
VC [V]
VO,NL
VC = f(IL1)
VC = f(IL2)
IL [A]
IL2,max IL1,max
Buck PSCI #2 PL
L2 i
L2
C2
d2
E2
Controller
Figure 3.9 Two parallel-connected buck PSCIs sharing a CPL with a droop
controller in each PSCI
V0;NL Rd;1 i L1
d1 ¼ (3.14)
E1
and
V0;NL Rd;2 i L2
d2 ¼ (3.15)
E2
Stability analysis and stabilization of DC microgrids 53
imply the insertion of virtual resistances Rd,1 and Rd,2 in series with each of the
inductors L1 and L2, respectively. Since the average current across these compo-
nents equal the output current of the PSCIs, and the steady state average output
voltage of a buck converter equals the product of the input voltage and the duty
cycle, (3.14) and (3.15) imply a static linear droop relationship between the output
voltage and output current. Dynamically, [9] shows that the addition of a droop
resistance allows limit cycle oscillations to be damped provided that the following
conditions obtained by studying the characteristic equation of the linearized system
are met:
Rd;1 Rd;2 PL
þ > (3.16)
L1 L2 ðC1 þ C2 ÞVO2
L1 þ L2 þ Rd;1 Rd;2 ðC1 þ C2 Þ PL
> 2 (3.17)
Rd;2 L1 þ Rd;1 L2 VO
V2
Rd;1 jjRd;2 < O (3.18)
PL
" #
P2L Rd;1 Rd;2 Rd;1 Rd;2 1 Rd;1 Rd;2
þ þ þ þ 2
VO4 ðC1 þ C2 Þ2 L1 L2 L1 L2 C1 þ C2 L21 L2
" 2 # " #
PL Rd;1 Rd;2 PL 1 1
2 þ þ >0
VO ðC1 þ C2 Þ L1 L2 VO2 ðC1 þ C2 Þ2 L1 L2
(3.19)
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
VO;NL þ 2
VO;NL 4PL Rd;1 jjRd;2
VO ¼
2
The effectiveness of the droop resistances insertion in order to damp the effect of
the CPL was verified experimentally with a reduced scale microgrid in [9] by using
a CPL of 100 W and two parallel buck PSCIs with E1 ¼ 35 V, E2 ¼ 30 V, C1 ¼
C2 ¼ 1,000 mF, L1 ¼ 640 mH, L2 ¼ 630 mH and VO,NL ¼ 25.7 V. Initially, the PSCIs
were controlled in open loop with a constant duty cycle equal to 0.734 for the buck
converter #1 and equal to 0.857. As Figure 3.10 shows, the result of such operation
is limit cycle oscillations for all state variables. Then, a primary droop controller as
the one represented by (3.15) and (3.14) with Rd1 ¼ 2.075 W and Rd2 ¼ 1.709 W is
enabled causing, as Figure 3.10 shows, damping of the oscillations and verifying
reaching a stable equilibrium point.
Although the insertion of virtual droop resistance through controller action
damps the limit cycle oscillations, they also naturally cause voltage deviations,
which in some cases may be undesirable. The conventional solution to this issue is
to add a so-called secondary droop control that regulates the main bus voltage by
shifting vertically the static droop lines in Figure 3.8. Thus, as indicated in [9],
54 DC distribution systems and microgrids
iL1
iL2
vC
Open loop Primary
control
Figure 3.10 Experimental results showing the transition from open loop
operation to primary droop control for two buck PSCIs with a CPL.
Image courtesy of Mahesh Srinivasan
!
1 ki Rd;1 ki Rd;2 Rd;1 Rd;2 PL
þ þ þ >0 (3.23)
C1 þ C2 L2 L1 L2 L1 L1 L2 L1 L2 VO;NL 2
ki Rd;1 þ Rd;2
>0 (3.24)
L1 L2 ðC1 þ C2 Þ
" #
P2L Rd;1 Rd;2 Rd;1 Rd;2 1 Rd;1 Rd;2
þ þ þ þ
4
VO;NL ðC 1 þ C 2 Þ2 L1 L2 L1 L2 C1 þ C2 L21 L22
" 2 #
PL Rd;1 Rd;2 1 1 1
2 þ þ >0 (3.25)
VO;NL ðC1 þ C2 Þ L1 L2 ðC1 þ C2 Þ L1 L2
Stability analysis and stabilization of DC microgrids 55
iL1
iL2
vC
Added
Primary secondary
control control
iL1
iL2
vC
P = 100 W P = 120 W
Figure 3.12 Experimental results showing droop control with load regulation for
two buck PSCIs with a CPL. Image courtesy of Mahesh Srinivasan
Output voltage regulation while still achieving a stable operating point was also
verified experimentally in [9]. As Figure 3.11 shows, the addition of integral reg-
ulators with ki ¼ 4.5 to the system used to produce Figure 3.10 compensate for the
primary droop controller voltage deviations and restores the main bus voltage to its
nominal value of 25.7 V. As it can be verified by comparing the current traces
before and after the secondary controller is enabled, the currents share the load
current maintaining a constant ratio of 1.22. As Figures 3.12 and 3.13 exemplify,
the addition of the integral stage of the controller allows to regulate the main bus
56 DC distribution systems and microgrids
iL1
iL2
vC
E1 = 35 V E1 = 30 V
Figure 3.13 Experimental results showing droop control with line regulation for
two buck PSCIs with a CPL. Image courtesy of Mahesh Srinivasan
voltage at a constant reference value even when the load’s power or one or more
input voltages change.
Although the discussion in this section still focused on buck PSCIs, virtual
droop resistances applied to other DC–DC converters will damp the limit cycle
oscillations caused by CPLs. In addition to the discussion in [9], the validity of this
conclusion and the possibility to achieve voltage regulation with an integral sec-
ondary droop controller was also described in [10,11].
Although the addition of a virtual resistance through controller action makes
the operating point stable, such controllers tend to have relatively slow dynamic
response. Moreover, in the case of the PD controller, the differential term intro-
duces noise susceptibility to the PSCIs. Geometric controllers provide another
approach to mitigate the destabilizing effect of CPLs and to achieve a
stable operating point but with a relatively fast dynamic response and without
introducing noise susceptibility. In geometric controllers, the state of the main
switch changes when the state space trajectory defined by the PSCI’s state variables
crosses a boundary. In practice, the boundary (also called switching surface) is
replaced by a hysteresis band containing such boundary in order to avoid chattering
when the trajectory tends to be confined along the switching surface. In general, it
is possible to distinguish three behaviors when a trajectory intersects a switching
surface [12]:
● Refractive behavior when the trajectories on one side of the boundary are
incident, but they are exiting on the other side of the boundary.
● Rejective behavior when trajectories on either side of the boundary are exiting.
● Reflective behavior when trajectories on both sides of the boundary are inci-
dent. Once a trajectory reaches a switching surface in a reflective behavior
region, it remains confined to the boundary. Hence, a reflective behavior could
Stability analysis and stabilization of DC microgrids 57
iL [A]
Linear
boundary
(k < 0)
Refractive area
Reflective
stable area
Reflective
unstable area
PL = constant
vC [V]
Equilibrium point
Figure 3.14 Boundary control behavior areas for a buck PSCI with a CPL
L iL
+
+ C vC PL
|vS|
–
–
Figure 3.16 Simplified model of a rectifier with a CPL in which |vS| represents a
rectified unfiltered sinusoidal voltage signal
vC(t) [V]
12
4
iL(t) [A]
t (ms)
0
0 20 40 60 80
Figure 3.17 Example of limit cycle period-1 behavior of a rectifier with a CPL
where C and L are the rectifier output filter capacitor and inductor, respectively,
and VS,O is an equivalent constant input voltage that equals |vs| when |vs| ¼ vC [14].
Condition (3.27) is necessary but not sufficient to observe a limit cycle behavior
considering the rectifier output capacitor voltage and inductor current as state
variables. In order to observe a limit cycle behavior, in addition to satisfy (3.27),
the initial capacitor voltage needs to be high enough. An approximate value for
such minimal initial capacitor voltage is given by [14]
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
" #ffi
u
u 1 P2 L1=4 2 1=4
P L P
v C ð0 Þ > t
L
L
2fVin L þ (3.28)
2fVin 4C 4C C
vC (t) [V]
12
iL(t) [A]
4
t (ms)
0
0 20 40 60 80
Figure 3.18 Example of limit cycle period-2 behavior of a rectifier with a CPL
This voltage variation is typically higher than conventional voltage ripple in rec-
tifiers with an equivalent resistive load [14].
3.6 Summary
This chapter has discussed the effects of CPLs in DC-microgrids. Such loads are
commonly observed in DC microgrids because of the extensive use of power
electronic circuits to interface loads. As it is explained, due to their nonlinear nat-
ure, CPL creates a complex dynamic system that present two possible behaviors
with an unstable equilibrium point unless PSCIs are properly controlled. In this
chapter, both time domain and geometric control approaches that achieve
stable equilibrium points in the presence of CPLs are explained. Passive approa-
ches to compensate the destabilizing effect of CPLs are also discussed, but these
approaches often lead to costly and bulky designs. Complex dynamics are also
observed in rectifiers with CPLs, which often present excessively large output
voltage variations.
References
[1] A. Toliyat and A. Kwasinski, ‘‘Energy Storage Sizing for Effective Primary
and Secondary Control of Low-Inertia Microgrids,’’ in Proc. 2015 IEEE
Power Electronics for Distributed Generation Conference, pp. 1–7, Aachen,
Germany, June 2015.
[2] N. Sony, S. Doolla and M. C. Chandorkar, ‘‘Improvement of Transient
Response in Microgrids Using Virtual Inertia,’’ IEEE Transactions on Power
Delivery, vol. 28, no. 3, pp. 1830–1838, July 2013.
Stability analysis and stabilization of DC microgrids 61
[3] X. Zhu, J. Cai, Q. Yan, J. Chen and X. Wang, ‘‘Virtual Inertia Control of
Wind-Battery-Based Islanded DC Micro-Grid,’’ in Proc. International
Conference on Renewable Power Generation (RPG 2015), pp. 1–7, October
2015.
[4] T. L. Vandoorn, B. Meersman, J. D. M. De Kooning and L. Vandevelde,
‘‘Analogy Between Conventional Grid Control and Islanded Microgrid
Control Based on a Global DC-Link Voltage Droop,’’ IEEE Transactions on
Power Delivery, vol. 27, no. 3, pp. 1405–1414, July 2012.
[5] A. Kwasinski, W. Weaver, and R. Balog, ‘‘Micro-Grids in Local Area Power
and Energy Systems,’’ Cambridge University Press, Cambridge, U.K., 2016.
[6] A. Kwasinski and C. N. Onwuchekwa, ‘‘Dynamic Behavior and Stabilization
of DC Microgrids with Instantaneous Constant-Power Loads,’’ IEEE
Transactions on Power Electronics, vol. 26, no. 3, pp. 822–834, March 2011.
[7] A. Kwasinski and P. T. Krein, ‘‘Passivity-Based Control of Buck Converters
with Constant-Power Loads,’’ in Proc. Power Electronics Specialist Con-
ference 2007 (PESC), pp. 259–265, Orlando, FL, USA, June 17–21, 2007.
[8] A. Kwasinski and P. T. Krein, ‘‘Stabilization of Constant Power Loads in
DC–DC Converters Using Passivity-Based Control,’’ in Proc. 2007 Inter-
national Telecommunications Energy Conference (INTELEC), pp. 867–874,
Rome, Italy, September 30–October 4, 2007.
[9] M. Srinivasan, ‘‘Hierarchical Control of DC Microgrids with Constant
Power Loads,’’ Ph.D. Dissertation, The University of Texas at Austin,
Austin, Texas, USA, 2017.
[10] M. Srinivasan and A. Kwasinski, ‘‘Decentralized Control of a Vehicular
Microgrid with Constant Power Loads,’’ in Proc. 2014 IEEE International
Electric Vehicle Conference, pp. 1–8, Florence, Italy, December 2014.
[11] M. Srinivasan and A. Kwasinski, ‘‘A Linear Damping Scheme for Higher
Order dc–dc Converters Supplying Constant-Power Loads in a dc Micro-
grid,’’ in Proc. 19th European Conference on Power Electronics and Appli-
cations (EPE) – Energy Conversion Congress and Exposition (ECCE)
Europe, pp. 1–10, Warsaw, Poland, September 2017.
[12] P. T. Krein, ‘‘Elements of Power Electronics,’’ Oxford University Press,
New York, NY, 1998.
[13] C. N. Onwuchekwa and A. Kwasinski, ‘‘Analysis of Boundary Control for
Buck Converters with Instantaneous Constant-Power Loads,’’ IEEE Trans-
actions on Power Electronics, vol. 25, no. 8, pp. 2018–2032, August 2010.
[14] C. N. Onwuchekwa and A. Kwasinski, ‘‘Dynamic Behavior of Single-Phase
Full-Wave Uncontrolled Rectifiers with Instantaneous Constant-Power
Loads,’’ in Proc. 2011 IEEE ECCE, pp. 3472–3479, Phoenix, AZ, USA,
September 17–22, 2011.
This page intentionally left blank
Chapter 4
Coordinated protection of DC microgrids
Jae-Do Park1, Md Habib Ullah1,
and Bhanu Babaiahgari1
4.1 Introduction
In the 2000s, the development of renewable energy resources and the advance of
power electronics technology, as well as the issues of fossil-fuel depletion and their
impact on environment, have driven extensive research on distributed power sys-
tems. The energy policy of many governments in the world competitively increases
the requirement of the penetration of renewable energy resources and distributed
generation. DC power systems have received renewed attention as a solution for the
distributed generation to replace or augment AC systems in various applications in
transmission and distribution systems. The amount of equipment that operates on
DC power is increasing, including photovoltaic (PV) panels, batteries, fuel cells,
and even AC motors driven by adjustable speed drives. Improved power electronics
converters allow DC power to be converted easily and efficiently. Unlike AC
systems, multiple sources can readily operate attached to a single DC bus because
they do not need to be synchronized. The system efficiency and reliability can be
improved with DC power delivery, because the DC-to-AC power conversion layer
can be eliminated from the back-to-back conversion scheme. Moreover, DC sys-
tems have an advantage in cable size because they do not suffer from skin effect,
and there is no need for reactive power. Multiterminal DC (MTDC) systems have
been investigated for high-voltage DC (HVDC) transmission systems and low-
voltage DC (LVDC) distribution systems as well. DC power systems have been
investigated for various microgrid systems, such as naval ships, traction systems,
data centers, buildings, aircrafts, and microgrids.
However, compared to the AC power systems that have been in operation as the
standard power system for the last two centuries after the ‘‘war of current’’ in the
early 1800s, the DC power systems have much shorter operation history. Standards
and accumulated experience are substantially insufficient. Moreover, it has been
applied to the systems in confined areas in many cases. Lack of knowledge base is
true as well when it comes to the system protection, especially for large systems.
1
Department of Electrical Engineering, University of Colorado Denver, USA
64 DC distribution systems and microgrids
Although the AC system protection technology is very mature, most of their devices
cannot be directly applied in DC systems. Typically, DC switchgear is more
expensive and may not always be available for certain systems.
The protection of electrical power systems is fundamentally essential for safe
and stable operation; it prevents personnel injury and minimizes the service inter-
ruption and system damage from equipment failure, human error, and disastrous
natural events. And this is one of the biggest challenges of DC power systems,
because of the nonalternating nature and fast dynamics of DC power. Furthermore,
although there are various devices that monitor the power system, detect abnorm-
alities, and perform automatic actions to protect the system and sustain the opera-
tion, development of a well-coordinated protection scheme for DC power system is
not a trivial task. As DC power system design is not quite a standardized procedure
yet, its protection system is developed on the case-by-case basis and designer’s
empirical techniques and methods would play an important role in design.
Typically, a DC microgrid has multiple terminals connected to the bus, to which
individual systems such as load, source, and bidirectional devices can interface
through power converters, such as voltage-source converters (VSCs) and line-
commutated converters (LCCs). Thyristors in LCCs have high-voltage and power
capability, but lower control performance compared to IGBTs in VSCs. LCCs have
been used for simpler PtP configurations. Recently, multimodular converter systems
have been widely investigated for increased voltage and power capacity of DC bus
systems. One of the challenges associated with the protection of VSC-based DC
microgrid systems is that the fault current must be detected and extinguished quickly
as the fault current withstand rating of typical VSCs is only twice the converter full-
load current. Unlike AC systems, the typical sources in DC microgrid interfaced
through VSCs do not have enough inertia to supply fault current while maintaining
bus voltage. And the VSCs will trip and disconnect themselves from the bus when
the bus voltage drops below a certain threshold. These make it difficult to locate the
fault because the transient that contains fault circuit information might not be
available to protection devices. Recently, semiconductor-based and hybrid circuit
breakers (CBs) and fault current limiters have been investigated, and they have been
used for fast DC current interruption and overcurrent limitation. Typically, CBs were
proposed for converter and battery protection, while fuses and molded-case CBs
(MCCBs) were used for feeder protection [1].
As well as the detection and interruption where traditional differential and
overcurrent relaying techniques and main/back-up scheme can be applied, location
and isolation of the fault in the microgrid are important tasks of a protection system
for fast recovery and impact minimization, which requires cooperation between
protective devices in the system. Although the line impedance method and travel-
ing wave method have been adopted as industry standard for AC systems [2], it is
difficult to directly apply to DC systems due to the inherent absence of frequency
and phase parameters and much faster current dynamics.
In this chapter, nature of the faults in DC microgrid due to lack of inductance
in the lines, typical DC power system configurations, coordinated protection
schemes, and their applications for DC power systems will be reviewed.
Coordinated protection of DC microgrids 65
(a) (b)
Figure 4.1 Types of faults in typical bipolar DC bus: (a) line-to-ground fault and
(b) line-to-line fault
66 DC distribution systems and microgrids
CDC
CDC
If 1
R L if
If 2
Es + C
–
t
(a) (b)
Figure 4.3 (a) Line-to-line fault equivalent circuit model. The fault current
consists of two current components from the source and bus
capacitors, respectively. (b) Typical fault current waveform.
The transient can be seen at the peak of the fault current
the grounded pole will not have an impact on the system, but a fault on the other
pole will cause a line-to-line short circuit through ground.
where Es is the line voltage, R, L, and C is the equivalent resistance, inductance, and
capacitance in the fault path, respectively. The equivalent impedance of the fault path,
pffiffiffiffiffiffiffi
including fault and line impedance, determines the natural frequency wo ¼ 1= LC
and damping factor of the fault current z ¼ a=wo , where a ¼ R=ð2LÞ [3]. A typical
waveform of fault current with this model is shown in Figure 4.3(b), and the transient
can be seen at the peak of the fault current. However, in case the fault resistance or
bus capacitance is large enough, such as the line-to-ground fault where the ground
resistance is considerably high, the fault current can be modeled as a first-order
damping thus having no oscillation.
It is very important to study short-circuit currents in a DC power system and
how it can be calculated. It has to be noted here that the load current is not taken
into consideration while calculating short-circuit current. A standard approximation
function of short-circuit fault current proposed by International Electrotechnical
Commission (IEC) [4] is shown in Figure 4.4, and the equations governing the
function are given by
1 et=t1
i1 ðtÞ ¼ ip ; 0 t tp (4.2)
1 etp =t1
where ip is peak short-circuit current, tp is the peak time, Ik is the quasi-steady-state
short-circuit current, t1 is the rise time constant, t2 is the decay time constant, and
Tk is the short-circuit duration.
i
τ1
ip
i2(t)
Ik
τ2
i1(t)
0 tP Tk t
Zc Zc
Γ Rf
Figure 4.5 Reflection of voltage surge at the fault location. As the surge arrives at
the fault location from VSC terminal, some parts of it are reflected
according to the reflection coefficient G as a function of fault
resistance Rf
For a HVDC system that has a long transmission line, a line-to-ground or line-
to-line fault can cause a voltage reflection, especially when the fault resistance is
small [5,6]. When a short-circuit fault happens, the voltage at the fault location
decreases and the fault current rapidly increases. Immediately negative voltage sur-
ges start to flow from the fault location into the VSC terminals, where a part of the
negative voltage surge is reflected back as a positive surge due to the termination of
cable capacitance. This traveling wave that is reflected between the fault location
and VSC terminal could increase to a very high magnitude. Figure 4.5 shows the
fault resistance Rf , reflection of surges, reflection coefficient G, and characteristic
impedance Zc , which can be given as follows with the line impedance parameters.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 R þ jwL
G¼ ; Zc ¼ (4.4)
1 þ 2 Rf =Zc G þ jwC
Converter 2 Converter 2
Converter 1 Converter 1
(a) (b)
Converter 3
Converter 1
Earthed pole
Converter 4
Converter 2
– Pole – Pole
(c)
in series between two poles at each terminal. The polarities of the poles are positive
and negative similar to the symmetric monopolar system, but there is an earthing or a
neutral point between the poles. The neutral points in the system can either be
connected to or disconnected from each other. No current flows through these points
in balanced operation. One of the poles can maintain the normal system operation
while there is disturbance in the other pole, and the AC side protection can be used to
ensure the converter protection, as well as the DC side protection schemes.
The possible faults in PtP configurations are line-to-ground fault, line-to-line
fault, and AC side faults. The monopolar system is highly vulnerable to a line-to-
ground fault as the power transmission gets fully interrupted unless a parallel path is
available, but normal system operation can be maintained with other poles in bipolar
configurations and it turns into an asymmetrical configuration; however, 50%
reduced power capacity can cause an overload to the sending side converters.
Although less likely to happen, the line-to-line fault causes very high current in the
system. As well as the direct contact between lines, two line-to-ground faults in both
poles can cause line-to-line fault if the system is ungrounded in symmetrical
monopolar system. The unit protection such as distance, directional, and differential
relaying can be effective solutions against faults in PtP configurations [8–10].
4.2.3.2 Multiterminal
In multiterminal configurations, a number of DC terminals are combined to form a
power system that has improved efficiency, reliability, and flexibility, especially
70 DC distribution systems and microgrids
during fault conditions as power can be transmitted by using multiple paths. The
system can withstand a single point of failure and effectively operate without affecting
the whole system. Various network configurations are possible for MTDC systems.
In the ring configuration shown in Figure 4.7(a), a DC microgrid system uses a
closed-loop bus. It is suitable for high-voltage operation due to less voltage fluctuation
[11,12]. This type of system can be operated either in closed or open loop depending
on operational conditions. Typically, converters are connected at each terminal
through CBs to control the power flow between the bus and connected branches. The
CBs can isolate faulted devices without shutting down the whole system. Because
the load can be supported from the either side of the connection, the system can
continue to operate even if some faulted bus segments are separated. The bus can
typically be configured with multiple zones or sections with static switches or CBs;
hence, the unit protection or directional protection techniques can be readily applic-
able for fault detection, which can be implemented by a number of intelligent CBs
placed in the bus to protect the different zones during fault conditions.
Generally, radial configuration is utilized in low-voltage and low-power sys-
tems [13]. In this structure, separate distribution feeders from a single DC grid are
fed by distributed generators. The feeders and distributed generators are connected
in the grid with interface converters and protected by CBs or other protective
devices. Figure 4.7(b) shows a radial DC power system. It is widely used in power
distribution systems due to its major advantages such as low initial cost, simplicity,
1 2
n 2
DC bus DC bus
1 3
5 3
2 n
4
DC bus DC bus
(d) (e)
Figure 4.7 Multiterminal configurations: (a) ring, (b) radial, (c) ladder, (d) star,
and (e) mesh. Rectangular boxes denote circuit breakers
Coordinated protection of DC microgrids 71
87 87
CT CT
CT CT
CT CT
+ + – +
– – + –
(a) (b)
Figure 4.9 LCC operation modes: (a) normal operation, (b) forced retard
operation. The rectifier polarity will be reversed so that the bus
current can decrease, while the inverter bypasses the DC current
through bypass pairs
has to be able to coordinate to properly discharge surge arresters and capacitors in the
bus. The system can enter the reclosing cycle, and the voltage level can be coordinated
in the restart mode (e.g., full or reduced voltage).
For MTDC systems, a ‘‘hand-shaking’’ method using fast-acting DC switches
as well as AC CBs has been proposed [23] (Figure 4.10). Although switches can
only open when there is no current, they can provide a lower cost solution com-
pared to those with DC CBs. Downside is that the DC bus has to be deenergized
because all of the AC CBs trip when a fault occurs. Using AC CBs would be one of
the economical ways to protect DC systems, but the slow interruption speed may be
an issue as the fastest AC CBs can interrupt the current in two cycles [24]. Fur-
thermore, it would completely shut down the whole system, which could be inef-
ficient especially for multiterminal systems. However, AC CBs can be readily used
as a back-up protection measure.
VSC#3 VSC#3
AC#3 AC#3
(a) (b)
AC#1 AC#2
VSC#1 VSC#2
VSC#3
(c) AC#3
Figure 4.10 Handshaking method [23]: (a) fault current flows to the fault point. (b)
All AC CBs open followed by appropriate DC switches after the fault
current is extinguished. Each VSC opens one of the associated DC
switches that had outgoing current at the time of fault. If all of them
are outgoing, the one with highest magnitude is opened. (c) Reclose
the switch if it sees the bus voltage after the AC CBs are reclosed
monitoring bus undervoltage and overcurrent, and various detection schemes have
been proposed, including large current change, rise time, and oscillation pattern
methods [25], dv=dt [5,26], di=dt and d 2 i=dt2 measurement [27], traveling wave
[28], wavelet-based [29,30], and model-based [31] detection. Using the first- and
second-order differentiation of voltage or current measurement can reduce detec-
tion time so that a smaller capacity CB can be used, but it is susceptible to signal
noise and filtering could decrease the detection speed.
Short-circuit detection of line-to-ground fault, which is most common, is
relatively straightforward unless the bus is ungrounded or high-resistance grounded
as can be seen in traction and shipboard power systems, or the fault resistance is
large. In the large fault resistance cases, the fault current is difficult to distinguish
from a normal load current or detect it with ground fault relay. For an ungrounded
bus that is adopted to avoid problems like corrosion by ground leakage current,
there is no fault current when just one line is short-circuited to ground. This is an
advantage because the system can keep operating with one line short-circuited to
ground, but the impact is catastrophic if the first fault is not cleared before the
Coordinated protection of DC microgrids 75
second line-to-ground fault happens in another pole. Although a ground fault can be
detected by measuring the voltage between the pole and ground, it is challenging to
extract more information for protection because of the low fault current level.
Unit protection techniques that compare the magnitudes or directions of
quantities at each end of a defined zone or unit, e.g., current differential protection,
can provide faster detection and better selectivity and have been widely used
typically for the primary protection method for AC systems. The current differ-
ential methods are readily applicable to DC systems and actually suitable to meet
the much faster protection requirements. Although it will require communication
between relays which increases the cost, the development of modern power systems
with increased number of sensors and communication infrastructure can help the
expansion of the usage of unit protection schemes. Differential detection schemes
have been suggested for various types and applications of DC power systems, and
many of them require high-speed data communication between protective devices.
Although it provides higher selectivity and flexibility, the communication-based
approach needs some independent back-up plan because communication error or
glitch needs to be taken into consideration. Distance protection techniques are also
widely used for AC systems to detect the fault and coordinate CBs, but small
impedance of DC cables would cause significant problems using distance relaying
methods for coordination (Figure 4.11) [5].
300
No fault detection
1st Zone 2nd Zone
250 U<,1
Voltage [kV]
dU
200
dt <,1
150
100
F1 (10 Ω)
50 F2 (0 Ω)
Figure 4.11 A dv=dt-based detection scheme [5]: protection zones and thresholds
in the voltage/voltage derivative plane is shown. Locus of the voltage
U and voltage derivative dU =dt is plotted for two line-to-ground
faults F1 (near) and F2 (far) from the measuring point. The system is
assumed to be asymmetric monopolar with pole voltage 320 kV. It
can be seen that the fault can be detected using the undervoltage
threshold, e.g., U<;1 and derivative threshold ðdU =dtÞ<;1
76 DC distribution systems and microgrids
Short-circuited
power pole
Figure 4.12 Fault current flows through short-circuited power pole via
antiparallel diode of a DC CB
Most of the devices participating in a DC microgrid are interfaced with the bus
through active power converters, such as LCCs and VSCs, which will disconnect
from the bus to protect themselves by blocking all the switches to avoid overload or
malfunctions from DC link undervoltage when a bus fault occurs. Diode rectifiers
are passive, but they typically have contactors and fuses for separation. Therefore,
it would be necessary for the protection system to interrupt the fault current and
isolate the faulted section as soon as possible to minimize the impact of the fault
and downtime of the system. However, if the DC bus is interfaced to AC grid
through a VSC, it should be noted that the VSC cannot block the fault current from
AC side. This happens because the bus voltage rapidly decreases due to the dis-
charge of DC bus capacitors and sources cannot supply enough current to sustain
the bus voltage. Hence, the protection system in DC side has to manage fault
conditions during this quite short critical time period or should coordinate with the
AC CBs to achieve a back-up protection from AC side. Similarly, if a fault happens
in VSC’s internal power poles, a single solid-state DC CB in the line cannot
block the fault current coming from the DC bus side through antiparallel diode
(Figure 4.12). Bidirectional solid-state CBs can be used, but high conduction loss
can be an issue.
medium-voltage level. DC/DC converters and choppers have been investigated for
DC fault current interruption as well. Conventional protective device coordination
techniques in AC systems, such as short-circuit calculations, main/back-up pro-
tection, and current-time coordinated settings, can be applied to the solid-state DC
CBs, fast-acting fuses, and no-fuse mechanical CBs in the DC system as well.
However, the conventional time-current coordination between CBs could be chal-
lenging in DC systems due to the fast dynamics, bidirectional nature of power flow,
and very short time interval that is allowed between CBs.
To effectively interrupt the bus from the AC and DC feeder, power converters
such as thyristor rectifiers and DC/DC converters have been suggested to limit and
block the fault current [32,33]. Fault current interruption can be coordinated with
current limiter to reduce the interruption time [34]. Modern protection systems
have adopted communication-based schemes to increase selectivity and flexibility.
It is relatively easy to isolate the faulted segment when the differential or direc-
tional detection is used (Figure 4.13) [35].
Fault location is an important task for protection systems to reduce the repair
time and system downtime. In AC systems, the fault location has been typically
performed by microcontroller-based relays that calculate the location of the fault
using short-circuit study data, traveling wave measurements, and impedance-based
methods. For the impedance-based algorithms, the related voltage and current
phase information is required, and this can be provided by phasor measurement
unit. The impedance-based distance calculation concept can readily be applied to
20
Probe voltage [V]
10
0
–10
–20
Power line 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Probe unit Time [s]
100
Probe current [A]
controller iP
Measurements 50
Control signals
Bypass switch 0
CP
for DC probing –50
–100
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time [s]
+ 2.5
VP
–
Magnitude
2
Current sensor 1.5
1
0.5
Switch-mode power supply 0
0 50 100 150 200 250 300 350
for DC/AC probe power Frequency [Hz]
(a) (b)
Figure 4.13 Sinusoidal DC fault probe unit [35]: (a) probe unit. (b) Probe
voltage, current, and FFT result (from top). The probe unit applies
voltage with sweeping frequency and the current response shows the
characteristic of fault circuit
78 DC distribution systems and microgrids
the fault location in DC microgrids. Techniques that extract information from the
instantaneous measurement have been suggested [36–38]; however, due to the
quite lower reactance and shorter length of the cables compared to those of AC
systems, it is not trivial to measure from the transient current especially when the
fault resistance is low. Furthermore, the transient happens at the top of the fault
current; however, the protection system must interrupt the fault current as
quickly as possible before the transient can be measured because the converters
typically cannot sustain the large fault current. Some off-line measurement
approaches using signal analysis have been proposed, e.g., traveling-wave-based
methods [39,40], wavelet [30,41], and probe signal injection [3,42]. Although
these can concurrently achieve fast fault current interruption and location, they
require additional hardware equipment and can be time-consuming or inaccurate
for complex configurations.
Isolation of the fault will minimize the affected area and maximize operation
continuity once the fault current is interrupted. The isolation of the faulted device,
e.g., a converter, from the bus is relatively straightforward because it typically trips
itself out, but it can generate some disturbance on the bus. For example, when a
source converter is lost from a DC microgrid system, the bus voltage will be main-
tained by other source converters if the load is sustainable with remaining sources;
however, the voltage will overshoot if that converter is back on line. Chopper circuits
[43,44], snubber circuits [45,46], and controller-based approaches [47,48] have
been suggested to suppress the overvoltage transient (Figure 4.14). Compared to the
typical control scheme that a single controller supplies the control actions for normal
and fault conditions, these multiple controller approaches have separate controllers
DC chopper
REC SEC
(a)
Ds Ll Ds Ll
Cs Cs
Rs
Rg Rs
Rg Rs
Rs
Cs Ds Cs
Ll Ll
Ds
(b) (c)
Load
AC AC
microgrid 1 microgrid 2
VSC VSC
A F A
Bus 2 Bus 3 Bus 4
20 kV DC B B B B B B F B
Bus 1
B
VSC A
B
F1 Load
AC
network
B Bus 5
B B
A F
A Bus 6
VSC
VSC
PV Load
Wind system
turbine
that regulate steady-state variables such as power and bus voltage, and transient
disturbances, e.g., loss of a converter, respectively. In a fault condition, the fault
controller takes control to maintain the best possible performance and continuous
operation by redistributing setpoints or shutting down load converters. The idea of
multiple hierarchical controllers is similar to the coordination of protective devices
in AC systems.
In case the fault happens in the bus, the faulted segment has to be isolated so
that the other intact parts of the microgrid continue to operate. This task is espe-
cially important for the mission-critical power systems. For example, the zones of
DC shipboard power systems are implemented with DC CBs, static switches and
redundant buses, so that the load can be continuously supported after the fault is
isolated. The system can be reconfigured to operate without the faulted section and
coordination between protection relays and source converters may be required in
this reconfiguration or self-healing process. In the system shown in Figure 4.15
[49], a fault can be isolated by stopping the source converters, opening the CBs,
locate the fault, and open associated static switches. Then, the system can resume
operation by restarting the source converters; however, load shedding or load
redistribution would be necessary because of the reduced capacity.
80 DC distribution systems and microgrids
4.3.3 Applications
4.3.3.1 High-voltage DC transmission
An example can be found in wind farms collector bus system where long distance
power transmission is required from off-shore wind farms to on-shore grid. Gen-
erally, wind turbines are connected to a collector bus via step-up transformer and
high-capacity converters through collector feeders. The collector bus is equipped
with protective devices to respond during faults in both collection and transmission
side. For the faults in HVDC link, unit protection schemes such as current differ-
ential, directional, and distance relaying are mostly applicable. Since it is long-
distance transmission, nonunit protection such as traveling-wave and wavelet
transformation schemes can also be potential solutions. All of these protection
schemes should have primary and back-up protection capabilities.
However, any fault in collection side also causes problem in operation of DC
collector bus. Generally, the wind turbine generators are protected by unit protec-
tion schemes with individual generator protection relay and CBs, and the trans-
former has either a fuse or separate protection relay based on the capacity. If any
fault happens near the wind turbine, the generator relay coordinates with the
transformer or converter protection devices to isolate the faulty section from the
collector feeder which is connected to the collector bus. If this primary protection
fails, then the collector bus relay is activated as a back-up protection measure.
However, to enhance the reliability of the protection, effective coordination and
communication of the protective devices need to be ensured.
DC DC DC
S-324 load S-224 load S-124 load
HV HV HV
input input input
PCR PCR PCR
AC AC AC
S-314 load S-214 load S-114 load
ring bus configuration by connecting the longitudinal buses with solid-state CBs
which allows to isolate the faulty section without deenergizing the entire bus. In a
fault condition, DC CBs should react very fast to extinguish the fault current to zero
before the upstream converters shut down by themselves. However, to coordinate
the fault isolation and system reconfiguration, effective communication is required
between the converters, CBs, and switches, and at the same time, fault location
schemes are also required. Traveling-wave-based techniques have been suggested
for locating fault as it offers sufficient accuracy [52]. Artificial neural network and
wavelets have been suggested as well [53].
i ∆t
Setting duration Trip
∆t1 Trip
(2)
(4)
(3)
(1)
Setting ∆i
∆tre
Setting di/dt Protection reset t
Figure 4.17 A di=dt protection scheme for traction power system [55]. If di=dt
rises and exceeds the set point, tripping decision will be taken only
when both Di and Dt are higher than the setting limits [curves (2) and
(3)]. However, if there is any momentary reduction in rising di=dt but
Dt is higher than the protection reset interval (Dtre ), then there will
be no tripping [curve (4)]
track feeders to see if the fault is temporary. However, typically, the traction power
system is ungrounded or grounded with high resistance to mitigate ground leakage
current and corrosion, which makes it difficult to detect line-to-ground fault. The
system experiences a line-to-line fault if ground fault happens at the same time on
another pole. Hence, it is required to detect and locate the line-to-ground fault
quickly to avoid unwanted discontinuity in the service due to a line-to-line fault.
Techniques such as off-line impedance probe unit [35], traveling wave-based
methods [56], superconducting fault current limiter [57], and high-speed digital relay-
based protection [58] have been suggested for traction power system fault protection.
IPV
Blocking
I1 diode I2 In
OCPD
GFPD Converter
PV PV PV
F1: L-G
F2: L-L
PV PV PV
PV PV PV DC bus
at low solar irradiation specially at cloudy day or evening, fault current level is
considerably low. If the fault remains undetected, it could potentially damage the
system, and furthermore, it consumes power from the healthy section as a load.
Although a blocking diode can be placed on each string to cut off the back-feeding
current, it challenges the OCPDs in fault interruption [59]. Furthermore, GFPDs in a
traditional PV system may trip the whole system for a line-to-ground fault. Recent
studies have been mainly focused on fault location and detection strategies. A vol-
tage transducers-based scheme has been proposed in [61] to protect against line-to-
ground as well as line-to-line fault, in which the top and bottom end string voltage is
compared with predefined threshold to classify and locate the fault. Differential
protection [62] and machine learning-based protection [63] strategies have been
suggested as well for PV system fault location and detection.
During a fault, the capacitor discharges so that the voltage falls down rapidly and the
amplitude of the fault current rises to a very high level within a very short time.
Moreover, this current can be detected by protective devices other than those located
on the fault path and unnecessarily trips the intact sections in the system. A relay
should be installed on the bus with coordinated fast reacting protective devices, so
that the faulty segment of the bus can be isolated without interrupting the healthy
sections.
IDG 1
External 2 IDG
AC
APU 2 Receptacle
GCU 1 GAPCU
GCU 2
APU line cont EXT power cont
Gen line cont 1 Gen line cont 2
CB CB
Main AC bus 1 Main AC bus 2
CSM/
G
CSM/G control Vital AC bus Semi-vital AC bus
FCLCB unit FCLCB
SHUNT
Figure 4.19 Single line diagram of the j150þ aircraft restructured distribution
network for a more electric aircraft (MEA) [65]
Coordinated protection of DC microgrids 85
current protection and OCPDs are mainly used for single DC bus-based system [65].
In case of the ring-bus aircraft, directional protection scheme can be applicable,
because once a fault happens at a point, then current will flow to that point from both
sides. Fault current can be interrupted by fuses, but replacement would be difficult for
aircraft onboard power systems. Resettable CBs can be used, but it may not be an
effective solution to arc fault detection and interruption. Semiconductor-based pro-
tective devices have been suggested for arc fault, which also provide protection
against short-circuit fault and overloads using I 2 t operating characteristics [68].
Recently, fault current limiting CBs (FCLCBs) based on series resonance with
superior current limiting and breaking advantages have been proposed [65]. A single
line diagram of a more electric aircraft power system is shown in Figure 4.19.
4.4 Summary
DC power systems have been investigated for transmission and distribution systems
due to the advantages such as improved efficiency, reliability, and simpler con-
figuration, as well as increasing DC-based distributed generation, energy storage,
and loads. Along with comparatively short operating history and insufficient
guidelines and standards, one of the current challenges is system protection tech-
niques to deal with the highly dynamic and nonzero crossing nature of DC power.
In this chapter, various coordinated protection schemes and applications for DC
power systems have been reviewed in conjunction with the analysis of DC fault
currents dynamics and topological bus configurations.
Acknowledgment
This work was supported by the National Science Foundation (NSF) under Award
ECCS-1554626.
References
[1] T. Dragičević and F. Blaabjerg, Microgrid: advanced control methods and
renewable energy system integration. Waltham, MA: Butterworth-Heinemann,
2016, pp. 263–279.
[2] IEEE Power and Energy Society, ‘‘IEEE guide for determining fault location
on AC transmission and distribution lines,’’ IEEE Std C37.114-2004, pp. 1–36,
2005.
[3] J. D. Park, J. Candelaria, L. Ma, and K. Dunn, ‘‘DC ring-bus microgrid fault
protection and identification of fault location,’’ IEEE Transactions on Power
Delivery, vol. 28, no. 4, pp. 2574–2584, October 2013.
[4] A. Berizzi, A. Silvestri, D. Zaninelli, and S. Massucco, ‘‘Short-circuit
current calculations for DC systems,’’ IEEE Transactions on Industry
Applications, vol. 32, no. 5, pp. 990–997, 1996.
86 DC distribution systems and microgrids
[46] F. Liu, W. Liu, X. Zha, H. Yang, and K. Feng, ‘‘Solid-state circuit breaker
snubber design for transient overvoltage suppression at bus fault interruption
in low-voltage DC microgrid,’’ IEEE Transactions on Power Electronics,
vol. 32, no. 4, pp. 3007–3021, April 2017.
[47] W. Lu and B. Ooi, ‘‘DC overvoltage control during loss of converter in
multiterminal voltage sourced converter based HVDC (M-VSC-HVDC),’’
IEEE Transactions on Power Delivery, vol. 18, no. 3, pp. 915–920, July
2003.
[48] L. Lamont, D. Jovcic, and K. Abbot, ‘‘VSC transmission control under
faults,’’ Universities Power Engineering Conference, vol. 3, pp. 1209–1213,
2004.
[49] M. Monadi, C. Koch-Ciobotaru, A. Luna, J. I. Candela, and P. Rodriguez,
‘‘Multi-terminal medium voltage DC grids fault location and isolation,’’ IET
Generation, Transmission Distribution, vol. 10, no. 14, pp. 3517–3528,
2016.
[50] R. M. Cuzner and V. Singh, ‘‘Future shipboard MVdc system protection
requirements and solid-state protective device topological tradeoffs,’’ IEEE
Journal of Emerging and Selected Topics in Power Electronics, vol. 5, no. 1,
pp. 244–259, 2017.
[51] M. W. Rose and R. M. Cuzner, ‘‘Fault isolation and reconfiguration in a
three-zone system,’’ in 2015 IEEE Electric Ship Technologies Symposium
(ESTS), 2015, pp. 409–414.
[52] E. Schweitzer, A. Guzman, M. Mynam, V. Skendzic, B. Kasztenny, and
S. Marx, ‘‘Protective relays with traveling wave technology revolutionize fault
locating,’’ IEEE Power and Energy Magazine, vol. 14, no. 2, pp. 114–120,
2016.
[53] Y. M. Yeap and A. Ukil, ‘‘Wavelet based fault analysis in HVDC system,’’
in 2014 40th Annual Conference of the IEEE Industrial Electronics Society
(IECON), 2014, pp. 2472–2478.
[54] R. Hill, ‘‘Electric railway traction. part 3. traction power supplies,’’ Power
Engineering Journal, vol. 8, no. 6, pp. 275–286, 1994.
[55] M. X. Li, J. H. He, Z.-q. Q. Bo, H. Yip, L. Yu, and A. Klimek, ‘‘Simulation
and algorithm development of protection scheme in DC traction system,’’ in
PowerTech, 2009 IEEE Bucharest. IEEE, 2009, pp. 1–6.
[56] C. Chang, T. Feng, A. Khambadkone, and S. Kumar, ‘‘Remote short-circuit
current determination in DC railway systems using wavelet transform,’’ IEE
Proceedings-Electric Power Applications, vol. 147, no. 6, pp. 520–526,
2000.
[57] H.-S. Shin, S.-M. Cho, and J.-C. Kim, ‘‘Protection scheme using SFCL for
electric railways with automatic power changeover switch system,’’ IEEE
Transactions on Applied Superconductivity, vol. 22, no. 3, pp. 1–4, 2012.
[58] E. Cinieri, A. Fumi, V. Salvatori, and C. Spalvieri, ‘‘A new high-speed
digital relay protection of the 3-kVdc electric railway lines,’’ IEEE Trans-
actions on Power Delivery, vol. 22, no. 4, pp. 2262–2270, 2007.
90 DC distribution systems and microgrids
5.1 Introduction
Over the past few years, large-scale utilization of renewable energy sources
together with the high penetration of low-to-medium voltage dc sources into the
generation mix of energy networks have increased attention toward deployment
of dc microgrids for improving system operation, reducing environmental foot-
prints, mitigating congestion and power losses through the network, increasing
energy efficiency and potentially boosting system economics. Moreover, with the
increasing share of dc appliances and loads, especially in residential and office
buildings, as well as the technological advancements in ac-to-dc rectifiers, dc
microgrids have become the point of interest for many research scientists and
engineers worldwide and have become a hot subject area for a large number of
research, development and demonstration activities. On the other hand, average
energy consumption distribution in different sectors at worldwide level illustrates
that more than one-third of the total consumption in each society is used by the
residential sector [1,2]. Although the increasing rate of energy consumption in
this sector has decelerated recently thanks to energy-aware solutions, energy
prices are still rising [3]. In this regard, many ongoing research activities are
directed toward energy efficiency in residential and office buildings and its
impact on the environment and climate [4–10]. Accordingly, a large number of
initiatives have been taken to promote sustainable consumption and green living
patterns as well as to promote delivery of affordable, clean and reliable power to
customers that, in turn, necessitates the use of smart automation and energy
management solutions. Although ‘‘energy management’’ is a term that could
mean differently, in this chapter, we are mainly concerned about the process of
energy saving and optimal scheduling of energy sources in a given system here
called microgrids.
1
Department of Energy Technology, Aalborg University, Denmark
2
Department of Energy and Power Systems, Faculty of Electrical Engineering and Computing,
University of Zagreb, Croatia
92 DC distribution systems and microgrids
Centralized P/EMS
Distributed control
Centralized control
Centralized P/EMS Centralized P/EMS
frequency of each controllable unit to control its output [41]. In the case of a dc
microgrid, bus voltages, and in the case of an ac microgrid, the system voltage and
frequency are the information sensed by each local droop controller and used
subsequently to adjust output active (and/or reactive) power as a generation unit.
Since there is no communication requirement for fulfilling the control objectives,
this control strategy is highly reliable. Moreover, this control structure could be
easily extended to different energy sources while enabling true plug-and-play fea-
tures [39]. Apart from these benefits, there are several issues in such a power/
energy management strategy. First, nonlinear current sharing among parallel units
cannot be addressed directly by this method. Second, low X/R line impedance ratio
may result in active and reactive power coupling and instability issues in low-
voltage microgrid systems and cause power sharing errors for generation units
[44,45]. Recently, work has been carried out to improve the performance of a
conventional droop-based control method by implementing the droop in virtual
frames [44, 46], adding virtual impedance in control loops [47,48], or adjusting the
output voltage bandwidth [44]. However, without a coordinating unit, such as a
central controller (CC) or a system optimizer, it would be a challenging task to
optimally manage the operation of a microgrid system with PP/EMS.
As another type of PP/EMS, maximum power point tracking (MPPT) control
methodology is also applied in microgrids to maximize power extraction from
renewable energy sources (mainly wind turbines (WTs) and photovoltaic (PV)
solar systems) under all conditions [49,50]. In such a power management techni-
que, the unit’s voltage and current are sampled frequently, and the duty ratio of the
interfaced converter is adjusted accordingly. However, it should be noted that in
islanded renewable-based microgrids, which are controlled based on MPPT prin-
ciples, energy storage systems (ESSs) must also be dispatched to provide voltage
and frequency regulation services [51].
Energy management systems for dc microgrids 97
Small wind PN
turbine Boost
converter
PV
AC
load
Battery
stack
DC
load
DC/DC Filter
Coordinated
WT Local control control
commands
Local consumers
AC/DC Converter to converter
communication DC/AC
Local control
LC 1 LC 2 LC n
…
DG 1 DG 2 DG n
Physical connection
CC
LC 1 LC 2 LC n LC 1 LC 2 LC n
… …
DG 1 DG 2 DG n DG 1 DG 2 DG n
Local terminal n
+
Power
DG n
– converter n
vdci Power-based droop
+
Power mp2
DG 2 2 vdc*
– converter 2
s
bu
ad
+
mp1
Lo
Power
DG 1
converter 1 1
–
vo1 io1 Po1 Po2 Po
PWM
Power calculation Low-pass
d1 + low-pass filter filter
vdci
Current Po1 Io1 Current-based droop
control Inner loops Power-based Current-based vdc* mc2
– droop droop
iin1 – iin1*
Voltage
control Droop controller mc1
vo1*
Converter control 1
Io1 Io2 Io
complex. It should be noted that alternative control methods for dc microgrids have
recently drawn a lot of attention in academic circles. However, their practical
application should be elaborately justified by performing modeling, analysis,
simulation, implementation, as well as a full cost-benefit analysis. For instance,
increased production cost and lead time often proves to be too large an obstacle for
their deployment.
Droop control is commonly installed on top of inner loops, primarily for cur-
rent sharing purposes. Figure 5.8 demonstrates that either output power or output
current can be selected as the feedback signal in droop control [60,61]. For dc
microgrids with a power-type load, output power can be used as droop feedback, as
shown in (1). On the other hand, when current signal is used, as shown in (2), droop
coefficient mc can be regarded as a virtual internal resistance. In this case, the
implementation and design of the parallel converter system in a dc microgrid can
be simplified to some extent as the control law is linear [60]. The principle of
current-based droop control was also extensively used in distributed power systems
for paralleling multiphase converters that supply computer CPUs. Here, droop
control is commonly known as adaptive voltage positioning [62–64]. The calcula-
tions of references for voltage controller, in the two aforementioned cases, are as
follows:
where vdci is the output of the droop controller, i.e. the reference value of the dc
output voltage of converter i, vdc is the rated value of the dc voltage; mp and mc are
the droop coefficients in power- and current-based droop controllers, while Poi and
ioi are the output power and current, respectively, of converter i.
