Higher-Order Averaging in Time-Periodic Systems
Higher-Order Averaging in Time-Periodic Systems
https://doi.org/10.1007/s11071-019-05085-4
ORIGINAL PAPER
Abstract In this paper we show how higher-order Time-periodic systems · Vibrational stabilization ·
averaging can be used to remedy serious technical Vibrational control
issues with the direct application of the averaging the-
orem. While doing so, we reconcile two higher-order
averaging methodologies that were developed indepen- 1 Introduction
dently using different tools and within different com-
munities: (i) perturbation theory using a near-identity Nature is teeming with periodically forced creatures
transformation and (ii) chronological calculus using (e.g., insects, birds, fish). With the current active
Lie algebraic tools. We provide the underpinning con- research trend of mimicking nature’s designs, there
cepts behind each averaging approach and provide a are plenty of bio-inspired engineering systems that are
mathematical proof for their equivalence up to the represented by nonlinear time-periodic (NLTP) mod-
fourth order. Moreover, we provide a higher-order aver- els such as swimming robots [1–4] and flapping-wing
aging study and analysis for two applications: the clas- micro-air-vehicles [5,6]. Additionally, high-frequency
sical problem of the Kapitza pendulum and the modern periodic control is a common approach for stabilization
application of flapping flight dynamics of micro-air- and path planning for under-actuated mechanical sys-
vehicles and/or insects. tems (with control inputs less than degrees of freedom)
[2,7–9]. Vibrational control is an open-loop stabiliza-
Keywords Averaging theory · Higher-order aver- tion technique via the application of sufficient high-
aging · Chronological calculus · Lie transform · frequency, high-amplitude, periodic inputs. It was first
introduced by Meerkov [10] for linear time-varying
This work was supported by the National Science Foundation, systems. It is well known for its robustness [11] and
Grant CMMI-1709746 and its continuation CMMI-1846308. elegance (stabilization without feedback). By defini-
M. Maggia · S. A. Eisa · H. E. Taha (B)
tion, vibrationally controlled systems are time-periodic
Department of Mechanical and Aerospace Engineering, systems.
University of California, Irvine, Irvine, CA 92697, USA Analyzing the dynamics of these time-varying sys-
H. E. Taha tems is quite more challenging than that of autonomous
e-mail: hetaha@uci.edu (time-invariant) ones. For example, a NLTP system
M. Maggia usually attains equilibrium when all of its states
e-mail: mmaggia@uci.edu undergo periodic solutions. That is, its equilibria may
S. A. Eisa not be represented by fixed points but rather by periodic
e-mail: seisa@uci.edu orbits. Moreover, even when the system is linear and
123
M. Maggia et al.
NLTP
cal implementation of the Floquet theorem precludes
Model scrutinizing the dynamical behavior of the system on
an analytical level. The second approach is to use the
averaging theorem to transform the NLTP system into a
nonlinear, time-invariant (NLTI), i.e., an autonomous,
Compute system in which the periodic orbit of the original sys-
Averaging Periodic Orbit tem corresponds to a fixed point. Then, the averaging
then Linearize
theorem guarantees that, for high enough frequency,
exponential stability of this fixed point implies expo-
NLTI LTP nential stability of the corresponding periodic orbit.
Model Model Thus, the averaging approach dodges the determination
of the periodic orbit under study and, hence, provides
Apply a compact, constructive technique to ensure the peri-
Find Fixed Point Floquet odic orbit and to analyze its stability (by ensuring the
then Linearize Decomposition corresponding fixed point and studying its stability). In
fact, averaging has been the standard and most common
technique to design vibrational control systems [7,20].
LTI Despite its convenience to analyze NLTP systems,
Model the application of averaging should be performed care-
fully. In fact, there might not be a more expressive
statement about the subtleties of averaging than that by
Spectral Sanders and Verhulst in their book on the topic [17]:
Analysis
To many physicists and astronomers averaging
seems to be such a natural procedure that they do
Fig. 1 The two main approaches used to analyze NLTP systems
not even bother to justify the process. However,
it is important to have a rigorous approximation
theory, since it is precisely the fact that averaging
seems so natural that obscures the pitfalls and
has a fixed point (ẋ = [A(t)]x), its stability cannot be
restrictions of the method.
assessed through the eigenvalues of the system matrix
[A(t)]. That is, in contrast to linear time-invariant sys- Sticking to the standard form of averaging, its direct
tems ẋ = [A]x where a Hurwitz matrix (its eigenval- application may not be admissible and/or practically
ues have strictly negative real parts) implies exponen- useful in many cases. First, the averaging theorem
tial stability, if all the eigenvalues of the time-varying necessitates a special structure (weakly forced sys-
matrix [A(t)] lie in the open left half of the complex tems). Therefore, it cannot be applied directly to
plane for all times, the system may still be unstable; vibrational control systems, which are typically high-
Markus and Yamabe [12] provided a counterexample. amplitude, periodically forced systems. The standard
In general, the stability analysis of NLTP systems remedy for this case is to apply the nonlinear variation
is conducted using one of two main approaches: a of constants formula [7,21] to transform the multi-scale
numerical approach based on the Floquet theorem [13– system into two companion systems, each of which is
15] and an analytical approach based on the averag- amenable to the averaging theorem (see [7,8,22,23]).
ing theorem [16–19]. As shown in Fig. 1, the first It is interesting to note that the use of the variation of
approach requires determination of the periodic orbit constants (VOC) formula before averaging goes back
(i.e., solving the dynamic equations). Then, the Flo- to Lagrange in his analysis of the perturbed two-body
quet theorem is used to analyze the stability of the lin- problem (see [24, pp. 181–184]). However, the VOC
ear, time-periodic (LTP) system obtained by lineariz- formula, which requires computation of flow maps (i.e.,
ing the dynamics about the concerned periodic orbit. solution of part of the differential equation), may not
For non-trivial applications, this approach cannot be be analytically tractable for complex systems, hence
performed analytically. Hence, the inevitable numeri- frittering away the analytical advantage of the averag-
123
Reconciliation of two higher-order averaging techniques
ing analysis. Second, even when the system structure colleagues Agrachev and Gamkrelidze [28,29] devel-
permits analytical application of the VOC (e.g., sim- oped a new branch of calculus to study time-varying
ple mechanical systems [7]), a major issue with the dynamical systems: the chronological calculus. The
averaging theorem is that it is applicable only for high main objective was to provide an exponential repre-
enough oscillation frequency. That is, it only guaran- sentation of the flow along a time-varying vector field,
tees the existence of a threshold frequency ω∗ above in a way that resembles the exponential representation
which stability of the averaged dynamics implies sta- of the flow of autonomous vector fields given by the
bility of the corresponding NLTP system. The predica- Fliess functional expansion [30]. Utilizing these math-
ment is that almost none of the averaging theorems tell ematical tools, Sarychev [31] and Vela [27] developed
how much ω∗ is. As such, vibrational control is typi- higher-order averaging techniques for time-periodic
cally achieved by ensuring that the averaged dynamics systems, generalizing the classical averaging theorem
is stable and then sweeping the frequency (e.g., by sim- to cases where the excitation frequency is not high
ulation) until hopefully the threshold is hit. Then, the enough. Inspired by the Floquet theorem, the main
averaging theorems guarantee stability for the origi- objective was to determine a time-invariant (averaged)
nal time-periodic system for all frequencies above this system whose flow after one period matches the flow
threshold. The main issue stressed in the recent work of the time-periodic system after one period. Hence,
of Berg and Wickramasinghe [20] is that finding a sta- the stability characteristics of the NLTP system can
ble response of the time-periodic system at a certain be deduced from the stability characteristics of the
frequency does not necessarily mean that the threshold averaged system, thereby extending the Floquet the-
is reached. That is, the time-periodic system may be orem to nonlinear systems. Because this approach is
stable at some ω1 and unstable at ω2 > ω1 (clearly, developed using differential geometric techniques, the
ω1 < ω2 < ω∗ ). This finding also points to the pos- higher-order averaged dynamics are given in terms of
sibility for vibrational stabilization at a frequency less Lie brackets between the time-periodic vector field and
than the threshold value, hence relaxing the averaging its integrals.
unfeasible requirements. Throughout several discussions between the third
In this paper, we propose higher-order averaging author and Prof. Ali Nayfeh, the latter always sug-
techniques as potential remedy for the two issues raised gested that the two averaging approaches are just the
above regarding direct (first-order) averaging. In par- same. Yet, it is not clear in the literature whether such
ticular, we show how higher-order averaging captures approaches are equivalent or not. Therefore, we find it
more dynamical features that are typically neglected a good opportunity to reconcile the two higher-order
by direct averaging. Moreover, the threshold frequency averaging techniques in this special issue in the mem-
may decrease as the order of averaging increases. While ory of Prof. Nayfeh, particularly because he made
doing this, we find it necessary to reconcile differ- several important contributions to the first classical
ent averaging approaches in the literature. Two higher- approach [13–15,32].
order averaging techniques can be found in the litera- The main contribution of this paper is to recon-
ture, which are developed separately employing differ- cile these two averaging techniques, under a unified
ent tools. The first technique is based on the classical notation, providing explicit formulas for higher-order
averaging theory developed by Krilov et al. [25,26]; averaging terms. Moreover, we prove that the two
some authors such as Vela [27] refer to it as the approaches are equivalent up to the fourth order. It is
KBM method. This classical approach has its roots noteworthy to mention that, unlike the classical averag-
in the perturbation theory; the main objective is to ing approach, the one based on chronological calculus
find a near-identity transformation that transforms the does not provide a recursive formula for the compu-
time-periodic system into a time-invariant system after tation of higher-order terms. Hence, the equivalence
neglecting terms of arbitrary small order of magnitude. between the two averaging techniques cannot be sys-
The other higher-order averaging technique is based tematically carried out for an arbitrarily high order. In
on the so-called chronological calculus, exploiting this paper, the equivalence is mathematically proven
Lie algebraic tools. In honor of the 70th birthday of only up to the fourth order. Moreover, we show a
the great Russian mathematician Lev Pontryagin, the “conceptual” equivalence between the two approaches
founder of the optimal control theory, his students and which suggests that the mathematical equivalence can
123
M. Maggia et al.
be extended for terms beyond the fourth order. This ẋ = f(x, t, ), (1)
equivalence will allow each approach to benefit from where x ∈ Rn , ẋ denotes the derivative of x with respect
tools and results developed in the other branch. After to t, and is a small parameter such that 0 < 1.
proving such an equivalence, we apply these tech- The vector field f is smooth in x, analytic in , and T -
niques to two applications: (i) a classical application of periodic in t. The averaged system corresponding to
the Kapitza pendulum (inverted pendulum subjected to (1) is
vertical oscillation) and (ii) a modern application on the
ẋ = f(x) (2)
hovering flight of insects and flapping-wing micro-air- T
vehicles. In these examples, we show how higher-order where f(x) = 1
T 0 f(x, τ, 0) dτ .
averaging captures more dynamical features than direct 1. If x(0) − x(0) = O(), then there exist b, ∗ ∈ R+
averaging and how it can mitigate the technical issues such that x(t) − x(t) = O() for all t ∈ [0, b/]
associated with the direct application of the averaging and for all ∈ (0, ∗ ).
theorem. 2. If the origin x = 0 is an exponentially stable equi-
The paper is organized as follows: In Sect. 2, the librium of system (2) and if x(0) − x(0) = O(),
two approaches (classical averaging and chronological then there exists an ∗ such that x(t)−x(t) = O()
calculus) are described separately. The reconciliation for all t ≥ 0 and for all ∈ (0, ∗ ). Furthermore,
of the two averaging approaches is given in detail in system (1) has a unique, exponentially stable, T -
Sect. 3. Lastly, Sect. 4 discusses the two applications periodic solution xT (t) such that x T (t) ≤ k for
mentioned above. some k ∈ R+ .
2 Two higher-order averaging techniques Sometimes, the first-order averaging approximation
is too poor and the averaged system (2) fails to cap-
NLTP systems are usually characterized by two ture some important characteristics of the NLTP sys-
timescales: a fast timescale and a slow timescale. The tem. In particular, Theorem 1 is valid only for weakly
goal of averaging is to filter the slow behavior by cap- forced systems or for systems with large separation
turing the mean effect of the dynamics over a period of between the two time scales (e.g., the ratio of the body’s
the fast timescale. During a period, the slowly vary- flight natural frequency to the flapping frequency for
ing states (that do not explicitly depend on the fast the hummingbird, ≈ ωn /ω, 1). Also, no infor-
time) are approximated as constants. As an example, mation is given about how large this separation should
let us consider a hummingbird flapping its wings. The be for the theorem to be valid. For these reasons, higher-
wings move periodically at a frequency ω that is much order averaging terms are introduced. The higher-order
higher than the natural frequency associated with the terms are also time-invariant and are scaled by powers
bird’s body flight ωn (i.e., ω ωn ). During one flap- of the small parameter 1. In the averaging the-
ping cycle, the body oscillates in all directions. Nev- ory, an r th-order averaged system implies neglecting
ertheless, on the average, it is hovering in the same terms of order r . As such, this truncation becomes
place (e.g., over a flower). Moreover, the oscillation of more accurate as the order increases and/or the param-
the body around the mean position/attitude is minimal. eter decreases.
Therefore, averaging seems so natural to analyze such In this section we briefly summarize the two higher-
a dynamical system. Analyzing the averaged dynamics order averaging approaches that are found in the liter-
can be extremely beneficial because the complexity of ature. We start with the classical averaging approach
the system can be noticeably reduced. In particular, the [15,17,18,33] and proceed with the description of the
explicit dependence on time is eliminated and the aver- less known approach based on chronological calculus
aged system becomes autonomous (i.e., it is fast time- [27,31,34].
invariant). The approximation by averaging is justified
by the following theorem, better known simply as the 2.1 Higher-order averaging via a near-identity
averaging theorem. transformation (lie transform)
Theorem 1 Averaging Theorem (Theorem 10.4 in
[16]). Consider the following NLTP system written in This higher-order averaging approach utilizes a partic-
averaging-canonical form ular change of coordinates to approximate the NLTP
123
Reconciliation of two higher-order averaging techniques
system by a NLTI one. The original NLTP system will 2.1.1 First-order averaging
be at times referred to as the exact system. The method-
ology has been developed from perturbation theory (see Let us rewrite Eqs. (1), (5) and (6) as
[13,15,17,18]).
ẋ = f1 (x, t) + 2 f̂(x, t, ), (7)
We begin our analysis by considering a NLTP
dynamical system written in the averaging-canonical x = y + w1 (y, t) + ŵ(y, t, ),
2
(8)
form1 (1). Since f is smooth in x and analytical in , it ẏ = g1 (y, t) + 2 ĝ(y, t, ), (9)
can be expanded in power series of as where f̂, ĝ and ŵ denote remainders. Differentiating
∞
m−1 Eq. (8) with respect to t and substituting by Eq. (9) in
f(x, t, ) = fm (x, t) = f1 (x, t)
m! the resulting equation, we obtain
m=1
2 ẋ = g1 (y, t) + 2 ĝ(y, t, )
∂w
+ f2 (x, t) + f3 (x, t) + · · ·
2! 3!
(3) ∂w1 1
+ g1 (y, t) + ĝ(y, t, ) +
2
where the functions fm are computed as, ∂y ∂t
∂ m−1 f + O( 2 ). (10)
fm (x, t) = m m−1 . (4)
∂ =0 Substituting Eq. (8) into the argument of f1 in Eq. (7),
The objective is to find a transformation that con- we obtain
verts the x-dynamics into a time-invariant system in the
transformed coordinates y. Given a transformation w, ẋ = f1 (y + w1 (y, t), t) + O( 2 ). (11)
we perform the following change of coordinates x → y Equating the terms in Eqs. (10) and (11) of the same
x = y + w(y, t, ), (5) -power, the 1 equality implies
where y ∈ Rn is the transformed state. For = 0, ∂w1 (y, t)
g1 (y, t) = f1 (y, t) − . (12)
Eq. (5) reduces to the identity x = y; because is small, ∂t
the transformation in (5) is a quasi-identity, hence the The vector field g1 (y, t) in Eq. (12) is still time-varying.
name near-identity transformation. This transforma- The only way to make g1 time-independent, while w
tion is sometimes referred to as a Lie transform. When is T -periodic, is to take g1 equal to the average, over a
working with NLTP systems, the function w is also period T , of the vector field f1 [17,18]
required to be T -periodic in t. Furthermore, we assume 1 T
w to be smooth in y and analytic in , so we can expand g1 (y) = f 1 (y) = f1 (y, t) dt. (13)
T 0
it in an asymptotic power series of as
∞ Combining Eqs. (12) and (13) we find an expression
m−1
w(y, t, ) = wm (y, t) for the first term of the expansion of the transformation
m! w(y, t) as
m=1
When transformation (5) is applied to system (1), we t
obtain a new nonlinear smooth system in the form w1 (y, t) = (f1 (y, τ ) − g1 (y)) dτ + C1 (y), (14)
0
∞
m where C1 (y) is an arbitrary t-independent and y-
ẏ = gm (y, t), (6)
m! dependent constant of integration.
m=1
where the functions gm (y, t) depend on the particular
choice of the near-identity transformation (5) and are 2.1.2 Higher-order averaging
in general time-varying vector fields. We now demon-
strate the procedure to make g1 (y, t) time-invariant, The previous reasoning can be easily extended to higher
and we then extend the reasoning to higher-order terms orders. In fact, the terms gm (y, t) are obtained by differ-
gm (y, t) with m > 1. entiating both sides of Eq. (5) with respect to time, sub-
stituting the -expanded Eqs. (1) and (6) in the resulting
1 It is important to notice that, if direct averaging is applied to expression, and finally equating terms of like power of
a system that is not written in the averaging-canonical form, the
following analysis might not be correct. Averaging a system not
. We thus obtain expressions equivalent to Eq. (12)
written in canonical form is known as crude averaging and is (m) ∂wm (y, t)
discouraged as pointed out in [17]. gm (y, t) = f0 (y, t) − . (15)
∂t
123
M. Maggia et al.
Equation (15) is also known as the homological equa- To compute a certain gm , one only needs to know
tion [17]. Proceeding analogously, we make gm (y, t) the power expansions f1 , f2 , . . . , fm in Eq. (4) and
time-invariant by taking to set the integration constants C1 , C2 , . . . , Cm−1 in
(m)1 T (m) Algorithm 1. The constants of integration Cm (y) are
gm (y) = f 0 (y) = f (y, t) dt, (16) independent of time, but are dependent on the state y;
T 0 0
therefore, different choices of such constants can lead
and the higher-order terms of the transformation w are
to significantly different results. The selection of Cm
given, similarly to Eq. (14) as
is arbitrary; there are typically three main choices that
t
wm (y, t) = f0(m) (y, τ ) − gm (y) dτ + Cm (y). are found in the literature:
0
1. Stroboscopic Averaging:
(17)
It is important to highlight that in Eqs. (15), (16) Cm = 0. (19)
(m)
and (17), the functions f0 (y, t) are not to be con- By choosing the stroboscopic averaging, the near-
fused with the functions fm (y, t) previously defined identity transformation (5) reduces to the identity
in Eq. (4). In fact, as previously described, the func- when t = nT , with n ∈ Z. In other words, if one
(m)
tions f0 (y, t) stem from equating coefficients of like were to observe, using a stroboscopic light flashing
power of and can be computed recursively using the at these times, the solutions of the exact system (1)
following algorithm. Given f0(1) = f1(0) = f1 and g1 and the transformed system (21) (with the same ini-
computed as in Eq. (13), then the higher-order aver- tial conditions), they would appear to coincide. For
aged vector fields gm for m > 1 can be computed elegance and simplicity of computations, this is the
following Algorithm 1. In Algorithm 1, we dropped most common choice for higher-order averaging.
the argument of f, g and w for the sake of notation 2. Zero-mean wm :
clarity. Moreover, Aik = k!/((k − i)!i!) are the bino- τ
1 T
mial coefficients and [·, ·] represents the Lie bracket Cm = − f̃m dτ dt. (20)
between vector fields. The standard results in the liter- T 0 0
ature [15,33] are written in this section in terms of Lie By choosing the constants of integration as in
brackets to facilitate the reconciliation with the other Eq. (20), the periodic functions wm will have
averaging approach based on chronological calculus. zero
mean over the period T . In other words,
1 T
The Lie bracket [X, Y] between the two vector fields T 0 wm dt = 0. This choice is preferable when
X(x, t), Y(x, t) ∈ Rn is defined as (see [21,35–37]) fm are written as Fourier series so that wm maintain
∂Y ∂X the same structure (see [18, Ch. 6]).
[X, Y] = X− Y. (18) 3. Reduced Averaging: Cm are chosen so that all the
∂x ∂x
functions gm vanish for m > 1. In other words, the
Algorithm 1 Computing gm for m > 1. dynamics is not captured by the averaged system,
For k = 2, . . . , m: but only via the near-identity transformation. (see
[17, Ch. 3]).
(0)
fk = fk ,
The transformed system (6), when all the vector
t
(k−1) fields gm are computed as in Eq. (13) and Algorithm 1
wk−1 = f0 − gk−1 dτ + Ck−1 ,
0 (i.e., gm are all time-invariant), becomes a NLTI system
k−2
(1)
fk−1 = fk +
(0) (0)
Aik−1 wi+1 , fk−i−1 , ∞
m
i=0 ẏ = gm (y). (21)
m!
( j) ( j−1)
k− j
k− j
m=1
fk− j = fk− j+1 + Ai
After proving equivalence with the complete averaging
i=0
( j−1) (k−i−1)
of Sarychev [38] and Vela [27] in the next section, it will
× wi+1 , fk− j−i − f0 + gk−i−1 , j = 2, . . . , k, be clearly seen that system (21) represents the complete
1 T averaging of the NLTP system (1). That is, system (21)
(k)
gk = f0 dt.
T 0 is an exact transformation of the original system (1):
No approximation has been introduced yet; if y(t) is
123
Reconciliation of two higher-order averaging techniques
123
M. Maggia et al.
We can write the solution of Eq. (29) using the To better understand the concept of GAT, we now
Volterra series expansion as recall some key results from the linear Floquet theory.
Consider the homogeneous LTP system,
x(t) = x0
∞
t τ1 τm−1 ẋ = A(t)x, (32)
+ ··· Y(x0 ) dτm . . . dτ1
with A ∈ Rn×n and T -periodic in t. The flow asso-
m=1 0 0 0
(30) ciated with (32) can be represented by the so-called
fundamental solution matrix, written as
where Y = Yτ1 ◦ Yτ2 ◦ · · · ◦ Yτm . Here, Yτ = t
∂Y →
f(x, τ ) and Y τ1 ◦ Y τ2 = ∂ xτ2 Y τ1 . The condi- Φ t = exp
A
A(τ )dτ ∈ Rn×n .
0
tions of convergence of series (30) are provided by
Agrachev and Gamkrelidze [29]. In a similar fash- The fundamental solution matrix of system (32), com-
ion to the exponential representation of solutions of puted after one period T , is known as the monodromy
time-invariant systems [30,40], Agrachev and Gamkre- matrix M = Φ A T.
lidze [29] expressed the flow of the time-varying vector Theorem 3 Linear Floquet Theorem [41, p. 11] Every
field f(x, t) as fundamental matrix solution Φ A
t t of the LTP system (32)
f → can be represented as the product Φ A t = P(t)e
Rt of a
φ t = exp f(x, τ ) dτ . (31) T -periodic matrix (i.e., P(t) = P(t + T )) and R is a
0
constant matrix given by R = T1 ln M.
In the chronological calculus, this flow is also known
as the right-chronological exponential. Theorem 4 System (32) is uniformly stable if and only
Then, the logarithm of (31) is defined as if all Floquet multipliers (eigenvalues of M) have mod-
uli less than 1.
Vt = ln(φ f
t ).
In other words, the flow along the time-invariant vector Based on Theorem 3, one can use the transformation
field Vt for a unit time is equivalent to the flow along x = Py to transform the LTP system (32) to the LTI
the time-varying vector field f(x, t) for a time t. system
It is noteworthy to mention that although Vt is an ẏ = Ry (33)
autonomous vector field, it is parametrized by time;
if the final time t is changed, the vector field will whose flow after one period is exactly equivalent to the
change. Furthermore, in Ref. [29] it was shown that flow of the original LTP system (32) after one period:
(m) eRT = Φ A
Vt = ∞ m=1 V t , with T . That is, based on our discussion in the pre-
vious section on classical averaging, system (33) rep-
t τ1
Vt(m) = resents a stroboscopic averaged dynamics of the LTP
0 0 system (32). Moreover, one can determine the stabil-
τm−1
ity of the LTP system (32) from the Floquet multipliers
··· Gm (Yτ1 , . . . , Yτm )dτm . . . dτ1 ,
0 according to Theorem 4, or from the averaged dynamics
(33): The solution x(t) of (32) converges to the origin if
and Gm are the commutators polynomials. For instance,
and only if the solution y(t) of (33) converges to zero.
for m = 1, 2, 3, we obtain
Therefore, the LTP system (32) is stable if and only if
1 the averaged dynamics (33) is stable (i.e., the eigen-
G1 (ζ1 ) = ζ1 , G2 (ζ1 , ζ2 ) = [ζ2 , ζ1 ]
2 values of the R lie in the open left half of the complex
1 plane); there is no approximation here. It is complete
G3 (ζ1 , ζ2 , ζ3 ) =([ζ3 , [ζ2 , ζ1 ]] + [[ζ3 , ζ2 ], ζ1 ]) ,
6 averaging. The eigenvalues of R are called the Floquet
where [·, ·] denotes the commutator, which, for vec- exponents, or the characteristic exponents.
tor fields, coincides with the Lie bracket. There- Sarychev [38] utilized the Floquet concept of aver-
fore, chronological calculus provides an algorithmic aging without approximation to extend the Floquet
approach to determine the logarithm of time-periodic theory to the nonlinear case, equivalently extending
vector fields analytically. the averaging theorem to higher order. Similar to the
123
Reconciliation of two higher-order averaging techniques
Table 1 Linear versus nonlinear Floquet theory The vector field Λ in Eq. (34) can be written as a
Linear Nonlinear power series expansion in such that
∞
ẋ = A(t)x ẋ = f(x, t) m
ẋ = Λ(x) = Λm (x), (35)
→ → t m!
φ f
t
t = exp
ΦA 0 A(τ )dτ t = exp 0 f(x, τ ) dτ m=1
M = ΦA
T M = φTf where x is the average of x. According to Theorem 5,
R= 1
T ln M VT = 1
T ln M the solution of (29) oscillates around the solution of
ẏ = Ry ẋ = (x) = ∞ m=1
m
m! m (x)
the averaged NLTI system (35). Finally, if the series in
Eq. (35) is truncated after r terms, one arrives at the
r th-order averaged version
123
M. Maggia et al.
3 Reconciliation of the two higher-order averaging Lemma 1 Given two smooth vector fields X(x, t), Y
approaches (x, t) ∈ Rn , then
∂ ∂Y ∂X
In this section we will show that, under some key [X, Y] = X, + ,Y , (41)
∂t ∂t ∂t
assumptions, the two higher-order methodologies can
be proven equivalent despite the fact that they were and if X is time-invariant, the following are true
developed following very different and independent
∂ ∂Y
derivations. We prove equivalence mathematically up [X, Y] = X, , (42)
∂t ∂t
to the fourth-order averaging (r = 4) under the follow-
ing assumptions: and
Assumption 1 Assume that the NLTP system is writ- [X, Y]dt = X, Ydt .
ten in the form (29)
ẋ = f(x, t), Proof Using Eq. (18), it follows that
with f smooth in x and T -periodic in t. ∂
[X, Y]
∂t
Assumption 2 (Stroboscopic averaging) Assume that
∂ ∂Y ∂X
the arbitrary constants of integration in Eq. (17) and = X− Y ,
∂t ∂x ∂x
Algorithm 1 are taken to be null
∂ ∂Y ∂Y ∂X ∂ ∂X ∂X ∂Y
= X+ − Y− ,
Cm = 0, for m ≥ 1. ∂t ∂x ∂x ∂t ∂t ∂x ∂x ∂t
∂ ∂Y ∂X ∂Y ∂Y ∂X ∂ ∂X
With Assumption 1, we reduce the analysis to the = X− + − Y,
∂x ∂t ∂x ∂t ∂x ∂t ∂x ∂t
class of systems in which the parameter does not
appear explicitly in f. As a result, Eq. (3) becomes ∂Y ∂X
= X, + ,Y .
f(x, t, ) = f(x, t) = f1 (x, t) (i.e., fm (x, t) = 0 for ∂t ∂t
m > 1) and system (1) is in the form (29). Other-
If X is time-invariant, then ∂X
∂t = 0 and Eq. (41) reduces
wise, the chronological calculus technique would not
to (42). Moreover,
be even applicable. The second assumption renders the
first approach a fundamental concept similar to the ∂Y ∂X
[X, Y]dt = X− Y dt,
chronological calculus approach: The solution of the ∂x ∂x
averaged system coincides with that of the original sys- ∂Y ∂X
= Xdt − Ydt,
tem stroboscopically (after each cycle). This concept ∂x ∂x
was the basis of the chronological calculus averaging:
∂ Ydt ∂X
The goal was to find an autonomous vector field whose = X− Ydt,
∂x ∂x
flow after one period matches that of the time-periodic
vector field after one period. Therefore, while we pro- = X, Ydt .
vide below a mathematical proof for the equivalence
between the two approaches up to fourth order, this
conceptual matching suggests that the equivalence can
be extended to all higher orders. Theorem 7 Given Assumptions 1 and 2, the following
First, we state the following properties of Lie are true: g1 = Λ1 , g2 = Λ2 , g3 = Λ3 , and g4 = Λ4 .
brackets which will be useful in reconciling the
Proof Given in the following Subsect. 3.1–3.4.
two higher-order averaging techniques. Let X(x, t),
Y(x, t), Z(x, t) ∈ Rn and α ∈ R. The Lie brackets,
as defined in Eq. (18), satisfy the following properties: 3.1 First order
1. [X, Y] = −[Y, X] (skew-symmetry),
2. [X + Y, αZ] = α[X, Z] + α[Y, Z] (linearity), By comparing Eqs. (24) and (37), it trivially follows
3. [[X, Y], Z]+[[Y, Z], X]+[[Z, X], Y] = 0 (Jacobi that g1 = Λ1 . This equivalence holds also without
identity). Assumption 2.
123
Reconciliation of two higher-order averaging techniques
For the second-order terms, we start by expanding g2 In a similar fashion to the second-order averaging, let
given in Eq. (25) as us expand g3 by substituting the values of w1 and w2
t t
1 T found in Eq. (28), into Eq. (26).
g2 = f2 + f1 dτ, f1 − g1 dτ , f1
t
T 0 0 0 1 T t
t t g3 = f3 +
f1 dτ, f2 − g1 dτ, f2
T 0 0 0
+ f1 dτ , g1 − g1 dτ , g1 (43) t t
0
0 +2 f2 dτ, f1 + f2 dτ, g1
0 0
+ [C1 , f1 ] + [C1 , g1 ] dt. t τ
+2 f1 dτ1 , f1 dτ, f1
From Assumption 1, we have f2 = 0 and f1 = f, 0
t τ
0
and from Assumption 2, C1 = 0. Equation (43) then + f1 dτ1 , f1 dτ, g1 +
becomes
t t 0
t τ
0
1 T
g2 = f dτ, f − g1 dτ , f −2 g1 dτ1 , f1 dτ, f1
T 0 0 0 0 0
t t t τ
+ f dτ , g1 − g1 dτ , g1 dt. − g1 dτ1 , f1 dτ, g1
0 0
0 0 t τ
Since g is time-invariant, Lemma 1 implies +2 f1 dτ1 , g1 dτ, f1
t 1
0 0
g1 dτ , g1 = t g1 , g1 = t g1 , g1 = 0. t τ
0 + f1 dτ1 , g1 dτ, g1
To recover the second-order averaging using chrono- 0
0
t T
logical calculus, Λ2 in Eq. (38), it remains to show that −2
1
t t T
f2
1 T
0 0
f dτ , g1 − g1 dτ , f dt = 0. t
T 0 0 0 + f1 dτ, f1 dt dτ, f1
(44)
0
t 1 T
Using the skew-symmetry property of the Lie bracket − f2
(18), the left-hand side of Eq. (44) becomes 0 T 0
t t
1 T t
+ f1 dτ, f1 dt dτ, g1
f dτ, g1 + f, g1 dτ, dt. 0t t
T 0 0 0
Applying property (41), we obtain − f1 dτ, f1 dτ, g1
t 0
t
0
t
1 T ∂ t
f dτ, g1 dτ, dt. + g1 dτ, f1 dτ, g1
T 0 ∂t 0 0 0 0
Using the fundamental theorem of calculus, the expres- t T t
1
sion above simplifies to +2 f1 dτ, f2 + f1 dτ, f1 dt
T
T T
0
0
0
1 t 1 t T
f dt, g1 dt, . −2 f2 +
g1 dτ, f1 dτ, f1 dt
T 0 0 0 0 T 0
T T t t
Finally, noting that 0 f dt = T g1 , and 0 g1 dt =
+2 [C1 , f1 ] dτ, f1 − 2 [C1 , g1 ] dτ, f1
T g1 , 0 0
t
1
T g1 , T g1 = 0. + [C1 , f2 ] + (f1 − g1 ) dτ + 3[C1 , g1 ]
T 0
Hence, the second-order averaging via Lie transform t
given in Eq. (43) reduces to + [C1 , g1 ] dτ , g1
t
0
t
1 T
g2 = f dτ , f dt. (45) + C1 , − (f1 − g1 ) dτ, g1
T 0 0 0
Comparing Eqs. (45) and (38), it is easy to see that T t
1
g2 = Λ2 . + 2 C1 , f2 + f1 dτ , f1 dt
T 0 0
123
M. Maggia et al.
t t
+ [C1 , 3[C1 , g1 ]] dt + 3[C2 , g1 ]. (46) 2 T
+ f dτ , f dτ, f dt.
T 0 0 0
According to Assumption 1, f = f1 and f2 = f3 = In order to prove that g3 = Λ3 , we now need to show
0. Assumption 2 leads to C1 = C2 = 0. Therefore, that
Eq. (46) simplifies to t τ
T 2 T
t τ g1 , g2 − f, f dτ1 , f dτ dt =
2 T 2 T 0
g3 = f dτ1 , f dτ, f dt 0
t
0
t
T 0 0 0 3T 2 T
T t τ − g1 , g2 + f dτ, f dτ, f dt,
2 2 T 0
− g1 dτ1 , f dτ, f dt 0 0
T 0 0 0 or equivalently, that
t τ
1 T 2 T t τ
+ f dτ1 , f dτ, g1 dt 2T g1 , g2 = f, f dτ1 , f dτ dt
T 0 0 0 T 0
t τ
0 0
t (49)
1 T 2 T t
− g1 dτ1 , f dτ, g1 dt + f dτ, f dτ, f dt.
T 0 0 0 T 0
T t τ 0 0
2 Let us consider the right-hand side of Eq. (49)
+ f dτ1 , g1 dτ, f dt
T 0 0 0 2 T t τ
t τ f, f dτ1 , f dτ dt
1 T T 0
+ f dτ1 , g1 dτ, g1 dt 0
t
0
t
T 0 0 0 2 T
(47) + f dτ, f dτ, f dt.
2 T t
T 0
− g2 dτ, f dt 0 0
T 0 0 Using the linearity of integration and the Lie bracket
t
1 T property (41), we obtain
− t t τ
g2 dτ, g1 dt
T 0 0 2 T ∂
T t t f dτ, f dτ1 , f dτ dt,
1 T 0 ∂t 0
− f dτ, f dτ, g1 dt 0 0
T 0 0 0 which is further simplified, by applying the fundamen-
t t
1 T tal theorem of calculus, to
+ g1 dτ, f dτ, g1 dt T T t
T 0 0 0 2
t f dτ, f dτ, f dt .
2 T T 0 0 0
+ f dτ, g2 dt
T 0 0 Finally, recalling the expressions for g1 and g2 in
T t Eqs. (24) and (45), respectively, we obtain
2
− g1 dτ, g2 dt,
T 0 0 2T g1 , g2 .
where g2 is given by Eq. (45). Using the Lie brackets Thus, g3 = Λ3 and the equivalence of the third-order
properties and some other algebraic manipulations,3 terms of the two higher-order averaging techniques is
Eq. (47) can be further simplified to proved.
T
g3 = g1 , g2
2
t τ 3.4 Fourth order
2 T
− f, f dτ1 , f dτ dt. (48)
T 0 0 0 According to Assumption 1, f = f1 and f2 = f3 = f4 =
Using the fact that g1 = Λ1 and g2 = Λ2 , we rewrite 0. Assumption 2 leads to C1 = C2 = C3 = 0. Sub-
Eq. (39) as, stituting the expressions for w1 , w2 and w3 in Eq. (28)
3T into Eq. (27) and after some algebraic manipulations,
Λ3 = − g1 , g2 the expression for g4 in Eq. (27) reduces to
2
T t τ τ2
3 6
The process of reducing Eqs. (47)–(48) is rather lengthy and g4 = f dτ2 , f dτ1 , f dτ, f dt
would take several pages. Thus, the authors decided not to include T 0 0 0 0
(50)
this part in this paper and refer the reader to our Mathematica® T2
file [42] for the complete proof. − g1 , g1 , g2 + T g1 , g3 ,
2
123
Reconciliation of two higher-order averaging techniques
123
M. Maggia et al.
x(t)
g f +g g
l
θ F y(0) = z(t)
x(0) = z(0) = x0
B1 cos(ωt)
Fig. 3 The flow along f + g is equivalent to the flow along the
vector fields F and then g
123
Reconciliation of two higher-order averaging techniques
0.5
0.4
0.3
0.2
0.1
0
-0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0
system (53). The approximations are computed up to as the frequency is increased, which matches the intu-
fifth order using the algorithm described in Sect. 2.1, ition behind averaging. These two conclusions indi-
and by choosing the stroboscopic averaging, namely cate that, for a required level of accuracy, the thresh-
in Algorithm 1, we set Ck = 0 for k = 1, 2, 3, 4. old frequency decreases with the averaging order. That
Hence, the two averaging techniques will lead to iden- is, higher-order averaging allows stabilization at fre-
tical results due to the proved equivalence in this case quencies lower than that demanded by direct averaging,
of ∂F/∂ = 0. Figure 4 illustrates the stability map hence making vibrational control more feasible.
of the Kapitza pendulum in a similar fashion to the Figure 6 shows the angular displacement and veloc-
work of Berg and Wickramasinghe [20]: The plotted ity (θ , θ̇) of the Kapitza pendulum with initial con-
quantities are defined as α = −Ω/ω2 and β = B/ω, ditions θ0 = 30◦ and θ̇0 = 0◦ /s and for Ω = 2,
and the region colored in yellow represents the Floquet B = 3, and ω = 20 rad/s. The three curves repre-
stable region. In addition, the figure shows the approx- sent the integration of the complete LTP dynamics (57),
imations of the lower stability boundary obtained via the first-order and third-order averaged dynamics. The
higher-order averaging. Although only the lower sta- solutions of the averaged dynamics have been trans-
bility boundary is captured, it is evident that the accu- formed according to Eq. (23). While both the averaged
racy of the approximations increases with the averag- solutions correctly capture the stability of the origi-
ing order. That is, the higher-order averaging stabil- nal system, the third-order approximation is practically
ity boundary converges asymptotically to the Floquet indistinguishable from the solution of the full LTP sys-
boundary as the averaging order increases. tem.
Figure 5 shows the percent error between the Floquet We also present in Figs. 7 and 8 the results for a non-
lower stability boundary and the higher-order averag- stroboscopic averaging case. In particular, the integra-
ing approximations as the forcing frequency ω is var- tion constants Ck are computed using Eq. (20). Similar
ied. Two important conclusions can be drawn from this conclusions to the stroboscopic case can be drawn.
figure. First, for a fixed ω, the error decreases as the
averaging order is increased, which matches the intu-
ition behind the asymptotically convergent series as 4.1.2 Higher-order averaging without VOC
more terms are taken [15]. Second, the error associated
with all the averaging order approximations decreases As shown above, vibrational control systems are typ-
ically forced by high-amplitude periodic inputs and
123
M. Maggia et al.
0.5
Error (%)
0.4
0.3
0.2
0.1
0
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
30
the time-periodic terms, which may not be analytically
20
tractable for many systems. Here we show that higher-
10 order averaging alone (without the VOC) may capture
0 some of the dynamical features that are lost by disre-
-10 garding the VOC. In order to demonstrate this point,
-20 consider again the system in Eq. (53) after the transfor-
mation τ = ωt = t/
-30
0 5 10 15 20 25 30
x1 x2
= , (58)
x2 (Ω − ωB cos(τ ))x1
50
Figure 9 shows the percent error in the stability
boundaries obtained by applying higher-order averag-
ing directly (without VOC) to system (58). Similar con-
0 clusions to those of Fig. 5 (with the VOC formula) can
be drawn: The accuracy increases as ω increases (i.e.,
decreases) and/or the averaging order is increased.
-50 However, in comparison with the higher-order averag-
0 5 10 15 20 25 30
ing with VOC, the accuracy of the approximations in
Fig. 9 is noticeably lower; a significantly higher-order
Fig. 6 Integration of the full LTP dynamics, first- and third-order averaging may be required to attain an acceptable level
averaged dynamics of the Kapitza pendulum, after applying the of accuracy. Finally, Fig. 10 shows the angular displace-
VOC formula. The simulations are obtained for Ω = 2, B = 3, ment and velocity of the Kapitza pendulum with initial
ω = 20 rad/s, and initial conditions θ0 = 30◦ and θ̇0 = 0◦ /s
conditions θ0 = 30◦ and θ̇0 = 0◦ /s and for Ω = 2,
B = 3, and ω = 20 rad/s. The four curves represent
the integration of the complete LTP dynamics (58), the
therefore are not in the averaging-canonical form; the first-order, third-order and fifth-order averaged dynam-
VOC formula is necessary to ensure a rigorous appli- ics. The solutions of the averaged dynamics have been
cation of the averaging theorem. However, as shown adjusted according to Eq. (23). In this case, the solution
above, the VOC requires computing the flow map along of the first-order averaged dynamics provides wrong
123
Reconciliation of two higher-order averaging techniques
0.4
0.3
0.2
0.1
0.075
0.05
0.025
information about the stability of the original system, when the application of the VOC formula is not possi-
while the stability characteristics are correctly captured ble, such as in the modern example shown below.
by the higher-order averaged dynamics. As expected,
when the VOC formula is not applied, a higher aver-
4.2 Hovering flight of insects and flapping-wing
aging order (fifth order versus third order for the VOC
micro-air-vehicles (FWMAVs)
case) may be required to obtain a close approximation
of the solution of the original LTP system.
Flapping flight dynamics is a very rich dynamical sys-
In conclusion, higher-order averaging might be ben-
tem because it is typically represented by a multi-body,
eficial to the stability analysis of complex systems even
nonlinear, time-varying model. To simplify the analy-
123
M. Maggia et al.
Error (%)
6
0
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
150
flapping flight dynamics in the longitudinal plane (x −z
plane in Fig. 11) will be governed by the exact same
100
set of equations governing conventional airplanes [49]
⎛ ⎞
50
⎛ ⎞ ⎛ ⎞ 1
X (x, t)
u̇ −qw − g sin θ ⎜ m ⎟
0 ⎜ ẇ ⎟ ⎜ qu + g cos θ ⎟ ⎜ m1 Z (x, t) ⎟
⎜ ⎟=⎜ ⎟+⎜ ⎟
⎝ q̇ ⎠ ⎝ 0 ⎠ ⎜ 1 M(x, t) ⎟ , (59)
-50 ⎝ Iy ⎠
0 1 2 3 4 5 θ̇ q 0
200
where I y and m are the pitch inertia and the body
100 mass, respectively, and g is the gravitational accel-
eration. Figure 11 illustrates a flapping-wing micro-
0 air-vehicle (FWMAV) in the longitudinal plane (x − z
plane) where the state vector x includes the forward and
-100
normal velocity components u, w, the pitching angular
velocity q, and the pitching angle θ . In Eq. (59), X , Z
-200
0 1 2 3 4 5 are the forward and normal aerodynamic forces, respec-
tively, and M is the aerodynamic pitching moment.
Fig. 10 Integration of the full LTP dynamics, first-, third- and Details of the aerodynamic model (dependence of
fifth-order averaged dynamics of the Kapitza pendulum, without the aerodynamic forces X , Z and M on the state vari-
applying the VOC formula. The simulations are obtained for ables u, w and q) are taken from Ref. [50] and given in
Ω = 2, B = 3, ω = 20 rad/s, and initial conditions θ0 = 30◦
and θ̇0 = 0◦ /s
“Appendix A.”
The main distinction between conventional air-
planes and FWMAVs or insect flight is that the aero-
sis of such a complicated system, it may be reasonable dynamic forces (X , Z and M) are essentially time-
to neglect the wing flexibility and inertial effects, even varying in the latter case, which renders system (59)
though it is a controversial assumption among the flap- time-varying nature. That is, system (59) can be writ-
ping flight dynamics community [6,48]. Doing so, the ten as
123
Reconciliation of two higher-order averaging techniques
123
M. Maggia et al.
Table 2 Ratios of flapping frequency to natural frequency ω/ωn and eigenvalues assessing stability (using first- and second-order
averaging) for five insects
ω
Insect ωn λ1st λ2nd
Hawkmoth 28.78 [−11.89, − 3.30, 0.19 ± 5.74i] [−10.40, − 3.09, − 0.66 ± 3.72i]
Crane fly 50.62 [−47.71, − 17.31, − 1.13 ± 5.53i] [−45.76, − 16.26, − 13.16, 7.90]
Bumblebee 144.46 [−11.63, − 4.39, 1.58 ± 6.55i] [−11.26, − 4.37, 1.38 ± 6.17i]
Dragonfly 145.50 [−13.11, − 7.03, 1.34 ± 6.65i] [−12.56, − 6.98, 1.04 ± 5.99i]
Hoverfly 113.98 [−14.01, − 7.27, 2.13 ± 8.56i] [−13.37, − 7.24, 1.79 ± 7.92i]
As explained by Khan and Agrawal [60], the time systems are qualitatively different. Moreover, we can
variable in Eq. (60) is scaled as τ = ωωn t, where ωn is the even identify the vibrational stabilization mechanism
natural frequency of the body flight dynamics (short- comparing the matrices of the linearized direct (first-
period mode) and ω is the flapping (forcing) frequency. order) and second-order averaged systems for one of
When ωωn is large enough and assuming that all the the insects (e.g., hawkmoth)
terms in f and ga are equally scaled (which is not true in
general), the dynamics expressed in terms of τ are in the ⎡ ⎤
−3.59 0 0 −9.81
averaging-canonical form with = ωωn . It is important ⎢ 0 −3.30 0 0 ⎥
D(Λ1 )(0) = ⎢
⎣ 39.95
⎥,
to note that, for hovering insects with relatively low 0 −7.92 0 ⎦
flapping frequencies (e.g., for the hawkmoth, ωωn ≈ 30) 0 0 1 0
⎡ ⎤
the ratio ωωn has usually been considered high enough to −3.58 0 0 −9.81
⎢ 0 −3.09 0 0 ⎥
justify direct averaging (see, for example, [50,52,55]). D(Λ1 + 2 Λ2 )(0) = ⎢
⎣ 29.98
⎥.
0 −8.13 −28.45 ⎦
The averaged dynamics of Eq. (60) are written (see
−2.90 0 0.96 0
[50]) as
ẋ = f(x) + ga (x) (61)
As can be seen, the element (3, 4), which corresponds
where x is the averaged state vector and ga (x) repre- to pitching acceleration (or pitching moment) due to
sents the ga (x, t) averaged aerodynamic loads: ga (x) = a pitching angle, is changed from 0, in the crude-
1 T
T 0 ga (x, τ ) dτ . averaged system, to −28.45 in the second-order aver-
After applying second-order averaging to system aged system. That is, a significant pitch stiffness (whose
(60), we obtain a significant change in the stability lack was the main reason behind the claimed instabil-
characteristics in comparison with crude averaging. ity) is created due to high-frequency, high-amplitude,
Table 2 shows the ratios ωωn for five different insects and zero-mean terms in ga which is a typical vibrational
the eigenvalues λ1st and λ2nd , computed after apply- stabilization behavior. Figure 12 shows a simulation
ing first- and second-order averaging, respectively. The of the hawkmoth dynamics for an initial θ disturbance
eigenvalues are used to assess stability. of 30◦ . The curves represent the simulation of the full
Some important observations can be made about NLTP system, the first-order averaged dynamics and
the results in Table 2. First, for very-high-frequency the second-order averaged dynamics. The solution of
cases, first-order (direct averaging) and second-order the second-order averaged dynamics greatly improves
averaging give the same information about the aver- the approximation obtained by applying direct averag-
aged system and its stability such as in the cases of the ing only.
bumblebee, the dragonfly and the hoverfly. In fact, for This example shows the power of higher-order aver-
these insects the ratio ωωn is very large (larger than 100). aging in capturing hidden stabilization mechanisms in
Also, this result is consistent with that obtained by the dynamically rich systems in an analytic manner that
method of multiple scales [68]. On the other hand, when allows deep study of the underpinning physics. This
the ratio ωωn is relatively lower, such as in the case of scrutiny of the physics behind vibrational stabilization
the crane fly and of the hawkmoth, the stability char- in flapping flight and its experimental validation using
acteristics of the first-order and second-order averaged motion data from a real hawkmoth is presented in [69].
123
Reconciliation of two higher-order averaging techniques
1 0.4
to a natural stabilizing action without feedback. In addi-
0.5 0.2 tion to these applications which demonstrate the power
w (m/sec)
u (m/sec)
123
M. Maggia et al.
123
Reconciliation of two higher-order averaging techniques
11. Baillieul, J., Lehman, B.: Open-loop control using oscilla- 35. Nijmeijer, H., Van der Schaft, A.: Nonlinear Dynamical
tory inputs. In: CRC Control Handbook, pp. 967–980 (1996) Control Systems, vol. 175. Springer, Berlin (1990)
12. Markus, L., Yamabe, H.: Global stability criteria for differ- 36. Isidori, A.: Nonlinear Control Systems. Springer, Berlin
ential systems. Osaka Math. J. 12(2), 305–317 (1960) (2013)
13. Nayfeh, A.H., Mook, D.T.: Nonlinear Oscillations. Wiley, 37. Sastry, S.: Nonlinear Systems: Analysis, Stability, and Con-
Hoboken (2008) trol, vol. 10. Springer, Berlin (2013)
14. Nayfeh, A.H., Balachandran, B.: Applied Nonlinear 38. Sarychev, A.: Stability criteria for time-periodic systems via
Dynamics: Analytical, Computational, and Experimental high-order averaging techniques. In: Nonlinear Control in
Methods. Wiley, Hoboken (2008) the Year 2000, Lecture Notes in Control and Information
15. Nayfeh, A.H.: Perturbation Methods. Wiley, Hoboken Sciences, vol. 2, pp. 365–377. Springer (2001)
(2008) 39. Agrachev, A.A., Gamkrelidze, R., Sarychev, A.: Local
16. Khalil, H.K.: Nonlinear Systems. Prentice-Hall, New Jersey invariants of smooth control systems. Acta Appl. Math.
2(5), 1–5 (1996) 14(3), 191–237 (1989)
17. Sanders, J.A., Verhulst, F., Murdock, J.: Averaging: the peri- 40. Grobner, W., Knapp, H.: Contributions to the method of Lie
odic case. In: Averaging Methods in Nonlinear Dynamical series (1967)
Systems, pp. 21–44. Springer (2007) 41. Hale, J.: Ordinary Differential Equations. John Wiley, New
18. Murdock, J.A.: Perturbations: Theory and Methods, vol. 27. York (1969)
SIAM, Philadelphia (1999) 42. Mathematica file associated with the third and fourth order
19. Guckenheimer, J., Holmes, P.: Nonlinear Oscillations, proof. http://taha.eng.uci.edu/biblio-html/index.html
Dynamical Systems, and Bifurcations of Vector Fields, vol. 43. Kapitza, P.: Dynamic stability of a pendulum with an oscil-
42. Springer, Berlin (2013) lating point of suspension. J. Exp. Theor. Phys. 21(5), 588–
20. Berg, J.M., Wickramasinghe, I.M.: Vibrational control with- 597 (1951)
out averaging. Automatica 58, 72–81 (2015) 44. Taha, H.E., Tahmasian, S., Woolsey, C.A., Nayfeh, A.H.,
21. Bullo, F., Lewis, A.D.: Geometric control of mechanical M.R, Hajj: The need for higher-order averaging in the sta-
systems. In: Applied Mathematics. Springer, Berlin (2004) bility analysis of hovering, flapping-wing flight. Bioinspir.
22. Hassan, A.M., Taha, H.E.: Combined averaging-shooting Biomim. 10(1), 016,002 (2015)
approach for the analysis of flapping flight dynamics. J. 45. Stephenson, A.: Xx. On induced stability. Lond. Edinb.
Guid. Control Dyn. 41(2), 542–549 (2017) Dublin Philos. Mag. J. Sci. 15(86), 233–236 (1908)
23. Hassan, A.M., Taha, H.E.: Differential-geometric-control 46. Kapitza, P.L.: Dynamical stability of a pendulum when its
formulation for flapping flight multi-body dynamics. J. Non- point of suspension vibrates, and pendulum with a vibrating
linear Sci. 1–39 (2018) suspension. Collect. Pap. PL Kapitza 2, 714–737 (1965)
24. Sanders, J.A., Verhulst, F.: Averaging Methods in Nonlinear 47. Stabilization of an inverted pendulum under high-frequency
Dynamical Systems. Springer, Berlin (1985) excitation (Kapitza pendulum). https://www.youtube.com/
25. Bogoliubov, N.N., Mitropolsky, Y.A., Gillis, J.: Asymptotic watch?v=is_ejYsvAjY
methods in the theory of non-linear oscillations. Phys. Today 48. Sun, M.: Insect flight dynamics: stability and control. Rev.
16, 61 (1963) Mod. Phys. 86(2), 615 (2014)
26. Mitropolsky, I.A.: Averaging method in non-linear mechan- 49. Nelson, R.C.: Flight Stability and Automatic Control.
ics. Int. J. Non-linear Mech. 2(1), 69–96 (1967) McGraw-Hill, New York (1989)
27. Vela, P.A.: Averaging and Control of Nonlinear Systems. 50. Taha, H.E., Hajj, M.R., Nayfeh, A.H.: On the longitudinal
Ph.D. Thesis, California Institute of Technology (2003) flight dynamics of hovering MAVs/insects. J. Guid. Control
28. Agrachev, A., Gamkrelidze, R.: Chronological algebras and Dyn. 37(3), 970–978 (2014)
nonstationary vector fields. J. Math. Sci. 17(1), 1650–1675 51. Alexander, D.E.: Nature’s Flyers: Birds, Insects, and the
(1981) Biomechanics of Flight. JHU Press (2002)
29. Agračev, A., Gamkrelidze, R.V.: The exponential represen- 52. Taylor, G.K., Thomas, A.L.R.: Animal flight dynamics II.
tation of flows and the chronological calculus. Math. USSR- Longitudinal stability in flapping flight. J. Theor. Biol. 214,
Sbornik 35(6), 727 (1979) 351–370 (2002)
30. Fliess, M., Lamnabhi, M., Lamnabhi-Lagarrigue, F.: An 53. Taylor, G.K., Thomas, A.L.: Dynamic flight stability in the
algebraic approach to nonlinear functional expansions. IEEE desert locust schistocerca gregaria. J. Exp. Biol. 206(16),
Trans. Circuits Syst. 30(8), 554–570 (1983) 2803–2829 (2003)
31. Sarychev, A.: Stability criteria for time-periodic systems via 54. Sun, M., Xiong, Y.: Dynamic flight stability of a hovering
high-order averaging techniques. In: Nonlinear Control in bumblebee. J. Exp. Biol. 208(3), 447–459 (2005)
the Year 2000, vol. 2, pp. 365–377. Springer (2001) 55. Sun, M., Wang, J., Xiong, Y.: Dynamic flight stability of
32. Nayfeh, A.H.: Introduction to Perturbation Techniques. hovering insects. Acta Mech. Sin. 23(3), 231–246 (2007)
Wiley, Hoboken (1981) 56. Xiong, Y., Sun, M.: Dynamic flight stability of a bumblebee
33. Yagasaki, K., Ichikawa, T.: Higher-order averaging for peri- in forward flight. Acta Mech. Sin. 24(1), 25–36 (2008)
odically forced weakly nonlinear systems. Int. J. Bifurc. 57. Gao, N., Aono, H., Liu, H.: A numerical analysis of dynamic
Chaos 9(03), 519–531 (1999) flight stability of hawkmoth hovering. J. Biomech. Sci. Eng.
34. Sarychev, A.V.: Lie-and chronologico-algebraic tools for 4(1), 105–116 (2009)
studying stability of time-varying systems. Syst. Control 58. Faruque, I., Humbert, J.S.: Dipteran insect flight dynam-
Lett. 43(1), 59–76 (2001) ics. Part 1. Longitudinal motion about hover. J. Theor. Biol
264(2), 538–552 (2010)
123
M. Maggia et al.
59. Cheng, B., Deng, X.: Translational and rotational damping 66. Popovic, P., Nayfeh, A.H., Oh, K., Nayfeh, S.A.: An
of flapping flight and its dynamics and stability at hovering. experimental investigation of energy transfer from a high-
IEEE Trans. Robot. 27(5), 849–864 (2011) frequency mode to a low-frequency mode in a flexible struc-
60. Khan, Z.A., Agrawal, S.K.: Control of longitudinal flight ture. Modal Anal. 1(1), 115–128 (1995)
dynamics of a flapping-wing micro air vehicle using time- 67. Oh, K., Nayfeh, A.: High-to low-frequency modal interac-
averaged model and differential flatness based controller. In: tions in a cantilever composite plate. J. Vib. Acoust. 120(2),
American Control Conference, 2007. ACC’07, pp. 5284– 579–587 (1998)
5289. IEEE (2007) 68. Taha, H.E., Nayfeh, A.H., Hajj, M.R.: Effect of the
61. Schenato, L., Campolo, D., Sastry, S.: Controllability issues aerodynamic-induced parametric excitation on the lon-
in flapping flight for biomimetic micro aerial vehicles gitudinal stability of hovering MAVs/insects. Nonlin-
(MAVs). In: Proceedings of the 42nd IEEE Conference on ear Dyn. 78, 2399–2408 (2014). https://doi.org/10.1007/
Decision and Control, vol. 6, pp. 6441–6447. IEEE (2003) s11071-014-1596-6
62. Nayfeh, S., Nayfeh, A.: Nonlinear interactions between two 69. Taha, H.E., Kiani, M., Hedrick, T.L., Greeter, J.S.M.: Vibra-
widely spaced modes—external excitation. Int. J. Bifurc. tional control: A hidden stabilization mechanism in insect
Chaos 3(02), 417–427 (1993) flight (under review). Nat. Commun
63. Nayfeh, S., Nayfeh, A.: Energy transfer from high-to low- 70. Taha, H.E., Hajj, M.R., Beran, P.S.: Unsteady nonlinear
frequency modes in a flexible structure via modulation. J. aerodynamics of hovering MAVs/insects. In: AIAA-Paper
Vib. Acoust. 116(2), 203–207 (1994) 2013–0504 (2013)
64. Nayfeh, A., Mook, D.: Energy transfer from high-frequency
to low-frequency modes in structures. J. Vib. Acoust.
117(B), 186–195 (1995) Publisher’s Note Springer Nature remains neutral with regard
65. Nayfeh, A.H., Chin, C.M.: Nonlinear interactions in a para- to jurisdictional claims in published maps and institutional affil-
metrically excited system with widely spaced frequencies. iations.
Nonlinear Dyn. 7(2), 195–216 (1995)
123