Skip to main content

Advertisement

Log in

Linear stability of piezoelectric-controlled discrete mechanical systems under nonconservative positional forces

  • Advances in Dynamics, Stability and Control of Mechanical Systems
  • Published:
Meccanica Aims and scope Submit manuscript

Abstract

Discrete models of piezoelectric controlled mechanical systems, subjected to follower forces, are derived in this paper. A targeted strategy of control is first discussed, leading to detect three novel linear controllers resembling, in the order, the Tuned Mass Damper with large and small mass, respectively, and Energy Sink, used in the literature for controlling linear and nonlinear oscillations. Here, however, and differently from those systems, coupling between the controller and the main structure is of gyroscopic type, whose magnitude is governed by an electro-mechanical parameter. Based on an eigenvalue sensitivity analysis, carried out via suitably selected perturbation methods, the effectiveness of the three controllers is investigated. As an application, the two-degree-of-freedom Ziegler column, undergoing Hopf bifurcation triggered by a follower force, equipped with piezoelectric controllers, is studied, and the different strategies proposed numerically compared.

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

Similar content being viewed by others

References

  1. Ziegler H (1952) Die stabilitätskriterien der elastomechanik. Ing Arch 20(1):49–56

    Article  MATH  MathSciNet  Google Scholar 

  2. Bolotin VV (1963) Nonconservative problems of the theory of elastic stability. Macmillan, New York

    MATH  Google Scholar 

  3. Seyranian AP, Mailybaev AA (2003) Multiparameter stability theory with mechanical applications, vol 13. World Scientific, Singapore

    MATH  Google Scholar 

  4. Kirillov ON (2007) Gyroscopic stabilization in the presence of nonconservative forces. Dokl Math 76(2):780–785

    Article  MATH  MathSciNet  Google Scholar 

  5. Hagedorn P (1970) On the destabilizing effect of non-linear damping in non-conservative systems with follower forces. Int J Non-linear Mech 5(2):341–358

    Article  MATH  MathSciNet  Google Scholar 

  6. Kirillov ON, Seyranian AP (2005) The effect of small internal and external damping on the stability of distributed non-conservative systems. J Appl Math Mech 69(4):529–552

    Article  MathSciNet  Google Scholar 

  7. Kirillov ON (2005) A theory of the destabilization paradox in non-conservative systems. Acta Mech 174(3–4):145–166

    Article  MATH  Google Scholar 

  8. Luongo A, D’Annibale F (2014a) A paradigmatic minimal system to explain the Ziegler paradox. Contin Mech Thermodyn. doi:10.1007/s00161-014-0363-8

  9. Luongo A, D’Annibale F (2014b) On the destabilizing effect of damping on discrete and continuous circulatory systems. J Sound Vib. doi:10.1016/j.jsv.2014.07.030

  10. Seyranian AP, Mailybaev AA (2011) Paradox of Nicolai and related effects. Zeitschrift für angewandte Mathematik und Physik 62(3):539–548

    Article  ADS  MATH  MathSciNet  Google Scholar 

  11. Seyranian AP, Di Egidio A, Contento A, Luongo A (2014) Solution to the problem of Nicolai. J Sound Vib 333(7):1932–1944

    Article  ADS  Google Scholar 

  12. Luongo A, Ferretti M, Seyranian AP (2014) Damping effects on stability of the compressed Nicolai beam. Math Mech Complex Syst. (in press)

  13. Beards CF (1996) Structural vibrations: analysis and damping. Butterworth, Heinemann, London

    Google Scholar 

  14. Blevins RD (1990) Flow-induced vibration. Van Nostrand Reinhold, New York

    Google Scholar 

  15. Abe M, Fujino Y (1994) Dynamic characterization of multiple tuned mass dampers and some design formulas. Earthq Eng Struct Dyn 23(8):813–835

    Article  Google Scholar 

  16. Den Hartog J (1956) Mechanical vibrations. McGraw-Hill, New York

    MATH  Google Scholar 

  17. Soong TT, Dargush GF (1997) Passive energy dissipation systems in structural engineering. Wiley, New York

    Google Scholar 

  18. Rana R, Soong TT (1998) Parametric study and simplified design of tuned mass dampers. Eng Struct 20(3):193–204

    Article  Google Scholar 

  19. Gattulli V, Di Fabio F, Luongo A (2001) Simple and double hopf bifurcations in aeroelastic oscillators with tuned mass dampers. J Frankl Inst 338(2–3):187–201

    Article  MATH  Google Scholar 

  20. Gattulli V, Di Fabio F, Luongo A (2003) One to one resonant double Hopf bifurcation in aeroelastic oscillators with tuned mass damper. J Sound Vib 262(2):201–217

    Article  ADS  Google Scholar 

  21. Vakakis AF, Bergman LA, Gendelman OV, Gladwell GM, Kerschen G, Lee YS, McFarland DM (2009) Nonlinear targeted energy transfer in mechanical and structural systems, vol 156. Springer, Netherlands

    Google Scholar 

  22. Luongo A, Zulli D (2012) Dynamic analysis of externally excited NES-controlled systems via a mixed Multiple Scale/Harmonic Balance algorithm. Nonlinear Dyn 70(3):2049–2061

    Article  MathSciNet  Google Scholar 

  23. Luongo A, Zulli D (2013) Aeroelastic instability analysis of NES-controlled systems via a mixed Multiple Scale/Harmonic Balance method. J Vib Control. doi:10.1177/1077546313480542

  24. Hagood NW, von Flotow A (1991) Damping of structural vibrations with piezoelectric materials and passive electrical networks. J Sound Vib 146:243–268

    Article  ADS  Google Scholar 

  25. Baz A, Poh S (1988) Performance of an active control system with piezoelectric actuators. J Sound Vib 126(2):327–343

    Article  ADS  Google Scholar 

  26. Elliott SJ, Gardonio P, Sors TC, Brennan MJ (2002) Active vibroacoustic control with multiple local feedback loops. J Acoust Soc Am 111(2):908–915

    Article  ADS  Google Scholar 

  27. Kurnik W, Przybyłowicz PM (2003) Active stabilisation of a piezoelectric fiber composite shaft subject to follower load. Int J Solids Struct 40(19):5063–5079

    Article  MATH  Google Scholar 

  28. Alessandroni S, Andreaus U, dell’Isola F, Porfiri M (2004) Piezo-ElectroMechanical (PEM) Kirchhoff–Love plates. Eur J Mech A/Solids 23:689–702

    Article  MATH  Google Scholar 

  29. Alessandroni S, dell’Isola F, Porfiri M (2002) A revival of electric analogs for vibrating mechanical systems aimed to their efficient control by PZT actuators. Int J Solids Struct 39:5295–5324

    Article  MATH  Google Scholar 

  30. dell’Isola F, Vidoli S (1998) Continuum modelling of piezoelectromechanical truss beams: an application to vibration damping. Arch Appl Mech 68(1):1–19

    Article  MATH  MathSciNet  Google Scholar 

  31. dell’Isola F, Henneke EG, Porfiri M (2002a) Synthesis of electrical networks interconnecting PZT actuators to damp mechanical vibrations. Int J Appl Electromagn Mech 14(1–4):417–424

    Google Scholar 

  32. dell’Isola F, Vestroni F, Vidoli S (2002b) A class of electro-mechanical systems: linear and nonlinear dynamics. Int J Appl Electromagn Mech 40(1):47–71

    MATH  Google Scholar 

  33. dell’Isola F, Maurini C, Porfiri M (2004) Passive damping of beam vibrations through distributed electric networks and piezoelectric transducers: prototype design and experimental validation. Smart Mater Struct

  34. Maurini C, dell’Isola F, Del Vescovo D (2002) Comparison of piezoelectronic networks acting as distributed vibration absorbers. Mech Syst Signal Process 18(5):1243–1271

    Article  ADS  Google Scholar 

  35. Porfiri M, dell’Isola F, Santini E (2005) Modeling and design of passive electric networks interconnecting piezoelectric transducers for distributed vibration control. Int J Appl Electromagn Mech 21(2):69–87

    Google Scholar 

  36. Rosi G, Pouget J, dell’Isola F (2010) Control of sound radiation and transmission by a piezoelectric plate with an optimized resistive electrode. Eur J Mech A-Solids 29(5):859–870

    Article  MathSciNet  Google Scholar 

  37. Crandall SH, Karnopp DC, Kurtz EF, Pridmore-Brown DC (1968) Dynamics of mechanical and electromechanical systems. Mc Graw-Hill, New York

    Google Scholar 

  38. Wang Y, Wang Z, Zu L (2012) Stability of viscoelastic rectangular plate with a piezoelectric layer subjected to follower force. Arch Appl Mech 83(4):495–507

    Article  Google Scholar 

  39. Belokon AV, Eremeyev VA, Nasedkin AV, Solov’yev AN (2000) Partitioned schemes of the finite-element method for dynamic problems of acoustoelectroelasticity. J Appl Math Mech 64(3):367–377

