Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Discontinuous maneuver trajectory prediction based on HOA-GRU method for the UAVs


  • Received: 02 April 2022 Revised: 15 May 2022 Accepted: 23 May 2022 Published: 10 June 2022
  • With the rapid development of artificial intelligence technology, the intelligence and autonomy of Unmanned Aerial Vehicles (UAVs) have been significantly improved. Because the real trajectory data is often discontinuous and random, the current aircraft maneuver trajectory prediction methods are far from meeting the practical requirements of the autonomous air tasks. Especially, in order to occupy a better position rapidly where it is easier to attack the enemy, a fast and accurate maneuver trajectory prediction method for the UAVs is proposed in this paper. Firstly, the prediction model of aircraft maneuvering trajectory is built by extracting characteristic information from the historical trajectory. Aiming at the problem of slow optimization speed and easy to fall into local optimization, a global aircraft maneuver trajectory prediction method based on the Hummingbird Optimization Algorithm (HOA) and Gated Recurrent Unit (GRU) is proposed. Then, the implementation process of the maneuver trajectory prediction method based on the above HOA-GRU network for the UAVs is presented. Finally, the aircraft maneuver trajectory prediction method is applied to a simulation training system with the discontinuous and random air task data. The simulation results show that the proposed method can predict the maneuver trajectory of the UAVs with discontinuous data in real time with less error and less time.

    Citation: Zhizhou Zhang, Zhenglei Wei, Bowen Nie, Yang Li. Discontinuous maneuver trajectory prediction based on HOA-GRU method for the UAVs[J]. Electronic Research Archive, 2022, 30(8): 3111-3129. doi: 10.3934/era.2022158

    Related Papers:

    [1] Jun Moon . The Pontryagin type maximum principle for Caputo fractional optimal control problems with terminal and running state constraints. AIMS Mathematics, 2025, 10(1): 884-920. doi: 10.3934/math.2025042
    [2] Yuna Oh, Jun Moon . The infinite-dimensional Pontryagin maximum principle for optimal control problems of fractional evolution equations with endpoint state constraints. AIMS Mathematics, 2024, 9(3): 6109-6144. doi: 10.3934/math.2024299
    [3] Ruiqing Shi, Yihong Zhang . Dynamic analysis and optimal control of a fractional order HIV/HTLV co-infection model with HIV-specific antibody immune response. AIMS Mathematics, 2024, 9(4): 9455-9493. doi: 10.3934/math.2024462
    [4] Jun Moon . A Pontryagin maximum principle for terminal state-constrained optimal control problems of Volterra integral equations with singular kernels. AIMS Mathematics, 2023, 8(10): 22924-22943. doi: 10.3934/math.20231166
    [5] Irmand Mikiela, Valentina Lanza, Nathalie Verdière, Damienne Provitolo . Optimal strategies to control human behaviors during a catastrophic event. AIMS Mathematics, 2022, 7(10): 18450-18466. doi: 10.3934/math.20221015
    [6] Xiangyun Shi, Xiwen Gao, Xueyong Zhou, Yongfeng Li . Analysis of an SQEIAR epidemic model with media coverage and asymptomatic infection. AIMS Mathematics, 2021, 6(11): 12298-12320. doi: 10.3934/math.2021712
    [7] Xuefeng Yue, Weiwei Zhu . The dynamics and control of an ISCRM fractional-order rumor propagation model containing media reports. AIMS Mathematics, 2024, 9(4): 9721-9745. doi: 10.3934/math.2024476
    [8] Sayed Saber, Azza M. Alghamdi, Ghada A. Ahmed, Khulud M. Alshehri . Mathematical Modelling and optimal control of pneumonia disease in sheep and goats in Al-Baha region with cost-effective strategies. AIMS Mathematics, 2022, 7(7): 12011-12049. doi: 10.3934/math.2022669
    [9] Asaf Khan, Gul Zaman, Roman Ullah, Nawazish Naveed . Optimal control strategies for a heroin epidemic model with age-dependent susceptibility and recovery-age. AIMS Mathematics, 2021, 6(2): 1377-1394. doi: 10.3934/math.2021086
    [10] Cuifang Lv, Xiaoyan Chen, Chaoxiong Du . Global dynamics of a cytokine-enhanced viral infection model with distributed delays and optimal control analysis. AIMS Mathematics, 2025, 10(4): 9493-9515. doi: 10.3934/math.2025438
  • With the rapid development of artificial intelligence technology, the intelligence and autonomy of Unmanned Aerial Vehicles (UAVs) have been significantly improved. Because the real trajectory data is often discontinuous and random, the current aircraft maneuver trajectory prediction methods are far from meeting the practical requirements of the autonomous air tasks. Especially, in order to occupy a better position rapidly where it is easier to attack the enemy, a fast and accurate maneuver trajectory prediction method for the UAVs is proposed in this paper. Firstly, the prediction model of aircraft maneuvering trajectory is built by extracting characteristic information from the historical trajectory. Aiming at the problem of slow optimization speed and easy to fall into local optimization, a global aircraft maneuver trajectory prediction method based on the Hummingbird Optimization Algorithm (HOA) and Gated Recurrent Unit (GRU) is proposed. Then, the implementation process of the maneuver trajectory prediction method based on the above HOA-GRU network for the UAVs is presented. Finally, the aircraft maneuver trajectory prediction method is applied to a simulation training system with the discontinuous and random air task data. The simulation results show that the proposed method can predict the maneuver trajectory of the UAVs with discontinuous data in real time with less error and less time.



    Optimal control (OC) emerged in the 1950s as an extension of the calculus of variations. A significant breakthrough in this field was the development and proof of Pontryagin's maximum principle (PMP) by Pontryagin and his collaborators [34]. This principle established the necessary conditions for an optimal solution and has become one of the most powerful tools for addressing OC problems. This formalization of OC theory raised several new questions, particularly in the context of differential equations. It led to the introduction of new generalized solution concepts and produced significant results concerning the existence of trajectories. The applications of OC are extensive, encompassing fields such as mathematics [41], physics [9], engineering [17], robotics [11], biology [24], and economics [39], among others. The primary objective of an OC problem is to identify, from a set of possible solutions (referred to as the admissible set) the one that minimizes or maximizes a given functional. That is, the goal is to determine an OC trajectory and the corresponding state trajectory. The system's dynamics, modeled by state variables, are influenced by control variables that enter the system of equations, thereby affecting its behavior.

