Next Article in Journal
Solutions of the Yang–Baxter Equation and Automaticity Related to Kronecker Modules
Previous Article in Journal
Composite Mould Design with Multiphysics FEM Computations Guidance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low-Voltage Photovoltaic System Based on a Continuous Input/Output Current Converter

by
Carlos Andres Ramos-Paja
1,*,
Juan David Bastidas-Rodriguez
2 and
Andres Julian Saavedra-Montes
1
1
Facultad de Minas, Universidad Nacional de Colombia, Medellin 050041, Colombia
2
Facultad de Ingeniería y Arquitectura, Universidad Nacional de Colombia, Manizales 170003, Colombia
*
Author to whom correspondence should be addressed.
Computation 2023, 11(2), 42; https://doi.org/10.3390/computation11020042
Submission received: 18 November 2022 / Revised: 31 January 2023 / Accepted: 7 February 2023 / Published: 20 February 2023

Abstract

:
Low-voltage photovoltaic systems are being widely used around the world, including their introduction into the power grid. The development of these systems requires the adaptation of several power converters, their static and dynamic modeling, the design of passive elements, and the design of the controller parameters, among other actions. Today, power converters are key elements in the development of photovoltaic systems, and classical power converters such as buck converters produce discontinuous input and output currents, requiring a high input capacitance and impacting the output power quality of these systems. This paper presents a proposal for a low-voltage photovoltaic system that uses a continuous input/output current buck converter, which enhances the operation of the classical buck converter in photovoltaic systems. The methodology describes the proposed photovoltaic system, including the power converter, its detailed operation, and the analysis of its waveforms. Moreover, the methodology includes a mathematical model of the photovoltaic system’s dynamic behavior and the design of a sliding-mode controller for maximum power extraction and perturbation rejection. The photovoltaic system is validated in two ways: first, a comparison with the classical buck converter highlighting the advantages of continuous input/output currents is presented; then, an application example using commercial devices is described in detail. The application example uses a flowchart to design the power converter and the sliding-mode controller, and a circuit simulation confirms the advantages of the continuous input/output current buck converter with its controller. In the circuit simulation, the control strategy is formed by a perturb and observe algorithm that generates the voltage reference for the sliding-mode controller, which guarantees the system stability, tracks the maximum power point, and rejects the double-frequency oscillations generated by an intended microinverter.

1. Introduction

Photovoltaic (PV) systems continue to be established as one of the most important renewable energy sources all around the world, since the installed capacity grows year after year. In 2021, about 175 GW were installed and commissioned, reaching a global installed capacity of 942 GW [1]. Moreover, according to the IEA, the Net Zero Emissions by 2050 scenario requires a significant increment in annual PV generation in the next years [2], which suggests that PV systems will continue growing in the future.
Typically, a grid-connected PV system is formed by a PV array, a capacitor ( C p v ), a dc/dc converter, a dc-link capacitor, a dc/ac converter, and a control system [3]. The PV array transforms the light power into electric power and is connected to the dc/dc converter through C p v , while the dc/dc converter modifies the array operating voltage to extract the maximum power from the array. Moreover, the dc/dc converter delivers its power to a dc-link capacitor (or a battery [4,5]), which forms a dc bus used by the dc/ac converter, whose main function is to deliver the power to ac loads, the grid, or both. The control system has three main objectives: The first one is to extract the maximum power from the PV array by using a maximum power point tracking (MPPT) technique, which generates a reference for the PV array voltage (or current). The second objective is to track the voltage (or current) reference generated by the MPPT with a controller to reject disturbances that may produce oscillations in the PV power or malfunction of the MPPT algorithm. The third objective of the control system is to inject the PV power to the grid, which requires synchronization with the grid voltage and injecting ac current to keep the dc-link capacitor voltage approximately constant [3].
Step-up dc/dc converters are widely used in grid-connected PV systems since the dc-link voltage is usually greater than the PV array voltage (hundreds of volts [6,7]). However, there are commercial inverters to supply ac loads where the dc-link voltage may be between 12V and 70V, such as off-grid PV inverters (e.g., [8,9,10,11]), PV inverters that can operate on-grid or off-grid (e.g., hybrid PV inverters [12,13,14]), and inverters supplied by a dc bus or a battery (e.g., [15,16,17,18]). In these applications, the inverter’s nominal input voltage corresponds to the battery’s nominal voltage (between 12V and 70V), while the PV array voltage is greater than those values due to the typical connection of PV modules in series and parallel to form the array. Therefore, step-down converters are required to couple the PV array to the dc bus.
In the literature, the buck converter is the most widely used step-down topology for PV systems to extract the maximum power from the PV generator, due to its simplicity, low number of components, and simple controllers [19]. However, the buck converter may be used in different forms depending on the PV system’s structure to couple the PV array to a dc bus [4,20], charge a battery that feeds a stand-alone inverter [4], feed a boost converter that elevates the voltage for a grid-connected inverter [20], regulate the dc bus voltage [21] or current [22,23] that feeds an inverter, and implement multilevel inverters [5,24], among others.
There are also different MPPT strategies in the PV systems discussed before. Although the two main implemented MPPT algorithms are perturb and observe (P&O) [4,5,22] and Incremental Conductance (IC) [20,23,24], some authors do not include an MPPT technique in the control system [21]. Moreover, some MPPT techniques generate the duty cycle of the buck converter with P&O [4,5] or IC [20,24] algorithms, while, in other systems, the output of the MPPT techniques is a reference of the PV array voltage [22] or current [23] with P&O or IC algorithms, respectively. At this point, it is important to mention that the MPPT algorithms that generate the converter’s duty cycle are simpler to implement, but they cannot reject disturbances in the PV array voltage or current, which may produce unexpected variations in the power harvested from the PV array and, in some cases, the malfunction of the MPPT [25]. In a PV array connected to an inverter, one of the main perturbations is the 100 Hz or 120 Hz oscillations in the PV array voltage produced by the 50 Hz or 60 Hz ac voltage generated by the inverter [25]. That is why it is common to implement compensators for the PV array voltage or current in the dc/dc converter connected to the array.
Moreover, most of the discussed PV systems use linear compensators [22,26], while others propose more complex alternatives [27,28]. Linear compensators are implemented to track a voltage [22] or current [23] reference generated by the MPPT technique; nevertheless, although the linear compensators described before can reject perturbations in the PV array voltage [22] or current [23], only the compensator proposed in [22] rejects the 100 Hz or 120 Hz oscillations in the PV array voltage produced by the inverter. Regarding the more complex controllers, it is possible to find a deadbeat controller of the PV array current that generates the converter’s duty cycle with an explicit expression, which can be easily implemented on a microcontroller [27]. Moreover, in [28], the authors use a cascade controller, where the inner loop is a sliding-mode controller of the inductor current and the outer loop is a P controller that tracks the PV array voltage reference generated by the MPPT technique.
In the PV systems described before, it is important to highlight that the input current of the buck converter connected to the PV arrays is discontinuous, since it corresponds to the MOSFET current [29]; therefore, it is necessary to use an input capacitor ( C p v ) in the range of hundreds of μ F (e.g., [21,23]) or even thousands of μ F (e.g., [26,30]) to obtain a continuous current in the PV array. Such capacitance ranges can be obtained with electrolytic capacitors [31], which not only have a shorter lifetime than other capacitors and affect converter reliability [31,32], but also have increased cost and weight [33].
One feasible option to reduce the capacitance of C p v and avoid the use of electrolytic capacitors is to replace the buck converter with the topology denominated, from here on, as a continuous input/output current (CIOC) buck converter (also known as superbuck converter), which is formed by two inductors, one capacitor, one MOSFET, and one diode [34,35]. This converter provides a continuous input current, which allows a significant reduction in the C p v capacitance, with the same input–output voltage and current ratios of the buck converter [34,35]. Although the CICO buck converter was originally proposed for a rectifier with power factor correction [36], it has also been used for different applications, such as dc voltage regulation for resistive loads [37,38] or generic loads [39,40,41], as well as the battery interface for a dc bus [42,43]. Nevertheless, to the best knowledge of the authors, it has only been used to extract the power from a PV array in [44].
The PV system proposed in [44] is formed by a PV module, a CICO buck, a battery, and an output capacitor (between the converter’s output and the battery). The paper proposes a cascade controller, where the inner loop tracks the PV array’s current reference generated by an MPPT algorithm, and the outer loop regulates the output voltage; however, the outer loop is only active when the output voltage surpasses the maximum voltage defined for the system. The controllers proposed in [44] are based on the small-signal model of the system and include stability analysis; additionally, the paper includes experimental results with the controller implemented with analog circuits. However, the proposed PV system uses two additional capacitors regarding the basic CIOC buck topology (i.e., output capacitor and C p v ), and C p v is in the range of electrolytic capacitors ( C p v = 1 mF), which reduces the converter reliability and increases its cost and weight. Additionally, the MPPT algorithm is not implemented, and the proposed PV array current controller is not clearly explained. Moreover, the paper does not include a procedure or guideline to design the PV array current controller, and it is not able to reject 100 Hz or 120 Hz oscillations in the PV array voltage if the proposed system is connected to an inverter.
Therefore, this paper proposes a PV system based on the CICO buck converter along with a control strategy and a design procedure for PV systems, where the array voltage is greater than the dc bus voltage and the power is delivered to the grid or ac loads. The adopted converter does not require an output capacitance, and the input capacitance is significantly lower than the one required by a PV system based on a buck converter; hence, it can be implemented with non-electrolytic capacitors. Furthermore, the proposed control strategy is inspired by the one proposed in [45], and it is formed by a P&O algorithm that generates the PV array voltage reference for a sliding-mode controller (SMC) that guarantees the system stability for any operating point and rejects the 100 or 120 Hz oscillations produced by the inverter. Finally, the paper also includes a design procedure of the proposed converter’s inductors and capacitors and SMC’s parameters considering the maximum voltage and current ripples, as well as the system’s stability. The proposed system is evaluated for a realistic study case using a professional software simulation of power electronics and commercial devices as a reference. In summary, the main contributions of the paper are (1) a PV system based on the CIOC buck with an MPPT technique formed by a P&O algorithm that generates the reference of the PV array voltage and an SMC that tracks such a reference, (2) a detailed design procedure of the SMC’s parameters to guarantee the system’s stability for any operation, and (3) a design procedure of the converter’s storage elements to guarantee the maximum ripples defined by the designer.
The rest of the paper is organized as follows: Section 2 describes the CIOC buck converter, Section 3 introduces the model of the proposed PV system that uses the CIOC buck converter, Section 4 shows a comparison of a PV system based on buck converter and the proposed PV system, Section 5 presents the proposed SMC controller for MPPT, Section 6 contains the design procedure along with an application example, Section 7 introduces the circuital implementation and the simulation results, and finally, Section 8 presents the conclusions.

