Ubc 2023 May Hsu Jhih-Da PDF
Ubc 2023 May Hsu Jhih-Da PDF
by
Jhih-Da Hsu
Doctor of Philosophy
in
November 2022
High-Performance Resonant Converters for Battery Chargers: Efficiency and Dynamics Improvement
Examining Committee:
Supervisor
Dr. Wilson Eberle, Associate Professor, Electrical and Computer Engineering, UBC Okanagan
Co-supervisor
Dr. Yu Christine Chen, Associate Professor, Electrical and Computer Engineering, UBC
University Examiner
University Examiner
ii
Abstract
As the requirement for clean energy grows, the demand for high-performance power conversion for energy
storage and battery charging applications has been soaring. Resonant converters, in particular, LLC or CLLC
converters, have been broadly adopted for high-power battery chargers. The purpose of this work is to further
improve the performance of the resonant converters from both the efficiency and dynamics aspects.
In terms of improving the efficiency, this work focuses on reducing the conduction losses of the output
rectifiers using Synchronous Rectification (SR). Conventional SR controllers detect the drain-source voltage
during the SR turn-on phase (vds.on ) as the control input. However, vds.on is low-magnitude and sensitive to
the voltage noise caused by parasitic elements. The distorted vds.on causes SR mis-triggering, undermining
the efficiency. With the focus on mainstream LLC resonant converters, this work first introduces a new SR
driving strategy based on the resonant capacitor voltage (RCV). Next, a simplified SR method is proposed;
it is based on the Volt-Second Product (VSP) of SR drain-source blocking voltage and rectifier current con-
duction time. Both methods employ large-magnitude voltages, which are insensitive to the noise generated
by parasitic components, reducing SR on-time error. The proposed SR methods are compared with the
Regarding the dynamics aspect, this work focuses on improving the small-signal dynamic model for
charge-controlled resonant converters. Charge mode control has been applied to resonant converters to im-
prove the system dynamics, and yet the conventional small-signal model emphasizes only the low-frequency
region, which is not suitable for high-bandwidth designs. This work establishes the small-signal modeling
methodology based on Extended Describing Functions (EDF) and phasor analysis, which successfully pre-
dicts the system frequency response across low- to high-frequency regions, enabling high-bandwidth designs.
As the proposed noise-tolerant SR methods improve the efficiency performance, the enhanced small-signal
model assists to achieve wide loop bandwidth, improving the dynamic performance. This work provides
iii
Lay Summary
Power converters are core elements interfacing electricity and batteries and they shape the modern world.
While high-performance converter design involves many aspects, this work emphasizes improvements to
efficiency and dynamic performance. In terms of efficiency, this work proposes two advanced solutions
to controlling the active switches that replace the lossy and passive components. Traditional methods for
controlling the switches are easily mistriggered by non-ideal behaviors of circuit elements. The proposed
solutions successfully increase the efficiency compared with the traditional methods. Regarding the system
dynamics aspect, this work establishes an enhanced dynamic model of the converter which predicts the ac-
tual behavior better than the conventional model. This model assists to achieve high-bandwidth designs, so
that the converter output recovers quickly after external disturbances. This work provides advanced solu-
tions to improving efficiency and dynamic performance of battery chargers. All the results are validated by
iv
Preface
This work is based on research performed at the Department of Electrical and Computer Engineering of the
University of British Columbia by Jhih-Da Hsu, under the supervision of Prof. Martin Ordonez and Prof.
Wilson Eberle. Chapter 1 contains modified portions of text from all the below-listed publications. Portions
of Chapters 2, 3, and 4 have been published in IEEE Transactions on Power Electronics and IEEE Journal
• J. -D. Hsu, M. Ordonez, W. Eberle, M. Craciun and C. Botting, “LLC Synchronous Rectification
Using Resonant Capacitor Voltage,” in IEEE Transactions on Power Electronics, vol. 34, no. 11, pp.
• J. -D. Hsu, M. Ordonez, W. Eberle, M. Craciun and C. Botting, “Noise-Tolerant LLC Synchronous
Rectification Using Volt-Second Product,” in IEEE Journal of Emerging and Selected Topics in Power
• J. -D. Hsu, M. Ordonez, W. Eberle, M. Craciun and C. Botting, “Enhanced Small-Signal Modeling
for Charge-Controlled Resonant Converters,” in IEEE Transactions on Power Electronics, vol. 37, no.
I was the lead investigator of the above papers, developed the proposed concepts, built simulation models,
performed experimental verifications, and wrote the manuscripts. Prof. Martin Ordonez and Prof. Wilson
Eberle were the supervisory coauthor on the above papers and were involved throughout the project and
provided technical advice and edited manuscripts. Mr. Marian Craciun and Mr. Chris Botting were industry
partners (Delta-Q Technologies, Vancouver, BC) in this project and valuable comments on the research from
v
Table of Contents
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
Lay Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xviii
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3.2 SR Based on the Drain-Source Voltage During the Turn-on Phase (vds.on based) . . . 5
vi
2 LLC Synchronous Rectification Using Resonant Capacitor Voltage . . . . . . . . . . . . . . . 16
2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
vii
5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
viii
List of Tables
ix
List of Figures
Figure 1.1 a) The block diagram of a two-stage on-board battery charger. b) The block diagram of
Figure 1.2 Conventional LLC SR control methods: current-driven method with the CT mounted on
the secondary side for sensing the rectifier current, which occupies space and introduces
Figure 1.3 Conventional LLC SR control methods: current-driven method with the CT located on
the primary side for sensing the resonant current, which exchanges space with increased
circuit complexity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Figure 1.4 Conventional LLC SR control methods: vds.on sensing method. vds.on is the voltage
across the SR switches in the on-state for controlling the SR. The magnitude is small
and can be easily affected by the parasitic components, such as the leak inductance (Llk )
Figure 1.5 Typical waveforms of the vds.on based adaptive SR tracking the rectifier current conduc-
Figure 1.6 Potential challenges of the vds.on based adaptive SR: early turn-on in the below-resonance
Figure 1.7 Potential challenges of the vds.on based adaptive SR: premature turn-off in the above-
resonance region, due to drastic change in the current slope and voltage across the leak-
age inductance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
x
Figure 1.8 a) Schematic diagram of a charge-controlled CLLC resonant converter. b) Conventional
small-signal model of the charge loop, where the charge feedback is approximated by
using irec , overlooking the effect of magnetizing current (im ). c) Conceptual illustration
of the frequency response of the conventional charge-loop model, which indicates the
Figure 2.1 Schematic diagram of a half-bridge center-tapped LLC resonant converter with syn-
chronous rectification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Figure 2.2 a) The key waveforms and b) the corresponding equivalent models of the resonant tank
Figure 2.3 a) The key waveforms and b) the corresponding equivalent models of the resonant tank
Figure 2.4 Light-load conditions: a) the key waveforms and b) the corresponding equivalent models
Figure 2.5 Light-load conditions: a) the key waveforms and b) the corresponding equivalent models
Figure 2.6 The proposed RCV strategy senses the large voltage signals ➊ and controls the SR by
simply integrating and comparing the voltage signals ➋. The typical waveforms show
the below-resonance operation for SR1 as an example to visualize the concept of the
proposed RCV strategy, where v p1.int and vCr .int are the voltage integrations of linear
Figure 2.7 The key switching sequence of the proposed RCV SR method in the below-resonance
Figure 2.8 The key switching sequence of the proposed RCV SR method in the above-resonance
Figure 2.9 The key waveforms in light-load conditions for a) below resonance, and b) above reso-
nance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Figure 2.10 The geometric representation of sensitivity analysis using the first harmonic approximation. 31
xi
Figure 2.12 Simulation results for the constant current mode operation. The converter operates at
27 A, a) 250 kHz, b) 230 kHz, c) 220 kHz, and d) 200 kHz; the RCV-controlled SR on-
time (tRCV ) is very close to the ideal rectifier current conduction time (tCOND ) as shown
in each condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Figure 2.13 Simulation results for the constant power mode operation. The converter operates at
650 W, a) 180 kHz, b) 175 kHz, c) 173 kHz, and d) 167 kHz. In each condition, the
RCV strategy accurately enables the SR during the ideal rectifier current conduction time. 37
Figure 2.14 Simulation results for the constant voltage mode operation, where the LLC converter
the load change, the ideal rectifier current conduction time is accurately predicted by the
RCV strategy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Figure 2.15 The experimental setup: the prototype of the RCV SR controller on the 650 W/ 24 V
Figure 2.16 Experimental results at the maximum power and below resonance, 650 W/ 170 kHz:
a) the vds.on sensing method turns off the SR early due to the inductive voltage offset
induced from the stray inductance, whereas b) the RCV strategy effectively estimates
the zero current crossing point by comparing the integrals of large and stable voltage
signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Figure 2.17 Experimental results at the maximum current, 27 A/ 250 kHz: a) because of the parasitic
effects, the traditional vds.on sensing method turns off the SR prematurely, while b) the
RCV strategy effectively estimates the zero current crossing by comparing the integrated
voltage signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Figure 2.18 Experimental results at light-load conditions, 5 A/ 30 V: a) the traditional vds.on sensing
method turns off the SR early due to the effect of parasitic elements. b) By using the
integrated voltage signals, the on-time error at light load is short in comparison. . . . . . 40
Figure 2.19 The ideal, RCV-controlled and traditional vds.on sensing-based SR on-time are compared.
b) the constant power mode, and c) the constant current mode. The RCV strategy is
effective in multiple operating conditions by using the integrations of the large voltage
signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
xii
Figure 2.20 Comparison of the effect of the SR on-time error on a) the efficiency and b) the SR
MOSFET temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Figure 2.21 The dynamic test waveforms under a large frequency step change, a) from 235 kHz to
190 kHz, and b) from 190 kHz to 235 kHz: the RCV-controlled SR tracks the actual rec-
tifier current conduction time instantaneously under severe voltage and current dynamics. 43
Figure 3.1 Typical waveforms of a half-bridge LLC resonant converter in the above-resonance region. 46
Figure 3.2 Typical waveforms of a half-bridge LLC resonant converter in the below-resonance region. 49
Figure 3.3 a) a typical LLC resonant converter with synchronous rectification, including circuit
parasitic components. b) block diagram of the conventional vds.on based SR and the
proposed Volt-Second Product (VSP) based SR; the former depends on the drain-source
voltage during the turn-on phase, while the latter employs the drain-source voltage dur-
ing the turn-off phase. c) vds.on1 and vds.on2 are in millivolt range and are easily affected
by the parasitic effects, whereas the proposed VSP SR uses large-magnitude voltage
Figure 3.4 Functional block diagram of the proposed Volt-Second Product (VSP) SR. . . . . . . . . 52
Figure 3.5 Typical waveforms of the proposed VSP SR: a) above-resonance and b) below-resonance
operations. The Tsc time window allows for correct setting of the SR for both cases. . . . 54
Figure 3.6 System block diagram of a typical closed-loop controlled LLC resonant converter with
Figure 3.7 The relationship between the change in the switching period (∆ts ) and the closed-loop
controller gain defines the rate of change in the rectifier current conduction time in the
above-resonance region. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Figure 3.8 Block diagram for implementation of the proposed VSP SR using TMS320F28379D
Figure 3.9 Timing sequence of the VSP SR controller for below- and above-resonance operations.
tSR1 [k] is computed at ADC1 where vDp1 [k] is sampled. Sync1 resets the time counter,
Figure 3.10 Flowchart of ADC1 interrupt service routine (ISR). In the kth cycle, it executes the core of
the VSP SR algorithm to determine the SR on-time tSR1 [k] based on tcnd1 [k − 1], vDp1 [k],
and vT [k]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
xiii
Figure 3.11 Plot of the theoretical, normalized rectifier current conduction time, tA , and time window
Tsc , considering 25 %, 50 %, 75 %, and 100 % load. The selection of Tsc is around the
center with sufficient margin between the rectifier current conduction time and tA . . . . 61
Figure 3.12 Simulation results of the VSP SR and the conventional adaptive SR under closed-loop
regulation: a) vo rising and b) vo falling. In both cases, the VSP SR maintains the dead
time δ T above the preset safe margin DT , whereas the conventional adaptive SR has
multiple cycles with insufficient safe margin during frequency rise, increasing the risk
of cross conduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Figure 3.13 The prototype of the 650 W/ 24 V LLC resonant converter with the proposed VSP SR
controller. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Figure 3.14 Steady-state test results in constant voltage regulation at 30 V/ 650 W, fs = 167 kHz: a)
the conventional adaptive SR turns on the switch early due to voltage ringing-induced SR
body diode conduction, introducing reverse current; b) the proposed VSP SR controls
SR on-time properly. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Figure 3.15 Steady-state test results in constant current regulation at 27 A/ 540 W, fs = 238 kHz:
a) the conventional adaptive SR turns off the switch prematurely due drastic change in
the current slope and voltage across the leakage inductance; b) the proposed VSP SR is
Figure 3.16 Efficiency test results under: a) constant voltage regulation, Vo = 30 V , and b) constant
current regulation, Io = 27 A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Figure 3.17 The dynamic test waveforms under a large command change, a) from Vmin to Vmax , and
b) from Vmax to Vmin : the VSP SR tracks the rectifier current conduction time properly
Figure 4.2 The AC equivalent circuit of the resonant tank using a transformer T-model. . . . . . . . 69
Figure 4.3 Extended Describing Functions (EDF) model of an open-loop, direct frequency-controlled,
half-bridge CLLC resonant converter. Input current is proportional to the sine compo-
nent of the primary resonant current, which forms the basis of the proposed enhanced
charge-control model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
xiv
Figure 4.4 a) Schematic diagram of a CLLC resonant converter with charge mode control. b) Block
diagram of the charge loop: the conventional model approximates the charge loop by
using irec , overlooking the effect of magnetizing current (im ). c) Conceptual illustration
of the frequency response of the charge loop: the proposed model removes the prediction
Figure 4.5 Block diagram of the conventional charge model. The equivalent feedback loop from
the output does not fully represent the actual charge loop. . . . . . . . . . . . . . . . . . 79
Figure 4.6 Timing sequence diagram for charge control without slope compensation along with the
Figure 4.7 Relationship between the sine and cosine components of the resonant current using Ex-
tended Describing Functions (EDF) in: a) time domain and b) complex plane. . . . . . . 81
Figure 4.8 Proposed block diagram of the small-signal model of a charge-controlled resonant con-
verter: a) direct b) cascade form. The feedback signal is the sine component of the
Figure 4.9 Timing sequence diagram for charge control with slope compensation along with the
Figure 4.10 Proposed equivalent small-signal circuit of a charge-controlled CLLC converter: focused
Figure 4.11 Comparison of the conventional and proposed analytical models (charge control-to-
output transfer function, Gvc (s)) with circuit simulation results at a) 234 V / 3.3 A; b)
320 V / 3.3 A; c) 400 V / 2.5 A; d) 400 V / 0.8 A. Estimation errors from the conventional
sign. The proposed model successfully predicts the frequency response across all the
conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Figure 4.12 The effect of slope compensation on the charge loop gain Gvc (s) predicted by the pro-
Figure 4.13 Simplified schematic diagram and key components of the test setup for verifying the
Figure 4.14 Analytical and experimental loop frequency response under a) constant-voltage and
b) constant-current regulation; the analytical results using the proposed model closely
xv
Figure 4.15 Experimental transient response under closed-loop regulation: a) 100 % – 80 % load; b)
xvi
List of Abbreviations
AC Alternating Current
CT Current Transformer
DC Direct Current
EV Electric Vehicle
IC Integrated Circuit
RC Resistor-Capacitor
SR Synchronous Rectification
xvii
Acknowledgments
I would like to take this opportunity to express my sincere gratitude to my supervisor, Prof. Martin Ordonez,
for all his support, guidance, and mentoring on my Ph.D. study throughout the past five years, during which
I have gained so much experience as well as technical and soft skills from him. This is the most precious
period in my career.
Special thanks to my committee members, Prof. William Dunford, and Prof. Wilson Eberle, for their
My appreciation also goes to the University of British Columbia (UBC), the Natural Sciences and En-
gineering Research Council (NSERC), and Delta-Q Technologies for supporting my research project. In
particular, Mr. Marian Craciun and Mr. Chris Botting from Delta-Q Technologies generously shared their
Many thanks to my colleagues, especially Mr. Franco Degioanni, Mr. Abbas Arshadi, Mr. Ettore Scabeni
Glitz, Dr. Ali Saket, Dr. Ignacio Galiano, Dr. Francisco Paz, and Dr. Ion Isbasescu, for all the informative
Finally but importantly, I thank my family, especially my beloved wife, Bonnie, for all their unconditional
xviii
Chapter 1
Introduction
1.1 Motivation
In recent years, the demand for electric vehicles (EVs) has been growing as many countries are aiming for
lower carbon emissions. To charge EV batteries, battery chargers are required to convert alternating current
(AC) from the utility grid to direct current (DC). Therefore, the power conversion performance is critical.
High-performance battery chargers convert power efficiently, which has environmental and financial benefits.
Moreover, they regulate the charging voltage in a certain range to avoid battery overcharge and undercharge,
which extends the battery life and makes the batteries and EVs more sustainable.
To convert AC to DC and facilitate EV battery charging, on-board battery chargers are equipped not only
on passenger EVs [4–6] but also on utility vehicles, such as golf carts and fork lifts [7, 8]. On-board battery
chargers have size and weight constraints, which makes the power conversion efficiency and loss reduction
even more critical. Conventionally, on-board battery chargers are composed of two stages: an AC/DC power
converter followed by a DC/DC converter, as shown in Fig. 1.1 (a). The first stage is mainly responsible
for maintaining power quality, which benefits the utility grid in many ways, such as maximizing real power
distribution. It also provides a constant, direct voltage to the second stage. The second stage takes the DC
output from the first stage and converts it to the required output voltage or current to charge the battery. It
responds to the command and regulate the output dynamically. In this research, the efficiency performance
and dynamic response of the DC/DC converter are the main focus.
In terms of efficiency improvement, a common approach is to reduce switching and conduction losses
of the power converter. Switching losses are caused by the switching of active components in the circuit.
These losses are proportional to the switching frequency and the overlap area of voltage and current of the
1
On-Board Battery Charger
Iout
(a)
Iout
Vin Vout
Figure 1.1: a) The block diagram of a two-stage on-board battery charger. b) The block diagram of a
typical DC/DC resonant converter.
switch during the transition. When the switches are switched at zero voltage or current, this is considered
soft-switching, and the switching losses are small. To achieve soft switching, advanced power converter
designs employ a resonant topology. A typical DC/DC resonant converter is composed of active switches
followed by a resonant tank, a transformer for isolation, a rectifier stage, and a output filter stage, as shown
in Fig. 1.1 (b). It achieves zero-voltage or zero-current switching using the resonant behaviors among the
capacitors and inductors in the resonant tank [9–11]. Among various topologies of resonant converters, the
LLC resonant converter, of which the resonant tank is composed of two inductors and one capacitor, has been
broadly applied to power converter designs [12–14]. LLC resonant converters are suitable for the DC/DC
stage of battery chargers primarily due to the ability to step up and down the output voltage in a certain
range [15–18].
LLC resonant converters first convert the DC input to AC to enable the resonance in the resonant tank.
The resonant voltage and current are high-frequency AC, which facilitates the power transfer using a com-
pact, high-frequency transformer. To charge batteries with DC power, the AC power inevitably needs to be
rectified. In LLC converters, the rectifier stage after the transformer is conventionally composed of rectifier
diodes. When conducting, the forward voltage drop of the diode and the average current flowing through
cause conduction losses. For high-power and high-current applications, the high current stress results in high
To reduce the conduction loss of the output rectifiers and further boost the power conversion efficiency,
2
synchronous rectification (SR) technologies are developed [22–24]. SR replaces the rectifier diodes with
active switches. The SR switches have low resistance in the on state and have the same characteristics as
diodes in the off state. Due to the small on-state resistor, the conduction loss of SR is very little. Therefore,
SR is particularly effective to reduce the conduction loss and improve the efficiency of high power, high
current battery chargers. From the efficiency point of view, to benefit the most from SR, it is ideal to make
the SR on-time the same as the rectifier current conduction time. The efficiencies of different SR methods
are mainly determined by the difference between the SR on-time and the actual rectifier current conduction
time. The larger the on-time error is, the less efficient the SR becomes. There are other aspects such as noise
tolerance and transient response, which opens the door to research opportunities.
In terms of the dynamic performance, to make the resonant converter in the DC/DC stage respond to the
battery charging command quickly, high loop bandwidth is desirable. Conventionally, when only the output
voltage is sensed and controlled, the control scheme is categorized as voltage mode control [25–27]. Since
the output voltage is the only variable that is controlled, the dynamics of current can complicate the design of
the voltage mode controller. Consequently, current mode control is applied to overcome this limitation. With
current mode control, it is common to use the inductor current as the feedback. Together with the output
voltage feedback, the control system forms a cascade loop architecture, where the current feedback is the
inner loop and the voltage feedback is the outer loop [28–30], and the compensator for each of the loops can
be designed individually. With this design strategy, both the current and voltage are controlled, and with a
For resonant converters, because the resonant inductor current is alternating, it is convenient to use the
integration of the current as the variable to be controlled. Since the integration of current over time is the
amount of electric charge, this control scheme is named charge mode control [31–33]. In this case, the charge
control loop is the inner loop and the output voltage loop is the outer loop. Similar to current mode control,
the charge inner loop also changes the dynamics of the system, and it makes the design of the outer loop
compensator easier to achieve high loop bandwidth. Since charge mode control also benefits high-bandwidth
design but differs from current mode control, it is essential to explore the small-signal dynamic model and
1.2 Objectives
The objectives of this work are to explore new Synchronous Rectification (SR) technologies that further
boost the efficiency of LLC resonant converters and to find the dynamic model of the resonant converter
3
to achieve high-bandwidth design, so that both the efficiency and dynamic performance improvements are
covered. SR is effective for reducing the conduction loss in the rectifier stage, particularly for high power and
high current applications. One of the major challenges of the conventional SR for LLC resonant converters
is to overcome the voltage noise introduced by the parasitic components, for the parasitic effects reduce the
efficiency. In this work, two SR solutions will be introduced. The control of the SR in both cases are based
on large-magnitude voltage signals, rejecting the parasitic effects and improving the efficiency.
Regarding the dynamic performance aspect, this work will look into the drawbacks of existing small-
signal models and propose an enhanced model that unlocks high-bandwidth designs. In particular, this work
develops the model for resonant converters under charge mode control. A systematic methodology will
be introduced. It is applicable to the mainstream resonant topologies, such as LLC and CLLC resonant
converters. The following literature review starts with investigations in existing SR technologies, followed
In current-driven methods, current transformers (CT) are widely applied as the current sensor, and in this
work, it is categorized as the first group of the existing Synchronous Rectification (SR) technologies. The
CT monitors the rectifier current for controlling the SR switches, as illustrated in Fig. 1.2. Since the CT
on the secondary side is in series with the SR switches, the sensed current signal can be directly used as
the SR control signal. The SR switches are on and off when the sensed current is higher or lower than
a threshold [34, 35]. Yet, the main challenge of CT is the space occupied, which causes constraints for
high-power density design. In addtion, it introduces extra conduction losses. In the research [36], an energy
recovery current-driven SR is proposed. A set of energy recovery winding is integrated with the CT for
recycling the current sensing energy, improving the efficiency of the CT. However, each SR requires one
CT for gate driving; two CTs for center-tapped full-wave rectification are needed, which occupy even more
space. For full-bridge rectification, the transformer structure is even more complicated. The high volume
occupied by the CT is unfavorable for a compact design. A current-driven SR for center-tapped full-wave
rectifiers in [37] requires only one CT. The CTs are integrated into one, and the utilization of the CT is
improved. In the publications [38–40], the application targets toward voltage-doubler and full-bridge rectifier
topology, and the current sensor count for each topology is reduced to one as well. Another idea to reduce the
space occupied by the CTs is to integrate them with the main transformer, as investigated in the literature [41].
4
Gate Driver Circuit
extra conduction
vgs.SR1
losses
HS
n::
iD1
vgs.HS
ir Lr SR
Vin CT
LS irec
vgs.LS vinv Lm
Cr im CT Co Ro vo
iD SR
vCr
vgs.SR2
Figure 1.2: Conventional LLC SR control methods: current-driven method with the CT mounted on
the secondary side for sensing the rectifier current, which occupies space and introduces extra
conduction losses.
Intrinsically, placing the CT in series with the power loop introduces extra losses. The peak rectifier current
of LLC converters is high, and the CT in series with the power loop has to be well-designed to prevent
saturation.
For avoiding high current sensing, the other group of the current-driven methods places the CT in series
with the resonant tank on the primary side [42–45], as illustrated in Fig. 1.3. The current stress of this kind is
less than that of secondary-side CTs for most applications using LLC converters. Only one CT is needed for
sensing the resonant current. However, the zero-current crossing of the resonant current (ir ) is different from
that of the rectifier current (irec ) due to the component of magnetizing current (im ). Therefore, additional
circuits for canceling im are needed. Since the slope of im is proportional to the output voltage (Vout ) when
the transformer is coupled, the output voltage is sensed and converted to cancel im . As a result, as long as
the portion of im is completely subtracted from ir , the remaining portion is in-phase and proportional to the
rectifier current and can be applied to generate the SR gate driving signal.
