You are on page 1of 11

3028

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 28, NO. 6, JUNE 2013

Photovoltaic Generator as an Input Source for Power Electronic Converters


Lari Nousiainen, Student Member, IEEE, Joonas Puukko, Student Member, IEEE, Anssi M ki, Student Member, IEEE, a Tuomas Messo, Student Member, IEEE, Juha Huusari, Student Member, IEEE, Juha Jokipii, Student Member, IEEE, Jukka Viinam ki, Diego Torres Lobera, Student Member, IEEE, Seppo Valkealahti, Member, IEEE, a and Teuvo Suntio, Senior Member, IEEE

AbstractA photovoltaic (PV) generator is internally a powerlimited nonlinear current source having both constant-currentand constant-voltage-like properties depending on the operating point. This paper investigates the dynamic properties of a PV generator and demonstrates that it has a profound effect on the operation of the interfacing converter. The most important properties an input source should have in order to emulate a real PV generator are dened. These properties are important, since a power electronic substitute is often used in the validation process instead of a real PV generator. This paper also qualies two commercial solar array simulators as an example in terms of the dened properties. Investigations are based on extensive practical measurements of real PV generators and the two commercial solar array simulators interfaced with dcdc as well as three- and single-phase dcac converters. Index TermsConverter, inverter, photovoltaics (PVs), solar array simulator, validation.

I. INTRODUCTION

NSTALLED capacity of photovoltaic generator (PVG)based energy systems is rapidly growing due to advantageous public and political climates [1]. Such systems are interfaced to dc or ac loads with power electronic devices, which have been shown to cause, e.g., harmonic distortion, reduce damping in the utility grid, and suffer from reliability problems [2][5]. These phenomena can even lead to instability or production outages and are expected to increase as the penetration depth of distributed generation grows [6], [7]. Therefore, an extensive validation process, which characterizes dynamic properties of the proposed interfacing converters, is of utmost importance to overcome or minimize the problems. The input source has a signicant effect on converter dynamics, as discussed in detail in [8][10]. A PVG is internally a

power-limited nonlinear current source having both constantcurrent (CC) and constant-voltage (CV) like properties depending on the operating point [11], which implies that the dynamics of a photovoltaic interfacing converter (PVIC) cannot be validated solely by using a voltage or current source as the input source. Therefore, the validation should be performed using a real PVG as the input source. If a real PVG is to be used in the PVIC validation process, an articial light source providing controllable illumination should be used to guarantee the repeatability of the measurements. This can be accomplished cost effectively in small scale, e.g., for a single PV module, but is impractical for larger systems. Therefore, a PVG is usually replaced with a power electronic substitute, i.e., a solar array simulator, so that time-invariant conditions can be guaranteed in the validation. This paper presents the dynamic properties of a real PVG, denes the most signicant parameters that will have an effect on PVIC dynamics, and demonstrates these effects by experimental measurements based on dcdc as well as singleand three-phase dcac converters. This paper also qualies two commercial solar array simulators as an example in terms of the dened properties and analyzes the differences between the solar array simulators and real PVGs both in time and frequency domains. The rest of the paper is organized as follows. The dynamic properties of PVGs are reviewed in Section II. The effects of PVG on the interfacing converter dynamics are presented in Section III. Section IV compares the dynamic properties of the commercial solar array simulators with real PVGs and denes the characteristics an input source should have in order to emulate a real PVG. Conclusions are drawn in Section V. II. DYNAMIC PROPERTIES OF A PV GENERATOR

Manuscript received April 10, 2012; revised June 12, 2012; accepted July 10, 2012. Date of current version December 7, 2012. Recommended for publication by Associate Editor Q.-C. Zhong. L. Nousiainen and J. Puukko were with the Department of Electrical Energy Engineering, Tampere University of Technology, Finland. They are now with ABB Drives, Helsinki, Finland (e-mail: lari.nousiainen@.abb.com; joonas.puukko@.abb.com). A. M ki, T. Messo, J. Jokipii, J. Viinam ki, D. T. Lobera, a a S. Valkealahti, and T. Suntio are with the Department of Electrical Energy Engineering, Tampere University of Technology, Tampere, Finland (e-mail: anssi.maki@tut.; tuomas.messo@tut.; juha.jokipii@tut.; jukka.viinamaki@ tut.; diego.torres@tut.; seppo.valkealahti@tut.; teuvo.suntio@tut.). J. Huusari is with ABB Corporate Research, Baden-D ttwil, Switzerland a (e-mail: juha.huusari@ch.abb.com). Digital Object Identier 10.1109/TPEL.2012.2209899

A simplied electrical equivalent circuit of a PV cell composes of a photocurrent source with parallel-connected diode and parasitic elements, as depicted in Fig. 1 [11], [12]. In Fig. 1, ipv and upv are the output current and voltage of the PV cell, respectively, iph is the photocurrent, which is linearly proportional to the irradiance, icpv is the current through the shunt capacitance cpv , and irsh is the current through the shunt resistance rsh . The shunt and series resistances rsh and rs represent various nonidealities in a real PV cell. The relation between diode current id and voltage ud can be modeled with an exponential equation, yielding a nonlinear resistance rd that can be used instead of the diode symbol in Fig. 1 [11], [13]. The

0885-8993/$31.00 2012 IEEE

NOUSIAINEN et al.: PHOTOVOLTAIC GENERATOR AS AN INPUT SOURCE FOR POWER ELECTRONIC CONVERTERS

3029

ipv

Magnitude (dB)

60 40 20 0 20 1 10 90

SC MPP OC 10
2

id
iph

icpv

irsh rsh

rs
upv

ud

rd

cpv

10

10

10

10

Phase (deg)

Fig. 1.

Simplied electrical equivalent circuit of a photovoltaic cell.

45 0 45 90 1 10 10
2

OC MPP SC 10 10 Frequency (Hz)


3 4 5 6

CC

1.2
Current, power, resistance and capacitance (p.u.)

CV ipv MPP

10

10

1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 Voltage (p.u.)
rpv ppv

Fig. 3.

PVG impedances at short circuit (SC), MPP, and open circuit (OC).

Fig. 3) as cpv 1 2rpv f3dB (2)

cpv

1.0

1.2

Fig. 2.

Static and dynamic terminal behavior of a PVG.

one-diode model can also be used to model the operation of a PV module, i.e., a series connection of PV cells, by scaling model parameters, as presented in [12]. The measured static and dynamic characteristics of a PV module are shown in Fig. 2 as normalized (p.u.) values. The measurement setup has been reported earlier in detail in [14]. The static currentvoltage and powervoltage characteristics show that the PV module is a highly nonlinear current source having limited output voltage and power. In order to maximally utilize the energy of solar radiation by using a PVG, the operating point has to be kept at the maximum power point (MPP) in which dipv d(upv ipv ) dppv = = Ipv + Upv =0 dupv dupv dupv (1)

where f3dB is the cutoff frequency of the impedance magnitude curve. Equation (2) gives a good estimate for cpv in the rs . In CV region, (2) slightly unCC region, since rd rsh derestimates cpv since rd rsh and rs are in the same order of magnitude, but is still sufciently accurate. According to Kirchhoffs laws, the analyzed or measured output impedance of a device is Z, if the direction of positive current ow is dened out of the device as in Figs. 1 and 2. Thus, the dynamic resistance rpv (or the incremental resistance as named in [15]) is in fact positive since upv /ipv in Fig. 2 is negative and has to be multiplied by 1 to obtain the correct impedance. III. DYNAMIC EFFECTS ON THE CONVERTER Dynamic properties of a power electronic converter are determined not only by the power stage but also by the type of source and load subsystems, which in turn dene the possible feedback variables. It has been shown in detail in [8][10] that the input source has a profound effect on the dynamics of the converter connected to it. The same power stage can be supplied either by a source that has current or voltage-source-like properties. The converter dynamics are completely different in these two cases. According to circuit theory and control engineering principles, the input variables are uncontrollable and only the output variables can be controlled. In practice, this means that an input-voltage-controlled converter (as usually adopted in PV applications for enabling maximum power transfer [16]) has to be analyzed as a current-fed system (i.e., input current is an uncontrollable input variable and input voltage is a controllable output variable) as will be done in the following sections. A. DCAC Interfacing Figs. 4 and 5 show conventional single- and three-phase VSItype PV inverters usually applied in interfacing PVGs to the

where Ipv and Upv are the MPP current and voltage, respectively. The dynamic behavior of the PV module is shown in Fig. 2 in terms of its dynamic resistance rpv = rd rsh + rs and capacitance cpv , which are nonlinear and dependent on the operating point. The dynamic resistance represents the low-frequency value of the impedance shown in Fig. 3 and is the most significant variable that will have an effect on the PVIC dynamics, as presented in more detail in Section III. The dynamic capacitance, in turn, can be approximated from PVG impedance (see

3030

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 28, NO. 6, JUNE 2013

iin iC

iL

rL

iP

iin
+
io
Z in Toi-d uod

iod

uL

uod
Gio-d iin
Gcr-qd uoq Gco-d d d Gco-qd d q Yo-d

uin uC

rC uo

uin
iinS

Toi-q uoq

C N

YS

ioq
Gci-d d d

uoq
Gio-q iin Gcr-dq uod Gco-dq d d Gco-q d q Yo-q

_ Gci-q dq

Fig. 4.

VSI-type single-phase PV inverter.

iin iC

iP

dd
La iLa rLa iLb rLb iLc rLc ioa iob ioc ua ub

dq

Fig. 7. Linear small-signal model of a three-phase inverter in synchronous reference frame [10].

uin uC

Lc

n
uc

Magnitude (dBA)

rC

Lb

C C
uL

40 20 0 20 40 10
0

CV CC MPP

N
Fig. 5. VSI-type three-phase PV inverter.

10

10

10

10

Z in Toi uo
Gci d Gio iin Gco d

Phase (deg)

iin

io

uin
iinS

uo Yo
_

YS

uo

270 180 90 0 90 180 270 360 0 10

CC MPP CV

10

10 10 Frequency (Hz)

10

d
Fig. 6. Linear small-signal model of a single-phase inverter [17].

Fig. 8. Three-phase inverter control-to-output-current transfer function G c o -d . Measurements with solid lines, prediction using measured source impedance with dotted lines, and prediction using (5) with dashed line.

utility grid. Dynamic properties of these inverters can be analyzed by constructing small-signal models that describe the dynamics between the uncontrollable input variables and the controllable output variables, as presented in Figs. 6 and 7, as linear small-signal models. Detailed modeling procedures for the single- and three-phase inverters can be found, e.g., from [10] and [17]. The operating-point-dependent dynamic effect of a PVG can be taken into account by considering the source as a parallel connection of a current source iinS and source admittance YS . According to Fig. 3, PVG impedance behaves as an RC circuit up to typical converter switching frequencies (c.a. 1 . . . 100 kHz). The source impedance ZS can be given according to Fig. 1 by
rp v

which can be approximated by considering rs = 0 and rpv = rd rsh + rs as ZS rd rsh 1 1 rpv scpv scpv (4)

and further at low frequencies by ZS rpv . (5)

ZS = rs + rd rsh

rs + rd rsh +scpv rs (rd rsh ) 1 = scpv 1 + scpv rs (rd rsh ) (3)

The use of (5) instead of (4) is justied if the interfacing converter input capacitance Cin cpv , which typically applies. Fig. 8 shows measured three-phase inverter control-to-outputcurrent transfer functions Gco-d compared with predictions obtained using either the measured source impedances or their low-frequency values, i.e., (5), to model the effect of the source. The results show high correlation, which conrms the analysis. According to the notations of Figs. 4 and 5 as well as using YS = 1/ZS = 1/rpv , the control-to-output transfer functions Gco and Gco-d without the parasitic elements for

NOUSIAINEN et al.: PHOTOVOLTAIC GENERATOR AS AN INPUT SOURCE FOR POWER ELECTRONIC CONVERTERS

3031

single- and three-phase inverters can be given by Gco =


Uin L

iin

iL

rL iC2

io

s s2 +

1 C

Ii n Uin

1 rp v

iC1

+ uL

1 rp v C

s+

D2 LC 1 C Ii n Uin

(6)
uin

Gco-d =

Uin L

s s s+

1 rp v 2 rp v C

s s2 +

1 rp v C

2 2 3 D d +D q 2 LC

. (7)
Fig. 9.

+ 2 +

+ uC1

rC1 C1

+ uC2

rC2 C2

uo

Boost-type dcdc converter with an input capacitor.

Analysis of (6) and (7) reveals that a right-half-plane (RHP) zero appears and the sign of the transfer function changes when the operating point moves from the CV to the CC region of a PVG. This is due to the fact that at MPP, the dynamic resistance of the PVG rpv coincides with the equivalent static loading resistance (i.e., the static input impedance of the PVIC) according to the maximum power transfer theorem [18] and (1). The dynamic resistance rpv is greater in magnitude than the static resistance in the CC region (CCR) and smaller in the CV region (CVR) of a PVG, which will change the location of the Gco and Gco-d zero in the complex plane as presented in (8) and veried by the experimental measurements from three-phase inverter presented in Fig. 8 CVR: MPP: CCR: rpv < Uin /Iin z < 0 rpv = Uin /Iin z = 0 rpv > Uin /Iin z > 0 LHP Origin RHP. (8)

ipv + DC upv _ upv

DC
u
ref pv

iinv + uinv _

iac

DC
uac

AC
uinv iac

AC grid

MPPT
Fig. 10.

Typical two-stage grid interface for a PVG.

Appearance of the RHP zero and the change of sign of the control-to-output-current transfer function means that the output current control cannot be stable both in the CC and CV regions of a PVG. Therefore, a cascaded control scheme (input-voltage output-current) is needed to enable the operation at all PVG operating points and to transfer maximum power in a reliable manner, as shown in detail in [8][10], and [17]. The RHP zero in the output-current-control loop will actually turn into an RHP pole in the input-voltage-control loop when the operating point is in the CC region of a PVG. This will naturally cause design constraints in the input-voltage control, as discussed in [19] and [20]. B. DCDC Interfacing Fig. 9 shows a dcdc converter based on a conventional boost topology with an additional input capacitor. The dcdc converter can be used as an upstream converter between a PVG and an inverter resulting in a two-stage conversion scheme, as presented in Fig. 10. The dcdc converter is responsible for the maximumpower-point-tracking (MPPT) function by controlling its own input voltage. The VSI-type inverter has a similar cascaded control structure as in the single-stage conversion scheme, thus enabling maximum power transfer. Accordingly, both the dc dc and the dcac converters in PV applications control its own input voltage. Therefore, they must be analyzed as current-fed current-output converters as previously discussed.

The additional dcdc stage enables the use of less seriesconnected PV modules compared to the single-stage inverters (see Figs. 4 and 5), which may be benecial in case of partial shading conditions [21]. The dcdc stage can also regulate its input voltage to practically pure dc (i.e., provides perfect power decoupling), which can increase the energy yield compared to single-stage single-phase inverters where the dc power uctuates at twice the grid frequency [22]. The dynamics of the dcdc converter, when loaded by a voltage-type load (i.e., an inverter controlling its input voltage, uo in Fig. 9), can be obtained by linearizing the switchingfrequency-averaged model presented in [23] and solving the transfer functions between the input and output variables in the frequency domain. The control-to-input-voltage transfer function for the converter in Fig. 9 without the parasitic elements can be given by Gci-dc = Uo LC1 s2 + 1
1 rp v C 1 1 LC1

s+

(9)

According to (9) and Fig. 11, there exist no operating-pointdependent phase shift or zeros that would move between the left and right halves of the complex plane causing control system design constraints as discussed earlier. The only difference between different PVG operating points is the change in the damping of the resonance of Gci-dc , as can be seen from Fig. 11. The input-voltage control should be designed so that the bandwidth exceeds the MPPT-algorithm execution frequency. In addition, a bandwidth over twice the grid fundamental frequency would be benecial in single-phase applications in order to prevent the low-frequency dc-link-voltage (i.e., inverter input voltage) ripple from reecting to the input terminals of the dcdc converter and, therefore, to the PVG.

3032

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 28, NO. 6, JUNE 2013

Magnitude (dBV)

Magnitude (dBA)

40 20 0
CC CV pred.

60 40 20 0 20 40 0 10 180 90 0 90 180 270 10


0

20 0 10 180 90 0 90 10
0

10

10

10

10

10

10

10

10

Phase (deg)

Phase (deg)

CC CV pred.

10

10 10 Frequency (Hz)

10

10

10 10 Frequency (Hz)

10

Fig. 11.

Control-to-input-voltage transfer function.

Fig. 12.

Control-to-output-voltage transfer function.

Some researchers claim that the dcdc converter in the twostage conversion scheme could be used to control the dc-link voltage [24][26], which means that the inverter controls only its output current, because two converters cannot control the same voltage. This control scheme does not work when the intention is to deliver maximum power from the input source without compromising the stability of the system, which will be proven next. The transfer functions presented in [23] can be used to compute the control-to-output-voltage transfer function using the method presented in [27] as GZ co-dc = o /d i uo GH = = co-dc H o /o Yo-dc i u d (10)

where the superscript H denotes an H-parameter model (current-to-current converter) and Z denotes a Z-parameter model (current-to-voltage converter) [28]. According to (10), the control-to-output-current and voltage transfer functions share the same zeros and the following analysis is valid for both of the transfer functions. The source-affected control-to-outputvoltage transfer function Gco-dc can be given according to [23] by GZ co-dc = I i n s2 + C2 s3 +
1 C 1 rp v 1 C 1 rp v

Uin Ii n L

s+

1 C1 L

Uin Ii n C 1 L rp v

s2 +

C 1 D 2 +C 2 C1 C2 L

s+

D C 1 C 2 L rp v

. (11)

This implies that if the output of the dcdc converter is to be controller, it needs to control also its input. But a cascaded input-voltage output-voltage control scheme for the dcdc converter cannot guarantee proper dc-link voltage for the VSI-type inverter. Also, determining the grid current reference will be problematic. Therefore, it can be concluded that each converter in the power processing chain have to control its input terminals if maximum power is to be supplied into the utility grid. In PV applications, the input-current control is prone to saturation, since the MPP current is close to the short-circuit current that is dependent on the irradiation level with relatively fast dynamics. The open-circuit voltage, on the other hand, is dependent mostly on the temperature with negligibly slow dynamics; thus, input-voltage control can be realized in a reliable manner [16]. Based on the previous analysis regarding dcac and dcdc interfacing, few key points can be outlined: 1) the dcdc converter in the two-stage conversion scheme is responsible for the MPPT, where input-voltage control is preferred over the open-loop-based MPPT [23]; 2) the dcac converter has a cascaded input-voltage output-current control structure both in the single- and two-stage conversion schemes; 3) therefore, the dc dc and dcac converters in PV applications have to be analyzed as current-fed current-output converters, where the input current is the source-side input variable and the input voltage is the source-side output variable; and 4) the practical tests have to be carried out by using input source emulating properly the behavior of real PVG. IV. EXAMPLE SOLAR ARRAY SIMULATORS Properties of two different commercial solar array simulators from two different manufacturers were evaluated. Two articial light units were used to illuminate the reference PV modules. The rst light unit is a based on uorescent lamps and is designed to produce radiation intensity of 500 W/m2 for a 30-W PV module. The second light unit based on halogen lamps is designed to produced the same intensity for a 190-W PV module. Both of the modules operate at half the nominal power. The

Analyzing (11) and Fig. 12 reveals that the transfer function has a low-frequency zero located on the RHP when the PVG is operating at voltages lower than the MPP. Accordingly, the output control (whether output voltage or current) of the interfacing dcdc boost converter cannot be designed to be stable at all PVG operating points, because the Gco-dc changes its sign between different operating regions and a low-frequency RHP zero appears.

NOUSIAINEN et al.: PHOTOVOLTAIC GENERATOR AS AN INPUT SOURCE FOR POWER ELECTRONIC CONVERTERS

3033

70 60

0.8

Resistance (dB)

50
Current (A)
PVG SAS Table

40 30 20 10 0

0.6

0.4

0.2
SAS Table

10 Voltage (V)

15

20

10 Voltage (V)

15

20

Fig. 13.

Measured dynamic resistances.

Fig. 15.

Solar array simulator IV curve comparison.

30 20 Capacitance (dBF) 10 0 10 20 30
PVG SAS Table

10 Voltage (V)

15

20

Fig. 14.

Measured dynamic capacitances.

rst solar array simulator, device A, was evaluated by using the 30-W and 190-W modules as references. The second solar array simulator, device B, was evaluated by using the 190-W module as a reference. A. 30-W Module Emulation The evaluated commercial power electronic substitute (device A) has three different modes of operation: SAS, table and xed modes. In the SAS mode, the simulator is programmed using three reference points: short-circuit current, open-circuit voltage, as well as current and voltage at the MPP. In the table mode, the IV curve is represented by voltagecurrent pairs with a limitation that the voltage points must be ascending and the current points descending. In the xed mode, a maximum voltage is given and the simulator operates as a voltage limited current source having rectangular IV curve characteristics. Measured dynamic resistances and capacitances from the commercial PVG and device A are shown in Figs. 13 and 14. Later on in the gures, the device A at the two different oper-

ating modes will be referred to as Table and SAS. Based on rpv of the real PVG, it can be seen that the curve has two distinct slopes. Because rpv is presented in dB, it is clear that rpv would be best approximated with a two-diode (i.e., double exponential) model as opposed to the one-diode model of Fig. 1. The same double exponential characteristics are also visible in Fig. 14, which presents the measured dynamic capacitances. Based on Fig. 13, the device A produces similar double exponential characteristics in respect to rpv when it is used in the table mode. However, the dynamic resistance of the device A used in the SAS mode is constant in the CV region, as can be seen from Fig. 13. This indicates that the dynamic resistance of a real PVG can be emulated with higher precision by using the table mode. The operating mode of device A does not affect the emulated dynamic capacitance, as can be seen from Fig. 14. The capacitance of the device A is considerably higher (up to 30 dBF higher) than the capacitance of the real PVG and it does not show double exponential characteristics. Fig. 15 presents the IV curves of solar array simulator in SAS and table modes. It was stated earlier that the dynamic resistance of device A in the SAS mode is constant in the CV region. Constant dynamic resistance implies that the IV curve is a straight line in the CV region, as can be noticed from Fig. 15. In the table mode, the device A produces IV curve correctly, and thus also the dynamic resistance as was analyzed in Section II. It is worth noting that a solar array simulator can produce the dynamic resistance of a PVG correctly only if the simulator produces the IV curve correctly when loaded with the specic power electronic device under test. Figs. 16 and 17 present step-like load change from CV to CC region so that the power at initial and nal operating points is the same. The short-circuit current for the PVG and the device A model was 1.0 A. The generator current exceeds the shortcircuit current value (which should be impossible based on the static IV curve) because of the stored energy in the dynamic capacitance. The overshoot is considerably smaller with the real PVG (peak current 1.222 A) than with the device A (peak current

3034

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 28, NO. 6, JUNE 2013

5 V div

0V

60 40 20 0 20 40 10
0

Magnitude (dBA)

CV CC

0.5 A div

10

10

10

10

0A

100 s

Fig. 16.

Load step change in case of the PVG.

180 90 0 90 180 270 10


0

Phase (deg)

CC CV

PVG Table

10

10 10 Frequency (Hz)

10

5 V div

Fig. 19.

Control-to-output current transfer function comparison.

0V

iin
i 0.5 A div

L1 iCd iC1 Rd rC1 uC1 Cd C1


uL1

iL1 rL1 iC2 rC2 uC2

L2
uL2

iL2 rL2 io uo

uin
0A
100 s

uCd

C2

Fig. 17.

Load step change in case of the device A in the table mode. Fig. 20. Boostbuck-type dcdc converter.

Magnitude (dB)

60 40 20 0 10 90
1

Table SAS

CC

PVG

CV 10
2

10

10

10

10

45 0 45 90 1 10 10
2

10 10 Frequency (Hz)

10

10

Fig. 18.

Impedance comparison.

1.526 A) since the dynamic capacitance of PVG is considerably smaller than that of the device A as discussed earlier. This effect should also be taken into account when validating time-domain behavior of interfacing converters. The high-frequency ripple in PVG current (see Fig. 16) is due to the resonant inverters driving the uorescent lamp unit. Figs. 1316 compared the PVG and the electronic substitute in terms of rpv , cpv , and time-domain responses. Fig. 18 presents the measured PVG and device A impedances in CC and CV regions of the IV curve. By considering the impedance in the

CC region, device A shows PVG-like characteristics but with a higher capacitance up to c.a. 2 kHz. The impedance in the CV region correlates up to the same frequency range. After the c.a. 2-kHz frequency, the device A impedance does not show similar resistivecapacitive characteristics as the real PVG does. This difference is visible in Fig. 19, which presents the measured control-to-output-current transfer function for the converter in Fig. 9. The measurements are performed at the same operating points as the impedances in Fig. 18. It can be concluded by considering Figs. 18 and 19 that the low-frequency value of the PVG impedance rpv is the major factor in determining the dynamic properties of the converter connected to a PVG as was discussed in Section III-A. However, if the input capacitance of the converter and the dynamic capacitance of the solar array simulator are in the same order of magnitude or smaller, the considerably higher capacitance of the solar array simulator has to be taken into account. Figs. 3 and 18 show that the phase of the PVG impedance lies between 90 . A solar array simulator should have similar passive-circuit-like characteristics, as in Fig. 18, in order to be justied as a substitute for the real PVG. B. 190-W Module Emulation The 190-W module and the devices A and B were interfaced with a boostbuck-type dcdc converter shown in Fig. 20, operating under input-voltage-feedback control. A low-bandwidth (75 Hz) integral controller was used yielding a phase margin of

Phase (deg)

NOUSIAINEN et al.: PHOTOVOLTAIC GENERATOR AS AN INPUT SOURCE FOR POWER ELECTRONIC CONVERTERS

3035

Magnitude (dB)

60 40 20 0 20 10 540
1

CC
u 5 V div

CV
i 1 A div

u
2 3 4 5 6

5 V div

10

10

10

10

10

Phase (deg)

360 180 0 180 1 10 10

PVG A B

CV for device B
0A 0V
200 ms
i 1 A div

10 10 Frequency (Hz)

10

10

Fig. 23.

Input-voltage-reference ramp with the device A.

Magnitude (dB)

Fig. 21.

Impedance comparison.

40 CC 20 CV 0 Zin > ZoB 20 1 10 450 360 270 180 90 0 90 10


1

u 5 V div u i 1 A div 5 V div

10

10

10

Phase (deg)

ZoB Zin 180 < 180 10


2

CV

0A 0V

200 ms

i 1 A div

CC

Fig. 22.

Input-voltage-reference ramp with the PVG.

88 in the input-voltage-control loop. A parallel damping circuit comprising series connection of a capacitor Cd and a resistor Rd was connected in parallel with the converter input in order to minimize the resonant behavior of the converter. The measured output impedances of the PVG and the devices A and B are shown in Fig. 21. It can be observed that both of the devices reproduce the dynamic resistance accurately (i.e., low-frequency impedance). The capacitance of the device B is higher than the capacitance of device A, which again is higher than the capacitance of the real PVG, as was the case also earlier. It can be also observed that the phase of the PVG and device A impedances lies between 90 within the whole operating region, as in Fig. 18. The phase of the device B impedance is similar to the CC region. However, the behavior of the device B changes dramatically when the operating point moves to the CV region. A resonance with 270 phase shift occurs approximately at 250 Hz, which can lead to impedance-based stability problems when the interfacing converter is connected. It is known that the stability and load interaction issues in a voltage-fed system can be studied by applying Nyquist stability criterion to the impedance ratio of the load and source subsystems known as minor-loop gain (Zo /Zin ) [29]. In current-fed systems, the stability is studied based on inverse minor-loop gain

10 Frequency (Hz)

10

Fig. 24. B.

Input impedance of the converter and output impedance of the device

(Zin /Zo ) [30]. If the impedance ratio of an interconnected stable source and load subsystems does not satisfy the Nyquist stability criterion, the interconnected system is unstable. It can be deduced that the stability criterion will be violated in a PV system if |Zin /Zo | 1, while the phase difference exceeds 180 . The concept of minor-loop gain is typically used to study grid interactions between the inverters and the utility grid [6], but applies also in the PVG/converter interface. A triangular input-voltage-reference ramp sweeping the operating points between the CC and CV regions was applied in the control system of the converter in Fig. 20. In Figs. 22 and 23, the ramp was applied with the PVG and device A. As can be seen, the system is stable within the whole operating range and the reference ramp is reproduced nicely. This is due to the fact that both the source and load impedance (the converter input impedance is shown in Fig. 24) phase lies between 90 . Accordingly, a phase difference of 180 , and thus, violation of the stability criterion does not take place.

3036

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 28, NO. 6, JUNE 2013

u 5 V div u i 1 A div 5 V div

0A 0V

200 ms

i 1 A div

Fig. 25.

Input-voltage-reference ramp with the device B.

Fig. 24 shows the converter input impedance Zin and the output impedance of the device B Zo-B in the CC and CV regions. The magnitude of the converter input impedance exceeds the source impedance magnitude after 100 Hz in the CV region; thus, instability is predicted to occur in the CV region since phase difference of 180 is found at a higher frequency. A frequency range is found where Zin > Zo also in the CC region, but a phase difference of 180 does not exist, indicating stable operation in the CC region. This information is veried in the time domain in Fig. 25. It can be seen that the system is stable in the CC region but oscillates at the CV region when the same input-voltage-reference ramp was applied, as shown in Figs. 22 and 23. V. CONCLUSION In this paper, the properties of a PV generator have been analyzed analytically and with experimental measurements. A PV generator is internally a power- and voltage-limited nonlinear current source having both CC- and CV-like properties depending on the operating point. The dynamic properties of the generator, which include the dynamic resistance and capacitance, are also operating-point-dependent nonlinear quantities. The dynamic resistance of a PV generator has profound effects on dynamic behavior of the interfacing converter. The dynamic resistance is known to equal the static resistance at the MPP. The current- and voltage-source properties of the PV generator can also be justied by considering the dynamic resistance. The dynamic resistance is greater in magnitude than the static resistance in the current region and lower in magnitude in the voltage region. This property moves operating-pointdependent zeros and poles in the converter dynamics between the right and left halves of the complex plane according to the operating point of the PV generator. The appearance of RHP zeros and poles and their effect on the control dynamics of PV converters are important to be taken into account if robust and stable PV power systems are to be designed. Key points for the analysis, design, and testing of PV converters can be summarized as follows: 1) in order to transfer maximum power, an input-side variable (current or voltage) must be controlled, or the converter has to operate at open loop; 2) input-voltage-feedback control is the most feasible method

to control a PV converter; 3) in case of input-voltage-controlled converter, the converter must be analyzed in such a way that it is fed by a current source; and 4) the nonideal PV generator source impedance has to be considered because of its profound effect on the interfacing converter dynamics. A real PV generator is usually replaced with a power electronic substitute, known as a solar array simulator, so that timeinvariant and controlled testing conditions can be guaranteed in the converter validation process without the need to invest in a large PV power plants. This paper also qualied two commercial solar array simulators as an example in terms of the dened dynamic properties and analyzed the differences both in time and frequency domains between the substitutes and real PV generators. The measured substitutes were shown to reproduce the dynamic resistance accurately although the dynamic capacitances were considerably higher and did not show similar properties as the PV generator capacitance. It was also noticed that in order to reproduce the PV generator characteristics with the highest precision, the IV curve of the electronic substitute should be programmed using several currentvoltage pairs if possible. The output impedance of one of the substitutes was such that it made the solar array simulator/converter interface unstable. The properties of electronic substitutes should always be tested properly so that the same converter dynamics is reproduced with the substitute and the real PV generator. Based on the investigations presented in this paper, the most important properties the solar array simulator shall have in order to properly emulate a real PVG can be summarized as follows: 1) The low-frequency impedance, i.e., the dynamic resistance and the IV curve have to be emulated as accurately as possible. 2) The dynamic capacitance should be also emulated accurately enough though it is not as important as the emulation of the dynamic resistance. The capacitance of the simulators are typically higher than the capacitance of the PV generator. This can affect the behavior of the converters having small input capacitor. 3) The real PV generator shows passive-circuit-like behavior, which means that the phase of its output impedance stays within 90 . The emulator has to follow the same behavior. REFERENCES
[1] B. K. Bose, Global warming: Energy, environmental pollution, and the impact of power electronics, IEEE Ind. Electron. Mag., vol. 4, no. 1, pp. 617, Mar. 2010. [2] M. A. Eltawil and Z. Zhao, Grid-connected photovoltaic power systems: technical and potential problems: A review, Renew. Sust. Energy Rev., vol. 14, no. 1, pp. 112129, Jan. 2010. [3] G. Petrone, G. Spagnuolo, R. Teodorescu, M. Veerachary, and M. Vitelli, Reliability issues in photovoltaic power processing systems, IEEE Trans. Ind. Electron., vol. 55, no. 7, pp. 25692580, Jul. 2008. [4] J. H. R. Enslin and P. J. M. Heskes, Harmonic interaction between a large number of distributed power inverters and the distribution network, IEEE Trans. Power Electron., vol. 19, no. 6, pp. 15861593, Nov. 2004. [5] M. Cespedes and J. Sun, Modeling and mitigation of harmonic resonance between wind turbines and the grid, in Proc. IEEE Energy Convers. Congr. Expo., Phoenix, AZ, Sep. 2011, pp. 21092116. [6] J. Sun, Impedance-based stability criterion for grid-connected inverters, IEEE Trans. Power Electron., vol. 26, no. 11, pp. 30753078, Nov. 2011. [7] M. Cespedes and J. Sun, Renewable energy systems instability involving grid-parallel inverters, in Proc. 24th Annu. IEEE Appl. Power Electron. Conf. Expo., Palm Springs, CA, Feb. 2009, pp. 19711977.

NOUSIAINEN et al.: PHOTOVOLTAIC GENERATOR AS AN INPUT SOURCE FOR POWER ELECTRONIC CONVERTERS

3037

[8] T. Suntio, J. Huusari, and J. Lepp aho, Issues on solar-generator intera facing with voltage-fed MPP-tracking converters, EPE J., vol. 20, no. 3, pp. 4047, Sep. 2010. [9] T. Suntio, J. Lepp aho, J. Huusari, and L. Nousiainen, Issues on solara generator interfacing with current-fed MPP-tracking converters, IEEE Trans. Power Electron., vol. 25, no. 9, pp. 24092419, Sep. 2010. [10] J. Puukko, T. Messo, and T. Suntio, Effect of photovoltaic generator on a typical VSI-based three-phase grid-connected photovoltaic inverter dynamics, in Proc. IET Renewable Power Generation Conf., Edinburgh, U.K., Sep. 2011, pp. 16. [11] S. Liu and R. A. Dougal, Dynamic multiphysics model for solar array, IEEE Trans. Energy Convers., vol. 17, no. 2, pp. 285294, Jun. 2002. [12] M. G. Villalva, J. R. Gazoli, and E. R. Filho, Comprehensive approach to modeling and simulation of photovoltaic arrays, IEEE Trans. Power Electron., vol. 24, no. 5, pp. 11981208, May 2009. [13] D. Chenvidhya, K. Kirtikara, and C. Jivacate, PV module dynamic impedance and its voltage and frequency dependencies, Sol. Energy Mater. Sol. Cells, vol. 86, no. 2, pp. 243251, Mar. 2005. [14] A. M ki, S. Valkealahti, and T. Suntio, Dynamic terminal characteristics a of a photovoltaic generator, in Proc. 14th Int. Power Electron. Motion Control Conf., Ohrid, Macedonia, Sep. 2010, pp. T12-76T12-80. [15] J. Thongrpon, K. Kirtikara, and C. Jivacate, A method for the determination of dynamic resistance of photovoltaic modules under illumination, Sol. Energy Mater. Sol. Cells, vol. 90, no. 1819, pp. 30783084, Nov. 2006. [16] W. Xiao, W. G. Dunford, P. R. Palmer, and A. Capel, Regulation of photovoltaic voltage, IEEE Trans. Ind. Electron., vol. 54, no. 3, pp. 1365 1374, Jun. 2007. [17] L. Nousiainen, J. Puukko, and T. Suntio, Simple VSI-based single-phase inverter: dynamical effect of photovoltaic generator and multiplier-based grid synchronization, in Proc. IET Renewable Power Generation Conf., Edinburgh, U.K., Sep. 2011, pp. 16. [18] J. Wyatt and L. Chua, Nonlinear resistive maximum power theorem, with solar cell application, IEEE Trans. Circuits Syst., vol. 30, no. 11, pp. 824828, Nov. 1983. [19] L. Nousiainen and T. Suntio, Dc-link voltage control of a single-phase photovoltaic inverter, in Proc. 6th IET Int. Conf. Power Electron., Mach. Drives, Bristol, U.K., Mar. 2012, pp. 16. [20] J. Puukko, L. Nousiainen, and T. Suntio, Effect of minimizing input capacitance in VSI-based renewable energy source converters, in Proc. 33rd IEEE Int. Telecommun. Energy Conf., Amsterdam, The Netherlands, Oct. 2011, pp. 19. [21] A. M ki and S. Valkealahti, Power losses in long string and parallela connected short strings of series-connected silicon-based photovoltaic modules due to partial shading conditions, IEEE Trans. Energy Convers., vol. 27, no. 1, pp. 173183, Mar. 2012. [22] J.-F. Wu, C.-H. Chang, L.-C. Lin, and C.-L. Kuo, Power loss comparison of single- and two-stage grid connected photovoltaic systems, IEEE Trans. Energy Convers., vol. 26, no. 2, pp. 707715, Jun. 2011. [23] T. Messo, J. Puukko, and T. Suntio, Effect of MPP-tracking dc/dc converter on VSI-based photovoltaic inverter dynamics, in Proc. 6th IET Int. Conf. Power Electron., Mach. Drives, Bristol, U.K., Mar. 2012, pp. 16. [24] Y.-H. Liao and C.-M. Lai, Newly-constructed simplied single-phase multistring multilevel inverter topology for distributed energy resources, IEEE Trans. Power Electron., vol. 26, no. 9, pp. 23862392, Sep. 2011. [25] S. Vighetti, J.-P. Ferrieux, and Y. Lembeye, Optimization and design of a cascaded dc/dc converter devoted to grid-connected photovoltaic systems, IEEE Trans. Power Del., vol. 27, no. 4, pp. 20182027, Apr. 2012. [26] J.-M. Kwon, K.-H. Nam, and B.-H. Kwon, Photovoltaic power conditioning system with line connection, IEEE Trans. Ind. Electron., vol. 53, no. 4, pp. 10481054, Aug. 2006. [27] J. Lepp aho, L. Nousiainen, J. Puukko, J. Huusari, and T. Suntio, Implea menting current-fed converters by adding an input capacitor at the input of voltage-fed converter for interfacing solar generator, in Proc. 14th Int. Power Electron. Motion Control Conf., Ohrid, Macedonia, Sep. 2010, pp. T12-81T12-88. [28] C. K. Tse, Linear Circuit Analysis. Edinburgh, UK: Addison Wesley Longman, 1998, p. 307. [29] R. D. Middlebrook, Input lter considerations in design and applications of switching regulators, in Proc. IEEE Ind. Appl. Soc. Annu. Meeting Rec., Chicago, IL, Oct. 1976, pp. 366382. [30] J. Lepp aho, J. Huusari, L. Nousiainen, J. Puukko, and T. Suntio, Dya namic properties and stability assessment of current-fed converters in photovoltaic applications, IEEJ Trans. Ind. Appl., vol. 131, no. 8, pp. 976 984, Aug. 2011.

Lari Nousiainen (S09) was born in Nurmes, Finland, in 1984. He received the M.Sc. (Tech) degree in electrical engineering from the Tampere University of Technology, Tampere, Finland, in 2009. He is currently working toward the Ph.D. degree in the Department of Electrical Energy Engineering, Tampere University of Technology. He completed writing his doctorate thesis on the analysis and design of gridtied photovoltaic inverters in June 2012. In July 2012, he started working at ABB Drives in Helsinki, Finland, as a Design Engineer. Dr. Nousiainen is a Member of the IEEE Power Electronics Society, the IEEE Industrial Electronics Society, and the IEEE Power and Energy Society.

Joonas Puukko (S10) was born in Helsinki, Finland, in 1983. He received the M.Sc. (Tech.) degree (with distinction) in electrical engineering from the Tampere University of Technology, Tampere, Finland, in 2008, where he also received the Ph.D. degree from the Department of Electrical Energy Engineering. He completed writing his doctorate thesis in May 2012. In July 2012, he started working at ABB Drives in Helsinki, Finland, as a Design Engineer. His research interests included power electronics, threephase dc/ac power supplies, dynamic modeling, and interfacing of renewable energy systems. Dr. Puukko is a Member of the IEEE Power Electronics Society, the IEEE Industrial Electronics Society, and the IEEE Power and Energy Society.

Anssi M ki (S09) was born in Kauhava, Finland, a 1985. He received the M.Sc. (Tech.) degree in electrical engineering from the Tampere University of Technology, Tampere, Finland, in 2010, where he has been working toward the Ph.D. degree in the Department of Electrical Energy Engineering as a Researcher. His current research interests include the operation of photovoltaic power generators and the development of maximum-power-point-tracking algorithms. Mr. M ki is a Member of the IEEE Power and a Energy Society.

Tuomas Messo (S11) was born in Helsinki, Finland, in 1985. He received the M.Sc. (Tech.) degree in electrical engineering from the Tampere University of Technology, Tampere, Finland, in 2011, where he is currently working toward the Ph.D. degree in the Department of Electrical Energy Engineering, as a Researcher. His research interests include power electronics, three-phase dc/ac power supplies, dynamic modeling, control design, and interfacing of photovoltaic energy systems.

3038

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 28, NO. 6, JUNE 2013

Juha Huusari (S09) received the M.Sc. (Tech.) degree in electrical engineering from the Tampere University of Technology, Tampere, Finland, in 2009. He is currently working toward the Ph.D. degree in the Department of Electrical Energy Engineering, Tampere University of Technology. As of August 2012, he will be with ABB Corporate Research in Baden-D ttwil, Switzerland, working as a a Scientist with photovoltaic electricity systems. He has authored and co-authored 12 conference publications and ve journal publications. He also has one international patent application. His current research interests include analysis and design of distributed maximum-power-point-tracking dcdc converters, issues related to interfacing of photovoltaic generators, as well as practical switching-power-supply design issues. Dr. Huusari is a Member of the IEEE Power Electronics Society, the IEEE Industrial Electronics Society, and the IEEE Power Engineering Society.

Seppo Valkealahti (M10) was born in Alavus, Finland, 1955. He received the M.Sc. and Ph.D. degrees in physics from the University of Jyv skyl , a a Jyv skyl , Finland, in 1983 and 1987, respectively. a a From 1982 to 1997, he was a Teacher and Researcher of physics at the University of Jyv skyl , in a a the Riso National Laboratory in Denmark, and in the Brookhaven National Laboratory in NY. From 1997 to 2004, he worked in ABB heading research and product development activities. In the beginning of 2004, he joined the Tampere University of Technology, Tampere, Finland, where he is currently a Professor in the Department of Electrical Energy Engineering. His research interests include electric-powerproduction- and consumption-related technologies, solar energy, and multiscientic problems related to power engineering. Prof. Valkealahti is a Member of the IEEE Power and Energy Society.

Juha Jokipii (S11) received the M.Sc. (Tech.) degree in electrical engineering from the Tampere University of Technology, Tampere, Finland, in 2011, where he is currently working toward the Ph.D. degree in the Department of Electrical Energy Engineering. His research interests include power electronics, three-phase dc/ac power supplies, dynamic modeling, and interfacing of renewable energy systems.

Jukka Viinam ki received the B.Sc. (Tech.) degree a in embedded systems from the Tampere University of Applied Sciences, Tampere, Finland, in 2009. He is currently working toward the M.Sc. in electrical engineering at the Tampere University of Technology. Since 2012, he has been a Research Assistant in the Department of Electrical Energy Engineering, Tampere University of Technology. His research interests include dc/dc converters and dc/ac inverters in photovoltaic applications.

Teuvo Suntio (M98SM08) received the M.Sc. (Tech) and D.Sc. (Tech) degrees in electrical engineering from the Helsinki University of Technology, Espoo, Finland, in 1981 and 1992, respectively. From 1977 to 1991, he worked at Fiskars Power Systems as a Design Engineer and R&D Manager. From 1991 to 1992, he worked at Ascom Eergy Systems Oy as an R&D Manager. From 1992 to 1994, he was an entrepreneur in power electronics design consultancy, and from 1994 to 1998 he worked at Efore Oyj as a Consultant and Project Manager. Since 1998, he has been a Professor specializing in switched-mode power converter technologies rst at the University of Oulu, Electronics Laboratory, and from August 2004 in the Department of Electrical Energy Engineering, Tampere University of Technology, Tampere, Finland. He holds several international patents and has authored about 180 international scientic journal and conference papers, the book Dynamic Prole of Switched-Mode ConverterModeling, Analysis and Control (Weinhein, Germany: Wiley-VCH, 2009) as well as two book chapters. His current research interests include dynamic modeling, control design, optimal electromagnetic interference design of switched-mode power converters, as well as interfacing of renewable energy sources. Prof. Suntio is a Member of the IEEE Power Electronics Society, the IEEE Industrial Electronics Society, and the IEEE Circuits and Systems Society as well as a Member of the European Power Electronics and Drives Association. From the beginning of 2010, he has served as an Associate Editor for the IEEE TRANSACTION ON POWER ELECTRONICS. He has also served as a Guest EditorIn-Chief of the special issue on power electronics in photovoltaic applications in the IEEE TRANSACTION ON POWER ELECTRONICS.

Diego Torres Lobera (S10) was born in Palma de Mallorca, Spain, in 1985. He received the M.Sc. degree in industrial engineering from the University of Zaragoza, Zaragoza, Spain, in 2010. Since the beginning of 2011, he has been a Researcher with the Department of Electrical Energy Engineering, Tampere University of Technology, Tampere, Finland. His current research interests include operation and modeling of photovoltaic power generators and maximum-power-point-tracking techniques.

You might also like