1
For the exact definition and a more in-depth discussion of the connectivity of communication net-
works, please refer to [83].
104 DC distribution systems and microgrids
where xi(t) and xj(t) are the values of variables of interest in LC i and LC j,
respectively. Here, j is iterated through all of the neighbors of LC i, represented
by Ni. Finally, bi(t) is an optional input bias of LC, which can be used to
declare it as a virtual leader. It can be seen from (5.3) that xi(t) is interactively
adapted with respect to the values of its neighboring units. Likewise, variables
in any other controller adapt with respect to the values of their own neighbors.
Consequently, it can be analytically proved that, if the communication network
is connected, all variable values will converge to a common average after a
certain amount of time [82,83]. Another option is to use a nonzero input bias in
one of the LCs. In that case, variables of all other LCs will converge to their
respective bias [84]. In either case, the ability of a consensus to share infor-
mation in such a manner has wider applicability than simple data averaging.
For instance, if every LC has information on the number of other active LCs,
an exact value of any specific variable can be calculated directly from the
average.
The collective dynamics of a communication system realized via consensus
protocol can be represented by the following equation:
where Q ¼ [qij] is the graph Laplacian of the network whose elements are
defined as follows:
1 j 2 Ni
qij ¼ (5.5)
jNi j j ¼ i
where |Ni| denotes the number of neighbors of node i. The topology of the
communication network is explicitly reflected by graph Laplacian, and it is
also possible to design weights of the respective matrix to control the con-
vergence speed [85]. Figure 5.9 shows the configurations of an exemplary
power electronics-based microgrid and its sparse communication network
In summary, it can be concluded that distributed control can achieve information
awareness comparable to that of centralized control. Therefore, objectives such as
output current sharing, voltage restoration, global efficiency enhancement, SoC
balancing and others can be easily realized. In this sense, distributed control offers
much wider functionalities than decentralized control, but remains protected from
the single point of failure. Its main limitation is the complexity of analytical per-
formance analysis, i.e. assessment of convergence speed and stability margins,
Energy management systems for dc microgrids 105
Converter n
Converter i
Converter 2
Converter 1
Converter n
.
.
. DC
microgrid Converter i
Converter 2
Converter 1
vi = E∠φ
Zl = Z∠θ vi = V∠0
Grid
Sij = Pij + jQij
DG unit PCC
EV X Pij
Pij ffi sinðjÞ ! j ffi (5.7)
X EV
E2 E V X Qij
Qij ffi cosðjÞ ! E V ffi (5.8)
X X E
meaning that the active and reactive power outputs of a given DG unit are propor-
tional to the phase angle difference j and the voltage magnitude difference (EV),
respectively. In other words, by controlling the real power, it would be possible to
adjust the power angle and, through reactive power control, voltage regulation can be
achieved. In a droop-based control, the same procedure is adopted with the exception
that each unit uses the frequency (instead of the power angle or phase angle), for
controlling active power flow.
Figure 5.11 shows the block diagram of the VCM control strategy which can
be applied in both grid-tied and islanded operating modes of microgrids and makes
the operation mode transition easy and smooth. As observed from the same figure,
active and reactive power references are provided by the P/EMS to meet certain goals
such as voltage profile smoothing. The lower level controllers embedded in the VCM
structure (named as active and reactive power controllers in Figure 5.11) assure that
the output voltage phase angle and magnitude are of the right orders and that they can
be designed as proportional controllers for realizing active power–frequency droop
(P–w) and reactive power–voltage magnitude droop (Q–E) as follows:
w ¼ wref m Pmeas Pref (5.9)
V ¼ Vref n Qmeas Qref (5.10)
In this strategy, inner current loops can be assigned to the voltage closed-loop
control system to improve the transient and stability performance [87].
Energy management systems for dc microgrids 107
Vref + Iref
Voltage + Current
_ controller _ controller
Vmeas Imeas
Qmeas
Vdc,ref
+ DC voltage Id,ref
_ controller
Vdc,meas
PCC
Pref + Active Id,ref + Current
power _ controller dq
_ controller PWM ZL Grid
abc
Pmeas
Qmeas Id
dq
Ig
abc
Iq
More complex controllers can also be integrated into the VCM to closely mimic
the behavior of a synchronous machine equipped with governor and excitation
systems [88].
Residential microgrid
Energy Information
RSEMS
Smart Home
meter scheduler Local
controllers
Distributed MGCC
generation
Local controller
µCHP
PCC
Power grid
E_*
ON/OFF
Qmaster q Vc
Voltage E + nd PQ
LPF Power
reference Pmaster io
θ md p calculation
generator 1/S _ LPF
+
ω*
Droop control
Schedulable
load
Slave unit (PCM) iL Lf Vg
Vdc
ωg
Vg
Iref + + + dq d
_ PI PWM
abc
abc
dq ωg
PLL
Inner-loop control
ωg
Vgd Vgq E*
Current Qslave + _ Eg E'g
1/nr LPF
reference
RSEMS
generator Pslave = P*
Reverse droop control
It should be mentioned that within the mentioned structure, the inner loop control
of the VCM aims to achieve good output voltage regulation with respect to a
determined capacitor voltage reference while the inner loop control of the PCM
regulates the output power. For both inner loops, typical PI controllers are appro-
priately designed and used. In the primary control loops, low-pass filters are also
applied to limit the loop bandwidth, so that the primary control can be separately
designed and the inner loop of the VCM and CCM can be considered as ideal
voltage and current sources, respectively.
5.5 Conclusions
With the application of dc microgrids at different scales and topologies, it is pos-
sible to integrate different energy sources (such PV systems, fuel cells and bat-
teries) into the energy mix of a larger system with lower conversion requirements.
Energy management systems for dc microgrids 111
Moreover, it is much easier and cheaper to convert ac power into dc power which,
in turn, helps economic operation of future energy systems. By using dc microgrids,
it is also possible to achieve energy savings up to a great extent (~15%) and to
improve system reliability by reducing the number of devices required and the total
points of failure. The dc architecture also enables cost-effective and green solutions
for operation and control of zero-net energy residential/office buildings as well as
data centers. However, such optimal performance mainly depends on the proper
design and application of EMSs which effectively manage the process of energy
production and consumption based on predefined objectives and constraints. There
are, definitely, a number of challenges in this regard which have to be suitably
addressed. Currently, there is a lack of approved standards and technical codes for
dc equipment and distribution networks at low voltage. There is also a lack of
approved and recognized dc architectures at low-to-medium voltage levels which in
turn necessitates different safety and protection practices in comparison with con-
ventional ac systems. Last but not least, there is a strong need for upgrading the
existing infrastructure to accommodate dc systems and interfaces.
References
[1] American Physical Society, ‘‘Buildings,’’ in ENERGY FUTURE: Think
Efficiency. American Physical Society, 2008, pp. 52–85.
[2] International Energy Agency, Transition to Sustainable Buildings – Strategies
and Opportunities to 2050. Paris: OECD/IEA, 2013.
[3] N. Dempsey, C. Barton, and D. Hough, ‘‘Energy Prices,’’ in The House of
Commons Library, 2016, no. 4153, pp. 1–18.
[4] A. Mousavi, C. W. Yang, C. Pang, and V. Vyatkin, ‘‘Energy efficient
automation model for office buildings based on ontology, agents and IEC
61499 function blocks,’’ in 19th IEEE Int. Conf. Emerg. Technol. Fact.
Autom. ETFA 2014, 2014, pp. 1–7.
[5] C. Delmastro, G. Mutani, M. Pastorelli, and G. Vicentini, ‘‘Urban mor-
phology and energy consumption in Italian residential buildings Urban
morphology and energy consumption in Italian residential buildings,’’ in 4th
International Conference on Renewable Energy Research and Applications,
2015, vol. 5, no. Jan. 2016, pp. 1603–1608.
[6] S. I. Symposium, ‘‘Simulation-based optimization in energy efficiency ret-
rofit for office building,’’ in Proc. 2014 IEEE/SICE Int., 2014, pp. 222–227.
[7] K. Vatanparvar, Q. Chau, A. Faruque, et al., ‘‘Home energy management as
a service over networking platforms,’’ in IEEE PES Conference on Innova-
tive Smart Grid Technologies (ISGT), Washington, DC, 2015, pp. 1–5.
[8] A. Gligor, H. Grif, and S. Oltean, ‘‘Considerations on an intelligent buildings
management system for an optimized energy consumption,’’ in 2006 IEEE
Int. Conf. Autom. Qual. Testing, Robot. AQTR, 2006, no. 1, pp. 280–284.
[9] A. Anvari-Moghaddam, H. Monsef, and A. Rahimi-Kian, ‘‘Optimal Smart
Home Energy Management Considering Energy Saving and a
112 DC distribution systems and microgrids
Comfortable Lifestyle,’’ IEEE Trans. Smart Grid, vol. 6, no. 1, pp. 324–332,
2015.
[10] A. Anvari-Moghaddam, H. Monsef, and A. Rahimi-Kian, ‘‘Cost-Effective
and Comfort-Aware Residential Energy Management under Different Pricing
Schemes and Weather Conditions,’’ Energy Build., vol. 86, pp. 782–793,
2015.
[11] Y. Li and F. Nejabatkhah, ‘‘Overview of Control, Integration and Energy
Management of Microgrids,’’ J. Mod. Power Syst. Clean Energy, vol. 2,
no. 3, pp. 212–222, Aug. 2014.
[12] A. Ghasemkhani, A. Anvari-Moghaddam, J. M. Guerrero, and B. Bak-
Jensen, ‘‘An efficient multi-objective approach for designing of commu-
nication interfaces in smart grids,’’ in IEEE PES Innovative Smart Grid
Technologies (ISGT 2016), 2016, pp. 1–6.
[13] S. Z. Islam, N. Mariun, H. Hizam, et al., ‘‘Communication for distributed
renewable generations (DRGs): A review on the penetration to smart grids
(SGs),’’ in PECon 2012 – 2012 IEEE Int. Conf. Power Energy, 2012,
no. December, pp. 870–875.
[14] K. M. Liyanage, M. A. M. Manaz, A. Yokoyama, Y. Ota, H. Taniguchi, and
T. Nakajima, ‘‘Impact of communication over a TCP/IP network on the
performance of a coordinated control scheme to reduce power fluctuation
due to distributed renewable energy generation,’’ in 2011 6th Int. Conf. Ind.
Inf. Syst. ICIIS 2011 – Conf. Proc., 2011, pp. 198–203.
[15] A. Anvari-Moghaddam, G. Mokhtari, and J. M. Guerrero, ‘‘Coordinated
Demand Response and Distributed Generation Management in Residential
Smart Microgrids,’’ in Energy Management of Distributed Generation
Systems, 1st ed., Dr. Eng. L. Mihet, Ed. InTech, 2016, pp. 27–57.
[16] M. Sechilariu, B. Wang, and F. Locment, ‘‘Power management and optimi-
zation for isolated DC microgrid,’’in 2014 International Symposium on
Power Electronics, Electrical Drives, Automation and Motion, Ischia, 2014,
pp. 1284–1289.
[17] F. Nejabatkhah and Y. W. Li, ‘‘Overview of Power Management Strategies
of Hybrid AC/DC Microgrid,’’ IEEE Trans. Power Electron., vol. 30, no. 12,
pp. 7072–7089, 2015.
[18] D. E. Olivares, S. Member, and C. A. Cañizares, ‘‘A Centralized Energy
Management System for Isolated Microgrids,’’ IEEE Trans. Smart Grid,
vol. 5, no. 4, pp. 1864–1875, 2014.
[19] L. E. Zubieta, ‘‘Power Management and Optimization Concept for DC
Microgrids,’’ in 2015 IEEE First International Conference on DC Micro-
grids (ICDCM), 2015, pp. 81–85.
[20] M. Parvizimosaed, F. Farmani, and A. Anvari-Moghaddam, ‘‘Optimal Energy
Management of a Micro-Grid with Renewable Energy Resources and Demand
Response,’’ J. Renewable Sustainable Energy, vol. 5, no. 5, pp. 31–48, 2013.
[21] A. Anvari-Moghaddam, A. Seifi, and T. Niknam, ‘‘Multi-Operation Man-
agement of a Typical Micro-Grids Using Particle Swarm Optimization:
A Comparative Study,’’ Renewable Sustainable Energy Rev., vol. 16, no. 2,
pp. 1268–1281, 2012.
Energy management systems for dc microgrids 113
[36] T. Lu, Z. Wang, Q. Ai, and W.-J. Lee, ‘‘Interactive Model for Energy
Management of Clustered Microgrids,’’ IEEE Trans. Ind. Appl., vol. 53,
no. 3, pp. 1739–1750, 2017.
[37] E. Rodriguez, J. C. Vasquez, M. Josep, et al., ‘‘Multi-level energy manage-
ment and optimal control of a residential DC microgrid,’’ in 2017 IEEE
International Conference on Consumer Electronics (ICCE), Las Vegas, NV,
2017, pp. 312–313.
[38] A. Anvari-moghadam, Q. Shafiee, J. C. Vasquez, and J. M. Guerrero,
‘‘Optimal adaptive droop control for effective load sharing in AC micro-
grids,’’ in 42nd Annual Conference of the IEEE Industrial Electronics
Society (IECON’16), 2016, pp. 2–7.
[39] A. Tah and D. Das, ‘‘An Enhanced Droop Control Method for Accurate
Load Sharing and Voltage Improvement of Isolated and Interconnected DC
Microgrids,’’ IEEE Trans. Sustainable Energy, vol. 7, no. 3, pp. 1194–1204,
2016.
[40] P. H. Huang, P. C. Liu, W. Xiao, and M. S. El Moursi, ‘‘A Novel Droop-
Based Average Voltage Sharing Control Strategy for DC Microgrids,’’ IEEE
Trans. Smart Grid, vol. 6, no. 3, pp. 1096–1106, 2015.
[41] A. Maknouninejad, Z. Qu, F. L. Lewis, and A. Davoudi, ‘‘Optimal, Non-
linear, and Distributed Designs of Droop Controls for DC Microgrids,’’
IEEE Trans. Smart Grid, vol. 5, no. 5, pp. 2508–2516, 2014.
[42] A. P. N. Tahim, D. J. Pagano, E. Lenz, and V. Stramosk, ‘‘Modeling and
Stability Analysis of Islanded DC Microgrids under Droop Control,’’ IEEE
Trans. Power Electron., vol. 30, no. 8, pp. 4597–4607, 2015.
[43] K. J. Bunker and W. W. Weaver, ‘‘Multidimensional Droop Control for
Wind Resources in dc Microgrids,’’ IET Gener. Transm. Distrib., vol. 11,
no. 3, pp. 657–664, 2017.
[44] Y. Li and Y. W. Li, ‘‘Power Management of Inverter Interfaced Autonomous
Microgrid based on Virtual Frequency-Voltage Frame,’’ IEEE Trans. Smart
Grid, vol. 2, no. 1, pp. 18–28, 2011.
[45] Y. W. Li and C. N. Kao, ‘‘An Accurate Power Control Strategy for Power-
Electronics-Interfaced Distributed Generation Units Operating in a Low-
Voltage Multibus Microgrid,’’ IEEE Trans. Power Electron., vol. 24, no. 12,
pp. 2977–2988, 2009..
[46] Y. Guan, J. M. Guerrero, X. Zhao, J. C. Vasquez, and X. Guo, ‘‘A New
Way of Controlling Parallel-Connected Inverters by Using Synchronous-
Reference-Frame Virtual Impedance Loop – Part I: Control Principle,’’
IEEE Trans. Power Electron., vol. 31, no. 6, pp. 4576–4593, 2016.
[47] J. Matas, M. Castilla, L. G. de Vicuña, J. Miret, and J. C. Vasquez, ‘‘Virtual
Impedance Loop for Droop-Controlled Single-Phase Parallel Inverters Using
a Second-Order General-Integrator Scheme,’’ IEEE Trans. Power Electron.,
vol. 25, no. 12, pp. 2993–3002, 2010.
[48] Y. Zhang and Y. Wei Li, ‘‘Energy Management Strategy for Supercapacitor
Virtual Impedance,’’ IEEE Trans. Power Electron., vol. 32, no. 4, pp.
2704–2716, 2017.
Energy management systems for dc microgrids 115
Multiphase VRMs,’’ IEEE Trans. Power Electron., vol. 23, no. 4, pp. 1733–
1742, 2008.
[63] C.-J. Chen, D. Chen, C.-S. Huang, M. Lee, and E. K.-L. Tseng, ‘‘Modeling and
Design Considerations of a Novel High-Gain Peak Current Control Scheme
to Achieve Adaptive Voltage Positioning (AVP) for DC Power Converters,’’
IEEE Trans. Power Electron., vol. 24, no. 12, pp. 2942–2950, 2009.
[64] H.-H. Huang, C.-Y. Hsieh, J.-Y. Liao, and K.-H. Chen, ‘‘Adaptive Droop
Resistance Technique for Adaptive Voltage Positioning in Boost DC–DC
Converters,’’ IEEE Trans. Power Electron., vol. 26, no. 7, pp. 1920–1932, 2011.
[65] F. Valenciaga and P. F. F. Puleston, ‘‘Supervisor Control for a Stand-Alone
Hybrid Generation System Using Wind and Photovoltaic Energy,’’ IEEE
Trans. Energy Convers., vol. 20, no. 2, pp. 398–405, Jun. 2005.
[66] F. Valenciaga and P. F. Puleston, ‘‘High-Order Sliding Control for a Wind
Energy Convers. System Based on a Permanent Magnet Synchronous Gen-
erator,’’ IEEE Trans. Energy Convers., vol. 23, no. 3, pp. 860–867, 2008.
[67] T. Dragicevic, J. M. Guerrero, J. C. Vasquez, and D. Skrlec, ‘‘Supervisory
Control of an Adaptive-Droop Regulated DC Microgrid With Battery
Management Capability,’’ IEEE Trans. Power Electron., vol. 29, no. 2,
pp. 695–706, 2014.
[68] L. Che and M. Shahidehpour, ‘‘DC Microgrids: Economic Operation and
Enhancement of Resilience by Hierarchical Control,’’ IEEE Trans. Smart
Grid, vol. 5, no. 5, pp. 2517–2526, 2014.
[69] X. Lu, J. M. M. Guerrero, K. Sun, J. C. C. Vasquez, R. Teodorescu, and
L. Huang, ‘‘Hierarchical Control of Parallel AC–DC Converter Interfaces for
Hybrid Microgrids,’’ IEEE Trans. Smart Grid, vol. 5, no. 2, pp. 683–692,
2014.
[70] B. Wang, M. Sechilariu, and F. Locment, ‘‘Intelligent DC Microgrid With
Smart Grid Communications: Control Strategy Consideration and Design,’’
IEEE Trans. Smart Grid, vol. 3, no. 4, pp. 2148–2156, 2012.
[71] C. Jin, P. Wang, J. Xiao, Y. Tang, and F. H. H. Choo, ‘‘Implementation of
Hierarchical Control in DC Microgrids,’’ IEEE Trans. Ind. Electron., vol. 61,
no. 8, pp. 4032–4042, 2014.
[72] Q. Shafiee, T. Dragicevic, J. C. Vasquez, and J. M. Guerrero, ‘‘Hierarchical
Control for Multiple DC-Microgrids Clusters,’’ IEEE Trans. Energy Con-
version, vol. 29, no. 4, pp. 922–933, 2014.
[73] J. Schonberger, R. Duke, and S. D. Round, ‘‘DC-Bus Signaling: A Dis-
tributed Control Strategy for a Hybrid Renewable Nanogrid,’’ IEEE Trans.
Ind. Electron., vol. 53, no. 5, pp. 1453–1460, 2006.
[74] Y. Gu, X. Xiang, W. Li, and X. He, ‘‘Mode-Adaptive Decentralized Control
for Renewable DC Microgrid With Enhanced Reliability and Flexibility,’’
IEEE Trans. Power Electron., vol. 29, no. 9, pp. 5072–5080, 2014.
[75] A. Gkountaras, S. Dieckerhoff, and T. Sezi, ‘‘Performance analysis of hybrid
microgrids applying SoC-adaptive droop control,’’ in Power Electronics and
Applications (EPE’14-ECCE Europe), 2014 16th European Conference on,
2014, pp. 1–10.
Energy management systems for dc microgrids 117
6.1 Introduction
With the world-wide energy shortage and deterioration of existing power grid, micro-
grid becomes one of the hottest research directions in the power engineering area.
Considering DC nature of many key components in the smart grid, such as photovoltaic
(PV), battery, fuel cell, super capacitor, etc., as well as many DC type loads, such as
light-emitting diode, DC microgrid has received more attention recently since it brings
the opportunity for boosting the efficiency by eliminating the unnecessary power
conversion stages. However, the existing DC microgrid can only interface with the
distribution system by using a heavy and bulky passive line frequency transformer plus
a rectifier, which has large space and heavy weight. Developing a more compact and
active grid interface to enable an intelligent DC microgrid system is still a research
focus. In this chapter, the solid-state transformer (SST)-enabled DC microgrid is pre-
sented. In addition, two system control strategies, namely, the centralized power
management and hierarchical power management strategies, are proposed. In addition,
an improved control strategy is proposed for increasing the penetration of distributed
renewable energy resources (DRER) integration, which controls SST-enabled DC
microgrid as a solid-state synchronous machine (SSSM). With the proposed control
concept, frequency and voltage stability are improved in case of high-power inter-
mittence at either DC load or DRER side. Design examples are given to illustrate the
main characteristics of the presented system and control schemes.
1
Electric Power, GE Global Research, United States
2
Electric and Computer Engineering, University of Texas Austin, United States
3
Intersil Corporation, United States
4
Grid Bridge Inc., United States
120 DC distribution systems and microgrids
Switching Switching
Source network Transformer network Load
technique, which is represented in Figure 6.1 [1]. The basic operation of the SST is
firstly to convert the 50/60 Hz AC voltage to a high-frequency one (normally in the
range of several to tens of kilohertz), then this high-frequency voltage is stepped
up/down by a high-frequency transformer with significantly decreased volume and
weight, and finally shaped back into the desired 50/60 Hz voltage to feed the load.
Therefore, the first advantage that the SST can offer is its reduced volume and
weight compared with traditional transformers. It is further seen from the config-
uration of the SST that some other potential functionalities that are not owned by
the traditional transformers may be obtained. First, the use of solid-state semi-
conductor devices and circuits makes the voltage and current regulation a possi-
bility. This brings promising features such as power flow control, voltage sag
compensation, fault current limitation, and others, which are not possible for tra-
ditional transformers. Second, voltage source converters connected from the sec-
ondary terminal of the SST could readily support a regulated DC bus, which could
be connected to DC microgrid, enabling this new microgrid architecture.
Figure 6.2 demonstrates the proposed SST-based DC microgrid system [2].
The AC loads are connected to the AC mains of SST, and the DC sources and loads
are connected to the DC mains for minimizing the conversion stages. The high-
voltage side of the SST is connected to the distribution system; thus, the conventional
transformer is eliminated. The presence of SST can also isolate the distribution
system from the residential side; therefore, the fault on one side will not affect
another. Furthermore, the high-voltage AC/DC converter in the distribution system
side can realize power factor regulation, enabling the Var compensation capability.
Therefore, the SST-based microgrid is more compact and shows superior character-
istics over conventional AC and DC microgrids.
Similar like the traditional microgrids, different power management strategies
could be applied. In the following sections, both the centralized and hierarchical
power management strategies are presented. In addition, SSSM control concepts
are also proposed to enhance the system stability at high-power intermittence
condition. Design examples are given to demonstrate the feasibility of the proposed
SST-enabled DC microgrid.
Control of solid-state transformer-enabled DC microgrids 121
Solid-state AC
transformer load
Grid
DC
load
AC
DESD DC/DC DC
SST-based
microgrid
DRER DC/DC
Passive grid
interaction
mode
Active grid
Islanding
interaction
mode
mode
If divided in detail, there are ten possible operating modes for the presented
system, and they are shown in Figure 6.4. The nomenclature in the flowchart is
defined in Table 6.1.
The division of the modes mainly depends on the operating status of the dis-
tribution system, battery state of charge (SOC), and battery power. These modes are
summarized and explained as below.
Mode 1: PV operates in maximum power point tracking (MPPT) mode, battery
stops working, and SST supplies the additional power for DC load.
Mode 2: PV operates in MPPT mode, battery operates in discharging power
limitation mode, and SST supplies the additional power for DC load.
Mode 3: PV operates in MPPT mode, battery balances the power within DC
microgrid, and SST only supplies power for AC load.
Mode 4: PV operates in MPPT mode, battery operates in charging power
limitation mode, and SST absorbs the additional power from DC microgrid.
Mode 5: PV operates in MPPT mode, battery stops working, and SST absorbs
the additional power from DC microgrid.
Mode 6: Part of loads is shed and the system operates in either Mode 8 or
Mode 9, depending on the operating condition.
Mode 7: Part of loads is shed and the system operates in Mode 8, Mode 9, or
Mode 10, depending on the operating condition.
Mode 8: PV operates in MPPT mode and battery operates in voltage tracking
mode.
Mode 9: PV operates in power tracking mode for supplying the total load
power. In this mode, the power reference of PV converter is set to
Ppv ¼ Pload DC þ Pload DC þ Pb max (6.1)
Intelligent energy
management
No Yes No Yes
P*>0 P*>0
Figure 6.4 Intelligent energy management system diagram of SST using PV and battery
124 DC distribution systems and microgrids
Ppv Power of PV
Pload_DC Power of DC load
Pload_AC Power of AC load
SOCmin Minimum allowed SOC
SOCmax Maximum allowed SOC
Pbmax Maximum allowed battery power
The battery operates in voltage tracking mode, and its absorbed power is
fixed to charging power limitation automatically.
Mode 10: PV operates in power tracking mode for supplying the total load power
and the battery controls the DC bus voltage while no power is delivered to the
load.
As it can be seen above, Modes 1–5 belong to the active grid interaction mode.
Mode 3 belongs to the passive grid interaction mode. Modes 6–10 belong to the
islanding mode. In the presented power management strategy, the matched battery
capacity needs to be designed, and the system should operate in Mode 3 for most
of the time. Only when the local power balancing capability is limited, a transition
to other modes occurs. Thus, the presented IEM system can maximize the utili-
zation of PV and battery and minimize the effect to the existing AC power grid
architecture.
1,200
1,100
1,000
Irradiation (W/m2)
900
800
700
600
500
400
300
200
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(a) Time (s)
15,000
DC load power
5,000
Battery power
–5,000
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(b) Time (s)
600
240 V AC voltage 400 V DC voltage 120 V AC voltage
400
Voltage (V)
200
–200
–400
2.5 2.55 2.6 2.65 2.7
(c) Time (s)
Figure 6.5 Key operating waveforms in Mode 3 of the system in Figure 6.2:
(a) irradiation of PV panel, (b) power distribution of the system,
(c) low-voltage terminal waveforms, (d) high-voltage terminal
waveforms, (e) SOC of the battery
126 DC distribution systems and microgrids
1.5
Distribution voltage (7.2 kV base) Input current (3 A base)
1
Voltage (V) Current (A)
0.5
–0.5
–1
–1.5
2.5 2.55 2.6 2.65 2.7
(d) Time (s)
80.04
80.03
State of charge (%)
80.02
80.01
80
79.99
79.98
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(e) Time (s)
The power of the PV has the same trend with the irradiation. SST only supplies
the AC load power, which is 10 kW. Since the power of the battery is within its
limitation, it can balance the power between DC load and PV. Figure 6.5(c)
shows the regulated low-voltage terminal waveforms, including 400 V DC and
120/240 V AC. In Figure 6.5(d), the SST input voltage and current are depicted.
The current is in phase with voltage, indicating a unity power factor operation.
The battery SOC is described by Figure 6.5(e), where the initial SOC is set to
80%. The trend of SOC depends on the direction of current. A positive battery
current discharges the battery and a negative battery current charges the battery in
the presented system in Figure 6.2.
Voltage (V)
Power (W)
0.5 200
DC load power
0 0
Battery power
–0.5 –200
Passive grid interaction mode Active grid interaction mode Passive grid interaction mode
–1 –400
0.5 1 1.5 2 2.5 3 3.5 4 4.5 2.5 2.55 2.6 2.65 2.7
(a) Time (s) (b) Time (s)
1.5 80.1
1 80.08
80.06
0
80.04
–0.5
80.02
–1
Passive grid interaction mode Active grid interaction mode Passive grid interaction mode
–1.5 80
0.5 1 1.5 2 2.5 3 3.5 4 4.5 0.5 1 1.5 2 2.5 3 3.5 4 4.5
(c) Time (s) (d) Time (s)
Figure 6.6 Key operating waveforms during mode transition (Mode 2 to Mode 3): (a) power distribution of the system,
(b) low-voltage terminal waveforms, (c) high-voltage terminal current (3 A base), (d) SOC of the battery pack
128 DC distribution systems and microgrids
are shown in Figure 6.6. Figure 6.6(a) shows the power distribution of the system.
At the beginning, the power of battery pack is within the limitation and the system
operates in the passive grid interaction mode (Mode 3). With the irradiation of PV
panel increasing, the charging power of the battery also increases. Then the system
transits into the active grid interaction mode (Mode 2) when battery power reaches
the 6 kW limitation. The additional power of DC microgrid flows into the SST and
supplies to the AC load. After that, the PV power decreases to a certain value and
the system operates back to passive grid interaction mode. Figure 6.6(b) depicts the
well-regulated AC and DC voltages. Figure 6.6(c) demonstrates the input current of
SST, which shows the same trend with the SST power since the input voltage is
fixed. Figure 6.6(d) shows the SOC of the battery and it increases continuously.
As it can be seen, the transition between these two modes is smooth and occurs
automatically.
6.3.3 Summary
A centralized power management strategy is presented for SST-enabled microgrid
system with PV and battery. It renders the coordinate management of power
Control of solid-state transformer-enabled DC microgrids 129
10,000
PV power
8,000
0
Battery power
–2,000
–4,000
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(a) Time (s)
600
240 V AC voltage 400 V DC voltage 120 V AC voltage
400
Voltage (V)
200
–200
–400
2.5 2.55 2.6 2.65 2.7
(b) Time (s)
80.01
80.005
State of charge (%)
80
79.995
79.99
79.985
79.98
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(c) Time (s)
Figure 6.7 Mode 8 key operating waveforms: (a) power distribution of the system,
(b) low-voltage terminal waveforms, (c) SOC of battery pack
× 104
1
PV power
0.5
DC load power
Power (W)
0
AC load power
–0.5
Battery power
Mode 8 Mode 9
–1
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(a) Time (s)
500
450
Voltage (V)
400
300
2.6 2.65 2.7 2.75 2.8 2.85 2.9 2.95 3
(b) Time (s)
80.08
80.06
State of charge (%)
80.04
80.02
80
79.98
0.5 1 1.5 2 2.5 3 3.5 4 4.5
(c) Time (s)
Figure 6.8 Key operating waveforms during mode transition (Mode 8 and
Mode 9): (a) power distribution of the system, (b) 400-V DC bus
voltage dynamics, (c) SOC of battery pack
Control of solid-state transformer-enabled DC microgrids 131
Tertiary control
Battery charge
and discharge
control at Secondary control
hours level
Voltage recover
second level
Primary control
Individual converter
Since PPV is equal or larger than zero, the power direction of battery is determined
by the difference between the PPV and PDC_Load. For example, the power of the
battery will be larger than zero when PDC_Load is larger than PPV, while the power
of the battery will be less than zero when PDC_Load is smaller than PPV.
Basically, the primary control algorithm is based on DC bus signal control, and
therefore, the distributed control can be achieved. In this control algorithm, the DC bus
voltage is no longer a fixed value, but varies in a certain range in the islanding mode.
The DC bus voltage range is defined as from 360 to 400 V in the presented system
design. In islanding mode, only the battery is used to regulate the DC bus voltage.
Droop control is adopted for battery control with the expression shown in (6.3):
V b ¼ V 0 R b Ib (6.3)
where Vb is the reference for the DC bus voltage, Vo is bus voltage value without
load and it is set to 380 V, Rb is the virtual output impedance, and Ib is battery
output current. For the PV module, it always operates in MPPT mode to deliver
power to the system. Noted here, the load management is involved in the primary
control. The load is divided into the critical load and the noncritical load. When the
PV has no output power to the system, the bus voltage will drop. The battery has to
source more power to the load and the battery SOC will decrease rapidly. To ensure
that the DC microgrid can operate as long as possible in islanding mode, the battery
SOC has to be considered. Therefore, to avoid battery SOC dropping fast, the
noncritical load will be shed from the system when the bus voltage decreases below
some preset threshold (370 V). When the bus voltage recovers to a certain value,
380 V for instance, the noncritical load can be reconnected to the system. The
primary control diagram is shown in Figure 6.10.
Battery control
v0 Vb jb
Droop Voltage Current
curve control control
Vbus ib
ib
PV control
MPPT mode
ipv
vref Voltage Current jPV
MPPT
control control
vpv
vbus
0 370 380
Secondary control
v0 Battery control
enable signal
VSST_dcout
Secondary Vb Current jb
Saturator Droop Voltage
control Vs curve control
control
ib
ib
Vbus
c
d V0 b
Vs < 0 Vs > 0
a
0 I
bus voltage, Vs is larger than zero and the droop curve is shifted upward (battery
operation point from a to b), and vice versa (battery operation point from c to d).
It is noticed that since the low bandwidth communication is adopted, the
secondary controller’s time step should agree with that of the communication to
prevent secondary controller saturation.
Secondary control v0
enable signal Battery control
VSST_dcout
Secondary Vb Voltage jb
Saturator Droop Current
control curve control control
Vs Vbus ib
Vbus
V V
g
V0 VBus_ref
V0
VBus_ref e
e
f
0 0 I
I
Figure 6.14 SST DC voltage shift in the tertiary control shown in Figure 6.13:
(a) SOC to higher boundary and (b) SOC to lower boundary
where kcp and kci are the tertiary control parameters. The bus voltage reference can
be as shown in (6.7):
VBus ref ¼ VSST dcref Vc (6.7)
As shown in Figure 6.14(a), for example, if Iref equals Id, the bus voltage operation
point will shift from point e to point f. Here, the bus voltage is less than the V0
for battery droop curve. Thus, the battery will switch to discharging mode auto-
matically. When SOC reaches LSOC, the SST DC output voltage will be higher than
V0 (point e to point g) and the battery will be charged [see Figure 6.14(b)].
Shifting battery droop curve
As in the previous tertiary control scheme, the current reference is determined by
the current reference selection block based on the SOC. If the SOC reaches the
HSOC, Iref equals Id. Or if the SOC reaches the LSOC, Iref equals Ic. In contrast to
the aforementioned tertiary control scheme, the battery droop curve is changed
instead of the SST DC bus voltage. Then the tertiary controller can be represented
as following:
V0
0 I
Figure 6.15 Droop curve shift with second tertiary control shown in Figure 6.13
Secondary control v0
enable signal Battery control
VSST_dcout
Secondary Vb jb
Saturator Droop Voltage Current
control curve control control
Vs
Vbus ib
Vbus Vt
I
Id I
SOC SOC ref Tertiary Saturator
LSOC HSOC control
Ic
ib Tertiary control
Current reference selection enable signal
Noncritical
load shedding
Noncritical
PV on load shedding
Noncritical
load on
power is larger than the load, the battery switches to the charging mode auto-
matically and the bus voltage is higher than 380 V. When the noncritical load
(1 kW) connects to the system, the bus voltage drops and the battery switches back
to the discharging mode. Since the bus voltage exceeds the 370 V (value for load
shedding), the noncritical load remains connected to the bus. After the PV dis-
connects from the system, the battery has to supply the power to the load alone and
the bus voltage decreases. After 0.5 s when the bus voltage drops below 370 V, the
noncritical load is shed from the system, and the bus voltage can recover. When
the PV is connected back to the system the bus voltage increases. After 0.5 s when
the bus voltage again exceeds 380 V (the value for noncritical load back), the
noncritical load is reconnected to the system automatically.
(b)
(c)
Figure 6.18 Waveforms with secondary control for the system in Figure 6.2:
(a) SST waveforms, (b) battery current, DC microgrid bus voltage,
and SST AC output, (c) PV waveform
Control of solid-state transformer-enabled DC microgrids 139
microgrid connects to the SST, and a seamless transfer is achieved. Furthermore, the
battery droop curve is lifted based on the secondary control loop as previously
described. Therefore, when the system is in SST-enabled mode, the battery still
outputs current to the load. For the PV module, there is no change in its output
voltage and current because it always operates in the MPPT mode as shown in
Figure 6.18(c).
(a)
(c)
Figure 6.19 Waveforms with first tertiary control principal: (a) SST waveform,
(b) battery current, DC microgrid bus voltage, and SST AC output,
(c) PV current and voltage
1) Input current 4 A/div
2) Input voltage 5 kV/div
3) VhDC 500 V/div Time 1 s/div
4) VlDC 500 V/div
(a)
Figure 6.20 Waveforms with second tertiary control: (a) SST waveform,
(b) battery current, DC microgrid bus voltage, and control flag,
(c) PV waveform
142 DC distribution systems and microgrids
by the SST output voltage range. When the charging and discharging current
references are large, the SST output DC voltage might reach its limitation and then
the battery’s output current cannot match with the references. Compared to the first
tertiary control method, the second tertiary control method does not have these two
problems since it just changes the battery’s droop curve. However, its disadvantage is
that communication ports need to be involved because the control output commands
need to be sent from the SST to the battery module, which will degrade the system
reliability. For the first tertiary method, the communication is exclusive since the
control method is only implemented on the SST side. Which method to select will
depend on the designer and the system requirements.
6.4.3 Summary
A hierarchical power management strategy for the SST-based DC microgrid is
proposed, which includes primary control, secondary control, and tertiary control.
The DC microgrid not only can operate more reliably in islanding mode by primary
control, but can also seamlessly transfer the DC microgrid from islanding mode to
SST-enabled mode via secondary control. The battery SOC is involved in the ter-
tiary control for battery management. Experimental results are presented to verify
the proposed schemes.
ωg, δg, vg
PCC
AC grid Line impedance
Pout, Q ω, δ, E
Q ω
Frequency
Volt-Var
regulation
regulation
vg vDC (first droop)
vDC_ref
E
SST
Pin vDC DC microgrid bus
Pbat PPV Pload
vDC
Battery Second
droop PV
Load
OR unit
Pbat_ref
SOC/
Battery unit economic
ωR mω – 1
+ Jvirs s
–
2E–d
Gpss(s)
Volt-Var +
QR mvg GLPF3(s) + E
regulation +
– +
Q GLPF2(s) Qf ER
Figure 6.22 Detailed control strategy diagram of the SST controlled as SSSM
144 DC distribution systems and microgrids
Pin vDC_R
+ –
– 1 vDC vDC_ f + 1
Pout GLPF1(s) w
CDCvDCs Jvir S + Kw
+
mwwR
a new balanced working point for Pin, Pout, and d is reached, while w synchronizes
with wg again. The microgrid’s total inertia is determined by CDC, VDC, wc1, and
Jvir. When CDC is designed normally, Jvir will be an extra control parameter to
adjust the inertia to meet the stability requirement in the power system.
According to the analysis above, any power from the DC microgrid can be
transmitted to the AC grid through the SST at a given AC frequency. The dynamics
of AC output active power is slowed down and completely decoupled from the
intermittent, variable DRERs and loads at the DC microgrid. So this microgrid’s
model is simple and promising for studying the scalability and stability of the
paralleled systems. Due to its swing characteristic, the microgrid is able to respond
to the short-term requirements of the grid frequency regulation.
vDC vDC
SSSM Battery
220 units
200
180
ω Pbat
0 wmin ωR ωmax –Pbat_max 0 Pbat_max
Figure 6.24 Detailed diagram of dual droop control for SSSM and battery units’
coordination
146 DC distribution systems and microgrids
where Prate is the converters’ rated power, and kw and kvDC are the droop slopes
for AC and DC systems. Combining (6.13) and (6.14), the first droop in the
SST can be rewritten as
kvDC
vDC vDC R ¼ ðw wR Þ ¼ mw ðw wR Þ (6.15)
kw
In (6.15), mw is determined by the allowable ranges of w and VDC. Typically,
1/kw is proportional to the SST’s rated power. The DC voltage ranges are
designed to be the same for different SSTs, then 1/kvDC is also proportional to
their rated power, and the resulting mw will be the same for different SSTs. For
the frequency regulation control in Figure 6.23, this droop control is naturally
integrated in it, and mw is also the coefficient for the part of the damping.
2. Second droop in energy storage units: The second droop in the energy storage
units is a traditional V–P droop, and mvDC is its droop slope. According to the
first droop in the SST, Pbat should be controlled based on VDC. This control is
shown in the right column in Figure 6.24. The hierarchical power management
discussed in Section 6.4 for battery is still applicable as an offset power
reference.
3. Coordination of dual droop: Figure 6.25 shows the coordination diagram of the
dual droop control. VDC is controlled according to w through the first droop.
This voltage is a command, which indicates the energy storage units how much
power SST requires corresponding to the second droop curve. Then energy
storage units will output the corresponding power through the second droop
and inject it into the AC grid via the SST. This is similar to the traditional
active power control in the SM, and finally, the SST with the DC microgrid
will have the w–P droop characteristic.
The dual droop control is distributed and autonomous without communication. In
steady state, the SSTs connected to the same AC grid will synchronize with it, share
and balance the power via the AC frequency. So the SST can provide up/down
reserve to the grid when wg varies in a desirable region, with the support from
energy storages. The power in the DC microgrid is also shared and balanced based
on the second droop among paralleled energy storage units. So the power man-
agement of the AC grid and DC microgrid are achieved simultaneously either in
PPV + Pload
+ Pin
Pbat + Power flow
Battery Pout
grid-connection or islanding modes. During the dynamics, the dual droop control
will also support the demanded inertia at a given wg. It is achieved by commanding
energy storages to provide temporary energy to the SST through VDC, when w and
VDC vary due to the disturbances from the AC grid or DC microgrid.
Q1 (V1,Q1)
Dead band
Capacitive
V1 100%
(nominal voltage)
% Available VARS (Q)
Percent
(V3,Q3) voltage
(V2,Q2) Increase
Inductive
V4
Q4 (V4,Q4)
Pout Pout
CH1: 0 W CH1: 0 W
Load reconnected
CH2: ω CH2: ω
120π rad/s 120π rad/s
vDC vDC
CH3: 200 V CH3: 200 V
Load disconnected
CH4: 0 A CH4: 0 A
ibat ibat
(a) (b)
Time (s) Time (s)
Pout Pout
CH1: 0 W CH1: 0 W
Load reconnected
CH2: ω CH2: ω
120π rad/s 120π rad/s
vDC vDC
CH3: 200 V CH3: 200 V
Load disconnected
CH4: 0 A CH4: 0 A
ibat ibat
Figure 6.27 Experimental results of Case I: (a) Jvir ¼ 1.06, 80-W load changes to
no load; (b) Jvir ¼ 1.06, no load changes to 80-W load; (c) Jvir ¼
4.24, 80-W load changes to no load; (d) Jvir ¼ 4.24, no load changes
to 80-W load. CH1: Pout (500 W/div); CH2: w (0.08 p rad/s/div);
CH3: vDC (10 V/div); CH4: ibat (2 A/div); X-axis: time t (400 ms/div)
Control of solid-state transformer-enabled DC microgrids 149
CH1: 0 W CH1: 0 W
Pout Pout
CH2: CH2:
120π rad/s ω 120π rad/s ω
vDC vDC
CH3: 200 V
CH3: 200 V
Source reconnected
CH4: 0 A CH4: 0 A
Source disconnected ibat ibat
CH1: 0 W CH1: 0 W
Pout Pout
CH2: CH2:
120π rad/s ω 120π rad/s ω
vDC vDC
CH3: 200 V CH3: 200 V
Source reconnected
CH4: 0 A CH4: 0 A
Source disconnected ibat ibat
Figure 6.28 Experimental results of Case II: (a) Jvir ¼ 1.06, 2.5-A source changes
to 0 A; (b) Jvir ¼ 1.06, 0-A source changes to 2.5 A; (c) Jvir ¼ 4.24,
2.5-A source changes to 0-A; (d) Jvir ¼ 4.24, 0-A source changes
to 2.5 A. CH1: Pout (500 W/div); CH2: w (0.08 p rad/s /div); CH3:
vDC (10 V/div); CH4: ibat (2 A/div); X-axis: time t (400 ms/div)
Pout
CH1: 0 W
CH2:
120π rad/s ω
vDC
CH3: 200 V
CH4: 0 A
ibat
Time (s)
Figure 6.29 Experimental results of Case III: CH1: Pout (500 W/div); CH2:
w (0.08 p rad/s /div); CH3: vDC (10 V/div); CH4: ibat (2 A/div);
X-axis: time t (20 s/div)
150 DC distribution systems and microgrids
CH4: 0 A CH4: 0 A
Soft-start ibat ibat
Figure 6.30 Experimental results of Case IV: (a) reconnecting with battery
unit and load and (b) reconnecting with battery unit and current
source. CH1: Pout (500 W/div); CH2: w (0.08 p rad/s /div); CH3:
VDC (10 V/div); CH4: ibat (2 A/div); X-axis: time t (400 ms/div)
it is the reason why sometimes w and VDC do not work at the rated values in
Figures 6.27 and 6.28. Since the proposed system can synchronize its own fre-
quency with wg with a high degree of accuracy, it can also work as a phasor
measurement unit in the power systems. In Figure 6.29, VDC and ibat vary and
follow w according to the dual droop control. This verifies the effectiveness and
functions of the dual droop control in the proposed concept, supplying the power
system with up/down reserve when the frequency deviates from the rated 60 Hz.
6.5.6 Summary
An SSSM concept is proposed to integrate the DC microgrid into legacy AC grid
through an SST. The benefits are (1) integrates the DERs, loads, and energy
storages in the DC microgrid into SSSM; (2) achieves power management for the
AC grid and DC microgrid, and enable the DERs and loads to output any power at a
given AC frequency; (3) introduces inertia, provides up-reserve and down-reserve
for the grid with support from practical energy storages, and decouples the AC
output from the fast responding DERs and loads. Experimental results verify the
effectiveness and the benefits of the presented method.
Control of solid-state transformer-enabled DC microgrids 151
6.6 Conclusion
The SST-enabled DC microgrid concept is introduced in this chapter. Compared to
the traditional microgrids, the presented system enables a compact and integrated
interface for the DC and AC grids. Centralized power management strategy, hier-
archical power management strategy, and SSSM concept are proposed to control
the operation of the presented system. Simulation and experimental results are
given to demonstrate the feasibility of the proposed system.
References
[1] X. She, A. Q. Huang, and R. Burgos, ‘‘Review of Solid-State Transformer
Technologies and Their Application in Power Distribution Systems,’’ IEEE
J. Emerg. Sel. Topics Power Electron., vol. 1, no. 3, pp. 186–198, September
2013.
[2] X. She, A. Q. Huang, S. Lukic, and M. Baran, ‘‘On Integration of Solid-State
Transformer with Zonal DC Microgrid,’’ IEEE Trans. Smart Grid, vol. 26,
no. 12, pp. 3778–3789, June 2012.
[3] J. M. Guerrero, J. C. Vasquez, J. Matas, L. Vicuna, and M. Castilla,
‘‘Hierarchical Control of Droop-Controlled ac and dc Microgrids – A General
Approach Toward Standardization,’’ IEEE Trans. Ind. Electron., vol. 58,
no. 1, pp. 158–172, May 2011.
[4] X. Yu, X. She, X. Ni, and A. Q. Huang, ‘‘System Integration and Hierarchical
Power Management Strategy for a Solid-State Transformer Interfaced
Microgrid System,’’ IEEE Trans. Power Electron., vol. 29, no. 8, pp.
4414–4425, August 2014.
[5] M. Ashabani, and Y. A.-R. I. Mohamed, ‘‘Novel Comprehensive Control
Framework for Incorporating VSCs to Smart Power Grids Using Bidirectional
Synchronous-VSC,’’ IEEE Trans. Power System, vol. 29, no. 2, pp. 805–814,
March 2014.
[6] Q. Zhong, and G. Weiss, ‘‘Synchronverters: Inverters that Mimic Synchro-
nous Generators,’’ IEEE Trans. Ind. Electron., vol. 58, no. 4, pp. 1259–1267,
April 2011.
[7] D. Chen, Y. Xu, and A. Q. Huang, ‘‘Integration of DC Microgrids as Virtual
Synchronous Machines into the AC Grid,’’ IEEE Trans. Ind. Electron.,
vol. 64, no. 9, 7455–7466, September 2017.
This page intentionally left blank
Chapter 7
The load as a controllable energy
asset in dc microgrids
Robert S. Balog1,2, Morcos Metry1,2,
and Mohammad Shadmand1,2
7.1 Introduction
The electric power grid is the backbone of modern societies and their economies by
interconnecting electrical energy generation to loads via a vast transmission and
distribution network. Heralded as the greatest engineering achievement of last
century [1], the power grid in the United States and other countries is deteriorating
due to age and usage stresses [2]. In fact, it experiences a major outage every
decade that costs more than $2 billion, and on any given day, 500,000 customers
are without power for 2 h or more [3]. The traditional grid was not designed with
the renewable energy sources in mind and the reverse power flows that occur and
thus does not easily accommodate the transformation from consumer to prosumer.
Simply patching the old system beyond its original design and safety margins will
not alleviate the problems of power delivery congestion, feeder system voltage
profile regulation, reactive power, harmonic power, and the other challenges of the
way in which the grid is now being used.