1.3.2 SR Based on the Drain-Source Voltage During the Turn-on Phase (vds.on based)
Recent research and development of SR technologies has shifted from incorporating current sensors or cur-
rent transformers to current sensor-less methods to reduce the associated size and conduction losses. The
mainstream of current sensor-less methods focuses on the drain-source voltage (vds.on ) sensing method,
where the SR drain-source voltage during the turn-on phase is used as the feedback signal. The primary
benefit compared with the current-driven methods is the removal of the current sensors, which eliminates the
5
vgs.SR1
HS
n::
iD1 SR
vgs.HS
ir Lr
Vin
LS irec
vgs.LS vinv Lm
vCr CT
im Co Ro vo
iD
Cr
SR
Figure 1.3: Conventional LLC SR control methods: current-driven method with the CT located on
the primary side for sensing the resonant current, which exchanges space with increased circuit
complexity.
The first group of this category, which is also widely adopted in the industry, directly senses the voltage
across the SR MOSFETs during the turn-on phase (vds.on ) without any compensation [46–48]. Due to the
small turn-on resistance of the SR MOSFETs, the sensed voltage is minimal. As a result, high accuracy,
small offset, and fast response comparator design is needed. Besides, the small vds.on is easily affected by
the voltage drops on the parasitic components. As illustrated in Fig. 1.4, the sensed voltage signal contains
the voltage across the leakage inductance (Llk ). The inductive voltage offsets the sensed vds.on , which causes
The leakage inductance exists in both the signal sensing loop and the MOSFET package. The leakage
inductance in the signal sensing loop can be reduced by taking extra care in the PCB (printed circuit board)
layout design, such as placing the SR controller as close to the SR switch as possible, which creates PCB
layout constraints. The leakage inductance in the MOSFET package, however, is still significant [49]. Pre-
mature turn-off of the SR leads to on-time error; a larger portion of the rectifier current is conducted by the
MOSFET body diode, which sacrifices the power conversion efficiency and increases the device temperature.
Therefore, system design effort is spent on reducing the parasitic effects. For instance, the voltage sensing
circuitry must be placed as close as possible to the SR switch to create a Kelvin connection; this introduces
design constraints.
The second group of the vds.on sensing method is to add an Resistor-Capacitor (RC) compensation net-
work in parallel with the MOSFET to compensate for the inductive voltage offset and further reduce the SR
on-time error. In the literature [50], a resettable RC circuit is introduced to compensate for the voltage offset
caused by the leakage inductance. In order to achieve good performance, the time constant of the RC circuit
6
vds.on sensing &
vgs.SR1
L lk compensation
vds.on vds.on is small and noisy;
HS Cds
i
n:: D1 vlk vulnerable to parasitic effects
vgs.HS
ir Lr SR Llk
Vin
LS irec vgs.SR1
vgs.LS vinv Lm
Cr vgs.SR2
im Co Ro vo
iD SR Llk
vCr
vlk
Cds
vds.on
vds.on sensing &
L lk compensation vgs.SR2
Figure 1.4: Conventional LLC SR control methods: vds.on sensing method. vds.on is the voltage across
the SR switches in the on-state for controlling the SR. The magnitude is small and can be easily
affected by the parasitic components, such as the leak inductance (Llk ) and the SR output capaci-
tance (Cds ).
must be carefully selected to match the impedance ratio of the leakage inductance and the MOSFET on-state
resistance. However, the leakage inductance varies with the MOSFET package and is usually not a confined
specification in datasheets. Besides, the MOSFET on-state resistance varies with manufacturing tolerances
and drifts with temperature. Therefore, it is challenging to match the impedance ratio. Another RC compen-
sation circuit to tackle the inductive voltage offset issue is proposed in [51, 52]. Compared with the prior
research, the compensation circuit is simplified, and yet accurate impedance matching is still required.
Considering the uncertainty of the leakage inductance and the associate voltage offset, vds.on based adaptive
SR has been developed to reduce the SR on-time error [53–56], and this is categorized as the third group
in the vds.on based methods. Adaptive SR is achieved by adjusting the SR on-time according to the time
difference between the rectifier conduction time and the SR on-time from the previous switching cycle. The
adaptive algorithm is purely based on time; first, the on-time error is detected, and then the SR conduction
time is changed by a time step in the next cycle. The adaptive SR is initially for adjusting the SR turn-off
instant; this concept is also used for reducing the SR turn-on error in the follow-up research [55, 57]. Some
SR methods deliberately leave a dead-time period for safe operation margin to prevent reverse current or
even cross-conduction of the SR switches [58–60]. The objective is to track the rectifier current conduction
time while maintaining the safe margin; however, following the adaptive algorithm, the safe margin can be
smaller in transient response, particularly when frequency increases. Other research in this field integrates
7
vds1
vds.on1
tcnd1[k-1] tcnd1[k]
vth.rev
vth.off
vth.on
iD1
iD2
vgs.SR1
tSR1[k-1] tSR1[k]
t0 t1 t2 t3 t4 t5 t6 t
Figure 1.5: Typical waveforms of the vds.on based adaptive SR tracking the rectifier current conduction
time by comparing vds.on1 with the internal voltage thresholds.
the adaptive SR method with the primary closed-loop controller [61–63]. Nevertheless, as the adaptive SR
methods still rely on measuring vds.on , improving the noise immunity remains a challenge. More details are
explained as follows.
Fundamentally, vds.on based SR contains a high-precision comparator that compares vds.on with two
threshold voltages (vth.on and vth.o f f ) to determine the SR on-time. As in the typical operating waveforms
shown in Fig. 1.5, the drain-source voltage of SR1 , vds1 , is illustrated to give a holistic view. The SR1 drain-
source voltage during the turn-on phase, vds.on1 , controls SR1 . The SR is on when vds.on1 is lower than vth.on
and off when it is higher than vth.o f f . Note that vds.on1 is not a perfect replica of the inverse of the current
waveform due to the voltage drops on the leakage inductance in the package of the switch as well as in the
voltage sensing loop. This leads to the premature turn-off of SR, as the time difference t1 to t2 shows in
Fig. 1.5.
For an adaptive SR, generally, there is a third threshold level (vth.rev ) to detect the time when the body
diode is reverse-biased, which is used as a reference for the next switching cycle to adjust the SR on-time. For
example, in Fig. 1.5, by comparing vds1 with vth.o f f and vth.rev , the time difference from t1 to t2 is detected.
If the time difference in the k − 1th cycle is sufficiently large, the controller increases tSR1 in the kth cycle
(from t4 to t5 ) to compensate for the time difference. Conversely, if the time difference is small, the controller
Although the adaptive SR effectively compensates for the on-time loss caused by the leakage inductance,
there are several challenges in practical applications. The first challenge is filtering the sub-resonance caused
by the parasitic components during the discontinuous conduction intervals. Figure 1.6 illustrates the typical
waveforms for below-resonance operation considering the effect of leakage inductance and parasitic capaci-
8
vds1
vds.on1
vth.rev
vth.off
vth.on
iD1
iD2
vgs.SR1
(ideal)
(non-ideal)
t0 t1 t2 t3 t4 t5 t6 t7 t
Figure 1.6: Potential challenges of the vds.on based adaptive SR: early turn-on in the below-resonance
region due to voltage ringing-induced SR body diode conduction.
vds1
vds.on1
vth.rev
vth.off
vth.on
iD1
iD2
vgs.SR1
(ideal)
(non-ideal)
t0 t1 t2 t3 t4 t5 t6 t7 t
Figure 1.7: Potential challenges of the vds.on based adaptive SR: premature turn-off in the above-
resonance region, due to drastic change in the current slope and voltage across the leakage in-
ductance.
tance (as Llk and Cds shown in Fig. 1.4). During the discontinuous conduction interval, the lumped parasitic
capacitors and the equivalent inductor Lr ∥ Lm cause the sub-resonance, the amplitude of which varies with
the initial voltage and current. As the initial current in the equivalent inductor is load-dependent, the voltage
amplitude can be larger in heavy loads, forcing the body diode to conduct.
The waveforms in Fig. 1.6 illustrate a scenario where the amplitude is large enough to force the body
diode of the SR MOSFET to conduct. Note that from t0 to t1 , when vds1 is lower than the turn-on threshold
(vth.on ), it meets the criteria to turn on the SR, which consequently causes early turn-on. Turning on the SR
MOSFET during this interval forces current to flow from the output to the input side; this reverse current
then circulates between the input and the output, which reduces the system efficiency. A common approach
to avoid mis-triggering is to set a blanking time that neglects any turn-on triggering. And yet, the parasitic
component resonant frequency depends on the system parameters, which makes the blanking time difficult
to determine.
9
The second challenge is premature turn-off due to the voltage drop on the leakage inductance in the
above-resonance region. As shown in Fig. 1.7, vds1 increases instantaneously at t2 , which is caused by the
drastic change in the current slope, as well as in the voltage across the leakage inductance, known as Llk didtD1 .
This voltage increase can reach the vth.rev level, causing false detection of the body diode conduction time
Since the vds.on method is sensitive to noise and parasitic effects, other SR control methods based on high-
magnitude voltage signals have been developed to avoid noise issues. These methods are categorized in the
alternative SR category; they require neither lossy current sensors nor sensitive voltage detection. A fixed
on-time approach for approximating the SR on-time is proposed in [64], while the SR on-time is estimated
using the output voltage in [65]. Nevertheless, both of methods apply only to the below-resonance operation,
which is not suitable for battery charging applications where the above-resonance operation exists. To cover
both below- and above-resonance operations, a method using the polarity of the inverter and the rectifier
voltage is introduced in [66], which is also based on high-magnitude voltage signals. However, the SR
on-time error increases with the switching frequency in the above-resonance region. For battery chargers,
this region corresponds to the state of charge where the battery is charged with high current at low voltage.
Another publication [67] proposes a dual edge tracking control method for the SR. This method also utilizes
large voltage signals to determine the SR conduction time. Yet, the SR on-time tracking method is similar
to conventional adaptive SR, creating opportunities for further improvement. Recent research on SR for
resonant converters has focused on computing the SR on-time using time-domain models [68–70]. Since SR
on-time solutions depend on the system parameters, the challenges are to overcome the parameter deviation
Table 1.1 summarizes the existing LLC SR control technologies in terms of the amount of current sensors,
noise tolerance, operating range, circuit complexity, and system parameter dependency. The current-driven
methods employ CTs as the current sensor introduces more power losses than the others, especially for CTs
mounted on the secondary side. As opposed to the current-driven method, vds.on sensing methods do not
require extra sensing components, and the induced extra power losses are minimized. Yet, the parasitic
effects have a significant impact on the conventional vds.on sensing method, causing premature turn-off of the
SR and affecting the operating range. With the SR on-time compensation circuit, due to the improvement in
noise tolerance, the on-time error can be reduced. However, it increases the circuit complexity and the SR
10
on-time variation range depends on the system parameters. The alternative methods attempt to avoid using
current sensors and the sensitive vds.on signal, and yet most of them are applicable only to a narrow operating
range or are highly dependent on the system parameters. Overall, the adaptive vds.on based method has good
characteristics from the listed aspects; however, there remain some challenges that will be discussed in the
Chapter 2 and 3.
Charge control for hard-switching topologies along with the small-signal model was introduced in [31–33].
In contrast with the well-known ripple-based control methods, particularly peak current mode control [71–
73], charge control first integrates the switching current and then compares the resulting charge signal with
the command to determine the duty cycle, switching frequency, or other control signals cycle-by-cycle. The
integrated current information is particularly useful for non-monotonic current that does not necessarily
peak at the end of switching. As this behavior is commonly seen in resonant topologies, charge control has
been applied to zero-voltage-switching multi-resonant forward converters [74] and was later developed for
Originally, charge controllers contain an integrator to integrate the sensed resonant current in every half
switching cycle [75, 76]. Recent publications studying LLC resonant converters utilize the resonant capacitor
resonant converter is shown in Figure 1.8 (a). The resonant converter under charge control has an inner
charger loop and an outer voltage loop. In the charge loop, the command vch is the result of the voltage loop
controller, and the feedback signal V1 is the resonant capacitor voltage. To determine the primary gate driving
signals, Bang-Bang Charge Control (BBCC) proposed in [77] compares the resonant capacitor voltage with
11
vch Conventional ws Converter vo
Charge Model Model
irec
equivalent feedback: irec
vHS Cr1 Cr2 secondary current
L1 i t n:1 i2 overlooks the
Cf Ro effect of im
Vg vLS Lm im L2 vo
i1 (b)
Cr1
Cr2 conventional
V1 V2 measured
vo ws 100 ws 10 ws
ws Charge Loop Voltage Loop vch
prediction error
V1 vch votlage loop
Latch vsaw
controller Vref
vo
vch
prediction
error
(a) (c)
the command directly, whereas in Hybrid-Hysteretic Control (HHC) proposed in [78], a ramp signal (vsaw )
served as the slope compensation, is injected to the feedback or the command as illustrated in Fig. 1.8 (a).
Modeling charge-controlled resonant converters requires modeling the power stage. Due to the nature
of large resonant current and voltage swing around the operating point, the linear approximation technique
for hard-switching topologies is not applicable [80]. To tackle small-signal modeling of resonant converters,
numerous methods have been researched [81–87]; these include sampled-data modeling methods, phasor
transformation, Extended Describing Functions (EDF), and averaged modeling methods, to name a few.
Among these methods, the EDF model using the first harmonic approximation is able to predict the fre-
quency response up to the Nyquist frequency, considering a reasonable amount of deviation from the reso-
The EDF method has been widely used due to its systematic approach and relatively low computational
complexity. Modern resonant topologies such as LCC, LLC [89–92], and bidirectional resonant convert-
ers [93–96] have applied the EDF method for modeling. One drawback of the EDF method is the large size
of the derived state-space matrix, due to splitting the harmonic approximated signals into sine and cosine
12
components. Yet, thanks to the systematic procedure, the computation can be quickly executed, and the
The analytical model for charge-controlled resonant converters, however, has been rarely discussed. In
the literature [76, 78], the behavior of charge control is described based on circuit simulation results, while
the full analytical model is absent. Bang-bang Charge Control (BBCC) proposed by [77] employs an input-
output energy balance concept to derive the model, which results in a single-pole system that oversimpli-
fies the loop response. Follow-up studies modeling Hybrid-Hysteretic Control (HHC) used the same ap-
proach [97, 98]. As shown in Fig. 1.8 (b), the conventional model derived by input-output energy balance
leads to an equivalent feedback loop from the AC component of the rectifier current (irec ). Nevertheless, this
path overlooks the dynamics of the resonant tank, particularly the effect of magnetizing current (im ) that con-
tributes to the charge accumulated in the primary side. This omission causes estimation errors in forecasting
Charge control is known as a type of current control that simplifies the system dynamics. For a cascade
control design, a well-compensated inner charge or current loop leads to a system with two real poles in
the frequency range of interest [28–30]. The compensation for the outer loop is thus simplified, and a high-
bandwidth design can be easily achieved. The charge or current control loop is also beneficial to multiphase
or paralleled converters that require current balancing [99–101]. Other advantages include fast overcurrent
comprehensive analysis with a particular focus on conduction loss reduction using novel Synchronous Recti-
fication (SR) technologies and achieving high-bandwidth design using an enhanced small signal model. Two
novel SR methods, namely Resonant Capacitor Voltage based SR (RCV SR) and Volt-Second Product based
SR (VSP SR) for LLC resonant converters, are proposed in this work. Both are based on large-magnitude
voltage signals that reject the parasitic effects. Thus, the SR on-time errors due to parasitic components are
reduced, improving the efficiency. The RCV SR covers all the operating modes of LLC resonant convert-
ers, whereas the VSP SR offers a low component count solution, with a focus on the efficiency improve-
ments in the most lossy operating conditions. On the other hand, regarding the improvement of dynamic
performance, the enhanced small-signal modeling is developed for a dual loop, charge-controlled resonant
converter. Compared with the conventional model, this enhanced model successfully predicts the frequency
13
response in both low-frequency and high-frequency ranges, which enables a high-bandwidth design. The
developed methodology applies to both LLC and CLLC resonant converters. Detailed contributions for each
1. Resonant Capacitor Voltage based SR (RCV SR): In contrast to traditional current-driven methods,
the RCV SR controls the SR on-time without any current sensing components. The bulky and lossy current
transformer (CT) can be removed, simplifying the converter design and eliminating related losses. Since
the RCV SR uses integrations of large-magnitude voltage signals, it is insensitive to noise on parasitic com-
ponents. Compared with conventional vds.on sensing methods, which rely on small-magnitude drain-source
voltage during the on-state, the RCV SR significantly reduces the on-time error, improving the efficiency.
As opposed to other alternative methods such as the fixed on-time approximation, the RCV SR driving strat-
egy can be applied to multiple operating modes regardless of load and frequency variations. This work will
present the detailed operating sequence and prove that the proposed RCV SR applies to both below- and
above-resonance regions and covers heavy-load to light-load conditions, improving the efficiency perfor-
mance.
2. Volt-Second Product based SR (VSP SR): With the focus on industrial battery charging applications,
this work proposes another noise-tolerant SR method to improve the efficiency performance. The SR on-
time is determined based on the product of SR drain-source blocking voltage and rectifier current conduction
time. This Volt-Second Product based SR (VSP SR) employs the high-magnitude voltage signal instead of
the low-magnitude, noise-sensitive vds.on and so is more tolerant of the parasitic effects that cause SR on-
time error, improving the efficiency. The simplicity in the voltage sensing also demands fewer PCB (printed
circuit board) layout constraints. In steady-state operation, the VSP SR tracks the rectifier conduction time
following the drain-source blocking voltage; in transient operation, the volt-second product modulates the SR
on-time and maintains the safe operation margin, preventing reverse current and potential cross-conduction.
Moreover, the VSP SR can be implemented using a microcontroller with only a pair of voltage dividers and
3. Enhanced Small-Signal Modeling for Charge-Controlled Resonant Converters: With the focus
on dynamic performance improvement, this part of the work introduces a small-signal modeling method-
ology for charge-controlled resonant converters to enable high-bandwidth design. The systematic approach
is applicable to both LLC and CLLC resonant converters. The proposed model corrects the conventional
charge mode control model and reveals that charge control forms an equivalent feedback path from the sine
component of the resonant current. This equivalent feedback loop directly corresponds to the actual feedback
14
path of the charge loop, which is not shown in the conventional model, and hence it provides comprehensive
and enhanced prediction of the system frequency response. In particular, the proposed model corresponds to
the actual frequency response across the frequency range of interest, whereas the conventional model starts
to deviate from the mid-frequency range. Due to the improvement in the mid- to high-frequency range, the
first; a half-bridge, center-tapped LLC resonant converter is used as an example. It is followed by the
introduction of the proposed Resonant Capacitor Voltage based SR (RCV SR), where the operating principles
for below- and above-resonance regions are explained in detail, as well as the consideration in light-load
conditions. In addition, the sensitivities regarding system parameter deviation, especially the resonant tank
parameters, are analyzed and discussed. At the end of this chapter, the efficiency performance of RCV SR is
In Chapter 3, the LLC resonant converter is first analyzed using time-domain equations, which establishes
the fundamentals of the proposed Volt-Second Product based SR (VSP SR). Next, the operating principles of
VSP SR are introduced, followed by the transient analysis. Details of the implementation using a microcon-
troller are given. This chapter finishes with the efficiency performance comparison with an industrial vds.on
Chapter 4 discusses the small-signal modeling of resonant converters. It starts with the modeling of
an open-loop, five-order CLLC resonant converter as an example and continues with a brief review of the
charge mode control and the conventional modeling method. Next, the proposed modeling methodology is
explained; the developed high-order CLLC resonant converter model can be directly reduced and applied to
LLC resonant converters. Finally, the proposed model is compared with the conventional model, with both
simulation and experimental results. A closed-loop design example using the proposed model is also given
to demonstrate that high loop bandwidth can be achieved with a simple compensator.
Chapter 5 concludes the contributions of this work and suggests potential research topics for the future.
This work contributes to the field of design of high-performance resonant converters for battery chargers in
terms of the improvement in efficiency and dynamics. The research outcomes are published in three IEEE
Transaction journals.
15
Chapter 2
The essence of synchronous rectification (SR) drive is to synchronize the SR driving signal with the rectifier
current. Ideally, the SR MOSFET is on when the rectifier current starts conduction and off when the rectifier
current stops. When the SR MOSFET is on, it is equivalent to a resistor (Rds.on ); intrinsically, the voltage
drop on Rds.on is much less than the body diode forward voltage drop. Therefore, to minimize the conduction
loss is to maximize the SR MOSFET turn-on time. Late turn-on or early turn-off result in longer body-
diode conduction time, which sacrifices the efficiency performance. Early turn-on or late turn-off lead to
reverse current or even catastrophic shoot-through, which should be avoided. Since the SR drive relates to
the rectifier current, it is worthwhile to first define the operating modes of LLC resonant converters. In this
chapter, a half-bridge, center-tapped LLC resonant converter is studied; the schematic diagram is shown in
Fig. 2.1.
the switching frequency, the operating regions can be divided into below-resonance and above-resonance
region, corresponding to the frequencies below and above the resonant frequency, respectively. Depending
on the voltage applied to the resonant tank and the transformer, each switching cycle can be divided into
four subsequences or operating modes in heavy-load conditions, despite the frequency region. In light-load
conditions, the subsequences are extended to six. This section begins with the heavy-load conditions.
In the below-resonance region, the key waveforms and the equivalent circuit of the resonant tank in each
16
vgs.SR1
HS
n::
iD1 SR
vgs.HS
ir Lr
Vin
LS irec
vgs.LS vinv Lm
Cr im Co Ro vo
iD
vCr
SR
vgs.SR2
Figure 2.1: Schematic diagram of a half-bridge center-tapped LLC resonant converter with syn-
chronous rectification.
subsequence are illustrated in Fig. 2.2 (a) and (b). In Mode 1, from t0 to t1 , the transformer is decoupled
because vLm is smaller than the reflected output voltage nVo . In this mode, since the transformer is decoupled,
the resonant current and the magnetizing current must be equal. In this mode, Lm and Lr connected in series
starts to resonate with the resonant capacitor (Cr ). Since no current is transferred to the secondary side, the
Mode 2 starts at t1 and ends at t2 when the transformer is once again decoupled. At t1 , when the inverter
voltage (vinv ) switches to Vin , as long as vLm is greater than nVo , the secondary rectifier is forward biased and
the transformer is coupled, clamping the vLm at nVo . As a result, Lr and Cr starts resonate. The difference
between the resonant current ir and the magnetizing current im is transferred to the secondary side as irec .
Mode 2 ends at t2 when ir decreases and reaches im . At this moment, the rectifiers are reverse biased and the
transformer is decoupled, which completes the first half of the switching cycle.
The general form of the differential equation describing this circuit based on KVL is
dir dim
vinv (t) = vCr + Lr + Lm . (2.1)
dt dt
dvCr
ir = im = Cr , (2.2)
dt
and
dir
0 = vCr + (Lr + Lm ) . (2.3)
dt
17
vinv vLr ir vLr ir irec n
Lr Lm Vin Lr Lm
Vp2
vCr vLr ir vLr ir irec n
Vp1 V in
Vin Lr Lm Lr Lm
vLm vLm nVo
irec Cr Cr
im im
vCr vCr
t0 t1 t2 t3 t4 t Mode 3: t2 ~ t3 Mode 4: t3 ~ t4
(a) (b)
Figure 2.2: a) The key waveforms and b) the corresponding equivalent models of the resonant tank in
the below-resonance region.
During this period, only Lr resonates with Cr . The other half cycle is composed of Mode 3 and 4, which are
similar to Mode 1 and 2, having the same KCL equations as (2.2) and (2.4), respectively. The differences are
the voltage across the resonant tank and the polarity of the voltage across Lm . In Mode 3,
dir
Vin = vCr + (Lr + Lm ) , (2.6)
dt
whereas in Mode 4,
dir
0 = vCr + Lr − nVo . (2.7)
dt
In the above-resonance region, the key waveforms and the equivalent circuit of the resonant tank in each
subsequence and are illustrated in Fig. 2.3 (a) and (b). Unlike the below-resonance region, the transformer
is continuously coupled among the four modes in heavy-load conditions; the voltage across the magnetizing
Mode 1 starts at t0 when vinv switches to Vin and ends at t1 when vLm changes its polarity from negative to
18
vLr ir irec n vLr ir irec n
vinv
Vin Lr Lm Vin Lr Lm
vCr vCr
t0 t1 t2 t3 t4 t Mode 3: t2 ~ t3 Mode 4: t3 ~ t4
(a) (b)
Figure 2.3: a) The key waveforms and b) the corresponding equivalent models of the resonant tank in
the above-resonance region.
positive nVo . Mode 2 is between t1 and t2 , where the behavior is the same as Mode 2 in the below-resonance
region. Mode 3 and 4 in the second-half cycle is similar to Mode 1 and 2, respectively; the differences are
vinv and the inversed voltage across Lm . It is worthwhile to notice that in Mode 1 and 3, the slope of ir is
steeper than the prior mode. It is because the voltage applied to the resonant tank has the maximum absolute
The KVL equation in the above-resonance region is the same as (2.1), with
dim nVo
Mode 2, 3
Lm = (2.8)
dt
−nVo
Mode 1, 4.
The KCL equation for all the four modes is the same as (2.4), since the transformer is continuously coupled.
