Vous êtes sur la page 1sur 13

Design and implementation of the low cost and fast solar charger

with the rooftop PV array of the vehicle


Thanh-Tuan Nguyen, Hyung Won Kim, Geun Hong Lee, Woojin Choi

Department of Electrical Engineering, Soongsil University, 1-1 Sangdo 5-Dong, Dongjak-Gu, 156-743 Seoul, Republic of Korea
Received 8 April 2013; received in revised form 26 June 2013; accepted 8 July 2013
Available online 2 August 2013
Communicated by: Associate Editor Elias Stefanakos
Abstract
In this paper, a comprehensive and detail design procedure for the low cost and fast solar charger with the rooftop PV array of the
vehicle is presented. A simplied maximum power point tracking technique is adopted to take advantages of the simpler structure of the
circuit and the shorter charge time than those of the conventional solar charger. In the small-signal modeling of the converter, linearized
equivalent circuit models of the PV array and the leadacid battery are included for the accurate charge control all over the operating
range. An experimental prototype is implemented by using a buck converter with a digital signal processor as a controller and tested by
using the Elgar TerraSAS solar array simulator to verify the feasibility and validity of the system and its control algorithm.
2013 Elsevier Ltd. All rights reserved.
Keywords: Solar charger; Simplied maximum power point tracking; Low cost; Shorter charge time
1. Introduction
As the concerns of conventional energy source exhaus-
tion and environmental pollution increase, attempts are
being made to replace fossil fuels with non-conventional
energy sources in various sectors. In vehicular transporta-
tion eld, automotive industry is undergoing a revolution
in the design of its electrical system. This is the result of
increasingly sophisticated engine as well as the introduction
of new electrically controlled functions (Kassakian et al.,
1996). Using rooftop photovoltaic (PV) arrays is an eec-
tive way to aid in providing additional power which can
power the ventilation system or air conditioning system
in vehicles (Giannouli and Yianoulis, 2012). The advantage
of using rooftop PV arrays is that even when the car is
parked under sunlight, the arrays can produce electric
energy to charge the battery and then power the systems
to cool down the atmosphere inside the car. However, there
are several drawbacks in using PV arrays as an energy
source for the charger such as high fabrication cost, low
energy conversion eciency and long charge time. Thus,
when the PV array is used for charging the battery as an
auxiliary energy source, the PV system need to be low in
cost and it is important to charge the battery within a short
period of time by optimizing the power generated by the
PV arrays at certain operating conditions under the varying
irradiance and temperature level. An intelligent battery
charger for hybrid electric vehicle has been proposed to
show that the solar energy can be used as an additional
energy source (Lu et al., 2007). However, only the design
concept was presented and thus the details of the system
remained unknown. In (Pires et al., 2012; Traube et al.,
2013), the electric vehicle charger which is able to charge
the battery of the vehicle from both PV array and the ac
grid are presented, but it is an o-board type system thus
is not applicable to the on-board charger. An analog max-
0038-092X/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.solener.2013.07.006

Corresponding author. Tel.: +82 2 820 0652; fax: +82 2 817 7961.
E-mail addresses: thanhtuan.cn@gmail.com (T.-T. Nguyen),
cwj777@ssu.ac.kr (W. Choi).
www.elsevier.com/locate/solener
Available online at www.sciencedirect.com
Solar Energy 96 (2013) 8395
imum power tracker which is simple in hardware imple-
mentation and fast in tracking time was implemented in
Ji et al. (2012). However, since the boost converter topol-
ogy is used, the topology is not suitable for low voltage
application such as the battery charger. Also it requires
the voltage and current sensors to measure the peak power
of the PV array. In (Ahmed and Shoyama, 2011), an
attempt has been made to improve the performance of
the maximum power point tracking (MPPT) algorithm in
terms of the number of required sensors and the tracking
time. The proposed MPPT algorithm requires only a cur-
rent sensor associated with the PV array to measure the
output current of it, however, additional sensors are
required to implement the charge control algorithm for
the battery. The numerous kinds of maximum power point
tracking (MPPT) methods were summarized in Esram and
Chapman (2007). Some of them suggest that it is adequate
to maximize the load current to maximize the load power
when the load is of voltage source type. Thus the load cur-
rent can be used as single control variable for the MPPT,
thereby making it possible to reduce the number of sensors.
In this paper, a solar charger for vehicles which use roof-
top PV arrays as an auxiliary energy source is presented.
This charger employs a MPPT technique to force the PV
array generate maximum power and then deliver to charge
a leadacid battery. The MPPT technique is implemented
by controlling output current without sensing the voltage
and current of the PV array thereby reducing the complex-
ity and the cost of the system. Also the constant current
control and the constant voltage control are used together
with the MPPT method to reduce the charge time and to
prolong the battery life.
2. Low cost and fast solar charger and its charge algorithm
Fig. 1 shows the block diagram of the solar charger
with reduced number of sensors. A PV array is used as
a power source which generates power to charge the bat-
tery. A leadacid battery is used to store the energy from
the PV array and a DC/DC buck converter is used to
charge it. The charger operates one of three charge
modes: MPPT charge, constant current (CC) charge
and constant voltage (CV) charge depending both on
the available energy from the PV array and the state of
the charge in the leadacid battery. When the voltage
of the battery is lower than the full charge voltage and
the charge current is lower than the rated charge current
of the battery, the MPPT mode is employed to charge the
battery with maximum power. In this mode, if the charge
current exceeds the rated charge current (0.15C in this
case), the controller changes its control method to current
control mode (CCM) to limit the charge current. When
the battery voltage reaches to its limit value, the constant
voltage mode (CVM) charge is performed until the bat-
tery is fully charged.
The MPPT method can be implemented by maximizing
the charge current I
out
at the output of the buck converter
as shown in Fig. 1. With this method, since two sensors for
detecting the voltage and current of the PV array are not
required, the cost of the proposed solar charge system
may be 510% lower than that of the system presented in
Ke and Makaran, 2009. Further, it helps reduce the com-
plexity of the system and also provides the exibility in
selecting the PV array for the charger since no sensors asso-
ciated with PV array are employed.
3. Principle of the maximum power point tracking technique
with reduced number of sensors
The purpose of MPPT techniques is to automatically
nd the voltage or current value at maximum power point
at which the PV array should operate and generate its max-
imum available power, even in varying operating condition
(Esram and Chapman, 2007).
Among the various maximum power point tracking
techniques the P&O (Perturbation and Observation)
method has been most widely used in practice due to its
simplicity of the implementation (Enrique et al., 2010). In
P&O algorithm, the output voltage of the PV array which
represents the operating point of the PV array is perturbed
by changing the duty cycle of the power converter switch
(Esram and Chapman, 2007). It can be seen from Fig. 2
that when the operating point of the PV array is on the left
of the MPP and moving towards the MPP, then the output
Fig. 1. Block diagram of the proposed solar charger with reduced number
of sensors. Fig. 2. The characteristic curves of photovoltaic array.
84 T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395
voltage of the PV array is increased and thus the output
power of the PV array increases accordingly. This
relationship between the output voltage and the output
power of the PV array is reversed when the operating point
is on the right of the MPP. Therefore, if there is an increase
in power due to the duty cycle variation, the subsequent
variation should be kept the same in direction to reach
the MPPT and if there is a decrease in power, the variation
should be reversed. This process is repeated periodically
until the MPP is reached. The operating point of the PV
array will oscillate around the MPP until there is a change
in irradiance or temperature resulting in a movement of the
MPP.
The conventional P&O MPPT method requires the
information of the change in the output power of the
PV array which changes due to the duty cycle variation.
In order to see the change in the PV array output power,
it is necessary to measure the output voltage and the
output current of the array. However, it can be done
by only observing the change in the output current of
the buck converter due to an assumption that the battery
voltage changes slowly during the charge process
(Shmilovitz, 2005; Sullivan and Powers, 1993; Senjyu
et al., 2002).
By assuming no loss in the buck converter in Fig. 1, we
have
P
out
P
in
1
It is clear that the changes in the output power of the
converter in a short period of time (MPPT period) are
the same as the changes in the input power. Thus the fol-
lowing Eq. (2) becomes valid.
dP
out
dt

dP
in
dt
2
The change in the output power of the converter can be
rewritten as (3).
dP
out
dt

d V
out
I
out

dt

V
out
dI
out
dt

I
out
dV
out
dt
3
Since the terminal voltage of the battery is slowly
increased during the charge, the voltage variation of the
battery during an MPPT period can be assumed to be zero.
Thus the Eq. (3) can be rewritten as (4).
dV
out
I
out

dt

V
out
dI
out
dt
4
From (2) and (4) we have
dP
in
dt

V
out
dI
out
dt
5
By dividing the both side of the Eq. (5) by the duty var-
iation of the switch in an MPPT period (dD/dt), the small
variation in the input power to the converter due to the
variation of the duty can be derived as (6) and (7).
dP
in
=dt
dD=dt
V
out
dI
out
=dt
dD=dt
6
dP
in
dD
V
out
dI
out
dD
7
The Eq. (7) shows that the small variation of the input
power (output power of the PV array) due to the small var-
iation of the duty can be simply detected by observing the
variation of the output current of the buck converter. It
suggests that the voltage and the current information of
the PV array are not necessary for the P&O MPPT tech-
nique in the charge application. Hence the system cost
can be signicantly reduced by removing the sensors for
the measurement of the PV output power. The owchart
for this algorithm is shown as in Fig. 3.
4. Design of the reactive components of the buck converter
considering current and voltage ripple limits of the leadacid
battery
Based on the rated power of the charge system, the char-
acteristics of the PV array and the voltage and current rip-
ple limits of the leadacid battery, reactive components of
the buck converter can be designed by using a conventional
method as followings.
4.1. Design of the inductor of the buck converter
Inductor is designed based on the maximum allowable
ripple value in the charge current of the leadacid battery.
The charge current ripple should be lower than the charge
current at the end of charge which is 0.02C to maintain the
continuous conduction mode (Trojan Battery Users
Guide, 2012). Thus the inductor can be designed by the fol-
lowing equation.
L P
V
o max
1 D
f
sw
jDI
L
j
8
where f
sw
is the switching frequency (Hz), D the duty cycle
at the end of the charge process (D = V
o_max
/V
pv
, V
pv
is the
output voltage of the PV array), DI
L
is the desirable ripple
value of the inductor current (A) (DI
L
= 0.02C rate in this
case), and V
o_max
is the maximum value of the dc compo-
nent of output voltage (V) (Battery voltage at the end of
charge).
4.2. Design of the input capacitor of the buck converter
In the conventional buck converter design, the output
capacitor is sized to limit the output voltage ripple. How-
ever, it is not necessary in the leadacid battery charge
application since the huge capacitance already exists in
the equivalent circuit model of the leadacid battery. While
the output capacitor can be omitted in this design, the
input capacitor has to be included to limit the PV output
current ripple less than 2% of its mean value (Koutroulis
et al., 2001). The input capacitance value can be calculated
by (9).
T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395 85
C
i
>
1 D
max
I
o max
D
max
0:02I
pv max
R
pv max
f
sw
9
where f
sw
is the switching frequency (Hz), D
max
is the duty
cycle at maximum output power, I
o_max
is the charge cur-
rent when the PV array works at the MPP (A), I
pv_max
is
the PV array output current when the PV array works at
the MPP (A), and R
pv_max
is the equivalent resistance of
the PV array at the MPP (O).
All the system parameters of the solar charger are listed
in Table 1.
5. Design of the charge controller including the linearized
equivalent circuit models of the PV array and the leadacid
battery
5.1. Linearized equivalent circuit model of the PV array
The characteristic equation of a PV array can be repre-
sented by (10).
i
pv
I
ph
I
s
exp
v
pv
R
s
i
pv
V
T
g
_ _
1
_ _

v
pv
R
s
i
pv
R
p
10
where i
pv
is the output current of the PV array (A), v
pv
is the
output voltage of the PV array (V), R
s
is the series resis-
tance (X), R
p
is the shunt resistance (X), I
ph
is the light in-
duced current (A), g is the diode ideality factor, I
s
is the
saturation current of the PV array (A), and V
T
is the ther-
mal voltage (V).
The equivalent circuit of a PV array can be illustrated as
Fig. 4(a).
As shown in Fig. 2, when the output voltage of the PV
array is lower than V
mp
the array behaves like a current
source and when the output voltage is higher than V
mp
it
acts like a voltage source. Practically, the series resistance
R
s
has dominant inuence on the operation of the PV mod-
ule when it operates in voltage source region, while the par-
allel resistance R
p
has stronger inuence than series
resistance in the current source region (Villalva et al.,
2009). However, this non-linear equivalent circuit of the
PV array is rarely used for the circuit analysis. In order
to reduce the complexity of the circuit analysis, lineariza-
tion technique can be used to derive a further simplied
equivalent circuit of the PV array at a certain operating
point (Villalva et al., 2010).
In order to obtain the slope of the linearized character-
istic curve of the PV array at a certain operating point, a
partial derivative of with respect to the PV voltage needs
to be calculated.
@i
pv
@v
pv

V ;I
R
s

1
I
s
gV
T
e
V RsI
gV
T

1
Rp
_
_
_
_
1
11
The linear model of the PV array can be described by the
tangent at an operating point (V, I) as (12).
Fig. 3. Proposed maximum power point tracking algorithm for the solar charger.
Table 1
System parameters of the solar charger.
Switching frequency f
s
60 (kHz)
Inductor L 100 (lH)
Inductor current I
L
6.0 (A)
Input capacitor C 220 (lF)
Input voltage at MPP V 17.0 (V)
Input voltage at point A V 12.0 (V)
Input voltage at point B V 20.5 (V)
Equivalent resistance of the PV array at MPP R
eq
1.25 (X)
Equivalent resistance of the PV array at point A R
eq
31.18 (X)
Equivalent resistance of the PV array at point B R
eq
0.39 (X)
Equivalent capacitance of the battery C
b
90,000 (F)
Equivalent resistance of the battery R
b
0.02 (X)
86 T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395
i
pv
g v
pv
V I 12
where g
@ipv
@vpv

V ;I
.
Thus the simplied equivalent circuit model of the PV
array can be represented by a voltage source and a series
resistor as shown in Fig. 4(b). (where R
eq
= 1/g and
V
eq
= VI/g).
However, since the above linearized model is only valid
at a certain operating point, it is necessary to linearize the
PV model at each operating point all over the operating
range of the buck converter on the IV characteristics
curve.
5.2. Modeling of the buck converter including linearized PV
array model and battery model
In this section, the control-to-inductor current and the
control-to-output voltage transfer functions are derived
by using small-signal modeling technique for the charge
control.
The linearized equivalent circuit model of the PV array
in Fig. 4(b) is combined with a buck converter and an
equivalent circuit model of the leadacid battery. The bat-
tery is modeled with an RC series circuit, where R
b
and C
b
represent the equivalent series resistance and the equivalent
capacitance of the battery, respectively (Jossen, 2006).
These two parameters of the equivalent circuit of the bat-
tery can be obtained by using current interruption test
and coulomb counting method. In this case, the equivalent
resistance R
b
and the equivalent capacitance C
b
of the
leadacid battery are equal to 0.02 (X) and 90,000 (F),
respectively.
The PWM converter can be modeled by the circuit aver-
aging technique. Thus, the averaged model of the solar
charger can be developed as shown in Fig. 5.
The control variable is the duty cycle d and the MOS-
FET switches at the frequency of f
s
= 1/T
s
. The MOSFET
is on during the interval dT
s
and o during the interval
(1 d) T
s
. By applying the Kirchhos voltage law
(KVL) for the secondary loop and the Kirchhos current
law (KCL) at node N, the voltage equation of the second-
ary loop and the current equation at the node N can be
written in (13) and (14).
v

d L
d
dt

i
L

i
L
Z
o
0 13
V
eq
v
R
eq
C
d
dt
v

i
L

d 0 14

i
L
; v and

d denote the average values of the inductor cur-
rent, the input voltage of the buck converter and the duty
cycle of the switching, respectively. In order to construct a
small-signal model, it is assumed that the averaged output
voltage v, the averaged inductor current

i
L
and the aver-
aged duty cycle

d waveforms will be equal to the corre-
sponding dc value V, I
L
and D, plus some superimposed
small ac variation ~v,
~
i
L
and
~
d (Erickson and Maksimovic,
2001). Hence, we have
v V ~v;

i
L
I
L

~
i
L
;

d D
~
d 15
where ~v << V ;
~
i
L
<< I
L
and
~
d << D.
By substituting (15) into (13) and (14), and then remov-
ing dc components, the non-linear products and by apply-
ing Laplace transformation, the control to output current
and the control to voltage transfer functions are obtained
in Laplace domain as follows.
G
id
s
sC
b
V 1 sCR
eq
I
L
R
eq
D
_
s
2
LC
b
sC
b
R
b
1 sCR
eq
1
_ _
sD
2
R
eq
C
b
16
G
vd
s
1sR
b
C
b
V 1sCR
eq
DI
L
R
eq
_
s
2
LC
b
1sCR
eq
1sR
b
C
b
1sCR
eq
sC
b
D
2
R
eq
17
The transfer functions of G
vd
(s) and G
id
(s) are investi-
gated at three certain operating points on the IV curve
of the PV array, which are point A in the current source
region, point B in the voltage source region and the MPP
as shown in Fig. 2. Point A and point B are chosen accord-
ing to the limit of the PV array operating point in this
study. Thus, point A in the current source is selected in
order to have the PV output voltage equal to 12.0 (V)
which is corresponding to the leadacid battery voltage.
Also, point B in the voltage source region is selected at
the point where the PV output current is 0.52 (A). This is
the operating point of the PV array at the time when the
charge current decreases to about 0.8 (A) indicating that
the leadacid battery is fully charged.
Fig. 4. Equivalent circuit of a PV array (a), and its linearized equivalent circuit (b).
T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395 87
5.3. Design of the current and voltage controller
According to the charge algorithm of the solar charger,
the constant voltage mode and the constant current mode
can be performed separately. The current control mode is
employed if the charge current exceeds the rated charged
current (0.15C) to limit the charge current at the rated
charged value. In the other case, the voltage control mode
is activated when the battery voltage reaches to the limit
voltage 14.0 (V) to charge the battery in the constant volt-
age mode (Dunlop, 1997).
Fig. 6 shows the frequency responses of the control to
output current and the control to output voltage of the
converter obtained by the PSIM simulation when the oper-
ating point of the PV array is at MPP. In this gure, the
frequency responses of the output current and voltage are
drawn by using ac sweep method for the solar charger cir-
cuit and the transfer functions, G
id
(s) and G
vd
(s) obtained
by the small-signal modeling. Since both gain response
and phase response in each case are matched each other,
the derived transfer functions can be considered valid.
Thus it can be used for investigating the dynamic charac-
teristics of the converter and designing the controllers.
Fig. 7 shows that the Bode plots of the G
id
(s) at three
dierent operating points of the PV array. It is noticed that
when the operating point moves from point A in the cur-
rent source region to point B in the voltage source region,
the resonant frequency of the system varies slightly around
350 (Hz) while the dc gain and the damping factor which
aect to the dynamics of the system vary signicantly. In
the design of the current controller a zero will be located
at the resonant frequency of the converter to avoid the
interaction in phase. Since the small change in the resonant
frequency makes it easier to design the PI controller for the
Fig. 5. The averaged model of the solar charger.
Fig. 6. Validation of G
id
(s) and G
vd
(s) transfer function at the MPP using AC sweep.
88 T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395
current control for all three cases, it can be considered as
an advantage in designing the current controller.
Similarly, by investigating the G
vd
(s) transfer function at
point A, MPP and point B, the Bode plots are drawn as
shown in Fig. 8. The resonant frequencies also slightly vary
in three cases at about 350 (Hz). However, the dc gains of
the Bode plots at MPP and point A are negative, thus it is
required to increase the low frequency gain of the system in
these cases by the controller.
It is clearly shown in Figs. 7 and 8 that the phase mar-
gins in the three cases are more than 90, which make the
converter operation stable enough. Since the gain at low
frequency of the open-loop system (both of G
id
(s) and
G
vd
(s) transfer function) in the current source region is
Fig. 7. Bode plot of G
id
(s) at three certain operating points of the PV array.
Fig. 8. Bode plot of G
vd
(s) at three certain operating points of the PV array.
T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395 89
low, a PI controller is employed to achieve a higher dc gain
at the low frequency range in the closed loop control.
The transfer function for the closed-loop control sys-
tems can be represented as (18).
Ts
G
pi
sGs
1 G
pi
sGsHs
18
where G
pi
(s) is the PI controller for the output current or
output voltage, G(s) is the control to output current or con-
trol to output voltage transfer function and H(s) is the gain
of the sensing circuit.
As mentioned above each PI controller is designed to
have a zero frequency equal to one-tenth of the resonant
frequency of the converter to avoid the interaction in the
Fig. 9. Bode plots of the closed-loop system for output current control at three certain operating points.
Fig. 10. Bode plots of the closed-loop system for output voltage control at three certain operating points.
90 T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395
phase response while still providing higher gains
(Tymerski, 2009). Hence, the PI controller having zero
frequency at 35 (Hz) is designed as (19) and (20) for the
current control and voltage control, respectively.
G
pi i
s
0:43s 95
s
19
G
pi v
s
17:85s 3925
s
20
Fig. 9 shows the Bode plots of the transfer functions of
the closed-loop current control with a PI controller at three
dierent operating points, point A, MPP and point B. It
can be seen in Fig. 9 that with the PI compensator G
pi_i
(s)
shown in (19), the closed-loop system is stable in the range
of PV operation with sucient phase margins. The Bode
plot of the transfer function of the closed-loop voltage con-
trol is also shown in Fig. 10. It can be seen that with the PI
Fig. 11. PSIM simulation results of the low cost MPPT algorithm.
Fig. 12. System conguration.
T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395 91
controller designed as (20) the stability of the closed-loop
voltage control system is ensured with high phase margin
for all three cases.
6. PSIM simulation result of the low cost MPPT algorithm
PSIM simulation was performed to verify the proposed
MPP tracking algorithm. The circuit for the simulation
consists of a PV model which is available in the PSIM soft-
ware as a library, a buck converter and an RC equivalent
circuit model of the leadacid battery. The MPPT algo-
rithm was composed by the C language and put in a C
block. The MPPT cycle was set at 30 ms. The duty cycle
was changed at each MPPT cycle by a constant step of
0.5% in this study. The test condition was set at 1200
(W/m
2
) of irradiance and 30 C of temperature. Under this
condition, the maximum generated power from the PV
array consisted of 36 cells in series is 115.6 (W) at the PV
array voltage 17.0 (V) and its current 6.8 (A).
The Fig. 11 shows the simulation results. When the
MPPT starts to work, as the output current of the PV array
increases to 6.8 (A), the output voltage of the PV array
decreases from the open-circuit voltage (21.2 (V)) to the
MPP voltage (17.0 (V)). At the same time the battery
charge current increases until it reaches to its maximum
value (7.6 (A)). The actual output power of the PV array
increases to its maximum available output power indicat-
ing that the MPP has been tracked successfully under the
test condition. After the MPP is tracked, the output current
of the PV array oscillates around the maximum value. It
takes about 1.8 s to track the MPP in this case. The MPP
tracking performance will be veried by the experiments
in the following section.
7. Experimental results
The MPPT and the battery charge algorithm described
above are experimentally tested with the Elgar TerraSAS
solar array simulator and a 12 V 40Ah leadacid battery.
The digital signal processor (DSP) TMS320F28335 from
Texas Instrument (TI) was used for full digital control of
the proposed MPPT algorithm and the charge method.
Fig. 13. The PV array outputs in MPPT mode.
Fig. 14. Battery voltage and current in MPPT mode.
92 T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395
The conguration of the system is shown in Fig. 12. The
MPPT cycle was set at 30 ms while the duty step (dD)
was 0.005.
The tracking algorithm was rst tested under a constant
irradiance and temperature level which can be set by soft-
ware installed on a PC. The irradiance level was set at
1200 (W/m
2
) while the temperature was set at 30 C. Under
this operating condition, the PV array generated a maxi-
mum power of 115.6 (W). At the beginning the PV output
voltage was equal to the open-circuit voltage 21.2 (V) and
the charge current was zero. Duty cycle was then adjusted
by the MPPT algorithm and consequently the PV array
output current started to increase while the PV output volt-
age decreased as shown in Fig. 13, hence the charge current
increased during this process. After the charge current
reached its maximum value at 7.6 (A) indicating that the
MPP had been tracked, the algorithm then kept the system
working around it to charge the battery with maximum
power from the PV array (Fig. 14).
Fig. 15 shows the operation of the solar charger under
the varying irradiance condition. During MPPT mode, if
the charge current exceeded the rated charge current (6.0
(A), 0.15C), the mode was immediately switched to CC
mode to keep the charge current at that value. As shown
Fig. 15. Mode transition from MPPT mode to CC mode due to the increase of the irradiance level.
Fig. 16. Irradiance proles in sunny and cloudy condition.
T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395 93
in Fig. 15 under the lower irradiance level, 700 (W/m
2
), the
charger performed MPPT because the charge current was
lower than the rated charge current. When the irradiance
level changed to 1200 (W/m
2
), which could deliver 7.6
(A) maximum charge current to the battery as shown in
Fig. 14, the charger switched its mode to CC mode to limit
the charge current at the rated value, 6.0 (A). The eciency
of the MPPT algorithm was measured with the support of
PV simulator as 99.4%.
In order to prove the superiority of the proposed charge
method in terms of charge time, total charge time was mea-
sured under the sunny and cloudy condition, respectively
and compared to the CVM charge method, which is often
adopted for the conventional solar charger. Before the start
of each test, the SOC of the battery was carefully adjusted
to provide the same condition for each test by using an
electronic load (Kikusui PLZ 1004 W). The test was started
under a sunny condition of which irradiance level was 850
(W/m
2
) (t = 14,000 (s)). It was increased to the maximum
value of 1050 (W/m
2
) (t = 22,000 (s)) and decreased to
930 (W/m
2
) at the end of the test (t = 27,500 (s)) as shown
in Fig. 16(a). The test was also started under a cloudy con-
dition of which irradiance level was 625 (W/m
2
) (t = 9,800
(s)). It was varied according to the prole inside the instru-
ment and nally ended up with 300 (W/m
2
) (t = 22,800 (s))
as shown in Fig. 16(b).
Fig. 17 shows that at the beginning (from t = 0 (s) to
t = 3,500 (s)) charge current by the proposed method is
higher than that of the conventional CVM charge method
due to the MPPT being performed. The charge current by
MPPT method increases gradually due to the increase in
the irradiance level. The maximum power point is continu-
ously tracked in this duration. Meanwhile, the charge cur-
rent by the conventional CVM reduces gradually. Once the
battery voltage increases to 14.0 (V), the proposed solar
charger switches to CVM to charge the battery with a con-
stant voltage. During the CVM charge, the charge current
is reduced according to the state of charge of the battery. In
both of these methods, the charge processes were stopped
as soon as the charge current decreased to 0.8 (A)
(0.02C) indicating that the battery is fully charged.
Fig. 17. Batter charge proles in two charge methods (proposed and
conventional CVM charge algorithm) under sunny condition.
Fig. 18. Battery charge proles in two charge methods (proposed and
conventional CVM charge algorithm) under cloudy condition.
94 T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395
Fig. 17 also shows that while the conventional charge time
takes 3.9 h, proposed charge method with MPPT technique
takes only 3.3 h which is about 85% of the required charge
time by the conventional CVM charge method. Under the
cloudy condition, as shown in the Fig. 18, the MPP track-
ing was performed properly and the proposed charger was
able to deliver higher current to charge the leadacid bat-
tery than that by the conventional CVM charger. The
charge times of the conventional and proposed method
were 3.6 h and 3.3 h, respectively. With the proposed
method the battery can be charged 0.3 h faster than that
by the conventional method and it is suggested that the
7.6% charge time can be saved by the proposed charge
method.
8. Conclusion
In this paper, a low cost and fast solar charger has
been proposed, and its validity and feasibility were pro-
ven by the simulation and experiments. The MPPT tech-
nique was achieved only with the battery current and
voltage, thereby reducing the number of the sensors hence
the cost of the charger. It is expected that the cost of the
proposed charger can be 510% lower than that of the
existing solar charger of the same power ratings. Also,
the MPPT technique was used in combination with the
CC and CV charge method in order to obtain a maxi-
mum charge eciency thereby reducing the charge time.
It has been proven by the experiments that the proposed
solar charger can save the charge time by 15% and 7.6%
compared to the conventional CVM charger under the
sunny and cloudy condition, respectively. The proposed
solar charger and its control algorithm can be used not
only for the rooftop PV array of the vehicle but also
for other stand-alone photovoltaic systems. However, in
order to further improve the performance of the solar
charger under the rapidly varying irradiance condition,
the MPPT cycle and the size of the duty step need to
be optimized based on the analysis on the step response
characteristics of the current.
Acknowledgment
This work was supported by the Human Resources
Development program 20124030200070 of the Korea Insti-
tute of Energy Technology Evaluation and Planning (KE-
TEP) grant funded by the Korea government Ministry of
Trade, Industry and Energy.
This work (research) was nancially supported in part
by the Ministry of Knowledge Economy (MKE) and Kor-
ea Institute for Advancement in Technology (KIAT)
through the Workforce Development Program in Strategic
Technology.
References
Ahmed, E.M., Shoyama, M., 2011. Variable step size maximum power
point tracker using a single variable for stand-alone battery storage pv
systems. Journal of Power Electronics 11, 218227.
Dunlop, J.P., 1997. Batteries and charge control in stand-alone photo-
voltaic system: fundamental and application. Florida Solar Energy
Center, 31.
Enrique, J.M., Andu jar, J.M., Boho rquez, M.A., 2010. A reliable, fast and
low cost maximum power point tracker for photovoltaic applications.
Solar Energy 84, 7989.
Erickson, R.W., Maksimovic, D., 2001. Fundamental of Power Electron-
ics, second ed. Kluwer Academic.
Esram, T., Chapman, P.L., 2007. Comparison of photovoltaic array
maximum power point tracking techniques. IEEE Transactions on
Energy Conversion 22, 439449.
Giannouli, M., Yianoulis, P., 2012. Study on the incorporation of
photovoltaic systems as an auxiliary power source for hybrid and
electric vehicles. Solar Energy 86, 441451.
Ji, S.K., Jang, D.H., Hong, S.S., 2012. Analog control algorithm for
maximum power trackers employed in photovoltaic applications.
Journal of Power Electronics 12, 503508.
Jossen, A., 2006. Fundamentals of battery dynamics. Journal of Power
Sources 154, 530538.
Kassakian, J.G., Wolf, H.C., Miller, J.M., Hurton, C.J., 1996. Automo-
tive electrical systems circa 2005. Spectrum, IEEE 33, 2227.
Ke, L., Makaran, J., 2009. Design of a solar powered battery charger. In:
Electrical Power & Energy Conference (EPEC), 2009 IEEE, pp. 15.
Koutroulis, E., Kalaitzakis, K., Voulgaris, N.C., 2001. Development of a
microcontroller-based, photovoltaic maximum power point tracking
control system. IEEE Transactions on Power Electronics 16, 4654.
Lu, W., Yamayee, Z.A., Melton, A., Realica, A., Turner, J., 2007. Work
in progress a solar powered battery charger for a hybrid electric
vehicle. In: Frontiers In Education Conference Global Engineering:
Knowledge Without Borders, Opportunities Without Passports. FIE
07. 37th Annual, pp. 1314.
Pires, V.F., Roque, A., Sousa, D.M., Marques, G., 2012. Photovoltaic
electric vehicle chargers as a support for reactive power compensation.
In: Renewable Energy Research and Applications (ICRERA), 2012
International Conference on, pp. 16.
Senjyu, T., Shirasawa, T., Uezato, K., 2002. A maximum power point
tracking control for photovoltaic array without voltage sensor. Journal
of Power Electronics 2, 155166.
Shmilovitz, D., 2005. On the control of photovoltaic maximum power
point tracker via output parameters. Electric Power Applications, IEE
Proceedings 152, 239248.
Sullivan, C.R., Powers, M.J., 1993. A high-eciency maximum power
point tracker for photovoltaic arrays in a solar-powered race vehicle.
In: Power Electronics Specialists Conference, PESC 93 Record., 24th
Annual IEEE, pp. 574580.
Traube, J., Fenglong, L., Maksimovic, D., Mossoba, J., Kromer, M.,
Faill, P., Katz, S., Borowy, B., Nichols, S., Casey, L., 2013. Mitigation
of solar irradiance intermittency in photovoltaic power systems with
integrated electric-vehicle charging functionality. Power Electronics,
IEEE Transactions on 28, 30583067.
Trojan Battery Users Guide, 2012. Trojan Battery.
Tymerski, R., 2009. Compensators for the Buck Converter: Design and
Analysis. Portland State University.
Villalva, M.G., Gazoli, J.R., Filho, E.R., 2009. Modeling and circuit-
based simulation of photovoltaic arrays. In: Power Electronics
Conference, 2009. COBEP 09. Brazilian, pp. 12441254.
Villalva, M.G., de Siqueira, T.G., Ruppert, E., 2010. Voltage regulation of
photovoltaic arrays: small-signal analysis and control design. Power
Electronics, IET 3, 869880.
T.-T. Nguyen et al. / Solar Energy 96 (2013) 8395 95

Vous aimerez peut-être aussi