MOSFET Selection to Minimize Losses in Low-Output

advertisement
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
MOSFET Selection to Minimize Losses in
Low-Output-Voltage DC-DC Converters.
Jon Gladish, Fairchild Semiconductor
Abstract — This paper focuses on the role of the power
MOSFET in achieving high-efficiency converter design. It
provides a brief overview of current low-voltage MOSFET
trench technologies, along with a discussion about onresistance versus gate charge trade-offs for MOSFETs
optimized for use as control or synchronous switches. It covers
the importance of the integrated Schottky diode (SyncFET™
MOSFET) in synchronous rectification and the necessity for
packaging technologies with low parasitic inductance and
resistance.
The MOSFET-to-circuit interaction is discussed in detail,
with TCAD mixed-mode simulations. All relevant MOSFET
switching events are analyzed: common-source inductance
versus drain current rise and fall, body diode conduction and
reverse recovery, external Schottky diode layout challenges
versus SyncFET MOSFET advantages, and elimination of
shoot-through currents from gate bounce. The simulated
MOSFET power losses are compared for various circuit
inductance cases and used for background in discussing
measured converter efficiency data. A review of popular
MOSFET loss equations is also discussed.
I. INTRODUCTION
The low-output-voltage DC-DC regulator is the
basic building block for essentially all CPUs,
memory, chipsets, and auxiliary supplies. Many of
the CPU or GPU ICs are demanding DC-DC
converters deliver a very high output current at a
very low output voltage with ever-increasing load
current slew rate requirements. There is also a
simultaneous drive to minimize both printed circuit
board (PCB) temperatures and converter sizes as
many PCBs are already fully populated with ICs
that can do without the heat coupling arising from
nearby inefficient power converters. Finally, the
converter must not induce excessive conducted or
radiated EMI into the surroundings, which requires
special attention to layout paths and proper selection
of components. For such switching converters, the
power MOSFET silicon and packaging technology
play important roles in realizing these design goals.
The process of selecting power MOSFETs for a
DC-DC converter design often begins with a
designer narrowing down a selection of MOSFETs
based upon a few key parameters. Parameters (or
features) such as the minimum guaranteed drainsource breakdown voltage (BVDSS), package type
(i.e. SO-8, TO-252, etc.), on-resistance (RDS(ON)),
and Figure-of-Merit (FOM), or [RDS(ON)] x Total
Gate Charge [QG(TOT)], typically give significant
insight into the expected MOSFET performance.
These parameters, combined with various other
datasheet parameters, are typically used within a
spreadsheet loss analysis to predict efficiencies
based on conduction and switching losses.
MOSFET loss equations formulated around piecewise linear approximations of switching waveforms
are found frequently in MOSFET supplier
application notes. A review of application
notes[1][2][3] from some power MOSFET and power
management suppliers reveals that there is industry
agreement for generalized MOSFET loss equations.
The set of loss equations generally gives the
converter designer an idea of the MOSFET
conduction and switching losses, but may fall short
of actual measured data. Loss calculators often
underestimate MOSFET switching losses since they
omit the influence of parasitic circuit inductance.
This paper takes an in-depth look into the losses
associated with power MOSFET switching
transitions and, through simulations, compares ideal
and non-ideal cases. While the discussion is
centered around the synchronous buck converter,
many of the parametric selection criteria also
applies to isolated DC-DC converters where a
primary-side MOSFET closely resembles the highside control MOSFET and the output or secondaryside rectifiers resemble the low side synchronous
MOSFET.
II. DISCRETE POWER MOSFETS
Discrete power MOSFETs are available within a
vast combination of RDS(ON), BVDSS and packaging
options. The multiple combinations are typically
1
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
Packaged RDSON (mΩ )
6
5
VGS = 10V
4
3
2
1
0
1998
2003
Year Introduced
2008
Fig. 1. RDS(ON) vs. Time [30V BVDSS MOSFET in an SO-8 footprint].
100
FOM QGD x RDS(ON)
1. Very low on-resistance is essential for
minimizing
synchronous
rectifier
and
controlling MOSFET conduction losses. Low
RDS(ON) in a discrete MOSFET implies that the
package and the silicon resistive contributions
are very low. Modern low-voltage discrete
power MOSFETs give special attention to
minimizing package resistance, since a
packaged MOSFET approaching the one
milliohm level may have 30% or more of the
total resistance as package bond wire and lead
resistance. Figure 1 provides curve-fitted
industry data for a typical 30V packaged
MOSFET RDS(ON) versus time (SO-8 footprint).
2. Low FOM is essential for optimizing control
MOSFET switching and conduction losses and
is typically needed for the prevention of
synchronous rectifier [CGD x dVDS/dt] induced
turn-on. While there are many metrics used to
grade switching MOSFETs, the RDS(ON) x
QG(TOT) (or QGD) FOM are the most common
and typically correlate with high performance.
While it can be argued that there are subtle
differences in each FOM, a lower RDS(ON) x QGD
FOM typically signifies a higher switching
grade MOSFET and often correlates to lower
RDS(ON) x QG(TOT). Figures 2 and 3 plot some
industry averages for MOSFET figure of merit
versus time.
3. Low internal series gate resistance (RG or gate
ESR) is important. A discrete power MOSFET
is often depicted (or modeled) as a lumped
circuit consisting of parasitic capacitance and
resistance of the active cell scaled for area,
where the MOSFET internal gate resistance and
internal capacitance determine the input
impedance and switching speeds.
VGS
4.5V
10V
50
0
1998
2003
2008
Year Introduced
Fig. 2. QGD x RDS(ON) FOM vs. Time.
300
FOM (QTOTAL x RDSON )
made available by a supplier to suit the needs of a
large array of applications, ranging from lower
performance, cost-sensitive designs to harsh
environment, high-reliability designs to highperformance designs where optimized packaging
and silicon need to be fully utilized.
Typically, these higher performance converter
designs push MOSFET silicon and packaging
technologies to strive for smaller, more efficient
products. It is essential that MOSFETs designed for
high-efficiency,
high-switching-frequency
applications (> 300 kHz) have a few key attributes
for meeting the ever-increasing demand for high
power densities with high converter efficiency. The
key attributes are:
10V RDSON x QTOT(10)
200
4.5V RDSON x QTOT(4.5)
100
0
1998
2003
Year Introduced
2008
Fig. 3. QG(TOT) x RDS(ON) FOM vs. Time.
4. Low parasitic package inductance, which is
important for optimizing MOSFET switching
speed and is required for minimizing the voltage
stresses associated with L x di/dt during
switching transitions.
5. Low thermal resistance (junction-to-case, ΘJC
and junction-to-ambient, ΘJA) for removal of
2
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
self-generated heat from the MOSFET silicon
and package increases reliability and provides
for minimized power losses with higher system
efficiencies.
6. Robust Forward Biased Safe Operating Area
(FBSOA) and Unclamped Inductive Switching
(UIS) are typically highly correlated and provide
insurance for surviving high-energy switching
spikes within the converter.
A. Packaging Technologies
As previously mentioned, power MOSFET silicon
is supplied within various packaging technologies to
accommodate the vast number of applications
where these products may be used. For example,
very small footprint Chip-Scale (CSP) or Ball Grid
Array (BGA) package, shown in Figure 4, provides
a very low height profile, along with exceptional
die-to-footprint ratio, optimizing space-constrained
systems. They also tend to offer the lowest parasitic
resistance and inductance due to their direct
connection to the PCB. They can be found in ultraportable electronics operating as load switches or
used in low-current switching converters. However,
larger high-current BGA packages have also been
demonstrated as excellent candidates for realizing
high efficiencies in very high-frequency switching
converters[4][5].
copper clip bonded. Each of these techniques,
especially source and gate bonding, influences the
packaging parasitic inductance and resistance.
While packaging resistance is part of the aggregate
RDS(ON) reported on datasheets, packaging
inductance is rarely stated. This value can be
approximated through lab measurement or can
sometimes be found in a MOSFET supplier’s
SPICE or SABER® model[6]. PQFN and MLP also
offer very low junction-to-case thermal resistance
due to the exposed copper header.
Drain
Source Bond Wires
Fig. 5. - MLP 5mm x 6mm.
The larger TO-220 or TO-252 (D-PAK) package,
shown in Figure 6, is typically found in automotive,
industrial, or computer applications, such as desktop
mother- and daughter-board voltage regulators
where PCB real estate is larger.
Source Bond Wires
Source
Drain
BGA 1.5mm x 1.5mm
CSP 1.0mm x 1.5mm
Fig. 4. - BGA and CSP packages.
The somewhat larger, but still very efficient, fully
encapsulated Power Quad (and Dual) Flat Package
(PQFN or DFN), Micro Lead-frame Package (MLP)
shown in Figure 5, or SO-8, are some of the most
popular packages found in DC-DC converter
designs (for discussion purposes in this paper,
PQFN and MLP are treated as similar packages).
All three packages provide good die-to-footprint
ratios and can provide low parasitic resistance and
inductance. Internally, the die attach (typically the
MOSFET drain) can be soldered or epoxy bonded,
while the source and gate connections can be
copper/gold/aluminum wire or ribbon bonded, or
Gate Bond Wire
Fig. 6. - TO-252 (D-PAK).
These packages offer excellent junction-to-case
thermal resistance with large copper headers that
provide for heat-sinking. They also typically have
the highest maximum rated operational temperature
range, a feature highly desirable in automotive and
industrial applications. However, the long lead and
bond wire lengths tend to create much higher
parasitic inductance compared to PQFN or SO-8,
shown in Figure 7. Packaging parasitics are
discussed in detail in reference [7]. Another
drawback is that the die-to-footprint ratio is less
desirable than PQFN or SO-8, which unnecessarily
occupies valuable PCB space.
3
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
Inductance ( nH )
4
SOURCE METAL
LGate (D-PAK)
P+ contact
N+ Source
3
dielectric
LSource (D-PAK)
2
LGate (PQFN)
1
Gate
CDS
LSource (PQFN)
Drain
P type
Body
CGS
CGD
0
1
10
Frequency ( MHz )
Channel
N- Epi
100
Fig. 7. Package parasitic inductance versus frequency.
B. Silicon Technologies
The active cell structure of a low-voltage discrete
power MOSFET is often described by the two
dimensional cross section of the cell or cells, shown
in Figure 8 (not drawn to scale - depicts the typical
repetitive nature of the cell). While MOSFET
technologies and cell structures have evolved
dramatically through the years, the MOSFET cell
structure can be segmented into three basic
categories: planar, trench or lateral. Of the three
structures, trench-gated MOSFETs are most
common for high-performance discrete power
MOSFETs with BVDSS < 200 V. They are chosen
primarily for their exceptionally low specific onresistance (product of resistance times area of
silicon – measured in milliohm-mm2 (or cm2)), and
are a technology capable of excellent RDS(ON) x
QG(TOT) (QGD) FOM across the BVDSS spectrum.
They also tend to provide for a very robust FBSOA
and UIS for surviving harsh switching events.
A compelling advantage of the trench structure is
in the ability to reduce on-resistance by providing
the shortest possible current path (vertical) from
drain to source through the lowest possible
resistance. As shown in Figure 8, the major
contributors to silicon resistance typically arise from
the channel (RCHANNEL), epitaxial (drift region) REPI,
and substrate regions RSUBSTRATE.
N+ Substrate
Drain Metal
Fig. 8. Trench Power MOSFET active cell cross section.
The substrate and epi-resistances are often
controlled by utilizing the thinnest, highest doped
silicon possible. While the ability to achieve tight
trench-to-trench cell pitch allows for an extremely
high channel width-to-length ratio (W/L), this
results in a very low channel resistance.
Drain
Gate
CGD
CDS
RG
Gate
Thick oxide for low CGD
CGS
Source
Fig. 9. Trench MOSFET cell with Thick Bottom Oxide (TBO).
The percentage of resistance associated with each
region varies dramatically, depending on design and
BVDSS. While RDS(ON) is vital for low conduction
losses, considerations must be made for enhanced
FOM, where trade-offs in trench depths and widths
exist to optimize the structure.
Variations to the standard trench cell of Figure 8
are often designed with the intention of preserving
low resistance, while enhancing the FOM.
Structures such as the shielded gate or thick bottom
4
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
oxide trench, shown in Figure 9, are two examples
that limit the gate-to-drain overlap capacitance, thus
lowering QGD and providing faster switching and
increased dv/dt immunity.
C. Integrated Schottky Diode (SyncFET MOSFET)
The monolithically integrated Schottky diode can
be one of the most important features added to the
low-side MOSFET die, especially as the input DC
input voltage increases. A well-designed Schottky
diode can simultaneously decrease the dead time
diode conduction losses and dramatically reduce
switching losses attributed to QRR. This provides
added output capacitance to reduce the recovery
dVDS/dt, assisting in the prevention C x dVDS/dt
turn-on.
Schottky contact
Drain
Source
Gate
Schottky diode cell
Source
Fig. 10. Trench Metal Barrier Schottky (TMBS) cell (SyncFET)
A typical Trench Metal Barrier Schottky (TMBS)
active cell is shown in Figure 10. This type of
Schottky diode structure integrates well into the
trench MOSFET, as it utilizes a trench structure as
an integral part of the overall Schottky diode
structure. The Schottky diode cell can be distributed
through the MOSFET active area or placed
separately in a dedicated area on the die, providing
an extremely small physical separation from the
MOSFET body diode. Both methods provide
reduced body diode injection and QRR[8]. The
Schottky diode contact is typically formed on the
topside of the structure where the N-type silicon and
metal form the Schottky diode barrier, while the
trench poly-silicon is often tied directly to source
metal that aids in providing a high breakdown in the
cell, along with enhanced robustness.
III. SYNCHRONOUS BUCK CONVERTER
The non-isolated synchronous buck converter,
shown in Figure 11, is widely used throughout
electronic systems to step down an intermediate DC
bus voltage to a logic level voltage powering a
CPU, GPU, memory, or other integrated circuits.
LF
Q1
VIN
Controller
and Driver
Q2
CF
VO
RLOAD
Fig. 11. Synchronous buck schematic.
The synchronous buck converter is used in this
paper as an evaluation platform for discussing
power MOSFET losses due to the popularity and
relative ease and convenience of evaluating a
control MOSFET (hard–switched) and synchronous
MOSFET in one circuit.
The MOSFET loss equations presented
correspond to the synchronous buck converter
operating waveforms in the steady-state, continuous
conduction mode (CCM). The equations are
presented with the corresponding power MOSFET
switching waveforms, which aid in loss
explanations. For this discussion, it is assumed the
reader has a basic knowledge of the buck converter
operation and is encouraged to read references [9]
and [10] for a more in-depth description of the
theory behind the synchronous buck operation.
The MOSFET loss discussions are intended to
review the ideal switching waveforms (no circuit
inductance) and transition into the more realistic
waveforms that include the effects from parasitic
package and circuit inductance. Then, with the aid
of simulation and measured waveforms, the
discussion explores each switching transition and
points out potential issues that may arise, causing
MOSFET switching to deviate from predicted
results. Finally, various converter efficiency curves
are shown to correlate to the switching waveforms.
5
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
IV. MOSFET LOSSES
Generally, MOSFET losses can be categorized as
either switching or conduction losses, where the
total loss across one switching cycle equals the sum
of the switching and conduction losses. A survey of
typical loss equations from references [1], [2], and
[3] are summarized in the following sections.
A. High Side MOSFET (Q1) Loss
From reference [1], the high-side MOSFET of the
synchronous buck operates with an on-time equal to
duty cycle, multiplied by switching period [D x TS].
The turn-off and turn-on switching currents are
equal to the inductor ripple current, plus or minus
the DC load current, respectively. Steady-state
losses incurred by the high MOSFET are as follows:
PLOSS (Q 1) = PCOND + PSW (ON ) + PSW (OFF ) + PGATE
(1)
PGATE (Q 1) = QG (TOT )VG fs
(2)
where:
QG(TOT) = total gate charge of Q1;
VG = Gate Drive DC voltage;
fS = switching frequency, TS = 1/fS.
PCOND (Q 1) = IQ2 1( RMS ) R DS (ON )( Q 1)
VIN IDS (OFF )tOFF
2
]fs
Q1 Turn-Off Loss
IDS(Q1)
VDS(Q1)
VGS(Q1)
VPlateau
VTH
t0
t1 QGD QGS2
Q2 VF Loss
ISD(Q2)
VDS(Q2)
2
+
QOSSVIN
+ QRRVIN ]fs
2
where:
IDS(ON) = inductor current at MOSFET turn-on;
tON = (QGS2+QGD)/iG(ON);
iG(ON) = (VG-VPLATEAU)/( RG_HS + RDRV_HSON);
QOSS = output charge of Q1[3];
QRR = diode recovery charge of Q2.
VDSON(Q2)
t3
t4
t5
t6
Fig. 12. MOSFET Q1 turn-off and Q2 turn-on loss waveforms.
(3)
(4)
1. Turn-off delay time (t1 – t0). During this phase,
only the MOSFET RDS(ON) is affected as it rises
in response to the lowering VGS.
2. Drain voltage rise time (t2 – t1), also known as
the gate plateau or “Miller” region. During this
switching event, VGS of Q1 is at a level where
the MOSFET can no longer conduct the drain
current at low VDS levels. Use Equation 6 for a
linearly approximated value for VPLATEAU:
VPLATEAU (OFF )Q1 =
VIN IDS(ON )tON
VTH
t2
where:
VIN = Converter DC input voltage;
IDS(OFF) = inductor current at MOSFET Q1 turn-off;
tOFF = (QGS2+QGD)/iG(OFF);
iG(OFF) = VPLATEAU/( RG_HS + RDRV_HSOFF).
PSW (ON )Q1 = [
VGS(Q2)
Q2 COSS
displacement
current
Diode VF
where:
RDS(ON)(Q1) = MOSFET on resistance;
IQ1(RMS) = RMS drain current.
PSW (OFF )Q1 = [
The switching characteristics of a control (hardswitching) MOSFET are a function of the MOSFET
input impedance (capacitance CISS and resistance,
RG) combined with the output impedance of the gate
drive. The MOSFET turn-off process can be
segmented in time into three phases shown in
Figure 12.
(5)
I DS (OFF )
gm
+ VTH
(6)
where gm = MOSFET transconductance.
In response, VDS rises while VGS essentially
stalls as the gate current charges the gate to
drain capacitance, CGD. Instantaneous power
losses can be large during this transition; there is
simultaneous high drain current with rising
drain voltage.
6
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
PSW (OFF _ QGD )Q1 = [
VIN I DS (OFF ) (t 2 − t 1 )
2
]fs
(7)
where:
(t2 - t1) = QGD/ iG(OFF);
QGD = gate to drain charge.
3. Current fall time (t3 – t2), gate discharge from
VPLATEAU to VTH. During this transition, the
channel of MOSFET Q1 is shut off while the
inductor current is transferred to MOSFET Q2
body diode. The drain voltage of Q1 is clamped
to the DC input bus by the body diode of
MOSFET Q2 as the drain current begins to fall
as VGS discharges.
PSW (OFF _ QGS2)Q1 = [
VIN I DS(OFF ) (t 3 − t 2 )
2
]fs
the diode tA phase (time t12 to t13) is large
compared to the tB phase (time t13 to t14). During
diode recovery, instantaneous power losses
across the high-side MOSFET are large, since
both the DC input voltage and diode recovery
current, plus output inductor load current,
remain across the high-side MOSFET until the
low-side MOSFET can begin to block voltage
(end of the tA phase).
PSW (ON _ QGS 2 )Q1 = [
VDSIDS(ON ) (t12 − t11)
2
(9)
+ QRRVIN ]fs
where:
(t12 - t11) = QGS2/ iG(ON);
QGS2 = gate charge from VTH to VPLATEAU.
(8)
Q1 Turn-On Loss
where:
(t3 - t2) = QGS2/ iG(OFF);
QGS2 = gate charge from VPLATEAU to VTH.
IDS(Q1)
VDS(Q1)
The ideal high-side MOSFET turn-on process can
also be segmented in time into three phases, shown
in Figure 13.
1. Turn-off delay time (t11 – t10). During this phase
the MOSFET is off with VGS rising toward VTH.
2. Current rise time (t12 – t11). The gate charges
from VTH to VPLATEAU. During this switching
event, the drain voltage of Q1 is clamped to the
DC input bus since the body diode of Q2 is
forward biased and conducting the inductor
current. The drain current of Q1 begins to rise as
VGS surpasses VTH. The inductor current is
being transferred from Q2 to Q1. The current
rise time ends once Q1 current equals the
inductor current plus the peak reverse recovery
current of Q2, IRR. For simplicity in loss
calculations, the diode reverse recovery current
is temporarily ignored to calculate the turn-on
losses with an ideal diode. QRR losses are treated
in a separate calculation.
Figure 13 also shows the impact of diode
reverse recovery on switching waveforms. The
diode recovery time (tRR) and reverse recovery
charge (QRR) specified on datasheets are
generally used by loss calculators as a straightforward (QRR x fSW x VIN) switching losses.
Referring to Figure 13, the assumption is that
VGS(Q1)
VPlateau
VTH
QGS2 QGD
ISD(Q2)
VGS(Q2)
VTH
VDS(Q2)
t7
t8
t9
t10
t11
t12
t13 t14
tA phase tB phase
Fig. 13. MOSFET Q1 turn-on and Q2 turn-off loss waveforms.
3. The drain voltage fall time (t13 – t12) is the turnon gate plateau, or “Miller” region, where there
is simultaneous high drain current with falling
drain voltage. During this switching event, VGS
of Q1 is at a level where the MOSFET conducts
the entire inductor load current, plus the diode
reverse recovery current. MOSFET (Q1) VDS
begins to fall as the body diode enters the
reverse blocking mode. Similar to the turn-off
Miller region, VGS essentially stalls out as the
gate current is used to discharge the gate to
drain capacitance, CGD.
7
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
The drain voltage fall time also creates an
internal MOSFET channel current due to the
discharging output capacitance COSS. This
current does not show in lab measurements,
however, it is a switching loss and is typically
treated with a separate loss contribution. An
output charge, QOSS, which takes into account
the non-linear effects of COSS as a function of
VDS, is often used to calculate losses:
PSW (ON _ QGD )Q1 = [
(IDS(ON ) (t13 − t12 ) + QOSS (Q1+Q 2) )
2
]VIN fs
(10)
B. Low-Side MOSFET (Q2) Loss
Typically, the low-side MOSFET is conductionloss dominated with additional diode conduction
losses for dead times. Turn-on losses for Q2 is
typically considered lossless since the transition is
considered highly capacitive, as shown in Figure 12.
The turn-off (diode recovery) event, shown in
Figure 13, is approximated as near lossless because
the QRR and QOSS are assumed to be dissipated in
the high-side MOSFET turn-on. A more
conservative technique is to add QOSS and QRR
losses to Q2. Another actual switching event for Q2
discussed in reference [1] is switching between the
MOSFET channel to body diode and vice versa.
Losses are omitted here as the equations assume fast
gate edges, however, for high-frequency operation,
this switching event must be considered. Q2 losses
are as follows:
P SWITCHING (OFF ) = [Q RR V IN +
Q OSS
( Q 2 )V IN
2
Note that both QRR and QOSS(Q2) are accounted for
as losses in Q1 turn-on from Equation 10, however,
a portion of these losses appears in Q2, with
Equation 14 as guidance for distributing the losses.
V. PARASITIC INDUCTANCE EFFECTS
The loss equations presented above are generally
used to estimate expected MOSFET performance,
but often fall short of predicting actual performance.
While there are numerous reasons this may occur,
more often than not, the culprit is the parasitic
circuit inductance, as shown in Figure 14. For lowvoltage MOSFETs, the influence of parasitic
inductance has been studied rather intensely in
recent years[11-13] and it has become general
knowledge that inductance can strongly influence
MOSFET switching characteristics, usually causing
increased switching losses and deviations from the
expected performance.
Ld_HS
Lg_HS
VIN
Rdrv_HS
Rg_HS
Control
MOSFET
Ls_HS
L_f
Ld_HS
PLOSS (Q 2) = PCOND(Q 2) + PSWITCHING (OFF ) + PGATE
(11)
PGATE ( Q 2 ) = Q G (TOT )( Q 2 )V G Fs
QG(TOT)(Q2) = Total gate charge for MOSFET Q2.
(12)
PCOND (Q 2) = PMOS _ COND + PDIODE _ COND
(13)
_ COND
PDIODE
(Q 2 ) = I Q 2 ( RMS ) 2 R DS (ON )( Q 2 )
_ COND
(Q 2 ) = [I Q 2 (ON )V F t DEAD (ON ) +
R_snubber
Lg_LS Rg_LS
Rdrv_LS
PMOS
(14)
]fs
R_load
C_f
Synchronous
MOSFET
C_snubber
Ls_HS
(13a)
(13b)
I Q 2 (OFF )V F t DEAD (OFF ) ]
where:
PMOS_COND (Q2) = MOSFET Q2 channel conduction;
PDIODE_COND (Q2) = Q2 body diode conduction;
VF = Q2 body diode forward conduction voltage;
TDEAD(ON) = dead time from Q1 (HS) off to Q2 (LS) on;
IQ2(ON) = inductor current at Q2 (LS) turn on ( = IDS(OFF) Q1);
TDEAD(OFF) = dead time from Q2 (LS) off to Q1 (HS) on;
IQ2(OFF) = inductor current Q2 (LS) turn off (= IDS(ON) Q1).
Fig. 14. Synchronous buck with parasitic inductance.
Parasitic inductance arising from both component
packaging and circuit layout is a reality of any
circuit. The above packaging section shows that a
package, such as the D-PAK, can have up to 2.5nH
of inductance from bond wire and leads, which is
added directly to the high-current / high-frequency
AC loop inductance being minimized (the highcurrent AC loop is defined in Figure 20). Worse yet,
when this inductance is common to both the gate
and power loop (common source inductance
8
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
described in the following section), it tends to
dictate the high-side MOSFET switching
characteristics, causing slower switching and higher
power losses.
Zero Circuit Inductance
Turn on loss
VDS
Turn off loss
IDS
VGS
tON1
tOFF1
tON2
tOFF2
Including Circuit Inductance
Turn on loss
Turn off loss
IDS
VGS
tON1
tON2
tOFF1
tOFF2
Fig. 15. High-side switching waveform comparison - zero parasitic
inductance (top) versus “typical” parasitic circuit inductance (lower).
Generic high-side MOSFET switching waveforms
are shown in Figure 15. The waveforms for the
parasitic inductance case are shown in side-by-side
comparison to the zero parasitic inductance (ideal)
case to clearly show the difference that arises in the
switching edges and times. Often, loop inductance
adds significant voltage stresses to the MOSFET,
arising from the fast circuit di/dt during turn on and
turn off.
Note that unless stated otherwise, all MOSFET
gate-to-source VGS and drain-to-source VDS
waveforms are defined at the silicon level,
excluding parasitic resistance and inductance.
While closed-form equations have been derived
for MOSFET switching losses, including the effects
of circuit inductance[11][12][13], this paper relies on
advanced simulations to quantify circuit-toMOSFET interactions.
A. TCAD Mixed-Mode Simulations
Technology Computer Aided Design (TCAD)
mixed-mode simulations are arguably the most
accurate method of modeling power MOSFET
losses[14]. TCAD software is typically used by
semiconductor devices and process engineers for
device development and modeling and is extremely
useful in modeling MOSFET silicon and package
interactions. The concept is to utilize a highly
accurate and calibrated physical MOSFET model in
combination with behavioral circuit elements to
accurately model MOSFET switching behavior. A
benefit of this simulation over SPICE is the very
accurate modeling for diode QRR, dynamic
avalanche, and other minority carrier effects.
For this paper, a TCAD mixed-mode simulation
was set up to predict losses for a single-phase, wideinput voltage converter (i.e. 7-22VIN for a typical
Notebook computer). While all reported simulation
losses are shown at 19VIN, output filter components,
parasitic inductance, and gate drive resistances were
chosen to give realistic ripple current and voltage
and accurate representation of the gate drive current
capability. Circuit parasitic inductances were chosen
for a “typical” 5x6mm PQFN layout.
The essence of this paper is summarized in the
next four plots. The plots summarize TCAD
simulated MOSFET losses for the same high-side
and low-side silicon die, altering circuit inductance
to model the effects on high-side and low-side
switching losses. The plots show MOSFET
switching loss trends where, many times, even
modest circuit inductance, due to careful layout and
well-chosen packages yield total MOSFET losses
that exceed the ideal or calculated expectations.
Figure 14 provides a simplified version of the
simulated TCAD circuit, which lumps many of the
circuit inductances into MOSFET drain and source
inductances, and is used to quantify the simulated
circuit values for Figures 17-19.
9
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
B. Simulated Cases
The four simulated cases describe the loss
behavior for similar MOSFETs (FDMS8680x1 HS
with FDMS8660ASx1 LS) operating around
varying circuit inductances. Test conditions are:
VIN = 19V, fSW = 300kHz, VOUT = 1.2V, VG = 5V.
DEVICE
FDMS8692
FDMS8660AS
RDSON
QG(5)
QGD
QGS
RG
Vt
10VGS 4.5VGS
(mΩ) (mΩ)
(nC)
(nC)
(nC)
(Ω)
(V)
8
30
2.1
5.2
2.7
12
1.0
1.2
1.8
1.7
Package
5x6 PQFN
5x6 PQFN
7.0
1.7
10.5
2.3
Case 1.) Zero Parasitic Inductance
An initial simulation provides a baseline case for
comparing losses. This case sets all parasitic
inductance to zero value with MOSFET loss
predictions summarized in Figure 16. MOSFET
switching waveforms for this case resemble Figure
15 (zero parasitic inductance).
