Next Article in Journal
A Highly Accurate Forest Fire Prediction Model Based on an Improved Dynamic Convolutional Neural Network
Next Article in Special Issue
Nonlinear Impact Force Reduction of Layered Polymers with the Damage-Trap Interface
Previous Article in Journal
Error-Bounded Learned Scientific Data Compression with Preservation of Derived Quantities
Previous Article in Special Issue
Vibration and Wave Analyses in the Functionally Graded Graphene-Reinforced Composite Plates Based on the First-Order Shear Deformation Plate Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Approximate Solution of Nonlinear Flexure of a Cantilever Beam with the Galerkin Method

1
Piezoelectric Device Laboratory, School of Mechanical Engineering and Mechanics, Ningbo University, 818 Fenghua Road, Ningbo 315211, China
2
Institute of Applied Mechanics, Ningbo Polytechnic, 1069 Xinda Road, Beilun District, Ningbo 315800, China
3
School of Mathematics and Statistics, Ningbo University, 818 Fenghua Road, Ningbo 315211, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(13), 6720; https://doi.org/10.3390/app12136720
Submission received: 24 May 2022 / Revised: 28 June 2022 / Accepted: 29 June 2022 / Published: 2 July 2022

Abstract

:
For the optimal design and accurate prediction of structural behavior, the nonlinear analysis of large deformation of elastic beams has broad applications in various engineering fields. In this study, the nonlinear equation of flexure of an elastic beam, also known as an elastica, was solved by the Galerkin method for a highly accurate solution. The numerical results showed that the third-order solution of the rotation angle at the free end of the beam is more accurate and efficient in comparison with results of other approximate methods, and is perfectly consistent with the exact solution in elliptic functions. A general procedure with the Galerkin method is demonstrated for efficient solutions of nonlinear differential equations with the potential for adoption and implementation in more applications.

1. Introduction

Differential equations are widely used for the description of scientific and engineering phenomena with frequent appearances of nonlinear types [1,2,3,4]. Consequently, differential equations and solution methods have a major role in modeling transmission dynamics and rendering control strategies for communicable diseases such as tuberculosis and COVID-19 [5,6]. There have been tremendous efforts to solve differential equations with generalized and ad hoc methods, but there are always problems that cannot be effectively solved for many reasons [7,8,9]. Some known solutions of some specialized differential equations, such as those related to Bessel functions and hypergeometric functions, can be effectively used, but there are still differential equations that cannot be solved with existing methods and known functions [10,11].
With classical methods of solutions failing in these contexts, in recent decades, some special methods with gradual sophistication have been developed, such as the multiscale method [10], homotopy analysis method [12,13,14,15], homotopy perturbation method [16], and others dealing with emerging differential equations in many fields [17,18]. Generally, most of these methods are asymptotic in nature, and can find solutions in series form with an improved converging rate and numerical stability over a relatively large interval. In combination with the algorithm development and computing tools for symbolic computation, these methods can powerfully and efficiently solve many practical and classical nonlinear equations. As a result, many techniques have been applied to the latest problems, and the improved solution techniques have been widely promoted and utilized.
Studies on the large deflection of beams mostly focused on different boundary conditions, cross-section characteristics, and loading conditions [19], among others. Formulated with the angle enclosed by the axes of the deformed and undeformed configurations, the nonlinear equation of an elastic beam subjected to concentrated forces or couples resembles the equation of motion of a rigid pendulum according to Kirchhoff’s kinetic analogy [20]. Large deflections of an end-supported beam subjected to a point load are solved using both the elliptic integral method and the shooting-optimization technique [21]. The homotopy analysis method was used to solve this problem with enough accuracy [13]. In recent attempts to find an optimal method for nonlinear vibrations of structures, it was discovered that the Galerkin method and its alternative form, the Rayleigh-Ritz method, can be effective for approximate analysis of nonlinear vibrations in certain circumstances [22,23,24,25]. The Legendre–Galerkin matrix (LGM) method is used to solve the vibration equations of single (SDOF) and multiple degree-of-freedom (MDOF) structural systems [26]. The nonresonance and primary resonance of microbeams with consideration of small-scale effects and nonlinear terms were obtained by the Galerkin and multiple-scale methods [27].
With the objective of finding an effective method and taking the advantage of the widespread availability of powerful symbolic computing tools and resources, it is thought that the Galerkin method can be used for obtaining relatively accurate solutions to nonlinear differential equations with a simple procedure [28,29]. The procedure can be developed with the appropriate trial solutions satisfying essential boundary and initial conditions, making the computing process more efficient for accelerated convergence and improving accuracy [30,31]. Furthermore, if the orthogonality of trial solutions is guaranteed, the calculation will be further optimized. These concepts are already practiced in linear problems with the Galerkin method and its modifications, and further improvements can be made with nonlinear problems as an innovative implementation. The basic ideas outlined here were applied to the nonlinear equation of a thin beam, and the results showed that the analysis can be used, and the results are quite accurate. The demonstration of the use of the Galerkin method for approximate solutions to nonlinear differential equations is novel and enlightening in the promotion of this classical and practical method for new objectives and applications. The robustness of this approach is more apparent when applied to this historical elastica problem, which has been used as a benchmark for various methods and techniques. This application presents an ideal case for employing the Galerkin method to take advantage of its efficiency and simplicity in dealing with increasing nonlinear differential equations from various applications in the absence of analytical or other approximate solutions.