    Fractional calculus (FC) is a well-known theory that extends the ideas of integration and differentiation to non-integer orders, providing a more flexible approach to modeling complex systems with non-local memory. Numerous publications have been devoted to exploring FC and fractional differential equations (FDEs) from various perspectives, covering both its fundamental definitions and principles (see, e.g., [22,38] and the references therein), and its diverse applications in areas such as biology [6], medicine [23], epidemiology [28], electrochemistry [31], physics [33], mechanics [35], and economy [37]. Since the begining of fractional calculus in 1695, numerous definitions of fractional derivatives have been introduced, including the Riemann-Liouville, Caputo, Hadamard, Gr¨unwald-Letnikov, Riesz, Erdélyi-Kober, and Miller-Ross derivatives, among others. Every form of fractional derivative comes with its own benefits and is selected based on the specific needs of the problem being addressed.

    In this work, we will explore recent advancements in fractional derivatives as introduced by [10]. Specifically, we will focus on the novel concepts of distributed-order fractional derivatives with respect to an arbitrary kernel in the Riemann-Liouville and Caputo senses. In their study, the authors established necessary and sufficient conditions within the context of the calculus of variations and derived an associated Euler–Lagrange equation. Building upon their results, we extend the analysis to the setting of optimal control theory and derive a corresponding generalization of the Pontryagin maximum principle.

    A key advantage of distributed-order fractional derivatives [8] lies in their ability to describe systems where the memory effect varies over time. Unlike classical or constant-order fractional derivatives, distributed-order derivatives are defined through an integral over a range of orders, weighted by a given distribution function. This provides a more flexible and realistic modeling tool for systems whose dynamic behavior cannot be adequately captured by a single, fixed fractional order. Distributed-order fractional derivatives have proven especially effective in modeling systems characterized by a broad spectrum of dynamic behaviors. This includes, for instance, viscoelastic materials exhibiting a wide range of relaxation times, transport phenomena influenced by multiple temporal and spatial scales, and control systems affected by diverse time delays. Notable applications span across fields such as viscoelasticity, anomalous diffusion, wave propagation, and the design of fractional-order Proportional-Integral-Derivative (PID) controllers [13,18].

    In parallel, the concept of fractional derivatives with respect to arbitrary kernels [3,38] further improves the model. These operators generalize classical definitions (such as the Riemann–Liouville or Caputo derivatives) by introducing a kernel function that dictates the memory behavior of the system. By selecting an appropriate kernel, one can recover various classical forms or construct entirely new operators. This kernel-based approach allows for incorporating diverse memory effects, singularities, or fading influence in time, which are often observed in real-world processes but difficult to model with standard tools.

    These two features—variable order and general kernels—collectively enable the modeling of complex phenomena with nonlocal and history-dependent behavior, particularly in contexts where the memory characteristics are not uniform or stationary. This is especially relevant in systems governed by internal friction, relaxation, hereditary stress-strain relations, or anomalous transport, where the effects of the past on the present are not constant and may depend on the nature of past events in intricate ways.

    Building on these two ideas, we consider OC problems that involve a generalized fractional derivative constraint. OC provides a rigorous mathematical framework for determining the best possible strategy to steer a dynamical system from an initial state to a desired final state while minimizing (or maximizing) a given cost functional. The inclusion of fractional operators—particularly of distributed-order or arbitrary-kernel types—within this framework enhances the ability to model systems with complex dynamics and long-range temporal dependencies. It enables decision-makers to formulate strategies that are optimal with respect to cost, energy, time, or other performance metrics, all while satisfying dynamic constraints such as differential equations, control bounds, and state limitations. This is crucial for achieving efficiency, such as reducing fuel consumption in aerospace trajectories, minimizing energy use in industrial processes, or shortening recovery time in medical treatments.

    Pioneering works, such as those in [1,2], developed necessary conditions for optimality with respect to the classical Caputo fractional derivative. These studies were further expanded in [14,15] with the formulation of the fractional Noether's theorem. Since then, numerous studies have emerged on fractional OC problems, addressing various fractional operators such as the Caputo [5,20], Riemann–Liouville [19], and distributed-order derivatives [30], as well as aspects like delays in systems [16] and fuzzy theory [32]. Numerical methods for solving fractional optimal control problems are available in the literature. For example, [21] uses a generalized differential transform method, [25,42] employ second-order and third-order numerical integration methods, and [27] applies a Legendre orthonormal polynomial basis.

    The primary objective of this paper is to establish a generalized fractional PMP applicable to OC problems involving the left-sided distributed-order Caputo fractional derivative with respect to an arbitrary smooth kernel. Additionally, we will present sufficient conditions for optimality based on PMP.

    The rest of the paper is structured as follows. Section 2 introduces some fundamental concepts and results from fractional calculus required for our work. In Section 3, we present the OC problem (POC) under study, some fundamental lemmas, PMP, and sufficient conditions for the global optimality of the problem (POC). Section 4 provides two examples that illustrate the practical relevance of our research. Finally, in Section 5, we summarize our findings and present ideas for future work.

    This section provides essential definitions and results with respect to the distributed-order Riemann-Liouville and Caputo derivatives depending on a given kernel [10]. We assume that readers are already acquainted with the definitions and properties of the classical fractional operators [22,38].

    Throughout the paper, we suppose that a<b are two real numbers, J=[a,b], γR+, and [γ] denotes the integer part of γ. First, we introduce essential concepts relevant to our work.

    Definition 2.1. [38] (Fractional integrals in the Riemann–Liouville sense) Let zL1(J,R), and let σC1(J,R) be a continuously differentiable function with a positive derivative.

