Home Technology Variable sampling time discrete sliding mode control for a flapping wing micro air vehicle using flapping frequency as the control input
Article Open Access

Variable sampling time discrete sliding mode control for a flapping wing micro air vehicle using flapping frequency as the control input

  • Joshua Hill and Farbod Fahimi EMAIL logo
Published/Copyright: August 9, 2024
Become an author with De Gruyter Brill

Abstract

In this article, two main assumptions commonly used in the reported literature are relaxed. First, instead of a continuous time controller, a discrete time controller is proposed. Second, the controller uses the flapping frequency as opposed to flapping force/moments as the control input commonly assumed by other research groups. A discrete time controller is more suitable for implementation and deployment to most real flapping wing micro aerial vehicles (FWMAVs), in which a computer controls a sampled system and the flapping wings are driven by servomotors. A robust reaching law based on the discrete sliding mode control method is proposed to stabilize the vertical altitude of a simulated FWMAV. The continuous time model of the FWMAV is converted to a discrete time model, and the discrete time sliding mode controller is developed for the new form of the model. Using the variable sampling time, the robust discrete sliding mode controller developed in this article is able to control a simulated FWMAV to successfully stabilize its position in spite of a 16% difference in model parameters, as well as a 0.7 m/s wind gust. The controller was able to withstand the wind gust with negligible deviation from its desired position. These simulation results further show an input frequency range similar to that of the experimentally derived model previously reported in the literature.

1 Introduction

1.1 Literature survey

Micro aerial vehicles (MAVs) are a class of miniature air vehicles that have a wingspan of less than 15 cm. These vehicles show significant promise in sensing and information gathering capabilities in fields such as environmental monitoring and homeland security. Inspired by nature, flapping wing micro aerial vehicles (FWMAVs) have the potential to augment the current capabilities of MAVs in the military and civilian field due to their capacity to be more aerodynamically agile and efficient, while appearing benign. There has been significant research over the years to study and create robotic representations of what is found in nature, such as the herring gull inspired Festo SmartBird, the hummingbird inspired aerovironment nano Hummingbird, and the dragonfly inspired Festo BionicOpter. The aerodynamics of these FWMAVs are complex and nonlinear, and their small size makes them highly susceptible to external disturbances such as wind gusts. To overcome these challenges, robust control methods are necessary to overcome these complexities and achieve the desired flight characteristics.

Over the past decade, there has been much attention given to FWMAVs and flight control methodologies. Simple proportional-integral-derivative (PID) controllers designed as proof of concepts for new FWMAV designs to prove out attitude control and hovering [13], but require the system to train in an idealized environment with the inability to compensate for disturbances or uncertainties. A nonlinear proportional-derivative (PD) controller that uses flapping angle amplitude for attitude control was explored by Wang [4]. A linear Proportional-integral (PI) controller is presented by Gayango et al. [5]. Trajectory tracking control [6] requires much knowledge of the systems states.

The disturbance observer based control [7], while it can handle limited disturbance and uncertainty, needs a nominal PID controller. Adaptive control methods are attractive in that they can balance controller parameters to provide optimal performance based on particular situations, such as reaching and settling time. Neural networks [810] and other iterative learning controllers [11,12] require a high computational cost and require adequate training. Active disturbance rejection controllers, such as those presented by Liang et al. [13] and Feng et al. [14] are cumbersome to implement, requiring much rigor to derive. Model-free methods have advantages over model-based strategies, in that they require less information about the FWMAV model, and are improved on with adaption methods [15]. However, they also have disadvantages, such as higher computational loads due to less transparent fuzzy logic rules [16]. Sliding mode control (SMC) is one of the best known methods for handling disturbances and uncertainties and has been successfully used with a variety of adaptation methods to further improve the performance of FWMAVs [1719]. In addition, a discrete time model predictive control was simulated by Zheng et al. [20] and a continuous time observer-based adaptive control was simulated by Meng et al. [21].

In the study by Khosravi and Novinzadeh [22], a model-free adaptive control is augmented with a discrete sliding mode controller (DSMC) with good results. Others have also reported research on robust DSMC for tracking periodic signals and repetitive tasks [23,24], but have used different applications than the FWMAV application.

1.2 Contributions

The control modes, methods, inputs, and the test methods used in the cited literature are listed in Table 1 as a quick reference. It can be seen that the majority of the research work done on FWMAVs assume that (1) the moments can be used as the control input, and (2) the system is operating in the continuous time domain. Most work, that make the aforementioned two assumptions use simulations to demonstrate the effectiveness of their proposed method.

Table 1

Methodology comparison

Citation Control mode Control method Control input Test method
[1] Continuous PID Servo speed Experiment
[2] Continuous PID Servo speed Experiment
[3] Continuous PID Moments Experiment
[4] Continuous PD Flapping amplitude Simulation
[5] Continuous PI Thrust/pitch Simulation
[6] Continuous Linear state feedback Wing angles Simulation
[7] Continuous PID Servo speed Experiment
[8] Continuous NN adaptive Servo speed Simulation
[9] Continuous NN adaptive Servo speed Simulation
[10] Continuous PD Flapping amplitude Simulation
[11] Discrete Learning NN Moments Simulation
[12] Continuous LQR method Wind angles Simulation
Iterative learning tuning
[13] Continuous State feedback Moments Simulation
[14] Continuous Dynamic inversion Moments Simulation
[15] Discrete Model free adaptive Moments Simulation
[16] Continuous Adaptive fuzzy Flapping amplitude Simulation
[17] Continuous Adaptive sliding mode Moments Simulation
[18] Continuous Adaptive sliding mode Voltage Experiment
[19] Continuous Adaptive Voltage Experiment
[20] Discrete Predictive control Force/moments Simulation
[21] Continuous Adaptive Forces Simulation
[22] Discrete Sliding mode Amplitude/phase shift Simulation

These two assumptions break down for testing a majority of real FWMAVs. The reason for the breakdown of these assumptions is as follows. The means of the real world control input to most FWMAVs is changing the flapping frequency of the wings. There is no real FWMAV in which the flapping force and moment can be specified in continuous time as the control input because that is practically impossible.

Most FWMAV hardware platforms designed for size and weight reduction are incapable of controlling moments, but can control the flapping frequency by manipulating the speed (or input voltage) of a servomotor. This fact is supported by studying Table 1. It can be seen that the majority of the experimental work use servo speed or voltage (which changes the servo speed) as the control input.

In this work, first, a discrete time SMC is proposed using a linear plant. Second, the flapping frequency is proposed as the input to the control system via a virtual control that uses the nonlinear conversion to update the sampling time. The combination of the aforementioned two features is the core of the contribution of this work.

The proposed method in this publication eliminates the need for the aforementioned assumptions (i.e., using force/moment as control input and using a continuous time domain for control). First, we use the flapping frequency as the control input, which can be easily specified by changing the speed of the motor that generates the flapping motion. Second, we keep the frequency of flapping constant for a full flapping cycle. This allows for the accurate use of averaging the force/moment generated by a full cycle flapping with a constant frequency. The state sampling is done at the beginning of a flapping cycle. The controller determines the flapping frequency using a discrete-time feedback. This frequency is applied for a full flapping cycle, which constitutes a sampling duration. Then, at the end of the flapping cycle, the states and control are updated again, and the process continues. Third, since the frequency is kept constant during the complete cycle, we have to use a discrete-time controller, and a continuous-time controller is rendered useless!

Using a continuous time system means that the flapping frequency (related to servo speed) can change in the middle of a flapping cycle. As a consequence, the commonly used “averaging of the aerodynamic forces/moments” in FWMAV modeling (which is normally done for a full cycle) breaks down. With our proposed discrete time control method, the flapping frequency will change at the start of each cycle. Once a cycle starts, it will complete with a constant frequency. Then, the next cycle will have a new frequency determined by the feedback control. As a consequence, the assumption of “averaging the aerodynamic forces/moments” holds, which reduces the uncertainty terms in the model.

It should be noted that once the flapping frequency for the next step is determined by the discrete time SMC, the duration of the full flapping cycle is determined by T = 1 f , where T is the full cycle duration and f is the flapping frequency. Then, the sampling time for the current step is T , which varies as the controller calculates the input frequency through time. That is how a variable sampling time system is implemented.

In Section 2, the detailed derivation of the proposed discrete time SMC is presented. In Section 3, the implementation of the DSMC on the 2D flight of a FWMAV is demonstrated. In Section 4, the robustness in the face of system uncertainties and disturbances are discussed. In Section 5, some simulations results are presented. Finally, in Section 6, the article is concluded.

2 Derivation of the DSMC

The discrete time SMC is derived using the reaching law approach presented in [25]. Consider a discrete plant:

(1) X ( k + 1 ) = A X ( k ) + b u ( k ) ,

where X is a vector of length n , u is a scalar representing the control input, A is a n × n state matrix, and b is a n × 1 control parameter vector.

2.1 Reaching law design

The following are necessary for designing a reaching law that achieves a stable error behavior around the discrete sliding surface:

Requirement 1: The system trajectory shall move monotonically toward a switching plane s and cross it in finite time, regardless of the initial state. Requirement 2: After the system has crossed the switching plane, it shall cross the plane in each successive sampling step, resulting in a zigzagging motion centered on the switching plane. Requirement 3: The size of each successive zigzagging step shall not increase and shall remain within a bounded region around the switching plane.

For a discrete time system, the reaching law from Gao et al. [25] is considered

(2) s ( k + 1 ) s ( k ) = q T s ( k ) ε T sgn ( s ( k ) ) ,

where 0 < 1 q T < 1 , ε > 0 , and T > 0 is the sampling time.

This equation can be broken into its two parts. The first, ε T sgn ( s ( k ) ) , ensures that the switching variable, s ( k ) , reaches the switching manifold at a constant rate given by ε T . Additionally, due to the signum function, sgn , it will reverse, or switch, the direction on crossing the switching manifold. The second part, q T s ( k ) , is a proportional term that will force the state to approach the switching manifold faster when s ( k ) is larger. As long as the inequality 0 < 1 q T < 1 holds, it guarantees that the first requirement is satisfied. The latter two requirements are guaranteed by the signum properties in the discrete time domain.

Until now, the reaching law (2) guarantees that the surface variable s ( k ) approaches the surface and stays in its vicinity as time passes. While this is an excellent way to reach the surface, there must be guarantees that the states slide to the desired point after the surface has been reached. The next section discusses how this can be achieved.

2.2 Switching surface design

Consider a linear switching surface.

(3) s ( k ) = c T X ( k ) = 0 ,

where c is a n × 1 vector and c T b 0 . The surface Eq. (3) is selected linear in terms of the control gains c i . This is because a linear surface Eq. (3) leads to a linear Eq. (9) in terms of gains c i , which is easier to solve in terms of the gains.

As mentioned in the previous section, the reaching law (2) guarantees that the surface variable stays in the vicinity around the surface, which is referred to as the quasi-sliding mode (QSM). The ideal QSM satisfies the following condition:

(4) s ( k + 1 ) = s ( k ) = 0 , k = 0 , 1 , 2 .

This condition means that if s ( k ) is zero, then s ( k + 1 ) must also be zero, ensuring that s ( k ) remains zero through time. Merging (3) and (4), and noting (1), one can obtain:

(5) s ( k + 1 ) = c T X ( k + 1 ) = c T [ A X ( k ) + b u e ( k ) ] = 0 .

Here, u e ( k ) , the equivalent control, is used instead of u ( k ) , indicating the control effort needed at the ideal QSM.

Solving (5) for the equivalent control u e can be determined as follows:

(6) u e ( k ) = ( c T b ) 1 c T A X ( k ) .

The dynamic equation while on the ideal sliding mode can be found by plugging in the equivalent control (6) into the following dynamic Eq. (1):

(7) X ( k + 1 ) = A ˆ X ( k ) ,

where

(8) A ˆ = [ I b ( c T b ) 1 c T ] A .

The vector c needs to be determined to guarantee the stability of the ideal QSM. Note that the desired behavior of the system requires that the system states error be decreasing. This can be achieved by ensuring that the roots of the characteristic polynomial of the system in (7) are between 0 and 1. The characteristic polynomial of the system in (7) is found as follows:

(9) λ I [ I b ( c T b ) 1 c T ] A = 0 ,

where the gain c is designed such that the roots of the above polynomial are between 0 and 1.

2.3 Demonstration of the achievement of the QSM

The switching law is defined in Eq. (3) in Section 2.1. However, it is important to know the bounds of s ( k ) , or the width of the stabilization band achieved by the switching law. In this section, the bounds of s ( k ) or width of the stabilization band is determined.

Theorem 1

Consider the sliding mode reaching law (3). After the sliding variable s ( k ) crosses 0, it crosses 0 repeatedly in all future sampling times (defined as QSM), while it remains bounded within a band with a width equal to Δ = 2 ε T 1 q T around 0.

Proof

According to the definition of the QSM, the sign of s ( k + 1 ) must be opposite to that of s ( k ) , indicating the s trajectory crosses 0 at each sampling time. The band around the surface s ( k ) = 0 , where every state s ( k ) satisfies this condition, makes up the QSM band. The control law (21) is now substituted into the original system dynamics (1), and both sides of the resulting equation is multiplied by c T , giving the response of the discrete variable structure control (VSC):

(10) c T X ( k + 1 ) = c T [ A X ( k ) b ( c T b ) 1 [ c T A X ( k ) c T X ( k ) + q T c T X ( k ) + ε T sgn ( c T X ( k ) ) ] ] .

Noting that s ( k ) = c T X ( k ) , (10) becomes:

(11) s ( k + 1 ) = ( 1 q T ) s ( k ) ε T sgn ( s ( k ) ) , where 1 q T > 0 .

With the response of the controlled system found, the width of the QSM band needs to be determined. To determine the width of this band, consider the case, where s ( k ) > 0 , and the s trajectory crosses the surface s = 0 . It follows that s ( k + 1 ) < 0 . Using (11) and noting that sgn ( s ( k ) ) = 1 :

(12) ( 1 q T ) s ( k ) ε T < 0 .

Solving the aforementioned for s ( k ) gives:

(13) 0 < s ( k ) < ε T 1 q T .

Similarly, consider the case where s ( k ) < 0 , and the s trajectory crosses the surface s = 0 . It follows that s ( k + 1 ) > 0 . Using (11) and noting that sgn ( s ( k ) ) = 1 :

(14) ( 1 q T ) s ( k ) + ε T > 0 .

Solving the aforementioned for s ( k ) gives:

(15) ε T 1 q T < s ( k ) < 0 .

Finally, combining the results (13) and (15), the QSM region can be defined as follows:

(16) ε T 1 q T < s ( k ) < ε T 1 q T

or

(17) s ( k ) < ε T 1 q T .

The aforementioned equation proves the bounds of s ( k ) below and above zero. It can be seen that the width of this band around 0 is:

(18)□ Δ = ε T 1 q T ε T 1 q T = 2 ε T 1 q T .

The QSM bandwidth shown in (18) is the size of the region around the surface, where the s ( k ) dynamics will remain and chatter. Note that the size of the QSM band Δ can be arbitrarily reduced by choosing a very small ε , q , and sampling time T , such that Δ 0 . Since s ( k ) remains within Δ 0 , then:

(19) s ( k ) 0 .

Thus far, it has been shown that the chosen reaching law is stable around the discrete sliding surface and is bounded within a specific QSM band. The stability of the whole control law must be studied, which is discussed in the next section.

2.4 VSC design

In this section, the reaching law (2) and the surface stabilized with the specially designed c in Section 2.2 will be used to achieve an effective controller. The time step change of the sliding surface is used as a start point in the derivations:

(20) s ( k + 1 ) s ( k ) = c T X ( k + 1 ) c T X ( k ) .

The term X ( k + 1 ) is expanded using the plant’s dynamics (1). Then, the right-hand side of the resulting equation is replaced by the reaching law (2). Finally, solving for the control u ( k ) the following variable structure control law is synthesized:

(21) u ( k ) = ( c T b ) 1 [ c T A X ( k ) c T X ( k ) + q T s ( k ) + ε T sgn ( s ( k ) ) ] .

The VSC law presented in (21) stabilizes the system (1) in the QSM domain around the surface. This is proven in the following theorem.

Theorem 2

Consider a dynamic system represented by (1). If the switching law (3) is used and if c is designed such that the system in (7) is stable, the control law (21) stabilizes system (1) in the QSM domain around the surface.

Proof

Substituting control law (3) in system (21), we obtain:

X ( k + 1 ) = A X ( k ) + b ( c T b ) 1 [ c T A X ( k ) c T X ( k ) + q T s ( k ) + ε T sgn ( s ( k ) ) ] .

Multiplying both sides of the aforementioned equation by c T and rearranging some terms, we have:

c T X ( k + 1 ) = c T A X ( k ) c T [ b ( c T b ) 1 c T A X ( k ) ] + c T X ( k ) q T s ( k ) ε T sgn ( s ( k ) ) .

By using the definition of s ( k ) from (3), we write:

c T X ( k + 1 ) = c T [ I b ( c T b ) 1 c T ] A X ( k ) + s ( k ) q T s ( k ) ε T sgn ( s ( k ) ) .

According to the result of Theorem 2 shown in (18), presented in section 2.3, the steady-state value of s ( k ) can be arbitrarily reduced very close to zero by the choice of a very small ε , q , and T (i.e., s ( k ) 0 ). The aforementioned equation reduces to:

X ( k + 1 ) = [ I b ( c T b ) 1 c T ] A X ( k ) .

If c is designed correctly such that the aforementioned linear system is stable, X ( k + 1 ) 0 , when k . The correct design of c is discussed later in Section 4.3. This concludes the proof.□

The aforementioned results are valid only for an ideal system that is perfectly known, which is unreasonable in an actual system. Model parameter uncertainties and disturbances would cause the idealized controller to become unstable, and thus, a robust controller must take these factors into consideration to provide the desired system performance.

3 DSMC for robust control

A system with parameter perturbations and external disturbances will now be considered. Assume the previously described nominal system (1) but introduce system parameter uncertainty and a disturbance term:

(22) X ( k + 1 ) = A X ( k ) + Δ A X ( k ) + b u ( k ) + Δ b u ( k ) + g ( k ) .

Here, Δ A and Δ b represent the system parameter uncertainty, and g ( k ) is the disturbance. The terms Δ b u ( k ) and g ( k ) can be combined into a single uncertainty term f ( k ) , which gives:

(23) X ( k + 1 ) = A X ( k ) + Δ A X ( k ) + b u ( k ) + f ( k ) .

For simplicity, notations A ˜ and f ˜ are defined such that:

(24) Δ A = b A ˜ , f = b f ˜ ,

where A ˜ is a row vector and f ˜ is a scalar. From (24), (22) becomes:

(25) X ( k + 1 ) = A X ( k ) + b [ A ˜ X ( k ) + u ( k ) + f ˜ ( k ) ] .

The equation of the ideal QSM is unchanged from the nominal case despite the presence of perturbations and the disturbance. The control law from (21) is amended to cancel out the uncertain terms A ˜ X ( k ) + f ˜ ( k ) :

(26) u ( k ) = ( c T b ) 1 [ c T A X ( k ) c T X ( k ) + q T c T X ( k ) + ε T sgn ( c T X ( k ) ) ] A ˜ X ( k ) f ˜ ( k ) .

Here, A ˜ and f ˜ are not known a priori. Therefore, the aforementioned control (26) cannot be implemented in its current form. To create a usable control, assume that A ˜ and f ˜ can be replaced with sufficiently conservative terms A c and f c to maintain the discrete reaching condition. Eq. (26) will thus become:

(27) u ( k ) = ( c T b ) 1 [ c T A X ( k ) s ( k ) + q T s ( k ) + ε T sgn ( s ( k ) ) ] A c X ( k ) f c ( k ) .

The effect of the controller (27) with conservative uncertain terms on the system (25) with the actual uncertain terms is studied next.

Consider the surface:

(28) s ( k + 1 ) = c T X ( k + 1 ) .

We substitute X ( k + 1 ) using (25) while using the control u ( k ) determined from (27). This yields:

(29) s ( k + 1 ) = s ( k ) q T s ( k ) ε T sgn ( s ( k ) ) + c T b [ A c X ( k ) f c ( k ) + A ˜ X ( k ) + f ˜ ( k ) ] .

To capture the unknown term f ˜ and A ˜ and their conservative counterparts f c and A c in single terms, the following notations are defined:

(30) S ˜ ( k ) = c T b A ˜ X ( k ) S c ( k ) = c T b A c X ( k ) , F ˜ ( k ) = c T b f ˜ ( k ) F c ( k ) = c T b f c ( k ) .

Eq. (29) can now be represented succinctly as follows:

(31) s ( k + 1 ) s ( k ) = q T s ( k ) ε T sgn ( s ( k ) ) + S ˜ ( k ) S c ( k ) + F ˜ ( k ) F c ( k ) .

It is reasonable to assume that the unknown terms S ˜ and F ˜ are upper and lower bounded such that:

(32) S L S ˜ S U , F L F ˜ F U .

The choice of S c and F c is done in such a way to ensure that the incremental change in s ( k ) (i.e., s ( k + 1 ) s ( k ) ) has the opposite sign as that of s ( k ) . A practical choice for this condition is:

(33) if s ( k ) > 0 , set S c = S U and F c = F U if s ( k ) < 0 , set S c = S L and F c = F L .

To streamline the use of Eq. (33) and avoid “if clauses” in the control law, the following new notations are defined:

(34) S 1 = S U + S L 2 S 2 = S U S L 2 F 1 = F U + F L 2 F 2 = F U F L 2 .

Now, instead of using Eq. (33) to determine S c and F c at any given control sampling time, the following equations can be used, which have the same effect as Eq. (33).

(35) S c = S 1 + S 2 sgn ( s ( k ) ) , F c = F 1 + F 2 sgn ( s ( k ) ) .

Finally, substituting (35) into the control (27), the robust control can be expressed as follows:

(36) u ( k ) = ( c T b ) 1 [ c T A X ( k ) c T X ( k ) + q T s ( k ) + ε T sgn ( s ( k ) ) + S 1 + F 1 + ( S 2 + F 2 ) sgn ( s ( k ) ) ] .

