
Citation: Beyza Billur İskender Eroǧlu, Dilara Yapışkan. Generalized conformable variational calculus and optimal control problems with variable terminal conditions[J]. AIMS Mathematics, 2020, 5(2): 1105-1126. doi: 10.3934/math.2020077
[1] | Ricardo Almeida . Variational problems of variable fractional order involving arbitrary kernels. AIMS Mathematics, 2022, 7(10): 18690-18707. doi: 10.3934/math.20221028 |
[2] | 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 |
[3] | Fátima Cruz, Ricardo Almeida, Natália Martins . Optimality conditions for variational problems involving distributed-order fractional derivatives with arbitrary kernels. AIMS Mathematics, 2021, 6(5): 5351-5369. doi: 10.3934/math.2021315 |
[4] | Tuba Gulsen, Sertac Goktas, Thabet Abdeljawad, Yusuf Gurefe . Sturm-Liouville problem in multiplicative fractional calculus. AIMS Mathematics, 2024, 9(8): 22794-22812. doi: 10.3934/math.20241109 |
[5] | Ahu Ercan . Comparative analysis for fractional nonlinear Sturm-Liouville equations with singular and non-singular kernels. AIMS Mathematics, 2022, 7(7): 13325-13343. doi: 10.3934/math.2022736 |
[6] | 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 |
[7] | Taewan Kim, Jung Hoon Kim . A new optimal control approach to uncertain Euler-Lagrange equations: H∞ disturbance estimator and generalized H2 tracking controller. AIMS Mathematics, 2024, 9(12): 34466-34487. doi: 10.3934/math.20241642 |
[8] | Rabha W. Ibrahim, Jay M. Jahangiri . Conformable differential operator generalizes the Briot-Bouquet differential equation in a complex domain. AIMS Mathematics, 2019, 4(6): 1582-1595. doi: 10.3934/math.2019.6.1582 |
[9] | 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 |
[10] | Mohra Zayed, Gamal Hassan . On the approximation of analytic functions by infinite series of fractional Ruscheweyh derivatives bases. AIMS Mathematics, 2024, 9(4): 8712-8731. doi: 10.3934/math.2024422 |
The conformable derivative has been proposed as an alternative to fractional derivatives by Khalil et al. [1] who aim to solve fractional order differential equations in analytical ways. Despite most of fractional derivative definition defined by the convolution integral with singular (i.e. Riemann–Liouville or Caputo [2]) or nonsingular (i.e. Atangana-Baleanu [3]) kernels, the conformable derivative of a real function f of order 0<α≤1 was suggested via the following limits:
dαf(t)dtα=f(α)(t)=limε→0f(t+εt1−α)−f(t)ε, t>0 | (1.1) |
dαf(t)dtα|t=0=f(α)(0)=limt→0f(α)(t). | (1.2) |
Since the conformable derivative has the limit based definition, it is a local operator like the classical derivative. Therefore, it has not the properties of memory and hereditary which are valuable for usual fractional derivatives. However, the conformable derivative supplies many rules of the classical derivative, the most important one of them is Leibniz rule which cannot be provided by usual fractional derivatives [4]. But, these mentioned points have led to a controversy in fractional scientific groups, that the conformable or any other local derivatives named with "fractional derivative", such as Katugampola, Kolwankar-Gangal, M-derivative, cannot be classified as fractional derivative [4,5,6]. A short time ago, Teodoro et al. have reviewed the definitions linking with the fractional derivatives and they have classified local and nonlocal operators into four classes to clarify the properties of each operator in detail [7]. According to the their classification, they have concluded that the local derivatives are not fractional derivatives. On the other hand, the conformable derivative has been recently covered into the deformed derivative concept which also covers q-derivative, fractal derivative and Haussdorf derivative, and deals with complex dynamical systems with the mathematics of local operators [8,9]. Therefore, it is still significant to explore the conformable or any other local derivatives in the view of their mathematical and physical properties [10]. This idea is also supported by the latest experimental work on anomalous diffusion by Zhou et al. [11] who show the conformable diffusion model is better to fit the experimental data than normal diffusion. Moreover, they indicate that the best fitting order of the conformable derivative is obtained by the short-term experimental data which means the model has the benefit of predicting the long-term subdiffusion process effectively. Another work that exhibits the effectiveness of the conformable derivative in the advection-diffusion process has been presented by Avcı et al. [12]. First of all, they propose a local description of matter flux by introducing parameters of matter diffusion and flow velocity via a power-law time scaling proportional with tα where α is the magnitude of anomalous transport corresponds to order of the conformable derivative, and then obtained the conformable advection-diffusion equation. Furthermore, they demonstrate as a biological application that the proposed conformable model coincides to anomalous diffusion of proteins owing to molecular crowding.
Due to simplicity and potential applications of the conformable derivative, mathematical properties of the operator have been studied by a considerable number of scientists. Abdeljawad defined left, right and sequential conformable derivatives, also gave chain rule, partial integration, conformable Taylor expansion and conformable Laplace transform [13]. The conformable Laplace transform was studied with its deeply details by Silva et al. [14]. The conformable partial derivative was defined by Atangana et al. [15]. Furthermore, complex conformable derivative was suggested and the basic properties of the operator were investigated by [16,17]. Analytical solutions of the conformable differential equations were investigated by Hammad and Khalil [18], Anderson and Ulness [19], Ünal and Gökdoğan [20], and so on [21,22,23]. Although the conformable differential equations can be solved analytically, the numerical solutions were also proposed as expected [24,25,26].
Recently, as a result of successful applications of the conformable derivative into science and engineering problems, analysis of the conformable derivative in system theory, optimization and control areas has been needed. Evirgen [27] modeled a conformable gradient based dynamical system to solve constraint optimization problems. Conformable variational calculus problem (CVCP) was introduced by Chung [28] who obtained the conformable Euler-Lagrange equations and one-dimensional conformable Newtonian mechanics. Optimal boundary control of a heat equation defined by the conformable derivative was proposed by İskender Eroğlu and Yapișkan [29]. The necessary optimality conditions for conformable optimal control problem (COCP) were given by Lazo and Torres [30] via Hamiltonian formalism. General transversality conditions for CVCP and COCPs were explored and the special cases of the general transversality conditions were proposed by us [31]. Optimal control of a conformable diffusion equation was also given as an application problem in [31]. At the same time, Chiranjeevi and Biswas [32] proposed closed form solution of COCP with fixed or variable final conditions. For the case of fixed final time and variable final state COCP, the proposed transversality conditions in [31] and [32] overlap.
The above works except [30] considered the problems with performance index defined by the conformable integral. However, the performance index defined with the classical integral, classified as generalized problem by [30], is seen to be more often used in application problems. Therefore, the current work focuses on the generalized conformable variational calculus problem (GCVCP) and the generalized optimal control problem (GCOCP) with variable terminal point, and aims to present the transversality conditions of the problems which mentioned in this text as generalized transversality conditions. Organization of the paper is planned as follows. The needed definitions and features of conformable derivative are collected in the consecutive section. After that, the generalized transversality conditions are achieved for problems of variational calculus in Section 3 and optimal control in Section 4. The special cases of generalized transversality conditions are also proposed in their related sections and depicted with application problems supported with graphics that are plotted using MATLAB program. Finally, all results obtained throughout the paper are summarized in conclusions.
In this chapter, we give only the needed definitions and theorems to construct the other parts of the paper.
Definition 2.1. ([13]) The left conformable derivative of a real valued function f on a closed interval [a,b] with the order of 0<α≤1 is defined as
dαadtαaf(t)=f(α)a(t)=limε→0f(t+ε(t−a)1−α)−f(t)ε. | (2.1) |
If the limit exists on the open interval (a,b), then
f(α)a(a)=limt→a+f(α)a(t) |
and
f(α)a(b)=limt→b−f(α)a(t). |
It is obvious that the conformable derivative coincides with the classical derivative for α=1, with another words, the conformable derivative is conformable for α→1. Furthermore, the conformable derivative is equivalent to f(α)a(t)=(t−a)1−αf′(t) for differentiable functions. By using this property, one can easily claim that the conformable derivative is not conformable for α→0
limα→0f(α)a(t)=tf′(t)≠f(t) |
which is also observed for most of local and nonlocal derivatives. Due to the similarities between the conformable and the classical derivative definitions, the conformable derivative has the advantageous properties such as Leibniz, quotient and chain rules (see [1,13]).
Definition 2.2. ([13]) Let f∈Cn{[a,b],R} for n∈N+ and 0<α≤1 is the order of conformable derivative. The left sequential conformable derivative of order n is defined by
(dαadtαa)nf(t)=dαadtαadαadtαa⋯dαadtαa⏟n−timesf(t). |
The following definition is adopted from [15].
Definition 2.3. For a real function f with m variables in which all xi∈[ai,bi],i=1,2,…,m, the left conformable partial derivative of order 0<α≤1 with respect to xi is defined by
∂αai∂aixαif(x1,x2,…,xm)=limε→0f(x1,…,xi−1,xi+ε(xi−ai)1−α,…,xm)−f(x1,…,xm)ε. | (2.2) |
Definition 2.4. ([13]) The left conformable integral of order 0<α≤1 for a real function f defined on a closed interval [a,b] is defined by
Iαaf(t)=t∫af(x)dαax=t∫af(x)(x−a)α−1dx. | (2.3) |
Theorem 2.1. ([13]) Let f,g:[a,b]→R be functions such that fg is differentiable. Then the conformable partial integration formula is given by
b∫af(t)g(α)(t)dαat=f(t)g(t)b|a−b∫ag(t)f(α)a(t)dαat. | (2.4) |
The conformable variational calculus is an expansion of the variational calculus that aims to minimize (or maximize) a performance index containing at least one conformable derivative term defined via the conformable or classical integral [31]. CVCP with classical performance index is classified as GCVCP by Lazo and Torres [30] who investigate the generalized conformable Euler-Lagrange equation of GCVCPs that contains both the conformable and classical derivatives. In this section, we aim to present the generalized transversality condition for GCVCPs. Therefore, we consider the following GCVCP
J[x]=te∫t0L(t,x(t),x(α)t0(t))dt, | (3.1) |
where 0<α≤1 is the order of conformable derivative, L∈Cα{[t0,te]×R2,R} is the Lagrangian and x(t)∈Cα{[t0,te],R} is an unknown function. Since the problem (3.1) does not contain the classical derivative term, it is a special case of the problem introduced in [30], so we first need to obtain the generalized conformable Euler-Lagrange equation of GCVCP (3.1). For this purpose, we temporarily assume that the problem (3.1) has fixed endpoint conditions also named separately as initial and terminal conditions presented as
x(t0)=x0 and x(te)=xe, where x0,xe∈R. | (3.2) |
Before going further, we give the basic concepts needed to solve the GCVCPs.
Definition 3.1. ([31]) For the problem (3.1) the functions x(t)∈Cα{[t0,te],R} satisfying the endpoint conditions (3.2) are called admissible functions.
Definition 3.2. ([31]) Assume x∗(t) is a minimum and x(t) is an admissible function of the problem (3.1). For an arbitrary function η∈Cα{[t0,te],R} provided η(t0)=η(te)=0, the weak variation is defined by
x(t)=x∗(t)+εη(t), | (3.3) |
where ε is a positive small quantity independent from t, x∗ and η.
The following lemma is the extended version of the fundamental lemma of variational calculus to the conformable variational calculus.
Lemma 3.1. ([30]) Assume 0<α≤1 and μ,η∈C{[t0,te],R} are arbitrary functions. If the following equation holds
te∫t0μ(t)η(t)dαt0t=0 | (3.4) |
for any continuous η satisfying η(t0)=η(te)=0, then
μ(t)=0. | (3.5) |
for all t∈[t0,te].
The following theorem gives the necessary optimality condition for the GCVCP given by (3.1)−(3.2). It is worth to note that the theorem will be proven via the variation method which differs from the proof given by [30] via Gateaux derivative.
Theorem 3.1. (Generalized Conformable Euler–Lagrange Equation) If x(t) is a minimum (or maximum) of the GCVCP defined by (3.1)−(3.2), then x(t) provides the generalized conformable Euler-Lagrange equation given as follows
(t−t0)1−α∂L∂x−dαt0dtαt0((t−t0)1−α∂L∂x(α)t0)=0. | (3.6) |
Proof. Let x∗(t) be the extremum of the performance index J and η(t) be an arbitrary function from the class of Cα{[t0,te],R} that satisfies η(t0)=η(te)=0. Then the weak variations can be written as
x(t)=x∗(t)+εη(t), | (3.7) |
x(α)t0(t)=x∗(α)t0(t)+εη(α)t0(t), | (3.8) |
for |ε|≪1. To find the extremum x∗(t), the variation of the performance index should be examined according to the weak variations of (3.7) and (3.8) as
ΔJ=te∫t0L(t,x∗(t)+εη(t),x∗(α)t0(t)+εη(α)t0(t))dt−te∫t0L(t,x∗(t),x∗(α)t0(t))dt. |
After that, Taylor expansion of the function L near the point (x∗,x∗(α)t0) for t∈[t0,te] via the pair of variables (εη,εη(α)t0) is replaced in ΔJ:
ΔJ=te∫t0{L(t,x∗(t),x∗(α)t0(t))+∂L∂xεη(t)+∂L∂x(α)t0εη(α)t0(t)+12(∂2L∂x2(εη(t))2+2∂2L∂x∂x(α)t0εη(t)εη(α)t0(t)+∂2L∂x(α)2t0(εη(α)t0(t))2)}dt−te∫t0L(t,x∗(t),x∗(α)t0(t))dt+O(ε3). | (3.9) |
Equation (3.9) can be written
ΔJ=εδJ+ε2δ2J+O(ε3), |
in which
δJ=te∫t0(∂L∂xη(t)+∂L∂x(α)t0η(α)t0(t))dt | (3.10) |
is the first and
δ2J=te∫t012(∂2L∂x2η(t)2+2∂2L∂x∂x(α)t0η(t)η(α)t0(t)+∂2L∂x(α)2t0(η(α)t0(t))2)dt | (3.11) |
is the second variations. Suppose x∗(t) is the minimum of J, then
ΔJ=εδJ+ε2δ2J+O(ε3)≥0 | (3.12) |
must be satisfied for all admissible η(t). If both sides of the inequality (3.12) are divided by ε, since ε may be positive or negative, the following inequalities arise
ΔJ=δJ+εδ2J+O(ε2)≤0,for ε>0,ΔJ=δJ+εδ2J+O(ε2)≥0,for ε<0. |
When the limit of the inequalities are taken for ε→0, it is held both δJ≤0 and δJ≥0 which give the first variation must be zero for all admissible η(t). Therefore, the necessary condition for the minimum is expressed as follows:
δJ=te∫t0(∂L∂xη(t)+∂L∂x(α)t0η(α)t0(t))dt=0. | (3.13) |
Note that, the same relation is also obtained if x∗(t) is assumed to be a maximum of J. To achieve a more useful result, we use the conformable partial integration formula and remove the conformable derivative term η(α)t0(t) from the equation (3.13). Therefore, we define the following transformation
∼L(t,x(t),x(α)t0(t))=(t−t0)1−αL(t,x(t),x(α)t0(t)) | (3.14) |
which transforms the classical integral of the first variation to a conformable integral as below
δJ=te∫t0(∂∼L∂xη(t)+∂∼L∂x(α)t0η(α)t0(t))dαt0t=0. | (3.15) |
Applying the conformable partial integration formula (2.4) to (3.15) gives
te∫t0∂∼L∂x(α)t0η(α)t0(t)dαt0t=−te∫t0dαt0dtαt0(∂∼L∂x(α)t0)η(t)dαt0t | (3.16) |
since η(t0)=η(te)=0. Substituting (3.16) into (3.15) and arranging the first variation according to the arbitrary function η(t) leads to
δJ=te∫t0(∂∼L∂xη(t)−dαt0dtαt0(∂∼L∂x(α)t0))η(t)dαt0t=0. |
By using the Lemma (3.1), the generalized conformable Euler-Lagrange equation is finally acquired from the first variation as
(t−t0)1−α∂L∂x−dαt0dtαt0((t−t0)1−α∂L∂x(α)t0)=0, |
that completes the proof.
It is known that the problems encountered in physics and engineering may not include the sufficient number of endpoint conditions. However, the prior properties of the variational calculus is that the methodology always supply the right number of conditions named as transversality conditions. Therefore, variational calculus problems whose one or both endpoint conditions are missing, can be solved with Euler-Lagrange equation and transversality conditions. Here, we propose the generalized transversality condition for a GCVCP in which the performance index has fixed initial condition x(t0)=x0 and has variable terminal condition x(τ)=xτ. For generality, the variable terminal condition is assumed to lie on a given curve. It is useful to indicate that the initial condition can be also assumed to be free, but we prefer to skip this assumption for brevity of the proof procedure which can be easily extended to the variable initial condition. So, we consider the GCVCP defined by the performance index
J[x]=τ∫t0L(t,x(t),x(α)t0(t))dt | (3.17) |
with fixed initial condition
x(t0)=x0 is fixed, | (3.18) |
and variable terminal condition
x(τ)=xτ is free which lies on a differentiable curve ν(t), | (3.19) |
where 0<α≤1 is the order of conformable derivative, L∈Cα{[t0,τ]×R2,R} is the Lagrangian and x=x(t)∈Cα{[t0,τ],R} is an unknown function.
Theorem 3.2. (Generalized Transversality Condition for GCVCP) If a GCVCP is defined by (3.17)-(3.19), then the generalized transversality condition is
L(τ,x(τ),x(α)t0(τ))Δτ+(τ−t0)1−α∂L∂x(α)t0|τεη(τ)=0, | (3.20) |
where η∈Cα{[t0,τ],R}) and Δτ are arbitrary functions. Additionally, it is also assumed that Δτ is from the remainder class of O(ε).
Proof. Assume x∗(t) be an extremum of J and intersects with the target curve ν(t) at t=τ∗. Consider the following weak variations for |ε|≪1
x(t)=x∗(t)+εη(t), x(α)t0(t)=x∗(α)t0(t)+εη(α)t0(t),τ=τ∗+εΔτ, | (3.21) |
in which η(t0)=0 and η(τ)≠0. The variation of the performance index is calculated via these weak variations as
ΔJ=τ∗+εΔτ∫t0L(t,x∗(t)+εη(t),x∗(α)t0(t)+εη(α)t0(t))dt−τ∗∫t0L(t,x∗(t),x∗(α)t0(t))dt | (3.22) |
and organized in the following form
ΔJ=τ∗+εΔτ∫τ∗L(t,x∗(t)+εη(t),x∗(α)t0(t)+εη(α)t0(t))dt+τ∗∫t0(L(t,x(t),x(α)t0(t))−L(t,x∗(t),x∗(α)t0(t)))dt. | (3.23) |
For t∈[t0,τ], expanding the function L to a Taylor series near the point (x∗,x∗(α)t0) via the pair of variables (εη,εη(α)t0) and substituting the expansion in ΔJ gives
ΔJ=τ∗+εΔτ∫τ∗(L(t,x∗(t),x∗(α)t0(t))+∂L∂xεη(t)+∂L∂x(α)t0εη(α)t0(t))dt+τ∗∫t0(L(t,x∗(t),x∗(α)t0(t))+∂L∂xεη(t)+∂L∂x(α)t0εη(α)t0(t)−L(t,x∗(t),x∗(α)t0(t)))dt+O(ε2). | (3.24) |
When the first integral is calculated via a numerical method, the following equation is held
ΔJ=(L(τ∗,x∗(τ∗),x∗(α)t0(τ∗))+∂L∂x|τ∗εη(τ∗)+∂L∂x(α)t0|τ∗εη(α)t0(τ∗))εΔτ+τ∗∫t0(∂L∂xεη(t)+∂L∂x(α)t0εη(α)t0(t))dt+O(ε2). | (3.25) |
Again, we can separate the variation to the first and second variations marked respectively with δJ and δ2J. Thus, the first variation is equal to zero due to the requirement to be an extremum:
δJ=L(τ∗,x∗(τ∗),x∗(α)t0(τ∗))Δτ+τ∗∫t0(∂L∂xη(t)+∂L∂x(α)t0η(α)t0(t))dt=0. | (3.26) |
If the transformation (3.14) is applied to (3.26), the first variation is expressed via the following conformable integral
δJ=L(τ∗,x∗(τ∗),x∗(α)t0(τ∗))Δτ+τ∗∫t0(∂∼L∂xη(t)+∂∼L∂x(α)t0η(α)t0(t))dαt0t=0, |
then the conformable integration by part for formula (2.4) is used to remove η(α)t0 term as below
δJ=L(τ∗,x∗(τ∗),x∗(α)t0(τ∗))Δτ+∂∼L∂x(α)t0|τ∗η(τ∗)+τ∗∫t0η(t)(∂∼L∂x−dαt0dtαt0(∂∼L∂x(α)t0))dαt0t=0. | (3.27) |
Since the term inside of the above integral is generalized conformable Euler-Lagrange equation, the generalized transversality condition of GCVCP is achieved as
L(τ∗,x(τ∗),x(α)t0(τ∗))Δτ+(τ∗−t0)1−α∂L∂x(α)t0|τ∗η(τ∗)=0. | (3.28) |
The generalized transversality condition (3.28) depends on the unknown arbitrary functions η and Δτ. To identify these functions and get more clear generalized transversality conditions, we examine the special cases considering for the terminal condition (3.19) by the below corollary.
Corollary 3.1. (Specialized Transversality Conditions for GCVCP) Taylor series expansion of the weak variation x(t)=x∗(t)+εη(t) at the point t=τ=τ∗+εΔτ according to εΔτ near the point τ∗ as
x(t)=x∗(τ∗+εΔτ)+εη(τ∗+εΔτ)=x∗(τ∗)+˙x∗(τ∗)εΔτ+εη(τ∗)+O(ε2). | (3.29) |
Intersection of the curve x(t) and the target curve ν(t) at the point t=τ leads also the requirement of Taylor expansion for ν(τ∗+εΔτ) via εΔτ about the point τ∗ as
ν(τ∗+εΔτ)=ν(τ∗)+˙ν(τ∗)εΔτ+O(ε2). | (3.30) |
Equating (3.29) with (3.30) and ignoring the remainder terms give the perturbation function as
η(τ∗)=(˙ν(τ∗)−˙x∗(τ∗))Δτ. | (3.31) |
Therefore, substituting the equation (3.31) into the equation (3.28) leads to the following transversality condition:
L(τ∗,x(τ∗),x(α)t0(τ∗))+(τ∗−t0)1−α∂L∂x(α)t0|τ∗(˙ν(τ∗)−˙x∗(τ∗))=0. | (3.32) |
This equation can be more simplified if either τ or xτ is specified while the other value is completely free. Therefore, we examine special forms of (3.32) by the following situations.
● Terminal Curve: If τ is free and xτ belongs to a differentiable target curve ν(t), then the transversality condition is
L(τ,x(τ),x(α)t0(τ))+(τ−t0)1−α∂L∂x(α)t0|τ(˙ν(τ)−˙x(τ))=0. | (3.33) |
● Vertical Terminal Line: If τ is fixed and xτ is free which coincides to a straight line perpendicular to the x-axis, then the transversality condition (3.32) can be rewritten by assuming ν(τ)≠0 as
1˙ν(τ)(L(τ,x(τ),x(α)t0(τ))−(τ−t0)1−α˙x(τ))+(τ−t0)1−α∂L∂x(α)t0|τ=0. |
For infinite ˙ν(τ), the transversality condition is reduced to
(τ−t0)1−α∂L∂x(α)t0|τ=0. | (3.34) |
Since this condition naturally arise in variational formulation, it is also referred as "natural boundary condition".
● Horizontal Terminal Line: If τ is free and xτ is fixed which coincides to a straight line parallel to x-axis, then for ˙ν(τ)=0 the transversality condition is obtained as
L(τ,x(τ),x(α)t0(τ))−(τ−t0)1−α˙x(τ)∂L∂x(α)t0|τ=0. | (3.35) |
Example 1. Find the minimum of
J[x]=τ∫0(1−x(α)0(t))2dt | (3.36) |
for each of the following cases:
● x(0)=1, x(τ)=2
● x(0)=1, τ=2.
● x(0)=1 and x(τ) lies on the differentiable curve ν(t)=2+(tα−1)2.
Solution 1. The generalized conformable Euler–Lagrange equation of the problem is
x(α)0(t)x(α)0(t)+(1−α)t−αx(α)0(t)=(1−α)t−α. |
Using the method for conformable linear differential equations with variable coefficients given by [18], the solution is achieved as
x(t)=c1+c2t2α−12α−1. |
To get the minimum solution of the problem, we need to determine the unknown coefficients. From the initial condition x(0)=1, the first coefficient is achieved as c1=1 which leads to
x(t)=1+c2t2α−12α−1. |
The other coefficient is specified via transversality conditions, so that we examine the each cases separately
● First Case: Since τ is free and x(τ)=2, we obtain
c2τ2α−1=2α−1. | (3.37) |
Writing the functions L and x into the transversality condition (3.35) gives
(1−x(α)0(τ))2−2c2τα−1(1−x(α)0(τ))=0. | (3.38) |
If the equations (3.37) and (3.38) are solved, then the minimum solution is identified. To illustrate the minimum solution, we choose α=0.7 and find the parameters asτ=0.2701 and c2=0.6752 via Symbolic Toolbox of MATLAB. According to these parameters, the minimum solution is plotted in Figure 1.
● Second Case: Since τ=2 and x(τ) is free, the transversality condition (3.34) gives
2τα−1(1−c2τα−1)=0. |
For α=0.7, the unknown coefficient is determined as c2=2.4623 and then the minimum function is plotted via these parameters in Figure 2.
● Third Case: Since τ and xτ are free but xτ lies on the curve ν(t)=2+(tα−1)2, the first equation needed to solve is obtained by equating x(τ)=ν(τ) as
c2τ2α−12α−1−(τα−1)2−1=0. | (3.39) |
The second equation arising from the transversality condition (3.33) is
(1−c2τα−1)(1−c2τα−1−(4α(τα−1)−2c2τα−1))=0. | (3.40) |
Solving the equations (3.39)−(3.40) gives the minimum solution which is illustrated in Figure 3 for the chosen value of α=0.7 corresponds to τ=0.3715 and c2=0.7430.
Conformable optimal control aims to find the state and control functions which minimize (or maximize) a given performance index defined with classical or conformable integrals subjected to a conformable dynamic constraints, [31]. In this section, we consider COCPs with classical performance index which we call generalized conformable optimal control problems (GCOCPs) and analyze the generalized transversality conditions for GCOCPs by utilizing the generalized transversality conditions for GCVCPs. For this purpose, firstly we need to know the necessary optimality conditions for GCOCPs. When these conditions are examined, it is concluded that the GCOCPs have the same optimality conditions with COCPs which are proposed by [30] via Hamiltonian formalism. To remind these conditions, we consider the following COCP which aims to minimize the performance index
J[x,u]=te∫t0L(t,x(t),u(t))dt |
with fixed initial x(t0) and fixed terminal x(te) conditions and subjected to a conformable dynamic constraint
x(α)t0(t)=f(t,x(t),u(t)) |
in which 0<α≤1 is the order of conformable derivative, L,f are the functions from the class of Cα{[t0,te]×R2,R}, x is the state from the class of Cα{[t0,te],R} and u is the control function. The Hamiltonian formulation for GCOCP is
H(t,x,u,λ)=−L(t,x,u)+λ(t)f(t,x,u) | (4.1) |
where λ∈Cα{[t0,te],R} is a Lagrange multiplier. Then the conformable Euler–Lagrange equations also known as necessary conditions of optimality are as follows
{x(α)t0(t)=∂H∂λ(t,x,u,λ),λ(α)t0(t)=−∂H∂x(t,x,u,λ),∂H∂u(t,x,u,λ)=0. | (4.2) |
As to be for GCVCPs, one or both endpoint conditions may be missing for GCOCPs. Here, we again assume that the initial condition is fixed while the terminal condition is variable via the same reason explained in the previous section. Therefore, we consider the most general form of GCOCPs that aims to minimize the performance index
J[x,u]=τ∫t0L(t,x(t),u(t))dt | (4.3) |
with fixed initial condition
x(t0)=x0 | (4.4) |
and variable terminal condition
xτ=x(τ) lies on a differentiable curve ν(t) | (4.5) |
which is subjected to a conformable dynamic constraint
x(α)t0(t)=f(t,x(t),u(t)), | (4.6) |
in which 0<α≤1 is the order or the conformable derivative, L,f are the functions from the class of Cα{[t0,tf]×R2,R}, x is the state from the class of Cα{[t0,tf],R} and u is the control function. The transversality conditions of the problem is given by the below theorem.
Theorem 4.1. (Transversality Conditions for GCOCP) If a GCOCP is defined by (4.3)−(4.6), then the generalized conformable transversality condition is
[−H(τ,x(τ),u(τ))+λ(τ)x(α)t0(τ)]Δτ+(τ−t0)1−αλ(τ)η(τ)=0, | (4.7) |
where η∈Cα{[t0,τ],R} and Δτ are arbitrary functions. Additionally, it is also assumed that Δτ is from the remainder class of O(ε).
Proof. Let x∗(t) and u∗(t) are the optimum functions in which x∗(t) intersects with the target curve ν(t) at t=τ∗. To identify optimum functions, we consider arbitrary functions η(t), ζ(t), Λ(t), Δτ that have α−order conformable derivative on [t0,τ], and define the weak variations at t∈[t0,τ] for |ε|≪1 as
x(t)=x∗(t)+εη(t), x(α)t0(t)=x∗(α)t0(t)+εη(α)t0(t), u(t)=u∗(t)+εζ(t),λ(t)=λ∗(t)+εΛ(t),τ=τ∗+εΔτ, | (4.8) |
Note that, η(t0)=ζ(t0)=Λ(t0)=0 and η(τ)≠0,ζ(τ)≠0,Λ(τ)≠0 since τ is unspecified. To use the Lagrange multiplier technique we define
Φ[x,u,λ]=τ∫t0λ(t)(x(α)t0(t)−f(t,x(t),u(t)))dt |
and then give the following functional
I[x,u,λ]=J[x,u]+Φ[x,u,λ] | (4.9) |
which obviously coincides to the performance index J since Φ is identically zero. To construct a clear proof, we examine the variations of the functionals J and Φ separately as in the following
ΔJ=τ∗+εΔτ∫τ∗L(t,x(t),u(t))dt+τ∗∫t0{L(t,x(t),u(t))−L(t,x∗(t),u∗(t))}dt | (4.10) |
ΔΦ=τ∗+εΔτ∫τ∗λ(t)(x(α)t0(t)−f(t,x(t),u(t)))dt+τ∗∫t0{λ(t)(x(α)t0(t)−f(t,x(t),u(t)))−λ∗(t)(x∗(α)t0(t)−f(t,x∗(t),u∗(t)))}dt. | (4.11) |
When the functions L and f are expanded to the Taylor series near the point (x∗,u∗) for t∈[t0,τ] via the pair of variables (εη,εζ), we get
ΔJ=τ∗+εΔτ∫τ∗{L(t,x∗(t),u∗(t))+∂L∂xεη(t)+∂L∂uεζ(t)}dt+τ∗∫t0(∂L∂xεη(t)+∂L∂uεζ(t))dt+O(ε2), | (4.12) |
ΔΦ=τ∗+εΔτ∫τ∗(λ∗(t)+εΛ(t)){x∗(α)t0(t)+εη(α)t0(t)−f(t,x∗(t),u∗(t))−∂f∂xεη(t)−∂f∂uεζ(t)}dt+τ∗∫t0(λ∗(t)+εΛ(t)){x∗(α)t0(t)+εη(α)t0(t)−f(t,x∗(t),u∗(t))−∂f∂xεη(t)−∂f∂uεζ(t)−λ∗(t)(x∗(α)t0(t)−f(t,x∗(t),u∗(t)))}dt+O(ε2). | (4.13) |
Then, calculating the first integrals in the above variations via a numerical method gives the first variations respectively as
δJ=L(τ∗,x∗(τ∗),u∗(τ∗))Δτ+τ∗∫t0(∂L∂xη(t)+∂L∂uζ(t))dt=0, | (4.14) |
δΦ=λ∗(τ∗)(x∗(α)t0(τ∗)−f(τ∗,x∗(τ∗),u∗(τ∗)))Δτ+τ∗∫t0λ∗(t)(η(α)t0(t)−∂f∂xη(t)−∂f∂uζ(t))dt+τ∗∫t0Λ(t)(x∗(α)t0(t)−f(t,x∗(t),u∗(t)))dt=0. | (4.15) |
To use the conformable partial integration formula (2.4) for the term τ∗∫t0λ∗(t)η(α)t0(t)dt in (4.15), we define the following transformation
˜λ∗(t)=(t−t0)1−αλ∗(t) |
which converts the classical integral to the conformable one. If the conformable integral is solved by the formula (2.4), then the integral is written be aware of η(t0)=0 as
τ∗∫t0λ∗(t)η(α)t0(t)dt=˜λ∗(τ∗)η(τ∗)−τ∗∫t0λ∗(α)t0(t)η(t)dαt0t. | (4.16) |
Substituting (4.16) into (4.15) leads to
δΦ=λ∗(τ∗)(x∗(α)t0(τ∗)−f(τ∗,x∗(τ∗),u∗(τ∗)))Δτ+˜λ∗(τ∗)η(τ∗)−τ∗∫t0{η(t)(˜λ∗(t)∂f∂x+˜λ∗(α)t0)+ζ(t)˜λ∗(t)∂f∂u}dαt0t+τ∗∫t0Λ(t)(x∗(α)t0(t)−f(t,x∗(t),u∗(t)))(t−t0)1−αdαt0t=0. |
From (4.9), the first variation δI corresponds to the sum of δJ and δΦ, equation (4.9) must be written in sense of conformable integral which gives the first variation of I is achieved via Hamiltonian formalism as
δI=(−H(τ∗,x∗(τ∗),u∗(τ∗))+λ∗(τ)x∗(α)t0(τ∗))Δτ+˜λ∗(τ∗)η(τ∗)−τ∗∫t0η(t)(˜λ∗(α)t0(t)+(t−t0)1−α∂H∂x)dαt0t−τ∗∫t0ζ(t)(t−t0)1−α∂H∂udαt0t+τ∗∫t0Λ(t)(x∗(α)t0(t)−f(t,x∗(t),u∗(t)))(t−t0)1−αdαt0t=0. | (4.17) |
Because the necessary optimality conditions are equal to zero, the above integrals vanish, so the transversality condition of GCOCP is acquired as
(−H(τ∗,x∗(τ∗),u∗(τ∗))+λ∗(τ∗)x∗(α)t0(τ∗))Δτ+(τ∗−t0)1−αλ∗(τ∗)η(τ∗)=0. | (4.18) |
More clear formula for the generalized transversality condition of GCOCP is investigated by the following corollary which removes the unknown arbitrary functions of η, Δτ from (4.18) and specializes the transversality conditions for particular cases.
Corollary 4.1. (Specialized Transversality Conditions for GCOCP) Since the arbitrary function η(τ∗) and the value Δτ arise in (4.18) are unknown, the relation (3.31) given by Corollary 3.1 is valid for η(τ∗) and Δτ. Substituting (3.31) into (4.18), and arranging (4.18) according to Δτ gives
(−H(τ∗,x∗(τ∗),u∗(τ∗))+λ∗(τ∗)x∗(α)t0(τ∗)+(τ∗−t0)1−αλ∗(τ∗)(˙ν(τ∗)−˙x∗(τ∗)))Δτ=0. | (4.19) |
Since Δτ is arbitrary, the generalized conformable transversality condition is finally achieved as
−H(τ∗,x∗(τ∗),u∗(τ∗))+λ∗(τ∗)x∗(α)t0(τ∗)+(τ∗−t0)1−αλ∗(τ∗)(˙ν(τ∗)−˙x∗(τ∗))=0 | (4.20) |
● Terminal Curve: The differentiable target curve ν(t) defines the behavior of the terminal point such that xτ=ν(τ). Therefore, the generalized transversality condition in the most general form must be
−H(τ,x(τ),u(τ))+λ(τ)x(α)t0(τ)+(τ−t0)1−αλ(τ)(˙ν(τ)−˙x(τ))=0 | (4.21) |
● Vertical Terminal Line: For fixed τ and variable xτ the target curve corresponds to a straight line perpendicular to the x-axis whose slope at τ is infinite. Therefore, the transversality condition for infinite ˙ν(τ) is obtained as
(τ−t0)1−αλ(τ)=0. | (4.22) |
● Horizontal Terminal Line: If τ is free and xτ is fixed, then ˙ν(τ)=0. Thus, the transversality conditions is to be
−H(τ,x(τ),u(τ))+λ(τ)x(α)t0(τ)−(τ−t0)1−αλ(τ)˙x(τ)=0. | (4.23) |
The following example is adopted from the example given in [33] by replacing the Riemann-Liouville fractional derivative with conformable derivative.
Example 2. Find the pair of state x(t) and control u(t) functions which minimizes the quadratic performance index
J[x,u]=12τ∫0(x2(t)+u2(t))dt, | (4.24) |
subjected to dynamic constraint
x(α)0(t)=−x(t)+u(t), | (4.25) |
with the following endpoint conditions
x(0)=10, x(τ)=1. | (4.26) |
Solution 2. The necessary conditions for optimality (4.2) via the Hamiltonian function of the problem
H(t,x,u,λ)=−12(x2(t)+u2(t))+λ(t)(−x(t)+u(t)). |
is calculated as
x(α)0(t)=−x(t)+u(t),λ(α)0(t)=x(t)+λ(t),u(t)=λ(t). | (4.27) |
These equations are solved by the analytical method proposed in [19] and the state function is obtained as
x(t)=c1e√2tαα+c2e−√2tαα. | (4.28) |
The fixed initial condition x(0)=10 and fixed terminal condition x(τ)=1 for free τ give
c1+c2=10, x(τ)=c1e√2ταα+(10−c1)e−√2ταα, |
respectively. From (4.27), the control function u(t) is obtained as
u(t)=(1+√2)(10−c1)e√2tαα+(1−√2)c1e−√2tαα. | (4.29) |
To identify the unknown coefficient c1, the generalized transversality condition (4.23) is used which gives
x2(τ)+u2(τ)−2u(τ)τ1−α˙x(τ)=0. | (4.30) |
Substituting (4.28), (4.29) into (4.30) and solving (4.30) present the optimal control. To illustrate the optimal control law, the solution is achieved by Symbolic Toolbox of MATLAB for chosen value of α=0.7 resulted as τ=0.1307, c1 =−5.0903 and c2 =15.0903. Thus, the state and control functions are plotted in Figure 4.
Now, we consider a GCOCP in the case of the terminal curve.
Example 3. Consider the terminal curve problem
J[u]=τ∫0√(1+u2(t))dt, | (4.31) |
subjected to dynamic constraint
x(α)0(t)=u(t), | (4.32) |
with the following endpoint conditions
x(0)=0, x(τ)=10−τ2 | (4.33) |
and find the optimal control law which minimize the GCOCP.
Solution 3. The necessary conditions for the optimality via the Hamiltonian function H(t,x,u,λ)=√(1+u2(t))−λ(t)u(t) are determined as
−λ(α)0(t)=0,x(α)0(t)=u(t),λ(t)=u(t)√(1+u2(t)). |
These equations can be solved by the analytical method proposed in [19] which gives the optimal state function as
x(t)=tααu(t)+c1. |
The initial condition gives c1=0 and so the terminal curve leads to
10−τ2=τααu(τ). |
This equation with the generalized transversality condition (4.21) calculated as
τ1−α(−2τ−u(τ)τα−1)u(τ)√(1+u2(τ))+√(1+u2(τ))=0 |
are solved together in which the solution presents the optimal control law. For chosen value of α=0.7, the solutions are achieved as τ=0.01230 and u(t)=152.1364 via MATLAB. Figure 5 depicts the optimal state and control functions, respectively.
The paper focuses on a pretty new area which is called conformable optimal control involved also conformable variational calculus as expected. Although optimal control problems with local or nonlocal operators have been investigated via performance index defined by the classical or own integral operators, classical performance indexes are more often used in application problems. Therefore, the conformable variational calculus and optimal control problems defined by the classical performance indexes are taken under consideration. The generalized Euler-Lagrange equations and the transversality conditions of the problems are obtained. Moreover, generalized transversality conditions are specialized to the terminal curve, fixed time horizon and fixed endpoint problems. Related examples for each problem have been solved and illustrated for a specific value of the order of conformable derivative.
This research is supported by a fund from the Scientific Research Projects Department of Balıkesir University under the grant number BAP 2018/022.
The authors declare that there is no conflicts of interest in this paper.
[1] |
R. Khalil, M. A. Horani, A. Yousef, A new definition of fractional derivative, J. Comput. Appl. Math., 264 (2014), 65-70. doi: 10.1016/j.cam.2014.01.002
![]() |
[2] | A. Kilbas, H. Srivastava, J. Trujillo, Theory and Applications of Fractional Differential Equations, North-Holland Mathematics Studies, New York, 2006. |
[3] |
A. Atangana, D. Baleanu, New fractional derivatives with nonlocal and non-singular kernel: theory and application to heat transfer model, Therm. Sci., 20 (2016), 763-769. doi: 10.2298/TSCI160111018A
![]() |
[4] |
V. E. Tarasov, No violation of the Leibniz rule. No fractional derivative, Commun. Nonlinear Sci., 18 (2013), 2945-2948. doi: 10.1016/j.cnsns.2013.04.001
![]() |
[5] |
V. E. Tarasov, No nonlocality. No fractional derivative, Commun. Nonlinear Sci., 62 (2018), 157-163. doi: 10.1016/j.cnsns.2018.02.019
![]() |
[6] |
M. D. Ortigueira, J. T. Machado, What is a fractional derivative?, J. Comput. Phys., 293 (2015), 4-13. doi: 10.1016/j.jcp.2014.07.019
![]() |
[7] |
G. S. Teodoro, J. T. Machado, E. C. De Oliveira, A review of definitions of fractional derivatives and other operators, J. Comput. Phys., 388 (2019), 195-208. doi: 10.1016/j.jcp.2019.03.008
![]() |
[8] |
J. Weberszpil, J. A. Helayël-Neto, Variational approach and deformed derivatives, Physica A, 450 (2016), 217-227. doi: 10.1016/j.physa.2015.12.145
![]() |
[9] |
W. Rosa, J. Weberszpil, Dual conformable derivative: Definition, simple properties and perspectives for applications, Chaos Soliton. Fract., 117 (2018), 137-141. doi: 10.1016/j.chaos.2018.10.019
![]() |
[10] | D. R. Anderson, E. Camrud, D. J. Ulness, On the nature of the conformable derivative and its applications to physics, Journal of Fractional Calculus and Applications, J. Fract. Calc. Appl., 10 (2019), 92-135. |
[11] |
H. W. Zhou, S. Yang, S. Q. Zhang, Conformable derivative approach to anomalous diffusion, Physica A, 491 (2018), 1001-1013. doi: 10.1016/j.physa.2017.09.101
![]() |
[12] |
D. Avcı, B. B. İskender Eroǧlu, N. Özdemir, The Dirichlet problem of a conformable advection-diffusion equation, Therm. Sci., 21 (2017), 9-18. doi: 10.2298/TSCI160421235A
![]() |
[13] |
T. Abdeljawad, On conformable fractional calculus, J. Comput. Appl. Math., 279 (2015), 57-66. doi: 10.1016/j.cam.2014.10.016
![]() |
[14] | F. Silva, D. Moreira, M. Moret, Conformable Laplace transform of fractional differential equations, Axioms, 7 (2018), 55. |
[15] | A. Atangana, D. Baleanu, A. Alsaedi, New properties of conformable derivative, Open Math., 13 (2015), 889-898. |
[16] | R. Khalil, A. Yousef, M. Al Horani, et al. Fractional analytic functions, Far East J. Math. Sci., 103 (2018), 113-123. |
[17] | S. Uçar, N. Yılmaz Özgür, B. B. İskender Eroǧlu, Complex conformable derivative, Arab. J. Geosci., 12 (2019), 201. |
[18] | M. A. Hammad and R. Khalil, Abel's formula and Wronskian for conformable fractional differential equations, Int. J. Differ. Equ. Appl., 13 (2014), 177-183. |
[19] | D. R. Anderson, D. J. Ulness, Results for conformable differential equations, preprint, 2016. |
[20] | E. Ünal, A. Gökdoǧan, E. Çelik, Solutions around a regular α singular point of a sequential conformable fractional differential equation, Kuwait J. Sci., 44 (2017), 9-16. |
[21] | N. Sene, Solutions for some conformable differential equations, Progr. Fract. Differ. Appl. 4 (2018), 493-501. |
[22] |
Z. Hammouch, T. Mekkaoui, P. Agarwal, Optical solitons for the Calogero-Bogoyavlenskii-Schiff equation in (2 + 1) dimensions with time-fractional conformable derivative, Eur. Phys. J. Plus, 133 (2018), 248-253. doi: 10.1140/epjp/i2018-12096-8
![]() |
[23] |
H. Bulut, T. A. Sulaiman, H. M. Başkonuş, Dark, bright optical and other solitons with conformable space-time fractional second-order spatiotemporal dispersion, Optik, 163 (2018), 1-7. doi: 10.1016/j.ijleo.2018.02.086
![]() |
[24] | M. Yavuz, Novel solution methods for initial boundary value problems of fractional order with conformable differentiation, An International Journal of Optimization and Control: Theories & Applications (IJOCTA), 8 (2017), 1-7. |
[25] |
E. Ünal, A. Gökdoǧan, Solution of conformable fractional ordinary differential equations via differential transform method, Optik, 128 (2017), 264-273. doi: 10.1016/j.ijleo.2016.10.031
![]() |
[26] | M. Yavuz, N. Özdemir, A different approach to the European option pricing model with new fractional operator, Math. Model. Nat. Pheno., 13 (2018), 12. |
[27] |
F. Evirgen, Conformable fractional gradient based dynamic system for constrained optimization problem, Acta Phys. Pol. A, 132 (2017), 1066-1069. doi: 10.12693/APhysPolA.132.1066
![]() |
[28] |
W. S. Chung, Fractional Newton mechanics with conformable fractional derivative, J. Comput. Appl. Math., 290 (2015), 150-158. doi: 10.1016/j.cam.2015.04.049
![]() |
[29] |
B. B. İskender Eroǧlu, D. Avcı, N. Özdemir, Optimal control problem for a conformable fractional heat conduction equation, ACTA Phys. Pol. A, 132 (2017), 658-662. doi: 10.12693/APhysPolA.132.658
![]() |
[30] |
J. M. Lazo, D. F. M. Torres, Variational calculus with conformable fractional derivatives, IEEE/CAA J. Autom. Sin., 4 (2017), 340-352. doi: 10.1109/JAS.2016.7510160
![]() |
[31] | B. B. İskender Eroǧlu, D. Yapişkan, Local generalization of transversality conditions for optimal control problem, Math. Model. Nat. Pheno., 14 (2019), 310. |
[32] | T. Chiranjeevi, R. K. Biswas, Closed-form solution of optimal control problem of a fractional order system, Journal of King Saud University-Science, 2019. |
[33] |
R. K. Biswas, S. Sen, Free final time fractional optimal control problems, J. Franklin. I., 351 (2014), 941-951. doi: 10.1016/j.jfranklin.2013.09.024
![]() |
1. | Derya Avcı, Beyza Billur İskender Eroğlu, Optimal control of the Cattaneo–Hristov heat diffusion model, 2021, 232, 0001-5970, 3529, 10.1007/s00707-021-03019-z | |
2. | Beyza Billur İskender Eroğlu, Dilara Yapışkan, Comparative analysis on fractional optimal control of an SLBS model, 2023, 421, 03770427, 114840, 10.1016/j.cam.2022.114840 | |
3. | Süleyman ÖĞREKÇİ, Serkan ASLIYÜCE, Fractional variational problems on conformable calculus, 2021, 70, 1303-5991, 719, 10.31801/cfsuasmas.820580 | |
4. | Maria N. F. Barreto, Gastão S. F. Frederico, José Vanterler da C. Sousa, Juan E. Nápoles Valdés, CALCULUS OF VARIATIONS AND OPTIMAL CONTROL WITH GENERALIZED DERIVATIVE, 2023, 53, 0035-7596, 10.1216/rmj.2023.53.1337 | |
5. | Dilara Yapışkan, Mine Yurtoğlu, Derya Avcı, Beyza Billur İskender Eroğlu, Ebenezer Bonyah, A Novel Model for Monkeypox Disease: System Analysis and Optimal Preventive Strategies, 2023, 47, 2731-8095, 1665, 10.1007/s40995-023-01525-4 | |
6. | Tharmalingam Gunasekar, Shanmugam Manikandan, Vediyappan Govindan, Piriadarshani D, Junaid Ahmad, Walid Emam, Isra Al-Shbeil, Symmetry Analyses of Epidemiological Model for Monkeypox Virus with Atangana–Baleanu Fractional Derivative, 2023, 15, 2073-8994, 1605, 10.3390/sym15081605 | |
7. | Beyza Billur İskender Eroğlu, Dilara Yapışkan, 2023, Chapter 3, 978-3-031-33182-4, 39, 10.1007/978-3-031-33183-1_3 | |
8. | Adina Veronica Crişan, Cresus Fonseca de Lima Godinho, Claudio Maia Porto, Ion Vasile Vancea, Conformable Lagrangian Mechanics of Actuated Pendulum, 2025, 13, 2227-7390, 1634, 10.3390/math13101634 | |
9. | Murugan Suba, Arikrishnan Venkatesan, Tharmalingam Gunasekar, Shanmugam Manikandan, Prabakaran Raghavendran, Shyam Sundar Santra, Shubhankar Karmakar, 2025, Chapter 11, 978-3-031-90916-0, 160, 10.1007/978-3-031-90917-7_11 |