This case clearly demonstrates the necessity for
low RDS(ON) for an optimized low-side switch as
channel conduction losses dominate losses. It also
shows that a well-designed high-side MOSFET can
be optimized to equally distribute conduction and
switching losses across load current, however, highside turn-on losses are nearly double turn-off due to
diode QRR, which must be minimized.
Case 2.) Typical 5x6mm PQFN Inductance
Figure 17 resembles a “typical” DC-DC point-ofload (POL) circuit using packaged PQFN
MOSFETs with well-designed (low inductance)
power and gate loops. All parameters are identical
to the circuit of Case 1, with the addition of
parasitic inductance: Ls_HS=0.4nH, Ls_LS=0.4nH,
Ld_HS=1.2nH,
Ld_LS=0.3nH,
Rg_HS=1Ω,
Rg_LS=1Ω, Lg_HS=9nH, Lg_Ls=6.5nH.
2.5
0.5
0.0
25 amps
HS.drv
HS.Toff
HS.Ton
HS.Cond
LS.drv
7 amps
HS.Packg
Loss Mechanism
15 amps
LS.Swch
Iout ( A )
25 amps
1.0
LS.diode cond
HS.drv
HS.Toff
HS.Ton
HS.Cond
25 amps
20 amps
15 amps
10 amps
7 amps
HS.Packg
Loss Mechanism
LS.drv
LS.Cond
LS.Packg
LS.diode cond
LS.Swch
LS.Loss
HS.Loss
0.0
20 amps
LS.Cond
0.5
15 amps
1.5
LS.Packg
1.0
10 amps
2.0
LS.Loss
HS.Loss
1.5
7 amps
Power Loss ( W )
Power Loss ( W )
2.5
7 amps
10 amps
15 amps
20 amps
25 amps
2.0
Iout ( A )
Fig. 16. MOSFET simulated loss - zero circuit inductance.
HS.Loss = total high-side silicon plus package loss
LS.Loss = total low-side silicon plus package loss
LS.Cond = low-side channel conduction loss
LS.diode_cond = diode dead-time conduction loss
LS.Packg = low-side package resistive loss
LS.Swch = low-side loss associated with drain voltage and
current during switching
HS.Cond = high-side channel conduction loss
HS.Toff = high-side turn-off switching loss due to the overlap
of VDS and IDS
HS.Ton = high-side turn-on switching loss due to the overlap
of VDS and IDS
HS.drv = high-side gate losses from QG(TOT)
LS.drv = low-side gate losses from QG(TOT)
Fig. 17. MOSFET simulated loss - including circuit inductance
(HS Lsource = 0.4nH).
The most notable loss difference from case 1 to
case 2 is in the high-side MOSFET. Case 2 predicts
close to 20% higher total high-side losses. Of that
loss, turn-off losses have increased the most, due to
added drain voltage stress from ringing and slower
edges rates due to source inductance. This case
describes the situation where only modest circuit
inductance tends to shift higher losses towards the
high-side turn-off transient, while lowering the turnon losses.
10
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
Case 3.) 5x6mm PQFN with Added High-Side
Source Inductance
Case 3 follows case 2, where an additional 0.4nH
is added to Ls_HS (Ls_LS = 0.8nH versus 0.4nH).
This case is typical of a situation where internal
package bond wiring is not fully optimized, adding
extra package inductance.
2.5
10 amps
15 amps
20 amps
1.5
25 amps
2.5
1.0
7 amps
0.5
HS.drv
HS.Toff
HS.Ton
HS.Cond
25 amps
20 amps
15 amps
10 amps
7 amps
HS.Packg
Loss Mechanism
LS.drv
0.0
LS.Swch
The additional package inductance for this case
adds extra inductance in the high-current AC loop,
as shown in Figure 20, which tends to slow
switching. Moreover, the package inductance shows
up as extra high-side common source inductance,
which has strong effects on high-side losses.
Common source inductance, often referred to as
“source inductance,” is the inductance shared by
both the gate and high-current loop, shown in
Figure 20. It effects switching by generating a
voltage between the MOSFET source and the gate
drive return during drain current rise and fall. The
source inductance voltage is actively (and
negatively) fed back to the MOSFET VGS (measured
at the silicon), slowing switching transients. High
source inductance can significantly degrade
MOSFET switching and is discussed in more detail
in the following section.
Comparing HS.Loss for Figure 18 versus 17,
notice that the high-side losses are approximately
30% higher (1.6W vs. 1.2W). The extra loss is
incurred by the high-side MOSFET in both the turnon and turn-off switching events, while low-side
MOSFET switching losses actually decrease.
25 amps
1.0
LS.diode cond
Fig. 18. MOSFET simulated loss - including circuit inductance
(HS Lsource = 0.8nH).
20 amps
LS.Cond
Iout ( A )
15 amps
1.5
LS.Packg
HS.drv
HS.Packg
HS.Toff
25 amps
20 amps
15 amps
10 amps
7 amps
HS.Ton
HS.Cond
LS.drv
LS.Swch
LS.diode cond
LS.Cond
Loss Mechanism
LS.Packg
LS.Loss
HS.Loss
0.0
10 amps
2.0
LS.Loss
HS.Loss
0.5
Power Loss ( W )
Power Loss ( W )
7 amps
2.0
Case 4.) 5x6mm PQFN with Added Low-Side
Source Inductance
The final case, shown in Figure 19, adds 0.4nH
into the low-side source (Ls_LS = 0.8nH). This is
common source inductance for the low-side (which
actually helps hold the low-side gate off and is
addressed in the shoot-through section of this
paper), which adds additional loop inductance for
added high-side losses. The trend is higher total
losses with lower low-side losses and greater highside losses.
Iout ( A )
Fig. 19. MOSFET simulated loss - including circuit inductance
(HS and LS Ls_LS = 0.8nH).
One clear trend is that increasing the power loop
inductance, either through packaging or layout,
severely impacts MOSFET power losses. The
RDS(ON) x QG(TOT) FOM becomes less effective for
selecting highly efficient MOSFETs for cases where
high inductance completely dominates switching.
C. High-Side Common Source Inductance
Often, higher than expected losses in the highside switch arise from common source inductance,
shown in Figure 20. As described, common source
inductance (Ls_LS) is the inductance shared by both
the gate loop and high-current AC loop. In general,
for a control switch, the rapidly changing drain
current (dIDS/dt) during MOSFET switching induces
a source voltage (L x dIDS/dt) with a polarity always
working against the gate drive action, i.e. negative
feedback, shown in Figures 20 and 21. Both the
MOSFET turn-on and turn-off currents are affected
with slower switching speed, causing increased
power losses.
11
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
LD_HS
Q1
LS_HS
92
90
Efficiency (%)
Figure 20 provides an example for the source
inductance voltage generated during turn-off (i.e.
falling drain current).
L x di/dt
VIN
86
84
Lsource = 0.4nH
82
Lsource = 0.8nH
5
Q2
10
15
20
25
Iout (A)
Fig. 22. Simulated efficiency for various high-side MOSFET source
inductance cases.
RDRIVE_HS
Ls_LS
High Current AC Loop
Gate Loop
Lsource = 2.3nH
78
LG_HS
LG_HS
Lsource = 1.2nH
80
Ld_LS
RG_HS
88
An analysis of VGS and VDS switching waveforms
displays the impact on switching times as the source
inductance increases, as shown in Figure 23.
Fig. 20. Schematic with high-side MOSFET source inductance.
VDS_Si
(LS_HS = 0.4 nH)
(LS_HS = 2.3 nH)
27.5
7.0
6.0
20.0
4.0
5.0
VgsHS (V)
17.5
VdsHS (V)
12.5
15.0
3.0
2.0
10.0
7.5
1.0
5.0
0.0
7.0
2.0
2.5
3.82
3.83
3.84
3.85
-6
time (seconds) *10
3.86
3.87
3.88
3.89
0.0
-1.0
-2.5
0.0
2.5
1.0
-2.5
5.0
VGS
VGS_Si
(LS_HS = 0.4 nH)
(LS_HS = 2.3 nH)
-1.0
7.5
3.0
10.0
4.0
5.0
VgsHS (V)
VdsHSpac (V)
12.5
15.0
17.5
6.0
20.0
22.5
VDS
0.0
8.0
22.5
8.0
25.0
VgsHS
9.0
27.5
VgsHS
9.0
10.0
30.0
IdsHS
VDS and VGS (2V/div)
25.0
IDS
VdsHS
VdsHSpac
VdsHSpac
VgsHS
10.0
30.0
FDMS8680_FDMS8660AS_20A:0_0.1cycle.ivl<MEDICI_DATA>
FDMS8680_FDMS8660AS_20A:0_0.1cycle.ivl<MEDICI_DATA>
3.82
3.83
3.84
3.85
-6
time (seconds) *10
3.86
3.87
3.88
3.89
Time (10ns/div)
Fig. 23. High-side MOSFET source inductance waveforms.
Fig. 21. Simulated high-side MOSFET waveforms for source inductance
= 2.3nH.
The increased source inductance gives rise to a
perceived increase in Miller time, shown in Figure
23 (0.4nH vs. 2.3nH). What is actually occurring is
a trickle discharge of the gate voltage as the source
inductance potential opposes the gate drive action.
Figure 22 shows simulated efficiencies for varying
high-side source inductance, where higher values
(i.e. 2.3nH for D-Pak) show unacceptable
performance for optimizing efficiency. The
efficiency difference is over 5% points at full load
versus 0.4nH (PQFN). For this simulation, VIN=19V,
VOUT=1.3V, and fS = 300kHz.
The plot provides a comparison of the high-side
gate and drain voltage waveforms, comparing
conditions where Ls_HS=0.4nH vs. 2.3nH with the
same silicon.
Identifying high source inductance through lab
measurement can be accomplished through VGS and
VDS oscilloscope measurements using high
bandwidth probes (500MHz). However, many
times, the measured waveforms from packages,
such as the D-Pak or PQFN, do not allow for clear
measurements across the MOSFET silicon since
packaging parasitic inductance is also measured.
Figure 24 provides a simulated example showing
the VGS measurement at the package gate-source
pins compared to the much different VGS measured
at the MOSFET die. Care must be exercised when
interpreting the lab-measured waveforms.
12
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
Q1
LS_HSpack
RG_HS
VGS_si
LG_HSpack
VGS_pkg
LG_HSbrd
RDRIVE_HS
FDMS8680_FDMS8660AS_Ls1nH_20A:0_0.1cycle.ivl<MEDICI_DATA>
10.0
30.0
VdsHS
VgsHS
9.0
27.5
VgsHSpac
7.0
6.0
20.0
4.0
5.0
VgsHS (V)
17.5
VdsHS (V)
12.5
15.0
3.0
10.0
VGS_Si
2.0
-2.5
0.0
0.0
2.5
1.0
5.0
7.5
VGS_pkg
VGS ( 1 V/div)
VDS ( 2.5 V/div)
22.5
8.0
25.0
VDS
3.810
3.815
3.820
3.825
3.830
3.835
3.840
-6
time (seconds) *10
3.845
3.850
3.855
3.860
3.865
Time (5ns/div)
Fig. 24. High-side MOSFET source inductance measurement.
D. Dead Times
1) High-Side Off to Low-Side On
The high-side off to low-side on transition is
typically free of issues like cross-conduction,
however, the goal should be to minimize the body
diode conduction time, allowing enough time for
the phase node to drop below ground potential (lowside body clamps this node to approximately -0.6V)
before releasing the low-side gate drive to turn on
the MOSFET channel. This ensures crossconduction is eliminated since the gate driver does
not respond until body diode conduction has been
initiated. This technique is typically used for
synchronous buck (or half-bridge)-type gate drivers
that utilize anti-cross-conduction circuitry. From
Figure 12, this is approximately time (t4 - t2).
2) Low-Side Off to High-Side On Switching
Transition (Break-before-Make)
The low-side off to high-side on transition is
usually the switching event that requires the most
attention when selecting a compatible MOSFET for
a given gate driver. This is also the switching event
that can be the most problematic when dealing with
optimizing efficiency. The goal of this transition is
to quickly switch current from the channel to the
body diode, which must be minimized to optimize
losses, yet long enough to prevent cross-conducting
currents from allowing the high- and low-side
MOSFET
channels
from
conducting
simultaneously. This is also when both diode QRR
and any potential [CGD x dVDS/dt] induced shootthrough currents occur. All of these conditions
require special care in selecting a low-side
MOSFET.
This event begins when the low-side gate drive
pulls low, providing a low resistance path
discharging the low-side MOSFET gate-to-source
voltage shifting current from the MOSFET channel
to the body diode. After another short delay, the
gate of the high-side MOSFET is charged, initiating
turn-on. The more popular synchronous buck gate
drivers typically utilize an adaptive gating procedure
where the low-side off to high-side on transition are
performed by monitoring the low-side gate-tosource signal until a preset threshold level (~ 1V) is
reached. After a short delay (from Figure 13, t10 –
t8) the high-side gate driver is released to charge the
high-side MOSFET. This type of gating attempts to
ensure that any low-side logic-level MOSFET used
has fully turned off the channel before the high-side
MOSFET is gated on, since the gate drive is
monitoring the gate signal.
a)
Diode QRR
The diode reverse recovery time (tRR) and reverse
recovery charge (QRR) specified on datasheets are
generally used by loss calculators as straightforward (QRR x fSW x VIN) switching losses. A word
of caution on using datasheet QRR numbers for loss
calculations: The reverse recovery current of a diode
is a function of many parameters, such as forward
current IF, reverse recovery diF/dt, DC bus voltage,
and junction temperature TJ. An increase in any one
of these conditions generally results in increased
QRR. Datasheet test conditions are usually lower
than typical converter operating conditions. The low
test conditions typically arise for manufacturing test
reasons (i.e. TJ=25°C, IF=1A, dIF/dt=100A/µs).
13
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
Since switching converters try to switch the
MOSFET as fast as possible, edge rates, such as
diF/dt, can be up to ten to twenty times faster than
the datasheet test conditions, increasing diode QRR
for DC-DC.
To further complicate the issue, diode recovery
times and charge reported on datasheets are often
the sum of COSS displacement current and the
recovered minority carrier current, QRR, and the
reactive currents arising from test circuit loop
inductance and capacitance, as shown in Figure 25.
As a result, the “QRR” number reported on the
datasheet is heavily dependent on the influence of
currents arising from a non-ideal testing
environment. Test circuits tend to have higher
inductance values compared to a well-designed DCDC PCB, since accommodations are made for
current sensing and test sockets.
dIF/dt
tA phase w/
inductance
tA phase
tB phase w/
inductance
tB phase
ISD(Q2)
VDS(Q2)
tB3
tB
QRR showing
QOSS
contribution
A common method to reduce both QRR (and diode
conduction losses) is the insertion of a Schottky
diode placed in parallel to the body diode. This
technique, while solid in theory, rarely gives the
optimal benefit of reducing both conduction and
QRR losses because the diode is physically separated
from the body diode with parasitic package and
wiring inductance. Moreover, gains in efficiency
quickly diminish as dead times reduce, since the
Schottky diode becomes less efficient at transferring
load current.
b)
SyncFET MOSFET
The most efficient method of minimizing QRRrelated switching losses is with a monolithically
integrated Schottky diode on the low-side MOSFET
die (SyncFET MOSFET). This can be one of the
most important features added to the low-side
MOSFET. A well-designed Schottky diode features
decreased dead-time diode conduction losses,
dramatically reduced switching losses attributed to
QRR, softened diode recovery resulting in lower
drain voltage stress and ring energy, and additional
output capacitance (COSS), further reducing the
recovery dVDS/dt and aiding the prevention of C x
dVDS/dt turn-on. In many circuits, the SyncFET
MOSFET provides much higher efficiencies at
lighter loads where QRR related losses dominate
total losses, as shown in Figure 26.
QRR with
circuit
inductance
tA1 tA2 tA3
Fig. 25. Diode QRR showing COSS displacement current.
While there are many factors to consider when
using datasheet QRR for MOSFET loss estimates,
TCAD simulations show that the combined effect
from diode QRR plus the stored energy from the
parasitic loop inductance generally equates to a
perceived two-to-three-times increase in QRR (from
reactive currents) when used in a typical DC-DC
converter, as simulated in Figure 14. This increase
is compared to an in-house QRR test circuit. A
conservative estimate for QRR losses typically takes
into account the ½LI2 losses from loop inductance.
14
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
HS.Packg
0.7%
HS.drv
1.9%
HS.Ton
6.7%
IOUT =
10A
LS.Cond
20.5%
HS.Toff
8.5%
LS.Packg
5.2%
HS.Cond
6.6%
LS.diode cond
3.4%
LS.drv
4.6%
LS.Swch
42.1%
HS.Packg
1.4%
HS.drv
0.5%
HS.Ton
10.0%
IOUT =
25A
HS.Toff
11.8%
LS.Cond
37.5%
Typically, a “shoot-through” condition arises from
capacitive feedback current through CGD into CGS
inducing a gate-bounce-induced channel turn-on of
the synchronous MOSFET, as shown in Figures 27
and 28. Holding the gate below threshold is
challenging because the high-frequency capacitive
displacement current from CGD (due to dVDS/dt)
couples back to circuit ground through the gate
electrode. The gate-to-ground impedance is the
parallel combination of the gate drive (ZG_DRV) and
the MOSFET gate-to-source (ZMOS_Gate) paths. As
dVDS/dt increases, the more favorable path for
displacement current is through the capacitive gatesource (CGS) path versus the highly inductive and
resistive gate drive loop.
HS.Cond
12.9%
LS.drv
1.3%
Ld_LS
LS.Packg
LS.Swch
LS.diode cond 10.5%
9.2%
4.9%
RDRIVE_LS
LG_brd LG_pack RG
CGD
CGS
Fig. 26. TCAD simulated switching loss.
ZGate_Drv ~ R + ωL
Another advantage of the SyncFET MOSFET is
that the monolithic nature of the Schottky diode cell
creates a high-frequency path from the MOSFET
channel to Schottky diode, which essentially
guarantees the PN body diode never completely
turns on (at reasonable currents). This situation is
much different from the case where an external
Schottky diode is placed in parallel to the MOSFET
body diode through an inductive loop created from
packaging and layout inductance. With the
SyncFET MOSFET, the reduction in QRR is
accomplished by adding Schottky diode area into
the MOSFET by an amount much smaller in area
than the discrete Schottky diode. Moreover, reduced
switching losses from the SyncFET MOSFET can
also be balanced with a higher RDS(ON) MOSFET to
achieve very high light-to mid-load efficiency while
still attaining similar heavy-load converter
efficiency compared to a non-SyncFET MOSFET.
c)
Shoot-Through and
Cross-Conduction
Another common loss encountered in high-speed
DC-DC circuits is an unwanted CGD x dVDS/dt (C x
dv/dt) induced turn-on of the channel[15][16].
ZMOS_Gate ~ 1/ωC
Q2
Ls_LS
Fig. 27. Synchronous rectifier switching waveforms showing [C x dv/dt]
induced turn-on.
Often, the most notable feature about C x dv/dt
turn-on is the changing (shallower) slope of the
drain voltage waveform as the channel turns on. It is
rarely a destructive event and more of a nuisance
causing increased power losses. The event is also
self-limiting since, as the channel turns on, dVDS/dt
decreases, which allows for the gate voltage to
discharge slowly, turning the channel back off. C x
dv/dt induced turn-on is encountered frequently in
synchronous buck designs and can actually aid in
limiting VDS stress during diode recovery when
parasitic inductance is included. However, it is
typically recommended to design a PCB and select
compatible MOSFETS / gate drivers to avoid shootthrough for maximum performance.
It is strongly recommended to select a MOSFET
with low internal gate resistance (RG) and lay out a
PCB with low gate loop inductance to maximize
low-side gate drive peak current capability, which
15
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
minimizes gate bounce and shoot-through currents.
Using MOSFETs with low QGD/QGS ensures that
shoot-through is minimized. Selecting low-QGD/QGS
MOSFETs provides minimizing gate bounce since
the total charge delivered to CGS from CGD results in
a lower gate bounce for a given gate impedance.
Low Side Channel Turn-On
the preset threshold (~1V), while the MOSFET VGS
is at a much higher potential, where the channel is
still gated on. Unlike C x dv/dt induced turn-on,
cross-conduction due to gate overlap can cause
excessive power dissipation that can damage the
MOSFETs. This is a situation that must be avoided
to attain high manufacturing reliability.
VGS(Q1)
VGS(Q1)
Ld_LS
CGD
VDS(Q2)
VDS(Q2)
ZG
CGS
VGS(Q2)
VDRV
Q2
VGS(Q2)
VGS
1V
VDRV
Ls_LS
VGS(Q2)
w/ Lsource
Fig. 29. Cross-conduction.
VgsHS
30.0
30.0
VdsLS
VdsLS
18.0
27.5
VgsLS
18.0
27.5
VgsLS
16.0
25.0
3.580
3.585
3.590
3.595
3.600
3.605
-6
time (seconds) *10
3.610
3.615
3.620
3.625
3.630
8.0
10.0
VgsHS (V)
6.0
2.0
0.0
-2.0
0.0
-2.5
2.5
0.0
-2.0
2.0
5.0
4.0
7.5
4.0
7.5
VGS(HS)
VGS(LS)
3.575
14.0
12.0
20.0
17.5
VdsLS (V)
12.5
15.0
10.0
8.0
10.0
VgsHS (V)
VGS(HS)
VGS(LS)
6.0
10.0
VdsLS (V)
12.5
15.0
17.5
12.0
20.0
14.0
22.5
16.0
25.0
VDS(Pk) = 29V
VDS(Pk) = 24V
22.5
20.0
FDMS8680_FDMS8660AS_Ls0p1_20A:0_0.1cycle.ivl<MEDICI_DATA>
VgsHS
20.0
FDMS8680_FDMS8660AS_Ls1nH_20A:0_0.1cycle.ivl<MEDICI_DATA>
5.0
Another approach to eliminate gate bounce is to
consider the low-side source inductance[10], which
can aid in preventing C x dv/dt shoot-through.
During high turn-on (low-side diode recovery
dIF/dt), the L x dIDS/dt voltage generated across
Ls_LS actually drives the low-side MOSFET source
node positive with respect to ground, which acts to
charge VGS negative, as shown in Figure 27. Driving
the gate negative allows added headroom for gate
bounce since the ∆VGS from C x dv/dt is applied to
a negative gate potential.
A shoot-through or “cross-conducting” situation
can also arise when an overlap of gate signals
causes both MOSFETs to simultaneously conduct,
as seen in Figure 29. Since adaptive dead-time
algorithms are usually used, cross-conduction
generally arises due to a layout-related issue or an
interaction with a MOSFET parameter, such as RG.
Figure 29 depicts a common situation where both
RG and LG (impedance ZG) cause the MOSFET
gate-to-source signal to differ dramatically from the
gate
driver
signal,
causing
errors
in
[15]
measurement . This typically reduces the dead
time since the gate drive is sensing a voltage below
2.5
Fig. 28. Synchronous rectifier switching waveforms showing [C x
dVDS/dt] induced turn-on.
Figure 30 compares TCAD simulated shootthrough to cross-conduction for the same MOSFETs
and circuit inductance from Figure 15, where the
difference is the dead time.
0.0
t0 tA tB
-2.5
IDS
3.55
3.56
3.57
3.58
3.59
-6
time (seconds) *10
3.60
3.61
3.62
3.63
Fig. 30. TCAD simulated synchronous rectifier switching waveforms
showing cross-conduction vs. dv/dt induced shoot-through.
The plot shows that the two events are much
different in nature, where the cross-conduction often
times can allow a very large cross-conducting
current to create a large drain voltage overshoot. It
also limits dv/dt, as the switching is slowed due to
the simultaneous high-side and low-side MOSFET
conduction. In contrast, C x dv/dt typically has a
fast initial dv/dt, but tends to limit the peak VDS as
the channel turns on.
16
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
A. Die Size vs. Efficiency
Generally, in a low-inductance circuit with low
packaged inductance power MOSFETs, the RDS(ON)
x QG(TOT) (QGD) FOM set certain expectations and
trends for efficiencies. For example, it is common
knowledge that a control MOSFET needs to be
optimized for both switching and conduction losses
if high converter efficiency is desired. It should also
be expected (from loss equations) that, for a family
of die sizes (for a given FOM silicon technology in
similar packages), an optimal die size exists where
efficiencies can be maximized across the useable
load current range. This reasoning holds for both
high- and low-side MOSFET selections.
1) High-Side Die Size for 12VIN and 19VIN V-Core
Figure 31 provides a comparison of 3x3mm
PQFN high-side MOSFETs for efficiency at 12VIN
and 19VIN. The test platform is a dual-phase
300kHz operating notebook V-Core regulator. Part
names with typical specifications are listed below.
For 12VIN, as expected, the smaller lower QG(TOT)
MOSFET (FDMC8296) excels at lighter load
efficiency where switching losses dominate, while
the lower RDS(ON) (FDMC8676) becomes more
efficient at heavier loads where RDS(ON) losses
dominate. At 19VIN, where high-side switching
losses can be very large, the FDMC8296 gives
higher efficiency across the load current range.
DEVICE
FDMC8676
FDMC8296
RDSON
QG(5) QGD QGS
RG
Vt
10VGS 4.5VGS
(mΩ) (mΩ)
(nC) (nC) (nC)
(Ω)
(V)
10
7.6
0.8
0.9
1.8
1.9
Package
3x3 PQFN
3x3 PQFN
4.7
6.5
7.1
9.5
3
3
4
2.5
91
Efficiency (%)
87
83
FDMC8296 - FDMS8660AS - 12Vin
FDMC8676 - FDMS8660AS - 12Vin
79
FDMC8296 - FDMS8660AS - 19Vin
FDMC8676 - FDMS8660AS -19Vin
75
0
10
20
30
40
50
Iout (A)
Fig. 31. Efficiency with varying die-size high-side MOSFETs; two-phase
notebook V-Core [VIN=12&19V, VOUT=1.3V, fSW=300kHz, VG=5V,
L=0.56µH, HS RDRV (source / sink = 0.8Ω
Ω ), LS RDRV (sink = 0.5Ω
Ω / source
= 1.0Ω
Ω )].
2) Low Side Die Size for 19VIN V-Core
A similar efficiency versus IOUT trade-off exists
for the low-side MOSFET die size shown in Figure
32. This case shows the smaller die, higher RDS(ON)
MOSFETs reduce switching losses at lower currents
due to lower QRR and QG(TOT), but become less
efficient at higher currents compared to larger die as
RDS(ON) losses dominate. The test platform is a
single-phase point of load (POL).
RDSON
QG(5) QGD QGS
RG
10VGS 4.5VGS
(mΩ) (mΩ)
(nC) (nC) (nC)
(Ω)
(V)
1.0
1.4
1.1
1.5
1.5
1.5
DEVICE
FDMS8660S
FDMS8670S
FDMS8672S
5x6 PQFN
5x6 PQFN
5x6 PQFN
1.9
2.8
4.0
2.6
3.6
5.2
44
24
16
16
10
6
11
8
5
Vt
90
Efficiency (%)
VI. EFFICIENCY MEASUREMENTS
This section provides and reviews measured
efficiency curves covering many of the topics
discussed throughout this paper. The examples
encompass many synchronous buck applications
from 12V input desktop computer voltage core (VCore) regulators (D-PAK MOSFETs) to 19.5VIN
Notebook V-Core with both SO-8 and PQFN
MOSFETs to generic synchronous buck POLs.
Many of the effects are studied with the aid of
measured switching waveforms to determine
efficiency trends. Cases reviewed are: A.) die size
versus efficiency, B.) packaging effects, C.) C x
dv/dt induced shoot-through, D.) cross-conduction
between high- and low-side MOSFETs, E.)
SyncFET MOSFET versus externally placed
Schottky diode.
85
FDMS8680 - FDMS8660S - 19Vin
80
FDMS8680 - FDMS8670S - 19Vin
FDMS8680 - FDMS8672S - 19Vin
75
0
5
10
15
20
25
Iout (A)
Fig. 32. Efficiency with varying die-size low side MOSFETs; single-phase
POL - [VIN=19V, VOUT=1.3V, fSW=500kHz, VG=5V, HS RDRV (pull up /
down = 1Ω
Ω ), LS RDRV (pull down = 0.5Ω
Ω / pull up = 1Ω
Ω )].
Typically, when trends such as these exist, the
power MOSFETs are operating as expected, where
gate drive and MOSFET RG / CISS time constants
17
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
are dictating the switching behavior. However, these
trends change quickly when very low impedance
gate drivers are used in combination with slight
increases in source inductance.
B. Packaging Effects
As noted throughout this paper, high packaging
parasitics combined with a less than ideal layout can
degrade (or severely degrade) the converter
efficiency. A few cases presented below compare
various packaging technologies and their influence
on efficiencies.
1) 5mm x 6mm PQFN High-Side vs. Die Size
Selecting a high-side MOSFET for optimized
efficiency usually requires a brute-force method of
narrowing down enhanced FOM MOSFETs
combined with a considerable amount of lab data.
This brute-force method is a common technique
since, more often than not, datasheet parameters and
FOM don’t fully correlate with lab data. One reason
for unexpected efficiency trends or lower efficiency
is package inductance and at times, very modest
inductance.
DEVICE
FDMS8692
FDMS8680
RDSON
QG(5) QGD QGS
RG
Vt
10VGS 4.5VGS
(mΩ) (mΩ)
(nC) (nC) (nC)
(Ω)
(V)
1.0
0.8
1.8
1.8
Package
5x6 PQFN
5x6 PQFN
7.0
5.5
10.5
8.5
8
10
2.1
2.7
2.7
3.2
Efficiency (%)
91
advantage across the entire load current range for
the FDMS8680, which seems to contradict the
example shown in Figure 30. What should be noted
here is that random selection of MOSFETs can yield
unexpected efficiency trends as manufacturing
tolerances are factored in (i.e MOSFET FOM has
variance influencing efficiency data). For this test
case, efficiency trends were verified over numerous
testing of random MOSFET date codes. The reason
behind the increased efficiency is package
inductance. The larger die FDMS8680 is
constructed with slightly lower package inductance,
giving rise to lower switching losses.
2) 3mm x 3mm vs. 5mm x 6mm PQFN High Side
The 3mm x 3mm PQFN has quickly emerged as
an excellent contender for high-efficiency high-side
MOSFET in high-current (>25A) POLs. 30V BVDSS
rated trench MOSFETs have pushed the typical
10VGS RDS(ON) of the packaged device to 4mΩ and
will continue to achieve unprecedented onresistance in the near future, where a single 3x3mm
packaged device begins to realize high-current
operation when used as a synchronous rectifier.
Efficiency curves in Figure 34 compare similar
MOSFET die packaged in both 3x3mm PQFN
(FDMC8296) and 5x6mm PQFN (FDMS8692 –
used in previous example), using the same low-side
MOSFET. The MOSFET parameters are shown
below.
87
DEVICE
83
FDMS8692
FDMC8296
FDMS8680 - FDMS8660AS - 12Vin
Package
5x6 PQFN
3x3 PQFN
RDSON
QG(5) QGD QGS
RG
10VGS 4.5VGS
(mΩ) (mΩ)
(nC) (nC) (nC)
(Ω)
(V)
8
7.6
1.0
0.9
1.8
1.9
7.0
6.5
10.5
9.5
2.1
2.5
2.7
3
Vt
FDMS8692 - FDMS8660AS - 12Vin
79
FDMS8680 - FDMS8660AS - 19Vin
FDMS8692 - FDMS8660AS - 19Vin
75
0
10
20
30
40
50
Iout (A)
Fig. 33. 5mm x 6mm versus 3mm x 3mm high-side comparison; twophase notebook V-Core [VIN=12&19V, VOUT=1.3V, fSW=300kHz, VG=5V,
L=0.56µH, HS RDRV (source / sink = 0.8Ω
Ω ), LS RDRV (sink = 0.5Ω
Ω / source
= 1.0Ω
Ω )]
Again, this efficiency trend indicates that circuit
inductance is beginning to influence switching. For
this comparison, the benefit of the FDMC8296
(3x3mm) over the FDMS8692 (5x6mm) is a
combination of enhanced package inductance and
resistance, which works together to increase
efficiency across the entire load current.
Figure 32 provides an efficiency example
comparing two high-side MOSFETs of similar
silicon technology (FDMS8680 vs 8692) scaled for
die size and packaged in a 5mm x 6mm PQFN. The
curve shows a clear and distinct efficiency
18
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
90
91
83
Efficiency ( % )
FDMC8296 - FDMS8660AS - 12Vin
85
FDD6296x1 HS, FDD8896x2 LS
80
FDMS8690x1 HS, FDD8896x2 LS
FDMS8692 - FDMS8660AS - 12Vin
79
FDD8880x1 HS, FDD8896x2 LS
FDMC8296 - FDMS8660AS - 19Vin
75
FDMS8692 - FDMS8660AS - 19Vin
0
75
10
20
30
40
50
60
70
80
Iout ( amps )
20
30
40
50
Iout (A)
Fig. 34. 5mm x 6mm high-side die size comparison; two-phase notebook
V-Core [VIN=12&19V, VOUT=1.3V, fSW=300kHz, VG=5V, L=0.56µH, HS
RDRV (source / sink = 0.8Ω
Ω ), LS RDRV (sink = 0.5Ω
Ω / source = 1.0Ω
Ω )]
It is worth noting that the FDMC8296 (3x3mm
PQFN) rivals the larger FDMS8680 (5x6mm PQFN
in Figure 32) for efficiency. This is a very
compelling advantage for converter designers
wishing to simultaneously increase power density
with efficiency.
DEVICE
Package
RDSON
10VGS 4.5VGS
(mΩ) (mΩ)
FDD8880
FDD6296
FDMS8690
D-PAK
D-PAK
5x6 MLP
7.0
7.0
7.4
9
9
9.9
QG(5) QGD QGS
RG
Vt
C. CGD x dVDS/dt Shoot-Through
Shoot-through arising from CGD x dVDS/dt
induced turn-on is one of the most common reasons
for decreased efficiency, especially at lighter load
currents. Waveform measurement of the low-side
gate and drain are typically the easiest method to
check for shoot-through[15], as shown in Figure 36.
Low Side Vds during diode recovery
Comparison waveforms of C dVds/dt induced shoot through versus no C dVds/dt.
35
7
VDS with C dv/dt
VDS without C dv/dt
Vds high Qgd device - C dv/dt shoot thru
30
6
Vds typical device - no C dv/dt shoot thru
Vgs high Qgd device - C dv/dt shoot thru
Vgs typical device - no C dv/dt shoot thru
25
5
20
4
15
3
10
2
5
1
0
0
-5
-1
-10
20
40
60
80
100
time ( nsec )
(nC) (nC) (nC)
(Ω)
(V)
13
5
12.2 3.5
10 2.9
1.3
1.1
1.7
1.6
3.8
4
3.5
While the enhanced FOM FDD6296 outperforms
the higher gate charge FDD8880, the advantage
goes to the lower inductance MLP (FDMS8690).
VGS (1V/div)
3) D-PAK vs. PQFN
Figures 33 and 34 describe cases where mild
package parasitics modestly influence MOSFET
switching. When dealing with packages such as the
D-PAK, the inductive influence on switching
behavior can be severe. Figure 35 shows an
example comparing efficiencies when using two
different FOM D-PAK high-side devices along with
a comparison to 5x6mm MLP.
Fig. 35. D-PAK vs. MLP high-side efficiency; [VIN=12V, VOUT=1.3V,
fSW=300kHz, VG=12V, RDRV HS (pull up / down = 3.8Ω
Ω / 1.4Ω
Ω ), RDRV LS
(pull up / down = 3.4Ω
Ω / 1.4Ω
Ω )]
Vds ( volts )
10
VDS (5V/div)
0
Vgs ( volts )
Efficiency (%)
87
VGS with C dv/dt
VGS without C dv/dt
120
-2
140
Time ( 20ns/div)
Fig. 36. Low-side switching waveforms for C x dv/dt
The waveforms of Figure 36 combined with an
efficiency test from similar FOM low die
MOSFETs with varying RDS(ON) (scaled die sizes)
provides for a clear picture of the problem, as
shown in Figure 37. The curves show a seven-point
efficiency difference for the FDMS8660S to the
FDMS8672S at 7A output current, which is a sign
of excessive switching losses. This situation
typically indicates that the layout and driver are not
working well with the selected MOSFETs.
19
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
RDSON
QG(5) QGD QGS
RG
Vt
10VGS 4.5VGS
(mΩ) (mΩ)
(nC) (nC) (nC)
(Ω)
(V)
1.0
1.4
1.1
1.5
1.5
1.5
DEVICE
FDMS8660S
FDMS8670S
FDMS8672S
5x6 PQFN
5x6 PQFN
5x6 PQFN
1.9
2.8
4.0
2.6
3.6
5.2
44
24
16
16
10
6
11
8
5
90
Efficiency [%]
85
80
FDMS8680 +FDMS8660S
75
to the gate loop layout. Specifically, oversized wide
gate traces and small gate loop area for minimizing
gate inductance, along with low MOSFET RG, are
recommended for the low-side MOSFET.
Figure 39 is an efficiency comparison depicting a
cross-conducting situation. This example presents
two low-side SyncFET MOSFETs that are diescaled for size. In this example, the FDS6699S
(larger die) shows a much lower than expected
efficiency, while the FDS6688S (smaller die)
performs well.
FDMS8680 +FDMS8670S
90 90
FDMS8680 +FDMS8672S
5
10
15
Load Current [A]
20
25
Fig. 37. Efficiency with higher QGD low-side MOSFETs; one-phase POL
[VIN=19V, VOUT=1.3V, fSW=500kHz, VG=5V, L=0.22µH, HS RDRV (pull
up /down=1Ω
Ω ), LS RDRV(pull down/up=0.5Ω
Ω /1Ω
Ω )]
For this example, it is recommended to use
MOSFETs with lower QGD (or QGD/QGS), which can
alleviate the problem shown in Figure 38.
DEVICE
FDMS8660AS
FDMS8670AS
FDMS8672AS
Package
5x6 PQFN
5x6 PQFN
5x6 PQFN
RDSON
QG(5) QGD QGS
RG
Vt
10VGS 4.5VGS
(mΩ) (mΩ)
(nC) (nC) (nC)
(Ω)
(V)
1.2
0.9
0.8
1.7
1.7
1.8
1.7
2.4
4.0
2.3
3.5
5.2
30
20
15
5.2
4
3.4
12
7.2
5.6
Efficiency [%]
85
80
FDMS8680 +FDMS8660AS
FDMS8680 +FDMS8670AS
FDMS8680 +FDMS8672AS
70
0
5
10
15
Load Current [A]
20
80 80
I
FDS6294x2 HS, FDS6688Sx2
LS
FDS6294x2
– FDS6688Sx2
FDS6294x2
–
FDS6699Sx2
FDS6294x2 HS, FDS6699Sx2 LS
75 75
70 70 0
0
(A)
5
10
15
20
25
30
5
10
15
20
25
30
Fig. 39. Efficiency comparison (cross-conduction); [VIN=19V,
VOUT=1.3V, fSW=300kHz, VG=5V, L=0.7µH, RDRV HS (pull up / down =
1.0Ω
Ω , RDRV LS pull down= 0.5Ω
Ω / pull up= 1Ω
Ω )]
90
75
85 85
Efficiency ( % )
0
Efficiency (%)
70
25
Fig. 38. - Efficiency with lower QGD low-side MOSFETs; one-phase POL
[VIN=19V, VOUT=1.3V, fSW=500kHz, VG=5V, L=0.22µH, HS RDRV (pull
up /down=1Ω
Ω ), LS RDRV (pull down/up=0.5Ω
Ω /1Ω
Ω )]
D. Cross-Conduction
Cross-conduction typically creates a clear and
distinct overlap in high-side and low-side gate
voltages. Since most popular gate drivers use an
adaptive dead time, this is often considered a nonissue. However, as many datasheets and application
notes point out; for the gate drive adaptive dead
time to operate correctly, attention needs to be paid
The first step in diagnosing the problem is to
measure low-side VDS and VGS waveforms. These
two waveforms contain enough information to
reveal whether the problem is C x dv/dt or crossconduction related. For this example, the measured
waveforms of Figure 40 show the limited (near nonexistent) body diode conduction, which is a signal
of high- and low-side gate overlap. For the body
diode measurement, the low-side VDS (or phase
node) waveform requires a zoom-in on the region of
interest (body diode conduction).
The gate driver used has an adaptive dead time of
20ns, which should yield distinct (and intended)
diode conduction shown in Figure 41 (same gate
driver used in combination with a low impedance
gate loop layout).
20
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
Fig. 42. Body diode conduction for FDS6688S vs. FDS6699S.
FDS6699S
diode
beginning
conduction
VGS = 1V/div, VDS = 5V/div
VGS = 1V/div, VDS = 0.1V/div
One of the common reasons for cross-conduction
arises from MOSFET RG in combination with high
gate loop inductance from layout (as is the case for
the efficiency curve, see Figure 39). This situation
causes dramatic signal differences from the low-side
MOSFET VGS versus the low-side drive-to-ground
voltage, shown in Figure 43.
Time (10 ns/div)
Fig. 40. Zoom of VDS for body diode voltage.
The diode conduction of the FDS6688S, shown in
Figure 42, is still not in the full body diode
conduction stage (channel partially on), but the
channel current is low enough to nearly prevent
cross-conduction. This is a situation where any
cross-conduction currents would be equal to or less
than diode related losses (VF and QRR). It is
important to note that any changes in MOSFET or
driver parameters can cause dramatic changes in
efficiency.
In addition, operating on the edge of crossconduction can be desirable for eliminating diode
related losses, but typically requires a more
advanced anti-cross-conduction algorithm.
VGS = 1V/div, VDS = 0.1V/div
VIN = 19V
VOUT = 1.2V
IOUT = 15A
L = 0.45µH
Diode
conduction
Time (10 ns/div)
Fig. 41. - MOSFET with ample diode conduction.
VGS = 1V/div, VDS = 0.1V/div
FDS6688S
FDS6699S
Low-side
driver
voltage
measured
at package
pins
(1V/div)
VGS
measured
at
package
pins
(1V/div)
LS FET VDS (zoom)
1V/div
HS FET VGS
(1V/div)
Time (10 ns/div)
Fig. 43. MOSFET VGS vs. LS gate drive signal.
The severe differences in signals can “trick” the
low-side gate anti-cross-conduction circuitry into
detecting the low-side MOSFET gate as below the
preset threshold of (~1V), basically defeating the
adaptive nature of the circuit.
E. SyncFET MOSFET
SyncFET MOSFET devices have been very
successful in power-sensitive applications, such as
computer notebooks and server regulators. High
efficiencies at light and mid loads are becoming
increasingly important as designs strive to increase
battery life, while also meeting Energy Star
requirements[17]. The advantages of SyncFET
MOSFETs are increased efficiency across the
usable load current range, while using a lower
component count and decreased switch-node
ringing, as shown in Figure 44.
Figure 45 is useful for comparing efficiencies of
the SyncFET MOSFET (FDS6299S) to an
equivalent non-SyncFET MOSFET (FDS6299) with
and without an external Schottky diode. The
SyncFET MOSFET advantage is clear (dead times ~
20ns).
Time (10 ns/div)
21
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
90
Efficiency (%)
VSW-NODE with SyncFET
85
FDS6694x1 HS, FDS6688Sx2 LS
FDS6694x1 HS, FDS6688x2 LS2 LS
80
0
5
10
15
Iout (amps)
20
25
Fig. 46. SyncFET vs. non-SyncFET comparison; [VIN=12V, VOUT=1.5V,
fSW=300kHz, VG=5V, L=1µH, RDRV HS (pull up / down = 1.5Ω
Ω , RDRV LS
pull down = 1Ω
Ω / pull up = 1.5Ω
Ω )].
Fig. 44. SyncFET impact on switch node voltage.
Efficiency ( % )
90
FDS6298x1 HS,
FDS6299Sx1 LS
88
86
84
FDS6298x1 HS,FDS6299
eqvlnt. LS no ext Schottky
82
FDS6298x1 HS, FDS6299
eqvlnt. LS with ext Schottky
80
78
0
5
10
15
20
IOUT ( Amps )
Fig. 45. SyncFET vs. external Schottky; one-phase POL - [VIN=19V,
VOUT=1.3V, fSW=340kHz, VG=5V, L=0.7µH, HS RDRV (pull up / down =
1Ω
Ω ), LS RDRV (pull down / up = 0.5Ω
Ω / 1Ω
Ω )].
While higher VIN (19.5V) battery applications are
known to highlight the SyncFET MOSFET
advantages, the actual benefit of the SyncFET
MOSFET is seen across many 12VIN (and even
5VIN) generic Point-of-Load (POL) converters.
Figure 46 provides an example comparing the
FDS6688 (non-SyncFET MOSFET) to the
FDS6688S (SyncFET MOSFET) in a generic 12VIN
POL. In this example, the efficiency gains actually
outweigh the previous 19VIN application. These
situations occur because the QRR-related switching
losses are a strong function of layout and gate drive
impedance.
VII. CONCLUSION
Power MOSFETs used as DC-DC converter
switches are often selected based on the RDS(ON) x
QG(TOT) (or QGD) FOM. Enhanced FOM typically
correlates well with high efficiency due to fast
switching control switches and high dVDS/dt
immunity for synchronous switches. While this
FOM combined with RG, QRR, and VTH (VPLATEAU)
provides insight into MOSFET performance,
calculated losses solely associated with these
parameters typically underestimate measured
MOSFET losses.
Additional switching losses often arise from
parasitic package or PCB layout inductance.
Excessive common-source inductance is often
encountered. This can significantly increase control
switch losses, slow switching speeds, and increase
voltage stresses during transients. For modern
MOSFETs with a low FOM, a reduction in source
inductance from (0.8nH to 0.4nH) results in a onepercentage improvement in efficiency at 25A.
In power-sensitive converters, care should be
taken in choosing low-inductance packages, such as
PQFN and MLP over packages such as the TO-252
(D-PAK). Parasitic inductance also increases
voltage and current stresses that can translate into
the need for higher BVDSS-rated MOSFETs with a
poorer Figure of Merit. To combat these stresses
and allow for the lowest possible BVDSS, additional
MOSFET features, such as the integrated Schottky
diode in the SyncFET MOSFET, alleviate the
voltage and current stresses, while providing
enhanced performance.
22
FAIRCHILD SEMICONDUCTOR POWER SEMINAR 2008 - 2009
REFERENCES
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
J. Klein, “Synchronous Buck MOSFET Loss Calculator with Excel
Model”, AN-6005, Fairchild Semiconductor.
ISL6227 Datasheet, Intersil, August 2007.
IRF7832 Datasheet, International Rectifier, June 2005.
Alan Elbanhawy, “Segmented Voltage Regulator Modules (VRM) as a
solution for CPU Core Voltage, AN-7018, Fairchild Semiconductor.
Alan Elbanhawy, “The Road to 200 Ampere VRM”, AN-7016,
Fairchild Semiconductor.
FDD8880 PSPICE Model, Fairchild Semiconductor, May 2003.
Mark Pavier, Andrew Sawle, Arthur Woodworth, Ralph Monteiro, Jason
Chiu, Carl Blake, “High-Frequency DC:DC Conversion : The Influence
of Package Parasitics”, Proc. APEC 2003.
Dan Calafut, “Trench Power MOSFET Low-Side Switch with
Optimized Integrated Schottky Diode SyncFET”, proc. ISPSD 2004.
Brian Lynch and Kurt Hesse, “Under the Hood of Low-Voltage DC/DC
Converters” ,Texas Instruments 2002 Power Supply Design Seminar.
Donald Schelle and Jorge Castorena, “Buck-Converter Design
Demystified”, Power Electronics Technology, June 2006.
Alan Elbanhawy, “Effect of Parasitic Inductance on Switching
Performance”, Proc. PCIM Europe, pp. 251-255.
Alan Elbanhawy, “Mathematical Treatment for HS MOSFET Turn Off”,
Proc PEDS 2003.
Alan Elbanhawy, “Effect of Parasitic Inductance on Switching
Performance of Synchronous Buck Converter”, Proc Intel Technology
Symposium 2003.
Chris Kocon, Jon Gladish, and Ashok Challa, “Advanced Physics-Based
Modeling of Power MOSFET Device Performance in the Synchronous
Buck Converter”, Proc. PCIM Europe 2006.
Jon Klein, “Shoot-Through in Synchronous Buck Converters”, AN6003, Fairchild Semiconductor.
Arthur Black, Jon Gladish, and Young-Sub Jeong, “Practical, Hands-on
Lab Experience in Addressing Shoot-Through in Synchronous Buck
Regulators”, Proc. PCIM Europe 2006.
Intel Corporation, “Energy Star System Implementation”, February 2007
revsion-001.
Jon Gladish is an application engineer at Fairchild
Semiconductor responsible for notebook power product
development. Prior to Fairchild, Jon worked at Harris
Semiconductor (Intersil) as an application engineer
focusing on the development of high voltage IGBTs, MCTs and
diodes for various AC-DC and DC-DC topologies. Jon’s professional
interests include developing high performance MOSFET and multichip modules (MCM) solutions for low voltage DC-DC converters.
Saber is a registered trademark of SabreMark Limited Partnership
and is used under license by Synopsys, Inc. All rights reserved.
23
Download