In the light-load conditions, there are two more subsequences with transformer decoupled due to the
nature of decreased output current. In the below-resonance region, as shown in Fig. 2.4, the transformer is
not coupled in Mode 2 when vinv switches to Vin at t1 . At this moment, the resonant voltage vCr is close to
Vin /2, which is relatively high in the first half cycle. Consequently, vLm is relatively small. If vLm is less than
nVo , the rectifier diodes are reverse biased, prohibiting the transformer from coupling. Mode 1 ends at t2
19
vLr ir vLr ir
Lr Lm Vin Lr Lm
vCr vCr
Mode 1: t0 ~ t1 Mode 2: t1 ~ t2
ir
im vLr ir irec n vLr ir
Vin Lr Lm Vin Lr Lm
(a) (b)
Figure 2.4: Light-load conditions: a) the key waveforms and b) the corresponding equivalent models
of the resonant tank in the below-resonance region.
where
Lr
Vp1 = Vin − nVo (1 + ). (2.10)
Lm
Similarly, in Mode 5, after vinv switches to zero, the transformer keeps decoupled until the amplitude of vCr
is large enough to forward bias the rectifier diodes. The criteria to end Mode 5 is
where
Lr
Vp2 = nVo (1 + ). (2.12)
Lm
In the above-resonance region, there are also two more subsequences with the transformer decoupled in
light-load conditions. As shown in Fig. 2.5, Mode 2 starts at t1 when ir reaches im and the transformer is
decoupled. At this moment, the transformer is not coupled as in the heavy-load condition, which is also
because vCr is larger than Vp1 . It results in insufficient voltage across the transformer to forward bias the
20
vLr ir irec n vLr ir
Vin Lr Lm Vin Lr Lm
vCr vCr
ir Mode 1: t0 ~ t1 Mode 2: t1 ~ t2
im
vLr ir irec n vLr ir irec n
Vin Lr Lm Lr Lm
vCr vCr
Mode 5: t4 ~ t5 Mode 6: t5 ~ t6
(a) (b)
Figure 2.5: Light-load conditions: a) the key waveforms and b) the corresponding equivalent models
of the resonant tank in the above-resonance region.
rectifier diodes. Similarly, in Mode 5, the transformer keeps decoupled until vCr is larger than Vp2 . It is
worthwhile to notice that when the load further decreases, ir can reach im before vinv switches. In this case,
the transformer is decoupled in both Mode 1 and 4, which makes the sequence the same as the below-
resonance condition.
In summary, for a frequency-modulated LLC resonant converter, the operating modes for heavy-load and
light-load conditions are discussed. The modes are divided by the voltages applied to the resonant tank and
the transformer. When the transformer is coupled, ±nVo is applied to the magnetizing inductor, and only
Lr resonates with Cr . When the transformer is decoupled, Lr and Lm are equivalently in series and resonate
with Cr . These operating modes affects the conduction angle of the rectifier diode, so they are essential to
the design of the synchronous rectification, which will be discussed in the following section.
tification (RCV SR) algorithm. The RCV SR driving strategy is simple and lossless while covering a wide
21
vgs.SR1
HS
n::
iD1
vgs.HS
ir Lr SR
Vin
LS irec
vgs.LS vinv Lm
Cr
im Co Ro vo
iD SR
vCr
vgs.SR2
irec
Proposed RCV Strategy
vCr voltage
Integration/ integrations
vCr
Comparison
Vin low pass filter vgs.SR1
vgs.SR2
voltage vp.int
Vout
integrations
vCr .int
Simple and lossless
Large and stable signals involved; vgs.SR1
high noise immunity
Wide operating range covered
t
Figure 2.6: The proposed RCV strategy senses the large voltage signals ➊ and controls the SR by
simply integrating and comparing the voltage signals ➋. The typical waveforms show the below-
resonance operation for SR1 as an example to visualize the concept of the proposed RCV strategy,
where v p1.int and vCr .int are the voltage integrations of linear combinations of the sensed voltages,
defined in the following section.
operating range. It uses the integration of large voltage signals, which is noise-tolerant compared with con-
ventional vds.on based methods. Also, current sensors are not needed, which saves the associated conduction
losses and cost comparing with conventional current-transformer (CT) based methods. To apply the RCV
strategy, the inverter of the LLC converter can be half-bridge or full-bridge with symmetric duty ratio, and
the rectifier can be center-tapped or full-bridge. Among those combinations, the topology shown in Fig. 2.6
is considered as an example. Figure 2.6 also depicts the simplified block diagram of the proposed RCV
SR control method, which involves three large voltage signals: the input voltage (Vin ), the output voltage
(Vout ), and primarily, the resonant capacitor voltage (vCr ). By comparing the integrations of the large voltage
22
2.2.1 Below-Resonance Region
Following the KVL equation (2.1), in Mode 2 and 3, the inverter voltage vinv is a constant Vin . Integrating
Z t3 Z t3 Z t2 Z t2 Z t3 Z t3
Vin = vCr dt + Lr dir + Lm dim + Lr dir + Lm dim . (2.13)
t1 t1 t1 t1 t2 t2
Since the difference between ir (t1 ) and ir (t2 ) is equal to the difference of im (t1 ) and im (t2 ), and the magne-
Z t3 Z t3
Lr
Vin dt = vCr dt + nVo (1 + ) · (t2 − t1 ) + (Lr + Lm ) · [ir (t3 ) − ir (t2 )]. (2.14)
t1 t1 Lm
The difference between ir (t3 ) and ir (t2 ) is not necessarily zero and cannot be ignored. Furthermore, in order
to avoid current sensing, this term must be replaced by voltage signals. In Mode 3, as (2.6) indicates, Lr
and Lm in series resonate with Cr . By integrating both sides of (2.6), the term for the change in the resonant
Z t3
(Lr + Lm ) · [ir (t3 ) − ir (t2 )] = Vin · (t3 − t2 ) − vCr dt. (2.15)
t2
Therefore, (2.14) can be fully expressed as integrals of voltage signals by replacing the current terms using
By applying the same procedure to the subsequence Mode 3 and 4, the RCV strategy for the other SR
As a result, the RCV strategy to drive the SR MOSFET SR1 and SR2 in the below-resonance region is
expressed by (2.16) and (2.17), respectively. The time to turn off SR1 is at t2 when the integrations of the DC
voltage Vp1 and the resonant capacitor voltage vCr are equal. As shown in Fig. 2.7, v p1.int and vCr .int represent
the time integrals of Vp1 and vCr , respectively; the two voltage integrations starts at t1 and cross over at t2 ,
which corresponds to the instant when the magnetizing current reaches the resonant current and irec reaches
23
vinv
ir
im
vCr
V in
irec
vp1.int
vCr.int
vp2.int
vCr.int
vgs.SR1
vgs.SR2
t0 t1 t2 t3 t4 t
Figure 2.7: The key switching sequence of the proposed RCV SR method in the below-resonance re-
gion, heavy-load conditions.
zero. This is exactly the time for the gate drive signal vgs.SR1 to be terminated. Similarly, in the second half
of the cycle, vCr .int and the integration of Vp2 (denoted by v p2.int ), start at t3 and intersect at t4 . This is exactly
the time when irec reaches zero, which is the instant to switch off the vgs.SR2 .
Note that the DC voltage and the resonant capacitor voltage in (2.16) and (2.17) are large voltage signals.
Therefore, the RCV strategy is more robust than traditional vds.on sensing methods which rely on small and
sensitive signals. In addition, the action of integration also filters out high frequency noise, which further
In heavy-load conditions, the turn-on of SR1 and the start of the integration are both synchronous with
the rising edge of the inverter voltage at t1 , defined by vinv > Vin /2; whereas the corresponding instant for
SR2 is synchronous with the falling edge of the inverter voltage at t3 , defined by vinv < Vin /2. In terms of
light-load conditions, there is another constraint that needs to be considered. However, it will be proved later
that by using the resonant capacitor voltage, the turn-on triggering for different loading conditions can be
24
2.2.2 Above-Resonance Region
In the above-resonance region, although the operating sequence is different from the below-resonance, the
SR conduction time can be found by comparing the integrated voltage signals. In Mode 1, the vinv is Vin and
vLm is −nVo ; integrating the KVL equation (2.1) in this time interval yields
Z t1 Z t1 Z t1
Vin dt = vCr dt − nVo dt + Lr [ir (t1 ) − ir (t0 )], (2.18)
t0 t0 t0
Z t3 Z t3
0= vCr dt + nVo dt + Lr [ir (t3 ) − ir (t2 )]. (2.19)
t2 t2
If the turns ratio of the secondary windings of the center-tapped transformer is unity, and the duty ratio is
symmetric, then the magnetizing and demagnetizing of Lm is also symmetric. It leads to the fact that the sum
of ir (t0 ) and ir (t2 ), and the sum of ir (t1 ) and ir (t3 ) are zero, and that the time intervals from t0 to t1 and from
t2 to t3 are equal. Therefore, by adding (2.18) and (2.19), the terms for the resonant current and the time
Z t1 Z t1 Z t3
Vin dt = vCr dt + vCr dt. (2.20)
t0 t0 t2
The magnetizing inductor Lm is charged from t1 to t3 , from which there is an important characteristic: the
resonant current and the magnetizing current are the same at the beginning and at the end, and the voltage
cross Lm is nVo . The voltage integrations during this interval are therefore represented as
Z t2 Z t3
Lr
Vin dt = vCr dt + nVo (1 + )(t3 − t1 ). (2.21)
t1 t1 Lm
As a result, for the operation above resonance, the RCV strategy for SR1 is obtained by adding (2.20) and
According to (2.22), the integral of Vp1 − vCr over the time interval t0 to t2 (v p.int1 ) is equal to twice the
integral of vCr from t2 to t3 (vCr .int1 ); by comparing the two integrals, the instant t3 to switch off SR1 can be
obtained. Note that t0 and t2 are synchronous with the inverter voltage and hence can be obtained. As shown
in Fig. 2.8, the first integral, v p.int1 , starts at t0 and stops at t2 ; the integration result is retained and compared
25
with the second integral, vCr .int1 . The SR driving signal vgs.SR1 is terminated at t3 , the instant when the two
integrals are equal. This moment exactly corresponds to the time when irec is zero.
By applying the same analysis on the second half of the cycle, the voltage integrations to determing the
Z t4 Z t5
[vCr −Vp2 ]dt = 2 [Vin − vCr ]dt. (2.23)
t2 t4
As (2.23) indicates, the integral of vCr − Vp2 over the time interval t2 to t4 (v p.int2 ), is equal to twice the
integral of Vin − vCr from t4 to t5 (vCr .int2 ). Hence, the turn-off instant for the present SR switching cycle is
at t5 where the two integrals are equal; this is also the moment to turn on the other SR MOSFET, SR1 . As
a result, the RCV strategy in the above-resonance region is described by (2.22) and (2.23). Similar to the
below-resonance region, the RCV strategy is in the form of large voltage integrations, and it is dominated
by the resonant capacitor voltage. By applying the integrations of large voltage signals, the RCV strategy
is more reliable than traditional vds.on sensing methods in terms of noise immunity. In addition, considering
that the turns ratio of the secondary windings of the center-tapped transformer is unity, (2.22) is equivalent
to
Z t2 Z t1
[Vp1 − vCr ]dt = 2 [Vin − vCr ]dt, (2.24)
t0 t0
In light-load conditions, according to the analysis in Section 2.1, there are two additional modes with trans-
former decoupled in both below- and above-resonance regions (Mode 2 and 5). After these two modes, the
transformer is coupled and the rectifier current (irec ) starts conduction. From the perspective of SR, it means
that the SR turn-on timing is delayed due to the existence of these two modes. Therefore, (2.9) – (2.12) that
describe the conditions for the transformer to be coupled have to be considered as the criteria to turn on the
SR switches. As shown in Fig. 2.9 (a) and (b), vgs.SR1 is on when vCr falls below Vp1 ; vgs.SR2 is on when vCr
rises above Vp2 . These are the instants when the transformer is coupled and irec starts conduction.
Regarding the turn-off instant of SR, it is also determined by comparing the same voltage integration
signals because the operating modes in the time intervals when the transformer is coupled are identical. As
26
vinv
ir
im
vCr
V in
irec
vp.int1
vCr.int1
vp.int2
vCr.int2
vgs.SR1
vgs.SR2
t0 t1 t2 t3 t4 t5 t
Figure 2.8: The key switching sequence of the proposed RCV SR method in the above-resonance re-
gion, heavy-load conditions.
shown in Fig. 2.9 (a), for example, the transformer is coupled from t2 to t3 and decoupled from t3 to t4 , the
equivalent circuits of which are the same as Mode 2 and 3 in the heavy-load conditions. Therefore, the same
In the above-resonance region, as illustrated in Fig. 2.9 (b), the criteria for turning on the SR MOSFETs
(2.9) – (2.12) are also valid. Thus, vgs.SR1 and vgs.SR2 are turned on at t2 and t5 , where vCr intersects Vp1 and
Vp2 , respectively. From t2 to t4 and from t5 to t7 , the transformer is coupled, despite the switch of the inverter
Z t3 Z t4
Lr
Vin dt = [vCr + nVo (1 + )]dt, (2.26)
t2 t2 Lm
and
Z t7 Z t7
Lr
Vin dt = [vCr − nVo (1 + )]dt. (2.27)
t6 t5 Lm
Z t3 Z t4
[Vp1 − vCr ]dt = [vCr +Vp2 ]dt, (2.28)
t2 t3
27
vinv vinv
ir ir
im im
Vp2
vCr Vp2
Vp1 V in vCr
Vp1 V in
irec
irec
vp1.int
vCr.int vp.int1
vCr.int1
vp2.int vp.int2
vCr.int vCr.int2
vgs.SR1 vgs.SR1
vgs.SR2 vgs.SR2
(a) (b)
Figure 2.9: The key waveforms in light-load conditions for a) below resonance, and b) above reso-
nance.
and
Z t6 Z t7
[vCr −Vp2 ]dt = [Vin − vCr +Vp2 ]dt. (2.29)
t5 t6
From t3 to t4 , vCr is close to Vp2 due to the small time difference between t3 and t5 , and hence, (2.28) can be
written as
Z t3 Z t4
[Vp1 − vCr ]dt = 2 vCr dt, (2.30)
t2 t3
which is in the same form as (2.22). Similarly, Vin − vCr is close to Vp2 from t6 to t7 , and (2.29) can be
replaced by
Z t6 Z t7
[vCr −Vp2 ]dt = 2 [Vin − vCr ]dt, (2.31)
t5 t6
As the loading current further decreases, if irec reaches zero before t3 in the first half cycle or before
t6 in the second half cycle, the operating modes are equivalent to the below-resonance region, despite the
frequency. In this case, (2.30) and (2.31) remains valid because the right-hand side of the equations are zero
(from the time interval t2 to t3 and t5 to t6 ), and (2.30) and (2.31) are essentially equivalent to (2.16) and
28
(2.17), the SR turn-off criteria for below-resonance operation.
In summary, for light-load operations, the criteria (2.9) and (2.11) indicates the instant to turn on SR and
to start the integration. As for the turn-off of SR, the voltage integrations employed for both heavy-load
and light-load conditions are identical; in the below-resonance region, they are (2.16) and (2.17), and in the
above-resonance region, they are (2.22) and (2.23). Under very light-load conditions, the modes for below-
and above-resonance operation are the same, and so are the SR turn-off criteria.
Although the SR driving strategies for below and above the resonant frequency are derived individually, the
two sets of the control laws are equivalent at the boundary where the switching frequency is at the resonant
frequency. At the boundary, the right-hand side of the above-resonance equations (2.22) and (2.23) are zero,
and on the left-hand side, t0 and t2 are replaced with t1 and t3 . As a result, they become equivalent to the
below-resonance equations (2.16) and (2.17) at the boundary. Therefore, the transition of the two modes at
Considering the light-load condition, the criteria described in (2.9) and (2.11) delay the start of integration
and the turn-on of the SR, but the forms of the voltage integrations for below and above resonance are
not changed as discussed in Section 2.2.3. The reason that the forms are kept the same is that when the
criterion are satisfied, the transformer is coupled, and the equivalent circuits for light-load and heavy-load
conditions are identical. When (2.9) and (2.11) take place prior to the start of integration defined in heavy-
load conditions, the latter becomes the more stringent criterion for defining the SR turn-on instant, and the
former has no effect on the control of SR. Therefore, the transition of light-load and heavy-load conditions
The SR on-time is obtained simply by comparing the voltage integrations in RCV SR, but the closed-form
expression of SR on-time is complicated. In order to give the insight into the impact of SR on-time error
versus the mismatch of the inductance ratio (Lr /Lm ) and simplify the sensitivity analysis procedure, the first
ir = i pk sin(ωt − φ ), (2.32)
29
where the angle φ defines the phase delay of the resonant current referring to the resonant tank voltage, and
i pk is the peak resonant current; both parameters are related to loading conditions. The integral of vCr is the
integration of the input voltage subtracts the total voltage across the resonant and magnetizing inductance,
which yields
Z t
vCr dt = (Vin − nVo )t − Lr i pk [sin(ωt − φ ) + sinφ ]. (2.33)
0
Combining (2.33) with (2.16) and (2.17), the SR on-time can be approximated by solving the equation:
where λ is the inductance ratio, Lr /Lm . Let the expression of the left side of (2.34) be a linear function fc (t)
As illustrated in Fig. 2.10, the geometric representation indicates that the crossover of fc (t) and g(t) corre-
sponds to tc , which is the cut-off of the rectifier current and the turn-off instant of the SR.
Considering that there is a mismatch of the inductance ratio between the actual value and the one set by
the RCV SR controller, the inductance ratio set by the controller is hence defined as
λ ′ = (1 + ε)λ , (2.37)
where ε is the mismatch of the inductance ratio. ε equals zero for no mismatch, and 0.2 for 20 % mismatch
in the inductance ratio. The linear curve with the mismatched inductance ratio is defined as fe (t):
The crossover of fe (t) with g(t) at te is the displaced SR turn-off instant, and the difference between the
actual cut-off of the rectifier current is ∆t, which defines the SR on-time error. The slope determined by the
30
y g(t)
te tc g ’(tc )
-nVo l Dt t
C A
-nVo l tc
-nVo l' Dy fc(t)
-nVo l' te
B
fe(t)
Figure 2.10: The geometric representation of sensitivity analysis using the first harmonic approxima-
tion.
As the slope of fc (t) is relatively low, the approximation is performed by using the distance from B to C
for simplifying the analysis. The slope ∆y/∆t is close to the differential of g(t) at the instant of tc , which is
defined by:
Solving for ∆t using (2.39) and (2.40), the relationship between the mismatch of the inductance ratio (ε) and
−nVotc ε
∆t = (2.41)
i pk Lm ωcos(π − φ ) − nVo ε
Based on the relationship defined in (2.41), the SR on-time error is less sensitive to the mismatch of
the inductance ratio at heavy-load conditions as the peak current (i pk ) is larger and the phase delay (φ ) is
shorter. This characteristic is important to the efficiency performance as the on-time error is more critical at
According to the discussions above, the resultant SR driving strategy using RCV is described as below with
31
on,
for (vinv < Vin /2) & (vCr > Vp2 ) & (vCr .int1 > v p.int1 )
SR2 : (2.43)
off,
for (vCr .int < v p2.int ) ∥ (vCr .int2 > v p.int2 )
By combining the SR turn-on and turn-off criteria with elementary logic operations, the RCV strategy ac-
commodates itself to each operating mode without having to detect the information of loading current and
resonant frequency. Take the turn-on of SR1 for example, at heavy-load conditions, the criterion vCr < Vp1 is
already satisfied prior to the rising edge of vinv , and hence the other two criteria are dominant. In this con-
dition, for above resonance, vCr .int2 is zero and less than v p.int2 at the rising edge of vinv , so vCr .int2 > v p.int2
naturally becomes the most strict criterion and determines the turn-on instant. As a result, the RCV strategy
adapts itself to different switching frequencies and load conditions. Moreover, the RCV SR driving strategy
is developed based on the integration and comparison of large voltage signals. In contrast to traditional vds.on
sensing methods which rely on small and noise-sensitive signals, the RCV strategy is less subject to noise
Although this work focuses on the half-bridge LLC topology, the same procedure can be applied to full-
bridge LLC converters with the consideration of the differences between the two inverter voltages. For the
half-bridge topology, the input is switching between Vin and zero; whereas for the full-bridge topology, the
input is between Vin and −Vin . By following the same procedure for the half-bridge topology, the signal Vp2
defined in (2.12) and the signal vCr .int1 defined in (2.22) are redefined as follows:
Lr
Vp2 = nVo (1 + ) −Vin , (2.44)
Lm
and
Z t3
vCr .int1 = 2 [Vin + vCr ]dt. (2.45)
t2
As a result, the other definitions for the voltage integrations are kept the same, and the RCV SR driving
32
on,
for (vinv < 0) & (vCr > Vp2 ) & (vCr .int1 > v p.int1 )
SR2 : (2.47)
off,
for (vCr .int < v p2.int ) ∥ (vCr .int2 > v p.int2 )
LLC resonant converter based on the conclusion of Section 2.2.6. It is possible to implement the control
strategy directly using (2.42) and (2.43). And yet, the turn-off criteria for below- and above-resonance
According to the vgs.SR1 turn-off criteria (2.16) for the below-resonance operation, moving all the terms
to the left-hand side of the equation results in the same form as the turn-off criteria (2.22). On the right-
hand side of (2.22), the integration starts after the switch of vinv , so as long as it is set as zero before the
integration, (2.22) is the same as (2.16) in the below-resonance region. Therefore, the combined turn-off
criteria is expressed as
Z t 0,
for t0 ≤ t ≤ t2
[Vp1 − vCr ]dt = , (2.48)
t0
tt 2vCr dt, for t > t2
R
2
where t0 indicates the start of integration and t2 refers to the falling edge of vinv .
The implementation is shown in Fig. 2.11. The constant gain K at the Vo input terminal represents
n(1 + Lr /Lm ), and the output of the amplifier is Vp2 as defined in (2.12). The sum of the input voltages is
fed into the corresponding integrator one to four. The first integrator represents the right-hand side of (2.48).
The reset signal of the first integrator, RES, is synchronous with the inverter voltage, vinv . Hence, the output
is zero during the time interval t0 ≤ t ≤ t2 , and the integration of 2vCr starts after t2 expires. The second
resettable integrator implements the left-hand side of (2.48). As shown in Fig. 2.11, the second integrator
integrates Vp1 − vCr , and the output (v p.int1 ) is compared with the output of the first integrator (vCr .int1 ). The
reset signal of the second integrator is determined by the inverse of the intersection of RES and the output of
the comparator A. Thus, the second integrator starts integration when the inverter voltage is high (controlled
by RES), and the criterion Vp1 > vCr for light-load SR turn-on control has to be satisfied before the start of
integration.
As discussed in Section 2.2.3, the rectifier current starts conduction when vCr is lower than Vp1 . This
criterion is met before the inverter voltage is high for heavy-load conditions, and hence the start of the
33
vCr vCr.int1
RES
C
Vin
Vp1 vCr vp.int1 R Q SR1
S
RES S Q
vCr A
B R Q SR2
Vin vCr D
S vCr.int2
RES RES
v inv
second integrator is synchronous with the rising edge of RES. However, for light-load conditions, Vp1 is
larger than vCr after the rising of vinv , and the criterion Vp1 > vCr becomes more strict than RES to start
the integration. Therefore, by taking the intersection of the two criteria, the start of the second integrator
naturally accommodates both of the light-load and heavy-load conditions. The comparator C compares the
output of the first and the second integrator, which connects the left- and right-hand side of (2.48). As a
result, the crossover of v p.int1 and vCr .int1 determines the turn-off of the SR MOSFET SR1 .
Similarly, the turn-off criteria of vgs.SR2 described in (2.17) and (2.31) can be combined, which leads to
Z t 0,
for t2 ≤ t ≤ t4
[vCr −Vp2 ]dt = (2.49)
t2
tt 2[Vin − vCr ]dt, for t > t4
R
4
34
where t2 indicates the start of integration and t4 refers to the rising edge of vinv . As illustrated in Fig. 2.11,
the third integrator implements the left-hand side of (2.49), and the fourth integrator implements the right-
hand side. The reset signal of the fourth integrator is the inverse of RES, denoted as RES. It makes the
output of the fourth integrator zero when the inverter voltage is low, which corresponds to the time interval
t2 ≤ t ≤ t4 . The fourth integrator starts integrating Vin − vCr at t4 , which is controlled by the falling edge of
RES. Regarding the third integrator, the reset signal is the inverse of the intersection of RES and the output
of the comparator B. This logic circuit is symmetrical to the reset signal of the second integrator, and the
purpose is also to control the start of integration in heavy-load and light-load conditions. In the last stage, the
output of the two integrators, vCr .int2 and v p.int2 , are fed to the comparator D, which completes the turn-off
The implementation of the turn-on of SR1 and SR2 follows (2.42) and (2.43), respectively. In order to
prevent the two SR switches from cross-conduction, the inverse of each gate driving signal is fed into the
logic gate that controls the set signal of the other SR switch to prohibit the turn-on triggering as shown in
Fig. 2.11. It generates the dead-time between the two SR switches when the switching frequency is nearby
or above the resonant frequency. As a result, the input voltages are integrated and compared to determine
the turn-on and turn-off of the SR switches. With simple logic combinations, the controller naturally adapts
The block diagram shown in Fig. 2.11 can be implemented by analog, digital, or mixed-signal circuitry.
In this work, the prototype is built using analog circuits, in which the key components are listed in Table 2.1.
