High Frequency Isolated Bidirectional Dual Active Bridge DC-DC Converters and Its Application To Distributed Energy Systems: An Overview

You are on page 1of 23

International Journal of Power Electronics and Drive Systems (IJPEDS)

Vol. 14, No. 2, June 2023, pp. 969~991


ISSN: 2088-8694, DOI: 10.11591/ijpeds.v14.i2.pp969-991  969

High frequency isolated bidirectional dual active bridge


DC-DC converters and its application to distributed energy
systems: an overview

Kiran Bathala, Dharavath Kishan, Nagendrappa Harischandrappa


Department of Electrical and Electronics Engineering, National Institute of Technology Karnataka, Surathkal, India

Article Info ABSTRACT


Article history: Among the DC-DC converters, an isolated bidirectional dual active bridge
converter is a core circuit for high-frequency power converters in distributed
Received Nov 1, 2022 energy system applications. These high-frequency power conversion systems
Revised Jan 19, 2023 attract academia and industry due to various advantages, such as high-power
Accepted Feb 1, 2023 density, less weight, reduced noise, high efficiency, low cost and high
reliability. First, the importance of power electronic converters in modern-
day life is introduced. Second, a topological overview of voltage-fed and
Keywords: current-fed isolated bidirectional dual active bridge converters is presented
with their importance in integrating hybrid power sources. Third, switching
Dual active bridge modes of isolated bidirectional DC-DC converters are also presented with a
Energy storage system degree of freedom of control. Forth, performance evaluation of voltage-fed
Isolated converters and current-fed isolated bidirectional converters has been presented with an
Phase shift control example. Their suitability in integrating fuel cells and photovoltaics with
Voltage fed-current fed energy storage systems in low to medium-power applications is presented.
This is an open access article under the CC BY-SA license.

Corresponding Author:
Kiran Bathala
Department of Electrical and Electronics Engineering, National Institute of Technology Karnataka
Surathkal, India
Email: [email protected]

1. INTRODUCTION
Renewable energy resources are intermittent, so it is necessary to provide auxiliary energy storage
systems for these resources to increase their reliability [1], [2]. In the energy sector, distributed energy
systems (DES) will occupy a significant role in the future. Power from solar, windmills, microturbines and
fuel cells (FC) will play a crucial role in DES. These non-conventional energy sources have no dispatch
capability on their own. Adding an energy storage system as a backup source in the distributed generation
system increases the reliability of the overall distributed energy resources. In the coming years, using a hybrid
electric vehicle (HEV) along with the utility grid in the form of plug-in hybrid electric vehicles is guaranteed
by introducing an energy storage system with long-term or short-term energy buffering capability [3]–[5].
DES requires specific PE topologies to directly convert the power to interconnect to the utility grid or
consumer applications.
Power electronics can occupy a considerable part of the total capital cost of a typical distributed
energy system, as shown in Figure 1 [6], [7]. The investment in PE and other capital costs occupy the same
percentage, i.e., 20% and 80% of the total cost in PV and Fuel Cell applications, respectively. In wind power
applications, PE and other capital investments occupy 30% and 70% of the total cost, respectively. In the
micro-turbine application, the investment cost on PE and other capital is 40% and 60%, respectively. The
input fuel cost is zero for PV and wind power applications among the four energy sources. The input fuel and
PE costs are the lowest for PV power generation. Whereas for the fuel cell application, the additional cost for

Journal homepage: http://ijpeds.iaescore.com


970  ISSN: 2088-8694

the input fuel is high compared to the PV. The PV and FC systems produce DC power, whereas, for utility
purposes, the generated power has to be converted to 1-ɸ or 3-ɸ AC power. The output power from the
windmills and micro-turbine systems is AC, and the frequency is not constant, which has to be converted to
50Hz AC for utility connection.
The PE conversion systems mainly employ power frequency transformers for achieving galvanic
isolation and impedance matching. However, the line or power frequency transformers reduce power
conversion systems' efficiency and power density [8]. The rapid growth of distributed energy technologies and
battery storage systems has raised the potential of power conversion systems as an ever-lasting key interface
[9]–[15]. High-frequency transformers in place of regular low-frequency transformers have become a new
trend in recent years. These HF transformers are taking power conversion to the next level with advantages
such as being compact, less weight and cost. HF power conversion systems with HF transformers can also
avert current and voltage waveform distortions.
Power conversion systems noise can be significantly reduced if the switching frequency is
maintained more significantly than 20 kHz. The distinct feature of power electronics converters for energy
storage systems (ESS) is that they must allow power flow in both directions, for both receiving powers from
the grid for charging and delivering power to the grid while discharging. DC-DC power converters play a
significant role in power distribution when the power is generated from renewable energy resources. In
searching for HF bidirectional power flow converters, isolated bidirectional DC-DC converters (IBDCs)
usually act as a core (mains) circuit. The demand that has been increasing for intermediate energy storage
systems in battery units has raised the demand for the bi-directional DC-DC converter, which gives galvanic
isolation apart from bidirectional power flow. Isolated bidirectional dual active bridge converters are the most
suitable to transfer a large amount of power. HF isolation transformer can prevent the direct current flow path
between the bridge circuits through the isolation transformer.

Figure 1. Investment on PE costs out of total capital costs in a DE systems [16]

Dual active bridge (DAB) isolated bidirectional DC-DC converters (IBDCs) have the highest power
transmission capacity because the power transmitted is directly proportional to the number of switches in the
converter. These converters can achieve soft switching easily, and power can be transmitted in both directions
[17]. The circulating current is a major disadvantage and decreases the converter's efficiency [18], [19]. The
decrease in efficiency increases with an increase in the difference between the switching frequency and
resonant frequency. A bidirectional DC-DC converter, succeeded by an inverter, is the most general choice
for storage devices. The bidirectional converters are fed from either a voltage or current source, depending on
the input available and output requirement. In voltage-fed IBDCs, non-resonant IBDCs [20], [21] resonant
IBDCs [22], [23] and similarly in current-fed, non-resonant [24], [25], resonant network [26]–[28] both this
type of converters are discussed respectively with their merits and demerits. The resonant network helps
voltage-fed, current-fed IBDC converters improve performance. Soft switching is explained with waveforms
and factors influencing soft switching, and its positive and negative impacts are presented. Various types of
resonant converters are presented; in terms of duty cycle, soft switching, range, extra components required
and their bidirectional transition speed. Depending on the input, it needs to be pre-regulated before supplying
to achieve the desired output; in such cases, a two-stage converter is needed. In two-stage converters, the
importance of interleaved converters is also presented [29], [30]. IDBC converters with fewer switches are

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  971

also discussed along with applications [31], [32]. The cost of various IBDC converters is compared along with
the applications.
The focus of this paper has been on various voltage and current-fed isolated BDCs topologies for low
to medium-power applications and their suitability in integrating renewable energy sources with battery
energy storage systems. The suitability of IBDC converters for low voltage input, such as fuel cells and
photovoltaic applications, has been discussed in the following sections. For fuel cell applications, for
example, two converters, i.e., voltage-fed and current-fed of exact specifications, are compared and
investigated for the low voltage input applications. The impact of the degree of freedom of control (DFC) is
presented by giving the power transmission and circulating power of various switching modes. DFC's
influence on losses and efficiency is also presented [33], [34]. An overview of energy storage systems is also
discussed [35]–[37]. This paper is scripted as follows: section 2 explains the broad classification of isolated
bidirectional converters. Switching modes are explained in section 3, followed by a discussion on the
importance of power electronics in distributed energy systems (DES) and the cost of various IBDC converters
in section 4, and the conclusion is presented in section 5.

2. CLASSIFICATION OF ISOLATED BIDIRECTIONAL DUAL ACTIVE BRIDGE DC-DC


CONVERTER TOPOLOGIES
The general block diagram with various stages of IBDC converters is shown in Figure 2, its
classification details based on the type of source are shown in Figure 3 and are explained in subsections 2.1
and 2.2 respectively.

2.1. Block diagram representation of configuration


The block diagram representation of the isolated DAB bidirectional DC-DC converter is shown in
Figure 2. Port-1 and port-2 can act as a source or sink depending on the direction of power flow. During the
forward power flow, a DC voltage source is connected to port 1. The voltage output of this source is
smoothened with the help of filters (L/C). The selection of filters is also made to convert the source as a
voltage type of current type. In the case of a current source, this constant DC voltage or current is converted
into AC by using a high frequency (HF) inverter. HF resonant elements are used to aid the bridge in
achieving soft switching. The output of the DC-AC converter is applying to the primary of the high-
frequency transformer via resonant network circuit elements. HF transformer is used to provide galvanic
isolation and to scale the voltage levels up or down. The transformer's output is given to the AC-DC
converter through the HF secondary side resonant network. This resonant network helps the AC-DC
converter to achieve soft switching. The AC-DC converter converts HF AC into DC voltage. The secondary
side filter network comprising C/L can act as a voltage/current source load at port 2. A capacitor is connected
across the output terminals of the controlled (AC-DC converter) rectifier. The capacitor gets charged during
the conduction period and offers low impedance to the ripple components. This capacitor discharges the
stored energy during the non-conduction period; it maintains a constant voltage across the output terminals of
the rectifier. So, this is named a voltage source type load.

HF
+
+ HF HF
Filter Resonant Trans- Resonant Filter Port-2
Port-1 DC-AC AC-DC
Network Network former Network Network
Converter Converter
- n1:n2 -

Figure 2. Block diagram representation of the isolated DAB bidirectional DC-DC converter

An inductive filter is connected in series between the AC-DC converter and the load at the filter
network shown in Figure 2. Since the frequency and its inductive reactance are directly proportional, DC
components offer zero impedance, and the rest are blocked. This filter stores the energy during the
conduction period, and it does not allow sudden changes in the current. During the non-conduction period,
the stored energy discharges through the body diode and maintains a constant current. So, it is called the
current source type load. During the reverse/ discharge operation, when stored energy is fed back, or a DC
supply is given at port-2, the same sequence of operation happens from port-2. In the HF AC part on either
side of the HF transformer, reactive HF networks enable energy storage capability, which helps change the
shapes of the switch current waveforms to reduce switching losses. The HF resonant networks are not
necessarily required for a full bridge isolated bidirectional DC-DC converter because of inherent parasitic

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
972  ISSN: 2088-8694

components such as leakage, magnetizing inductance and a parasitic capacitance of practical HF transformer.
These resonant networks can aid the converter in achieving higher efficiency [38], [39].

2.2. Classification
The DAB DC-DC converters, in general, are classified as one-stage and two-stage BDCs, as shown in
Figure 3. Furthermore, the isolated converters are classified into non-resonant and resonant IBDCs. These
converters may be fed from a voltage or current source depending on the applications and kind of input supply
available. These converters are explained in this in terms of their contribution. IBDCs converters are further
classified as one-stage and two-stage converters based on the number of power conversion stages. The single-
stage converter is further classified as a non-resonant and load-resonant converter. These resonant converters
are classified as voltage source type and current source type converters based on the type of source used. The
magnetically coupled with and without DC link converters are further classified as voltage-source and current-
source converters. The magnetically coupled interleaved converter has only the current source as input. The
dual active bridge (DAB) converters initially started with non-resonant converters, and non-resonant DAB
(NRDAB) converters with a voltage source can enable the converter to achieve high power density, reduce the
switching loss through soft switching [40], [41].