    For t>a, the left-sided σ-R-L fractional integral of the function z of order γ is given by

    Iγ,σa+z(t):=taσ(τ)z(τ)Γ(γ)(σ(t)σ(τ))1γdτ,

    and for t<b, the right-sided σ-R-L fractional integral is

    Iγ,σbz(t):=btσ(τ)z(τ)Γ(γ)(σ(τ)σ(t))1γdτ.

    Definition 2.2. [38] (Fractional derivatives in the Riemann–Liouville sense) Let zL1(J,R), and let σCn(J,R) be such that σ(t)>0, given tJ.

    The left-sided σ-R-L fractional derivative of the function z of order γ is given by

    Dγ,σa+z(t):=(1σ(t)ddt)nInγ,σa+z(t),

    and the right-sided σ-R-L fractional derivative is

    Dγ,σbz(t):=(1σ(t)ddt)nInγ,σbz(t),

    where n=[γ]+1.

    Definition 2.3. [3] (Fractional derivatives in the Caputo sense) For a fixed γR+, define nN as: n=[γ]+1 if γN, and n=γ if γN. Moreover, consider z,σCn(J,R) where σ(t)>0, tJ. The left- and right-sided σ-C fractional derivatives of the function z of order γ are given by

    CDγ,σa+z(t):=Inγ,σa+(1σ(t)ddt)nz(t),

    and

    CDγ,σbz(t):=Inγ,σb(1σ(t)ddt)nz(t),

    respectively.

    We remark that, for 0<γ<1:

    Dγ,σa+z(t):=1Γ(1γ)(1σ(t)ddt)taσ(τ)(σ(t)σ(τ))γz(τ)dτ,
    Dγ,σbz(t):=1Γ(1γ)(1σ(t)ddt)btσ(τ)(σ(τ)σ(t))γz(τ)dτ,
    CDγ,σa+z(t):=1Γ(1γ)ta(σ(t)σ(τ))γz(τ)dτ,

    and

    CDγ,σbz(t):=1Γ(1γ)bt(σ(τ)σ(t))γz(τ)dτ.

    If γ=1, we get

    Dγ,σa+z(t)=CDγ,σa+z(t)=z(t)σ(t)

    and

    Dγ,σbz(t)=CDγ,σbz(t)=z(t)σ(t).

    Next, we recall some properties that will be useful in our proofs. For γ(0,1] and zC1(J,R), the following relations hold:

    CDγ,σa+Iγ,σa+z(t)=z(t)

    and

    Iγ,σa+CDγ,σa+z(t)=z(t)z(a).

    For further details, we refer the reader to [3].

    Since the fractional OC problem examined in this paper involves a fractional derivative of order γ within the range (0, 1], we will henceforth assume γ(0,1].

    To define the distributed-order derivatives, we need to fix the order-weighting function, denoted by Φ. Here, Φ is a continuous function defined on [0,1] such that Φ([0,1])[0,1] and 10Φ(γ)dγ>0.

    Definition 2.4. [10] (Distributional-order fractional derivatives in the Riemann–Liouville sense) Let zL1(J,R). The left- and right-sided σ-D-R-L fractional derivatives of a function z with respect to the distribution Φ are defined by

    DΦ(γ),σa+z(t):=10Φ(γ)Dγ,σa+z(t)dγandDΦ(γ),σbz(t):=10Φ(γ)Dγ,σbz(t)dγ,

    where Dγ,σa+ and Dγ,σb are the left- and right-sided σ-R-L fractional derivatives of order γ, respectively.

    Definition 2.5. [10] (Distributional-order fractional derivatives in the Caputo sense) The left- and right-sided σ-D-C fractional derivatives of a function zC1(J,R) with respect to the distribution Φ are defined by

    CDΦ(γ),σa+z(t):=10Φ(γ)CDγ,σa+z(t)dγandCDΦ(γ),σbz(t):=10Φ(γ)CDγ,σbz(t)dγ,

    where CDγ,σa+ and CDγ,σb are the left- and right-sided σ-C fractional derivatives of order γ, respectively.

    In what follows, we introduce the concepts of σ-distributional-order fractional integrals:

    I1Φ(γ),σa+z(t):=10Φ(γ)I1γ,σa+z(t)dγandI1Φ(γ),σbz(t):=10Φ(γ)I1γ,σbz(t)dγ,

    where I1γ,σa+ and I1γ,σb are the left- and right-sided σ-R-L fractional integrals of order 1γ, respectively.

    It is evident from the definitions that distributed-order operators are linear.

    In the following, we present a generalized fractional integration by parts formula which is useful to demonstrate some of our results.

    Lemma 2.6. (Generalized fractional integration by parts) [10] Given wC(J,R) and zC1(J,R), the following holds:

    baw(t)CDΦ(γ),σa+z(t)dt=baz(t)(DΦ(γ),σbw(t)σ(t))σ(t)dt+[z(t)(I1Φ(γ),σbw(t)σ(t))]t=bt=a.

    We now present a general form of Gronwall's inequality, a crucial tool for comparing solutions of FDEs that involve σ-fractional derivatives.

    Lemma 2.7. (Generalized fractional Gronwall inequality) [40] Let u,vL1(J,R), σC1(J,R) with σ(t)>0 for all tJ, and hC(J,R). Assume also that u,v,h are nonnegative and h is nondecreasing. If

    u(t)v(t)+h(t)taσ(τ)(σ(t)σ(τ))γ1u(τ)dτ,tJ,

    then

    u(t)v(t)+tai=1[Γ(γ)h(t)]iΓ(iγ)σ(τ)(σ(t)σ(τ))iγ1v(τ)dτ,tJ.

    Remark 2.8. In Lemma 2.7, the assumption that h is nondecreasing is often satisfied in practical contexts where h represents quantities such as cumulative costs, energy consumption, or memory effects that naturally increase or remain constant over time.

    Next, we recall the definition of the Mittag-Leffler function (with one parameter), which generalizes the standard exponential function and plays a fundamental role in the study of differential equations of fractional order.

    Definition 2.9. The Mittag-Leffler function with parameter γ>0 is defined by

    Eγ(t):=i=0tiΓ(γi+1),tR.

    To conclude this section, we present the following result, which is a corollary of Lemma 2.7 and will be useful in the proof of Lemma 3.3.

    Corollary 2.10. (cf. [40]) Under the hypotheses of Lemma 2.7, if v is a nondecreasing function, then, for all tJ,

    u(t)v(t)Eγ(Γ(γ)h(t)[σ(t)σ(a)]γ).

    The aim of this section is to establish necessary and sufficient optimality conditions for a fractional OC problem involving the left-sided Caputo distributed-order fractional derivative with respect to a smooth kernel σ.

    In what follows, we use the standard notations: PC(J,R) denotes the set of all real-valued piecewise continuous functions defined on J, and PC1(J,R) denotes the set of piecewise smooth functions defined on J. The fractional OC problem is defined by the following formulation:

    Problem (POC): Determine xPC1(J,R) and uPC(J,R) that extremizes the functional

    J(x,u)=baL(t,x(t),u(t))dt,

    under the following restrictions: the FDE

    CDΦ(γ),σa+x(t)=f(t,x(t),u(t)),tJ,

    and the initial condition

    x(a)=xa,

    where LC1(J×R2,R) and fC2(J×R2,R).

    Remark 3.1. There are several equivalent ways to formulate an optimal control problem, notably in the Mayer, Lagrange, and Bolza forms. Through appropriate changes of variables and auxiliary state transformations, one can show that these formulations are theoretically equivalent (see, e.g., Chapter 3 in [26]). In this paper, we adopt the Lagrange form for consistency and simplicity.

    We will now prove a result that establishes a relationship between an optimal state trajectory of the OC problem (POC) and the state solution of a perturbed version of (POC).

    Lemma 3.2. Let the pair (¯u,¯x) be an optimal solution to (POC). For tJ, consider a variation of ¯u of the form ¯u+ξη, where ηPC(J,R) and ξR. Denote by uξ(t) such variation and xξ the solution of

    CDΦ(γ),σa+y(t)=f(t,y(t),uξ(t)),y(a)=xa.

    Then, as ξ tends to zero, xξ¯x.

    Proof. By hypothesis, we know that

    CDΦ(γ),σa+xξ(t)=f(t,xξ(t),uξ(t))andCDΦ(γ),σa+¯x(t)=f(t,¯x(t),¯u(t)),tJ,

    where xξ(a)=¯x(a)=xa. From the linearity and the definition of the operator CDΦ(γ),σa+, we may conclude that

    10Φ(γ)CDγ,σa+(xξ(t)¯x(t))dγ=f(t,xξ(t),uξ(t))f(t,¯x(t),¯u(t)).

    Using the mean value theorem in the integral form, we conclude that there exists β[0,1] such that

    CDβ,σa+(xξ(t)¯x(t))10Φ(γ)dγ=f(t,xξ(t),uξ(t))f(t,¯x(t),¯u(t)).

    Denoting ω:=10Φ(γ)dγ, we conclude that

    CDβ,σa+(xξ(t)¯x(t))=f(t,xξ(t),uξ(t))f(t,¯x(t),¯u(t))ω.

    Thus,

    xξ(t)¯x(t)=Iβ,σa+(f(t,xξ(t),uξ(t))f(t,¯x(t),¯u(t))ω).

    Let k1 and k2 be two positive real numbers such that

    |f(t,xξ(t),uξ(t))f(t,¯x(t),¯u(t))|k1|xξ(t)¯x(t)|+k2|uξ(t)¯u(t)|,tJ,

    and denote by k their maximum. Therefore,

    |xξ(t)¯x(t)|Iβ,σa+(|f(t,xξ(t),uξ(t))f(t,¯x(t),¯u(t))|ω)kω(Iβ,σa+(|xξ(t)¯x(t)|+|uξ(t)¯u(t)|)),tJ.

    Since uξ(t)¯u(t)=ξη(t), we have

    |xξ(t)¯x(t)|kω(|ξ|Iβ,σa+(|η(t)|)+Iβ,σa+(|xξ(t)¯x(t)|))=kω|ξ|Iβ,σa+(|η(t)|)+kω1Γ(β)taσ(τ)(σ(t)σ(τ))β1|xξ(τ)¯x(τ)|dτ,

    for all tJ.

    Now, applying Lemma 2.7, we obtain that

    |xξ(t)¯x(t)|kω|ξ|Iβ,σa+(|η(t)|)+tai=1(kω)i+1Γ(iβ)σ(τ)(σ(t)σ(τ))iβ1|ξ|Iβ,σa+(|η(τ)|)dτ,tJ.

    Thus, by applying the mean value theorem once more, we deduce that there exists some ¯τJ such that

    |xξ(t)¯x(t)|kω|ξ|Iβ,σa+(|η(t)|)+|ξ|Iβ,σa+(|η(¯τ)|)i=1(kω)i+1Γ(iβ+1)(σ(t)σ(a))iβ=kω|ξ|[Iβ,σa+(|η(t)|)+Iβ,σa+(|η(¯τ)|)(Eβ(kω(σ(t)σ(a))β)1)],tJ.

    So, the proof is finished by taking the limit when ξ0.

    Before presenting our next result, we extend the one presented in [12, Theorem 3.4] for FDEs with the σ-Caputo fractional derivative.

    Lemma 3.3. Consider the following two FDEs of order γ(0,1]:

    CDγ,σa+x(t)=f(t,x(t))andCDγ,σa+x(t)=˜f(t,x(t)),tJ,

    with the same initial condition. Suppose the functions f and ˜f are Lipschitz continuous with respect to x, with Lipschitz constant L>0. If y and z are the unique solutions to the first and second equations, respectively, then there exists a constant C>0, independent of y and z, such that

    yzCf˜f.

    Proof. Given tJ, we have:

    |y(t)z(t)|=|Iγ,σa+(f(t,y(t))˜f(t,z(t)))||Iγ,σa+(f(t,y(t))f(t,z(t)))|+|Iγ,σa+(f(t,z(t))˜f(t,z(t)))|Iγ,σa+(L|y(t)z(t)|)+f˜f1Γ(γ)taσ(τ)(σ(t)σ(τ))γ1dτLΓ(γ)taσ(τ)(σ(t)σ(τ))γ1|y(τ)z(τ)|dτ+f˜f(σ(b)σ(a))γΓ(γ+1).

    Applying Corollary 2.10, with

    u(t):=|y(t)z(t)|,v(t):=f˜f(σ(b)σ(a))γΓ(γ+1),h(t):=LΓ(γ),

    we obtain

    |y(t)z(t)|f˜f(σ(b)σ(a))γΓ(γ+1)Eγ(L(σ(t)σ(a))γ),tJ.

    Thus, we conclude that

    yzCf˜f,whereC=(σ(b)σ(a))γΓ(γ+1)Eγ(L(σ(b)σ(a))γ),

    which proves the desired result.

    Throughout the following, we use the standard Big-O notation: a(x)=O(b(x)) means that there exists a constant C>0 such that |a(x)|C|b(x)| for all x sufficiently close to 0.

    Lemma 3.4. Suppose we are in the conditions of Lemma 3.2. Then, there exists a mapping ν:JR such that

    xξ(t)=¯x(t)+ξν(t)+O(ξ2).

    Proof. Given that fC2, the function f can be expanded as

    f(t,xξ(t),uξ(t))=f(t,¯x(t),¯u(t))+(xξ(t)¯x(t))f(t,¯x(t),¯u(t))x+(uξ(t)¯u(t))f(t,¯x(t),¯u(t))u+O(|xξ(t)¯x(t)|2,|uξ(t)¯u(t)|2),tJ.

    For all tJ, we have uξ(t)¯u(t)=ξη(t). Additionally, as proven in Lemma 3.2, xξ(t)¯x(t)=ξχ(t), for some finite function χ. Thus, the term O(|xξ(t)¯x(t)|2,|uξ(t)¯u(t)|2) simplifies to O(ξ2). Hence, we can express the fractional derivative of xξ as:

    CDΦ(γ),σa+xξ(t)=CDΦ(γ),σa+¯x(t)+(xξ(t)¯x(t))f(t,¯x(t),¯u(t))x+ξη(t)f(t,¯x(t),¯u(t))u+O(ξ2).

    Therefore, for ξ0 and tJ,

    CDΦ(γ),σa+xξ(t)¯x(t)ξ=xξ(t)¯x(t)ξf(t,¯x(t),¯u(t))x+η(t)f(t,¯x(t),¯u(t))u+O(ξ2)ξ.

    Consider the following two FDEs:

    CDΦ(γ),σa+y(t)=y(t)f(t,¯x(t),¯u(t))x+η(t)f(t,¯x(t),¯u(t))u+O(ξ2)ξ (3.1)

    and

    CDΦ(γ),σa+y(t)=y(t)f(t,¯x(t),¯u(t))x+η(t)f(t,¯x(t),¯u(t))u, (3.2)

    subject to y(a)=0. The existence and uniqueness of solutions for distributed-order FDEs can be ensured using standard results for FDEs involving the σ-Caputo fractional derivative (see, e.g., [4]). This is achieved by using the relation

    CDΦ(γ),σa+y(t)=ωCD¯γ,σa+y(t),

    for some ¯γ[0,1] and with ω=10Φ(γ)dγ. Equation (3.1) has the solution y1:=xξ¯xξ, and let y2 be the solution of Eq (3.2). By Lemma 3.3, we obtain that y2(t)=limξ0y1(t), proving the desired result with ν:=y2.

    We are now ready to present the main result of our work:

    Theorem 3.5. (PMP for (POC)) Let (¯x,¯u) be an optimal pair to problem (POC). Then, there exists a mapping λPC1(J,R) such that:

    The two following FDEs hold:

    Lu(t,¯x(t),¯u(t))+λ(t)fu(t,¯x(t),¯u(t))=0,tJ; (3.3)

    and

    (DΦ(γ),σbλ(t)σ(t))σ(t)=Lx(t,¯x(t),¯u(t))+λ(t)fx(t,¯x(t),¯u(t)),tJ; (3.4)

    The transversality condition

    I1Φ(γ),σbλ(b)σ(b)=0. (3.5)

    Proof. Suppose that (¯x,¯u) is a solution to problem (POC). Consider a variation of ¯u of the form ¯u+ξη, denoted by uξ, where ηPC(J,R) and ξR. Also, let xξ be the state variable satisfying

    {CDΦ(γ),σa+xξ(t)=f(t,xξ(t),uξ(t)),tJ,xξ(a)=xa. (3.6)

    Observe that, as ξ goes to zero, uξ goes to ¯u on J, and that

    uξ(t)ξ|ξ=0=η(t). (3.7)

    By Lemma 3.2 we can conclude that xξ¯x on J when ξ0. Furthermore, by Lemma 3.4, the partial derivative xξ(t)ξ|ξ=0=ν(t) exists for each t. We are considering the following functional

    J(xξ,uξ)=baL(t,xξ(t),uξ(t))dt.

    Let λPC1(J,R), to be explained later. From Lemma 2.6, we conclude that

    baλ(t)CDΦ(γ),σa+xξ(t)dtbaxξ(t)(DΦ(γ),σbλ(t)σ(t))σ(t)dt[xξ(t)(I1Φ(γ),σbλ(t)σ(t))]t=bt=a=0.

    Then,

    J(xξ,uξ)=ba[L(t,xξ(t),uξ(t))+λ(t)f(t,xξ(t),uξ(t))xξ(t)(DΦ(γ),σbλ(t)σ(t))σ(t)]dtxξ(b)(I1Φ(γ),σbλ(b)σ(b))+xa(I1Φ(γ),σbλ(a)σ(a)).

    Since (¯x,¯u) is a solution to problem (POC), we conclude that

    0=ddξJ(xξ,uξ)|ξ=0=ba[Lxxξ(t)ξ|ξ=0+Luuξ(t)ξ|ξ=0+λ(t)(fxxξ(t)ξ|ξ=0+fuuξ(t)ξ|ξ=0)]dtba[(DΦ(γ),σbλ(t)σ(t))σ(t)xξ(t)ξ|ξ=0]dt(I1Φ(γ),σbλ(b)σ(b))xξ(b)ξ|ξ=0,

    where the partial derivatives of L and f are evaluated at the point (t,¯x(t),¯u(t)). Now replacing (3.7), we get

    0=ba[Lxxξ(t)ξ|ξ=0+Luη(t)+λ(t)(fxxξ(t)ξ|ξ=0+fuη(t))]dtba(DΦ(γ),σbλ(t)σ(t))σ(t)xξ(t)ξ|ξ=0dt(I1Φ(γ),σbλ(b)σ(b))xξ(b)ξ|ξ=0.

    Rearranging the terms we have

    0=ba[(Lx+λ(t)fx(DΦ(γ),σbλ(t)σ(t))σ(t))xξ(t)ξ|ξ=0+(Lu+λ(t)fu)η(t)]dt(I1Φ(γ),σbλ(b)σ(b))xξ(b)ξ|ξ=0.

    Introducing the Hamiltonian function

    H(t,x(t),u(t),λ(t))=L(t,x(t),u(t))+λ(t)f(t,x(t),u(t)),tJ,

    the expression simplifies to:

    0=ba[(Hx(DΦ(γ),σbλ(t)σ(t))σ(t))xξ(t)ξ|ξ=0+Huη(t)]dt(I1Φ(γ),σbλ(b)σ(b))xξ(b)ξ|ξ=0,

    where H/x and H/u are evaluated at (t,¯x(t),¯u(t),λ(t)). Choosing the function λ as the solution of the system

    (DΦ(γ),σbλ(t)σ(t))σ(t)=Hx,withI1Φ(γ),σbλ(b)σ(b)=0,

    we get

    baHu(t,¯x(t),¯u(t),λ(t))η(t)dt=0.

    Since η is arbitrary, from the Du Bois-Reymond lemma (see, e.g., [7]), we obtain

    Hu(t,¯x(t),¯u(t),λ(t))=0,

    for all tJ, proving the desired result.

    Definition 3.6. The Hamiltonian is defined as H(t,x(t),u(t),λ(t))=L(t,x(t),u(t))+λ(t)f(t,x(t),u(t)), where L is the Lagrange function, f is the system dynamics, and λ is the adjoint variable.

    Remark 3.7. We observe that:

    (1) Taking σ as the identity function, we recover the results presented in [29], while also correcting minor inaccuracies in their proofs. This allows us to refine and extend the results previously established for the classical distributed-order fractional derivative. Furthermore, by introducing a general kernel, we not only broaden the scope of the original framework but also provide a more general formulation that encompasses and extends the findings of the aforementioned study.

    (2) If f(t,x(t),u(t))=u(t),tJ, our problem (POC) reduces to the following fractional problem of the calculus of variations:

    J(x)=baL(t,x(t),CDΦ(γ),σa+x(t))dtextr,

    subject to the initial condition x(a)=xa, where LC1. From Theorem 3.5 we deduce that if ¯x(t) is an extremizer, then there exists λPC1(J,R) such that:

    λ(t)=3L(t,¯x(t),CDΦ(γ),σa+¯x(t)),tJ,

    (DΦ(γ),σbλ(t)σ(t))σ(t)=2L(t,¯x(t),CDΦ(γ),σa+¯x(t)),tJ,

    I1Φ(γ),σbλ(b)σ(b)=0,

    where iL denotes the partial derivative of L with respect to its ith coordinate. Hence, we obtain the Euler-Lagrange equation:

    2L(t,¯x(t),CDΦ(γ),σa+¯x(t))+(DΦ(γ),σb3L(t,¯x(t),CDΦ(γ),σa+¯x(t))σ(t))σ(t)=0,tJ,

    and the tranversality condition:

    I1Φ(γ),σb3L(b,¯x(b),CDΦ(γ),σa+¯x(b))σ(b)=0,

    proved in [10].

    We conclude this section by establishing sufficient optimality conditions for our OC problem.

    Definition 3.8. We say that a triple (¯x,¯u,λ) is a Pontryagin extremal of Problem (POC) if it satisfies conditions (3.3)–(3.5).

    Theorem 3.9. (Sufficient conditions for global optimality I) Let (¯x,¯u,λ) be a Pontryagin extremal of Problem (POC) with λ(t)0, for all tJ.

    If L and f are convex functions, then (¯x,¯u) is a global minimizer of functional J;

    If L and f are concave functions, then (¯x,¯u) is a global maximizer of functional J.

    Proof. We will only prove the case where L and f are convex; the other case is analogous. If L is a convex function, then (see, e.g., [36])

    L(t,¯x(t),¯u(t))L(t,x(t),u(t))Lx(t,¯x(t),¯u(t))(¯xx)(t)+Lu(t,¯x(t),¯u(t))(¯uu)(t),

    for any control u and associate state x. Hence,

    J(¯x,¯u)J(x,u)=ba(L(t,¯x(t),¯u(t))L(t,x(t),u(t)))dtba(Lx(t,¯x(t),¯u(t))(¯xx)(t)+Lu(t,¯x(t),¯u(t))(¯uu)(t))dt.

    Using the adjoint equation (3.4) and the optimality condition (3.3), we obtain

    J(¯x,¯u)J(x,u)ba[(DΦ(γ),σbλ(t)σ(t))σ(t)λ(t)fx(t,¯x(t),¯u(t))](¯xx)(t)dtbaλ(t)fu(t,¯x(t),¯u(t))(¯uu)(t)dt.

    Rearranging the terms, we have

    J(¯x,¯u)J(x,u)ba(¯xx)(t)(DΦ(γ),σbλ(t)σ(t))σ(t)dtba(λ(t)fx(t,¯x(t),¯u(t))(¯xx)(t)+λ(t)fu(t,¯x(t),¯u(t))(¯uu)(t))dt.

