Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter Open Access October 16, 2018

Analysis of projectile motion in view of conformable derivative

  • Abraham Ortega Contreras , J. Juan Rosales García EMAIL logo , Leonardo Martínez Jiménez and Jorge Mario Cruz-Duarte
From the journal Open Physics

Abstract

This paper presents new solutions for twodimensional projectile motion in a free and resistive medium, obtained within the newly established conformable derivative. For free motion, we obtain analytical solutions and show that the trajectory, height, flight time, optimal angle, and maximum range depend on the order of the conformable derivative, 0 < γ ≤ 1. Likewise, we analyse and simulate the projectile motion in a resistive medium by assuming several scenarios. The obtained trajectories never exceed the ordinary ones, given by γ = 1, unlike results reported in other studies.

1 Introduction

In the last three decades, much work has been performed on fractional calculus (FC). FC is the natural generalization of ordinary derivatives and integrals. It deals with operators having a non-integer (arbitrary) order. Today, there are excellent books and reviews on fractional calculus and its applications in science and engineering [1234567]. FC has been employed in numerous practical and theoretical problems, but, for the sake of brevity, we mention a few of them as follows: the design of optimal control systems [8]; the analysis of anomalous relaxation and diffusion processes [9], and the regularized long-wave equation [10]; the study of a bead sliding on a wire [11]; and the investigation of chaotic systems [12], heat conduction [13], nerve impulse transmission [14], chemical kinetic systems [15], and oscillating circuits [16]. However, this generalization of the ordinary calculus is still inconsistent in physical and geometric interpretation, because several definitions of fractional derivatives have been proposed. These definitions include the Riemann-Liouville, Grunwald-Letnikov, Weyl, Caputo, Marchaud and Riesz fractional derivatives [12345]. These fractional derivatives seldom satisfy the well-known formulae, e.g., the product, quotient, and chain rules. As a result, researchers have been trying to construct new definitions of fractional derivatives and integrals. As an illustrative example, helpful work on the theory of derivatives and integrals is done in [17].

Definition

Let f : [0, ∞) → ℝ be a function. Then, the γ-th order conformable derivative of f is defined by [17] as

Tγ(f)(t)=limϵ0f(t+ϵt1γ)f(t)ϵ,t>0,(1)

where γ ∈ (0,1]. If f is γ-differentiable in some (0,a), a > 0, and limt0+fγ(t) exists, then, it is defined as

fγ(0)=limt0+fγ(t),(2)

and the conformable integral of a function, Iγa(f)(t), starting from a ≥ 0, is defined as,

Iγa(f)(t)=atf(x)x1γdx.(3)

The integral is the usual Riemann improper integral, and γ ∈ (0,1]. The most important properties of this conformable derivative are given in the following theorem.

Theorem

Let γ ∈ (0,1], and f and g be γ-differentiable at any point t > 0. Then,

  1. Tγ(af + bg) = aTγ(f) + bTγ(g) for all a,b ∈ ℝ,

  2. Tγ (tp) = ptp-γ for all p ∈ ℝ,

  3. Tγ(λ) = 0 for all constant function f(t) = λ,

  4. Tγ (fg) = fTγ (g) + gTγ(f),

  5. Tγ(fg)=gTγ(f)fTγ(g)g2,

  6. Tγ(f)(t)=t1γdfdt, if f is differentiable.

This conformable derivative has attracted the interest of researchers in recent years since it seems to satisfy all the requirements of the standard derivative [18], [19]. Hence, there is a large amount of work done in this area at the present time [2021222324].

We would like to point out that the term “fractional” is not the one used in the notation of fractional calculus [25]. However, the properties of this conformable derivative make it suitable for investigating real systems and to get new insights due to the presence of the fractional parameter 0 < γ ≤ 1.

In this paper, we use the conformable derivative to discuss two-dimensional projectile motion. First, we consider free two-dimensional projectile motion. Subsequently, new formulae for the trajectory, height, range, flight time, optimal angle, and maximum range are obtained. We show that these formulae depend on the conformable derivative order 0 < γ ≤ 1, and in the particular case of γ = 1, they become ordinary. Finally, we consider two-dimensional projectile motion in a resistive medium.

This paper is organized as follows. In Section 2, the main results of ordinary free projectile motion are reviewed. In Section 3, new analytical solutions for this motion are obtained. In Section 4, we consider motion in resistive media. Finally, the conclusions are given in Section 5.

2 Ordinary free projectile motion

We study the motion of a projectile moving in a free two-dimensional space. The projectile is treated as a particle of mass (m), under a uniform gravitational force (g), and disregarding any other resisting or external force. Assuming the particle starts from rest (i.e., x0 = y0 = 0 m), with an initial velocity of modulus v0 and angle θ with respect to the horizontal axis x, then, the classical equations of motion for this system, in the x, y plane, are,

md2xdt2=0,(4)
md2ydt2=mg,(5)

with the initial conditions,

x0=x(0)=0,    x˙(0)=v0cosθ,(6)
y0=y(0)=0,    y˙(0)=v0sinθ.(7)

The solutions, given in parametric form, are

x(t)=v0tcosθ,    y(t)=v0tsinθ12gt2,(8)

and the corresponding velocities are

vx=v0cosθ,    vy=v0sinθgt.(9)

By eliminating the time t in the expressions (8), we obtain the trajectory equation, given by the parabola

y(x)=xtanθg2v02x2cos2θ.(10)

Moreover, three interesting features for studying two-dimensional projectile motion are described as follows. Range (R) is defined as the horizontal distance traveled by the particle, from start to end. It can be obtained with (10) assuming the landing condition, y(x = R) = 0 m,

R=2v02gsinθcosθ.(11)

Maximum altitude (ymax) is the highest point reached by the particle. In this point, vy = 0 m/s and, from (9), we have tmax = v0 sinθ/g. Substituting tmax in the second equation of (8), we have

ymax=v02sin2θ2g.(12)

Flight time (Tflight) is the amount of time that the projectile spends in the air, i.e., from its shooting to its landing [26]. It is determined as

Tflight=2tmax=2v0sinθg.(13)

3 Free projectile motion in view of conformable derivative

Usually, authors replace integer derivative operators with fractional ones on a purely mathematical basis [262728]. However, this practice is not completely correct, from the physical and engineering point of view, and a dimensional correction in the new equation is necessary. Having this in mind, in [29] has been proposed a systematic way to construct fractional differential equations; this has been successfully applied in [30313233]. This procedure considers the introduction of parameters such as σt and σx with an appropriate dimensionality. In other words,

ddt1σt1γdγdtγ,    ddx1σx1γdγdxγ,(14)

where γ is an arbitrary parameter which represents the order of the derivative 0 < γ ≤ 1. σt is a parameter representing the fractional time components in the system with dimensionality in time unit [t]; likewise, the parameter σx has dimension of length unit [l], and represents the spatial fractional components [29]. These are the only conditions imposed on parameters σt and σx.

Therefore, we fractionalize the equations (4) and (5), considering the first relation in (14), and using the conformable derivative, in particular the property 6 from the previous theorem

dγdtγf(t)=t1γddtf(t),(15)

and arrive at [34]

ddtt1γσ1γddt,    0<γ1,(16)

where we have omitted the subscript t in σ. Substituting into (9), we obtain the corresponding conformable differential equations

dxdt=v0σ1γtγ1cosθ,(17)
dydt=v0σ1γtγ1sinθgσ1γtγ.(18)

We can integrate them directly

x(t;γ)=σ1γγtγv0cosθ,(19)
y(t;γ)=gσ1γγ+1tγ+1+σ1γγtγv0sinθ.(20)

We also determine the components of the conformable velocity of the system by differentiating with respect to time as

vx(t;γ)=v0σ1γtγ1cosθ,(21)
vy(t;γ)=v0σ1γtγ1sinθgσ1γtγ.(22)

Unlike expressions (8) and (9), the coordinates and velocities in (19)-(22) depend on the fractional parameter 0 < γ ≤ 1. In the particular case γ = 1, those are reduced to (8) and (9). We want to emphasize that the solutions given by (19) and (20) are qualitatively different from the solutions previously obtained in [26], [27] and [32]. Therefore, in the projectile motion under the conformable derivative sense, we can achieve new formulae for the trajectory, maximum height, range, flight time, maximum range, and optimal angle, as will be shown:

  1. Trajectory (y = y(x)): This is determined by eliminating t from (19) and replacing it in (20), for an arbitrary value of γ ∈ (0,1]), giving

    y(x;γ)=xtanθgσ1γγ+1(γxσ1γv0cosθ)1γ+1.(23)

    Note that (23) becomes (10), when γ = 1.

  2. Maximum height (H): This is obtained by using the condition vy(t;γ) = 0 m/s, and employing (22), to find

    tmax=v0sinθg,(24)

    since tmax indicates the time when the particle has reached its maximum height; then, substituting this time in (20), we obtain the maximum height H, as

    H=ymax(θ;γ)=σ1γ(v0sinθ)γ+1(γ+1)gγ.(25)

    We have the ordinary maximum height (12), when γ = 1.

  3. Range (R): The maximum displacement is calculated from the trajectory equation, and it occurs when y (x = R) = 0 m. That is,

    R=xmax(θ;γ)=σ1γγ(1+γγg)γv0γ+1cosθ(sinθ)γ.(26)

    Note again, if γ = 1, we have (11).

  4. Flight time (Tflight): This is defined as the time value T at which the projectile hits the ground, y(t = T) = 0 m, as

    Tflight(θ;γ)=(γ+1γ)v0sinθg.(27)

    In the case γ = 1, we have (13).

  5. Maximum range (Rmax): This feature is considerably interesting for both practical and theoretical studies. It can be determined by finding the optimal projection angle θop, and evaluating it in Rmax = R(θop). The necessary condition to maximize the range, RRmax is given by [dR/dθ]θop=0. Thus, from (26) we have

    dRdθ|θop=ddθ[cosθ(sinθ)γ]θop=(sinθop)γ+1+γ(cosθop)2(sinθop)γ1=0.(28)

    By solving this equation with respect to θop, we obtain

    θop=tan1γ,    0<γ1.(29)

    In the case γ = 1, the optimal angle is θop = π/4 (as expected). Hence, using trigonometric formulae,

    sinθ=tanθ1+(tanθ)2=γ1+γ,(30)
    cosθ=11+(tanθ)2=11+γ,(31)

    and, by replacing this expression in (26), we obtain the maximum projectile range Rmax as

    Rmax=σ1γγ(1+γγg)γ(γ1+γ)γ/2v0γ+11+γ.(32)

    From this expression, in the case γ = 1, we get the classical maximum range Rcl,maximum=v02/g.

These new formulae have the same dimensionality as in the ordinary case, because σ has dimension of seconds. Furthermore, for demonstrative purposes, Figure 1 shows the trajectory (Figure 1a), the maximum height (Figure 1b), and the range (Figure 1c) of a particle in projectile motion with different values of γ.

Figure 1 Plots of trajectory, maximum height and range for different values of γ, with v0 = 100 m/s, g = 9.81m/s2, and θ = 40°
Figure 1

Plots of trajectory, maximum height and range for different values of γ, with v0 = 100 m/s, g = 9.81m/s2, and θ = 40°

4 Resistive projectile motion in view of conformable derivative

The motion equations of a projectile going through an isotropic medium in two-dimensional space with a resistive force proportional to its velocity, have the form [35]:

dvxdt+kvx=0,(33)
dvydt+kvy=g,(34)

where k is a positive constant and its dimensionality is the inverse of seconds [k] = s-1. The initial conditions are given by

x0=x(0)=0m,x˙=v0cosθ,(35)
y0=y(0)=0m,y˙=v0sinθ,(36)

namely, the projectile starts from rest, with an initial velocity of module v0 and an angle θ with respect the x-axis. The solutions of these equations (33) and (34) and satisfying the initial conditions (35) and (36) are given in [35]

vx(t)=v0cosθekt,(37)
vy(t)=gk+(v0sinθ+gk)ekt,(38)

and

x(t)=v0cosθk(1ekt),(39)
y(t)=gkt+1k(v0sinθ+gk)(1ekt).(40)

Now, we fractionalize the equations (33) and (34). In the case of motion in a resistive medium, we can use k instead of σ in (36) [33], namely

ddtk1γt1γddt,    0<γ1.(41)

Substituting the operator in (41) into equations (33) and (34), we have the conformable differential equations

dvxdt+kγtγ1vx=0,(42)
dvydt+kγtγ1vy=gkγ1tγ1.(43)

These expressions are ordinary linear, homogeneous and non-homogeneous differential equations, hence, they can be solved easily using the ordinary methods. The analytical solutions that satisfy initial conditions (35) and (36) are:

vx(t;γ)=v0cosθe(kγγtγ),(44)
vy(t;γ)=gk+(v0sinθ+gk)e(kγγtγ).(45)

In the case γ = 1, these equations become (37) and (38). Once again, replacing (41) in the equations (44) and (45), and integrating them, we get the parametric formulae as:

x(t;γ)=v0cosθk(1ekγγtγ),(46)
y(t;γ)=gkγ2γtγ+1k(v0sinθ+gk)(1ekγγtγ).(47)

Moreover, if γ = 1, then equations (46) and (47) become the ordinary equations (39) and (40). For the particular case of angle θ = 90°, we have

x(t;γ)=0,(48)
y(t;γ)=gkγ2γtγ+1k(v0+gk)(1ekγγtγ).(49)

Therefore, we only have a vertical conformable trajectory [34].

As an illustrative example, Figure 2 presents the projectile motion for a defined medium with k = 0.1 s-1, in terms of x(t;γ), y(t;γ) and t. The main tri-dimensional view of this behavior is displayed with its planes.

Figure 2 Projectile motion in terms of x(t; γ), y(t; γ) and t, within a resistive medium (k = 0.1 s−1) and for different values of γ (i.e., γ = 1.000, 0.875, and 0.750), with v0 = 22 m/s, g = 9.81 m/s2, and θ = 60°
Figure 2

Projectile motion in terms of x(t; γ), y(t; γ) and t, within a resistive medium (k = 0.1 s−1) and for different values of γ (i.e., γ = 1.000, 0.875, and 0.750), with v0 = 22 m/s, g = 9.81 m/s2, and θ = 60°

In addition, the vertical displacement for different values of γ, k, velocities and angles are shown in the Figure 3. It is observed that the trajectory varies with γ (cf. Figure 1), but there are negligible changes for different values of k.

Figure 3 The vertical displacement of a body for different values of γ, k, and initial conditions
Figure 3

The vertical displacement of a body for different values of γ, k, and initial conditions

5 Conclusion

In this work, we have considered free and resistive two-dimensional projectile motion in the conformable derivative sense. New formulae for the trajectory (23), height (25), range (26), flight time (27), optimal angle (29), and maximum range (32) were obtained. Also, calculations were performed in the case of a resistive medium (46) and (47). The obtained formulae are qualitatively different from the solutions previously obtained in [26], [27], [32] and [33]. Moreover, we found that the new formulae depend on the conformable derivative order 0 < γ ≤ 1, and they became ordinary in the particular case γ = 1. We noticed that the obtained trajectories never exceeded the ordinary ones where γ = 1, unlike the results obtained in other studies [27].

Acknowledgement

The authors acknowledge the fruitful discussion with I. Lyanzuridi and D. Baleanu. This work was supported by CONACyT and DICIS-University of Guanajuato.

References

[1] Oldham K.B., Spanier J., The Fractional Calculus. Theory and Applications of Differentiation and Integration of Arbitrary Order, 1974, Academic Press, New YorkSearch in Google Scholar

[2] Podlubny I., Fractional Differential Equations, 1999, Academic Press, New YorkSearch in Google Scholar

[3] Kilbas A., Srivastava H., Trujillo J., Theory and Applications of Fractional Differential Equations, 2006, Math. Studies, North-Holland, New YorkSearch in Google Scholar

[4] Hilfer R., Applications of Fractional Calculus in Physics, 2000, World Scientific, Singapore10.1142/3779Search in Google Scholar

[5] Caputo M.,Mainardi F., A new dissipation model based on memory mechanism, Pure Appl. Geophys. 1971, 91, 134-147.10.1007/BF00879562Search in Google Scholar

[6] Baleanu D., Guvenc Z.B., Machado J.A.T., New Trends in Nanotechnology and Fractional Calculus Applications, 2010, Springer New York.10.1007/978-90-481-3293-5Search in Google Scholar

[7] Uchaikin V., Fractional Derivatives for Physicists and Engineers, 2013, Springer H.D., London, New York.10.1007/978-3-642-33911-0Search in Google Scholar

[8] Baleanu D., Jajarmi A., Hajipour M., A New Formulation of the Fractional Optimal Control Problems Involving Mittag-Leffler Nonsingular Kernel, J. Optim. Theory Appl., 2017, 175(3), 718-737.10.1007/s10957-017-1186-0Search in Google Scholar

[9] Sun H.G., Hao X., Zhang Y., Baleanu D., Relaxation and diffusion models with non-singular kernels,Phys. A Stat. Mech. Appl., 2017, 468(1), 590-596.10.1016/j.physa.2016.10.066Search in Google Scholar

[10] Kumar D., Singh J., Baleanu D., Sushila, Analysis of regularized long-wave equation associated with a new fractional operator with Mittag-Leffler type kernel,Phys. A Stat. Mech. Appl., 2018, 492, 155-167.10.1016/j.physa.2017.10.002Search in Google Scholar

[11] Baleanu D., Jajarmi A., Asad J.H., Blaszczyk T., The motion of a bead sliding on a wire in fractional sense, Acta Phys. Pol. A, 2017, 131(6), 1561-1564.10.12693/APhysPolA.131.1561Search in Google Scholar

[12] Hajipour M., Jajarmi A., Baleanu D., An Efficient Non-standard Finite Difference Scheme for a Class of Fractional Chaotic Systems, J. Comput. Nonlinear Dyn., 2018, 13(2), 021013, 1-9.10.1115/1.4038444Search in Google Scholar

[13] Zhao D., Singh J., Kumar D., Rathore S., Yang X., An efficient computational technique for local fractional heat conduction equations in fractal media, J. Nonlinear Sci. Appl., 2017, 10(4), 1478-1486.10.22436/jnsa.010.04.17Search in Google Scholar

[14] Kumar D., Singh J., Baleanu D., A new numerical algorithm for fractional Fitzhugh-Nagumo equation arising in transmission of nerve impulses, Nonlinear Dyn., 2018, 91(1), 307-317.10.1007/s11071-017-3870-xSearch in Google Scholar

[15] Singh J., Kumar D., Baleanu, D., On the analysis of chemical kinetics system pertaining to a fractional derivative with Mittag-Leffler type kernel, Chaos, 2017, 27(10), 103113.10.1063/1.4995032Search in Google Scholar PubMed

[16] Kumar D., Agarwal R.P., Singh J., A modified numerical scheme and convergence analysis for fractional model of Lienard’s equation, J. Comput. Appl. Math.,2018, 339, 405-413.10.1016/j.cam.2017.03.011Search in Google Scholar

[17] Khalil R., Al Horani M., Yousef A., Sababheh M., A new definition of fractional derivative, J. Comput. Appl. Math., 2014, 264, 65-70.10.1016/j.cam.2014.01.002Search in Google Scholar

[18] Katugampola U.N., A new fractional derivative with classical properties, 2014, arXiv:1410.6535v1.Search in Google Scholar

[19] Abdeljawad T., On conformable fractional calculus, J. Comput. Appl. Math., 2015, 279, 57-66.10.1016/j.cam.2014.10.016Search in Google Scholar

[20] Cenesiz Y., Baleanu D, Kurt A., Tasbozan O., New exact solutions of Burger’s type equations with conformable derivative, Waves in Random and complex Media, 2017,27(1),103-116.10.1080/17455030.2016.1205237Search in Google Scholar

[21] Zhao D., Li T., On conformable delta fractional calculus on time scales, J. Math. Computer Sci., 2016, 16, 324-335.10.22436/jmcs.016.03.03Search in Google Scholar

[22] Al Horani M., AbuHammad M., Khalil R., Variation of parameters for local fractional non-homogeneous linear differential equations, J. Math. Computer Sci., 2016, 16, 147-153.10.22436/jmcs.016.02.03Search in Google Scholar

[23] Abu Hammad I, Khalil R., Fractional Fourier series with applications, Am. J. Comput. Applied Math., 2014, 4(6), 187-191.Search in Google Scholar

[24] Atangana A., Baleanu D., Alsaedi A., New properties of conformable derivative, Open Math., 2015, 13, 889-898.10.1515/math-2015-0081Search in Google Scholar

[25] Tarasov V.E., No violation of the Leibniz rule. No fractional derivative, Commun. Nonlinear Sci. Numer. Simulat., 2013, 18, 2945-2948.10.1016/j.cnsns.2013.04.001Search in Google Scholar

[26] Ahmad B., Batarfi H., Nieto J.J., Otero-Zarraquinos O., Shammakh W., Projectile motion via Riemann-Liouville calculus, Adv. Diff. Equations, 2015, 63(1), 1-14.10.1186/s13662-015-0400-3Search in Google Scholar

[27] Ebaid A., Analysis of projectile motion in view of fractional calculus, Appl. Math. Modelling, 2011, 35, 1231-1239.10.1016/j.apm.2010.08.010Search in Google Scholar

[28] Sau Fa K., A falling body problem through the air in view of the fractional derivative approach, Physica A, 2005, 350, 199-206.10.1016/j.physa.2004.11.041Search in Google Scholar

[29] Rosales J.J., Gomez J.J., Guía M., Tkach V.I., Fractional Electromagnetic Waves, 2011, LFNM2011 International Conference on Laser and Fiber-Optical Networks Modeling 4-8 September, Kharkov, Ukraine.10.1109/LFNM.2011.6144969Search in Google Scholar

[30] Rosales J.J., Guía M., Gomez J.F., Tkach V.I., Fractional Electromagnetic Wave,Disc. Nonl. Compl.,2012, 1(4), 325-335.10.1109/LFNM.2011.6144969Search in Google Scholar

[31] Gomez-Aguilar J.F., Rosales-Garcia J.J., Bernal-Alvarado J.J., Cordova-Fraga T., Guzman-Cabrera R., Fractional Mechanical Oscillator, Rev. Mex. Fis., 2012, 58, 348-352.Search in Google Scholar

[32] Rosales García J.J., Guía M., Martínez J., Baleanu D., Motion of a particle in a resistive medium using fractional calculus approach, Proceed. Rom. Acad. Ser. A, 2013, 14, 42-47.Search in Google Scholar

[33] Rosales J.J., Guía M., Gomez F.,Aguilar F., Martínez J., Twodimensional fractional projectile motion in a resisting medium, Cent. Eur. J. Phys., 2014, 12(7), 517-520.10.2478/s11534-014-0473-8Search in Google Scholar

[34] Ebaid A., Masaedeh B., El-Zahar E., A new fractional model for the falling body problem, Chin. Phys. Lett., 2017, 34(2), 02020110.1088/0256-307X/34/2/020201Search in Google Scholar

[35] Thornton S.T., Marion J.B., Classical dynamics of particles and systems, 2004, Thomson Brooks/ColeSearch in Google Scholar

Received: 2017-05-10
Accepted: 2018-04-11
Published Online: 2018-10-16

© 2018 A. Ortega Contreras et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 20.4.2024 from https://www.degruyter.com/document/doi/10.1515/phys-2018-0076/html
Scroll to top button