Voltage Source
Non-Resonant
Converters I-V Source
One-Stage
Converters Voltage Source
Resonant
IBDC Converters
Converters I-V Source

Two-Stage Voltage DC Link


Converters Current DC Link

Figure 3. Classification of dual active bridge (DAB) DC-DC converters

2.3. Voltage-fed non-resonant IBDC converters


A typical structure of an isolated bidirectional dual active bridge DC-DC converter (IBDC) is shown
in Figure 4. The NRDAB converter can be effectively utilized because of the symmetrical structure, and a
simple control technique can be implemented. Even though the current and voltage rating of each switch is the
same, the power transmission capacity of the IBDC converter is directly proportional to the number of
switches. Eight switch IBDC converter power transmission capacity is double that of four switches (half
bridge) IBDC converter [42]. Soft switching, zero voltage switching (ZVS) and zero current switching (ZCS)
are shown in Figures 5(a) and 5(b) respectively, for the NRDAB IBDC ZVS is achieved by the junction
capacitance and the magnetizing inductance of the HF transformer. NRDAB fed voltage source converter has
circulating current when it is controlled by using the single-phase shift control (SPS) technique. The circulating
current in NRDAB due to SPS increases with input voltage. It further leads to an increase in the peak and RMS
current of the switching devices, which results in an increase in the conduction losses; hence, the efficiency of
the converter reduces. To overcome the drawbacks of the SPS technique, the dual-phase shift technique (DPS)
is proposed for a voltage-fed non-resonant DAB (VNRDAB) converter [42].
Conventional isolated DAB converter uses eight switches, and it needs a separate gate drive circuit. By
reducing the number of switch counts, the cost of a switch and its gate driver can be saved. In the literature, a
DAB converter with a push-pull topology on the primary side of the HF transformer is presented. To achieve
zero voltage switching (ZVS) for the push-pull topology, an additional switch is connected on the primary side,
which is fed from low voltage, and it is shown in Figure 6. On the primary side, only three switches are used
when compared to conventional DAB, and this will reduce the cost of driving the circuit. Whereas, on the
secondary side full bridge is used. To achieve ZVS for the converter, extended phase shift (EPS) control is used
[43]. In the DAB converter with the SPS control scheme, two-level square wave voltage appears across the
isolation transformer, which causes the circulating current. This circulating current can be reduced by increasing
the levels of the voltage across the isolation transformer. A hybrid full-bridge isolated buck-boost converter
offers three-level voltages at the primary side of the isolation transformer, and it has five variable voltage levels.
The increase in voltage levels reduces the circulating current. This leads to wider voltage control in the
integrated buck-boost converters. This converter offers low voltage stress, which is half of the input voltage on
transformer primary side switches and offers less peak and current stress of the switching devices.

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  973

In the place of switches Q1 to Q8 are Various HF networks that can be


either MOSFETs/IGBTs made up of used depends on the application
Ge/Si/GaAs/SiC/GaN material which are given in Fig. 9
I1 I2
I1 < 0 and I2 > 0
Forward Mode
Q1 D1 Q3 D3 Q5 D5 Q7 D7

High frequency resonant/


V1 c
a non-resonant network V2
C1 Vab along with isolation Vcd C2
transformer with n1:n2
b d
turns ratio

Q2 Q4 D4 Q6 Q8 D8
D2 D6

Reverse Mode

H Bridge-1 I1 > 0 and I2 < 0 H Bridge-2

Control Strategy (FFPWM, VFPWM,


SPS,EPS,DPS and TPS)

Figure 4. Typical structure of isolated bidirectional dual active bridge DC-DC converter (IBDC)

(a) (b)

Figure 5. Typical waveforms depicting the feature of soft switching of (a) zero voltage switching and
(b) zero current switching

Figure 6. Improved push–pull topology [43]

A DAB converter with a hybrid bridge on the primary side and a half-bridge on the secondary side of
the converter is shown in Figure 7 [44], controlled with a voltage match control technique. It is controlled
with the DAB converter's extended phase shift control (EPS). This control technique is used when the overall
voltage gain is double compared to its minimum input voltage to output voltage gain. This control technique
offers ZVS for the six main switching devices and for the two auxiliary switches [44]. In order to
accommodate many sources, the converter discussed in [45] is an appropriate choice. In an integrated buck-
boost converter, to avoid the circulating current, the converter operates in boundary current mode or
discontinues conduction mode [46], [47].
The DAB converters, also found in traction applications, have disadvantages such as switching noise,
voltage ripple at the DC input side, quantization noise in analogue to digital converters and noise in the
measuring circuits. The mechanism behind generating these noises and their impact on the DAB converter is
High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
974  ISSN: 2088-8694

analyzed. To reduce the influence of these noises, the sampling frequency of the DAB converter must be
different from the control frequency. To increase the performance and the dynamic behavior of power
electronic traction transformer, the control frequency must be kept much higher when the sampling frequency
of the DAB converter is higher than the switching frequency [48]. The IBDC converter can also be fed from
multiple input sources for energy storage applications. This converter can operate with a single independent or
combinational source, as shown in Figure 8. In multiple-input IBDC configuration, the multiple energy
storage systems are connected for transferring the power in charging/discharging directions. This multiple-
input DAB converter offers enhanced efficiency under light load conditions compared to the conventional
DAB converter. This converter reduces the circulating power flow and current stress on switches, and the
power transfer range can be extended easily [49]. Dual-phase shift (DPS) control scheme is used for the
VNRDAB converter to improve efficiency, and the switching characteristics of VNRDAB converters are
presented. The power loss model to estimate the energy loss for each power device is developed. Analysis to
have optimized efficiency characteristics across the selected range is also presented [50].

Figure 7. DAB with an auxiliary half bridge on Figure 8. Multiple-input configuration of


primary side [44] isolated bidirectional DC–DC converter [45]

Single phase VNRDAB has a lesser number of switches than three-phase VNRDAB. The current
rating of switching devices in three-phase VNRDAB converter is lower. The currents in each phase of the
three-phase VNRDAB converter are low compared to the single phase VNRDAB, which makes this converter
more efficient. The DAB converter is provided with average current control and load current feed forward
control technique to protect the circuit from overcurrent which is obvious because of voltage mode control
technique [42].
Three-phase isolated bidirectional DAB converter with voltage source is presented [51]–[54]. In
literature, three phase IBDC converter is compared with the single-phase voltage fed IBDC converter.
Compared to single phase topology, three-phase VNRDAB converter offers higher power density. On the
other side, control circuit of single phase VNRDAB is simpler compared to the three- phase topology. Three-
phase topology maintains constant power between the two ports; this feature will reduce the filter size. The
same three-phase converter can be operated as single-phase topology, since it can provide redundancy [55].
A modulation strategy is presented to reduce the conduction losses for the total load range of the
three-phase VNRDAB converter. It increases the range of soft switching of the converter even under a large
variation of voltage operations. Broad analysis of the duty cycle control is presented to improve the
performance of the converter. Fast transient current control method is presented for the three-phase VNRDAB
converter; it balances the transformer currents for any abrupt change in duty cycles and reduces the oscillation
in DC currents [56]. The disadvantages of the VNRDAB converters, HF transformer gets saturated due to the
DC component of an asymmetrical waveform of the transformer current, in order to avoid the disadvantage
additional components are required for these converters. In VNRDAB converters the effect of dead time on
voltage sag, polarity reversal and unexpected phase shift in the converter is presented [43].

2.4. Voltage fed resonant IBDC converters


In voltage source resonant isolated bidirectional DC-DC converters, apart from the magnetizing
inductance, a capacitor is connected in series with the transformer. In these resonant converters, along with
junction capacitance, a series LC resonant network is also present, as shown in Figure 9(a). Two methods are
proposed to analyze a dual active bridge series resonant converter. AC equivalent circuit analysis is carried out
based on the type of load. In the first method, a voltage-source type of load is taken.
In the second method, a controlled rectifier with resistive load is considered. For these two methods,
a detailed analysis is presented. Soft switching is achieved for both sides of the dual active bridge, and the

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  975

switches on the Primary bridge have ZVS, whereas the switches on the secondary side have zero current
switching’s (ZCS) [23]. Compared to the non-resonant IBDC, the resonant IBDC resonant converter has
additional resonant elements. These resonant elements are connected to the HF transformer, and the various
resonant configurations are shown in Figure 9. These resonant elements are used to achieve soft switching for
the primary and/or secondary side bridges of the HF transformer. Since the resonant elements change the
waveforms of the switch currents, this feature can be used to attain reduced switching losses [50]. IBDC
converters with series resonant networks [54], [60] with switchable resonance frequency and converters with
resonant filters are presented [61].
A DAB converter with a symmetric LC resonant network is presented, as shown in Figure 9(b). A
decoupled control scheme with a PI controller is designed to control both deep currents and resonant IBD
converters. These currents, in turn, control the magnitude of the power transfer and Phase angle between the
full bridges being controlled to reduce losses by improving the soft switching of the converter. Even though
the load increases due to the series capacitor, RMS currents remain low in switches and resonant tanks.
Because of the two voltage source ports, bidirectional power flow control has no restrictions. At no load
conditions, currents become exceedingly high. Output RMS currents become high for low-voltage
applications. It is not suitable for this application because of the voltage source output. It is unsuitable for
wide-range operation of voltage and power values [62].
Lee et al. [63] the advantages of LLC resonant converters are presented. It ensures ZVS from light
load to full load, switch turn off current and switching loss are low. This resonant network also offers high
voltage gain and reduces the necessity for bulk capacitors. High efficiency is possible at high input voltage
and ZCS for the secondary side switches. Demerits of the LLC resonant converter include a reduction in
power conversion efficiency because of the circulating power. A multi-element resonant converter can
overcome the demerits of the LLC converter [64]. In symmetric resonant LLC, a network is used, as shown in
Figure 9(b), for the LV DC distribution system. Inductors are placed on a single core, reducing the
converter's size. It offers ZVS for the primary switches and soft commutation for the secondary switches. The
voltage stress on the switches was reduced without any snubber components. Intelligent digital control
systems regulate the voltage at the output side and simultaneously control the power in both directions [57].
Figure 9(c) shows the type-11 LLC resonant tank, sharing features closer to the CL parallel-type resonant
converter. The load comes in parallel with the resonant network Cs-Lm-Ls during the resonant stage. The main
difference between Type-4 and type-11 is that the latter type offers higher voltage gain than the former at the
same switching frequency. A modified LLC resonant topology is presented with a control scheme, and all
switches achieve soft switching. This LLC topology, shown in Figure 9(c), offers better performance
compared to the series resonant converter, has reduced circulating current, turn-off losses and improved
conversion efficiency compared to the conventional IBDCs. The control scheme used for modified LLC
topology automatically changes the direction of power at the output without interruption. This feature makes
this converter more suitable for energy storage applications [39]. Two different LLC resonant networks are
presented, as shown in Figures 9(d) and 9(e). Figure 9(d) shows that it is closer to regular LC series resonant
converter because of the LsCs1 resonant stage. Type-4 LLC resonant network shown in Figure 9(e) is
extensively viewed as a band filter with the series-parallel type of resonant network. Figure 9(f) shows LCL
resonant network, an extra inductor (Ls) connected across the HF transformer's secondary side. This inductor
is used in the place magnetizing inductor, whereas leakage inductance of the transformer is observed as part
of the series resonant inductor (Lp) on the transformer's primary [59].
This LCL bidirectional DAB converter used a ZVT circuit to achieve soft switching for one of its
switching devices under light load conditions. This LCL DAB converter is controlled by using a modified
gating scheme. Table 1 shows different converters and their soft-switching solutions. Voltage source dual
active bridge series resonant converter with a variable frequency power control scheme is presented. An out-
of-phase relationship between tank current and bridge voltages causes the switching loss (due to hard
switching) in converters. This out-of-phase relationship also induces conduction loss due to increased
circulating current, which increases the root-mean-square value of the tank current. To achieve the soft
switching, minimum tank current operation is given to establish an in-phase relationship between resonant tank
current and secondary voltage of the transformer. A series resonant DAB converter is presented to achieve
minimum-tank-current operation along with full-range soft-switching. In the design process, parasitic
capacitance is considered due to the non-ideal behavior of switching devices. Dead time is estimated for
achieving complete soft switching [65]. DAB converter with various operation modes is given with their
respective boundary conditions. These conditions are distinguished by phase-shift angle and load conditions.
The expression for ripple at the output voltage was derived. The dead-band effect and safe operational area are
further investigated. The relations between output power, leakage inductance, and switching frequency are also
presented [66].

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
976  ISSN: 2088-8694

a Ls Cs c a Ls1 Ls2 c
Cs1 Cs2
Vab Vcd Vab Vcd
b d b d
1:n 1:n
(a) (b)
c a
a Lm1 Ls c
Cs1 Cs2
Vab Lm Vcd Vab Lm Vcd
b Ls Cs d
d b
1:n 1:n
(c) (d)

a Ls2 c a Lp c
Cs1 Cs2 Cp
Vab Vcd Vab Ls Vcd
Ls1
b d b d
1:n 1:n
(e) (f)

Figure 9. Various resonant networks of IBDCs (a) series resonant (SR) [23], (b) symmetric SR [57], (c) LLC
resonant tank [39], (d) type-4 CLLC [58], (e) type-11 CLLC resonant tank [58], and (f) LCL resonant tank [59]

Table 1. Different soft switching solutions [17]


S. Type of Resonant Switching Duty ratio Soft-switching SS Additional Bidirectional
No. converter Network Scheme (SS) range component transit-on
RN) speed
1 Conventional No Phase-shift 50% duty ratio ZVS for few Narrow No Fast
type modulation for all switches switches
2 LC-RN Series RN Phase shift 50% duty ratio ZVS for few Narrow A capacitor Fast
modulation for all switches switches, ZCS
for secondary
switches
3 CLLC- Series- Frequency 50% duty cycle ZVS for inverter Wide Two Slow
asymmetric parallel modulation for inverter switches, ZCS capacitors
RN RN switches, for rectifier and an
additional switches inductor
resonant signals
for rectifier
switches
4 CLLC- Series RN Frequency 50% duty ratio ZVS for inverter Wide Two Slow
symmetric modulation for inverter switches, soft capacitors
RN switches, turn- commutation for and an
off for rectifier rectifier switches inductor
switches

In the literature, the forward and reverse modes of the converters are fed from the voltage source. But,
depending upon the application, IBDCs can be fed from the current source/voltage source. This type of single-
phase current source converter offers a high step-up ratio because of its integral boost operation [20], [67]
which is presented in the following subsection.

2.5. Current fed non-resonant IBDC converters


A Current source converter can have high power transfer capacity and requires low gate drive
requirement. The highly efficient current-fed IBDC converter is presented for the FC EV driving system, and
it is shown in Figure 10. This topology offers a smaller number of switches. Hence, the cost of the gate drive
requirement is reduced to half compared to a conventional DAB converter. Coupled inductors reduce the
converter's size, increasing the converter's power density [24]. A snubber less line commutated current fed
isolated bidirectional dual active bridge (CIBDC) converter is presented for fuel cell applications. The
modulation technique used in this CIBDC offers a natural commutation, ZCS, as shown in Figure 5 for the
primary and ZVS for the secondary switches. Because of the natural commutation and soft switching, losses
are reduced significantly. These features make this topology suitable for fuel cell-based EV applications. The
modulation approach suggested is simple and easy to implement.

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  977

HV DC Bus

L1
Cs
Q1 DX
D1 Q3 D3

Battery
Traction Traction
+ Inverter Motor
VBus
Lr _
1:n
DY
Fuel Cell Q2 D2
Q4
Stack D4

L2

Figure 10. A bidirectional DC–DC converter for fuel cell electric vehicle driving system [24]

The circuit is modular and easily interleaved to extend the power range for high-power applications
[68]. Using the merits of topology presented, interleaved topology with two cells for ESS as voltage doubler
in FC electric vehicles applications is shown in Figure 11. This topology offers fewer switching losses and
allows HF switching with compact size and higher power density. The voltage and current-fed IBDCs are full
bridge converters [69]–[71], it [71] has two cells with four bridges and two HF isolation transformers, one cell
can transfer 50% of power in the case of failure of another cell. Voltage-fed current double converters [72],
[73] current source push-pull converters [58], IBDC converters using an asymmetrical PWM modulation
scheme can be seen with more than one switch [74]–[76]. These converters can be applied for high-power
applications. A current source IBDC converter is presented with inductors on the DC low voltage side. These
inductors are placed on a single core to reduce the core losses and size and increase the power density [77],
[78]. A modulation scheme controls the switches to achieve soft switching throughout the load range. This
work also presents the ZVS for the LV side bridge when the duty cycle is less than 50% [77].
In the literature, a 1kW current source isolated bidirectional converter is presented. In this work, dual
PWM plus double phase shift control technique is used to reduce circulating current. So, the RMS leakage
current, peak current and conduction losses are reduced. To achieve soft switching, respective conditions for
the switching devices are also derived. CIBDC converter is presented for low-voltage battery applications. A
modulation scheme is used to control the converter to overcome the low-efficiency performance, especially at
light loads [79]. This control scheme also improves efficiency for a wide load range on a variable input
voltage. The proposed control scheme offers the ZVS and flexibility in the degree of freedom of control,
making these more suitable for battery storage applications. A comparative analysis is carried out among the
conventional and proposed control schemes, and the benefits of the control scheme with respect to conduction
loss of the switches and core loss of the inductor are presented. To reduce the magnetizing inductance of an of
an isolation transformer for a CIBDC converter, a mathematical design approach is presented. This design
shows the least possible magnetizing inductance, which is sufficient to maintain the circulating current to
achieve ZVS operation [28].
The voltage source and current source type under non-resonant converters, the voltage source type
achieve more effective converter utilization. The advantage of current-fed converters is less peak switch
current because of the reduced circulating current. Hence, the transformer KVA rating and turn ratio can be
minimized. In VNRDAB converters, the power transfer capability is limited by increasing the switching
frequency and leakage inductance. The current source can enable the parallelization of converters easily, and
the current ripple becomes low. Other advantages of current source converters are high step-up ratio and
multiport interface capability, and wide input voltage range operations. One of the disadvantages of CIBDC
converter is the unequal distribution of the current stress within the common legs due to the unsymmetrical
structure of the converter [73], [80].
A three-phase current-fed isolated bidirectional DC-DC converter (CIBDC) is compared with a
voltage-fed non-resonant dual active bridge (VNRDAB), earlier offers low RMS current, zero voltage
switching (ZVS) and high efficiency for the total operating range of voltage. The input ripple current is
maintained low by using the interleaved topology. This configuration uses Y-Y connected transformers, which
distribute the current uniformly among the three phases compared to any other transformer connections. To
overcome the disadvantages of the non-resonant IBDC converters, resonant IBDC converters are proposed [81].

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
978  ISSN: 2088-8694

L1 La

Full Bridge Half Bridge


V1 Converter Converter RL
n:1

L2 Lb
Full Bridge Half Bridge
Converter Converter
n:1

Figure 11. Interleaved current-fed IBDC with FB on primary and HB on secondary [71]

2.6. Current fed resonant IBDC converters


In resonant current source type converters, the parallel resonant converter is suitable for the low value
of output voltages and high value of output currents. When the load decreases, the current through the resonant
tank and switches does not reduce. Therefore, it is less suitable for wide-range voltage and power applications.
The efficiency of the converter reduces because of the hard-switching operation of the switches. A modulation
scheme for the series load-resonant current fed isolated bidirectional converter (CIBDC) is presented, which
achieves soft switching and reduces the circulating power. This modulation scheme produces a voltage drop
across the capacitor, which is connected in series with an isolation transformer to achieve soft switching even
when the input or output voltage doubles comes into the scenario whenever the input or output voltage
becomes [82]. A current-fed dual active bridge converter with a series resonant converter for low voltage
applications, i.e., PV and FC, are presented, and it has partial soft switching. For energy storage applications,
LLC resonant current-fed DAB converter is presented with complete soft switching, and inductive elements
are mounted on a single core. This makes the whole converter compact because of utilizing a single core [73].
A current fed with PWM controlled bidirectional series resonant converter is presented for energy
storage applications, and it is shown in Figure 12. Voltage regulation for a wide range of voltage is achieved
with the PWM control technique. Soft switching (SS) for all the switches is achieved for a wide range. The
design of inductive elements is made easy by choosing the switching frequency nearer to the resonant
frequency. Fixed frequency-controlled voltage and current source dual active bridge converter achieved wide
range soft-switching while maintaining higher efficiency compared to voltage-fed PWM converters [78]. The
value of snubber capacitance across the switch influences the value of minimum current flowing through the
leakage inductance of the transformer for achieving SS when the switch turns off. The value of snubber
capacitance and minimum current through the leakage inductor required to achieve soft switching are directly
proportional. When the winding resistance of the transformer is negligible, if the magnetizing inductance
(Lmag) decreases, then it results in low transformer utilization and loads current will be lagging, which is a
precondition for zero voltage switching (ZVS) [83]. The influence of Lmag on the SS region of operation is
shown in Figure 13.

3.0 K=1
Q3 K=
Q1 Q7
Q5 D7 2.5
D1 D3 D5
Input-Bridge
Ls 2.0 Boundary
L1
(pu)

a c 1.5
CC
Vo'

La
CC Vp Vs VH
L2 b d
Soft switching
1.0 region
1:n
Q2 Q4 Q8 0.5
D2 D4 Q6 D6 D8 Output-Bridge
VL Boundary
0.2 0.4 0.6 0.8
Io' (pu)

Figure 12. Current fed series Figure 13: Impact of transformer Lmag on
resonant IBDC [78] the SS region of operation

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  979

2.7. Two stage converters


In two-stage converters, the supply voltage is pre-regulated or transformed with the help of switched
capacitor at the first stage. In the second stage, fine regulation of output voltage takes place with a high-
frequency magnetic element; the block diagram of the two-stage converter is shown in Figure 14. They are the
voltage DC-link on the low voltage side and the current DC-link on the high voltage side used to further
differentiate the converter topologies. Two circuit configurations, parallel and series-parallel, are investigated
with advantages and disadvantages. An intermediate DC voltage link is used to the energy storage system
interface, which must have the functions of a bi-directional (two quadrant) converter connecting two voltage
type sources with significantly different voltage values [84].

Filter network
Fine
Pre-regulated regulation of
Supply (or) output voltage Load
Transformed stage
stage with a HF
magnetic stage

Figure 14. Block diagram representation of two stage converter shown on the output voltage-output current
plane [85]

The available designs of hybrid vehicles include a BDC between the voltage bus and the energy
storage system. A bidirectional DC-DC converter is used between the DC voltage bus and the Energy storage
system (ESS). DC voltage is used to run the motor in hybrid vehicles, and a current-fed full bridge IBDC with
an integrating buck converter is given. This topology is useful where the capacitors are used to hold the high
voltage [86]. A three-phase interleaved converter is presented to integrate a low-voltage fuel cell, ultra-
capacitor and load, as shown in Figure 15. This converter is a solution to convert energy by integrating a fuel
cell (FC) and an ultra-capacitor (UC). Since it is a bidirectional power converter, it is possible to consume and
supply ripple-free current from FC and UC. Because of the interleaved feature of the components, the current
rating at input increases without parallel connection. Input filter inductors are coupled with a single core,
reducing the core number and current ripple and increasing the converter's efficiency [80].

Q3a Q3c Q3e


Q1c Q1e Q2a Q2c Q2e
Q1a D1e D3a D3c D3e
D2a D2c D2e
D1a D1c

L2a L2a L2a


Ls1

L2b L2b L2b


Ls2 C1 C2

Ls3 L2c L2c L2c


Load
Q1d Q1f Q2f
Q1b Q2b Q2d Q3b Q3d Q3f
D1f D2b D2d D3b
D1b D1d V2 D2f D3d D3f
V1

Figure 15. Interleaved three-port three phase converter for high power applications [80]

2.8. Isolated bidirectional DC-DC converters with low switch count


Isolated converters with a lower number of switches are shown in Figure 16. The converter shown in
Figure 16(a) is a hard-switched fly-back converter. This operates in both forward and reverses modes of
operation in continuous conduction mode. This converter is observed from the conventional fly-back converter
[31]. The fly-back converter is equipped with an auxiliary circuit and is operated with soft switching. It has
three switches, as shown in Figure 16(b) and is operated in zero current switching to reduce the losses, which
ultimately increases the converter's efficiency [31]. This converter has input, output voltage and power of
24 V, 12 V and 50 W, respectively. This converter achieves 83% efficiency; it is 10% less than the hard-
switched converter shown in Figure 16(a). The integrated Cuk converter can be seen in Figure 16(c). This
magnetic integration gives various advantages; by optimizing the switching frequency. Other advantages of
magnetic integration are more power density, less cost, low noise, capacitive filter requirements at both ends
High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
980  ISSN: 2088-8694

can be minimized, and operating frequency can also be minimized [87]. This converter achieves an efficiency
of 86% in forward mode and 76.74% in the reverse mode of operation.
A Zeta converter with galvanic isolation and two switches is shown in Figure 16(d). This converter
overcomes the disadvantage of traditional Zeta converter; high voltage stress on the switches caused by the
transformer leakage inductance and switching device body capacitance. This converter has an input, output
voltage and power rating of 60V, 10V and 30W, respectively [88]. This converter achieves a peak efficiency of
88%. An IBD current-fed converter is presented in Figure 16(e). This converter consists of a flyback
transformer with coupled inductors LFBp and LFBs. A push-pull transformer with coupled magnetics and LFBs.
A push-pull transformer with coupled magnetics LPPp1, LPPs1 and LPPp2, LPPs2. Four bidirectional current-
controlled, unidirectional voltage switching devices are present in the converter. This converter is fed from
80V input and has 160V output with a power rating of 800W. This current-fed converter offers less input
current ripple on either side of the converter. It offers high efficiency with fewer passive components [89].
A push-pull converter derived from the dual active bridge is shown in Figure 16(f). This converter
uses coupled inductors instead of separate inductors. It also consists of a forward circuit with a half-bridge. It
requires less gate drive circuitry because it uses a smaller number of switches. The input, output voltage and
power ratings are 40V, 200V and 500W, respectively. All switches are operated with soft switching, i.e., zero
voltage switching [32]. This converter has a maximum efficiency of 92.5%. Isolated bidirectional DC-DC
converters with low switch counts have a narrow voltage gain range. These converters are applicable for low
and medium-power applications.