One solution that does not require a $5 trillion complete overhaul of the
existing $1.5–$2 trillion [4] US electricity grid is to separate the functions of bulk
power delivery from point-of-load (POL) control. In this paradigm, we interconnect
loads and distributed energy sources and storage into a local area power and energy
system (LAPES) [5], sometimes called a microgrid [6], which is then connected to
the utility power grid or to other microgrids [7]. The Department of Energy defines
a microgrid as ‘‘a group of interconnected loads and distributed energy resources
within clearly defined electrical boundaries that acts as a single controllable entity
with respect to the grid’’ [8]. A classical perspective of ac power systems is a top-
down, generator-to-load flow of power in which utilities control the generation and
infrastructure and are required by a regulatory body to deliver power to the load.
1
Department of Electrical and Computer Engineering Texas, A&M University, USA
2
Department of Electrical and Computer Engineering, Kansas State University, USA
154 DC distribution systems and microgrids
The utility has no control of the load, which imposes operational constraints on the
systems. In a LAPES, the load as well as distributed sources and storage can all be
controlled, which changes the paradigm of generation following load into a para-
digm in which control of the load itself can be thought of as a controllable energy
asset [9]. The paradigm further changes as the control moved from centralized to
distributed.
In the past, some utilities have had load-shed programs in which customers
relinquish some limited control of their load. Such a system was most commonly
used for peak load shaving in which customer-owned air-conditioning systems were
fitted with cut-off switches that responded to a command from the grid or system
operator, as explained by Commonwealth Edison Company of Illinois [10]. Thus,
shedding individual load to match available generation capabilities or congestion
constraints to avert a blackout is one way the load can be considered as an energy
asset [11]. Load shed has also been considered for ac systems where both voltage and
frequency stability were a concern and implemented with frequency-activated load
shedding [12]. Less aggressive approaches that have been considered include con-
servation voltage reduction in which the voltage on a feeder system is lowered to
reduce power consumption, which can work to some degree when the loads are
mainly resistive in nature [13]. Market-based incentive programs take a slightly
different approach and attempt to encourage customers to modify their power con-
sumption based on adjusting the price of electricity. In some cases, a smartphone app
is used to signal cost savings that may be possible by voluntarily reducing con-
sumption, or offer an incentive, such as a coupon, for volunteering to shed load
[14,15]. However, such initiatives depend on the active participation of the customer.
While it raises public awareness of energy consumption, it may not be reliable for
peak-time load shaving to avert a crisis if the customer does not behave as expected.
Fundamentally, the challenge in any electrical power system is that of just-in-
time delivery: supply must balance consumption. If the system is congested, or
generation is already at peak, one way to relax the constraint is to introduce energy
storage. Pumped hydro can be cost effective but depends on geography and thus is
not a universal solution, just like compressed air storage which depends on vast
underground storage. Battery-based storage can be more arbitrarily located in the
utility system as needed. Examples of grid-scale storage include the Younicos cube
or the Tesla power pack [16,17]. Another approach is to tap into a growing market of
residential-scale storage element such as the Tesla power wall located in individual
houses [18]. However, while adding storage can decouple supply from demand, it
also adds cost.
Electric
power grid
High-level Load and PV module
Wind PCC
controller
generators
(e.g., Energy PEI
management)
PEI
PV PEI Local area
modules power and
energy system PEI
Source-side
energy storage
Electric PEI
vehicles PEI PEI
Figure 7.1 One-line system diagram for a local area power electronic system
(LAPES) that integrates community-scale energy storage with
distributed solar generation colocated with the electrical load. The
system is interconnected with the electric utility at a single point of
common coupling (PCC). Sizing and control of the energy storage
unit, the ‘‘Central Energy Storage’’ can enable arbitrary power flow
profiles at the PCC to the grid
would not merely supply power to the load but would also interact with the
loads, sources, and storage. In such a paradigm, the operation of the load, dis-
tributed sources, and distributed storage can respond to the LAPES system to
ensure stable operation and even participate in energy market opportunities
upstream in the utility power grid by controlling power flow at the PCC. Without
loss of generality, examples of a LAPES include a neighborhood or subdivision of
houses, an industrial park, a shopping center, a school campus. In these examples,
the central energy storage shown in Figure 7.1 would be a common resource used
by the entire system, much like a community lake is a common resource used for
firefighting or a retention basin is used for flood control.
In this chapter, we consider a LAPES as a dc microgrid system. Virtually, all
modern loads are dc, and dc systems offer efficiency and control advantages over
ac for distribution networks and integration of renewable sources and distributed
storage [5]. The focus of this chapter is on control of the load so that the system can
view the load as an energy asset rather than a liability. In other words, view the load
as an integral participant in the overall power and energy balance of the microgrid.
This is done by exploring how POL controllers combined with modulation of the
bus voltage can control the power and energy consumption of the load in the system
without needing expensive communications.
156 DC distribution systems and microgrids
30%
20%
10%
91 91
30%
20%
10%
0%
0%
Reduction in
92 92 energy availability
due to failure in
93 93 portion of energy
Energy availability (%)
40
Load modulation (%)
30 D
10 B
0 A
0.50 0.55 0.60 0.65 0.70 0.75 0.80 0.85 0.90 0.95 1
(c) Capital expense (p.u.)
Figure 7.2 Pareto-frontier reveals the set of optimal solutions for the hybrid
system [22]. (a) Load control benefits to design: energy storage size/
cost reduction by moving from point A to D (modulating the load from
0% to 30% while maintaining the energy availability of the
microgrid), (b) load control benefit to control: portion of the energy
storage bank is failed, thus energy availability will be decreased by
moving from point A to B due to this failure; the energy availability for
the reduced available energy storage size can be increased by moving
from point B to B000 (load modulation from 0% to 30%), (c) load
control benefit to capital cost of the energy storage system
10% to 30%. This can be accomplished by sending a signal to each dc/dc or dc/ac POL
converter. Two scenarios are considered that reveal the benefit of load control
(modulation) in a microgrid system: one scenario encompasses the gained benefits
from proper storage and load modulation ratios of the microgrid systems, while the
other displays strategies for failures in portion of energy storage bank.
The first scenario is illustrated in Figure 7.2(a). This figure shows that if
point A is selected as the final optimal solution of the base-case based on particular
preferences to achieve 98% energy availability, the required energy storage will be
0.85 p.u. This configuration requires bulky and expensive energy storage to com-
pensate the stochastic behavior of renewable energy resources and/or to meet the
peak load at 0% load modulation (an uncontrolled load). The load modulation
could refer to the fraction of total load that is shed or the percentage of power
reduced in individual load for a period of time. For example, dimming a light or
changing the set point on a thermostat represents load power modulation that is a
different control paradigm than simply turning off the load or HVAC system which
may balance providing benefit to stabilize the grid while also minimizing impact to
the end-user. By modulating the load from 0% to 30% (i.e., moving from point A to
D), the required energy storage system requirement is reduced significantly from
0.85 to 0.55 p.u. as shown in Figure 7.2(a), while maintaining the same energy
availability levels. This scenario demonstrates the benefits of the load control to the
design and sizing for the microgrid system.
The second scenario is illustrated in Figure 7.2(b) where the system was
initially designed for operation at point A. Then, portion of the energy storage bank
is failed, causing reduction of energy availability from 98% to 97% (i.e., moving
from point A to B). The advantage of the load modulation comes into play in this
scenario; by modulating the load at 10%, the same energy availability of 98% could
be maintained with a failed storage element. Further energy availability improve-
ment could be achieved by modulating the load further. For example, at 30%
modulation, the energy availability reaches 99.99% (shown by moving from point B
to B000 ) at an energy storage size of 0.72 p.u. Thus, the function of load modulation, in
this example, increased the energy availability by 1.9% and decreased the energy
storage size by 0.13 p.u. in comparison with the originally designed system that did
not have controllable load considered initially. Thus, the load control can improve
the system controllability for maximizing the energy availability as well as com-
pensating the consequences of energy storage device failures in a typical microgrid
system.
It is worth mentioning that moving from point A to D in Figure 7.2(a) is a
capital expense saving at the design stage because less energy storage is required.
The effect of load modulation on the capital cost of the energy storage system is
illustrated in Figure 7.2(c). As is illustrated, for example, if lights are dimmed to
modulate the load to 30%, the capital cost of the system will be reduced to 0.55 p.u.
at point D. However, if the loads are not modulated, the capital cost of the system is
0.85 p.u. at point A. The capital cost of 0.85 p.u. demonstrates the cost of an
optimal system design at 0% load modulation, and the capital cost of 0.55 p.u.
demonstrates the cost of an optimal system design at 30% load modulation. This
The load as a controllable energy asset in dc microgrids 159
demonstrates the impact of load modulation on the end user in terms of pro-
ductivity, capability, and economic model because the end user would expect
financial compensation for the reduced performance or reduced capability.
2016 April
Inertia provided by Nuclear Inertia provided by Coal Inertia provided by Simple Cycle Inertia provided by Combined Cycle Load
300,000 60,000
250,000 50,000
200,000 40,000
Inertia (MW*s)
Load (MW)
150,000 30,000
100,000 20,000
50,000 10,000
0 0
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
-0 01 00
16 0
00
20 :0
03 4-2 6 0:
04 4-2 6 0:
05 4-2 6 0:
06 4-2 6 0:
07 4-2 6 0:
08 4-2 6 0:
09 4-2 6 0:
10 4-2 6 0:
11 4-2 6 0:
12 4-2 6 0:
13 4-2 6 0:
14 4-2 6 0:
15 4-2 6 0:
16 4-2 6 0:
17 4-2 6 0:
18 4-2 6 0:
19 4-2 6 0:
20 4-2 6 0:
21 4-2 6 0:
22 4-2 6 0:
23 4-2 6 0:
24 4-2 6 0:
25 4-2 6 0:
26 4-2 6 0:
27 4-2 6 0:
28 4-2 6 0:
29 4-2 6 0:
30 4-2 6 0:
0:
4- 6 0
-0 01
02 4-2
-0
01
Figure 7.3 Example of generation inertia and load consumption data available
from ERCOT for the state of Texas. The figure shows the daily cycle
which changes over the month of April in 2014 [24]
160 DC distribution systems and microgrids
Consider the peak-time load consumption which occurs during daylight hours.
Since lighting loads represent 30% of the electrical consumption in commercial
buildings, modern buildings are designed with controllable lighting systems for
energy conservation and adjust artificial lighting based on the amount of sunlight
entering the building [25], occupancy, and usage requirements. More windows
unshaded means a lower illuminosity foot-candles (ftcd) is required from the artifi-
cial lighting system. Standards recommend office spaces to have an illuminosity of
46 ftcd [26]. Hence, sunlight exposure from the office windows could reduce light
consumption. While this might seem like a good idea in principle, tenants of the
office complain about sudden switch off of lighting during direct sunlight, delayed
turn on of the lighting when the room is darker, and a continuous switching on and
off of lights on cloudy days [27]. Such concerns mean that in many workspaces,
these ‘‘intelligent’’ systems are bypassed; hence, lights remain at their full illu-
minosity the whole day, impeding the energy saving opportunity and missing the
opportunity to control the load as an energy source. As many lightbulbs come now
with dimming features, lighting modulation could be achieved by adjusting the
brightness rather than turning on or off certain light fixtures. This is achieved by
giving some load modulation control to the grid. After all, a user can barely distin-
guish the difference between a 46-ftcd and a 35-ftcd room [28] yet when taken at the
scale of an entire building over a period of time can provide enough energy to have a
meaningful impact on the stability and energy availability of the microgrid. Hence,
lighting in commercial buildings is well suited as a controllable energy asset.
Consider a building heating, ventilation, and air conditioning system (HVAC).
A building has a large thermal mass so adjusting blower speeds or temperature set
points on a short enough time bases may have minimal impact on the occupancies
perception but frees up energy. Typical HVAC systems have a control approach
similar in principle to the bang-bang controller, in that there is a 2 C upper and lower
limits around the set point and the HVAC compressor and fan will work when the
temperature exceeds the upper limit and turns off when the temperature goes under the
lower limit. Such systems could be thought of as a manageable load asset, in that
the occupants of a room would barely be able to tell if the temperature went up to 3 C
above the upper limit during times of contingencies and would not notice the lower fan
speed intensity. Such initiative takes place in the middle of the day, during peak
consumption time, assuming the HVAC system has been on all day and has already
cooled the space. Many companies have considered demand side energy management
in which the HVAC unit could be controlled by the grid. Edison Commonwealth of
Illinois [10] has the initiative of Central AC Cycling for homeowners, which offers the
installation of direct load control switches or utilizing a smart thermostat to cycle the
load during peak time by turning off the compressor when needed, up to 15 min every
30-min interval. Applying such system on large commercial and government build-
ings such as schools and offices could add a flexibility of manageable load assets.
Consider electric vehicles (EV) or plug in hybrids (PHEV): if all the owners
care about is having a fully charged battery ready for the next trip, he or she could
set the expected time the car will next be needed by using a smart phone app [29]
when parking the car. From the driver’s perspective, the rate at which energy is
The load as a controllable energy asset in dc microgrids 161
delivered to the car battery is meaningless as long the battery is fully charged and
ready for service when the user returns [30]. Thus, EVs become a controllable
energy asset by scheduling unnecessary EV charging off peak times. In 20 years or
so, when EV become the mainstream technology, utilizing such load could be
significant. For example, a parking garage next to a hospital with a 1,000 cars
capacity could have enough energy storage, using vehicle-to-grid protocols [31],
for one of the hospital’s building segments (e.g., operation rooms or intensive care
units) during emergency power deficiencies.
Numerous application examples can be envisioned, ranging from civilian to
military, terrestrial to outer space, to personal power networks, in which load can
participate in the overall control and operation of power and energy management of
the dc microgrid system. It may be easy to visualize the imperative for load control
if the microgrid is stand-alone and energy constrained. The need also exists, how-
ever, for grid-connected dc microgrids, like the one in Figure 7.1, in which the
operator desires to shape the load profile at the PCC for reasons of economics or
utility-side stability. For example, the microgrid can take advantage of market
clearing prices to control local energy assets [32]. This may be obvious for the
cycling of energy stored in battery, but we can also modulate a load to achieve
similar effects. Therefore, there are many considerations to be taken into account
during the design of the microgrid. During the design stage for the energy storage,
sizing of the storage is determined by assuming fully modulated load to size the
storage based on worst case availably requirement. During operation, energy sto-
rage must be first used as primary energy asset and then the load modulation is
secondary. Energy tapping opportunities from the load should be categorized in a
way that ensures the minimum cost. For example, tapping into EV batteries could
potentially result in higher battery maintenance cost than reducing the fan speed of
an HVAC system which could only cost a little convenience.
Automatic
generator control
of grid Diurnal load
and
PV curves kW h/day
for PV
Economic
Electrical stability dispatch
of system of grid HVDC
load
Power Energy
management management
Power buffer
Zbus
Point
+ + of
Vbus Rin – Load
– load
converter
Estorage
the load. After the transient passes, the buffer returns to a power regulation mode
and draws additional incremental power to recharge the buffer capacitor. In effect,
a power buffer stretches the time scale of the transient, diminishing the impact of
tight converter regulation.
A power buffer is limited by the amount of stored energy available in the bus
capacitor. The time that a power buffer can maintain constant input impedance
while continuing to supply full power to the load is called its sustaining time and is
defined as
2 2
Cbuffer Vload0 Vload1
2
Vbus0
Tsustain ¼ 2 (7.1)
2Plaod Vbus0 Vbus1
2
where Vload0 is the nominal load voltage, Vbus0 is the nominal input voltage, Vbus1 is
the sag voltage, and Pload1 is the load power. The buffer design parameters in the
sustaining time are Vload1 (minimum allowable load voltage) and Cbuffer (energy
storage capacity).
A plot of a typical sustaining time versus voltage sag for a buffer supplying a
100 W, 400 V load from a 100 V distribution system is shown in Figure 7.6. As
long as the voltage sag magnitude and duration fall above the curve, the power
buffer can successfully ride through the transient, presenting constant impedance to
the power system bus while maintaining constant power to the load. If the transient
event begins to approach the sustaining time limit, then the local load control needs
to switch strategies to maintain stability.
1
Ride through capability
0.8
Bus voltage [p.u.]
0.6
Insufficient energy storage
0.4
220 µF
0.2 470 µF
1,000 µF
0
10–2 10–1 100 101 102
Time (s)
7.4.1 Control
The generally adopted control hierarchy of the dc microgrid is of close resemblance
to that of the conventional grid, in that it is stratified into tertiary, secondary, and
primary control levels [36]. The highest level controller, the tertiary level,
responsible for economic dispatch and coordination, assigns the microgrid voltage
that is appropriate for power exchange between the microgrid and the main grid
[37]. Such function is communicated to the PCC, as shown in Figure 7.1. The
secondary controller updates the voltage set points for the primary controllers to
match the voltage demands of the secondary controller. Like the tertiary controller,
the secondary controller is carried out at the high-level controller. Then, the signals
The load as a controllable energy asset in dc microgrids 165
from the secondary controller are sent to all the power electronics interface (PEI)
units as illustrated in Figure 7.1.
Implemented locally at each PEI, the primary controller regulates the voltages
for the individual converters to facilitate load sharing among sources. The control
structure of the primary controller based on distributed control [35] is inherently
different from the secondary and tertiary controllers that are based on centralized
control. Accordingly, secondary and tertiary controller units are prone to higher
risks of failure and instability in the event of losing any links in the topology [34].
On the other hand, distributed control provides higher reliability as each PEI unit
controller is independent in operation from failures in other units. It also means that
systems based on distributed control are easily scalable [38]. A detailed illustration
on the differences between centralized and distributed control is illustrated when
studying the difference between active current sharing (centralized control) and
droop control (distributed control) in Table 7.2.
7.4.2 Architecture
As has been shown in Figure 7.1, a LAPES microgrid system incorporates renew-
able energy generation systems such as solar photovoltaics and wind along with
storage elements and EV. A LAPES microgrid is not merely a supplier of load to its
load but goes beyond by using residential scale generation on the load when
available. Without loss of generality, the loads could be residential houses, office
buildings, hospitals, university campuses, etc.; within each building, there could be
several distribution and load systems as illustrated in Table 7.1.
A high-level energy management system provides signals to the buildings on
power and energy usage/setting, and then the PEI for the building turns that setting
into a control response. A case scenario could be to balance generation to meet the
load; however, there is a point when generator control is no longer possible to
supply loads from generation alone. Hence, the need to tap into the energy
resources of the load itself. Consider a case study of a residential house with an ac
power system and that has different types of load as shown in Figure 7.7. One could
observe the energy storage potential in some of the loads (i.e., the UPS of a com-
puter system) and utilize that for the overall system stability. Moreover, a load
like residential lighting may be modulated/dimmed to reduce load demands while
keeping the lights to a comfortable illumination level. Such control actions may
Fluorescent
Micro M lighting
turbine
Consumer
AC
ac electronics AC
ac AC
ac
Water DC
dc HVAC DC
dc DC
dc
Electronic
heater motor ballast
DC
dc DC
dc drive DC
dc DC
dc
AC
ac AC
ac AC
ac AC
ac
60 Hz
ac
AC
ac AC
ac AC
ac AC
ac
DC
dc DC
dc DC
dc DC
dc
PCC
UPS
AC DC
dc
grid dc
DC
Solar Solid-state
Electronics
lighting
dc
DC
AC
ac
AC
ac
DC
dc
Computer
be achieved using a configuration like that of the single bus radial system of Fig-
ure 7.8 connected to the LAPES microgrid of Figure 7.1. The details of the control
methods are explained in the following subsections.
+ Input
bus
V dc–dc
– bus
R
filter
us
POL converter #3
Lb
+ Input
bus
V dc–dc
– bus
R
filter
us
POL converter #2
Lb
+ Input
bus
V dc–dc
us
R – bus filter
POL converter #1
Lb
Dc source +
converter –
Figure 7.8 Single bus radial system with three loads [39]
Normal operation
High Low
Load priority
Medium
Supply the load until the Supply the load until the Monitor the bus to see if the
local energy reserve is local energy reserve is load can be turned back on
depleted depleted or the bus
continues to deteriorate
Startup
During normal operation, the buffer transfers the load power demand from the
bus, while maintaining its internal energy storage. This mode completely couples
the load and bus dynamics. When the bus experiences a transient that triggers a
protection event, load priority determines the course of action to maintain system
168 DC distribution systems and microgrids
stability. The highest priority loads remain connected to the bus until the energy is
depleted. Lesser priority loads continue to monitor the bus and disconnect if the bus
is sensed to become worse according to some metric.
For high-priority buffer loads, the wellbeing of the load is favored; therefore,
the buffer will attempt to maintain the load until its internal energy storage is
depleted, while presenting a constant input impedance to the bus. When the internal
energy reaches a given set-point, the POL switches to a load-shed strategy. For a
medium-priority buffer load, the loads welfare is favored less. Throughout a pro-
tection event, the input impedance remains constant, while stored energy supple-
ments the load power. However, if during the protection event, the bus condition
becomes worse, the strategy is switched to a load shed.
When the bus recovers from a transient, the high and medium priority buffers
change to replenish energy storage mode in anticipation of the next event. While
drawing full load power from the bus to supply the needs of the load, additional power
is drawn to recharge the buffer energy storage capacitor. If during the replenish cycle
the bus experiences another transient, then the buffer reenters a constant input impe-
dance mode. Since the replenish cycle was interrupted, the sustaining time for the
latest transient will be diminished as there is less stored energy. If the replenish cycle
finishes uninterrupted by a bus transient, then the buffer returns to normal operation.
When the load is low priority, or the buffer energy storage has been depleted in
a high or medium buffer, load shed is implemented. This entails shutting down the
load in a manner that will cause the least inconvenience for startup and minimize
the impact on the load. The strategy is a function of the nature of the load.
When the protection event has cleared, a load-appropriate startup strategy is
implemented. To minimize the chance of triggering further system transients, the bus
power to the load should be minimized during startup, so-called soft-start. When the
load is high priority, power is immediately delivered to the load. When a low or
medium-priority load is started, the buffer energy storage capacitor is precharged
before turning on the load. During startup, load priority determines if the load is
immediately connected or if the power buffer starts up first. Highest priority loads
immediately connect to the bus, while lower priority loads first allow their power
buffer to soft-start, eliminating inrush current and gracefully loading the bus with an
initial constant impedance to help stabilize the system. After a time, the power buffer
will revert to normal operation. If a bus transient occurs during charging of the
buffer, then the charging cycle ends until the bus returns to its nominal state.
when significant, could result in overload and thermal stresses that jeopardize the
system reliability [40]. The common methods of current sharing used in the literature
fall into two general groups: active sharing and droop control [41].
Active current-sharing techniques involve a control structure and a method of
programming individual converters with a reference current. One implementation is
to use a master/slave configuration such that one dc source is designated as the
master and is used to control the bus voltage. The remaining dc sources, designated
as slaves, operate as current sources. This strategy produces a stiff bus voltage and
controlled load dispatch at each source. There are two main limitations of this
technique: high-speed communication is required and a single point failure can
disable the entire system. In practice, active current sharing techniques are best
suited for physically small systems, such as paralleled voltage regulator module
applications. If the topology were fixed and known a priori, more sophisticated
controls such as interleaving can be used to reduce ripple. In droop control [42], the
output voltage of the source drops as current increases. This is a form of local control
since converters autonomously share load current by sensing the local bus voltage.
Droop control can be as simple as a series resistance or a more efficient closed-loop
controller such as a phase-angle controller in a rectifier source converter [32]. This
scheme has been proposed for use in large-scale distributed systems with dynami-
cally changing topologies since it supports plug-and-play reconfiguration and system
scaling and is robust to component failures. A summary of the differences between
active current sharing and droop control techniques is as shown in Table 7.2.
Table 7.2 Comparison between load sharing techniques: active current sharing
and droop control
sources, loads, and storage devices. In fact, switching dc power converters is inher-
ently nonlinear, that interconnecting a large number of them poses stability issues to
the whole microgrid. POL converters act as instantaneous CPLs, as they regulate the
load voltage too tightly [43]. The tight POL voltage regulation results in negative
dynamic input resistance which leads to destabilizing effects during a voltage sag.
CPLs have become very common in various applications related to microgrid inte-
gration such as in data centers, telecommunication, wireless communication base
stations. Hence, the need for an effective solution to the destabilizing effect during
voltage sags, which are quite common due to the events of loss of generation or
increase in load. Since finite dc systems are inherently weak, such CPL systems are
subject to extreme voltage sags or even voltage collapse.
Many solutions are proposed in the literature [43,44], which include the addition
of filters, the addition of bulk energy storage devices that are directly connected to
the system main bus, load shedding, the addition of linear controllers and the addi-
tion of boundary controllers. This section reviews the benefits of droop control and
load modulation to remedy the destabilizing effects that could accompany CPLs.
1
Zbus + Zload
Is,ref Is –
Converter 1
Vref + K + Vs
– Gconv(s) sCS
ωLP
s + ωLP
V V V V V
Voc
Vop1
Vop2
Vop3
I I I I I
Iop1 Iop2 Iop3 Imax Iop1 Iop2 Iop3 Imax Iop1 Iop2 Iop3 Imax Iop1 Iop2 Iop3 Imax Iop1 Iop2 Iop3 Imax
Number of sources
5 4 2
Vop (V) 47.75 46.54 38.97
Iop (A) 5.672 7.273 17.38
Ibus (A) 28.36 29.09 34.75
Pload (W) 1,354 1,354 1,354
the available converters without the need for a central controller to redispatch the
source converters. If a converter turns off or fails, the remaining converters sense a
decrease in bus voltage and increase their respective output current to compensate
for the lost source.
Consider a system with five sources on a common dc bus supplying 1,354 W of
total load. Each dc source converter has a load-line that describes the v–i terminal
characteristics, as shown in Figure 7.11. Assuming negligible bus impedance
between the five converters, the solution to the base case (where all converters are
operational) results in the bus voltage Vop1 with each converter supplying Iop1
current. The analytical solution for the operating point is found by solving the load-
flow equations for n source converters and m constant-power loads (Table 7.3):
1
Voc;n In ¼ Vbus ; 8n
Kn
X X Pm (7.3)
In ¼
n m
Vbus
It is observed that the droop gain is the slope of the v–i curve and modifies the
actual source impedance. Thus, a simple model for a source converter under droop
control is a constant voltage behind impedance:
1
Vs ¼ Voc is (7.4)
K
172 DC distribution systems and microgrids
Although droop control can be as simple as a series resistance, a more energy effi-
cient choice is a closed-loop controller such as a phase-angle controller in a rectifier
source converter. For an arbitrary source converter, the permissible droop resistance
is lower bounded by the actual source resistance of the converter:
Rs Rdroop (7.6)
In the previous example, the source converters are assumed to be identical with
identical droop characteristics. Thus, the total load current is shared equally.
In general, however, each converter can have an arbitrary droop characteristic
representing its operating parameters, power limits, or preferred dispatch:
1
Rdroop ¼ Rs ; where 0 l < 1 (7.7)
1l
Thus, droop control programs the effective output impedance of the source-converter
and has been shown to result in current sharing. It also has direct implications to
system dynamic behavior.
This analysis suggests that droop control can be used to dynamically stabilize a
system [45]. Consider the three-bus system from Figure 7.8. The system is simu-
lated in DYMOLA using linearized averaged-model buck converters with damped
input filters. The source is modeled as an ideal voltage. The system is stable,
albeit with very light damping, z ¼ 107 106, as shown by a transient response in
Figure 7.12. The system eventually reaches steady state with each load-converter
supplying 144 W of load (about 3 A of current from the bus).
The system has reached steady state prior to 5 s when the load on bus 1
experiences a step-change in output power to 576 W. Figures 7.13 and 7.14 show
that the increase in current at bus 1 depresses the voltages at each bus in the system
51
50.5
Voltage [V] 50
49.5
49
48.5
48
Figure 7.12 Initial system transient response for a lightly damped system
30
25
Bus 1
20
Current [A]
15
10 Bus 1
Bus 2 Bus 2
5
Bus 3 Bus 3
0
4.8 4.9 5 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 6
Time [s]
Figure 7.13 Bus current of each POL converter. At 5 s, the load increases at
bus 1. The previously stable system becomes unstable with growing
oscillations. At 5.5 s, the droop resistance at the source converter
is adjusted to restabilize the system
174 DC distribution systems and microgrids
70
60
Bus1 [V]
50
40
30
70
60
Bus2 [V]
50
40
30
70
60
Bus3 [V]
50
40
30
4.8 4.9 5 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 6
Time [s]
Figure 7.14 Bus voltage at each POL converter. At 5 s, the load increases at
bus 1. The previously stable system becomes unstable with growing
oscillations. At 5.5 s, the droop resistance at the source converter
is adjusted to restabilize the system
and that the previously stable system becomes unstable with growing oscillations.
At 5.5 s, droop control is enabled on the source converter. The droop control results
in an equivalent resistance of 0.10 W. The droop control increases the effective
system damping to z ¼ 5.64 103 and the unstable system is stabilized.
A potential drawback to droop control is that there is a stationary error intro-
duced in the bus voltage as seen in Figure 7.14. In the literature, a second control
loop with low-pass filtering and PI control is proposed to increase the internal
open-circuit source voltage VOC and compensate for this stationary error. Since in
this example, droop control is only used to stabilize the system, not share load, the
stationary error can be minimized by adaptively programming the minimum droop
necessary to achieve the desired damping.
The drop in bus voltage, however, carries useful information about the health
of the system. The drop can be due to a partial loss of generation and indicate that
The load as a controllable energy asset in dc microgrids 175
the system may be energy constrained, as shown in Figure 7.11, or the drop can be a
result of needing increased damping which would indicate that the system is
nearing a region of instability as shown in Figure 7.14. Load-side controls that
sense this bus voltage drop will not know why the drop occurred but can take
mitigating action such as switching supply buses, activating a power buffer, or
invoking load interruption to alleviate system stress.
Autonomous local control performs dynamic load interruption based on the sensed
bus voltage at the POL converter and the priority of that load. Each of the loads in
the radial system shown in Figure 7.8 is assigned a priority based on the explana-
tion of Section 7.4.3. Table 7.4 gives an example of the priority assignment that
will be used in this section. By selecting the highest priority load to be the one
furthest from the source, the example will demonstrate that priority need not be
constrained by the topology.
The control strategy in (7.10) is to turn off a particular load when the sensed
bus voltage drops below a lower threshold and turn the converter back on when the
sensed voltage exceeds an upper threshold. The control is complicated when the on
and off control action of the converter causes its input voltage to drop below lower
threshold during turn-on and to rise above the upper threshold during turn-off.
Hysteresis is useful to provide a dead band to prevent this type of chattering. In a
larger system, however, the possibility arises of not having a large enough hyster-
esis band. In this situation, additional information and controls are needed to pre-
vent the chattering (Figure 7.15).
One method to obtain these set points is to perform an exhaustive contingency load
flow analysis. Each contingency represents a particular combination of operating
points for the system POL converters. For a small system, this offline process is
straightforward and computationally fast. The resulting set of power flow solutions
is processed by a search algorithm to find the upper and lower voltage limits for
each load converter.
Converter Priority
POL 1 Semiessential (SE)
POL 2 Nonessential (NE)
POL 3 Essential (E)
176 DC distribution systems and microgrids
49
48
47
46
Vbus
45
44
43
42
Load shed
1
POL 1 0
1
POL 2 0
1
POL 3 0
1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4
Time [s]
I1 P1 I2 P2 I3 P3
Vs +
−
POLC1 POLC2 POLC3
L1 L2 L3
Figure 7.16 Model of a radial dc power system with a single source and three
point-of-load converters
Start
Compute efficiency
for each converter
Updated bus
voltages
No
Bus voltages
converged?
Yes
End
good line and load regulation to ensure constant output power given a fixed resis-
tive load. Input power to the converter depends on the efficiency of the converter,
which is a function of the input voltage and the output power. Thus, the power flow
equations are complicated by the nonlinear efficiency of the dc-to-dc converters.
To improve the power flow results, the efficiency of the converters was
included in the power flow algorithm, shown in Figure 7.17. This added an outer
iteration loop to the power flow algorithm although it could be incorporated
directly into a power flow routine to increase computation efficiency. The power
flow algorithm begins by assuming that there is no voltage drop in the system and
that the voltage at each node on the bus is identical to the open-circuit source
voltage. These node voltages are then used to calculate the initial efficiency for
each dc–dc converter and hence the initial power withdrawn from each node on the
bus. The PST power flow program is then run. The result of the power flow, the
new system node voltages, is used to update the efficiency of each converter, and
the process is repeated until node voltages converge to within an acceptable bound.
178 DC distribution systems and microgrids
N jstatej (7.15)
The three-converter example in Figure 7.16 will therefore require 27 power flow
computations. The parameter limit identifies the type of voltage limit. It takes the
value
limit 2 fmin; max; UVPg (7.16)
The minimum limit is the lowest allowed input voltage before the converter turns
off, maximum is the highest input voltage before a converter turns on, and UVP
refers to the hardware UVP limit designed to self-protect.
The results of the exhaustive contingency analysis for the system in Figure 7.16
are graphically presented in Figure 7.18. The results of the 27 power flows are shown
for each of the three POL converters. Horizontal lines indicate the open circuit supply
voltage, VS, and the UVP limit, VUVP, where the converter turns off to self-protect.
Since POL3 is furthest from the source, it has the lowest bus voltages for each con-
tingency. Solid bars indicate contingencies of interest where one or more converters in
that contingency would trip off-line due to UVP.
23
VS 22
21
20
19
18
Input voltage
17
16
15
14
VUVP 13
12
11
10
9
1 POL1 27 1 POL2 27 1 POL3 27
27 contingencies at each point-of-load converter
Figure 7.18 Results of exhaustive contingency analysis on the radial test system
with a single source and three point-of-load converters. The load on
each converter can be either off, base load, or overload. Solid bars
indicate contingencies where a bus voltage in the system is below the
UVP for that converter
analysis. The process starts with the lowest priority converter, POL2. The lower
bound on the lower-limit is
8 9
>
> V ðNEÞUVP ; >
>
< =
on
on
V ðNEÞmin > max V ðNEÞ ; V ðEÞ V ðEÞUVP (7.17)
>
> >
>
: on ;
V ðNEÞ ; V ðSEÞ V ðSEÞUVP
The algorithm finds the voltage POL2 for all contingencies where the higher
priority loads are on and their input voltage is below the UVP limit. In this exam-
ple, this occurs for five contingencies at POL3, shown as the dark bars in
Figure 7.18. The algorithm also considers the possibility that the voltage at POL2
can fall below the UVP limit.
The upper-limit has both an upper and lower bound. Only, the essential and
nonessential loads are considered here for simplicity.
( )
high off off V ðEÞoff ;
V ðEÞ ; V ðNEÞ < V ðNEÞmax < min (7.18)
V ðEÞbase ; V ðNEÞoff
The lower bound for restoring the nonessential load occurs when the essential load
is at the maximum power and the nonessential load is off. The upper bound is found
180 DC distribution systems and microgrids
In a power system in steady state, the supply is matched to the load and the system is
stable. However, many dc distribution systems are electrically weak and do not have
the spinning reserves or other stability mechanisms. The bus voltage can sag for many
reasons such as partial loss of generation, increase in load, or topological reconfi-
guration. Further, tight voltage regulation in dc–dc converters makes them operate as
CPL, as demonstrated in Section 7.5.1, which draw increasing current for decreasing
bus voltage, possibly leading to further voltage sag or even voltage collapse [43].
Demand-side management is a suite of techniques that control the loads so that
they become integral components in system stability. Interruptible load is one
method that provides curtailment of demand to promote system security. Autono-
mous local control is investigated to perform this load-side control and improve
system reliability.
The P–V curve is a useful tool to visualize the operation of a power system.
Figure 7.19 illustrates a family of the familiar P–V system curve. Maximum power
transmission occurs at the nose were the source impedance and load impedance are
equal. In a dc system, the bus voltage drops as the load increases due to voltage-
divider action of the source impedance and the load impedance.
A system is initially in steady state with voltage V(t1) delivering total load
power of P(t1). The system impedance suddenly increases, perhaps due to a partial
loss of generation or topological reconfiguration, and the operating point moves to
a new P–V curve at time t2. However, the voltage V(t2) is below the under-voltage
Vbus
Voc
t6 t3 t1
VUVP
t5
t2
MPT
Figure 7.19 P–V Curve showing operating points as the system impedance
increases and loads are interrupted
The load as a controllable energy asset in dc microgrids 181
V(t3)
V(t6)
V(t1)
VUVP
V(t5)
V(t2)
(a)
Pload(t1)
Pload(t3)
Pload(t6)
t1 t2 t3 t4 t5 t6
(b)
Figure 7.20 Ideal bus voltage and load power as system impedance increases and
loads are interrupted to prevent voltage collapse. (a) Bus voltage
decreases in response to increased system impedance at t1 to reach
the operating point on the new P–V curve at t2. The new bus voltage
is below the UVP limit, so control action causes load to be shed,
moving to a mew operating point on the same P–V curve at t3 with a
higher bus voltage. The cycle repeats at t4. (b) Load power in the
system changes as point-of-load converters are turned-off to reduce
total system load when the bus voltage drops below the UVP
limit and load is shed, moving to a new operating point on the same P–V curve at t3.
The time-domain waveforms in Figure 7.20 reveal that these changes in operating
points do not occur instantaneously. The trajectories on the two figures, however,
are idealized to improve clarity of the system response and do not include the
dynamics associated with the inductance of the bus, the input filter, and the con-
stant-power dc–dc converters.
Load type
Status
Hotel (domestic) Lifeboats Propulsion Weaponry
In port Essential Nonessential Nonessential Nonessential
Patrol Semiessential Semiessential Essential Semiessential
General quarters Nonessential Nonessential Essential Essential
Abandon ship Nonessential Essential Nonessential Essential
on the operation of the system. In a naval ship, each load is classified as either
nonessential, semiessential, or essential. A general framework for organizing the
load priorities is a two-dimensional matrix as shown in Table 7.5. In one example,
propulsion is considered the highest priority, while hotel loads such as lighting in
crew quarters and power in the galley are less important and can be sacrificed
depending on the threat level. In another example, the launch equipment for life-
boats has a higher priority under ‘‘Patrol’’ status for safety reasons but yields
priority to other systems such as weapons and propulsion under ‘‘General Quar-
ters’’ status. If local energy is available in a power buffer, load priority is useful to
determine if the load simply turns off or the power buffer operates when trouble is
sensed on the dc bus.
Fine-tuning the performance of the system requires that each POL converter
has some information about the entire system. Low-bandwidth signaling from the
command and control center can broadcast the current state of the system, but each
POL converter ultimately decides how to use the information—unlike in a cen-
tralized control scheme where each load is directly controlled. The distributed
control strategy is inherently fault-tolerant because each controller acts indepen-
dently. If the low-bandwidth communication is compromised, each controller can
continue to operate using the last-known state or revert to fail-safe operation as
determined by the load-priority table.
7.9 Summary
In a traditional power grid system, the operator had total control of generation and
distribution assets while the load was viewed as a disturbance. Thus, planning
and operation necessitated always being prepared for unforeseen changes in the load
consumption. The result is that the US power grid is amazingly resilient, robust, and
expensive. As we consider a new paradigm of dc microgrid systems, overcapacity
may not be feasible for technological and/or economic reason. Yet, high power
quality and availability is more important than ever particularly to support the digital
economy and information age. This is compounded as renewable sources become
increasingly utilized, and the system operator no longer has total and arbitrary
control of the generation.
The load as a controllable energy asset in dc microgrids 183
Acknowledgments
The authors wish to acknowledge the various funding sources which supported the
research behind this content of this chapter. Dr. Balog acknowledges support
received from the Qatar Foundation. This publication was made possible by NPRP
grant no. 9-204-2-103 from the Qatar National Research Fund (a member of Qatar
Foundation). The statements made herein are solely the responsibility of the
authors. This work was also supported in part by the Grainger Center for Electric
Machinery and Electromechanics at the University of Illinois. Dr. Mohamad
Shadmand acknowledges support from the Qatar Foundation. This publication was
made possible by NPRP-EP grant no. X-033-2-007 from the Qatar National
Research Fund (a member of Qatar Foundation). The statements made herein are
solely the responsibility of the authors. Mr. Morcos Metry acknowledges the
sponsorship of the Qatar Research Leadership Program (QRLP) under the Qatar
National Research Fund.
184 DC distribution systems and microgrids
References
[1] N. A. Armstrong, ‘‘The Engineered Century: A Century Hence, 2000 May
Be Viewed as Quite a Primitive Period in Human History. It’s Something to
Hope For,’’ The Bridge, vol. 30, no. 1, pp. 14–18, Spring 2000.
[2] C. W. Gellings, ‘‘New Products and Services for the Electric Power Industry,’’
The Bridge, vol. 40, no. 1, pp. 21–28, Spring 2010.
[3] S. M. Amin, ‘‘Securing the Electricity Grid,’’ The Bridge, vol. 40, no. 1,
pp. 13–20, Spring 2010.
[4] J. D. Rhodes and The Conversation. (May 2017). The Outdated US Electric
Grid is Going to Cost $5 Trillion to Replace. Available: http://www.
businessinsider.com/replacing-us-electrical-grid-cost-2017-3.
[5] A. Kwasinski, W. Weaver, and R. S. Balog, Microgrids and Other Local
Area Power and Energy Systems. Cambridge University Press, Cambridge,
UK, 2016.
[6] R. H. Lassetter, ‘‘Smart Distribution: Coupled Microgrids,’’ Proceedings of
the IEEE, vol. 99, no. 6, pp. 1074–1082, Jun 2011.
[7] Y. Zhang, L. Xie, and Q. Ding, ‘‘Interactive Control of Coupled Microgrids
for Guaranteed System-Wide Small Signal Stability,’’ IEEE Transactions on
Smart Grid, vol. 7, no. 2, pp. 1088–1096, Nov 2015.
[8] D. T. Ton and M. A. Smith, ‘‘The U.S. Department of Energy’s Microgrid
Initiative,’’ The Electricity Journal, vol. 25, no. 8, pp. 84–94, Oct 2012.
[9] R. S. Balog, W. W. Weaver, and P. T. Krein, ‘‘The Load as an Energy Asset
in a Distributed DC SmartGrid Architecture,’’ IEEE Transactions on Smart
Grid, vol. 3, no. 1, pp. 253–260, March 2012, Art. no. TSG-00146-2011.
[10] Commonwealth Edison Co. (May 2017). Central AC Cycling. Available: https://
www.comed.com/WaysToSave/ForYourHome/Pages/CentralACCycling.aspx.
[11] J. O. Swanson and J. P. Jolliffe, ‘‘Load Shedding Program in the Pacific
Northwest,’’ Transactions of the American Institute of Electrical Engineers.
Part III: Power Apparatus and Systems, vol. 73, no. 2, pp. 1655–1668, Jan 1954.
[12] R. M. Maliszewski, R. D. Dunlop, and G. L. Wilson, ‘‘Frequency Actuated
Load Shedding and Restoration Part I—Philosophy,’’ IEEE Transactions on
Power Apparatus and Systems, vol. PAS-90, no. 4, pp. 1452–1459, Jul 1971.
[13] Z. Wang, B. Chen, J. Wang, and M. M. Begovic, ‘‘Stochastic DG Placement
for Conservation Voltage Reduction Based on Multiple Replications
Procedure,’’ IEEE Transactions on Power Delivery, vol. 30, no. 3, pp.
1039–1047, Jun 2015.
[14] H. Chen, Y. Li, R. H. Y. Louie, and B. Vucetic, ‘‘Autonomous Demand Side
Management Based on Energy Consumption Scheduling and Instantaneous
Load Billing: An Aggregative Game Approach,’’ IEEE Transactions on
Smart Grid, vol. 5, no. 4, pp. 1744–1754, Jun 2014.
[15] H. Zhong, L. Xie, and Q. Xia, ‘‘Coupon Incentive-Based Demand Response:
Theory and Case Study,’’ IEEE Transactions on Power Systems, vol. 28,
no. 2, pp. 1266–1276, May 2013.
The load as a controllable energy asset in dc microgrids 185
[16] Younicos. (May 2017). Y.Cube: Our Plug-and-Play Energy Storage Solution
[Online]. Available: https://www.younicos.com/products/y-cube/.
[17] Tesla. (May 2017). PowerPack: Utility and Business Energy Storage.
Available: https://www.tesla.com/powerpack.
[18] Tesla. (May 2017). Powerwall: Reliable Power Day and Night. Available:
https://www.tesla.com/powerwall.
[19] L. Poulin, Reducing Risk with Software Process Improvement. Boca Raton,
FL: Auerbach Publications, Taylor & Francis Group, 2005.
[20] A. V. Lotov and K. Miettinen, ‘‘Visualizing the Pareto Frontier,’’ in Multi-
objective Optimization: Interactive and Evolutionary Approaches, J. Branke,
K. Deb, K. Mirettinen, and R. Slowinski, Eds.: Springer-Verlag Berlin
Heidelberg, 2008, pp. 213–243.
[21] N. Barr, Economics of the Welfare State, 5 ed. United Kingdom: Oxford
University Press, 2012.
[22] M. B. Shadmand and R. S. Balog, ‘‘Multi-Objective Optimization and
Design of Photovoltaic-Wind Hybrid System for Community Smart dc
Microgrid,’’ IEEE Transactions on Smart Grid, vol. 5, no. 5, pp. 2635–2643,
Sep 2014.
[23] J. W. Kimball, B. T. Kuhn, and R. S. Balog, ‘‘A System Design Approach
for Unattended Solar Energy Harvesting Supply,’’ IEEE Transactions on
Power Electronics, vol. 24, no. 4, pp. 952–962, Apr 2009.
[24] ERCOT. (Sep 2016, May 2017). Renewable Integration at ERCOT. Avail-
able: http://www.cigre.cl/seminarios/wp-content/uploads/2016/09/ERCOT-
DAN-WOODFIN.pdf.
[25] Lighting Research Center (LRC). (April 2010, May 2017). Demonstration
and Evaluation of Lighting Technologies and Applications: Daylight-
Harvesting Switch. Available: http://www.lrc.rpi.edu/programs/DELTA/pdf/
FTDELTA_DaylightHarvestingSwitch.pdf.
[26] The National Optical Astronomy Observatory. (May 2017). Recommended
Light Levels (Illuminance) for Outdoor and Indoor Venues. Available:
https://www.noao.edu/education/QLTkit/ACTIVITY_Documents/Safety/
LightLevels_outdoorþindoor.pdf.
[27] Occupational Safety and Health Branch. (Dec 2008, May 2017). Lighting
Assessment in the Workplace. Available: http://www.labour.gov.hk/eng/
public/oh/Lighting.pdf.
[28] R. J. Sledz, ‘‘Control Room Lighting: An Application of Human Factors Engi-
neering,’’ IEEE Power Engineering Review, vol. PAS-101, no. 8, pp. 2755–2761,
Aug 1982.
[29] J. Voelcker. (Oct 2007, May 2017). Can Plug-In Hybrid Electric Vehicles Keep
the Electric Grid Stable? Available: http://spectrum.ieee.org/transportation/
advanced-cars/can-plugin-hybrid-electric-vehicles-keep-the-electric-grid-
stable.
[30] A. Vaughan. (Mar 2017, May 2017). Charge Electric Cars Smartly to Take
Pressure off National Grid—Minister. Available: https://www.theguardian.
186 DC distribution systems and microgrids
com/environment/2017/mar/20/electric-cars-uk-power-grids-charging-peaks-
sse-demand-side-response.
[31] S. Chakraborty, W. Kramer, B. Kroposki, et al., ‘‘Interim Test Procedures
for Evaluating Electrical Performance and Grid Integration of Vehicle-to-Grid
Applications,’’ National Renewable Energy Lab (NREL), Golden, Colorado.
Jun 2011, Available: http://www.nrel.gov/docs/fy11osti/51001.pdf.
[32] I. U. Nutkani, P. C. Loh, P. Wang, and F. Blaabjerg, ‘‘Cost-Prioritized Droop
Schemes for Autonomous AC Microgrids,’’ IEEE Transactions on Power
Electronics, vol. 30, no. 2, pp. 1109–1119, Feb 2015.
[33] L.-L. Fan, V. Nasirian, H. Modares, F. L. Lewis, Y.-D. Song, and A. Davoudi,
‘‘Game-Theoretic Control of Active Loads in DC Microgrids,’’ IEEE
Transactions on Energy Conversion, vol. 31, no. 3, pp. 882–895, Sep 2016.
[34] J. M. Guerrero, P. C. Loh, T. L. Lee, and M. Chandorkar, ‘‘Advanced
Control Architectures for Intelligent Microgrids Part II: Power Quality,
Energy Storage, and ac/dc Microgrids,’’ IEEE Transactions on Industrial
Electronics, vol. 60, no. 4, pp. 1263–1270, Apr 2013.
[35] J. M. Guerrero, M. Chandorkar, T.-L. Lee, and P. C. Loh, ‘‘Advanced
Control Architectures for Intelligent Microgrids—Part I: Decentralized and
Hierarchical Control,’’ IEEE Transactions on Industrial Electronics, vol. 60,
no. 4, pp. 1254–1262, Apr 2013.
[36] V. Nasirian, A. Davoudi, F. L. Lewis, and J. M. Guerrero, ‘‘Distributed
Adaptive Droop Control for DC Distribution Systems,’’ IEEE Transactions
on Energy Conversion, vol. 29, no. 4, pp. 944–956, Dec 2014.
[37] M. D. Cook, G. G. Parker, R. D. Robinett, and W. W. Weaver, ‘‘Decen-
tralized Mode-Adaptive Guidance and Control for DC Microgrid,’’ IEEE
Transactions on Power Delivery, vol. 32, no. 1, pp. 263–271, Feb 2017.
[38] J. He, Y. W. Li, J. M. Guerrero, F. Blaabjerg, and J. C. Vasquez, ‘‘An
Islanding Microgrid Power Sharing Approach Using Enhanced Virtual
Impedance Control Scheme,’’ IEEE Transactions on Power Electronics,
vol. 28, no. 11, pp. 5272–5282, Nov 2013.
[39] R. S. Balog, ‘‘Autonomous Local Control in Distributed DC Power
Systems,’’ PhD Dissertation, Department of Electrical and Computer Engi-
neering, University of Illinois at Urbana-Champaign, 2006.
[40] F. Gao, S. Bozhko, G. Asher, P. Wheeler, and C. Patel, ‘‘An Improved
Voltage Compensation Approach in a Droop-Controlled DC Power System
for the More Electric Aircraft,’’ IEEE Transactions on Power Electronics,
vol. 31, no. 10, pp. 7369–7383, Oct 2016.
[41] M. B. Shadmand, R. S. Balog, and H. Abu-Rub, ‘‘Model Predictive Control
of PV Sources in a Smart dc Distribution System: Maximum Power Point
Tracking and Droop Control,’’ IEEE Transactions on Energy Conversion,
vol. 29, no. 4, pp. 913–921, Dec 2014.
[42] X. Lu, J. M. Guerrero, K. Sun, and J. C. Vasquez, ‘‘An Improved Droop
Control Method for dc Microgrids Based on Low Bandwidth Communica-
tion with dc Bus Voltage Restoration and Enhanced Current Sharing
The load as a controllable energy asset in dc microgrids 187
1
Department of Electrical and Computer Engineering, North Carolina State University, USA
190 DC distribution systems and microgrids
dc Fast CHAdeMO GB/T CCS Type 1 (US) CCS Type 2 (EU) Tesla
charging IEEE 2030.1.1, (China) SAE J1772, IEC 62196-3 (US, EU)
system IEC 62196-3 GB/T 20234.3, IEC 62196-3 (Configuration FF)
(Configuration AA) IEC 62196-3 (Configuration EE)
(Configuration BB)
Charge coupler
inlet
Supercharger
Modified Mennekes
Type 2 (EU)
exchange of maximum charging parameters between a car and a charger. The car
would also perform the insulation test before the charging can begin. If all the
required criteria are met, the car closes its dc contactor and the charging may begin.
During the charging process, the charger follows the voltage and current commands
from the car’s Battery Management System (BMS). Each charging event typically
consists of two charging modes (regimes): a ‘‘Constant Current’’ (CC) mode fol-
lowed by a ‘‘Constant Voltage’’ (CV) mode. During the CC mode, the charger
follows a slow varying current reference set by the BMS. After a certain time spent
in the CC mode (i.e., when the battery voltage reaches a certain value), the BMS
switches to CV mode and sends a voltage reference command, which increases
slowly as the charging progresses. The charging voltage increases much faster in
the CC mode, than in the CV mode. During the CV mode, the charging current
decreases. The time spent in CC and CV modes depends mainly on the applied
charging current and the initial battery State of Charge (SOC). When the battery
reaches a certain preset SOC, the car signals the charger to terminate the charging
by reducing the charging current to zero. The car then disconnects itself from the
charger by opening its dc contactor. The example of a charging profile of a Tesla
model S85 and the corresponding charging power and the battery SOC are shown
in Figure 8.1 [19]. The battery was charged from 12% SOC to 94% SOC, with a
peak power of 117.2 kW occurring at 16% SOC (2 min after the charging was
started). In this case, the CC mode lasted for approximately 2 min. During the
charging process, the car and the charger typically exchange the information on
current and voltage reference, the battery SOC and some status information. In case
of Tesla Superchargers, the car will also periodically send its Vehicle Identification
Number (VIN) to identify itself as a Tesla vehicle.
194 DC distribution systems and microgrids
Battery current Battery voltage Charging power Battery SOC
350 410 120 100
110 90
300 400
Figure 8.1 The charging profile of a Tesla model S85 and the corresponding
charging power and the battery SOC
A B C
Isolated dc/dc stage
PFC
Input Rectifier Output
circuit (dc/ac) (ac/dc)
filter (ac/dc) filter
(dc/dc)
HF
transformer
Figure 8.2 The simplified block diagram of a conventional dc fast charger power
conversion system
Figure 8.3 The boost-type six-switch PFC converter using IGBTs with input
LLC filter
Figure 8.4 The buck-type six-switch PFC converter with input and output filters
series with the capacitors) can be used instead of the LLC filter to make the system
more efficient [23,24]. The six-switch PWM converter shown in Figure 8.3 has a
boost-type characteristic (the output dc voltage must be higher than the peak input
ac line-to-line voltage), enables sinusoidal input currents, bidirectional power flow
and can operate with high efficiency and arbitrary phase difference between the
input voltage and the input current fundamental [25,26].
The sinusoidal waveform of the input current is achieved by a closed-loop
current control, which requires sensing of the input ac currents. The high turn-on
switching loss in each IGBT as a result of a reverse recovery of the antiparallel
diode of the other IGBT in the same phase leg can be avoided by using SiC
MOSFETs with antiparallel Junction Barrier Schottky (JBS) diodes which practi-
cally do not exhibit any reverse recovery.
If only unidirectional power flow is required, a buck-type six-switch converter
shown in Figure 8.4 can be used. This converter has some advantages over the
196 DC distribution systems and microgrids
Figure 8.7 (a) Unidirectional dc/dc converter with active snubber and LC output
filter, (b) passive RCD snubber and (c) passive CDD snubber
sources: the transformer leakage inductance and the output filter inductance) and
severe ringing at the diode bridge output due to the interaction of the transformer
leakage inductance and the parasitic capacitance of the reverse biased diodes. To
reduce the voltage overshoot and the ringing, the different types of snubber circuits
are proposed and applied. The ones that are typically used due to their simplicity and
effectiveness are shown in Figure 8.7. The active snubber circuit shows the best
performance since it can completely eliminate the voltage ringing and the overshoot
[31]. However, that comes with a price of the increased control complexity. The
passive RCD snubber is a simple but effective alternative, and the CDD snubber is a
cost-effective and low-loss version of the passive snubber with better clamping
ability than the RCD snubber [32].
If bidirectional power flow is required, an active rectifier can be used instead
of the diode bridge at the transformer’s secondary. The snubber would still be
required to reduce the ringing across the secondary-side switches.
A dual active bridge (DAB) converter (shown in Figure 8.8) is another bidir-
ectional converter that has been proposed for EV charging applications mainly due
to its high power density and high efficiency, buck-boost operation, low device
stress, small filter components and low sensitivity to system parasitics [33–40].
When introduced in 1991 [40], the DAB converter initially did not receive a
wider adoption due to the high power losses and relatively low switching frequency
198 DC distribution systems and microgrids
(a) (b)
Figure 8.9 Dual half-bridge converter: (a) with voltage-fed output and (b) with
current-fed output
be sized for the average power demand, rather than the peak demand [49]. In this
configuration, an energy storage system is connected to the dc bus to supply power
when the demand exceeds the average that can be provided from the grid. It is
shown that over 98% of power demand could be satisfied even with a relatively
small storage system that is rated based on the average value of the needed storage
power. Even though the charging would be delayed for over 60% of the customers,
maximum delay time would be approximately 2 min, and the average delay time
would not be longer than 10 s.
800 A
250 A 250 A 250 A 250 A 15 A
CT 5 5 5 5 3
MV/LV transformer
As already mentioned [49], adding the energy storage to the dc fast charging
system could minimize the peak demand charges and therefore reduce the operat-
ing costs. The added energy storage needs to be sized properly to be capable of
buffering the power and energy demands of the charging station during the peak
demand period. During the off-peak hours, the stored energy can be used to provide
ancillary services for improved grid stability and security. The onsite photovoltaic
energy generation system can also be added to the charging station to improve
decoupling of power provided to the vehicles from the power drawn from the grid.
State-of-the-art systems are typically coupled on the LV ac side through ac-to-dc or
dc-to-ac converters, as illustrated in Figure 8.11(a). Even though bidirectional
operation is enabled by charging standards, majority of modern dc fast chargers
operate in unidirectional mode.
One significant advantage of using the ac bus is the availability and maturity of
the rectifier and inverter technology, availability of ac switchgear and protection
devices, and well-established standards and practices for the ac power distribution
systems. On the other hand, having more conversion stages in the ac-coupled
202 DC distribution systems and microgrids
Energy
PV storage
dc/dc dc/dc
MV
grid
MV/LV
dc/ac dc/ac
ac/dc ac/dc
dc/dc dc/dc
(a)
PV Energy
storage
MV
grid
MV/LV
dc/dc dc/dc
ac/dc
dc/dc dc/dc
(b)
system (to interface dc generation, energy storage and dc loads to the ac system)
increases the system complexity and cost and decreases the system efficiency.
Also, dealing with reactive power, inverter synchronization, and voltage and fre-
quency control during islanded operation of the ac-connected systems makes the
ac systems more complicated to control than the dc-connected systems. A dc-
connected system, illustrated in Figure 8.11(b), features one centralized ac/dc
converter and provides more convenient and more energy efficient way of con-
necting the energy storage and the renewable energy sources to the central dc bus.
The test results of a simple ac-coupled system with 20 kW (16 kWh) lithium–
polymer battery storage and an EV charger commercial prototype with maximum
50 kW dc and 22 kW ac output power are reported in [54]. The battery voltage was
first boosted to 600 V and then fed to an inverter which was connected to the 400-V
ac grid through the low-frequency isolation transformer. The peak power shaving
ability is demonstrated for two different cases: charging Nissan Leaf from 65%
Electric vehicle charging infrastructure and dc microgrids 203
SOC for approximately 10 min, and from 40% to 100% SOC. During the second
test, the energy delivered from the energy storage was 4.5 kWh, which was 36% of
the total energy absorbed by the charger (12.4 kWh). An implementation of
dc-coupled system is reported in [55], where 288 V/40 Ah (11.5 kWh) lead-acid
battery is used as a storage to reduce the power drawn from the ac grid during the
fast charging of the 50 V/40 Ah (2 kWh) LiFePO4 battery used in two wheelers.
Energy
PV
storage
MV grid
SST
dc/dc dc/dc
HF TR
ac/dc dc/dc
dc/dc dc/dc
Service transformer
Figure 8.13 Comparison of the SOA EV charger system and the SST-based
MV solution from [60]
single-phase line and a 25-kW prototype that connects to the 7.2-kV ac single-phase
line. In both the cases, the rated output voltage was 400 V and the grid-to-vehicle
efficiency was higher than 96%. Another, more recent effort, has demonstrated a
modular 50-kW dc fast charger that uses 1.2 kV SiC devices, connects to 2.4 kV ac
single-phase line, outputs 400 V and has the efficiency of over 96% [60]. The ben-
efits of the MV approach with SiC devices are illustrated in Figure 8.13.
As previously mentioned, a dedicated service transformer is typically required
to step down the distribution system medium voltage and provide three-phase supply
to the commercial dc fast charger. With the service transformer’s efficiency of
98.5% (at 50 kVA level), the grid-to-battery efficiency of the 50-kW transformer-
and-charger systems is less than 93%. The efficiency of higher power chargers
(135 kW Tesla Supercharger or 150 kW Delta Stromtankstelle 4.0) do not exceed
93% while still requiring bulky low-voltage service transformers to provide
208/480 V three-phase input, which itself has an efficiency below 99% (at 150 kVA
power level). The efficiency of the entire system with the 150-kV charger is therefore
approximately 92%. Compared to the SOA system, the MV dc fast charger prototype
reported in [60] reduces the size and the weight by ten-fold and improves the system
efficiency from 93% to over 96%—a reduction in losses by approximately 50%.
Additionally, by connecting directly to the MV line, the installation costs can be
reduced by at least 40%, compared to the conventional solution [61].
The MV EV fast chargers are designed to be modular, using identical modules
as building blocks to reach the desired power and voltage levels. The modules are
typically connected in series at the input and in parallel at the output, as illustrated
in Figure 8.14 [58,60,62,63]. The series connection at the input allows the rela-
tively low-voltage off-the-shelf silicon IGBTs or SiC MOSFETs to serve the
medium-voltage application. The use of SiC devices substantially increases the
Module 1 Module 1 200 – 450 V
800 V 625 V
2.4 kV 2.4 kV
Module 2 Module 2
Module 3 Module 3
(a) (b)
Module 1 Module 1
VOUT VOUT
VIN VIN
Module 2 Module 2
Module N Module N
(c) (d)
Figure 8.14 Input-series–output-parallel MV charger topologies: (a) multicell boost (MCB) topology [60], (b) UDFC topology [58],
(c) cascaded H-bridge topology with integrated split battery [62], (d) cascaded H-Bridge with current-fed DAB [63]
206 DC distribution systems and microgrids
efficiency of the system compared to the use of the silicon equivalents, due to
the lower switching and conduction losses. Additionally, the high switching fre-
quency operation of SiC MOSFETs enables significant reduction in size of the
inductors and the transformers used to provide the required galvanic isolation, thus
reducing the overall system size and weight.
With the input-series connection, a special care must be taken to balance the
voltage of dc link capacitors in each module, while ensuring the low total harmonic
distortion (THD) of the input current [63,64]. The high isolation of gate drive
power supplies is required for safe operation at high common-mode voltage. Low
coupling capacitance of gate drive power supplies and power transformers is
essential for reduction of the conducted EMI, especially if high-speed WBG devi-
ces are used. Another significant technical challenge is related to the design and
implementation of the fast-acting system protections and safety features at medium
voltage level. The system must be able to withstand Basic Insulation Level tests
and low-frequency-voltage insulation tests, where the converter is exposed to vol-
tages significantly higher than its rated voltage. Therefore, additional circuitry is
required to clamp the voltage seen by the converter in order to meet these
requirements. Furthermore, the converter must have an internal protection
mechanism that clears any system fault in a few microseconds if the converter
devices are to be protected and the system is to continue to operate after the fault is
cleared. Such fast clearing of a fault cannot be achieved using conventional MV
mechanical switches, and solid-state solutions are required.
The design of a 16.7-kW isolated dc/dc MCB module used in a 50-kW MV fast
charger prototype from [60] is illustrated in Figure 8.15 [65]. The module topology
uses low number of active semiconductor switches, has high utilization of the
switches, and enables independent control of input and output stages and simple dc
bus capacitor voltage balancing control [66]. The switching frequency of the three-
level boost stage is set to 25 kHz and the minimum switching frequency is limited
by the required THD of the input current. The switching frequency of the NPC
converter is set to 50 kHz.
The SST-based MV EV charger system can service many other applications,
including industrial applications, data centers, and aerospace and military appli-
cations. As an example, the reduction of weight and volume is particularly sig-
nificant in naval shipboard applications, where the transformers for a redundant
propulsion system can weigh as much as 35 t. The power converters for shipboard
applications need to have high power density in order to save space and fuel, and dc
MV systems with 100 kHz switching frequencies are seen as an emerging bench-
mark for these systems in the 2025–30 time frame [67].
The proposed high-power medium-voltage modular rectifier can also serve as a
power supply for data centers. According to DOE’s Center of Expertise for Energy
Efficiency in Data Center’s recent report [68], data centers in the United States
consumed an estimated 70 billion kWh in 2014, representing about 1.8% of the
total US electricity consumption. Additionally, based on current trends, US data
centers are projected to consume approximately 73 billion kWh in 2020. Most of
these data centers (except high-end and hyperscale centers) operate with the Power
Electric vehicle charging infrastructure and dc microgrids 207
DC link capacitors
HF power transformer
Output inductor
Figure 8.15 16.7 kW isolated dc/dc module as a building block for MV EV fast
charger [65]
Usage Effectiveness of approximately 2, which means that half of the used energy
is consumed by the data center’s supporting infrastructure (cooling equipment,
uninterrupted power supplies, lighting, etc.). As a result, any power savings that can
be made in the operation of the data center will have significant environmental
implications as well as impacts on the system cost of operation. Recently, data
centers with 380 V dc distribution systems have been proposed, with their inher-
ently higher efficiency due to the fewer conversion stages. The SST-based topology
used for MV EV chargers can be used to efficiently supply 380 V dc to the data
centers. The benefits of higher system efficiency, better power conditioning and
better quality power can all be achieved with the SST-based design [69].
208 DC distribution systems and microgrids
MV 3Φ distribution
SST
ac
dc ac
dc
dc
Breaker dc
dc
dc
DESD
Net meter dc Battery
dc
Utility owned
Breaker
dc Battery
dc
Breaker Breaker
dc dc
800 V
dc dc
Breaker 400 V
Net meter dc
dc
dc bus ac bus
the power distribution system by completely decoupling the medium voltage dis-
tribution grid from the low voltage feeder sections. The bidirectional power flow
capability of the SST provides possibilities to integrate the distributed energy
storage devices and to feed locally generated power back to the grid in a controlled
way. Recent technology advances in WBG semiconductor devices, such as high-
voltage SiC MOSFETs and diodes, serve as the enabling technology for the SST.
Despite its unparalleled capabilities, significant challenges stand in the way of
the complete adoption of the SST by the electric utilities today. The key barriers
include (1) inherent reliability concerns of replacing the passive transformer with a
power electronic device, (2) relatively low penetration of renewable resources seen
in the system today, which can mostly be handled by the existing infrastructure,
(3) the large inertia still present in the distribution system, supplied by the legacy
generators and (4) limited ability to monetize the perfect power quality supplied by
the SST. As a result, the other venues to demonstrate the SST technology can be
explored and showcase the benefits of direct connection of power electronics to the
MV grid. Medium-voltage SSTs using WBG devices can serve many other emer-
ging and established applications including EV fast chargers, data center power
supplies and military power conversion stages, to name a few. The EV fast chargers
present the largest opportunity, given the size and the rate of growth of the market.
Further, the use of a medium-voltage dc fast charger has clear monetary advantages
since this solution will substantially reduce the installation cost, which makes up to
half of the price of deploying a commercial dc fast charger, and it is particularly
suitable for deployment in densely populated urban areas. Overall, the SST-based
approach has potential to help reduce the EV charging time and bring a gas station
experience to the EV users, while providing the integration of renewable resources
and energy storage in a cost-competitive way.
References
[1] International Energy Agency (IEA) and the Electric Vehicles Initiative of the
Clean Energy Ministerial (EVI), ‘‘Global EV Outlook 2017,’’ OECD/IEA,
Paris, France, Jul. 2017.
[2] ‘‘Monthly Plug-In Sales Scorecard’’. [Online]. Available: https://insideevs.
com/monthly-plug-in-sales-scorecard/. Accessed on Jan. 5, 2018.
[3] ‘‘Compare Electric Cars Side-by-Side’’. [Online]. Available: https://www.
fueleconomy.gov/feg/evsbs.shtml. Accessed on Jan. 5, 2018.
[4] D. Howell, S. Ahmed, R. B. Carlson, et al., ‘‘Enabling Fast Charging: A
Technology Gap Assessment,’’ U.S. Department of Energy, Office of Energy
efficiency & Renewable Energy, Report INL/EXT-17-41638, Oct. 2017.
[Online]. Available: https://energy.gov/sites/prod/files/2017/10/f38/XFC%
20Technology%20Gap%20Assessment%20Report_FINAL_10202017.pdf.
[5] ‘‘Largest Global Charger Coverage’’. [Online]. Available: https://www.
chademo.com/. Accessed on Jan. 5, 2018.
210 DC distribution systems and microgrids
[34] S. Han and D. Divan, ‘‘Bi-directional DC/DC converters for plug-in hybrid
electric vehicle (PHEV) applications,’’ 2008 Twenty-Third Annual IEEE
Applied Power Electronics Conference and Exposition, Austin, TX, 2008,
pp. 784–789.
[35] C. Mi, H. Bai, C. Wang and S. Gargies, ‘‘Operation, Design and Control of
Dual H-Bridge-based Isolated Bidirectional DC–DC Converter,’’ in IET
Power Electronics, vol. 1, no. 4, pp. 507–517, Dec. 2008.
[36] R. J. Ferreira, L. M. Miranda, R. E. Araújo and J. P. Lopes, ‘‘A new
bi-directional charger for vehicle-to-grid integration,’’ 2011 2nd IEEE PES
International Conference and Exhibition on Innovative Smart Grid Tech-
nologies, Manchester, 2011, pp. 1–5.
[37] H. van Hoek, M. Neubert and R. W. De Doncker, ‘‘Enhanced Modulation
Strategy for a Three-Phase Dual Active Bridge—Boosting Efficiency of an
Electric Vehicle Converter,’’ in IEEE Transactions on Power Electronics,
vol. 28, no. 12, pp. 5499–5507, Dec. 2013.
[38] L. Xue, Z. Shen, D. Boroyevich, P. Mattavelli and D. Diaz, ‘‘Dual Active
Bridge-Based Battery Charger for Plug-in Hybrid Electric Vehicle with
Charging Current Containing Low Frequency Ripple,’’ in IEEE Transactions
on Power Electronics, vol. 30, no. 12, pp. 7299–7307, Dec. 2015.
[39] L. Xue, M. Mu, D. Boroyevich and P. Mattavelli, ‘‘The optimal design of
GaN-based Dual Active Bridge for bi-directional Plug-IN Hybrid Electric
Vehicle (PHEV) charger,’’ 2015 IEEE Applied Power Electronics
Conference and Exposition (APEC), Charlotte, NC, 2015, pp. 602–608.
[40] R. W. A. A. De Doncker, D. M. Divan and M. H. Kheraluwala, ‘‘A Three-
Phase Soft-Switched High-Power-Density DC/DC Converter for High-
Power Applications,’’ in IEEE Transactions on Industry Applications,
vol. 27, no. 1, pp. 63–73, Jan./Feb. 1991.
[41] H. Li, D. Liu, F. Z. Peng and G.-J. Su, ‘‘Small signal analysis of a dual half
bridge isolated ZVS Bi-directional DC–DC converter for electrical vehicle
applications,’’ 2005 IEEE 36th Power Electronics Specialists Conference,
Recife, 2005, pp. 2777–2782.
[42] M. Vasiladiotis, B. Bahrani, N. Burger and A. Rufer, ‘‘Modular converter
architecture for medium voltage ultra fast EV charging stations: Dual half-
bridge-based isolation stage,’’ 2014 International Power Electronics Conference
(IPEC-Hiroshima 2014 – ECCE ASIA), Hiroshima, 2014, pp. 1386–1393.
[43] W. Kempton, J. Tomić, ‘‘Vehicle-to-Grid Power Fundamentals: Calculating
Capacity and Net Revenue,’’ in Journal of Power Sources, vol. 144, no. 1,
pp. 268–279, 2005.
[44] J. A. P. Lopes, P. M. R. Almeida and F. J. Soares, ‘‘Using vehicle-to-grid to
maximize the integration of intermittent renewable energy resources in
islanded electric grids,’’ 2009 International Conference on Clean Electrical
Power, Capri, 2009, pp. 290–295.
[45] P. Kadurek, C. Ioakimidis and P. Ferrao, ‘‘Electric vehicles and their impact
to the electric grid in isolated systems,’’ 2009 International Conference on
Power Engineering, Energy and Electrical Drives, Lisbon, 2009, pp. 49–54.
Electric vehicle charging infrastructure and dc microgrids 213
[58] EPRI, Utility Direct Medium Voltage DC Fast Charger Update: DC Fast
Charger Characterization. EPRI, Palo Alto, CA: 2012. 1024106.
[59] S. Rajagopalan, A. Maitra, J. Halliwell, M. Davis and M. Duvall, ‘‘Fast
charging: An in-depth look at market penetration, charging characteristics,
and advanced technologies,’’ 2013 World Electric Vehicle Symposium and
Exhibition (EVS27), Barcelona, 2013, pp. 1–11.
[60] S. Srdic, X. Liang, C. Zhang, W. Yu and S. Lukic, ‘‘A SiC-based high-
performance medium-voltage fast charger for plug-in electric vehicles,’’
2016 IEEE Energy Conversion Congress and Exposition (ECCE), Milwaukee,
WI, 2016, pp. 1–6.
[61] A. Maitra, ‘‘EPRI’s utility direct DC fast charger–development, testing,
demonstration,’’ EPRI IWC Meeting, Atlanta, GA, Mar. 28, 2012.
[62] M. Vasiladiotis and A. Rufer, ‘‘A Modular Multiport Power Electronic
Transformer with Integrated Split Battery Energy Storage for Versatile
Ultrafast EV Charging Stations,’’ in IEEE Transactions on Industrial Elec-
tronics, vol. 62, no. 5, pp. 3213–3222, May 2015.
[63] D. Sha, G. Xu and Y. Xu, ‘‘Utility Direct Interfaced Charger/Discharger
Employing Unified Voltage Balance Control for Cascaded H-Bridge Units
and Decentralized Control for CF-DAB Modules,’’ in IEEE Transactions on
Industrial Electronics, vol. 64, no. 10, pp. 7831–7841, Oct. 2017.
[64] X. Liang, C. Zhang, S. Srdic and S. Lukic, ‘‘Predictive Control of a
Series-Interleaved Multi-Cell Three-Level Boost Power Factor Correction
Converter,’’ in IEEE Transactions on Power Electronics, vol. 33, no. 10,
pp. 8948–8960, Oct. 2018.
[65] S. Srdic, C. Zhang, X. Liang, W. Yu and S. Lukic, ‘‘A SiC-based power
converter module for medium-voltage fast charger for plug-in electric
vehicles,’’ 2016 IEEE Applied Power Electronics Conference and Exposi-
tion (APEC), Long Beach, CA, 2016, pp. 2714–2719.
[66] D. Rothmund, G. Ortiz and J. W. Kolar, ‘‘SiC-based unidirectional solid-
state transformer concepts for directly interfacing 400 V DC to medium-
voltage AC distribution systems,’’ 2014 IEEE 36th International
Telecommunications Energy Conference (INTELEC), Vancouver, BC,
2014, pp. 1–9.
[67] F. Wang, Z. Zhang, T. Ericsen, R. Raju, R. Burgos and D. Boroyevich,
‘‘Advances in Power Conversion and Drives for Shipboard Systems,’’ in
Proceedings of the IEEE, vol. 103, no. 12, pp. 2285–2311, Dec. 2015.
[68] A. Shehabi, S. Smith, D. Sartor, et al., ‘‘United States Data Center Energy
Usage Report,’’ Lawrence Berkeley National Laboratory, Jun. 2016. [Online].
Available: https://eta.lbl.gov/sites/all/files/publications/lbnl-1005775_v2.pdf.
[69] G. Zhabelova, A. Yavarian, V. Vyatkin and A. Q. Huang, ‘‘Data center
energy efficiency and power quality: An alternative approach with solid state
transformer,’’ IECON 2015 – 41st Annual Conference of the IEEE Industrial
Electronics Society, Yokohama, 2015, pp. 1294–1300.
[70] A. Huang, ‘‘FREEDM system – a vision for the future grid,’’ IEEE PES
General Meeting, Minneapolis, MN, 2010, pp. 1–4.
Chapter 9
Overview and design of solid-state transformers
Levy Costa1, Marco Liserre1, and Giampaolo Buticchi1
The concept of electronic transformer was first introduced by [1] in the end of
1960s, where the main concept would be to adjust the output voltage by using the
power electronics, while contributing to size and volume reduction. The apparatus
in question isolated the primary and secondary side in frequency higher than the
grid frequency and for that reason the size, weight and volume could be con-
siderably reduced. At that time, there was no real applicability for such apparatus,
mainly because of the performance limitation of the available semiconductors on
the market.
However, with the advancement of the power electronics and semiconductors
technology, new faster devices with lower switching energy were developed,
enabling faster switching frequency operation. As a result, the volume and size of
the power converters were considerably reduced, bringing advantages for those
applications in which the power density is very important. At the same time, the
classic solution used in traction for the electric locomotive was very heavy, bulky
and inefficient, because of the low frequency transformer (LFT) employed [2]
associated to the power electronics converters required for providing the suitable
dc voltage to the variable speed drive on the locomotive. In this sense, the power
electronics transformer or solid-state transformer (SST), as usually called, could
provide the suitable controlled and isolated dc voltage in a unique solution, offering
either volume and weight reduction (around 20%–50%) and efficiency improve-
ment (from 93% to 96%) [2,3]. Thus, the SST became a reality in traction
applications.
Figure 9.1 illustrates the conventional solution (state of the art) used on the
locomotives, as well as the modern solution based on SST. The benefits offered by
the SST in traction application are mainly on volume and weight reduction. Thus,
the advantages of the system are only related to its hardware. The SST has been
frequently conceptualized as presented below.
1
Chair of Power Electronics Faculty of Engineering, Christian-Albrechts-University of Kiel, Germany.
216 DC distribution systems and microgrids
MVAC MVAC
Conventional solution SST-based solution
LFT Modernization
Traction
ac dc Solid-
M state M
trans
dc ac
MVAC MVAC
Services MVDC
MV side
3 3
Efficiency Volume
Reliability Weight
Functionality Cost
Figure 9.3 Requirements and performance of the SST and ST for traction and
distribution grid applications
device, leading to the concept of smart transformer (ST). Then the ST can be
conceptualized as described below.
Figure 9.3 shows the main requirements and performance of the SST for both
described applications. While the volume and weight reduction are the most
218 DC distribution systems and microgrids
SST architecture
Power stages
1: MV ac-dc
1 Stage 2 Stages 3 Stages 2: Isolated dc-dc
3: LV dc-ac
Direct matrix converter Indirect matrix converter MV and LV decoupling
High power density Partial functionalities More functionalities
No extra functionalities
Scalability in
Modularity
9.3.2 Modularity
The next classification for the ST architecture concerns the degree of modularity.
It can be classified as modular, nonmodular or semimodular. The nonmodular system
is based on a single power converter, as exemplified in Figure 9.5(a), and usually
takes the advantage of high voltage wide-bandgap (WBG) semiconductors [4,5,8].
Special devices with high blocking voltage capabilities, such as 10 kV Silicon Car-
bide (SiC) MOSFETs or 15 kV SiC IGBTs [9–11] are needed to handle the MV level
in the power converter. Because these devices are not commercially available, but
only for research purposes, no currently available products are using this technology.
A standard approach to handle the MV in a nonmodular solution used in industrial
applications is to employ semiconductor connected in series; however, this demands
an additional control effort to balance the voltage over the semiconductors.
Conversely, the modular approach consists of several basic converters rated for
LV or low current, which are used as building blocks for the entire system, as
depicted in Figure 9.5(b). In this solution, the basic building block shares the voltage
dυ
dt
Nonmodular
MVDC LVDC
link link
Esw
(a)
Module
dυ
dt
LVDC
Modular
link
Esw
(b)
on the MV side and the current in the LV among them, enabling the use of LV rated
devices. In spite of the fact that more components are used compared to the non-
modular solution, the modular approach is more economically advantageous, as
demonstrated in [12]. Additionally, modular architectures bring several advantages to
the power and voltage scalability, maintenance and the implementation of fault-
tolerant strategies. In comparison to the nonmodular, the modular architectures have
reduced electromagnetic interference (EMI) emissions [due to the low dv=dt and
di=dt, as illustrated in Figure 9.5(a) and (b)] and the possibility to use standard
LV rated devices that perform well, benefiting both efficiency and cost.
Some features from the modular and nonmodular concepts can be combined to
form the semimodular concept, as an alternative for both previous approaches. In
this case, the architecture is implemented using basic modules sharing the power,
similar to the modular approach, but a nonreplaceable component processing the
total system power is part of the architecture. It can also be called as a partially
modular approach, in which the modules on the primary side divide the power
among them, while the high frequency transformer (HFT) and the secondary side
bridge processes all the power.
Basic unit
Basic unit
Cell level
(a)
System 1
Converter level
System level
(b) (c)
Figure 9.6 Modularization level: (a) cell level, (b) converter level and (c) system level
Overview and design of solid-state transformers 221
Cell level:
This is the most basic level, in which the basic unit is the cell used inside the
converter, as exemplified in Figure 9.6(a). Another example of a cell-level modular
system is the modular multilevel converter that is a single converter using several
cells as basic units.
Converter level:
The next level is the converter level. In this case, the power converter is the basic
unit, and they are combined in series and/or parallel to implement the entire system,
as shown in Figure 9.6(b). This is the most employed modularization level and the
input-series output-parallel connection is very popular. Interleaved converter is
another example of the converter level modular system, in which several converters
are connected in parallel.
System level:
Finally, the system level is presented in Figure 9.6(c). In this case, the full system
is used as the basic unit. It is often used in application where fault-tolerance or
scalability is highly desired.
All these levels can be freely combined, resulting in a modular system in
multiple levels. In modular ST architectures, the converter level is the most com-
mon one, although the cell level can also be used.
MV MV MV MV
AC AC AC AC
LVDC
(a) (b)
MV MV MV MV
AC AC AC AC
LVDC
MVDC MVDC
(c) (d)
connected. This strategy does not bring any advantages, and for this reason it has
not being adopted [3]. On the other hand, the distributed dc links on the MV side
provides advantages to the system, because the MV level involved in this side can
be distributed among the modules. For this reason, this concept has been inten-
sively used [3,4,12–16] and the most adopted structure is the one presented in
Figure 9.7(b).
Recently, the availability of the MVDC link has been discussed by the research
community, but there is no solution in this regard. The few amount of dc loads
applicable for MV level compared to the LV one (which is already well estab-
lished) makes the MVDC availability still controversial. Nevertheless, new dc
loads in LV level are emerging (e.g., electric fast charging station), supporting
research on the architectures with MVDC link. Consequently, the architecture with
MVDC link as shown in Figure 9.7(c) and (d) is not well adopted yet, but it is a
high potential structure and deserves attention.
In summary, the availability of the LVDC link is of paramount importance,
and then two configurations with respect to the dc link connectivity are highlighted:
the available LVDC link [Figure 9.7(b)] and the available LVDC and MVDC links
[Figure 9.7(d)].
Overview and design of solid-state transformers 223
In the last two decades, many different SST and ST architectures have been pro-
posed and discussed in literature for traction and distribution grid applications. The
most relevant architectures used in both application are discussed.
In traction applications, the architectures are characterized to use commercia-
lized HV Si-IGBT, two power processing stages, as well as modular approach. The
efficiencies for each architecture for traction system are presented in Table 9.1. All
of them provide higher efficiency and power density, when compared to the con-
ventional solution using the LFT.
The three-stage architecture offers decoupling between the input and output,
besides the availability of at least one dc link. Consequently, it is the preferable
structure for grid applications, as seen in Table 9.1. To handle the MV level, two
approaches are commonly used: modular with commercial LV devices and non-
modular with customized HV devices. Regardless the adopted concept, most of
these architecture takes advantages of the high performance WBG SiC devices.
They offer lower conduction and switching losses compared to the standard devi-
ces, and they are very promising in ST application.
Figure 9.8(a) shows the efficiency values available in the literature according
to power of the system, as denoted in Table 9.1. For distribution system, very few
ST demonstrators with solid results have been presented in the literature so far. The
Table 9.1 Comparison of the ST and SST architectures from different research
groups/companies
Distribution application
98 25
Distribution F K Distribution
H D F
96 20
C B G
94
C
I 15
B Traction
92 Traction 10 J
90 5 A I
Experimental E
Theoretical
H K
88 A 0 G
101 102 103 101 102 103 104
(a) Power rating (kVA) (b) Power rating (kVA)
available data shows that the efficiency achieved by this kind of system is around
94.5%–97%. For traction systems, the ST efficiency is in the range of 94.5%–96%,
according to the available data provided experimentally (excluding the theoretical
one). Although the ST architectures in the distribution use more power stages
compared to those used in Traction, their performance is potentially higher,
because the utilized WBG semiconductors that plays a very important role in this
regards. Figure 9.8(b) shows the isolation frequency for each ST architecture
shown in Table 9.1, according to the power level. In traction, the isolation fre-
quency is normally below 10 kHz, whereas in distribution grid is mostly around
10–20 kHz. The limitation of switching frequency is ST for traction is explained by
the fact of HV Si-IGBT are normally used to save cost. It results in more switching
losses or switching frequency limitations.
9.5.1 Requirements
The requirements for the dc–dc stage of the ST, independently of the adopted
approach (modular or nonmodular), are summarized as follows:
Overview and design of solid-state transformers 225
● High voltage capability in the MV side: Handle voltage level of around 6–25 kV.
● High current in the LV side: Current level in the range of 100–2,000 A is
expected [3].
● High voltage isolation: The MF/HF transformer employed has an isolation
requirement defined by the own MVAC level, i.e., between 10 and 15 kV,
regardless of the solution adopted for the dc–dc stage.
● Decoupling degree between the MV dc side and the LV dc side: Ability to
provide total decoupling between the MVAC grid and the LVAC grid, so that
the MV stage and LV stage converters can operate independently from each
other. It means that the MV ac–dc stage can operate and provide the required
ancillary services to the MV grid (e.g., provide reactive power), without dis-
turbing or being disturbed by the LV dc–ac stage operation. The same remark
is also valid for the LV dc–ac stage operation. To achieve such decoupling
degree, both dc links (MV and LV) must be maintained constant with minimal
oscillation voltage (i.e., 5% of the LVDC [3]), and the dc–dc converter should
guarantee the voltage regulation.
● Control of the power flow: Ability to control the power flow between the
LVDC link and MVDC link and manage the connection of loads of both sides
[4,5,8].
● Bidirectionality: Many publications agree that the bidirectional power flow
capability is required for the dc–dc stage [3–5,8,33]. On the other hand, this
requirement is questionable, because the reverse power is an exceptional case
in the distribution system. If economically viable, unidirectional solution could
also be accepted. Despite that, bidirectionality is considered as a requirement
in modern ST.
● dc Breaker function: Overload and short-circuit protection (working as a
dc breaker) of the possible load/source/microgrid connected to the LVDC link)
[3–5,8,33].
● High efficiency: The ST is not only intended to replace the LFT of the dis-
tribution system but also solve problems coming from the grid modernization
[3]. However, high efficiency is expected for the ST, in order to compete with
the LFT. The highest efficiencies obtained from an SST in traction applications
are around 95%–97%, for two processing stages (i.e., ac–dc and dc–dc) [14].
Taking this fact into account, an expected efficiency of 96% is defined for
the whole ST system, as shown in Figure 9.9(b). The MV stage can achieve
an efficiency around 98.6%–99.3% [14,34], when multilevel converters are
used, while the LV stage can offer an efficiency in the range of 98.8%–99.2%
[35–37]. The dc–dc stage should provide an efficiency between 97.5% and
98.6%, so that the entire system achieves the desired efficiency, as shown in
Figure 9.9. Thus, an efficiency goal above 97.5% can be defined.
dc–dc module
dc–dc module dc
Figure 9.9 (a) Equivalent single phase modular ST architecture adopting the
defined CHB, (b) expected efficiency for the whole system, as well as
for each power conversion stage ([Ref MV] ¼ [14,34], [Ref LV] ¼
[35–37]), (c) implementation of the basic module of the modular
dc–dc stage
before, this approach consists of several basic modules combined to build the entire
architecture, as shown in Figure 9.9(a). The basic modules are isolated dc–dc
converters, as depicted in Figure 9.9(c), and many topology options are available.
The topology choice of the basic module plays an important role in the ST design,
once the performance of the entire system depends on the efficiency of the indi-
vidual module. The specification of the dc–dc basic module is usually in the range
shown in Table 9.2. Considering this and the requirements described earlier, an
overview of the most suitable topologies for the basic modules is presented as
follows.
Among several options of the dc–dc converters, two families of converters are
pointed out as the most promising: the SRCs and the active bridges converters.
s1 s3 s1,s s3,s
1:n
iL Cr Rp
leak
Lp
leak
Lsleak Rsleak
Ci Co
Vi Vo
υp Lm Rm υs
s2 s4 s2,s s4,s
Parameter Definition
Angular switching frequency ws ¼ 2pfs
Angular resonant frequency wo ¼ 2pfo
Relative angular frequency w ¼ ws =wps ffiffiffiffiffiffiffiffiffiffi
Resonant frequency fo ¼ 1=2p Lr Cr
Quality factor Q ¼ 2pfo Lr =Ra ¼ 1=2pfo Cr Ra
Vi
υp 1
Q=1
2/f s
Vi 0.8
fo<fs
Gain (Vo/Vi)
iLr fo=fs
0.6
γ Q=2
0.4
Vcpk
ΔυCr = 2Vcpk 0.2
Q=3
υCr Q=4
fs = fo
Q=5
0
Vcpk 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Normalized frequency ( f = fo/ fs)
(a) (b)
Figure 9.11 Series resonant converter: (a) main waveforms of the SRC and
(b) normalized voltage gain (Vo =Vi ) of the SRC as a function of the
normalized operation frequency (fo =fs ), according to the quality
factor Q
and usually the switching frequency is used to control the output voltage of
the SRC. However, this strategy implies a complex control system and poor
utilization of the HFT, once it must be designed to operate for the minimum
frequency. Thus, it is not suitable for ST applications, in which the power level
is relatively high.
The most efficient operating point of the SRC is when it operates in the
DCM with the switching frequency (fs ) equal or slightly below the resonance
frequency (fo ), i.e., with unity gain. In this operation mode, the primary side
switches achieve ZVS and the output diodes achieve zero-current-switching.
Moreover, the converter has good transformer utilization and also low EMI
emission due to the smooth current shape (low di=dt). In this operation point, the
power flow is not directly controlled, and the power conversion between input
and output is defined by (9.2), where n is the transformer turns ratio. On the
other hand, it is not a disadvantage in ST application, once the system has
enough degrees of freedom to control the input voltage by using the front-end
rectifier (first stage).
V o ¼ n Vi (9.2)
MV side LV side
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffi
nPo =4Vo p=ðfs Lr Cr Þ Po =4Vo p=ðfs Lr Cr Þ
The most popular converter of the active bridge family is the dual active bridge
(DAB) and its topology is presented in Figure 9.12. The DAB converter is com-
posed of two active bridges, connected through the inductance LDAB and a trans-
former. Even though the DAB topology is similar to the SRC, its operation is
totally different. Other topologies of the active bridge converter family, such as the
three-phase DAB and the multiple-active-bridge (MAB), can also be adopted as
a basic module of the ST. In fact, the MAB converter has been intensively
230 DC distribution systems and microgrids
S1 S3 S1,s S3,s
Primary Secondary
bridge
LDAB bridge
iL
iL LDAB 1:n
Ci Co
Vi Vo
υp υs υp υs
S2 S4 S2,s S4,s
(a) (b)
Figure 9.12 Topology of the dual active bridge (DAB) dc–dc converter (a) and its
equivalent circuit (b)
s1 s1 s1
s3 s3 s3
S1s S1s S1s
S3s S3s S3s
υs υp
υp υp υp
υs υs υs j
iL iL iL
j D1 D2
D2
is1 is1 is1
υs1 υs1 υs1 D1
investigated for ST applications like in [42], and good results have been presented.
However, the DAB converter is analyzed next for the sake of simplicity.
When the input voltage and/or output voltage of the converter deviate sig-
nificantly from their nominal values (around 20%, according to [46,49]), the soft-
switching operation is normally lost. In this condition, the PSM is not advantageous
anymore [46]. To overcome this problem, alternative modulation strategies have
been proposed, like the triangular current mode (TCM) and the trapezoidal mod-
ulation (TPM), as shown in Figure 9.13(b) and (c), respectively. The TCM imposes
a triangular shape current on the transformer, while the TPM generates a trape-
zoidal current waveform on the transformer, as shown in (b) and (c), respectively.
In contrary to the PSM, the TCM and TPM control the duty-cycle for transferring
power between the input and output of the converter. These strategies can operate
with soft-switching independently from the input and/or output voltage deviation,
but with penalty on the peak value of the currents. To transfer the same amount of
power, the DAB converter presents higher peak values of currents, when TCM and
TPM are adopted in comparison with the PSM, because of the triangular or rec-
tangular shape of the current. It means that the rms values on the semiconductors
and transformers are also expected to be higher, implying on possible increase of
the conduction losses.
In ST application, the input and output voltage are typically constant without
large deviation of the nominal value. In this sense, the PSM is very advantageous,
and therefore the DAB converter is analyzed next using this strategy.
Considering the current waveforms shown in Figure 9.14, the inductor current
can be described by (9.5). In this equation, the initial values are defined by (9.6)
and (9.7), respectively.
8 vp þ nvs
>
< iL ðto Þ þ L
> t; to < t < t1
DAB
iL ¼ (9.5)
>
> v þ nvs Ts
: iL ðt1 Þ þ p ðt t1 Þ t1 < t < t0 þ
LDAB 2
232 DC distribution systems and microgrids
S1
S3
S1s
S3s
φ υs
υp
υp
υs
iL(to) iL(π)
iL
iL(0)
is1
υs1
Ts / 2
t0 t1 t2 t3 Ts
(π) (2π)
Figure 9.14 Voltage and current waveforms of the DAB converter operating with
PSM
1.2
dup d = 1.5
Hard switching
1
d = 1.25
0.6
d = 0.75
0.4 d = 0.5
0.2
ddown
0
0 π/6 π/4 π/3 π/2
Phase – shift (ϕ)
Figure 9.15 Output power of the DAB converter in function of phase-shift angle
j, considering several input–output voltage relation d
indicates the soft-switching region of the DAB converter, which is the desired
operation region to reduce losses.
8
> 3j
< dZVSðupÞ ¼ 1
2p (9.9)
> 1
: dZVSðdownÞ ¼
1 ð3j=2pÞ
Figure 9.15 shows the variation of the normalized power as a function of the
phase-shift angle for different values of d, in which the soft-switching boundaries
are highlighted. Note that the DAB converter operates in soft-switching for the
entire range of load if d ¼ 1, i.e., Vi ¼ nVo . Consequently, the DAB converter can
be properly designed, so that it operates with soft-switching for the whole load
range.
Figure 9.16 Dual active bridge: (a) topology of the active bridge and the current
and voltage waveforms on the semiconductors of the (b) MV bridge
and (c) LV bridge
Table 9.6 Current stresses on the semiconductor S1b (MV side) of the DAB
converter
Table 9.7 Current stresses on the semiconductor S1a (LV side) of the DAB
converter
average and rms currents on the channel and body diode of the semiconductor S1 a,
while Table 9.6 shows the current effort on the semiconductor of the LV side S1 b of
the DAB converter. The current stresses on the inductor regardless it is connected
to the LV or MV side are presented in Table 9.8.
As observed in these equations, the current stresses of all components are
somehow proportional to defining j. Furthermore, when the power flows from
the MV to the LV, most of the current flows through the channel of the MV side
semiconductors, whereas it flows mostly through the body diode on the LV side.
Therefore, it is very important to select a power semiconductor for the LV side with
low forward drop voltage for the body diode.
s1b s3b
iLb Lb 1 : n
Ci
Vb
υb
s2b s4b
s1a s3a
s1c s3c
iLc Lc La
Ci Co
Vc Va
υc υa
s2c s4c
s2a s4a
s1d s3d
iLd Ld
Ci
Vd
υd
s2d s4d
DC link control
d Phase
Vdc Σ PI shift
modulator
V*dc
The objective of the control is to ensure a stable bus regulation, with a good
immunity from the load variations. Figure 9.18 shows a simplified diagram of the
control. The normalized phase shift d ¼ ðj=2pÞ is the output of a PI regulator that
controls the dc voltage.
Considering the equation that regulates the power transfer (9.12), the current
that flows into the output port can be obtained by Idc ¼ Pdc =Vdc .
nVin Vdc
Pdc ¼ d ð1 2jdjÞ (9.12)
fsw Llk
Overview and design of solid-state transformers 237
Figure 9.19 Simplified model for the voltage control of the dc bus
–100
–90
Phase (°)
–135
–180
10–1 10 0 101 10 2 10 3 10 4 10 5
Frequency (Hz)
Output impedance
50
Magnitude (dB)
–50
–100
–150
90
45
Phase (°)
–45
–90
10–1 100 101 10 2 10 3 10 4 10 5
Frequency (Hz)
Figure 9.20 Frequency response of the voltage control and of the output
impedance
Figure 9.20 shows the frequency responses of the voltage control and of the
output impedance with the voltage control. As can be seen, the crossover of the
open-loop transfer function always happens at the maximum of the phase. Of
course, choosing a higher crossover frequency reduces the phase margin.
9.9 Summary
The SST has several advantages when applied in the modern electric distribution
grid. Among them, the availability of the dc link for connection of dc microgrid is
Overview and design of solid-state transformers 239
mentioned in this chapter. There are many different possibilities to implement the
architecture of the SST, and an overview and classification have been presented.
For electric distribution, the three power processing stages architecture is cited to
be the best choice, because it provides the decoupling between the MV side and LV
side and also providing dc link connectivity. Furthermore, the dc–dc stage plays an
important role on the SST architecture design, because it is responsible for the
major losses of the system, besides to be in charge for controlling the LVDC link.
Therefore, the SRC and the DAB have been pointed out as the most promising
choice. The SRC can provide a very high efficiency, but it is not able to control
the power flow and the regulate the LVDC link properly. The DAB converter, on
the other hand, can provide a high efficiency solution and provide full control. The
design aspects of the power stage and control stage of the DAB converter focusing
on the SST application has also been discussed in this chapter.
References
[22] M. K. Das, C. Capell, D. E. Grider, et al., ‘‘10 kV, 120 a SiC half H-bridge
power MOSFET modules suitable for high frequency, medium voltage
applications,’’ in 2011 IEEE Energy Conversion Congress and Exposition,
Sep 2011, pp. 2689–2692.
[23] L. Meysenc, P. Stefanutti, P. Noisette, N. Hugo, and A. Akdag, ‘‘Multilevel
ac/dc converter for traction application,’’ 2009, U.S. Patent US7558087.
[24] N. Hugo, P. Stefanutti, M. Pellerin, and A. Akdag, ‘‘Power electronics
traction transformer,’’ in 2007 European Conference on Power Electronics
and Applications, Sep 2007, pp. 1–10.
[25] D. Dujic, C. Zhao, A. Mester, et al., ‘‘Power electronic traction transformer-
low voltage prototype,’’ IEEE Transactions on Power Electronics, vol. 28,
no. 12, pp. 5522–5534, Dec 2013.
[26] J. Taufiq, ‘‘Power electronics technologies for railway vehicles,’’ in 2007
Power Conversion Conference – Nagoya, Apr 2007, pp. 1388–1393.
[27] J. Martin, P. Ladoux, B. Chauchat, J. Casarin, and S. Nicolau, ‘‘Medium fre-
quency transformer for railway traction: soft switching converter with high
voltage semi-conductors,’’ in 2008 International Symposium on Power Elec-
tronics, Electrical Drives, Automation and Motion, Jun 2008, pp. 1180–1185.
[28] M. Steiner and H. Reinold, ‘‘Medium frequency topology in railway
applications,’’ in 2007 European Conference on Power Electronics and
Applications, Sep 2007, pp. 1–10.
[29] M. Glinka and R. Marquardt, ‘‘A new ac/ac multilevel converter family,’’
IEEE Transactions on Industrial Electronics, vol. 52, no. 3, pp. 662–669,
Jun 2005.
[30] M. Glinka, ‘‘Prototype of multiphase modular-multilevel-converter with
2 MW power rating and 17-level-output-voltage,’’ in 2004 IEEE 35th
Annual Power Electronics Specialists Conference (IEEE Cat. No.
04CH37551), vol. 4, 2004, pp. 2572–2576.
[31] M. Liserre, M. Andresen, L. Costa, and G. Buticchi, ‘‘Power routing in
modular smart transformers: active thermal control through uneven loading
of cells,’’ IEEE Industrial Electronics Magazine, vol. 10, no. 3, pp. 43–53,
Fall 2016.
[32] L. F. Costa, G. Buticchi, and M. Liserre, ‘‘Quad-active-bridge as cross-link
for medium voltage modular inverters,’’ in IEEE Energy Conversion Con-
gress and Exposition (ECCE), Sep 2015, pp. 645–652.
[33] J. E. Huber and J. W. Kolar, ‘‘Solid-state transformers: on the origins and
evolution of key concepts,’’ IEEE Industrial Electronics Magazine, vol. 10,
no. 3, pp. 19–28, Sep 2016.
[34] L. Schrittwieser, M. Leibl, M. Haider, F. Thöny, J. W. Kolar, and T. B.
Soeiro, ‘‘99.3% Efficient three-phase buck-type all-SiC SWISS rectifier
for dc distribution systems,’’ in 2017 IEEE Applied Power Electronics
Conference and Exposition (APEC), March 2017, pp. 2173–2178.
[35] B. Benkendorff, F. W. Fuchs, and M. Liserre, ‘‘Simulated and measured
efficiency verification power circulation method of a high power low voltage
242 DC distribution systems and microgrids
10.1 Introduction
The implementation of active distribution networks aims to modernize and improve
the sustainability of power systems, in order to include distributed energy genera-
tion resources and also to address the growing presence of power electronics in
different applications. This will be done through the integration of communication
and control technologies for effective control and monitoring of the system, in the
form of active and flexible DC distributed networks. The deployment of these
active distribution networks improves the performance of the traditional electric
system in terms of cost, efficiency, space and reliability. Different factors moti-
vated studies and development of a wide variety of DC distribution systems during
the last years [1], and led to a regained popularity in DC energy systems. Some of
these reasons are the growing presence of DC-based consumer equipment, the
significant increase in the installation of renewable energy conversion systems and
their detrimental effects on the power quality in conventional AC grids.
Moreover, the successful introduction of transmission of electrical power using
high-voltage DC (HVDC) systems has further validated their enhanced efficiency,
higher power capacity and controllability when compared to their AC counterparts,
especially in long distance and underground or submarine transmission systems.
These features allowed important developments in HVDC transmission config-
urations over the last years [2–6]. Consequently, these improvements later extended
to distribution level, enabled by advances in power electronics, the relative
maturity of DC-based consumer appliances and their subsequent widespread
growth [7,8]. Currently, DC-based equipment is increasing its penetration in low-
voltage kilowatt-scale applications in many sectors: aerospace, automotive, data
1
Department of Electronic Engineering, Universidad Tecnica Federico Santa Maria, Chile
2
Faculty of Engineering and Applied Sciences, Universidad de los Andes, Chile
3
Faculty of Engineering, Universidad Catolica de la Santisima Concepcion, Chile
4
Department of Electrical and Computer Engineering, Ryerson University, Canada
246 DC distribution systems and microgrids
power transfer when a line fails or that allows the connection of the loads to two
regulated voltages [13]. In addition, the converters in bipolar systems can either
operate with relatively low-rated voltage, thereby securing converter cost compe-
titiveness, or be connected to higher AC voltages while maintaining the step-down
effort of the DC–DC stages. Efficiency is also enhanced in three-wire systems, as
the currents are reduced for a system with the same power rating [14].
On the other hand, grounding of DC systems is a complex issue and its design
involves several factors [11]. One of the features of a bipolar network is that it has
higher safety levels since the grounding connection is easier and simpler using the
neutral conductor [1]. Unlike unipolar systems, the lack of uncertainty in the pole-
to-ground voltages allows to clear faults easily and quickly [1,8,15]. However, if
bipolar systems present differences in the generation or power consumption of the
loads connected to each pole, voltage imbalance can occur, which can adversely
affect the receiving voltage end or the loads connected to it, if not addressed
properly [8,16,17]. The current circulating through the neutral conductor affects the
voltage balance of the system and also its efficiency, as during such scenarios the
losses in the feeder are increased [7].
For these reasons, balance control becomes essential in bipolar systems [10], as
there is no way to guarantee the loads will be identical in both of the buses of the
distributed system. Aiming in this direction, different balancing methods have been
proposed in the literature, which will be discussed in detail in the upcoming sec-
tions. Considering the tremendous potential that bipolar DC microgrids hold for the
development of DC distribution systems, the remainder of this chapter will provide
an overall view in the efforts being made in this architecture and the current
technologies existing in the literature.
The bipolar architecture for DC distribution networks offers interesting features over
the conventional unipolar counterpart. This structure is exhibited in Figure 10.1, where
it can be seen that the utility AC voltage is converted to DC with the use of a dis-
tribution transformer and an active rectifier, also called distribution converter. Then, at
DC level the system adopts a three-wire system that consists of the positive conductor
(p), the negative conductor (n) and the neutral conductor (z). As it has been demon-
strated at transmission level for HVDC, and despite being a more complex technical
solution, the aforementioned configuration presents clear advantages in terms of effi-
ciency, reliability, safety and transmission capacity when compared to conventional
two-wire systems. Advances in power electronics and their decreasing costs have
allowed an increased penetration of power converters in different applications, thus
enabling the consequent expansion of DC systems at distribution level [1].
Overall, the system exhibited in Figure 10.1 features several benefits. First, the
three-wire configuration resembles the traditional AC system in the sense that it
provides the connection to two different regulated voltages. The voltage between
the positive and negative poles is analogous to the line-to-line voltage, while the
248 DC distribution systems and microgrids
Wind
PV
LED lighting
Heating
DC AC DC
DC DC DC
p
MVAC Distribution n
transformer
DC DC DC DC
Distribution DC AC DC DC
converter
voltage of a pole with respect to the neutral connector is analogous to the phase-to-
neutral one [8]. This allows to accommodate a wide set of DGs and loads with
different voltage and power ratings combinations in a single DC network [18–20].
From an economical point of view, bipolar systems secure the cost competi-
tiveness of the converters interfacing the different stages of the network. This is
related with the reduction of the voltage ratings of the power electronic compo-
nents, the efficiency improvement due the reduction of the rated current, besides
the elimination of unnecessary DC–AC stages [14]. The latter is because most of
the stages involved in DGs are either DC-based, e.g., photovoltaic (PV) panels, fuel
cells and batteries or generate outputs with variable voltage/frequency, e.g., wind,
small hydro, wave/tidal energy conversion systems, hence require power electro-
nics devices to accommodate their output to network conditions [14]. In addition,
the presence of batteries and storage stages further enhances the benefits of DC
distribution, as it yields to greater efficiency improvements [21,22]. The result is a
substantial enhancement in the quality of the electric system and at the same time,
reduction of the costs when compared to conventional AC solutions [13].
The aforementioned resemblance with three-phase AC systems is also bene-
ficial for the migration process to bipolar DC networks. Considering a conventional
grounded three-phase system, which requires five conductors for its realization
(three-phase conductors, one neutral and the protective earth or grounding con-
ductor). These installed cables can serve in a retrofitted DC system that will be
superior in terms of power ratings capability and efficiency than the original one
[14]. Typically, low-voltage power cables can be found in different sizes and
ratings. As an example, multi-wire 230 V AC installations use 300 V/500 V 10 A
cables, following the structure in Figure 10.2(a). The system is then rated for
6.21 kW (assuming a power factor of 0.9). If such system would be replaced by
(a) a 325 V DC network or (b) a 325 V DC network, using the configuration
Bipolar-type DC microgrids for high-quality power distribution 249
S p p
R T PE n PE p z PE z
N n n
Figure 10.2 Use of cables in retrofitted systems or new DC systems: (a) AC cable;
(b) unipolar DC cable; and (c) bipolar DC cable
proposed in Figure 10.2(b) and (c), the power transfer capability will be enhanced
to 6.5 kW in both cases, with no considerations for reactive power and using the
same cables. However, bipolar DC system cables only have to withstand 10 A while
the unipolar one should be rated for 20 A. This illustrates some of the benefits of
implementing a bipolar DC microgrid from an existing AC installation.
Ip
Sa Sb Sc
Lg iga Vd1 C1
Iz
MVAC Vd2 C2
Sa Sb Sc
In
Ip
S Sb Sc
Lg iga a
Vd 1 C1
- - -
Sa Sb Sc
Iz
MVAC S S b̃ S c̃
L g igã ã
Vd2 C2
- - -
S ã S b̃ S c̃
In
exhibited in Figure 10.3 does it by connecting to the neutral point of the transfor-
mer with the DC bus midpoint. This additional current path allows the regulation of
the midpoint current, in order to prevent DC bus voltages from drifting and main-
tain their proper regulation. However, the generation of non-zero DC currents at the
AC side can lead to transformer saturation and thereby it should be strictly limited.
To avoid this issue, different approaches have been proposed in the literature
[7,8,16,23,31,32], both based in passive or active methods, the latter ones enabled
by power electronics. The main solutions will be covered in the upcoming sections.
Ip
Sa1 Sb1 Sc1
Vd1 C1
Sa2 Sb2 Sc2
Lg iga Iz
z
MVAC Sa1 Sb1 Sc1
Vd2 C2
Sa2 Sb2 Sc2
In
(a)
1
0.8
Load ratio ε
0.6
0.4
from the unbalanced loads. In this way, the 3L-NPC focuses on regulating the input
currents and power quality while the DC voltages remain unaltered. This leads to
higher current stresses on the balancing stages, so interleaving channels can be
promoted in order to reduce their ratings [37–39].
A different concept is to implement a coordinated balancing scheme, where
both the central converter and the balancing stages share the current redistribution
efforts [17,20]. The balancing principle is the same as with the 2L-VSC, and the
exceeding current should be redistributed by the means of a power converter stage.
The DC-link structure opens the possibility to three-level DC–DC stages as it will
be covered in the upcoming section.
Ip
Sa Sb Sc Sd
Lg iga Vd1 C1
Iz
MVAC
S¯a S¯b S¯c S¯d Vd2 C2
In
+ + + +
S1 Lb1 S1 S1
Vd1+ C1 Vd1+ C1 Lb1 Vd1+ C1 Vd1+ C1
Lb Cb Lb2
Vd Vd Cb + Vd +
Vd +
Cb S2 S2
S2 +
Vd2 + S1 Lb2 Vd2+ C2 Lb1 S2 Vd2+ C2
C2
Lb2 S2 Vd2 C2
– – – –
I pc
+ Sa Sb Sc
Vd1 C1 Lb
Izc +
Vo Co
+
Vd2 C2 S¯a S¯b S¯c
Inc
Ip
Sa Sb Sc +
Lg iga Vd1 C1
Iz
MVAC +
Lc S¯a S¯b S¯c Vd2 C2
In
the voltage balancing stage is to use additional channels and operate them interleaved.
This connection will reduce the high current ripple exhibited in the DC inductor Lb of
the buck/boost stage shown in Figure 10.6, as the current will be equally shared by the
additional channels. By means of this configuration it is possible to reduce the inductor
size, as the interleaved operation increases the output equivalent frequency. This leads
to a smooth output current, which allows the reduction in the output inductor by a factor
of 1/n, for an n-channel converter [24].
The circuit in Figure 10.8 can also serve towards the power balance of the
buses. The topology proposed in [7] regulates the positive and negative bar currents
in such a way that the neutral current Izc is almost zero. However, this is not
sufficient to guarantee a stable operation, as the exceeding power must be relocated
to the opposite pole and returned to the grid. The result is a control scheme based
on the sum–difference domain, that ensures balance for asymmetrical operation in
split-DC networks. As it will be shown in the next section, the resemblance with a
three-phase system motivated a control scheme based on the symmetrical compo-
nent method applied for grid-connected converters.
Ip I pc
S1 S1
Lb
Vd1 C1 Vd1 C1
S2 S2
Iz ib Izc
Vo Co
Lb S S3
3
Vd2 C2 Vd2 C2
S4 S4
In Inc
(a) (b)
The design of the core leads to zero-sequence inductances that are substantially
lower than the stationary frame ones, which are desired to be large in order to
reduce sustained AC currents given the connection. The structure of the inductor
enables a viable path for the returning DC currents, enhancing the controllability of
the converter and regulating the DC bus voltages regardless the load condition.
This approach offers a trade-off between having less power electronics and
increasing the losses in the coupled inductor showing promising results.
exists, the modulator takes the exceeding current from the bus lightly loaded and
supplies it to the other bus through the DC inductor.
The works in [38,39] use a different three-level DC–DC converter for feeding
high-power loads and also contribute to minimize power fluctuations in bipolar DC
systems. The power circuit is presented in Figure 10.10(b), where it can be seen
that the basic unit is composed by four switching devices and their corresponding
free-wheeling diodes, along with an output inductor and the common output filter
capacitor. Depending on the power level, more units can be connected in inter-
leaved configuration to minimize current stress in the switches and inductors [38].
One bipolar application of the aforementioned topology is interfacing an ESS
[39]. By replacing the additional balancing leg, this converter allows to accomplish
two simultaneous objectives: the main one is to perform the charge and discharge
of the storage element according to the energy management supervisor controller
of the DC network, but at the same time uses this process to relocate the power
consumption at DC level and complement the balancing task performed by the
distribution converter. Given the nature of the topology, the system is able to
guarantee stable operation with an ESS with reduced power ratings.
Alternatively, in [38], two interleaved channels of the converter in Figure 10.10(b)
are used to feed high-power loads. Besides enhancing the output power quality, reduce
the current stress and the filter size and its modularity, this three-level topology can also
assist the distribution converter to perform DC power balancing, allowing to improve
the overall system efficiency and the quality of the AC input currents. From the con-
trollability point of view, this stage can be either complementing the balancing efforts
carried by the central converter, or exclusively perform the balance in a distributed
manner between the remaining stages as suggested by the authors.
Go(S)
r1 + u1 r2 + u2 y2 y1
C1(S) C2(S) Go2(S) Go1(S)
– –
Inner loop
Outer loop
in such a way that it mitigates the effects of disturbances before they significantly
affect the system output y1 [40].
The main benefits of the cascaded control are obtained when when Go2 ðsÞ
presents significant non-linearities that limit loop performance or when Go1 ðsÞ
limits the bandwidth in a basic control loop (i.e., the presence of non-minimum
phase zeros and/or pure time delays). The first benefit, is explained because the
inclusion of C2 ðsÞ allows a high gain control, which tends to reduce the effect of
nonlinearities. The latter one, is because this inner controller pre-compensates the
effect of the disturbances that C1 ðsÞ has to deal with [40].
( )2
Vd
Notch +
filter
vgd
igd*+ Vd
Vd*+ – + + +
– +
PI PI vcd*
wLg dq
igd Vref S
igq vcq* PWM 3f -VSC
wLg abc Gating
signals
igq*+ – + – Lg
Vg
+ θg
PI vgq PLL
vgd
dq axis vgq dq
decoupling igd ig
igq abc
Grid
controller in order to enhance the DC-link voltage dynamics, which are regulated
by an outer proportional integral (PI) controller. Note that the feedback of the DC
voltage is filtered with a notch in 2 wg to minimize the impact of the rectification
ripple in the grid current. For the synchronization of the scheme, a phase-locked
loop (PLL) is employed to guarantee the unity power factor operation.
Vd*+ – Pg*+ hP
– Switching S
PI Pg Qg*+ hQ 3f -VSC
table Gating
–
qk signal
Qg
Lg
qg vg
PLL
Power ig
estimator
Grid
dq synchronous frame. The result is fast transient response and high static perfor-
mance, with no steady-state error, as a result of the DC nature of the controlled
quantities [29].
From Figure 10.13, it can be seen that the measured signals are converted to
the dq frame, which is aligned with the grid voltage using a PLL, hence the name of
the scheme. This allows to decouple the regulation of the active and reactive power
through the orthogonal components of the current, igd and igq , respectively.
For the regulation of the current components and the DC-link voltage, PI
controllers are used, leading to zero steady-state tracking errors and good dynamic
performance. Finally, their actuations are transformed back to a three-phase refer-
ence frame and passed to the modulator. This modulation stage can be either PWM
or SVM, for the generation of the gating signals. This control scheme can also be
applied to the NPC converter, by modifying the modulation stage accordingly and
also considering an midpoint balancing mechanism. Conventionally, a PI controller
is implemented in order to perform the partial balancing strategy, which regulates
the midpoint deviations depending on the selected modulation strategy [17].
2ε
Ip
i*b + db Sd
In Max PWM
–
Iz ib PI
sgn
Figure 10.15 Voltage balancing circuit controller for the additional NPC leg
Dp
Iz* + + I∑
* Ipc* + + +
– + –
Iz PID Ipc PI S
damp. T–1 + –
1+Do PWM VSC
–
Vo* + + I∆* *
Inc + +
– + – +
Vo PI Inc PI Dn
damp.
10.6 Summary
During the last decades, AC systems dominated the power transmission and dis-
tribution applications almost exclusively. However, a recent convergence of needs
originated in different sectors (renewable energy conversion, information tech-
nology and transportation) have accelerated the development of DC systems. Nowa-
days, DC systems are present at both transmission and distribution levels, offering
high-performance solutions with enhanced efficiency and reliability, besides reducing
the number of power conversion stages involved and uninterrupted power delivery.
For LVDC active networks, two kinds of architectures are possible: unipolar and
bipolar. Despite being a more sophisticated and technically complex solution, bipolar
structure provide several advantages over conventional unipolar ones. Higher avail-
ability, efficiency and flexibility are just a few advantages featured by bipolar systems.
A typical bipolar grid has certain similarities with a three-phase AC system, where
the phase voltage in an AC system is the analog to the voltage between one pole and the
neutral point in the bipolar LVDC grid, and the line-to-line AC voltage can be con-
sidered as the voltage between the positive and negative poles of the bipolar LVDC
grid. This resemblance with AC systems is beneficial for the operation, retrofitting and
also for developing the control schemes.
Bipolar-type DC microgrids for high-quality power distribution 263
However, if the loads connected to each pole of this LVDC grid are unba-
lanced, the system operates asymmetrically. This uneven power distribution along
with the dependency between the poles can be problematic if not taken care
properly. The main issues are the imbalances in the DC voltages, higher losses and
a decreased power quality of the system.
In order to solve these imbalance problems additional circuitry is required.
Voltage balancers and/or current redistributors are the key players that allow to deal
with asymmetrical operation and keep the system in a balanced and stable operation.
As it was discussed in the chapter, the converter topologies employed in bipolar
LVDC systems need to be bidirectional, in order to allow the reversible power flow
given the flexible structure of distributed networks.
This chapter presented a brief overview covering the different aspects of
bipolar LVDC networks. Distribution converter topologies, balancing stages and
also their control schemes are discussed in order to highlight the efforts being made
in this growing architecture.
Acknowledgement
The authors acknowledge the support provided by the CONICYT projects FON-
DECYT 11170774, AC3E (Basal/FB0008) and SERC Chile (FONDAP/15110019).
References
[1] D. Salomonsson. Modeling, Control and Protection of Low-Voltage DC
Microgrids. PhD thesis, Royal Institute of Technology, Sweden, Apr. 2009.
[2] C. W. Taylor and S. Lefebvre. HVDC Controls for System Dynamic
Performance. IEEE Trans. Power Syst., 6(2):743–752, May 1991.
[3] B. H. Bakken and H. H. Faanes. Technical and Economic Aspects of Using a
Long Submarine HVDC Connection for Frequency Control. IEEE Trans.
Power Syst., 12(3):1252–1258, Aug. 1997.
[4] L. Zhang, L. Harnefors, and H. P. Nee. Interconnection of Two Very Weak
AC Systems by VSC-HVDC Links Using Power-Synchronization Control.
IEEE Trans. Power Syst., 26(1):344–355, Feb. 2011.
[5] L. Zhang, L. Harnefors, and H. P. Nee. Modeling and Control of VSC-
HVDC Links Connected to Island Systems. IEEE Trans. Power Syst., 26(2):
783–793, May 2011.
[6] G. P. Adam, K. H. Ahmed, S. J. Finney, K. Bell, and B. W. Williams.
New Breed of Network Fault-Tolerant Voltage-Source-Converter
HVDC Transmission System. IEEE Trans. Power Syst., 28(1):335–346,
Feb. 2013.
[7] J. Lago and M. L. Heldwein. Operation and Control-Oriented Modeling of a
Power Converter for Current Balancing and Stability Improvement of DC
Active Distribution Networks. IEEE Trans. Power Electron., 26(3):877–885,
Mar. 2011.
264 DC distribution systems and microgrids
[8] Y. Gu, W. Li, and X. He. Analysis and Control of Bipolar LVDC Grid With
DC Symmetrical Component Method. IEEE Trans. Power Syst., 31(1):
685–694, Jan. 2016.
[9] H. Kakigano, Y. Miura, and T. Ise. Distribution Voltage Control for DC
Microgrids Using Fuzzy Control and Gain-Scheduling Technique. IEEE
Trans. Power Electron., 28(5):2246–2258, May 2013.
[10] H. Kakigano, Y. Miura, and T. Ise. Low-Voltage Bipolar-Type DC Micro-
grid for Super High Quality Distribution. IEEE Trans. Power Electron.,
25(12):3066–3075, Dec. 2010.
[11] T. Dragicevic, X. Lu, J. C. Vasquez, and J. M. Guerrero. DC Microgrids –
Part II: A Review of Power Architectures, Applications, and Standardization
Issues. IEEE Trans. Power Electron., 31(5):3528–3549, May 2016.
[12] Directive 2014/35/EU of the European Parliament and of the Council of
26 February 2014 on the Harmonisation of the Laws of the Member States
Relating to the Making Available on the Market of Electrical Equipment
Designed for Use within Certain Voltage Limits (recast) (Text with
EEA Relevance). Official Journal of the European Union, L 96:357–374,
Mar. 2014.
[13] T. Kaipia, P. Salonen, J. Lassila, and J. Partanen. Possibilities of the Low
Voltage DC Distribution Systems. In Nordic Distribu. and Asset Manag.
Conference (NORDAC), Stockholm, Sweden, Aug. 2006.
[14] D. Salomonsson and A. Sannino. Low-Voltage DC Distribution System for
Commercial Power Systems With Sensitive Electronic Loads. IEEE Trans.
Power Del., 22(3):1620–1627, Jul. 2007.
[15] H. Kakigano, Y. Miura, T. Ise, and R. Uchida. DC Voltage Control of the
DC Micro-grid for Super High Quality Distribution. In Power Conversion
Conference – Nagoya, 2007. PCC ’07, pp. 518–525, Apr. 2007.
[16] Byung-Moon Han. A Half-Bridge Voltage Balancer with New Controller for
Bipolar DC Distribution Systems. Energies, 9(3): 182, Mar. 2016.
[17] S. Rivera, B. Wu, S. Kouro, V. Yaramasu, and J. Wang. Electric Vehicle
Charging Station Using a Neutral Point Clamped Converter with Bipolar DC
Bus. IEEE Trans. Ind. Electron., 62(4):1999–2009, Apr. 2015.
[18] J. Lago, J. Moia, and M. L. Heldwein. Evaluation of Power Converters to
Implement Bipolar DC Active Distribution Networks – DC-DC converters.
In 2011 IEEE Energy Conversion Congr. and Expo. (ECCE), pp. 985–990,
Sep. 2011.
[19] X. Zhang, C. Gong, and Z. Yao. Three-Level DC Converter for Balancing DC
800-V Voltage. IEEE Trans. Power Electron., 30(7):3499–3507, Jul. 2015.
[20] F. Wang, Z. Lei, X. Xu, and X. Shu. Topology Deduction and Analysis of
Voltage Balancers for DC Microgrid. IEEE Trans. Emerg. Sel. Topics Power
Electron., 5(2):672–680, Jun. 2017.
[21] E. Rodriguez-Diaz, M. Savaghebi, J. C. Vasquez, and J. M. Guerrero.
An Overview of Low Voltage DC Distribution Systems for Residential
Applications. In 2015 IEEE 5th International Conference on Consumer
Electronics – Berlin (ICCE-Berlin), pp. 318–322, Sep. 2015.
Bipolar-type DC microgrids for high-quality power distribution 265
11.1 Introduction
Similar to utility grid and other terrestrial microgrids, aircraft electrical power system
(EPS) has its own power generation, distribution, utilization and energy storage. The
rapid development of power electronics technology has allowed the converters to
operate at DC voltage levels required for transmission, distribution and consumption.
Today there is a clear tendency that high voltage (HV) and low voltage have seen the
proliferation of DC systems and its implementation for transmission and distribution
of electricity in aircraft electrical systems.
This chapter will provide an overview of aircraft DC microgrids. Section 11.2
introduces the aircraft EPS, covering the topics from power generation, distribution
and utilization. Section 11.3 reviews aircraft electrical system standards and high-
lights the power quality and power factor requirement for aircraft applications.
Section 11.4 presents the concept of aircraft electric starter/generator system.
Control and stability analysis of aircraft DC microgrids are performed in Sections
11.5 and 11.6, respectively.
1
Department of Engineering Science, University of Oxford, UK
2
Department of Electrical and Electronic Engineering, The University of Nottingham, UK
268 DC distribution systems and microgrids
Jet fuel
Propulsion thrust
Jet fuel
Propulsion thrust
Engine-driven
generators
in the utility grid, the EPS architecture with DC distribution is considered as one of the
most promising candidates for MEA due to the potential advantages such as higher
efficiency compared to AC, less weight of harness, absence of reactive power related
issues and also reduced costs. Compared with existing DC distribution systems in
aircraft, the future architectures will move towards HV distribution.
Table 11.1 Power generation type for recent civil and military aircrafts [4]
operated from a fixed frequency supply. Having many distributed power converters
gives more options for a safe aircraft system design as redundancy can be built-in at
the systems level, avoiding any single points of failure within the design [1]. In new
aircrafts such as the Boeing B787, Airbus A380 and A350, this VSVF architecture
has been employed [5]. In these aircrafts, while the voltage is regulated at either
115 V (A380) or 230 VAC (B787), the frequency changes with the engine speed
and typically varies between 360 and 800 Hz [6].
Figure 11.3 shows the distribution system in B787. The 230 VAC bus,
generated by the variable speed synchronous generator, is then transformed and
rectified by an auto-transformer rectifier unit (ATRU) to supply 270 VDC bus.
The distribution buses also supply 115 VAC bus bars via auto transformer units,
and 28 VDC buses via transformer rectifier units (TRUs). This permits more con-
ventional consumers such as window heat, APU fuel pump and avionics compo-
nents, to be powered from less conventional 230 VAC supply. The availability of
the 115 VAC and 28 VDC buses therefore avoids the need to develop specialized
power conversion for each of the client loads. The distribution architecture also
furnishes 230 VAC direct to certain users for which no transformation is necessary.
These loads include the wing anti-ice heater mats, environment control system
(ECS) recirculation fans and cargo heaters. To provide a temporary source of
emergency power for essential loads, and to provide a source of energy for starting
the APU, the B787 has two 28 V batteries.
Aircraft DC microgrids 271
28 V
BAT
28 VDC bus
28 V BAT
EMERG LV
Emergency
loads
28 V
TRU loads
115 VAC
bus AC
ATU
loads
230 VAC bus
± 270 V DC
bus
ATRU
G1 ECS1
WIPS1
WIPS2 ± 270 V DC
G2 bus
ECS2
ATRU
28 VDC bus
28 V 28 V
TRU loads ESS
Figure 11.3 Power generation and distribution system for Boeing 787 having twin
generators
272 DC distribution systems and microgrids
AC
Power electronic loads 28 V BAT
converter 1
G1
WIPS1 28 V BAT
EMERG LV
ECS1
Emergency
loads
270 VDC
PM EMA6 28 V
ESS
270 VDC loads
bus
28 V ESS
Power electronic
PM EMA2
converter 2
G2 ECS2
PM EMA1
WIPS2
115 VAC
bus
AC
loads
approaches mainly for safety reasons. However, if the approach is properly estab-
lished, one can assume that the appropriate safety requirements can be met in
future. In return, sources paralleling idea will lead to (1) possibility of power
sharing among different sources with minimization of total weight of generation
systems, (2) ease of energy management and power flow control, (3) convenient
integration with power sources of different nature (fuel cells, energy storage devi-
ces, batteries, etc.), (4) improved availability of electrical energy onboard.
Fuel Power
pumps 10 kW+ consumption
Flight in an aircraft
controls 3–40 kW
Landing
gear 25–70 kW
Air conditioning
4 × 60 kW+
(ECS)
250 kW+
WIPS
Loads
Table 11.2 Current harmonic limits for single phase and balanced three phase
system in aircrafts [13]
order to meet the power quality requirement. Recent research showed that, to meet
the low input harmonic current requirements, the converters with active front-end
are attractive solutions for interfacing with the AC bus [16]. However, to effec-
tively control the input current harmonics to meet stringent specifications remains a
major challenge. It would be more difficult to meet the entire requirement simul-
taneously without reducing the other performances, such as converter’s voltage
transfer ratio, input power factor and robustness against unbalanced and/or
distorted voltage input. On the other hand, the requirements also present a major
design constraint for the filter design of the converters. A relatively small
second-order input filter is typically employed in the converters to smooth out the
switching ripple current at the power input lines. Alternative topologies for active
front-side control and input power filter designs to further reduce the weight and
size are also a big challenge.
MIL-STD-407F, US Military Standard, defines a standardized power interface
between a military aircraft and its equipment and carriage stores, covering such
topics as voltage, frequency, power factor, electrical noise and abnormal conditions
(overvoltage and undervoltage), for both AC and DC systems. Figure 11.6 shows
the limit for 115 V AC, 400 Hz and 270 V DC, 28 V DC. The envelope of the
nominal voltage transients is shown in Figure 11.6. It can be seen that in 270 V DC
system, the tolerable steady-state range is between 250 and 280 V, which indicates
the possibility of droop control (will be discussed in the following section).
The voltage at the transient can go up to 330 V and down to 200 V. Similarly, the
steady-state limit of 28 VDC system is 22–29 V. The overvoltage and under voltage
for 270 and 28 V DC system are illustrated in Table 11.3.
The high fundamental frequency (400 Hz in CF EPS and 360–800 Hz in
VF EPS) is also a challenge for aircrafts. Achieving low input current distortion and
unity input power factor at such high line frequencies require much wider control
bandwidth compared to the industry 50/60 Hz systems. Moreover, VF EPS will
350 330 50
50
300
280
DC voltage, V
0.0125 s
DC voltage, V
40
250
250 29
30
0.0825 s
0.015 s 22
200 20 18
150 10
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.0025 0.05 – 0.075 0.1 0.125 0.15 0.175
(a) Time, s (b) Time, s
Figure 11.6 Envelope of normal voltage transients for (a) 270 V DC system and
(b) 28 V DC system
276 DC distribution systems and microgrids
Table 11.3 Voltage limits for 270 and 28 V DC system in aircraft DC MGs
have more challenges for power factor correction (PFC). It is because that, the
cable impedance will vary with the AC frequency in VF EPS. As a result, the PFC
needs to provide reactive power varying with the frequency. Another reason is that,
because the frequency range of VF EPS is typically 360–800 Hz, most of which is
higher than CF, the voltage drop of cables in VF EPS is higher than in the system of
CF EPS.
VAC,
Control constant frequency
system
Speed
VAC Control
variable frequency signals
Bus
Load
Bi-directional
Aircraft Starter/
power Load
engine generator Variable converter
frequency Load
Variable speed
VC Idc
Cdc
SM PMM DC bus
–
Iqref
PI
VC mag
Idref –
PI
Figure 11.8 Inner current loop control of the PMM-based S/G [19]
modulation indexes for pulse width modulation. Then the core system can be fully
controlled by using both dq current demands.
Figure 11.9 shows the outer loop of the S/G. Since the designed PMM is for
high speed operations, the flux weakening controller needs to be taken into account
when the machine speed exceeds the base speed. For the starter mode, speed con-
troller and flux weakening controllers are employed to generate the q-axis current
reference and d-axis current reference. When the stator voltage is less than its
278 DC distribution systems and microgrids
Flux weakening
VC max controller
Idref
1/√3
PI
VC –
wm Inner loop
Speed
controller max = √Imax–Id
Iqref 2 ref 2 as shown in
Figure 13.8
wmref –
PI S Iqref
G
Imax Idcref – Iq max TSR
Edc
–Imax PI limiter
Idc
Idc
controller
Figure 11.9 Outer loop control for Aircraft Electrical Generation with Active
Rectification Technology (AEGART) S/G system
maximum value, the output of the flux weakening controller is saturated to zero.
The controller starts to work when the stator voltage exceeds the maximum voltage
limit. With the outer voltage regulation loop (Vc loop), the flux-weakening control
is fulfilled by injecting the negative d-axis current. For the generator mode, the
conventional PI controllers are used to deflux the machine (d-axis) at high speed
and control the output DC current (q-axis). The stator current references in the d
and q axes are obtained from the output of the flux weakening controller and IDC
controller respectively. The reference of the AC voltage (Vc) is dependent on the
ref
DC voltage. The DC current reference (IDC ) is determined by the desired droop
characteristic.
Figure 11.10 shows a lab prototype for the electric S/G developed in Not-
tingham [20].
Aircraft DC microgrids 279
Power
Power electronic
G1 electronic
converter 1 HVDC bus G3
converter 3
Cable 1
Cable 3
C1
Power electronic C3
G2 converter 2 RL
Cable 2
Convention
C2 al loads
C4
Battery-converter
system
Network
{ {
(a) (b)
V–P droop
V–I droop I0
I0 characteristic
characteristic
*
V *
V* * P V dc
Io V dc V dc
V dc P
I
(c) (d)
Figure 11.12 Droop characteristic employed in active front ends (AFEs). (a) I–V
droop, (b) P–V droop, (c) V–I droop, (d) V–P droop [19]
Aircraft DC microgrids 281
1 1
0.8 0.8
Voltage
Voltage
0.6 0.6
0.4 0.4
V–I curve for P–V droop V–P curve for I–V droop
0.2 0.2
-----Equivalent I–V droop -----Equivalent P–V droop
Figure 11.13 Similarity validation of P–V droop and I–V droop [26] in DC MGs.
(a) Power voltage droop; (b) Current voltage droop
The voltage-mode approach employs the V–I droop characteristic which uses
the measured branch current to generate the terminal voltage reference. The control
scheme of the voltage-mode droop-controlled AFE is shown in Figure 11.14. As
expressed in (1), the DC voltage reference is generated according to the branch
output DC current using the V–I droop characteristic.
VDC ¼ Vo kIDC (11.1)
The current-mode approach uses the measured voltage to calculate the desired
injecting DC current. The current-mode droop control scheme is shown in
Figure 11.15, with the current reference derived from the I–V droop characteristic,
based on the DC voltage measurement:
Vo VDC
IDC ¼ (11.2)
k
where Vo is the nominal bus voltage; k is the droop gain; VDC is the DC voltage
measurement; IDC is the generated DC current reference.
Figure 11.16(a) and (b) shows the primary control for current-mode and
voltage-mode droop controlled systems, respectively.
I–V droop Idc
characteristic
Idc controller
I
Vdc Idcref Iqref Inner loop
V PI Figure 13.8
Vdc
V0 Idc —
DC bus
Module 1 I1
Cable
Idc controller
Vo I1ref Inner current
1/k1 PI ref
loop V1
I1 Iq1 Cable
Module n In
Cable
Idc controller
Inref Inner current
1/kn PI loop Vn
Vo In Iqnref Cable
{
(a)
DC bus
Module 1 I1
Cable
Vdc controller
ref Inner current
Vo V1
PI
loop V1
V1 Iq1ref Cable
k1
{
Module n In
Cable
Vdc controller
Vo Vnref Inner current
PI loop Vn
Vn Iqnref Cable
kn
{
(b)
Figure 11.16 Primary control for AFEs-based aircraft DC MGs (a) current-mode
droop control and (b) voltage-mode droop control
Aircraft DC microgrids 283
vdc
OP 1_ne w
OP2_new
vo
Secondary
Primary OP1
OP2
V1 I1
Cable
Vb
Ii IL
Cable
Vn In
Load
Cable
As one can see from Figure 11.18, at steady state, the bus voltage can be
expressed as follows:
V b ¼ V o I 1 ðk 1 þ R 1 Þ ¼ V o I 2 ðk 2 þ R 2 Þ ¼ ¼ V o I n ðk n þ R n Þ
(11.3)
The total load current can be written as
X
n
1
I L ¼ I 1 þ I 2 þ þ I n ¼ ðV o V b Þ (11.4)
i¼1
ki þ R i
V0 kt
ki
k1 k2
I
Figure 11.19 Relationship between global droop gain and individual droop gain
in DC MGs
Aircraft DC microgrids 285
added to the terminal voltage reference for each module which can be expressed as
follows
DV ¼ ILt kt (11.7)
where ILt is the total load current and kt is the global droop gain. It is easily
implemented since only the load current needs to be measured and no DC voltage
controller is in need. If multiple loads are in presence and distributed among the
EPS, the total load current measurement can be obtained at the main feeders (if this
exists) which supplies the power to all loads.
However, the application is limited to small-scale DC MGs, i.e., the converters
need to be close to the main bus and an additional current sensor need to be
installed at the main bus to measure the total load current. It is not applicable for
large-scale MGs, where the converters and loads are distributed among the net-
work, and it is not practical to install a current sensor at the main bus for the total
load measurement. In this sense, a communication line is still required for the
information exchange, as shown in Figure 11.20.
Control and stability are two important topics that interest researchers in aerospace
engineering. Typical loads within aircraft DC microgrids are tightly controlled by
power electronic converters and can often behave like constant-power loads
(CPLs).
Figure 11.21 shows an example of a tightly controlled motor drive system,
which behaves as a CPL. A DC/AC inverter drives an electric machine and tightly
regulates the rotating speed. Therefore the speed (w) is almost constant under this
rigorous regulation. Since the rotating load has a one-to-one torque–speed char-
acteristic, for every speed, there is only one corresponding torque. Hence, for a
given constant speed (w), torque (T) is constant and hence power (the multi-
plication of speed and torque) is also constant. Thus, the tightly controlled DC/AC
inverter and its associated motor drive present a CPL characteristic.
CPLs have negative incremental impedance characteristics which can lead to
interaction problems among subsystems, deteriorating system stability and also
power quality [35,36]. Thus, this undesirable destabilizing effect induced by CPLs
is a great concern for the MEA EPS. The candidate architecture should be carefully
examined for stability in order to guarantee safe EPS operation for a wide range of
operation scenarios.
Two types of stability analysis can be remarked here: large signal and small
signal analysis. Large signal stability analysis investigates the ability of a power
system to maintain stable operation and reach acceptable steady-state operating
points when subjected to large disturbances. The Lyapunov-based method has been
popular in terms of large signal stability analysis, where non-linear mathematical
modelling is used instead of small signal linearization [37,38]. The main advantage
of Lyapunov-based approaches is that a developed Lyapunov function allows the
286 DC distribution systems and microgrids
DC bus
ILt Module 1
I1
kb Cable
Vdc controller
Vo V1ref Inner
PI current loop V1
Iq1ref Cable
Droop k V1
1 {
ILt Module 2 I2
kb Cable
Vdc controller
ref Inner
Vo V2
PI ref
current loop V
Iq2 2 Cable
Droop k2 V2
{
ILt Module n In
kb Cable
Vdc controller
ref Inner
Vo Vn
PI current loop Vn
Iqnref Cable
Droop kn Vn
{
(a)
DC bus
n Module 1
idc1 ∑ idcj ILt kb Vdc controller
I1
Cable
j=1
Vo V1ref Iq1ref Inner
PI current loop V1
idcn Cable
Droop k1 V1
{
Communication line
idc1
n Module 2
∑ idcj ILt kb Vdc controller
I2
Cable
Cable
j=1
idcn Vo V2ref Iq2ref Inner
PI current loop V2
Cable
Droop k2 V2
{
idc1
n
ILt k Module n In
∑ idcj b
Vdc controller Cable
idcn
j=1
Vo Vnref Iqnref Inner
PI current loop Vn
Cable
Droop kn Vn
{
(b)
Speed
Speed Regulator to
measurement
reference control the
speed
DC bus
ω
DC/AC Motor
inverter
TL = constant
ω = constant
Pin = constant
P = ω TL = constant
Vbus
Source Load
Vin Vo
converter converter
Zo_s Zin_L
source Zo_s subsystems. The input of the full system is Vin and output is Vo. The
voltage at the breaking point for the subsystem is Vbus. The definition of the source
converter is the converter which regulates the bus voltage, and the load converter is
responsible of current control.
The total input-to-output transfer function of the cascaded subsystems can be
written by
Zin L Vin 1
Vo ¼ Vin ¼ (11.8)
Zin L þ Zo s Zin L 1 þ TMLG
where the ratio Zo_s/Zin_L, which is often referred to as the minor loop gain, TMLG
can be given by
Zo s
TMLG ¼ (11.9)
Zin L
If each of the two subsystems is individually properly designed with good dynamic
performance, the influence of their interaction can then be studied by looking into
the minor loop gain. In particular, in order to preserve the stability, it is mandatory
that minor loop gain meets the Nyquist stability criterion [45]. It should also be
noted that if the detailed information about source and load systems is not available
and the respective impedances cannot be analytically constructed, then they should
be measured online. Since the source and load converter are assumed to be
stable standalone, the minor loop gain term in (11.9) is the one responsible for
stability.
In the specific context of aircraft power systems, so far, several publications
have discussed the system stability in the MEA electric power systems. The sta-
bility of a switched reluctance motor-based 270 V DC power system has been
analysed in [46]. A permanent magnet synchronous generator-based aircraft DC
MG is investigated in [47–49], and the influence of parameter variation on system
stability is presented. Components (R, L, C parameters), operating points (voltage,
power) and control parameters (controller gains, droop gains, etc.) will influence
the system stability. Take the droop gain as an example, the Bode plot of the source
and load impedance in the voltage-mode droop-controlled system is shown in
Figure 11.23. The source impedance magnitude in the low-frequency range
increases with an increase in the droop gain, whilst the load impedance magnitude
decreases. Therefore, the droop gain not only affects the power sharing ratio but
also the system stability.
This chapter provides an overview of architecture, power quality, control and sta-
bility in aircraft DC microgrids. First the evolution of aircraft DC microgrids
including power generation, distribution, utilization and energy storage is
reviewed. Then the chapter presents the power quality requirement in aircraft DC
microgrid including the AC current harmonics and DC voltage envelope. As an
Aircraft DC microgrids 289
Bode diagram
40
Load impedance
20
Magnitude (dB)
0
k=2
–20 k=1
k = 0.5
–40
Source impedance
–60
360
Source impedance
Phase (deg)
0
Load impedance
–360 –2
10 10–1 100 101 102 103 104
Frequency (Hz)
Acknowledgement
The authors acknowledge the support from CleanSky JTI Projects, FP7 European
Integrated Projects-http://www.cleansky.eu.
References
[1] P. Wheeler and S. Bozhko, ‘‘The more electric aircraft: technology and
challenge,’’ IEEE Electrif. Mag., vol. 2, no. 4, pp. 6–12, Dec. 2014.
290 DC distribution systems and microgrids
power loads and variable power loads—a review,’’ IEEE Trans. Power
Electron., vol. 27, no. 4, pp. 1773–1787, Apr. 2012.
[43] M. Cespedes and J. Sun, ‘‘Impedance modeling and analysis of grid-
connected voltage-source converters,’’ IEEE Trans. Power Electron., vol. 29,
no. 3, pp. 1254–1261, Mar. 2014.
[44] A. Radwan and Y. Mohamed, ‘‘Assessment and mitigation of interaction
dynamics in hybrid AC/DC distribution generation systems,’’ IEEE Trans.
Smart Grid, vol. 3, no. 3, pp. 1382–1393, Sep. 2012.
[45] R. D. Middlebrook, ‘‘Input filer consideration in design and application of
switching regulators,’’ in Proc. IEEE IAS Annu. Meet., Chicago, USA, Oct.
1976, pp. 336–382.
[46] L. Han, J. Wang, and D. Howe, ‘‘Small-signal stability studies of a 270V DC
more-electric aircraft power system,’’ in Proc. Power Electronics Machines
and Drives 2006. The 3rd IET International Conference (PEMD), Apr. 2006,
pp. 197–201.
[47] K.-N. Areerak, T. Wu, S. V. Bozhko, G. M. Asher, and D. W. P. Thomas,
‘‘Aircraft power system stability study including effect of voltage control
and actuators dynamic,’’ IEEE Trans. Aerosp. Electron. Syst., vol. 47, no. 4,
pp. 2574–2589, Oct. 2011.
[48] K.-N. Areerak, S. V. Bozhko, G. M. Asher, L. De Lillo, and D. W. P. Thomas,
‘‘Stability study for a hybrid AC–DC more-electric aircraft power system,’’
IEEE Trans. Aerosp. Electron. Syst., vol. 48, no. 1, pp. 329–347, Jan. 2012.
[49] F. Gao and S. Bozhko, ‘‘Modeling and impedance analysis of a single DC
bus based multiple-source multiple-load electrical power system,’’ IEEE
Trans. Transp. Electrif., vol. 2, no. 3, pp. 335–346, Sep. 2016.
This page intentionally left blank
Chapter 12
Shipboard MVDC microgrids
Dong-Choon Lee1, Yoon-Cheul Jeung1, Dinh Du To1,
and Duc Dung Le1
12.1 Introduction
Electric power was successfully applied to marine vessels in 1880s. Due to the
advance of technology in batteries, motors, and engines, the primitive electric
devices were installed on the ship, and in the early twentieth century the electric
propulsion system was developed [1]. The electric propulsion has a lot of advan-
tages such as better boarding comport, design flexibility, low vibration and noises,
low fuel consumption and emission, etc.
In electric propulsion systems, the propulsion motors are supplied by the dedi-
cated generators and the other loads are supplied by the auxiliary generators. To avoid
the inefficient structure of system, multiple gen-sets are connected to one common
bus, which supply propulsion and service loads together. This system is called
as integrated propulsion system (IPS) [2]. When the load power demand in the IPS is
low, some of the prime movers are operated at high output power region and the others
are deactivated, which increases the overall system efficiency. Queen Elizabeth II
launched in 1987 was the first diesel-electric ship that embeds the IPS [3].
Due to the evolution of the power electronics technology, the concept of all-
electric ship (AES) has been established, where, needless to say about propulsion
and service loads, even the loads using hydraulic and compressed air power are
replaced by electric devices. In 2015, the world’s first purely battery-driven car and
passenger ferry Ampere was launched in Norway [4].
So far, the AC power distribution system has been commonly used in ship-
board power systems since the AC motors are mostly employed for electric pro-
pulsion. Recently, however, the shipboard medium voltage DC (MVDC) power
system is attracting a great attention since it has lots of advantages such as smaller
high-frequency transformer, easy connection and disconnection, no harmonic and
unbalance problems, no need of synchronization, no reactive power, high effi-
ciency, flexible design in space, easy monitoring of power quality, fuel saving by
variable speed operation, and so on [5].
1
Department of Electrical Engineering, Yeungnam University, South Korea
296 DC distribution systems and microgrids
In this chapter, the shipboard MVDC microgrid system is introduced. First, the
architecture and components of the system are described, and then the modeling
and control are dealt with. Next, power management based on hierarchical control
is described and then fault and protection schemes are discussed. Some simulation
results for the case study are provided.
Energy Active
Fuel cells filter Pulsed load
storage Ship to grid
PM
SG1 GT1 DE1
SG1 ESS
M1
CB CB CB
Medium voltage DC bus CB
CB CB CB CB CB CB
CB CB CB CB CB
CB
Zone 1 Zone 2 Zone 3 Zone 4 Zone 5 CB
load load load load load
CB CB CB CB CB
CB CB CB CB CB CB
CB
CB CB CB
M2 Pulsed
PM load
DE2 GT2 SG2
SG2
Figure 12.2 Ring bus structure of shipboard MVDC power system with zonal
distribution [5]
expected that fuel consumption and emission can be reduced by up to 20% and that
footprints and weight of electrical equipment can be reduced by up to 30% com-
pared with the AC distribution systems [6].
The SG is the most popular machine of which power rating is ranged from a
few kW to tens of MW. The stator of the SG has three-phase windings and the rotor
has a field winding for DC excitation. The SG produces AC voltages at 50 or 60 Hz
frequency at the rated speed. The shaft speed is typically ranged from 600 or
720 rpm to 3,000 or 3,600 rpm, respectively [8].
The stator of the PMSG is similar to that of the SG, but the rotor has permanent
magnets instead of windings. So, the efficiency of the PMSG is higher since the
copper loss of the rotor is eliminated. Furthermore, the weight of the PMSG is
reduced by up to 30%. There is a commercial product which has 3,000 rpm and
8 MW [9].
Recently, application of superconductor SGs is gaining attention due to high
efficiency and compact size [2].
● AC/DC converters
In shipboard MVDC power systems, the output voltage of the main generators is
connected to the DC bus through the AC/DC converters that are diode rectifiers or
phase-controlled converters with silicon-controlled rectifier (SCR). In the case of diode
rectifiers, if the SG is used, its output voltage is controlled by an exciter, as shown in
Figure 12.3(a). If the PMSG is used, however, the back-end DC/DC converter is
required to match the MVDC bus voltage, as shown in Figure 12.3(b). If the SCR
rectifier instead of diode rectifier is employed, the back-end converter is not necessary.
On the other hand, multilevel converters have been so far used for medium
voltage applications, in which neutral-point clamped (NPC), flying capacitor,
cascade-H bridge converters are included. Recently, the modular multilevel converter
(MMC) topology has been applied to the high voltage DC (HVDC) and MVDC
systems, as shown in Figure 12.4. Due to the modular structure, the voltage rating of
MMC can be expanded easily.
● DC/DC converters
The DC/DC converter is required for connecting the ESS to the MVDC bus. A dual
active bridge (DAB) DC/DC converter is popularly used since it has advantages of
bidirectional power flow, high efficiency, modular structure, isolation, and soft
switching. By a multiple connection of the single DAB, the power capacity of the
converter can be expanded easily, which is shown in Figure 12.5 [13].
Similarly, modular multilevel DC/DC converters are also a candidate for high
power applications [14].
● DC/AC converters
In electric propulsion ships, DC/AC converters, that is, inverters are employed to
feed the propulsion motors and other AC loads. The two-level inverter is sufficient
for small pump and fan motor drives and hotel loads. For propulsion drives,
DC-DC
Rectifier converter DC bus
Rectifier DC bus P
6-Phase
wound-field 12-Phase
P
SG PMSG
+ Prime PM
+
Prime
SG
mover
Gear − mover SG −
box
N
Exciter
N
(a) (b)
Figure 12.3 AC/DC converter for MVDC shipboard system: (a) Wound-field SG
and (b) PMSG
300 DC distribution systems and microgrids
DC bus
+ +
− − +
+
LVDC MVDC
12.3.1 Gen-sets
The gen-set consists of a prime mover and a SG. A classical gen-set is operated at a
constant speed to generate a fixed electrical frequency and output voltage. Although
such a gen-set is of low cost, its fuel efficiency is low at light load conditions, and the
output voltage and frequency are fluctuated when the load varies. If a variable speed
operation scheme is applied, a significant amount of fuel saving is achieved. The
optimal speeds of DEs can be derived from the specific fuel consumption curve [18].
Figure 12.6 shows the control block diagram of the gen-set system, which
consists of a GT, a wound-field SG with an exciter, and a diode rectifier. The
governor can utilize conventional PI or lead-lag controllers to regulate the speed of
the prime mover. Recently, some advanced control techniques like sliding mode
control, fuzzy logic control, and model predictive control have been applied [19].
In the wound-field SG, the magnitude of output voltage is controlled by an exciter,
so that the DC output voltage rectified by the diode rectifier is regulated indirectly.
In the case of PMSG, the thyristor phase-controlled rectifier or the diode rectifier
with back-end DC/DC converter are usually used.
302 DC distribution systems and microgrids
Speed
detection
+ – Gas τout Vout
ω* Governor
turbine Vdc
Synchronous Diode
generator rectifier
+ Excitation Vf Iout
Vdc* Exciter
– controller
Ky tout
w* PI e−ts
1 + tcs
min
To simplify the design of the speed control loop, the dynamics of prime
movers are often modeled as a first-order or second-order delay. Figure 12.7 shows
the speed control block diagram of the prime mover with PI regulators, where t is a
time delay, tc is the time constant, and Ky is the torque constant. For better control
performance, the characteristics of the engine and turbine, effects of temperature,
shaft speed, combustion, and cooling air flow need to be included in the detailed
modeling [20].
For the long-term system analysis, the generator model is effectively simpli-
fied in a series circuit of variable AC voltage source, resistance, and inductance
[21]. The exciter system is often modeled as a first-order or second-order time
delay, which is usually controlled by thyristor phase-controlled converters [22].
scheme which increases the output voltage of the ESS should be applied. Finally,
the power management controller produces a power reference for the ESS [23].
vref
i ¼ vnom Rii ii (12.1)
Tertiary control
Central control
Information
of energy sources
and loads
Secondary control
Communication line
Local control
where vref
i is the voltage reference for ith unit, vnom is the nominal voltage, ii is the
output current, and Rii is the droop coefficient.
The droop gain can be obtained from the maximum output current (idc_max) and
the allowable voltage variation (Dvdc), as [25]
Dvdc
Rii ¼ : (12.2)
idc max
When the gen-sets and ESSs supply power to the DC grid, each local controller
regulates its output voltage and performs the power sharing. In the case of battery
ESS, the management of the SOC can be included. If only one generator is operated
or if the generator is operated at rated power, the droop control is not applied. When
switching the control modes, the higher-level controller sends a control mode
selection signal to the primary one via communication as shown in Figure 12.9.
Rated power
Tertiary V *DC
mode
control VDC + Mode
selection
+
Output
Secondary voltage
control Power sharing V *DC reference
mode
+
+
–
Rd Output current
Nominal voltage
Pri δvi
ma Se
ry con
dar
y
Rii
Output current
output current in the primary level is increased according to the load, the output
voltage is decreased due to droop control. Then, the secondary control yields a
DC-bus voltage compensation term, dvi , by which the modified voltage reference is
obtained as [26]
vref
i ¼ vnom Rii ii þ dvi : (12.3)
The tertiary control, which is the highest-level control, decides whether the gen-sets
and ESSs are operated or not, depending on the load power demand. Furthermore,
the tertiary controller can perform the economic power dispatch control and energy
balancing control.
Parameters Values
MVDC-bus voltage 6 kV
GT-SG set 36 MW 2
DE-PMSG set 4 MW 2
Propulsion (M1, M2) 36 MW 2
Zonal loads 5 MW
Pulsed power load 4 MW for 1 s
Energy storage system 4 MW h
306 DC distribution systems and microgrids
The secondary controller regulates the MVDC bus voltage with information of
gen-sets. The output voltage references of each gen-set system are calculated in
(12.3), which will compensate for the variations of the MVDC bus voltage. Finally,
the tertiary controller selects the operating mode of the gen-sets according to the
load profile.
● Load profile
The load profile has four operating modes, as shown in Figure 12.11, which are
classified by the power demand of ship loads.
Mode I: The consumption of the load power is lower than 5 MW. The gener-
ated power is supplied only to services loads since the ship is in port. In this
mode, only DEs except GTs are operated.
Mode II: The consumption of the load power is ranged between 5 and 36 MW.
In this mode, only one GT can support the whole power demand for pro-
pulsion and service loads.
Mode III: The consumption of the load power is ranged between 36 and
72 MW. In this mode, both GTs are operated and DEs are not operated. The
PPL is applied in this mode.
Mode IV: The consumption of the load power is higher than 72 MW. The four
gen-sets should be operated to satisfy the power demand while the ship is on
the voyage at the maximum speed.
The operating modes of the four gen-sets are summarized in Table 12.2.
● Simulation results
Figure 12.12 shows the system performance in operating modes I and II. Initially,
the system is operating in mode I, where two DEs supply power to loads. At 10 s,
the propulsion drives are activated. Then, the load power is increased and the
operating mode is changed to mode II. The central control sends a start command to
the GT1. At the same time, the governor adjusts the mechanical torque to increase
80
Mode IV
70 72 MW
60
Load power (MW)
Mode III
50
40
36 MW
30
20 Mode II
10
5 MW
Mode I
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300 320
Time (s)
40
Pload Pgen PESS (MW)
Pgen
DE and GT1 speed
30 1
ωr_DE (1 and 2)
20 ωr_GT1
(p.u.)
Pload
10 PESS 0.5
0
–10 0
(a) 10 20 30 40 50 60 (d) 10 20 30 40 50 60
40 6.3
DC bus voltage (kV)
GT1 power (MW)
6.2
30
6.1
20 6.0
10 5.9
5.8
0
5.7
(b) 10 20 30 40 50 60 (e) 10 20 30 40 50 60
5 4
DE power (MW)
4
Operating mode
PDE (1 and 2) 3
3
2
2
1 1
0
0
10 20 30 40 50 60 10 20 30 40 50 60
(c) Time (s) (f) Time (s)
Figure 12.12 System performance in operating modes I and II: (a) generator,
ESS, and load powers, (b) output power of GT1, (c) output power
of DE1 and DE2, (d) speed of DE1, DE2, and GT1, (e) DC bus
voltage, (f) operating mode
308 DC distribution systems and microgrids
the speed. During this interval, the load power exceeds the maximum power of two
DEs (DE1 and DE2). The central controller requests the ESS to support the
demanded power at 13 s. When the GT reaches to the optimal speed at 15 s, the
output voltage is regulated by the exciter system and the GT begins to support
the power. The three generators are still operated together until the load power
stops increasing at 25 s. After that, the central controller sends a stop command to
the DE1 and DE2 at 27 s. At 28 s, the charging operation of the ESS is activated.
At 45 s, the propulsion load begins to decrease (not shown). As it reaches 5 MW
at 55 s, the operating mode is changed back to mode I. The GT1 continues sup-
plying power to the load. If the propulsion stops completely, then, only zonal loads
consume the power. The DE1 and DE2 are requested to supply the power. When
the DE1 and DE2 can afford to supply the load demand, the central controller gives
a stop command to the GT1 at 57 s.
Figure 12.13 shows the system performance in operating modes II and III.
Initially, the system is operating in mode II, where one GT1 supplies the power of
25 MW to the loads. At 110 s, the propulsion loads begin to increase to accelerate
the ship. At 118 s, the load power demand becomes higher than 36 MW, from
80 Pgen
Pload Pgen PESS (MW)
1
60
GT speed (p.u.)
ω r_GT1
40 PLoad ω r_GT2
0.5
20 PESS
0
0
(a) 110 120 130 140 150 160 170 180 (d) 110 120 130 140 150 160 170 180
50 6.2
DC bus voltage (kV)
GT1 power (MW)
40 PGT1 6.1
30 6.0
5.9
20
5.8
10 PGT2 5.7
0 5.6
(b) 110 120 130 140 150 160 170 180 (e) 110 120 130 140 150 160 170 180
5 4
Operating mode
DE power (MW)
4 3
3
2
2
1 1
0
0
(c) 110 120 130 140 150 160 170 180 (f) 110 120 130 140 150 160 170 180
Figure 12.13 System performance in operating modes II and III: (a) generator,
ESS, and load powers, (b) output power of GT1 and GT2, (c) output
power of DE1 and DE2, (d) speed of GT1 and GT2, (e) DC bus
voltage, (f) operating mode
Shipboard MVDC microgrids 309
which the operating mode is changed to mode III. The central controller requests
the ESS to support the demanded power at 118 s and at the same time sends a start
command to the GT2. After some delay, the output power of GT2 begins to
increase and the support of the ESS decreases. At 133 s, two GTs begin to operate
at a balanced power sharing mode without the support of the ESS. While the system
is operating in steady states, the ESS energy is recovered by a charging operation.
At 150 s, the propulsion load begins to decrease (not shown). As it reaches
36 MW at 159 s, the operating mode is changed back to mode II. Two GTs are still
operating in parallel until the load demand stops decreasing at 165 s. Then, the
central controller sends a stop command to the GT2. At 175 s, the GT2 is deacti-
vated completely and the GT1 alone supplies the power.
Figure 12.14 shows the system performance in operating modes III and IV.
Initially, the system is operating in mode III, where two GTs supply the power of
55 MW to the loads. At 230 s, the propulsion loads begins to increase for accel-
erating the ship. At 242 s, the load power demand becomes higher than 72 MW, DE and GT speed (p.u.)
100
Pload Pgen PESS (MW)
Pgen ωGT1
80 1
60 ωGT2
Pload
40 0.5
PESS ωDE
20
0
0
230 240 250 260 270 280 230 240 250 260 270 280
(a) (d)
40 6.3
DC bus voltage (kV)
PGT2
GT power (MW)
35 6.2
PGT1 6.1
30 6.0
25 5.9
5.8
20 5.7
(b) 230 240 250 260 270 280 (e) 230 240 250 260 270 280
5 4
DE power (MW)
Operating mode
4
3
3 PDE (1 and 2)
2 2
1 1
0 0
230 240 250 260 270 280 230 240 250 260 270 280
(c) Time (s) (f) Time (s)
Figure 12.14 System performance in operating modes III and IV: (a) generator,
ESS, and load powers, (b) output power of GT1 and GT2, (c) output
power of DE1 and DE2, (d) speed of DE1 and 2 and GT1 and 2,
(e) DC bus voltage, (f) operating mode
310 DC distribution systems and microgrids
5 PPLL PESS
6.1
6.0
5.9
5.8
5.7
285 290 295 300 305 310 315
(b) Time (s)
Figure 12.15 System responses for pulsed power load: (a) pulsed power load
and ESS power, (b) DC bus voltage
from which the operating mode is changed to mode IV. The central controller
requests the ESS to support the demanded power at 243 s and at the same time
sends a start command to the DE1 and DE2. At 247 s, both of the DE1 and DE2
generate their rated powers, from which the output power of the ESS is decreased.
During this interval, the output power of GTs varies to make a balance between the
total supplied power and the load power demand.
At 260 s, the propulsion load begins to decrease. As it reaches 72 MW at 263 s,
the operating mode is changed back to mode III. Two GTs and two DEs are still
operating in parallel until the load demand stops decreasing at 55 MW, at 271 s.
After a delay, the central controller sends a stop command to the DE1 and DE2.
Then, only the GT1 and GT2 supply the power to the loads.
Figure 12.15 shows the responses in the case of applying the PPLs while the
system is operating in mode III. The PPL consumes a high power of 4 MW for 1 s.
The central controller sends a command to the ESS, which supplies the demand of
the PPL quickly. It is seen that the MVDC bus voltage has a little fluctuations of
about 150 V, which is within the allowable range of 10%.
For the linear time-invariant (LTI) system, the system is stable if and only if
the following conditions are satisfied:
(i) Zbus(s) includes no right-half plane poles.
(ii) Re{Zbus(jw)} 0, 8w
Trip signal
generated
Voltage
Fault
Normal current System
operation limited restored
Current
Fault
t0 t1 t2 t3 t4 t5 t6 t7 Time
6 kV
DC/AC
MVDC bus 2,300 V AC
+ – converter
Port propulsion
A1 B
AC/DC Icm_A
1 Sensing
DC/DC
converter Icm_B station 4
converter 1 kV DC
R2 Zone 1 load
Vn A A2 C
n R3
Icm_A Sensing
Icm_A 2 DC/DC Icm_C
R4 Icm_A station 5
3
converter 1 kV DC
R1 Vcm_off Sensing Zone 2 load
station 1 A3 D
C1
Sensing
GFDL Sensing Icm_D station 6
CR Sensing
controller station 3
Local station 2
area
network
of which are 1 kV DC for zone 1 and 2 loads and 2,300 V AC for port propulsion
drives, respectively.
Depending on the location of the ground fault, the voltages at all interfaces
with respect to the ground shift by different levels [36]. For instance, if a ground
fault occurs at positive or negative DC buses, the voltages at all the interfaces shift
negatively or positively by a half of DC bus voltage, respectively. However, if the
ground fault occurs at AC bus, the voltages at all the interfaces shift positively and
negatively by a peak value of phase voltages in AC bus.
Figure 12.18 shows the line-to-ground voltages at each interface when the
ground fault occurs in the system of Figure 12.18. If the ground fault occurs at the
negative bus of 1 kV DC in zone 1, all the output voltages of power converters
are shifted by 500 V, as shown in Figure 12.18(a). For example, the positive
MVDC bus voltage shifts from 3 to 3.5 kV DC and the negative bus voltage shifts
from 3 to 2.5 kV DC. In the case of ground fault at one phase of 2,300 V AC
bus, the voltages at all the interfaces are shifted by a peak phase voltage, 1,878 V,
shown in Figure 12.18(b).
To protect the equipment, a strategy for detecting and locating the ground fault
is needed, which is discussed as follows:
● Ground-fault detection
A ground fault condition is checked by measuring a common-mode offset voltage,
Vcm_off, at the output terminal of the AC/DC converter. Vcm_off is the voltage across
a resistor, R4, which is connected from the mid-point of two identical resistors, R2
and R3, to the ground, as shown in Figure 12.17.
Shipboard MVDC microgrids 315
5k 5k
Voltage (V)
Voltage (V)
0 0
–5k –5k
Propulsion motor 2,300 V AC bus voltage Propulsion motor 2,300 V AC bus voltage
2k 2k
Voltage (V)
Voltage (V)
0 0
–2k –2k
1.5k 1.5k
Voltage (V)
Voltage (V)
0 0
–1.5k –1.5k
1.5k 1.5k
Voltage (V)
Voltage (V)
0 0
–1.5k –1.5k
0.76 0.78 0.8 0.82 0.84 0.76 0.78 0.8 0.82 0.84
Time (s) Time (s)
(a) (b)
Figure 12.18 Line-to-ground voltages when a ground fault occurs at (a) negative
bus of the 1 kV DC zone 1 load, (b) one phase of 2,300 V AC bus
In the ground-fault detection and location (GFDL) controller, the rms and
average values of Vcm_off are calculated. If both of them are higher than the
thresholds for an interval long enough not to be considered as a transient state, it is
concluded that a DC ground fault has occurred. On the other hand, if only its rms
value is higher than the threshold, it is known that an AC ground fault occurs.
● Ground-fault location
To locate the ground fault, the series R1C1 circuit which is connected to the neutral
point of the three-phase power source is utilized. If the line-to-ground fault is
detected, the controller gives a command to close the contactor relay (CR). Then, a
third-order harmonic current is conducted from the chassis ground to the fault point
through the CR. At all the interfaces of power converters, the common-mode cur-
rents, Icm_x, which are the sum of currents through positive and negative buses are
measured. Then, each sensing station calculates the rms value of its Icm_x and sends
to the GFDL controller through communication links in a delay time.
316 DC distribution systems and microgrids
If the DC ground fault occurs, the common-mode currents of AC/DC and DC/
DC converters are examined. Then, it is identified that the ground fault occurs in
the converter where the highest common-mode current flows. On the other hand, if
the AC ground fault occurs, it is decided that the inverter where the highest com-
mon-mode current flows has a ground fault.
Figure 12.19 shows the waveforms of Vcm_off and Icm_x at the same fault con-
dition as in Figure 12.19. If there is no ground fault in the system, the common-
mode voltage and current are zero. If the ground fault occurs at the negative DC bus
of zone 1 load (point C), the rms and average values of Vcm_off reach to 500 V, from
which the DC ground fault is detected, as shown in the top of Figure 12.19(a).
After a delay, the fault location is found from the Icm_A and Icm_C. If the
ground fault occurs at one phase of inverter (point B), the rms and value of Vcm_off
is 500 V, but its average value is almost zero, as shown in Figure 12.19(b). In this
Voltage (V)
1k 1k
0 0
–1k –1k
Icm_A Icm_A
0.6 0.6
Current (A)
Current (A)
0.4 0.4
0.2 0.2
0 0
Icm_B Icm_B
0.6 0.6
Current (A)
Current (A)
0.4 0.4
0.2 0.2
0 0
Icm_C Icm_C
0.6 0.6
Current (A)
Current (A)
0.4 0.4
0.2 0.2
0 0
Icm_D Icm_D
0.6 0.6
Current (A)
Current (A)
0.4 0.4
0.2 0.2
0 0
0.76 0.78 0.8 0.82 0.84 0.76 0.78 0.8 0.82 0.84
Time (s) Time (s)
(a) (b)
Figure 12.19 Common-mode offset voltages (in rms) and currents (in rms) when a
ground fault occurs at (a) negative DC bus of zone 1 (b) one phase
of 2,300 VAC bus at B
Shipboard MVDC microgrids 317
case, it is diagnosed that the AC ground fault has occurred. It is seen that the
common-mode currents of Icm_A and Icm_B have a significant value.
12.6 Summary
As the environmental regulations have become stricter and stricter recently, the
demand on higher fuel efficiency and lower emission for the ship is more pressing.
Therefore, the AES, which has high survivability and reliability, is predicted to be a
future trend of the ship. Since the MVDC IPS well fits with the AES, more research
efforts should be devoted to the development of this technology up to the maturity
stage.
References
[1] E. Skjong, R. Volden, E. Rødskar, M. Molinas, T. A. Johansen, and
J. Cunningham, ‘‘Past present and future challenges of the marine vessel’s
electrical power system,’’ IEEE Trans. Transport. Electrific., vol. 2, no. 4,
pp. 522–537, 2016.
[2] M. Patel, Shipboard propulsion and power electronics and ocean energy,
USA, FL, Boca Raton, CRC Press, 2012.
[3] ‘‘Queen Elizabeth 2,’’ 2002. Available from: https://en.wikipedia.org/w/
index.php?title=Queen_Elizabeth_2&offset=20041114142442&limit=500&
action=history.
[4] O. Alnes, S. Eriksen, and B.-J. Vartdal, ‘‘Battery-powered ships: A class
society perspective,’’ IEEE Electrific. Mag., vol. 5, no. 3, pp. 10–21, 2017.
[5] IEEE Std. 1709-2010, ‘‘IEEE recommended practice for 1 kV to 35 kV
medium voltage DC power systems on ships,’’ in IEEE Std. 1709-2010, Nov.
2010, pp. 1–54.
[6] J.-F. Hansen, J. Lindtjoern, T. Myklebust, and K. Vanska, ‘‘Onboard DC
grid: The newest design for marine power and propulsion systems,’’
ABB Rev., no. 2, 2012. Available from: https://library.e.abb.com/public/
b4f3f099e9d21360c1257a8a003beac2/ABB%20Generations_20%20Onboard%
20DC%20grid.pdf.
[7] D. Li, R. A. Dougal, E. Thirunavukarasu, and A. Ouroua, ‘‘Variable speed
operation of turbogenerators to improve part-load efficiency,’’ in Proc. IEEE
Electr. Ship Technol. Symp., 2013, pp. 353–359.
[8] Energy and Environmental Analysis, Inc. an ICF Company, Technology
characterization: Reciprocating engines, Environmental Protection Agency,
Combined Heat and Power Partnership Program, 2008.
[9] ABB, Generators for wind power, proven generators – reliable power, ABB
Motors and Generators, 2010.
[10] Y. Tang and A. Khaligh, ‘‘Bidirectional hybrid battery/ultracapacitor energy
storage systems for next generation MVDC shipboard power systems,’’ in
Proc. IEEE VPPC, 2011, pp. 1–6.
318 DC distribution systems and microgrids
[11] S. Kim, S. Choe, S. Ko, and S. Sul, ‘‘A naval integrated power system with a
battery energy storage system: Fuel efficiency reliability and quality of
power,’’ IEEE Electrific. Mag., vol. 3, no. 2, pp. 22–33, 2015.
[12] J. F. Hansen and R. Lysebo, Comparison of electric power and propulsion
plants for LNG carriers with different propulsion systems, ABB AS, 2007.
[13] M. Stieneker and R. W. D. Doncker, ‘‘Dual-active bridge dc-dc converter
systems for medium-voltage DC distribution grids,’’ in Proc. IEEE Power
Electronics Conference and 1st Southern Power Electronics Conference
(COBEP/SPEC), 2015, pp. 1–6.
[14] R. Mo and H. Li, ‘‘Hybrid energy storage system with active filter function
for shipboard MVDC system applications based on isolated modular multi-
level DC/DC converter,’’ IEEE J. Emerg. Sel. Top. Power Electron., vol. 5,
no. 1, pp. 79–87, 2017.
[15] P. Manuelle, B. Singam, and S. Siala, ‘‘Induction motors fed by PWM
MV7000 converters enhance electric propulsion performance,’’ in Proc.
IEEE Power Electronics and Applications, 2009. EPE’09. 13th European
Conference on, 2009, pp. 1–9.
[16] M. Spichartz, V. Staudt, and A. Steimel, ‘‘Modular multilevel converter for
propulsion system of electric ships,’’ in Proc. IEEE Electric Ship Technol.
Symp, 2013, pp. 237–242.
[17] X. Pei, O. Cwikowski, M. Barnes, A. C. Smith, and R. Shuttleworth, ‘‘A
review of technologies for MVDC circuit breakers,’’ in Proc. Industrial
Electronics Society, IECON 2016—42nd Annual Conference of the IEEE,
2016, pp. 3799–3805.
[18] S.-H. Lee, J.-S. Yim, J.-H. Lee, and S.-K. Sul, ‘‘Design of speed control loop
of a variable speed diesel engine generator by electric governor,’’ in Proc.
IEEE IAS Annu. Meet., 2008, pp. 1–5.
[19] S. Di Cairano, D. Yanakiev, A. Bemporad, I. V. Kolmanovsky, and
D. Hrovat, ‘‘Model predictive idle speed control: Design analysis and
experimental evaluation,’’ IEEE Trans. Control Syst. Technol., vol. 20, no. 1,
pp. 84–97, 2012.
[20] S. K. Yee, J. V. Milanovic, and F. M. Hughes, ‘‘Overview and comparative
analysis of gas turbine models for system stability studies,’’ IEEE Trans.
Power Syst., vol. 23, no. 1, pp. 108–118, 2008.
[21] M. Farasat, A. S. Arabali, and A. M. Trzynadlowski, ‘‘A novel control
principle for all-electric ship power systems,’’ in Proc. IEEE Electric Ship
Technologies Symposium (ESTS), 2013, pp. 178–184.
[22] S. D. Umans and D. J. Driscoll, ‘‘Excitation system for rotating synchronous
machines,’’ U.S. Patent 6 362 588, Mar. 26, 2002.
[23] O. Palizban, K. Kauhaniemi, and J. M. Guerrero, ‘‘Microgrids in active
network management—Part I: Hierarchical control energy storage virtual
power plants and market participation,’’ Renewable Sustainable Energy
Rev., vol. 36, pp. 428–439, 2014.
[24] Z. Jin, G. Sulligoi, R. Cuzner, L. Meng, J. C. Vasquez, and J. M. Guerrero,
‘‘Next-generation shipboard DC power system: Introduction smart grid and
Shipboard MVDC microgrids 319
13.1 Introduction
1
Electrical and Computer Engineering Department, McMaster Automotive Resource Centre (MARC),
McMaster University, Canada
322 DC distribution systems and microgrids
140
120
100
Thousand vehicles
80
60
40
20
0
2011 2012 2013 2014 2015
Year
Volvo XC90 Honda Fit EV Mercedes B-Class E BMW i3
Mercedes S550 Plug in Kia Soul EV BMW i8 Ford C-MAX Energi PHEV
Tesla Model X Porsche Cayenne S E Chevrolet Spark Ford Fusion Energi
Smart ED Porsche Panamera S E VW e-Golf Toyota Prius Plug-in
BMW X5 Mitsubishi i-MiEV Smart for Two EV Tesla Model S
BMW Active E Cadillac ELR Fiat 500E Chevrolet Volt PHEV
Honda Accord Toyota RAV4 EV Ford Focus EV Nissan LEAF
Figure 13.1 2011–2015 U.S. plug-in electric vehicle sales by model [2]
Off-board
HV battery
charger
Charge port
LV DC-link
LV battery APM dc
DC
Grid
LV On-board ac
AC
accessories HV
DC/DC boost HV DC-link battery
charger
Electric HV battery
motor DC/AC inverter
Inductive charger
Inductive charger
and transfers power magnetically. The AC/DC power factor correction (PFC)
converter, high-frequency DC/AC primary side, and the primary transducer are all
located away from the vehicle. The secondary transducer and AC/DC rectifier are
installed inside the vehicle. The conductive charging systems use direct contact
DC-based EVs and hybrid EVs 323
between the connector and the charge inlet [4]. These can be categorized as
onboard and off-board battery chargers. The onboard charger is a must for elec-
trified vehicles needing external charging. It has power limitations because of cost,
weight, and space constraints. The onboard charger connects to the grid through
the AC charge port from the vehicle and creates the AC/DC power conversion
inside the vehicle. The off-board charger has higher power rating and thus reduces
the charging time. It has the AC/DC power conversion away from the vehicle and
charges the HV battery directly through the DC charge port of the vehicle.
Typically, the propulsion system is an electric motor with a three-phase DC/
AC inverter as the power processing unit between the HV battery and the motor.
Some electrified powertrains apply an additional DC/DC boost converter in
between the HV battery and the inverter to improve the system’s efficiency by
increasing the DC-link voltage and providing an adjustable DC-link voltage for the
traction inverter.
The LV vehicular system voltage is typically 12 V; therefore, the HV battery
cannot be utilized to supply power directly to the vehicular loads. For this reason,
power converters are required, which convert HV from the traction battery to a
lower voltage in order to supply power to the LV vehicular loads, and to charge the
LV battery, which is usually a lead-acid battery. This power converter is called an
auxiliary power module (APM).
In this chapter, several converter topologies of PFC converter, isolated DC/DC
converter, APM, and active power filter (APF) in electrified vehicle applications
are presented. The practical design considerations, including power switch,
DC-link capacitor, and DC bus bar, are given. Finally, the state-of-the-art topolo-
gical system reconfigurations in electrified vehicles are presented.
Boost
DC DC
HV Traction motor
DC AC
battery
DC LV auxiliary system
DC
APM 12 V LV battery
Auxiliary loads
11%
Lighting
32% Wiper and window system
Other loads
Autonomous system
14%
20%
Figure 13.4 Typical LV auxiliary load power distribution in the electrified vehicle
lamps draw most of the power. Wiper and window system related loads draw
around 14% of the total power. Electronic accessory loads include the computer,
displays, power outlets, CD player, Bluetooth, and Global Positioning System.
Electric power steering and motor engaging park brake are some other loads. Other
than that, in vehicles where additional luxury loads are requested, such as power
sunroof, active suspension system, or entertainment systems, the total power level
could be higher.
Besides these conventional auxiliary loads, with the recent development on the
autonomous vehicle, the needed power from autonomous system in the vehicle appli-
cation has also been evaluated. The sensors could be used in the autonomous system are
DC-based EVs and hybrid EVs 325
camera, radio detection and ranging (RADAR), light detection and ranging (LiDAR),
and inertial measurement unit. LiDAR is a sensor used to detect the presence of objects
round the vehicle, so the computer can interpret the vehicle’s relation to its surround-
ings. Besides these sensors, extra computer processing is required for the autonomous
system. The total amount of power occupied by this system is around 200 W.
LH DH5
DH1 DH3 LH1 DH1 DH2
Load
Load
~ SH1
~ LH2
Cdc Cdc
DH2 DH4 SH1 SH2
DH3 DH4
(a) (b)
LH ~ LH
~
Cdc Cdc
(c) (d)
Figure 13.5 Single-phase PFC converter: (a) boost PFC converter. (b) 2-Phase
bridgeless PFC converter. (c) Bridgeless totem-pole PFC converter.
(d) GaN-based bridgeless totem-pole PFC converter with SR
326 DC distribution systems and microgrids
MOSFETs which further reduce the conduction loss. (2) Lower part counts, higher
power density, and lower bill of material cost. It uses fewer parts and has a simpler
circuit. It needs only one inductor and neither SiC diodes nor AC return diodes are
required. (3) Bidirectional power flow. BTPPFC is inherently capable of bidirec-
tional operation, which is ideal as the application requires power flow in both
directions. Therefore, the BTPPFC is a promising bidirectional PFC candidate for
achieving bidirectional power flow from grid to vehicle and from vehicle to grid.
The main function of the second-stage DC/DC converter is to offer regulated
voltage and current to better charge the HV battery. Two-stage topology provides
high-power factor, wide line regulation performance and clean charge current.
However, cost also needs to be taken into account, especially for the onboard
charger. The most widely used unidirectional isolated topologies are the phase-shift
full bridge and LLC resonant converter as shown in Figure 13.6. As a frequency-
modulated resonant converter, the LLC converter has a resonant tank between the
primary and secondary sides. The capacitor Cs and the inductors Ls and Lm form
the resonant tank. The Lm is the transformer magnetizing inductor and Ls can be the
transformer leakage inductor or an additional auxiliary inductor. The phase shift
full bridge is a PWM-modulated converter. It utilizes the transformer leakage
inductor and the parasitic capacitors from the primary switches to form a resonant
circuit and creates a switching pattern that utilizes the lagging current in the
inductor to fully discharge the parasitic capacitor before turning on the/a switch.
As shown in Figure 13.6, two contactors T1 and T2 need to be installed on the
DC terminals of a HV battery. They provide the isolation between the battery and
T1
S1 S3
D1 D3 +
Ls
HV battery
Vdc n:1
+
Cdc Co V
– Ho
–
S2 S4 D2 D4
T2
(a)
T1
S1 S3
D1 D3 +
Ls n:1
HV battery
Vdc
+
Cdc Lm Co
– Cs VHo
–
S2 S4 D2 D4
T2
(b)
remaining vehicle systems. Among them, T1 is the main contactor and T2 is the
backup contactor from a safety consideration.
9
8
7
Capacitance Cdc (mF)
6
5
4
3 5
2
10
V)
1
(
ge
ta
0
ol
15
kv
1,000
ea
2,000 3,000
-p
4,000
-to
5,000 20
6,000
ak
Figure 13.7 Relation among the power level, capacitance, and its voltage ripple
DC-based EVs and hybrid EVs 329
To increase the power density and reliability of the power electronic system
in vehicle applications, capacitor-less designs have been proposed. APF can be a
potential solution to reduce the required capacitance. The HV battery charger
structure with additional APF is shown in Figure 13.8.
Four conventional single-phase APF converter topologies are shown in
Figure 13.9. They are a voltage source inverter, a current source inverter, a
Idc+ir Idc
AC +
~
DC HV DC
AC
port ir port
APF
(a)
Idc+ir Idc
AC DC AC +
~
DC AC DC HV DC
AC
port ir port
APF
(b)
Figure 13.8 HV battery charger structures with additional APF: (a) single-stage
charger, (b) two-stage charger with isolated DC/DC converter
S1 S3 S1 S3
Lr D1 D3
Cr Cr Lr
S2 S4
S2 S4
D2 D4
(a) (b)
S1 S1
Lr Lr
Cr
Cr S2 S2
(c) (d)
Figure 13.9 Four typical conventional APFs. (a) Voltage source inverter.
(b) Current source inverter. (c) Bidirectional buck converter.
(d) Bidirectional boost converter
330 DC distribution systems and microgrids
bidirectional buck converter, and a bidirectional boost converter. With these addi-
tional APF circuits, the required capacitance can be reduced significantly. How-
ever, by adding an extra APF circuit, additional power electronic components are
needed, and thus the system’s complexity is also increased.
S5
LV battery and aux load
T1
S1 S3 +
+ Ls L1
n:1
HV battery
VHo Co
Cdc –
– L2
S2 S4 S6
T2
(a)
T1
LV battery and aux load
S1 S3 S5 S7
+ Ls n:1 +
HV battery
VHo
Cdc Co
–
– VLo
S2 S4 S6 S8
T2
(b)
Figure 13.10 LV APM isolated DC/DC converter candidates (a) phase-shift full
bridge with current doubler, (b) DAB converter
DC-based EVs and hybrid EVs 331
promising candidate. The DAB converter has active full bridges on both primary
and secondary sides. DAB can provide ZVS on both primary and secondary sides
by utilizing the leakage inductor. Its symmetry and fixed switching frequency make
it a robust converter.
Vce Ic
400 650
350 550
Switch voltage Vce (V)
350 350
300 300
250 250
200 200
150 150
100 100
50 50
0 0
0 100 200 300 400 500 600 700
(b) Time (ns)
Figure 13.11 Current through and voltage across an IGBT during turn-on and
turn-off transients: (a) during turn-on transient and (b) during
turn-off transient
332 DC distribution systems and microgrids
Figure 13.11(b) is generated by the voltage across the circuit stray inductance
during the switch turn-off period. Generally speaking, the voltage and current
ratings of the device are chosen to be about 1.5–2 times the values required by
the system.
In addition, other parameters of the switching devices must also be considered,
such as equivalent turn-on resistance, series equivalent inductance, switching time,
and switching energy loss. The large equivalent resistance, switching time, and
energy loss result in high conduction and switching power loss, while the large
series equivalent inductance contributes to the total circuit stray inductance and
leads to a HV spike that may damage the switching device [8].
l r
I cto r
ondu n laye
w C tio r
ula cto
t Ins ondu
C
d
t
Figure 13.13 A laminated bus bar example where t, w, l, and d are conductor
thickness, width, length along the current direction, and the
distance between two conductors
in order to prevent overheating and the lifetime degradation of the capacitor, the
rated current of the selected capacitor must be higher than the root mean square
(RMS) value of the DC-link ripple current. Finally, the rated voltage of the film
capacitor could be selected as the same as the DC-link voltage, since its over-
voltage capability is high and can reach up to twice the rated voltage.
Generally, there are no upper limits of the conductor and insulation layer
thickness; however, there are lower limits. For the thickness of the insulation layer,
in order to sustain the operating DC-link voltage, the value should be larger than a
minimum value. The dielectric strength of various insulation materials with dif-
ferent thicknesses is not a constant value, thus the insulation layer thickness should
be defined depending on the materials and the practical dielectric strength. For
conductor thickness, to avoid overheat on the entire bus bar or at some hot spots,
the current density in each conductor must meet the requirement of 5 A/mm2 [10].
The cross-section area A in the current direction is defined by (13.1), and the
current density J is the amount of current I flowing through the unit area of this
cross section, (13.2).
A ¼ w t mm2 (13.1)
I
J¼ A=mm2 (13.2)
A
There are two types of currents flowing in the DC-link bus bar conductors, DC
current and AC current. The DC current flows from the input terminals to the
switching device terminals, while the AC current is the current ripple that flows
through the DC-link capacitor. During the design, both DC and AC current
densities need to be considered, and the conductor thickness and width can be
determined by the required conducting current and the DC-link current ripple.
Finally, the total current density must satisfy the conductor current’s density
requirement [11].
DC current on the bus bar spreads out evenly in the conductor. As a result, DC
current flows through the entire cross section area of the conductor. Here, the
thickness of the current layer is equal to the thickness of the conductor. However,
the AC current tends to be concentrated near to the conductor’s surface due to the
skin effect that is shown in Figure 13.14. Thus, the thickness of the current layer,
J[A_per_m2]
4.9397e+006
4.4779e+006
4.1862e+006
3.8945e+006
3.6028e+006
3.3111e+006
3.0195e+006
2.7278e+006
2.4361e+006
2.1444e+006
1.8527e+006
1.5610e+006
1.2693e+006
9.7764e+005
6.8595e+005
3.9426e+005
d d
1.0257e+005
which is called skin depth d, is smaller than the thickness of the conductor. Hence,
adjusting the thickness of the conductor only might not be an effective solution to
improve the AC current density, which can be calculated by replacing t by d in
(13.1). To satisfy the current density requirement, the conductor width also needs to
be considered under AC current conditions.
In addition, when more than one DC-link capacitor is utilized, the current
distribution on the bus bar needs to be balanced to avoid imbalanced current
sharing among the capacitors. Unbalanced current sharing between DC-link capa-
citors may result in higher current stress and a shorter lifetime for some capacitors
than others. This will also impact the reliability of the system. To achieve identical
impedance from different current paths on the DC bus bar and balanced current
sharing between capacitors, a symmetric structure of the bus bar is usually
preferred.
In the bus bar design, besides the requirements of current density and current
distribution, the bus bar parasitic parameters should also be considered.
The parasitic parameters include bus bar resistance, stray inductance, and stray
capacitance. When current flows through a bus bar, power loss is mainly generated
by the resistance. Voltage spikes during switching device turn-off transients are
caused by stray inductance. Besides, high-frequency noise can be filtered by stray
capacitance. As a result, large stray capacitance, small resistance, and small stray
inductance are preferred. On the one hand, the bus bar resistance should be
designed to be as small as possible and is usually under 1 mW. Then, the amount of
power loss caused by this resistance could be several watts. On the other hand, the
thickness of dielectric material layer should be designed to be as thin as possible to
achieve relatively large stray capacitance. Nevertheless, the dielectric layer thick-
ness also affects the insulation ability of the bus bar; to sustain the required DC-link
voltage, the thickness of the dielectric layer must be above a minimum value. In
fact, other than the optimization of resistance and stray capacitance, most efforts
are usually focused on minimizing the bus bar’s stray inductance during the design;
this is because the ohmic losses generated by the bus bar resistance is negligible
compared to the total converter power losses, and there are limited improvements
that could be achieved in stray capacitance maximization.
Total bus bar stray inductance Ltotal can be estimated by the summation of
conductor self-inductance Lself and mutual-inductance LM. Self-inductance is
determined by the shape of the conductor, while mutual-inductance is influenced
by both the shape and the overlapping area of the two conductors. When the current
directions are opposite to each other in two DC-link bus bar conductors, mutual-
inductance is negative. Thus, the total stray inductance of the bus bar can be
obtained as (13.3). In order to minimize the bus bar’s stray inductance, the over-
lapping area of the two conductors should be maximized. By laminating two con-
ductors of the bus bar, the stray inductance can be reduced significantly. That is the
reason why the laminated bus bar structure is popular in various bus bar designs,
seen in Figure 13.13.
Ltotal ¼ 2 Lself LM (13.3)
336 DC distribution systems and microgrids
The designed bus bar can be analyzed by FEA simulations, which is a straight-
forward method for estimating the parasitic parameters. Then, the voltage spike
could be evaluated with the estimated bus bar stray inductance. An equivalent
circuit of a converter phase-leg during the switching transient is shown by Fig-
ure 13.15, where the upper switch S1 is being turned off, while its corresponding
lower diode is turning on. The direction of current i does not change during this
time period, and the current in the high-side switch is transferring to the low-side
diode. During this turn-off period, voltage spike occurs as it is plotted by Fig-
ure 13.11. It is obvious that the voltage spike across switch S1 is caused by the
released energy from the inductances in the circuit, including series equivalent
inductance Lc of the DC-link capacitor, series equivalent inductance L1 of the
switching device, and the bus bar stray inductance Lbusbar. Thus, the voltage spike
and the peak voltage across switch S1 can be calculated by (13.4)–(13.6).
+ Vl,busbar – ia,1
Lbusbar +
L1 Vl,1
– –
Vl,c +
Lc S1
+ Vce
–
i
+
–
Udc C
L2 Vl,2
– +
ia,2
Figure 13.15 The equivalent circuit of a converter phase-leg during the turn-off
transient of switch S1
DC-based EVs and hybrid EVs 337
The integration of power electronics is critical, as it can reduce the size, weight, and
cost of the overall system. A simple type of system-level integration combines
several parts to better utilize the space in the three dimensions. For example, the
APM converter and traction inverter in the vehicle can be placed together in one
enclosure to share the DC-link capacitors, controller, heat sink, and housing cover
as presented in [12]. More advanced system-level integrated converters also allow
sharing of the power switches and gate drivers. However, these converters usually
either employ a relatively complex control strategy or apply a complicated topol-
ogy to regulate the power flows.
Figure 13.16 shows an integrated converter topology that achieves the bidir-
ectional power flows among four DC terminals [13]. The converter has the func-
tions of one DAB and two two-phase bidirectional buck converters but only need
eight switches. Two coupled inductors are needed so that different inductances can
be obtained according to different current flow directions for different power
conversions. The duty cycle is used to control the buck converter, and the phase
shift angle between the primary and secondary side is applied to control the DAB.
Therefore, the integrated converter can work independently and simultaneously
with multiple functions. However, there is a trade-off between the duty cycle and
phase shift angle which might limit the performance of both power conversions,
especially for the conversion with wide voltage and wide power range, such as a
battery charger. In addition, those integrated multifunctional switches need to
carry the overall power for both power conversions, leading to a higher current
rating on the power switches. This might become an obstacle for the power
switches and cooling system to reach higher power. Eventually, the reduced part
might be the gate-driver circuit only, as the required die size in the power switch
devices and the amount of heat-dissipation are still the same compared to the
nonintegrated method.
The operation of electrified vehicle has two modes. These are HV battery
charging and vehicle running mode. This makes electrified vehicle a unique
application. In other words, the role of operating time can also be considered as the
S1 S3 S5 S7
La Lc
DC DC
port 1 Cin1 Co1 port 3
S2 S4 Lb Ld S6 S8
DC DC
Cin2 Co2
port 2 port 4
fourth dimension for the system-level integration. Therefore, in the recent years,
the more integrated power electronic systems in vehicle applications have been
evaluated and proposed with an attempt to further reduce the size, weight, and cost.
These systems are dual-mode, which are application-oriented and fully utilize the
characteristics of the vehicle application.
T1
Electric motor
T3
AC AC DC +
~ HV DC
port DC DC port
T2
(a)
T1
T3 PFC
AC AC DC AC +
~ HV DC
port DC
DC AC DC port
Electric AC
M T2
motor DC
(b)
build the boost converter. As with the previous case, contactors T1 and T2 are
employed for reconfiguration between the two functions and isolation between the
battery and remaining vehicle systems.
T1
T3 PFC
AC ~ AC DC AC +
HV DC
port DC AC DC port
T2
+ DC
LV DC
port AC
(a)
T1
T3
AC DC +
AC ~ HV DC
port DC DC port
T2
T4
AC +
LV DC
DC port
(b)
T1
T3 PFC
AC ~ AC DC +
HV DC
port DC DC port
T2
APF
+ DC
LV DC AC
AC DC
port
(c)
Figure 13.18(b) shows the second system which integrates the primary DC/AC
side of the APM with the first-stage AC/DC PFC converter of the HV battery
charger. An extra inductor is still needed for the PFC, and the extra relay T4 is
required to achieve the reconfiguration.
The third type is shown in Figure 13.18(c). It integrates the primary DC/AC
side of the APM with the APF to alleviate the second-order harmonic current from
the single-phase grid line for the HV battery charger. As a result, the additional
power switches and their gate drivers, heat sink, and extra inductors can be elimi-
nated. However, the extra second-order harmonic energy storage capacitor is still
required for the APF.
13.5 Conclusions
In this chapter, the power electronics converters and inverter in vehicle application
are introduced. Several possible topologies for PFC, isolated DC/DC converter,
APM, and APF are presented. The practical design of DC bus bars and the selection
of power switch devices and DC-link capacitors are given. Finally, the state-of-the-
art topological reconfigurations of DC systems in electrified vehicles are presented.
One of the biggest challenges for the next generation of power electronic
systems in vehicle application will be the cost reduction to provide more affordable
solutions. This has become one of the major barriers for electrified vehicle for mass
commercialization. These dual-mode integration approaches are dedicated to
vehicle application and can reduce the cost, size, and weight of the system.
Therefore, they will also help promoting and accelerating the paradigm shift to the
transportation 2.0.
References
[19] R. Hou and A. Emadi, ‘‘A primary full-integrated active filter auxiliary
power module in electrified vehicle applications with single-phase onboard
chargers,’’ IEEE Transactions on Power Electronics, vol. 32, no. 11,
pp. 8393–8405, Nov. 2017.
[20] Y. Kim, C. Oh, W. Sung, and B. K. Lee, ‘‘Topology and control scheme of
OBC–LDC integrated power unit for electric vehicles,’’ IEEE Transactions
on Power Electronics, vol. 32, no. 3, pp. 1731–1743, Mar. 2017.
Chapter 14
DC data centers
Enver Candan1 and Robert C.N. Pilawa-Podgurski2
14.1 Introduction
In data centers, since the major energy supply is the AC utility and the primary
power consumers (i.e., IT equipment) require low voltage DC, both AC and DC
power infrastructures are concurrent. As the power is delivered from the utility to
the low voltage DC loads, rectification (power conversion from AC to DC form)
1
Department of Electrical and Computer Engineering, University of Illinois at Urbana-Champaign, USA
2
Department of Electrical Engineering and Computer Sciences, University of California, USA
344 DC distribution systems and microgrids
(b)
(c)
distributed in high DC voltage form within the data center. Then, as shown in
Figure 14.1(b), a dedicated DC–DC converter per IT equipment steps down the
high DC voltage at the IT equipment input. Alternatively, the AC power can
be distributed to racks that host the IT equipment, and rectifiers that are in the same
rack as the IT equipment (or in another rack that is in close proximity) provide DC
power to the nearby IT equipment as shown in Figure 14.1(c). Simplified diagrams
of these common data center power distribution architectures are depicted in
Figure 14.1. For prioritizing transition from AC to DC power distribution in
Figure 14.1, isolation transformers, protection equipment, and cooling devices are
not depicted in these figures but of course exist in practical designs and may
introduce additional power conversion stages.
Because of the extensive background, acceptability and well-established
standardization of the AC distribution in many other applications, a high percen-
tage of the existing data centers use derivations of the power delivery architectures
depicted in Figure 14.1(a) or (c), which are fundamentally inherited from telecom
applications. Recently, DC power delivery architecture, as depicted in Figure 14.1
(b), has gained attention, mainly because it involves fewer conversion stages and
could potentially simplify the integration of ancillary distributed energy resources,
such as solar PV and fuel cells (FC).
market share, and compatibility with the IT equipment have prevented the wide-
spread adoption of high voltage DC in data centers, while already well-established
AC distribution architectures have kept developing to meet expectations. This
chapter assesses the state of high voltage DC power delivery architecture for data
centers and points out its unique advantages and challenges but does not attempt
to provide a quantitative comparison between the AC and DC distributions.
14.3 Efficiency
By far, the largest operating cost of data centers is the cost of electricity. For this
reason, improving the energy efficiency of the computing and power delivery is
critical. As shown in Figure 14.1, regardless of the preferred distribution archi-
tecture, the utility power has to go through several cascaded power conversion
stages before it reaches the IT equipment. Since the power consumed by the IT
equipment must be processed by each power converter, the overall power infra-
structure efficiency is limited by the stage with the lowest efficiency. Power
delivery studies must thus consider the entire power conversion stage, from the
high voltage AC input to the building, all the way down to the CPU and memory
voltages, around 1 V.
While at a high-level, it may appear that simply reducing the number of con-
version stages would yield increased efficiency. However, one must be careful to
consider that for a constant power converter volume, a high-step down power
converter generally has lower efficiency than a converter with a modest voltage
step-down. This is particularly apparent in the case of the recently proposed 48 V
to the point-of-load concept, where the conventional architecture shown in
Figure 14.2 involves first a 48 to 9–12 V conversion, followed by (typically sev-
eral) 9–12 V to the point-of-load (e.g., 1–2 V) converters. While it may be tempting
to simply eliminate the two-stage conversion and design a single 48–1 V converter,
such a converter is significantly more difficult to design to be highly efficient and
highly power dense. For example, consider the reference design of [21], which
represents a single-stage buck converter, achieving a peak efficiency of 84%.
DC 1V
DC
CPU
DC
DC 1.2 V
DC
Memory
To rack-level converter DC
( AC–DC or DC–DC ) 3.3 V
DC
Disk
DC
Point-of-load IT equipment
converter
● Similar to the twice-line energy buffering, power factor correction (PFC) cir-
cuitry is an essential requirement in both single and multiphase high power
rectifiers. The most common PFC architecture is the boost-type converter,
meaning the output of the PFC circuit is at a higher voltage than its input. Since
the IT equipment requires low voltage DC, employing PFC closer to the low
voltage load requires a back-end, high conversion ratio, voltage step-down
converter. Although the centralized three phase rectification does not eliminate
the PFC circuitry, it can remove a cascaded voltage step up and down con-
version at the IT equipment input, which is fundamentally counterintuitive
since the load is at low DC voltage. Here, recent developments on step-down
(e.g., buck-derived) PFC rectifiers [30] show promise to achieve increased
system-level efficiencies.
Another potential advantage of the reduced number of cascaded stages in DC
data centers is the increase in available volume inside the IT equipment blade or
rack for more data processing and/or storage units. For example, the AC to DC
conversion stage in the PSU of the IT equipment requires twice-line frequency
energy buffering and PFC circuitries (as explained above), which consist of bulky
passive components since they need to be designed for grid frequency. Removal of
the AC to DC stage from the PSU thus allows more space in the IT equipment blade
or rack.
Since all the power consumed by the IT equipment must be processed through
the cascaded power stages, the system-level efficiency is still limited by the lowest
efficient power stage. Thus, research efforts have focused on the efficiency
improvements of each major power electronics converter type for high voltage DC
data center applications. Development and commercial availability of wide band-
gap devices have been leveraged in converter designs. Consequently, high-90%
efficiencies were achieved in three-phase rectifiers with 400 V DC output by
eliminating the boost PFC stage [30,31], and mid-90% efficiencies in 400–12 V or
48 V DC to DC converters by using resonant topologies [32–34].
Entirely new power delivery architectures also hold great promise for dramatic
increase in power-delivery efficiency. Instead of focusing on the efficiency
improvements of the individual power converters in the DC to DC conversion stage,
a series-stacked architecture (Figure 14.3) that focuses on reducing the amount of
DC bus
Rack or blade
AC
or DC DC
AC or DC IT Eq.
DC
power supply
DC DC IT Eq.
DC
Central
converter
Differential DC IT Eq.
DC
converter
power processed between the high voltage DC bus and multiple IT equipment has
been proposed in [35, 36]. By electrically connecting IT equipment of the same type
(servers in these works), the series-stacked architecture greatly reduces the requisite
power processed while achieving inherent high voltage step-down. The bulk power to
the servers is delivered by the DC bus current that passes through all the servers
without being processed. Since the series connected loads (servers) must conduct the
same amount of current, their input voltages drift in the case of a power consumption
mismatch. Bidirectional differential converters are employed at intermediate nodes to
maintain voltage balance under all circumstances. These converters are able to
exchange the difference between the server currents in the case of a power mismatch
between the servers. Through series stacking and active voltage balancing using dif-
ferential power processing, only the difference power between the servers needs to be
processed. Therefore, the amount of processed power is decoupled from the total
power delivered, yielding greatly reduced power loss during the power conversion in
comparison to the conventional power-delivery systems where the full server power
must be processed by the converters. Experimental verification of the series-stacked
architecture has demonstrated higher than 99.8% peak efficiency for the steady-state
operation of the servers, corresponding to an up to 40 reduction in average power
loss in comparison to a state-of-the-art PSU to date [36]. In addition, practical chal-
lenges of a series-stacked architecture such as load balancing in software [37],
hot-swapping [38], and unregulated bus operation [39] have been analyzed and
experimentally addressed without sacrificing high power conversion efficiency.
Although experimental verification of this architecture is limited to a stack of four
servers to a 48 V, its superior performance has been analyzed in case studies for 32
series-stacked servers connected to a high voltage (380 V) DC bus [35,36,38].
14.4 Reliability
Maintaining high reliability in data centers is crucial because of our society’s
dependence on the uninterrupted IT services. Nowadays, any outage of the IT
services can have a large impact, both financial and in terms of societal impact.
A typical desired target reliability for a data center is 99.99% uptime (often called
as ‘‘four nines’’), which corresponds to 52.5 min downtime per year [40]. This
requirement, combined with maintaining high efficiency, makes data center power
delivery architecture design challenging. Fortunately, similar to improving the
power delivery architecture efficiency, reducing the number of cascaded power
stages typically reduces the overall risk of system failure and the mean time
between failures. Therefore, the opportunities to reduce the number of power
conversion stages in DC data centers have been also considered to facilitate higher
reliability and uptime. Elimination of the conversion stages such as power
distribution transformers and the inverter at the UPS output in high voltage DC
distribution has been considered as the main advantage for the DC data centers
[11,41,42]. However, the analysis in these works is qualitative, and the details are
unclear. A 2010 study quantitatively compared the reliability of AC and DC power
350 DC distribution systems and microgrids
distribution for data centers with emphasis on UPS and concluded that the DC
distribution would be more reliable than the AC distribution without supplementary
effect of redundancy [16]. To the best of our knowledge, a more recent quantitative
reliability analysis for the DC data centers, which considers UPS, power converter
failure mechanisms and inclusion of different redundancy options, is missing from
the literature on this topic. Nevertheless, reducing the number of conversion stages
alone is not sufficient to assure reliable power distribution; back-up power and fault
tolerance must be incorporated in the data center power architecture design to
achieve the desired reliability level.
14.4.1 Fault tolerance
Fault tolerance is the ability of a system to maintain its operation in the case of
component failures. A fault tolerant system must be able to detect faults, protect the
system from the failed component, and provide redundancy. Fault detection and
redundancy for DC data centers is explained here, while protection methods are
explained later in Section 14.7.
There are two main types of faults to be detected in DC power delivery
architectures: over voltage and over current. Over voltage detection can be rela-
tively easily implemented in DC systems using voltage dividers and comparators.
However, over current detection is nontrivial in DC architectures because of in-
rush current phenomena and limited converter power ratings. Moreover, in data
centers, the rapidly changing loads may appear as an over current situation if load-
scheduling causes many servers to initiate computing at the same time. In-rush,
current can be thought of as an instantaneous and unusual rise at the terminal
current of a DC source or load and is present in all DC systems. Fundamentally, the
in-rush current can occur when a DC source or load connects to another DC source
or load due to the initial voltage difference between their filter capacitances.
The magnitude of the in-rush current depends on the voltage difference, filter
capacitance size, series inductive elements, and the contact resistance. If not
managed, in-rush currents may lead to equipment damage and voltage droops,
which can interfere with the converter control algorithm. Leveraging the fact that
in-rush currents occur during physical connection of different DC sources or loads,
in-rush current limitation techniques such as precharge circuits and soft-start rou-
tines have been developed. However, from a fault detection perspective, differ-
entiating between in-rush currents and an over current due to a component failure is
the main challenge in over current detection in DC data centers. Similarly, the
limited converter power rating complicates over current detection due to the faults.
Because of the finite power rating of the converters in a DC bus, a fault current due
to a component failure is limited by the rated converter power. A fault current may
also result in reduced voltage in the DC bus. Especially in data centers where the
instantaneous load current changes dynamically, differentiating between the full
load current and fault current requires careful analysis and tuning.
Redundancy is typically achieved through the incorporation of additional and
separate power conversion stages, UPS and power distribution paths in the data
center power infrastructure. The redundant components may be operated at all
DC data centers 351
times (e.g., each running at partial load to increase peak efficiency) but, strictly
speaking, only needed to meet the power demand of the load in case of failures.
Typical redundancy levels for data centers are N þ 1, 2N, and 2N þ 1, where
N represents the exact number of power converters or UPS systems in parallel to
meet the load demand. Uptime Institute defines Tier Classification levels for data
centers depending on the redundancy level of the data center [43]. Tier I represents
basic data center infrastructure without any redundancy. Tier II certification
requires redundant power stages and UPS; however, the power distribution path is
not redundant. Tier III certification requires the data center to have both redundant
power stages and multiple independent power distribution paths, although only
one distribution path is actively used at any time, while the other is for maintenance
purposes. Tier IV certification requires both redundant power stages and multiple
active power distribution paths configured to serve the entire data center under any
infrastructure failure. Tier Certificate requirements are summarized in Table 14.1,
and the details can be found in [43].
The DC power distribution in a data center simplifies redundant power archi-
tecture design. Since there is no synchronization and frequency control needed,
replacing redundant power stages is straightforward, given in-rush currents are
managed. In addition, if modular and redundant design is preferred, DC–DC con-
verters, UPS, and even the centralized rectifiers can be replaced without disrupting
the IT equipment operation. In addition to contributing to the overall reliability
through improving redundancy level, such a feature intuitively leads to reduced
equipment repair time in data centers, which further improves availability.
Rack
AC DC IT Eq.
or DC DC
Blade
DC IT Eq.
DC DC
AC or DC Blade
source Rack-level
UPS DC IT Eq.
DC
(a) Blade
Rack Rack
AC DC
IT Eq. IT Eq.
or DC +
_ Blade
Blade
DC
IT Eq. IT Eq.
DC
AC or DC Blade Blade
source
Distributed UPS Facility-level
IT Eq. UPS IT Eq.
Blade Blade
(b) (c)
Figure 14.4 Various UPS configurations for data centers. (a) Rack-level UPS.
UPS is located inside the racks and provides back-up power for
multiple IT equipment. (b) Distributed UPS. Each IT equipment has a
dedicated UPS. This configuration offers uninterruptible power in
the case of any converter failure. (c) Facility-level UPS. UPS is
located outside of the rack and provides back-up power for
multiple racks
The high voltage DC bus is the prime location for UPS and backup generators
to interface. Due to the DC voltage output of UPS, its connection to the high
voltage DC distribution architecture is straightforward. However, any backup
generator requires a rectifier stage to be able to supply power to the high voltage
DC bus. Although this rectifier is an additional power conversion stage, since it
only operates during a utility level power loss, the system-level efficiency and
reliability are not considerably penalized. In addition, connection of backup gen-
erators to high voltage DC bus does not require synchronization and frequency
control which are essential in AC architectures. Control methodologies are studied
for the DC data centers that employ diesel generators and battery banks for utility
power loss scenarios as early as 2008 [44].
14.5.2 Cooling
In addition to the computational data storage and networking load, cooling requires
substantial electrical energy in data centers. Since the IT equipment mainly consists
of digital circuits, combined with the losses in the power distribution architecture,
almost all energy consumed in data centers results in heat that must be dissipated.
In order to maintain safe operation of the IT equipment, temperature control is
critical at all times. Therefore, the reliability discussion above also applies to the
cooling system, and back-up power should be designed to support the cooling load
in the case of a utility outage.
In order to provide sustainable heat removal from the IT equipment, a typical
data center cooling infrastructure includes air conditioning equipment and chilled
water system, which require pumps and fans. Such equipment involves electric
motors which are typically controlled by adjustable speed drives (ASD) for max-
imum efficiency and performance. Similar to UPS, DC data centers require only an
inverter stage for ASD, compared to AC-driven ASDs, which first perform recti-
fication, followed by the adjustable frequency and voltage inverter. Overall, one
354 DC distribution systems and microgrids
would expect increased higher efficiency and reliability [50] owing to the reduced
number of stages (and conversions) in the direct DC approach.
14.5.3 Lighting
Although the lighting load in data centers is small in comparison to the computa-
tional and cooling load, DC data centers present unique advantages if modern and
efficient lighting technologies such as LEDs are used. Because of the DC nature of
the modern lighting loads, power conversion stages like rectification, twice-line
energy buffering, and PFC are not required when DC power distribution archi-
tecture is employed. As discussed in the efficiency section, reducing the number of
power conversion stages reduces the power conversion losses for the lighting loads
as well. Recent literature has demonstrated LED drivers with 380 V input for DC
microgrids, featuring dimming and short circuit protection [51].
14.6 Installation
Early papers promoting high voltage DC distribution in data centers noted famil-
iarity of the IT equipment with high voltage DC since the PFC stage of the PSU
typically involves a 400 V DC intermediate voltage. However, the installation
of high voltage DC distribution in data centers requires more than familiarity with
400 V DC. Major installation considerations include isolation, grounding, wiring,
connectors, and total cost of ownership (TCO).
14.6.1 Isolation
Electrical (galvanic) isolation has been an essential part of data center power-
delivery architectures. Provided by transformer usage, the electrical isolation offers
to filter grid disturbances, harmonic currents, and electrical noise. Also, the elec-
trical isolation limits ground loops and circulation of DC currents between the IT
equipment and racks [2]. Because of these benefits, the electrical isolation through
isolation transformers is a recommended practice by IEEE STD 1100 in AC power
distribution architectures [52].
Inherited from the AC power delivery, the DC power delivery architecture in
data centers also employs electrical isolation between the high voltage DC bus and
the IT equipment input [11]. The electrical isolation in DC data centers can be
achieved through incorporation of transformers in the DC to DC conversion stage
as the high DC voltage is stepped down to 12 or 48 V for the IT equipment input.
Recent developments in transformer design and wide bandgap devices [33,34] and
optimization approaches [53] enabled high efficiency and compact converters for
this DC to DC conversion stage. Also, in order to provide only electrical isolation
(without voltage conversion), unity transformation ratio is demonstrated in DC to
DC converters for data center applications at 400 [54] and 48 V [55].
Although electrical noise, ground loops and circulation of DC currents may
still be present in DC data centers, enforcing the electrical isolation may not be the
DC data centers 355
only and must-have practice to overcome such issues. For example, modern com-
munication links typically provide inherent isolation, either through the medium
itself (fiber optics), signal isolation transformers (Ethernet), or AC coupling capa-
citors (high speed serial links). Provided adequate system grounding, high safety
may be obtainable without requiring inherent electrical isolation. An example of
this transition is the case of grid-tied PV inverters, which until recently were
required to have galvanic isolation in the United States, while transformer-less
inverters were adopted earlier in Europe since they offered higher power efficiency
and density at a lower cost [56]. Similar efforts may lead to the elimination of
electrical isolation from data center power distribution architectures in the future.
14.6.2 Grounding
Proper grounding is an essential requirement in power infrastructure both for safety
and signal integrity. Grounding contributes to the safety from electrical hazards by
routing damaging currents away from the IT equipment and personnel. Although
safety is the primary driver for choosing a grounding configuration, proper
grounding also enables a common voltage reference for the overall electrical sys-
tem in the data center including the power infrastructure and communication
equipment.
A review of grounding concepts for DC data centers can be found in [57], and
ETSI EN 301605 [13] defines the grounding standard for high voltage DC dis-
tribution in data centers. Both recommend high resistive mid-point ground as
depicted in Figure 14.5 for DC data centers. This configuration keeps the fault
current flow at a level that prevents both electric shock of the human body and fire
hazard to the IT equipment.
14.6.3 Wiring
Wiring is an important consideration when transferring energy from sources
to loads in data centers. Cable size for wiring in data centers is determined by
considering maximum expected current, allowed voltage drop between the source
and load, and target worst case power transfer efficiency. Moreover, electrical
Protective earthing
Line + conductor
Rack
DC
+ IT Eq.
_
Blade
DC
Line –
Main earthing terminal
{
{
{
insulation to protect from short circuit faults must be considered, with some dif-
ferences between AC and DC rating. Because of the difference in the peak and idle
power of the IT equipment, the maximum expected current in data centers results in
oversizing the cables for the average operation. In addition, allowed voltage drop is
restricted by standards and also depends on line regulation capabilities of the front-
end power electronics converter of the load. On the other hand, although the wiring
power efficiency has little effect on overall power infrastructure efficiency, the
target worst-case power transfer efficiency may result in increasing the cable size.
Here, the differences between AC and DC data center wiring are not viewed as
significant, and DC wiring does not appear to a significant hindrance to DC data
center adoption. A study of wiring practices for high voltage DC data centers can
be found in [58].
14.6.4 Connectors
Connectors are another key element for transferring energy from the cables to the
IT equipment or power conversion stages efficiently and reliably, which are the two
most important aspects of data center power infrastructure. The efficient energy
transfer through connectors requires minimal contact resistance; otherwise, sub-
stantial energy loss may occur which conflicts with the high-efficiency target of
data centers as mentioned in Section 14.3. In addition, the connectors must also be
able to sustain possible arcs and in-rush currents as the IT equipment or power
converters are connected and disconnected during faults or maintenance. The reli-
able connectors are thus important to meet high uptime expectations in data centers
as discussed in Section 14.4.
In [59], the basic characteristics of a 400 V 10 A connector are provided.
Ongoing research on connectors for high voltage DC data centers focuses on arc
and in-rush current limitation in plugs and sockets by employing permanent mag-
nets [60] and embedded solid state electronics [61,62] while providing mechanical
isolation. Here, it is likely that research and product development in other DC
power transmission applications, such as solar PV and electric vehicle charging
will help drive down costs and provide market acceptance.
center TCO and suggests measuring the TCO on a per-rack basis [64]. A three
phase rectifier for DC data centers is optimized for TCO in [30]. In [42], the cost
advantage of high voltage DC data centers is reported as 15% less in capital cost
and 36% less in lifetime cost, but the detailed analysis is missing. It should be noted
that improvements in efficiency and reliability do not translate to business deci-
sions unless the TCO analysis is incorporated into the benefits.
Critically, as TCO analysis requires accurate field data regarding reliability,
it is likely that estimates of hypothetical designs without empirical results vary
widely in their estimates. Hence, the TOC aspects of DC and AC data centers are
difficult to assess at this time. Moreover, the major corporation that carefully track
these metrics (e.g., Facebook, Amazon, Google, etc.) generally view these numbers
as key competitive features of their respective designs and are unlikely to share
them with researchers for dissemination.
14.7 Protection
The IT equipment and power infrastructure in data centers represent large invest-
ments; therefore, any damage due to the power system faults must be prevented by
protection equipment such as fuses, relays, and circuit breakers.
Following the fault detection as mentioned in Section 14.4.1, the protection is
activated to isolate the failed IT equipment or section of the power infrastructure
from the rest of the system. Therefore, the location of the protection equipment in
the power infrastructure is vital to enable isolation of any IT equipment and power
stage whenever needed. The lack of a periodic zero-crossing of DC voltage and
current complicates the protection equipment design for DC data centers. In addi-
tion, although sometimes overlooked, abrupt interruption of a DC current results
in high current slew rates (di/dt), which induces high voltages in any parasitic
inductive loop along the power path. This may endanger semiconductors and
capacitor voltage ratings if not considered.
The periodic zero-crossing is inherent in AC distribution architectures as the
polarity of voltage and current alternates 50 or 60 times per second; however, in
DC distribution, the voltage and current are controlled to be constant values and do
not naturally cross zero. Therefore, self-sustaining arcs are more likely to be pre-
sent in DC power architectures. The basic operation principle of the protection
equipment (i.e., moving electrical contacts away from each other when triggered) is
similar in AC and DC systems, but with added challenges for DC protection. In
order to successfully extinguish an arch in DC distribution architectures, the elec-
trical contacts must move not only further away from each other but also faster than
in AC distribution. Alternatively, protection equipment involving electronics to
force the DC current toward zero can also be used. Details of the operating prin-
ciples for DC protection equipment can be found in [65]. Various aspects of pro-
tection in DC microgrids, from modeling DC arcs [66] to implementation of hybrid
(both electronic and mechanical) circuit breakers [67], have been investigated in
the literature. We note that the availability of low-cost, standards-compliant DC
358 DC distribution systems and microgrids
Power quality issues such as phase imbalance, frequency drifting, and harmonic
currents at the multipliers of fundamental grid frequency are native to ac distribu-
tion architectures but are not present in the DC distribution architectures. However,
harmonics due to specifically ASD presence for cooling loads can be severe in DC
data centers. In addition, excessive use of filter capacitance to overcome electro-
magnetic interference in power electronics converters may result in increased
in-rush currents, which reduces the power quality in DC data centers. A detailed
discussion of such power quality issues in DC microgrids (with an example of DC
data centers) can be found in [68].
14.9 Stability
The DC power distribution architectures in data centers can also present stability
challenges due to rapid variation of the IT equipment workload. The IT equipment
in data centers interfaces with the power distribution system through point-of-load
converters and thus acts as constant power loads. Dynamic utilization of the IT
equipment at different workloads results in rapid transitions between idle and peak
power consumption, resulting in challenging the power electronics converter filter
and controller design for stable operation. In addition, since high efficiency is the
primary objective in DC data centers, every power electronics converter is designed
to be highly efficient which results in little damping in the power flow path.
Therefore, the stability of DC distribution architectures is an ongoing research topic
in the literature. A comprehensive examination of stability issues in DC microgrids
with a focus on constant power loads can be found in [69]. Recently, DC electric
springs have been proposed as an active suspension system to stabilize the DC bus
in microgrids. Theoretical analysis and an experimental proof-of-concept in a 48 V
DC electric spring implementation can be found in [70]. In addition, another stabi-
lity analysis with a focus on droop-controlled DC microgrids can be found in [71].
completely DC powered data centers. Table 14.2 summarizes the examples given in
this section.
● Duke Energy Data Center, Charlotte, NC, USA
Duke Energy Data Center in Charlotte, NC is one of the earliest demonstra-
tions of distributing 380 V DC to the IT equipment racks in the United States.
In [17], 15% reduction in energy consumption compared to the conventional
AC power distribution system is reported. The system includes AC to DC and
DC to DC converters from Delta Products Corporation for the power infra-
structure, and DL 385 G7 servers from Hewlett-Packard (HP), Power795
servers from IBM, and Clarion CX4-960 data storage units from EMC as the
IT equipment. Total compute power was 100 kW, and the power delivery
architecture achieved 72.04% efficiency in 2011 [17].
● NTT Cloud, Tokyo, Japan
NTT facilities has installed a 500 kW system with 380 V high voltage DC
distribution for 4-MW scale cloud facility in February 2014. In this system,
N þ 1 redundant rectifier modules (each rated at 100 kW) provide the high
voltage DC bus to 80 server racks. The system spans 250 m2, has 5 min of
backup power provided by 380 V batteries, and uses mid-point high impedance
grounding. Further details of the implementation including safety and protec-
tion practices can be found in [72].
● Green Datacenter, Zurich, Switzerland
Collaborating with ABB, Green Datacenter installed a 1 MW 380 V DC power
distribution architecture as an extension to an existing data center in May 2012
in Zurich, Switzerland [73]. They reported 15% investment cost reduction in
comparison to an AC architecture. The overall architecture, including both
power distribution and cooling, is estimated to provide up to 20% energy
savings. The IT equipment used is compatible with high voltage DC distribu-
tion and includes HP X1800 G2 Network Storage System, HP DL385 servers,
and the HP BladeSystem c3000.
● Texas Advanced Computing Center, Austin, TX, USA
A supercomputer called ‘‘Hikari’’, which consists of 432 HP Enterprise
Apollo 8000 XL730f servers, has been installed by NTT at Texas Advanced
Computing Center (TACC) in August 2016. The Hikari supercomputer is
energized through to a 380 V DC bus which can be supplied by the solar panels
360 DC distribution systems and microgrids
at the TACC parking lot up to 208 kW. The 380 V DC bus can also be con-
nected to the AC utility using rectifiers to be functional at night [74].
● Level 3, Broomfield, CO, USA
Level 3, a US-based service provider, has recently published their transition
from 48 to 380 V DC distribution [75]. The laboratory implementation is
considered as a proof-of-concept for 380 V DC distribution. The installed
capacity is 60 kW, although extension up to 288 kW is mentioned. Currently,
two 60 kW rectifiers supply the high voltage DC bus at 380 V, followed by two
30 kW 380 V to 48 V DC to DC converters. In addition, the current imple-
mentation has 15–30 min of backup power provided by 380 V batteries, and
preferred þ/190 V with mid-point high impedance grounding as recom-
mended by the ETSI standard. The system was put in service in June 2016.
necessarily be bulkier and more expensive. In addition to the increased cost asso-
ciated with more advanced circuit breaker technology, the cost for DC breakers is
also higher due to the smaller market size than that for AC breakers.
14.12 Conclusion
Data centers consume a significant amount of electric energy; therefore, their power
delivery architectures require critical consideration for their sustainable growth. In
this chapter, we discussed the efficiency, reliability, integration with renewable
resources, installation, and protection in the DC power delivery architecture in data
centers. We also provided a short list of some existing DC data center installations and
our thoughts on the critical obstacles for widespread acceptance of this architecture.
Although high voltage DC distribution throughout the entire data center facility is still
not widely utilized, we expect that the performance outcome of the early adopters of
this architecture will promote the use of entirely DC power distributed data centers.
References
[1] Shehabi A, Smith SJ, Sartor DA, et al. United states data center energy usage
report. Ernest Orlando Lawrence Berkeley National Laboratory; 2016.
LBNL-1005775.
[2] Rasmussen N. AC vs. DC power distribution for data centers [White Paper];
2011. Available from: http://www.apc.com/salestools/SADE-5TNRLG/
SADE-5TNRLG_R6_EN.pdf.
[3] Koomey JG. Estimating total power consumption by servers in the US and
the world; 2007. Available from: http://www-sop.inria.fr/mascotte/Contrats/
DIMAGREEN/wiki/uploads/Main/svrpwrusecompletefinal.pdf.
362 DC distribution systems and microgrids
[62] Tan K, Peng C, Liu P, Song X, Huang AQ. Zero standby power high effi-
ciency hot plugging outlet for 380 VDC power delivery system. In: Proc.
31st Annu. IEEE Appl. Power Electron. Conf. Expo.; 2016. p. 132–137.
[63] Koomey J. A simple model for determining true total cost of ownership
for data centers [White Paper]; 2007. Available from: https://www.
missioncriticalmagazine.com/ext/resources/MC/Home/Files/PDFs/(TUI3011B)
SimpleModelDetermingTrueTCO.pdf.
[64] Rasmussen N. Determining total cost of ownership for data center and net-
work room infrastructure [White paper]; 2011. Available from: http://www.
apc.com/salestools/CMRP-5T9PQG/CMRP-5T9PQG_R4_EN.pdf.
[65] Salomonsson D, Soder L, Sannino A. Protection of low-voltage dc micro-
grids. IEEE Trans Power Delivery. 2009 Jul;24(3):1045–1053.
[66] Uriarte FM, Gattozzi AL, Herbst JD, et al. A dc arc model for series faults in
low voltage microgrids. IEEE Trans Smart Grid. 2012 Dec;3(4):2063–2070.
[67] Meyer JM, Rufer A. A dc hybrid circuit breaker with ultra-fast contact
opening and integrated gate-commutated thyristors (IGCTs). IEEE Trans
Power Delivery. 2006 Apr;21(2):646–651.
[68] Whaite S, Grainger B, Kwasinski A. Power quality in dc power distribution
systems and microgrids. Energies. 2015 May;8(5):4378–4399. Available
from: http://dx.doi.org/10.3390/en8054378.
[69] Kwasinski A, Onwuchekwa CN. Dynamic behavior and stabilization of dc
microgrids with instantaneous constant-power loads. IEEE Trans Power
Electron. 2011 Mar;26(3):822–834.
[70] Mok KT, Wang MH, Tan SC, Hui SYR. Dc electric springs – a technology
for stabilizing dc power distribution systems. IEEE Trans Power Electron.
2017 Feb;32(2):1088–1105.
[71] Anand S, Fernandes BG. Reduced-order model and stability analysis of low-
voltage dc microgrid. IEEE Trans Ind Electron. 2013 Nov;60(11):5040–5049.
[72] Inamori J, Hoshi H, Tanaka T, Babasaki T, Hirose K. 380-VDC power
distribution system for 4-MW-scale cloud facility. In: Proc. IEEE Int.
Telecommun. Energy Conf.; 2014. p. 1–8.
[73] Schmidt T, Ligi A, Zangger E, Monachesi A. World’s most powerful dc
data center online [Press Release]. Available from: http://www04.
abb.com/global/CNABB/CNABB050.NSF!OpenDatabase&db=/global/cnabb/
cnabb054.nsf&v=BF6&e=us&url=/global/seitp/seitp202.nsf/0/C30D32AC816
B2D5BC1257A140026A31F!OpenDocument.
[74] Salazar J. New Hikari supercomputer starts solar HVDC; 2016. Online.
Available from: https://www.tacc.utexas.edu/-/new-hikari-supercomputer-
starts-solar-hvdc.
[75] Ambriz R, Kania M. A service provider’s decision to move from 48 V to
380 V powering: the problem statement, technical assessment, financial
analysis and practical implementation plan. In: Proc. IEEE Int. Tele-
commun. Energy Conf.; 2016. p. 1–7.
[76] Lei Y, Barth C, Qin S, et al. A 2-kW single-phase seven-level flying capa-
citor multilevel inverter with an active energy buffer. IEEE Trans Power
Electron. 2017 Nov;32(11):8570–8581.
Chapter 15
DC microgrid in residential buildings
Rajeev Kumar Chauhan1, Francisco Gonzalez-Longatt2,
Bharat Singh Rajpurohit1, and Sri Nivas Singh3
15.1 Introduction
A microgrid is an innovative control and management architecture at the distribu-
tion level, which makes it easy to implement smart grid techniques at power dis-
tribution level [1]. Depleting of fossil fuels, reducing the emission of greenhouse
gases and increasing energy demand are main factors responsible for the growing
penetration of photovoltaic (PV) generators to power distribution grids [2]. The
distribution generation (DG) has the potential to provide ancillary services under
these circumstances. For example, they may have many functions such as instan-
taneous power reserve, emergency supply and peak power saving. The PV system
with DG is a good solution for electrification to poorly grid connected or isolated
area. This concept easily works in both the AC and DC microgrid system [3].
However, the implementation of the DC microgrid simplifies and provides the
opportunity to integrate the renewable energy resource (RES) such as PV system,
and fuel cells (FC) without converting into AC [4]. These are intrinsically DC
power source at higher efficiency. Additionally, the loads used in buildings may be
DC or AC load which can be converted into DC. The rapid developments in the DC
technologies, the DC loads, are more energy efficient than the AC loads [5]. The
DC output power of PV systems can be used directly without conversion into the
low voltage direct current (LVDC) distribution system [6].
There are no inductive, capacitive and skin effects in the DC system. The DC
system has a better voltage regulation than the AC system because there is no
inductive effect on steady state [3]. Additionally, the power loss due to charging
and discharging of the capacitor is eliminated in the DC system. So overall, there is
less power loss in the DC system [7]. Due to the absence of the skin effect, the
entire cross-section area of the line conductor is used in the DC system. It means a
thinner conductor could be used for DC systems, and it reduces the line conductor
weight. Additionally, as the cross-sectional area is inversely proportional to the
1
School of Computing and Electrical Engineering, Indian Institute of Technology Mandi, India
2
Department of Electrical Engineering, University of Loughborough, United Kingdom
3
Department of Electrical Engineering, Indian Institute of Technology Kanpur, India
368 DC distribution systems and microgrids
conductor resistance, the DC system has less line resistance than the AC system [7].
In this way, the direct current distribution system improves the system efficiency.
Additionally, the local utilisation of the energy reduces the transmission and dis-
tribution losses and may reduce the shortage of electricity. In this chapter, analysis
and comparison between AC and DC microgrid in residential buildings have been
done based on appliances, converters and their power losses in both systems. The
layouts for LVDC distribution network have been discussed. Both unipolar and
bipolar layouts of LVDC system have been discussed. Two microgrid system
configurations have been discussed: AC residential building [i.e. AC distribution
system (ACDS) with DC appliances] and DC residential building (i.e. distribution
network with DC appliances).
MGCC
LC LC LC LC LC
AC-1
system
Hydro- DG DG PV
unit-4 unit-3 arrays
turbine
AC
loads
AC
loads
LVAC line LVAC line
DC ESS WECS
AC-2 loads
system
LVAC: Low voltage alternating current; ESS: Energy storage system;
WECS: Wind energy conversion system; PCC: Point of common coupling
Figure 15.2 The general structure of AC microgrid with DG units and mixed types
of loads
370 DC distribution systems and microgrids
the network consisting of the DG units and load circuits can form a small isolated
AC electric power system, i.e. an ‘AC microgrid’. During normal operating
conditions, the two networks are interconnected at the PCC, while the loads are
supplied from the local sources (e.g. the RES-based DG units) and if necessary
from the utility. If the load demand power is less than the power produced by DG
units, excess power can be exported to the public utility (PU).
AC DG
system unit-3 WECS
Hydro- DG
unit-4
turbine
AC
loads
PCC
AC
loads
LVAC line LVAC line
LVDC line
Sensitive
Public
DC load
utility
DG DG
unit-1 unit-2
DC ESS PV array
DC loads
system
LVAC: Low voltage alternating current; ESS: Energy storage system;
WECS: Wind energy conversion system; PCC: Point of common coupling
Figure 15.3 The general structure of hybrid AC–DC microgrid with DG units and
mixed types of loads and generators
DC microgrid in residential buildings 371
which are having different types of sources, and loads are the type of AC–DC
systems [9].
DC-1 DG
system unit-3 WECS
Hydro- DG
unit-4
turbine
DC
loads
PCC
DC
loads
LVDC line LVDC line
DG DG
unit-1 unit-2
DC ESS PV array
DC-2 loads
system
LVDC: Low voltage direct current; ESS: Energy storage system;
WECS: Wind energy conversion system; PCC: Point of common coupling
Figure 15.4 General structure of DC microgrid with DG units and mixed load
and generator types
372 DC distribution systems and microgrids
microgrid system interconnected with the main systems at PCC which can be a
medium voltage AC network from the conventional power plants or an HVDC
transmission line connecting an offshore wind farm.
There are several topologies used by the DC microgrid. However, the unipolar or
bipolar type structure is typically used to configure the DC microgrid.
DC/AC
converter 1
AC/DC Consumer 1
main
PCC converter
Consumer 2
DC/AC
Public converter 2
utility Consumer 3
DC/AC
converter 3
DC/AC
converter 1
AC/DC Consumer 1
main
converter
PCC
Consumer 2
DC/AC
Public converter 2
utility
Consumer 3
DC/AC
converter 3
LDC ¼ g c (15.2)
If there is a variable DC load connected to the system, the calculation of total daily
load is as follows:
X
W
LDC ¼ gj cj (15.3)
j¼1
where, g1, g2, . . . , gW and c1, c2, . . . , cW are variable DC loads in amperes and
different time instants at which DC loads are switched ‘ON’ in a day.
374 DC distribution systems and microgrids
If the AC load is connected to the system, then the AC voltage must be con-
verted to DC voltage and the inverter efficiency is also considered. The DC load of
the AC system can be expressed as:
k
g¼ (15.4)
VDC hinv
Table 15.1 shows the characteristic values of a DC system based on the type of load
(AC or DC load). If the DC load 2.4 kW is supplied at 24 and 48 V, the TDL is
2,400 and 1,200 A h, respectively. In the case of 2.4 kW AC load at 120 and 220 V
supplied by the same efficiency inverter (92%), then TDL is 2,609 A h remains the
same but higher than the TDL as obtained in the case of 2.4 kW DC load at 24 V. If
a 2.4 kW AC load supplied at 220 V by the 95% efficiency inverter, the obtained is
TDL 2,526 A h. The above calculation shows that TDL depends upon the DC
system voltage and inverter efficiency. If the inverter efficiency is high, the losses
in DC to AC conversion are less, and the system TDL will be reduced. As the
system voltage is high, the system TDL will also decrease, i.e. system losses may
decrease. On the other hand, if there is a lower inverter efficiency or DC supply
voltage, the TDL will be higher. It means the battery bank (BB) will discharge at
the fastest rate, which may decrease the battery lifetime and efficiency.
system and makes the system more economical. The power transfer in a DC system
can be expressed as
P¼V I (15.6)
The current in a DC system can be expressed as
P
I¼ (15.7)
V
where P is the transferred power, V and I are the system voltage and current,
respectively. As shown in (15.7), if DC voltage is applied instead of AC voltage
then the insulation can bear higher voltage stress, which allows applying higher DC
voltage for the same system. The current is inversely proportional to the voltage.
It means as the system voltage increase, their current will decrease. Therefore, the
system copper losses (p) will also decrease.
p ¼ I 2R (15.8)
where p represents power losses in the system and R is the resistance of the feeder
cable.
The AC and DC system comparison in terms of current and power losses can
be found in Table 15.2. The current and power losses in the 325 V DC system
(equivalent to 230 V AC system) are approximately one-third and a half respec-
tively as compared to the 230 V AC system (single-phase AC voltage level in
India). Additionally, it has around 2/3 times and 7/8 times less current flow and
power losses, respectively, as compared to the 110-V AC system (single-phase AC
voltage level in the United States).
Table 15.3 shows the energy savings, which can be obtained by switching from
AC technologies to the most energy efficient DC-internal technologies. It is seen
from Table 15.3 that the total energy saving using DC technology in the residential
loads is varying between 30% and 71%. Some rectifier losses have already existed
in a few cases like a refrigerator, cloth washer and fans, in AC system.
The integration of DC sources to conventional AC system necessitates the
introduction of DC–AC converter at the generation end, thereby adding conversion
losses and complexity [11]. In last two decays, the continued development of DC
technologies to produce an energy efficient DC appliance is a cause of significant
decrement in the building load but insists on introducing AC–DC converter and
Table 15.3 Possible energy saving using most efficient DC internal technology-
based appliances
increase the conversion loss and complexity of the system [12]. The details of the
voltage and power ratings of the DC-appliance and their AC–DC converters effi-
ciency to connect these appliances to the conventional AC system can be found in
Table 15.4. The AC–DC converter efficiencies vary from 78% to 90% according to
DC microgrid in residential buildings 377
Table 15.4. It can be noted that higher the converter power rating is, the higher
is the AC–DC efficiency, as the highest efficiency 90%, which is in the case
of a hybrid car with a converter power of rating 3,000 W. The cell phone
converter of 4-W power rating has the lowest efficiency of 78%, as mentioned
in Table 15.4.
R2 R4 R5
Consumer TV
portal
Car
SST Refrigerator
Main R6 Washing
AC–DC–AC machine
Transformer switch converter
11 kV / 440 V AC board
~
=
Battery
bank
R1 = Study room
Positive (+) R2 = Entertainment room
Solar
Neutral ( ) R3 = Bed room
panels
Negative (–) R4 = Kitchen
R5 = Garage
R1 R3 Charging pad for
R6 = Laundry, control room
Computer portable appliances
SB = Switch board
SB SST: Solid-state transformer
Garage
R2 R4 R5
Consumer TV
portal SB Car
SST Refrigerator
Main R6 Washing
AC–DC converter switch machine
230 VAC/24 V DC board SB
Battery
bank
PACB ¼ PA þ pC (15.9)
X
j¼na
PA ¼ Paj (15.10)
j¼1
The total power loss (pC) in the converters is the addition of power loss in the
internal converters of the appliances (pac) and the power loss in the source con-
verter (psc) and can be expressed as
X
j¼na X
j¼ns
pC ¼ pacj þ pscj (15.11)
j¼1 j¼1
X
j¼na X
j¼nc
PDCB ¼ paj þ pcj (15.12)
j¼1 j¼1
Central control
SCADA Web server station
Management
level
Primary network
Gate way
Secondary network
RTU RTU RTU RTU RTU
Automation level
RTU: Remote
terminal unit Sensor Actuator Field level
the device to the automated level, but these devices are working with a different
protocol compared to the devices connected to primary network devices directly.
2,000
DC residential building
AC residential building
1,500
Power (W)
1,000
500
0
0.00 4.00 8.00 12.00 16.00 20.00 23.59
Time (h)
2,000
PV: DC residential building
PV: AC residential building
1,500 PU: DC residential building
PU: AC residential building
Power (W)
1,000
500
0
0.00 4.00 8.00 12.00 16.00 20.00 23.59
Time (h)
in the DC compatible appliances of the building, power loss in the internal con-
verter of appliances, power loss in the BB and PV converter as represented in
Figure 15.10 by the red line. The building demand always remains higher for AC
residential building compared to the DC residential building.
The PU and PV power curves for AC and DC residential building are shown in
Figure 15.11. The PU and PV power curve remains always lower for the DC resi-
dential building compared to the AC residential building. Different operation
modes are discussed (Mode I–Mode III).
Mode I (PU as a source): The PU supplies the building load during the
0.00–03.00 and 21.30–23.59 h time interval. In the case of AC residential building,
all the building loads are DC compatible and connected via AC–DC internal
384 DC distribution systems and microgrids
converters. The converters loss is the combination of power loss in the internal
converters of the switch ‘ON’ appliances, and power equation can be expressed as
Xao
j¼n
Ppu ðtÞ ¼ paj ðtÞ þ pacj ðtÞ (15.13)
j¼1
where Ppu(t) is the PU power at the time instant t, Paj(t) is the power consumption
in the jth appliance, nao is the number of switched ‘ON’ appliances in the building
and pacj(t) is the power loss in the internal converter of the jth appliance.
Moreover, in the DC residential building, all the building loads are DC com-
patible and directly (without any converter) connected to the 12 V bipolar DC
distribution system. The load is connected to the PU via AC–DC converter. So the
power loss in the converter is the power loss in the PU converter. The power
equation can be expressed as
Xao
j¼n X
j¼ns
Ppu ðtÞ ¼ paj ðtÞ þ pscj ðtÞ (15.14)
j¼1 j¼1
where pscj(t) is the power loss in the jth source converter, ns is the number of power
source converters.
Mode II (BB as a source): The BB supplies the building load during the
03:01–07:21 and 19:51–20.00 h time interval as represented in Figure 15.12. In the
AC residential building, all the building loads are DC compatible and connected via
AC–DC internal converters. The converters losses are a combination of power loss
in the internal converters of a switch ‘ON’ appliances and the BB inverter. The
power equation can be expressed as
X
j¼na
Xj¼ns
Pbb ðtÞ ¼ Padj ðtÞ þ padcj ðtÞ þ pscj ðtÞ (15.15)
j¼1 j¼1
1,500
DC residential building
1,000 AC residential building
500
Power (W)
–500
–1,000
–1,500
0.00 4.00 8.00 12.00 16.00 20.00 23.59
Time (h)
In the case of DC residential building, all the building loads are DC compa-
tible. The appliances and DC power source (i.e. BB) are directly (without any
converter) connected to the DC bus. Therefore, the power loss in the sources con-
verters and internal converter of the appliances remain zero in this mode as
represented by a green line in Figures 15.13 and 15.14, respectively. The power
equation can be expressed as:
X
j¼na
Pbb ðtÞ ¼ Paj ðtÞ (15.16)
j¼1
Mode III (BB and PV): The PV generates the power during the 07.22–19.50 h
time interval. The BB balances the power by supplying the surplus load and
absorbing the surplus PV power. In the case of AC residential building, all the
400
DC residential building
AC residential building
300
Power (W)
200
100
0
0.00 4.00 8.00 12.00 16.00 20.00 23.59
Time (h)
200
DC residential building
AC residential building
150
Power (W)
100
50
0
0.00 4.00 8.00 12.00 16.00 20.00 23.59
Time (h)
building loads are DC compatible and connected via AC–DC internal converters.
The converter losses are the combination of power losses in the internal converters
of a switch ‘ON’ appliances including with PV and BB inverter. The power
equation can be expressed as:
X
j¼na
Xj¼ns
Ppv ðtÞ Pbb ðtÞ ¼ Padj ðtÞ þ padcj ðtÞ þ pscj ðtÞ (15.17)
j¼1 j¼1
In the case of DC residential building, all the building loads are DC compatible.
The DC appliances and DC power source (i.e. BB and PV) are directly connected to
the DC bus without any converter. Therefore, the power loss in the source con-
verters and internal converter of appliances remains zero in this mode as repre-
sented by the green line in Figures 15.13 and 15.14, respectively. The power
equation can be expressed as:
X
j¼na
Ppv ðtÞ Pbb ðtÞ ¼ Paj ðtÞ (15.18)
j¼1
15.12 Conclusions
In this chapter, the concept of DC microgrid for the residential buildings has been
discussed. Comparison of AC vs DC system and DC microgrids architecture has
been discussed. The distribution topologies discussed in this chapter are very
helpful to understand the most efficient way of interconnection. The data of energy
dissipated in DC appliances, and the cable cost data representing are used to show
the correlation of different parameters associated with the losses. This chapter
demonstrates different configurations for both the ACDS and DC distribution
system. A power system control strategy based approach is used for the voltage
DC microgrid in residential buildings 387
References
16.1 Introduction
Climate changes induced by the use of fossil fuels and also the shortage of such
energy have been the major driving forces to develop environmental-friendly
renewable energy systems [1,2]. Many attempts have been made to advance
renewable energies, two of which are wind power and solar photovoltaic (PV)
power that are still the most favorable, as it is seen in Figure 16.1. Depending on
the power ratings, PV systems are mostly installed at a residential-scale, while wind
power systems are connected to medium-voltage (MV) or high-voltage (HV) power
grids. Hence, for microgrid applications, PV energy is more feasible also due to its
high scalability. The still declining price of PV panels is another reason for a
massive deployment of PV microgrid systems. Notably, in those applications,
power electronic converters (e.g., dc–dc converters and dc–ac inverters) are the
key, which converts the dc power from PV panels to another type of source (e.g., dc
at a different level or ac power with a fixed amplitude and a constant frequency).
Additionally, the high penetration rate of dc loads including energy storage
(ES) systems (e.g., batteries and electric vehicles) makes dc microgrids a promising
solution for energy integration [3,4]. For one thing, in the context of a large-scale
utilization of dc loads, the dc microgrid architecture eliminates at least one con-
version stage (i.e., ac–dc rectification), and thus, the entire system efficiency
maybe is improved. For another thing, it simplifies the system control, and the
frequency stability is addressed to some extent, which, however, is a bigger chal-
lenge in ac power systems. Yet, despite the above compelling reasons to deploy
PV microgrids, issues are also encountered [4]. For instance, the PV power is
fluctuating, possibly leading to dc bus voltage variations, being harmful to sensitive
equipment. Nevertheless, employing dc microgrids has been increasingly
acknowledged in recent years, and a great deal of the focus has been put on relia-
bility, efficiency, and cost-effectiveness, i.e., reducing the cost of PV energy.
In practice, dc microgrids are utilized in various sectors, ranging from small-scale
1
Department of Energy Technology, Aalborg University, Denmark
390 DC distribution systems and microgrids
466.5 GW
Hydro
Marine energy
109.7 GW
Wind energy
Solar energy
Bioenergy
1,243 GW 295.7 GW Geothermal
Figure 16.1 Global renewable power capacity composition in 2016 (more than
2,000 GW in total) [2], where hydro power also includes pumped
storage and mixed plants; solar energy includes solar photovoltaic
(290.8 GW) and concentrated solar power (4.7 GW)
HVDC
ac Utility system Fuel cell PV panels Wind farms
ac dc dc dc ac
dc dc dc dc dc
LVDC–MVDC
dc dc dc dc dc dc
dc dc dc dc dc ac
Sensitive ac Loads
Energy loads
dc Industrial dc Data Electric
facility centers vehicle storage
dc dc
dc dc
+170 V 3-phase 200 V
ac Utility ac balancer dc
Voltage
0V
dc ac
–170 V
Single-phase
dc 100 V
dc dc
ac
dc 12-48 V
dc
dc
Fuel cell
LVDC 70 V Distribution
dc Power
dc and
control
protection
PV Storage Essential,
panels (battery) semi-essential,
nonessential
loads
reliability and flexibility are necessary for missions in space stations, reliable dc
microgrid with high robustness in both hardware and software is required and a
promising solution [9,10].
The loads in space crafts can be categorized into three groups: essential, semi-
essential, and nonessential loads. dc Microgrids in space applications should solely
supply the on-board loads. Therefore, nonessential loads are typically adopted to
be sheded during critical situations to regulate the voltage. It should be noted that
the dc microgrid employs special protection systems to maintain the stability in
space stations.
PV panels PV panels
dc
dc
dc
ac Link ac Link dc
Generator Generator
ac dc
dc dc
dc ac
ac ac
dc dc dc dc
ac ac ac ac
Motor Motor
Motor Motor
marine applications [14]. With such structures, not only the operation cost is
reduced but also the maintenance of engines is lowered due to more efficient
operation. Additionally, dynamic response enhancement and maneuverability are
advantages of dc microgrids for marine applications.
ac Utility
ac
Storage dc
(battery)
PV panel
Charge bus
Discharge bus
Storage LVDC 380 V
(battery)
PDU
MNS/RPP MNS/RPP
dc/dc dc/dc
PSU PSU
Load Load
Figure 16.6 LVDC microgrid architecture for data center. PDU, power
distribution unit; RPP, remote power panel; PSU, power supply unit;
MNS, switchgear
DC microgrids for photovoltaic powered systems 395
ac dc dc dc ac
dc dc dc dc dc
LVDC 380 V
dc dc dc dc dc
ac dc dc dc dc
ac Loads
TV,
LED light
Washer, computer
Stove dryer
Figure 16.7 A residential dc microgrid in a future smart home, where the arrow
indicates the power flow direction
396 DC distribution systems and microgrids
A residential dc microgrid typically has two voltage levels, i.e., low voltage at
24 and 380 V. The 24-V system powers LED lighting and small appliances, while
high-power heating equipment is supplied by the 380 V. In such a residential dc
microgrid, the power is fed from devices and controllable converters, which can
provide a wide range of voltages as well as current limitations (protections).
Therefore, electromechanical protection devices can be eliminated and the system
safety can be maintained.
ac dc ac dc ac
dc dc dc dc dc
Bidirectional dc–dc
converter
dc dc dc
ac ac ac
M M M
dc dc dc dc
dc dc dc dc
Local dc-bus
dc dc
dc dc
Common dc-bus
Loads
dc dc dc dc
dc dc dc dc
Loads
Iph D V
ID
Iph ID I
V V V
I
Rs I
Rs = 0
Iph D V Rs > 0
ID
(a) (b) V
Figure 16.12 Single-diode PV model with series resistance (Rs represents the
series resistance): (a) electrical circuit and (b) V–I characteristic
I is a reverse bias current of the diode, k is the Boltzmann’s constant, q is the charge
of an electron (¼ 1:602 1019 C), and T refers to the absolute p–n junction
temperature in Kelvin, and a is the diode ideality factor typically within 1–2 [20].
Nevertheless, this simplified model cannot demonstrate the real behavior of a PV
cell. Thus, a single-diode model with series resistance has to be introduced.
where Rs refers to the series resistance. In order to model the behavior of a PV cell
in high-temperature conditions and manufacturing defects, an additional shunt
resistor is also applied to the model [23].
I
Rs I
Iph D Rsh V
Rsh
ID IRsh Rsh finite
(a) (b) V
Figure 16.13 Single-diode model with shunt resistance (Rsh represents the shunt
resistance): (a) electrical circuit and (b) V–I characteristic
DC microgrids for photovoltaic powered systems 401
Rs I
Iph D1 D2 Rsh V
ID1 ID2 IR
sh
where a ¼ 1 and a ¼ 2. Hence, the diode (D1) in Figure 16.14 models the effect of
the HV and the diode (D2) models the effect of the low voltage. A multiple-diode
model is also presented in the literature which increases the model complexity.
As the power of a single PV cell is very low, a great deal of PV cells should be
connected in series and parallel to form a PV module/panel up to several hundred
watts. In a similar way, the PV modules/panels can be connected in series and
parallel to further achieve high-power and HV PV arrays. The number of series
and parallel panels in each PV array is dependent on the P–V characteristics of each
panel, the input voltage, and the rated current of the interfacing converter.
Furthermore, as the output power of PV panels varies following the solar
irradiance and ambient temperature, an interfacing converter is required to control
the output current of the PV panels to harvest the maximum power. This is called a
maximum power pointing tracking (MPPT) control. In the next section, different
dc–dc converters for PV applications are presented.
Generally, dc–dc converters can be implemented between PV panels and the dc-bus
or a dc–ac converter in order to control the output voltage/current of the PV panel
for power optimization. Thus, different power converter topologies for isolated and
nonisolated converters and various control strategies have been presented [24–26].
In the following, transformerless and transformer-based (typically, high-frequency
transformers) topologies are explained. Since this chapter focuses on dc microgrids,
some dc–dc converters are introduced in this section. Furthermore, some com-
mercial inverter topologies will be introduced, where the first conversion stage can
be used in dc microgrids as the interlinking between PV panels and the dc-bus.
IPV Lb
PV strings/modules
Cpv S D
Cdc Vdc
o
Cp
S Db
Lb Df
Cdc1
IPV
PV strings/modules
Vdc
Cpv Cdc2
o
Cp
Figure 16.16 Parallel input, series output bipolar dc–dc converter for PV
applications
input voltage. The input voltage depends on the PV characteristics and can be
determined by an MPPT control system to obtain the maximum power. Further-
more, the output voltage is almost a constant value, which can be controlled by the
microgrid.
In order to increase the efficiency of the boost converter, a dc–dc converter
with a series output and a parallel input can be used. This structure can be seen in
Figure 16.16. It is worth to note that the voltage step-up of this boost converter
structure is also enhanced [27].
Boost converters can be used as an interface between PV panels and dc
microgrids and/or between PV panels and inverters. Furthermore, a boost converter
is employed in industrial PV inverters as the first stage in the entire energy con-
version process to control the PV panels to operate at MPPT.
For example, Figure 16.17 shows a commercial topology of a boost converter,
introduced in SMA Sunny Boy 5000 TL.
In order to increase the voltage level and standardization of the output PV
voltage, safety, transformer-based topologies, which isolate the PV side-circuit
from the grid side, have been presented.
DC microgrids for photovoltaic powered systems 403
PV string 1 Vdc
IPV Lb Db
Cpv S Cb
IPV Lb Db
PV string 2
IGrid
S
Cpv S Cb
LGrid
Grid S
IPV Lb Db
PV string 3
Cpv S Cb
IPV
S S
PV string
Cpv
Vdc
HFT
S S
Vdc
IPV Cb
D1 D2
S1 S2
PV string 1
Cdc
Cpv
S4 HFT
S3
D3 D 4
IPV Cb D1 D2
S1 S2 S1 IGrid S3
PV string 2
Cdc LGrid
Cpv S4
S2 Grid
S4 HFT
S3 D3 D 4
Cb
IPV D1 D 2
S1 S2
PV string 3
Cdc
Cpv
HFT
S3 S4
D3 D 4
The full-bridge diode-based converter rectifies the output voltage of the trans-
former to a dc voltage.
Figure 16.19 shows a commercial high-frequency transformer-based dc–dc
converter designed by Powerlynx. In this application, dc–dc full-bridge transformer-
based converters are employed as an interface converter between PV modules/strings
and the dc–ac inverter.
Another transformer-based topology, i.e., the push–pull converter, is shown in
Figure 16.20, which is one of the most economic dc–dc converter configurations.
Minimum number of switching devices, the galvanic isolation, and the simple
design are the main reasons making this topology more popular.
D1 D3 D1 D3
PV string
PV string
HFT HFT
S1 S2 D4 D2 S1 S2 D4 D2
(a) (b)
Figure 16.20 Single-phase, push–pull topology converter: (a) with hard switching
operation and (b) with clamped snubber circuit
PV modules
ac Utility
1,500 V
dc
ac
Transformer
Combiner box
Voltage Voltage
SoC range
SoC 100%
SoC 0%
Rated current
Rated current
Rated current
MPPT range
SoC range
(a) 0 Irated (b) Icharge 0 Idischarge
units are operating under the MPPT power and share the load based on their droop
gains [33]. Furthermore, the maximum supplied power by the PV units are limited
by the MPPT power as shown in Figure 16.22(a). By decreasing the microgrid
demand, the dc voltage will be increased, and hence, the PV controller enters into
the droop-controlled mode.
The battery units also should be controlled in terms of power sharing and
energy level, i.e., SoC level. An autonomous droop characteristic for distributed
batteries in the microgrid can be shown as Figure 16.22(b), where the maximum
charging and discharging power can be set taking into account the SoC level of
batteries. Hence, by adapting the droop characteristics in the local controller, the
energy level of each battery can be controlled. Furthermore, the power sharing,
both charging and discharging powers, among the batteries can be controlled by
properly setting the droop gains shown in Figure 16.22(b).
Since the scope of this chapter focuses on the PV units in microgrids, possible
control structures of PV converters are shown in Figure 16.23. For simplicity,
a dc–dc boost converter is considered as shown in Figure 16.23(a). For the control
structures shown in Figure 16.23(b) and (c), a supervisory control changes the PV
operation mode following the system generation-load balance constraints. How-
ever, in the control structures shown in Figure 16.23(d) and (e), the operation mode
of the PV units will be determined by the local parameters, i.e., the MPPT power
and the droop-induced power. These approaches are suitable for autonomous droop
techniques.
PV converter
Ioutput
Iin
Voutput IL Vin
MPPT
(a)
VMPPT IMPPT
VMPPT –
PI PI
– IMPPT
Vin IL
PI
V* V* PI –
PI PI –
– – IL
IL Supervisory Voutput
Voutput Ioutput Supervisory
control Droop Rd
Rd Ioutput Droop control
(c)
(b)
Vinput
VMPPT –
IMPPT PI
MIN
PI
V* PI PI V* PI –
– – –
0 IL
Voutput IL Voutput
Droop Rd Ioutput Rd Ioutput
Droop
(d) (e)
Figure 16.23 Adjustable droop control for MPPT-based units (e.g., PV):
(a) dc–dc boost converter for PV, (b), (c) mode transition strategy,
and (d), (e) seamless strategy
batteries as shown in Figure 16.24. The parameters of the simulated dc grid are
given in Table 16.1. The droop characteristics of the units are set as shown in
Figure 16.23(a) and (b), where the droop gains of batteries are 2 and 1 W.
Furthermore, the MPPT currents of PV units are set to 2.25 and 2.75 A, and the
droop gains of PV units are 10 W. The batteries are considered to be fully charged
in the beginning. Simulation results are shown in Figure 16.25. The microgrid load
is 3.2 kW and the PV units are operating in MPPT mode with (2.25 þ 2.75)
400 ¼ 2 kW. The remaining load power is supplied by batteries, where the output
current of the first battery is two times that of the other one following their droop
gains as shown in Figure 16.25(b). Therefore, the batteries are operating in the
droop mode and forming the dc-link voltage, and the maximum power of the PV
units is injected into the microgrid. Once the load is decreased to 1.6 kW, the dc-
link voltage is increased as shown in Figure 16.25. Following Figure 16.22(b), by
increasing the dc voltage, the PV units should operate in the droop-controlled mode
and decrease the output current to balance the demand of the system. The load
sharing among the dc sources is shown in Figure 16.25(b). The PV units are
working in the droop-controlled mode and supplying the equal current following
their droop gains. Furthermore, the battery currents are zero, since they are con-
sidered as fully charged. In this mode, the dc voltage is formed by the PV units.
DC microgrids for photovoltaic powered systems 409
Ldc
Vi,PV1
Vo,PV1
PV 1
Cdc Cdc
MPPT
–
PI PI IMPPT
–
MIN
V*
1/R1
PV converter 2 Ii,PV2
Ldc
Vi,PV2
Vo,PV2
PV 2
Cdc Cdc
MPPT
–
PI PI IMPPT
–
MIN
V*
1/R2
Battery converter 1
Io,BT1 Ldc
Battery 1
Vo,BT1
Ii,BT1 Cdc
Cdc
Idch
VPCC V*
Battery converter 2
Io,BT2 Ldc
Battery 2
Vo,BT 2
Ii,BT2 Cdc
Cdc
Figure 16.24 A typical dc microgrid with two dispatchable units, i.e., battery
and two nondispatchable PV units
410 DC distribution systems and microgrids
Parameter Value
PV converter droop gains (W) 10, 10
Battery converter droop gains (W) 2, 1
Load power (kW) 1.6, 1.6
Operating MPPT current of PVs (A) 2.25, 2.75
Rated power of PV converter 2 kW
Rated power of battery converter 2 kW
dc Inductor of converters 2 mH
dc Capacitor of converters 500 mF
410
dc link voltage (V)
405
Battery controlled PV controlled
400
390
(a) 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2
Time (s)
4
MPPT mode Constant power mode
3
Output current (A)
1
PV 1 PV 2
0
Batt. 1 Batt. 2
–1
(b) 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2
Time (s)
16.8 Summary
dc Microgrids introduce an infrastructure for integrating dc-based energy resources,
storages and loads with high reliability, availability, and efficiency. Meanwhile, PV
systems are one of the most popular renewable resources due to its modularity as
well as lower costs. This chapter discussed different applications and structures of
DC microgrids for photovoltaic powered systems 411
References
[1] Blaabjerg, F., Yang, Y., Yang, D., Wang, X.: ‘‘Distributed power-generation
systems and protection’’ Proc. IEEE, 2017, 105, (7), pp. 1311–1331.
[2] International Renewable Energy Agency: ‘‘Renewable Energy Capacity
Statistics 2017,’’ 2017.
[3] Youichi, I., Yang, Z., Hirofmi, A.: ‘‘DC Microgrid Based Distribution
Power Generation System’’, in ‘‘Power Electronics and Motion Control
Conference, 2004’’ (IPEMC, 2004), the 4th International, 2004, 3, pp.
1740–1745. IEEE.
[4] Kumar, D., Zare, F., Ghosh, A.: ‘‘DC microgrid technology: system archi-
tectures, AC grid interfaces, grounding schemes, power quality, commu-
nication networks, applications, and standardizations aspects’’ IEEE Access,
2017, 5, pp. 12230–12256.
[5] Diaz, N.L., Dragicevic, T., Vasquez, J.C., Guerrero, J.M.: ‘‘Intelligent dis-
tributed generation and storage units for DC microgrids: a new concept on
cooperative control without communications beyond droop control’’ IEEE
Trans. Smart Grid, 2014, 5, (5), pp. 2476–2485.
[6] Elsayed, A.T., Mohamed, A.A., Mohammed, O.A.: ‘‘DC microgrids and
distribution systems: an overview’’ Electr. Power Syst. Res., 2015, 119,
pp. 407–417.
412 DC distribution systems and microgrids
[7] Dragicevic, T., Lu, X., Vasquez, J.C., Guerrero, J.M.: ‘‘DC microgrids—
part II: a review of power architectures, applications, and standardization
issues’’ IEEE Trans. Power Electron., 2016, 31, (5), pp. 3528–3549.
[8] Kakigano, H., Miura, Y., Ise, T.: ‘‘Low-voltage bipolar-type DC microgrid
for super high quality distribution’’ IEEE Trans. Power Electron., 2010,
25, (12), pp. 3066–3075.
[9] McNelis, A.M., Beach, R., Soeder, J.F., et al.: ‘‘Simulation and Control Lab
Development for Power and Energy Management for NASA Manned Deep
Space Missions,’’ in ‘‘12th International Energy Conversion Engineering
Conference’’ (American Institute of Aeronautics and Astronautics, 2014).
[10] Dever, T., Trase, L., Soeder, J.F., et al.: ‘‘Application of Autonomous
Spacecraft Power Control Technology to Terrestrial Microgrids,’’ in ‘‘12th
International Energy Conversion Engineering Conference’’ (American
Institute of Aeronautics and Astronautics, 2014).
[11] Khooban, M.-H., Dragicevic, T., Blaabjerg, F., Delimar, M.: ‘‘Shipboard
microgrids: a novel approach to load frequency control’’ IEEE Trans.
Sustain. Energy, 2017, 9, (2), pp. 843–852.
[12] Hebner, R.E., Uriarte, F.M., Kwaskinski, A., Gattozzi, A.L., Estes, H.B.,
Anwar, A.: ‘‘Technical cross-fertilization between terrestrial microgrids and
ship power systems’’ J. Mod. Power Syst. Clean Energy, 2016, 4, (2),
pp. 161–179.
[13] Im, W.-S., Wang, C., Tan, L., Liu, W., Liu, L.: ‘‘Cooperative controls for
pulsed power load accommodation in a shipboard power system’’ IEEE
Trans. Power Syst., 2016, 31, (6), pp. 5181–5189.
[14] Hansen, J.F., Lindtjørn, J.O., Vanska, K.: ‘‘Onboard DC Grid for enhanced
DP operation in ships’’ MTS Dyn. Position. Conf., Houston, 2011.
[15] Jiang, T., Yu, L., Cao, Y.: ‘‘Green Energy and Technology Energy Man-
agement of Internet Data Centers in Smart Grid.’’ Berlin Heidelberg:
Springer, 2016.
[16] Yu, L., Jiang, T., Zou, Y.: ‘‘Real-time energy management for cloud data
centers in smart microgrids’’ IEEE Access, 2016, 4, pp. 941–950.
[17] Rodriguez-Diaz, E., Vasquez, J.C., Guerrero, J.M.: ‘‘Intelligent DC homes
in future sustainable energy systems: when efficiency and intelligence work
together’’ IEEE Consum. Electron. Mag., 2016, 5, (1), pp. 74–80.
[18] Reed, G.F., Grainger, B.M., Sparacino, A.R., Zhi-Hong M.: ‘‘Ship to grid:
medium-voltage DC concepts in theory and practice’’ IEEE Power Energy
Mag., 2012, 10, (6), pp. 70–79.
[19] Kounev, V., Tipper, D., Grainger, B.M., Reed, G.: ‘‘Analysis of an Offshore
Medium Voltage DC Microgrid Environment; Part II: Communication
Network Architecture,’’ in ‘‘2014 IEEE PES T&D Conference and Exposi-
tion’’ (IEEE, 2014), pp. 1–5.
[20] Patel, H., Agarwal, V.: ‘‘MATLAB-based modeling to study the effects of
partial shading on PV array characteristics’’ IEEE Trans. Energy Convers.,
2008, 23, (1), pp. 302–310.
DC microgrids for photovoltaic powered systems 413
[21] Mette, A., Pysch, D., Emanuel, G., Erath, D., Preu, R., Glunz, S.W.: ‘‘Series
resistance characterization of industrial silicon solar cells with screen-
printed contacts using hotmelt paste’’ Prog. Photovoltaics Res. Appl., 2007,
15, (6), pp. 493–505.
[22] Pysch, D., Mette, A., Glunz, S.W.: ‘‘A review and comparison of different
methods to determine the series resistance of solar cells’’ Sol. Energy Mater.
Sol. Cells, 2007, 91, (18), pp. 1698–1706.
[23] Salam, Z., Ishaque, K., Taheri, H.: ‘‘An Improved Two-Diode Photovoltaic
(PV) Model for PV System,’’ in ‘‘2010 Joint International Conference on
Power Electronics, Drives and Energy Systems & 2010 Power India’’ (IEEE,
2010), pp. 1–5.
[24] Liu, K.-H., Lee, F.C.Y.: ‘‘Zero-voltage switching technique in DC/DC
converters’’ IEEE Trans. Power Electron., 1990, 5, (3), pp. 293–304.
[25] Lee, K.-H., Chung, E., Han, Y., Ha, J.-I.: ‘‘A family of high-frequency
single-switch DC–DC converters with low switch voltage stress based
on impedance networks’’ IEEE Trans. Power Electron., 2017, 32, (4),
pp. 2913–2924.
[26] Rehman, M.M.U., Zhang, F., Zane, R., Maksimovic, D.: ‘‘Control of Bidir-
ectional DC/DC Converters in Reconfigurable, Modular Battery Systems,’’
in ‘‘2017 IEEE Applied Power Electronics Conference and Exposition
(APEC)’’ (IEEE, 2017), pp. 1277–1283.
[27] Yang, Y., Blaabjerg, F.: ‘‘Overview of single-phase grid-connected
photovoltaic systems’’ Electr. Power Components Syst., 2015, 43, (12),
pp. 1352–1363.
[28] S. Moskowitz: ‘‘The Next Opportunity for Utility PV Cost Reductions:
1,500 Volts DC, Greentech Media,’’ 2018. https://www.greentechmedia.com/
articles/read/the-next-opportunity-for-utility-pv-cost-reductions-1500-volts-
dc, accessed February 2018.
[29] Serban, E., Ordonez, M., Pondiche, C.: ‘‘DC-bus voltage range extension in
1500 V photovoltaic inverters’’ IEEE J. Emerg. Sel. Top. Power Electron.,
2015, 3, (4), pp. 901–917.
[30] Inzunza, R., Okuyama, R., Tanaka, T., Kinoshita, M.: ‘‘Development of a
1500 Vdc Photovoltaic Inverter for Utility-Scale PV Power Plants,’’ in
‘‘2015 IEEE 2nd International Future Energy Electronics Conference
(IFEEC)’’ (IEEE, 2015), 2015, pp. 1–4.
[31] C. Crowell: ‘‘1,500-Volt Systems to Trend in 2017: Here’s What You Need
to Know – Yaskawa – Solectria Solar,’’ 2018. https://solectria.com/com-
pany/in-the-news/1-500-volt-systems-to-trend-in-2017-heres-what-you-need-
to-know/, accessed February 2018.
[32] Boroyevich, D., Cvetkovic, I., Burgos, R., Dong, D.: ‘‘Intergrid: a future
electronic energy network?’’ IEEE J. Emerg. Sel. Top. Power Electron.,
2013, 1, (3), pp. 127–138.
[33] Dragicevic, T., Guerrero, J.M., Vasquez, J.C., Skrlec, D.: ‘‘Supervisory
control of an adaptive-droop regulated DC microgrid with battery manage-
ment capability’’ IEEE Trans. Power Electron., 2014, 29, (2), pp. 695–706.
414 DC distribution systems and microgrids
[34] Gu, Y., Xiang, X., Li, W., He, X.: ‘‘Mode-adaptive decentralized control for
renewable DC microgrid with enhanced reliability and flexibility’’ IEEE
Trans. Power Electron., 2014, 29, (9), pp. 5072–5080.
[35] Peyghami, S., Mokhtari, H., Blaabjerg, F.: ‘‘Decentralized load sharing in an
LVDC microgrid with an adaptive droop approach based on a superimposed
frequency’’ IEEE J. Emerg. Sel. Top. Power Electron., 2017, 5, (3),
pp. 1205–1215.
Chapter 17
Demonstration sites of dc microgrids
Aditya Shekhar1, Laura Ramı́rez-Elizondo1,
Seyedmahdi Izadkhast1 and Pavol Bauer1
17.1 Introduction
In recent years, several demonstration projects using dc microgrids have been
implemented across the world due to some distinct advantages of dc over ac sys-
tems. The purpose of such initiatives is to validate the theoretically predicted
benefits of dc distribution in practical scenarios. This chapter presents a non-
comprehensive overview of existing demonstrations and pilot projects for a wide
range of applications such as off-grid microgrids, transportation electrification,
datacenters, residential and industrial purposes. For each application, key aspects
such as architecture, components, control, protection and socioeconomic impacts
are highlighted. A short discussion is offered on trends in voltage levels, capacity
and topology, progressing toward possible standardization approaches based on
recognized best practice.
17.2.1 Architecture
A typical architecture for off-grid dc solar home systems (SHS) is shown in
Figure 17.2. The structure is radial with a 48-V common dc bus interconnecting the
1
Department of Electrical Sustainable Energy, DC Systems, Energy Conversion and Storage (DCE&S)
Group, Delft University of Technology, The Netherlands.
416 DC distribution systems and microgrids
125 kW PV
5 W LED
bulb
Rooftop PV
Central
controller
with 48 V
dc bus
Solar panel
N: Normal line
E: Emergency line
48 V dc N
IRU E Home 1
48 V dc N
230 V ac IRU E Home 2
48 V dc N
IRU E Home 3
48 V dc N
IRU E Home 4
Grid
Bluetooth
low-energy
interface
Battery
Cloud
distributed sources, storage and loads. To guarantee power supply to critical loads,
connections are bifurcated into normal line (N) and the emergency line (E). Each
home has an inverter-less remote unit (IRU). The option of remote monitoring
using, for example, a Bluetooth interface is also available.
Demonstration sites of dc microgrids 417
17.2.2 Components
The minimum needs for a low-income household may at least consist of loads such
as a few fans, tube-lights, bulbs and mobile phone chargers. With available
opportunities, some households showed preference for entertainment loads such as
television and commercial loads such as tailoring [2–4], amounting to a demand
of 100–500 W. With use of dc, 2–2.5 times reduction in battery and solar module
size at half the cost and efficiency enhancement at upward of 15% have been
reported [1,2]. The use of dc-ready appliances is encouraged, claiming efficiency
gains of almost 50% with such practice [2], such as 1 18 W LED tube-light,
1 32 W brush-less dc fan, 1 5 W LED bulb and a mobile phone charger. The
125-Wp solar photovoltaic (PV) generation has a trade-off between cost and
availability. Storage capacity caters for load demands when neither PV nor grid is
available, designed to supply power for a downtime of 1–2 days [2]. Some ac loads
can be directly supplied from the grid or through the dc bus, even though, the use of
an additional inverter is discouraged.
17.2.5 Discussion
A summary of various dc offgrid applications worldwide with respect to voltage
levels, capacity, topology, and control is listed in Table 17.1. In most cases, the
demonstrations were dominated by low capacity (below 500 W) and low voltage
levels (48 V or below, in line with future IEEE P2030.10 Standard for rural and
remote dc), signifying that this technology is developed to meet the basic energy
418 DC distribution systems and microgrids
Figure 17.3 An illustration of a dc-based all electric ship (adapted from [14])
17.3.1.1 Architecture
Keeping in mind the redundancy requirements, two buses, the port and starboard
bus run through the length of the ship. An illustration of such a dc microgrid island
on ship is shown in Figure 17.3. Different shipboard dc electrical distribution
system architectures with trade-offs between reliability, complexity and efficiency
are discussed in [11]. An interesting compact design concept of ‘‘Power Corridor’’
toward the IPS keeping in mind the space reservations for AES is proposed in [13].
17.3.1.2 Components
As part of The Electric Ship Research and Development consortium, the University
of Texas at Austin center for electromechanics (UT-CEM) assembled a 2-kV
MVDC microgrid testbed as shown in Figure 17.4 [15].
The isolated naval microgrid is envisioned to operate at 80–100 MW with
different steady state and pulsed power loads. The following components are
highlighted: (a) gas turbine driven 3 MW high speed generator, (b) full bridge
passive diode rectifier, (c) bidirectional, active, IGBT controlled converter,
(d) 1.25 MW Toshiba model variable frequency drive capable of operating up to
300 Hz modified from its commercial version to operate as a ‘‘dc-ready’’ device
able to directly interface the load with the common dc bus, (e) Kahn model
hydraulic dynamometer capable of absorbing 4.77 MW at 18,000 rpm, that can be
controlled to mimic transient conditions in dc distribution systems.
480 V
(c)
utility (b)
(a) line (d)
ALPS
inductors
Field controllable 300 A 1,500 V
engine-driven ALPS ALPS
ac generator transformer passive rectifier
Toshiba Dyno
480 V 1.25 MW VFD
ALPS (e)
utility Kahn
motor/generator
line dynamometer
Figure 17.4 Testbed at UT-CEM for MVDC shipboard microgrid (courtesy: [15])
voltage control. The machine reacted to the power electronics and resulted in dis-
tortion in phase current, leading to flux pulsations and eddy current losses, an
important challenge to consider during design.
For stable ship operation, it is important to maintain power availability during
contingencies. It is identified that due to interconnectivity, the reconfiguration
capability maybe a potential advantage of MVDC IPS in AES [11]. While
acknowledging significant efficiency gains and lesser problems with harmonic
distortion, The ‘‘ABB’’ group of industries noted that the challenge in the design of
the marine on-board dc system was related to protection and selective localization
of faults. Their approach was a coordinated use of fuses and isolation switches
along with protective power electronic devices. Further, the controlled thyristor
rectifier connected to the generator played an added role of protection device. It was
claimed that dc fault currents can be controlled within 10–20 ms, providing a sig-
nificant reduction in fault energy as compared to the conventional ac system [16].
Demonstration sites of dc microgrids 421
17.3.2.1 Architecture
A typical dc traction system with rail return is shown in Figure 17.5. Considering
that catenaries of more than one conductors are usually not preferred [18], mono-
polar dc maybe advantageous as it simplifies the overhead catenary system for
collecting traction current and reduces costs.
dc busbar
(750 V/1,500 V)
dc feeder
cable
Upward catenary
Downward catenary
M Locomotive
M
Upward rail
Downward rail
With standard voltage levels of 750 V for urban metros and trams and 1.5–
3 kV for interurban and regional system [17], spacing between 1 and 10 MW
substations is selected between 3 and 20 km depending on the operating voltages
(higher the voltage, greater the allowable spacing). The system is sectionalized
using tie breakers to ensure isolation of faults to a limited zone of influence [23]. In
general, the zone of influence is about 8 km for a 1.5-kV dc substation with 4 kA
current [24].
17.3.2.2 Components
The key components of each dc substation are rectifiers, filters, transformers and
dc switchgear. With 4 12-car trains nearby (2 leaving and 2 approaching), each
drawing 4 kA, the demand could be as high as 25 MVA [25]. However, the nature
of the load is such that it is not expected to burden the substation for a long time
continuously. In order to minimize the related system costs, such substations
maybe designed to operate with overload capacity for a short while, typically
150%, 300% and 450% for 1 h, 1 min and 10 s, respectively, for a 3 MW, 750 V
supply [24]. The use of energy storage and tie-lines is employed in dc traction on
the Yamatoji line in Japan [19]. Apart from providing ancillary services such as
regenerative energy absorption, supplying power deficit and compensating
catenary voltage drop, at least 3% decrease in energy from dc substation feeders
was reported in the project.
17.3.2.3 Protection
Protection of 3 kV dc traction systems from lightening surges, and particularly, the
vulnerability of overhead catenary conductor to lightning strike is discussed in
[21,26]. For rail-return traction supply, issues associated with rail to earth potential
have been cautioned in [27]. Generally, a combination of ac and dc switchgear is
needed to isolate faults in traction supply systems [28]. High-speed dc circuit
breakers are employed to clear the faults within 15–30 ms. An intertripping
mechanism is incorporated while ensuring a coordinated discriminatory operation
between ac and dc side circuit breakers [29]. Further, a reverse current tripping is
incorporated in the dc switchgear to isolate faulty rectifiers from the common dc
bus. In [23], it is highlighted how dynamic nature of traction loads could mimic the
sharp rise in fault currents when train is accelerating. Therefore, the time setting of
the di/dt tripping relay should be carefully chosen based on the traction system
characteristics.
17.3.3 Discussion
Some examples of transportation electrification using dc microgrids are summar-
ized in Table 17.2. It is observed that the trend is toward high installed capacity of
operation and correspondingly medium voltage levels are chosen. The dc shipboard
IPS is controlled in a decentralized way to crucially increase the reliability, while
using radial or distributed topologies. Traction systems typically use hierarchical
control scheme in a sectioned, monopolar with rail return topology.
Demonstration sites of dc microgrids 423
17.4 Datacenters
Currently, many datacenters around the globe are tending to employ dc technology
due to its high reliability and efficiency. To highlight the importance of efficiency,
in 2006 1.5% of total US electricity consumption was from datacenters [37]. This
section provides an overview of dc datacenters at Lawrence Berkeley National
Laboratory (LBNL) and then reviews various dc datacenters around the world.
424 DC distribution systems and microgrids
380 V
480 V Transformer UPS Server 480 V UPS DC Bus Server
ac grid ac grid
ac dc ac dc dc ac dc dc
dc ac dc dc dc dc dc dc
+ dc + dc
–
dc 12 V dc
(a) (b)
Figure 17.6 Datacenter distribution system: (a) ac based and (b) dc based [38]
17.4.1 Architecture
The schematics of ac and dc datacenters studied by LBNL are shown in
Figure 17.6. The dc datacenter servers are connected to a 380 V dc bus, where from
the main ac grid is routed via uninterruptible power supply (UPS). It is indicated
that one dc/ac conversion step in UPS and one ac/dc conversion step in server
power supply is reduced for a dc based datacenter.
17.4.2 Components
The need for standardization of various dc-ready commercially available compo-
nents such as servers, connectors, power cords and in-rack distribution strips was
identified in [38]. The project compared the performance of ac and dc datacenters
with the same level of achievable functionality and claimed that the results
demonstrated an efficiency improvement by 7%–8% with dc. Out of this, it was
reported that the share of efficiency gain was 4% for UPS, 2% for transformer and
2% for server power supply. It was anticipated that with component optimization,
further improvements are possible.
17.4.4 Discussion
Table 17.3 lists the information of some dc-based datacenters from around the
world. The system level operating voltage of datacenters is standardizing at about
380 Vdc. Even the technical report of INTEL labs describing a 400 V datacenter
affirms its support to the standardization around 380 V [40]. Regarding the size,
there is a clear trend toward high power dc datacenters, as Intel Labs have reached
Demonstration sites of dc microgrids 425
up to 5.5 MW. The key challenges highlighted by the LBNL study include (a)
economics of scale may shift in favor of dc if knowledge on its benefits are prop-
erly disseminated. (b) Industry leadership in moving toward development of stan-
dardized practices is desirable for widespread market adoption on dc-based
datacenters. Demonstration sites for datacenters using dc distribution
17.5.1 dc House
17.5.1.1 Architecture
Many household appliances such as televisions, refrigerators, computers, LED
lighting, EVs as well as sources like solar PV and storage elements can be inter-
connected to form a dc nanogrid. A typical grid connected dc residential house is
shown in Figure 17.8.
17.5.1.2 Components
An example of a demo house from EPARC, Taiwan [45], is shown in Figure 17.9.
The focus of this demonstration was to adapt commonly used ac residential appli-
ances indicated in the figure to operate with dc input. Various components were
connected radially to a common dc bus of 380 20 V with about 5–10 kW
capacity.
The products were converted from ac source with power factor correction by
rerouting the protection and circuitry. It was highlighted that several appliance
manufacturers such SAMPO, TECO, Eulife, Jamicon and Fego supported this
modification of their products. It is encouraged to mark such products as ‘‘dc-
ready’’ to reinforce end-user confidence in adopting dc technologies [46]. Opera-
tional data collected showed a efficiency enhancement of 8.5% compared to ac.
426 DC distribution systems and microgrids
Traditional Proposed
PV PV
ac dc dc "DC Ready"
Loads loads
dc dc dc
ac dc dc
dc dc dc
ac dc dc
dc ac Heat pump ac Heat pump
ac dc dc
Fuel cell Fuel cell
dc dc dc
ac dc – + dc – +
dc dc dc
Drive Drive
Battery PV
ac grid
= =
MPPT
= =
Bidirectional
dc grid
System ~ voltage
monitoring regulator
=
and control
380 ± 20 V
+
–
17.5.1.3 Control
Heuristic controls try to minimize the energy exchange with the main grid using
storage elements while maximizing the use of distributed green energy resources in
the system [45]. The grid connected bidirectional inverter performs the energy
balance and voltage regulation of the dc bus. The role of this inverter can be
diversified, for example, to draw or store energy in the flywheel or operating it as a
high power rating charger/discharger for dc UPS [45].
Demonstration sites of dc microgrids 427
17.5.1.4 Protection
It was reported that a special dc plug was developed by Fujitsu to eliminate arcing.
Issues with series arcing are more relevant in residential dc nanogrids due to a
relatively high operating voltage where sustained arcs can be expected [5]. Con-
sidering the wear and tear issues as well as added costs of such components, a novel
selective series arc extinction scheme is developed in [47] with experimental proof of
concept. Challenges such as localized short circuit protection and standardization of
safety codes pertaining to different system components are important considerations.
17.5.2.2 Components
The front-end ac/dc conversion step in individual devices can be removed, com-
prising filter, rectifier, power factor correction and link capacitor. A central rectifier
428 DC distribution systems and microgrids
Single PV-panels
250 W micro MPP dc-lighting
PV-strings tracker (heliox)
String 1
String 2
dc Distribution cabinet dc+ M
24 V nanogrid
dc+ M dc+ M
dc 15 kW solar dc dc
dc MPPT (ENP) dc dc
Solar dc dc
inverters ac ac Central dc bus
L N L N 15 kW 10.5 kW 10 kW
dc dc 3 kW dc dc
central electronic micro-CHP unit bidirectional EV
rectifier ac load ac dccharger dc
(MTT)
(ENP) (heliox)
L1 dc+ M dc+ M
L2
L3
ac mains N
with common dc bus shared among dc-ready devices can offer size, cost and energy
savings [48]. This design concept can further benefit from emerging fast switching
power electronic devices based on materials such as silicon carbide (SiC) and
gallium nitride (GaN). The microgrid components illustrated in Figure 17.10 were
developed by partners Emerson Network Power, Philips, Heliox, MTT and/or were
retrofitted for 380 V, dc operation, from commercially available ac devices [48].
17.5.2.3 Control
Central energy management system (EMS) can optimize the operation of dc office
building. The central rectifier, for example in [48], can regulate the output dc bus
voltage. Therein, a star point programmable logic controller is employed to monitor
the system parameters and govern the power flow in each subsystem. Information
such as weather forecast and user behavior can be utilize by the EMS to fulfill
higher level control objectives such as building CO2 reduction. Specifically in [50],
it was shown that about 10% reduction in CO2 emissions is achieved just by con-
verting the system from ac to dc, while above 16% are achieved if dc EMS is used.
Heuristic control was used to minimize grid energy exchange, maximize the
available green energy resources and optimize the use of energy storage.
17.5.2.4 Protection
In [48], dc switch technology from ABB is employed to interrupt fault currents.
Fast responding power electronic based solutions may also be considered.
TN-S grounding scheme is used in the test bed, consistent with the ETSI EN 301
Demonstration sites of dc microgrids 429
605 standard for grounding systems in 400 V dc datacenters [51]. Special attention
must be paid to arc detection, for instance, based on the methods presented in
[47,52].
17.5.2.5 Economics
The study at Xiamen University showed that the installation cost of 150 kW dc
microgrid to power a building is about $2.2/W [53], having a payback of 9 years
and 5.5 years with and without incentives, respectively.
17.5.3.1 Architecture
Three nanogrids, each with 2–5 kW capacity and 51.2 V internal dc bus, were
interconnected by 380 V external dc bus as shown in Figure 17.11 [54]. The open
energy system (OES) was a multilayered, recursively scalable, bottom-up approach
as shown in Figure 17.12 and was realized for a cluster of 19 dc nanogrids [55].
17.5.3.2 Components
A full scale prototype of the idea was realized at the Okinawa Institute of Science
and Technology, Japan as shown in Figure 17.13.
ac grid
Windmill
PV PV
Communication
Breaker
dc power bus
Control Battery
Battery
dc/dc Battery
Battery
Local power
trade OES
Center for
open energy system research
Figure 17.13 Full scale demo prototype of the OES concept (courtesy: [54])
17.5.3.3 Control
The peer-to-peer (P2P) multiagent control for OES dc microgrids is described in
[56]. Highlighting that the standard hierarchical control in IEEE 1547 was limited
to large generation, top-down approach, it was argued that a grid independent
fully decentralized design was necessary for resilient, scalable and self-sufficient
operation. It was anticipated that the role of Information Communication
Demonstration sites of dc microgrids 431
Technologies for active, multilayered and collaborative P2P control in OES would
be significant.
Ensuring that the failure of any single entity would not result in complete loss
of network services, the 19 autonomous dc nanogrids at Okinawa, Japan were
operated as community-wide, interconnected P2P controlled microgrid [56]. It was
shown that the self-sufficiency ratio, representing minimization of external non-
renewable energy, needs improved by more than 4% when decentralized P2P dc
exchange was allowed as compared to a stand-alone nanogrid. The solar operation
ratio, signifying reduction in maximizing solar generation usage and minimizing
curtailment, improved by almost 10% with dc interconnections as compared to
stand-alone mode. A centralized control approach provided a further improvement
of 4% and 2%, respectively. While it was claimed that an improvement in resilience
could not be shown experimentally, its conceptual discussion was presented [56].
17.5.4 Discussion
A selection of demonstration projects of residential and commercial buildings,
transitioning toward campus dc and universal dc grids is presented in Table 17.4.
Demonstration sites for dc microgrids in residential, commercial buildings and
campus applications
The dc operating voltage was usually selected at 380 V with capacity as high
as 150 kW. Most small-scale projects used radial topology with centralized control.
A scalable, multilayered approach with multiagent control was suggested in [54].
There is potential for a universal dc distribution with meshed, multilayered and
bipolar topology employing distributed control. One important issue discussed in
many initiatives was the protection aspect. Specifically, the localization and
extinction of arcing was considered vital. Impact and reliability of dc switch based
protection is another important domain. Perhaps, a clear demonstration of the
interplay between reliability and availability, supported by a selective and recon-
figurable protection system, will strengthen the acceptability of a meshed universal
dc grid.
References
[1] Jhunjhunwala A, Lolla A, Kaur P. Solar-dc microgrid for Indian homes:
a transforming power scenario. IEEE Electrification Magazine. 2016
Jun;4(2):10–19.
[2] Kaur P, Jain S, Jhunjhunwala A. Solar-DC deployment experience in off-grid
and near off-grid homes: Economics, technology and policy analysis. In: IEEE
First International Conference on DC Microgrids; 2015. p. 26–31.
[3] Chakrabarty S, Islam T. Financial viability and eco-efficiency of the solar
home systems (SHS) in Bangladesh. Energy. 2011; 36(8): 4821–4827.
{PRES} 2010.
[4] Mondal MAH. Economic viability of solar home systems: case study of
Bangladesh. Renewable Energy. 2010;35(6):1125–1129.
[5] Liu Z, Shekhar A, Ramı́rez-Elizondo L, et al. Characterization of series arcs
in LVdc microgrids. In: IEEE Second International Conference on DC
Microgrids (ICDCM); 2017.
[6] IRENA. Solar PV in Africa: Costs and Markets. International Renewable
Energy Agency (IRENA). Available at: https://www.irena.org/-/media/
Files/IRENA/Agency/Publication/2016/IRENA_Solar_PV_Costs_Africa_
2016.pdf
[7] Shenai K, Jhunjhunwala A, Kaur P. Electrifying India: using solar dc
microgrids. IEEE Power Electronics Magazine. 2016 Dec;3(4):42–48.
[8] Paleta R, Pina A, Santos Silva CA. Polygeneration energy container:
designing and testing energy services for remote developing communities.
IEEE Transactions on Sustainable Energy. 2014 Oct;5(4):1348–1355.
[9] Sulligoi G, Tessarolo A, Benucci V, et al. Shipboard power generation:
design and development of a medium-voltage dc generation system. IEEE
Industry Applications Magazine. 2013 Jul;19(4):47–55.
[10] Jin Z, Sulligoi G, Cuzner R, et al. Design and control of hybrid power and
propulsion systems for smart ships: a review of developments. Applied
Energy. 2017;194:30–54.
[11] Shekhar A, Ramı́rez-Elizondo L, Bauer P. Next-generation shipboard DC
power system: introduction smart grid and dc microgrid technologies into
maritime electrical networks. IEEE Electrification Magazine. 2016 Jun;
4(2):45–57.
[12] IEEE Recommended Practice for 1 kV to 35 kV Medium-Voltage DC Power
Systems on Ships. IEEE Std 1709-2010. 2010 Nov; p. 1–54. doi: 10.1109/
IEEESTD.2010.5623440.
[13] Shekhar A, Ramı́rez-Elizondo L, Bauer P, et al. DC microgrid islands
on ships. In: IEEE Second International Conference on DC Microgrids
(ICDCM); 2017.
[14] Herbst JD, Gattozzi AL, Ouroua A, et al. Flexible test bed for MVDC and
HFAC electric ship power system architectures for Navy ships. In: 2011
IEEE Electric Ship Technologies Symposium; 2011. p. 66–71.
Demonstration sites of dc microgrids 433
[29] Goh EJ, Chu KN, Ng NK. 1500 V DC traction system for the North East
Line. In: 2004 International Conference on Power System Technology,
2004. PowerCon 2004. vol. 2; 2004. p. 1904–1909.
[30] Kumagai K, Fujita T, Nakahira M, et al. Study on train operation energy
between commuter train and traction substations in a Japanese urban Railway.
In: IEEE Electrical Power and Energy Conference (EPEC); 2015. p. 50–55.
[31] Black, Veatch. Analysis of Smart Grid Technologies. Hawai’i Natural
Energy Institute; 2013.
[32] Roose LR. Hawaii Smart Technology Demonstration for a 100% RE Future
[Presentation]. www.nedo.go.jp: Smart Community Summit 2016; Jun 2016
[cited July, 2017].
[33] Capasso C, Veneri O. Experimental study of a DC charging station for full
electric and plug in hybrid vehicles. Applied Energy. 2015;152:131 – 142.
[34] Powerwall 2 and Solar Roof Launch [Video]. www.tesla.com; Oct 2016
[cited July, 2017].
[35] Tesla and SolarCity [Web News]. www.tesla.com; Nov 2016 [cited July,
2017].
[36] Exploring Future Power Management [Technical News]. www.delta.tudelft.nl;
Apr 2016 [cited July, 2017].
[37] AlLee G, Tschudi W. Edison Redux: 380 Vdc brings reliability and effi-
ciency to sustainable data centers. IEEE Power and Energy Magazine. 2012
Nov;10(6):50–59.
[38] Ton M, Fortenbery B, Tschudi W. DC Power for Improved Data Center
Efficiency. Berkeley, CA: Lawrence Berkeley National Laboratory; 2008
[39] IEEE Recommended Practice for Powering and Grounding Electronic
Equipment - Redline. IEEE Std 1100-2005 (Revision of IEEE Std 1100-
1999) – Redline. 2006 May; p. 1–703.
[40] Aldridge T, Pratt A, Kumar P, et al. Evaluating 400V direct-current for data
centers: a case study comparing 400 Vdc with 480–208 Vac power dis-
tribution for energy efficiency and other benefits. Intel Labs.
[41] Pratt A, Kumar P, Aldridge TV. Evaluation of 400V DC distribution in telco
and data centers to improve energy efficiency. In: INTELEC 07 – 29th
International Telecommunications Energy Conference; 2007. p. 32–39.
[42] Yajima H, Babasaki T, Usui K, et al. Energy-saving and efficient use of
renewable energy by introducing an 380 VDC power-supply system in data
centers. In: IEEE International Telecommunications Energy Conference;
2015. p. 1–4.
[43] Kaga M, Noritake M, Hirose K, et al. Verification of container data center
using 380 V dc power distribution system. In: 2012 International Conference
on Renewable Energy Research and Applications (ICRERA); 2012. p. 1–5.
[44] Patterson BT. DC, come home: DC microgrids and the birth of the ‘‘Enernet’’.
IEEE Power and Energy Magazine. 2012 Nov;10(6):60–69.
[45] Wu TF, Chen YK, Yu GR, et al. Design and development of dc-distributed
system with grid connection for residential applications. In: 8th International
Conference on Power Electronics – ECCE Asia; 2011. p. 235–241.
Demonstration sites of dc microgrids 435
[46] Mackay L, van der Blij NH, Ramirez-Elizondo L, et al. Toward the universal
DC distribution system. Electric Power Components and Systems. 2017;45
(10):1032–1042.
[47] Shekhar A, Ramirez-Elizondo L, Bandyopadhyay S, et al. Detection of
series arcs using load side voltage drop for protection of low voltage DC
systems. IEEE Transactions on Smart Grid. doi: 10.1109/TSG.2017.2707438.
[48] Wunder B, Ott L, Szpek M, et al. Energy efficient DC-grids for commercial
buildings. In: 2014 IEEE 36th International Telecommunications Energy
Conference (INTELEC); 2014. p. 1–8.
[49] Environmental Engineering (EE); Power supply interface at the input to
telecommunications and datacom (ICT) equipment; Part 3: Operated by
rectified current source, alternating current source or direct current source up
to 400 V; [ETSI EN 300 132-3-1]. European Standards; 2012.
[50] Noritake M, Yuasa K, Takeda T, et al. The demonstration of the CO2 reduc-
tion effect of 400V class DC micro grid for offices. In: 2015 IEEE Interna-
tional Telecommunications Energy Conference (INTELEC); 2015. p. 1–6.
[51] Environmental Engineering (EE); Earthing and bonding of 400 VDC data and
telecom (ICT) equipment [ETSI EN 301 605]. European Standards; 2013.
[52] Strobl C. Arc fault detection in DC microgrids. In: 2015 IEEE First Inter-
national Conference on DC Microgrids (ICDCM); 2015. p. 181–186.
[53] Zhang F, Meng C, Yang Y, et al. Advantages and challenges of DC micro-
grid for commercial building a case study from Xiamen university DC
microgrid. In: IEEE First International Conference on DC Microgrids
(ICDCM); 2015. p. 355–358.
[54] Werth A, Kitamura N, Tanaka K. Conceptual study for open energy systems:
distributed energy network using interconnected DC nanogrids. IEEE
Transactions on Smart Grid. 2015 Jul;6(4):1621–1630.
[55] Werth A, Tokoro M, Tanaka K. Bottom-up and recursive interconnection for
multi-layer DC microgrids. In: 2016 IEEE 16th International Conference on
Environment and Electrical Engineering (EEEIC); 2016. p. 1–6.
[56] Werth A, André A, Kawamoto A, et al. Peer-to-peer control system for DC
microgrids. IEEE Transactions on Smart Grid. 2018 Jul;9(4):3667–3675.
[57] Hirose K, Reilly JT, Irie H. The Sendai microgrid operational experience in
the aftermath of the Tohoku earthquake: a case study. New Energy and
Industrial Technology Development Organization (NEDO); 2013.
[58] Fregosi D, Ravula S, Brhlik D, et al. A comparative study of DC and AC
microgrids in commercial buildings across different climates and operating
profiles. In: IEEE First International Conference on DC Microgrids; 2015.
p. 159–164.
[59] Noritake M, Yuasa K, Takeda T, et al. The demonstration of the CO2
reduction effect of 400V class DC micro grid for offices. In: 2015 IEEE
International Telecommunications Energy Conference (INTELEC); 2015.
p. 1–6.
This page intentionally left blank
Index
AC–DC converter control 258, 299 energy storage system (ESS) 274
balancing leg control 261 power distribution 269–73
direct power control (DPC) 260–1 power generation 268–9
single-phase control 258–9 power quality requirements in
symmetrical component aircrafts 274–6
control-based methods 261–2 power utilization 273
voltage-oriented control 259–60 stability analysis 285–8
AC generators 297–8 starter/generator control 276–8
AC microgrid system 1, 20, 368–70 Aircraft Electrical Generation with
AC power systems 1, 153, 301 Active Rectification
AC residential buildings 378–9 Technology (AEGART)
battery bank power in 384 project 276, 278
conceptual layout of 377 aircraft power systems 84–5
conversion losses in 385 all-electric ship (AES) 295
conversion power losses in internal architecture of dc microgrids 390
converters loads in 385 LVDC microgrid
and DC residential buildings, for data centers 394–5
comparison between 377, 382–6 for homes 395–6
active-bridge converter 236–8 for space applications 392–3
active current-sharing techniques 169 MVDC and LVDC microgrid for
active front end (AFE) rectifier dc distribution systems 390–1
276, 280 MVDC microgrid
active power filters in HV battery for marine applications 393–4
chargers 328–30 for oil-drilling applications 396
AC vs DC microgrid system, autonomous load control 164, 175
mathematical analysis of 373 architecture 165–6
total daily load (TDL) 373–4 control levels 164–5
voltage, current and power losses in strategy for controlling the load to
DC supply 374–7 be an energy asset 166–8
adjustable speed drives (ASD) 353 auto-transformer rectifier unit
Aggregators 199 (ATRU) 270
aircraft DC microgrids 267 auxiliary loads 323–4
control strategies in aircraft DC auxiliary power module (APM) 323
microgrids 279 auxiliary power unit (APU) 269
primary control 280–2 average current sharing (ACS)
secondary control 283–5 technique 15–17
438 DC distribution systems and microgrids