Vous êtes sur la page 1sur 26

Meccanica 34: 311–336, 1999.

c 2000 Kluwer Academic Publishers. Printed in the Netherlands.

Nonlinear Dynamics and Stability of a Two D.O.F.


Elastic/Elasto-Plastic Model System

S.V. SOROKIN1, A.V. TERENTIEV1 and B.L. KARIHALOO2


1 State Marine Technical University of St. Petersburg, Department of Engineering Mechanics, Lotsmanskaya
str.3; St. Petersburg 190008, Russia
2 University of Wales Cardiff, Cardiff School of Engineering, Queen’s Buildings, PO Box 686; Cardiff CF2 3TB,
United Kingdom

(Received: 15 March 1999; accepted in revised form: 25 June 1999)

Abstract. The local and global nonlinear dynamics of a two-degree-of-freedom model system is studied. The
undeflected model consists of an inverted T formed by three rigid bars, with the tips of the two horizontal bars
supported on springs. The springs exhibit an elasto-plastic response, including the Bauschinger effect. The vertical
rigid bar is subjected to a conservative (dead) or non-conservative (follower) force having static and periodic
components. First, the method of multiple scales is used for the analysis of the local dynamics of the system
with elastic springs. The attention is focused at modal interaction phenomena in weak excitation at primary res-
onance and in hard sub-harmonic excitation. Three different asymptotic expansions are utilised to get a structural
response for typical ranges of excitation parameters. Numerical integration of the governing equations is then
performed to validate results of asymptotic analysis in each case. A full global nonlinear dynamics analysis of the
elasto-plastic system is performed to reveal the role of plastic deformations in the stability of this system. Static
‘force-displacement’ curves are plotted and the role of plastic deformations in the destabilisation of the system is
discussed. Large-amplitude non-linear oscillations of the elasto-plastic system are studied, including the influence
of material hardening and of static and sinusoidal components of the applied force. A practical method is proposed
for the study of a non-conservative elasto-plastic system as a non-conservative elastic system with an ‘equivalent’
viscous damping.
Sommario. Viene studiata la dinamica nonlineare locale e globale di un sistema modello a due gradi di libertà.
Il sistema consiste in un elemento a T invertito formato da tre barre rigide, con gli estremi delle barre orizzo-
ntali supportate da molle che esibiscono una risposta elasto-plastica comprensiva dell’effetto Bauschinger. La
barra rigida verticale è soggetta ad una forza conservativa (dead) o non-conservativa (follower) con compon-
enti statica e variabile periodicamente. Si analizza dapprima la dinamica locale del sistema con molle elastiche
tramite il metodo delle scale multiple. L’attenzione viene concentrata sui fenomeni di interazione modale sotto
eccitazioni di limitata intensità in risonanza primaria o sotto eccitazioni di intensità elevata in risonanza subar-
monica. Vengono utilizzate tre differenti espansioni asintotiche per ottenere la risposta del sistema per valori tipici
dei parametri dell’eccitazione. Si effettua poi una simulazione numerica delle equazioni che governano il problema
per convalidare i risultati forniti dall’analisi asintotica in ciascun caso. Viene effettuata una analisi dinamica
nonlineare globale del sistema elasto-plastico: in tale contesto, si presentano curve ‘forza-spostamento’ statiche
e si discute il ruolo delle deformazioni plastiche nella stabilità del sistema. Si studiano oscillazioni nonlineari
di grande ampiezza includendo l’influenza dell’incrudimento del materiale e delle componenti statica e sinus-
oidale della forza applicata. Viene infine proposto un metodo pratico per lo studio di un sistema elasto-plastico
non-conservativo attraverso un sistema elastico non-conservativo dotato di smorzamento viscoso ‘equivalente’.

Key words: Plasticity and visco-plasticity, Stability, Vibrations (structures), Nonlinear dynamics.
312 S.V. Sorokin et al.

1. Introduction

The two-degree-of-freedom system considered in this paper (see Figure 1) has been introduced
by Sewell [1] to analyse elasto-plastic stability of a column by a static perturbation technique.
Although no discussion of practical relevance of this model has been given in [1], this system
may be thought of as, for example, a lumped parameters model of a fluid-conveying tube
built in an absolutely rigid plate (a platform) resting at a continuous foundation (a soil). The
modelling of a realistic fluid-conveying structure by an articulated system of hinged elements
[2–7] is relevant to the case of a long and flexible straight tube when the angular motion
is associated with its deformation. However, in some practical applications (in particular, in
some devices used in oil industry) a tube is rather short and not very flexible whereas a plate
is not completely fixed. Then it is more important to take into account the angular motion
of the whole system ‘a tube – a plate’, rather than to consider an articulated cantilevered
tube. The angular stiffness is produced by distributed reactive forces acting from a soil at the
foundation. In lumped parameters modelling, two resultants of these forces may be treated as
reactions in two spring elements shown in Figure 1. Moreover, the behaviour of the soil may
follow Hooke’s law only up to a certain level of loading. As this level is exceeded, inelastic
(plastic) effects may become dominant. In such a case, as a first rough approximation, fluid
loading may be modelled simply by a follower force acting at the upper end of a tube (as it
is introduced by Sewell [1]). Although a refined modelling of fluid loading should include
coriolis and centrifugal forces, the Sewell model is useful to explore elasto-plastic behaviour
of such a structure. Local nonlinear dynamics of two-degree-of-freedom linearly elastic model
systems at typical excitation conditions have been analysed by many authors with the use of
multiple scales method (see for example [2, 8–15] and references given there). The Sewell
model with ideally elastic springs falls into the class of systems considered in [10–15]. How-
ever, there are some specific features in its local dynamics controlled by static and periodic
components of conservative and non-conservative forces, which are explored in the first part of

Figure 1. The Sewell model for an elasto-plastic column.


Nonlinear Dynamics and Stability 313

the present paper as a meaningful preliminary step for the description of the system nonlinear
dynamics.
In the second part of the paper the attention is focused at global elastic and elasto-plastic
dynamics of Sewell’s model. Examples of similar elasto-plastic (piecewise linear) systems
subjected to periodic loading may be found in various engineering applications (see [16–21]).
The usual source of the nonlinearity is a discontinuity of the material stiffness depending on
state variables. We consider here material model with kinematic hardening and assume its
bilinear hysteretic behaviour. There are some studies dedicated to investigations of systems
having such a constitutive law using both analytical and numerical methods. Klyushnikov
[16] suggested the model of rod behaviour including plastic effects and performed quasi-static
analysis of its stability. Shaw and Holmes [17] studied a single-degree-of-freedom oscillator
including nonlinearity of the restoring force that is piecewise linear with a single change
in slope. Conditions of existence of harmonic, sub-harmonic and chaotic motions and the
corresponding bifurcation conditions are derived. Narayanan and Sekar [18] adopted multi-
harmonic balance cum Newton–Raphson procedure with an FFT algorithm to obtain the stable
and unstable periodic solutions corresponding to a single-degree-of-freedom vibrating system
with unsymmetric piecewise linear stiffness. Pratap et al. [19, 20] investigated the resulting
motion of a single-degree-of-freedom oscillator with kinematic hardening using eigenvalues
of the linearized Poincaré map. In their analysis two kinds of periodic impulse forcing are
suggested: direct and parametric. Chatterjee et al. [21] have extended the method of equivalent
linearization to analyse different oscillating systems having piecewise nonlinearity. Using this
method allows not only to improve numerical results but is suitable to the analysis of both
super-harmonic motions and sub-harmonic motions.
The above analyses are performed for single-degree-of-freedom models by using simpli-
fied analytical approaches. In the present paper, a more complicated system is considered
and attention is focused at bi-modal oscillations including pronouncedly developed plastic
deformations. The extensive analysis of global nonlinear dynamics of elasto-plastic model
is carried out to reveal some features of its behaviour not observed in conservative sys-
tems.

2. Governing Equations for the Linearly Elastic Sewell’s Model System

The Sewell model shown in its deflected state in Figure 1 consists of an inverted T in its
undeflected state formed by three rigid bars. The vertical bar of length l has its mass m,
lumped at its top, whereas the two horizontal bars, each of length c, have a mass m1 per
unit length. The tips of the horizontal bars rest at the springs, which exhibits a linear elastic
response. The vertical rigid rod is loaded at its top by a force P that may be either conservative
(dead) or non-conservative (follower) and have both static and periodic components. The two
reaction forces F1 and F2 from the springs are always directed in the vertical direction. The
root of the tee is constrained to move in a vertical line, i.e. parallel to the reaction forces.
Let this displacement u be measured from some arbitrary datum. The configuration of the
deflected system is thus completely specified when the two independent co-ordinates θ and
u are known. Specifically, the displacements of the ends of the arms will be 11,2 = u ±
c sin θ.
314 S.V. Sorokin et al.

Exact expressions for the potential energy, kinetic energy, work done by the driving force,
Lagrange function and Rayleigh dissipation function of the model system (see Figure 1) are
V = 12 E[(u − c sin θ)2 + (u + c sin θ)2 ],
K = 12 m[(u̇ + l θ̇ sin θ)2 + l 2 θ̇ 2 cos2 θ] + 13 m1 c3 θ̇ 2 + m1 u̇2 c,
A = R1 [u + l(1 − cos θ)] + R2 l sin θ,
(1)
L =K −V
= 12 [m(u̇ + l θ̇ sin θ)2 +ml 2 θ̇ 2 cos2 θ − 2Eu2 − 2Ec2 sin2 θ] + 13 m1 c3 θ̇ 2 +m1 u̇2 c,
R = βu u̇2 + βθ θ̇ 2
where
R1 = (1 − α)P cos θ + αP , R2 = (α − 1)P sin θ.
Then Lagrange equations of motion of the model system can now be written:
Mu ü + ml θ̈ sin θ + ml θ̇ 2 cos θ + 2Eu + 2βu u̇ = (1 − α)P cos θ + αP ,
(2)
ml ü sin θ + Mθ θ̈ + Ec2 sin 2θ + 2βθ θ̇ = αP l sin θ,
where
Mu = 2m1 c + m, Mθ = 23 m1 c3 + ml 2 ,
and a dot denotes the time derivatives d/dt.
In Equations (2) α = 1 corresponds to dead loading and α = 0 to the follower force. In the
case of dead force (α = 1) the right-hand side of (2) is easily obtained when the work done by
dead force is added with an opposite sign to potential energy of the system. In the case of the
follower force (α = 0) the right-hand side of (2) is obtained by considering the equilibrium
conditions in the disturbed state of the system.
Other parameters in Equations (1) and (2) are: E, the stiffness of the spring element (it is
equal to Young’s modulus if the cross-section of the spring element has a unit area); u, θ, the
principal co-ordinates (see Figure 1); βu , βθ , the coefficients of viscous damping for vertical
and angular oscillations.
The linearized equations of motion in the absence of damping follow from (2)
Mu ü + 2Eu = P0 ,
(3)
Mθ θ̈ + (2Ec2 − αP0 l)θ = 0
from which the eigenfrequencies of the vertical and angular oscillations can be easily found
to be
s s
2E 2Ec2 − αP0 l
ω1 = , ω2 = , (4)
Mu Mθ
where P0 is the static component of the applied compressive force, P = P0 + P1 cos t. In
(3) the variable component is put to zero, P1 = 0.
As is well known [10–15], nonlinear modal interaction effects arise when eigenfrequencies
are related in a definite manner. It is clearly seen from (4) that the two eigenfrequencies can be
Nonlinear Dynamics and Stability 315

easily matched by varying P0 in the case of α = 1, i.e. when the applied force is conservative.
For a system under a follower force (α = 0) this is only possible when the geometry of the
system is modified, e.g. by varying the length c of the horizontal rod.
An examination of modal interaction effects is performed within the classical theory of
local nonlinear dynamics by using the method of multiple scales on the time variable [10, 11,
15]. This method allows the solution of Equations (2) to be written as function of independent
time-variables (scales). Thus, if t is written as T0 = t, T1 = εt, where ε serves only to indicate
the level of an approximation [15], then the time derivative becomes

d ∂ ∂
= +ε .
dt ∂T0 ∂T1

Respectively, an asymptotic solution can be expressed in the form

u(t) = u0 (T0 , T1 ) + εu1 (T0 , T1 ),


(5)
θ(t) = θ0 (T0 , T1 ) + εθ1 (T0 , T1 ).

The fast scale describes oscillations in ‘real time’, while the slow time T1 accounts for slow
modulations of amplitudes and phases.
The governing equations of motion (2) for use in the method of multiple scales have to be
rewritten in the polynomial form:

Mu ü + 2Eu = ε n P0 + ε k P1 cos t +
+ ε[−2βu u̇ − ml θ̈θ − ml θ̇ 2 − 12 θ 2 (1 − α)(ε n P0 + ε k P1 cos t)], (6)
Mθ θ̈ + 2Ec2 θ = α(ε n P0 + ε k P1 cos t)l(θ − 16 θ 3 ) + ε(−2βθ θ̇ − ml üθ + 43 θ 3 Ec2 ).

As it is adopted in standard analysis by the method of multiple scales, modal damping terms
as well as nonlinear components of restoring and inertial forces are assumed to be of order ε.
Three different approximation levels are explored for the static and the periodic components
of a driving force, so that in Equation (6) parameters n, k may be put equal to 0 or 1. The
above formulation with two time scales is applied to the study of the motions of the system at
external resonance excitation conditions.

3. Weak Excitation

Assume that the time-dependent component of driving force is harmonic with a frequency
close to the resonant frequency of the vertical motions ω1 . Then a weak excitation (i.e. rather
small amplitude of the periodic driving force) is sufficient to produce large amplitudes of
oscillation. However, two cases should be distinguished depending upon ranging of the static
component of a driving force. The applied compressive force may be presented either as n =
0, k = 1, i.e., P = P0 + εP1 cos t or as n = k = 1, i.e., P = εP0 + εP1 cos t. Modal
interaction in dynamics of the model system in both the two cases is considered in this section.
The case P = P0 + εP1 cos t is believed to be the most realistic one. It is relevant to a
comparatively heavy structure excited by a slightly unbalanced motor or by small periodic
316 S.V. Sorokin et al.

flow pulsation. The following system of amplitude modulation equations is obtained (see
Appendix I):
 
0 P1 b2 1−α
ω1 Mu a = −βu ω1 α − sin 91 − 2mlω2 −
2
P0 sin(91 + 92 ),
2 4 2
 
P1 b2 1−α
aω1 Mu 910 = − cos 91 − 2mlω22 − P0 cos(91 + 92 ) − aω1 Mu σ1 ,
2 4 2
mlω12 αP1 l
ω2 Mθ b0 = −βθ ω2 b + ab sin(91 + 92 ) + b sin 92 , (7)
4 4
mlω12 αlP0 b3
bω2 Mθ 920 = ab cos(91 + 92 ) + Ec2 b3 − +
2 8
αP1 l
+ b cos 92 + bMθ ω2 (σ1 + σ2 ).
2
Here, a prime denotes ∂( )/∂T1 . We have introduced a set of phase angles as: 91 = χ − σ1 T1 ,
92 = 2β + (σ1 + σ2 )T1 . The derivatives of original phase angles become

χ 0 = σ1 + 910 , β 0 = 12 (−920 + σ1 + σ2 ). (8)

Solutions of the system of amplitude modulation equations (Equation (7)) as well as their
stability analysis are presented in Appendix I.
An alternative formulation of a driving force is P = ε(P0 + P1 cos t). This ranging
is explored here to make sure that no specific effects are missed when the static compon-
ent gradually increases from zero to some moderate value that (in the case α = 1) is still
sufficiently far from a critical force. The corresponding amplitude modulation equations are
P1 mlω22 b2
ω1 Mu a 0 = −βu ω1 a − sin 91 − sin(91 + 92 ),
2 2
P1 mlω22 b2
aω1 Mu 910 = − cos 91 − cos(91 + 92 ) − aω1 Mu σ1 ,
2 2
mlω12 αP1 l
ω2 Mθ b0 = −βθ ω2 b + ab sin(91 + 92 ) + b sin 92 , (9)
4 4
mlω12
bω2 Mθ 920 = ab cos(91 + 92 ) + Ec2 b3 + αlP0 b+
2
αP1 l
+ b cos 92 b + bMθ ω2 (σ1 + σ2 ),
2
where the phase angles are introduced in the same way as in Equation (7). The modification
in ranging of static component of a driving force results in apparent change in this system
as compared with the system of Equations (7) – all three terms containing static force in
Equations (7) are replaced by the single term αlP0 b in the fourth of Equations (9). Solutions of
the system of amplitude modulation equations (Equation (9)) are presented and their stability
analysis is also performed in Appendix I.
Let us compare nonlinear local dynamics of the model system in loading cases of (i) P =
P0 + εP1 cos t and (ii) P = εP0 + εP1 cos t with results of direct numerical integration
of original equations (Equation (2)). Amplitude response curves at loading case (i) are shown
in Figure 2 for perfectly tuned external and internal resonance (σ1 = σ2 = 0) at α = 0 (i.e.
for a follower force) and the parameters selected as: m = 1 Kg, m1 = 1 Kg m−1 , c = 1 m, l =
Nonlinear Dynamics and Stability 317

Figure 2. Forced response curves for weak excitation at parametric resonance. Solid curves case (i): (3) and (2)
are analytical solutions, case (ii): (1) and (5) are numerical solutions (amplitude a and b, respectively), dashed line
(4) – amplitude a.

3.366 m, βu = 0.1 Kg s−1 , βθ = 0.1 Kg m2 s−1 , E = 2000 Kg s−2 , P0 = 1.0 N. Solid lines
present the (dimensional) amplitude of vertical motion a (curve 3) and of angular amplitude
b (curve 2) as a function of the amplitude of excitation force in the stable bi-modal regime of
motions, and the dashed line 4 represents the unstable regime of purely vertical motions of the
system. The curves 1 and 5 are plotted after numerical integration of Equations (2). There is a
very good agreement between asymptotic predictions and numerical results. These amplitude
response curves are rather typical for 2:1 modal interaction in many other mechanical systems
(see [9–14]). The key feature of such a behaviour is the saturation of vertical motions at a
certain level of excitation, accompanied by the development of angular motions.
Similar calculations are repeated for a case of dead force acting at the model system, i.e.,
for a case when α = 0 is replaced by α = 1 with all other parameters unchanged. As it
follows from Equations (7) in this case forcing terms appear in equations of angular motions.
However, solution of amplitude modulation equations (Equation (7)) is not much affected by
these modifications, and there is less than 0.1% difference in amplitudes of both the vertical
and angular amplitudes in stable bi-modal motions of the system loaded by a dead force and
a follower force. This may be explained by large stiffness of springs so that the second term
Ec2 b3 in the fourth of Equations (7) in fact controls dynamics of the system.
Now we address to the case (ii), when static component of the force is of the same order
as its periodic component, namely, we replace P0 = 1.0 N by P0 = 0.01 N and retain all
other parameters unchanged. Then solution of amplitude modulation equations (Equation (9))
does not differ from the solution of Equations (7) displayed in Figure 2 as well as results
of numerical integration of Equations (2). A transition from dead loading to follower loading
also does not change stable solution of these equations. Thus, for the selected set of parameters
both asymptotic expansions predict the same stable bi-modal regime of motion for cases (i)
and (ii) that is not affected by the nature of applied driving force. The fourth of Equations
(7) contains a term proportional to the third power of angular amplitude 18 αlP0 b3 , while the
fourth of Equations (9) contains αlP0 b. Both these equations have the term Ec2 b3 . Apparently,
318 S.V. Sorokin et al.

the difference in solutions of Equations (7) and (9) should be clearly seen at small angular
amplitudes, when in Equations (9) the term proportional to the first power of b may be of
the same order as the one proportional to the third power of b, multiplied by a large stiffness
coefficient. However, in such a case a bi-modal regime of motion is unstable or this difference
manifests itself in a very narrow range of parameters (see Figure 2) just around bifurcation
point.
Possibly, there exist some combinations of parameters of the model which make the dif-
ference in local dynamics more pronounced, but this aspect is beyond the scope of the present
paper and, in fact, does not appear interesting. We note that ranging of static force of order
O(1) adopted hereafter is consistent as no instability effects occur at lower levels of its ap-
proximation. In principle, the role of static component may be explored further, but this is also
not within the framework of the present analysis.

4. Parametric Hard Excitation by a Follower Force

4.1. L OCAL DYNAMICS

The assumption of weak excitation adopted in the previous section permitted modal inter-
action analysis and resulted in the prediction of stable bi-modal regime of motion for the
model system in certain ranges of parameters. However, from the practical standpoint, it is
not quite clear how small the actual magnitude of the driving force ought to be in order for
these predictions to be valid. This question will be clarified below in Section 4.2, while here
the amplitude of driving force is treated as arbitrary (so-called hard excitation):
P = P0 + P1 cos t. (10)
The case of dead force (α = 1) is then related to the well-known Mathieu equation for angular
motions [22], and will not be considered further.
Unlike the case considered in Section 3, it is now assumed that the driving frequency is
sufficiently far from the eigenfrequency ω1 of vertical motions. Under the above assumptions
regarding , the secular terms do not show up in the equation describing the vertical motions
(see Appendix II), so that we use only the linear forced response
P0 P1 cos t
u0 = + . (11)
2E 2E − Mu 2
Thus, the further ε 1 -order analysis does not concern modal interaction effects and it is quite
close to parametric excitation of a single-degree-of-freedom model. Nevertheless, it is relevant
to explore this subject in some details here in view of numerical analysis of large-amplitude
motions of the elastic and the elasto-plastic system in similar excitation conditions performed
in Sections 4.2 and 5. Solutions of the amplitude modulation equations and their stability
analysis are presented in Appendix II.
There are four non-dimensional parameters controlling the motions of the model system
under hard excitation conditions:
P1 ml ¯ =  , k = βθ , ω̄ = ω2 .
p̄1 = , 
EMθ ω2 Mθ ω2 ω1
To depict in a two-dimensional plot the results of stability analysis, two parameters should be
fixed, say, k = 3.6 × 10−4 and ω̄ = 0.82. While k is selected rather arbitrarily, the choice of ω̄
Nonlinear Dynamics and Stability 319

Figure 3. (a) Stability diagrams for k = 3.6 × 10−4 and ω̄ = 0.82; (b) frequency response curves for
k = 3.6 × 10−4 , p̄1 = 0.09 and ω̄ = 0.82.

is dictated by the desire to estimate the range of validity of the asymptotic solution obtained
above under the assumption that the driving frequency is sufficiently far removed from the
resonant frequency of vertical oscillations. The stability (instability) regions are shown in
¯ The regions of stability for b1 (instability for b2 ) coincide with
Figure 3 as a function p̄1 ().
their existence regions. The following zones of stability or instability are identified in Figure 3:
320 S.V. Sorokin et al.
¯
Zone 1 (between the -axis and the curve designated 1): in this zone the ‘zero’-solution is
stable (i.e. stable vertical motion);
Zone 2: In this zone both the ‘zero’-solution (purely vertical motion) and the solution b1
(vertical and angular motions) are stable, but the solution b2 is unstable;
Zone 3: In this zone the solution b1 (bi-modal regime of motion) is stable but the ‘zero’-
solution is unstable.

An alternative way to present results isqto plot frequency response curves. Figure 3(b)
¯ where b̄1,2 = b1,2 Ec2 /Mθ ω22 , for k = 3.6 × 10−4 , p̄1 = 0.09 and
shows the curves b̄1,2 (),
ω̄ = 0.82. In the vicinity of parametric resonance, the solutions b1 and b2 are characterised
by a soft nonlinearity as in a broad class of mechanical systems. However, as  ¯ approaches
1/0.82 = 1.23, the curve b1 tends to infinity, while the solution b2 vanishes at  ¯ = 1.31.
The rapid growth of b1 in the vicinity of  ¯ = 1.23 (¯ ≈ ω1 ) indicates that the assumptions
made in constructing the asymptotic solutions are violated. To find an asymptotic solution in
this frequency band, it is, of course, necessary to formulate a problem of near-resonant weak
excitation of vertical motions and then to look at the modal interaction effects, as was done in
Section 3. In the frequency range 1.23 <  ¯ < 1.31 the solution b1 is stable, the ‘zero’ solution
is unstable and the solution b2 does not exist. In the range 1.31 <  ¯ < 1.94 the solutions
b = 0 and b1 are stable, while the solution b2 is unstable. In the range 1.94 <  ¯ < 2.04 the
solution b1 is stable, the solution b = 0 is unstable and the solution b2 does not exist. Finally,
for ¯ > 2.04, the only existing solution (namely the ‘zero’-solution) is stable.
To facilitate the use of solutions in the broad frequency band 0 <  ¯ < 2 it is important to
1 ¯
ensure that ω̄ < 2 , i.e.  is sufficiently lower than ω1 . The regions of stability/instability and
typical frequency response plots are shown in Figure 4(a,b), respectively, for the following
parameters k = 0.036, ω̄ = 0.3, p̄1 = 0.07. The frequency plots are typical for a ‘standard
softening’ nonlinearity discussed by Nayfeh and Mook [11] and Thomsen [15]. The designa-
tions of stability regions in Figure 4(a) are the same as in Figure 3(a). In the frequency range
0< ¯ < 0.56, the only existing solution is b = 0. In the range 0.56 <  ¯ < 1.93 the solutions
b = 0 and b1 are stable, while the solution b2 is unstable. In the range 1.93 <  ¯ < 2.08
the solution b1 is stable, the solution b = 0 is unstable and the solution b2 does not exist.
Finally, for ¯ > 2.08, the only existing solution (b = 0) is stable. The damping parameter
k in this example is two orders of magnitude larger than in the case relevant to Figure 3, so
¯
that the zone in which ‘zero-solution’ is stable broadens near the -axis. Thus, the damping
of angular motions produces a stabilising effect, as it was also observed by Bolotin [23] and
appears quite obvious. It should be noted that the case represented by Figure 4 is rather typical
for the order one-half subharmonic resonance in even single-degree-of-freedom systems [13,
14] as no modal interaction is involved at all, unlike the case considered in Section 3.
The obtained results are of an asymptotic nature because the solutions of the amplitude
modulation equations represent only principal part of the whole solution of Equations (A.I.2)
in the case of weak excitation and of Equation (A.II.2) in the case of hard excitation. It is
therefore important to verify the predictions of local dynamics asymptotic analysis by direct
integration of the exact equations of motion (2). This comparison is also important for a proper
choice of numerical integration algorithm for the case of the elasto-plastic constitutive law to
be studied here.
In Figure 5(a,b) curves 1 show amplitudes of vertical (a) and angular (b) motions cal-
culated by the fourth order Runge–Kutta formula. They refer to perfectly tuned external
Nonlinear Dynamics and Stability 321

Figure 4. (a) Stability diagrams for k = 0.036, ω̄ = 0.3; (b) frequency response curves for k = 0.036,
ω̄ = 0.3, p̄1 = 0.07.

resonance ( = 2ω2 ) under hard excitation (P = P0 + P1 cos t) conditions with the
following dimensionless parameters: k = 3.6 × 10−4 , ω̄ = 0.82. The dimensional para-
meters are:  = 58.554 s−1 , m = 1 kg, m1 = 1 kg/m, l = 2 m, c = 1 m, E = 2000 kg s2 ,
βu = 10.0 kg s−1 , βθ = 0.05 kg m 2 s−1 , P0 = 1 N. Curve 2 gives the periodic part of amplitude
a from formula (11) and the amplitude b found from the modulation equation (Equation
(A.II.4)). The difference between the amplitudes of vertical motion a given by curves 1 and
2 is not very large because formula (11) displays the whole solution of Equations (A.II.1) of
322 S.V. Sorokin et al.

Figure 5. Validation of asymptotic analysis: (a) amplitudes of vertical oscillations versus driving force, (b) amp-
litudes of angular oscillations versus driving force; curves 1 are analytical solutions, curves 2 are numerical
solutions.

order ε 0 . This difference is somewhat larger for the amplitudes of angular motion b because
the asymptotic formula for this solution does not represent the whole solution of order ε 1 , for
which an improved ε 1 -order would be needed, but there is a very good agreement between
the asymptotic and numerical solutions at the threshold value of P1 relevant to the onset of
angular motions.
Nonlinear Dynamics and Stability 323

4.2. OVERTURNING P HENOMENON

In the local dynamics reported in Section 4.1, a stable bi-modal regime of motions is predicted
for arbitrarily large amplitudes of the follower force (see Figures 3(a) and 4(a)). However, the
axial force (conservative or non-conservative) may be able to overturn the model system (i.e.
to flip it from the upright to the inverted position), see Figure 1. This cannot be attributed
to the effects of local dynamics. Hence, there is an upper limit to the validity of asymptotic
results obtained from the analysis of the local nonlinear dynamics by the method of multiple
scales. In the case of a conservative applied force this upper limit may be roughly estimated
by its static critical value, determined by the ordinary divergence condition
s
2Ec2 − αP0 l 2Ec2
ω2 = = 0, so that Pcr = .
Mθ l

As the static component P of the conservative force reaches this value, the system will flip
from its upright to the inverted position.
If the compressive force is of the follower type (α = 0), the analysis of linearized Equa-
tions (3) does not predict a critical value at which the system loses stability by divergence
or flutter. However, this does not mean that the upright position of the model system loaded
by the follower force is always dynamically stable. An examination of the excitation of the
system with a driving frequency close to the parametric resonance shows that the overturning
phenomenon is still possible. A stability check is performed by the direct integration of the
exact equations of motion (2) with initial conditions relevant to a small deviation of the model
system from its straight upright position, say θ0 = 10−3 . If this disturbance develops in
time and results in overturning the system, then the upright position is unstable. If not, the
amplitude of the driving force is incremented and the procedure repeated. The results of this
stability check are summarised in Figure 6 (k = 0 – solid curve and k = 3.6 × 10−4 – dashed
curve), which shows approximately the boundary of regions of stable solutions in the manner
that is customary to the depiction of the stable solutions of the Mathieu equation in a Ince-
Strutt diagram. Points below the curve correspond to the combinations of driving frequency
¯ = /ω2 and amplitude p̄ = p1 /Ec that do not overturn the system. As is seen from
Figure 6, overturning occurs in a narrow region of parameters in the vicinity of the parametric
resonance at large amplitudes of force compared to those used in the local dynamics analysis
(this follows from a comparison of the force scales in Figures 3 and 6) and viscous damping
produces stabilising effect. It may thus be concluded that the predictions of local dynamics
for this kind of loading are applicable to a rather broad range of amplitudes of the follower
force, whereas the overturning of the system is essentially a global dynamics phenomenon.

5. Dynamic Stability of the Elasto-Plastic Model System

In the previous sections, the simple model of linear elastic spring elements has been explored.
In particular, the modal interaction effects generated by the nonlinear geometry were revealed
under typical excitation conditions. A comparison of the asymptotic and numerical results
gave confidence in the choice of Runge–Kutta algorithm for numerical integration. This al-
gorithm will be used in the present section to study the global nonlinear dynamics of the model
system, taking into account plastic deformations in springs.
324 S.V. Sorokin et al.

Figure 6. Stability diagram for global nonlinear dynamics at parametric resonance: solid curve for k = 0 and
dashed curve for k = 3.6 × 10−4 .

Figure 7. Constitutive law for spring elements (a) as used in the numerical algorithm, (b) as used to explain ‘static
force-displacement’ curves.

A model of linear strain-hardening, including the Bauschinger effect (which takes place if
the amplitude of a driving force exceeds the threshold value FY ), is adopted (Figure 7(a)) for
each spring. The equations of motion (2) are re-written as

Mu ü + ml θ̈ sin θ + ml θ̇ 2 cos θ + F1 + F2 + 2βã u̇ = (1 − α)P cos θ + αP ,


ml ü sin θ + Mθ θ̈ + (F2 − F1 )c cos θ + 2βθ θ̇ = αP l sin θ,
Nonlinear Dynamics and Stability 325

where the forces F1 and F2 acting at the springs are related to their displacements 11,2 =
u ± c sin θ (signs plus and minus are chosen for the right and left spring, respectively) in a
rather complicated way, as described in Appendix III.

5.1. A NALYSIS OF S TATIC OVERTURNING

We begin with the numerical integration of Equation (12) for the case of static (dead or fol-
lower) load. Integration is performed in the time domain, with a stepwise increase in the static
force. At regular time intervals small angular perturbations are introduced and the motions
studied without incrementing the force. In a stable state, these disturbances die away due to
damping, but in an unstable state they result in the overturning of the system. The following
non-dimensional parameters are introduced:
P FY u l Ek βθ
P̄ = , f¯ = , ū = , l¯ = , Ē = and β̄θ = .
Ec Ec c c E ω2 Mθ
There is a principal difference in the behaviour of the system loaded by a dead or follower
static force. In the former case there is destabilisation (overturning) of the divergence type at
the critical static dead force. In the latter case, neither divergence nor flutter occurs, so that
the system does not overturn, regardless of the value of the follower force. This difference
is clearly observed in the static ‘force-displacement’ curves (Figure 8(a,b)). The relationship
between dead force and static displacement can be of three types. The first one is relevant
to overturning in the elastic state with no plastic deformation of the springs involved in
the transition from the upright to the inverted equilibrium position. In this case there is no
difference in the behaviour under the dead and follower forces (to say nothing of the change
of configuration of the system loaded by dead force at p̄st = p̄cr that occurs with no increment
in the vertical displacement). This case is not of any particular interest and will not be further
pursued here.
The second possibility is destabilisation of the system loaded by dead force of a level
such that both springs are in the plastic regime. Figure 8(a) shows the variation of the static
displacement ū (curve 1, the displacement under dead force; curve 2, displacements under
follower force) with the static force P̄st , for f¯ = 0.15, Ē = 0.5, l¯ = 5.0. Here f¯ is the
non-dimensional force relevant to the springs being in the plastic range (Figure 7(a)). As no
overturning occurs under a static follower force, the discussion below is confined to loading
by a dead force. The discontinuity in the derivative at p̄st = 0.30 (Figure 8(a)) corresponds to
the occurrence of plastic deformations in springs resulting from a change in the modules from
E to Ek (Figure 7(a)). The two curves diverge at p̄st ≈ 0.35 (point O in Figure 8(a)) and do
not merge again in the range 0.35 < P̄st < 0.73. Destabilisation occurs at p̄st ≈ 0.35 (i.e. at
the critical force for elasto-plastic buckling), so that the system changes its configuration from
θ = 0 to θ = π (vertical part of curve 1). The post-critical equilibrium is controlled by E, as
in the elastic case. This effect is also illustrated on Figure 7(b). At the pre-critical equilibrium
state, generic points of both springs are at position B. During the destabilisation process, one
of the springs exhibits a closed cycle of unloading and reloading thereafter, while the other
experiences overloading and unloading afterwards. In the final, overturned equilibrium state,
the forces and displacements acting at the springs are the same (see point C). If loading is
continued, then at a certain level of the force the loading paths for both the dead and the
follower forces merge at the point D. If the system is loaded by a follower force, then both the
generic points simply follow the path OABD. The merging point D is also found in Figure 8(a)
and occurs at p̄st ≈ 0.73.
326 S.V. Sorokin et al.

Figure 8. Axial displacement u versus the static force p̄st for (a) p̄cr = 0.40, f¯ = 0.15, Ē = 0.5, l¯ = 5.0; (b)
p̄cr = 0.22, f¯ = 0.15, Ē = 0.5, l¯ = 9.0.

The third possibility is relevant to the case when the destabilisation of the system begins in
the elastic zone of the springs (point O in Figure 8(b)) but during the process of destabilisation
the springs experience plastic deformations (Figure 8(b)). The only qualitative difference from
the second case discussed above is that the loading paths under the dead and follower forces
do not coincide already in the elastic (for the follower force) regime. The curves depicted in
Figure 8(b) were obtained for the following set of parameters f¯ = 0.15, Ē = 0.5, l¯ = 9.0.
Figure 9 shows the variation of the critical dead force p̄cr with l, ¯ for f¯ = 0.15. Curve
1 gives the critical force found from the analysis of the linear elastic problem, p̄cr = 2/l, ¯
Nonlinear Dynamics and Stability 327

Figure 9. Critical value of the dead force versus l¯ for f¯ = 0.15 and Ē = 1.0 (curve 1), Ē = 0.5 (curve 2),
Ē = 0.1 (curve 3).

whereas curves 2 and 3 give the critical forces for elasto-plastic destabilisation when Ē =
0.5 and 0.1, respectively. Elasto-plastic destabilisation is identified by the overturning phe-
nomenon detected during the direct numerical integration of Equation (26), for there is no
analytical solution for elasto-plastic stability problem. It is clearly seen that the plastic de-
formation of springs significantly decreases the critical force, compared with the predictions
from linear elasticity. This statement is also confirmed by Figure 10, which shows the variation
of the critical force p̄cr with Ē, for f¯ = 0.15, l¯ = 5.0. Curve 1 shows the critical force
found from numerical integration, whereas the horizontal line 2 corresponds to p̄cr = 2/l, ¯
i.e., the non-dimensional critical force in linear elastic approximation that does not depend
upon E. Judging from Figure 10, it may be concluded that the critical force p̄cr in the elasto-
plastic zone is most sensitive to variations in Ē in the neighbourhood of 1. The relatively low
sensitivity at small values of Ē is explained by the strong damping of the overturn-induced
vibrations when large plastic deformations have developed in the springs. Note that curve 1
does not merge smoothly into curve 2 at Ē = 1, because the numerical algorithm is based on
a discontinuity in the slope from E to Ek (Figure 7(a)).

5.2. DYNAMIC S TABILITY OF THE S YSTEM L OADED BY S TATIC AND P ERIODIC F ORCES

Finally, let us consider the dynamics of the elasto-plastic system loaded by the force p̄ =
p̄st + p̄1 cos t. As before, two loading types should be distinguished – loading by dead force
and loading by follower force. In the former case, there is a strong influence of p̄st that has
already been detected in the linear analysis, as well as in the analysis of static overturning
phenomenon. The periodic component p̄1 of the dead force stimulates overturning, but the
angular motions of the overturned system die away due to viscous damping and/or damping
328 S.V. Sorokin et al.

Figure 10. Critical value of the dead force versus Ē for f¯ = 0.15, l¯ = 5.0, curve 1, critical force found from
numerical integration; curve 2, critical force found from linear elastic approximation.

by plastic deformations of the springs. Here, the attention is focused on the more interesting
case of the follower force. It will be recalled that under a follower force an elastic system
overturns in a rather narrow band of frequencies and amplitudes of excitation (see Figure 6)
and that no elastic-plastic overturning occurs when a force is static (i.e. p̄1 = 0).
We begin with the identification of conditions for the overturning of a perfectly tuned
system (for the set of parameters given in Section 4, which is relevant to external parametric
resonance). Figure 11 shows combinations of values of β̄θ (viscous damping for angular oscil-
lations) and Ē below which there is the dynamic destabilisation (overturning) of the system for
three levels of excitation: p̄1 = 0.5 (curve 1), p̄1 = 0.85 (curve 2) and p̄1 = 2.0 (curve 3). The
static component is selected as p̄st = 0.85, whereas the threshold (yield) value is f¯ = 0.8 (see
Figure 7(a)). There is a qualitative difference between the cases 1, 2 and the case 3. When p̄1
is equal to 0.5 or 0.85, the spring response does not jump from line 1 to line 3, and vice versa,
at each cycle of loading, since p̄1 < 2f¯. By contrast, when p̄1 = 2.0, the springs experience
elasto-plastic deformation at each loading cycle. As seen from Figure 11, for small values of
Ē the system does not overturn due to the very strong influence of static plastic deformation
generated by p̄st . When p̄1 = 2.0 there is a closed cycle of plastic deformation at each loading
cycle which produces the same effect as does viscous damping. Nevertheless, as the ‘plastic
damping’ is generated by the large amplitude periodic component p̄1 , to prevent the system
from overturning at p̄1 = 2.0 it is necessary to have more viscous damping than in the case of
p̄1 = 0.85. The ‘plastic damping’ alone at p̄1 = 2.0 cannot stabilise the system, as it does at
p̄1 = 0.5 or p̄1 = 0.85.
The stability and instability (overturning) regions in Figure 12 are separated by the curve
plotted in co-ordinates (p̄st , p̄1 ) for f¯ = 0.8, Ē = 0.5 and βθ = 0. This figure illustrates
the role of static force p̄st in the global stability of the model system. As the static force p̄st
Nonlinear Dynamics and Stability 329

Figure 11. Combinations of β̄θ and Ē resulting in dynamic destabilisation (overturning) of the system for
p̄1 = 0.5 (curve 1), p̄1 = 0.85 (curve 2) and p̄1 = 2.0 (curve 3).

Figure 12. Stability (below the curve) and instability (overturning) regions (above the curve) for p̄0 = 0.8,
Ē = 0.5 and β̄θ = 0.
330 S.V. Sorokin et al.

Figure 13. ‘Equivalent’ viscous damping coefficients β̄θ versus the excitation amplitude p̄1 for elastic system
(Ē = 1) and two elasto-plastic systems (curve 1: Ē = 0.5 and curve 2: Ē = 0.2).

grows, the periodic component p̄1 that causes overturning also increases linearly for small
and moderate values of p̄st . The rate of growth of p̄1 , however, decreases when p̄st becomes
sufficiently large.
Figure 13 gives the values of ‘equivalent’ viscous damping coefficients β̄θ found for an
ideally elastic ‘equivalent’ system (having Ē = 1) as a function of the excitation amplitude
p̄1 (f¯ = 0.8, p̄st = 0.85). The approximate equivalence between the elasto-plastic and ideally
elastic systems is established in the simplest way by demanding that the amplitudes of their
stationary angular oscillations be approximately the same. The other conditions necessary for
the equivalence concern amplitude of excitation force: it should be the same for the elastic
and elasto-plastic systems and should correspond to ‘pre-overturning’ motion. Curves 1 and
2 give β̄θ for an ideally elastic system which is equivalent in the above sense to the elasto-
plastic system with Ē = 0.5 and Ē = 0.2, respectively. The difference in scales of damping
in Figures 11 and 13 can be used to estimate the contribution of plastic damping. Curves 1
and 2 in Figure 13 are very close to each other, meaning that the stationary amplitudes of
angular oscillations are not much dependent on Ē; they are independent of β̄θ but depend on
the excitation amplitude p̄1 . For large values of p̄1 the two curves diverge rapidly, since the
plastic effects are significantly developed, but the oscillating system exhibits closed cycles of
loading and unloading.

6. Conclusions

An investigation has been made of the nonlinear dynamics of Sewell’s two-degree-of-freedom


model system for an elasto-plastic column subjected to a dead or follower force having both
static and sinusoidal components.
Nonlinear Dynamics and Stability 331

The analysis of an ideally elastic system revealed that the static dead force determines the
eigenfrequency of angular oscillations and that destabilisation of divergence type occurs as the
dead force reaches its critical value. By contrast, a static follower force has no influence on
the eigenfrequencies and does not destabilise the system. The analysis of local dynamics was
performed for two typical excitation conditions: (i) weak excitation at parametric resonance
 = ω1 , ω1 = 2ω2 ; and (ii) parametric hard excitation. In the former case, the analytical
solution for the bi-modal regime of motion was obtained and typical modal interaction effects
were discovered. In the latter case, the existence and stability of asymptotic solutions were
studied and a ‘soft’ nonlinearity of the forced response was detected. The static component of
the follower force was found to have no influence upon the local dynamics under conditions of
hard parametric excitation. Special attention was paid to the verification of asymptotic results
by the direct numerical integration of the exact equations of motion. The numerical integration
of the equations of global nonlinear dynamics of an ideally elastic system loaded by a follower
force was performed in the vicinity of parametric resonance. A stability diagram was plotted
to identify the combinations of frequencies and amplitudes of the force that cause the model
system to lose stability by overturning.
For the elasto-plastic column numerical integration was performed in time domain. Con-
ditions of static stability were analysed first, and it was found that the critical (overturning)
value of the dead force is strongly influenced by parameter Ē and that the plastic deform-
ations markedly stimulate destabilisation. Standard ‘static force-displacement’ curves were
compared for loading by a dead force and a follower one. The curves were found to differ
for a certain range of forces. The equations of motion of the elasto-plastic system under
a follower force containing both static and sinusoidal components were integrated. It was
found that the hardening modulus Ek could completely prevent from dynamic overturning
of the system at small amplitudes of the sinusoidal component p1 . The static component of
the follower force has a strong stabilising influence upon the system. Finally, an ‘equiva-
lent’ viscous damping of an ideally linear system was calculated for the system with ‘plastic
damping’, in order to investigate a non-conservative system with an ‘equivalent’ viscous
damping.

Appendix I. Analysis of Weak Excitation

1. The case P = P0 + εP1 cos t (small harmonic forcing)

The original equations (Equation (6)) are conveniently split into two groups:

Order ε 0

∂ 2 u0
Mu + 2Eu0 = P0 ,
∂ 2 T02
(A.I.1)
∂ 2 θ0
Mθ 2 2 + (2Ec2 − αP0 l)θ0 = 0.
∂ T0
332 S.V. Sorokin et al.

Order ε 1
"  #
∂ 2 u1 ∂u0 ∂ 2 θ0 ∂θ0 2 ∂ 2 u0
Mu 2 2 + 2Eu1 = −2βu − ml θ 0 + − 2M u +
∂ T0 ∂T0 ∂T02 ∂T0 ∂T0 ∂T1
θ2
+ P1 cos T0 − P0 (1 − α) 0 ,
2
  (A.I.2)
∂ 2 θ1 ∂θ 0 ∂ 2
u0 4 2 αP0 l
Mθ 2 2 + (2Ec − αP0 l)θ1 = −2βθ
2
− ml θ0 + Ec − θ0 −
3
∂ T0 ∂T0 ∂T02 3 6
∂ 2 θ0
− 2Mθ + αP1 lθ0 cos T0 .
∂T0 ∂T1
The solution of Equation (A.I.1) is sought in the form
P0
u0 (T0 , T1 ) = + A(T1 )eiω1 T0 + Ā(T1 )e−iω1 T0 ,
2E (A.I.3)
θ0 (T0 , T1 ) = B(T1 )eiω2 T0 + B̄(T1 )e−iω2 T0 .
Here A and B are slowly-varying complex amplitudes of vertical and angular motions, and
Ā, B̄ are their complex conjugates. The static component of load generates a constant com-
ponent of axial displacement P0 /2E and in the case of loading by a dead force also produces
a change in eigenfrequency of angular motion. In addition to the external resonant excitation
 = ω1 + εσ1 , (A.I.4)
there is an internal parametric resonance to take into account
ω1 = 2ω2 + εσ2 , (A.I.5)
where σ1 , σ2 are detuning parameters, indicating how close the frequencies are to each other.
To obtain a uniformly valid asymptotic solution of Equations (A.I.2), the secular terms in
the fast time T0 must be eliminated from the right-hand side. Complex functions in the slow
time are sought in the form
A(T1 ) = 12 a(T1 ) exp(iχ(T1 )), B(T1 ) = 12 b(T1 ) exp(iβ(T1 )),
where a(T1 ) and b(T1 ) are their real-valued amplitudes and χ(T1 ), β(T1 ) their real-valued
phases.
Using (A.I.3–A.I.5), eliminating secular terms, and separating the real and imaginary parts
give the modulation (Equation (7)). Steady-state solution is sought by letting a 0 = b0 = aψ10 =
bψ20 = 0. Then there are two possible solutions: either the purely vertical uni-modal (b = 0)
oscillations or bi-modal oscillations (b 6 = 0). The uni-modal regime, usually referred to as the
‘zero-solution’ is defined by the following amplitude and phase of vertical oscillations:
P1 βu
a= q , tan ψ1 = . (A.I.6)
2Mu ω1 σ12 + (βu /Mu )2 σ1 Mu

In the case of a perfectly tuned resonance (σ1 = σ2 = 0) this solution is reduced to


P1 π mω1
a= , ψ1 = , b = 0, ψ2 = arccos q , (A.I.7)
2ω1 βu 2 4α 2 βu2 + m2 ω12
Nonlinear Dynamics and Stability 333

and it is stable if
q
8βu βθ ω2 4α 2 βu2 + m2 ω12
P1 < . (A.I.8)
(4α 2 βu2 + m2 ω12 )l
The bi-modal regime, when both amplitudes are not equal to zero: a 6 = 0 and b 6 = 0 is more
complicated. It is not possible in general to present the solution in an explicit analytical form,
as was done above, irrespective of whether α = 1 or α = 0. To judge upon stability of
this regime of motion the Jacobian of the amplitude modulation equations is constructed and
its eigenvalues are found numerically. If the inequality (A.I.8) is reversed, bi-modal solution
becomes stable and purely vertical motion loses its stability.
2. The case P = εP0 + εP1 cos t (comparably small static and harmonic forcing)
The equations of order ε 0 become
∂ 2 u0
Mu + 2Eu0 = 0,
∂ 2 T02
(A.I.9)
∂ 2 θ0
Mθ 2 2 + 2Ec2 θ0 = 0.
∂ T0
This system does not contain any loading terms and its solution describes linear vertical and
angular motions of the unloaded model
u0 (T0 , T1 ) = A(T1 )eiω1 T0 + Ā(T1 )e−iω1 T0 ,
(A.I.10)
θ0 (T0 , T1 ) = B(T1 )eiω2 T0 + B̄(T1 )e−iω2 T0 .

Respectively, equations of order ε 1 become


"  #
∂ 2 u1 ∂u0 ∂ 2 θ0 ∂θ0 2 ∂ 2 u0
Mu 2 2 + 2Eu1 = −2βu − ml θ 0 + − 2Mu +
∂ T0 ∂T0 ∂T02 ∂T0 ∂T0 ∂T1
+ P0 + P1 cos T0 ,
(A.I.11)
∂ 2 θ1 ∂θ0 ∂ 2 u0 4
Mθ 2 2 + 2Ec2 θ1 = −2βθ − ml 2
θ0 + Ec2 θ03 −
∂ T0 ∂T0 ∂T0 3
∂ 2 θ0
− 2Mθ + αlθ0 (P0 + P1 cos T0 ).
∂T0 ∂T1
‘Zero-solution’ (b = 0) for the modulation equations (Equation (9)) of this case is given by
the expression (A.I.6), and phase angle of angular motions is found from the equation
mlω12 αP1 l cos ψ2
a cos(ψ1 + ψ2 ) + αlP0 + + Mθ ω2 (σ1 + σ2 ) = 0. (A.I.12)
2 2
For the perfectly tuned resonance the phase angle of angular motions is
 q 
−8α 2 βu P0 + mω1 4α 2 βu2 (−4P02 + P12 ) + m2 ω12 P12
ψ2 = arccos  . (A.I.13)
(4α 2 βu2 + m2 ω12 )P1
334 S.V. Sorokin et al.

Unlike the previous case, it is not possible to formulate stability condition in a closed analy-
tical form. Solution b = 0 becomes unstable when the following condition is held:
−4βθ ω2 + αlP1 sin ψ2 + amlω12 sin(ψ1 + ψ2 ) > 0. (A.I.14)
In this case, bi-modal regime becomes stable.

Appendix II. Analysis of Hard Excitation

If the force is non-conservative (α = 0), then the equations of motion of order ε 0 are
∂ 2 u0
Mu + 2Eu0 = P0 + P1 cos t,
∂T02
(A.II.1)
∂ 2 θ0
Mθ + (2Ec2 − αP0 l)θ0 = 0,
∂T02
and of order ε 1 are
"  #
∂ 2 u1 ∂u0 ∂ 2 θ0 ∂θ0 2
Mu + 2Eu1 = −2βu − ml θ0 + −
∂T02 ∂T0 ∂T02 ∂T0
∂ 2 u0 θ2
− 2Mu − P0 0 , (A.II.2)
∂T0 ∂T1 2
∂ 2 θ1 ∂θ0 ∂ 2 u0 4 2 3 ∂ 2 θ0
Mθ + 2Ec 2
θ 1 = −2βθ − ml θ 0 + Ec θ0 − 2Mθ .
∂T02 ∂T0 ∂T02 3 ∂T0 ∂T1
The solution is sought in the form
P0 P1 cos t
u0 = + + Aeiω1 t + Āe−iω1 t , θ0 = Beiω2 t + B̄e−iω2 t . (A.II.3)
2E 2E − Mu 2
The complex function describing the angular motions in the slow time scale is sought in the
form B(T1 ) = 12 b(T1 ) exp(iβ(T1 )). The amplitude modulation equation is
mlqω2 βθ
b0 = b sin ψ − b,
Mθ Mθ
(A.II.4)
1 0 mlqω2 1 Ec2 3 1
bψ = b cos ψ + b + σ b.
2 Mθ 2 Mθ ω2 2
Here
P1
q= , ψ = σ T1 − 2β. (A.II.5)
2E[1 − (/ω1 )2 ]
Equation (A.II.4) shows that the static component of the follower force has no influence upon
the local dynamics of the system in conditions of hard parametric excitation.
Steady-state solution is sought by letting b0 = bψ 0 = 0. They are
1. ‘Zero-solution’:
σ Mθ
cos ψ = − , b = 0. (A.II.6a)
2mlqω2
Nonlinear Dynamics and Stability 335

2. Non-zero solution:
v s
r u    
Mθ ω2 u
t mlqω2 2 βθ 2
b1,2 = ±2 − − σ,
Ec2 Mθ Mθ (A.II.6b)
2βθ ω2
tan ψ = − .
ω2 Mθ σ + Ec2 b2
It follows that the ‘zero’-solution is stable, if the excitation parameter satisfies the follow-
ing condition:
q
Mθ σ 2 /4 + βθ2 /Mθ2
|q| < , (A.II.7)
mlω2
otherwise it is unstable.
The nonzero solution b1 (positive sign in A.II.6b) is stable, if
q
Mθ σ 2 /4 + βθ2 /Mθ2 βθ
|q| > and σ > 0, or |q| > and σ < 0. (A.II.8)
mlω2 mlω2
The nonzero solution b2 (negative sign in A.II.6b) is unstable, if
q
θ σ /4 + βθ /Mθ
2 2 2
βθ M
< |q| < and σ < 0. (A.II.9)
2mlω2 mlω2

Appendix III. Governing Relations for Plastic System

If the generic point X(F, 1) lies on the line OA (Figure 7(a)), then the Hooke’s law is valid
F1,2 = E11,2 up to threshold (yield) value of Fy . If X lies between lines 1 and 2 (Figure 7(a)),
i.e., if Ek 11,2 +FY (1−Ek /E) > F1,2 > Ek 11,2 −FY (1−Ek /E), then the force acting in each
spring element is related to its displacement as follows: F2,1 = ξ1 [E11,2 + (Ek − E)η1,2 +
FY (1 − Ek /E)] + ξ2 [E11,2 + (Ek − E)ζ1,2 − FY (1 − Ek /E)]. Here, Ek is the ‘hardening
modulus’ of the spring element, ξ1 is equal to 1, if X moves from line 1 (line 3) to line 2 (or
any parallel to it); otherwise it is equal to zero. η1,2 (ζ1,2 ) are the abscissas of the jumps from
line 1 (line 3) to line 2, respectively. If X at the preceding step of integration in time lay on
the line OA or any line parallel to it and if the condition F1,2 < Ek 11,2 − F (1 − Ek /E) is
satisfied, then F1,2 = Ek 11,2 − F (1 − Ek /E) (generic point passes to line 3). If X was on the
line OA or any line parallel to it and if condition F1,2 > Ek 11,2 + FY (1 − Ek /E) is met, then
F1,2 = Ek 11,2 + FY (1 − Ek /E) (i.e. X passes to line 1).
If X(F, 1) is on the line 1 and 1̇1,2 > 0, then F2,1 = Ek 11,2 + FY (1 − Ek /E). Al-
ternatively, if the generic point X(F, 1) is on the line 1 and 1̇1,2 6 0, then it returns
to a line parallel to OA, so that η1,2 = 11,2 , ζ1,2 = 0, ξ1 = 1, ξ2 = 0, and F2,1 =
E11,2 + (Ek − E)ζ1,2 + FY (1 − Ek /E).
If the generic point X(F, 1) lies on the line 3 and 1̇1,2 6 0, then F2,1 = Ek 11,2 − FY (1 −
Ek /E). Alternatively, if X is on the line 3 and 1̇1,2 > 0, then it returns to a line parallel to OA,
so that η1,2 = 0, ζ1,2 = 11,2 , ξ1 = 0, ξ2 = 1, and F2,1 = E11,2 +(Ek −E)x1,2 −FY (1−Ek /E).
336 S.V. Sorokin et al.

As there is a discontinuity in the first derivative of the function F (1), an asymptotic


analysis cannot be performed on Equations (12).

References

1. Sewell, M.J., ‘The static perturbation technique in buckling problems’, J. Mech. Physics of Solids 1 (1965)
264–287.
2. Paidoussis, M.P., Fluid-structure Interaction. Slender Structures and Axial Flow, Vol. 1, Academic Press,
1998.
3. Benjamin, T.B., ‘Dynamics of a system of articulated pipes conveying fluid’, Proc. Roy. Soc. Lond., Ser. A
261 (1961) 457–499.
4. Bishop, R.E.D. and Fawzy, I., ‘Free and forced oscillation of a vertical tube containing a flowing fluid’, Phil.
Trans. Roy. Soc. Lond., Ser. A 284 (1976) 1–47.
5. Paidoussis, M.P. and Li, G.X., ‘Pipes conveying fluid: a model dynamical problem’, J. Fluids Struc. 7 (1993)
137–204.
6. Paidoussis, M.P., Luu, T.P. and Laithier, B.E., ‘Dynamics of finite-length tubular beams conveying fluid’, J.
Sound Vibration 106 (1986) 311–331.
7. Langthjem, M.A., Dynamics, Stability and Optimal Design of Structures with Fluid Interaction, Ph.D.
Thesis, Technical University of Denmark, 1996.
8. Panovko, Ya.G. and Sorokin, S.V., ‘On quasi-stability of viscoelastic systems with follower forces’,
Mekhanika tverdogo tela (Mechanics of Solids) 22(5) (1987) 87–96.
9. Thomsen, J.J.,‘Chaotic dynamics of the partially follower-loaded elastic double pendulum’, J. Sound
Vibration 188(3) (1995) 385–405.
10. Nayfeh, A.H., Perturbation Methods, Wiley, New York, 1973.
11. Nayfeh, A.H. and Mook, D.T., Nonlinear Oscillations, Wiley, New York, 1979.
12. Nayfeh, A.H. and Balachandran, B., ‘Modal interactions in dynamical and structural systems’, Appl. Mech.
Rev. 42 (1989) 175–202.
13. Nayfeh, A.H., ‘The response of single-degree-of-freedom systems with quadratic and cubic nonlinearities to
a subharmonic excitation’, J. Sound Vibration 89 (1983) 457–470.
14. Mook, D.T., Plaut, R.H. and Haquang, C., ‘The influence of an internal resonance on non-linear structural
vibrations under subharmonic excitation conditions’, J. Sound Vibration 102 (1985) 473–492.
15. Thomsen, J.J., Vibrations and Stability: Order and Chaos, McGraw-Hill, London, 1997.
16. Klyushnikov, V.D., Stability of Elasto-plastic Systems, Nauka, Moscow, 1980.
17. Shaw, S.W. and Holmes, P.J., ‘A periodically forced piecewise linear oscillator’, J. Sound Vibration 90 (1983)
129–155.
18. Narayanan, S. and Sekar, P., ‘Periodic and chaotic responses of an sdf system with piecewise linear stiffness
subjected to combined harmonic and flow induced excitations’, J. Sound Vibration 184 (1995) 281–298.
19. Pratap, R., Mukherjee, S. and Moon, F.C., ‘Dynamic behaviour of a bilinear hysteretic elasto-plastic
oscillator, Part I: Free oscillations’, J. Sound Vibration 172 (1994) 321–337.
20. Pratap, R., Mukherjee, S. and Moon, F.C., ‘Dynamic behaviour of a bilinear hysteretic elasto-plastic
oscillator, Part II: Oscillations under periodic impulse forcing’, J. Sound Vibration 172 (1994) 339–358.
21. Chatterjee, S., Mallik, A.K. and Ghosh, A., ‘Periodic response of piecewise nonlinear oscillators under
harmonic excitation’, J. Sound Vibration 191 (1996) 129–144.
22. Bolotin, V.V., Non-conservative Problems of the Theory of Elastic Stability, Pergamon Press, Oxford, 1963.
23. Bolotin, V.V., The Dynamic Stability of Elastic Systems, Holden Day, San Francisco, 1964.

Vous aimerez peut-être aussi