Next Article in Journal
Prospective Analysis of Time-Fractional Emden–Fowler Model Using Elzaki Transform Homotopy Perturbation Method
Previous Article in Journal
Stability Analysis of a Fractional-Order Time-Delayed Solow Growth Model with Environmental Pollution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fractional-Order Modeling and Nonlinear Dynamic Analysis of Forward Converter

School of Mechanical and Electrical Engineering, Guangzhou University, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Fractal Fract. 2024, 8(6), 362; https://doi.org/10.3390/fractalfract8060362
Submission received: 23 April 2024 / Revised: 7 June 2024 / Accepted: 17 June 2024 / Published: 19 June 2024
(This article belongs to the Section Engineering)

Abstract

:
To accurately investigate the nonlinear dynamic characteristics of a forward converter, a fractional-order state-space averaged model of a forward converter in continuous conduction mode (CCM) is established based on the fractional calculus theory. And nonlinear dynamical bifurcation maps which use PI controller parameters and a reference current as bifurcation parameters are obtained. The nonlinear dynamic behavior is analyzed and compared with that of an integral-order forward converter. The results show that under certain operating conditions, the fractional-order forward converter exhibits bifurcations characterized by low-frequency oscillations and period-doubling as certain circuit and control parameters change. Under the same circuit conditions, there is a difference in the stable parameter region between the fractional and integral-order models of the forward converter. The stable zone of the fractional-order forward converter is larger than that of the integral-order one. Therefore, the circuit struggles to enter states of bifurcation and chaos. The stability domain for low-frequency oscillations and period-doubling bifurcations can be accurately predicted by using a small signal model and a predictive correction model of the fractional-order forward converter, respectively. Finally, by performing circuit simulations and hardware-in-the-loop experiments, the rationality and correctness of the theoretical analysis are verified.

1. Introduction

Fractional calculus refers to a field where the order of differentiation and integration can be any arbitrary or fractional number. Employing the theory of fractional calculus can greatly expand our ability to describe systems [1]. Given the fractional-order characteristics of dielectric materials, actual inductors and capacitors are fractional-order components [2,3,4,5]. Therefore, a forward converter that contains fractional-order inductors and capacitors is considered as a fractional-order system. Approximating actual systems as integer-order ones is only viable for systems where there is not a high requirement for accuracy [6]. With the rapid development of power electronic converters, there is a growing demand for high levels of stability, reliability, and accuracy [7,8]. A forward converter is a highly nonlinear time-variant system that presents rich nonlinear dynamics under certain working conditions, directly affecting the converter’s stability and reliability [9,10,11]. These nonlinear dynamic phenomena are mainly attributed to the nonlinearities introduced by switching devices and feedback loops [12]. In power converter design, particular attention must be paid to nonlinear dynamics to ensure system stability and performance. Typically, electronic engineers hope the converter operates in a stable state [13]. Therefore, predicting the system’s bifurcation boundaries to prevent such nonlinear dynamic behavior is a research priority. The dynamical analysis of fractional-order nonlinear systems is a hot research topic [14]. Many existing studies indicate that the models established using fractional calculus closely resemble actual situations [15] and can more accurately describe the physical phenomena of power electronic converters.
In recent years, research has demonstrated that fractional-order models of inductors and capacitors accurately represent their real electrical characteristics. Inductors and capacitors are essential components in power converters, making the modeling and nonlinear dynamic analysis of fractional-order converters practically significant. Studies on the fractional-order modeling of DC-DC converters have been reported in [16,17], mainly focusing on non-isolated converters. For example, the fractional-order modeling of boost converters was examined in [18], and the effect of fractional-order on harmonics was analyzed. A fractional-order model of the Boost converter was established in [19], and the effect of the fractional-order on harmonics was analyzed. Circuit and simulation models for fractional-order Buck–Boost converters were also developed in [20], revealing the influence of fractional-order on harmonic characteristics. These findings suggest that the fractional order can serve as an additional system parameter. However, research on fractional-order models of complex power converters, particularly those with transformers, remains limited. The inclusion of electromagnetic issues complicates their modeling process.
Power electronic converters exhibit rich and complex nonlinear dynamic behaviors, including various bifurcation phenomena and chaotic motions. When bifurcation and chaotic motion occur in a converter, they cause substantial noise and instability, which are detrimental to practical applications. Therefore, it is crucial to avoid bifurcation phenomena and chaotic motion as much as possible. However, designing the system parameters too far from the bifurcation boundary can degrade the system’s dynamic response and other performance metrics. Analyzing the nonlinear phenomena in converters is a significant research focus. The methods for studying bifurcation phenomena in DC-DC converters have been summarized in [21]. A novel discrete mapping method for analyzing nonlinear phenomena in DC-DC converters was proposed in [22]. Permutation entropy was used to characterize the nonlinear phenomena in nonlinear DC-DC converter systems in [23]. The dynamic behavior of a single-stage Boost converter system with memristive load peak current mode control was studied in [24]. Most of these studies focus on integral-order systems, while the nonlinear phenomena and analysis methods for fractional-order systems are considerably more complex. The dynamical behavior of Boost converter systems with memristor loads was analyzed in [25]. A fractional-order model for magnetically coupled Boost converters in CCM mode was established, and the nonlinear dynamic phenomena of this model were explored in [26]. A fractional-order Duffing system was established, and bifurcation diagrams, phase diagrams, Poincaré maps, and time domain waveforms were used to study the nonlinear dynamic behavior of the fractional-order Duffing system in [27]. These studies indicate that fractional-order systems exhibit richer dynamic characteristics, making it challenging to accurately describe the actual operating state of power electronic converters using integral-order models. Thus, studying fractional-order systems is crucial for establishing precise system models.
Much of the current research on fractional-order converters is mostly focused on certain singular nonlinear dynamic behaviors, lacking the systematic analysis of the nonlinear phenomena generated by converters. In fact, different control methods can lead to various nonlinear dynamic behaviors in power electronic converters, all of which can affect the system’s output variables. The analysis of the nonlinear dynamic behavior of fractional-order isolated DC-DC converters is even more complex. The fractional-order modeling of a flyback converter has been proposed in [28], but it did not delve deeply into the nonlinear dynamic characteristics of this model. Similar to the flyback converter, forward converters are also traditional isolated DC-DC converters. The fractional-order modeling and study of its nonlinear dynamic behavior also hold significant research value.
In this paper, a state-space averaged model of the fractional-order forward converter in continuous conduction mode (CCM) is established, and the profound study and analysis of its nonlinear dynamic behavior are conducted. The structure of this paper is as follows: Section 2 establishes the state-space averaged model of the fractional-order forward converter and derives the direct current component, the alternating current component, and the transfer function for this model in CCM. Section 3 conducts simulations of the fractional-order forward converter’s circuit model and compares the results with those of an integral-order forward converter. Section 4 focuses on discussing the low-frequency oscillation phenomena and period-doubling bifurcations of the forward converter, deeply analyzing the nonlinear dynamic behavior of the model, and accurately predicting the stable parameter domain of fractional-order forward converter using small signal model and predictor-corrector methods. In Section 5, the rationality and correctness of theoretical analysis are further validated through hardware-in-the-loop experiments. Finally, conclusions are drawn in Section 6.

2. Fractional-Order Modeling of Forward Converter

Fractional calculus can be viewed as the generalization of traditional calculus that extends ordinary integrations and differentiations to non-integral orders, thus demonstrating its wider applicability [29].
Fractional calculus can be defined through three main approaches: Riemann–Liouville (RL), Grünwald–Letnikov (GL), and Caputo [30]. The Caputo definition is particularly adept at addressing initial value problems in fractional differential equations. It is expressed as
d μ x t d t μ = D t μ a x t = 1 Γ m μ a t t ζ m μ 1 x m ζ d ζ
In this formula, m is an integer such that m − 1 < µ < m, where Γ(•) represents the gamma function, and α denotes the lower bound of the definition.
Using the Laplace transform, the Caputo fractional derivative is defined as
L D t μ 0 x t = 0 D t μ 0 x t × e s t d t = s μ X s k = 0 m 1 s k D t μ k 1 0 x k 0
In this paper, the Caputo derivative is employed for its acceptable error margin in all calculations [31]. It is essential for accurate results of the initial conditions that are explicitly known before the commencement of calculations.
With zero initial conditions, the Laplace transform of the Caputo fractional derivative is represented as
L D t μ 0 x t = 0 D t μ 0 x t × e s t d t = s μ X s
According to Westerlund’s 1994 linear capacitor model [3], the relationship between the current iC and the voltage vC in a real capacitor is given by
i C = C d α v C d t α         ,   0   <   α   <   1
where C represents the capacitance, and α denotes the fractional order related to the dielectric type.
Westerlund also asserted that a real inductor inherently possesses fractional properties [4]. Based on existing research, the voltage across a real inductor is
v L = L d β i L d t β   ,   0   <   β   <   1
in which, L is the inductance, and β is the fractional order associated with electromagnetism.
In continuous conduction mode (CCM), the forward converter has two switching states, as shown in Figure 1. The transformer performs voltage transformation, where n denotes the transformer ratio, n = N2/N1. A flux reset circuit is formed by inductor L0 and diode D0. S is an active switch, D1 and D2 are diodes, L is the filter inductor, and C is the filter capacitor, Vin is the input voltage, α is the order of L, and β is the order of C. In Figure 2a, S is closed, Vin provides power to the load resistor, the inductor is in a charging state. In Figure 2b, S is in an open state, the inductor releases energy and transfers it to the load resistor, and the inductor and capacitor maintain the power supply state.
State 1: When S is on (0 ≤ tdT), the diode is reverse-biased, and the inductor current flows through S, increasing iL. The state-space equations for State 1 using iL and vo as state variables are
d α i L d t α d β v o d t β = 0 1 L 1 C 1 R C i L v o + n L 0 v in
A 1 = 0 1 L 1 C 1 R C   ,   b 1 = n L 0
State 2: When S is off (dTt < T), D becomes forward-biased, and the inductor current decreases, while flowing through D. The state-space equations for State 2 using iL and vo as state variables are
d α i L d t α d β v o d t β = 0 1 L 1 C 1 R C i L v o + 0 0 v in
A 2 = 0 1 L 1 C 1 R C   ,   b 2 = 0 0
According to [32], the coefficient matrices and vectors for the complete state equations are
A = d A 1 + 1 d A 2 = 0 1 L 1 C 1 R C
b = d b 1 + 1 d b 2 = n d L 0
In power converters, the current iL and voltage vo exhibit high-frequency switching harmonics. To mitigate these harmonics, averaging the waveforms over a switching period is necessary [31]. This can be achieved by defining the sliding average value as
y = 1 T t t + T y τ d τ
in which, y represents the circuit variable of the converter.
The principles of fractional calculus lead to the derivation of the following equation.
d γ y d t γ = d γ 1 T t t + T y τ d τ d t γ = 1 T t t + T d γ y τ d t γ d τ = d γ y d t γ
where γ is the order, and it satisfies the condition 0 < γ < 1.
In state-space modeling, the input voltage vin and the duty ratio d are also considered. i L , v o , v in , and d , are the mean values of iL, vo, vin, and d, respectively, and are depicted as
i L = I L + i ^ L   ,   v o = V o + v ^ o   ,   v in = V in + v ^ in   ,   d = D + d ^
in the equations; the variables with uppercase letters indicate the DC components in i L , v o , v in , and d , while the AC components are represented by i ^ L , v ^ o , v ^ in , and d ^ represent the corresponding AC components. Assuming small signals, the AC components are significantly smaller in magnitude compared to the DC parts.
The state-space averaged model of the fractional-order forward converter can be derived as
d α i L d t α d β v o d t β = 0 1 L 1 C 1 R C i L v o + n d L 0 v in
= 0 1 L 1 C 1 R C I L + i ^ L V o + v ^ o + D + d ^ 0 V in + v ^ in
The quiescent operation point is formulated into
I L V o = nDV in R nDV in
According to (6), the increment of iL during (0, dT) can be calculated as ΔiL.
Δ i L = nV in V o D T α α L Γ α
where Γ(•) denotes the Gamma function.
According to (17), the current ripple ΔiL is dependent on variables such as the input voltage Vin, the duty cycle D, the switching period T, and the inductor order α. Incorporating (16) and (17), the peak current iLmax and minimum current iLmin can be obtained as
i Lmax = I L + 1 2 Δ i L = nDV in R + nV in V o D T α 2 α L Γ α
i Lmin = I L 1 2 Δ i L = nDV in R nV in V o D T α 2 α L Γ α
Likewise, based on (16), the reduction in capacitor voltage within the timeframe of (0, dT) can be computed. Δvo can be described as
Δ v o = 1 E β D β T β R C V OT
where VOT is the initial voltage value when the switch S is turned on, and Eβ(•) is a Mittag-Leffler function. VOT can be approximated as
V OT = 1 2 Δ v o + V o
The expression for voltage ripple can be derived from (14), (20), and (21) as
Δ v o = 2 nDV in R 1 E β D β T β R C 1 + E β D β T β R C
In order to ensure the stable operation of the forward converter in CCM mode, the value of the inductor current iL needs to be greater than or equal to 0 at the moment of switch cycle initiation or termination, namely
I L 1 2 Δ i L     0
When substituting (16) and (17) into (23), the result is
nDV in R nV in V o D T α 2 α L Γ α
As can be derived from (18), the conditions for the stable operation of the forward converter in CCM mode are related to the load resistance R, the duty cycle D, the period T, excitation inductance L, and the order α. The larger α is, the easier it is for the forward converter to operate in CCM mode. When α equals one, it corresponds to the same conditions for continuous operation in integral-order modeling as stated.
After isolating the AC component in (13) and ignoring the higher-order signals, the state-space equations for the fractional-order AC small signal model are obtained
d α i ^ L d t α d β v ^ o d t β = 0 1 L 1 C 1 R C i ^ L v ^ o + D L 0 v ^ in + d ^ L 0 V in
Through the application of the Laplace transform based on fractional calculus, the following is derived.
s α i ^ L s = 1 L v ^ o s + nD L v ^ in s + n V in L d ^ s
s β v o s = 1 C i ^ L s 1 R C v ^ o s
According to (26) and (27), the transfer function from v ^ in to v ^ o can be expressed as
G v o v in s = v ^ o s v ^ in s d ^ s = 0 = nD LC s α + β + 1 R C s α + 1 LC
The transfer function from d ^ to v ^ o can be expressed as
G v o d s = v ^ o s d ^ s v ^ in s = 0 = nV in LC s α + β + 1 R C s α + 1 LC
The transfer function from v ^ in to i ^ L can be denoted as
G i L v in s = i ^ L s v ^ in s d ^ s = 0 = nD L s β + nD LCR s α + β + 1 R C s α + 1 LC
The transfer function from d ^ to i ^ L can be indicated as
G i L d s = i ^ L s d ^ s v ^ in s = 0 = nV in L s β + nV in LCR s α + β + 1 R C s α + 1 LC
The relationship between the transfer functions in (28)–(31) and the fractional orders of the inductor and capacitor can be observed. By assigning a value of one to all the orders, the model transitions to integral-order transfer functions.

3. Simulation Results

To validate the fractional-order models mentioned above, this section simulates the fractional-order model of the forward converter and compares the simulation results with those of the integral-order model. The circuit parameters of the forward converter are as follows: Vin = 5 V, n = 1, D = 0.7, L = 0.125 mH, C = 200 μF, R = 5 Ω, f = 10 kHz, α = 0.95, and β = 0.95. In order to further validate the research results, this paper establishes an approximate circuit for fractional-order inductors and fractional order capacitors using the Oustaloup algorithm and employing resistor/capacitor or resistor/inductor networks. Fractional-order elements are not commercially available, but researchers have suggested constructing equivalent circuits using the fractional properties of elements [33,34].
The impedance chain of the fractional-order inductance is shown in Figure 3. Parallel impedance can be obtained from the parallel relationship between resistance and inductance, and the model structure of fractional-order inductance impedance chain can be obtained using (32).
  Ls α = i = 1 n R Li s s + R Li L Li + R Lin
The values of RLi and LLi for the inductor in the order α = 0.95 are shown in Table 1.
Similarly, for capacitance possessing fractional-order characteristics, the structure of their impedance chain is depicted in Figure 4.
1 C s β = i = 1 n 1 C Ci s + 1 R Ci C Ci + R Cin
The component parameters in the fractional-order capacitor impedance chain are also provided in Table 1.
Figure 5 presents the Bode plots for constructed fractional-order inductance and capacitance. As observed, the Oustaloup method of fitting the impedance chain is ideal, except at both ends of the fitting band. This is an inherent characteristic of Oustaloup approximation. Using too low an order for fitting might not offer the necessary accuracy, while too high an order can increase the scale of impedance network. Ninth-order fitting provides a good balance between accuracy and complexity, effectively capturing the characteristics of fractional-order elements. This paper uses ninth-order fitting for s0.95 and s−0.95, setting the switch frequency of the forward converter to 10 kHz. By appropriately configuring the fitting interval, it demonstrates favorable frequency characteristics at the switching frequency. From the diagram, within the defined frequency range, the magnitude lines fit well, and the phase diagram fluctuates within a certain band around the theoretical curve.
By replacing inductance L and capacitance C in Figure 1 with the approximate circuits for fractional-order inductance and capacitance shown in Figure 3 and Figure 4, we can observe from the circuit simulation results in Figure 6 that the ripples of iL and vo increase significantly as the orders of inductance and capacitance decrease. As it can be observed in Figure 6, the inductor current in the fractional-order model is higher than that in the integral-order model. This phenomenon is triggered by the fractional-order characteristic of the capacitors. When the capacitor’s order is 0.95, from Figure 4 it can be deduced that the angle between its voltage and current is approximately −85°. Consequently, the power factor cosφ is non-zero, with φ being the angle between the voltage and current, which expands the total output power of the circuit, and subsequently leads to an escalation in the input current, inducing an increase in the inductor current. The circuit simulation results are shown in Table 2.
The bode diagrams of G i L v in s and G v o v in s with (α, β) = (1, 1) and (α, β) = (0.95, 0.95), respectively, are shown in Figure 7. The transfer functions G i L v in s and G v o v in s derived from the state-space averaged model are provided in (27) and (34), respectively. The majority of the differences seen in the circuit models can be attributed to adjustments made to the fractional-order inductors and capacitors. When compared with the integral-order model of the forward converter, due to the fractional-order properties of the impedance chain, the amplitude-frequency characteristic curve of the fractional-order model shifts to the right, and the phase-frequency characteristic curve exhibits a leading phase angle, which suggests a superior dynamic performance and stability for the fractional-order model of the forward converter.
The dynamic characteristics of the fractional-order model and integral-order model of the forward converter can be observed in Figure 8. It can be found that when the orders of the energy storage components in the circuit transition to fractional orders, the circuit’s response is more immediate, the transient state fluctuations lessen, and it enters the steady-state phase more rapidly. This is due to the introduction of fractional-order chain, which alters the internal parameters of the circuit. As illustrated by the Bode plot in Figure 7, the change in the order of the energy storage elements modifies the system’s transfer function. Fractional-order circuits exhibit a better dynamic performance, which is reflected in the system’s regulation time, response speed, and overshoot. All these factors imply that relative to the integral-order model, the fractional-order model of the forward converter delivers a better dynamic performance and stability. The dynamic performance of the circuit is shown in Table 3. It can be observed that there is a significant difference in the overshoot between the two. This is because the peak voltage of the integral-order forward converter far exceeds the steady-state output voltage of the circuit. However, due to the reduction in the order of the energy storage components, the peak voltage of the fractional-order forward converter has significantly decreased, becoming relatively close to the steady-state voltage of the circuit. As a result, the system’s overshoot is significantly improved. The data in Table 3 further illustrate that compared to the integral-order forward converter, the fractional-order forward converter exhibits certain advantages.

4. Nonlinear Dynamic Analysis

As a nonlinear time-varying system, the forward converter encompasses a rich set of nonlinear dynamical behaviors. Low-frequency oscillations usually refer to oscillations within the frequency range of several tens of hertz to several thousand hertz. Such oscillations can lead to fluctuations in the output voltage, affecting the stability and reliability of the power supply system. Period-doubling bifurcation refers to the phenomenon where the oscillation frequency of the system’s motion doubles or halves as the system parameters change, which similarly affects the stability and reliability of power electronic converters. To explore the nonlinear dynamical behavior of the forward converter, MATLAB/Simulink (version: 2018a) circuit simulation is used to compare the nonlinear dynamical phenomena of fractional-order and integral-order models of the forward converter and to analyze their stable parameter domains. The state-space average model and predictive-correction model can accurately predict the stable parameter domain of the fractional-order forward converter.

4.1. Low-Frequency Oscillation Phenomena

4.1.1. Low-Frequency Oscillation Phenomena in Fractional-Order Forward Converter

To analyze the low-frequency oscillation phenomena of the fractional-order forward converter, this study establishes a fractional-order model of the voltage-mode-controlled forward converter. With PI controller parameters as bifurcation parameters, a comparison of low-frequency oscillation phenomena when the forward converter (α, β) = (1, 1) and (α, β) = (0.95, 0.95) is conducted. The phase diagrams of the model in stable and low-frequency oscillation states are plotted with inductor current and output voltage as coordinate axes. The system’s block diagram is illustrated in Figure 9.
Initially, Ki in the PI controller is used as the bifurcation parameter, and the bifurcation diagram of the forward converter for (α, β) = (1, 1) is shown in Figure 10a. When (α, β) = (0.95, 0.95), as illustrated in Figure 10b, similar to the integral-order forward converter, as Ki changes, the system exhibits lower-frequency oscillation. The system’s motion transitions from a fixed point to periodic oscillation. As the orders decrease, the bifurcation point shifts backward, indicating that the stable parameter domain enlarges.
To further describe the low-frequency oscillation phenomena of the fractional-order forward converter, Figure 11 presents the time-domain waveform of the inductor current at the bifurcation point. Figure 11a shows the time-domain waveform of the inductor current iL for a forward converter with (α, β) = (1, 1) when Ki = 320. Figure 11b shows the time-domain waveform of the inductor current iL for a forward converter with (α, β) = (0.95, 0.95) when Ki = 885. It can be observed that when the converter operates at the bifurcation point, due to the occurrence of low-frequency oscillations, the inductor current iL drops to zero during certain switching cycles, putting the system in a critical state of continuous inductor current mode. That is, the system alternates between continuous conduction mode (CCM) and Discontinuous Conduction Mode (DCM). When the converter experiences low-frequency oscillations, the ripple of the inductor current significantly increases, worsening the operating performance of the system.
Figure 12 shows the phase diagram formed by the inductor current and the output voltage. Using Ki as the bifurcation parameter for a forward converter with (α, β) = (1, 1), when Ki < 320, the converter operates in a period-1 stable manner, as illustrated in Figure 12a, where the phase diagram is a closed, smooth loop. When Ki ≥ 320, the system undergoes low-frequency oscillations, forming a limit cycle in the phase diagram, as shown in Figure 12b. Similarly, for a forward converter with (α, β) = (0.95, 0.95), when Ki < 885, the system operates stably, forming a closed, smooth loop in the phase diagram, as depicted in Figure 12c. When Ki ≥ 885, the converter exhibits low-frequency oscillation behavior, and the corresponding phase diagram shown in Figure 12d is a limit cycle. The appearance of limit cycles in the phase diagram of DC-DC converters is indicative of low-frequency oscillations in the circuit. It is observed that when the circuit undergoes low-frequency oscillations, both the inductor current and the output voltage ripple significantly increase. These observations are consistent with the conclusions drawn from the bifurcation diagrams and time domain graphs.

4.1.2. Predicting the Low-Frequency Oscillation Characteristics of the Fractional-Order Forward Converter

The small-signal averaging approach is a modelling strategy extensively adopted in power electronics systems. Despite the fact that the small-signal averaging model ignores the influence of switching characteristics in the converter, it considers the low-frequency oscillation dynamics of the system. Hence, the small-signal averaging model can be effectively used to analyze the low-frequency oscillation phenomena in the system. The small-signal block diagram of the system is shown in Figure 13.
Gv(s) represents
G v s = v ^ o d ^ = nV in LC s α + β + 1 R C s α + 1 LC
The transfer function of the PI voltage control loop is
G c s = v ^ con v ^ e = K p + K i 1 s
The transfer function of the PWM stage is
G d s = d ^ v ^ con = 1 V M
in which, VM = VUVV.
With the use of (31), (32), and (33), we can determine the open-loop and closed-loop transfer functions of the system, expressed as (34) and (35), respectively.
G k s = G v s G c s G d s
= b 0 s + b 1 a 0 s α + β + 1 + a 1 s α + 1 + a 2 s
in which, b0 = nVinKp/LC, b1 = nVinKi/LC, a0 = VM, a1 = VM/RC, and a2 = VM/LC.
Table 4 shows the gain margin, phase margin, and crossover frequency of the fractional-order forward converter as Ki gradually increases. It is observed that as Ki increases, both the gain margin and phase margin of the system progressively decrease until the system loses stability. For the forward converter with (α, β) = (1, 1), the system transitions from stable to unstable as Ki increases from 310 to 320. Similarly, for the forward converter with (α, β) = (0.95, 0.95), the system is on the verge of stable operation as Ki changes from 875 to 885. Notably, these critical points coincide with the bifurcation points obtained from the circuit model. This demonstrates that the small-signal model established based on the state-space averaging method can accurately predict the low-frequency oscillation boundaries of the fractional-order forward converter.

4.2. Period-Doubling Bifurcation

4.2.1. The Period-Doubling Bifurcation in Fractional-Order Forward Converter

In order to analyze the period-doubling bifurcations of the fractional-order forward converter, this study has established a bifurcation diagram for a current-controlled fractional-order forward converter. With the reference current as a bifurcation parameter, the comparison of period-doubling bifurcation when the forward converter (α, β) = (1, 1) and (α, β) = (0.95, 0.95) is conducted. The phase diagrams of the model in a stable, period-doubling bifurcation state and a chaotic state are plotted with the inductor current and output voltage as coordinate axes. The system’s block diagram is illustrated in Figure 14.
The reference current is used as a parameter to describe the bifurcation. As the Iref varies, the bifurcation diagram for a forward converter with (α, β) = (1, 1) under continuous inductor current mode is shown in Figure 15a. Figure 15b shows the bifurcation diagram for a forward converter with (α, β) = (0.95, 0.95) under continuous inductor current mode. It is observed that as Iref changes, the forward converter first passes through period-1, and then period-2, followed by period-4, and it eventually reaches the chaotic mode. When the system is in chaos, the inductor current iL no longer has periodicity and exhibits large fluctuations. Comparing Figure 15a and b, the bifurcation points move backward as the order of the fractional-order converter decreases. The fractional-order forward converter possesses a larger stable parameter domain.
To better observe the period-doubling bifurcation phenomenon of the fractional-order forward converter, Figure 16 illustrates the time-domain waveforms of the inductor current at the bifurcation points for the fractional-order forward converter. Figure 16a shows the inductor current waveform for a forward converter with (α, β) = (1, 1) at Iref = 5.7 A. At this point, the converter is at the bifurcation point, and the circuit operates in a period-2 state. Figure 16b depicts the time-domain graph of the inductor current at the bifurcation point for a forward converter with (α, β) = (0.95, 0.95). When Iref = 8.2 A, the circuit undergoes period-doubling bifurcation, transitioning the system’s motion from period T to period 2T. Period-doubling bifurcation is one of the pathways leading to chaotic behavior. As Iref gradually increases, the fractional-order forward converter transitions to a chaotic state, cascading through the period-doubling pathways.
To further observe the impact of the reference current on the system, phase diagrams are utilized to describe the four states of the fractional-order forward converter as Iref varies. As shown in Figure 17, with the increase in Iref, the forward converter with (α, β) = (1, 1) displays four different states. It is observed that when Iref = 5 A, the phase diagram exhibits a single closed, smooth loop, indicating that the system is in a stable operating condition, namely the period-1 state. When Iref increases to 6 A, the phase diagram still shows a closed loop, but with indentations, suggesting the occurrence of a period-doubling bifurcation. At Iref = 7 A, the heart-shaped closed loop in the phase diagram undergoes another period-doubling bifurcation, with the inductor current and output voltage transforming into signals four times the period of the ramp signal, indicating that the system operates in the period-4 state. When Iref reaches 8 A, the trajectories in the phase diagram become chaotic and disordered, signifying that the system has entered a chaotic state. Figure 8 presents the phase diagrams of the forward converter with (α, β) = (0.95, 0.95), illustrating the changes with varying Iref. Similar to the integral-order circuit, as Iref gradually increases, the system transitions from a stable period-1 operation to a period-2 state. Subsequently, the system undergoes another period-doubling bifurcation, moving into the period-4 state. Eventually, the trajectories in the phase diagram become chaotic, indicating the system’s entry into a chaotic state. Compared to the integral-order model, the fractional-order model with the same parameters exhibits different dynamical behaviors. For instance, at Iref = 7 A, the forward converter with (α, β) = (1, 1) is in the period-4 state, as shown in Figure 17c, whereas the forward converter with (α, β) = (0.95, 0.95) remains in the period-1 state, as depicted in Figure 18a. Similarly, it can be deduced from the phase diagrams that fractional-order forward converters are more resistant to entering a chaotic state.

4.2.2. Prediction of the Period-Doubling Bifurcation Boundary for Fractional-Order Forward Converter

The fractional-order predictor-corrector model is a method used for handling nonlinear systems. It accurately characterizes the dynamic properties of a system and can predict the period-doubling bifurcation boundaries of a fractional-order forward converter. This algorithm is a time-domain approximation method in the fractional-order domain for solving fractional calculus. It can discretize differential equations of any order to compute numerical solutions and conduct stability analysis of these solutions. Equations (6) and (8) describe two operating states of the fractional-order forward converter in continuous conduction mode. This model is applied to discretize the fractional-order forward converter operating in CCM. The discretized form of the fractional-order forward converter is shown in (38) and (39). Equation (40) represents the nonlinear switching function S(t), where S = 1 means the switch is on, and S = 0 means the switch is off. Equation (41) describes the switching conditions of the converter.
i m + 1 = i 0 + h q 1 Γ q 1 + 2 nV in L S u m + 1 p L + h q 1 Γ q 1 + 2 i = 0 m a i , m + 1 q 1 nV in L S u i L
u m + 1 = u 0 + h q 2 Γ q 2 + 2 i m + 1 p C u m + 1 p R C + h q 2 Γ q 2 + 2 i = 0 m a i , m + 1 q 2 i i C u i R C
S ( t ) = 1 ,   mT < t < m + d T   0 ,   m + d T < t < m + 1 T
σ = i L ( t ) I ref = 0
In the formula, q1 and q2 represent the orders of the inductor and capacitor, respectively. i0 and u0 are the initial values of the inductor current and capacitor voltage, respectively. The correction coefficients for the inductor current and capacitor voltage are denoted as ai,m+1q1 and ai,m+1q2, respectively. m represents the number of iterations. The initial estimate approximations for the inductor current and capacitor voltage are represented as im+1p and um+1p, and their expressions are provided in (42)–(45).
a i , m + 1 q 1 = m q 1 + 1 m q 1 m + 1 q 1 , i = 0 m i + 2 q 1 + 1 + m i q 1 + 1 2 m i + 1 q 1 + 1 , 1 i m i , i = m + 1
a i , m + 1 q 2 = m q 2 + 1 m q 2 m + 1 q 2 , i = 0 m i + 2 q 2 + 1 + m i q 2 + 1 2 m i + 1 q 2 + 1 , 1 i m i , i = m + 1
i m + 1 p = i 0 + 1 Γ q 1 i = 0 m b i , m + 1 q 1 nV in L S u i L
u m + 1 p = u 0 + 1 Γ q 2 i = 0 m b i , m + 1 q 2 i L C u i R C
The fractional-order predictor-corrector model is used to describe the states of the forward converter under different Iref values. After m iterations, the waveform of the inductor current can be obtained. Figure 19 shows the inductor current waveforms of the forward converter with (α, β) = (1, 1). As depicted in Figure 19a, when Iref = 5.4 A, the circuit operates in a stable state. When Iref increases to 5.7 A, the inductor current undergoes period-doubling bifurcation, as shown in Figure 19b, and the circuit enters the period-2 state. Figure 20 presents the inductor current waveforms for the forward converter with (α, β) = (0.95, 0.95). At Iref = 7.9 A, the circuit is in the period-1 state, as illustrated in Figure 20a. As shown in Figure 20b, when Iref reaches 8.2 A, the inductor current exhibits period-doubling bifurcation. Clearly, the bifurcation points obtained using the fractional-order predictor-corrector model are consistent with the results of circuit simulations. This indicates that the model can accurately predict the period-doubling bifurcation boundaries of a fractional-order forward converter.

5. Experimental Results

To further verify the validity of theoretical analysis, a hardware-in-the-loop experiment was conducted. The experimental system consists of an NI PXIe-1082 chassis (ModelingTech, Shanghai, China), a digital signal processor (TMS320F28335, Texas Instruments, Dallas, The USA) and an oscilloscope. The experimental circuit parameters of the forward converter are shown in Table 5. The experimental results are shown in Figure 21, Figure 22 and Figure 23; it is evident that as the order of the fractional-order forward converter decreases, the ripple of the inductor current and the output voltage ripple greatly increase. This aligns with the circuit simulation results of the forward converter. The order of the fractional-order component has a significant impact on the forward converter. As the order decreases, the ripple suppression capability of the energy storage component weakens.
The parameter Ki of the PI controller is used as the bifurcation parameter. By adjusting Ki, the low-frequency oscillation of the forward converter is observed. The experimental results are shown in Figure 21 and Figure 22. Figure 21 shows the experimental results of the forward converter with (α, β) = (1, 1). When Ki = 300, the circuit is in a stable state. As Ki increases to 340, the circuit begins to oscillate at a low frequency. Figure 22 shows the experimental results of the forward converter with (α, β) = (0.95, 0.95). As Ki increases from 870 to 910, the circuit transitions from a stable state to low-frequency oscillations. It can be observed that when the forward converter experiences low-frequency oscillations, the ripple of the circuit’s inductor current and output voltage significantly increase. Comparing the experimental results with the MATLAB/Simulink simulation results, the same conclusion can be drawn. Therefore, the reasonableness and correctness of the theoretical analysis are validated. When the order of the fractional-order forward converter decreases, the bifurcation point will move backward, and the parameter stability region of the system is greater, which means the circuit cannot easily enter the low-frequency oscillation state.
Utilizing the reference current as a parameter to describe bifurcations. By adjusting Iref, the period-doubling bifurcation phenomenon of the forward converter can be obtained. The experimental results are shown in Figure 23. Figure 23a,b shows the experimental results of the forward converter with (α, β) = (1, 1). When Iref = 5.6 A, the circuit is in a stable state. As Iref increases to 6.1 A, the circuit undergoes period-doubling bifurcation, transitioning from a period-1 state to a period-2 state. Figure 23c,d shows the experimental results of the forward converter with (α, β) = (0.95, 0.95). As Iref increases from 7.4 to 7.9 A, the circuit transitions from a stable state to period-doubling bifurcation. It can be observed that when the forward converter experiences period-doubling bifurcation, the inductor current undergoes fission, with the period changing from T to 2T. The experimental results closely correspond to the simulation outcomes, thereby affirming the accuracy and validity of theoretical analysis. As the order of the fractional forward converter decreases, the period-doubling bifurcation point moves backward. This implies that, compared to the integral-order forward converter, the fractional-order forward converter cannot easily enter into period-doubling bifurcation.

6. Conclusions

In this article, a fractional-order forward converter model is established based on the fractional calculus theory and state-space averaging method. The orders of the inductor and capacitor affect the performance of the system. As the orders decrease, the system responds faster and has smaller overshoot in transient processes, indicating a more better dynamic performance and stability. The reason for this is revealed in the right-shifted amplitude-frequency characteristic curve and the leading amplitude-frequency characteristic curve. As an integral-order forward converter, the fractional-order forward converter also exhibits nonlinear dynamical phenomena. However, the fractional-order forward converter cannot easily enter the states of bifurcation and chaos, and therefore has a wider range of control parameter values to stabilize the system. The bifurcation diagrams, time-domain simulation results, and phase diagrams all prove the correctness of analysis. The bifurcation boundary is predicted by using the fractional-order system stability determination method. The validity of theoretical analysis is further verified by a hardware-in-the-loop experimental study.

Author Contributions

X.W.: project administration, supervision, analysis and interpretation of results, visualization, investigation, analysis, revision, and editing. Z.Z.: methodology, study conception and design, software, analysis and interpretation of results, draft manuscript preparation, analysis, revision, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

Guangdong Basic and Applied Basic Research Foundation (2021A1515010616). Guangzhou Science and Technology Project (202102010404).

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. De Oliveira, E.C.; Tenreiro Machado, J.A. A review of definitions for fractional derivatives and integral. Math. Probl. Eng. 2014, 2014, 1–6. [Google Scholar] [CrossRef]
  2. Petráš, I. A note on the fractional-order Chua’s system. Chaos Solitons Fractals 2008, 38, 140–147. [Google Scholar] [CrossRef]
  3. Westerlund, S.; Ekstam, L. Capacitor theory. IEEE Trans. Dielectr. Electr. Insul. 1994, 1, 826–839. [Google Scholar] [CrossRef]
  4. Westerlund, S. Dead matter has memory! Phys. Scr. 1991, 43, 174–179. [Google Scholar] [CrossRef]
  5. Jesus, I.S.; Tenreiro Machado, J.A. Development of fractional order capacitors based on electrolyte processes. Nonlinear Dyn. 2009, 56, 45–55. [Google Scholar] [CrossRef]
  6. Jiang, Y.; Zhang, B. High-power fractional-order capacitor with 1<α<2 based on power converter. IEEE Trans. Ind. Electron. 2018, 65, 3157–3164. [Google Scholar]
  7. Wang, F.Q.; Ma, X.K. Fractional order buck–boost converter in CCM: Modelling, analysis and simulations. Int. J. Electron. 2014, 101, 1671–1682. [Google Scholar] [CrossRef]
  8. Sharma, M.; Rajpurohit, B.S.; Agnihotri, S.; Rathore, A.K. Development of fractional order modeling of voltage source converters. IEEE Access 2020, 8, 131750–131759. [Google Scholar] [CrossRef]
  9. Laugesen, J.L.; Mosekilde, E.; Zhusubaliyev, Z.T. Bifurcation structure of the -type period-doubling transition. Phys. D Nonlinear Phenom. 2012, 241, 488–496. [Google Scholar] [CrossRef]
  10. Xie, F.; Yang, R.; Zhang, B. Bifurcation and border collision analysis of voltage-mode-controlled Flyback converter based on total ampere-turns. IEEE Trans. Circuits Syst. I Regul. Pap. 2011, 58, 2269–2280. [Google Scholar] [CrossRef]
  11. Liu, J.; Xu, H. Nonlinear dynamic research of Buck converter based on multiscale entropy. In Proceedings of the 2015 Fifth International Conference on Instrumentation and Measurement, Computer, Communication and Control (IMCCC) 2015, Qinhuangdao, China, 18–20 September 2015; pp. 1042–1045. [Google Scholar]
  12. Aroudi, A.E.; Debbat, M.; Martinez-Salamero, L. Poincaré maps modeling and local orbital stability analysis of discontinuous piecewise affine periodically driven systems. Nonlinear Dyn. 2007, 50, 431–445. [Google Scholar] [CrossRef]
  13. Rodriguez-Amenedo, J.L.; Gomez, S.A. Damping low-frequency oscillations in power systems using grid-forming converters. IEEE Access 2021, 9, 158984–158997. [Google Scholar] [CrossRef]
  14. He, G.T.; Luo, M.K. Dynamic behavior of fractional order Duffing chaotic system and its synchronization via singly active control. Appl. Math. Mech. 2012, 33, 567–582. [Google Scholar] [CrossRef]
  15. Xie, L.; Liu, Z.; Zhang, B. A modeling and analysis method for CCM Ffractional order buck-boost converter by using R–L fractional definition. J. Electr. Eng. Technol. 2020, 15, 1651–1661. [Google Scholar] [CrossRef]
  16. Wang, F.Q.; Ma, X.K. Fractional order modeling and simulation analysis of boost converter in continuous conduction mode operation. Acta Phys. 2011, 60, 070506. [Google Scholar] [CrossRef]
  17. Tan, C.; Liang, Z.S. Modeling and simulation analysis of fractional-order boost converter in pseudo-continuous conduction mode. Acta Phys. 2014, 63, 070502. [Google Scholar]
  18. Kianpoor, N.; Yousefi, M.; Bayati, N.; Hajizadeh, A.; Soltani, M. Fractional order modelling of dc-dc boost converters. In Proceedings of the 2019 IEEE 28th International Symposium on Industrial Electronics (ISIE), Vancouver, BC, Canada, 12–14 June 2019; pp. 864–869. [Google Scholar]
  19. Chen, X.; Chen, Y.; Zhang, B.; Qiu, D. A modeling and analysis method for fractional-order dc–dc converters. IEEE Trans. Power Electron. 2017, 32, 7034–7044. [Google Scholar] [CrossRef]
  20. Fang, S.; Wang, X. Modeling and analysis method of fractional-order buck–boost converter. Int. J. Circuit Theory Appl. 2020, 48, 1493–1510. [Google Scholar] [CrossRef]
  21. El Aroudi, A.; Debbat, M.; Giral, R.; Olivar, G.; Benadero, L.; Toribio, E. Bifurcations in dc–dc switching converters: Review of methods and applications. Int. J. Bifurc. Chaos 2005, 15, 1549–1578. [Google Scholar] [CrossRef]
  22. Di Bernardo, M.; Garefalo, F.; Glielmo, L.; Vasca, F. Switchings, bifurcations, and chaos in dc/dc converters. IEEE Trans. Circuits Syst. I Fundam. Theory Appl. 1998, 45, 133–141. [Google Scholar] [CrossRef]
  23. Luo, Z.; Xie, F.; Zhang, B.; Qiu, D. Quantifying the nonlinear dynamic behavior of the dc-dc converter via permutation entropy. Energies 2018, 11, 2747. [Google Scholar] [CrossRef]
  24. Zhang, R.; Wu, A.; Zhang, S.; Wang, Z.; Cang, S. Dynamical analysis and circuit implementation of a dc/dc single-stage boost converter with memristance load. Nonlinear Dyn. 2018, 93, 1741–1755. [Google Scholar] [CrossRef]
  25. Wu, C.; Zhang, Q.; Yang, N.; Jia, R.; Liu, C. Dynamical analysis of a fractional-order boost converter with fractional-order memristive load. Int. J. Bifurc. Chaos 2022, 32, 2250032. [Google Scholar] [CrossRef]
  26. Jia, Z.; Liu, C. Fractional-order modeling and simulation of magnetic coupled boost converter in continuous conduction mode. Int. J. Bifurc. Chaos 2018, 28, 1850061. [Google Scholar] [CrossRef]
  27. Li, Z.; Chen, D.; Zhu, J.; Liu, Y. Nonlinear dynamics of fractional order Duffing system. Chaos Solitons Fractals 2015, 81, 111–116. [Google Scholar] [CrossRef]
  28. Yang, C.; Xie, F.; Chen, Y.; Xiao, W.; Zhang, B. Modeling and Analysis of the Fractional-order flyback converter in continuous conduction mode by Caputo fractional calculus. Electronics 2020, 9, 1544. [Google Scholar] [CrossRef]
  29. Uchaikin, V.V.; Cahoy, D.O.; Sibatov, R.T. Fractional processes: From Poisson to branching one. Int. J. Bifurc. Chaos 2008, 18, 2717–2725. [Google Scholar] [CrossRef]
  30. Podlubny, I. Fractional Differential Equations; Academic Press: San Diego, CA, USA, 1999. [Google Scholar]
  31. Yang, N.N.; Liu, C.X.; Wu, C.J. Modeling and dynamics analysis of the fractional-order Buck-Boost converter in continuous conduction mode. Chin. Phys. B 2012, 21, 080503. [Google Scholar] [CrossRef]
  32. Dwari, S.; Jayawant, S.; Beechner, T.; Miller, S.K.; Mathew, A.; Chen, M.; Riehl, J.; Sun, J. Dynamics characterization of coupled-inductor boost dc-dc converters. In Proceedings of the 2006 IEEE Workshops on Computers in Power Electronics, Troy, NY, USA, 16–19 July 2006; pp. 264–269. [Google Scholar]
  33. Charef, A.; Sun, H.H.; Tsao, Y.Y.; Onaral, B. Fractal system as represented by singularity function. IEEE Trans. Autom. Control 1992, 37, 1465–1470. [Google Scholar] [CrossRef]
  34. Valsa, J.; Vlach, J. RC models of a constant phase element. Int. J. Circuit Theory Appl. 2013, 41, 59–67. [Google Scholar] [CrossRef]
Figure 1. Circuit of forward converter and its operating principle. (a) Main circuit; (b) Waveforms of drive signal and inductor current.
Figure 1. Circuit of forward converter and its operating principle. (a) Main circuit; (b) Waveforms of drive signal and inductor current.
Fractalfract 08 00362 g001
Figure 2. Operating state diagram of the forward converter. (a) State 1, inductor charges; (b) State 2, inductor discharges.
Figure 2. Operating state diagram of the forward converter. (a) State 1, inductor charges; (b) State 2, inductor discharges.
Fractalfract 08 00362 g002
Figure 3. The structure of the fractional-order inductance impedance chain (L = 0.125 mH; α = 0.95).
Figure 3. The structure of the fractional-order inductance impedance chain (L = 0.125 mH; α = 0.95).
Fractalfract 08 00362 g003
Figure 4. The structure of the fractional-order capacitor impedance chain (C = 200 µF; β = 0.95).
Figure 4. The structure of the fractional-order capacitor impedance chain (C = 200 µF; β = 0.95).
Fractalfract 08 00362 g004
Figure 5. Bode diagrams of the constructed fractional-order components. (a) inductance with α = 0.95; (b) capacitance with β = 0.95.
Figure 5. Bode diagrams of the constructed fractional-order components. (a) inductance with α = 0.95; (b) capacitance with β = 0.95.
Fractalfract 08 00362 g005
Figure 6. Waveform diagrams of the simulation results for the forward converter. (a) iL; (b) vO.
Figure 6. Waveform diagrams of the simulation results for the forward converter. (a) iL; (b) vO.
Fractalfract 08 00362 g006
Figure 7. Bode diagram of the state-space average model of the forward converter with (α, β) = (1, 1) and (α, β) = (0.95, 0.95). (a) G i L v in s ; (b) G v o v in s .
Figure 7. Bode diagram of the state-space average model of the forward converter with (α, β) = (1, 1) and (α, β) = (0.95, 0.95). (a) G i L v in s ; (b) G v o v in s .
Fractalfract 08 00362 g007
Figure 8. Transient waveforms of circuit simulation results. (a) iL; (b) vo.
Figure 8. Transient waveforms of circuit simulation results. (a) iL; (b) vo.
Fractalfract 08 00362 g008
Figure 9. Block diagram of voltage-mode-controlled forward converter.
Figure 9. Block diagram of voltage-mode-controlled forward converter.
Fractalfract 08 00362 g009
Figure 10. Bifurcation diagrams (iL versus Ki) for the voltage-mode-controlled forward converter. (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Figure 10. Bifurcation diagrams (iL versus Ki) for the voltage-mode-controlled forward converter. (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Fractalfract 08 00362 g010
Figure 11. Waveforms of the forward converter at the bifurcation point (Ki as the bifurcation parameter). (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Figure 11. Waveforms of the forward converter at the bifurcation point (Ki as the bifurcation parameter). (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Fractalfract 08 00362 g011
Figure 12. Phase diagrams for the voltage-mode-controlled forward converter (Ki as the bifurcation parameter). (a,b) (α, β) = (1, 1); (c,d) (α, β) = (0.95, 0.95).
Figure 12. Phase diagrams for the voltage-mode-controlled forward converter (Ki as the bifurcation parameter). (a,b) (α, β) = (1, 1); (c,d) (α, β) = (0.95, 0.95).
Fractalfract 08 00362 g012
Figure 13. Block diagram of forward converter based on PI voltage feedback control.
Figure 13. Block diagram of forward converter based on PI voltage feedback control.
Fractalfract 08 00362 g013
Figure 14. Block diagram of current-mode-controlled forward converter.
Figure 14. Block diagram of current-mode-controlled forward converter.
Fractalfract 08 00362 g014
Figure 15. Bifurcation diagrams (iL versus Iref) for the current-mode-controlled forward converter. (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Figure 15. Bifurcation diagrams (iL versus Iref) for the current-mode-controlled forward converter. (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Fractalfract 08 00362 g015
Figure 16. Waveforms of the forward converter at the bifurcation point (Iref as the bifurcation parameter). (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Figure 16. Waveforms of the forward converter at the bifurcation point (Iref as the bifurcation parameter). (a) (α, β) = (1, 1); (b) (α, β) = (0.95, 0.95).
Fractalfract 08 00362 g016
Figure 17. Phase diagrams for the current-mode-controlled forward converter with (α, β) = (1, 1). (a) Iref = 5 A; (b) Iref = 6 A; (c) Iref = 7 A; (d) Iref = 8 A.
Figure 17. Phase diagrams for the current-mode-controlled forward converter with (α, β) = (1, 1). (a) Iref = 5 A; (b) Iref = 6 A; (c) Iref = 7 A; (d) Iref = 8 A.
Fractalfract 08 00362 g017
Figure 18. Phase diagrams for the current-mode-controlled forward converter with (α, β) = (0.95, 0.95). (a) Iref = 7 A; (b) Iref = 8.5 A; (c) Iref = 10 A; (d) Iref = 11 A.
Figure 18. Phase diagrams for the current-mode-controlled forward converter with (α, β) = (0.95, 0.95). (a) Iref = 7 A; (b) Iref = 8.5 A; (c) Iref = 10 A; (d) Iref = 11 A.
Fractalfract 08 00362 g018
Figure 19. Waveform diagram of the predictive correction model for the (α, β) = (1, 1) forward converter. (a) Iref = 5.4 A; (b) Iref = 5.7 A.
Figure 19. Waveform diagram of the predictive correction model for the (α, β) = (1, 1) forward converter. (a) Iref = 5.4 A; (b) Iref = 5.7 A.
Fractalfract 08 00362 g019
Figure 20. Waveform diagram of the predictive correction model for the (α, β) = (0.95, 0.95) forward converter. (a) Iref = 7.9 A; (b) Iref = 8.2 A.
Figure 20. Waveform diagram of the predictive correction model for the (α, β) = (0.95, 0.95) forward converter. (a) Iref = 7.9 A; (b) Iref = 8.2 A.
Fractalfract 08 00362 g020
Figure 21. Stable state and low-frequency oscillation of the forward converter with (α, β) = (1, 1). (a,b) (Ki = 300); (c,d) (Ki = 340).
Figure 21. Stable state and low-frequency oscillation of the forward converter with (α, β) = (1, 1). (a,b) (Ki = 300); (c,d) (Ki = 340).
Fractalfract 08 00362 g021
Figure 22. Stable state and low-frequency oscillation of the forward converter with (α, β) = (0.95, 0.95). (a,b) (Ki = 870); (c,d) (Ki = 910).
Figure 22. Stable state and low-frequency oscillation of the forward converter with (α, β) = (0.95, 0.95). (a,b) (Ki = 870); (c,d) (Ki = 910).
Fractalfract 08 00362 g022
Figure 23. Stable state and period-doubling bifurcation of the forward converter. (a,b) (α, β) = (1, 1); (c,d) (α, β) = (0.95, 0.95).
Figure 23. Stable state and period-doubling bifurcation of the forward converter. (a,b) (α, β) = (1, 1); (c,d) (α, β) = (0.95, 0.95).
Fractalfract 08 00362 g023
Table 1. The values of RLi, LLi, RCi, and CCi for fractional-order components with α = 0.95 and β = 0.95.
Table 1. The values of RLi, LLi, RCi, and CCi for fractional-order components with α = 0.95 and β = 0.95.
α/β0.950.95
iRLi (Ω)LLi (mH)RCi (Ω)CCi (mF)
in0.00000002/0.000000000399/
161.8750.065880.0000000005040.952
20.71250.009810.000000006141.021
30.058750.010380.00000007041.176
40.005130.011750.0000008041.296
50.000450.013380.000009021.492
60.000040.015130.0001041.694
70.00000350.017250.001161.924
80.00000030.019510.01422.041
90.0000000250.020631.240.304
Table 2. Simulation results of the forward converter.
Table 2. Simulation results of the forward converter.
ParameterSymbol(α, β) = (1, 1)(α, β) = (0.95, 0.95)
Maximum of inductor currentiL_max1.121 A4.204 A
Minimum of inductor currentiL_min0.274 A2.641 A
Maximum of output voltagevo_max3.517 V3.511 V
Minimum of output voltagevo_min3.464 V3.443 V
Table 3. The setting times, overshoot, peak time, and peak voltage of the output voltage.
Table 3. The setting times, overshoot, peak time, and peak voltage of the output voltage.
(α, β)Setting Time (ms)Overshoot (%)Peak Time (ms)Peak Voltage (V)
(1, 1)5.83480.910.48416.241
(0.95, 0.95)1.54136.310.48114.702
Table 4. The gain margin, phase margin, and crossover frequency of the (α, β) = (1, 1) and (α, β) = (0.95, 0.95) forward converter as Ki varies.
Table 4. The gain margin, phase margin, and crossover frequency of the (α, β) = (1, 1) and (α, β) = (0.95, 0.95) forward converter as Ki varies.
(α, β) = (1, 1)(α, β) = (0.95, 0.95)
KiA(fg) (dB)φ(fc) (°)fc(Hz)StateKiA(fg) (dB)φ(fc) (°)fc (Hz)State
2800.987.216251stable8450.334.6110,009stable
2900.684.936290stable8550.243.6310,080stable
3000.382.816331stable8650.132.2510,130stable
3100.111.246370stable8750.050.7410,195stable
320−0.23−1.256408unstable885−0.07−1.1210,260unstable
Table 5. Circuit parameters of the forward converter.
Table 5. Circuit parameters of the forward converter.
VariableSignificationValue
VinInput voltage10 V/5 V
fSwitching frequency10 kHz
LInductance0.125 mH
CCapacitance200 μF
RLoad resistance3.5 Ω/1 Ω
nTransformer ratio1
αInductor order0.95/1
βCapacitor order0.95/1
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

Wang, X.; Zhang, Z. Fractional-Order Modeling and Nonlinear Dynamic Analysis of Forward Converter. Fractal Fract. 2024, 8, 362. https://doi.org/10.3390/fractalfract8060362

AMA Style

Wang X, Zhang Z. Fractional-Order Modeling and Nonlinear Dynamic Analysis of Forward Converter. Fractal and Fractional. 2024; 8(6):362. https://doi.org/10.3390/fractalfract8060362

Chicago/Turabian Style

Wang, Xiaogang, and Zetian Zhang. 2024. "Fractional-Order Modeling and Nonlinear Dynamic Analysis of Forward Converter" Fractal and Fractional 8, no. 6: 362. https://doi.org/10.3390/fractalfract8060362

APA Style

Wang, X., & Zhang, Z. (2024). Fractional-Order Modeling and Nonlinear Dynamic Analysis of Forward Converter. Fractal and Fractional, 8(6), 362. https://doi.org/10.3390/fractalfract8060362

Article Metrics

Back to TopTop