    Using the generalized integration by parts formula (Lemma 2.6) in the first integral, we obtain

    J(¯x,¯u)J(x,u)baλ(t)CDΦ(γ),σa+(¯xx)(t)dt[(¯xx)(t)I1Φ(γ),σbλ(t)σ(t)]t=bt=aba(λ(t)fx(t,¯x(t),¯u(t))(¯xx)(t)+λ(t)fu(t,¯x(t),¯u(t))(¯uu)(t))dt.

    Since

    [(¯xx)(t)I1Φ(γ),σbλ(t)σ(t)]t=bt=a=0(by the transversally condition (3.5) and¯x(a)=x(a)=xa)

    and

    CDΦ(γ),σa+(¯xx)(t)=f(t,¯x(t),¯u(t))f(t,x(t),u(t)),

    we get

    J(¯x,¯u)J(x,u)baλ(t)(f(t,¯x(t),¯u(t))f(t,x(t),u(t)))dtbaλ(t)(fx(t,¯x(t),¯u(t))(¯xx)(t)+fu(t,¯x(t),¯u(t))(¯uu)(t))dt.

    Since f is convex, we have

    f(t,¯x(t),¯u(t))f(t,x(t),u(t))fx(t,¯x(t),¯u(t))(¯xx)(t)+fu(t,¯x(t),¯u(t))(¯uu)(t),

    for any admissible control u and its associated state x. Moreover, since λ(t)0, for all tJ, it follows that J(¯x,¯u)J(x,u)0, proving the desired result. Similarly, we can prove the following result.

    Theorem 3.10. (Sufficient conditions for global optimality II) Let (¯x,¯u,λ) be a Pontryagin extremal of Problem (POC) with λ(t)<0, for all tJ.

    If L and f are convex functions, then (¯x,¯u) is a global maximizer of functional J;

    If L and f are concave functions, then (¯x,¯u) is a global minimizer of functional J.

    We now present two examples in order to illustrate our main result.

    Example 4.1. Consider the following problem:

    J(x,u)=10((x(t)(σ(t)σ(0))2)2+(u(t)(σ(t)σ(0))2σ(t)+σ(0)ln(σ(t)σ(0)))2)dtextr,

    subject to

    CDϕ(γ),σ0+x(t)=u(t),t[0,1],andx(0)=0.

    The order-weighting function is ϕ:[0,1][0,1] defined by

    ϕ(γ)=Γ(3γ)2.

    The Hamiltonian in this case is given by

    H(t,x,u,λ)=(x(t)(σ(t)σ(0))2)2+(u(t)(σ(t)σ(0))2σ(t)+σ(0)ln(σ(t)σ(0)))2+λ(t)u(t).

    From (3.3)(3.5), we get

    λ(t)=2(u(t)(σ(t)σ(0))2σ(t)+σ(0)ln(σ(t)σ(0))),t[0,1],
    (Dϕ(γ),σ1λ(t)σ(t))σ(t)=2(x(t)(σ(t)σ(0))2),t[0,1],

    and

    I1ϕ(γ),σ1λ(1)σ(1)=0.

    Note that the triple (¯x,¯u,λ) given by:

    ¯x(t)=(σ(t)σ(0))2,¯u(t)=(σ(t)σ(0))2σ(t)+σ(0)ln(σ(t)σ(0)),λ(t)=0,t[0,1],

    satisfies the necessary optimality conditions given by the Pontryagin maximum principle. Since the Lagrangian L and f(t,x(t),u(t))=u(t) are convex, then (¯x,¯u) is a global minimizer of functional J.

    Example 4.2. Consider the optimal control problem:

    J(x,u)=10(sin(x(t))+cos(x(t))+x(t)u(t))dtextr,

    subject to the restriction

    CDϕ(γ),σ0+x(t)=x(t)u(t),t[0,1],withx(0)=0.

    The corresponding Hamiltonian is given by

    H(t,x,u,λ)=sin(x(t))+cos(x(t))+x(t)u(t)+λ(t)x(t)u(t).

    From the stationarity condition (3.3), we obtain

    x(t)+λ(t)x(t)=0λ(t)=1,t[0,1].

    The remaining necessary conditions, as given by (3.4) and (3.5), are

    (Dϕ(γ),σ11σ(t))σ(t)=cos(x(t))sin(x(t)),t[0,1],

    and

    I1ϕ(γ),σ11σ(1)=0.

    In this work, we explored necessary and sufficient conditions for optimality in fractional OC problems involving a novel fractional operator. Specifically, we addressed the challenge of determining the optimal processes for a fractional OC problem, where the regularity of the optimal solution is assured. The solution to this question is provided by an extension of the PMP, for which we have established the basic version applicable to fractional OC problems, involving a new generalized fractional derivative with respect to a smooth kernel. Additionally, we presented sufficient conditions of optimality for this class of problems using the newly defined fractional derivative. Future research will aim to extend these theorems to OC problems with bounded control constraints. Additionally, it is crucial to develop numerical methods for determining the optimal pair. One potential approach is to approximate the fractional operators using a finite sum, thereby transforming the fractional problem into a finite-dimensional problem, which can then be more effectively solved numerically.

    Fátima Cruz, Ricardo Almeida and Natália Martins: Formal analysis, investigation, methodology, and writing – review & editing, of this manuscript. All authors have read and approved the final version of the manuscript for publication.

    The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work is supported by Portuguese funds through the CIDMA - Center for Research and Development in Mathematics and Applications, and the Portuguese Foundation for Science and Technology (FCT-Fundação para a Ciência e a Tecnologia), within project UIDB/04106/2025.

    Professor Ricardo Almeida is an editorial board member for AIMS Mathematics and was not involved in the editorial review and/or the decision to publish this article.

    All authors declare no conflicts of interest.



    [1] J. L. Yepes, I. Hwang, M. Rotea, New algorithms for aircraft intent inference and trajectory prediction, J. Guid. Control Dyn., 30 (2012), 370–382. https://doi.org/10.2514/1.26750 doi: 10.2514/1.26750
    [2] L. Xie, Z. Wei, D. Ding, Z. Zhang, A. Tang, Long and short term maneuver trajectory prediction of UCAV based on deep learning, IEEE Access, 9 (2021), 32321–32340. https://doi.org/10.1109/ACCESS.2021.3060783 doi: 10.1109/ACCESS.2021.3060783
    [3] D. Ding, Z. Wei, S. Tang, Z. Huang, Robust maneuvering decision-making method for air combat using adaptive prediction weight, Syst. Eng. Electron., 42 (2020), 2275–2284.
    [4] T. Wang, B. Huang, 4D flight trajectory prediction model based on improved Kalman filter, J. Comput. Appl., 34 (2014), 1812–1815.
    [5] K. Zhang, J. Xiong, Fan. Li, T. Fu, Bayesian trajectory prediction for a hypersonic gliding reentry vehicle based on intent inference, J. Astronaut., 39 (2018), 1258–1265. https://doi.org/10.3873/j.issn.1000-1328.2018.11.008 doi: 10.3873/j.issn.1000-1328.2018.11.008
    [6] L. Wang, Q. Xing, Y. Mao, A track forecasting algorithm of boost-glide unpropulsive skipping vehicle, J. Air Force Eng. Univ. (Nat. Sci. Ed.), 16 (2015), 24–27.
    [7] Q. Wang, Z. Zhang, Z. Wang, Y. Wang, W. Zhou, The trajectory prediction of spacecraft by grey method, Meas. Sci. Technol., 27 (2016), 085011–085020. https://doi.org/10.1088/0957-0233/27/8/085011 doi: 10.1088/0957-0233/27/8/085011
    [8] C. G. Prevost, A. Desbiens, E. Gagnon, Extended kalman filter for state estimation and trajectory prediction of a moving object detected by an unmanned aerial vehicle, in 2007 American Control Conference, (2007), 1805–1810. https://doi.org/10.1109/ACC.2007.4282823
    [9] G. Li, H. Zhang, G. Tang, Typical trajectory characteristics of hypersonic glide vehicle, J. Astronaut., 36 (2015) 397–403. https://doi.org/10.3873/j.issn.1000-1328.2015.04.005 doi: 10.3873/j.issn.1000-1328.2015.04.005
    [10] C. Han, J. Xiong, K. Zhang, X. Lan, Decomposition ensemble trajectory prediction algorithm for hypersonic vehicle, Syst. Eng. Electron., 40 (2018), 151–158. https://doi.org/10.3969/j.issn.1001-506X.2018.01.22 doi: 10.3969/j.issn.1001-506X.2018.01.22
    [11] M. Q. Chen, Aircraft climb trajectory prediction using neural network, Appl. Mech. Mater., 373-375 (2013), 1247–1250. https://doi.org/10.4028/www.scientific.net/AMM.373-375.1247 doi: 10.4028/www.scientific.net/AMM.373-375.1247
    [12] X. Wang, R. Yang, J. Zuo, X. Xu, L. Yue, Trajectory prediction of target aircraft based on HPSO-TPFENN neural network, J. Northwest. Polytech. Univ., 37 (2019), 612–620. https://doi.org/10.1051/jnwpu/20193730612 doi: 10.1051/jnwpu/20193730612
    [13] H. Zhang, C. Huang, S. Tang, Y. Xuan, CNN-based real-time prediction method of flight trajectory of unmanned combat aerial vehicle, Acta Armamentarii, 41 (2020), 1894–1903. https://doi.org/10.3969/j.issn.1000-1093.2020.09.022 doi: 10.3969/j.issn.1000-1093.2020.09.022
    [14] Y. Lecun, Y. Bengio, G. Hinton, Deep learning, Nature, 521 (2015), 436–444. https://doi.org/10.1038/nature14539 doi: 10.1038/nature14539
    [15] S. L. Churchill, RNN composition of thematically diverse video game melodies, Comput. Games, 8 (2019), 41–58. https://doi.org/10.1007/s40869-018-0063-x doi: 10.1007/s40869-018-0063-x
    [16] S. Hochreiter, J. Schmidhuber, Long short-term memory, Neural Comput., 9 (1997), 1735–1780. https://doi.org/10.1162/neco.1997.9.8.1735 doi: 10.1162/neco.1997.9.8.1735
    [17] F. Rui, Z. Zuo, L. Li, Using LSTM and GRU neural network methods for traffic flow prediction, in 2016 31st Youth Academic Annual Conference of Chinese Association of Automation (YAC), (2016), 324–328. https://doi.org/10.1109/YAC.2016.7804912
    [18] Z. Zhang, C. Huang, D. Ding, S. Tang, B. Han, H. Huang, Hummingbirds optimization algorithm-based particle filter for maneuvering target tracking, Nonlinear Dyn., 97 (2019), 1227–1243. https://doi.org/10.1007/s11071-019-05043-0 doi: 10.1007/s11071-019-05043-0
    [19] Z. Zhang, C. Huang, H. Huang, S. Tang, K. Dong, An optimization method: hummingbirds optimization algorithm, J. Syst. Eng. Electron., 29 (2018), 386–404. https://doi.org/10.21629/JSEE.2018.02.19 doi: 10.21629/JSEE.2018.02.19
    [20] J. Y. Yoon, A. H. Lee, H. J. Lee, Rendezvous: opportunistic data delivery to mobile users by UAVs through target trajectory prediction, IEEE Trans. Veh. Technol., 69 (2020), 2230–2245. https://doi.org/10.1109/TVT.2019.2962391 doi: 10.1109/TVT.2019.2962391
    [21] B. Hu, H. Yang, L. Wang, S. Chen, A trajectory prediction based intelligent handover control method in UAV cellular networks, China Commun., 16 (2019), 1–14. Available from: https://ieeexplore.ieee.org/document/8633299.
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2557) PDF downloads(109) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog