Sei sulla pagina 1di 7

A Digital PFC Controller without

Input Voltage Sensing


Barry Mather, Bhaskar Ramachandran and Dragan Maksimovic
Colorado Power Electronics Center, ECE Department, University of Colorado, Boulder, CO 80309-0425
B arry.Mather, B haskar.Ramachandran, Dragan.Maksimovic I @ colorado.edu

L i(t)
Abstract- This paper introduces a novel digital PFC (DPFC)
control approach that requires no input voltage sensing or
current loop compensation, yet results in low-harmonic operation
over a universal input voltage range and loads ranging from
high-power operation in continuous conduction mode down to
near-zero load. The controller is based on low-resolution DPWM
and A/D converters, requires no microcontroller or DSP
programming, and is well suited for a simple, low-cost
integrated-circuit realization, or as an HDL core suitable for
integration with other power control functions. Experimental
verification results are shown for a 500 W boost PFC rectifier.

LINTRODUCTION
Fig. 1: DPFC controlled boost rectifier
Power factor correction (PFC) boost rectifiers are used in a
wide range of applications that are required to meet the The purpose of this paper is to introduce a novel digital PFC
EN61000-3-2 standard. At low-to-medium power levels, the (DPFC) control approach based on a simple current control
transition-mode control (i.e. critical conduction mode, or law that allows operation in CCM without input voltage
operation at the boundary of continuous conduction mode sensing. Further objectives are to show how low-harmonic
(CCM) and discontinuous conduction mode (DCM)), which operation over a universal input voltage range and a wide
offers simplicity and performance advantages, is widely used range in power can be achieved using low-resolution A/D
and supported by a range of commercially available low-cost converters, a low-resolution digital pulse-width modulator
controllers. At higher power levels (above several hundred (DPWM) and minimal digital hardware. Figure 1 shows a
Watts), CCM operation is often preferred because of lower block diagram of the boost converter with the proposed DPFC
conduction losses and reduced EMI filtering requirements. controller.
Averaged current mode control in combination with a slow The paper is organized as follows: Section II introduces the
voltage control loop and a multiplier, which is a well-known DPFC current control law. Section III describes the voltage
control approach for CCM PFC, requires a more complex regulation loop and addresses mitigation of limit-cycling
implementation compared to the transition-mode control. With oscillations related to quantization of power. Section IV
the motivation of simplicity comparable to transition-mode or discusses system implementation, quantization issues and
DCM operation, together with low-harmonic, low conduction experimental results for a 500 W boost DPFC rectifier.
loss, and low EMI performance in CCM, the nonlinear-carrier Conclusions are given in Section V.
(NLC) control technique was introduced in [1, 2]. This II. DIGITAL PFC CURRENT CONTROL LAW
technique eliminated the needs for input voltage sensing, With reference to Fig. 1, in a PFC rectifier the current
current loop compensation, and a precision analog multiplier.
control objective can be written as: ig= VgIRe, where vg is the
Variations in the implementation of the analog NLC control
rectified line voltage, ig is the low-frequency (average)
approach have been reported in [3-6], and are now used in
component of the inductor current, and Re is the emulated
commercially available controllers [7, 8].
Digital PFC controllers, offering improved system interface, resistance, Re = Vg,rmIs2P (P is the input power) [22]. Using the
power management features, support for multi-module
quasi-static relationship for the CCM boost converter,
operation, and improved voltage-loop dynamic responses, VOU(P-d) = vg, where d is the switch duty ratio, the current
control objective can be expressed as
have recently received increased attention. Most of the digital
PFC control techniques reported so far have been based on
9 (I d) Voutp
(I ) -(1-d)
=
(1)
DSP or microcontroller implementations (e.g. [9-19]), or have
relied on multiple current samples per switching period
Re V.,2 rms
u

[20, 21]. where

1-4244-0714-1/07/$20.00 C 2007 IEEE. 198


=
VoutP
2
grms =-

d[n] = 1- uig [n] = 1- ui[n]


R
VOut
e

takes the role of a power control signal. In analog NLC


approaches [1, 2], (1) is solved using a voltage comparator by
constructing a carrier waveform based on d -X t/T,. A different
approach, better suited for digital realization follows by
solving (1) for the duty cycle command d[n] directly as a
function of the current sample i[n]=ig[n], where i[n] represents
a sample of the inductor current ideally in the middle of the
switch on time or in the middle of switch off time, and u is the
power control signal:

Equation (3) is the basic version of the proposed DPFC


current control law. Note that (3) requires no input voltage
sensing or compensation in the current control loop. Figure 2
shows experimental waveforms illustrating operation of a
DPFC controller based on (3). The current A/D conversion
start (CONVST) and end of conversion (EOC) signals show
(2)

(3)
3-.

4-.

T(z=Gd
Ch3
..
...............OOV ..Ch2

.10VUo
SW- CH2
5.00 v w Ch4

(z) Gi,(z)=
5.006 V
.CONVh

Bw M 2. 00p's Ch3
5.00 V w

L2K1critz

where Kcrit = ReTJI2L is a parameter that determines the


operating mode (CCM or DCM) of the boost converter in the
fijI
Fig. 2: Experimental waveforms illustrating operation of the
boost PFC based on the DPFC current control law (3)
RT 1

PFC rectifier [22]. Based on (7), stability of the current control


loop can be examined using root locus with Kcrit as a gain
1
T-

2.2 v

I
I

(7)

how the inductor current is sampled in the middle of the parameter, as shown in Fig. 3(a). The root locus shows that the
switch off time (as in [18, 23]) and the duty cycle command current loop is stable (i.e. has a pole inside the unit circle) as
d[n] is then updated based on the current sample. A triangle- long as Kcrit < 1, which is the same condition that characterizes
wave DPWM facilitates current sampling in the middle of the the PFC boost operating in CCM over the entire line cycle
switch off time (or in the middle of the switch on time). [22]. In conclusion, the current loop based on (3) is small-
A. Stability of the current control loop signal stable at high power levels when the converter always
operates in CCM. At lower power levels, the boost converter
In this section, examine small-signal stability of the
we operates in DCM around the zero crossings of the ac line, and
current control law (3) starting from a discrete-time model in CCM around the peak of the ac line. Instability of the
based on the quasi-static approximation, i.e., assuming that the current loop in such CCM operation typically manifests itself
input voltage Vg and the output voltage Vou, can be considered as current period-doubling, which may be tolerated. It is
constant on the time scale of several switching periods T, nevertheless of interest to investigate modifications of the
From the large-signal relationship [23], current control law to achieve stable operation in CCM at
lighter loads, i.e. for larger values of Kcrit. In particular, we
i[n + 1] = i[n] + g
T - out (1 - d[n])T(4) considered a control law based on a weighted sum of two
current samples,
a small-signal discrete-time relation yields the control-to- d[n] 1-uig[n] =1 -u( i[n]+(1-a)i[n-1]) (8)
current transfer function:
The maximum stable Kcrit value is attained when cx = 0.75.
Gid (Z) -V L0 Ts (5) The corresponding root locus shown in Fig. 3(b) indicates
z-1
d() current loop stability for Kcrit < 2. This effectively extends the
range of stable operation from Kcrit < 1 to Kcrit < 2, or down to
Linearization of (3) gives the effective current-loop half of the power level at which the converter starts operating
compensator transfer function:

Gic(z)= d(z) -u (6)


i (z)

Combining (2), (5) and (6), we obtain the effective discrete-


time current control loop gain,

Fig. 3a. Fig 3b.


199 Fig. 3: Root-locus of T(z) for (3a) the current control law given in (3),
and (3b) the current control law in (8) for cx = 0.75
in DCM for a portion of the line period than previously. At the proposed DPFC controller to fast voltage control (as in [18]
even lighter loads, the converter is in DCM most of the time. or [ 19]), we sample the output voltage at the switching
B. Operation at light loads frequencyf, which is much higher than the line frequencyfine.
The large oversampling ratio fJfi,, is also very beneficial in
At very light loads the converter will operate in DCM during resolving an issue related to power quantization. Average
the entire input voltage line cycle. To maintain voltage power delivered to the load is determined by the value of the
regulation in light load conditions a further modification is power command u. In the interest of minimizing the hardware
added to the current control law, requirements, it is beneficial to reduce the word size for the
d[n] = dmax -u (a i[n] + (1 - a)i[n -1]) (9) power command. However, at high power levels the power
differential between a 1 LSB step in u may become
considerably large; meaning that the resolution of power
At light loads, the power control command u saturates at Umax
levels drops significantly compared to lower power levels.
which is set based on stability criteria. Beyond this point,
The quantization of u, however, means that the power
further reductions in power are accomplished by reducing the
delivered to the load may not match exactly the power
maximum duty cycle, dmax. At higher power, when u < umax, consumed by the load. As a result, the voltage loop enters a
dmax is equal to 1, thus reducing the control law (9) back to (8) low-frequency limit-cycling mode (subharmonic with respect
(or (3)). This modification makes voltage regulation possible to the line frequency) leading to highly undesirable operation
down to essentially zero load even for high input voltage for a PFC. As an example, Fig. 6(a) reveals low frequency
levels, at the expense of somewhat increased current distortion
oscillation near the apex of the current waveform.
at such light loads.
Next, let us consider how the power quantization problem
III. VOLTAGE REGULATION AND POWER CONTROL can be completely resolved in the DPFC controller. First,
A block diagram of the boost rectifier with the complete notice that the small-signal model in Fig. 5 includes an output
DPFC controller is shown in Fig. 4. Based on the sampled resistance equal to one half of the small-signal load resistance.
output voltage error, the voltage loop controller G, computes As a result, even in the case of an ideal power sink load (an
the power control signal u[n]. As described in the previous ideal downstream DC-DC regulator), there is a finite small-
section, as the power is reduced, the command u saturates at signal resistance loading the output, which helps stabilize the
Umax = 0.5 for (3). Control through dmax then enables power output voltage at zero average error even with finite power
control down to zero load. resolution of the controller. Second, given the large
Figure 5 shows a small-signal model of the power stage oversampling ratio fJfjine, XA modulation can be very effective
obtained by averaging over half line cycle [22], for the DPFC in improving effective resolution of the power command u,
current control law described in Section II. This model is used with minimal hardware overhead. This is even more so in
to design a standard slow voltage loop control where G, can digital PFC controllers than in digital controllers for DC-DC
be a simple linear PI compensator. To remove the need for converters where effective XA techniques were previously
synchronization to the ac line, and to facilitate extensions of described [24].
In the DPFC realization, we implemented a simple, first-
order XA modulator (XAu in Fig. 4) in the "error-feedback"
configuration [25], as shown in Fig. 7. The command with
word length of ni, bits is effectively reduced to a dithered
command with word length of n0u, < ni, bits. In the
experimental prototype, for u dithering, we have ni, = 16,
nout= 8. The result of u dithering can be observed in Fig. 6(b)
for the same operating conditions as in Fig. 6(a). In the
experimental results of Section IV we further show how u
d dithering through XA modulation removes undesirable
subharmonic limit-cycling under all operating conditions.
l l fii[n] ~~~~~~~~~~~~~Voltage-loopl
.
. ~~~~~~~~compensator .
d- DPFC Umx en
S ~~~~CDurrFent T9 7 \), CC
n
Control Law Umin| g2vg,rmsf j2U4 r2 C R \

P2 V2
I 1 Digital PFC Controller V
J2-v g2,rms ;r2 2P
Fig. 4: Complete DPFC controller. Experimental prototype parameters: universal
input voltage range, V0 = 380V, L = lmH, C = 230uF, transistor: FCP2ON60, Fig. 5: Averaged small signal model of the DPFC controlled
diode: 15ETX06, switching frequency: lOOkHz, maximum power: 500W boost for slow voltage loop design

200
Current ADC Resolution Vg = 120 Vrms Vg = 220 Vrms
(bits) mA/bit THD (%) THD (%)

:: ~~~Vg 8 33 4.3 2.3


7 65 4.4 2.4
6 130 4.7 2.7
..a
5 260 5.3 3.5
4 521 6.1 5.4
3 1042 8.1 9.1
Table 1: THD measurements for various current ADC resolutions, 4-bit DPWM,
*k
P= 500W

resolutions of the DPWM and the current sensing ADC were


investigated to this aim.
Chi 100 v. .
M2.OOms Ch2f
5.00 Mg 6OOmV
Fig. 6a.: no SA dithering implemented i. DPWM resolution

The resolution of the DPWM was variable from 1-bit to 9-


bits. During experimentation it was determined that the 1-bit
..
Vg and 2-bit settings were simply not feasible and represented
. .. .. .. extreme quantization effects. Implementing a first-order LA
............. .......................... ........ modulator (YA d in Fig. 4) to dither the duty-cycle comand d
allowed the DPWM resolution to be lowered to as low as 3-
.. .. .. .. ..
bits while maintaining harmonic limit standards at a 500 W
g.
.. .. .. .. .. ..
power level. A DPWM setting of 4-bits, with YA modulation
implemented, was found to be a suitable setting for the
operation of the boost PFC over the entire range of input
voltages and output power levels.
ii. Current sensing ADC resolution
Chi 100V S.OOVV M2.00ms
Fig. 6b: YA dithering only
. ,.
..... .Ch2f .. 6.lmV. .
The current sensing ADC had a variable resolution of 3-bits
to 8-bits. Table 1 shows how changing the current sense
... .. .... ... ....
resolution effects the total harmonic distortion (THD) of the
line current. The current sensing resolution in mA/bit is also
....
... Vg .. presented in Table 1. For this experiment the DPWM
resolution was set to 4-bits of resolution and YA modulation
... ... was enabled for both the u and d signals. The current LSB
quantization step is as high as 1 A in the case of a 3-bit current
........................... ......................
sense A/D, however, at 500W the harmonic current limits are
.. .. .
not exceeded.

iii. Combined quantization effects

Figure 9 shows three line current waveforms for different


YA modulation implementations. Real-time inspection of the
Cnh uuv 5.0uuV v2i.0uums Ouumv current waveform shown in Fig. 8a reveals low frequency
Fig. 6c: u and d SA dithering oscillation near the apex of the current waveform. Circles
Fig. 6: Converter waveforms for various SA dithering settings, P have been added to show the locations of the oscillations that
500W, Vg rws = 120 Vrms, 4-bit DPWM, 8-bit current sensing ADC lead to the aforementioned undesirable sub-harmonics.
Figure 8b shows the current waveform when IA, is
IV. SYSTEM IMPLEMENTATION implemented, the sub-harmonic steps are no longer present but
The experimental prototype consists of a boost rectifier a stair-step appearance still is present in the current waveform.
power stage and a digital controller implemented using a With YAd and YA, implemented, Fig. 8c produces a clean
Xilinx FPGA development platform. The digital controller waveform that easily meets harmonic current specifications.
allows experimentation with the resolutions of current sensing Figure 9 is a plot of the harmonic currents for the same
(up to 8-bit, 33 mA current LSB resolution), voltage sensing conditions as shown in Fig. 8. Harmonic current values are
(up to 8-bit, 3 V LSB output voltage resolution), and the shown for both the YAd and YA, enabled and IA, only settings.
201
Fig. 7: Error-feedback configuration of a first-order YA
modulator for dithering

The Class A harmonic current limits for the odd harmonics are
also shown. Without YAd active the 3-bit DPWM is not
capable of passing the harmonic current limits, with YA d
enabled the 3-bit DPWMV easily passes the harmonic current
limits. A total of six internal dithering bits (LSBs) were
utilized for the 3-bit DPWMV setting. Fig. 8a.: no YA dithering implemented

B. Experimental waveforms
Vg
i. Operation at high and moderate power

Table 2 shows the experimental line current (ia3, the I1-+I


rectified line voltage (vg) and the output voltage ripple (vrippie)
waveforms for different line voltages (220 Vrms and 120 Vrms)
and loads (500 W and 100 W) with a 4-bit DPWM and a 4-bit lac
current sensing ADC. The Class A harmonic current limits
are mieet fur all operating conditions.

ii. Operation at very low power

Figure 10 shows the converter waveforms at a line voltage Chi: 10''' .... S.OOVQ2 M2.OOms Ch2\-c 2.3 V
of 220 Vrms and a load of 20 W. The converter output is still
Fig. 8b: YA, dithering only
regulated and the line current meets harmonic current
standards although the THD is compromnised.
V. CONCLUSIONS Vg

This paper introduced a novel digital PFC (DPFC) control


approach that requires no input voltage sensing or current loop 14-+
compensation, yet results in stable, low-harmonic operation
over a universal input voltage range and load ranging from
lac
high-load operation in continuous conduction mode down to
near-zero load. The controller architecture, together with
simple, first-order IA modulation blocks for control signal
dithering, enable DPFC controller realization with a low-
resolution DPWM, low resolution A/D converters, and a low
clock rate. The DPFC controller is suitable for stand-alone
low-cost custom-IC implementation, or as a Verilog HDL .Chi' bOYv 5.00 VQ M2.00ms 'Chf2\-V 2.3 V

(hardware description language) module well suited for Fig. 8c: u and d YA dithering

integration with other power control functions. Experimental Fig. 8: Converter waveforms for various YA dithering settings, P=

verification results are shown for a 500 W boost DPFC 500W, Vgrn<, 120 Vrm,s, 4-bit DPWM, 8-bit current sensing ADC

rectifier.

202
0.6- u dithering only

0.5- ~u and d dithering


r

1. odd harmonic limits


E

3 7 11 15 19 23 27 1 35 39
Fig. 10: Converter waveforms for very light load
Hamoi Number
operation, P =20W, Vg, =220 Vrm,s, 4-bit DPWM

Fig. 9: Harmonic currents with and without XA d dithering implemented, P and current sensing ADC

500W, Vg, 120 Vrm,s, 3-bit DPWM, 4-bit current sensing ADC

Vg,rms =l12OVrms Vg,rms = 220 Vrms


t --.

1 -+
P = IOOW
I
'ac
KE

Vripple Vripple
3-+

.V .m.. ,. F. . -.' kuu0V


k ... .FS. OWL. -. 'k k L. -; Lf2
F. A-.
Lfl I IUU
V ~~~~~~f .UU ivtu m Lfl CflI ME 2.00UVS I I-
iM2.uums
. . I 11 11 I
52umV
I I 11 II

Ch3 2.SO 0V%l Ch3 2.SO 0V"%


I

Vg

.a
1 -+
P = 500W

.r

,pp le

3-'
RE

1.1-
Chll . .'.k
100 V
1. " ..- - - --
5.UU VS
11- ..-
M2.OUms
',..-A '-'
1h12 'L
I- -
uumv chii bO v M 2.OOYSI M42.OOms Ch2X- SSOm Y,
am 10.0V%\ Ch3 IO.O0V"%
Table 2: Experimental waveforms for 100W and 500W for input voltages of 12OVm, and 220Vm,

203
REFERENCES [14] W. Zhang, G. Feng, Y. Lui, B. Wu, "A digital power factor correction
(PFC) control strategy optimized for DSP," IEEE Trans. on Power
[1] D. Maksimovic, Y. Jang, R. W. Erickson, "Nonlinear-carrier control for Electron., vol. 19, no. 6, pp. 1474-1485, Nov. 2004.
high-power-factor boost rectifiers," IEEE Trans. on Power Electron., [15] P. Mattavelli, W. Stefanutti, G. Spiazzi, P. Tenti, "Digital control of
vol. 11, no. 4, pp. 578 - 584, July 1996. single-phase power factor preregulators suitable for smart-power
[2] R. Zane, D. Maksimovic, "Nonlinear-carrier control for high-power- integration," IEEE PESC 2004, pp. 3195-3201.
factor rectifiers based on up-down switching converters," IEEE Trans. [16] L. T. Jakobsen, N. Nielsen, C. Wolf, M. A. E. Andersen, "Hybrid
on Power Electron., vol. 13, no. 2, pp. 213 - 221, March 1998. control method for a single phase PFC using a low cost microcontroller"
[3] J. Rajagopalan, F. C. Lee, P. Nora, "A general technique for derivation IEEEAPEC2005, pp. 1710-1715.
of average current mode control laws for single-phase power-factor- [17] P. Mattavelli, G. Spiazzi, P. Tenti, "Predictive digital control of power
correction circuits without input voltage sensing," IEEE Trans. on factor preregulators with input voltage estimation using disturbance
Power Electron., vol. 14, no. 4, pp. 663 - 672, July 1999. observers,", IEEE Trans. on Power Electron., vol. 20, no. 1, pp. 140 -
[4] Z. Lai, K. M. Smedley, "A family of continuous-conduction-mode 147, Jan. 2005.
power-factor-correction controllers based on the general pulse-width [18] A. Prodic, J. Chen, D. Maksimovic, R. W. Erickson, "Self-tuning
modulator," IEEE Trans. on Power Electron., vol. 13, no. 3, pp. 501 - digitally controlled low-harmonic rectifier having fast dynamic
510, May 1998 response," IEEE Trans. on Power Electron., vol. 18, no. 1, part 2, pp.
[5] J. P. Gegner, C. Q. Lee, "Linear peak current mode control: a simple 420-428, Jan. 2003.
active power factor correction control technique for continuous [19] A. Prodic, D. Maksimovic, R. W. Erickson, "Dead-zone digital
conduction mode," IEEE PESC 1996, pp. 196-202. controllers for improved dynamic response of low harmonic rectifiers,"
[6] S. Ben-Yaakov, I. Zeltser, "PWM converters with resistive input," IEEE Trans. on Power Electron., vol. 21, no. 1, pp. 173 - 181, Jan.
PCIM 1998, pp. 87-95. 2006.
[7] Infineon ICEIPCSO1, PFC-CCM (continuous conduction mode) control [20] P. Zumel, A. de Castro, 0. Garcia, T. Riesgo, J. Uceda, "A simple
IC for SMPS, www.infineon.com. digital hardware to control a PFC converter," IECON 2001. vol. 2, pp.
[8] R. Brown, M. Soldano, "One cycle control IC simplifies PFC design," 943 -948.
IEEE APEC 2005, IRI 150, www.irf.com. [21] A. de Castro, P. Zumel, 0. Garcia, T. Riesgo, J. Uceda, "Concurrent and
[9] C. Spiazzi, P. Mattavelli, L. Rossetto, "Power factor preregulators with simple digital controller of an AC/DC converter with power factor
improved dynamic response," IEEE Trans. on Power Electron., vol. 12, correction based on an FPGA," IEEE Trans. on Power Electron., vol.
no. 2, pp. 343 - 349, March 1997. 18, no. 1, part 2, pp. 334 - 343, Jan. 2003.
[10] S. Buso, P. Mattavelli, L. Rossetto, G. Spiazzi, "Simple digital control [22] R. W. Erickson, D. Maksimovic, Fundamentals of Power Electronics,
improving dynamic performance of power factor preregulators," IEEE 2nd ed, Norwell, MA, Kluwer, 2001.
Trans. on Power Electron., vol. 13, no. 5, pp. 814 - 823, Sept. 1998. [23] H. Peng, D. Maksimovic, "Digital current-mode controller for DC-DC
[11] S. Bibian, H. Jin, "Digital control with improved performance for boost converters," APEC 2005, vol. 2, pp. 899 - 905.
power factor correction circuits," IEEE APEC 2001, pp. 137-143. [24] Z. Lukic, K. Wang, and A. Prodic, "High-Frequency Digital Controller
[12] K. De Gusseme, D. Van de Sype, A. P. Van den Bossche, J. A. for DC-DC Converters Based on Multi-Bit SD Pulse-Width
Melkebeek, "Sample correction for digitally controlled boost PFC Modulation," IEEE APEC 2005.
converters operating both in CCM and DCM," IEEE APEC 2003, pp. [25] D. A. Johns, K. Martin, Analog Integrated Circuit Design, New York:
389-395. Wiley, 1997, Ch.14, pp. 531-547.
[13] S. Kim, P. Enjeti, "Control of multiple single phase PFC modules with a
single low-cost DSP," IEEE Trans. on Ind. Appl., vol. 18, no. 5, pp.
1379-1385, Sep. 2003.

204

Potrebbero piacerti anche