2. Continuous Input/Output Buck Converter for PV Applications

The proposed PV system based on the continuous input/output current (CIOC) buck converter is depicted in Figure 1. Such a PV system description includes the controller required to mitigate environmental and load perturbations, which is discussed in Section 5; the algorithm to track the maximum power point (MPPT), which in this case is the perturb and observe algorithm (P&O); and a filter needed to ensure global stability, which is discussed in Section 5.6. The load of the PV system is modeled using the voltage source v o , which represents the input of an inverter. The PV source is connected at the CIOC buck converter, where the main advantage of the CIOC converter concerns the continuous input current requested to the PV source, which reduces the input capacitor C p v in comparison with the classical buck converter; such an advantage is demonstrated in Section 4. In addition, the CIOC converter also provides a continuous current i o to the load (inverter), thus providing a high power quality in comparison with converters with discontinuous output current.
The CIOC buck converter is formed by a MOSFET and a diode, where the control signal u provided by the controller defines the state of those semiconductors. Moreover, the converter has two inductors ( L 1 and L 2 ), an internal capacitor ( C i ) and an input capacitor to regulate the PV source ( C p v ). The P&O algorithm measures the PV current and voltage ( i p v and v p v ) to track the optimal PV voltage v p o , as it is described in [46,47]. Then, such an optimal value is processed by the stability filter to generate the controller reference v r .
The following subsections describe the behavior of the PV system, based on the CIOC buck converter, in the two operation topologies generated by u = 1 and u = 0 , i.e., when the MOSFET is closed and open, respectively.

2.1. Topology 1 ( u = 1 )

The fist topology occurs when the control signal is u = 1 , which closes the MOSFET and opens the diode, which results in the equivalent circuit reported in Figure 2. The topology shows that the input current of the CIOC buck converter is equal to the current of the L 1 inductor, named i 1 ; hence, it is a continuous input current. In this topology, the internal capacitor C i is discharged by the current of L 2 ( i 2 ), thus providing additional power to the load. Finally, the topology shows that the internal capacitor C i provides a path connecting L 1 and L 2 , while the MOSFET (closed) provides a path connecting both inductors with the load. Therefore, the output current i o is equal to the sum of both L 1 and L 2 currents, which enables the CIOC buck converter to provide a continuous current to the load.
The differential equations modeling this topology are
d i 1 d t = v p v v o L 1
d i 2 d t = v i v o L 2
d v i d t = i 2 C i
d v p v d t = i p v i 1 C p v
In addition, the MOSFET current ( i M ) and diode voltage ( v D ) are
i M = i o = i 1 + i 2
v D = v i

2.2. Topology 0 ( u = 0 )

The second topology occurs when the control signal is u = 0 , which opens the MOSFET and closes the diode. The equivalent circuit is reported in Figure 3, which also shows that the input current of the CIOC buck converter is i 1 . In this topology, the internal capacitor C i is charged by the current of L 1 ( i 1 ), and it also provides the path connecting L 1 and L 2 , while the diode (closed) provides a path connecting both inductors with the load. Therefore, as in the previous topology, the output current i o is continuous and equal to the sum of both i 1 and i 2 .
The differential equations modeling this topology are
d i 1 d t = v p v v o v i L 1
d i 2 d t = v o L 2
d v i d t = i 1 C i
d v p v d t = i p v i 1 C p v
In addition, the MOSFET voltage ( v M ) and diode current ( i D ) are
v M = v i
i D = i o = i 1 + i 2

2.3. Waveforms Analysis

Figure 4 shows the waveforms of the different variables in the PV system based on the CIOC buck converter, where T is the switching period. These waveforms are constructed using the equations previously obtained in Section 2.1 and Section 2.2 for the two topologies occurring during the PV system operation. The waveform of the L 1 current ( i 1 ), which is the input current of the converter, confirms that the CIOC buck converter requests continuous current to the input capacitor. This is observed in the waveform of input capacitor current ( i C p v ), which is also continuous, and it is centered in zero due to the charge balance principle [29]. This characteristic ensures a smaller input capacitance requirement in comparison with converters imposing discontinuous capacitor currents, such as the classical buck structure.
The waveform of the PV voltage is obtained by integrating i C p v , which results in a second-order ripple, as reported in [29], due to the triangular shape of i C p v . In contrast, the waveform of the voltage v i at the intermediate capacitor is triangular, this is because such a capacitor forms a first-order filter [29]. The current in L 2 ( i 2 ) is also triangular, and it is in phase with i 2 . Therefore, the output current i o = i 1 + i 2 is continuous, and it exhibits a ripple equal to the sum of both current ripples.
Finally, in topology 1 ( u = 1 ), the output current is transferred to the load using the MOSFET, thus the MOSFET current ( i M ) is a discontinuous waveform with a peak current equal to i o . Similarly, the output current is transferred to the load using the diode during topology 0 ( u = 0 ); thus, the diode current ( i D ) is a complementary discontinuous waveform with the same peak current. Finally, the MOSFET voltage ( v M ) is equal to v i when u = 0 , while the diode voltage ( v D ) is equal to v i when u = 1 .
In conclusion, both the input and output currents of the CIOC buck converter are continuous, which reduces the input capacitor of the PV system and provides a better power quality to the load in comparison with converters imposing discontinuous currents. This conclusion is further supported by a detailed comparison with a PV system based on a classical buck converter, which is given in Section 4.

3. Mathematical Model of the PV System Based on the CIOC Buck Converter

The elements and controller design of the PV system based on the CIOC buck converter require mathematical models with different characteristics. The first modeling approach, named switched model, is based on the differential equations obtained in Section 2.1 and Section 2.2.
The switched differential equations modeling the inductor currents and capacitors voltages include the binary control signal u, which defines the effective equations for each topology:
d i 1 d t = v p v v o v i · 1 u L 1
d i 2 d t = v i · u v o L 2
d v i d t = i 1 · 1 u i 2 · u C i
d v p v d t = i p v i 1 C p v
Similarly, the current and voltages of both the MOSFET and diode are
i M = i 1 + i 2 · u
v M = v i · 1 u
i D = i 1 + i 2 · 1 u
v D = v i · u
The previous model is useful to design nonlinear controllers with binary control signals, which is the case of the sliding-mode controller (SMC) designed in Section 5, to ensure a stable operation of the PV system.
A second modeling approach, named averaged model, is used to design controllers with continuous control signals, such as the duty cycle d. Moreover, the averaged model is used to analyze the closed-loop dynamics of dc/dc converters, which is the case of the closed-loop analysis performed in Section 5.4. The averaged model is obtained by averaging the switched differential equations within the switching period T, this taking into account that the duty cycle is the average value of the MOSFET control signal u, as reported in Equation (21).
d = 1 T · 0 T u d t
Then, applying the same averaging technique for Equations (13) to (16) results in the following averaged differential equations:
d i 1 d t = v p v v o v i · 1 d L 1
d i 2 d t = v i · d v o L 2
d v i d t = i 1 · 1 d i 2 · d C i
d v p v d t = i p v i 1 C p v
Applying the same averaging technique for (17) to (20) results in the averaged equations for the currents and voltages of both the MOSFET and diode:
i M = i 1 + i 2 · d
v M = v i · 1 d
i D = i 1 + i 2 · 1 d
v D = v i · d
The third modeling approach consists in obtaining the steady-state values of the inductors currents, capacitors voltages, and duty cycle. This approach is useful to design the passive elements of the PV system and to select the MOSFET and diode according the current and voltage stresses. The steady-state equations are calculated by considering the averaged differential equations equal to zero:
v i = v p v
i 2 = i 1 · 1 d d
i p v = i 1
v p v = v o d
d = v o v p v = i 1 i 1 + i 2
It is important to note that the duty cycle (34) is the same one required by the classical buck converter; thus, the CIOC buck converter can be used to replace classical buck converters in PV systems.
In addition, the ripple values are also needed to design both the converter and SMC. These ripples are obtained from the differential equations for each topology given in Section 2.1 and Section 2.2. For example, the current ripple in i 1 is obtained in topology 1 from (1), as given in (35). Similarly, the current ripple in i 2 is obtained in the same topology as given in (36). Taking into account that i o = i 1 + i 2 , and that i 1 and i 2 are in phase, the ripple in the output current is calculated as given in (37).
δ i 1 = v p v · d · 1 d · T 2 · L 1
δ i 2 = v p v · d · 1 d · T 2 · L 2
δ i o = v p v · d · 1 d · T 2 · 1 L 1 + 1 L 2
The voltage ripple of C i is also calculated in topology 0 from (9), as given in (38). Instead, the ripple at the PV voltage is calculated using the second-order filter procedure given in [29], where the current ripple in i 1 is integrated and the charge balance in C p v is applied, thus obtaining the expression given in (39).
δ v i = i p v · 1 d · T 2 · C i
δ v p v = v p v · d · 1 d · T 2 16 · C p v · L 1

