You are on page 1of 9

www.ietdl.

org
Published in IET Renewable Power Generation
Received on 27th April 2014
Revised on 24th August 2014
Accepted on 15th September 2014
doi: 10.1049/iet-rpg.2014.0149

ISSN 1752-1416

Comprehensive dynamic analysis of photovoltaic


generator interfacing DC–DC boost power stage
J. Viinamäki1, J. Jokipii1, T. Messo1, T. Suntio1, M. Sitbon2, A. Kuperman2
1
Department of Electrical Engineering, Tampere University of Technology, POB 692, Tampere FI-33101, Finland
2
Hybrid Energy Sources Laboratory, Department of Electrical Engineering and Electronics, Ariel University, POB 3,
Ariel 40700, Israel
E-mail: alonku@ariel.ac.il

Abstract: In transformer-less grid-connected renewable energy systems, interfacing of photovoltaic (PV) generators is typically
implemented by means of DC–DC boost-power-stage converter, mainly because of its voltage-boosting capability. In order to
track the maximum power point of a PV generator, input voltage of the converter is usually feedback controlled, forcing the
converter to operate as a current-sourced rather than voltage-sourced converter. Nevertheless, PV generator interfacing power
stage is commonly assumed to possess the same dynamic properties as corresponding voltage-sourced power stage.
Investigations presented in this study reveal explicitly that the dynamics of PV generator interfacing DC–DC boost power
stage resembles conventional buck power stage behaviour with duty-cycle independent resonance and additional right-half-
plane zeros. In addition, the duty cycle has to be decreased for increasing the corresponding output variables (i.e. input
voltage and output current). Extended experimental results are given to support the theoretical findings.

1 Introduction current) in PV applications [18]. Even though the PVG


influence on the steady state and dynamic behaviour is
Interfacing of photovoltaic generators (PVG) to the utility known for some time, analysis of PVG-interfacing
grid necessitates the use of single-stage or cascaded DC– converters is still performed in most cases as in the
AC power electronics converter structures [1–3], where the conventional voltage-sourced applications (e.g. [19, 20]).
number of stages corresponds to the number of cascaded Even though several recently published papers [21, 22]
converters in the system. In typical cascaded systems, DC– reveal some of the PVG-induced effects on boost converter
DC boost converter with an additional input capacitor is the operation (for reference, dynamic models of conventional
PVG-interfacing power stage [4–9] followed by a boost converter can be found e.g. from [23; pp. 77–79]), no
grid-interfacing inverter [9, 10]. Such a power stage single source comprehensively revealing coupled
resembles the pulse-width-modulated switching shunt unit PVG-boost power stage dynamics exists. Consequently, the
utilised in space applications [11–13] with an LC-type input main goal of this paper is to present a comprehensive
filter (cf. Fig. 1). Note that the PVG-interfacing converter is dynamic analysis of the boost power stage in
usually terminated by a constant-voltage (CV)-type load PVG-interfacing applications and clearly indicate the
such as storage battery (e.g. battery chargers in off-grid differences between conventional and PVG-interfacing
applications) or input-voltage-controlled grid-interfacing boost-power-stage dynamics. In addition, this paper
inverter. provides a methodology to analyse internally
It has been observed already in early 1970s that current-mode-controlled boost power stages discussed in
PVG-interfacing power stages exhibit quite different [22] as well.
steady-state and dynamic properties than conventional PVG is known to exhibit properties of both
voltage-sourced converters of the same topology [14–18] constant-current (CC) and CV sources depending on the
For example, it was shown that the output dynamics of a operating point [24, 25] as depicted in Fig. 2, where rpv, cpv
buck-converter-based power stage (which is minimum (both operation-point-dependent for given environmental
phase in the conventional applications) possesses a conditions), and ppv normalised PVG dynamic resistance,
right-half-plane (RHP) zero in photovoltaic (PV) capacitance and output power, respectively. At voltages less
applications [14]. Moreover, output-side feedback control than the maximum power point (MPP) voltage, the current
[15] and peak-current-mode-control [16, 17] make a PVG is approximately constant and the dynamic resistance rather
interfacing buck power stage unstable. In addition, unlike in high; hence PVG resembles a current source. Therefore this
conventional voltage-sourced applications, duty cycle of a characteristic curve section is often referred to as a constant
PVG interfacing converter has to be decreased for current (CC) region. On the other hand, at currents lower
increasing output variables (i.e. input voltage and output than the MPP current, the voltage is approximately constant

306 IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314
& The Institution of Engineering and Technology 2015 doi: 10.1049/iet-rpg.2014.0149
www.ietdl.org
2 Small-signal models of boost converter
Boost DC–DC converter under study with component
definitions is shown in Fig. 3. Small-signal models (referred
to as unterminated or internal) are derived first for voltage
and current mode controlled converters and then combined
with PVG model to obtain coupled dynamic behaviour
[23]. Coupled dynamic models are presented and discussed
Fig. 1 PVG interfacing by means of a DC–DC boost power stage in detail in Section 3.
Following guidelines in [31], one-switching-cycle (Ts)
averaged dynamics of the system in Fig. 3 is given by
and the dynamic resistance is rather low; hence PVG ⎧

