Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access July 28, 2020

Optimal control for generating excited state expansion in ring potential

  • Hoon Yu , Seung Jin Kim and Jung Bog Kim EMAIL logo
From the journal Open Physics

Abstract

We applied an optimal control algorithm to an ultra-cold atomic system for constructing an atomic Sagnac interferometer in a ring trap. We constructed a ring potential on an atom chip by using an RF-dressed potential. A field gradient along the radial direction in a ring trap known as the dimple-ring trap is generated by using an additional RF field. The position of the dimple is moved by changing the phase of the RF field [1]. For Sagnac interferometers, we suggest transferring Bose–Einstein condensates to a dimple-ring trap and shaking the dimple potential to excite atoms to the vibrational-excited state of the dimple-ring potential. The optimal control theory is used to find a way to shake the dimple-ring trap for an excitation. After excitation, atoms are released from the dimple-ring trap to a ring trap by adiabatically turning off the additional RF field, and this constructs a Sagnac interferometer when opposite momentum components are overlapped. We also describe the simulation to construct the interferometer.

1 Introduction

Atom interferometers are dynamic tools applied for precision measurements and for studying the wave nature of interacting matter over the past few decades. Measurements of inertial effects such as acceleration and rotation [2,3,4,5], physical constants [6,7], gravitational constant [8,9], and dark energy [10,11] in free space have been demonstrated. Interferometers in a confined trap using the wave nature of matter have also been constructed in various potentials such as multi-well [12,13], ring [14,15,16,17,18], optical lattice [19,20,21,22], and certain wave-guide traps [20,21,22,23,24,25,26,27]. Atom interferometers require splitting, reflecting, and recombining coherent sources just as an optical interferometer. Various ways have been adopted to realize interferometric components by using Raman pulse [2,3,4,5,7], Bragg diffraction [25,26,27], and adiabatically changing trapping potentials [11,12].

An optimal control algorithm is a process to find a way from the initial state to the desired state by perturbing systems. It was adopted to optimize atomic systems and manipulate the desired atomic states such as ± n k momentum states or excited states [19,20,28,29,30,31,32]. It is also used for machine-learning when it works with closed-loop control and constructing atom interferometers to find ways to get interferometric components.

In this article, we describe a process to demonstrate an atomic Sagnac interferometer using ultra-cold atoms by using a vibrational-excited state in a ring trap potential. The RF-dressed trapping potentials that we experimentally demonstrated for a ring and a dimple-ring trap in ref. [1] are reviewed in Section 2. In Section 3, we introduce a way to construct a Sagnac interferometer. First, we adiabatically load ultra-cold atoms from the initial trapping potential to the ground state of the dimple-ring potential. And then, we shake the dimple-ring potential to excite atoms. The way to shake the potential is obtained from an optimal control algorithm. We describe the details of the optimal control algorithm we used in Section 3. The excited atoms are released to a ring potential, and they start to expand. We also simulated the Sagnac interferometry when opposite momentum components are overlapped.

2 RF-dressed dimple-ring potential

We demonstrated dimple-ring potential near an atom chip by using an RF-dressed potential [1,33]. Figure 1 shows the pattern of wires on our atom chip. We used the center wire to create an Ioffe magnetic trap with a bias magnetic field and the dimple wire for enhancing the longitudinal confinement of the trap. Each side wire, RF A and B, is used to create an RF-dressed potential, and the RF field on the center wire creates asymmetry in the potential such as the tilted double-well and dimple-ring potential [1]. The magnetic field is given by static, B s , and oscillating, B RF , fields such as

(1) B s ( r ) = G x e x G y e y + B I e z

(2) B RF = B RF A e x cos ω RF A t + B RF B e y cos ω RF B t δ + B RF z e z cos ω RF z t δ z ,

where G is the gradient of the static trap and B I is the magnitude of the Ioffe trap-bottom. B RF A , B RF B , and B RF z represent the amplitude of each RF field. δ and δ z are the relative phases to the RF field on the A wire.

Figure 1 
               (a) Wire pattern for the dimple-ring potential; (b) dimple-ring potential as 
                     
                        
                        
                           
                              
                                 δ
                              
                              
                                 z
                              
                           
                        
                        {\delta }_{z}
                     
                  , where G = 50 T/m, 
                     
                        
                        
                           
                              
                                 B
                              
                              
                                 I
                              
                           
                        
                        {B}_{\text{I}}
                     
                   = 
                     
                        
                        
                           
                              
                                 B
                              
                              
                                 RF
                              
                           
                        
                        {B}_{\text{RF}}
                     
                   = 1 G, 
                     
                        
                        
                           
                              
                                 B
                              
                              
                                 RF
                              
                              
                                 Z
                              
                           
                        
                        {B}_{\text{RF}}^{Z}
                     
                   = 100 mG, 
                     
                        
                        
                           
                              
                                 ω
                              
                              
                                 RF
                              
                           
                        
                        {\omega }_{\text{RF}}
                     
                   = 2π × 650 kHz, and δ = π/2 [1].
Figure 1

(a) Wire pattern for the dimple-ring potential; (b) dimple-ring potential as δ z , where G = 50 T/m, B I = B RF = 1 G, B RF Z = 100 mG, ω RF = 2π × 650 kHz, and δ = π/2 [1].

RF-induced potential can be written as the following equation:

(3) V ( r ) = m F g F μ B Δ ( r ) 2 + Ω ( r ) 2 .

In equation (3), m F is the magnetic quantum number, g F is the Lande g-factor of the hyperfine structure, and μ B is the Bohr magneton [1,34]. The detuning term, Δ ( r ) , and the coupling term between static and oscillating fields, Ω ( r ) , can be given by the following equations:

(4) Δ ( r ) = B s ( r ) ω RF g F μ B ,

(5) Ω ( r ) = B eff ( r ) × B s ( r ) B s ( r ) .

The effective field in equation (5) is calculated as follows:

(6) B eff ( r ) = y [ β ( r ) ] z [ α ( r ) ] B RF A 0 0 + z g F g F δ y [ β ( r ) ] z [ α ( r ) ] 0 B RF B 0 + z g F g F δ z y [ β ( r ) ] z [ α ( r ) ] 0 0 B RF z ,

with a rotating matrix, i [ θ ] , along the i axis, where tan [ α ( r ) ] = B s ( r ) y / B s ( r ) x and tan [ β ( r ) ] = B s ( r ) x 2 + B s ( r ) y 2 / B s ( r ) z .

Since the detuning term is rotationally symmetric in the cylindrical coordinate, the coupling term determines the geometry of the RF-dressed potential. For simplicity, we assumed ω RF A = ω RF B = ω RF z = ω RF and B RF A = B RF B = B RF . When B RF z is not zero and its amplitude is much less than B RF , the coupling term difference between cases of B RF z 0 and B RF z = 0 can be written as follows:

(7) Ω ( r ) 2 B RF z 0 Ω ( r ) 2 B RF z = 0 = G B s ( r ) + B I 2 B s ( r ) 2 B RF × B RF z ( x cos δ z y sin δ z ) ,

where δ = π / 2 .

Figure 1 shows the wire pattern of our chip and the potential for the dimple-ring trap as δ z , described by equations (6) and (7).

When B RF z is zero, the effective potential has a ring shape on the XY plane. We can also generate a field gradient along the radial direction of the ring trap called the dimple-ring potential by using I RF D to generate B RF z , and the position of the trap minimum can be controlled by changing the phase difference between I RF A and I RF D . Experimental demonstration to generate condensation in the ring trap and the dimple-ring trap is described in ref. [1] and [33].

3 Optimal control algorithms for exciting atoms in a dimple-ring trap

We suggest the following method for constructing a Sagnac-type atom interferometer using a ring potential. First, we generated condensed atoms in an Ioffe trap, and then we transferred the atoms to a dimple-ring trap, i.e., to one of the dimples shown in Figure 1(b). We were able to add momentum to atoms by shaking the dimple-ring potential along the tangential direction to excite atoms in the trap. In order to make the excitation, we applied an optimal control algorithm to find a way of shaking the dimple-ring potential by changing δ z . After excitation, we transferred the excited atoms to a ring potential by adiabatically ramping down the AC current of the dimple wire, I RF z . Since there is no confinement along the radial direction in a ring trap as shown in Figure 1(b), atoms will freely expand in a ring trap along clock-wise and counter clock-wise directions. Atoms would overlap similar to an optical Sagnac interferometer finally.

For simplicity, we assume that the trap and the perturbation are in one dimension to apply an optimal algorithm. In Figure 2, blue shows the initial state as the ground state in a dimple-ring trap, and red shows the desired state, which is the first excited state. When the 1D trapping potential is in black (Figure 2), we solved the Gross–Pitaevskii equation (GPE) to find the ground state of condensed atoms, and the excited state is calculated by optimizing GPE orthogonality with the ground state.

Figure 2 
               Dimple-ring potential given in black, where 
                     
                        
                        
                           
                              
                                 δ
                              
                              
                                 Z
                              
                           
                           =
                           4
                           π
                           /
                           5
                        
                        {\delta }_{Z}=4\pi /5
                     
                  , blue is the ground state as the initial state.
Figure 2

Dimple-ring potential given in black, where δ Z = 4 π / 5 , blue is the ground state as the initial state.

The goal of the optimal algorithm is to find the appropriate shaking function of the dimple-ring trap to excite the atoms, and it runs by the following way [28,30].

First, the algorithm calculates the time evolution of the initial state with a given initial shaking, which can be any function. The initial function in our case is a sinusoidal function, which has the same initial and final conditions, λ ( 0 ) = λ ( T ) = 0 . After shaking, in a second, we calculate the cost function using equation (8), which says how the result is close to the desired state and how smooth the shaking is.

(8) J ( ψ ( T ) , λ ) = 1 2 1 ψ D ψ ( T ) 2 + γ 2 0 T [ λ ̇ ( t ) ] 2 d t .

In the above equation, ψ D is the desired state, ψ ( T ) is the wave function at the final time of shaking, γ is a weight function showing how smooth the shaking is, and λ is a control parameter related to shaking, which is the displacement of the dimple-ring trap or δ z in this case. The optimal algorithm seeks the control, λ , that minimizes the cost function, J ( ψ , λ ) . In order to minimize the cost function, ref. [28] and [30] introduce the Lagrange function with the constrain that the wave function has to fulfill GPE as shown in equation (9)

(9) L ( ψ , p , λ ) = J ( ψ , λ ) + Re o T p ( r , t ) i ψ ̇ ( t ) H λ + g ψ ( t ) 2 ) ψ ( t ) d t ,

where p ( r , t ) is the adjoint function of Lagrangian, H λ is the Hamiltonian of the shaken dimple-ring potential as a function of λ , and g is the coupling constant of GPE related to atom–atom interaction. At the minimum of J ( ψ , λ ) , L ( ψ , p , λ ) must fulfill equation (10),

(10a) t L ψ ̇ L ψ = 0 ,

(10b) t L p ̇ L p = 0 ,

(10c) t L λ ̇ L λ = 0 .

From equations (8) and (9), we can write equation (10) as equation (11):

(11a) i p t = ( H λ + 2 g ψ 2 ) p + g ψ 2 p ,

(11b) i ψ t = H λ + g ψ 2 ψ ,

(11c) γ 2 λ t 2 = Re ψ H λ λ p ,

at initial conditions ψ ( 0 ) = ψ 0 , λ ( 0 ) = λ 0 , and λ ( T ) = λ T , where ψ 0 is the ground state in the dimple-ring trap, and λ 0 and λ T are the first and the final control parameters. We can also find the final adjoint, p, by integrating equation (11a) with time such as i p = ψ D ψ ( T ) ψ D .

It is difficult for the initial guess of λ to fulfill equation (11) simultaneously. However, equation (11c) can tell us how the control goes to minimize J ( ψ , λ ) by adding a fraction term given as equation (12) to previous control when equations (11a) and (11b) are satisfied.

(12) L λ = γ λ ̈ Re ψ H λ λ p .

Therefore, it calculates p ( t ) from equation (11a) and p ( T ) as reverse in time for the third step. And then, a new λ ( t ) can be calculated from equation (12). By keeping these iterations, we could optimize the control to get to the desired state.

Besides calculating the time-evolution of the wave function and the way to minimize the cost function we used as described in ref. [30], there are other mathematical ways about fractional derivatives to calculate differential equations [36,37,38,39,40,41,42,43,44,45].

Figure 3(a) shows the atomic wave function as time when the dimple-ring potential is shaken, resulting in the optimized control to excite the ground state of the dimple-ring trap for 20 ms and hold the excited state in the dimple-ring trap for 20 ms. Figure 3(b) shows the optimized control in black and the initial guess in red. The cost function as a function of the iteration of the optimal algorithm is shown in the box. A smaller cost function means that the result is closer to the desired state and the control is smoother. We can see that the cost function is exponentially decreased as the iteration increased, and the decay time to the minimum depends on the initial guess of control. The time step of our optimal algorithm is set to 25 µs.

Figure 3 
               (a) Excitation of the ground state in the dimple-ring trap for 20 ms and holding it for 20 ms. (b) Initial guess of control in red and the optimized control in black, and cost function as a function of iteration is given in the box.
Figure 3

(a) Excitation of the ground state in the dimple-ring trap for 20 ms and holding it for 20 ms. (b) Initial guess of control in red and the optimized control in black, and cost function as a function of iteration is given in the box.

The range of phase for the optimized control is less than 3 deg, and the main frequency component is about 2.4 kHz in our case. Therefore, it requires a phase shifter that has 650 kHz in the main frequency, ±1.5 deg in the bandwidth, and 2.4 kHz in a dynamic range at least when we experimentally apply the result.

Figure 4 shows the next step of constructing the atomic Sagnac interferometer in a ring trap. After excitation, we suggest adiabatically transferring atoms from the dimple-ring trap to a ring trap by ramping down the RF field in the center wire. After that, we let atoms expand in a ring potential. Finally, we construct an atomic Sagnac interferometer in the ring trap when opposite momentum components are overlapped. The interferometric phase is obtained from the position of each atomic cloud and we expect that the system rotation changes the phase [35].

Figure 4 
               Expansion of the excited atoms in a ring trap. Time step is 800 µs.
Figure 4

Expansion of the excited atoms in a ring trap. Time step is 800 µs.

4 Conclusion

We simulated a method to construct an atomic Sagnac interferometer in a ring trap. In order to add a bi-directional momentum, we suggest using the excited state. First, we load condensed atoms in a dimple-ring trap. After loading, we shake the trap to excite the atoms by changing the center wire RF field. The shaking function is obtained by the optimal control algorithm. The advantage of the optimal algorithm is that we are able to shake the atoms to find a way and just use it experimentally even though we do not know the physics of the result. After excitation, atoms are adiabatically transferred to a ring trap, and they expand in the ring trap similar to an optical Sagnac interferometer. Our main idea is to apply an optimal algorithm to our trapping system for constructing an atomic Sagnac interferometer. This does not require other experimental devices such as lasers for Bragg diffractions or Raman pulses and perturbing tapping potentials during interferometric integration time. We expect that this type of interferometer measures system rotation.

Acknowledgments

This work was supported by the Rare Isotope Science Project of the Institute for Basic Science funded by the Ministry of Science and ICT and NRF of Korea (2013M7A1A1075764) and the National Research Foundation of Korea under NRF-2017R1A2B4008175.

References

[1] Kim SJ, Yu H, Gang ST, Anderson DZ, Kim JB. Controllable asymmetric double well and ring potential on an atom chip. Phys Rev A. 2016;93:033612.10.1103/PhysRevA.93.033612Search in Google Scholar

[2] Geiger R, Ménoret V, Stern G, Zahzam N, Cheinet P, Battelier B, et al. Detecting inertial effects with airborne matter-wave interferometry. Nat Commun. 2011;2:474.10.1038/ncomms1479Search in Google Scholar PubMed PubMed Central

[3] Wu X, Zi F, Dudley J, Bilotta RJ, Canoza P, Müller H. Multiaxis atom interferometry with a single-diode laser and a pyramidal magneto-optical trap. Optica. 2017;4:1545.10.1364/OPTICA.4.001545Search in Google Scholar

[4] Cheiney P, Fouché L, Templier S, Napolitano F, Battelier B, Bouyer P, et al. Navigation-compatible hybrid quantum accelerometer using a Kalman filter. Phys Rev Appl. 2018;10:034030.10.1103/PhysRevApplied.10.034030Search in Google Scholar

[5] Garrido Alzar CL. Compact chip-scale guided cold atom gyrometers for inertial navigation: Enableing technologies and design study. AVS Quantum Sci. 2019;1:014702.10.1116/1.5120348Search in Google Scholar

[6] Rosi G, Sorrentino F, Cacciapuoti L, Prevedelli M, Tino GM. Precision measurement of the Newtonian gravitational constant using cold atoms. Nature. 2014;510:518.10.1038/nature13433Search in Google Scholar PubMed

[7] Parker RH, Yu C, Zhong W, Estey B, Müller H. Measurement of the fine-structure constant as a test of the Standard Model. Science. 2018;360:6385.10.1126/science.aap7706Search in Google Scholar PubMed

[8] Aguilera DN, Ahlers H, Battelier B, Bawamia A, Bertoldi A, Bondarescu R, et al. STE-QUEST—test of the universality of free fall using cold atom interferometry. Classical Quantum Gravity. 2014;31:115010.10.1088/0264-9381/31/11/115010Search in Google Scholar

[9] Canuel B, Bertoldi A, Amand L, Borgo EP, Chantrait T, Danquigny C, et al. Exploring gravity with the MIGA large scale atom interferometer. Sci Rep. 2018;8:14064.10.1038/s41598-018-32165-zSearch in Google Scholar PubMed PubMed Central

[10] Hamilton P, Jaffe M, Haslinger P, Simmons Q, Müller H, Khoury J. Atom-interferometry constraints on dark energy. Science. 2015;349:849.10.1126/science.aaa8883Search in Google Scholar PubMed

[11] Burrage C, Copeland EJ, Hinds EA. Probing dark energy with atom interferometry. J Cosmol Astropart Phys. 2015;3:042.10.1088/1475-7516/2015/03/042Search in Google Scholar

[12] Shin Y, Sanner C, Jo G-B, Pasquini TA, Saba M, Ketterle W, et al. Interference of Bose–Einstein condensates split with an atom chip. Phys Rev A. 2005;72:021604(R).10.1103/PhysRevA.72.021604Search in Google Scholar

[13] Sewell RJ, Dingjan J, Baumgartner F, Llorente-Garcia I, Eriksson S, Hinds EA, et al. Atom chip for BEC interferometry. J Phys B At Mol Opt Phys. 2010;43:051003.10.1088/0953-4075/43/5/051003Search in Google Scholar

[14] Gupta S, Murch KW, Moore KL, Purdy TP, Stamper-Kurn DM. Bose–Einstein condensation in a circular waveguide. Phys Rev Lett. 2005;95:143201.10.1103/PhysRevLett.95.143201Search in Google Scholar PubMed

[15] Ryu C, Blinova AA, Blackburn PW, Boshier MG. Experimental realization of Josephson junctions for an atom SQUID. Phys Rev Lett. 2013;111:205301.10.1103/PhysRevLett.111.205301Search in Google Scholar PubMed

[16] Bell TA, Glidden JAP, Humbert L, Bromley MWJ, Haine SA, Davis MJ, et al. Bose–Einstein condensation in large time-averaged optical ring potentials. New J Phys. 2016;18:035003.10.1088/1367-2630/18/3/035003Search in Google Scholar

[17] Turpin A, Polo J, Loiko YV, Küber J, Schmaltz F, Kalkandjiev TK, et al. Blue-detuned optical ring trap for Bose–Einstein condensates based on conical refraction. Opt Express. 2015;23:1638.10.1364/OE.23.001638Search in Google Scholar PubMed

[18] Roy A, Angom D. Finite temperature expansion dynamics of Bose–Einstein condensates in ring traps. Phys Lett A. 2017;5:55.10.1016/j.physleta.2017.05.055Search in Google Scholar

[19] Weidner CA, Anderson DZ. Experimental demonstration of Shaken-lattice interferometry. Phys Rev Lett. 2018;120:263201.10.1103/PhysRevLett.120.263201Search in Google Scholar PubMed

[20] Weidner CA, Yu H, Kosloff R, Anderson DZ. Atom interferometry using a shaken optical lattice. Phys Rev A. 2017;95:043624.10.1103/PhysRevA.95.043624Search in Google Scholar

[21] Choi DI, Niu Q. Bose–Einstein condensates in an optical lattice. Phys Rev Lett. 1999;82:2022.10.1103/PhysRevLett.82.2022Search in Google Scholar

[22] Morsch O, Oberthaler M. Dynamics of Bose–Einstein condensates in optical lattices. Rev Mod Phys. 2006;78:179.10.1103/RevModPhys.78.179Search in Google Scholar

[23] Böhi P, Riedel MF, Hoffrogge J, Reichel J, Hänsch T, Treutlein P. Coherent manipulation of Bose–Einstein condensates with state-dependent microwave potentials on an atom chip. Nat Phys. 2009;5:592.10.1038/nphys1329Search in Google Scholar

[24] Berrada T, van Frank S, Bücker R, Schumm T, Schaff J-F, Schmiedmayer J. Integrated Mach–Zehnder interferometer for Bose–Einstein condensates. Nat Commun. 2013;4:2077.10.1038/ncomms3077Search in Google Scholar PubMed

[25] McDonald GD, Keal H, Altin PA, Debs JE, Bennetts S, Kuhn CCN, et al. Optically guided linear Mach–Zehnder atom interferometer. Phys Rev A. 2013;87:013632.10.1103/PhysRevA.87.013632Search in Google Scholar

[26] Burke JHT, Deissler B, Hughes KJ, Sackett CA. Confinement effects in a guided-wave atom interferometer with millimeter-scale arm separation. Phys Rev A. 2008;78:023619.10.1103/PhysRevA.78.023619Search in Google Scholar

[27] Wang YJ, Anderson DZ, Bright VM, Cornell EA, Diot Q, et al. Atom Michelson interferometer on a chip using a Bose–Einstein condensate. Phys Rev Lett. 2005;94:090405.10.1103/PhysRevLett.94.090405Search in Google Scholar PubMed

[28] Bücker R, Berrada T, van Frank S, Schaff JF, Schumm T, Schmiedmayer J, et al. Vibrational state inversion of a Bose–Einstein condensate: optimal control and state tomography. J Phys B At Mol Opt Phys. 2013;46:104012.10.1088/0953-4075/46/10/104012Search in Google Scholar

[29] Bücker R, Grond J, Manz S, Berrada T, Betz T, Koller C, et al. Twin-atom beams. Nat Phys. 2011;7:608.10.1038/nphys1992Search in Google Scholar

[30] Hohenester U. OCTBEC—A Matlab toolbox for optimal quantum control of Bose–Einstein condensates. Comput Phys Commun. 2014;185:194.10.1016/j.cpc.2013.09.016Search in Google Scholar

[31] Rohringer W, Bücker R, Manz S, Betz T, Koller CH, Göbel M, et al. Stochastic optimization of a cold atom experiment using a genetic algorithm. Appl Phys Lett. 2008;93:264101.10.1063/1.3058756Search in Google Scholar

[32] Palao JP, Kosloff R. Optimal control theory for unitary transformations. Phys Rev A. 2003;68:062308.10.1103/PhysRevA.68.062308Search in Google Scholar

[33] Kim SJ, Yu H, Gang ST, Kim JB. Matter-wave beam splitter on an atom chip for a portable atom interferometer. Appl Phys B. 2017;123:154.10.1007/s00340-017-6719-6Search in Google Scholar

[34] Lesanovsky I, Schumm T, Hofferberth S, Andersson LM, Krüger P, Schmiedmayer J. Adiabatic radio-frequency potentials for the coherent manipulation of matter waves. Phys Rev A. 2006;73:033619.10.1103/PhysRevA.73.033619Search in Google Scholar

[35] Li-Juan C, Shu-Juan L, Bao-Long L. The interference effect of a Bose–Einstein condensate in a ring-shaped trap. Chin Phys Lett. 2012;29:050305.10.1088/0256-307X/29/5/050305Search in Google Scholar

[36] Akgül A. Reproducing kernel method for fractional derivative with non-local and non-singular kernel. 10.1007/978-3-030-11662-0_1.Search in Google Scholar

[37] Akgül EK. Solutions of the linear and nonlinear differential equations within the generalized fractional derivatives. Chaos. 2019;29:023108.10.1063/1.5084035Search in Google Scholar PubMed

[38] Akgül A. A novel method for a fractional derivative with non-local and non-singular kernel. Chaos Solit Fract. 2018;114:478.10.1016/j.chaos.2018.07.032Search in Google Scholar

[39] Baleanu D, Mousalou A, Rezapour S. A new method for investigating approximate solutions of some fractional integro-differential equations involving the Caputo–Fabrizio derivative. Adv Differ Equ. 2017;2017:51.10.1186/s13662-017-1088-3Search in Google Scholar

[40] Akgül A, Akgül EK. A novel method for solutions of fourth-order fractional boundary value problems. Fract Fraction. 2019;3:33.10.3390/fractalfract3020033Search in Google Scholar

[41] Boutarfa B, Akgül A, Inc M. New approach for the Fornberg–Whitham type equations. J Comput Appl Math. 2017;312:13.10.1016/j.cam.2015.09.016Search in Google Scholar

[42] Akgül A. On the solution of higher order difference equation. Math Meth Appl Sci. 2017;40:6165.10.1002/mma.3870Search in Google Scholar

[43] Akgül A. Reproducing kernel Hilbert space method based on reproducing kernel functions for investigating boundary layer flow of a Powell–Eyring non-Newtonian fluid. J Taibah Univ Sci. 2019;13:858.10.1080/16583655.2019.1651988Search in Google Scholar

[44] Akgül A, Bonyah E. Reproducing kernel Hilbert space method for the solutions of generalized Kuramoto–Sivashinsky equation. J Taibah Univ Sci. 2019;13:661.10.1080/16583655.2019.1618547Search in Google Scholar

[45] Akgül A. A novel method for the solution of Blasius equation in semi-infinite domains. Int J Optim Control Theor Appl. 2017;7:225.10.11121/ijocta.01.2017.00363Search in Google Scholar

Received: 2019-10-18
Revised: 2020-04-19
Accepted: 2020-06-30
Published Online: 2020-07-28

© 2020 Hoon Yu et al., published by De Gruyter

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

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/phys-2020-0171/html
Scroll to top button