Skip to main content
Log in

Optimal configurations of circular bars under free torsional and longitudinal vibration based on Pontryagin’s maximum principle

  • Published:
Meccanica Aims and scope Submit manuscript

Abstract

Optimal configurations of circular bars under free torsional and longitudinal vibration are investigated using the Pontryagin’s maximum principle (PMP). The necessary optimality condition of PMP is analysed by considering the cross-sectional area of bars as control variable and by using Maier objective functional to control the final state of the objective functional. Optimal configurations associated to different orders of eigen frequencies, eigen vectors, and specific boundary conditions are demonstrated. Their relations are also explained qualitatively. Numerical results show the equivalent about optimal configurations and eigen frequencies under specific boundary conditions.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Abbreviations

A :

Cross-sectional area of the bar

c ωi :

Coefficient in the objective function

c W :

Coefficient in the objective function

d e :

Diameter of element e

d max :

Maximum diameter

d min :

Minimum diameter

E :

Elastic modulus

F :

Objective function

G :

Shear modulus

H :

Hamiltonian function

J p :

Polar moment of inertia of the cross-sectional area of the circular bar

k ωW :

Non-negative weight, k ωW  ∈ [0, 1)

L :

Length of the bar

L e :

Length of element e

M :

Internal moment

m :

A positive integer

N :

Internal force

N H :

Adjoint state variable (N H  = −p u )

n :

Number of nodes

n L :

A positive integer

p N :

Adjoint variable of N

p W :

Adjoint variable of W

p u :

Adjoint variable of u

p ω :

Adjoint variable of ω

W :

Total weight of the bar

W 0 :

The total weight of the initial bar (before optimization)

[k] e :

Stiffness matrix of the eth element

[m] e :

Mass matrix of the eth element

[K]:

Global stiffness matrix

[M]:

Global mass matrix

u :

Displacement

u H :

Adjoint state variable (u H  = p N )

ρ :

Mass density

ω i :

The ith natural frequency

ω 0i :

The ith natural frequency of the initial bar (before optimization)

References

  1. Hagedorn P, DasGupta A (2007) Vibrations and waves in continuous mechanical systems. Wiley, England

    Book  MATH  Google Scholar 

  2. Rao SS (2007) Vibration of continuous systems. Wiley, USA

    Google Scholar 

  3. Ma H, Lu Y, Wu Z et al (2015) A new dynamic model of rotor–blade systems. J Sound Vib. doi:10.1016/j.jsv.2015.07.036i

    Google Scholar 

  4. Sihler C (2006) A novel torsional exciter for modal vibration testing of large rotating machinery. Mech Syst Signal Process 20:1725–1740

    Article  ADS  Google Scholar 

  5. Murawski L, Charchalis A (2014) Simplified method of torsional vibration calculation of marine power transmission system. Mar Struct 39:335–349

    Article  Google Scholar 

  6. Astashev VK, Babitsky VI (2007) Ultrasonic processes and machines—dynamics, control and applications. Springer, USA

    Book  MATH  Google Scholar 

  7. Al-Budairi H, Lucas M, Harkness P (2013) A design approach for longitudinal–torsional ultrasonic transducers. Sens Actuators A 198:99–106

    Article  Google Scholar 

  8. Yi Y, Seemann W, Gausmann R, Zhong J (2005) Development and analysis of a longitudinal and torsional type ultrasonic motor with two stators. Ultrasonics 43:629–634

    Article  Google Scholar 

  9. Kubojima Y, Sonoda S (2015) Measuring Young’s modulus of a wooden bar using longitudinal vibration without measuring its weight. Eur J Wood Prod. doi:10.1007/s00107-015-0884-2

    Google Scholar 

  10. Pontryagin LS, Boltyanskii VG, Gamkrelidze RV, Mishchenko EF (1962) The mathematical theory of optimal processes. Wiley, England

    Google Scholar 

  11. Szymczak C (1983) Optimal design of thin walled I beams for extreme natural frequency of torsional vibrations. J Sound Vib 86:235–241

    Article  ADS  MATH  Google Scholar 

  12. Szymczak C (1984) Optimal design of thin walled I beams for a given natural frequency of torsional vibrations. J Sound Vib 97:137–144

    Article  ADS  MATH  Google Scholar 

  13. Atanackovic TM, Djukic DS (1991) The influence of shear on the stability of a Pflüger column. J Sound Vib 144:531–535

    Article  ADS  Google Scholar 

  14. Atanackovic TM, Simic SS (1999) On the optimal shape of a Pflüger column. Eur J Mech A Solids 18:903–913

    Article  MathSciNet  MATH  Google Scholar 

  15. Glavardanov VB, Atanackovic TM (2001) Optimal shape of a twisted and compressed rod. Eur J Mech A Solids 20:795–809

    Article  MATH  Google Scholar 

  16. Atanackovic TM, Braun DJ (2005) The strongest rotating rod. Int J Non-Linear Mech 40:747–754

    Article  MathSciNet  MATH  Google Scholar 

  17. Atanackovic TM, Novakovic BN (2006) Optimal shape of an elastic column on elastic foundation. Eur J Mech A Solids 25:154–165

    Article  MathSciNet  MATH  Google Scholar 

  18. Atanackovic TM (2007) Optimal shape of a strongest inverted column. J Comput Appl Math 203:209–218

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Atanackovic TM (2007) Optimal shape of a rotating rod with unsymmetrical boundary conditions. J Appl Mech 74:1234

    Article  Google Scholar 

  20. Jelicic ZD, Atanackovic TM (2007) Optimal shape of a vertical rotating column. Int J Non-Linear Mech 42:172–179

    Article  Google Scholar 

  21. Braun DJ (2008) On the optimal shape of compressed rotating rod with shear and extensibility. Int J Non-Linear Mech 43:131–139

    Article  MATH  Google Scholar 

  22. Atanackovic TM, Jakovljevic BB, Petkovic MR (2010) On the optimal shape of a column with partial elastic foundation. Eur J Mech A Solids 29:283–289

    Article  MathSciNet  Google Scholar 

  23. Glavardanov VB, Spasic DT, Atanackovic TM (2012) Stability and optimal shape of Pflu¨ger micro/nano beam. Int J Solids Struct 49:2559–2567

    Article  Google Scholar 

  24. Atanackovic TM, Novakovic BN, Vrcelj Z (2012) Shape optimization against buckling of micro- and nano-rods. Arch Appl Mech 82:1303–1311

    Article  MATH  Google Scholar 

  25. Le MQ, Tran DT, Bui HL (2012) Optimal design of a torsional shaft system using Pontryagin’s maximum principle. Meccanica 47:1197–1207

    Article  MathSciNet  MATH  Google Scholar 

  26. Bui HL, Tran DT, Le MQ, Tran MT (2015) Multi-objective optimal control for eigen-frequencies of a torsional shaft using Pontryagin’s maximum principle. Meccanica. doi:10.1007/s11012-015-0162-8

    MathSciNet  MATH  Google Scholar 

  27. Thorby D (2008) Structural dynamics and vibration in practice. Elsevier, USA

    Google Scholar 

  28. Geering HP (2007) Optimal control with engineering applications. Springer, Berlin

    MATH  Google Scholar 

  29. Lebedev LP, Cloud MJ (2003) The calculus of variations and functional analysis with optimal control and applications in mechanics. World Scientific, Singapore

    MATH  Google Scholar 

Download references

Acknowledgments

This research is funded by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under Grant Number 107.02-2012.03.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Duc-Trung Tran.

Appendix: Proof of the necessary optimality condition (11b) cited in the Sect. 2.2

Appendix: Proof of the necessary optimality condition (11b) cited in the Sect. 2.2

The natural frequency ω i is here considered as a state variable. It means that the role of ω i is equivalent to those of u and N in Eq. (2b). The total weight W is also a state variable. So, the state Eq. (2b) of the bar can be rewritten in the form:

$$\frac{du}{dx} = \frac{N}{EA}$$
(15a)
$$\frac{dN}{dx} = - \rho A\omega_{i}^{2} u$$
(15b)
$$\frac{{d\omega_{i} }}{dx} = 0$$
(15c)
$$\frac{dW}{dx} = \rho A$$
(15d)

The objective function, Eq. (10), can be rewritten in term of the Maier objective functional as follows:

$$F = (1 - k_{\omega W} )\frac{{c_{\omega i} \omega_{i} (L)}}{{\omega_{0i} }} + k_{\omega W} \frac{{c_{W} W(L)}}{{W_{0} }} \to \hbox{min} .$$
(16)

The Hamiltonian function H can be established as follows:

$$H = p_{u} \frac{N}{EA} - p_{N} \rho A\omega_{i}^{2} u + p_{\omega } \omega_{i}^{\prime} + p_{W} \rho A,(\omega_{i}^{{\prime }} = 0)$$
(17)

The adjoint equations can be expressed under the following forms:

$$\frac{{dp_{u} }}{dx} = - \frac{\partial H}{\partial u} = \rho A\omega_{i}^{2} p_{N}$$
(18a)
$$\frac{{dp_{N} }}{dx} = - \frac{\partial H}{\partial N} = - \frac{1}{EA}p_{u}$$
(18b)
$$\frac{{dp_{\omega } }}{dx} = - \frac{\partial H}{{\partial \omega_{i} }} = 2\rho_{i} A_{i} \omega_{i} up_{N}$$
(18c)
$$\frac{{dp_{W} }}{dx} = - \frac{\partial H}{\partial W} = 0$$
(18d)

The adjoint variables \(p_{\varphi } ,p_{M} ,p_{\omega } ,p_{w}\) are determined as below:

$$\sum\limits_{j = 1}^{m} {p_{j} (L)\delta y_{j} (L)} - \sum\limits_{j = 1}^{m} {p_{j} (0)\delta y_{j} (0)} + \delta F = 0$$
(19)

where m = 4; y 1 = u, y 2 = N, y 3 = ω i and y 4 = W are state variables and p 1 = p u , p 2 = p N , p 3 = p ω and p 4 = p W are adjoint variables. Thus,

$$\begin{aligned} p_{u} (L)\delta u(L) + p_{N} (L)\delta N(L) + p_{\omega } (L)\delta \omega_{i} (L) + p_{W} (L)\delta W(L){-}p_{u} (0)\delta u(0){-}p_{N} (0)\delta N(0) \hfill \\ {-}\,p_{\omega } (0)\delta \omega_{i} (0){-}p_{W} (0)\delta W(0)(1 - k_{\omega W} )\frac{{c_{\omega i} \delta \omega_{i} (L)}}{{\omega_{0i} }} + k_{\omega W} \frac{{c_{W} \delta W(L)}}{{W_{0} }} = \, 0 \hfill \\ \end{aligned}$$
(20)

Or,

$$\begin{aligned} p_{u} (L)\delta u(L) + p_{N} (L)\delta N(L) + \left[ {p_{\omega } (L) + \frac{{c_{\omega i} (1 - k_{\omega W} )}}{{\omega_{0i} }}\delta \omega_{i} (L)} \right] + \left[ {p_{W} (L) + \frac{{c_{W} k_{\omega W} }}{{W_{0} }}} \right]\delta W(L) \hfill \\ {-}\,p_{u} (0)\delta u(0){-}p_{N} (0)\delta N(0) - p_{\omega } (0)\delta \omega_{i} (0){-}\,p_{W} (0)\delta W(0) = 0 \hfill \\ \end{aligned}$$
(21)

We obtain:

$$\left\{ {\begin{array}{*{20}c} {p_{\omega } (L) = - \frac{{c_{\omega i} (1 - k_{\omega W} )}}{{\omega_{0i} }}} \\ {p_{W} (L) = - \frac{{c_{W} k_{\omega W} }}{{W_{0} }}} \\ {p_{\omega } (0) = 0} \\ {p_{W} (0) = 0} \\ \end{array} } \right.$$
(22)

and other expressions which are depended on the boundary conditions of the bar system as:

  1. (a)
    $${\text{free}}{\text{-}}{\text{free:}}\;N(0) = N(L) = 0\;(3{\text{b}}) \Rightarrow p_{u} (L) = 0;p_{u} (0) = 0$$
    (23a)
  2. (b)
    $${\text{fixed}}{\text{-}}{\text{free:}}\;u(0) = 0,N(L) = 0\; (4{\text{b)}} \Rightarrow p_{u} (L) = 0;p_{N} (0) = 0$$
    (23b)
  3. (c)
    $${\text{free}}{\text{-}}{\text{fixed:}}\;N(0) = 0,u(L) = 0\; (5{\rm b})\Rightarrow p_{N} (L) = 0;p_{u} (0) = 0$$
    (23c)
  4. (d)
    $${\text{fixed}}{\text{-}}{\text{fixed:}}\;u(0) = 0,u(L) = 0\; (6{\text{b)}} \Rightarrow p_{N} (L) = 0;p_{N} (0) = 0$$
    (23d)

Assigning:

$$\left\{ {\begin{array}{*{20}c} {p_{u} = - N_{H} } \\ {p_{N} = u_{H} } \\ \end{array} } \right.$$
(24)

Combining Eqs. (18a), (18b) and (24) yields:

$$\left\{ {\begin{array}{*{20}c} {\frac{{du_{H}^{{}} }}{dx} = \frac{{N_{H} }}{EA}} \\ {\frac{{dN_{H}^{{}} }}{dx} = - \rho A\omega_{i}^{2} u_{H} } \\ \end{array} } \right.$$
(25)

Equations (23a, 23b, 23c, 23d) can be rewritten as follows:

  1. (a)
    $$p_{u} (L) = 0;p_{u} (0) = 0\;(23{\rm a}) \Rightarrow N_{H} (0) = 0;N_{H} (L) = 0$$
    (26a)
  2. (b)
    $$p_{u} (L) = 0;p_{N} (0) = 0\; (23{\text{b)}} \Rightarrow u_{H} (0) = 0;N_{H} (L) = 0$$
    (26b)
  3. (c)
    $$p_{N} (L) = 0;p_{u} (0) = 0\;(23{\text{c}}) \Rightarrow N_{H} (0) = 0;u_{H} (L) = 0$$
    (26c)
  4. (d)
    $$p_{N} (L) = 0;p_{N} (0) = 0\;(23{\text{d}}) \Rightarrow u_{H} (0) = 0;u_{H} (L) = 0$$
    (26d)

It should be emphasized that Eqs. (25) exhibit a similar functional form as Eqs. (2b). Equations (3b, 4b, 5b, 6b) and (26a, 26b, 26c, 26d) take also the same functional forms, describing the boundary conditions of the adjoint system.

Therefore, it can be concluded that an analogy between the adjoint variables and the original variables is obtained. It allows yielding the following equations:

$$\left\{ {\begin{array}{*{20}c} {kN_{H} = N \, } \\ {ku_{H} = u} \\ \end{array} } \right.$$
(27)

Explicit expression of k can be determined by integrating the Eq. (18c) with appropriate conditions in Eq. (22):

$$\int\limits_{0}^{L} {p_{\omega }^{\prime} dx = p_{\omega }^{{}} } (L) - p_{\omega } (0) = - \frac{{c_{\omega i} (1 - k_{\omega W} )}}{{\omega_{0i} }} = \frac{{2\rho \omega_{i} }}{k}\int\limits_{0}^{L} {Au^{2} dx}$$
(28)

Hence,

$$k = \frac{{2\rho \omega_{i} \omega_{0i} }}{{c_{\omega i} (1 - k_{\omega W} )}}\int\limits_{0}^{L} {Au^{2} dx}$$
(29)

Thus, the analogy coefficient k is positive/negative for the case of maximizing/minimizing ω i . It was demonstrated by considering the natural frequency ω i as a state variable. The Hamiltonian function, Eq. (11b), will be maximized if:

$$H = \frac{1}{k}\left[ { - \frac{{N^{2} }}{EA} - \rho A\omega_{i}^{2} u^{2} } \right] - \frac{{c_{W} k_{\omega W} }}{{W_{0} }}\rho A \to \hbox{max} \,({\text{in}}\;A)$$
(11b)
$${\text{Or}}\;H = - \frac{{N^{2} }}{EA} - \rho A\omega^{2} u^{2} - k^{*} \rho A \to \hbox{max}$$
(30)

where, \(k^{*} = k.\) \(\frac{{c_{W} k_{\omega W} }}{{W_{0} }}\) is a constant, if we have no objective of weight, \(k^{*} = 0.\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bui, HL., Tran, MT., Le, MQ. et al. Optimal configurations of circular bars under free torsional and longitudinal vibration based on Pontryagin’s maximum principle. Meccanica 51, 1491–1502 (2016). https://doi.org/10.1007/s11012-015-0324-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11012-015-0324-8

Keywords

Navigation