(a) (b)

(c) (d)

(e) (f)

Figure 16. IBDCs with low switch count: (a) hard-switched bidirectional fly-back converter [31], (b) A ZCS
bidirectional fly-back converter [31], (c) integrated bidirectional cuk converter [87], (d) isolated two-transistor
zeta PWM dc-dc converter [80], (e) current fed fly-back push-pull DC-DC converter [89], and (f) push-pull
forward half bridge bidirectional DC-DC converter [32]

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  981

3. SWITCHING MODES OF IBDCS


This section, it is explained the results of the research and, at the same time, is given a comprehensive
discussion. Results can be presented in figures, graphs, tables and others that make the reader understand
easily. The discussion can be made in several sub-sections. Switching modes of IBDCs plays an important role
in controlling the power flow in both directions (i.e., forward/charging or backward/discharging). The degree
of freedom of control (DFC) decides the efficiency of the control scheme [90]. As DFC increases, efficient
control and conversion of power are possible. The few control schemes for IBDCs are described in subsections
3.1, 3.2, 3.3 and 3.4. The switching pulses, typical voltage, and current waveforms for the schemes are shown
in Figure 17.

3.1. Single phase shift control (SPS)


The single-phase shift control (SPS) control technique is the simple and most preferred technique for
isolated bidirectional dual active bridge converters. In this technique, diagonal switch pairs in two bridges are
turned on to produce square waveforms having a 50% duty ratio for the respective switches. The square
waveforms on either side of the isolation transformer have a phase difference; the leading voltage side bridge
delivers the power to the lagging voltage bridge side of the transformer. Only a phase-shift ratio (or angle) ’d’
is chosen as freedom of control by adjusting ‘d’ power between the bridges is controlled. Typical waveforms
of the DAB converter with SPS control are shown in Figure 17(a). For voltage source non-resonant converters
(VNRDAB), this SPS technique has the disadvantages such as circulating current and reactive power. These
problems cause high peak current and high-power loss [91]. The other disadvantages are that the range of ZVS
is limited, and current stress is more when the transformation ratio of the transformer is not near unity. This
control technique is less suitable for a wide range of voltage variation applications, and it gives a maximum
efficiency of 88 to 92% for various IBDCs. The expression for the power transmission and circulating power
under the SPS control scheme is given in (1) and (2).
𝑉1 𝑉2
Ptr = 𝑑(1 − 𝑑) (1)
2𝑛𝐿𝑓𝑠

𝑉1 𝑉2 [𝑚+(2𝑑−1)]2
Circulating Power (Pcp) = (2)
16𝑓𝑠 𝑛(𝑚+1)

Where ‘m’ is the ratio of V1 and nV2, n is the transformation ratio of the isolation transformer, and L is the
leakage inductance of the transformer. The maximum circulating power (Pcp) flow occurs at m =1, which is
approximately 25% of maximum power transmission (d = 0.5) CP flow will be more than 25% when m > 1 &
(𝑛𝑉2 )2
d =0.5; (Pbase = (3𝑚 − 1)). P(pu) = d2, concludes that the Pcp exists in both input and output voltage ends
8𝑓𝑠 𝐿
for the entire power transmission range [92].

3.2. Extended phase shift control (EPS)


The EPS control technique is the improved version of the SPS technique, shown in Figure 17(b).
This technique turns on the diagonal switch pair in one bridge, and another diagonal switch pair is turned on
with an inner phase shift ratio. The output voltage at the DC-AC converter is a three-level square wave voltage,
and the voltage at the input of the AC-DC converter is the two-level voltage with a 50% pulse width. This
technique offers zero-level voltage on the primary side and three-level voltage, as shown in Figure 17(b); this
reduces the circulating power by making the reverse flow of current zero. By comparing the EPS technique
with the SPS technique, only a phase-shift ratio (or angle) ‘d’ is present in SPS. In the EPS scheme, the outer
phase-shift ratio ‘d1’ and inner phase-shift ratio ‘d2’ are chosen as freedom of control. Outer phase shift ratio
‘d1’ controls the direction and magnitude of power.
Inner phase-shift ratio d2 is used to control the circulating power flow and increase the range of ZVS
[91]. This control technique in a non-resonant voltage source converter efficiently regulates the power level.
This scheme increases the efficiency and dynamic performance under a wide range of input voltage
variations and even under the disturbance in a load [93]. The power transmission and circulating power (Pcp)
under this scheme are expressed in (3) and (4).
𝑉1 𝑉2 𝑑1
Ptr = [𝑑2 (1 − 𝑑2 ) + (1 − 𝑑1 − 2𝑑2 ) (3)
2𝑛𝐿𝑓𝑠 2

𝑉1 𝑉2 [𝑚(1−𝑑1 )+(2𝑑2 −1)]2 (1−2𝑑2 )


Pcp = ;m> (4)
16𝑓𝑠 𝑛(𝑚+1) (1−𝑑1 )

(1−2𝑑2 )
when m ≤ , Pcp is zero.
(1−𝑑1 )

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
982  ISSN: 2088-8694

The Pcp in SPS and EPS is the same when d1 = 0, due to additional phase shift ‘d1’ EPS reduces the circulating
power by keeping the same power transmission as the SPS technique. The EPS control scheme can transfer
more power than SPS when 0 ≤ d2 < 0.5. This can achieve an efficiency of 92.5% [92].

3.3. Dual phase shift control (DPS)


The DPS control technique is an improved version of the EPS scheme; the operational waveforms of
the DPS scheme are shown in Figure 17(c). Compared to EPS, the diagonal switch pairs are switched with the
inner phase shift ratio, and the inner phase shift ratios are equal. This switching pattern results in three-level
AC waveforms on either side of the isolation transformer [13]. The implementation of the DPS technique for a
voltage source non-resonant converter is presented [93]. Three schemes of control algorithms, such as single-
phase shift scheme, dual phase-shift scheme and model-based phase shift control scheme, are implemented on
a full-bridge non-resonant voltage source converter. The experimental results of these control schemes are
compared; MPSC offers better dynamic performance, and the DPS scheme can eliminate the reactive power
under light load conditions. DPS offers various advantages compared to the SPS: reduced current stress and
steady-state current, an increase in efficiency and range of ZVS.
Compared to the SPS control scheme, this control scheme can reduce the voltage ripple in voltage-
sourced DAB [91]. For the non-resonant voltage source converter, both SPS and DPS schemes are
implemented and proposed a mathematical model to analyze current stress for the IBDC converter. This
analysis showed a reduction in current stress, improved converter efficiency and increased power transfer
capability. This control scheme makes the converter suitable for high voltage conversion ratio and light load
conditions [94], [95]. During specific operating periods, dead band compensation can be carried out without
difficulty in the DPS scheme without a current sensor. Hence, the DPS scheme is easier to carry out and
improves dynamic performance. The governing equation of power transmission and circulating or backflow
power are given in (5) to (13). This scheme can achieve maximum efficiency of 89 to 96.5% for IBDCs.

𝑉1 𝑉2 𝑑12
Ptr = [𝑑2 (1 − 𝑑2 ) − ] ; (0≤d1≤d2≤1) (5)
2𝑛𝐿𝑓𝑠 2

𝑛𝑉1 𝑉2 (3𝑚−1)
Pcp = [𝑚(1 − 𝑑1 ) + (2𝑑2 − 𝑑1 − 1)]2 at input; 0 ≤ d1 ≤ d2 ≤ 1 (6)
32𝑓𝑠 𝐿(𝑚+1)

𝑛𝑉1 𝑉2 (3𝑚−1) 1
Pcp = [𝑚(1 + 𝑑1 − 2𝑑2 ) + (𝑑1 − 1)]2 at output; d1≤1-2d2( ) (7)
32𝑓𝑠 𝐿(𝑚+1) 1+𝑚

1
Pcp = 0; d1 > 1-2d2 ( ) (8)
1+𝑚

𝑉1 𝑉2 𝑑2
Ptr = (1 − 𝑑1 − )𝑑2 ; (0≤d2< d1≤1) (9)
2𝑛𝐿𝑓𝑠 2

1
𝑛𝑉1 𝑉2 (3𝑚−1) 4𝑚𝑑22 + [𝑚(1 − 𝑑1 ) + 𝑑1 − 2𝑑2 − 1] 1
Pcp= [ 𝑘−1 ];d1≤1+2d2( )at input (10)
32𝑓𝑠 𝐿 1−𝑚
[𝑚(1 − 𝑑1 − 2𝑑2 ) + (𝑑1 − 1)]

𝑛𝑉1 𝑉2 (3𝑚−1) 1
Pcp = 𝑚𝐷22 ; d1 > 1+2d2 ( ) (11)
32𝑓𝑠 𝐿 1−𝑚

1
PCP = 0; d1 > 1+2d2 ( ) (12)
1−𝑚

1
𝑛𝑉1 𝑉2 (3𝑚−1) 4𝑚𝑑22 + [𝑚(1 − 𝑑1 ) + 𝑑1 − 2𝑑2 − 1] 1
PCP = [ 𝑚−1 ] d1 ≤ 1+2d2 ( ) at output (13)
32𝑓𝑠 𝐿(𝑚−1) 1−𝑚
[𝑚(1 − 𝑑1 − 2𝑑2 ) + (𝑑1 − 1)]

3.4. Triple phase shift control (TPS)


This control scheme is a better version of DPS control; the TPS scheme has three degrees of freedom
of control. The diagonal switch pairs are turned on for this control scheme with an inner phase shift ratio in
two full bridges. Here the inner phase shift ratios are unequal, and typical operational waveforms are shown in
Figure 17(d). Most of the research on the TPS scheme focused on utilizing the converter in an optimum
operation [96], [97]. This technique mainly focuses on the optimized zone of operation. Power transferred from
source to load [98], [99] is given by (14) to (17).

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  983
𝑛𝑉1 𝑉2
Ptr = [(1 − 𝑑1 − 𝑑3 ) + 𝑇(1 − 𝑑1 − 𝑑2 − 𝑑3 )] (14)
𝑛𝑓𝑟

Circulating power (Pcp)

𝑛𝑉1 𝑉2 𝑚(2𝑑3 +𝑑1 −1)+1


Pcp = [(1 − 2𝑑3 − 𝑑2 ) − (1 − 𝑑1 )]2 ;d3 ≤ min [m (1-𝑑1 )+2𝑑3 − 1, ](15)
16𝑓𝑠 𝐿(𝑚+1) (1+2𝑚)

𝑛𝑉1 𝑉2
Pcp = {4𝑚(𝑑 − 𝑑2 )[(1 − 𝑑2 ) + 𝑚(𝑑3 + 𝑑1 − 𝑑2 − 1)] − [(1 − 𝑑2 ) + 𝑚(2𝑑3 − 2𝑑2 + 𝑑1 −
3
16𝑓𝑠 𝐿𝑚
2 𝑚(2𝑑3 +𝑑1 −1)
1)] }; < d2 ≤ [1+m (2d3+ d1-1)] (16)
(1+2𝑚)

𝑉1 (𝑑2 +𝑑3 −2𝑑1 −1)


Pcp = 0; m = ≤− (17)
𝑛𝑉2 (1−𝑑1 )

Where d1, d2 and d3 are the phase shifts represented in Figure 17(d). Practical application point of view, the
SPS, EPS, DPS and TPS control scheme offers one, two, two and three control degrees, respectively. So, the
TPS scheme is a complex scheme to carry out. For the EPS scheme, the operating states of both full bridges
must be modified when the voltage conversion states or power flow directions are modified. SPS scheme has
low efficiency at light loads, low range of soft switching, and circulating current. Hence, the DPS scheme can
be chosen as an optimal scheme for big-scale practical implementation from the difficulty level and performance.
However, TPS can achieve an efficiency range of 96 to 97.8% for the IBDC converters [34], [49].

(a) (b)

(c) (d)

Figure 17. Typical operating waveforms of different control schemes: (a) SPS control, (b) EPS control, (c)
DPS control and (d) TPS control [17]

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
984  ISSN: 2088-8694

4. DISCUSSION
In non-resonant IBDC DC-DC voltage-fed converters, non-inductive power will appear when SPS
controls it. Even though the SPS technique offers less complexity in implementation, it is inefficient in
reducing reactive power. In place of SPS, DPS can reduce the circulating power. DPS have two degrees of
control (‘d1’ and ‘d2’), whereas it is one (‘d’) for the SPS technique and is shown in Figures 17(a) and 17(c).
The degree of freedom control can enable switches to operate in soft switching (ZVS/ZCS). The factor
affecting the ZVS can be seen in Table 2. ZVS and ZCS operate the converter in lagging and leading power
factor operation, respectively, as shown in Figure 5. Voltage-fed IBDC converters can maintain square wave
voltage across the isolation transformer without voltage spikes. The degree of freedom of control (‘d1’, ‘d2’,
and ‘d3’) is three for the TPS technique, which is complex for practical implementation, and it is shown in
Figure 17(d). This technique for voltage-fed IBDC converter gives less circulating power compared to the
former three techniques (SPS, EPS and DPS) [100], [101].
In voltage-fed IBDC converters, efficiency increases with the degree of freedom of control; TPS
switching technique offers high efficiency. The current carrying capacity in voltage-fed IBDC converters
increases with an increase in switch count, which is not valid with current-fed IBDC converters. However, in
current-fed IBDC converters, due to current source and inherent boosting ability, these are best suitable for low
voltage and high current applications [102].

Table 2. Factors affecting ZVS and their positive and negative effects [95]
S. No. Variable Positive effect Negative effect
1 Magnetizing current Support ZVS More current stress and conduction loss
2 Leakage inductance Support ZVS, by reducing Reduces the maximum effective duty ratio; hence poor VA
the rate of reversal of the utilization and more conduction loss. Results in higher ringing
primary current and dissipation in the secondary rectifiers
3 TDelay Large TDelay aids ZVS at light Large TDelay reduces the effective duty ratio and is particularly
loads and affects adversely at undesirable at very high switching frequencies
high loads.
4 Capacitance across the Large CDS aids in lossless Large CDS demands more energy to be stored in the transformer
MOSFET-CDS turn-off inductances, to be fully discharged, hence bad for ZVS

In resonant voltage-fed isolated bidirectional dual active bridge (IBDC) converters, soft switching is
achieved easily because of the resonance. The range of soft switching is narrow in the case of non-resonant
compared to the resonant IBDC converters. The magnetizing current in non-resonant converters results in more
current stress and losses in conducting devices. In resonant IBDCs, a resonant tank offers nearly sinusoidal
current, which results in lower current stress and losses of the conducting devices. Among the resonant IBDCs,
series LC resonant have a narrow soft-switching range compared to symmetric and asymmetric CLLC resonant
converters (LC, Symmetric CLLC and asymmetric CLLC are shown in Figures 9(a)-(c). In CLLC-type IBDCs,
asymmetric offers ZVS for inverter switches and ZCS for rectifier switches. Symmetric resonant tank gives
ZVS for inverter switches and soft commutation for rectifier switches.

4.1. Cost comparison of IBDCs


A few factors decide the quality of any power electronic converter; cost is one, along with the
converter's efficiency, size and power density. The cost of various IBDC converter topologies is given in
Table 3 (see Appendix), with state-of-the-art and advanced components. These advanced components (GaN
switches) offer ten times smaller turn-on resistance and turn-on and turn-off times; these features lead to very
less losses and allow higher energy density. Since these are designed at higher frequencies, i.e., 100kHz or
higher, the size of the magnetic components becomes smaller.

4.2. Example: comparison of VFC and CFC


The IBDC converters shown in Figures 18(a) and 18(b) are fed from a voltage source and current,
respectively [20]. These two converters, which are rated at 1kW of power and 288V of output voltage taken as an
example. Converters are investigated to check their suitability for the low voltage fuel cell and PV applications.
From Figure 19(a) and Table 4, suitable components are chosen for the converter. Since it's for LV high current
application, switches are selected with low Ron for the primary & secondary switches. The per-unit values of
Figure 19(a) are shown in Figure 19 (b). The complete losses and efficiency of the two converters are shown in
Figure 19(c). It is evident from Figure 19(a) that the current-fed converter (CFC) has less peak value of current
through the circuit devices, and it gives less circulating current [42], [103]. This results in less conduction loss.
Switch voltages of high frequency CFC and voltage-fed converter (VFC) are seen as the same from Figure 19(a)
[55], which are lower values. In the case of VFC, the voltage supplied to the battery is clamped by the capacitor.

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  985
Ls

Q1 Q3 Q1 Q3
D1 D3
Q5 D5
Q7 D7
D1 D3
Q5 D5
Q7 D7

Ls Ls
a c a c
RL CO RL
V1 Ci Vp Vs CO V1 Ci Vp Vs
b d b d
n:1 n:1
Q2 Q4 Q2 Q4
D4 Q6 D6 Q8 D8 D2 D4 Q6 D6 Q8 D8
D2

(a) (b)

Figure 18. IBDCs (a) voltage fed dual active bridge IBDC converter and (b) current fed dual active bridge
IBDC converter

(a) (b)

(c)

Figure 19. Comparison of VFC and CFC (a) comparison of various parameters of VSC and CSC, (b) per unit
values of figure, and (c) losses and efficiency comparison of two converters

Current-fed converter (CFC) sustains higher efficiency for the same battery application, even after
battery voltage fluctuates and load current decreases. It is evident from Figure 19(a) KVA rating and turns ratio
required for the CFC is small compared to the voltage-fed converter (VFC); this makes the volume of the CFC
converter very small compared to the VFC. Power transfer in HF Isolated power electronics systems depends
on leakage inductance (Llk) and switching frequency (fs). If the turns ratio is higher, it will be challenging to
design a transformer with less leakage inductance in VSC. An increase in Llk and fs can reduce the power

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
986  ISSN: 2088-8694

transfer capacity. At a given value of fs and rated power, CFC can have a higher value of Llk than VFC. For
example, P = 5 kW VFC needs Llk = 34 nH, whereas CFC needs 103 nH. Designing Llk = 34 nH inductance
with a higher turn’s ratio is very difficult. Therefore, for a power rating less than 5 kW, CFC can supply one
cell while VFC needs two or more cells with interleaving. This makes the volume of VFC higher compared to
CFC of the same power rating. Because of the DC current received from the battery, CFC has better battery
utilization compared to VFC. CFC can offer; high-power single cell, better battery utilization, lower peak
currents, lower isolation transformer kVA ratings, less circulating currents, and improved load efficiency,
which makes CFC a better choice for FC and PV applications.
In CFC, a secondary modulation scheme clamps the primary side switch voltage at a less reflected
voltage. This feature enables us to use semiconductor switches with low Ron and voltage ratings. Table 4
shows that primary Rds(on) is low for VFC compared to CFC; its conduction loss is still more, as shown in
Figure 19(c). It is due to the primary RMS current, which is higher than CFC, as shown in Figure 19(a).
Theoretical and simulation studies gave [20] semiconductor device RMS and maximum current rise due to the
increase in supply voltage, which in turn increases the conduction losses in primary side devices and lowers the
converter's efficiency. For battery application, the efficiency of VFC falls due to an increase in the maximum
and RMS current of the switch even after the battery is charged beyond the nominal value.

Table 4. Selected components of VFC and CFC [97]


Converter Voltage fed converter Current fed converter
IRFB3004PBF IRFB3006PBF
(Primary side) (Primary side)
V1 = 40 V, Id = 340 A, V1 = 60 V, Id = 195 A,
Rdson = 1.43 mΩ Rdson = 2.1 mΩ
HF Switches
IPA60R385CP IPA60R385CP
(Secondary side) (Secondary side)
V1 = 650 V, Id = 9 A, V1 = 650 V, Id = 9 A,
Rdson = 0.385 Ω Rdson = 0.385 Ω

5. CONCLUSION
Power converters have become the most critical and inevitable components of the power electronic
interface used in renewable energy generation. Isolated bidirectional dual active bridge DC-DC (IBDC) power
converter topologies have been discussed, and their performance has been compared to recommend one for
low input voltage distributed energy applications. Classification of voltage-fed and current-fed resonant and
non-resonant IBDC converters has been made and discussed. It is found that current-fed converters offer
inherent voltage-boosting capability, which makes it suitable for low-voltage PV and Fuel cell applications.
Current-fed resonant converters maintain soft switching for a wide range of input voltage and load variations.
A 12 V DC input, 288 V DC output and 1000 W of output power-rated voltage-fed and current-fed IBDC
converters have been investigated to determine their suitability in low-voltage renewable energy applications.
Based on performance parameters such as efficiency, current stress and loss of duty cycle, it is concluded that
the current-fed converter topology outperforms the voltage-fed one. Power control methods that include single
phase-shift, extended phase-shift, dual phase-shift and triple phase-shift controls have been discussed. Single
phase-shift control is easier to control and implement while triple phase-shift control is efficient but complex
to implement in IBDC converters in renewable energy applications.

APPENDIX

Table 3. Cost comparison of various topologies


No. of switched Passive components Gate driver (ICs Vin/Vout/Po/ Estimated Cost in $ Estimated cost in $ %η
(IRF640NPBF) capacitors(C), IR2110, HCPL2730 Application (with state- of-the-art (with Advanced
inductors (L) and capacitors) components) switches (GaN)
GAN063650WSA)
8 [6] C: 2 * 2200 μF / 3*ICs IR2110 3*ICs 350 V DC / 1.5 53.77 163.53 97
350 V DC L: 10.5 IR2110 3*10 μF, kV / 10 kW /
μH Electrolytic Electrical traction
7 [45] C: 2 * 880 μF, 970 3*ICs IR2110 3*ICs 60 V DC / 300 V 53.1 149.14 95
μF / 300 V DC L: IR2110 3*10 μF, / 500 W / Energy
187 μH Electrolytic storage

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  987

Table 3. Cost comparison of various topologies (continue)


No. of switched Passive components Gate driver (ICs Vin/Vout/Po/ Estimated Cost in $ Estimated cost in $ %η
(IRF640NPBF) capacitors(C), IR2110, HCPL2730 Application (with state- of-the- (with Advanced
inductors (L) and capacitors) art components) switches (GaN)
GAN063650WSA)
8 [44] C: 2 * 20μF, 3*ICs IR2110 3*ICs 120 - 240 V DC / 55.07 164.83 95
2 * 400 μF / 240 V IR2110 3*10 μF, 96 V /1 kW /
DC L: 20 μH Electrolytic Energy storage
6 [104] C: 2*10 μF L: 2*15 3*ICs IR2110 3*ICs 12 V DC / 150 - 51.78 134.1 96
μH + core IR2110 3*10 μF, 400 V DC / 2
Electrolytic kW/ FC-EV
8 [23] C: 32.01 nF / 75 V 3*ICs IR2110 3*ICs 110 - 130 V DC / 53.12 162.88 95
AC L: 95.5 μF IR2110 3*10 μF, 75 -100 V DC /
Electrolytic 200 DC / Energy
storage
8 [80] C: 1410 nF, 33 nF / 3*ICs IR2110 3*ICs 200 - 400 V / 48 54.42 164.18 96
630 V L: 228 μH, IR2110 3*10 μF, V DC / 500
76 μH Electrolytic W/Battery
8 [105] C: 200 nF / 120 V 3*ICs IR2110 3*ICs 380 V DC / 220 53.77 163.53 97.8
AC L: 130 μH, 30 IR21103*10 μF, V DC / 5 kW /
μH Electrolytic Energy storage
8 [2] C: 47 nF / 240 V 3*ICs IR2110 3*ICs 160 - 240 V DC / 55.72 164.48 96
AC L: 53.89 μH, IR2110 3*10 μF, 400 V DC / 1.6
2*400 μH, Electrolytic kW / ESS
2*300 μH
8 [69] C: 2200 μF / 50 V 3*ICs IR2110 3*ICs 40 - 56.4 V DC / 55.07 164.83 94
DC L: 2*164 μH, IR2110 3*10 μF, 150 -300 / 10 kW
2*45 μH Electrolytic / Hybrid ESS
9 [106] C: 22 μF / 100 V 3*ICs IR2110 3*ICs 8 V DC / 288 V / 57.04 180.52 95
DC, 3.9 μF / 630 V IR2110 3*10 μF, 1.6 kW / FC
DC L: 2*40 μH, 86 Electrolytic
μH, 36 μH
6 [107] C: 56 μF, 180 μH / 3*ICs IR2110 3*ICs 30 - 60 V DC / 53.73 136.05 96.69
63 V DC, 2*6.8 μF IR2110 3*10 μF, 400 V/ 1 kW /
/ 400 V DC L: Electrolytic Energy storage
10.54 μH, 0.74 μH,
18.56 μH, 18.41 μH
8 [106] C: 5.6 μF / 100 V 3*ICs IR2110 3*ICs 600 V DC/ 18 - 58.32 168.08 96.42
DC, 2*3.9 μF / 250 IR2110 3*10 μF, 28 V DC / 1 kW/
V DC, 10 μF / 250 Electrolytic Battery
V DC L: 1.5 μH, 11
μH, 500 μH, 2*10
μH, 60μH
8 [73] C : 0.18 μF / 900 V 3*ICs IR2110 3*ICs 160 - 240 V DC / 57.02 166.78 96.49
AC, 12 μF / 100 V IR2110 3*10μF, 400 V/ 1.6 kW /
DC, 180 μF / 63 V Electrolytic Battery
DC, 1.5 μF / 630 V
DC L : 150 μH,
3*300 μH
6 [79] C: 3.9 μF / 630 V 3*ICs IR2110 3*ICs 22 - 4 V DC / 350 51.13 133.45 97.48
DC L: 9.6 μH, IR2110 3*10μF, V 200W / FC
8.6 μH Electrolytic
8 [44] C: 4.7 μF / 630 V 3*ICs IR2110 3*ICs 12 V DC / 288 V 53.77 163.53 97.07
DC L: 1.5 μH, IR2110 3*10μF, / 250 W / FC
16.4 μH Electrolytic Vehicles

REFERENCES
[1] M. Liserre, T. Sauter, and J. Hung, “Future energy systems: integrating renewable energy sources into the smart power grid
through industrial electronics,” IEEE Industrial Electronics Magazine, 4(1), 18–37. doi:10.1109/mie.2010.935861, vol. 38, no.
12, pp. 1721–1729, 2016.
[2] T. Bi, B. Yang, K. Jia, L. Zheng, Q. Liu, and Q. Yang, “Review on renewable energy source fault characteristics analysis,” CSEE
Journal of Power and Energy Systems, vol. 8, no. 4, pp. 963–972, 2022, doi: 10.17775/CSEEJPES.2021.06890.
[3] S. Chakraborty, B. Kroposki, and W. Kramer, “Advanced power electronic interfaces for distributed energy systems part 2 :
modeling, development, and experimental evaluation of advanced control functions for single-phase utility-connected inverter
advanced power electronic interfaces for distribute,” no. November, p. 62, 2008.
[4] H. Wu, K. Sun, Y. Li, and Y. Xing, “Fixed-frequency PWM-controlled bidirectional current-fed soft-switching series-resonant
converter for energy storage applications,” IEEE Transactions on Industrial Electronics, vol. 64, no. 8, pp. 6190–6201, 2017, doi:
10.1109/TIE.2017.2682020.
[5] A. A. E. B. A. El Halim, E. H. E. Bayoumi, W. El-Khattam, and A. M. Ibrahim, “Electric vehicles: a review of their components
and technologies,” International Journal of Power Electronics and Drive Systems, vol. 13, no. 4, pp. 2041–2061, 2022, doi:
10.11591/ijpeds.v13.i4.pp2041-2061.

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
988  ISSN: 2088-8694

[6] Z. Tang, Y. Yang, and F. Blaabjerg, “Power electronics: The enabling technology for renewable energy integration,” CSEE
Journal of Power and Energy Systems, vol. 8, no. 1, pp. 39–52, 2022, doi: 10.17775/CSEEJPES.2021.02850.
[7] C. Katar and C. P. Uzunoglu, “A comparative study on AC/DC analysis of an operational low voltage distribution system,”
International Journal of Applied Power Engineering (IJAPE), vol. 10, no. 3, p. 193, 2021, doi: 10.11591/ijape.v10.i3.pp193-206.
[8] S. Inoue and H. Akagi, “A bidirectional isolated DC–DC converter as a core circuit of the next-generation medium-voltage power
conversion system,” IEEE Transactions on Power Electronics, vol. 22, no. 2, pp. 535–542, Mar. 2007, doi: 10.1109/TPEL.2006.889939.
[9] M. Berrehil El Kattel, R. Mayer, M. Douglas Possamai, and S. Vidal Garcia Oliveira, “Bidirectional isolated three-phase dc-dc
converter using coupled inductor for dc microgrid applications,” International Journal of Circuit Theory and Applications, vol.
48, no. 6, pp. 832–859, 2020, doi: 10.1002/cta.2795.
[10] N. Safitri, A. M. Shiddiq Yunus, F. Fauzi, and N. Naziruddin, “Integrated arrangement of advanced power electronics through
hybrid smart grid system,” Telkomnika (Telecommunication Computing Electronics and Control), vol. 18, no. 6, pp. 3202–3209,
2020, doi: 10.12928/TELKOMNIKA.vl8i6.13433.
[11] R. Muzzammel, R. Arshad, S. Bashir, U. Mushtaq, F. Durrani, and S. Noshin, “Comparative analysis of optimal power flow in
renewable energy sources based microgrids,” International Journal of Electrical and Computer Engineering, vol. 13, no. 2, pp.
1241–1259, 2023, doi: 10.11591/ijece.v13i2.pp1241-1259.
[12] J. A. Adebisi, I. H. Denwigwe, and O. M. Babatunde, “Hydrogen storage for micro-grid application: a framework for ranking fuel
cell technologies based on technical parameters,” International Journal of Electrical and Computer Engineering, vol. 13, no. 2,
pp. 1221–1230, 2023, doi: 10.11591/ijece.v13i2.pp1221-1230.
[13] F. A. Syam, M. I. Abu El-Sebah, K. S. Sakkoury, and E. A. Sweelem, “Operation of biogas-solar-diesel hybrid renewable energy
system with minimum reserved energy,” Indonesian Journal of Electrical Engineering and Computer Science, vol. 28, no. 3, pp.
1203–1213, 2022, doi: 10.11591/ijeecs.v28.i3.pp1203-1213.
[14] J. Hu, Y. Shan, K. W. Cheng, and S. Islam, “Overview of power converter control in microgrids - challenges, advances, and
future trends,” IEEE Transactions on Power Electronics, vol. 37, no. 8, pp. 9907–9922, 2022, doi: 10.1109/TPEL.2022.3159828.
[15] N. G. F. Dos Santos, J. R. R. Zientarski, and M. L. Da Silva Martins, “A review of series-connected partial power converters for
DC-DC applications,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 10, no. 6, pp. 7825–7838, 2022,
doi: 10.1109/JESTPE.2021.3082869.
[16] W. Kramer, S. Chakraborty, B. Kroposki, and H. Thomas, “Advanced power electronic interfaces for distributed energy systems
part 1: systems and topologies,” Golden, CO, Mar. 2008. doi: 10.2172/926102.
[17] B. Zhao, Q. Song, W. Liu, and Y. Sun, “Overview of Dual-Active-Bridge Isolated Bidirectional DC–DC Converter for High-
Frequency-Link Power-Conversion System,” IEEE Transactions on Power Electronics, vol. 29, no. 8, pp. 4091–4106, Aug. 2014,
doi: 10.1109/TPEL.2013.2289913.
[18] V. Karthikeyan and R. Gupta, “Zero circulating current modulation for isolated bidirectional dual-active-bridge DC-DC
converter,” IET Power Electronics, vol. 9, no. 7, pp. 1553–1561, 2016, doi: 10.1049/iet-pel.2015.0475.
[19] Z. Wang and H. Li, “An integrated three-port bidirectional DC-DC converter for PV application on a DC distribution system,”
IEEE Transactions on Power Electronics, vol. 28, no. 10, pp. 4612–4624, 2013, doi: 10.1109/TPEL.2012.2236580.
[20] A. K. Rathore and U. R. Prasanna, “Comparison of soft-switching voltage-fed and current-fed bi-directional isolated Dc/Dc
converters for fuel cell vehicles,” IEEE International Symposium on Industrial Electronics, pp. 252–257, 2012, doi:
10.1109/ISIE.2012.6237093.
[21] X. Liu et al., “Novel dual-phase-shift control with bidirectional inner phase shifts for a dual-active-bridge converter having low
surge current and stable power control,” IEEE Transactions on Power Electronics, vol. 32, no. 5, pp. 4095–4106, 2017, doi:
10.1109/TPEL.2016.2593939.
[22] D. Segaran, D. G. Holmes, and B. P. McGrath, “Comparative analysis of single- And three-phase dual active bridge bidirectional
DC-DC converters,” Australian Journal of Electrical and Electronics Engineering, vol. 6, no. 3, pp. 329–337, 2009, doi:
10.1080/1448837X.2009.11464251.
[23] X. Li and A. K. S. Bhat, “Analysis and design of high-frequency isolated dual-bridge series resonant DC/DC converter,” IEEE
Transactions on Power Electronics, vol. 25, no. 4, pp. 850–862, 2010, doi: 10.1109/TPEL.2009.2034662.
[24] H.-J. Chiu and L.-W. Lin, “A bidirectional DC–DC converter for fuel cell electric vehicle driving system,” IEEE Transactions on
Power Electronics, vol. 21, no. 4, pp. 950–958, Jul. 2006, doi: 10.1109/TPEL.2006.876863.
[25] D. Sha, X. Wang, K. Liu, and C. Chen, “A current-fed dual-active-bridge DC–DC converter using extended duty cycle control
and magnetic-integrated inductors with optimized voltage mismatching control,” IEEE Transactions on Power Electronics, vol.
34, no. 1, pp. 462–473, Jan. 2019, doi: 10.1109/TPEL.2018.2825991.
[26] Z. Qin, Y. Shen, P. C. Loh, H. Wang, and F. Blaabjerg, “A dual active bridge converter with an extended high-efficiency range by
DC blocking capacitor voltage control,” IEEE Transactions on Power Electronics, vol. 33, no. 7, pp. 5949–5966, 2018, doi:
10.1109/TPEL.2017.2746518.
[27] Y. Shi, R. Li, Y. Xue, and H. Li, “Optimized operation of current-fed dual active bridge DC-DC converter for PV applications,”
IEEE Transactions on Industrial Electronics, vol. 62, no. 11, pp. 6986–6995, 2015, doi: 10.1109/TIE.2015.2432093.
[28] K. Bathala, D. Kishan, and N. Harischandrappa, “Current source isolated bidirectional series resonant DC-DC converter for solar
power/fuel cell and energy storage application,” IECON Proceedings (Industrial Electronics Conference), vol. 2021-Octob, 2021,
doi: 10.1109/IECON48115.2021.9589693.
[29] P. J. Wolfs, “A current-sourced DC-DC converter derived via the duality principle from the half-bridge converter,” IEEE
Transactions on Industrial Electronics, vol. 40, no. 1, pp. 139–144, 1993, doi: 10.1109/41.184830.
[30] D. Liu and H. Li, “A three-port three-phase DC-DC converter for hybrid low voltage fuel cell and ultracapacitor,” IECON
Proceedings (Industrial Electronics Conference), pp. 2558–2563, 2006, doi: 10.1109/IECON.2006.347822.
[31] H. S. H. Chung, W. L. Cheung, and K. S. Tang, “A ZCS bidirectional flyback dc/dc converter,” IEEE Transactions on Power
Electronics, vol. 19, no. 6, pp. 1426–1434, 2004, doi: 10.1109/TPEL.2004.836643.
[32] Z. Zhang, O. C. Thomsen, and M. A. E. Andersen, “Optimal design of a push-pull-forward half-bridge (PPFHB) bidirectional
DC-DC converter with variable input voltage,” IEEE Transactions on Industrial Electronics, vol. 59, no. 7, pp. 2761–2771, 2012,
doi: 10.1109/TIE.2011.2134051.
[33] H. Bai, Z. Nie, and C. C. Mi, “Experimental comparison of traditional phase-shift, dual-phase-shift, and model-based control of
isolated bidirectional dc-dc converters,” IEEE Transactions on Power Electronics, vol. 25, no. 6, pp. 1444–1449, 2010, doi:
10.1109/TPEL.2009.2039648.
[34] S. S. Muthuraj, V. K. Kanakesh, P. Das, and S. K. Panda, “Triple phase shift control of an LLL tank based bidirectional dual active bridge
converter,” IEEE Transactions on Power Electronics, vol. 32, no. 10, pp. 8035–8053, 2017, doi: 10.1109/TPEL.2016.2637506.