The logic gates are general-purpose; conventional summing amplifier and inverting integrator circuits are
applied to the adder and the resettable integrator. The main design considerations for the adders and the
integrators are the bandwidth and the slew rate. The bandwidth of the adder is suggested to be at least ten
times of the maximum operational switching frequency to avoid the phase delay of the sensed vCr . The slew
rate of the operational amplifiers for the integrator one and four are recommended to be higher than the other
two, since the voltage integrations are doubled as indicated by (2.48) and (2.49).
650 W/ 24 V LLC converter for battery charging applications, with the system parameters given in Table 2.2.
The simulation and experiments are performed under the constant current, constant power, and constant volt-
age mode, which are essential operating modes for battery chargers. The simulation results for heavy-load
35
SR Triggering [V] Rectifier Current [A]
fS = 250 kHz fS = 230 kHz fS = 220 kHz fS = 200 kHz
40 Vout = 16.5 V Vout = 19.3 V Vout = 20.8 V Vout = 23.5 V
20
tCOND tCOND tCOND tCOND
0
2.5
tRCV tRCV tRCV tRCV
0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Time [ms] Time [ms] Time [ms] Time [ms]
(a) (b) (c) (d)
Figure 2.12: Simulation results for the constant current mode operation. The converter operates at
27 A, a) 250 kHz, b) 230 kHz, c) 220 kHz, and d) 200 kHz; the RCV-controlled SR on-time
(tRCV ) is very close to the ideal rectifier current conduction time (tCOND ) as shown in each con-
dition.
conditions shown in Fig. 2.12 and Fig. 2.13 are for the constant current mode and the constant power mode
operation respectively. In the constant current mode, the output current is fixed at its maximum value, and
the switching frequency is above the resonance. In this mode, (2.22) and (2.23) determines the turn-off of
SR. In Fig. 2.12, the simulation results indicate that the RCV controls the turn-off instant accurately at each
switching frequency, despite the change in the output voltage. The falling edge of the SR triggering signal
is exactly at the zero crossing of the rectifier current, and the RCV-controlled SR on-time (tRCV ) is close to
the ideal rectifier current conduction time (tCOND ). Note that the difference between tCOND and tRCV at the
beginning of each cycle is caused by the deadtime set to prevent cross conduction of the SR MOSFETs.
In the constant power mode operation, the LLC resonant converter operates in the below-resonance re-
gion, where the output voltage is above the nominal value. The output voltage and current change at the same
time according to the frequency, while the output power is kept at its maximum. In this operating mode, tRCV
Parameter Values
input voltage, Vin 400 V
nominal power, Pout 650 W
nominal output voltage, Vo 24 V
switching frequency, fsw 150–250 kHz
magnetizing inductance, Lm 103.4 µH
resonant inductance, Lr 37.7 µH
resonant capacitor, Cr 18.8 nF
transformer turns ratio, n 8.1
36
SR Triggering [V] Rectifier Current [A]
fS = 180 kHz fS = 175 kHz fS = 173 kHz fS = 167 kHz
40 i iout = 24.4 A iout = 24 A iout = 23 A
out = 25.7 A
20
tCOND tCOND tCOND tCOND
0
2.5
tRCV tRCV tRCV tRCV
0
0 2 4 6 0 2 4 6 0 2 4 6 0 2 4 6
Time [ms] Time [ms] Time [ms] Time [ms]
(a) (b) (c) (d)
Figure 2.13: Simulation results for the constant power mode operation. The converter operates at
650 W, a) 180 kHz, b) 175 kHz, c) 173 kHz, and d) 167 kHz. In each condition, the RCV
strategy accurately enables the SR during the ideal rectifier current conduction time.
30
SR Triggering [V] Rectifier Current [A]
10 tCOND tCOND
tCOND tCOND
0
2.5
tRCV tRCV tRCV tRCV
0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Time [ms] Time [ms] Time [ms] Time [ms]
(a) (b) (c) (d)
Figure 2.14: Simulation results for the constant voltage mode operation, where the LLC converter op-
erates at 30 V output and a) 60 %, b) 45 %, c) 23 %, and d) 5 % load. Regardless of the load
change, the ideal rectifier current conduction time is accurately predicted by the RCV strategy.
is automatically determined by (2.16) and (2.17) since they are the most strict criteria in (2.48) and (2.49).
As the simulation results illustrated in Fig. 2.13 (a) to (d), the SR on-time of employing the RCV strategy,
is exactly the same as tCOND at each switching frequency regardless of variations in the output voltage and
current.
As the constant voltage and constant power mode demonstrate the operations under heavy-load condi-
tions, the simulation results from medium-load to light-load conditions are demonstrated under the constant
voltage mode. The RCV-controlled SR on-time and the theoretical conduction time of the rectifier current
from 60 % to 5 % load are compared in Fig. 2.14 (a) to (d). The behavior in the heavy-load conditions is
similar to Fig. 2.13. As the output current decreases, the amplitude of vCr shrinks, and the start of the rectifier
37
current is delayed; the rectifier current starts conducting at the instant when (2.9) or (2.11) is satisfied, and
hence by using the information of vCr , the instant of triggering on the SR can be controlled by using the
RCV strategy. The turn-off of the SR follows (2.16) and (2.17) by comparing the integrated voltage signals.
According to simulation results, the switching of the SR is synchronous with the start and cutoff of the recti-
fier current in each load condition. As a result, the RCV-controlled SR on-time is equal to the ideal rectifier
current conduction time from below to above resonant frequency, as well as from heavy-load to light-load
conditions.
The 650 W/ 24 V LLC resonant converter shown in Fig. 2.15 is set up with the system parameters listed
in Table III. The DSP sets the frequency of the primary switches for the LLC resonant converter; the RCV
SR controller detects Vin , Vo , and vCr and determines the driving signal for the SR MOSFETs. The block
diagram and the key components used for implementing the RCV SR controller are shown in Fig. 2.11 and
In order to verify the functionality of the RCV strategy, the integrated voltages are displayed and com-
pared with the conduction time of the rectifier current and that of a conventional vds.on sensing SR controller
under the constant power, constant current, and constant voltage mode, respectively. In the constant power
mode, the output power is set at 650 W. As Fig. 2.16 shows, the LLC converter is below the resonant
frequency. Under the same condition, the result of the conventional vds.on sensing method is displayed in
Fig. 2.16 (a), whereas the proposed RCV strategy is presented in Fig. 2.16 (b). As shown in Fig. 2.16 (a), the
drain-source voltage of SR1 , vds.SR1 , is zoomed to show the behavior when SR1 conducts the rectifier current.
As the zoomed vds.SR1 waveform in the bottom part shows, vds.SR1 reaches the proximity of zero, and the SR
is switched off. Due to the inductive voltage offset across the stray inductance, the zero-voltage crossing
of vds.SR1 is much earlier than the actual zero-current crossing of iD1 . The SR turns off prematurely, and
the body diode of the MOSFET starts to conduct, which reduces the efficiency and increases the operating
As opposed to the vds.on sensing method, the proposed RCV SR driving strategy does not rely on the small
and sensitive vds.on signal. The input signals, Vin , Vo , and vCr , are large in magnitude and stable. Besides,
the integrators of the RCV SR controller are equivalent to low pass filters and resistive to high-frequency
noise. As a result, the estimated SR turn-off instant is very close to the zero current crossing. As depicted in
Fig. 2.16 (b), the crossover of the two integrated signals, v p1.int and vCr .int , corresponds to the moment when
When operating in the constant current mode, the output current is fixed at 27 A, and the switching
38
RCV SR SR Control
Controller Output
Resonant
Capacitor
Resonant
Inductor
Transformer
Figure 2.15: The experimental setup: the prototype of the RCV SR controller on the 650 W/ 24 V LLC
resonant converter.
vinv
iD1 32 A iD1
vds.SR1 Zero-current Crossing
(a) (b)
Figure 2.16: Experimental results at the maximum power and below resonance, 650 W/ 170 kHz: a)
the vds.on sensing method turns off the SR early due to the inductive voltage offset induced from
the stray inductance, whereas b) the RCV strategy effectively estimates the zero current crossing
point by comparing the integrals of large and stable voltage signals.
frequency of the LLC converter is above resonance. The critical waveforms of the traditional vds.on sensing
and the RCV method are displayed in Fig. 2.17 (a) and (b) in parallel. As Fig. 2.17 (a) shows, the slope of iD1
soars at the instant when the inverter voltage changes. As a result, the inductive voltage offsets vds.on even
more significantly. The zoomed vds.SR1 waveform shows that it reaches zero before the actual zero-current
Under the same condition, the RCV SR driving strategy is not affected by the inductive voltage due to
large magnitude input signals and the filtering effect of the voltage integration. The experimental results
shown in Fig. 2.17 (b) demonstrate that the crossover of the voltage integrations, representing the turn-off
instant of the SR, is very close to the zero current crossing point. Therefore, the proposed RCV strategy also
39
vinv
iD1 29 A
iD1
vds.SR1 Zero-current Crossing
(a) (b)
Figure 2.17: Experimental results at the maximum current, 27 A/ 250 kHz: a) because of the parasitic
effects, the traditional vds.on sensing method turns off the SR prematurely, while b) the RCV
strategy effectively estimates the zero current crossing by comparing the integrated voltage sig-
nals.
Vp1 V p1
vCr vCr
7.6 A Zero-current Crossing
iD1 iD1 2.6 A
vds.SR1 vinv
(a) (b)
Figure 2.18: Experimental results at light-load conditions, 5 A/ 30 V: a) the traditional vds.on sensing
method turns off the SR early due to the effect of parasitic elements. b) By using the integrated
voltage signals, the on-time error at light load is short in comparison.
Concerning light-load operations, the experiment is performed in the constant voltage mode, where the
output voltage is at 30 V. As illustrated in Fig. 2.18 (a) and (b), the conventional vds.on sensing and the
RCV-based method are compared at 5 A load. In this load condition, the rectifier current starts conducting
after the switching of the inverter voltage which is the rising edge of vinv in this case. As indicated in (2.9),
the rectifier current is not conducted until the criterion is satisfied, and this is verified in Fig. 2.18. At the
leftmost side of the waveform, the resonant capacitor voltage vCr crosses over Vp1 , and it is close to the start
of the rectifier current. As depicted in Fig. 2.18 (a), the conventional vds.on sensing method turns on the SR
MOSFET near the crossover of Vp1 and vCr ; however, regarding the turn-off instant, there exists a significant
discrepancy caused by the inductive voltage offset. For around 25 % of the conduction time, the rectifier
current is conducted through the body diode according to Fig. 2.18 (a). By comparison, the RCV strategy
40
determines the SR turn-off time slightly earlier than the zero current crossing, and yet, the on-time error is
merely around 6 % as indicated in Fig. 2.18 (b). Overall, by comparing the integrations of the large voltage
signals, the RCV strategy is less sensitive than the conventional vds.on sensing method, and the SR on-time
is closer to the actual rectifier current conduction time, particularly at heavy-load conditions.
The experimental results of the ideal SR on-time (corresponding to the actual rectifier current conduction
time), the RCV-controlled and the traditional vds.on sensing-based SR on-time in each operating mode are
summarized and compared in Fig. 2.19. In light-load conditions, as demonstrated in Fig. 2.19 (a), the output
current (Iout ) tapers from 7.5 A to 1 A as the output voltage is regulated at 30 V. According to the experimental
results, the on-time error of the vds.on sensing method is around 23 % to 30 % of the ideal SR on-time. In
comparison, the RCV-controlled SR on-time follows the trend of the ideal SR on-time over the load range.
In the constant power mode, as shown in Fig. 2.19 (b), the output power (Pout ) is limited at its maximum,
650 W, whereas the output voltage transits from 30 V to 24 V as the frequency increases; correspondingly,
the output current is increased from 21.7 A to 25.5 A. For each load condition, the RCV method estimates
the actual rectifier current conduction time effectively by using the large voltage signals, whereas the on-time
In the constant current mode, as presented in Fig. 2.19 (c), Iout is set at 27 A, and Vo is decreased from
23 V to 16.5 V as the frequency increases. The RCV-controlled SR on-time also closely follows the trend
of the ideal SR on-time in each condition, while by vds.on sensing, the error is about 25 % to 35 % of the
on-time. Therefore, the experimental results verified that by using the large voltage integrations, the RCV
strategy effectively covers a wide operating range from below to above resonance as well as from light-load
to heavy-load conditions.
By employing the integrations of the large magnitude voltage signals, the RCV SR driving strategy is less
sensitive to the parasitic effects over a wide operating range as opposed to the conventional vds.on sensing
method. The on-time error is reduced so that the power conversion efficiency and the thermal performance
of the SR MOSFET can be improved. Figure 2.20 (a) and (b) compare the LLC converter efficiency and the
SR MOSFET temperature under different loading conditions when the converter operates at 30 V. In light-
load conditions, due to lower current and fewer differences in SR on-time, the efficiency and the thermal
improvement is insignificant. However, as the load increases, the impact is more considerable. At the full-
load condition, the power loss is reduced by 6 %, which decreases the temperature of the SR MOSFET by
8.5 ◦ C. The reduction accounts for 28 % in the temperature rise relative to 25◦ C room temperature.
The experimental results verified that by using the large voltage integrations, the RCV strategy effectively
41
3 Vout : 30 V 3 Pout : 650 W 3 Iout : 27 A
SR on-time [µs]
SR on-time [µs]
2 30 % error 2 36 % error 2 35 % error
Figure 2.19: The ideal, RCV-controlled and traditional vds.on sensing-based SR on-time are compared.
Experiments are conducted in a) the constant voltage mode in light-load conditions, b) the con-
stant power mode, and c) the constant current mode. The RCV strategy is effective in multiple
operating conditions by using the integrations of the large voltage signals.
0.22 %
94.8 95.02 % 45
94.6 % 42 ℃ 7.5 ℃ 47.2 ℃
94.85 % 94.85 % 39.2 ℃
94.6 40 4.7 ℃
36.3 ℃ 4.3 ℃ 41.4 ℃
94.4 94.23 % 0.35 % 35 2.5 ℃ 37.3 ℃
RCV 34.9 ℃ RCV
94.2 0.09 % 30
Traditional 94.25 % 33.8 ℃ Traditional
94.0 94.14 % 25
150 275 400 650 150 275 400 650
Output Power [W] Output Power [W]
(a) (b)
Figure 2.20: Comparison of the effect of the SR on-time error on a) the efficiency and b) the SR MOS-
FET temperature.
covers a wide operating range from below to above resonance as well as from light-load to heavy-load
conditions. By reducing the on-time error from the conventional vds.on sensing method, the power conversion
The proposed RCV SR driving strategy employs the resonant capacitor voltage that contains the instan-
taneous information of input current and hence can respond to the dynamics of the system rapidly. The dy-
namic test result is demonstrated as shown in Fig. 2.21. The switching frequency steps down from 235 kHz
to 190 kHz within one cycle as shown in Fig. 2.21 (a) and steps up from 190 kHz to 235 kHz also within one
cycle as shown in Fig. 2.21 (b). The evaluated frequency variation range accounts for nearly half of the reg-
ular operating frequency range (150 kHz to 250 kHz). The step change is large and extremely fast compared
with a regular control loop design. The frequency change rate of a closed-loop controller is often limited
42
severe dynamics in
severe dynamics in
Vout voltage and current Vout voltage and current
iD1 iD1
vgs.SR1 vgs.SR1
zoom-in RCV tracks instantaneously zoom-in
RCV tracks instantaneously
235 kHz 190 kHz
190 kHz 235 kHz
(a) (b)
Figure 2.21: The dynamic test waveforms under a large frequency step change, a) from 235 kHz to
190 kHz, and b) from 190 kHz to 235 kHz: the RCV-controlled SR tracks the actual rectifier
current conduction time instantaneously under severe voltage and current dynamics.
by the loop bandwidth; a sudden frequency change is rare in regular operations. However, for validation
purposes, this test condition is applied to verify the proposed RCV SR control method can respond to the
extreme condition properly. The resonant frequency of this LLC converter is around 200 kHz; thus, the test
condition forces the converter switching between the two regions: below and above the resonant frequency.
Therefore, this test condition also validates the ability of RCV SR switching between different operating
modes.
As shown in Fig. 2.21 (a), the switching frequency jumps from 235 kHz down to 190 kHz in the fourth
switching cycle, and hence the on time is increased and so is the peak of the rectifier current (iD1 ). During
the transient response time, both of the output voltage and the rectifier current swings in large amplitudes
drastically. As the waveform shows, the SR gate (vGS.SR1 ) driven by the RCV strategy tracks the actual
rectifier current cycle-by-cycle under this severe dynamic test condition. The frequency step-up test result is
shown in Fig. 2.21 (b); as the zoom-in waveforms show, the switching frequency is increased from 190 kHz
to 235 kHz in the fourth cycle. During the transient, the output voltage dips from 26 V to 22 V, and the peak
of the rectifier current is suddenly cut in half, and yet, the RCV SR on-time tracks the actual rectifier current
conduction time very well. As the rectifier current reduces, the SR on-time shrinks accordingly because
the amplitude of vCr is decreased simultaneously as discussed in Section 2.2.3. At the sixth switching cycle
shown in Fig. 2.21 (b), when the rectifier current reaches zero, the SR switching automatically stops because
the amplitude of vCr is less than nVo (1 + Lr /Lm). In the following cycles, the rectifier current increases and
so does the SR on-time. Therefore, the extreme test condition validates that the proposed RCV SR responds
43
2.5 Summary
In this chapter, an LLC synchronous rectification (SR) strategy based on the Resonant Capacitor Voltage
(RCV) is presented. The RCV strategy does not require current sensors; the SR turn-on and turn-off instants
are determined by simply comparing the integrations of the large-magnitude signals: the resonant capacitor
voltage, the converter input voltage, and the output voltage. By integrating the large voltage signals, the
RCV strategy provides a more noise-immune solution than traditional vds.on sensing methods. Significantly,
it is effective for multiple operating modes, covering both the below- and above-resonance regions. The
functionality and the effectiveness of the RCV strategy are validated by the simulation and experiments of a
650 W/ 24 V LLC converter. The results prove that the RCV controls the SR effectively over a wide operating
range. The dynamic test demonstrates that the RCV strategy can track the actual rectifier current conduction
time cycle-by-cycle under severe frequency variations. Compared with the traditional vds.on sensing method,
the proposed RCV SR driving strategy significantly reduces the on-time error caused by the parasitic effects.
As a result, the power loss is reduced by 6 % in the full-load condition, and for the SR MOSFET, 28 %
in the temperature rise relative to 25◦ C room temperature is reduced. Therefore, the RCV strategy offers a
noise-insensitive and effective solution for the synchronous rectification of LLC resonant converters.
44
Chapter 3
In Chapter 2, the resonant capacitor voltage based synchronous rectification (RCV SR) is introduced. This
large-magnitude voltage based method effectively reduces the SR on-time error and improves the efficiency
performance compared with the conventional vds.on based, non-adaptive SR. However, the amount of reset-
table integrators needed can add complexity to the circuit. For simplifying the design, this chapter introduces
the other SR method of this work, which is also based on large-magnitude voltage signals, and the SR on-
time control is insensitive to the voltage across leakage inductance. This method controls the SR using the
product of the blocking voltage of the SR switch and the rectifier current conduction time, and thus it is called
Volt-Second Product (VSP) SR. This chapter covers the fundamentals of the VSP SR algorithm as well as
the implementation. The proposed VSP SR is compared with an vds.on based adaptive SR to demonstrate the
efficiency improvement.
the time domain, as synchronous rectification is essentially synchronization in time. Figure 3.1 illustrates
the typical above-resonance waveforms of a half-bridge LLC resonant converter, the schematic of which
is shown in Fig. 2.1. The Rectifier Current Equations [102] of a half-bridge LLC resonant converter are
established by solving the following differential equations. In the above-resonance region, defining t0 to be
zero, t2 simply equals half of the switching period Ts /2. For a given input and output voltage, the resonant
45
vLS
vHS Ts 2
vCr VP1
IP1 V in
ir VS1 IS1
im
-IS1
vD2 -IP1
tA
irec
t0 t1 t2 t3 t4 t5 t
Figure 3.1: Typical waveforms of a half-bridge LLC resonant converter in the above-resonance region.
where
1
Ωr = √ (3.3)
LrCr
Similarly, for the above-resonance operation, in the second time interval (t1 – t2 ),
ir.T 2 = Cr dvCr.T
2
dt
vCr.T 2 = Vin − nVo − Lr didtr.T 2 (3.4)
dim.T 2 nVo
=
dt Lm
46
q
Cr
i
r.T 2 = −I S1 cos(Ω r (t − t A )) − (VS1 −Vin + nVo ) Lr sin(Ωr (t − tA ))
q
−I Lr (3.5)
v Cr.T 2 = S1 Cr sin(Ωr (t − tA )) +Vin − nVo + (VS1 −Vin + nVo ) cos(Ωr (t − tA ))
im.T 2 = nVo (t − tA ) − IS1
Lm
In (3.5), tA is simply the interval from t0 to t1 and is dependent on the operating point Vin , Vo , and Ts ,
which will be defined later. In (3.2) and (3.5), the initial conditions VP1 , VS1 , IP1 , and IS1 are the four
independent unknowns. Therefore, four equations are needed to solve for the initial conditions. According
to the waveform of the magnetizing current im , it takes Ts /2 for im to rise from −IS1 to IS1 , and thus
Ts
im.T 2 (tA + ) = IS1 (3.6)
2
The other three equations are derived by considering the continuity of the resonant current and voltage
Ts
vCr.T 1 (t0 ) + vCr.T 2 ( ) = Vin . (3.9)
2
Applying the relationships (3.6) – (3.9) to (3.2) and (3.5), the initial conditions VP1 , VS1 , IP1 , and IS1 can be
where
Lm Ωr Ts
m= , φA = Ωr tA , φS = , (3.11)
Lr 4
47
and
φS sin(2φS )
Vin m cos φA + 2 cos φA − 4 cos φS cos(φA − φS )
Vm1 = − nVo (3.12)
2 cos φS (cos φS + sin φS tan φA ) 2 cos φS (cos φS + sin φS tan φA )
By taking the initial conditions (3.10) back to (3.5) and considering ir.T 2 ( T2s ) = IP1 , the steady-state gain
√
where Ωr is the short-circuit resonant frequency 1/ LrCr . As a result, the duration tA is explicitly expressed
as
Ts α
tA = − (3.14)
4 Ωr
where
n 2nV Ω T Ω T
Ωr Ts
o
−1 o r s r s
α = sin sin + cos (3.15)
Vin 4 4m 4
As for the below-resonance region, the typical operating waveforms are shown in Fig. 3.2. Defining t0
– t1 and t1 – t2 as the first and second interval, the resonant current and voltage in the first time interval are
expressed as
ir.T 1 = Cr dvCr.T 1
dt
(3.16)
vCr.T 1 = −(Lr + Lm ) dir.T 1
dt
where
1
Ωo = p (3.18)
(Lr + Lm )Cr
48
vLS
vHS Ts 2
vCr VS2
VP2 V in
IS2 IP2
ir
im
-IS2 -IP2
vD2
tB
irec
t0 t1 t2 t3 t4 t
Figure 3.2: Typical waveforms of a half-bridge LLC resonant converter in the below-resonance region.
Similarly, to solve for the four initial conditions, VP2 , VS2 , IP2 , and IS2 , four voltage and current equa-
tions have to be established. In the second time interval, the magnetizing current is IS2 at t2 , and this time
Ts
im.T 2 ( ) = IS1 (3.21)
2
The other three equations are derived by considering the continuity of the resonant current and voltage
49
vCr.T 1 (tB ) = vCr.T 2 (tB ). (3.23)
Ts
vCr.T 1 (t0 ) + vCr.T 2 ( ) = Vin . (3.24)
2
Applying the relationships (3.21) – (3.24) to (3.17) and (3.20), the initial conditions VP2 , VS2 , IP2 , and IS2 can
where
Ωr Ts
φB = ΩotB , β = ( − tB ), (3.26)
2 2
and
√
1 + m(1 + cos(φB ))((Vin − nVo ) cos(2β ) + nVo )
Vm2 = √ (3.27)
1 + m(1 + cos(2β ))(1 + cos(φB )) − sin(2β ) sin(φB )
2nβVo √
m ( 1 + m sin(2β ) − (1 + m) sin(φB ))
Vm3 = √ (3.28)
1 + m(1 + cos(2β ))(1 + cos(φB )) − sin(2β ) sin(φB )
Note that Ts /2 − tB corresponds to the rectifier conduction time in the below-resonance operation.