    Article  Google Scholar 

  40. Giorgio I, Culla A, Del Vescovo D (2009) Multimode vibration control using several piezoelectric transducers shunted with a multiterminal network. Arch Appl Mech 79(9):859–879

    Article  MATH  Google Scholar 

  41. Rosi G, Paccapeli R, Ollivier F, Pouget J (2012) Optimization of piezoelectric patches positioning for passive sound radiation control of plates. J Vib Control 19(5):658–673

    Article  MathSciNet  Google Scholar 

  42. Vidoli S, dell’Isola F (2000) Modal coupling in one-dimensional electromechanical structured continua. Acta Mech 141(1–2):37–50

    Article  MATH  Google Scholar 

  43. Luongo A (1993) Eigensolutions sensitivity for nonsymmetric matrices with repeated eigenvalues. AIAA J 31(7):1321–1328

    Article  ADS  MATH  Google Scholar 

  44. Luongo A (1995) Free vibrations and sensitivity analysis of a defective two degree-of-freedom system. AIAA J 33(1):120–127

    Article  ADS  MATH  Google Scholar 

  45. Luongo A, Ferretti M (2014) Can a semi-simple eigenvalue admit fractional sensitivities? Appl Math Comput. doi:10.1016/j.amc.2014.01.178

  46. Guckenheimer J, Holmes P (1983) Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, vol 42. Springer, New York

    MATH  Google Scholar 

  47. Luongo A, Paolone A, Di Egidio A (2003) Multiple timescales analysis for 1:2 and 1:3 resonant hopf bifurcations. Nonlinear Dyn 34(3–4 SPEC. ISS.):269–291

    Article  MATH  MathSciNet  Google Scholar 

  48. Luongo A, Di Egidio A, Paolone A (2004) Multiscale analysis of defective multiple-hopf bifurcations. Comput Struct 82(31–32):2705–2722

    Article  Google Scholar 

  49. Luongo A, Di Egidio A (2005) Bifurcation equations through multiple-scales analysis for a continuous model of a planar beam. Nonlinear Dyn 41(1–3):171–190

    Article  MATH  Google Scholar 

  50. Luongo A, Egidio A (2006) Divergence, hopf and double-zero bifurcations of a nonlinear planar beam. Comput Struct 84(24–25):1596–1605

    Article  Google Scholar 

  51. Di Egidio A, Luongo A, Paolone A (2007) Linear and non-linear interactions between static and dynamic bifurcations of damped planar beams. Int J Non-linear Mech 42(1):88–98

    Article  MATH  Google Scholar 

  52. Luongo A, D’Annibale F (2012) Bifurcation analysis of damped visco-elastic planar beams under simultaneous gravitational and follower forces. Int J Mod Phys B 26(25)

  53. Luongo A, D’Annibale F (2013) Double zero bifurcation of non-linear viscoelastic beams under conservative and non-conservative loads. Int J Non-linear Mech 55:128–139

    Article  Google Scholar 

  54. Bilharz H (1944) Bemerkung zu einem Satze von Hurwitz. Zeitschrift für Angewandte Mathematik und Mechanik 24:77–82

    Article  MATH  MathSciNet  Google Scholar 

  55. Perelmuter AV, Slivker V (2013) Handbook on mechanical stability in engineering. Institute of Electrical and Electronic Engineers, World Scientific, New Jersey

    Book  Google Scholar 

  56. Atluri SN, Cazzani A (1995) Rotations in computational solid mechanics. Arch Comput Methods Eng 2(1):49–138

    Article  MathSciNet  Google Scholar 

  57. Greco L, Cuomo M (2013) B-Spline interpolation of Kirchhoff–Love space rods. Comput Methods Appl Mech Eng 256:251–269

    Article  ADS  MathSciNet  Google Scholar 

  58. IEEE (1987) IEEE standard on piezoelectricity—IEEE Std 176-1987. Institute of Electrical and Electronic Engineers

Download references

Acknowledgments

This work was supported by the Italian Ministry of University (MIUR) through the PRIN cofinanced program no.2010MBJK5B.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Francesco D’Annibale.

Appendix: Discrete piezoelectro-mechanical model

Appendix: Discrete piezoelectro-mechanical model

We consider a discrete piezoelectro-mechanical system. In it, we distinguish three subsystems, each made of elementary components, namely: a structure, consisting of masses (M), springs (S) and dashpots (D); an electrical circuit, made of capacitors (C), inductors (L), resistors (R) and, occasionally, transformers; a set of piezoelectric devices (P).

The system is conveniently regarded as constituted by point joints, connected by one-dimensional devices. We consider the structure made by mechanical joints, at which masses are lumped, interconnected by S and D devices. Similarly, we consider the circuit as constituted by electrical joints, linked by C,L,R devices. Finally, each of the P devices links a couple of mechanical joints and a couple of electrical joints. The state of the system is defined by n m displacements \({\mathbf{X}}:=\left\{ X_{I}\right\} \) of the mechanical joints and n e flux linkages \({\mathbf{Y}}:=\left\{ \psi _{J}\right\} \) at the electrical joints, together with their time-derivatives \({\dot{\mathbf{X}}},{\dot{\mathbf{Y}}}\) (we remember that the flux linkage at a point is defined as the time-integral of the potential V J at that point, namely \({{\dot{\psi}}}_{J}:=V_{J}\)).

1.1 Constitutive laws

The linear constitutive laws of the structure components are:

$$f_{i}^{S}=k_{i}\varDelta l_{i}^{S},\qquad f_{i}^{D}=b_{i}\varDelta \dot{l}_{i}^{D}$$
(39)

where \(f_{i}^{H}\,\left( H=S,D\right) \) are forces, k i spring stiffnesses, b i damping coefficients and \(\varDelta l_{i}^{H}\) elongations.

The constitutive laws of the circuit components are:

$$Q_{j}^{C}=C_{j}\varDelta {\dot{\psi}}_{j}^{C},\qquad \dot{Q}_{j}^{L}=\frac{1}{L_{j}}\varDelta \psi _{j}^{L},\qquad \dot{Q}_{j}^{R}=\frac{1}{R_{j}}\varDelta {\dot{\psi}}_{j}^{R}$$
(40)

where \(Q_{j}^{\alpha }\,\left( \alpha =C,L,R\right) \) are electric charges, \(C_{j},\, L_{j},\, R_{j}\) capacitances, inductances and resistances, respectively, and \(\varDelta \psi _{j}^{\alpha }\) is the flux-linkage difference between the two terminals of the device. According to a (one of the two possible [37]) electro-mechanical analogy, the capacitance corresponds to the mass, the inverse of the inductance to the stiffness, the inverse of the resistance to the viscous damping, while the time-derivative of the charge corresponds to the force, and the flux linkage to the displacement.

The constitutive law of the piezoelectric components is taken in the following gyroscopic-type form [58]:

$$\left\{ \begin{array}{l} f_{k}^{P}\\ Q_{k}^{P} \end{array}\right\} =\left[ \begin{array}{ll} k_{k}^{P} &{} -g_{k}^{P}\\ g_{k}^{P} &{} C_{k}^{P} \end{array}\right] \left\{ \begin{array}{l} \varDelta l_{k}^{P}\\ \varDelta {\dot{\psi}}_{k}^{P} \end{array}\right\}$$
(41)

where \(k_{k}^{P}\), \(C_{k}^{P}\), \(g_{k}^{P}\) are stiffness, capacitance and electro-mechanical coupling, respectively.

By collecting forces \({\mathbf{f}}_{\alpha }:=\left\{ f_{i}^{\alpha }\right\} \) \(\left( \alpha =S,D,P\right) \) as well charges \({\mathbf{Q}}_{\alpha }:=\left\{ Q_{j}^{\alpha }\right\} \) \(\left( \alpha =C,L,R,P\right) \), the constitutive laws (39), (40), (41) also read:

$$\begin{aligned} \begin{aligned} {\mathbf{f}}_{S}&={\mathbf{k}}{\mathbf{\varvec{\Updelta}}}\varvec{l}_{S},\,\,\,\,\,\,\,\,{\mathbf{f}}_{D}={\mathbf{b}}{\mathbf{\varvec{\Updelta}}}{\dot{\varvec l}}_{D}\\ {\mathbf{Q}}_{C}&={\mathbf{C}}\varvec{\Updelta }{\dot{{\varvec{\psi}}}}_{C},\,\,{\dot{\mathbf{Q}}}_{L}={\mathbf{L}}^{-1}\varvec{\Updelta }\varvec{\psi }_{L},\;{\dot{\mathbf{Q}}}_{R}={\mathbf{R}}^{-1}\varvec{\Updelta }{\dot{{\varvec{\psi}}}}_{R}\\ \left\{ \begin{array}{l} {\mathbf{f}}_{P}\\ {\mathbf{Q}}_{P} \end{array}\right\}&=\left[ \begin{array}{ll} {\mathbf{k}}_{P} &{} -{\mathbf{g}}_{P}\\ {\mathbf{g}}_{P} &{} {\mathbf{C}}_{P} \end{array}\right] \left\{ \begin{array}{l} {\mathbf{\varvec{\Updelta}}}\varvec{l}_{P}\\ \varvec{\Updelta }{\dot{{\varvec{\psi}}}}_{P} \end{array}\right\} \end{aligned} \end{aligned}$$
(42)

where:

$$\begin{aligned} \begin{aligned}{\mathbf{k}}&:={\rm diag}\left[ k_{i}\right] ,&{\mathbf{b}}&:={\rm diag}\left[ b_{i}\right] \\ {\mathbf{C}}&:={\rm diag}\left[ C_{j}\right] ,&{\mathbf{L}}^{-1}&:={\rm diag}\left[ 1/L_{j}\right] ,&\,{\mathbf{R}}^{-1}&:={\rm diag}\left[ 1/R_{j}\right] \\ {\mathbf{k}}_{P}&:={\rm diag}\left[ k_{k}^{P}\right] ,&{\mathbf{C}}_{P}&:={\rm diag}\left[ C_{k}^{P}\right] ,&\,{\mathbf{g}}_{P}&:={\rm diag}\left[ g_{k}^{P}\right] \end{aligned} \end{aligned}$$
(43)

1.2 Topology

Accounting for topology (and geometry) of the structure, the set of the spring elongations \({\mathbf{\varvec{\Updelta}}}\varvec{l}_{S}:=\left\{ \Updelta l_{i}^{S}\right\} \) and of the dashpot elongation-rates \({\mathbf{\varvec{\Updelta}}}{\dot{\varvec l}}_{D}:=\left\{ \Updelta \dot{l}_{i}^{D}\right\} \) is expressed in terms of nodal displacements and velocities:

$$\begin{aligned} {\mathbf{\varvec{\Updelta}}}\varvec{l}_{S}={\mathbf{D}}_{S}{\mathbf{X}},\qquad {\mathbf{\varvec{\Updelta}}}{\dot{\varvec l}}_{D}={\mathbf{D}}_{D}{\dot{\mathbf{X}}} \end{aligned}$$
(44)

where \({\mathbf{D}}_{H}\,\left( H=S,D\right) \) are kinematic operators. Topology of the circuit provides the differences of the flux linkages and their derivatives, \(\varvec{\Delta }{\dot{{\varvec{\psi}}}}_{C}:=\left\{ \Delta {\dot{\psi}}_{j}^{C}\right\} \), \(\varvec{\Delta }\varvec{\psi }_{L}:=\left\{ \Delta \psi _{j}^{L}\right\} \), \(\varvec{\Delta }{\dot{{\varvec{\psi}}}}_{R}:=\left\{ \Delta {\dot{\psi}}_{j}^{R}\right\} \), expressed in terms of the homologous quantities at the joints:

$$\begin{aligned} \varvec{\Delta }{\dot{{\varvec{\psi}}}}_{C}={\mathbf{D}}_{C}{\dot{\mathbf{Y}}},\qquad \varvec{\Delta }\varvec{\psi }_{L}={\mathbf{D}}_{L}{\mathbf{Y}},\qquad \varvec{\Delta }{\dot{{\varvec{\psi}}}}_{R}={\mathbf{D}}_{R}{\dot{\mathbf{Y}}} \end{aligned}$$
(45)

For the piezoelectric arrangement, for which \({\mathbf{\varvec{\Delta}}}\varvec{l}_{P}:=\left\{ \Delta l_{k}^{P}\right\} \), \(\varvec{\Delta }{\dot{{\varvec{\psi}}}}_{P}:=\left\{ \Delta {\dot{\psi}}_{k}^{P}\right\} \), we similarly have:

$$\begin{aligned} {\mathbf{\varvec{\Delta}}}\varvec{l}_{P}={\mathbf{D}}_{PM}{\mathbf{X}},\qquad \varvec{\Delta }{\dot{{\varvec{\psi}}}}_{P}={\mathbf{D}}_{PE}{{{\dot{\mathbf{Y}}}}} \end{aligned}$$
(46)

All the \({\mathbf{D}}_{H}\) will be referred as topological operators.

1.3 Hamilton principle

The evolution equations are derived by the extended Hamilton principle:

$$\begin{aligned} \delta H:=\int _{t_{1}}^{t_{2}}\left( \delta T-\delta U+\delta W\right) {\text {d}}t=0,\quad \forall \,\delta {\mathbf{X}},{\mathbf{\delta Y}} \end{aligned}$$
(47)

where T is kinetic energy, U the potential energy, and \(\delta W\) is the work of the nonconservative forces done in the virtual variation of the state, \(\delta {\mathbf{X}},{\mathbf{\delta Y}}.\) Moreover \(\left[ t_{1},t_{2}\right] \) is an arbitrary interval of time, at whose ends motion is prescribed.

Concerning the kinetic energy, this is the sum of two contributions, relevant to the masses of the structure and of the piezoelectric devices, respectively; this latter, however, is usually negligible, so that:

$$\begin{aligned} T=\frac{1}{2}{\dot{\mathbf{X}}}^{T}{\mathbf{M}}_{M}{\dot{\mathbf{X}}} \end{aligned}$$
(48)

where \({\mathbf{M}}_{M}\) is the structure inertia matrix.

The potential energy of the system is the sum of four contributions, namely \(U:=U_{S}+U_{C}+U_{L}+U_{P}\), relevant to springs, capacitors, inductors and piezoelectric devices, respectively. Accounting for the constitutive laws (42) and for the topology (44)-(46), they read:

$$\begin{aligned} \begin{aligned}&U_{S}=\frac{1}{2}{\mathbf{f}}_{S}^{T}{\mathbf{\varvec{\Delta}}}\varvec{l}_{S}=\frac{1}{2}{\mathbf{X}}^{T}{\mathbf{D}}_{S}^{T}{\mathbf{k}}{\mathbf{D}}_{S}{\mathbf{X}}=\frac{1}{2}{\mathbf{X}}^{T}{\mathbf{K}}_{S}{\mathbf{X}}\\&U_{C}=-\frac{1}{2}{\mathbf{Q}}_{C}^{T}\varvec{\Delta }{\dot{{\varvec{\psi}}}}_{C}=-\frac{1}{2}{\dot{\mathbf{Y}}}^{T}{\mathbf{D}}_{C}^{T}{\mathbf{C}}{\mathbf{D}}_{C}{\dot{\mathbf{Y}}}=-\frac{1}{2}{\dot{\mathbf{Y}}}^{T}{\mathbf{M}}_{C}{\dot{\mathbf{Y}}}\\&U_{L}=\frac{1}{2}{\dot{\mathbf{Q}}}_{L}^{T}\varvec{\Delta }\varvec{\psi }_{L}=\frac{1}{2}{\mathbf{Y}}^{T}{\mathbf{D}}_{L}^{T}{\mathbf{L}}^{-1}{\mathbf{D}}_{L}{\mathbf{Y}}=\frac{1}{2}{\mathbf{Y}}^{T}{\mathbf{K}}_{L}{\mathbf{Y}} \end{aligned} \end{aligned}$$
(49)

while:

$$\begin{aligned} \begin{aligned}&U_{P} = \;\;\frac{1}{2}\left( {\mathbf{f}}_{P}^{T}{\mathbf{\varvec{\Delta}}}\varvec{l}_{P}-{\mathbf{Q}}_{P}^{T}\varvec{\Delta }{\dot{{\varvec{\psi}}}}_{P}\right) \\&= \;\;\frac{1}{2}\left( {\mathbf{X}}^{T}{\mathbf{D}}_{PM}^{T}{\mathbf{k}}_{P}{\mathbf{D}}_{PM}{\mathbf{X}}-{\dot{\mathbf{Y}}}^{T}{\mathbf{D}}_{PE}^{T}{\mathbf{C}}_{P}{\mathbf{D}}_{PE}{\dot{\mathbf{Y}}}\right) \\&\;\; -\frac{1}{2}\left( {\dot{\mathbf{Y}}}^{T}{\mathbf{D}}_{PE}^{T}{\mathbf{g}}_{P}{\mathbf{D}}_{PM}{\mathbf{X}}+{\mathbf{X}}^{T}{\mathbf{D}}_{PM}^{T}{\mathbf{g}}_{P}{\mathbf{D}}_{PE}{\dot{\mathbf{Y}}}\right) \\&= \;\;\frac{1}{2}\left( {\mathbf{X}}^{T}{\mathbf{K}}_{P}{\mathbf{X}}-{\dot{\mathbf{Y}}}^{T}{\mathbf{M}}_{P}{\dot{\mathbf{Y}}}\right) -{\dot{\mathbf{Y}}}^{T}{\mathbf{G}}{\mathbf{X}} \end{aligned} \end{aligned}$$
(50)

Here, the following positions hold for matrices:

$$\begin{aligned} \begin{aligned}{\mathbf{K}}_{S}&:={\mathbf{D}}_{S}^{T}{\mathbf{k}}{\mathbf{D}}_{S},&\,{\mathbf{K}}_{L}&:={\mathbf{D}}_{L}^{T}{\mathbf{L}}^{-1}{\mathbf{D}}_{L},&{\mathbf{K}}_{P}&:={\mathbf{D}}_{PM}^{T}{\mathbf{k}}_{P}{\mathbf{D}}_{PM}\\ {\mathbf{M}}_{C}&:={\mathbf{D}}_{C}^{T}{\mathbf{C}}{\mathbf{D}}_{C},&\,{\mathbf{M}}_{P}&:={\mathbf{D}}_{PE}^{T}{\mathbf{C}}_{P}{\mathbf{D}}_{PE},&{\mathbf{G}}&:={\mathbf{D}}_{PE}^{T}{\mathbf{g}}_{P}{\mathbf{D}}_{PM} \end{aligned} \end{aligned}$$
(51)

to be referred, in the order, as: the spring, inductance and piezoelectric stiffnesses; the capacitor and piezoelectric masses; the electro-mechanical gyroscopic coupling. All stiffness and mass matrices are squared and symmetric; the gyroscopic matrix is generally non-squared and, even when it is squared, is generally non-symmetric. Since, usually, \(\left\| {\mathbf{K}}_{P}\right\| \ll \left\| {\mathbf{K}}_{S}\right\| \), the piezoelectric contribution to the system stiffness is neglected. Note that the capacitor potential formally appears as (changed of sign) kinetic energy, since it depends on \({\dot{\mathbf{Y}}}\).

By collecting the previous contributions, and differently grouping them, we have:

$$\begin{aligned} \begin{aligned}&T=\frac{1}{2}{\dot{\mathbf{X}}}^{T}{\mathbf{M}}_{M}{\dot{\mathbf{X}}}+\frac{1}{2}{\dot{\mathbf{Y}}}^{T}\left( {\mathbf{M}}_{C}+{\mathbf{M}}_{P}\right) {\dot{\mathbf{Y}}}\\&U=\frac{1}{2}{\mathbf{X}}^{T}\left( {\mathbf{K}}_{S}+{\mathbf{K}}_{P}\right) {\mathbf{X}}+\frac{1}{2}{\mathbf{Y}}^{T}{\mathbf{K}}_{L}{\mathbf{Y}}-{\dot{\mathbf{Y}}}^{T}{\mathbf{G}}{\mathbf{X}} \end{aligned} \end{aligned}$$
(52)

or:

$$\begin{aligned} \begin{aligned}&T-U&= \;\;\frac{1}{2}{\dot{\mathbf{X}}}^{T}{\mathbf{M}}_{m}{\dot{\mathbf{X}}}+\frac{1}{2}{\dot{\mathbf{Y}}}^{T}{\mathbf{M}}_{e}{\dot{\mathbf{Y}}}-\frac{1}{2}{\mathbf{X}}^{T}{\mathbf{K}}_{m}{\mathbf{X}}\\&&\;\; -\frac{1}{2}{\mathbf{Y}}^{T}{\mathbf{K}}_{e}{\mathbf{Y}}+{\dot{\mathbf{Y}}}^{T}{\mathbf{G}}{\mathbf{X}} \end{aligned} \end{aligned}$$
(53)

where we defined:

$$\begin{aligned} \begin{aligned} {\mathbf{M}}_{m}&:={\mathbf{M}}_{M},&\,{\mathbf{M}}_{e}&:={\mathbf{M}}_{C}+{\mathbf{M}}_{P}\\ {\mathbf{K}}_{m}&:={\mathbf{K}}_{S}+{\mathbf{K}}_{P},&\,{\mathbf{K}}_{e}&:={\mathbf{K}}_{L} \end{aligned} \end{aligned}$$
(54)

as the mechanical (m) and the electrical (e) mass and stiffness matrices, respectively. Usually, since piezoelectric devices already behave as capacitors, no further capacitors are needed in the circuit, so that \({\mathbf{M}}_{C}={\mathbf{0}}\).

Concerning nonconservative actions, they consist of: (a) the virtual work of the damping forces \({\mathbf{f}}_{D}\) over the elongation-rate \(\delta \left( {\mathbf{\varvec{\Delta}}}\varvec{l}_{D}\right) \) of the dashpots; (b) the complementary virtual work expended by the charge-rates at resistors \({\dot{\mathbf{Q}}}_{R}\) over the virtual variation of the flux linkage differences \(\delta \left( {\mathbf{\varvec{\Delta}}}\varvec{\psi }{}_{R}\right) \) of the resistors; (c) the virtual work expended by the external forces acting on the structure, \({\mathbf{f}}_{nc}:=-{\mathbf{H}}{\mathbf{X}}\), here assumed to be linear and homogeneous in the displacements (circulatory forces, with \({\mathbf{H}}\) the circulatory matrix, [2]), over the virtual displacements \(\delta {\mathbf{X}}\). They, respectively, read:

$$\begin{aligned} \begin{aligned}\delta W_{D}&=-{\mathbf{f}}_{D}^{T}\delta \left( {\mathbf{\varvec{\Delta}}}\varvec{l}_{D}\right) =-\delta {\mathbf{X}}^{T}{\mathbf{D}}_{D}^{T}{\mathbf{b}}{\mathbf{D}}_{D}\dot{\mathbf{X}}=-\delta {\mathbf{X}}^{T}{\mathbf{B}}_{m}{\dot{\mathbf{X}}}\\ \delta W_{R}&=-{\dot{\mathbf{Q}}}_{R}\delta \left( {\mathbf{\varvec{\Delta}}}\varvec{\psi }{}_{R}\right) =-\delta {\mathbf{Y}}^{T}{\mathbf{D}}_{R}^{T}{\mathbf{\mathbf{R}}}^{-1}{\mathbf{D}}_{R}\dot{\mathbf{Y}}=-\delta {\mathbf{Y}}^{T}{\mathbf{B}}_{e}{\dot{\mathbf{Y}}}\\ \delta W_{nc}&=-{\mathbf{f}}_{nc}^{T}\delta {\mathbf{X}}=-\delta {\mathbf{X}}^{T}{\mathbf{H}}{\mathbf{X}} \end{aligned} \end{aligned}$$
(55)

where:

$$\begin{aligned} {\mathbf{B}}_{m}:={\mathbf{D}}_{D}^{T}{\mathbf{b}}{\mathbf{D}}_{D},\qquad {\mathbf{B}}_{e}:={\mathbf{D}}_{R}^{T}{\mathbf{R}}^{-1}{\mathbf{D}}_{R} \end{aligned}$$
(56)

are the mechanical and electrical damping matrices, respectively.

By substituting Eq. (53) and (55) in the Principle (47), this latter reads \(\delta H_m+\delta H_e+\delta H_{em}=0\), \(\forall \,\delta {\mathbf{X}},\,\delta {\mathbf{Y}}\), where:

$$\begin{aligned} \begin{aligned}&\delta H_{m}\ \; := \int _{t_{1}}^{t_{2}}\left( \delta \dot{\mathbf{X}}^{T}{\mathbf{M}}_{m}\dot{\mathbf{X}}-{\mathbf{\delta X}}^{T}{\mathbf{K}}_{m}{\mathbf{X}}-\delta {\mathbf{X}}^{T}{\mathbf{B}}_{m}\dot{\mathbf{X}}\right) {\rm d}t\\& \; \quad -\int _{t_{1}}^{t_{2}}{\mathbf{\delta X}}^{T}{\mathbf{H}}{\mathbf{X}}{\rm d}t\\&\delta H_{e} := \; \int _{t_{1}}^{t_{2}}\left( \delta \dot{\mathbf{Y}}^{T}{\mathbf{M}}_{e}\dot{\mathbf{Y}}-{\mathbf{\delta Y}}^{T}{\mathbf{K}}_{e}{\mathbf{Y}}-{\mathbf{\delta Y}}^{T}{\mathbf{B}}_{e}\dot{\mathbf{Y}}\right) {\rm d}t\\&\delta H_{em} := \;\int _{t_{1}}^{t_{2}}\left( \delta {\mathbf{X}}^{T}{\mathbf{G}}^{T}{\dot{\mathbf{Y}}}+{\delta \dot{\mathbf{Y}}}{}^{T}{\mathbf{G}}{\mathbf{X}}\right) {\rm d}t \end{aligned} \end{aligned}$$
(57)

After integration by parts, the Eq. (1) is finally obtained.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

D’Annibale, F., Rosi, G. & Luongo, A. Linear stability of piezoelectric-controlled discrete mechanical systems under nonconservative positional forces. Meccanica 50, 825–839 (2015). https://doi.org/10.1007/s11012-014-0037-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11012-014-0037-4

Keywords

Navigation