4 DSMC design for the FWMAV

4.1 The FWMAV continuous time dynamic model

A theoretical FWMAV based on an experimentally derived dynamic model of a butterfly [26,27] is used to demonstrate the derived DSMC. A brief derivation of the FWMAV system and the conversion to discrete time will provide context for the results.

The free-body diagram of the FWMAV is shown in Figure 1. The x z coordinates are the body axes of the FWMAV. In Figure 1, U is the forward velocity, L is the lift generated by the flapping wings, m g is the weight of the FWMAV, and θ is the angle between the lift and the weight forces.

Figure 1 
                  Forces acting on a FWMAV in a climbing trajectory.
Figure 1

Forces acting on a FWMAV in a climbing trajectory.

According to Figure 1, the balance of forces along the z axis is written as follows:

(37) m z ¨ = L m g cos ( θ ) ,

where g is the gravitational constant, z ¨ is the acceleration of the body, m is the mass of the butterfly, and θ is the climb angle with respect to the horizon. The lift, L , is modeled with Theodorsen’s lift equation in the QSM, and is expanded as follows:

(38) L = 1 2 ρ S r ˆ 2 2 C L = L = 1 2 ρ S r 2 ˆ 2 × π c m 2 U 2 ( z ¨ h ¨ ) + 2 π U ( z ˙ h ˙ ) + C L , α .

This lift term shows the coupling between the body’s velocity z ˙ and acceleration z ¨ , and the wing’s velocity h ˙ and acceleration h ¨ . It includes, as constants, the fluid density, ρ , mean velocity magnitude, U , area of the left and right forewings, S , nondimensional second moment of inertia of the wing area, r ˆ 2 2 , mean chord c m , and the lift coefficient due to a mean angle of attack, C L , α .

The motion of the wing is given as follows:

(39) h ( t ) = h a cos ( 2 π f t ) ,

where h a is the wing plunge amplitude, f is the flapping frequency, and t is time. For the sake of simplicity, the constant terms are consolidated, and the lift equation now becomes:

(40) L = C 1 ( C 2 ( z ¨ h ¨ ) + C 3 ( z ˙ h ˙ ) + C L , α ) ,

where

C 1 = 1 2 ρ U 2 S r ˆ 2 2 , C 2 = π c m 2 U 2 , C 3 = 2 π U ,

Substituting the lift equation back into the force balance equation and solving for z ¨ yields

(41) z ¨ = C 1 C 3 m C 1 C 2 z ˙ C 1 C 2 m C 1 C 2 h ¨ C 1 C 3 m C 1 C 2 h ˙ + C 1 C L , α m g cos ( θ ) m C 1 C 2 .

There is a lift component C L , α ρ U 2 S r ˆ 2 2 2 that always offsets the weight component, m g cos θ . This lift component is modeled as C L , α = 2 π α , where α is the mean angle of attack, and is equal to α = m g cos θ ( ρ U 2 S r ˆ 2 2 ) [27]. Substituting the derivatives of (39) into the aforementioned force balance equation gives:

(42) z ¨ = C 5 z ˙ + 4 W 1 π 2 f 2 h a cos ( 2 π f t ) + 2 W 2 π f h a sin ( 2 π f t ) + C 8 + C 9 ,

where

C 5 = C 1 C 3 m C 1 C 2 .

W 1 = C 1 C 2 m C 1 C 2 , W 2 = C 1 C 3 m C 1 C 2 .

C 8 = C 1 C L , α m C 1 C 2 , C 9 = m g cos θ m C 1 C 2 .

For simplicity, the frequency control terms are grouped into a single variable, w ( t ) .

(43) w ( t ) = 4 W 1 π 2 f 2 h a cos ( 2 π f t ) + 2 W 2 π f h a sin ( 2 π f t ) + C 9 .

This produces the continuous time linear plant that serves as the starting point for discrete time conversion.

(44) z ¨ = C 5 z ˙ + w ( t ) + C 8 .

To simplify the notation, a virtual control v ( t ) = w ( t ) + C 8 is introduced. Eq. (44) becomes:

(45) z ¨ = C 5 z ˙ + v ( t ) .

4.2 Discretizing the FWMAV’s dynamic model

The linear plant (45) is discretized using the Euler method. The position, velocity, and acceleration terms are discretized as follows:

(46) z ( k ) z d ( k ) = x 2 ( k ) z ˙ ( k ) z ˙ d ( k ) = x 2 ( k ) x 2 ( k 1 ) T z ¨ ( k ) z ¨ d ( k ) = 1 T x 2 ( k + 1 ) x 2 ( k ) T x 2 ( k ) x 2 ( k 1 ) T ,

where z d ( k ) is the desired position of the FWMAV at instance k . Plugging equations (46) into the system (42) and solving for x 2 ( k + 1 ) yields:

(47) x 2 ( k + 1 ) = ( 2 + T C 5 ) x 2 ( k ) ( 1 + T C 5 ) x 2 ( k 1 ) + T 2 u ( k ) ,

where

(48) u ( k ) = v ( k ) + C 5 z ˙ d ( k ) z ¨ d ( k ) .

Set

(49) x 1 ( k ) = x 2 ( k 1 ) .

The aforementioned equation is substituted into (47), and the result is written in the matrix form:

(50) x 1 ( k + 1 ) x 2 ( k + 1 ) = 0 1 ( 1 + T C 5 ) ( 2 + T C 5 ) × x 1 ( k ) x 2 ( k ) + 0 T 2 u ( k ) ,

which is the standard form of

(51) X ( k + 1 ) = A X ( k ) + b u ( k ) .

4.3 Designing the surface gain vector c

Following Section 2.2, the vector c needs to be determined to guarantee the stability of the ideal QSM. Recall the dynamic Eq. (7) while on the ideal QSM. The vector c can be determined by first defining the plant’s parameters on the ideal QSM as A ˆ , as in (7), but with the FWMAV system as follows:

(52) A ˆ = I 0 T 2 c 1 c 2 0 T 2 1 c 1 c 2 0 1 ( 1 + T C 5 ) ( 2 + T C 5 ) .

With A ˆ defined, the dynamics of the FWMAV on the ideal sliding mode can be described as follows:

(53) X ( k + 1 ) = A ˆ X ( k ) = 0 1 0 c 1 c 2 X ( k ) .

The characteristic equation of the system to-be designed is determined.

(54) λ I 0 1 0 c 1 c 2 λ 1 0 λ + c 1 c 2 λ 2 + c 1 c 2 λ + 0 .

To ensure that the characteristic equation in (54) results in a stable system (53), we design the elements of the c matrix as follows:

(55) c 1 = L 1 and c 2 = 1 ,

where 0 < L 1 < 1 . The gains c 1 and c 2 in (55) have been designed to form a stable sliding mode for the FWMAV system. The stability of the system with the aforementioned gains is proven in the following theorem.

Theorem 3

Consider the dynamics of the FWMAV in the ideal sliding mode, as described in (53). If the gains c 1 and c 2 are determined by (55), where 0 < L 1 < 1 , the vector X ( k + 1 ) approaches a zero vector from any initial condition X ( 0 ) as k increases.

Proof

Substituting (55) in (53) results in:

X ( k + 1 ) = 0 1 0 L 1 X ( k ) .

Substituting X ( k ) = [ x 1 ( k ) , x 2 ( k ) ] T in the aforementioned equation yields:

(56) x 1 ( k + 1 ) = x 2 ( k ) x 2 ( k + 1 ) = L 1 x 2 ( k ) .

The second equation in (56) ensures that all x 2 ( k + 1 ) are smaller than x 2 ( k ) , because 0 < L 1 < 1 . This means that x 2 is reduced at any step and approaches zeros as k increases. The first equation in (56) shows that x 1 ( k + 1 ) is reduced to zero as x 2 approaches zero. So, both x 1 and x 2 approach zero, and consequently, the vector X ( k ) approaches a zero vector as k increases.□

Finally, according to Theorem 3, the coefficients of c 1 and c 2 have been designed to form a stable sliding mode for the FWMAV system.

4.4 Bounds of the FWMAV system uncertainty and disturbance

Next, the control parameters that sufficiently bound the state parameter uncertainty and disturbance terms need to be determined. In Section 3, state parameter uncertainty and a disturbance term were incorporated in (23):

(57) X ( k + 1 ) = A X ( k ) + Δ A X ( k ) + b u ( k ) + f ( k ) .

For the FWMAV system, the previous equation is explicitly:

(58) x 2 ( k + 1 ) = ( 2 + T ( C 5 + Δ C 5 ) ) x 2 ( k ) ( 1 + T ( C 5 + Δ C 5 ) ) x 1 ( k ) + T 2 u ( k ) + d ( k ) ,

which can be re-written as follows:

(59) x 2 ( k + 1 ) = ( 2 + T C 5 ) x 2 ( k ) ( 1 + T C 5 ) x 1 ( k ) + T 2 [ A 1 ˜ x 1 ( k ) + A 2 ˜ x 2 ( k ) + u ( k ) + d ˜ ( k ) ] ,

with

T Δ C 5 = T 2 A 1 ˜ , T Δ C 5 = T 2 A 2 ˜ , d ( k ) = T 2 d ˜ ( k ) .

Following the procedure laid out in Section 3, the control can be ascertained:

(60) u ( k ) = 1 T 2 ( 1 q T ) s ( k ) ε T sgn ( s ( k ) ) c 1 x 2 ( k ) c 2 ( 2 + T C 5 ) x 2 ( k ) + ( 1 + T C 5 ) x 1 ( k ) ] A 1 c x 1 ( k ) A 2 c x 2 ( k ) D c ( k ) ,

where A ˜ 1 , A ˜ 2 , and d ˜ ( k ) are replaced by their conservative counterparts, A 1 c , A 2 c , and D c ( k ) . Also, the values of S ˜ ( k ) and F ˜ ( k ) are as follows:

(61) S ˜ ( k ) = c 2 T 2 ( A 1 ˜ x 1 ( k ) + A 2 ˜ x 2 ( k ) ) , F ˜ ( k ) = c 2 T 2 d ˜ ( k ) .

The aforementioned equations are used to determine the bounds of uncertainty and disturbance.

(62) S L c 2 T 2 ( A 1 ˜ x 1 ( k ) + A 2 ˜ x 2 ( k ) ) S U F L c 2 T 2 d ˜ ( k ) F U ,

where S 1 , S 2 , F 1 , and F 2 are determined from Eqs (34) and (35). Finally, the control for the FWMAV system is expressed as follows:

(63) u ( k ) = 1 T 2 c 2 [ ( 1 q T ) s ( k ) ε T sgn ( s ( k ) ) c 1 x 2 ( k ) + c 2 [ ( 1 + T C 5 ) x 1 ( k ) ( 2 + T C 5 ) x 2 ( k ) ] + ( S 1 + F 1 ) + ( S 2 + F 2 ) sgn ( s ( k ) ) ] .

As this represents a “virtual control” that does not directly translate to the flapping frequency, a conversion must be formulated to realize this physical control input.

4.5 Virtual control to physical control conversion

The nonlinear control terms are grouped as shown in (43), which is repeated here for simplicity.

(64) w ( t ) = 4 W 1 π 2 f 2 h a cos ( 2 π f t ) + 2 W 2 π f h a sin ( 2 π f t ) + C 9 .

In the discrete model, v ( k ) is constant during each sampling period, T . As such, v ( k ) must have the same effect as the right-hand side of (64) during the sampling time on the FWMAV’s body displacement, z .

As shown in (44) and (45), z ¨ is a function of v ( t ) , and therefore, a function of w ( t ) . From this, a relationship for the flapping frequency f as a function of the control v in discrete time can be surmised by twice integrating (64).

After the first integration of (64), it is noted that w ( t ) at t = 0 is C 9 , which yields:

(65) ( w C 9 ) t = 2 W 1 π f h a sin ( 2 π f t ) W 2 h a cos ( 2 π f t ) + C 9 t + W 2 h a .

The second integration is performed over the sampling time T :

(66) 1 2 ( w 2 C 9 2 ) T 2 = W 1 h a cos ( 2 π f T ) + W 1 h a W 2 h a 2 π f sin ( 2 π f T ) + 1 2 C 9 T 2 + W 2 h a T .

Since T = 1 f , Eq. (66) can be reduced and the frequency term f can be solved for:

(67) f ( t ) = w ( t ) 2 C 9 2 W 2 h a .

Noting that w ( t ) = v ( t ) C 8 , Eq. (67) can finally result in the direct relation between the virtual control u and the flapping frequency f .

(68) f ( t ) = v ( t ) C 8 2 C 9 2 W 2 h a .

5 Numerical simulation and results

For the simulation, the system will begin with an initial condition of x 2 ( 0 ) = 0.2 m with a variable time step. The simulated FWMAV has the nominal system parameters shown in Table 2.

Table 2

Nominal system parameters

Parameter Value
Mass (m) 0.415 g
Wing area (A) 19 mm2
Drag coefficient ( C D ) 1.28
Air density ( ρ ) 1.205 kg/m3
Wing plunge amplitude ( h a ) 0.0324 m
System constant C 5 10.124 s 1
System constant C 8 9.191 m/s 2
System constant C 9 9.191 m/s2
System constant W 2 10.125 s 1

The control parameters selected for the DSMC are shown in Table 3.

Table 3

DSMC parameters

Parameter Value
L 1 0.4
L 2 0
c 1 1
c 2 0.4
ε 0.005
q 3
S U 0.0008 m
S L 0.0008 m
F U 0.002 m
F L 0.002 m

To simulate system parameter variation, an additive system perturbation of 16% above nominal system parameters is included. The simulated disturbance is the resulting acceleration due to drag on the FWMAV wings due to a gust of wind.

(69) a D = 1 2 m ρ C D A v 2 .

This gust of wind has a maximum velocity of 0.7 m/s, beginning 10 s into the simulation, and ending 5 s later, as shown in the following equation and graphically in Figure 2:

(70) v ( t ) = 0.28 t 2.8 10 t < 12.5 0.28 t + 4.2 12.5 t 15 .

Figure 2 
               Wind gust profile.
Figure 2

Wind gust profile.

The FWMAV has a mass of m = 0.415 g, with wing area of A = 19 mm 2 , treated as a flat plate with a drag coefficient of C D = 1.28 , in air with density ρ = 1.205 kg m 3 .

Two controllers are tested via simulations. The first one is the robust DSMC proposed in this article (control law (35)). The second one is a nonrobust (standard) DSMC (control law (21)).

In Figures 3 and 4, it can be seen that the FWMAV begins at an initial height of 0.2 m and flies to the desired sinusoidal trajectory with amplitude 0.15 m and period of 30 s. At t = 10 s , the disturbance begins and the FWMAV using the robust DSMC in Figure 3 maintains the trajectory, with little to no trajectory deviation. This is in contrast to the FWMAV using the nonrobust DSMC in Figure 4, where the FWMAV clearly deviates from the desired trajectory before returning to the desired trajectory after the disturbance ends at t = 15  s.

Figure 3 
               Robust DSMC: Results for vertical displacement of the FWMAV 
                     
                        
                        
                           z
                        
                        z
                     
                  .
Figure 3

Robust DSMC: Results for vertical displacement of the FWMAV z .

Figure 4 
               Nonrobust DSMC: Results for vertical displacement of the FWMAV 
                     
                        
                        
                           z
                        
                        z
                     
                  .
Figure 4

Nonrobust DSMC: Results for vertical displacement of the FWMAV z .

Similarly, Figures 5 and 6 show the corresponding trajectory tracking error for the robust and nonrobust DSMC controllers, respectively. In these figures, it can be seen that the maximum tracking error for the worst case disturbance for the nonrobust and robust DSMC are 0.049 m and 0.027 m, respectively. This demonstrates the superiority of the proposed robust DSMC, which has reduced the maximum tracking error by approximately 45%.

Figure 5 
               Robust DSMC: Results for vertical displacement error of the FWMAV 
                     
                        
                        
                           z
                        
                        z
                     
                  –
                     
                        
                        
                           
                              
                                 z
                              
                              
                                 d
                              
                           
                        
                        {z}_{d}
                     
                  .
Figure 5

Robust DSMC: Results for vertical displacement error of the FWMAV z z d .

Figure 6 
               Nonrobust DSMC: Results for vertical displacement error of the FWMAV 
                     
                        
                        
                           z
                        
                        z
                     
                  –
                     
                        
                        
                           
                              
                                 z
                              
                              
                                 d
                              
                           
                        
                        {z}_{d}
                     
                  .
Figure 6

Nonrobust DSMC: Results for vertical displacement error of the FWMAV z z d .

The error reduction is supported by observing the performance of the switching variable s in Figure 7, noting that it very quickly transitions from the reaching phase to the sliding phase and remains stable despite the presence of the disturbance. Note that, as proved in Section 2.3, the sliding variable s remains in a bound around the surface s = 0 . The sliding variable’s band is in the range of [ 0.004 , 0.018 ] m.

Figure 7 
               Robust DSMC: Results for the hyperplane, 
                     
                        
                        
                           s
                        
                        s
                     
                  .
Figure 7

Robust DSMC: Results for the hyperplane, s .

Looking at the control effort in Figure 8, one can see that the FWMAV control input smoothly and quickly reaches a steady-state response, increasing its relative effort to compensate for the disturbance during the disturbance presence period. The steady-state control input fluctuates within the range of [ 6.1 , 2.5 ] m/s 2 .

Figure 8 
               Robust DSMC: Results for control effort, 
                     
                        
                        
                           v
                        
                        v
                     
                  .
Figure 8

Robust DSMC: Results for control effort, v .

Converting the control effort to flapping frequency using Eq. (68), one can show that the flapping frequency maintains a frequency range of [ 10.7 , 23.4 ] Hz. The resulting input frequency range is similar to that of the experimentally derived model by Kang et al. [27]. This is an expected result. The FWMAV dynamic model used in our simulation is adopted from system identification of a monarch butterfly presented by Kang et al. [27]. Our defined desired flight regime is defined to closely match the flights used for system identification by Kang et al. [27]. Therefore, it is expected that the frequency control effort calculated by the DSMC controller must match the open-loop flapping frequency measured by experiments by Kang et al. [27]. The matching frequencies between the two publications further verifies the validity of the proposed DSMC controller.

Note that the input chatter (flapping frequency f ) that is shown in the result of the application of the DSMC on the FWMAV (Figure 9) is misleading. In each sampling time, an input frequency f is calculated, with its duration of application T = 1 f . Then, one cycle of wing flap following Eq. (39) is applied for T seconds. The wing flap h ( t ) starts at 0 ( h ( t ) = 0 ) at the current time t , it lasts T seconds, and at the end of T seconds, it ends a 0 ( h ( t + T ) = 0 ). Then, at time t + T , the same flapping pattern repeats itself. So, there is no discontinuity in the wing flapping motion. In consequence, there is no chatter in the wing flapping motion. The wing flapping motion consists of a series of continuous curves in the harmonic shape of Eq. (39) with different frequencies.

Figure 9 
               Robust DSMC: Results for flapping frequency, 
                     
                        
                        
                           f
                        
                        f
                     
                  .
Figure 9

Robust DSMC: Results for flapping frequency, f .

As a direct result of the changing flapping frequencies, the sampling time must change throughout the simulation. For higher flapping frequencies, the time between subsequent wing flaps naturally shortens. Conversely, for lower flapping frequencies, the time between wing flaps lengthens. The change in the sampling time T is shown in Figure 10. At the steady-state, the sampling time fluctuates in the range of [ 0.043 , 0.097 ] s. However, the sampling time trends lower as the flapping frequency increases in the duration of disturbances.

Figure 10 
               Robust DSMC: Results for variable sampling time, 
                     
                        
                        
                           T
                        
                        T
                     
                  .
Figure 10

Robust DSMC: Results for variable sampling time, T .

6 Conclusion

In this article, two main assumptions commonly used in the reported literature are relaxed. First, instead of a continuous time controller, a discrete time controller is proposed. Second, the controller uses the flapping frequency as opposed to flapping force/moments as the control input commonly assumed by other research groups. A discrete time controller is more suitable for implementation and deployment to most real FWMAVs, in which a computer controls a sampled system and the flapping wings are driven by servomotors.

A robust discrete sliding mode reaching controller was proposed to stabilize the vertical altitude of a FWMAV. The continuous time plant of the FWMAV was converted to a discrete time plant and the controller was developed for this new form. As can be seen by the results presented in this article, the FWMAV was able to successfully stabilize its position with negligible error despite a plant uncertainty of 16% and a wind gust with a maximum peak of 0.7 m/s. These trajectory and error results are compared to that of a nonrobust DSMC in Figures 36. The robust controller performance is further bolstered by Figure 9, which shows an input frequency range of [10.7, 23.4] Hz during the steady state, which are similar to that of the experimentally derived model by Kang et al. [27], thus validating the controller.

  1. Funding information: This research was not funded by any funding agency.

  2. Author contributions: The research idea and the overall methodology were formulated by J.H. The detailed derivations of the approach and the creation of the computer simulation code were performed by J.H., with suggested corrections in the derivations by F.F and suggested improvements to the simulations. The draft manuscript was authored by J.H., which went through several revisions according to feedback from F.F. All authors have accepted responsibility for the entire content of this manuscript and approved its submission. The first author was a PhD Candidate who conducted this research. The second author was the first author’s dissertation research advisor.

  3. Conflict of interest: The authors state that there is no conflict of interest.

  4. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] Benyamin RSB, Nugroho MAA, Sepri AY, Trilaksono BR, Moelyadi A. Dynamic modeling using system identification and attitude control design of GaneFly flapping wings micro aerial vehicle (FWMAV). In: 2017 7th IEEE International Conference on System Engineering and Technology (ICSET); 2017. p. 200–5. 10.1109/ICSEngT.2017.8123446Search in Google Scholar

[2] Xiao S, Deng H, Hu K, Zhang S. Design and control of hoverable bionic flapping wing micro air vehicle. In: 2021 IEEE International Conference on Electrical Engineering and Mechatronics Technology (ICEEMT); 2021. p. 224–7. 10.1109/ICEEMT52412.2021.9601914Search in Google Scholar

[3] Phan HV, Kang T, Park H. Controlled hovering flight of an insect-like tailless flapping-wing micro air vehicle. In: 2017 IEEE International Conference on Mechatronics (ICM); 2017. p. 74–78. 10.1109/ICMECH.2017.7921083Search in Google Scholar

[4] Wang L, Jiang W, Zhao L, Jiao Z. Modeling and Hover flight control of a micromechanical flapping-wing aircraft inspired by Wing-Tail interaction. IEEE/ASME Trans Mechatron. 2023;28(6):3132–42. 10.1109/TMECH.2023.3254921Search in Google Scholar

[5] Gayango D, Salmoral R, Romero H, Carmona JM, Suarez A, Ollero A. Benchmark evaluation of hybrid fixed-flapping wing aerial robot with autopilot architecture for autonomous outdoor flight operations. IEEE Robot Autom Lett. 2023;8(7):4243–50. 10.1109/LRA.2023.3280753Search in Google Scholar

[6] Wissa BE, Youssef AW, El-Badawy AA. Two-DoF controller design for a six-DoF flapping wing micro aerial vehicle. In: 2020 2nd Novel Intelligent and Leading Emerging Sciences Conference (NILES); 2020. p. 308–14. 10.1109/NILES50944.2020.9257957Search in Google Scholar

[7] Lee J, Ryu S, Kim HJ. Stable flight of a flapping-wing micro air vehicle under wind disturbance. IEEE RA-L. 2020;5(4):5685–92. 10.1109/LRA.2020.3009064Search in Google Scholar

[8] Mousavi SMS, Pourtakdoust SH. Improved neural adaptive control for nonlinear oscillatory dynamic of flapping wings. J Guid Control Dyn. 2023;46(1):97–113. 10.2514/1.G006478Search in Google Scholar

[9] He W, Mu X, Zhang L, Zou Y. Modeling and trajectory tracking control for flapping-wing micro aerial vehicles. IEEE/CAA JAS. 2021;8(1):148–56. 10.1109/JAS.2020.1003417Search in Google Scholar

[10] Mu X, He X, Zhang L, Fu Q, He W. Modeling and trajectory generation of bird-like flapping-wing micro air vehicle. In: 2020 Chinese Automation Congress (CAC); 2020. p. 3832–7. 10.1109/CAC51589.2020.9327362Search in Google Scholar

[11] He W, Meng T, He X, Sun C. Iterative learning control for a flapping wing micro aerial vehicle under distributed disturbances. IEEE Trans Cybern. 2019;49(4):1524–35. Search in Google Scholar

[12] Du C. Application of iterative learning tuning to a dragonfly-like flapping wing micro aerial vehicle. IEEE Trans Cybern. 2019;49(4):1524–35. 10.1109/TCYB.2018.2808321Search in Google Scholar PubMed

[13] Liang S, Song B, Xuan J. Active disturbance rejection attitude control for a bird-like flapping wing micro air vehicle during automatic landing. IEEE Access. 2020;8(1):171359–72. 10.1109/ACCESS.2020.3024793Search in Google Scholar

[14] Feng C, Wang Q, Hu C, Gong L. Active disturbance rejection attitude control for flapping wing micro aerial vehicle with nonaffine-in-control characteristics. IEEE Access. 2020;8(1):20013–27. 10.1109/ACCESS.2020.2968482Search in Google Scholar

[15] Wang T, Jin S, Hou Z. Model free adaptive pitch control of a flapping wing micro aerial vehicle with input saturation. In: 2020 IEEE 9th Data Driven Control and Learning Systems Conference (DDCLS); 2020. p. 627–32. 10.1109/DDCLS49620.2020.9275203Search in Google Scholar

[16] Ferdaus M, Anavatti SG, Al-Mahasneh AJ, Pratama M, Garratt MA. Development of hyperplane-based adaptive T-S fuzzy controller for micro aerial robots. In: 2019 IEEE International Conference on Industry 4.0, Artificial Intelligence, and Communications Technology (IAICT); 2019. p. 50–6. 10.1109/ICIAICT.2019.8784726Search in Google Scholar

[17] Banazadeh A, Taymourtash N. Adaptive attitude and position control of an insect-like flapping wing air vehicle. Nonlin Dyn. 2016;85(7):47–66. 10.1007/s11071-016-2666-8Search in Google Scholar

[18] Tu Z, Fei F, Liu L, Zhou Y, Deng X. Flying with damaged wings: the effect on flight capacity and bio-inspired coping strategies of a flapping wing robot. IEEE Robot Autom Lett. 2021;6(2):2114–21. 10.1109/LRA.2021.3059626Search in Google Scholar

[19] Jimbo T, Ozaki T, Amano Y, Fujimoto K. Flight control of flapping-wing robot with three paired direct-driven piezoelectric actuators. IFAC-Papers OnLine. 2020;53(2):9391–8. 10.1016/j.ifacol.2020.12.2408Search in Google Scholar

[20] Zheng H, Chen W, Xie F. Control simulation of flapping-wing micro aerial vehicle based on multi-level optimization model predictive control. IEEE Access. 2024;12(1):40700–9. 10.1109/ACCESS.2024.3376646Search in Google Scholar

[21] Meng T, Zhang Y, Fu Q, Wang J. Observer-based adaptive control for a coupled PDE-ODE system of a flexible wing. IEEE Trans Syst Man Cybern Syst. 2024;1(1):1–11. 10.1109/TSMC.2024.3376459Search in Google Scholar

[22] Khosravi M, Novinzadeh AB. A multi-body control approach for flapping wing micro aerial vehicles. Aerosp Sci Technol. 2021;112(1):106525. 10.1016/j.ast.2021.106525Search in Google Scholar

[23] Kurniawan E, Wang H, Sirenden BH, Prakosa JA, Adinanta H, Suryadi S. Discrete-time modified repetitive sliding mode control for uncertain linear systems. Int J Adapt Control. 2021;35(11):2245–58. 10.1002/acs.3316Search in Google Scholar

[24] Mitrevska M, Cao Z, Zheng J, Kurniawan E, Man Z. Discrete terminal sliding mode repetitive control for a linear actuator with nonlinear friction and uncertainties. Int J Robust Nonlin Control. 2019;29(13):4285–97. 10.1002/rnc.4639Search in Google Scholar

[25] Gao W, Wang Y, Homaifa A. Discrete-time variable structure control systems. IEEE Trans Ind Electron. 1995;42(2):117–22. 10.1109/41.370376Search in Google Scholar

[26] Sridhar M, Kang CK, Landrum DB. Beneficial effect of the coupled wing-body dynamics on power consumption in butterflies. In: 2019 AIAA Scitech Forum; 2019. p. AIAA–2019–0566. 10.2514/6.2019-0566Search in Google Scholar

[27] Kang CK, Cranford J, Sridhar MK, Kodali D, Landrum DB, Slegers N. Experimental characterization of a butterfly in climbing flight. AIAA J. 2018;56(1):15–24. 10.2514/1.J055360Search in Google Scholar

Received: 2024-02-12
Revised: 2024-05-21
Accepted: 2024-06-07
Published Online: 2024-08-09

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Editorial
  2. Focus on NLENG 2023 Volume 12 Issue 1
  3. Research Articles
  4. Seismic vulnerability signal analysis of low tower cable-stayed bridges method based on convolutional attention network
  5. Robust passivity-based nonlinear controller design for bilateral teleoperation system under variable time delay and variable load disturbance
  6. A physically consistent AI-based SPH emulator for computational fluid dynamics
  7. Asymmetrical novel hyperchaotic system with two exponential functions and an application to image encryption
  8. A novel framework for effective structural vulnerability assessment of tubular structures using machine learning algorithms (GA and ANN) for hybrid simulations
  9. Flow and irreversible mechanism of pure and hybridized non-Newtonian nanofluids through elastic surfaces with melting effects
  10. Stability analysis of the corruption dynamics under fractional-order interventions
  11. Solutions of certain initial-boundary value problems via a new extended Laplace transform
  12. Numerical solution of two-dimensional fractional differential equations using Laplace transform with residual power series method
  13. Fractional-order lead networks to avoid limit cycle in control loops with dead zone and plant servo system
  14. Modeling anomalous transport in fractal porous media: A study of fractional diffusion PDEs using numerical method
  15. Analysis of nonlinear dynamics of RC slabs under blast loads: A hybrid machine learning approach
  16. On theoretical and numerical analysis of fractal--fractional non-linear hybrid differential equations
  17. Traveling wave solutions, numerical solutions, and stability analysis of the (2+1) conformal time-fractional generalized q-deformed sinh-Gordon equation
  18. Influence of damage on large displacement buckling analysis of beams
  19. Approximate numerical procedures for the Navier–Stokes system through the generalized method of lines
  20. Mathematical analysis of a combustible viscoelastic material in a cylindrical channel taking into account induced electric field: A spectral approach
  21. A new operational matrix method to solve nonlinear fractional differential equations
  22. New solutions for the generalized q-deformed wave equation with q-translation symmetry
  23. Optimize the corrosion behaviour and mechanical properties of AISI 316 stainless steel under heat treatment and previous cold working
  24. Soliton dynamics of the KdV–mKdV equation using three distinct exact methods in nonlinear phenomena
  25. Investigation of the lubrication performance of a marine diesel engine crankshaft using a thermo-electrohydrodynamic model
  26. Modeling credit risk with mixed fractional Brownian motion: An application to barrier options
  27. Method of feature extraction of abnormal communication signal in network based on nonlinear technology
  28. An innovative binocular vision-based method for displacement measurement in membrane structures
  29. An analysis of exponential kernel fractional difference operator for delta positivity
  30. Novel analytic solutions of strain wave model in micro-structured solids
  31. Conditions for the existence of soliton solutions: An analysis of coefficients in the generalized Wu–Zhang system and generalized Sawada–Kotera model
  32. Scale-3 Haar wavelet-based method of fractal-fractional differential equations with power law kernel and exponential decay kernel
  33. Non-linear influences of track dynamic irregularities on vertical levelling loss of heavy-haul railway track geometry under cyclic loadings
  34. Fast analysis approach for instability problems of thin shells utilizing ANNs and a Bayesian regularization back-propagation algorithm
  35. Validity and error analysis of calculating matrix exponential function and vector product
  36. Optimizing execution time and cost while scheduling scientific workflow in edge data center with fault tolerance awareness
  37. Estimating the dynamics of the drinking epidemic model with control interventions: A sensitivity analysis
  38. Online and offline physical education quality assessment based on mobile edge computing
  39. Discovering optical solutions to a nonlinear Schrödinger equation and its bifurcation and chaos analysis
  40. New convolved Fibonacci collocation procedure for the Fitzhugh–Nagumo non-linear equation
  41. Study of weakly nonlinear double-diffusive magneto-convection with throughflow under concentration modulation
  42. Variable sampling time discrete sliding mode control for a flapping wing micro air vehicle using flapping frequency as the control input
  43. Error analysis of arbitrarily high-order stepping schemes for fractional integro-differential equations with weakly singular kernels
  44. Solitary and periodic pattern solutions for time-fractional generalized nonlinear Schrödinger equation
  45. An unconditionally stable numerical scheme for solving nonlinear Fisher equation
  46. Effect of modulated boundary on heat and mass transport of Walter-B viscoelastic fluid saturated in porous medium
  47. Analysis of heat mass transfer in a squeezed Carreau nanofluid flow due to a sensor surface with variable thermal conductivity
  48. Navigating waves: Advancing ocean dynamics through the nonlinear Schrödinger equation
  49. Experimental and numerical investigations into torsional-flexural behaviours of railway composite sleepers and bearers
  50. Novel dynamics of the fractional KFG equation through the unified and unified solver schemes with stability and multistability analysis
  51. Analysis of the magnetohydrodynamic effects on non-Newtonian fluid flow in an inclined non-uniform channel under long-wavelength, low-Reynolds number conditions
  52. Convergence analysis of non-matching finite elements for a linear monotone additive Schwarz scheme for semi-linear elliptic problems
  53. Global well-posedness and exponential decay estimates for semilinear Newell–Whitehead–Segel equation
  54. Petrov-Galerkin method for small deflections in fourth-order beam equations in civil engineering
  55. Solution of third-order nonlinear integro-differential equations with parallel computing for intelligent IoT and wireless networks using the Haar wavelet method
  56. Mathematical modeling and computational analysis of hepatitis B virus transmission using the higher-order Galerkin scheme
  57. Mathematical model based on nonlinear differential equations and its control algorithm
  58. Bifurcation and chaos: Unraveling soliton solutions in a couple fractional-order nonlinear evolution equation
  59. Space–time variable-order carbon nanotube model using modified Atangana–Baleanu–Caputo derivative
  60. Minimal universal laser network model: Synchronization, extreme events, and multistability
  61. Valuation of forward start option with mean reverting stock model for uncertain markets
  62. Geometric nonlinear analysis based on the generalized displacement control method and orthogonal iteration
  63. Fuzzy neural network with backpropagation for fuzzy quadratic programming problems and portfolio optimization problems
  64. B-spline curve theory: An overview and applications in real life
  65. Nonlinearity modeling for online estimation of industrial cooling fan speed subject to model uncertainties and state-dependent measurement noise
  66. Quantitative analysis and modeling of ride sharing behavior based on internet of vehicles
  67. Review Article
  68. Bond performance of recycled coarse aggregate concrete with rebar under freeze–thaw environment: A review
  69. Retraction
  70. Retraction of “Convolutional neural network for UAV image processing and navigation in tree plantations based on deep learning”
  71. Special Issue: Dynamic Engineering and Control Methods for the Nonlinear Systems - Part II
  72. Improved nonlinear model predictive control with inequality constraints using particle filtering for nonlinear and highly coupled dynamical systems
  73. Anti-control of Hopf bifurcation for a chaotic system
  74. Special Issue: Decision and Control in Nonlinear Systems - Part I
  75. Addressing target loss and actuator saturation in visual servoing of multirotors: A nonrecursive augmented dynamics control approach
  76. Collaborative control of multi-manipulator systems in intelligent manufacturing based on event-triggered and adaptive strategy
  77. Greenhouse monitoring system integrating NB-IOT technology and a cloud service framework
  78. Special Issue: Unleashing the Power of AI and ML in Dynamical System Research
  79. Computational analysis of the Covid-19 model using the continuous Galerkin–Petrov scheme
  80. Special Issue: Nonlinear Analysis and Design of Communication Networks for IoT Applications - Part I
  81. Research on the role of multi-sensor system information fusion in improving hardware control accuracy of intelligent system
  82. Advanced integration of IoT and AI algorithms for comprehensive smart meter data analysis in smart grids
Downloaded on 8.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/nleng-2024-0017/html
Scroll to top button