By taking the initial conditions 3.25 back to 3.20 and considering ir.T 2 ( T2s ) = IS2 , the gain equation in the
nVo m
= √ , (3.29)
Vin /2 β cot(β ) − β 1 + m tan( φB ) + m
2
Although tB or β can only be expressed implicitly according to (3.29) due to the coexistence of trigonometric
and non-trigonometric terms, the critical information to the proposed VSP SR algorithm is the range of tA
50
vgs.SR1
HS
vHS iD1 Llk
n::
Vin vlk irec
ir Lr
Cds
LS irec vds.on1
vLS Lm
Co vD1
Cr vds.on2 vD2
im Cds vo
iD vlk Ro Plateau Voltage Plateau Voltage
vCr vDp1 vDp2
Ra Ra Llk
Rb R vgs.SR2
vD1 b vD2 vds.on1
vds.on2
(a) vth.rev
millivolt range
vth.on vth.off
Conventional vgs.SR1 plateau volt.
vds.on1 vgs.SR1
vds.on based vgs.SR2 sampling Proposed vth.on
vgs.SR2
vds.on2 vD1 VSP SR Proposed:
SR logic Conventional:
vD2 rising edge logic based on vds.on1,2 based on vDp1,2
( vth.rev for detection vgs.SR1
vth.off
vgs.SR2
adaptive SR) Premature Premature
vHS , vLS turn-off turn-off
small and sensitive signal simple implementation
easily affected by parasitic large voltage signal sensing, tolerant of
Improved SR on-time
components parasitic effects
(b) (c)
Figure 3.3: a) a typical LLC resonant converter with synchronous rectification, including circuit par-
asitic components. b) block diagram of the conventional vds.on based SR and the proposed Volt-
Second Product (VSP) based SR; the former depends on the drain-source voltage during the turn-
on phase, while the latter employs the drain-source voltage during the turn-off phase. c) vds.on1
and vds.on2 are in millivolt range and are easily affected by the parasitic effects, whereas the pro-
posed VSP SR uses large-magnitude voltage signals and improves the SR on-time.
stage for communication purposes, this part of the work proposes a noise-tolerant SR method that can be
easily integrated with the microcontroller. The SR on-time is determined based on the product of SR drain-
source blocking voltage during the turn-off phase and rectifier current conduction time. This Volt-Second
Product based SR (VSP SR) employs the large-magnitude voltage signal instead of the low-magnitude,
noise-sensitive vds.on used by the conventional SR controllers, and hence the VSP SR is more tolerant of
parasitic effects, reducing the SR on-time error. As illustrated in Fig. 3.3 (a) and (b), the proposed VSP
SR simply uses a pair of voltage dividers to sense the drain-source voltage during the turn-off phase (vD1
and vD2 ). Along with the primary switch driving signals, the VSP SR algorithm determines the SR on-
time. Figure 3.3 (c) shows an example of the parasitic effects on the SR gating signals. The small vds.on
voltage is susceptible to the effects of voltage across the stray inductance, which can cause premature turn-
off even with the conventional adaptive scheme. The proposed VSP SR, on the other hand, improves the
51
plateau voltage sampling
vD1
vDp1 CMPB1
ADC1
CMPB2 vgs.SR1
vDp2
CMPA1 Saw1 VSP SR PWM vgs.SR2
Sync 1
vD1.logic Cap 1 algorithm generation
tcnd1 tcnd1 Sync 2
Saw1
vref sampling tcnd2
Sync 1
Saw1 Saw2
rising edge vLS.sync vLS.sync T sc (digital counter)
detection generation
Figure 3.4: Functional block diagram of the proposed Volt-Second Product (VSP) SR.
SR efficiency performance by employing large-magnitude voltage signals. More details are provided in the
following sections.
A more detailed functional block diagram of the proposed VSP SR control logic is shown in Fig. 3.4. The
sensed (vD1 , vD2 ) are fed into the plateau voltage sampling block, where the plateau of the blocking voltage
vDpx (x = 1, 2) are sampled. This information is used as the voltage control signal of the Volt-Second Product
(VSP) SR. The time control signal (tcndx ), on the other hand, comes for the tcndx sampling block. The tcndx
sampling block is controlled by the time capture signal (Capx ), the synchronization signal (Syncx ) and the
sawtooth digital counter signal (Sawx ). The Capx signal captures the instantaneous value of Sawx , which
corresponds to the rectifier current conduction time (namely the time control signal tcndx ), while the Syncx
Based on the voltage and time control signals, the VSP SR algorithm determines the SR on-time control
signal CMPBx . The VSP SR algorithm also includes the calculation of the SR control time window, Tsc ,
which is a period of time that allows turn-on trigger of the SR. This Tsc time window starts at the falling
edge of the primary switches control signals (vLS , vHS ) and finishes when the time expires, which will be
discussed in the following section. As shown in the block diagram, vLS.sync is a pulsating signal that is
synchronized with the falling edge of vLS and has a pulse width of Tsc ; the other corresponding signal is
vHS.sync . These two signals and the logic-level signals of vDx , namely vDx .logic , determine the synchronization
signals (Syncx ). The logic-level signals are generated simply by comparing vDx with a threshold voltage
52
(vre f ), which is commonly around 1 to 2 V . Detailed operating waveforms and sequence are explained in the
following sections.
The turn-on of SR using the proposed VSP method is determined by the logic level of the drain-source
voltage (vD1 , vD2 ) and the primary gate-drive signals (vLS , vHS ). As illustrated in Fig. 3.5, vD1.logic and
vD2.logic are the logic levels of the corresponding drain-source voltages, which are determined by comparing
vD1 and vD2 with a reference voltage, vre f . The other pair of control signals is vLS.sync and vHS.sync , which
is respectively set at the falling edge of vLS and vHS , and is reset after a SR control time window, Tsc . The
SR driving signal vgs.SR1 is set when both vD2.logic and vLS.sync are high; correspondingly, vgs.SR2 is set when
both vD1.logic and vHS.sync are high. Since the operation principles for vgs.SR1 and vgs.SR2 are symmetrical, the
For the above-resonance operation, as shown in Fig. 3.5 (a), the primary low-side switch vLS turns off at
t0 , and vLS.sync is set at the same time. From t0 to t1 , because the rectifier current is conducting, the rectifier
voltage vD2 is clamped low until the current iD2 reaches zero, which means that vD2.logic changes its state after
the rise of vLS.sync . Therefore, vgs.SR1 is set at the rising edge of vD2.logic in the above-resonance region. Note
that the minimum duration of the SR control time window Tsc is determined by the maximum time interval
from t0 to t1 , which takes place at the maximum switching frequency. Ideally, the SR on-time tSR1 is equal to
the rectifier current conduction time tcnd1 ; in practice, a safe operation margin is reserved at the falling edge
For the below-resonance operation, as shown in Fig. 3.5 (b), from t0 to t1 , the transformer is decoupled
despite the residual energy circulating in the loop. The voltage ringing causes vD2.logic to change its states
frequently and the rising edges are not necessarily synchronous with the conduction of iD1 as in the case of
above-resonance. Rather, when the input voltage changes its polarity, the transformer is coupled, and the
rectifier current iD1 starts to conduct. This instant is synchronous with the falling edge of vLS or the rising
edge of vLS.sync . Because vLS.sync rises when the transformer is coupled, where vD2.logic is in the high state,
vLS.sync determines the turn-on of vgs.SR1 in this case. Note that the pulse width of vLS.sync , namely Tsc , cannot
The selection of the time window Tsc is based on the range of tA and Ts /2 −tB . To unify the SR algorithm
across above- and below-resonance regions, it is necessary to select Tsc such that the time interval applies to
both operation modes. In the above-resonance region, Tsc must be larger than tA . According to (3.14) and
53
vD1
vD2 tcnd1[k-1] vDp1 [k] tcnd1[k]
vDp1 [k-1]
vref
vLS
vHS
vD1.logic
vD2.logic
tA
vLS.sync
vHS.sync T sc T sc
iD1
iD2
vgs.SR1
tSR1[k-1] tSR1[k]
t0 t1 t2 t3’ t3 t4 t5 t6 t7’ t7 t
(a)
vD1
vD2 tcnd1[k-1] vDp1 [k] tcnd1[k]
vDp1 [k-1]
vref
vLS
vHS
vD1.logic
vD2.logic
tB
vLS.sync
vHS.sync T sc T sc
iD1
iD2
vgs.SR1
tSR1[k-1] tSR1[k]
t0 t1 t2’ t2 t3 t4 t5 t6’ t6 t7 t
(b)
Figure 3.5: Typical waveforms of the proposed VSP SR: a) above-resonance and b) below-resonance
operations. The Tsc time window allows for correct setting of the SR for both cases.
(3.15), because α is a positive real number, and the switching period Ts is less than the resonant period Tr
In the below-resonance region, Tsc must be less than Ts /2 − tB , otherwise, the rising edges of vD2.logic at
the end of rectifier current can mis-trigger vgs.SR1 . According to (3.29), since the gain is greater or equal to
If Ts /2 − tB ≤ Tr /4, then tan β ≤ 1 following the definition of β (3.26). Based on (3.31), tan(ΩotB /2) then
54
√
should be greater or equal to 1/ 1 + m, which can be easily disproved by the instance of tB equal to zero.
Therefore,
Ts Tr
− tB > . (3.32)
2 4
As a result, the time window Tsc for vLS.sync and vHS.sync is selected as Tr /4, which is applicable to both
Based on the operating principles of the resonant tank and the logic signals generated from drain-source
voltage and the primary gate-drive signals, the criteria to turn on the SR in above- and below-resonance
regions are determined respectively. The turn-on of vgs.SR1 is synchronous with vD2.logic for the above-
resonance, while it is synchronous with vLS.sync for the below-resonance operation. By simply taking the
logic AND operation of vD2.logic and vLS.sync , these two signals take over the control of SR seamlessly with-
The sensed drain-source voltage not only relates to the turn-on of SR but also contains the information
to control the turn-off instant. The core VSP SR algorithm determines the SR on-time for both below- and
above-resonance conditions using the sampled drain-source blocking voltage from the kth cycle (vDP1 [k])
and the rectifier current conduction time from the previous cycle (tcnd1 [k − 1]); both are extracted from the
drain-source voltage during the SR turn-off phase. The core VSP SR algorithm is a linear combination of
where DT is the safe operation margin, vcm is a constant for compensation, and
(vDp1 [k], vT [k]), if vDp1 [k] ≤ vT [k];
(vnum , vden ) = (3.34)
(vT [k], vDp1 [k]), otherwise.
For a center-tapped rectifier, vDp1 [k] and vT [k] are respectively proportional to the output voltage and target
with the same ratio. According to the VSP SR core equation, in steady states, vnum and vden are equal,
so the SR on-time tSR1 [k] reaches the rectifier conduction time with a constant safe operation margin DT .
In transient response, the sampled drain-source blocking voltage is the key to adjusting the SR on-time
dynamically. With a proper selection of vcm , the proposed VSP SR algorithm maintains the safe margin
55
above DT during the transient, which will be analyzed in the following subsection.
For a frequency modulated LLC resonant converter, the switching frequency is modulated by the closed-loop
controller. When the frequency increases, the SR on-time must reduce accordingly, otherwise, the rectifier
current can flow reversely and cause large voltage spikes at the turn-off of the SR. It is particularly critical in
the above-resonance region because the rectifier current conduction time changes with the switching period.
As the system block diagram shows in Fig. 3.6, the closed-loop controller outputs the control signal vc
based on ∆vo , the error of the output voltage. The frequency modulator determines the switching frequency
of the primary driving signals vHS and vLS by comparing vc with a sawtooth signal vsaw . Considering a typical
PI controller with the proportional and integral gain Kp and KI , the relationship between vc and vsaw is shown
in Fig. 3.7. Based on the geometry, the change in the switching period (∆ts ) per half-cycle is defined as
where Sc and Ssaw are the slopes of the controller output and the internal sawtooth signal, respectively. In
the above-resonance region, ∆ts equals the change in the rectifier current conduction time, as the conduction
time is one-half of the switching period. Therefore, to maintain the safe operation margin (DT ) during the
dynamic response, as illustrated in Fig. 3.7, the time difference between the rectifier current conduction time
δ T [k] := tcnd1 [k] − tSR1 [k] = tcnd1 [k − 1] − tSR1 [k] − ∆ts ≥ DT. (3.36)
∆vo · tcnd1 [k − 1]
≥ ∆ts [k] (3.37)
vden + vcm
For the above-resonance operation, the inequality (3.37) is more critical than below-resonance because
the reduced switching period directly leads to a reduction of rectifier conduction time. Combining (3.35) and
56
HS SR
Vin n::
Lr
Lm
LS SR Co Ro vo
Signal Cr
Isolation vD1
vD2
vHS vLS
Logic VSP SR
Gates Controller
Figure 3.6: System block diagram of a typical closed-loop controlled LLC resonant converter with the
proposed VSP SR controller.
vc
vsaw D vc
Sc
D ts
Ssaw
T s [k-1] 2
vHS
vLS T s [k-1] 2 T s [k] 2 T s [k+1] 2
Figure 3.7: The relationship between the change in the switching period (∆ts ) and the closed-loop
controller gain defines the rate of change in the rectifier current conduction time in the above-
resonance region.
(3.37) gives the criterion for selection of the compensation voltage vcm :
Ssaw · tcnd1 [k − 1]
vcm ≤ − vden (3.38)
Kp + KI · tcnd1 [k − 1]
The larger the compensation vcm is, the less the VSP SR is sensitive to the frequency change. To ensure
a sufficient safe margin, the maximum of vden and the minimum of tcnd1 can be considered. Due to the
symmetry, similar analyses can be applied to the other SR. As a result, the complete VSP SR algorithm is
summarized as follows:
57
set,
vD2.logic & vLS.sync
vgs.SR1
reset, tSR1 reached
(3.39)
set,
vD1.logic & vHS.sync
vgs.SR2
reset, tSR2 reached
The system block diagram is shown in Fig. 3.6, where both the closed-loop controller and the VSP SR are
in the single microcontroller. It clearly shows that the voltage sensing of the VSP SR requires only a pair
of voltage dividers for each of the SR switches. Due to the nature of large-magnitude voltage sensing, the
PCB (printed circuit board) layout design has less constraints than conventional vds.on based methods. The
SR controller, that is, the microcontroller, does not have to be placed as close to the SR switches as possible,
This section focuses on the VSP SR controller part, of which the implementation is shown in Fig. 3.8.
According to the VSP SR algorithm (3.39), vgs.SR1 is set by taking the logical AND operation of vD2.logic and
vLs.sync , which is indicated as Sync1 in Fig. 3.8. vgs.SR1 is reset according to the core VSP SR algorithm (3.33)
and (3.34), where the blocking voltage vDp1 and the rectifier current conduction time tcnd1 are the required
input signals. The sampling of vDp1 is implemented by feeding vD1 to the ADC module ADC1 , whereas tcnd1
is captured at the rising edge of vD1.logic , denoted as Cap1 . The implementation of the other SR is mirrored.
The detailed timing sequence for below- and above-resonance operations is shown in Fig. 3.9. As shown
in the timing sequence, each SR is respectively controlled by an internal time-based counter signal, Saw1 and
Saw2 . These time-based counters are used for recording the rectifier conduction time, triggering the turn-off
of SR, and initiating ADC sampling in each cycle. Note that the VSP SR algorithm (3.39) applies to both
below- and above-resonance conditions; the following description focusing on the first SR is applicable to
both conditions.
In the timing sequence, the k − 1th cycle starts at t0 , and the kth cycle starts at t4 ; each cycle is initiated
by the ADC1 interrupt service routine (ISR), of which the flowchart is shown in Fig. 3.10. At t1 and t5 ,
Sync1 resets the counter Saw1 and sets vgs.SR1 . When the counter Saw1 reaches CMPB1 , a reset signal is
sent to vgs.SR1 at t2 and t6 , respectively. The threshold CMPB1 corresponds to the SR on-time calculated by
58
vD1 ADC1
vgs.SR1 ,
vD1.logic Cap 1 vgs.SR2
vref PWM
vLS , vHS
Sync 1 Module
vLS.sync vLS.sync ,
vHS.sync
vD2 ADC2
vD2.logic Cap 2
vref
Sync 2
vHS.sync
Microcontroller
Figure 3.8: Block diagram for implementation of the proposed VSP SR using TMS320F28379D with
only a few external logic gates.
vD1
vD2 v vDp1 [k] vDp1 [k-1] vDp1 [k]
Dp1 [k-1]
vref
Sync 1
Cap 1
ADC1
tcnd1[k-1] tcnd1[k] tcnd1[k-1] tcnd1[k]
CMPA1
Saw1 tSR1[k-1] tSR1[k]
CMPB1 tSR1[k-1] tSR1[k]
Sync 2
Cap 2
ADC2
CMPA2
Saw2
CMPB2
vgs.SR1
vgs.SR2 tSR1[k-1] tSR1[k] tSR1[k-1] tSR1[k]
t0 t1 t2 t3 t4 t5 t6 t7 t t0 t1 t2 t3 t4 t5 t6 t7 t
Below-Resonance Region Above-Resonance Region
Figure 3.9: Timing sequence of the VSP SR controller for below- and above-resonance operations.
tSR1 [k] is computed at ADC1 where vDp1 [k] is sampled. Sync1 resets the time counter, while Cap1
captures tcnd1 .
(3.33), which is updated in the ADC1 ISR of each cycle. At the rising edge of vD1 , correspondingly t3 and
t7 in the k − 1th and kth cycle, the time-interval capture signal Cap1 is triggered, and the value of Saw1 is
captured. This value corresponds to tcnd1 , and the captured value in the k − 1th cycle is used to determine
tSR1 [k] following (3.33). Note that the triggering of Cap1 is latched until the following ADC2 to prevent false
triggering. This is particularly important for the below-resonance mode to avoid false triggering caused by
voltage ringing on vD1 during the zero-current time intervals. Since each tcnd1 is captured following an ADC2
event, the reset of Cap1 latch is handled by ADC2 ISR; the reset of Cap2 latch is handled by ADC1 ISR.
As shown by the flowchart of ADC1 ISR in Fig. 3.10, the blocking voltage of the present cycle and the
rectifier conduction time from the previous cycle are acquired. The acquired vDp1 [k] and tcnd1 [k − 1] are for
calculating tSR1 [k]. In addition, the latch of Cap2 is reset and ready for capturing the conduction time of the
59
ISR: ADC1 [k]
Yes
vnum = vT [k]
vden = vDp1 [k] Return
Figure 3.10: Flowchart of ADC1 interrupt service routine (ISR). In the kth cycle, it executes the core
of the VSP SR algorithm to determine the SR on-time tSR1 [k] based on tcnd1 [k − 1], vDp1 [k], and
vT [k].
other SR at the next rising edge of vD2 . By the end of the ISR, tSR1 is assigned to CMPB1 , which compares
under closed-loop control. The system block diagram is shown in Fig. 3.6 with the system parameters listed
in Table 3.1. To examine the selection of the time window Tsc , Fig. 3.11 shows the theoretical rectifier
current conduction time for below-resonance and tA for above-resonance operations using (3.13) and (3.29).
Considering the SR operating load range from 25 % to 100 %, Fig. 3.11 demonstrates that the selected time
Parameter Values
input voltage, Vin 400 V
nominal power, Pout 650 W
maximum output current, Imax 27 A
nominal output voltage, Vnom 24 V
maximum output voltage, Vmax 30 V
minimum output voltage, Vmin 20 V
switching frequency, fs 150–250 kHz
resonant frequency, Fr 189 kHz
magnetizing inductance, Lm 103.4 µH
resonant inductance, Lr 37.7 µH
resonant capacitor, Cr 18.8 nF
transformer turns ratio, n 8.1
primary switches IPA65R400CE
SR switches STP240N10F7
60
fs F r (below resonance)
0.75 0.8 0.85 0.9 0.95 1
0.5
Ro Decreases ( t s 2 - t B) T r
0.375
T sc T r
t Tr
0.25
0.125
tA T r
Ro Decreases
0
1 1.05 1.1 1.15 1.2 1.25
fs F r (above resonance)
Figure 3.11: Plot of the theoretical, normalized rectifier current conduction time, tA , and time window
Tsc , considering 25 %, 50 %, 75 %, and 100 % load. The selection of Tsc is around the center
with sufficient margin between the rectifier current conduction time and tA .
window Tsc meets the criteria (3.30) and (3.32) with sufficient margin.
In order to verify that the proposed VSP algorithm effectively adjusts the SR on-time and maintains the
safe margin in transient response, the difference between the SR on-time and the rectifier current conduction
time (δ T ) is simulated. The above-resonance operation is focused as the SR on-time changes significantly
with frequency in this region. As shown in Fig. 3.12 (a), the output voltage (vo ) starts rising due to a step-up
vo command, and the frequency decreases following the control loop. In this case, both the conventional
and the proposed SR keep the safe operation margin above DT . When vo command steps down, as shown
in Fig. 3.12 (b), the frequency increases as a result of closed-loop regulation. In this case, however, several
cycles of the conventional adaptive SR have insufficient safe margin, since the detection of δ T < DT has to
take place first before the SR on-time can be reduced in the next cycle. In contrast, the proposed VSP SR
algorithm uses both the voltage and time information to change the SR on-time dynamically, maintaining the
The 650 W battery charger prototype for validating the VSP SR is shown in Fig. 3.13. The VSP SR
daughterboard located on the secondary side detects the drain-source voltages, synthesizes the SR triggering
signals, and transmits them to the microcontroller board. For the below-resonance operation, Fig. 3.14 shows
the converter operating at the maximum voltage and maximum power condition, where Fig. 3.14 (a) is the
result using a commercial vds.on based adaptive SR controller, and Fig. 3.14 (b) is the result of the proposed
VSP SR. Figure 3.14 (a) shows that the adaptive SR controller turns on the switch early. It is mis-triggered
when the SR body diode starts conduction due to the sub-resonance between the parasitic capacitors and the
equivalent inductor. Consequently, the rectifier has reverse current flow, which lowers the efficiency.
61
24 25
vo [V]
vo [V]
22
23
20 21
40 40
iD1 [A]
iD1 [A]
20 20
0 0
dT [Clock] vgs.SR1[V]
dT [Clock] vgs.SR1[V]
10 10
0 0
80 Proposed VSP SR Adaptive SR 80 Proposed VSP SR Adaptive SR
60 60
40 40
20 20
DT dT > DT DT dT < DT
0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
t [ms] t [ms]
(a) (b)
Figure 3.12: Simulation results of the VSP SR and the conventional adaptive SR under closed-loop
regulation: a) vo rising and b) vo falling. In both cases, the VSP SR maintains the dead time δ T
above the preset safe margin DT , whereas the conventional adaptive SR has multiple cycles with
insufficient safe margin during frequency rise, increasing the risk of cross conduction.
Primary Side
VSP SR
Daughterboard
Secondary Side
Microcontroller
Board
Figure 3.13: The prototype of the 650 W/ 24 V LLC resonant converter with the proposed VSP SR
controller.
In contrast, the proposed VSP SR, as shown in Fig. 3.14 (b), is not affected by the parasitic effects and
turns on the switch properly after the rectifier current starts conduction. The turn-on delay of the prototype is
around 240 ns, which can be further improved by reducing the propagation delay. By removing the reverse
current, the VSP SR improves the efficiency by 0.2%, which is around 3% of the total loss.
For above-resonance operation, the results are compared in Fig. 3.15 (a) and (b), where the operating
condition is at the minimum output voltage and maximum current. As shown in Fig. 3.15 (a), the vds.on
based adaptive SR turns off the SR prematurely due to drastic change in the current slope and voltage across
the leakage inductance, and the dead time is nearly as much as twice that in the below-resonance operation.
62
vgs.SR1 vgs.SR2 (10V/ Div.) irec (20A/ Div.) vDS(SR1) (20V/ Div.) vgs.SR1 vgs.SR2 (10V/ Div.) irec (20A/ Div.) vDS(SR1) (20V/ Div.)
1 ms / Div. 1 ms / Div.
vDS(SR1) (2V/ Div.) Zoom-in vDS(SR1) (2V/ Div.) Zoom-in
inversion current 180 ns 110 ns
(a) (b)
Figure 3.14: Steady-state test results in constant voltage regulation at 30 V/ 650 W, fs = 167 kHz: a)
the conventional adaptive SR turns on the switch early due to voltage ringing-induced SR body
diode conduction, introducing reverse current; b) the proposed VSP SR controls SR on-time
properly.
vgs.SR1 vgs.SR2 (10V/ Div.) irec (20A/ Div.) vDS(SR1) (20V/ Div.) vgs.SR1 vgs.SR2 (10V/ Div.) irec (20A/ Div.) vDS(SR1) (20V/ Div.)
1 ms / Div. 1 ms / Div.
vDS(SR1) (2V/ Div.) Zoom-in vDS(SR1) (2V/ Div.) Zoom-in
330 ns 120 ns
(a) (b)
Figure 3.15: Steady-state test results in constant current regulation at 27 A/ 540 W, fs = 238 kHz: a) the
conventional adaptive SR turns off the switch prematurely due drastic change in the current slope
and voltage across the leakage inductance; b) the proposed VSP SR is tolerant to the parasitic
effects, maintaining consistent dead time.
Contrary to the conventional method, the proposed VSP SR is not affected by the parasitic effect, and it
maintains a consistent safe margin for both operating modes. As shown in Fig. 3.15 (b), the SR tracks the
rectifier current conduction time and turns off the SR 120 ns before the current reaches zero, which is close
to the margin left in the below-resonance operation. Compared with Fig. 3.15 (a), the dead time is reduced
The efficiency results of the 650 W prototype using the conventional vds.on method and the proposed
VSP SR are compared in Fig. 3.16 (a) and (b) considering constant-voltage and constant-current regulation,
respectively. For constant-voltage regulation, the output voltage is regulated at Vmax (30 V ), and the operating
63
96.2
Proposed VSP SR Proposed VSP SR
93.2 93.13
95.80 Adaptive SR Adaptive SR
95.8 92.92
93.11
95.76
Efficiency [%]
Efficiency [%]
95.32 92.8
95.4
92.49
95.32 92.52
94.79 92.4
95.0
92.00 92.28
94.6 94.76 More loss due to 94.43 92.0
Efficiency improvement due
reverse current
to reduction of on-time error
91.77
94.2 94.25 91.6
25 50 75 100 20.5 21.5 22.5 23.5
Load [%] Output Voltage [V]
(a) (b)
Figure 3.16: Efficiency test results under: a) constant voltage regulation, Vo = 30 V , and b) constant
current regulation, Io = 27 A.
vgs.SR1 vgs.SR2 (10V/ Div.) irec (20A/ Div.) vo (10V/ Div.) vgs.SR1 vgs.SR2 (10V/ Div.) irec (20A/ Div.) vo (10V/ Div.)
2 ms / Div. 2 ms / Div.
(a) (b)
Figure 3.17: The dynamic test waveforms under a large command change, a) from Vmin to Vmax , and b)
from Vmax to Vmin : the VSP SR tracks the rectifier current conduction time properly under severe
voltage and current dynamics.
frequencies are below the resonance. The efficiency test results are comparable because the SR on-time of
the proposed VSP SR and the conventional adaptive SR are similar. In the full-load condition, the adaptive
SR is less efficient due to the reverse current shown in Fig. 3.14 (a).
For constant-current regulation, the output current is regulated at Imax (27 A) and the efficiency is mea-
sured above the minimum output voltage Vmin (20 V ). As shown in Fig. 3.16 (b), the efficiency of the
conventional adaptive SR is lower, and the gap increases with frequency. It is because the conventional
method turns off the SR prematurely, and the dead time increases with frequency.
The closed-loop dynamic test results of the proposed VSP SR are shown in Fig. 3.17. The command
changes from Vmin to Vmax and vice versa, which emulates the largest step change in the system, covering
the full operating frequency range. Under such a severe transient, both the rectifier current (irec ) and the
output voltage (vout ) respond dramatically. The VSP SR successfully tracks the switching frequency and the
64
rectifier current conduction time in both the output voltage step-up and step-down conditions. As shown
in Fig. 3.17 (a), the output voltage increases as the frequency decreases; the operating mode transits from
safe margin. In Fig. 3.17 (b), the output command steps down and the operating frequency increases from
the minimum to the maximum. The SR on-time also tracks the rectifier current conduction and maintains
3.5 Summary
This chapter introduces a Volt-Second Product (VSP) based SR control method for LLC resonant converters
in high-power, high-current battery charging applications. The VSP SR employs the SR drain-source voltage
during the turn-off phase, which is large-magnitude and noise-tolerant in contrast to the drain-source voltage
in the turn-on phase (vds.on ). The VSP SR is implemented digitally with only a few external logic gates.
It covers both below- and above-resonance operations, which is essential to battery charging applications.
Compared with the conventional vds.on based adaptive SR, the VSP SR reduces the on-time error. In the
below-resonance region, the VSP SR turns on the switch properly without any reverse current, which reduces
the loss by 3%. In the above-resonance region, the VSP SR is independent of the voltage drop on the stray
inductance and hence reduces the on-time error by 64%. In transient response, the VSP SR algorithm
actively adjusts the SR on-time to maintain a safe operation margin, which prevents potential reverse current
or cross-conduction. Therefore, the VSP SR offers a simple, noise-tolerant and effective solution to LLC SR
Since Chapter 2 compares RCV SR with vds.on based non-adaptive SR and this chapter discusses the
comparison between VSP SR and vds.on based adaptive SR, it is worthwhile to summarize them. Table 3.2
compares the SR methods in terms of noise tolerance, operating range, circuit complexity, and SR on-time
error. The vds.on based non-adaptive SR is greatly affected by the noise across the parasitic components, and
thus the SR on-time error is the largest among the four, reducing the suitable operating range. The vds.on
65
based adaptive SR is more noise-tolerant than the non-adaptive, although the on-time error caused by the
parasitic effects remains significant in the above-resonance region. Commercialized integrated circuits (ICs)
for the vds.on based methods are available, and the application circuit is simple due to the highly integrated
circuits. The proposed RCV and VSP SR use discrete components for prototyping, which makes the circuit
appear to be more complex. However, the circuit complexity can be reduced by employing a customized IC.
The proposed RCV SR and VSP SR are both based on large voltage signals, and hence they are insensitive
to the parasitic effects, reducing the on-time error. RCV SR requires a couple of voltage integrators for each
of the SR, which complicates the circuit. VSP SR greatly simplifies the implementation compared with
RCV SR. It has fewer input signals and does not require voltage integrators. Although RCV SR has more
complex circuits, it closely tracks the rectifier current conduction time in transient as well as in steady-state.
Compared with VSP SR, RCV SR is more suitable for applications that continuously operate with dynamic
loads, such as computing products. For EV battery charging applications, where the load changes slowly
over time, VSP SR is simple and effective. Overall, compared with vds.on based methods, both the proposed
RCV SR and VSP SR improve the efficiency by reducing the SR on-time error, which makes them strong
66
Chapter 4
While Chapters 2 and 3 covered the efficiency aspect, one major indicator of steady-state performance, this
chapter will discuss the dynamics of resonant converters, particularly resonant converters with charge mode
control. Charge mode control refers to the control of the integration of inductor current in every switching
period of the converter, which equivalently controls the amount of electric charge per switching cycle. The
command of the charge control loop is from the output of the voltage loop controller that controls the output
voltage. In this way, a cascade loop architecture is formed, where the charge feedback is the inner loop and
the voltage feedback is the outer loop. The inner charge loop shapes the plant of the outer voltage loop, and
a well-designed inner loop makes the outer loop easy to achieve higher loop bandwidth.
Although charge mode control has been developed for resonant converters, the existing small-signal
dynamic model for a charge-controlled resonant converter is oversimplified, which leads to prediction errors,
especially in high-frequency regions. This chapter develops a new dynamic model for charge-controlled
resonant converters that improves the prediction of the dynamics, assisting high-bandwidth design. This
chapter uses a CLLC resonant converter as the case study, since this topology is also becoming popular in
battery charging applications, and it has a higher order resonant tank, which is more complex than LLC
converters. In fact, LLC resonant converters can be considered as a special case of CLLC converters, and
the modeling methodology developed for CLLC converters is applicable to LLC converters. This chapter
starts with the open-loop model of a half-bridge CLLC resonant converters, of which the schematic diagram
67
Cr1 Cr2
irec
L1 i1 it n:1 i2 L2
Vg Cf Ro Vo
Lm
Cr1 Cr2
im
V1 V2
Describing Functions (EDF) are broadly-accepted techniques due to the systematic procedures. This section
first reviews the open-loop EDF model of a CLLC resonant converter, which is used for developing the
The study of the small-signal model of resonant converters using EDF focuses on modeling the slowly
time-varying envelope of the resonant current and voltage. The procedure involves 1) Deriving the state
equations of the resonant circuit based on KVL and KCL; 2) Applying harmonic approximation to the
resonant current and voltages and deriving the corresponding envelope terms of the sine and cosine parts;
3) Deriving the EDF of the inverter and rectifier voltages and finding the corresponding envelope terms of
the sine and cosine parts; 4) Taking the results from the second and third steps into the equations defined in
the first step and equating the sine and cosine coefficients respectively, with the new set of state equations
being called modulation equations; 5) Solving for the state variables at the operating point; 6) Perturbing and
Referring to the AC equivalent circuit shown in Fig. 4.2, a transformer T-mode is applied to facilitate
the analysis. The secondary side components are referred to the primary side. In the equivalent circuit, vab
and vcd represents the AC component of the inverter and rectifier voltage, respectively. For a half-bridge
inverter, vab is a square wave with amplitude ±Vin /2, and for a half-bridge rectifier, vcd is a square wave with
and
dim di1 dim
Lm = n2 L2 − n2 L2 + vt + vcd , (4.2)
dt dt dt
68
L1 i1 it n2L2 i2 n
im
vab Lm vcd
C1 C2 n2
v1 nv2
Figure 4.2: The AC equivalent circuit of the resonant tank using a transformer T-model.
By KCL,
dv1
i1 = C1 , (4.5)
dt
and
i2 C2 dvt
i1 − im = = 2 . (4.6)
n n dt
The First Harmonic Approximation (FHA) is applied in this case. The first harmonic of i1 and im are
69
represented as the sum of their sine and cosine components, and the general form is represented as follows:
i1 = i1s sin ωst + i1c cos ωst
(4.9)
im = ims sin ωst + imc cos ωst
Also, the secondary resonant current i2 is dependent on i1 and im , which is equal to n(i1 − im ) or nit . Hence,
the FHA of it is
it = its sin ωst + itc cos ωst = (i1s − ims ) sin ωst + (i1c − imc ) cos ωst. (4.11)
Note that i1s , i1c , ims , imc , its , and itc are functions of time. Since its and itc are the sine and cosine components
Until now, each item in the KVL and KCL equation (4.3) and (4.8) can be replaced by the corresponding
first harmonic, except vab and vcd . The first harmonic of the inverter voltage vab is
2
vab1 = Vg sin ωst (4.15)
π
70
Note that the cosine part of vab1 is zero because vab is selected as the reference. As for the rectifier voltage vcd
(referred to the primary side), considering that the transformer is continuously coupled, vcd is nvo /2 when i1
is larger than im and −nvo /2 when i1 is less than im , and thus vcd is expressed as
nvo
vcd = sign(i1 − im ) . (4.16)
2
The modulation terms are the coefficients of sin ωst and cos ωst in the previous step. Those terms mod-
ulates the amplitude of the first harmonic components sin ωst and cos ωst, and change in lower frequency
than ωs . The small-signal modeling of the resonant converter focuses on the dynamics of the modulation
terms because the dynamics of the system is determined by the dynamics of the amplitude of the voltages
and currents.
To establish the state equations of the modulation terms, replace i1 , im , it , v1 , vt , vab , and vcd in (4.3)
and (4.8) with the corresponding FHA in (4.9) – (4.11), (4.13) – (4.15), and (4.17). And then, equate the
modulation terms of the sin ωst and cos ωst, respectively. The resultant state equations are
′ Lm +L2′
di1s 2 Lm +L2 Lm 2 Lm n2 (i1s −ims )
dt = ωs i1c + π Lµ vg − Lµ v1s − Lµ vts − π Lµ vo
i pk
Lm +L2′ 2 Lm n2 (i1c −imc )
di1c Lm
dt = −ω s i 1s − L µ
v1c − L µ
vtc − π Lµ i pk vo
′ L2′ 2 L1 n2 (i1s −ims )
dims 2 L2 L1
dt = ωs imc + π Lµ vg − Lµ v1s + Lµ vts + π Lµ vo
i pk
L2′ 2 L1 n2 (i1c −imc )
dimc L1
dt = −ω s i ms − L µ
v 1c + L µ
vtc + π Lµ i pk vo
dv1s i1s (4.18)
dt = ωs v1c + C1
dv1c
= −ωs v1s + Ci1c1
dt
dvts i1s ims
dt = ωs vtc + C2′ − C2′
dvtc i1c imc
dt = −ωs vts + C2′ − C2′
dvo = 1 i pk − vo
dt π Cf Ro C f
This set of differential equations can be represented in a matrix form in the following large-signal and small-
71
signal state equations.
Based on the differential equations (4.18), to solve for state variables at the operating point, replacing
each state variable x with X and set dX/dt to be zero. The steady-state equations can be simplified by
replacing the output voltage (Vo ) and the first harmonic component of secondary-side resonant peak current
(Ipk ) with the equivalent output load resistance. The equivalent load resistance is derived based on FHA as
√
2
Vcd1,rms = Vo . (4.19)
π
The output voltage (Vo ) is the product of the average output current (Io ) and output load resistor (Ro ), and Io
1
Io = Ipk Kbr , (4.20)
π
where Kbr is the coefficient that represents the effect of transformer decoupling time on the magnitude of the
envelope terms [85]. It equals one if the transformer is continuously coupled and is less than one if there are
where Ωs is the steady-state switching frequency, and Ωo is the resonant frequency. Therefore, the RMS
π
I2,rms = √ Io . (4.22)
2Kbr
According to (4.19) and (4.22), the equivalent load resistance Re is derived by dividing Vcd1,rms by I2,rms ,
which results in
2n2 Ro Kbr
Re = . (4.23)
π2
Note that Re is already reflected to the primary side, considering the turns ratio n. Thus, the term Vo /Ipk is
72
replaced by
Vo πRe
= 2. (4.24)
Ipk 2n
As a result, the steady-state equation is downsized to eight independent equations. Define Xst as the state
T
Xst = I1s I1c Ims Imc V1s V1c Vts Vtc , (4.25)
where each of the state variables represents the steady-state peak value. When the transformer is contin-
uously coupled, the peak values are not changed. If there are decoupling periods in one switching cycle,
depending on the steady-state switching frequency (Ωs ) and the resonant frequency (Ωo ), the average value
T
B st = 2 ′
π Vg (Lm + L2 ) 0 2 ′
π Vg L2 0 0 0 0 0 , (4.27)
The peak values of the state variables: I1s , I1c , Ims , Imc , V1s , V1c , Vts , and Vtc , can be derived by solving
(4.26) for Xst . The peak current in the secondary side (Ipk ) is solved by
q
Ipk = (I1s − Ims )2 + (I1c − Imc )2 , (4.29)
73
which leads to the solution of the average output current Io following (4.20). Finally, the steady-state output
Vo = Io Ro (4.30)
As a result, the voltage and current at the operating point can be solved with the given system parameters
along with the input voltage, operating frequency, and output load. This information is then applied to derive
The linearization of (4.18) is achieved by taking partial derivatives of each term with respect to each state
∂ ωs i1c ∂ ωs i1c ˆ ˆ,
d(ωs i1c ) = ω̂s + i1c = I1c ω̂s + Ωs i1c (4.31)
∂ ωs ∂ i1c
and
2n2 vo i1s − ims I2 ˆ I2 ˆ Its Itc ˆ Its Itc ˆ Re Its
d( ) = n2 Re ( 2tc i1s − 2tc ims − 2 i1c + 2 imc )+ vˆo (4.32)
π i pk Ipk Ipk Ipk Ipk Vo
ẋ = Aol x + Bol u
(4.33)
y = Cvo x,
T
ˆ i1c
x = i1s ˆ ims
ˆ imc
ˆ vˆ1s vˆ1c vˆts vˆtc vˆo
T
u = ω̂s vˆg (4.34)
y = vˆo .
74
The state matrix is defined as
Aol =
−Krm Itc Krm Itc −(Lm +L2′ ) −Lm −Krs Lm
Its Ωs + Krm Its −Krm Lµ 0 Lµ 0 Lµ
−(Lm +L2′ )
−Ω + K −Krm Its Krm Its −Lm −Krc Lm
s rm Itc −Krm Itc 0 Lµ 0 Lµ Lµ
Krl Itc −Krl Itc L2′ L1 Krs L1
−Krl Ωs + Krl 0 0
Its Its Lµ Lµ Lµ
−Krl Krl Its −Krl Its −L2′ L1 Krc L1
Itc −Ωs + Krl Itc 0 Lµ 0 Lµ Lµ
(4.35)
1
C1 0 0 0 0 Ωs 0 0 0
1
0 0 0 −Ωs 0 0 0 0
C1
1 −1
C2′ 0 C2′ 0 0 0 0 Ωs 0
1 −1
0 C2′ 0 C2′ 0 0 −Ωs 0 0
n2 Its n2 Itc −n2 Its −n2 Itc −1
πI pkC f πI pkC f πI pkC f πI pkC f 0 0 0 0 Ro C f
where
n2 Lm Its Itc Re Krm L1
Krm = 2
, Krl = (4.36)
Lµ Ipk Lm
where
T
Bωs = I1c −I1s Imc −Ims V1c −V1s Vtc −Vts 0 , (4.40)
T
2(Lm +L2′ ) 2L2′
Bvg = πLµ 0 πLµ 0 0 0 0 0 0 . (4.41)
For the study of control-to-output transfer function, the output matrix is selected as
Cvo = 0 0 0 0 0 0 0 0 1 . (4.42)
As a result, the transfer function of an open-loop, direct frequency-controlled CLLC resonant converter is
75
n2Its n2Itc
L1I1c ws WsL1 i1c v1s its L2’ Itc ws WsL2’ itc vts i i
L2’ p Ipk ts p Ipk tc
C1 Lm ims C2’ Rs
vg i1s L1 Cf Ro
p C1 V 1c ws C2’ Vtc ws
i1s LmImcws
Rtitc vo
2vg Ws C1v1c C2’ Wsvtc
p WsLmimc
Krsvo
Krcvo
Ws C1v1s WsLmims C2’ Wsvts
Rtits
L1 C1 V1s ws LmImsws C2’ Vts ws
i1c C1 imc C2’ Rc
Lm
L1 I1s ws WsL1 i1s v1c itc L2’ L2’ Its ws WsL2’ its vtc
Figure 4.3: Extended Describing Functions (EDF) model of an open-loop, direct frequency-controlled,
half-bridge CLLC resonant converter. Input current is proportional to the sine component of the
primary resonant current, which forms the basis of the proposed enhanced charge-control model.
derived:
vˆo
Gv f (s) := = Cvo (sI − Aol )−1 Bωs . (4.43)
ω̂s
The small-signal equivalent circuit of an open-loop half-bridge CLLC converter can be obtained by ana-
lyzing the linearized modulation equations (4.33). The result is shown in Fig. 4.3, where the inverter stage,
the resonant tank, and the rectifier stage are included. The resonant tank model shows that each of the
resonant inductor and capacitor is composed of three elements. The EDF model of a resonant inductor is
composed of the inductor in series with two voltage sources; one is coupled from the switching frequency,
and the other is coupled from the sine or cosine part of the inductor current. As for the EDF model of a
resonant capacitor, it consists of the capacitor and two current sources in parallel; one is coupled from the
switching frequency, and the other is coupled from the sine or cosine part of the capacitor voltage. The in-
verter stage model shows that the input current is proportional to the sine component of the primary resonant
current (i1s ). This information is essential to the proposed charge-control model that will be discussed in
Section 4.2.
procedure from the previous section. The next step is to derived the charge-controlled model by closing the
charge loop. Charge control for resonant converters has been discussed in [75–79], however, the small signal
76
ws vo ws vo
Converter Converter
Model irec Model i1s
vo
Figure 4.4: a) Schematic diagram of a CLLC resonant converter with charge mode control. b) Block
diagram of the charge loop: the conventional model approximates the charge loop by using irec ,
overlooking the effect of magnetizing current (im ). c) Conceptual illustration of the frequency
response of the charge loop: the proposed model removes the prediction error, which is useful for
high-bandwidth design.
model developed since then, including [97, 98] works only for the low-frequency range. As the advancement
in high-performance resonant converter accelerates, it is worthwhile to explore the model that predicts the
Charge mode control regulates the charges in each switching cycle. Fundamentally, charges are the
integration of current. In the context of resonant converters, it refers to the integration of the primary-
side resonant current. Regarding the implementation of the charge sensing circuit, the mainstream charge
controllers either use a resistive current sensor and a resettable integrator, or directly sense the resonant
capacitor voltage since the voltage is the integration of resonant current in a series resonant tank. In this
work, a commercial charge controller using the resonant capacitor voltage is employed for the analysis of
the small-signal modeling. As shown in Fig. 4.4 (a), the charge controller senses the resonant capacitor
voltage V1 and compares it with the charge command Vch to control the switches. The sawtooth signal vsaw
The charge command Vch limits the charges accumulated in every half-switching cycle. The switch turns
off when V1 reaches Vch , and it initiates the other complementary switch. The charge command comes from
77
the voltage loop controller; with the inner charge loop and outer voltage loop, the converter is controlled
by a dual loop, cascade control system. The existence of this inner charge loop changes the open-loop
model derived in Section 4.1, which changes the design of the voltage loop controller. The conventional
modeling method concludes an equivalent feedback loop from irec , as shown in Fig. 4.4 (b), which is counter
intuitive. The proposed model, on the other hand, models the charge feedback as i1s , the sine component of
the resonant current, which is related to the actual feedback loop in the schematic diagram. Figure 4.4 (c)
illustrates a conceptual presentation of typical frequency response with charge control. The conventional
model deviates from the measured response from the mid- to high-frequency range, whereas the proposed
This section investigates the modeling of the charge controlled resonant converter, with the focus on the
control-to-output (Vch to Vo ) transfer function. First, a review of the conventional model is given, followed
The conventional methodology for modeling charge-controlled resonant converters [77, 98] is based on en-
ergy balance of the system input and output. The procedure includes: 1) Defining system energy conservation
by equalizing the input and output power; 2) Replacing the average input current with the charge control in-
put and switching frequency; 3) Replacing the output current using the output voltage and impedance; 4)
Following this procedure, the transfer function from the control input to output voltage yields
vˆo KA Gv f (s)
= , (4.44)
vˆch 1 + KA KB Gv f (s)Gy (s)
where KA and KB are constants determined by the operating point, and Gy (s) is a transfer function related
to the output admittance. The transfer function Gv f (s) in [77] is replaced with the DC gain of the transfer
function, whereas in [98], the complete frequency-to-output transfer function is applied. The block diagram
that interprets the conventional model (4.44) is shown in Fig. 4.5. In this block diagram, an equivalent
ˆ represents the AC component of the rectifier current.
feedback loop is formed from the output, where irec
This equivalent charge feedback loop can be considered as an estimation of the actual charge feedback
signal, that is, the resonant capacitor voltage. However, it is not the full representation because the amplitude
of irec is proportional to it , which excludes im . According to Fig. 4.4 (a), the charge controller senses V1 as
the feedback signal, which is the integration of i1 , containing the information of im . Therefore, the effect of
78
vch ws vo
KA Gvf (s)
equivalent charge
feedback loop
irec
KB Gy(s)
Figure 4.5: Block diagram of the conventional charge model. The equivalent feedback loop from the
output does not fully represent the actual charge loop.
the magnetizing inductor current is overlooked. As a result, prediction errors occur in the frequency domain,
Since the essence of charge control is to regulate the accumulated resonant current, which is equal to the
charge stored in the resonant capacitor in every half switching cycle, the small-signal modeling methodol-
ogy proposed in this work focuses on linearizing the dynamics of these two parts. To simplify the problem,
this section first considers Bang-Bang Charge Control (BBCC), which has no slope compensation, and sec-
ond Hybrid-Hysteretic Control (HHC), which includes slope compensation. As illustrated in Fig. 4.6, the
where v1a is the attenuated resonant capacitor voltage; Katt v1a equals v1 . In a practical implementation, the
voltage sensing gain serves as the attenuation. The charge stored in C1 , or the accumulated i1 in each half
cycle, is controlled by ±vch . Referring to Fig. 4.6, let t0 be zero; the charge stored in C1 from zero to t equals
Z t
2Katt C1 vch = i1 (t)dt (4.46)
0
79
i1pk
i1
Ts / 2 Ts / 2 + t
vHS Ws Ws+ ws
t0 t1 t2 t3 t
Figure 4.6: Timing sequence diagram for charge control without slope compensation along with the
perturbation applied for linearization.
so (4.46) leads to
2Katt C1 vch =
−i1pk (4.48)
[cos(ωst) cos φ − sin(ωst) sin φ − cos φ ] ,
ωs
which is the nonlinear equation of charge control.
As the sine and cosine components of the resonant current intrinsically determines the phase angle φ ,
of which the relationship in time domain is illustrated in Fig. 4.7 (a). In the phasor diagram, as Fig. 4.7 (b)
shows,
iis iic
cos φ = , sin φ = , (4.50)
i1pk i1pk
and thus,
Next, to linearize (4.51), apply Taylor series expansion and neglect the higher order terms, which yields
Figure 4.6 shows the operating points (Vch , Ωs , Ts /2) and the perturbation of each variable (vˆch , ω̂s , tˆ). The
80
i1 i1pk ( sin wst f ) Im
i1ssin wst
i1s Re
f
t
i1c coswst
i1c
f i1pk
(a) (b)
Figure 4.7: Relationship between the sine and cosine components of the resonant current using Ex-
tended Describing Functions (EDF) in: a) time domain and b) complex plane.
As a result, the cosine term I1c is eliminated, and the large-signal part of (4.52) is
I1s
Ωs = , (4.54)
MVch
1 ˆ I1s
ω̂s = i1s − vˆ .
2 ch (4.55)
MVch MVch
Starting with the definition of the charge in the resonant tank, the small-signal modeling procedure in-
corporates EDF, phasor analysis, and linearization to derive the relationship between the charge command
and the frequency; these results are useful for both large and small-signal analyses of a charge-controlled
resonant converter. Taking the computed result of I1s from (4.26) into (4.54) yields the charge command
Vch at the designated operating point. This is essential for engineers to design and evaluate the controller in
open-loop tests.
Based on this analysis, the constructed charge-control model is related to frequency ω̂s , the sine compo-
ˆ , and the control input vˆch . To illustrate the block diagram of the proposed
nent of the resonant current i1s
ˆ:
model, first define the open-loop transfer function from ω̂s to i1s
ˆ
i1s
Gi1s f (s) := = Ci1s (sI − Aol )−1 Bωs , (4.56)
ω̂s
81
open-loop
converter model vo
Gvf (s)
using EDF
vch ws i1s
Fm Gi1sf (s)
Ri
(a)
vch ws i1s vo
Fm Gi1sf (s) Gvi1s(s)
(b)
Figure 4.8: Proposed block diagram of the small-signal model of a charge-controlled resonant con-
verter: a) direct b) cascade form. The feedback signal is the sine component of the primary
resonant current.
where
Ci1s = 1 0 0 0 0 0 0 0 0 . (4.57)
Combining with the converter open loop model derived using EDF, the block diagram of the proposed model
loop system model is completed by closing the vˆo feedback loop with an outer-loop controller that gives vˆch
command. The outer loop is not the focus of this paper; however, it will be shown in the next section for
validation purposes.