⎪ dki l ku l r + drds + d ′ rd
resembles a voltage source. Consequently, this characteristic ⎪ L1 Ts = in Ts − L1
⎪ kiL1 lTs

⎪ dt L1 L1
curve section is often referred as constant voltage (CV) ⎪


⎪ d ′ (kuo lTs + UD )
region. The dual nature of PVG implies that the device is a ⎪
⎪ −
non-linear source, typically represented by Norton/Thevenin ⎪


⎪ L1
equivalent circuits in small-signal linearised dynamic ⎪


⎪ dku l ki l ki l
analysis [26]. Nevertheless, the PVG is inherently a current ⎪

C1 Ts
= − L1 Ts + in Ts
source; therefore the only feasible input-side-controlled dt C1 C1
′ (1)
variable is PVG voltage [10, 18, 25, 27–30]. ⎪
⎪ dku l d ki l ki l


C2 Ts
= L1 Ts
− o Ts
This paper investigates the dynamic properties of boost ⎪
⎪ dt C2 C

⎪  2
power stages used in PV applications, applying small-signal ⎪


⎪ d
analysis to the coupled PVG – boost converter system [31]. ⎪
⎪ kuin lTs = 1 + rC1 C1 kuC1 lTs

⎪ dt
It is revealed that input-side control dynamics does not ⎪
⎪  

⎪ d
incorporate any RHP zeros, which would affect the ⎪
⎩ kuo lTs = 1 + rC2 C2 kuC2 lTs
input-voltage controller design. Nevertheless, it is shown dt
that the LC-type input filter creates two different RHP zeros
in the output control dynamics of the converter if the PVG Small-signal dynamics of (1) is then obtained as
is operated in CC region. Moreover, it is shown that a
switched-mode converter under direct duty cycle control ⎧

⎪ dîL1 ûin rL1 + Drds + D′ rd
behaves as a DC transformer, automatically adopting input ⎪
⎪ = − îL1

⎪ dt L1 L1
source properties. Hence, it is quite important to realise the ⎪
⎪   

⎪ Uo + UD rds − rd D′
correct dynamic behaviour of the converter to assure ⎪
⎪ + − −
reliability as discussed in [32]. Moreover, small-signal ⎪
⎪ I d̂ û

⎪ L1 L1
L1 L1 o
models for average-current-mode-controlled boost power ⎪


⎪ dûC1 1 1
stage are developed as well, justifying the use of proposed ⎪
⎨ = − îL1 + îin
adaptive control proposed in [22]. dt C1 C1
(2)
The rest of the paper is organised as follows. Small-signal ⎪ dûC2 D′
⎪ I 1

⎪ = îL1 − L1 d̂ − îo
modelling of the boost-power-stage converter is described ⎪
⎪ dt C C C

⎪  2 2  2
in Section 2. PVG-induced effects on the converter ⎪


⎪ d
dynamics are discussed in Section 3. Experimental ⎪
⎪ û = 1 + rC1 C1 û
⎪ in
⎪ dt C1
validation of PVG-converter coupled system small-signal ⎪
⎪  

⎪ d
behaviour is presented in Section 4. Conclusions are drawn ⎪
⎩ ûo = 1 + rC2 C2 û
in Section 5. dt C2

linearised around
⎧ 

⎪ Uin = (rL + Drds + D′ rd )IL1 + D′ Uo + UD


⎨ IL1 = Iin
Io = D′ Iin (3)



⎪ UC1 = Uin

UC2 = Uo

Fig. 2 Normalised terminal characteristics of a PVG Fig. 3 Unterminated boost DC–DC power stage

IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314 307
doi: 10.1049/iet-rpg.2014.0149 & The Institution of Engineering and Technology 2015
www.ietdl.org
2.1 Voltage mode open-loop converter conventional boost converter, where such the dependence
exists [23; pp. 77–79]. The numerator of the
Voltage mode open-loop converter representation (in which v
control-to-output-current transfer function (Gco in (5)) is a
duty cycle is the control input and input voltage is the second-order polynomial possessing two distinct RHP roots,
controlled variable, see Fig. 4; single-loop control structure which can be approximated as
will be used with no additional feedbacks) can be derived
from (2) by applying Laplace transform and performing −I pv (s + vLow−freq )(s + vMedium−freq )
matrix manipulations. Resulting plant (transfer function
v
Gco ≃   
matrix form) may be represented in general form (Laplace s2 + s Req /L1 + 1/L1 C1
variable is omitted) as
1
⎡ ⎤ ⎡ ⎤ vLow−freq ≃ − (7)
îin îin C1 (D′ Ueq /Ipv − Req )
ûin Zinv Toiv Gciv ⎣ û ⎦ = H v ⎣ û ⎦
= (4) D′ Ueq /Ipv − Req
îo v
Gio −Yov v
Gco o o
d̂ d̂ vMedium−freq ≃ −
L1
where d̂ is the control input (generated by the loop controller);
The low-frequency RHP zero resembles the zero present in
îin and ûo are source and load imposed external disturbances,
the output-control dynamics of PVG interfacing buck power
respectively, and superscript ‘v’ denotes voltage mode
stage [30]. The medium-frequency RHP zero resembles the
converter. Individual open-loop transfer functions in (4) are
zero present in the output control-dynamics of a
⎧ conventional boost converter [23; pp. 77–79]. At the same
⎪ 1 time, control-to-input-voltage transfer function (Gciv in (5))


⎪ Zinv = D−1 (Req − rC1 + sL1 )(1 + srC1 C1 )

⎪ L does not contain any RHP zeros. Negative sign of


1


⎪ D control-related transfer functions dictates interchanging the

⎪ T v
= D −1
(1 + srC1 C1 )

⎪ oi
L1 C1 signs of feedback and reference signals in order to ensure



⎪ stable operation (or using negative controller gain,
⎪ v

Ueq

⎪ G = −D−1 (1 + srC1 C1 ) alternatively). It should be mentioned that
⎨ ci L1 C1 control-to-output-current transfer function is usually
′2 (5)
⎪ −1 D sC2 irrelevant in PVG interfacing. Nevertheless, DC-link

⎪ Y v
= D s+

⎪ o
L1 1 + srC2 C2 voltage has to be limited to a certain maximum value (for



⎪ ′ reliability/safety reasons), depending on the nature of the

⎪ D

⎪ Giov
= D−1 (1 + srC1 C1 ) device connected at the output terminal (capacitor/battery).

⎪ L1 C1 The limitation is realised by providing an extra feedback

⎪   ′  

⎪ loop from the output voltage. As well as for protection,

⎪ G = −D I s − s
−1 D U eq Req 1

⎩ co
v
pv
2
− + output voltage control may also be used during stand-alone
L1 Ipv L1 L1 C1
operation when the inverter controls the ac voltage.
The control-to-output-voltage transfer function is Gco/Yo;
with hence the above mentioned RHP zeros will affect the
⎧ output-voltage feedback loop design [28].

⎪ Ueq = Uo + UD + (rd − rds )Iin According to the boost-power-stage dynamic

⎨ R = r + r + Dr + D′ r representation in (5)–(7), modelling assuming ideal voltage
eq L1 C1 ds d
(6) source as an input source would yield quite different

⎪ R 1

⎩ D = s2 + s
eq
+ dynamic representation as given in [23; pp. 77–79], which
L1 L1 C1 does not describe the dynamic behaviour of the
PVG-interfacing converter. As discussed in [25], the PVG
has significant impact on the dynamic behaviour of
Note that Δ is similar to the denominator of a conventional interfacing converter. These issues are discussed more
buck-type converter transform functions, where the resonant thoroughly in Section 3.
frequency is duty cycle independent as opposed to the
2.2 Current mode open loop converter

In [22], an application of cascaded controlled


boost-power-stage as PVG power processing stage was
described, where inductor-current feedback and PVG
voltage feedback formed inner and outer loops,

Fig. 5 Open loop plant of boost power stage with closed inner
Fig. 4 Voltage mode open loop plant of boost power stage inductor current loop

308 IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314
& The Institution of Engineering and Technology 2015 doi: 10.1049/iet-rpg.2014.0149
www.ietdl.org
respectively. In case of cascaded control, inner current loop
alters dynamic behaviour of the converter, yielding different
set of transfer functions.
Open loop block diagram with closed inner inductor
current loop of a boost power stage is shown in Fig. 5, where
⎡⎤
îin
 ⎢ ⎥
îL = GiL GoL GcL ⎣ ûo ⎦ (8)

with GiL, GoL and GcL denoting transfer functions from input
variables to inductor current, derived from (3) as [23, pp. 19–21]

1 + srC1 C1 −1 D′ s −1 Fig. 6 Open loop plant of boost power stage with closed inner
GiL = Dc GoL = D
L1 C1 L1 c inductor current loop
(9)
Ueq s −1 Req 1
GcL = D Dc = s + s 2
+
L1 c L1 L1 C1 transfer functions in (10) reduce then to

1 + rC1 C1 s D′ IL1 sC2


Furthermore, Gcc denotes the transfer function of inductor Zinc = Yoc = +
C1 s Ueq 1 + srC2 C2
current loop controller and îREF is current loop reference
signal. Note that control-to-inductor-current transfer IL1 1 + rC1 C1 s
function GcL does not contain any RHP zeros and is Toic = 0 c
Gio =
Ueq C1 s
positive. Consequently, conventional negative feedback is
IL1 C1 L1 s + C1 Req s + 1
2
used. 1 + rC1 C1 s
In case of current mode converter, current loop reference Gcic = − C1 s
c
Gco = D′ −
Ueq C1 s
îREF serves as control input; hence the plant (see Fig. 6)
may be represented as (12)
⎡ ⎤ ⎡ ⎤
î îin
ûin Zinc Toic Gcic ⎣ in ⎦ 2.3 Input-voltage feedback
= ûo = H c ⎣ ûo ⎦ (10)
îo c
Gio −Yoc c
Gco
îREF îREF General control block diagram facilitating derivation of
closed-loop transfer functions under input voltage closed
where îREF is the control input (generated by the outer loop loop control is presented in Fig. 7 with x = ‘v’, ĉ = d̂ in
controller) and superscript ‘c’ denotes current mode case of voltage mode converter and x = ‘c’, ĉ = îREF in case
converter. Individual open-loop transfer functions in (10) are of a current mode converter. The applicable open-loop
transfer function set is defined either by voltage mode
GiL Gv + Lc ZC1 converter Hv (4) or current mode converter Hc (10). The
Zinc = (1 − ) · ZC1 = io input-voltage feedback loop gain is then Lxin = Gcv x
Gcix ,
1 + Lc 1 + Lc x
where Gcv is the controller transfer function (accompanied
v
Lc GiL Gco by negative sign to ensure stability) and Gcix the
c
Gio = Gio
v

1 + Lc GcL mode-dependent control-to-input-voltage transfer function.
(11) According to Fig. 7, the set of closed-loop transfer
G Z Toiv v
Lc GoL Gco functions is obtained as
Toic = oL C1 = Yoc = Yov −
1 + Lc 1 + Lc 1 + Lc GcL
v Zinx x
Gio − Lxin Gio−1
x
Lc Lc Gco x
Zin−c = x
Gio−c =
Gcic =− ·Z c
Gco = 1 − Lxin 1 − Lxin
1 + Lc C1 1 + Lc GcL
Toix Yox − Lxin Yo−1
x

where LC = GccGcL denotes inductor-current-loop gain and


x
Toi−c = x
Yo−c = (13)
1 − Lxin 1 − Lxin
ZC1 is input capacitor (C1) impedance. The negative sign of
control-to-input-voltage transfer function (Gcic in (10)) −Lxin x
Gco Lxin
x
Gci−c = x
Gco−c =−
indicates that low-frequency phase of the transfer function 1 − Lxin x
Gse−in Gcix 1 − Lxin
starts at 180°, necessitating either interchanging the signs of
feedback and reference signals of the outer loop or
choosing negative controller gain, similarly to the outcome
of Section 1. In addition, same RHP zeros as in voltage
mode converter exist in the output control dynamics of
current mode converter.
In case inner current loop is properly closed and its
bandwidth is much higher than outer loop bandwidth,
open-loop converter representation can be derived from (2)
by applying Laplace transform, setting îL1 = îREF , Fig. 7 Control-block diagram for deriving the set of closed-loop
eliminating d̂ and manipulating. Individual open-loop transfer functions under input-voltage control

IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314 309
doi: 10.1049/iet-rpg.2014.0149 & The Institution of Engineering and Technology 2015
www.ietdl.org
where so-called ‘ideal’ input-to-output transfer function
x
Gio−1 and ‘ideal’ output admittance Yo−1
x
are defined as

Zinx Gco
x
Toix Gco
x
x
Gio−1 = Gio
x
− x
x
Yo−1 = Yox + x (14)
Gci Gci

3 PVG Effect
Fig. 8 presents linear network model of a general PVG Fig. 8 A two-port model of a PVG interfacing boost power stage
interfacing boost converter, terminated by a voltage source
(typically representing regulated DC bus of a
grid-interfacing DC/AC converter). Small-signal behaviour
of PVG is represented by its Norton equivalent, derived
from physical principles, as shown in Fig. 9. It consists of a
photocurrent source iPV, recombination current iD,
equivalent shunt capacitance CPV and equivalent shunt and
series resistances RSH and RS, respectively [24].
Based on the equivalent circuit, the output current is given by
Fig. 9 PVG equivalent circuit
ipv = iPV − iD − vPV R−1
SH − CPV v̇PV (15)

where recombination losses are usually modelled by one or where so-called ‘ideal’ input impedance Zin−1x
and input
x
more semi-conductor diodes connected in parallel, drawing impedance at open-circuit output port Zin−oco are defined as
current attaining general form of
x x x x
  Gio Gci Gio Toi
x
= Zinx − x
= Zinx +
I0k e(vPV /ak VT ) − 1
Zin−1 Zin−oco (22)
iD = (16) Gcox Yox
k
The low-frequency impedance of the PV generator (cf. Fig. 2)
with I0 k, αk and VT denoting reverse saturation current,
ideality factor of kth diode and thermal voltage, rpv = RS + RSH  RD (23)
respectively. The parameters of Norton equivalent circuit
are thus given by
was shown to cause significant changes in the dynamic
 −1 behaviour of the converter in [25]. Since PVG capacitance
Ypv = RS + RSH  RD  (CPV s)−1 (17) is rather small and is usually offset by the converter input
capacitance, the most significant influence PVG imposes on
and the converter dynamic behaviour can be approximated by
−1
substituting Ypv in (21) with rpv .
iph = iPV RSH  RD  (CPV s)−1 Ypv (18)
3.1 Voltage mode open-loop system
respectively, with RD representing dynamic resistance of (16).
Consequently, linearised characteristics of a PVG are given by According to the described procedure, PVG-affected
control-to-input-voltage transfer function of a voltage mode
îpv = îph − Ypv ûpv (19) converter can be approximated (neglecting parasitics for
simplicity) as
Therefore when connected to a converter represented by a
general matrix form Gciv
GcivS =
⎡ ⎤ 1 + Ypv Zinv
îin 1 Uo
ûin Zinx Toix Gcix ⎣ û ⎦ ≃−  
= (20) L1 C1 s2 + s 1/r C + 1/L C
(24)
îo x
Goi −Yox x
Gco o
ĉ pv 1 1 1

with x = ‘v’, ĉ = d̂ in case of voltage mode converter and indicating that PVG only affects the damping of
x = ‘c’, ĉ = îREF in case of a current mode converter such corresponding resonant circuit as follows. In the CC region
that îin = îpv and ûpv = ûin (cf. Fig. 8), the PVG-converter rpv is high, implying low-damped resonant behaviour. On
coupled dynamics is obtained as [27, 29, 30] (see (21)) the other hand, rpv is low in the CV region, implying
highly damped resonant behaviour. This behaviour is quite

⎡ ⎤
⎡ ⎤ Zinx Toix Gcix ⎡ ⎤
  î îph
GcixS ⎢ ⎥ ⎢ ⎢ 1 + Ypv Zin 1 + Ypv Zinx 1 + Ypv Zin ⎥
ph x x
ûpv ZinxS ToixS ⎥⎢ ⎥
= ⎣ û ⎦ = ⎢ ⎥⎣ ûo ⎦ (21)
îo xS
Goi −YoxS xS
Gco
o
⎣ x
Gio 1 + Ypv Zin−oco
x
1 + Ypv Zin−1
x
x ⎦
ĉ − Yox G ĉ
1 + Ypv Zinx 1 + Ypv Zinx 1 + Ypv Zinx co

310 IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314
& The Institution of Engineering and Technology 2015 doi: 10.1049/iet-rpg.2014.0149
www.ietdl.org
predictable since rpv is connected in parallel with the 3.2 Current mode open-loop system
input-terminal capacitor and the smallest impedance
dominates in the parallel connection. Similarly, In case of current mode converter, corresponding
PVG-affected control-to-output-current transfer function of PVG-affected control-to-input-voltage transfer function is
a voltage mode controlled converter can be approximated as (neglecting parasitics for simplicity)
(neglecting parasitics for simplicity)
Gcic 1
GcicS = c ≃ −rpv (29)
1 + Ypv Zin C1 rpv s + 1
1 + Ypv Zin−1
v
vS
Gco = v
Gco
1 + Ypv Zinv specifying that PVG affects both pole location and gain of the
−Ipv (s + vLow−freq )(s + vMedium−freq ) transfer function, justifying the use of the adaptive controller
≃    (25) introduced in [22] because of their dependence on both
s2 + s 1/rpv C1 + 1/L1 C1 operating point and environmental conditions. As
mentioned above, in case of voltage mode converter
damping only is operating point and environmental
with conditions dependent.
Regarding output-control dynamics, conclusions drawn for
1 the voltage mode converter are valid in case of current-mode
vLow−freq ≃ (26) controlled converter as well.
L1 /rpv − C1 Upv /Ipv
3.3 Closed-loop system
and
When input-voltage control is applied, closed-loop output
admittance is given by
1 Upv
vMedium−freq ≃ − (27)
rpv C1 Ipv L1 Yox Lxin
x
Yo−c = x − Yx (30)
1 − Lin 1 − Lxin o−1
In addition to resonant damping effects discussed above, the
control-to-output-current transfer function experiences At low frequencies (where input-voltage-feedback-loop gain
changes affecting its zeros since Upv/Ipv in (26) and (27) is usually high), ‘ideal’ output admittance Yo−1
x
dominates
denotes PVG static resistance. Its behaviour can be deduced in the behaviour of the closed-loop output admittance. For
from Fig. 2 with respect to dynamic resistance rpv when the case of e.g. voltage mode converter, substituting (5) into
operating point moves along the I–V curve. Based on the (14), ‘ideal’ output admittance is obtained as
resistances behaviour, the following can be concluded:
D′ Ipv sC2
x
Yo−1 = + (31)
(1) ωMedium-freq always resides in RHP, similarly to Ueq 1 + srC2 C2
conventional voltage-fed boost converter;
(2) ωLow-freq is located in RHP in CC region and moves to the and its low-frequency value is hence D′ Ipv /Ueq ≃ Io /Uo (i.e.
left-half-plane when the operating point enters PVG CV positive incremental resistance) [9, 33]. Positive sign of
region. Consequently, output-side feedback control cannot output terminal incremental resistance is a consequence
be used without compromising the stability, as explained in of output current direction, flowing out of the terminal
[27]. Therefore changes in output-control dynamics are (see e.g. [34, 35] for detailed discussion). Such a property is
irrelevant from the practical point of view unless typical for constant-power-output or power-limited devices
output-voltage limiting control is applied, as discussed in such as input-side-controlled PVG interfacing converters.
Section 2. Similar effect is observed in case of
current-mode controlled converter.
4 Experimental validation
In double-stage PV systems, where PVG interfacing
To validate the derived theory, voltage mode controlled boost
DC–DC converter is followed by grid-interfacing DC–AC
converter was built as shown in Fig. 10. Control algorithm
converter, output impedance of PVG interfacing converter
affects the dynamic behaviour of the grid interfacing
converter, as discussed in [9, 33]. The most significant
changes to the DC–AC converter are caused by
low-frequency behaviour of output admittance YoxS , given
by (neglecting the parasitics for simplicity)

D′2 1 + srpv C1
YovS ≃   + sC2
L1 C1 rpv s2 + s 1/r C + 1/L C
(28)
pv 1 1 1

in case of voltage mode operation. Note that the


low-frequency admittance is given by D′2 /rpv , denoting
reflected low-frequency operating point and environmental
conditions dependent PVG impedance. Fig. 10 Experimental setup

IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314 311
doi: 10.1049/iet-rpg.2014.0149 & The Institution of Engineering and Technology 2015
www.ietdl.org

Fig. 13 Measured time-domain response of PVG voltage and


current to a step change in the input-voltage reference from 23 V
Fig. 11 Measured output impedance of the PVG in the CC region (CV operating point) to 15 V (CC operating point) and back
(CC, solid line),
MPP (dash-dotted line) and CV region (dashed line)
Measured input-voltage loop gains of the converter at
above defined three operating points are presented in
(integrative controller) was executed by Texas Instruments Fig. 12. The results indicate that crossover frequencies are
28335-series digital signal processor. Such a controller is quite close to each other, residing in 50–70 Hz range.
usually employed in low cost commercial designs, where Corresponding phase margins are approximately 90°. The
high output loop bandwidth is not required. The controller effect of the PVG on the input-voltage loop gain is quite
is sufficient in PV applications due low dynamics of the insignificant within the chosen control bandwidth.
environmental phenomena. Agilent Technologies E4360A Time-domain responses of PVG voltage and current to a
solar array emulator, proven to imitate the behaviour of a step change in the input-voltage reference of 23 V (CV) to
real PVG quite well [25] was adopted. The power stage was 15 V (CC) and back are shown in Fig. 13, complying with
terminated by a Chroma 63103 electronic load operating in the information provided by loop gains in Fig. 12.
CV mode. Frequency responses given next were measured Measured and predicted PVG-affected control-to-input-
by Model 3120 Venable Instruments frequency-response voltage transfer functions at CC and CV operating points
analyser with an impedance measurement kit. Solar array are shown in Fig. 14, demonstrating good match. The
emulator was configured to yield the following operating source effect on resonant damping and phase behaviour are
regime: short-circuit current of 1.65 A, open-circuit voltage as discussed in Sections 2 and 3.
of 25 V and MPP of 20 V/1.5 A. CC and CV-region operating Measured and predicted PVG-affected control-to-output-
points were selected as 15.0 V/1.6 A and 23.0 V/0.7 A, current transfer functions are shown in Fig. 15. Two RHP
respectively. zeros are clearly observed in the CC region response,
Measured frequency responses of the PVG at CC, MPP and leading to the phase change of more than 360°. The results
CV operating points are presented in Fig. 11. Corresponding also indicate that the phase changes by 180° and one of the
values of rpv are 66.0 Ω (CC), 10.6 Ω (MPP) and 1.4 Ω (CV). RHP zeros disappears when the converter enters the CV
region, well corresponding to the model prediction.
Measured PVG-affected open-loop output impedances are
shown in Fig. 16, where the low-frequency impedance

Fig. 12 Measured input-voltage loop gains in the CC region (solid


line) Fig. 14 Measured (CC: solid lines; CV: dashed lines) and
at the MPP (dash-dotted line) and in the CV region (dashed line) predicted (dotted lines) control-to-input-voltage transfer functions

312 IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314
& The Institution of Engineering and Technology 2015 doi: 10.1049/iet-rpg.2014.0149
www.ietdl.org
corresponds to rpv/D′2 (i.e. CC: 66 Ω 0.49; 49 dBΩ, MPP:
10.6 Ω 0.63; 29 dBΩ, CV: 1.4 Ω 0.74; 8 dBΩ).
Corresponding PVG-affected closed-loop output
impedances are shown in Fig. 17. Resistive behaviour of
the closed-loop output impedances at CC and MPP
operating points supports theoretical findings of Section 3.
Slight reduction of loop gain and crossover frequency in the
CV region as well as the resistive open-loop behaviour
affects the closed-loop output impedance such that resistive
behaviour is visible at low frequencies only.

5 Conclusions
The paper investigated dynamic properties of a
PVG-interfacing boost-power-stage by developing analytic
small-signal model and verifying the derived dynamics by
Fig. 15 Measured (CC: solid lines; CV: dashed lines) and experimental frequency response measurements.
predicted (dotted lines) control-to-output-current transfer functions Investigations clearly show that steady-state and dynamic
behaviour of a conventional voltage-fed boost converter
significantly differs from the steady-state and dynamic
behaviour of the PVG-interfacing boost-power-stage even
though the power hardware is the same. It was shown that
the steady-state behaviour of PVG-converter coupled system
resembles buck converter behaviour, and its dynamics
possesses properties resembling both buck converter
behaviour (i.e. the resonant frequency independency on the
duty cycle) and boost converter behaviour (i.e.
non-minimum phase control dynamics). The converter was
stated to resemble the pulse width modulated switching
shunt unit used in space applications with additional input
LC filter. RHP zeros in control-to-output dynamics were
shown to be similar to the zero in voltage-fed boost
converter and the zero in current-fed buck converter.
In addition to the above, useful control-block diagrams
were also given for facilitating the analysis and design
of cascaded control system, where the inductor-current-
feedback loop forms the inner loop, and the PVG-
voltage-feedback loop forms the outer loop. Pure integrative
controller was shown to be sufficient to serve as outer loop
compensator in voltage mode converter application because
Fig. 16 Measured PVG-affected open-loop output impedances of low bandwidth of environmental variables (irradiation
at CC, MPP and CV operating points and temperature). For the current mode power stage it was
shown that input-voltage loop gain is significantly
dependent on the PVG dynamic resistance, calling for
adaptive controller design.

6 References
1 Blaabjerg, F., Chen, Z., Kjaer, S.B.: ‘Power electronics as efficient
interface in dispersed power generation’, IEEE Trans. Power
Electron., 2004, 19, (5), pp. 1184–1194
2 Kroposki, B., Pink, C., DeBlasio, R., Thomas, H., Simôes, M., Sen, P.K.:
‘Benefits of power electronic interfaces for distributed energy systems’,
IEEE Trans. Energy Convers., 2010, 25, (3), pp. 901–908
3 Meneses, D., Blaabjerg, F., García, Ó., Cobos, J.A.: ‘Review and
comparison of step-up transformerless topologies for photovoltaic
AC-module applications’, IEEE Trans. Power Electron., 2013, 28, (6),
pp. 2649–2663
4 Ho, C., Breuninger, H., Pettersson, S., Escobar, G., Serpa, L., Coccia,
A.: ‘Practical design and implementation procedure of an interleaved
boost converter using SiC diodes for PV applications’, IEEE Trans.
Power Electron., 2012, 27, (6), pp. 2835–2845
5 Ho, C., Breuninger, H., Pettersson, S., Escobar, G., Canales, F.: ‘A
comparative performance study of an interleaved boost converter
using commercial Si and SiC diodes for PV applications’, IEEE
Trans. Power Electron., 2013, 28, (1), pp. 289–299
Fig. 17 Measured PVG-affected closed-loop output impedances 6 Konstantopoulos, G.C., Alexandridis, A.T.: ‘Non-linear voltage
at CC, MPP and CV operating points regulator design for DC/DC boost converters used in photovoltaic

IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314 313
doi: 10.1049/iet-rpg.2014.0149 & The Institution of Engineering and Technology 2015
www.ietdl.org
applications: analysis and experimental results’, IET Renew. Power. 21 Messo, T., Jokipii, J., Suntio, T.: ‘Steady-state and dynamic properties of
Gener., 2013, 7, (3), pp. 296–308 boost-power-stage converter in photovoltaic applications’. Proc. IEEE
7 Zainuri, M., Radzi, M., Soh, A., Rahim, N.: ‘Development of adaptive PEDG, 2012, pp. 34–40
perturb and observe-fuzzy control maximum power point tracking for 22 Urtasun, A., Sanchis, P., Marroyo, L.: ‘Adaptive voltage control of the
photovoltaic boost dc-dc converter’, IET Renew. Power Gener., 2014, DC/DC boost stage in PV converters with small input capacitor’,
8, (2), pp. 183–194 IEEE Trans. Power Electron., 2013, 28, (11), pp. 5038–5048
8 Urayai, C., Amaratunga, G.A.J.: ‘Electronic load interface for improving 23 Suntio, T.: ‘Dynamic profile of switched-mode converter – modeling,
PV dual-converter system operational margin’, IET Renew. Power analysis and control’ (Wiley-VCH, Weinheim, Germany, 2009)
Gener., 2014, 7, (1), pp. 10–17 24 Liu, S., Dougal, R.A.: ‘Dynamic multiphysics model for solar array’,
9 Messo, T., Jokipii, J., Puukko, J., Suntio, T.: ‘Determining the value of IEEE Trans. Energy Convers., 2002, 17, (2), pp. 285–294
DC-link capacitance to ensure stable operation of a three-phase 25 Nousiainen, L., et al.: ‘Photovoltaic generator as an input source for
photovoltaic inverter’, IEEE Trans. Power Electron., 2014, 29, (2), power electronic converters’, IEEE Trans. Power Electron., 2013, 28,
pp. 665–675 (6), pp. 3028–3038
10 Puukko, J., Suntio, T.: ‘Dynamic properties of a voltage source 26 Chen, Y.-M., Huang, A.Q., Yu, X.: ‘A high step-up three-port DC-DC
inverter-based three-phase inverter in photovoltaic application’, IET converter for stand-alone PV/battery power systems’, IEEE Trans.
Renew. Power Gener., 2012, 6, (6), pp. 381–391 Power Electron., 2013, 28, (11), pp. 5049–5062
11 Lee, D.H., Cho, B.H., Patil, A.R., Lee, F.C.: ‘Design of a single PWM 27 Suntio, T., Leppäaho, J., Huusari, J., Nousiainen, L.: ‘Issues on
section shunt switching unit for high power space systems’. Proc. IEEE solar-generator-interfacing with current-fed MPP-tracking converters’,
PESC, 1995, pp. 145–149 IEEE Trans. Power Electron., 2010, 25, (9), pp. 2409–2419
12 Carrigós, A., Carrasco, J.A., Blanes, J.M., Sanchis, E.: ‘Modeling the
28 Suntio, T., Huusari, J., Leppäaho, J.: ‘Issues on solar-generator-
sequential switching shunt series regulator’, IEEE Power Electron.
interfacing with voltage-fed MPP-tracking converters’, EPE J., 2010,
Lett., 2005, 3, (1), pp. 7–13
20, (3), pp. 40–47
13 Anil Kumar, R., Suresh, M.S., Nagaraju, J.: ‘Effect of solar array
29 Leppäaho, J., Suntio, T.: ‘Dynamic characteristics of current-fed
capacitance on the performance of switching shunt voltage regulator’,
superbuck converter’, IEEE Trans. Power Electron., 2011, 26, (1),
IEEE Trans. Power Electron., 2006, 21, (2), pp. 543–548
14 Glass, M.C.: ‘Advancements in the design of solar array to battery pp. 200–209
charge current regulators’. Proc. IEEE PESC, 1977, pp. 346–350 30 Leppäaho, J., Huusari, J., Nousiainen, L., Puukko, J., Suntio, T.:
15 Capel, A., Marpinard, J.C., Jalade, J., Valentin, M.: ‘Current fed and ‘Dynamic properties and stability assessment of current-fed converters
voltage fed switching DC/DC converters - steady state and dynamic in photovoltaic applications’, IEEJ Trans. Ind. Appl., 2011, 131, (8),
models, their applications in space technology’. Proc. IEEE pp. 976–984
INTELEC, 1983, pp. 421–430 31 Middlebrook, R.D., Ćuk, S.: ‘A general unified approach to modelling
16 Siri, K.: ‘Study of system instability in solar-array-based power systems’, switching power stages’, Int. J. Electron., 1977, 42, (6), pp. 521–550
IEEE Trans. Aerosp. Electron. Syst., 2000, 36, (3), pp. 957–964 32 Shmilovitz, D., Singer, S.: ‘A switched-mode converter suitable for
17 Leppäaho, J., Suntio, T.: ‘Dynamic properties of PCM-controlled superconductive magnetic energy storage (SMES) systems’. Proc.
current-fed boost converter in photovoltaic system interfacing’. Proc. IEEE APEC, 2002, pp. 630–634
EPE-ECCE, 2011, pp. 1–10 33 Messo, T., Puukko, J.P., Suntio, T.: ‘Effect of MPP-tracking DC/DC
18 Xiao, W., Dunford, W.G., Palmer, P.R., Capel, A.: ‘Regulation of converter on VSI-based photovoltaic inverter dynamics’. Proc. IET
photovoltaic voltage’, IEEE Trans. Ind. Electron., 2007, 54, (3), PEMD, 2012, pp. 1–6
pp. 1365–1374 34 Suntio, T., Viinamäki, J., Jokipii, J., Messo, T., Kuperman, A.:
19 Villalva, M.G., de Siqueira, T.G., Ruppert, E.: ‘Voltage regulation of ‘Dynamic characterization of power electronic interfaces’, IEEE
photovoltaic arrays: small-signal analysis and control design’, IET J. Emerg. Sel. Topics Power Electron., DOI: i11.1109/
Power Electron., 2010, 3, (6), pp. 869–880 JESTPE.2014.2313704 (accepted)
20 Danyali, S., Hosseini, S.H., Gharehpetian, G.B.: ‘New extendable 35 Huusari, J., Suntio, T.: ‘Dynamic properties of current-fed quadratic
single-stage multi-input DC-DC/AC boost converter’, IEEE Trans. full-bridge buck converter for distributed photovoltaic MPP-tracking
Power Electron., 2014, 29, (2), pp. 775–788 systems’, IEEE Trans. Power Electron., 2012, 27, (11), pp. 4681–4689

314 IET Renew. Power Gener., 2015, Vol. 9, Iss. 4, pp. 306–314
& The Institution of Engineering and Technology 2015 doi: 10.1049/iet-rpg.2014.0149

You might also like