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  989

[35] S. Habib et al., “Contemporary trends in power electronics converters for charging solutions of electric vehicles,” CSEE Journal
of Power and Energy Systems, vol. 6, no. 4, pp. 911–929, 2020, doi: 10.17775/CSEEJPES.2019.02700.
[36] X. Li and S. Wang, “Energy management and operational control methods for grid battery energy storage systems,” CSEE
Journal of Power and Energy Systems, vol. 7, no. 5, pp. 1026–1040, 2021, doi: 10.17775/CSEEJPES.2019.00160.
[37] R. B. Schainker, “Executive overview: Energy storage options for a sustainable energy future,” 2004 IEEE Power Engineering
Society General Meeting, vol. 2, pp. 2309–2314, 2004, doi: 10.1109/pes.2004.1373298.
[38] R. L. Steigerwald, “A comparison of half-bridge resonant converter topologies,” APEC 1987 - 2nd Annual IEEE Applied Power
Electronics Conference and Exposition, Conference Proceedings, pp. 135–144, 2015, doi: 10.1109/APEC.1987.7067142.
[39] T. Jiang, J. Zhang, X. Wu, K. Sheng, and Y. Wang, “A bidirectional LLC resonant converter with automatic forward and
backward mode transition,” IEEE Transactions on Power Electronics, vol. 30, no. 2, pp. 757–770, 2015, doi:
10.1109/TPEL.2014.2307329.
[40] P. Xuewei and A. K. Rathore, “Novel interleaved bidirectional snubberless naturally clamped zero current commutated soft-
switching current-fed full-bridge voltage doubler for fuel cell vehicles,” 2013 IEEE Energy Conversion Congress and Exposition,
ECCE 2013, pp. 3615–3622, 2013, doi: 10.1109/ECCE.2013.6647177.
[41] N. Hou and Y. W. Li, “Overview and comparison of modulation and control strategies for a nonresonant single-phase dual-active-bridge
DC-DC converter,” IEEE Transactions on Power Electronics, vol. 35, no. 3, pp. 3148–3172, 2020, doi: 10.1109/TPEL.2019.2927930.
[42] S. Shao, H. Chen, X. Wu, J. Zhang, and K. Sheng, “Circulating current and ZVS-on of a dual active bridge DC-DC converter: a
review,” IEEE Access, vol. 7, pp. 50561–50572, 2019, doi: 10.1109/ACCESS.2019.2911009.
[43] Y. Lu, Q. Wu, Q. Wang, D. Liu, and L. Xiao, “Analysis of a novel zero-voltage-switching bidirectional DC/DC converter for energy
storage system,” IEEE Transactions on Power Electronics, vol. 33, no. 4, pp. 3169–3179, Apr. 2018, doi: 10.1109/TPEL.2017.2703949.
[44] G. Xu, D. Sha, Y. Xu, and X. Liao, “Hybrid-bridge-based DAB converter with voltage match control for wide voltage conversion gain
application,” IEEE Transactions on Power Electronics, vol. 33, no. 2, pp. 1378–1388, 2018, doi: 10.1109/TPEL.2017.2678524.
[45] V. Karthikeyan and R. Gupta, “Multiple-input configuration of isolated bidirectional DC-DC converter for power flow control in
combinational battery storage,” IEEE Transactions on Industrial Informatics, vol. 14, no. 1, pp. 2–11, 2018, doi:
10.1109/TII.2017.2707106.
[46] Y. Li, F. Li, F. W. Zhao, X. J. You, K. Zhang, and M. Liang, “Hybrid three-level full-bridge isolated buck-boost converter with
clamped inductor for wider voltage range application,” IEEE Transactions on Power Electronics, vol. 34, no. 3, pp. 2923–2937,
2019, doi: 10.1109/TPEL.2018.2845308.
[47] F. Li, Y. Li, and X. You, “Optimal dual-phase-shift control strategy of an isolated buck-boost converter with a clamped inductor,”
IEEE Transactions on Power Electronics, vol. 33, no. 6, pp. 5374–5385, 2018, doi: 10.1109/TPEL.2017.2732439.
[48] J. Yang, J. Liu, J. Zhang, N. Zhao, Y. Wang, and T. Q. Zheng, “Multirate digital signal processing and noise suppression for dual
active bridge DC-DC converters in a power electronic traction transformer,” IEEE Transactions on Power Electronics, vol. 33,
no. 12, pp. 10885–10902, 2018, doi: 10.1109/TPEL.2018.2803744.
[49] V. Karthikeyan and R. Gupta, “FRS-DAB converter for elimination of circulation power flow at input and output ends,” IEEE
Transactions on Industrial Electronics, vol. 65, no. 3, pp. 2135–2144, 2018, doi: 10.1109/TIE.2017.2740853.
[50] B. Zhao, Q. Song, and W. Liu, “Efficiency characterization and optimization of isolated bidirectional DC-DC converter based on
dual-phase-shift control for DC distribution application,” IEEE Transactions on Power Electronics, vol. 28, no. 4, pp. 1711–1727,
2013, doi: 10.1109/TPEL.2012.2210563.
[51] 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 Conference Record of the 1988 IEEE Industry Applications Society Annual Meeting,
Sep. 1991, vol. 27, no. 1, pp. 796–805, doi: 10.1109/IAS.1988.25153.
[52] J. E. Ochoa Sosa, R. O. Nunez, G. G. Oggier, and G. O. Garcia, “Open transistors fault-tolerant schemes of three-phase dual active bridge
DC-DC converters,” IEEE Latin America Transactions, vol. 19, no. 3, pp. 385–395, 2021, doi: 10.1109/TLA.2021.9447587.
[53] J. Walter and R. W. De Doncker, “High-power bi-directional DC/DC converter topology for future automobiles,” EPE Journal
(European Power Electronics and Drives Journal), vol. 14, no. 2, pp. 28–33, 2004, doi: 10.1080/09398368.2004.11463558.
[54] H. Krishnaswami and N. Mohan, “Three-port series-resonant DC-DC converter to interface renewable energy sources with
bidirectional load and energy storage ports,” IEEE Transactions on Power Electronics, vol. 24, no. 10, pp. 2289–2297, 2009, doi:
10.1109/TPEL.2009.2022756.
[55] E. E. Henao-Bravo, C. A. Ramos-Paja, A. J. Saavedra-Montes, D. González-Montoya, and J. Sierra-Pérez, “Design method of dual active
bridge converters for photovoltaic systems with high voltage gain,” Energies, vol. 13, no. 7, 2020, doi: 10.3390/en13071711.
[56] J. Huang, Z. Li, L. Shi, Y. Wang, and J. Zhu, “Optimized modulation and dynamic control of a three-phase dual active bridge converter
with variable duty cycles,” IEEE Transactions on Power Electronics, vol. 34, no. 3, pp. 2856–2873, 2019, doi:
10.1109/TPEL.2018.2842021.
[57] J.-H. Jung, H.-S. Kim, M.-H. Ryu, and J.-W. Baek, “Design methodology of bidirectional CLLC resonant converter for high-
frequency isolation of dc distribution systems,” IEEE Transactions on Power Electronics, vol. 28, no. 4, pp. 1741–1755, Apr.
2013, doi: 10.1109/TPEL.2012.2213346.
[58] W. Chen, P. Rong, and Z. Lu, “Snubberless bidirectional DC-DC converter with new CLLC resonant tank featuring minimized switching
loss,” IEEE Transactions on Industrial Electronics, vol. 57, no. 9, pp. 3075–3086, 2010, doi: 10.1109/TIE.2009.2037099.
[59] H. Chen and A. K. S. Bhat, “A bidirectional dual-bridge LCL-type series resonant converter controlled with modified gating
scheme,” 2016 IEEE 8th International Power Electronics and Motion Control Conference, IPEMC-ECCE Asia 2016, pp. 3036–
3042, 2016, doi: 10.1109/IPEMC.2016.7512780.
[60] I. W. Hofsajer, J. A. Ferreira, and J. D. van Wyk, “A new manufacturing and packaging technology for the integration of power
electronics,” in IAS ’95. Conference Record of the 1995 IEEE Industry Applications Conference Thirtieth IAS Annual Meeting,
1995, vol. 1, pp. 891–897, doi: 10.1109/IAS.1995.530392.
[61] A. Isurin and A. Cook, “A novel resonant converter topology and its application,” PESC Record - IEEE Annual Power
Electronics Specialists Conference, vol. 2, pp. 1039–1044, 2001, doi: 10.1109/PESC.2001.954256.
[62] W. L. Malan, D. M. Vilathgamuwa, and G. R. Walker, “Modeling and control of a resonant dual active bridge with a tuned CLLC
network,” IEEE Transactions on Power Electronics, vol. 31, no. 10, pp. 7297–7310, 2016, doi: 10.1109/TPEL.2015.2507787.
[63] F. C. Lee, S. Wang, P. Kong, C. Wang, and D. Fu, “Power architecture design with improved system efficiency, EMI and power density,”
in PESC Record - IEEE Annual Power Electronics Specialists Conference, Jun. 2008, pp. 4131–4137, doi: 10.1109/PESC.2008.4592602.
[64] B. Yang and F. C. Lee, “LLC resonant converter for front end DC / DC conversion alpha J . Zhang and Guisong Huang,” vol. 00,
no. c, pp. 1108–1112, 2002.
[65] M. Yaqoob, K. H. Loo, and Y. M. Lai, “Modeling the effect of dead-Time on the soft-switching characteristic of variable-
frequency modulated series-resonant DAB converter,” 2017 IEEE 18th Workshop on Control and Modeling for Power
Electronics, COMPEL 2017, 2017, doi: 10.1109/COMPEL.2017.8013306.
High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)
990  ISSN: 2088-8694

[66] C. Mi, H. Bai, C. Wang, and S. Gargies, “Operation, design and control of dual H-bridge-based isolated bidirectional DC-DC
converter,” IET Power Electronics, vol. 1, no. 4, pp. 507–5017, 2008, doi: 10.1049/iet-pel:20080004.
[67] P. Xuewei and A. K. Rathore, “Comparison of bi-directional voltage-fed and current-fed dual active bridge isolated dc/dc
converters low voltage high current applications,” IEEE International Symposium on Industrial Electronics, pp. 2566–2571, 2014,
doi: 10.1109/ISIE.2014.6865024.
[68] P. Xuewei and A. K. Rathore, “Novel bidirectional snubberless naturally commutated soft-switching current-fed full-bridge
isolated DC/DC converter for fuel cell vehicles,” IEEE Transactions on Industrial Electronics, vol. 61, no. 5, pp. 2307–2315,
2014, doi: 10.1109/TIE.2013.2271599.
[69] Z. Ding, C. Yang, Z. Zhang, C. Wang, and S. Xie, “A novel soft-switching multiport bidirectional dc-dc converter for hybrid energy
storage system,” IEEE Transactions on Power Electronics, vol. 29, no. 4, pp. 1595–1609, 2014, doi: 10.1109/TPEL.2013.2264596.
[70] Z. Guo, K. Sun, T.-F. Wu, and C. Li, “An improved modulation scheme of current-fed bidirectional DC–DC converters for loss
reduction,” IEEE Transactions on Power Electronics, vol. 33, no. 5, pp. 4441–4457, May 2018, doi: 10.1109/TPEL.2017.2719722.
[71] P. Xuewei and A. K. Rathore, “Novel interleaved bidirectional snubberless soft-switching current-fed full-bridge voltage doubler for fuel-
cell vehicles,” IEEE Transactions on Power Electronics, vol. 28, no. 12, pp. 5535–5546, 2013, doi: 10.1109/TPEL.2013.2252199.
[72] J. Riedel, D. G. Holmes, B. P. McGrath, and C. Teixeira, “Maintaining Continuous ZVS Operation of a Dual Active Bridge by
Reduced Coupling Transformers,” IEEE Transactions on Industrial Electronics, vol. 65, no. 12, pp. 9438–9448, 2018, doi:
10.1109/TIE.2018.2815993.
[73] Y. Li, Y. Xing, Y. Lu, H. Wu, and P. Xu, “Performance analysis of a current-fed bidirectional LLC resonant converter,” IECON
Proceedings (Industrial Electronics Conference), pp. 2486–2491, 2016, doi: 10.1109/IECON.2016.7793163.
[74] Z. Shen, R. Burgos, D. Boroyevich, and F. Wang, “Soft-switching capability analysis of a dual active bridge dc-dc converter,”
IEEE Electric Ship Technologies Symposium, ESTS 2009, pp. 334–339, 2009, doi: 10.1109/ESTS.2009.4906533.
[75] E. T. H. Zurich and F. Krismer, “Modeling and optimization of bidirectional dual active bridge DC – DC converter topologies,”
no. 19177, 2010, doi: 10.3929/ethz-a-006395373.
[76] A. K. Jain and R. Ayyanar, “PWM control of dual active bridge: Comprehensive analysis and experimental verification,” IEEE
Transactions on Power Electronics, vol. 26, no. 4, pp. 1215–1227, 2011, doi: 10.1109/TPEL.2010.2070519.
[77] D. Sha, Y. Xu, J. Zhang, and Y. Yan, “Current-fed hybrid dual active bridge DC–DC converter for a fuel cell power conditioning
system with reduced input current ripple,” IEEE Transactions on Industrial Electronics, vol. 64, no. 8, pp. 6628–6638, Aug.
2017, doi: 10.1109/TIE.2017.2698376.
[78] H. Wu, S. Ding, K. Sun, L. Zhang, Y. Li, and Y. Xing, “Bidirectional soft-switching series-resonant converter with simple PWM
control and load-independent voltage-gain characteristics for energy storage system in DC microgrids,” IEEE Journal of
Emerging and Selected Topics in Power Electronics, vol. 5, no. 3, pp. 995–1007, 2017, doi: 10.1109/JESTPE.2017.2651049.
[79] A. K. Rathore and U. R. Prasanna, “Analysis, design, and experimental results of novel snubberless bidirectional naturally
clamped ZCS/ZVS current-fed half-bridge DC/DC converter for fuel cell vehicles,” IEEE Transactions on Industrial Electronics,
vol. 60, no. 10, pp. 4482–4491, 2013, doi: 10.1109/TIE.2012.2213563.
[80] D. Liu and H. Li, “A ZVS bi-directional DC-DC converter for multiple energy storage elements,” IEEE Transactions on Power
Electronics, vol. 21, no. 5, pp. 1513–1517, 2006, doi: 10.1109/TPEL.2006.882450.
[81] Z. Wang and H. Li, “A soft switching three-phase current-fed bidirectional DC-DC converter with high efficiency over a wide input
voltage range,” IEEE Transactions on Power Electronics, vol. 27, no. 2, pp. 669–684, 2012, doi: 10.1109/TPEL.2011.2160284.
[82] H. Pinheiro and P. K. Jain, “Series-parallel resonant UPS with capacitive output DC bus filter for powering HFC networks,” IEEE
Transactions on Power Electronics, vol. 17, no. 6, pp. 971–979, Nov. 2002, doi: 10.1109/TPEL.2002.805595.
[83] M. H. Kheraluwala, R. W. Gascoigne, D. M. Divan, and E. D. Baumann, “Performance characterization of a high-power dual active
bridge dc-to-dc converter,” IEEE Transactions on Industry Applications, vol. 28, no. 6, pp. 1294–1301, 1992, doi: 10.1109/28.175280.
[84] M. Nowak, J. Hildebrandt, and P. Luniewski, “Converters with AC transformer intermediate link suitable as interfaces for
supercapacitor energy storage,” PESC Record - IEEE Annual Power Electronics Specialists Conference, vol. 5, pp. 4067–4073,
2004, doi: 10.1109/PESC.2004.1355196.
[85] 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,” IEEE Transactions on Industry Applications, vol. 27, no. 1, pp. 63–73, 1991, doi: 10.1109/28.67533.
[86] O. García, L. A. Flores, J. A. Oliver, J. A. Cobos, and J. De La Peña, “Bi-directional dc-dc converter for hybrid vehicles,” PESC Record -
IEEE Annual Power Electronics Specialists Conference, vol. 2005, pp. 1881–1886, 2005, doi: 10.1109/PESC.2005.1581888.
[87] A. A. Aboulnaga and A. Emadi, “Performance evaluation of the isolated bidirectional ćuk converter with integrated magnetics,” PESC
Record - IEEE Annual Power Electronics Specialists Conference, vol. 2, pp. 1557–1562, 2004, doi: 10.1109/PESC.2004.1355657.
[88] D. Murthy-Bellur and M. K. Kazimierczuk, “Isolated two-transistor zeta converter with reduced transistor voltage stress,” IEEE
Transactions on Circuits and Systems II: Express Briefs, vol. 58, no. 1, pp. 41–45, 2011, doi: 10.1109/TCSII.2010.2092829.
[89] E. V. De Souza and I. Barbi, “Bidirectional current-fed fly back-push-pull DC-DC converter,” pp. 8–13, 2011.
[90] F. D. Esteban, F. M. Serra, and C. H. De Angelo, “Control of a DC-DC dual active bridge converter in DC microgrids
applications,” IEEE Latin America Transactions, vol. 19, no. 8, pp. 1261–1269, 2021, doi: 10.1109/TLA.2021.9475856.
[91] H. Bai and C. Mi, “Eliminate reactive power and increase system efficiency of isolated bidirectional dual-active-bridge dc-dc
converters using novel dual-phase-shift control,” IEEE Transactions on Power Electronics, vol. 23, no. 6, pp. 2905–2914, 2008,
doi: 10.1109/TPEL.2008.2005103.
[92] B. Zhao, Q. Yu, and W. Sun, “Extended-phase-shift control of isolated bidirectional DC-DC converter for power distribution in
microgrid,” IEEE Transactions on Power Electronics, vol. 27, no. 11, pp. 4667–4680, 2012, doi: 10.1109/TPEL.2011.2180928.
[93] B. Zhao, Q. Song, and W. Liu, “Power characterization of isolated bidirectional dual-active-bridge dc-dc converter with dual-phase-shift
control,” IEEE Transactions on Power Electronics, vol. 27, no. 9, pp. 4172–4176, 2012, doi: 10.1109/TPEL.2012.2189586.
[94] B. Zhao, Q. Song, W. Liu, and W. Sun, “Current-stress-optimized switching strategy of isolated bidirectional DC-DC converter with dual-
phase-shift control,” IEEE Transactions on Industrial Electronics, vol. 60, no. 10, pp. 4458–4467, 2013, doi: 10.1109/TIE.2012.2210374.
[95] K. Rajapandian, V. Ramanarayanan, and R. Ramkumar, “A 250 kHz/560 W phase modulated converter,” in Proceedings of International
Conference on Power Electronics, Drives and Energy Systems for Industrial Growth, vol. 1, pp. 20–26, doi:
10.1109/PEDES.1996.537276.
[96] F. Krismer and J. W. Kolar, “Accurate small-signal model for the digital control of an automotive bidirectional dual active
bridge,” IEEE Transactions on Power Electronics, vol. 24, no. 12, pp. 2756–2768, 2009, doi: 10.1109/TPEL.2009.2027904.
[97] A. Tong, L. Hang, G. Li, X. Jiang, and S. Gao, “Modeling and analysis of a dual-active-bridge-isolated bidirectional DC/DC
converter to minimize RMS current with whole operating range,” IEEE Transactions on Power Electronics, vol. 33, no. 6, pp.
5302–5316, 2018, doi: 10.1109/TPEL.2017.2692276.

Int J Pow Elec & Dri Syst, Vol. 14, No. 2, June 2023: 969-991
Int J Pow Elec & Dri Syst ISSN: 2088-8694  991

[98] F. Krismer and J. W. Kolar, “Closed form solution for minimum conduction loss modulation of DAB converters,” IEEE
Transactions on Power Electronics, vol. 27, no. 1, pp. 174–188, 2012, doi: 10.1109/TPEL.2011.2157976.
[99] C. Wang and Q. Song, “The research on the triple phase shift control of the isolated bidirectional DC-DC converter,” Proceedings
of 2019 IEEE 3rd International Electrical and Energy Conference, CIEEC 2019, 2019, doi: 10.1109/CIEEC47146.2019.CIEEC-
201985.
[100] D. Mou, Q. Luo, J. Li, Y. Wei, and P. Sun, “Five-degree-of-freedom modulation scheme for dual active bridge DC-DC
converter,” IEEE Transactions on Power Electronics, vol. 36, no. 9, pp. 10584–10601, 2021, doi: 10.1109/TPEL.2021.3056800.
[101] X. Chen, G. Xu, H. Han, D. Liu, Y. Sun, and M. Su, “Light-Load Efficiency Enhancement of High-frequency Dual-Active-Bridge
Converter under SPS Control,” IEEE Transactions on Industrial Electronics, vol. 68, no. 12, pp. 12941–12946, 2021, doi:
10.1109/TIE.2020.3044803.
[102] J. Guacaneme, G. Garcerá, E. Figueres, I. Patrao, and R. González-Medina, “Dynamic modeling of a dual active bridge DC to DC
converter with average current control and load-current feed-forward,” International Journal of Circuit Theory and Applications,
vol. 43, no. 10, pp. 1311–1332, 2015, doi: 10.1002/cta.2012.
[103] X. Pan, H. Li, Y. Liu, T. Zhao, C. Ju, and A. K. Rathore, “An Overview and comprehensive comparative evaluation of current-
fed-isolated-bidirectional DC/DC converter,” IEEE Transactions on Power Electronics, vol. 35, no. 3, pp. 2737–2763, 2020, doi:
10.1109/TPEL.2019.2931739.
[104] G. J. Su and L. Tang, “A multiphase, modular, bidirectional, triple-voltage dc-dc converter for hybrid and fuel cell vehicle power
systems,” IEEE Transactions on Power Electronics, vol. 23, no. 6, pp. 3035–3046, 2008, doi: 10.1109/TPEL.2008.2005386.
[105] R. Lenke, F. Mura, and R. W. De Doncker, “Comparison of non-resonant and super-resonant dual-active ZVS-operated high-
power DC-DC converters,” 2009 13th European Conference on Power Electronics and Applications, EPE ’09, 2009.
[106] K. Wang, C. Y. Lin, L. Zhu, D. Qu, F. C. Lee, and J. S. Lai, “Bi-directional dc to dc converters for fuel cell systems,” IEEE
Workshop on Power Electronics in Transportation, pp. 47–51, 1998, doi: 10.1109/pet.1998.731056.
[107] X. Sun, X. Wu, Y. Shen, X. Li, and Z. Lu, “A Current-Fed Isolated Bidirectional DC-DC Converter,” IEEE Transactions on
Power Electronics, vol. 32, no. 9, pp. 6882–6895, 2017, doi: 10.1109/TPEL.2016.2623306.

BIOGRAPHIES OF AUTHORS

Kiran Bathala (Student Member, IEEE) received a Bachelor’s Degree in


Electrical and Electronics Engineering from Jawaharlal Nehru Technological University
Anantapur (SVCET), India, in 2009. He completed his Master’s Degree from National
Institute of Technology Calicut, India, in 2012. Currently continuing his Ph.D. degree in
Electrical & Electronics Engineering in National Institute of Technology, Karnataka, India
from 2017. Before joining his Ph.D., he was working as Asst. Professor in SVCET Chittoor,
A.P, India from 2012-2016. His research interests are Power electronics in integrating
renewable energy resources with energy storage systems. Soft switching Converters. He can
be contacted at email: [email protected].

Dharavath Kishan (Senior Member, IEEE) received a Bachelor Degree in


Electrical and Electronics Engineering from Jawaharlal Nehru Technological University,
Hyderabad (MRET), 2011. He completed his Master’s Degree from degree in Electrical &
Electronics Engineering from Jawaharlal Nehru Technological University, Hyderabad
(CMRCET), India, in 2013. He received his Ph.D., Degree from National Institute of
Technology Tiruchirapalli, India, in 2018. Currently serving as Assistant Professor with
Department of Electrical & Electronics Engineering, National Institute of Technology
Karnataka, India. His research interests are Wireless Battery Charging, Smart transportation
and electrification and Electric energy storage systems and motor drives for Transportation
Electrification. He can be contacted at email: [email protected].

Nagendrappa Harischandrappa (Senior Member, IEEE) received the B.E.


Degree in Electrical and Electronics Engineering and the M.Tech. degree in Power and
Energy Systems from the National Institute of Technology Karnataka, India, in 1999 and
2002, respectively, and the Ph.D. degree in electrical engineering from the University of
Victoria, Victoria, BC, Canada, in 2015. He is currently working as an Assistant Professor
with Department of Electrical & Electronics Engineering, National Institute of Technology
Karnataka, India. For a short period, he was an Assistant Engineer (operation and
maintenance) with the Power Distribution Utility, Mangalore Electricity Supply Company
(MESCOM) Ltd., Ajjampura, India. His research interests include high frequency soft-
switching converters for power generation from renewable energy sources and their grid-
interfacing applications. He can be contacted at email: [email protected].

High frequency isolated bidirectional dual active bridge DC-DC converters … (Kiran Bathala)

You might also like