The small-signal equation (4.55) and the proposed charge control block diagram reveal that the charge
controller is essentially a proportional controller (P-controller), where the gain is defined by the modulation
gain (Fm ). Furthermore, the equivalent feedback signal is the sine component of the resonant current, even
though the charge control loop ostensibly feeds back the resonant capacitor voltage or the integrated resonant
current. The physical meaning of the sine component of the primary resonant current, based on the analysis
82
in Section 4.1, is a current that is proportional to the average input current. The derived block diagram of
the proposed charge-control model is therefore consistent with the essence of charge control: to control the
Having established the model for charge control without slope compensation, also known as Bang-Bang
Charge Control (BBCC), we can go one step further by adding slope compensation. In this case, also known
as the Hybrid-Hysteretic Control (HHC), a sawtooth signal vsaw is added to the sensed resonant capacitor
voltage, or, equivalently, subtracted from the charge command ±vch . As indicated in Fig. 4.9, the slope of
I1s πSe
Ωs = + , (4.60)
MVch 2Vch
1 ˆ I1s πSe
ω̂s = i1s − 2
+ 2 vˆch . (4.61)
MVch MVch 2Vch
Comparing (4.62) with (4.58), Se provides the ability to adjust the charge loop gain. Without slope
compensation, as in BBCC, the modulation gain and the current sensing gain are basically determined by
the system. In (4.58), Vch and I1s depend on the system operating conditions, and the gain M is determined
by the attenuation of the resonant capacitor voltage sensing. Therefore, the charge controller design for
83
vch
vch vsaw
vch Se
v1a Vch Vch vch
Se
vch
vch vsaw
Se
vsaw
Se
Ts 2 Ts 2 t
vHS Ws Ws+ ws
t0 t1 t2 t3 t
Figure 4.9: Timing sequence diagram for charge control with slope compensation along with the per-
turbation applied for linearization.
BBCC is almost fixed by the system parameters. With slope compensation, as in HHC, Se becomes the
parameter to adjust the charge controller. When Se equals zero, (4.54) − (4.58) are equivalent to (4.60) −
(4.62), respectively, so the former can be considered as a special case of the latter. When Se increases such
that Se ≫ 2I1s /πM, the fixed slope compensation (vsaw ) dominates the switching frequency, and the effect
of the charge control feedback (i1s ) is negligible. Hence, the controller becomes direct frequency control as
voltage mode. In this case, Fm is −Ωs /Vch , and the operating frequency Ωs is approximately πSe /2Vch . By
rearranging (4.60), it is easily derived that in a half switching cycle, the current-sense voltage is I1s /MΩs ,
while the slope compensation magnitude is πSe /2Ωs ; the ratio of these two components indicates whether
The full small-signal model considering the converter model can be derived by replacing ω̂s in (4.33)
with (4.61). As a result, the state-space equations for a charge-controlled resonant converter is defined as
ẋ = Ah x + Bh uh
(4.63)
y = Cvo x.
The state and the output vectors remain the same as (4.33), while the input vector becomes
uh = vˆch vˆg . (4.64)
84
The new state matrix derived as
where
−I1c 0 0 · · · 0 0
I
1s 0 0
−Imc
Ims
.. .. ..
Ah0 = Fm Ri
−V1c . . ., (4.66)
V1s
−Vtc
V
ts 0 0
0 0 0 ··· 0 0
where
Subsequently, the charge control-to-output transfer function for a charge-controlled resonant converter is
expressed as
vˆo
Gvc (s) := = Cvo (sI − Ah )−1 Bh0 . (4.69)
vˆch
ˆ feedback loop modifies the resonant tank model, while leaving the inverter and
The presence of the i1s
rectifier stages unaffected, remaining the same as the open-loop model. The equivalent circuit of the resonant
tank is illustrated in Fig. 4.10. With charge control, an equivalent resistor Rch appears in series with L1 and C1 ,
which implies that a damper is generated in the resonant circuit. The dependent voltage and current sources
containing the Fm or Ri coefficient are affected by Se if slope compensation is implemented. The lump
effects of introducing Rch and Fm -Ri dependent sources are difficult to see from the equivalent circuit model
alone but can be easily understood by plotting the frequency response of Gvc (s), which will be analyzed and
Although charge control with and without slope compensation, such as BBCC and HHC, has been pre-
sented before, the exact model of the feedback loop was not discovered. So far, a new mathematical model of
85
WsL1 i1c v1s its Fm L2’ Itc vch WsL2’ itc vts
FmRiL1 I1c L2’
Rch C1 C2’
FmL1 I1c vch Lm ims FmRiL2’ Itc i1s Rs
L1 FmRiC1V1ci1s FmRi C2’ Vtci1s
FmRiLmImc i1s
i1s
FmC1V1cvch Fm C2’ Vtc vch Rtitc
2vg FmLmImcvch
p
WsC1v1c WsC2’ vtc
WsLmimc Krsvo
WsLmims Krcvo
i1c WsC1v1s WsC2’ vts
L1 FmLmImsvch
FmC1V1svch FmC2’ Vtsvch Rtits
FmRiLmIms i1s
FmRi C1V1si1s FmRiC2’ Vtsi1s
FmRiL2’ Its i1s Rc
C1 Lm imc C2’
FmL1 I1s vch ( Ws FmRiI1s( L1 i1s v1c itc L2’ FmL2’ Itsvch WsL2’ its vtc
Figure 4.10: Proposed equivalent small-signal circuit of a charge-controlled CLLC converter: focused
on the resonant tank that is affected by the charge loop.
charge control is derived using EDF, phasor analysis, and linearization, and the full equivalent small-signal
circuit is proposed.
In the case of LLC resonant converters, for example, the half-bridge center-tapped LLC resonant converter
shown in Fig. 2.1, the same modeling procedure can be applied with some adjustments in the coefficients.
First, considering that for a center-tapped transformer, the reflected output voltage is nVo instead of nVo /2.
Second, in the differential equation (4.18), since the LLC resonant converter does not have the secondary-
side resonant elements, L2 and C2 , the parameter L2′ and the voltage across C2 are set to be zero. As a result,
86
di1s 2 1 1 4 1 n2 (i1s −ims )
dt = ωs i1c + π L1 vg − L1 v1s − π L1 vo
i pk
4 1 n2 (i1c −imc )
di1c 1
dt = −ωs i1s − L1 v1c − π L1 vo
i pk
4 1 n2 (i1s −ims )
dims
dt = ω s i mc + π Lm i pk vo
dimc 4 1 n2 (i1c −imc ) (4.70)
dt = −ωs ims + π Lm i pk vo
dv1s i1s
dt = ωs v1c + C1
dv1c i1c
dt = −ωs v1s + C1
dvo = 2 i pk − vo
dt π Cf Ro C f
It is clear that (4.70) shares the same form as (4.18) with reduced size (L2 and C2 eliminated). The
coefficients of vo and i pk are doubled due to the center-tapped configuration. Therefore, from the modeling
Following (4.70), the state vector in the steady state is defined as:
T
Xst = I1s I1c Ims Imc V1s V1c . (4.71)
Then, Xst can be derived by solving (4.26) with the steady-state input and state matrices for LLC converters,
defined as
T
st
B = 2
π Vg Lm 0 0 0 0 0 , (4.72)
Re −Ωs L1 −Re 0 1 0
Ω L
s 1 Re 0 −Re 0 1
−Re 0 Re −Ωs Lm 0 0
Ast = . (4.73)
0
R e −Ω L
s m −R e 0 0
1/C 0 0 0 0 Ωs
1
0 1/C1 0 0 −Ωs 0
8n2 Ro Kbr
Re = . (4.74)
π2
It is derived from the term 4n2 vo /(πi pk ) in (4.70), where Ipk is related with Io with a coefficient 2Kbr /π for
87
a center-tapped rectifier:
2 2nKbr
q
Io = Ipk Kbr = (I1s − Ims )2 + (I1c − Imc )2 , (4.75)
π π
For the small-signal state equations, similarly, it can be solved by linearizing (4.70) or simply apply-
ing zero L2 , vts , and vtc in the small-signal state equations of the CLLC converter. For the LLC resonant
T
ˆ i1c
x = i1s ˆ ims
ˆ imc
ˆ vˆ1s vˆ1c vˆo
T
u = ω̂s vˆg (4.76)
y = vˆo .
where
n2 Its Itc Re Krm L1
Krm = 2
, Krl = (4.78)
L1 Ipk Lm
Rt Itc Rt Its
Rt = Krm L1 , Rs = , Rc = , (4.79)
Its Itc
Its Re Itc Re
Krs = , Krc = . (4.80)
Vo Vo
88
where
T
Bωs = I1c −I1s Imc −Ims V1c −V1s 0 , (4.82)
T
Bvg = 2
πL1 0 0 0 0 0 0 . (4.83)
The results (4.77) – (4.83) are obtained by letting L2 , vts , and vtc in the small-signal state equations of the
CLLC converter (4.35) – (4.41) be zero, which is the same as taking the partial derivatives of (4.70) for
linearization. For the study of control-to-output transfer function, the output matrix is selected as
Cvo = 0 0 0 0 0 0 1 , (4.84)
and the transfer function Gv f of an open-loop, direct frequency-controlled LLC resonant converter can be
Based on the analysis of charge control loop in Section 4.2.2, the charge loop essentially takes i1s , the
sine component of the primary resonant current as the feedback signal. It means that the loop is closed in the
primary side, and the secondary-side circuit or configurations do not affect the modeling of the charge loop.
Therefore, the conclusions from Section 4.2.2 remains valid for a half-bridge LLC resonant converter. As a
result, the small-signal model of a charge-controlled half-bridge, center-tapped LLC resonant converter can
be derived following the procedure (4.45) – (4.69), with the state matrix for the LLC converter defined as:
where
−I 0 0 ··· 0 0
1c
I1s 0 0
−I
mc
Ah0 = Fm Ri Ims , (4.86)
.
−V1c .. . .. .
..
V1s
0 0 0 ··· 0 0
while Fm and Ri are the same as (4.58) or (4.62) for the case without or with slope compensation. Again,
the matrix Ah0 can also be derived by directly letting the Vts and Vtc terms in (4.66) be zero. Since the small-
89
signal model of the LLC resonant converter is a special case of the CLLC resonant converter, the following
simulation and experiments focus on the verification of the CLLC resonant converter.
with system parameters as shown in Table 4.1. This converter demonstrates power supply operation with
constant-voltage and constant-current regulation. With operating modes covering both below- and above-
resonance regions, this prototype is used for validation of the proposed model and demonstration of high-
bandwidth design. As shown in Fig. 4.11, the control-to-output frequency response, Gvc (s), of the proposed
model is compared with the conventional model and the circuit simulation as a tool for validation.
For operation above the resonant frequency, shown in Fig. 4.11 (a) and (b), the converter is biased at the
maximum current. The test condition for Fig. 4.11 (a) is the minimum output voltage with the maximum
switching frequency, while the condition for Fig. 4.11 (b) is around the nominal output voltage. The oper-
ating conditions in Fig. 4.11 (c) and (d) respectively correspond to full-load and light-load operation at the
maximum output voltage, where the switching frequencies are below resonance. Across all the operating
conditions, estimation errors appear in the conventional model; the red areas highlight the errors under the
frequency of concern. On the other hand, the proposed model significantly reduces the discrepancies, which
Parameter Value
input voltage, Vin 400 V
nominal power, Pout 1 kW
maximum output current, Imax 3.3 A
nominal output voltage, Vnom 330 V
maximum output voltage, Vmax 400 V
switching frequency, fsw 155 − 180 kHz
primary resonant inductance, L1 45 µH
primary resonant capacitor, C1 20 nF
magnetizing inductance, Lm 36 µH
secondary resonant inductance, L2 31 µH
secondary resonant capacitor, C2 29 nF
output capacitor, C f 1 µF
transformer turns ratio, n 1.2
voltage attenuation, Katt 303
slope compensation, Se 0.184 V/µs
90
40 overestimated 40
gain
Gain [dB]
Gain [dB]
20 20
0 0
Conventional Model Conventional Model
Proposed Model Proposed Model
-20 Simulation -20 Simulation
0 0
Phase [degree]
Phase [degree]
-90 -90
40 40 underestimated
gain
Gain [dB]
20 20 Gain [dB]
0 0
Conventional Model Conventional Model
Proposed Model Proposed Model
-20 Simulation -20 Simulation
0 0
Phase [degree]
Phase [degree]
-90 -90
Figure 4.11: Comparison of the conventional and proposed analytical models (charge control-to-output
transfer function, Gvc (s)) with circuit simulation results at a) 234 V / 3.3 A; b) 320 V / 3.3 A;
c) 400 V / 2.5 A; d) 400 V / 0.8 A. Estimation errors from the conventional model render it
unsuitable for control-loop analyses especially in a high-bandwidth design. The proposed model
successfully predicts the frequency response across all the conditions.
The frequency response shows that from the mid-frequency range and above, the conventional model
starts to deviate from both the proposed model and the circuit simulation results. This behavior is evident in
the above-resonance conditions seen in Fig. 4.11 (a) and (b). These discrepancies hinder determination of the
crossover frequency and phase margin for the outer loop. In the light-load condition, shown in Fig. 4.11 (d),
the proposed model successfully predicts the resonant peak caused by the complex pole. This resonant peak
is particularly important for a high-bandwidth design as it inevitably constrains the gain margin of the outer
91
40
10dB
20
Gain [dB]
Se = Se0 10dB
9 Se0
0 27Se0 wc.max wc.max
81Se0 Se = Se0 Se = 27Se0
-20
-270
-360
10 2 10 3 10 4 10 5
Frequency [Hz]
Figure 4.12: The effect of slope compensation on the charge loop gain Gvc (s) predicted by the proposed
model assists rapid selection of slope compensation.
loop. Overall, the proposed model successfully predicts the charge loop response, providing an analytical
The proposed analytical model is useful for quickly determining the slope compensation. By plotting the
frequency response with a set of slope compensation Se , the appropriate value can be selected, facilitating
the subsequent design of the outer loop compensation. Performing a parameter sweep is time-consuming for
circuit simulation, whereas by applying the proposed analytical model, it can be done in a few seconds.
Applying the proposed model, Fig. 4.12 shows a set of Se at light-load operation, where the high-
frequency complex pole is the main constraint for a high-bandwidth design. The base slope compensation
Se0 is selected as 0.184 V/µs in this case study. As Se increases, the system becomes closer to a direct
frequency-controlled system as predicted by the model, where the dominant pole and the second pole move
towards each other and tend to form a low-frequency resonant peak: around 6 kHz in this case.
Proper slope compensation design suppresses the high-frequency resonant peak using the second pole.
As shown in Fig. 4.12, selecting 27Se0 can potentially achieve higher bandwidth than the case with Se0 ,
considering a PI compensator for the outer loop. To achieve higher bandwidth for designs with low values of
slope compensation, a high-frequency pole can be considered to suppress the resonant peak. This prediction
cannot be done by the conventional model, as it does not predict the complex pole in the high-frequency
range. The proposed analytical model facilitates rapid selection of an appropriate slope compensation for
92
vHS Cr1 Cr2
Vg L1
Cf Ro
vLS Lm L2 vo
Cr1
Cr2 Rcs
Cd1
Cd2 Cp1
Vss Cz2
R1
Charge Rb Cz1 Rv
Controller
UCC
256302
1.2 V R2
0.2 V
R3
Cz3 Ri
Cp2 Cz4
Figure 4.13: Simplified schematic diagram and key components of the test setup for verifying the pro-
posed model using a commercial charge controller.
In order to verify the proposed model, a 400 V / 1 kW prototype is built based on the system parameters
shown in Table 4.1. The charge controller employed is TI UCC256302; this device has slope compensation
for the charge loop, which is suitable for verifying the proposed model. The selection of the operational
amplifiers for closing the outer loop is flexible; ST TSM1052 is applied in this paper as an example. Table 4.2
lists the key component values. Figure 4.13 illustrates the simplified schematic diagram, where the outer
loop is closed with two operational amplifiers for constant voltage and current regulation, respectively. Both
constant voltage and current regulation are included to validate the proposed model at operating modes below
The charge loop compensation is determined by the capacitive voltage divider formed by Cd1 and Cd2 .
The ratio determines the attenuation of resonant capacitor voltage sensing, Katt , whereas Cd1 and the internal
sourcing current of the device determine the slope compensation. Therefore, Cd1 is selected considering the
93
20
Gain [dB]
0
-20
-40
0
Phase [degree]
-90
20
Gain [dB]
-20
-40
0
Phase [degree]
-90
234 V/ 3.3 A Experimental
-180
234 V/ 3.3 A Analytical
320 V/ 3.3 A Experimental
-270
320 V/ 3.3 A Analytical
-360
5 x 10 2 10 3 10 4 5 x10 4
Frequency [Hz]
(b)
Figure 4.14: Analytical and experimental loop frequency response under a) constant-voltage and b)
constant-current regulation; the analytical results using the proposed model closely match the
experimental measurements.
effect of slope compensation shown in Fig. 4.12, and Cd2 is selected following Katt .
The outer loop compensator selection depends on individual design objectives. The proposed model,
however, successfully predicts the frequency response and can be integrated into the design procedure to
achieve high-bandwidth design, improving the dynamic performance. In this work, the compensator for the
constant-voltage regulation is designed based on the frequency response shown in Fig. 4.11 (c) and (d). In
particular, the resonant peak at light load causes a constraint in the gain margin, limiting the bandwidth. The
As a result, the outer loop compensator is designed based on the proposed model, and a PI with lead
compensator is chosen for a high-bandwidth design. Along with the model of the outer loop compensator
and the proposed model for the charge loop, the loop gain is plotted and compared with the experimen-
94
vo (5V/ Div.)
1 ms
io (1A/ Div.)
(a)
vo (5V/ Div.)
1 ms
io (1A/ Div.)
(b)
Figure 4.15: Experimental transient response under closed-loop regulation: a) 100 % – 80 % load; b)
60 % – 40 % load. The time domain response demonstrates stable operation, as predicted in the
frequency domain.
tal results. Figure 4.14 compares the analytical and experimental frequency response in constant-voltage
and constant-current regulation that respectively correspond to below- and above-resonance operation; the
proposed model successfully predicts the loop gain in both operating modes. Figure 4.15 shows the time
domain transient response of the converter under closed-loop regulation. This result demonstrates that the
closed-loop regulation is stable as predicted by the model. Therefore, the proposed model is practical for the
4.4 Summary
In this chapter, a new, enhanced analytical model for a charge-controlled resonant converter is proposed. To
derive the small-signal model, the integration of the resonant current is first described using the Extended
Describing Functions (EDF). Then, through phasor analyses, the phase angle of the resonant current is
transformed, which enables linearization. The proposed model reveals that the charge controller essentially
controls the sine component of the primary resonant current, which is proportional to the average input
95
current. The proposed charge-control model successfully predicts the frequency response across multiple
operating modes, including the high-frequency complex pole neglected by the conventional model, enabling
a high-bandwidth control loop design. The frequency response is plotted without needing to run a time-
consuming parameter sweep with circuit simulation tools. The loop gain predicted by the proposed model
is validated by experimental results from a 400 V / 1 kW CLLC resonant converter, and the results are
consistent. Lastly, the dynamic load test results demonstrate that the converter under closed-loop regulation
is stable as predicted. Therefore, the proposed analytical model improves the conventional one, providing
practical and rapid frequency-domain evaluation for the design of charge control.
96
Chapter 5
Conclusion
for improving the efficiency. Synchronous Rectification (SR) is an effective solution to reducing the recti-
fier conduction loss, particularly for high-current power converters. Conventional vds.on based methods are
sensitive to parasitic effects, which results in more SR on-time error and reduced efficiency performance.
In terms of the system dynamic performance, charge-mode control has been established as a replacement of
current-mode control for resonant converters. However, the existing model only works for the low-frequency
range, which opens a window for the research in a more complete model for high-bandwidth design.
This work focuses on the efficiency and dynamics improvement. It proposes two advanced SR solutions
to LLC resonant converters and a small signal modeling methodology for charge-controlled resonant con-
verters. The Resonant Capacitor Voltage based SR (RCV SR) is developed by analyzing the relationships
of the resonant capacitor voltage and the input and output voltages. By comparing the voltage integrations,
the SR on-time can be determined. The RCV SR covers below- and above-resonance regions as well as
Next, the Volt-Second Product based SR (VSP) is proposed. The VSP SR algorithm determines the SR
on-time based on the product of rectifier current conduction time and the blocking voltage of SR MOSFETs.
The VSP SR tracks the rectifier current conduction time from the previous cycle in steady-state while main-
taining a safe margin during the transient. Both the RCV and VSP SR are large-magnitude voltage based and
have less on-time error and better efficiency performance, compared with conventional vds.on based methods.
Regarding the system dynamics aspect, the proposed small-signal dynamic model for charge-controlled
97
resonant converters concludes that the equivalent charge feedback signal is the sine component of the reso-
nant current, which is essentially the average input current. The average input current contains the dynamics
of the magnetizing current, and yet it is neglected by the conventional model, leading to prediction error.
The proposed dynamic model takes that part into consideration and successfully predicts the frequency re-
sponse from the low-frequency to high-frequency regions, which enables high-bandwidth design, improving
1. Large-Magnitude Voltage Based, Noise-Tolerant SR: Both of the proposed RCV and VSP SR are
based on large-magnitude voltages as opposed to the conventional vds.on based methods. Since the sensed
voltages are much larger than the voltage drops on the parasitic components, both proposed methods are
insensitive to the parasitic effects. As a result, the SR on-time error caused by the parasitic effects is removed,
improving the efficiency. The proposed SR and the conventional vds.on based SR methods are tested on a
650W/ 24V battery charger. Compared with the vds.on based non-adaptive SR, the RCV SR reduces the loss
by 6 % in the full-load condition. The VSP SR is further compared with the advanced, adaptive version of
the vds.on based SR and shows loss reduction by 3 % in the full-load condition.
2. Simplified, Highly-Integrated SR: The proposed VSP SR offers a simple solution to LLC SR using
the output stage for communication with the battery management system, and potentially, to function as
the closed-loop controller. With only a few basic logic gates, the VSP SR can be easily integrated into the
microcontroller, achieving a highly-integrated design. In addition, VSP SR has less printed-circuit board
(PCB) layout constraints. Conventional vds.on based SR controllers are supposed to be placed as close to the
SR MOSFETs as possible to reduce the stray inductance in the voltage sensing loop. The proposed VSP SR
senses the blocking voltage of the SR MOSFET simply using a pair of voltage dividers, which makes the
small-signal modeling methodology based on Extended Describing Functions (EDF) and phasor analysis
is applicable to both CLLC and LLC resonant converters. This model not only successfully predicts the
frequency response, but also reveals the essence of the charge loop. This part of the work concludes that
the charge loop controls the sine component of the resonant current, which is essentially the average input
current. As a result, the charge loop has two split real poles in the mid- to low-frequency range, and their
positions are affected by the slope compensation. For a high-bandwidth design, the effect of the second
98
pole must be taken into consideration. This enhanced small-signal model predicts the frequency response
very well from the low-frequency to high-frequency regions, and thus is useful for a high-bandwidth design,
The research outcomes of this work have been published in IEEE Transactions on Power Electronics and
• J. -D. Hsu, M. Ordonez, W. Eberle, M. Craciun and C. Botting, “LLC Synchronous Rectification
Using Resonant Capacitor Voltage,” in IEEE Transactions on Power Electronics, vol. 34, no. 11, pp.
• J. -D. Hsu, M. Ordonez, W. Eberle, M. Craciun and C. Botting, “Noise-Tolerant LLC Synchronous
Rectification Using Volt-Second Product,” in IEEE Journal of Emerging and Selected Topics in Power
• J. -D. Hsu, M. Ordonez, W. Eberle, M. Craciun and C. Botting, “Enhanced Small-Signal Modeling
for Charge-Controlled Resonant Converters,” in IEEE Transactions on Power Electronics, vol. 37, no.
signals as well as established the fundamentals of small signal modeling for charge-controlled resonant
converters. The methodologies cover both steady-state and transient analyses for resonant converters. The
• Integrated VSP SR and charge-mode controller: In Chapter 3, this work has demonstrated the in-
located on the secondary side. It is possible to integrate the charge-loop compensator with the same
microcontroller. The challenge is to pass the AC component of the resonant capacitor voltage to the
secondary side. For a half-bridge topology, since the resonant capacitor voltage contains a DC com-
ponent, a pulse transformer with a DC blocking capacitor can be used to transfer the AC component.
The transferred resonant capacitor voltage on the secondary side is then compared with the output of
the voltage loop compensator to control the primary switches. Details about the operating sequence
99
• SR for bidirectional resonant converters: Bidirectional resonant converters are popular topologies
for high-efficiency energy recycling or backward power transfer. Due to the active switching devices
on both the primary and the secondary sides, it is reasonable to enable SR to improve the efficiency
performance. The proposed VSP SR can be the first candidate for its simplicity. The operating prin-
ciples of VSP SR for bidirectional resonant converters such as CLLC converters are similar to those
of LLC resonant converters. Hence, VSP SR for bidirectional resonant converters can be a potential
research topic. For charge-controlled bidirectional resonant converters, since the resonant capacitor
voltages from both sides and the input and output voltages are all available, RCV SR can be consid-
ered to make the best use of those signals. The resettable integrators to perform RCV SR can be shared
by both sides to maximize the utilization. Detailed control logics need further investigation.
• Charge control for three-phase resonant converters: Three-phase resonant converters are known
for its capacity for high-power, high-voltage applications. By interleaving the three phases, the output
current ripple can be reduced, which is beneficial to the power density. However, due to the unavoid-
able component tolerance, the three-phase currents are rarely perfectly balanced, which deteriorates
the performance. Since charge control intrinsically controls the average input current, potentially it
can achieve current balancing. The phase-shift of the three legs can be adjusted according to the charge
feedback. The impact of the phase shift on the dynamics of the charge-controlled three-phase resonant
100
Bibliography
[1] J.-D. Hsu, M. Ordonez, W. Eberle, M. Craciun, and C. Botting, “LLC synchronous rectification
using resonant capacitor voltage,” IEEE Trans. Power Electron., vol. 34, no. 11, pp. 10 970–10 987,
2019. → page v
[2] J.-D. Hsu, M. Ordonez, W. Eberle, M. Craciun, and C. Botting, “Noise-tolerant LLC synchronous
rectification using volt-second product,” IEEE J. Emerging Sel. Top. Power Electron., pp. 1–1, 2022.
[3] J.-D. Hsu, M. Ordonez, W. Eberle, M. Craciun, and C. Botting, “Enhanced small-signal modeling for
charge-controlled resonant converters,” IEEE Trans. Power Electron., vol. 37, no. 2, pp. 1736–1747,
2022. → page v
[4] C. Shi, H. Wang, S. Dusmez, and A. Khaligh, “A sic-based high-efficiency isolated onboard pev
charger with ultrawide dc-link voltage range,” IEEE Trans. Ind. Appl., vol. 53, no. 1, pp. 501–511,
2017. → page 1
[5] H. V. Nguyen, D.-D. To, and D.-C. Lee, “Onboard battery chargers for plug-in electric vehicles with
dual functional circuit for low-voltage battery charging and active power decoupling,” IEEE Access,
vol. 6, pp. 70 212–70 222, 2018.
[6] R. Hou and A. Emadi, “Applied integrated active filter auxiliary power module for electrified
vehicles with single-phase onboard chargers,” IEEE Trans. Power Electron., vol. 32, no. 3, pp.
1860–1871, 2017. → page 1
[7] J. Gupta, R. Kushwaha, and B. Singh, “Improved power quality transformerless single-stage
bridgeless converter based charger for light electric vehicles,” IEEE Trans. Power Electron., vol. 36,
no. 7, pp. 7716–7724, 2021. → page 1
[8] J.-Y. Lee, Y.-D. Yoon, and J.-W. Kang, “A single-phase battery charger design for lev based on dc-src
with resonant valley-fill circuit,” IEEE Trans. Ind Electron., vol. 62, no. 4, pp. 2195–2205, 2015. →
page 1
[9] R. L. Steigerwald, “A comparison of half-bridge resonant converter topologies,” IEEE Trans. Power
Electron., vol. 3, no. 2, pp. 174–182, Apr. 1988. → page 2
[12] B. Yang, F. C. Lee, A. J. Zhang, and G. Huang, “LLC resonant converter for front end DC/DC
conversion,” in Proc. IEEE Appl. Power Electron. Conf. Expo., vol. 2, pp. 1108–1112, 2002. → page
2
101
[13] F. Musavi, M. Craciun, D. S. Gautam, W. Eberle, and W. G. Dunford, “An LLC resonant DC-DC
converter for wide output voltage range battery charging applications,” IEEE Trans. Power Electron.,
vol. 28, no. 12, pp. 5437–5445, Dec. 2013.
[14] X. Sun, Y. Shen, Y. Zhu, and X. Guo, “Interleaved boost-integrated LLC resonant converter with
fixed-frequency PWM control for renewable energy generation applications,” IEEE Trans. Power
Electron., vol. 30, no. 8, pp. 4312–4326, Aug. 2015. → page 2
[15] D. Kim, M. Kim, and B. Lee, “An integrated battery charger with high power density and efficiency
for electric vehicles,” IEEE Trans. Power Electron., vol. 32, no. 6, pp. 4553–4565, Jun. 2017. →
page 2
[16] Z. Zhao, Q. Xu, Y. Dai, and A. Luo, “Minimum resonant capacitor design of high-power LLC
resonant converter for comprehensive efficiency improvement in battery charging application,” IET
Power Electron., vol. 11, no. 11, pp. 1866–1874, 2018.
[17] R. Pandey and B. Singh, “A power-factor-corrected LLC resonant converter for electric vehicle
charger using cuk converter,” IEEE Trans. Ind. Appl., vol. 55, no. 6, pp. 6278–6286, Nov. 2019.
[18] N. Shafiei, M. Ordonez, M. Craciun, C. Botting, and M. Edington, “Burst mode elimination in
high-power LLC resonant battery charger for electric vehicles,” IEEE Trans. Power Electron.,
vol. 31, no. 2, pp. 1173–1188, Feb. 2016. → page 2
[19] M. Li, Q. Chen, X. Ren, Y. Zhang, K. Jin, and B. Chen, “The integrated LLC resonant converter
using center-tapped transformer for on-board EV charger,” in Proc. IEEE Energy Convers. Congr.
Expo., pp. 6293–6298, 2015. → page 2
[20] C. Duan, H. Bai, W. Guo, and Z. Nie, “Design of a 2.5-kW 400/12-V high-efficiency DC/DC
converter using a novel synchronous rectification control for electric vehicles,” IEEE Trans. Transp.
Electrific., vol. 1, no. 1, pp. 106–114, Jun. 2015.
[21] X. Wang, C. Jiang, B. Lei, H. Teng, H. K. Bai, and J. L. Kirtley, “Power-loss analysis and efficiency
maximization of a silicon-carbide MOSFET-based three-phase 10-kW bidirectional EV charger
using variable-DC-bus control,” IEEE J. Emerging Sel. Top. Power Electron., vol. 4, no. 3, pp.
880–892, Sep. 2016. → page 2
[22] M. Jovanovic, M. Zhang, and F. Lee, “Evaluation of synchronous-rectification efficiency
improvement limits in forward converters,” IEEE Trans. Ind. Electron., vol. 42, no. 4, pp. 387–395,
1995. → page 3
[23] Honglin Pan, Y. C. Liang, and R. Oruganti, “Design of smart power synchronous rectifier,” IEEE
Trans. Power Electron., vol. 14, no. 2, pp. 308–315, Mar. 1999.
[24] Y. Zhang, Y. Gao, L. Zhou, and M. Sumner, “A switched-capacitor bidirectional DC-DC converter
with wide voltage gain range for electric vehicles with hybrid energy sources,” IEEE Trans. Power
Electron., vol. 33, no. 11, pp. 9459–9469, Nov. 2018. → page 3
[25] K. Yao, Y. Ren, and F. Lee, “Critical bandwidth for the load transient response of voltage regulator
modules,” IEEE Trans. Power Electron., vol. 19, no. 6, pp. 1454–1461, 2004. → page 3
[26] Y. Panov and M. Jovanovic, “Small-signal analysis and control design of isolated power supplies
with optocoupler feedback,” IEEE Trans Power Electron, vol. 20, no. 4, pp. 823–832, 2005.
[27] E. Figueres, G. Garcera, J. Benavent, M. Pascual, and J. Martinez, “Adaptive two-loop voltage-mode
control of DC-DC switching converters,” IEEE Trans. Ind. Electron., vol. 53, no. 1, pp. 239–253,
2006. → page 3
102
[28] C. W. Deisch, “Simple switching control method changes power converter into a current source,” in
Proc. IEEE Power Electron. Spec. Conf., pp. 300–306, 1978. → pages 3, 13
[29] R. B. Ridley, “A new, continuous-time model for current-mode control (power convertors),” IEEE
Trans. Power Electron., vol. 6, no. 2, pp. 271–280, 1991.
[30] J. Leyva-Ramos and J. A. Morales-Saldana, “A design criteria for the current gain in
current-programmed regulators,” IEEE Trans. Ind. Electron., vol. 45, no. 4, pp. 568–573, 1998. →
pages 3, 13
[31] W. Tang, F. C. Lee, R. B. Ridley, and I. Cohen, “Charge control: modeling, analysis, and design,”
IEEE Trans. Power Electron., vol. 8, no. 4, pp. 396–403, 1993. → pages 3, 11
[32] F. Huliehel, W. Tang, F. Lee, and B. Cho, “Modeling, analysis, and design of the quasi-charge
control,” IEEE Trans. Power Electron., vol. 10, no. 5, pp. 597–604, 1995.
[33] V. Vorperian, “The charge-controlled PWM switch,” in Proc. IEEE Power Electron. Spec. Conf.,
vol. 1, pp. 533–542 vol.1, 1996. → pages 3, 11
[35] G. K. Y. Ho, R. Yu, and B. M. H. Pong, “Current driven synchronous rectifier with saturable current
transformer and dynamice gate voltage control for LLC resonant converter,” in Proc. IEEE Appl.
Power Electron. Conf. and Expo., pp. 2345–2351, 2012. → page 4
[36] X. Xie, J. C. P. Liu, F. N. K. Poon, and M. H. Pong, “A novel high frequency current-driven
synchronous rectifier applicable to most switching topologies,” IEEE Trans. Power Electron., vol. 16,
no. 5, pp. 635–648, Sep. 2001. → page 4
[37] X. Guo, W. Lin, and X. Wu, “A novel current driven method for center-tapped synchronous rectifier,”
in Proc. IEEE ICPE-ECCE Asia, pp. 449–454, 2010. → page 4
[38] G. Zhang, J. Zhang, Z. Chen, X. Wu, and Z. Qian, “LLC resonant DC/DC converter with
current-driven synchronized voltage-doubler rectifier,” in Proc. IEEE Energy Convers. Congr. Expo.,
pp. 744–749, 2009. → page 4
[39] J. Zhang, J. Wang, G. Zhang, and Z. Qian, “A hybrid driving scheme for full-bridge synchronous
rectifier in llc resonant converter,” IEEE Trans. Power Electron., vol. 27, no. 11, pp. 4549–4561,
Nov. 2012.
[40] J. Zhang, J. Wang, G. Zhang, and Z. Qian, “A hybrid driving scheme for full-bridge synchronous
rectifier in LLC resonant converter,” IEEE Trans. Power Electron., vol. 27, no. 11, pp. 4549–4561,
Nov. 2012. → page 4
[41] G. K. Y. Ho, R. Yu, and B. M. H. Pong, “Current driven synchronous rectifier for LLC resonant
converter with a novel integrated current transformer,” in Proc. IET Int. Conf. Power Electron.,
Machines and Drives, pp. 1–5, Mar. 2012. → page 4
[42] X. Wu, B. Li, Z. Qian, and R. Zhao, “Current driven synchronous rectifier with primary current
sensing for LLC converter,” in Proc. IEEE Energy Convers. Congr. Expo., pp. 738–743, 2009. →
page 5
[43] X. Wu, G. Hua, J. Zhang, and Z. Qian, “A new current-driven synchronous rectifier for series-parallel
resonant LLC DC-DC converter,” IEEE Trans. Ind. Electron., vol. 58, no. 1, pp. 289–297, Jan. 2011.
103
[44] B.-C. Kim, H.-S. Park, S. C. Moon, Y.-D. Kim, D.-Y. Kim, and G. W. Moon, “The novel
synchronous rectifier driving method for LLC series resonant converter,” in Proc. IEEE IECON, pp.
810–813, 2012.
[45] C. Zhao, B. h. Li, J. Cao, Y. Chen, X. Wu, and Z. Qian, “A novel primary current detecting concept
for synchronous rectified LLC resonant converter,” in Proc. IEEE Energy Convers. Congr. Expo., pp.
766–770, 2009. → page 5
[47] STMicroelectronics, “SRK2000A: Synchronous rectifier smart driver for LLC and resonant
converters ,” STMicroelectronics, Geneva, Switzerland, May. 2017, [Online]. Available:
st.com/resource/en/datasheet/srk2000a.pdf.
[49] A. Lokhandwala, M. Salato, and M. Soldano, “Dual SmartRectifier and DirectFET chipset
overcomes package source inductance effects and provides accurate sensing for synchronous
rectification in DC-DC resonant converters,” in Proc. IEEE Appl. Power Electron. Conf. Expo., pp.
1559–1562, 2007. → page 6
[50] D. Fu, Y. Liu, F. C. Lee, and M. Xu, “A novel driving scheme for synchronous rectifiers in LLC
resonant converters,” IEEE Trans. Power Electron., vol. 24, no. 5, pp. 1321–1329, May. 2009. →
page 6
[51] D. Wang, L. Jia, J. Fu, Y. Liu, and P. C. Sen, “A new driving method for synchronous rectifiers of
LLC resonant converter with zero-crossing noise filter,” in Proc. IEEE Energy Convers. Congr.
Expo., pp. 249–255, 2010. → page 7
[52] D. Wang and Y. Liu, “A zero-crossing noise filter for driving synchronous rectifiers of LLC resonant
converter,” IEEE Trans. Power Electron., vol. 29, no. 4, pp. 1953–1965, Apr. 2014. → page 7
[53] W. Feng, D. Huang, P. Mattavelli, D. Fu, and F. C. Lee, “Digital implementation of driving scheme
for synchronous rectification in LLC resonant converter,” in Proc. IEEE Energy Convers. Congr.
Expo., pp. 256–263, 2010. → page 7
[54] W. Feng, F. C. Lee, P. Mattavelli, and D. Huang, “A universal adaptive driving scheme for
synchronous rectification in LLC resonant converters,” IEEE Trans. Power Electron., vol. 27, no. 8,
pp. 3775–3781, Aug. 2012.
[55] F. Wang, B. A. McDonald, J. Langham, and B. Fan, “A novel adaptive synchronous rectification
method for digitally controlled LLC converters,” in Proc. IEEE Appl. Power Electron. Conf. Expo.,
pp. 334–338, 2016. → page 7
[56] Q. Qian, S. Xu, J. Yu, W. Sun, and H. Li, “A digital detecting method for synchronous rectification
based on dual-verification for LLC resonant converter,” in Proc. IEEE Appl. Power Electron. Conf.
Expo., pp. 2091–2097, 2018. → page 7
[57] M. Sato, G. M. Dousoky, and M. Shoyama, “Improved digital control scheme of synchronous
rectification for resonant converter at light load conditions,” in Proc. IEEE Int. Telecommun. Energy
Conf., pp. 1–5, 2015. → page 7
104
[58] ON Semiconductor, “FAN6248HC/HD/LC/LD: Advanced synchronous rectifier controller for LLC
resonant converter,” ON Semiconductor, Phoenix, Arizona, US, Jan. 2018, [Online]. Available:
http://www.onsemi.com/pub/Collateral/FAN6248HC-D.PDF. → page 7
[59] STMicroelectronics, “SRK2001: Adaptive synchronous rectification controller for LLC and resonant
and converter ,” STMicroelectronics, Geneva, Switzerland, Feb. 2017, [Online]. Available:
st.com/resource/en/datasheet/srk2001.pdf.
[60] S. Moon, C. Chen, and R.-J. Wang, “A new dead time regulation synchronous rectification control
method for high efficiency LLC resonant converters,” IEEE Trans. Power Electron., vol. 36, no. 9,
pp. 10 673–10 683, 2021. → page 7
[61] Texas Instruments, “UCD3138: Highly integrated digital controller for isolated power,” Texas
Instruments, Dallas, Texas, US, Mar. 2012, [Online]. Available:
http://www.ti.com/lit/ds/symlink/ucd3138.pdf. → page 8
[62] C. Fei, F. C. Lee, and Q. Li, “Digital implementation of adaptive synchronous rectifier (SR) driving
scheme for LLC resonant converters,” in Proc. IEEE Appl. Power Electron. Conf. Expo., pp.
322–328, 2016.
[63] M. S. Amouzandeh, B. Mahdavikhah, A. Prodic, and B. McDonald, “Digital synchronous
rectification controller for LLC resonant converters,” in Proc. IEEE Appl. Power Electron. Conf.
Expo., pp. 329–333, 2016. → page 8
[64] J. Wang and B. Lu, “Open loop synchronous rectifier driver for LLC resonant converter,” in Proc.
IEEE Appl. Power Electron. Conf. and Expo., pp. 2048–2051, 2013. → page 10
[65] S. Abe, S. Yang, M. Shoyama, T. Zaitsu, J. Yamamoto, S. Ueda, and T. Ninomiya, “Adaptive driving
of synchronous rectifier for LLC converter without signal sensing,” in Proc. IEEE Appl. Power
Electron. Conf. Expo., pp. 1370–1375, 2013. → page 10
[66] M. Mohammadi and M. Ordonez, “Synchronous rectification of LLC resonant converters using
homopolarity cycle modulation,” IEEE Trans. Ind. Electron., vol. 66, no. 3, pp. 1781–1790, Mar.
2019. → page 10
[67] H. Choi, “Dual edge tracking control for synchronous rectification (SR) of LLC resonant converter,”
in Proc. IEEE Appl. Power Electron. Conf. Expo., pp. 15–20, 2015. → page 10
[68] B. Li, M. Chen, X. Wang, N. Chen, X. Sun, and D. Zhang, “An optimized digital synchronous
rectification scheme based on time-domain model of resonant CLLC circuit,” IEEE Trans. Power
Electron., vol. 36, no. 9, pp. 10 933–10 948, 2021. → page 10
[69] H. Li, S. Wang, Z. Zhang, J. Zhang, M. Li, Z. Gu, X. Ren, and Q. Chen, “Bidirectional synchronous
rectification on-line calculation control for high voltage applications in sic bidirectional llc portable
chargers,” IEEE Trans. Power Electron., vol. 36, no. 5, pp. 5557–5568, 2021.
[70] X. Zhu, H. Li, Z. Zhang, Z. Li, Y. Yang, Z. Gu, S. Wang, X. Ren, and Q. Chen, “A sensorless
model-based digital driving scheme for synchronous rectification in 1-kv input 1-MHz GaN LLC
converters,” IEEE Trans. Power Electron., vol. 36, no. 7, pp. 8359–8369, 2021. → page 10
[71] J. Sun, “Characterization and performance comparison of ripple-based control for voltage regulator
modules,” IEEE Trans. Power Electron., vol. 21, no. 2, pp. 346–353, 2006. → page 11
[72] G. Zhou, G. Mao, H. Zhao, S. Zhou, and P. Mattavelli, “Digital average-ripple-based control
techniques for switching converters with fast transient performance,” IEEE J. Emerg. Sel. Topics
Power Electron., vol. 9, no. 1, pp. 89–101, 2021.
105
[73] M. Leng, G. Zhou, Q. Tian, G. Xu, and X. Zhang, “Improved small-signal model for switching
converter with ripple-based control,” IEEE Trans. Ind. Electron., vol. 68, no. 1, pp. 222–235, 2021.
→ page 11
[74] W. Tang, C. S. Leu, and F. C. Lee, “Charge control for zero-voltage-switching multiresonant
converter,” IEEE Trans. Power Electron., vol. 11, no. 2, pp. 270–274, 1996. → page 11
[75] C. Adragna, A. V. Novelli, and C. L. Santoro, “Charge-mode control device for a resonant converter,”
Apr. 15 2014, US Patent 8,699,240. → pages 11, 76
[76] H. Choi, “Charge current control for LLC resonant converter,” in Proc. IEEE Appl. Power Electron.
Conf. Expo., pp. 1448–1452, Mar. 2015. → pages 11, 13
[77] Z. Hu, Y. Liu, and P. C. Sen, “Bang-bang charge control for LLC resonant converters,” IEEE Trans.
Power Electron., vol. 30, no. 2, pp. 1093–1108, Feb. 2015. → pages 11, 13, 78
[78] B. McDonald and Y. Li, “A novel LLC resonant controller with best-in-class transient performance
and low standby power consumption,” in Proc. IEEE Appl. Power Electron. Conf. Expo., pp.
489–493, Mar. 2018. → pages 12, 13
[79] S. Kang and B. Cho, “Digitally implemented charge control for LLC resonant converters,” IEEE
Trans. Ind. Electron., vol. 64, no. 8, pp. 6159–6168, Aug. 2017. → pages 11, 76
[80] V. Vorperian and S. Cuk, “Small signal analysis of resonant converters,” in Proc. IEEE Power
Electron. Spec. Conf., pp. 269–282, 1983. → page 12
[81] A. F. Witulski and R. W. Erickson, “Small signal ac equivalent circuit modelling of the series
resonant converter,” in Proc. IEEE Power Electron. Spec. Conf., pp. 693–704, 1987. → page 12
[82] S. R. Sanders, J. M. Noworolski, X. Z. Liu, and G. C. Verghese, “Generalized averaging method for
power conversion circuits,” in Proc. IEEE Power Electron. Spec. Conf., pp. 333–340, 1990.
[83] C. Rim and G. Cho, “Phasor transformation and its application to the DC/AC analyses of frequency
phase-controlled series resonant converters (SRC),” IEEE Trans. Power Electron., vol. 5, no. 2, pp.
201–211, 1990.
[84] Y. Yin, R. Zane, J. Glaser, and R. Erickson, “Small-signal analysis of frequency-controlled electronic
ballasts,” IEEE Trans. Circuits Syst. I, Fundam. Theory Appl., vol. 50, no. 8, pp. 1103–1110, 2003.
[85] E. X. Yang, F. C. Lee, and M. M. Jovanovic, “Small-signal modeling of series and parallel resonant
converters,” in Proc. IEEE Appl. Power Electron. Conf. Expo., pp. 785–792, 1992. → page 72
[86] J. Sun and H. Grotstollen, “Averaged modeling and analysis of resonant converters,” in Proc. IEEE
Power Electron. Spec. Conf., pp. 707–713, 1993.
[87] F. Degioanni, I. G. Zurbriggen, and M. Ordonez, “Dual-loop controller for LLC resonant converters
using an average equivalent model,” IEEE Trans. Power Electron., vol. 33, no. 11, pp. 9875–9889,
2018. → page 12
[88] E. X.-Q. Yang, “Extended describing function method for small-signal modeling of resonant and
multi-resonant converters,” Ph.D. dissertation, Dept. Elect. Eng., Virginia Polytechnic Inst. State
Univ. Blacksburg, VA, USA, Feb. 1994. → page 12
[89] J. A. Martin-Ramos, A. M. Pernia, J. Diaz, F. Nuno, and J. M. Alonso, “A circuit for the large and
small signal dynamic modeling of the PRC-LCC resonant topology with a capacitor as output filter,”
in Proc. IEEE Power Electron. Spec. Conf., pp. 635–641, 2005. → page 12
106
[90] M. Hu, N. Frohleke, and J. Bocker, “Small-signal model and control design of LCC resonant
converter with a capacitive load applied in very low frequency high voltage test system,” in Proc.
IEEE Energy Convers. Congr. Expo., pp. 2972–2979, 2009.
[91] C. Chang, E. Chang, C. Cheng, H. Cheng, and S. Lin, “Small signal modeling of LLC resonant
converters based on extended describing function,” in Proc. IEEE Int. Symp. Comput., Consum.
Control, pp. 365–368, 2012.
[92] C. Buccella, C. Cecati, H. Latafat, P. Pepe, and K. Razi, “Linearization of LLC resonant converter
model based on extended describing function concept,” in Proc. IEEE Int. Workshop Intell. Energy
Syst., pp. 131–136, 2013. → page 12
[93] Z. U. Zahid, Z. M. Dalala, R. Chen, B. Chen, and J. Lai, “Design of bidirectional DC-DC resonant
converter for vehicle-to-grid (V2G) applications,” IEEE Trans. Transport. Electrific., vol. 1, no. 3,
pp. 232–244, Oct. 2015. → page 12
[94] Z. U. Zahid, Z. M. Dalala, C. Zheng, R. Chen, W. E. Faraci, J. J. Lai, G. Lisi, and D. Anderson,
“Modeling and control of series-series compensated inductive power transfer system,” IEEE J.
Emerg. Sel. Topics Power Electron., vol. 3, no. 1, pp. 111–123, Mar. 2015.
[95] 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 Trans. Power Electron., vol. 31, no. 10, pp.
7297–7310, Oct. 2016.
[96] C. Xu, S. Wang, X. Zhang, X. Wu, B. Wei, J. Xu, T. Cai, H. Hu, and Z. Li, “Dynamic modeling of
bidirectional inductive power transfer systems,” in Proc. IEEE Conf. Ind. Electron. Appl., pp.
452–456, 2019. → page 12
[97] S. S. Shah, U. Raheja, and S. Bhattacharya, “Input impedance analyses of charge controlled and
frequency controlled LLC resonant converter,” in Proc. IEEE Energy Convers. Congr. Expo., pp.
1–5, Sep. 2018. → pages 13, 77
[98] R. Yang, B. A. McDonald, and Y. Li, “Investigation on the small signal characteristic based on the
LLC hybrid hysteretic charge control,” CPSS Trans. Power Electron. Appl., vol. 4, no. 2, pp.
128–142, Jun. 2019. → pages 13, 77, 78
[99] Z. Hu, Y. Qiu, Y. Liu, and P. C. Sen, “A control strategy and design method for interleaved LLC
converters operating at variable switching frequency,” IEEE Trans. Power Electron, vol. 29, no. 8,
pp. 4426–4437, 2014. → page 13
[101] S. S. Shah, S. K. Rastogi, and S. Bhattacharya, “Paralleling of LLC resonant converters,” IEEE
Trans. Power Electron., vol. 36, no. 6, pp. 6276–6287, 2021. → page 13
[102] E. S. Glitz, J.-D. Hsu, and M. Ordonez, “Power loss estimation in LLC synchronous rectification
using rectifier current equations,” IEEE Trans. Ind. Electron., vol. 67, no. 5, pp. 3696–3704, 2020. →
page 45
107