2. Galerkin Method for Solutions of Nonlinear Differential Equations

To implement the Galerkin method to solve nonlinear differential equations, the fundamental concepts and procedures are introduced to provide an essential understanding. The rigorous mathematical procedure and associated assumptions and proofs are not included in this paper, but they can be found in many mathematical textbooks and monographs [10,32]. A general nonlinear differential second-order equation is
N ( x , y , y k , y , y , p ) = 0 ,   x [ a , b ] ,
where y , y ( y ) , p , x , a , and b are the function, derivatives with respect to the spatial variable, nonlinear term with the power of k, parameter, spatial variable, lower bound, and upper bound, respectively. With prescribed boundary and initial conditions, the approximate solution, also generally asymptotic, can be assumed as
y = n = 0 M A n Y n x ,
where the chosen function Y n x satisfies the boundary conditions, and An are coefficients to be determined with the limitations by an integer M. As a general rule of thumb, the trial solutions Y n x can be conveniently chosen from the solutions of a simplified linear equation, so calculations will be simple and efficient. Furthermore, it is important to expect that these solutions are also orthogonal. There is abundant literature on the optimal selections of trial functions, and one must deal with some typical equation types and appropriate boundary conditions.
Applying the Galerkin method, the problem in Equation (1) is transformed to
a b N x , y , y k , y , y , p n = 0 M δ A n Y n x   d x = 0 .
Because the Galerkin method has proven its capability to provide accurate solutions to linear differential equations as a powerful and robust approximate procedure, it is also thought the nonlinear equations can be solved in a similar manner with equal strength. With this hypothesis in mind, the arbitrary variation in independent coefficients in Equation (3) results in a series of nonlinear algebraic equations of unknown amplitudes after integration as
N ¯ n A m ,   A m q , p = 0 ,   n , m = 1 , 2 , , M ,
using the equation and trial functions with parameters p and q. By solving the system of nonlinear algebraic equations for amplitude Am, the approximate solution can be obtained. It is better to have orthogonal functions in the domain of integration for a simple form and least couplings of unknowns in Equation (4). After obtaining solutions of the coefficients, the solution is in Equation (2) for manipulations in relation to the problem.
If a simplified procedure is sought, with a careful selection of trial functions satisfying the boundary conditions, the procedure starts with
y = A 0 Y 0 x ,
Then, the first approximate equation is
a b N x , A 0 Y 0 x , A 0 k Y 0 k x , A 0 Y 0 x , A 0 Y 0 x , p Y 0 x d x = 0 .
The coefficient A0 is to be solved from the above equation. To improve the solution with an additional function, it is further assumed that the updated solution is
y = A 0 Y 0 x + A 1 Y 1 x ,     δ y = δ A 1 Y 1 x .  
Then, the following Equation is for the unknown amplitude A1
a b N x , y , y k , y , y , p Y 1 x d x = 0 ,
by using known A0 from the procedure above in Equation (6). By repeating this procedure with the successive addition of higher-order solutions, other coefficients of higher-order approximations can be successively obtained for an improved solution.
With the two steps for the approximate solution in the above demonstration, the results have to be compared before a conclusion can be drawn on the advantage of the Galerkin method. If a good trial solution is selected from the beginning, the accuracy of solutions will be satisfactory based on the proven trust in the Galerkin method. Further proof and validation can be made with some typical examples. These steps are simple with given equations that are asymptotic in nature. The amplitudes can be improved through an iterative procedure from the nonlinear algebraic equations.

3. Equation of Flexure of an Elastic Beam and Its Solution

With the procedure and algorithm outlined above, a typical nonlinear equation is needed to test the procedure and accuracy. Among the many typical nonlinear differential equations, the large deformation of a beam under a concentrated force at the free end, also known as the elastica, is [14,15,31,32,33]
θ + α cos θ = 0 ,   θ 0 = 0 ,       θ 1 = 0 ,
where θ = θ x is the rotation of the cross-section of beam, and α = P l 2 / E I is a constant of the property of the beam problem, with P ,   l , and E I as the point load at the free tip, beam length, and bending stiffness of the beam, respectively.
This is widely known as the elastica problem, which has been extensively studied [32,33,34]. The elastica of nonprismatic cantilever beams with rectangular cross-sections that are subjected to combined loading was studied by Lee [35]. Wang et al. used the homotopy analysis method (HAM) for this problem with approximate solutions in power series with a large radius of convergence [15]. An elastic beam with relatively large deformation under a concentrated load is a common structural problem in both civil and mechanical engineering [10,11,14,15,32,33,34]. The analytical methods and exact solution to this problem are very important for the design and optimization of these structures. Therefore, a new solution technique with improved results is important for validating the proposed approach with the Galerkin method. In previous studies, the traditional Galerkin method was first used for analysis of wave propagation in finite solids with nonlinear complications [36]. By assuming the spatial variation in displacements in the exact form of the known linear mode shape, Wang et al. used the orthogonal properties of the linear mode shapes to obtain the equations to approximately determine the temporal behavior [37]. Wang further expanded the application scope of the Galerkin method by integrating the weighted equation of motion over the time of one period of vibrations to eliminate the harmonics from the deformation function [31]. This is an extension of Galerkin method that has broad applications in approximate solutions, particularly for solid mechanics problems [29]. This extended Galerkin method can also be utilized for the analysis of nonlinear differential equations of free and forced nonlinear vibrations of structures as a new technique with many advantages [28,29,30,31]. The procedure outlined in this paper for the determination of displacement coefficients from nonlinear algebraic equations is new, but it is a natural extension of the popular Galerkin method meriting wide publicization and promotion.
By examining Equation (9), it is clear that the trial solution with trigonometric functions satisfying the boundary conditions is in the form of
θ x = n = 0 A n sin 2 n + 1 2 π x .
The amplitudes A n are determined with the procedure outlined above. One must substitute Equation (10) into Equation (9) with a standard Galerkin formulation
0 1 θ + α cos θ δ θ d x = 0 ,
where
δ θ = n = 0 δ A n sin 2 n + 1 2 π x ,
with arbitrary variation in amplitudes δ A n .
For the evaluation of the integral, there are several choices of analytical expressions and power series expressions. It is more favorable to use
cos θ = n = 0 1 n θ 2 n 2 n ! = 1 θ 2 2 + θ 4 24 θ 6 720 + .  
Then, the problem will be the evaluation of individual equations without coupling as
0 1 θ + α cos θ sin 2 n + 1 2 π x   d x = 0 ,   n = 1 ,   2 ,   3 ,   , M .
With Equation (4), Equation (6) can be written as
0 1 n = 0 N 2 n + 1 2 π 2 A n sin 2 n + 1 2 π x + α   cos n = 0 N A n sin 2 n + 1 2 π x ] sin 2 n + 1 2 π x   d x = 0 n = 1 , 2 , 3 , ,   M ,
for the coupled amplitudes with the choice of order M . The procedure outlined above is followed to obtain the solutions of amplitudes.
To obtain the amplitude A 0 , the approximate solution is taken as
θ = A 0 sin π 2 x ,
and the integration of Equation (7) yields
575 1575 π 3 A 0 16 α 525 A 0 2 + 35 A 0 4 A 0 6 = 0 .
To find appropriate solution of A 0 α , the curves from Equation (17) are plotted in Figure 1 to help the solution procedure of this equation.
As shown in Figure 1, there are two real solutions of A0 with α from the nonlinear algebraic equation in Equation (17). The amplitude A0 can be given with α, but the standard Maclaurin expansion only offers expression with a relatively small convergent radius. To obtain a power series solution with relatively larger convergent radius, an auxiliary expansion scheme with Chebyshev polynomial can be utilized. The procedure starts with the expression of A 0 α , and the expansion is
A 0 α = n = 0 C n T n α 5 1 , C n = 2 5 π 0 10 A 0 α T n α 5 1 1 α 5 1 2 d α ,  
where T n   is the n th-order Chebyshev polynomial, and C n is the coefficient of expansion. In case an explicit expression of A 0 α is not available, the symbolic expression from mathematical tools such as Mathematica® and MATLAB® can be utilized for the evaluation in Equation (18). Previous studies have shown that Equations (17) and (18) here have good stability and convergence [15,32,33,34]. In this study, from Equations (17) and (18), the explicit expression of A 0 α with approximation is
A 0 = 1.12113113903689 × 10 7 + 0.5160330121688547 α   0.00009479584375745703 α 2 0.045486335180154686 α 3   + 0.000012757993686000196 α 4 + 0.006466636289057949 α 5   + 0.006958737479441875 α 6 0.012229157808245239 α 7   + 0.008664382051570434 α 8 0.0039340381580180665 α 9           + 0.0012851620189843023 α 10 0.00031666548778055986 α 11   + 0.00006014045641459205 α 12   0.000008877160503395703 α 13   + 0.000001017529293530474 α 14   8.973639340531093 × 10 8 α 15   + 5.973823554191147 × 10 9 α 16   2.903563905834224 × 10 10 α 17   + 9.721628341667188 × 10 12 α 18   2.004035813851492 × 10 13 α 19   + 1.916801557948531 × 10 15 α 20
Then, from Equation (10) with the known value of A 0 , the approximate solution is modified to
θ = A 0 sin π 2 x + A 1 sin 3 π 2 x ,  
where A 1 is solved. Following the same procedure as demonstrated for A 0 , the nonlinear algebraic equation for amplitude A 1 is
2 α 3 π 9 π 2 A 1 8 + 1 654729075 π 2 α [ 230945 A 0 2 189 27 A 0 2 + A 0 4 25194 A 0 13365 1430 A 0 2 + 59 A 0 4 A 1 + 14535 5005 1118 A 0 2 + 161 A 0 4 A 1 2 6460 A 0 6561 + 695 A 0 2 A 1 3 + 285 ( 17017 + 3947 A 0 2 ) A 1 4 1771470 A 0 A 1 5 138567 A 1 6 ] = 0 .
Then, from Equation (21) with an approximate procedure for an explicit expression, the convergent solution is
A 1 = 2 α 3 π + 2 α A 0 2 189 27 A 0 2 + A 0 4 2835 π 9 π 2 8 + 4 α A 0 13365 1430 A 0 2 + 59 A 0 4 51975 π   ,
noting the next approximate solution is
θ = A 0 sin π 2 x + A 1 sin 3 π 2 x + A 2 sin 5 π 2 x ,  
and the algebraic equation for A 2 is
104187267600 α A 0 6 + 403446387600 α A 1 6 3205762080 α A 0 5 ( 535 A 1 751 A 2 ) + 193391815584 α A 1 5 A 2 + 2042887600 α A 1 4 ( 6561 + 2437 A 2 2 ) + 93000000 A 1 3 A 2 ( 124729 + 11855 A 2 2 ) + 24812400 α A 0 4 ( 46189 + 132753 A 1 2 206834 A 1 A 2 + 17313 A 2 2 ) + 1708100000 α A 1 A 2 116127 14850 A 2 2 + 625 A 2 4 + 35377680 α A 1 2 5009445 1890690 A 2 2 + 131671 A 2 4 + 8595449577 ( 25200 α + 196875 π 3 A 2     + 8400 α A 2 2 560 α A 2 4 + 16 α A 2 6 ) 4315200 α A 0 3 [ 1099929 A 1 3 3598787 A 1 2 A 2     625 A 2 22287 + 2785 A 2 2 + 23 A 1 365313 + 137021 A 2 2 ]     + 71920 α A 0 2 [ 50284785 A 1 4 115877220 A 1 3 A 2         + 90 A 1 2 6589523 + 2468591 A 2 2 + A 1 512155260 A 2 64424868 A 2 3   209 686205 254610 A 2 2 + 17599 A 2 4 ]       160 α A 0 [ 5067002235 A 1 5 55775187135 A 1 4 A 2   41354 A 1 2 A 2 20529405 + 2545379 A 2 2       + 16182 A 1 3 7763305 + 2927733 A 2 2 9530625 A 2 358701 44950 A 2 2 + 1875 A 2 4     + 3875 A 1 277272567 103551750 A 2 2 + 7178125 A 2 4 ] = 0 ,        
with the explicit expression of A 2 as
A 2 = ( 216605329340400 α 10314539492400 α A 0 2   1146059943600 α A 0 4 + 104187267600 α A 0 6   171908991540000 α A 0 A 1 + 36257169124800 α A 0 3 A 1   1715082712800 α A 0 5 A 1 + 177222542187600 α A 1 2   + 42652664474400 α A 0 2 A 1 2 + 3293920537200 α A 0 4 A 1 2   + 20100128241600 α A 0 A 1 3 4746413620800 α A 0 3 A 1 3   13403385543600 α A 1 4 + 3616481737200 α A 0 2 A 1 4   810720357600 α A 0 A 1 5   + 403446387600 α A 1 6 ) / ( 1692229135471875 π 3   + 546983154900000 α A 0 60108039000000 α A 0 3   + 2407527322080 α A 0 5 + 198356528700000 α A 1   + 36834206299200 α A 0 2 A 1 5132047941600 α A 0 4 A 1   135835682299200 α A 0 A 1 2 + 15529485662400 α A 0 3 A 1 2   11599797000000 α A 1 3 8333889662400 α A 0 2 A 1 3   + 8924029941600 α A 0 A 1 4 + 193391815584 α A 1 5 ) .
Now with all three coefficients known, the approximate solution is given as in Equation (23), which is quite different from the power series solution produced by other methods [15].
To verify this solution and make comparisons with other known results, the rotation angle at the free end is obtained by setting x = 1 in Equation (24) as
θ B = A 0 A 1 + A 2 .
For comparison, the exact solution from Equation (9) is [15]
α = K μ 2 F ϕ , μ ,     μ = 1 + sin θ B 2       ,     ϕ = arcsin 1 2 μ .
Here, K μ is the complete elliptic function of the first kind, and F ϕ , μ is the elliptic function of the second kind. Then, the relationship between α and θ B can be analytically obtained from these equations.
The optimal approximate solution by the HAM is given as [15]
θ B = α 2 f α g α ,  
with
f α = 1 + 3.98575 × 10 2 α 2 5.41174 × 10 2 α 4   + 5.72575 × 10 3 α 6 + 3.79533 × 10 4 α 8   1.51429 × 10 11 α 14 2.29142 × 10 15 α 16     8.87896 × 10 6 α 10 +   2.63041 × 10 8 α 12     3.45006 × 10 21 α 18     7.00678 × 10 28 α 20 , g α = 1 +   0.131524 α 2     5.99231   × 10 2 α 4   +   2.34466 × 10 3 α 6     +   9.90299 × 10 4 α 8     1.37001 × 10 5 α 10 +   3.44172 × 10 8 α 12     1.45098 × 10 11 α 14   2.26721 × 10 15 α 16     3.50731 × 10 21 α 18     7.00678 × 10 28 α 20 .
To visualize the solutions and make comparisons, different approximate solutions and the exact solution are plotted in Figure 2 with parameter α .
It is clear from Figure 2 that the approximations are very good with just two terms. With three terms, it is almost the exact solution in the range of parameter α , as shown in the plot. The excellent accuracy is a little surprising given the fact the procedure and technique are relatively simple, as shown in this paper. As for a comparison with the rigorous and sophisticated HAM [15], it also shows that for this particular example, this method provides a relatively accurate approximation with a much simpler procedure. We can plot the rotation angle at the free end θ B   vs. α with solutions of different orders in Figure 3.
From Figure 3, it can be found that there is still a noticeable gap between the first-order and the exact solutions. On the other hand, the second-order solution gradually approaches to the exact solution. The third-order solution is almost exact in a large interval in comparison with the closed-form solution with elliptic functions. It is evident now that the Galerkin method is extremely efficient in obtaining an accurate solution to the nonlinear differential equation of the elastica. The excellent choice of the trigonometric sine function as the basis of trial solutions makes the procedure ideal, and the accuracy is also surprisingly good with just the third-order solution. It is evident that the Galerkin method can be equally powerful for nonlinear problems arising from applications in many fields. The elastica problem has a long and rich history, but it is still a hot topic related to the flexure and buckling of beams and rods; accurate results are essentially obtained from the elliptic functions [38,39]. The procedure and solution from this study show that with the Galerkin method, a new technique to solve similar nonlinear differential equations is available, and it can be much simpler to obtain reasonably accurate solutions. This was the objective of this study: to demonstrate the usefulness and efficiency of the Galerkin method with nonlinear differential equations. It is a neglected technique worth being promoted, as demonstrated by the popular example of the elastica problem.
The choice of trial functions is of great importance in the simplification of calculation and determination of accuracy, but the choice should be made from experience in dealing with approximate solutions for the Galerkin and Rayleigh–Ritz methods, which are equivalent in most cases. Furthermore, the solution of coupled nonlinear equations of amplitudes can be hard to acquire, but the utilization of symbolic computing tools such as Mathematica® and MATLAB® can provide much better solutions with explicit expressions that can only be given and meaningful for the lower orders. For practical problems such as the elastica problem in this study, the combination of ideal trial functions, optimal procedure, iteration, and symbolic calculation produces sufficiently accurate results to meet application requirements. If a systematic procedure with symbolic computing is developed from the approximate solutions of nonlinear algebraic equations, the efficiency and accuracy can be greatly increased to achieve better solutions of nonlinear differential equations. Further studies of and improvement in the Galerkin method will add powerful tools to the collection of techniques for nonlinear differential equations. The power and capabilities of symbolic computation can make the Galerkin method favorable for some nonlinear differential problems with less numerical computation, as shown with the elastica problem in this study.

4. Conclusions

By utilizing the Galerkin method with a proper choice of trial functions of the approximate solution, a novel procedure to obtain asymptotic solutions of nonlinear differential equations was presented and validated. The method and procedure were verified by solving a beam flexure problem with an enhanced solution in comparison with the solution from the homotopy analysis method, with satisfaction. The procedure is straightforward for mathematical implementation because the calculation is simple, but the optimal choice of the trial function is critical due to the simplification of calculation with the orthogonality through integration. The further refinement of coefficients can be achieved with iterative procedure, which is also easy to implement in the calculation procedure. The accurate results in such a simple procedure demonstrated that similar nonlinear problems can be solved with this new and efficient method, achieving relatively accurate solutions with a limited number of terms. While the nonlinear problems are receiving attention for important applications in many natural phenomena and engineering problems, innovative methods for solving nonlinear differential equations including this Galerkin method should also be examined and refined to enrich the toolbox for addressing greater challenges posed by analytical requirements. The use of the traditional Galerkin method for approximate solutions to nonlinear differential equations in this study is only an example to show the advantage of a simple computational procedure for a seemingly difficult problem with exact solutions in elliptic functions and a large number of numerical computation processes.

Author Contributions

Conception, revision, and submission of the paper, J.W.; mathematical formulation, procedure, derivation, calculation, and check, J.Z. and R.W., with help from L.W.; reviews and discussions of the study and drafting were also completed with the participation of T.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (Grants 11672142, 11772163, and 12172183) with additional support from the Technology Innovation 2025 Program (Grant 2019B10122) of the Municipality of Ningbo, Zhejiang Province, China.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data from this study are available from the authors and their website.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Nayfeh, A.H.; Mook, D.T. Nonlinear Oscillations; Wiley: New York, NY, USA, 1979. [Google Scholar]
  2. Khan, H.; Tunc, C.; Khan, A. Green function’s properties and existence theorems for nonlinear singular-delay-fractional differential equations. Discret. Contin. Dyn. Syst. 2020, 13, 2475–2487. [Google Scholar] [CrossRef] [Green Version]
  3. Khan, H.; Chen, W.; Khan, A.; Khan, T.S.; Al-Madlal, Q.M. Hyers-Ulam stability and existence criteria for coupled fractional differential equations involving p-Laplacian operator. Adv. Differ. Equ. 2018, 2018, 455. [Google Scholar] [CrossRef] [Green Version]
  4. Khan, H.; Khan, A.; Chen, W.; Shah, K. Stability analysis and a numerical scheme for fractional Klein-Gordon equations. Math. Method Appl. Sci. 2019, 42, 723–732. [Google Scholar] [CrossRef]
  5. Ullah, I.; Ahmad, S.; Al-Mdallal, Q.; Khan, Z. Stability analysis of a dynamical model of tuberculosis with incomplete treatment. Adv. Differ. Equ. 2020, 2020, 499. [Google Scholar] [CrossRef]
  6. Sher, M.; Shah, K.; Khan, Z.A.; Khan, H.; Khan, A. Computational and theoretical modeling of the transmission dynamics of novel COVID-19 under Mittag-Leffler power law. Alex. Eng. J. 2020, 59, 3133–3147. [Google Scholar] [CrossRef]
  7. Jin, G.; Ye, T.; Su, Z. Structural Vibration; Science Press and Springer: Beijing, China, 2015. [Google Scholar]
  8. Nayfeh, A.H.; Balachandran, B. Applied Nonlinear Dynamics: Analytical, Computational, and Experimental Methods; Wiley: New York, NY, USA, 2008. [Google Scholar]
  9. Chen, H.B.; Li, Y. Bifurcation and stability of periodic solutions of Duffing equations. Nonlinearity 2008, 21, 2485. [Google Scholar] [CrossRef]
  10. Hu, H.Y. Applied Nonlinear Mechanics; Aviation Industry Press: Beijing, China, 2000. (In Chinese) [Google Scholar]
  11. Bayat, M.; Pakar, I.; Domairry, G. Recent developments of some asymptotic methods and their applications for nonlinear vibration equations in engineering problems: A review. Lat. Am. J. Solids Struct. 2012, 9, 1–93. [Google Scholar] [CrossRef] [Green Version]
  12. Banerjee, J.R.; Su, H.; Jackson, D.R. Free vibration of rotating tapered beams using the dynamic stiffness method. J. Sound Vib. 2006, 298, 1034–1054. [Google Scholar] [CrossRef]
  13. Liao, S.J. Beyond Perturbation: Introduction to the Homotopy Analysis Method; Chapman & Hall/CRC Press: Boca Raton, FL, USA, 2003. [Google Scholar]
  14. Li, Y.S.; Li, X.Y.; Xie, C.; Huo, S.H. Explicit solution to large deformation of cantilever beam by improved homotopy analysis method II: Vertical and horizontal displacements. Appl. Sci. 2022, 12, 2513. [Google Scholar] [CrossRef]
  15. Wang, J.; Chen, J.K.; Liao, S.J. An explicit solution of the large deformation of a cantilever beam under point load at the free tip. J. Comput. Appl. Math. 2008, 212, 320–330. [Google Scholar] [CrossRef] [Green Version]
  16. He, J.H. Homotopy perturbation technique. Comput. Methods Appl. Mech. Eng. 1999, 178, 257–262. [Google Scholar] [CrossRef]
  17. Amore, P.; Aranda, A. Improved Lindstedt-Poincaré method for the solution of nonlinear problem. J. Sound Vib. 2005, 283, 1115–1136. [Google Scholar] [CrossRef] [Green Version]
  18. Chowdhury, M.S.H.; Hosen, M.A.; Ahmad, K.; Ali, M.Y.; Ismail, A.F. High-order approximate solutions of strongly nonlinear cubic-quantic Duffing oscillator based on the harmonic balance method. Results Phys. 2017, 7, 3962–3967. [Google Scholar] [CrossRef]
  19. Wang, C.M.; Lam, K.Y.; He, X.Q.; Chucheepsakul, S. Large deflections of an end supported beam subjected to a point load. Int. J. Non-Linear Mech. 1997, 32, 63–72. [Google Scholar] [CrossRef]
  20. Love, A.E.H. A Treatise on the Mathematical Theory of Elasticity; Dover Publications Inc.: New York, NY, USA, 1944. [Google Scholar]
  21. Humer, A. Elliptic integral solution of the extensible elastica with a variable length under a concentrated force. Acta Mech. 2011, 222, 209–223. [Google Scholar] [CrossRef]
  22. Oliveri, V.; Milazzo, A. A Rayleigh-Ritz approach for postbuckling analysis of variable angle tow composite stiffened panels. Comput. Struct. 2018, 196, 263–276. [Google Scholar] [CrossRef]
  23. Zhao, Q.; Ye, T.Q. Hybrid changeable basis Galerkin technique for nonlinear analysis of structures. Appl. Math. Mech. 1995, 16, 667–674. [Google Scholar]
  24. Zhang, G.C.; Ding, H.; Chen, L.Q.; Yang, S.P. Galerkin method for steady-state response of nonlinear forced vibration of axially moving beams at supercritical speeds. J. Sound Vib. 2012, 331, 1612–1623. [Google Scholar] [CrossRef]
  25. Ghafari, E.; Rezaeepazhand, J. Two-dimensional cross-sectional analysis of composite beams using Rayleigh-Ritz-based dimensional reduction method. Compos. Struct. 2018, 184, 872–882. [Google Scholar] [CrossRef]
  26. Momeni, M.; Beni, M.R.; Bedon, C.; Najafgholipour, M.A.; Dehghan, S.M.; Sharifi, B.J.; Hadianfard, M.A. Dynamic response analysis of structures using Legendre-Galerkin matrix method. Appl. Sci. 2021, 11, 9307. [Google Scholar] [CrossRef]
  27. Huang, K.; Li, T.P.; Xu, W.; Cao, L. Effects of nonlinear damping on vibrations of microbeam. Appl. Sci. 2022, 12, 3206. [Google Scholar] [CrossRef]
  28. Wang, J. The extended Rayleigh-Ritz method for an analysis of nonlinear vibrations. Mech. Adv. Mater. Struct. 2021. [CrossRef]
  29. Shi, B.Y.; Yang, J.; Wang, J. Forced vibration analysis of multi-degree-of-freedom nonlinear systems with an extended Galerkin method. Mech. Adv. Mater. Struct. 2022. [CrossRef]
  30. Jing, H.M.; Gong, X.L.; Wang, J.; Wu, R.X.; Huang, B. An analysis of nonlinear beam vibrations with the extended Rayleigh-Ritz method. J. Comput. Appl. Mech. 2022, 8, 1299–1306. [Google Scholar] [CrossRef]
  31. Wang, J.; Wu, R.X. The extended Galerkin method for approximate solutions of nonlinear vibration equations. Appl. Sci. 2022, 12, 2979. [Google Scholar] [CrossRef]
  32. Chen, S.H. Quantitative Analytical Methods of Strong Nonlinear Vibrations; Science Press: Beijing, China, 2009. (In Chinese) [Google Scholar]
  33. Gere, J.M.; Timoshenko, S.P. Mechanics of Materials; PWS Publishing Company: Boston, MA, USA, 1997. [Google Scholar]
  34. Pagani, A.; Carrera, E. Unified formulation of geometrically nonlinear refined beam theories. Mech. Adv. Mater. Struct. 2018, 25, 15–31. [Google Scholar] [CrossRef] [Green Version]
  35. Lee, J.K.; Lee, B.K. Elastica of non-prismatic and nonlinear elastic cantilever beams under combined loading. Appl. Sci. 2019, 9, 877. [Google Scholar] [CrossRef] [Green Version]
  36. Abd-alla, A.; Maugin, G.A. Nonlinear phenomena in magnetostrictive elastic resonators. Int. J. Eng. Sci. 1989, 27, 1613–1619. [Google Scholar] [CrossRef]
  37. Wu, R.X.; Wang, J.; Du, J.K.; Huang, D.J.; Hu, Y.T. The nonlinear thickness-shear vibrations of quartz crystal plates under an electric field. Int. J. Non-linear Mech. 2014, 61, 32–38. [Google Scholar] [CrossRef]
  38. Barbieri, E. Analytical solution of the cantilevered elastica subjected to a normal uniformly distributed follower load. Int. J. Solids Struct. 2020, 202, 486–494. [Google Scholar] [CrossRef]
  39. Misseroni, D.; Barbieri, E.; Pugno, N.M. Extreme deformations of the cantilever Euler elastica under transverse aerodynamic load. Extreme Mech. Lett. 2021, 42, 101110. [Google Scholar] [CrossRef]
Figure 1. The solutions of A0 vs. parameter α.
Figure 1. The solutions of A0 vs. parameter α.
Applsci 12 06720 g001
Figure 2. The rotation angle at the free end θB vs. α with different solutions.
Figure 2. The rotation angle at the free end θB vs. α with different solutions.
Applsci 12 06720 g002
Figure 3. The rotation angle at the free end θB vs. α with solutions of different orders.
Figure 3. The rotation angle at the free end θB vs. α with solutions of different orders.
Applsci 12 06720 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.; Wu, R.; Wang, J.; Ma, T.; Wang, L. The Approximate Solution of Nonlinear Flexure of a Cantilever Beam with the Galerkin Method. Appl. Sci. 2022, 12, 6720. https://doi.org/10.3390/app12136720

AMA Style

Zhang J, Wu R, Wang J, Ma T, Wang L. The Approximate Solution of Nonlinear Flexure of a Cantilever Beam with the Galerkin Method. Applied Sciences. 2022; 12(13):6720. https://doi.org/10.3390/app12136720

Chicago/Turabian Style

Zhang, Jun, Rongxing Wu, Ji Wang, Tingfeng Ma, and Lihong Wang. 2022. "The Approximate Solution of Nonlinear Flexure of a Cantilever Beam with the Galerkin Method" Applied Sciences 12, no. 13: 6720. https://doi.org/10.3390/app12136720

APA Style

Zhang, J., Wu, R., Wang, J., Ma, T., & Wang, L. (2022). The Approximate Solution of Nonlinear Flexure of a Cantilever Beam with the Galerkin Method. Applied Sciences, 12(13), 6720. https://doi.org/10.3390/app12136720

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop