You are on page 1of 9

Nonlinear modal decomposition using normal form

transformations
Simon A. Neild

, Andrea Cammarano and David J. Wagg


Department of Mechanical Engineering
University of Bristol
Queens Building, University Walk,
Bristol, UK, BS8 1TR
ABSRACT
In this paper we discuss a technique for decomposing multi-degree-of-freedom weakly nonlinear
systems into a simpler form. This type of decomposition technique is an established cornerstone
of linear modal analysis. Extending this type of technique to nonlinear multi-degree-of-freedom
systems has been an important area of research in recent years. The key result in this work is that a
theoretical transformation process is used to reveal both the linear and nonlinear system resonances.
For each resonance, the parameters which characterise the backbone curves and higher harmonic
components of the response, can be obtained. The underlying mathematical technique is based on
a near identity normal form transformation for systems of equations written in second-order form.
This is a natural approach for structural dynamics where the governing equations of motion are
written in this form as standard practice. The example is a system with cubic nonlinearities, and
shows how the transformed equations can be used to obtain a time independent representation of
the system response. It is shown that when the natural frequencies are close to an integer multiple
of each other, the backbone curve bifurcates. Examples of the predicted responses are compared to
time-stepping simulations to demonstrate the accuracy of the technique.
Keywords: normal form; resonances; backbone curves; nonlinear; multi-degree-of-freedom
Introduction
The concept of normal modes of vibration is fundamental to the analysis of linear multi degree-of-
freedom systems, with each mode relating to a physical conguration of the system and a natural
frequency. For linear systems, the governing equations of motion can be decomposed into a modal
model by applying a linear modal transform. This results in a modal model of the system consisting
of a series of independent oscillators that capture the complete system behaviour via superposition
of the modal responses. The ability to create a modal modal from a structural model is key to the
experimental identication of dynamic properties known as modal testing [1] and model updating
[2]. While a linear model can capture the key characteristics of many systems, nonlinear behaviour is
encountered in a large number of applications, such as structures that exhibit large deections resulting
in geometric nonlinearities.
The idea of using a modal approach to analysing the behaviour of nonlinear systems has many
challenges. The concept of nonlinear normal modes was introduced in the 1960s by Rosenberg [3] and
many researchers have pursued this idea see for example [4, 5, 6] and references therein.
In this paper we consider a dierent approach, in that rather than looking to dene modes, we
attempt to decompose the nonlinear system into its simplest (aka normal) form. In the context of
multi degree-of-freedom nonlinear structural systems, simplest form means elimination of as many

Address all correspondence to this author simon.neild@bristol.ac.uk.


cross-coupling and nonlinear terms as possible. In fact the method we use will nd this simplied form
and also includes expressions for the modeshapes as a by product of the process as well.
The method of normal forms has a long history (see [7] for example), and involves applying trans-
formations to the governing equations of motion with the aim of nding a simplied form. Detailed
discussions on the theory of normal forms are given in [8, 9] among others. Of particular interest here
is the application of normal forms to nd periodic steady-state system response solutions, as consid-
ered by Jezequel and Lamarque [10] and Nayfeh [11]. Potential advantages of using normal forms over
other perturbation techniques for this type of problem include the ease in which it can be extended
to consider multi-degree-of-freedom systems and non-autonomous systems [12] and its suitability for
analysis using symbolic manipulation programs [13]. Relating to nonlinear normal modes, Jezequel
and Lamarque [10] considered a multi-mode system using normal forms. Touze et al. [14] linked the
near-identity transform, which provides an asymptotic non-linear change of coordinates for the system,
to the normal modes of the system.
The normal form transformation is normally applied to dynamic equations expressed in their state-
space, or rst-order dierential equation, form. Recently a method of applying the normal form tech-
nique directly to second-order nonlinear oscillators has been reported [15, 16]. As linear modal analysis
is based on second-order dierential equation representation of the equations of motion, this second-
order normal form technique has potential to provide a more natural nonlinear extension to the linear
problem. Here, we use the second-order normal form technique to nd the backbone curves for a two
degree-of-freedom system.
We present results for a two degree-of-freedom system with cubic spring nonlinearities. The back-
bone curves are found using the second-order normal form technique and it is shown that when the
natural frequencies are close to an integer multiple of each other, one of the backbone curves bifur-
cates. The curves are obtained by considering the initial conditions which results in one of four types of
modeshape for the system. Examples of the resulting motion from a selection of the initial conditions
are also shown. These predicted responses are compared to time-stepping simulations to demonstrate
the accuracy of the technique.
Modal decomposition for a nonlinear system
The normal forms technique as dened by [19] involves applying a series of transformations resulting
in an equation of motion using the following steps:
apply a linear transform to decouple the linear terms,
apply a forcing transformation,
apply a nonlinear near-identity transform.
The aim of the forcing and the near-identity transforms is to remove the non-resonant terms for each
mode. The non-resonant terms are dened as those in the equation of motion that result in harmonics
of the natural frequency (in the case of the unforced system) or of the dominant response frequency (in
the case of forced systems). The forcing transformation targets the non-resonant forcing terms, and
the near-identity transform removes the nonlinear non-resonant terms. Note that the response due to
the non-resonant terms is not lost, instead it is captured in the transform equations. Transforming
these non-resonant terms terms out of the equations of motion, for, say, the n
th
mode, allows the use
of a trial solution of the form U
n
cos(
rn
t
n
) to solve the equation exactly, thereby removing the
need for a harmonic balance type approximation.
Full details of the method can be found in [19, 16], here we consider a two mass oscillator with
three springs, one connecting each mass to ground and one between the masses. The springs connecting
the masses to ground have a force-deection relationship F = k +
3
and the spring between the
masses has the relationship F = k
2
+
2

3
, where is the length change of the spring. Linear
viscous dampers are also connected between the masses and ground (damping c) and between the
masses (damping c
2
). The conguration of the system is such that the resulting system has symmetric
mass, damping and stiness matrices. Single frequency forcing P
1
cos(t) and P
2
cos(t) is applied to
the two masses. The resulting equation of motion is
_
m 0
0 m
_ _
x
1
x
2
_
+
_
c +c
2
c
2
c
2
c +c
2
_ _
x
1
x
2
_
+ (1)
_
k +k
2
k
2
k
2
k +k
2
_ _
x
1
x
2
_
+
_
x
3
1
+
2
(x
1
x
2
)
3
x
3
2
+
2
(x
2
x
1
)
3
_
=
1
2
_
P
1
P
1
P
2
P
2
_
r
where r = {r
p
, r
m
}
T
= {e
it
, e
it
}
T
is used to represent the harmonic forcing.
Applying the linear modal transform x = q, where is a matrix of eigenvectors of M
1
K results
in
q + q +N
q
(q) = P
q
r, (2)
where,
=
_
1 1
1 1
_
, =
_
k/m 0
0 (k + 2k
2
)/m
_
,
N
q
(q) =

m
_
q
3
1
+ 3q
1
q
2
2
3q
2
1
q
2
+q
3
2
_
+
_
2
1

n1
0
0 2
2

n2
_
q,
P
q
=
1
4m
_
P
1
+P
2
P
1
+P
2
P
1
P
2
P
1
P
2
_
. (3)
Also = 1 + (8
2
/), the linearized natural frequencies
n1
= k/m and
n2
= (k + 2k
2
)/m and
damping relationships 2
1

n1
= c/m and 2
2

n2
= (c + 2c
2
)/m.
The forcing transform q = v+[e]r is now applied and considering each mode in turn, this transform
removes forcing terms that are away from resonance while retaining those that are near-resonant. In
this paper we will not consider forced equations of motion in detail, and as a result we write q = v
and P
q
= 0. For the case where q = v, the resulting equation of motion is
v + v +N
v
(v, v, r) = P
v
r (4)
where P
v
= P
q
= 0. The resulting nonlinear term, without damping, can written as
N
v
(v) = N
q
(v) =

m
_
v
3
1
+ 3v
1
v
2
2
3v
2
1
v
2
+v
3
2
_
(5)
Now the equation of motion is in a form in which the near-identity nonlinear transform can be
applied. This transform is written as
v = u +h(u, u, r) (6)
The aim of the transform is to simplify the dynamic equation, Eqn. (4), into a form that can be solved
exactly using a single frequency trial solution for each mode. The resulting transformed (i.e. simplied)
normal form equation is expressed as
u + u +N
u
(u, u, r) = P
u
r (7)
where currently N
u
, and the transform h, are the unknown.
To keep track of the relative sizes of terms we introduce , a small parameter, which can be viewed
as a bookkeeping aid. Using this bookkeeping notation, the nonlinear vectors and the transform vectors
are expressed as a power series in
N
v
(v, v, r) = n
v1
(v, v, r) +
2
n
v2
(v, v, r) +
N
u
(u, u, r) = n
u1
(u, u, r) +
2
n
u2
(u, u, r) + (8)
h(u, u, r) = h
1
(u, u, r) +
2
h
2
(u, u, r) +
Note that there are no
0
terms as N
v
, N
u
and h are all small compared to the linear terms.
Using the method described in [19, 16]
2
and higher terms are assumed to be zero, and the
coecients of the
1
terms can be written as
n
v1
(u, u, r) = [n
v
]u

(u
p
, u
m
, r)
n
u1
(u, u, r) = [n
u
]u

(u
p
, u
m
, r) (9)
h
1
(u, u, r) = [h]u

(u
p
, u
m
, r)
where the [] matrices are N L where N is the number of degrees-of-freedom and L is the length of
u

which contains all the nonlinear terms. The expressions in (9) can be used to obtain a relationship
for the coecients of the form
[

h] = [n
v
] [n
u
] (10)
where the (n, )
th
element in [

h],

h
n,l
, is related to the corresponding element in [h], h
n,l
, using

h
n,
=
_
_
_
(m
p
m
m
) +
N

n=1
(s
np
s
nm
)
rn
_
2

2
rn
_
_
h
n,
=
n,
h
n,
. (11)
Here we use the matrix [] in which the (n, )
th
element is
n,
see [15] or [16] for full details of this
derivation.
For the near-identity transform, v u, the nonlinear term is rewritten in terms of u, the substi-
tution u = u
p
+u
m
is made and nonlinear term is written in matrix form
N
v
(u
p
+u
m
) = [n
v
]u

, (12)
using Eqns. (9) and (8), where for each element u
n
in u we have
u
n
= u
np
+u
nm
: u
np
=
U
n
2
e
in
e
irnt
, u
nm
=
U
n
2
e
in
e
irnt
. (13)
Using these relationships for our current example we obtain
u

=
_

_
u
3
1p
u
2
1p
u
1m
u
1p
u
2
1m
u
3
1m
u
1p
u
2
2p
u
1p
u
2p
u
2m
u
1p
u
2
2m
u
1m
u
2
2p
u
1m
u
2p
u
2m
u
1m
u
2
2m
u
2
1p
u
2p
u
2
1p
u
2m
u
1p
u
1m
u
2p
u
1p
u
1m
u
2m
u
2
1m
u
2p
u
2
1m
u
2m
u
3
2p
u
2
2p
u
2m
u
2p
u
2
2m
u
3
2m
_

_
, [n
u
]
T
=

m
_

_
1 0
3 0
3 0
1 0
3 0
6 0
3 0
3 0
6 0
3 0
0 3
0 3
0 6
0 6
0 3
0 3
0
0 3
0 3
0
_

_
[]
T
=
2
r1
_

_
8
0
0
8
4(n
2
+n)
0
4(n
2
n)
4(n
2
n)
0
4(n
2
+n)
4(1 +n)
4(1 n)
0
0
4(1 n)
4(1 +n)
8n
2
0
0
8n
2
_

_
(14)
where
r2
= n
r1
and, in [], a dash has been used where the corresponding value in [n
v
] is zero and
hence the value in [] is of no importance. Note that as
ri

ni
the linear natural frequencies must
be approximately related by
n2
n
n1
.
We are now in a position to select both the transform terms and the nonlinear terms in the dynamic
equation for u. Considering element (n, ), if possible we wish to remove the nonlinear term from the
dynamic equation and place it in the transform equation. However this is not possible for all terms.
From Eqn. (11) it can be seen that h
n,
can be large compared to

h
n,
if
n,
is close to zero. Also from
Eqn. (10), it can be seen that [

h] is of a similar size to [n
u
] and [n
v
]. Hence for the terms where
n,
is close to zero, this would suggest that if the element in the nonlinear term in matrix [n
v
] is placed in
the transform matrix [h], the transform can not be said to be near-identity. This results in two options
for the (n, ) element, that satisfy Eqn. (10) and ensure the near-identity condition it met. These are
the preferred option in which the term is removed from the equation of motion
Preferred Option: n
u,n,
= 0, h
n,
= n
v,n,
/
n,
(15)
and, to ensure that the transform is near-identity, the near-resonant option in which the nonlinear term
is kept in the equation of motion
Near-resonant Option: n
u,n,
= n
v,n,
, h
n,
= 0 (16)
Using these transforms the equation of motion can be solved, for each mode in turn, using a trial
solution in the form u
n
=
Un
2
e
in
e
irnt
+
Un
2
e
in
e
irnt
(for the n
th
mode) without the need for a
harmonic balance approximation. Information regarding the harmonics of the response are captured
in the transform equation
v = u +[h]u

+ (17)
This equation can also be used to identify the nonlinear normal modes of a system.
From [] it can be seen that regardless of the value of n the following terms are resonant; [1, 2],
[1, 3], [1, 6], [1, 9], [2, 13], [2, 14], [2, 18] and [2, 19]. In addition if n = 1 the terms [1, 7], [1, 8], [2, 12] and
[2, 15] are resonant. For the resonant terms we set the relevant terms in [n
u
] to equal those in [n
v
] and
set the corresponding terms in [h], the transform matrix, to equal zero. Provided n = 1, this results in
the following normal form for the system
u
1
+
2
n1
u
1
+
3
m
u
1p
u
1m
u
1
+
6
m
u
2p
u
2m
u
1
= 0 (18)
u
2
+
2
n2
u
2
+
3
m
u
2p
u
2m
u
2
+
6
m
u
1p
u
1m
u
2
= 0 (19)
where u
ip
+u
im
= u
i
has been used.
Backbone curves
To understand the modal response of the system using second-order normal forms we will concentrate
on the underlying modal dynamics, i.e. looking at the system response when the damping and forcing
are zero. Taking the n = 1 case, and making the substitutions u
ip
= (U
i
/2)e

ri
t
and u
im
= (U
i
/2)e

ri
t
,
such that u
i
= U
i
cos(
ri
t), and noting that
r2
= n
r1
results in the time-independent relationships
_

2
r1
+
2
n1
+
3
4m
(U
2
1
+ 2U
2
2
)
_
U
1
= 0, (20)
_
n
2

2
r1
+
2
n2
+
3
4m
(2U
2
1
+U
2
2
)
_
U
2
= 0. (21)
Note that for the case where n = 1, the [1, 7], [1, 8], [2, 12] and [2, 15] terms will be included in the
normal form equations, which results in time-independent relationships
_

2
r1
+
2
n1
+
3
4m
(U
2
1
+ 3U
2
2
)
_
U
1
= 0, (22)
_

2
r1
+
2
n2
+
3
4m
(3U
2
1
+U
2
2
)
_
U
2
= 0. (23)
From here, for clarity when deriving the equations we will concentrate on the n = 1 case, however
the technique is the same for both. There are two straightforward solutions to Eqns. (20) and (21) and
to Eqns. (22) and (23), these are
S1: U
2
= 0,
2
r1
=
2
n1
+
3
4m
U
2
1
(24)
S2: U
1
= 0, n
2

2
r1
=
2
n2
+
3
4m
U
2
2
(25)
These represent two backbone curves for the two degree-of-freedom system with S1 and S2 correspond-
ing to a response in the rst and second modes respectively.
In addition to these, under certain conditions there are two further solutions in which both U
1
and
U
2
are non-zero. Taking Eqns. (22) and (23), and setting the terms in the square brackets of both
equations to zero gives

2
r1
=
2
n1
+
3
4m
(U
2
1
+ 3U
2
2
) =
2
n2
+
3
4m
(3U
2
1
+U
2
2
) (26)
Rearranging this equation, and using = 1 + (8
2
/), gives
U
2
1
= (1 4

)U
2
2

2m
3
(
2
n2

2
n1
). (27)
It can be seen that for a real solution to U
1
two conditions must be met
4
2
and U
2
2

2m
3( 4
2
)
(
2
n2

2
n1
). (28)
This indicates that unless
n2
=
n1
, for this solution to be valid U
2
cannot be zero. However, U
1
= 0
is a valid solution, and in this case comparing Eqns. (25) and (26) we can see that the solution lies on
the S2 backbone curve, and in fact these two solutions branch out of the S2 backbone.
If we eliminate U
1
from Eqn. (26) we obtain the response frequency equation

2
r1
=
2
r2
=
3
2
n1
+
2
n2
2
+
3(
2
)
m
U
2
2
(29)
Taking Eqns. (27) and (29) we can summarise the two further solution as
S3: U
1
=
_
(1 4

)U
2
2

2m
3
(
2
n2

2
n1
),

2
r2
=
3
2
n1
+
2
n2
2
+
3(
2
)
m
U
2
2
,
S4: U
1
=
_
(1 4

)U
2
2

2m
3
(
2
n2

2
n1
),

2
r2
=
3
2
n1
+
2
n2
2
+
3(
2
)
m
U
2
2
. (30)
These two solutions indicate that four modal solutions exist despite the system just having two degree-
of-freedom.
Modeshapes
Using these backbone solutions we can generate the modeshapes for the system. Here, for the unforced,
undamped system we will plot the initial displacement conditions (assuming zero initial velocities) at
which a mode is observed in the x
1
x
2
plane. Then we consider the resulting modeshapes from these
initial conditions. The forced response is considered in [22].
Firstly we must consider the near-identity transform, Eqn. (17), to calculate the response in x
x = q = v = (u +[h]u

) (31)
1 0.5 0 0.5 1
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
initial condition plot X1
X
2
1 0.5 0 0.5 1
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
initial condition plot X1
X
2
Figure 1: Initial displacement conditions at which a modal response is observed, assuming the initial velocities are zero
(shown as thin lines), for (a) n=1 system and (b) n=3 system. The thick line gives an example of the resultant oscillatory
motion from a point on the S3 solution branch.
Again considering the n = 1 case, using Eqns. (14) and (15) we can write
x
1
= (U
1
+U
2
) cos(
r1
t) +

m
U
3
1
+ 3U
2
1
U
2
+ 3U
1
U
2
2
+U
3
2
32
2
r1
cos(3
r1
t)
x
2
= (U
1
U
2
) cos(
r1
t) +

m
U
3
1
3U
2
1
U
2
+ 3U
1
U
2
2
U
3
2
32
2
r1
cos(3
r1
t). (32)
For backbone solutions S1 and S2, where U
2
= 0 and U
1
= 0 respectively, these equations show that
the corresponding modeshapes (x
1
= x
2
and x
1
= x
2
respectively) are independent of amplitude and
response frequency
r1
. For solutions S3 and S4 the modeshapes are more complex. For these solutions
the ratio x
2
/x
1
is a function of amplitude and hence response frequency. The initial displacements that
result in a modal response may be written as
x
1
|
initial
=
2
r1
(U
1
+U
2
) + (/32m)(U
3
1
+ 3U
2
1
U
2
+ 3U
1
U
2
2
+U
3
2
)
x
2
|
initial
=
2
r1
(U
1
U
2
) + (/32m)(U
3
1
3U
2
1
U
2
+ 3U
1
U
2
2
U
3
2
) (33)
with
r1
and the relationship between U
1
and U
2
are dened by Eqn. (30) (noting
r1
=
r2
as n = 1).
Simulation Results
Consider the two degree-of-freedom system with parameters; m = 1, k = 1, k
2
= 0.005, = 0.4 and

2
= 0.05 such that
n1
= 1 and
n2
= 1.005. Using Eqn. (33), the initial conditions which result in
a modal vibration are shown as thin lines in Figure 1(a). The thick line shows the resulting motion if
released from a point on the S3 solution branch, calculated using Eqn. (32).
In the previous gure the oscillatory motion is almost an exact straight line, hence the mode shape is
essentially xed during the oscillation and is just a function of initial displacement conditions. However,
for the case where n = 1 this is not the case, the modeshape varies during oscillations if the system is
released from branches S3 or S4. To see this we consider the case where n = 3 using a system with
parameters m = 1, k = 1, k
2
= 4.015, = 0.05 and
2
= 10 such that
n1
= 1 and
n2
= 3.005. Figure
1(b) shows the initial conditions which result in a modal response (thin lines). It can be seen that for
this system the bifurcation occurs on the S1 solution. The thick lines show a series of responses for
the system released from the S3 initial conditions. It can be seen that in this case the shape of the
response varies over the oscillation.
To assess the accuracy of the normal form analysis, Figure 2(a) shows the system response calculated
using time-stepping simulation for two periods of oscillation (thick dashed line). It can be seen that it
agrees well with the normal form prediction. Figure 2(b) shows the corresponding response of x
1
and x
2
over time, conrming the behaviour is a vibration in unison as dened by [3]. To further highlight the
vibration-in-unison behaviour, Figure 3 shows approximately two cycles of time simulation response of
the system for initial conditions that do not lie on the S1-4 solution curves.
1 0.5 0 0.5 1
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
initial condition plot X1
X
2
0 2 4 6 8 10 12 14
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
time (s)
x
1
,

x
2
Figure 2: (a) comparison between time-stepping simulation (thick dashed line) over two periods of oscillation and the
normal form predictions for an initial displacement condition lying on the S3 solution and (b) the corresponding response
in x1 (solid line) and x2 (dashed line) over time.
1 0.5 0 0.5 1
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
initial condition plot X1
X
2
Figure 3: Approximately two cycles of time-stepping simulation response for initial conditions that do not lie on the
S1-S4 modal lines.
Conclusion
In this paper, we demonstrate how the second-order normal form method can be used to understand the
nonlinear modal response of a system. This is done by considering a two degree-of-freedom nonlinear
system with closely matching linear natural frequencies. We show how the system backbone curves
can be found using the second-order normal form technique. These curves are plotted in the x
1
x
2
plane as initial displacement conditions (assuming no initial velocities) from which a modal response is
observed. These solutions are more complex than the expressions for the underlying linear system as
for some x
1
values there are four x
2
values at which a modal response is observed due to a bifurcation
in one of the backbone curves. The responses over a cycle of oscillation are also shown for a number of
initial conditions, highlighting that the mode shape can vary over each cycle of oscillation. The normal
form predictions are shown to have high accuracy using time-stepping simulation results.
References
[1] Ewins, D.J., 2000. Modal Testing: Theory, Practice and Application, second edition. Research
Studies Press Ltd. Hertfordshire, UK.
[2] Friswell, M.I., Mottershead, J.E., 1995. Finite Element Model Updating in Structural Dynamics.
Kluwer Academic Publishers, London, UK.
[3] Rosenberg, R.M., 1960. Normal modes of nonlinear dual-mode systems. Journal of Applied Me-
chanics, 27, pp. 263268.
[4] Shaw, S.W., Pierre, C., 1991. Non-linear normal modes and invariant manifolds. Journal of Sound
and Vibration, 150, pp. 170-173.
[5] Vakakis, A.F. (Ed.), 2002. Normal Modes and Localization in Nonlinear Systems. Kluwer Academic
Publishers, London, UK.
[6] Kerschen, G., Peeters, M., Golinval, J.C., Vakakis, A.F., 2009. Nonlinear normal modes, part I:
a useful framework for the structural dynamicist. Mechanical Systems and Signal Processing, 23,
pp. 170-194.
[7] Kahn, P.B., Zarmi, Y, 2000. Nonlinear dynamics: a tutorial on the method of normal forms.
American Journal of Physics, 68(10), pp. 907919.
[8] Arnold, V.I., 1988. Geometrical methods in the theory of ordinary dierential equations. Springer.
[9] Murdock, J., 2002. Normal forms and unfoldings for local dynamical systems. Springer.
[10] Jezequel, L., Lamarque, C.H., 1991. Analysis of nonlinear dynamic systems by the normal form
theory. Journal of Sound and Vibration, 149, pp. 429459.
[11] Nayfeh, A.H., 2000. Nonlinear Interactions, Wiley series in nonlinear science.
[12] Kahn, P.B., Zarmi, Y., 1991. Minimal normal forms in harmonic oscillations with small nonlinear
perturbations. Physica D, 54, pp. 6574.
[13] Rand, R.H., Ambruster, D., 1987. Perturbation Methods, Bifurcation Theory and Computer Alge-
bra, Springer.
[14] Touze, C., Thomas, O., Huberdeau, A., 2004. Asymptotic non-lnear normal modes for large-
amplitude vibrations of continuous structures, Computers and Structures, 82, pp. 26712682.
[15] Neild, S.A., Wagg, D.J., 2010. Applying the method of normal forms to second-order nonlinear
vibration problems. Proceedings of the Royal Society, Part A, 467(2128), pp. 11411163.
[16] Neild, S.A., 2012. Approximate Methods for Analysing Nonlinear Structures, published in Exploit-
ing Nonlinear Behaviour in Structural Dynamics, Editors Wagg, D.J, Virgin, L.N., Springer.
[17] Vakakis, A.F., Rand, R.H., 1992. Normal modes and global dynamics of a 2-degree-of-freedom
nonlinear-system; part I: low energies. International Journal of Non-Linear Mechanics, 27, pp.
861-874.
[18] Vakakis, A.F., Rand, R.H., 1992. Normal modes and global dynamics of a 2-degree-of-freedom
nonlinear-system; part II: high energies. International Journal of Non-Linear Mechanics, 27, pp.
875-888.
[19] Wagg, D.J., Neild, S.A., 2009. Nonlinear Vibration with control, Springer-Verlag.
[20] Xin, Z., Neild, S.A., Wagg, D.J., 2011. The selection of the linearized natural frequency for the
second-order normal form method. Proceedings of the ASME International Design Engineering
Technical Conferences & Computers and Information in Engineering Conference IDETC/CIE
2011, Washington, DC, USA, paper DETC2011-47654.
[21] Doedel, E.J., Paenroth, R.C., Champneys, A.R., Fairgrieve, T.F., Kuznetsov, Y.A., Oldeman,
B.E., Sandstede, B., Wang, X.J., 2007. AUTO-07P: Continuation and bifurcation software for
ordinary dierential equations, URL http://indy. cs. concordia. ca/auto.
[22] Neild, S.A., Cammarano, A., Wagg, D.J., 2012. Towards a technique for nonlinear modal analy-
sis, Proceedings of IDETC/CIE ASME International Design Engineering Technical Conferences,
Chicago, USA, paper DETC2012-70322.

You might also like