4. Comparison with a PV System Based on the Classical Buck Converter

Taking into account that the proposed PV system based on the CIOC buck converter has the same duty cycle as the classical buck converter, it is a suitable candidate to reduce the input capacitor of step-down PV systems. However, the CIOC buck converter has two inductors and two capacitors; thus, it is required to compare the inductive and capacitive requirements of both buck converters to provide a selection criterion.
The scheme of a PV system based on the classical buck converter is reported in Figure 5, where the PV source is interfaced using a capacitor which must filter the discontinuous input current generated by the MOSFET. Moreover, in the classical buck converter, the inductor current is the output current provided to the load.
The mathematical model of this PV system based on the classical buck converter is given in (40) and (41), where u is the control signal of the MOSFET, i b is the current in the buck inductor L b , and C p v b is the input capacitance. The current and voltage of the MOSFET are i M b and v M b , respectively, while the current and voltage of the diode are i D b and v D b , respectively.
d i b d t = v p v · u v o L b
d v p v d t = i p v i b · u C p v b
i M b = i b · u
v M b = v p v · 1 u
i D b = i b · 1 u
v D b = v p v · u
Performing the averaging procedure and making the differential equations equal to zero leads to the steady-state relations given in (46) and (47). Moreover, the inductor current ripple δ i b and PV voltage ripple δ v p v are given in (48) and (49), respectively.
d = v o v p v
i b = i p v d
δ i b = v p v · d · 1 d · T 2 · L b
δ v p v = i p v · 1 d · T 2 · C p v b
In order to provide a fair comparison between both CIOC and classical buck converters, both converters must be designed to provide the same voltage ripple to the PV source and the same current ripple to the load (inverter). Considering both inductors of the CIOC converter are equal, thus L 1 = L 2 , the output current ripple of the PV system is given in (50). The output ripple current in a PV system based on the classical buck converter was previously reported in (48); to obtain the same ripple magnitude in both PV systems, the classical buck solution must have half the inductance of the CIOC converter, hence L b = L 1 / 2 .
δ i o = v p v · d · 1 d · T L 1
However, the inductor of the classical buck converter must support a higher current in comparison with the inductors of the CIOC converter, which is concluded from (31), (32), and (47). Therefore, a better measurement to compare the inductive requirements of both converters is the energy stored in the inductors: for example, for a duty cycle d = 0.5 , the inductor of the classical buck converter stores E L , b = 1 2 · L b · i b 2 = 2 · L b · i p v 2 , since i b = 2 · i p v , as given in (47). Taking into account that L b = L 1 / 2 to ensure the same output current ripple in both converters, the energy stored in the classical buck inductor becomes E L , b = L 1 · i p v 2 . For the same duty cycle d = 0.5 , the energy stored in each of the CIOC inductors is E 1 = 1 2 · L 1 · i p v 2 , since i 1 = i p v , as given in (32). Therefore, the total energy stored in both CIOC inductors is E L , C I O C = L 1 · i p v 2 , which is the same energy stored in the single inductor of the classical buck converter. In conclusion, both converters have the same inductive requirements (in terms of stored energy) to provide the same output current ripple.
Concerning the PV voltage ripple, making equal the ripple Equations (39) and (49) with L b = L 1 / 2 leads to the relation needed between the input capacitor of the classical buck converter C p v b and the input capacitor of the CIOC buck converter C p v to obtain the same ripple:
C p v b C p v = 8 · i p v · L 1 v o · T
To illustrate the previous relation, an inductor of L 1 = 38 μ H is considered. Moreover, an output voltage for both PV system v o = 24 V and a switching frequency equal to 93 kHz ( T = 10.8 μ s) are assumed for the comparison. Under those conditions, the classical buck converter requires an input capacitor almost 12 times higher than the CIOC converter. However, in this case, both capacitors have the same voltage, thus the energy stored in the input capacitor of the classical buck converter is 12 times higher than the energy stored in the input capacitor of the CIOC converter.
An additional comparison between both converters is provided in Figure 6, where variations on the inductance, switching period, and output voltage are considered. These comparisons show that reducing the inductance also reduces the difference in the capacitance requirements, but at the expense of increasing the output current ripple. Similarly, increasing the switching period also reduces the difference in the capacitance requirements and increases the output current ripple. Finally, increasing the output voltage reduces the difference in the capacitance requirements, but the output voltage is defined by the load, thus it is not a freedom degree. In any case, it is important to note that in small output voltages the proposed PV system requires a much smaller input capacitor.
In conclusion, for the same PV voltage ripple condition, the CIOC PV system has a much lower capacitance requirement in comparison with a PV system based on the classical buck converter. It is noted that the CIOC PV system requires an additional intermediate capacitor C i , but such a capacitor is smaller than the main capacitor C p v because the C i current is defined by the inductor currents; this condition is illustrated in Section 7.

5. Sliding-Mode Controller for Maximum Power Point Tracking

The correct operation of the PV system requires the tracking of the optimal operating condition for a particular irradiance and temperature reaching the PV source. Several algorithms have been designed to track the maximum power point (MPP), in which the PV source produces the maximum power [48]. The proposed PV system based on the CIOC converter can operate with any of these maximum power point tracking (MPPT) algorithms; however, the most widely adopted MPPT solution is the perturb and observe (P&O) algorithm, which is described in [49]. Therefore, this section considers the P&O algorithm parameters to design the controller for the CIOC converter, but any other MPPT algorithm can be used.
The P&O algorithm is a hill-climbing optimization technique [46]; thus, the manipulated variable is perturbed to detect the direction in which the objective variable is optimized. In the case of PV systems, the manipulated variable to be optimized is the voltage at the PV source v p v , and the objective variable to be maximized is the PV power p p v . Therefore, the input voltage v p v of the CIOC converter must be regulated to follow the reference provided by the P & O algorithm, which in this paper is named v M P P T . This regulation is performed using a sliding-mode controller (SMC) to provide global stability.

5.1. Definition of the Switching Function and Transversality Analysis

The first step to design an SMC is to design the desired dynamic of the system, which is expressed as a switching function defining the system trajectory. Taking into account the control objective v p v = v r , where v r is the controller reference, the obvious switching function Ψ v is given in (52), where the desired surface is Ψ v = 0 .
Ψ v = v r v p v
However, the switching function is usable only if three conditions are fulfilled, as demonstrated in [50,51]:
  • Transversality condition, which evaluates the presence of the control variable into the switching function derivative. This condition is mandatory, otherwise the SMC will not be able to modify the trajectory of the CIOC converter. In this case, the control signal u defines the activation/deactivation of the MOSFET and diode. The transversality condition is formalized as follows:
    d d u d Ψ d t 0
  • Reachability conditions, which evaluate the ability to reach the desired operating condition from any starting point. This condition is mandatory to compensate any perturbation of the system. In this case, the perturbations are introduced by changes on the solar irradiance, or by changes in the load voltage and/or impedance.
  • Equivalent control condition, which evaluates that the average value of the control signal u is always trapped inside the control variable limits; i.e., evaluating the saturation of the average value. In dc/dc converters, the average value of the control signal corresponds to the duty cycle (21), hence the equivalent control condition evaluates that the duty cycle is never saturated. This is important since a dc/dc converter with saturated duty cycle operates in open loop, thus nullifying the control action.
In addition to the previous three conditions, the closed-loop dynamics of the SMC must be analyzed to ensure that the desired performance is achieved.
Performing the transversality analysis to the switching function Ψ v requires the calculation of that function derivative, which is given in (54). Then, applying the transversality test defined in (53) results in d d u d Ψ d t = 0 , since u is not present in Equation (54); thus, the switching function (52) does not fulfill the transversality condition, and it is not suitable to design an SMC.
d Ψ v d t = d v r d t i p v i 1 C p v
The next option is to include the derivative of the objective variable into the switching function, which could enable to modify the system trajectory. In this case, the derivative of the PV voltage can be obtained from measuring the input capacitor current, since i C p v = C p v · d v p v d t . In addition, the new sliding function will include the integral of the voltage error to ensure a correct tracking of the reference. Finally, the components of the new switching function Ψ are weighted using the constants k p , k i , and k c as follows:
Ψ = k p · v r v p v + k i · v r v p v d t + k c · i C p v
The derivative of this new switching function is given in (56), and applying the transversality test (53) leads to expression (57), which confirms that the transversality condition is fulfilled, because k c 0 is needed to include the input current measurement.
d Ψ d t = k p · d v r d t k v · i p v i 1 C p v + k i · v r v p v + k c · d i p v d t k c · v p v v o v i · 1 u L 1
d d u d Ψ d t = k c · v i L 1 0
In conclusion, the switching function (55) is a suitable option to design the SMC for the PV system based on the CIOC converter, where the surface is Ψ = 0 . However, the sign of the transversality value (57) is needed to perform the analysis of the reachability conditions. In Section 5.4 it is demonstrated that parameter k c must be negative ( k c < 0 ) to ensure stable closed-loop dynamics; thus, the sign of the transversality value (57) is positive.

5.2. Reachability Conditions

The reachability conditions evaluate the system capability to reach the surface:
  • When the system is operating under the surface ( Ψ < 0 ), the switching function derivative must be positive d Ψ d t > 0 to reach the desired condition Ψ = 0 . Taking into account that the transversality value (57) is positive, this means that a positive change on the control signal u (from 0 to 1) produces a positive change on d Ψ d t ; thus, this reachability condition is formalized as follows:
    lim ψ 0 d Ψ d t u = 1 > 0
  • When the system is operating above the surface ( Ψ > 0 ), the switching function derivative must be negative d Ψ d t < 0 to reach the desired condition Ψ = 0 . Since (57) is positive, this means that a negative change on the control signal u (from 1 to 0) produces a negative change in d Ψ d t ; thus, this second reachability condition is formalized as follows:
    lim ψ 0 + d Ψ d t u = 0 < 0
Evaluating the previous theoretical reachability conditions, using the switching function derivative given in (56) leads to the following dynamic restrictions for the reference derivative d v r d t , which must be fulfilled to ensure both reachability conditions (58) and (59):
k c k p · d i p v d t v p v v o L 1 k i k p · Δ v p o < d v r d t < k c k p · d i p v d t + v o L 1 k i k p · Δ v p o
The previous expression takes into account that the average value of v p v is equal to the average value of v i , as demonstrated in (30); in addition, it also takes into account that the average value of i p v is equal to the average value of i 1 , as demonstrated in (32). Moreover, since the reference v r is defined by a P&O algorithm, each perturbation in the reference has a magnitude equal to the P&O perturbation parameter Δ v p o , thus Δ v p o = v r v p v .
In conclusion, the dynamic behavior of the reference signal must be constrained as given in (60) to ensure that both reachability conditions are always fulfilled.

5.3. Equivalent Control Condition

The analysis of the equivalent control condition requires to calculate the average value of the control signal u, which is equal to the duty cycle, as given in (21). This analysis, which evaluates the saturation of the duty cycle, is performed when the system is operating into the sliding surface, thus Ψ = 0 and d Ψ d t = 0 . Then, using the expression of d Ψ d t given in (56) at the condition d Ψ d t = 0 leads to the following expression for the duty cycle:
d = L 1 k c · v i · k p · d v r d t + k c · d i p v d t + k c v o L 1 + k i · Δ v p o
Then, the equivalent control condition is 0 < d < 1 , and introducing the d value obtained in (61) leads to the same dynamic restriction obtained from the reachability conditions, which was reported in (60). Therefore, fulfilling the dynamic restriction given in (60) ensures that both reachability and equivalent control conditions are true, hence ensuring the global stability of the PV system based on the CIOC converter.

5.4. Closed-Loop Dynamics and Controller Parameters

The closed-loop operation of the PV system is analyzed considering a correct operation of the SMC; hence, it is assumed that the switching function (55) is inside the sliding surface Ψ = 0 . Therefore, the SMC imposes the following current-to-input capacitor C p v :
i C p v = k p · v r v p v + k i · v r v p v d t k c
Considering that i C p v = C p v · d v p v d t and applying the Laplace transformation leads to the following expression for the closed-loop dynamics of the PV voltage:
v p v v r = k p · s + k i k c · C p v · s 2 + k p · s + k i
From the Routh–Hurwitz stability criterion [52], it is known that all coefficients of the previous transfer function must be positive to ensure that both equivalent poles are at the left-hand plane (LHP) of the Laplace space, thus ensuring a stable closed-loop operation, as anticipated in Section 5.1. Therefore, the following conditions must be fulfilled: k c < 0 , k p > 0 and k i > 0 . The equivalent poles of the closed-loop transfer function are given in (64), where the pole values depend on the relations k p / k c and k i / k c . Hence, the k c value given in (65) is defined to ensure the negative sign.
s = 1 2 · k p / k c C p v ± k p / k c C p v 2 + 4 · k i / k c C p v
k c = 1
The other two parameters of the SMC ( k p and k i ) must be designed to ensure stable operation of the P&O algorithm. In [46,47], it was demonstrated that the settling time t s of the PV voltage v p v must be shorter than the perturbation period T p o of the P&O algorithm, otherwise the system becomes unstable. This stability condition ( t s < T p o ) is simple to explain: the P&O algorithm perturbs the PV voltage v p v to observe the direction in which the PV power p p v is increased; therefore, the PV voltage must be stable to perform a correct observation of the PV power. Instead, if the PV voltage is not stable, the observed PV power will be a transient value; thus, it could produce a wrong decision, making the operation of the P&O algorithm unstable. In conclusion, the SMC must ensure that the settling time t s of the PV voltage is correct, i.e., t s < T p o .
The reference signal provided by the P&O algorithm is a series of perturbations with magnitude Δ v p o each T a seconds, which is explained in detail in [46,49]; hence, the reference signal v r of the SMC can be modeled as a step waveform with magnitude Δ v p o , which is represented in the Laplace domain as follows:
v r = Δ v p o s
The time response of the PV voltage to the perturbation of the P&O algorithm is obtained by calculating the inverse Laplace transformation of the multiplication of transfer function (63) by reference signal (66). For this process, it is imposed to (63) a damping factor ρ = 1 , which provides a PV voltage waveform with low overshoots. Contrasting the denominator coefficients of (63) with the canonical form s 2 + 2 · ρ ω n · s + ω n 2 , for a defined ρ = 1 condition, leads to the k i value given in (67). Using such a k i value, the time response of the PV voltage is given in (68).
k i = k p 2 4 · C p v
v p v = Δ v p o · 1 e k p · t / 2 · C p v + k p · t 2 · C p v · e k p · t / 2 · C p v
The settling time t s is calculated when the PV voltage enters an acceptable band, which in this case is selected as ϵ = 1 % . Then, solving Equation (68) for v p v = 1 + ϵ results in the settling time expression given in (69), which is used to obtain the value of k p , given in (70), which is needed to ensure the desired settling time t s value.
t s = 2 · C p v · 1 W ϵ · e 1 k p
k p = 2 · C p v t s · 1 W ϵ · e 1

5.5. Hysteresis Band and Switching Frequency

The main inconvenience of the sliding-mode control technique is the high switching frequency imposed by the controller:
  • Stage 1: When Ψ < 0 , the first reachability condition (58) imposes the control action u = 1 , which forces the increment of the switching function Ψ towards Ψ = 0 . However, when Ψ reaches 0, there is no change in the control action, thus Ψ is further increased into Ψ > 0 (Stage 2).
  • Stage 2: When Ψ > 0 , the first reachability condition (59) imposes the control action u = 0 , which forces the decrement in the switching function Ψ towards Ψ = 0 . However, as in Stage 1, there is no change in the control action when when Ψ = 0 , thus Ψ is further decreased into Ψ > 0 (Stage 1).
  • In this theoretical operation, transition time between Stage 1 and Stage 2 is only limited by the time required by the MOSFETs and diodes to open and close.
However, MOSFET and diodes have transition times recommended for correct operation which are much longer than the maximum transition times physically achievable. Thus, operating those semiconductors at higher switching frequencies (than the recommended ones) reduces the semiconductors lifetime and significantly increase the power losses.
Therefore, the practical implementation of SMC considers the introduction of hysteresis bands around the sliding surface, which makes it possible to limit the switching frequency to be in agreement with the requirements of the semiconductors. In this way, the practical sliding surface is formalized, as given in (71), and the associated practical reachability conditions are formalized in expression (72).
H Ψ + H
lim ψ H d Ψ d t u = 1 > 0 lim ψ + H + d Ψ d t u = 0 < 0
These practical reachability conditions impose the control law to be used in the SMC implementation, which is formalized as follows:
if Ψ < H set u = 1 if Ψ > + H set u = 0
The maximum switching frequency of this practical implementation is defined by the time that it takes the switching function to travel between the hysteresis limits, i.e., from H to + H , which corresponds to the ripple of the switching function. Therefore, the ripple of the switching function is equal to the hysteresis limit H.
The analysis of the ripple H is performed using the ripple of the variables forming the switching function Ψ = k p · v r v p v + k i · v r v p v d t + k c · i C p v , which are the PV voltage v p v and input capacitor current i C p v . Taking into account that the correct operation of the SMC ensures that v p v = v r , the ripple of the term k p · v r v p v is equal to the ripple of the PV voltage multiplied by k p , i.e., k p · δ v p v ( t ) . Moreover, the charge balance principle [29] ensures that the average current in the input capacitor is equal to zero, hence the ripple of the term k c · i C p v is equal to the ripple in the inductor current multiplied by k c , i.e., k c · δ i 1 ( t ) . Finally, since the ripple at the PV voltage is symmetrical with respect to the average value, the ripple of the term k i · v r v p v d t is equal to zero. The previous analyses are synthesized as follows:
H = k p · δ v p v ( t ) + k c · δ i 1 ( t )
From Figure 4, it is observed that the ripple waveforms of i C p v and v p v are phase-shifted in 90°, hence the maximum value of the voltage ripple in v p v occurs when the ripple of i C p v is zero, and the maximum value of the ripple in i C p v occurs when the ripple in v p v is zero ( v p v equal to the average value). In conclusion, the ripple H of Ψ is equal to the maximum component as follows:
H = max k p · δ v p v , k c · δ i 1
The previous equation is combined with the expressions for δ i 1 , given in (35), and δ v p v , given in (39), to calculate the H value that ensures the desired maximum switching frequency F = 1 / T , which is defined in agreement with the characteristics of the MOSFET and diode selected for the implementation.

5.6. SMC Implementation

The controller (SMC) implementation is divided in two parts: first, the switching function Ψ (55) must be calculated; second, the control law (73) must be implemented. Figure 7 shows the block diagram for the SMC implementation, where the calculation of the switching function is based on linear elements such as adders, static gains, and an integrator. These linear blocks are constructed using operational amplifiers, as it is shown in the circuital implementation described aftewards. The control law is implemented using two comparators for the Ψ < H and Ψ > + H operations, as well as an SR flip-flop for the set u = 1 and set u = 0 operations, where the last one corresponds to the reset condition of the flip-flop.
The block diagram also shows the connection of the P&O algorithm, where a filter is inserted between the MPPT block and the SMC. That filter is needed to ensure that the dynamic restriction (60), imposed by the reachability conditions, is fulfilled. The filter is implemented using the first-order transfer function given in (76), where the time constant of the filter τ f must be calculated in agreement with the restrictions for d v r d t .
v r v p o = 1 τ f · s + 1
The time response of the filter is (77), where the maximum derivative occurs at t = 0 . Therefore, the time constant of the filter needed to ensure a desired maximum derivative ( R a t e ) is given in (78).
v r = Δ v p o · 1 e t / τ f
τ f = Δ v p o R a t e
Finally, since this first-order filter imposes the same derivative limitation to both positive and negative changes, the R a t e value must be selected to be equal to the limit of (60) with lower magnitude. This will ensure that both dynamic restrictions are always fulfilled, thus ensuring the SMC stability.

6. Design Procedure and Application Example

This section synthesizes the design procedure of the proposed PV solution using an application example based on the SP500M6-96 PV panel [53] and the commercial inverter COTEK SP-700-124 [8], which is a 700 W inverter with a 24 V input and a single-phase ac output of 100, 110, 115, or 120 V and 50 or 60 Hz. However, both grid-connected inverters and off-grid inverters produce the inverter input oscillations at twice the output frequency. The characteristics of both the panel and pure sine wave inverter are given in Table 1. Although a pure sine wave off-grid inverter is used in the application example, a grid-connected inverter also could be used with the CIOC. A commercial example of a grid-connected inverter is presented in [11], where the inverter input voltage is between 19V and 33V, and the PV array maximum voltage is 250 V; moreover, this particular inverter also has an input for a battery bank with a nominal voltage of 24 V and can be connected to a grid of 230 V of 50 Hz.
The electrical characteristics of the SP500M6-96 PV panel are presented in Figure 8, where the current vs. voltage (I-V) and power vs. voltage (P-V) curves are observed for different irradiance (S) conditions. The figure also shows the trajectory of the maximum power points (MPP) of the PV panel, which corresponds to the trajectory where the PV system must be operated.
The first step is to design the inductors of the CIOC converter. Since the ripple δ i o of the output current (37) depends on the inverse of both inductors, these inductors are designed equally to share the ripple balanced between both i 1 and i 2 currents, thus L 1 = L 2 . In addition, the switching frequency is limited to 100 kHz to enable the use of standard (and cheap) power MOSFET and diodes, such as CSD19538Q3A [54] and V15P8 [55], respectively; hence, F 100 kHz. From the electrical characteristics reported in Figure 8, it is observed that the PV source provides 88.5 W at S = 200 W/m 2 , which imposes an output current of the CIOC converter (delivered to the load at 24 V) equal to 3.69 A; hence, the output current ripple at S = 200 W/m 2 must be δ i o 3.69 A to ensure continuous conduction mode (CCM) in the CIOC converter for S 200 W/m 2 . The design procedure could be performed for a lower irradiance limit, but the small current value, thus small current ripple, will introduce negligible harmonics that will not justify the increment in the inductances. Based on the previous values, Equation (75) is used to calculate the switching frequency for different inductance values for the irradiance range 200 W/m 2 S 1000 W/m 2 , where the data for S = 1000 W/m 2 corresponds to the MPP conditions reported in Table 1. Figure 9 depicts the results of Equation (75), where the value L 1 = L 2 = 38 μ H is selected to ensure the desired F 100 kHz condition, since the Earth irradiance is always lower than 1000 W/m 2 .
The next step is to design the input capacitor C p v in agreement with the requirements of the PV system. Taking into account that the P&O algorithm has an efficiency of 99% in tracking the MPP condition [49], the PV voltage ripple δ v p v must be lower than 1% of v m p p . Therefore, the peak-to-peak amplitude of the PV ripple is defined as the 20% of such a 1% limit, hence δ v p v = 50.87 mV. Then, using expression (39) and the previous value of L 1 leads to the input capacitance value 42.88 μ F, where the commercial value C p v = 47 μ F is selected. Concerning the internal capacitor ( C i ), it is designed to ensure a CCM operation for 80% of the converter conditions: taking into account that v i = v p v , as given in (30), and the peak-to-peak voltage ripple in C i is defined as 20% of v m p p , thus δ v i = 4.94 V. Then, using Equation (38), the internal capacitance 5.54 μ F is calculated, where the commercial value C i = 5.6 μ F is selected.
After the power stage parameters have been designed, the following step is to calculate the appropriate parameters for the controller (SMC). The first parameter was already designed in Equation (65), where k c = 1 . The other two parameters are calculated in agreement with the requirements of the P&O algorithm, which in turn is designed as described in [46,47]: the perturbation amplitude is designed as the 1% of the MPP voltage, thus Δ v p o = 500 mV; the perturbation period is designed as T p o = 500 μ s to provide a balance between processing requirements and tracking speed. Then, the settling time of the PV voltage is defined as 50% of the T p o parameter to ensure the P&O stability ( t s < T p o ), thus t s = 250 μ s. Using Equation (70) and both the C p v and t s values leads to the calculation of the parameter k p = 2.36[A/V]; the remaining SMC parameter is calculated using Equation (67) as k i = 29.5[kA/ (V · s)].
The last part of the design concerns the stability filter. This application is designed to ensure global stability even under fast perturbations on the irradiance conditions, which is considered as one sun per millisecond or d S d t = 1000 W/(m 2 · ms), which is translated into d i p v d t = 10.87 A/ms. This irradiance perturbation corresponds to a transition from full irradiance to complete shading (or vice-versa) in one millisecond, thus covering any practical perturbation. Using Equation (60) with the system parameters and the PV current perturbation leads to the maximum reference derivative R a t e = 0.257V/ μ s accepted to ensure the SMC stability. Finally, from Equation (78), the filter time constant is calculated as τ f = 1.95 μ s, which is less than 20% of a single switching period of the CIOC converter; thus, it has a negligible effect on the settling time. The last parameter of the PV system is the hysteresis width, which is calculated from Equation (75) as H = 1.67 A. Table 2 summarizes the parameters of this application example used to illustrate the proposed design process.
Finally, Figure 10 synthesizes the proposed design process for the PV system based on the CIOC converter: the first part describes the converter design based on both the source and load characteristics, the second part concerns the MPPT algorithm design, and the third part is devoted to the controller (SMC) design to guarantee global stability.

7. Circuital Implementation and Simulations

The circuital implementation of the complete PV system based on the CIOC converter and the SMC was performed in the power electronics simulator PSIM [56]. Such a commercial power electronics simulator takes into account the nonlinear and switched behavior of both the MOSFET and diode, provides realistic models for operational amplifiers and flip-flops, and also enables the emulation of microcontrollers using ANSI C code. Figure 11 shows the proposed PV system implemented in the schematic editor of PSIM.
The PV source is implemented using the ideal single-diode model reported in [25], where the model equation is i p v = i s c A · e B · v p v , and the parameters values for the SP500M6-96 PV panel are A = 642.9 nA and B = 0.2823 V 1 . The load is represented by a series connection of a DC source, which represents the average voltage imposed by the inverter, as well as a sinusoidal component at double the output frequency, which models the voltage oscillation at double of the output frequency caused by the ac connection of a single-phase inverter. The amplitude of such a low-frequency oscillation was discussed in [57], and the mathematical analyses provided in such a work lead to expression (79), where f g is the output frequency, C a c is the capacitor at the input of the inverter, and v o is the average voltage imposed by the inverter at the output of the CIOC converter. For the PV source and inverter parameters given in Table 1, a capacitor C a c = 5.7 mF will produce a voltage oscillation between 21.6 V and 26.4 V, which fits the input voltage range required by the COTEK SP-700-224 inverter to properly operate (see Table 1). Thus, the maximum amplitude of the load voltage perturbations is equal to 20% of the nominal value ( 24 V ).
Δ v o = p p v 4 · π · f g · C a c · v o
The parameters of the CIOC converter are the same ones designed in the previous section and reported in Table 2. However, the circuit of Figure 11 includes one voltage sensor to measure the PV voltage, which is needed to process both the P&O algorithm and the calculation of the switching function. In addition, two current sensors are also added: the first one measures the PV current, which is needed to process the P&O algorithm, and the second one measures the input capacitor current, which is needed for the calculation of the switching function.
The circuital implementation of the control stage is divided into four parts. First, the P&O algorithm is coded using the algorithm reported in [49], which is implemented with a C-block inside PSIM to emulate a microcontroller. The second part implements the stability filter using an operational amplifier, and the third part performs the calculation of the switching function with circuits based on operational amplifiers. Finally, the fourth part implements the control law using two comparators and an SR flip-flop.
The first simulation compares the operation of the designed PV system based on the CIOC converter with a PV system based on the classical buck converter. The design of the buck PV system was discussed in Section 4, where the buck inductor and capacitor needed to match the PV voltage, and the output current ripples of the CIOC solution are L b = 19 μ H and C p v b = 550 μ F, respectively. Figure 12 shows the comparison of the PV systems based on both the CIOC and the classical buck converters, where both solutions introduce the same PV voltage ripple (top traces), thus fulfilling the requirements of the MPPT algorithm. Moreover, both solutions also introduce the same current ripple into the load (second traces of Figure 12), hence providing the same power quality. However, the PV system based on the classical buck converter requires an input capacitor 11.7 times bigger than the capacitive requirement of the CIOC converter (the buck converter stores 11.7 times more energy), which introduces reliability problems, since bigger capacitances have higher failure rates [58]. This high capacitive condition of the buck converter is caused by the discontinuous input (MOSFET) current that must be filtered in the input capacitor: the current at the input capacitor i C p v = i p v i M is discontinuous for the buck-based PV system, which in this example corresponds to an RMS current equal to 10.51 A; instead, the CIOC input capacitor has a continuous current ( i C p v = i p v i 1 ), thus the input capacitor must filter much lower current harmonics (RMS current equal to 1.01 A). Such a condition is confirmed in the third traces of Figure 12. Finally, despite the inductors of the CIOC converter having higher inductance, the current processed by these inductors is much lower, and the total energy stored in the inductors of the CIOC converter is the same energy stored in the single inductor of the buck converter; thus, the inductive requirements are equivalent, as it was demonstrated in Section 4.
Figure 12 also confirms the correct L 1 , L 2 , and C p v design: the PV voltage ripple is 50.65 mV, which is lower than the design limit calculated in Section 6 ( δ v p v = 50.87 mV). Similarly, Figure 12 shows that the output current ripple is equal to 3.5 A, which is in agreement with the value predicted with Equation (50). In addition, the switching frequency in this steady-state condition is F = 92.6 kHz, thus fulfilling the design criterion imposed in Section 6 ( F 100 kHz). It must be noted that the switching frequency will never be higher than 100 kHz, hence the maximum switching losses can be estimated with such a limit frequency. In fact, the simulation reports that the switching frequency of both the CIOC and buck PV systems is the same, which is confirmed by the waveforms of the MOSFET and diode currents ( i M and i D , respectively). Since both PV systems operate with the same switching frequency under steady-state conditions ( 92.6 kHz), the waveforms of the MOSFETs’ and diodes’ currents are the same, just displaced in time. This is further verified by calculating the RMS currents of the MOSFETs and diodes in both PV systems, obtaining the same value. Finally, taking into account that the MOSFET and diode waveforms are equal in both PV systems, the switching losses are thus also the same; for the example in Figure 12, these losses are, on average, 1.5 W per period considering a MOSFET on resistor equal to 4.8 m Ω , which corresponds to 0.3% of the generated power ( 500 W).
Figure 12 also shows the voltage and current on the intermediate capacitor C i of the CIOC PV system, named v i and i C i , respectively. Such a C i capacitor was designed to ensure a small peak-to-peak ripple in v i equal to 20% of the steady-state value, resulting in the small capacitor of 5.6 μ F reported in Table 2. Therefore, the capacitive storage requirements of the CIOC are not significantly increased by C i , since, in this example, C p v + C i = 52.6 μ F, which is more than 10 times smaller than the capacitive requirement of the buck-based PV system. Figure 12 shows that the current in C i is discontinuous, thus producing a large RMS current that must be supported by the capacitor; nevertheless, it is possible to find commercial film capacitors that support these RMS currents [59,60]. In case it is needed, increasing C i to acquire a non-electrolytic capacitor with higher RMS current capability will reduce the v i ripple, which will increase the power range in which the inductor currents are continuous, thus not affecting the system performance.
The second set of circuit simulations concerns the validation of the mathematical model proposed in Section 3, including the action of the SMC designed in Section 5. This validation is performed in both frequency and time domains. Figure 13 shows the frequency response of the closed-loop system generated by both the circuit simulation and the theoretical model previously reported in Equation (63). These results evidence the accurate representation provided by the mathematical model proposed in this paper, since both the amplitude and phase of the frequency response are correctly predicted. Moreover, such a Bode diagram also confirms the correct tracking of the reference v r , where no error or phase shift is observed up to 1 kHz; thus, the step-like reference of the P&O will be tracked with null steady-state error.
In addition, the frequency response of the closed-loop circuit to perturbations in the load voltage is investigated in Figure 14. Such a Bode diagram confirms the correct mitigation of perturbations occurring in the load voltage, where the maximum perturbation transference is 0.62% (at 3.4 kHz). In fact, the perturbation transference at 120 Hz is 0.04%; thus, the voltage oscillations caused by the ac output of the inverter (load) will be satisfactorily rejected, and these oscillations will not interfere with the MPPT process.
The previous frequency-based results are complemented with the evaluation of the time-domain performance reported in Figure 15. Such a circuit simulation confirms the correct tracking of the reference provided by the SMC. Moreover, the circuital implementation of the SMC imposes the same dynamic performance predicted by the transfer function (63), thus confirming the correct implementation of both the power stage and control system. This circuit simulation also confirms the fulfillment of both the reachability and equivalent control conditions: First, the switching function Ψ is always trapped inside the hysteresis band ( H to + H ), thus the reachability conditions are fulfilled; therefore, the local stability is confirmed. Second, the duty cycle is always trapped inside the physical limits (0 and 1), hence no duty cycle saturation occurs. Both conditions confirm the global stability of the PV system.
The simulation results of Figure 15 also confirm that the switching frequency is always below the design limit ( 100 kHz), where high switching frequencies are imposed to provide fast tracking of the reference, returning to steady switching frequencies after the transient ends. Finally, these results also confirms the accurate settling time imposed by the SMC, which fits the design value ( 250 μ s). Therefore, it is confirmed that the proposed PV system (power stage and SMC) fulfills the requirements to extract the maximum power from the PV source, even with load voltage perturbations.
An additional simulation was conducted to evaluate the controller performance under different perturbations at the load voltage. It is clear that 60 Hz inverters will produce perturbations at 120 Hz, and 50 Hz inverters will produce perturbations at 100 Hz; however, inverter failures or nonmodeled grid behaviors could produce additional perturbations to other frequencies. Therefore, the proposed PV system was tested under load voltage perturbations at multiple frequencies: 60 Hz, 120 Hz, 240 Hz, 500 Hz, 1 kHz, and 10 kHz. The results of such a test are presented in Figure 16, which confirm that the proposed SMC ensures the correct regulation of the PV voltage under different load voltage perturbations, thus not only the classical 120 Hz perturbation is compensated.
The previous conclusions are tested by implementing a circuit simulation of the complete PV system, including the P&O algorithm. Such simulation results are presented in both Figure 17 and Figure 18. This final test considers a changing irradiance condition with fast transients of 1000 W/(m 2 · s) (one sun per millisecond), starting a full irradiance ( 1000 W/m 2 ), then falling to a medium irradiance ( 600 W/m 2 ), and later falling to a low irradiance ( 200 W/m 2 ), recovering the full irradiance at the end of the profile. In addition, this test also considers a 20% voltage oscillation at 120 Hz perturbing the output of the CIOC converter, which emulates the connection with the commercial inverter COTEK SP-700-124 [8].
The performance of the PV voltage confirms the correct tracking of the reference provided by the P&O, which is evident since both v p v and v p o waveforms are superimposed. In addition, the PV voltage waveform also shows the three-point behavior typical of stable P&O algorithms operating at the optimal condition [46,49], which ensures that the P&O algorithm reaches the maximum power points (MPP). Moreover, the tracking of the new MPP is observed in each irradiance transition. The maximum power extraction and tracking are confirmed by the trajectories reported in Figure 18, which shows the paths taken to track each MPP. Finally, the successful operation of the PV system under load voltage oscillations is explained by the fact that these oscillations are modulated into the duty cycle, thus compensating them without saturating the duty cycle. Finally, Figure 17 also shows the dynamic simulation of the PV system based on the buck converter previously designed for comparison purposes (red waveforms). To provide a fair comparison, such a buck-based PV system was complemented with the same P&O algorithm [61] and with a sliding-mode controller designed as described in [28] to provide a similar settling time. The results confirm that both the proposed CIOC and buck-based solutions provide the same PV power, which is expected since the buck converter was designed to provide the same ripples as observed in Figure 12, thus resulting in the same duty cycle and similar voltage profiles but requiring much higher capacitive storage for the classical buck solution.
In conclusion, the circuit simulations performed in this section confirm the following conditions:
  • The CIOC converter was accurately designed to impose the desired current and voltage ripples to both the PV source and the load.
  • The CIOC converter provides the same ripples in comparison with the buck converter, requiring the same inductive energy storage but with much lower capacitive energy storage, thus improving reliability.
  • The SMC provides global stability to the PV system in any operating condition.
  • The proposed PV system, based on the CIOC converter and SMC, ensures the extraction of the maximum power, even with high perturbations in the load voltage, which is a common case when inverters are added to the PV system to form a small-power microinverter.

8. Conclusions

A low-voltage photovoltaic system that uses a continuous input/output current buck converter was proposed in this paper. The design of the converter and its control systems included the definition of the maximum switching frequency of 100 kHz and the calculation of two inductances, L 1 = L 2 = 38 uH, two capacitances, C p v = 47 uF and C i = 5.6 μ F, P&O algorithm parameters, Δ v p o = 500 mV and T p o = 500 ms, three controller parameters k c = 1 , k p = 2.36 A/V, and k i = 29.5 kA/(V· s), and H = 1.67 A, which ensures the maximum switching frequency.
The main advantage of the proposed PV system is to impose desired current and voltage ripples to both the photovoltaic source and the load with lower input capacitance requirements than the classical buck converter. For instance, in a simulated application with the SP500M6-96 PV panel and the commercial inverter COTEK SP-700-124; the imposed PV voltage ripple was 50.87 mV, with an input capacitor of 47 μ F and an internal capacitor of 5.6 μ F. Moreover, the implemented P&O algorithm extracted the maximum power of the photovoltaic source, in this case 95 W, 297 W, and 500 W for three different irradiance values, and the proposed sliding-mode controller provided global stability to the system in any operating condition and rejected output perturbations, including the double-frequency oscillations of 120 Hz produced by an intended microinverter. To simulate the application example, circuit simulations, developed with the power electronics simulator PSIM, confirmed that the designed CIOC converter imposed the desired current and voltage ripple to both the PV source and the load. Moreover, the CIOC converter provided the same ripple compared with the classical buck converter, requiring the same inductive energy storage but with lower capacitive energy storage.
Despite the advantages of the proposed low-voltage photovoltaic system, some requirements were faced to operate the system satisfactorily: a complex mathematical model of the converter to design the controllers, which also requires a complex theory to be analyzed and designed, and the design of the converter is more complex than the classical buck to achieve the requirements of the PV system.
For the implementation stage, two clear disadvantages will have to be covered: a greater number of elements than the classical buck converter, reducing the reliability of the system (one more inductor), and the design and implementation of the MOSFET driver in the high-voltage side of the system.

Author Contributions

Conceptualization, methodology, software, and writing (review and editing): C.A.R.-P., J.D.B.-R. and A.J.S.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Universidad Nacional de Colombia and Minciencias (Fondo nacional de financiamiento para ciencia, la tecnología y la innovación Francisco José de Caldas) under the project “Estrategia de transformación del sector energético Colombiano en el horizonte de 2030—Energetica 2030”—“Supervisión, control y protecciones de sistemas eléctricos con alta participación de fuentes no convencionales de energía renovable y fuentes de generación distribuida con alta incertidumbre en el escenario Colombia 2030” (Contrato No. FP44842-210-2018, Hermes: 38945).

Data Availability Statement

The data used in this study are reported in the paper figures and tables.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IEA PVPS. Snapshot of Global PV Markets 2022; Technical Report; International Energy Agency: Paris, France, 2022. [Google Scholar]
  2. IEA. Solar PV-More Efforts Needed; IEA: Paris, France, 2022. [Google Scholar]
  3. Dogga, R.; Pathak, M.K. Recent trends in solar PV inverter topologies. Sol. Energy 2019, 183, 57–73. [Google Scholar] [CrossRef]
  4. Sikder, P.S.; Pal, N.; Patro, K.S. A modeling of stand-alone solar photo-voltaic system for rural electrification purposes. Rev. Roum. Des Sci. Tech. Ser. Electrotech. Energetique 2019, 64, 241–246. [Google Scholar]
  5. Pirooz, A.; Noroozian, R.; Rodriguez, J. Investigation of model predictive control for converter-based stand-alone DC distribution networks fed by PV units. Int. Trans. Electr. Energy Syst. 2017, 27, e2396. [Google Scholar] [CrossRef]
  6. Huawei. Smart Energy Controller SUN2000-3-10KTL-M1 (High Current Version); Huawei: Shenzhen, China, 2021. [Google Scholar]
  7. Huawei. Smart String Energy Storage System; Huawei: Shenzhen, China, 2022. [Google Scholar]
  8. COTEK. 700W Pure Sine Wave Inverter SP-700 Series; COTEK: Taoyuan City, Taiwan, 2021. [Google Scholar]
  9. Schneider Electric. Conext MPPT 60 150 Solar Charge Controller; Schneider Electric: Rueil-Malmaison, France, 2022. [Google Scholar]
  10. Steca. Sine Wave Inverters Steca Solarix 5000-48, 2400-24, 1000-12; Steca: Memmingen, Germany, 2023. [Google Scholar]
  11. Victron Energy. EasySolar-II 3kVA MPPT 250/70 GX, 5kVA MPPT 250/100 GX; Victron Energy: Almere, The Netherlands, 2023. [Google Scholar]
  12. MUst. Pure Sine Wave High Frequency Solar Inverter (450V) - PH1800 PRO Series (3KW, 5.2KW); MUst: Hongkong, China, 2022. [Google Scholar]
  13. INVT. BD 3-6kW-RL1 Single-Phase Hybrid Solar Inverter; INVT: Shenzhen, China, 2023. [Google Scholar]
  14. Solar-Inverter, S.I. Deye Hybrid 3 Phase Solar Inverter 10KW 12KW 380V 400V 60HZ Home Inverters. 2022. [Google Scholar]
  15. Schneider Electric. SW Solar Hybrid Inverter 120/240V; Schneider Electric: Rueil-Malmaison, France, 2022. [Google Scholar]
  16. Steca. Sine Wave InvertersSteca Xtender; Steca: Memmingen, Germany, 2023. [Google Scholar]
  17. Victron Energy. Phoenix Inverters Smart 1600 VA–5000 VA; Victron Energy: Almere, The Netherlands, 2022. [Google Scholar]
  18. Duracell. Duracell 1000/2000/3000 Watt High Power Inverters; Duracell: Bethel, CT, USA, 2022. [Google Scholar]
  19. Baharudin, N.H.; Mansur, T.M.N.T.; Hamid, F.A.; Ali, R.; Misrun, M.I. Topologies of DC-DC converter in solar PV applications. Indones. J. Electr. Eng. Comput. Sci. 2017, 8, 368–374. [Google Scholar] [CrossRef]
  20. Dahbi, S.; Mazozi, I.; El Ouariachi, M.; Messaoudi, A.; Aziz, A. Implementation of a multi-control architecture in a photovoltaic/Grid/Electrolysis system for usual use and clean storage by hydrogen production. Int. J. Renew. Energy Res. 2017, 7, 1825–1835. [Google Scholar] [CrossRef]
  21. Toghani Holari, Y.; Taher, S.A.; Mehrasa, M. Distributed energy storage system-based nonlinear control strategy for hybrid microgrid power management included wind/PV units in grid-connected operation. Int. Trans. Electr. Energy Syst. 2020, 30, e12237. [Google Scholar] [CrossRef]
  22. Kabalci, E. Design and analysis of a hybrid renewable energy plant with solar and wind power. Energy Convers. Manag. 2013, 72, 51–59. [Google Scholar] [CrossRef]
  23. Tahmasbi Fard, M.; Tarafdar Hagh, M. Current Source Inverter Based Grid Connected Hybrid PV-Wind Power Generation Unit. Int. J. Electron. 2020, 107, 839–857. [Google Scholar] [CrossRef]
  24. Prathap, J.A.; Anandhi, T.S. Field programmable gate array-based new fusion control for photovoltaic inverter. Comput. Electr. Eng. 2022, 98, 107738. [Google Scholar] [CrossRef]
  25. Petrone, G.; Ramos-Paja, C.A.; Spagnuolo, G. Photovoltaic Sources Modeling; John Wiley & Sons, Ltd.: Chichester, UK, 2017. [Google Scholar] [CrossRef]
  26. Zhao, Z.; Xu, M.; Chen, Q.; Lai, J.S.; Cho, Y. Derivation, Analysis, and Implementation of a Boost–Buck Converter-Based High-Efficiency PV Inverter. IEEE Trans. Power Electron. 2012, 27, 1304–1313. [Google Scholar] [CrossRef]
  27. Heydari-Doostabad, H.; Monfared, M. An Integrated Interleaved Dual-Mode Time-Sharing Inverter for Single-Phase Grid-Tied Applications. IEEE Trans. Ind. Electron. 2019, 66, 286–296. [Google Scholar] [CrossRef]
  28. Ramos-Paja, C.A.; Saavedra-Montes, A.J.; Bastidas-Rodríguez, J.D. Cargador de baterías fotovoltaico con control por modos deslizantes y limitación de la derivada de corriente de carga. TecnoLógicas 2018, 21, 129–145. [Google Scholar] [CrossRef]
  29. Erickson, R.W.; Maksimovic, D. Fundamentals of Power Electronics, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 2001. [Google Scholar]
  30. Srikanth, M.; Pakkiraiah, B.; Upadhyay, P.; Kalyani, S.T. Dual-mode photovoltaic bidirectional inverter operation for seamless power transfer to dc and AC loads with the grid interface. Int. J. Photoenergy 2019, 2019, 8498435. [Google Scholar] [CrossRef]
  31. Gunawardane, K. Capacitors as energy storage devices-simple basics to current commercial families. In Energy Storage Devices for Electronic Systems: Rechargeable Batteries and Supercapacitors; Number 0; Elsevier Inc.: Amsterdam, The Netherlands, 2015; pp. 137–148. [Google Scholar] [CrossRef]
  32. Irfan, M.S.; Shin, J.H.; Park, J.H. New control method for power decoupling of electrolytic capacitor-less photovoltaic micro-inverter with primary side regulation. J. Electr. Eng. Technol. 2018, 13, 677–687. [Google Scholar] [CrossRef]
  33. Hu, H.; Harb, S.; Kutkut, N.; Batarseh, I.; Shen, Z.J. A review of power decoupling techniques for microinverters with three different decoupling capacitor locations in PV systems. IEEE Trans. Power Electron. 2013, 28, 2711–2726. [Google Scholar] [CrossRef]
  34. Williams, B.W. DC-to-DC converters with continuous input and output power. IEEE Trans. Power Electron. 2013, 28, 2307–2316. [Google Scholar] [CrossRef]
  35. Zakzouk, N.E.; Ibrahim, R.A. Modelling and Performance Analysis of a Buck Converter Featuring Continuous Input/Continous Output Current for Wind Energy-DC Microgrid. In Proceedings of the 2022 6th International Conference on Green Energy and Applications, ICGEA 2022, online, 4–6 March 2022; IEEE: Piscataway, NJ, USA, 2022; pp. 127–132. [Google Scholar] [CrossRef]
  36. Grigore, V.; Kyyra, J. Step-down converter with low ripple input current for Power Factor Correction. In Proceedings of the Conference Proceedings-IEEE Applied Power Electronics Conference and Exposition-APEC, New Orleans, LA, USA, 6–10 February 2000; pp. 188–194. [Google Scholar] [CrossRef]
  37. Gitau, M.N.; Adam, G.P.; Masike, L.; Mbukani, M.W.K. Unified Approach for Synthesis and Analysis of Non-Isolated DC-DC Converters. IEEE Access 2021, 9, 120088–120109. [Google Scholar] [CrossRef]
  38. Chen, Z.; Chen, Y.; Yan, Z. Simplified Hysteresis Sliding-Mode Control for Superbuck Converter. IEEE Trans. Circuits Syst. II: Express Briefs 2020, 67, 3212–3216. [Google Scholar] [CrossRef]
  39. Karppanen, M.; Arminen, J.; Suntio, T.; Savela, K.; Simola, J. Dynamical modeling and characterization of peak-current-controlled superbuck converter. IEEE Trans. Power Electron. 2008, 23, 1370–1380. [Google Scholar] [CrossRef]
  40. Huusari, J.; Leppäaho, J.; Suntio, T. Dynamic properties of PCM-controlled superbuck converter Discrete vs. coupled inductor implementation. EPE J. (European Power Electron. Drives J.) 2010, 20, 8–15. [Google Scholar] [CrossRef]
  41. Abdulslam, A.; Mercier, P.P. A Passive-Stacked Third-Order Buck Converter with Inherent Input Filtering Achieving 0.7-W/mm2 Power Density and 94% Peak Efficiency. IEEE Solid-State Circuits Lett. 2019, 2, 240–243. [Google Scholar] [CrossRef]
  42. Saxena, A.R.; Kulshreshtha, A. A fourth-order bidirectional DC–DC converter for interfacing battery in a solar-photovoltaic-fed low-voltage residential DC nano-grid: Design and analysis. Int. J. Circuit Theory Appl. 2021, 49, 1–27. [Google Scholar] [CrossRef]
  43. Zhang, G.; Jin, N.; Qu, L.; Yu, S.S. Inherently Non-Pulsating Input Current DC-DC Converter for Battery Storage Systems. IEEE Access 2020, 8, 140293–140302. [Google Scholar] [CrossRef]
  44. Suntio, T.; Huusari, J.; Leppäaho, J. Issues on solar-generator interfacing with voltage-fed MPP-tracking converters. EPE J. (European Power Electron. Drives J.) 2010, 20, 40–47. [Google Scholar] [CrossRef]
  45. Ramos-Paja, C.A.; Danilo-Montoya, O.; Grisales-Noreña, L.F. Photovoltaic System for Microinverter Applications Based on a Non-Electrolytic-Capacitor Boost Converter and a Sliding-Mode Controller. Electronics 2022, 11, 2923. [Google Scholar] [CrossRef]
  46. Femia, N.; Petrone, G.; Spagnuolo, G.; Vitelli, M. Optimization of Perturb and Observe Maximum Power Point Tracking Method. IEEE Trans. Power Electron. 2005, 20, 963–973. [Google Scholar] [CrossRef]
  47. Femia, N.; Petrone, G.; Spagnuolo, G.; Vitelli, M. A technique for improving P&O MPPT performances of double-stage grid-connected photovoltaic systems. IEEE Trans. Ind. Electron. 2009, 56, 4473–4482. [Google Scholar] [CrossRef]
  48. Bastidas-Rodriguez, J.; Franco, E.; Petrone, G.; Ramos-Paja, C.; Spagnuolo, G. Maximum power point tracking architectures for photovoltaic systems in mismatching conditions: A review. IET Power Electron. 2014, 7, 1396–1413. [Google Scholar] [CrossRef]
  49. Femia, N.; Petrone, G.; Spagnuolo, G.; Vitelli, M. Power Electronics and Control Techniques for Maximum Energy Harvesting in Photovoltaic Systems; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar] [CrossRef]
  50. Sira-Ramirez, H. Sliding motions in bilinear switched networks. IEEE Trans. Circuits Syst. 1987, 34, 919–933. [Google Scholar] [CrossRef]
  51. González Montoya, D.; Ramos-Paja, C.A.; Giral, R. Improved design of sliding mode controllers based on the requirements of MPPT techniques. IEEE Trans. Power Electron. 2016, 31, 235–247. [Google Scholar] [CrossRef]
  52. Golnaraghi, F.; Kuo, B.C. Automatic Control Systems, 9th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010; pp. 1–772. [Google Scholar]
  53. SpolarPV. Datsheet SPM6-96 460-510 Watt Mono Crystalline Module; SpolarPV: Nanjing, China, 2021. [Google Scholar]
  54. Texas Instruments. Datasheet CSD19538Q3A 100-V N-Channel NexFET ™ Power MOSFET; Texas Instruments: Dallas, TX, USA, 2017. [Google Scholar]
  55. VISHAY. Datasheet V15P8 High Current Density Surface-Mount (TMBS®) Trench MOS Barrier Schottky Rectifier; VISHAY: Malvern, PA, USA, 2019. [Google Scholar]
  56. PSIM. PSIM: The Ultimate Simulation Environment for Power Conversion and Motor Control; PSIM: Troy, MI, USA, 2021. [Google Scholar]
  57. Ramos-Paja, C.A.; Saavedra-Montes, A.J.; Bastidas-Rodriguez, J.D. Co-Design of the Control and Power Stages of a Boost-Based Rectifier with Power Factor Correction Depending on Performance Criteria. Computation 2022, 10, 61. [Google Scholar] [CrossRef]
  58. Abdi, B.; Ghasemi, R.; Mirtalaei, S.M.M. The Effect of Electrolytic Capacitors on SMPS’s Failure Rate. Int. J. Mach. Learn. Comput. 2013, 3, 300–304. [Google Scholar] [CrossRef] [Green Version]
  59. KEMET. C4AQ, Radial, 2 or 4 Leads, 500 - 1,500 VDC, for DC Link (Automotive Grade); KEMET: Fort Lauderdale, FL, USA, 2022. [Google Scholar]
  60. VISHAY. THB Grade IIIB Class X2 Interference Suppression Film Capacitor Radial MKP 305 VAC-Across the Line; VISHAY: Malvern, PA, USA, 2022. [Google Scholar]
  61. Hussain, A.; Kumar, A.; Behera, L. Sliding mode control of a buck converter for maximum power point tracking of a solar panel. In Proceedings of the 2013 IEEE International Conference on Control Applications (CCA), Hyderabad, India, 28–30 August 2013; pp. 661–666. [Google Scholar] [CrossRef]
Figure 1. PV system based on the CIOC buck converter.
Figure 1. PV system based on the CIOC buck converter.
Computation 11 00042 g001
Figure 2. Topology 1 of the PV system based on the CIOC buck converter.
Figure 2. Topology 1 of the PV system based on the CIOC buck converter.
Computation 11 00042 g002
Figure 3. Topology 0 of the PV system based on the CIOC buck converter.
Figure 3. Topology 0 of the PV system based on the CIOC buck converter.
Computation 11 00042 g003
Figure 4. Waveforms of the PV system based on the CIOC buck converter.
Figure 4. Waveforms of the PV system based on the CIOC buck converter.
Computation 11 00042 g004
Figure 5. PV system based on the classical buck converter.
Figure 5. PV system based on the classical buck converter.
Computation 11 00042 g005
Figure 6. Input capacitor comparison between both buck converters.
Figure 6. Input capacitor comparison between both buck converters.
Computation 11 00042 g006
Figure 7. Block diagram of the proposed SMC (measured signals: blue lines; control signal: red line).
Figure 7. Block diagram of the proposed SMC (measured signals: blue lines; control signal: red line).
Computation 11 00042 g007
Figure 8. Electrical characteristics of the SP500M6-96 PV panel.
Figure 8. Electrical characteristics of the SP500M6-96 PV panel.
Computation 11 00042 g008
Figure 9. Design of the inductors in the CIOC converter.
Figure 9. Design of the inductors in the CIOC converter.
Computation 11 00042 g009
Figure 10. Flowchart of the proposed design process, where the P&O parameters are defined following [46,47].
Figure 10. Flowchart of the proposed design process, where the P&O parameters are defined following [46,47].
Computation 11 00042 g010
Figure 11. Circuital implementation of the PV system based on the CIOC converter and SMC.
Figure 11. Circuital implementation of the PV system based on the CIOC converter and SMC.
Computation 11 00042 g011
Figure 12. Performance comparison of PV systems based on both CIOC and classical buck converters.
Figure 12. Performance comparison of PV systems based on both CIOC and classical buck converters.
Computation 11 00042 g012
Figure 13. Frequency response of the PV system in closed loop.
Figure 13. Frequency response of the PV system in closed loop.
Computation 11 00042 g013
Figure 14. Perturbations mitigation in the closed-loop circuit.
Figure 14. Perturbations mitigation in the closed-loop circuit.
Computation 11 00042 g014
Figure 15. Performance of the PV system to changes on the reference value.
Figure 15. Performance of the PV system to changes on the reference value.
Computation 11 00042 g015
Figure 16. Performance of the PV system to different perturbations in the load voltage.
Figure 16. Performance of the PV system to different perturbations in the load voltage.
Computation 11 00042 g016
Figure 17. Maximum power point tracking using the CIOC converter with SMC and the classical buck converter with SMC.
Figure 17. Maximum power point tracking using the CIOC converter with SMC and the classical buck converter with SMC.
Computation 11 00042 g017
Figure 18. Maximum power point tracking trajectories.
Figure 18. Maximum power point tracking trajectories.
Computation 11 00042 g018
Table 1. Parameters of the SP500M6-96 PV panel and COTEK SP-700-124 inverter.
Table 1. Parameters of the SP500M6-96 PV panel and COTEK SP-700-124 inverter.
ParameterValue
Panel maximum Power ( P m p p )500 W
Panel maximum power voltage ( V m p p )48.63 V
Panel maximum power current ( I m p p )10.28 A
Panel short-circuit current ( I s c )10.87 A
Panel open-circuit voltage ( V o c )58.95 V
Inverter input voltage range ( v o )21–33 V
Inverter nominal output voltage (RMS)100/110/115/120 V
Inverter nominal output power700 W
Inverter output frequency50/60 Hz
Table 2. Design example of the PV system based on the CIOC converter.
Table 2. Design example of the PV system based on the CIOC converter.
ParameterValue
L 1 , L 2 38 [ μ H]
C p v 47 [ μ F]
C i 5.6 [ μ F]
max F 100 [kHz]
Δ v p o 500 [mV]
T p o 500 [ μ s]
t s 250 [ μ s]
k c −1 [−]
k p 2.36 [A/V]
k i 29.5 [kA/(V · s)]
R a t e 0.257 [V/ μ s]
τ f 1.95 [ μ s]
H1.67 [A]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ramos-Paja, C.A.; Bastidas-Rodriguez, J.D.; Saavedra-Montes, A.J. Low-Voltage Photovoltaic System Based on a Continuous Input/Output Current Converter. Computation 2023, 11, 42. https://doi.org/10.3390/computation11020042

AMA Style

Ramos-Paja CA, Bastidas-Rodriguez JD, Saavedra-Montes AJ. Low-Voltage Photovoltaic System Based on a Continuous Input/Output Current Converter. Computation. 2023; 11(2):42. https://doi.org/10.3390/computation11020042

Chicago/Turabian Style

Ramos-Paja, Carlos Andres, Juan David Bastidas-Rodriguez, and Andres Julian Saavedra-Montes. 2023. "Low-Voltage Photovoltaic System Based on a Continuous Input/Output Current Converter" Computation 11, no. 2: 42. https://doi.org/10.3390/computation11020042

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop