You are on page 1of 11

CHAOS 23, 023101 (2013)

Attracting and repelling Lagrangian coherent structures from a single


computation
Mohammad Farazmand1,2 and George Haller2
1
Department of Mathematics, ETH Z urich, 8092 Zurich, Switzerland
2
Institute for Mechanical Systems, ETH Zurich, 8092 Zurich, Switzerland

(Received 18 January 2013; accepted 19 March 2013; published online 12 April 2013)
Hyperbolic Lagrangian Coherent Structures (LCSs) are locally most repelling or most attracting
material surfaces in a finite-time dynamical system. To identify both types of hyperbolic LCSs at
the same time instance, the standard practice has been to compute repelling LCSs from future data
and attracting LCSs from past data. This approach tacitly assumes that coherent structures in the
flow are fundamentally recurrent, and hence gives inconsistent results for temporally aperiodic
systems. Here, we resolve this inconsistency by showing how both repelling and attracting LCSs
are computable at the same time instance from a single forward or a single backward run. These
LCSs are obtained as surfaces normal to the weakest and strongest eigenvectors of the Cauchy-
C 2013 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4800210]
Green strain tensor. V

Repelling and attracting Lagrangian coherent structures 2012; Haller and Beron-Vera, 2012) are now available for
(LCSs) are material surfaces that govern mixing patterns their extraction from flow data.
in complex dynamical systems. Recent developments All available hyperbolic LCS methods fundamentally
made the accurate computation of both types of struc- seek locations of large particle separation. They will high-
tures possible, but not for the same data set: repelling light repelling LCS positions at some initial time t a from
LCSs are invariably obtained from future data, and a forward-time analysis of the flow over a finite time-interval
attracting LCSs from past data. For temporally aperiodic [a, b]. Similarly, these methods reveal attracting LCSs at the
flows, this practice locates repelling and attracting LCSs final time t b from a backward-time analysis of the flow
for two different finite-time dynamical systems. Here, we over [a, b]. The complete hyperbolic LCS distribution at a
resolve this inconsistency by showing that both types of fixed time t 2 a; b# is, therefore, not directly available.
LCSs can be computed at the same time instance from
Two main approaches have been employed to resolve
the same data set.
this issue (see Figure 1 for an illustration):
1. Approach I: Divide the finite time interval of interest as
a; b# a; t0 # [ t0 ; b#. Compute repelling LCSs from a
I. INTRODUCTION
forward run over t0 ; b#, and attracting LCSs from the
The differential equations governing a number of physi- backward run over a; t0 # (see, e.g., Lekien and Ross
cal processes are only known as observational or numerical (2010); Lipinski and Mohseni (2010)). Both repelling and
data sets. Examples include oceanic and atmospheric particle attracting LCSs are then obtained at the same time slice
motion, whose velocity field is only known at discrete loca- t0. However, they correspond to two different finite-time
tions, evolving aperiodically over a finite time-interval of dynamical systems: one defined over a; t0 # and the other
availability. For such temporally aperiodic data sets, classic over t0 ; b#. This approach works well for a roughly T-per-
dynamical conceptssuch as fixed points, periodic orbits, sta- iodic system, when t0 $ a and b $ t0 are integer multiples
ble and unstable manifolds, or chaotic attractorsare either of T. In general, however, hyperbolic LCSs computed
undefined or nongeneric. over a; t0 # and over t0 ; b# do not evolve into each other as
Instead of relying on classic concepts, one may seek in- t0 is varied, and hence the resulting structures are not
fluential surfaces responsible for the formation of observed dynamically consistent. In addition, one cannot identify
trajectory patterns over a finite time frame of interest. Such a attracting LCSs at time a or repelling LCSs at time b from
surface is necessarily a material surface, i.e., a codimension- this approach.
one set of initial conditions evolving with the flow. Among 2. Approach II: Extract repelling LCSs at the initial time a
material surfaces, an attracting LCS is defined as a locally from a forward run over [a, b]; extract attracting LCSs at
most attracting material surface in the phase space (Haller the final time b from a backward run over [a, b]. Obtain
and Yuan, 2000; Haller, 2011). Repelling LCSs are defined repelling LCSs at any time t0 2 a; b# by advecting repel-
as locally most repelling material surfaces, i.e., attracting ling LCSs from a to t0 under the flow. Similarly, obtain
LCSs in backward-time. Repelling and attracting LCSs attracting LCSs at any time t0 2 a; b# by advecting
together are referred to as hyperbolic LCSs. Both heuristic attracting LCSs from b to t0 under the flow. This approach
detection methods (Peacock and Dabiri, 2010) and rigorous identifies LCSs based on the full available data, and pro-
variational algorithms (Haller, 2011; Farazmand and Haller, vides dynamically consistent surfaces that evolve into

1054-1500/2013/23(2)/023101/11/$30.00 23, 023101-1 C 2013 AIP Publishing LLC


V
023101-2 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

FIG. 1. Schematic illustration of


approach I (a) and approach II (b) in the
extended phase space.

each other as t0 varies (Haller, 2011; Farazmand and Haller, We demonstrate the results on three examples: an auton-
2012). Since the forward-time advection of a repelling LCS omous Duffing oscillator (Sec. V A), a direct numerical sim-
(as well as the backward-time advection of an attracting ulation of two-dimensional turbulence (Sec. V B), and the
LCS) is numerically unstable (see Figure 2), this approach three-dimensional classic ABC flow (Sec. V C).
requires extra care to suppress growing instabilities
(Farazmand and Haller, 2012). Even under well-controlled II. PRELIMINARIES AND NOTATION
instabilities, however, a further issue arises in near-
incompressible flows: repelling LCSs shrink exponentially Consider the dynamical system
under forward-advection, and attracting LCSs shrink expo- x_ ux; t; x 2 U ' Rn ; t 2 I a; b#; (1)
nentially under backward-advection. Therefore, while the
LCSs obtained in this fashion are dynamically consistent, where u : U ( I ! Rn is a sufficiently smooth velocity field.
they require substantial numerical effort to extract and may For t0 ; t 2 I, define the flow map
still reveal little about the dynamics.
Ftt0 : U ! U
Here, we develop a new approach that keeps the dynam- (2)
x0 7! xt; t0 ; x0 ;
ical consistency of approach II but eliminates the instability
and shrinkage of advected LCSs. Our key observation is that
as the unique one-to-one map that takes the initial condition
attracting LCSs can also be recovered as codimension-one
x0 to its time-t position xt; t0 ; x0 under system (1).
hypersurfaces normal to the weakest eigenvector field of the
The forward CauchyGreen strain tensor over the time
forward Cauchy-Green strain tensor. These stretch-surfaces
interval I is defined in terms of the flow gradient rFba as
are obtained from the same forward-time calculation that
reveals repelling LCSs as strain-surfaces, i.e., codimension- ! ">
Cf rFba rFba : (3)
one surfaces normal to the dominant eigenvector of the for-
ward Cauchy-Green strain tensor (Farazmand and Haller,
At each initial condition x0 2 U, the tensor Cf x0 is repre-
2012). The locally most compressing strain-surfaces and the
sented by a symmetric, positive definite, n ( n matrix with
locally most expanding stretch-surfaces then reveal repelling
and attracting LCSs at the same initial time a based on a sin- an orthonormal set of eigenvectors fnfk x0 g1)k)n , and with a
gle forward-time calculation over [a, b]. corresponding set of eigenvalues fkfk x0 g1)k)n satisfying

FIG. 2. The errors in the computation of


a repelling LCS grow exponentially as
the LCS is advected forwards in time.
The same statement holds for the
backward-time advection of an attracting
LCS.
023101-3 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

Cf x0 nfk x0 kfk x0 nfk x0 ; k 2 f1; 2; ; ng; (4a) time direction they are advected in. This instability can only
be controlled by employing a high-end numerical integrator
0 < kf1 x0 ) kf2 x0 ) * * * ) kfn x0 : (4b) which refines the advected surface when large stretching
develops. Even under high-precision advection, however, the
These invariants of the CauchyGreen strain tensor charac- end-result is an exponentially shrinking surface which only
terize the deformation experienced by trajectories starting captures subsets of the most influential material surfaces.
close to x0. If a unit sphere is placed at x0, its image under the
linearized flow map rFba will be an ellipsoid whose principal
axes align with the eigenvectors fnfk x0 g1)k)n and have IV. MAIN RESULT
corresponding lengths fkfk x0 g1)k)n . Here, we present a direct method to identify both attract-
Similarly, the backward CauchyGreen strain tensor ing and repelling LCSs at the same time instance, using the
over the time interval I is defined as same finite time-interval. These surfaces, therefore, are based
on the assessment of the same finite-time dynamical system,
Cb rFab > rFab : (5)
avoiding the dynamical inconsistency we reviewed for
approach I in the Introduction.
Its eigenvalues fkbk x0 g1)k)n and orthonormal eigenvectors
In particular, we show that the initial position of an
fnbk x0 g1)k)n satisfy similar properties as those in Eq. (4).
attracting LCS, Aa, is everywhere orthogonal to the weak-
Their geometric meaning is similar to that of the invariants
est eigenvector nf1 of the tensor Cf . This, together with the
of Cf, but in backward time.
orthogonality of the initial repelling LCS position Ra to
the dominant eigenvector nfn of Cf , allows for the simultane-
III. REPELLING AND ATTRACTING LCSs ous construction of attracting and repelling LCSs at time
A repelling LCS over the time interval I is a t a, utilizing the same time interval [a, b]. All this renders
codimension-one material surface that is pointwise more the computation of the backward CauchyGreen strain tensor
repelling over I than any nearby material surface. If Rt Cb unnecessary.
represents the time-t position of such a LCS, then the initial Similarly, if locating the hyperbolic structures based on
LCS position Ra must be everywhere orthogonal to the past flow data (now-casting) is of interest, current positions
most-stretching eigenvector nfn of the forward CauchyGreen of repelling and attracting LCSs can be located from a single
strain tensor Cf (Haller, 2011; Haller and Beron-Vera, backward-time integration starting from the present time. In
2012). Specifically, we must have this case, the current position of an attracting LCS is every-
where orthogonal to the dominant eigenvecor nbn of the back-
Txa Ra ? nfn xa ; (6) ward CauchyGreen strain tensor Cb. Likewise, the current
positions of repelling LCSs are everywhere orthogonal to the
for any point xa 2 Ra, where Txa Ra denotes the tangent weakest eigenvector nb1 of the tensor Cb.
space of Ra at point xa. We start with definitions of the surfaces involved in our
Similarly, an attracting LCS over the time interval I is a results:
codimension-one material surface that is pointwise more Definition 1 (Strain-surface). Let Mt be an (n $ 1)-
attracting over I than any nearby material surface. If At is dimensional smooth material surface in U, evolving under
the time-t position of an attracting LCS, its final position the flow map over the time interval I [a, b] as
Ab satisfies Mt Fta Ma. Denote the tangent space of M at a
point x 2 M by Tx M.
Txb Ab ? nbn xb ; (7)
(i) Mt is called a forward strain-surface if Ma is
for all points xb 2 Ab. That is, the time-b position of everywhere normal to the eigenvector field nfn , i.e.,
attracting LCS is everywhere orthogonal to the eigenvector Txa Ma ? nfn xa ; 8xa 2 Ma:
nbn of the backward CauchyGreen strain tensor Cb .
The relation (6) enables the construction of repelling (ii) Mt is called a backward strain-surface if Mb is
LCS candidates at time t a, while Eq. (7) enables the con- everywhere normal to the eigenvector field nbn , i.e.,
struction of attracting LCS candidates at the final time t b
(see, e.g., Farazmand and Haller (2012); Hadjighasem,
Txb Mb ? nbn xb ; 8xb 2 Mb:
Farazmand, and Haller (2012)). Since LCSs are constructed
as material surfaces, they move with the flow. Therefore,
Strain-surfaces are generalizations of the strainlines
LCS positions at an intermediate time t0 2 a; b# are, in prin-
introduced in Farazmand and Haller (2012) and Haller and
ciple, uniquely determined by their end-positions
Beron-Vera (2012) in the theory of hyperbolic LCSs for two-
Rt0 Fta0 Ra; At0 Ftb0 Ab: (8) dimensional flows. By contrast, the stretch-surfaces appear-
ing in the following definition have not yet been used even in
As discussed in the introduction, however, using the two-dimensional LCS detection.
advection formulae (8) leads to numerical instabilities. This Definition 2 (Stretch-surface). Let Mt be an (n $ 1)-
is because the material surfaces involved are unstable in the dimensional material surface as in Definition 1.
023101-4 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

(i) Mt is called a forward stretch-surface if Ma is (i) A repelling LCS, Rt, is a forward strain-surface,
everywhere normal to the eigenvector field nf1 , i.e., i.e., Ra is everywhere orthogonal to the eigenvector
field nfn . Furthermore, Rt is also a backward stretch-
Txa Ma ? nf1 xa ; 8xa 2 Ma: surface, i.e., Rb is everywhere orthogonal to the
eigenvector field nb1 .
(ii) Mt is called a backward stretch-surface if Mb is (ii) An attracting LCS, At, is a forward stretch-surface,
everywhere normal to the eigenvector field nb1 , i.e., i.e., Aa is everywhere orthogonal to the eigenvector
field nf1 . Furthermore, At is also a backward strain-
surface, i.e., Ab is everywhere orthogonal to the
Txb Mb ? nb1 xb ; 8xb 2 Mb:
eigenvector field nbn .
By definition, the local orientation of a forward strain- Among other things, the above corollary enables the vis-
surface is known at the initial time t a. The following ualization of attracting and repelling LCSs simultaneously at
theorem determines the local orientation of the same strain- the initial time t a of a finite time-interval [a, b] over
surface at the final time t b, rendering the forward- which the underlying dynamical system is known (see
advection of the surface unnecessary. The same theorem Sec. V below for examples). This only requires the computa-
provides the local orientation of backward strain-surfaces at tion of the forward-time CauchyGreen strain tensor Cf , ren-
the initial time t a (see Figure 3 for an illustration). dering backward-time computations unnecessary. Similarly,
Theorem 1. the simultaneous visualization of attracting and repelling
LCSs at the final time t b is possible by only one
(i) Forward strain-surfaces coincide with backward
backward-time computation.
stretch-surfaces.
We finally comment on the relationship between for-
(ii) Backward strain-surfaces coincide with forward
ward strain-surfaces and classic stable manifolds. If a stable
stretch-surfaces.
manifold exists for a trajectory through a given initial point
Proof. See Appendix A. w p, then the forward strain-surface at p aligns asymptotically
The following corollary summarizes the implications of with the t a slice of the stable manifold at p. This is
Theorem 1, along with known results from Haller (2011) and because the stable subspace at p becomes asymptotically the
Farazmand and Haller (2012). most repelling direction in forward time for all trajectories in
Corollary 1. Let Rt and At be, respectively, repel- the stable manifold. At the same time, trajectories in the
ling and attracting LCSs of the dynamical system (1). Then, unstable manifold through p have no known forward-time
the following hold: asymptotic behavior. Thus, in the general case, forward

FIG. 3. (a) A forward strain-surface


evolves into a backward stretch-surface.
(b) A forward stretch-surface evolves
into a backward strain-surface.
023101-5 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

stretch-surfaces will not align with t a slices of unstable r0 s nf1 rs; r0 r0 ; (10)
manifolds. A similar asymmetry holds in backward time:
backward strain-surfaces align with unstable manifolds, but where r : s 7! rs denotes parametrization by arc-length.
backward stretch-surfaces do not align with classic stable Similarly, forward stretchlines are trajectories of the ODE
manifolds. We will see illustrations of these asymmetries in
the examples below. p0 s nf2 ps; p0 p0 ; (11)

with p : s 7! ps denoting an arclength-parametrization.


V. EXAMPLES
Since we are interested in the de facto finite-time stable and
Here, we demonstrate the application of Corollary 1 on unstable manifolds passing through the hyperbolic fixed
three examples: the classic Duffing oscillator, a two- point (0, 0), we set r0 p0 0; 0.
dimensional turbulence simulation, and the classic ABC We observe that as the integration time T increases, the
flow. In the two-dimensional case (i.e., n 2), we refer to unique strainline and the unique stretchline through the ori-
strain- and stretch-surfaces as strainlines and stretchlines, gin converge to their asymptotic limits. Figure 5 shows the
respectively. convergence of these curves around the hyperbolic fixed
point (0, 0). For integration times T , 2, the computed
A. Duffing oscillator
strainlines and stretchlines are virtually indistinguishable
from their asymptotic limits. Therefore, in the following, we
Here, we show that even for a two-dimensional autono- fix the integration time T b a 2 with a 0 and b 2.
mous system, stretchlines and strainlines act as de facto sta- Note that while the strainline is indistinguishable from
ble and unstable manifolds over finite time intervals. Indeed, the stable manifold, the stretchline differs from the unstable
over such intervals, sets of initial conditions will be seen to manifold (see Figure 5(c)). Stretchlines as de facto finite-
follow stretchlines in forward time. Only asymptotically do time unstable manifolds define the directions along which
these initial conditions align with the well-known classic passive tracers are observed to stretch. To demonstrate this,
unstable manifolds.
Consider the unforced and undamped Duffing oscillator

x_ 1 x2 ;
(9)
x_ 2 4x1 $ x31 ;

whose Hamiltonian Hx1 ; x2 12 x41 $ 4x21 x22 is conserved


along the trajectories (see Figure 4). The hyperbolic fixed
point (0, 0) of the system admits two homoclinic orbits
(shown in red), which coincide with the stable and unstable
manifolds of the fixed point.
By Definition 1, forward strainlines over a finite time
interval are everywhere orthogonal to the eigenvector field
nf2 of the forward strain tensor Cf . As a result, strainlines are
trajectories of the autonomous ordinary differential equation
(ODE)

FIG. 5. (a) Forward stretchline through the origin for three integration times
T 0.5 ($ ( $), T 1 ($w$) and T 2 ($!$). (b) Forward strainline for
the same integration times, as in panel (a). (c) The asymptotic position of
the strainline ($- $) and the stretchline ($- $) compared to the classic stable
FIG. 4. Trajectories of system (9). The homoclinic orbits are shown in red. and unstable manifolds (black).
023101-6 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

in Figure 6, three disks with radii 10$3 , 5 ( 10$3 , and 10$2 The forcing f is random in phase and active over the wave
are initially centered at the origin. For short advection times, numbers 3:5 < k < 4:5. The initial condition u0 is the instan-
the tracers elongate in the direction of the stretchline, not the taneous velocity field of a decaying turbulent flow. We solve
unstable manifold. Unlike the classic unstable manifold, equations (12) by a standard pseudo-spectral method with
stretchlines evolve in time and only become invariant when 512 ( 512 modes. The time integration is carried out by a
viewed in the extended phase space of the (x, t) variables. 4th order RungeKutta method with adaptive step-size
For longer advection times (not presented here), the stretch- (MATLABs ODE45). Equation (12) is solved over the time
line converges to the unstable manifold and becomes virtu- interval I [0, 50].
ally indistinguishable from it. One can, in principle, compute an attracting LCS at the
beginning of a time interval I [a, b] by advecting the
B. Two-dimensional turbulence attracting LCS extracted at t b back to t a. As mentioned
in the Introduction, however, this process is numerically
We consider a two-dimensional velocity field unstable since attracting LCSs become unstable in backward
u : U ( R ! R2 , obtained as a numerical solution of the time. Their instability is apparent in Figure 7, where an
NavierStokes equations attracting LCS (red) is advected backwards from t 50 to
the initial time t 0. The advected curve is noisy and devi-
@t u u * ru $rp !Du f ;
ates from the true pre-image (blue curve). The true pre-
r * u 0; (12) image, the stretchline, is computed as a trajectory of the
ux; 0 u0 x: eigenvector filed nf2 of the forward CauchyGreen strain
tensor Cf .
The domain U 0; 2p# ( 0; 2p# is periodic in both spatial We now extract the set of attracting LCSs that shape
directions. The non-dimensional viscosity ! is equal to 10$5 . observed global tracer patterns in this turbulent flow.

FIG. 6. (a) Classical stable and unstable


manifolds (black) are shown together
with the stretchline through the origin
(magenta). Three blobs of tracers with
radii 10$3 (blue), 5 ( 10$3 (yellow),
and 10$2 (red) are centered at the origin.
The tracers and the manifolds are then
advected to time t 0.1 (b), t 0.2 (c),
and t 0.4 (d). Over the time interval
[0, 2], the stretchline is the de facto
unstable manifold for spreading tracers.
For larger advection times, this de facto
unstable manifold practically coincides
the classic unstable manifold of the
origin.
023101-7 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

Algorithm 1.
1. Compute the CauchyGreen strain tensor Cf over a uni-
form grid.
2. Locate elliptic barriers by the procedure described in
Haller and Beron-Vera (2012); Hadjighasem, Farazmand,
and Haller (2012).
3. Compute stretchlines as trajectories of Eq. (11). The ini-
tial conditions p0 are chosen from a uniform grid over the
phase space.
FIG. 7. Stretchline (blue) and the advected image of an attracting LCS (red)
at t 0. The exponential growth of errors in backward-time advection of the 4. Stop the stretchline integration once the stretchlines reach
LCS results in a jagged curve that deviates from the true attracting LCS. either a singular point or an elliptic region bounded by an
elliptic barrier.
Corollary 1 establishes that such LCSs are necessarily for- 5. For each stretchline so obtained, compute the relative
ward stretchlines, i.e., trajectories of Eq. (11). It then remains stretching (13).
to select the trajectories of this ODE that stretch more under 6. Locate attracting LCSs as the stretchlines with locally
forward advection than any neighboring stretchline (Haller maximal relative stretching.
and Beron-Vera, 2012).
The relative stretching of a material line is defined as To illustrate the defining role of stretchlines in the for-
the ratio of its length at the final time t b to its initial length mation of turbulent mixing patterns, we consider three con-
at time t a. For a forward-time stretchline c, one can show centric circles of tracers with radii 0.05, 0.1, and 0.2 at the
(see Appendix B) that the relative stretching is given by initial time t a 0 (see Figure 8). The circles are centered
qffiffiffiffi on a stretchline with locally largest relative stretching (black
1 curve). Then, the stretchlines and tracers are advected to
qc kf2 ds; (13)
c c times t0 10, t0 15, and t0 25. In each case, we find
that the tracer pattern stretches and aligns with the evolving
where c is the length of c at time t a. Note that no mate- stretchline, as expected.
rial line advection is required for computing the relative We now turn to the global geometry of the attracting
stretching in Eq. (13). LCSs. Figure 9(a) shows stretchlines computed from a uniform
In order to locate the stretchlines that locally maximize grid of 30 ( 30 points. Attracting LCSs at time t 0, extracted
the relative stretching (13), we adopt the numerical proce- as stretchlines with the locally largest relative stretching, are
dure outlined in Haller and Beron-Vera (2012) for locating highlighted in red. Also shown are the elliptic barriers (greed
the locally least-stretching strainlines. Specifically, we first closed curves), as well as a select set of blue tracer disks that
compute a dense enough set of stretchlines as the trajectories will be used to illustrate the role of attracting LCSs. The
of ODE (11). We stop the integration once the stretchline advected positions of attracting LCSs, elliptic barriers, and
reaches a singularity of the tensor field Cf or crosses an ellip- tracer disks are shown in Figure 9(b). Note how the attracting
tic transport barrier. LCSs govern the deformation of the tracer disks in the turbu-
A singularity of Cf is a point where Cf equals the iden- lent mixing region. Meanwhile, the elliptic barriers keep their
tity tensor, and hence its eigenvectors are not uniquely coherence by preserving their arclength and enclosed area.
defined (see Delmarcelle and Hesselink (1994) and Tricoche,
Scheuermann, and Hagen (2000) for more details). An ellip- C. ABC flow
tic barrier is the outermost member of a nested set of closed
curves that preserve their initial length (at time t a) under In two dimensions, stretchlines are constructed as trajec-
advection up to time t b (Haller and Beron-Vera, 2012). In tories of the eigenvector field nf2 . The resulting curves are, by
an incompressible flow, an elliptic barrier also preserves its construction, everywhere orthogonal to the eigenvector field
enclosed area under advection, and hence the elliptic domain nf1 . In higher dimensions, however, constructing stretch-
it encloses remains highly coherent. For this reason, elliptic surfaces that are everywhere orthogonal to the eigenvector
barriers can be considered as generalizations of outermost nf1 is nontrivial. In fact, for a given eigenvector field, such a
KolmogorovArnoldMoser (KAM) curves generically surface may only exists locally if a Frobenius condition is
observed in temporally periodic two-dimensional flows satisfied (Lee, 2009). This condition requires the eigenvec-
(Haller and Beron-Vera, 2012). tors spanning the tangent space of the manifold (here,
We locate elliptic barriers using the detection algorithm
developed in Haller and Beron-Vera (2012) and Hadjighasem, fnfk g2)k)n ) to be in involution, i.e., their Poisson brackets
Farazmand, and Haller (2012). With the location of these bar- nfi ; nfj # should be in the tangent space of the manifold for any
riers and of the singularities of Cf at hand, stretchlines are i; j 2 f2; 3; ; ng.
truncated to compact line segments, rendering the integral in Even when the subset of the phase space satisfying this
Eq. (13) well-defined. Attracting LCSs at t a are then located Frobenius condition is known, constructing stretch-surfaces
as stretchline segments that have higher relative stretching (13) globally as smooth parametrized manifolds normal to a spe-
than any of their C1-close neighbors. This process is briefly cific vector field is challenging (Palmerius, Cooper, and
summarized in the following algorithm: Ynnerman, 2009; Balzer, 2012). Here, we only illustrate that
023101-8 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

FIG. 8. (a) The concentric tracers with


radii 0.05 (blue), 0.1 (yellow), and 0.2
(red). The stretchline (black) passing
through the center is computed from the
time interval [0, 50] (i.e., a 0 and
b 50). The tracers and the stretchline
are then advected forward in time to t 10
(b), t 15 (c), t 25 (d).

FIG. 9. (a) Forward stretchlines at t 0.


The attracting LCSs (i.e., locally most-
stretching stretchlines) are highlighted in
red. The green closed curves show the
boundaries of elliptic regions. Tracers
(blue circles) are used to visualize the
overall mixing patterns. (b) Advected
image of the attracting LCSs, tracers,
and elliptic barriers at time t 50.

locally constructed stretch-surfaces do govern the formation advected images of the tracer and the plane at time t 4 are
of tracer patterns in three-dimensional flows as well. shown in Figure 10(b), demonstrating that the stretch-surface
We use the classic ABC flow (Arnold and Khesin, 1998) through the center of the tracer blob acts as a de facto unsta-
ble manifold in this three-dimensional example as well.
x_ 1 A sinx3 C cosx2 ;
x_ 2 B sinx1 A cosx3 ; (14) VI. CONCLUSIONS
x_ 3 C sinx2 B cosx1 ; We have shown that both repelling and attracting LCSs
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffi (finite-time stable and unstable manifolds) at a time instance
with A 1, B 2=3, and C 1=3. The Cf strain tensor t a can be extracted from a single forward-time computa-
is computed over the time interval I [0, 4] (i.e., a 0 and tion over a time interval I [a, b]. This extraction requires
b 4). We release a spherical blob of initial conditions cen- the computation of the eigenvectors of the forward
tered at p; p with radius 0.1. We approximate the stretch- CauchyGreen strain tensor Cf . It has been found previously
surface passing through this point by the plane normal to the (Haller, 2011; Haller and Beron-Vera, 2012) that at time
first eigenvector nf1 of Cf. Figure 10(a) shows this plane t a, the position of repelling LCSs are strain-surfaces, i.e.,
together with the sphere of tracers at time t 0. The are everywhere orthogonal to the dominant eigenvector of
023101-9 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

FIG. 10. (a) A spherical tracer surface


(blue) at time t 0 and the corresponding
approximate stretch-surface (red) passing
through its origin. (b) The advected posi-
tions of these surfaces at the final time
t 4.

Cf . Here, we proved that the t a positions of attracting reciprocal of the smallest eigenvalue kb1 of the backward-
LCSs are stretch-surfaces, i.e., are everywhere orthogonal to time strain tensor Cb at the point xb Fba xa , i.e.,
the weakest eigenvector of Cf .
The attracting LCSs obtained in this fashion are observed 1
kfn xa : (A1)
as centerpieces around which tracer patterns develop. Even in kb1 xb
autonomous dynamical systems, these evolving centerpieces
of trajectory evolution differ from classic unstable manifolds, Similarly, we have
forming de facto unstable manifolds over finite times.
1
In two-dimensional dynamical systems, stretchlines can kbn xb : (A2)
be directly computed as most-stretching trajectories of the kf1 xa
autonomous ODE (11). In higher dimensions, stretch-
surfaces satisfy linear systems of partial differential equa- Proof. This follows directly from Eq. (13) in Haller and
tions (PDEs), as any surface normal to a given vector field Sapsis (2011). w
does (Palmerius, Cooper, and Ynnerman, 2009). While a Lemma 2. For any xa 2 U, the following identities hold
self-consistent global solution of these PDEs remains for any k 2 f1; 2; ; ng
numerically challenging, here we have illustrated the local
organizing role of stretch-surfaces through the advection of hnfn xa ; rFab xb nbk xb i
their tangent spaces in the classic ABC flow. Results on the kfn xa kbk xb hnfn xa ; rFab xb nbk xb i; (A3)
construction of attracting LCSs from globally computed
stretch-surfaces will be reported elsewhere. hnbn xb ; rFba xa nfk xa i
kbn xb kfk xa hnbn xb ; rFba xa nfk xa i; (A4)
ACKNOWLEDGMENTS

G.H. acknowledges partial support by the Canadian where h*; *i is the Euclidean inner product between two
NSERC under Grant No. 401839-11. vectors.
Proof. We prove identity (A3). The proof of Eq. (A4) is
similar and will be omitted.
APPENDIX A: PROOF OF THEOREM 1
! First, "note that since the flow map is invertible, we have
In order to prove Theorem 1, we need two lemmas. The Fab Fba xa xa for any xa 2 U. Differentiating this identity
first lemma draws a connection between eigenvalues of the for- with respect to xa, we obtain
ward- and backward-time CauchyGreen strain tensors. The
second lemma establishes a relation between their eigenvectors. rFab xb rFba xa #$1 : (A5)
Lemma 1. The largest eigenvalue kfn of the forward-time
strain tensor Cf at a point xa 2 U coincides with the The result then follows from the identity

hnfn xa ; rFab xb nbk xb i hnfn xa ; rFab xb #$> rFab xb #> rFab xb nbk xb i
hrFab xb #$1 nfn xa ; Cb xb nbk xb i
kbk xb hrFba xa nfn xa ; nbk xb i
kbk xb hrFba xa #$> rFba xa #> rFba xa nfn xa ; nbk xb i
kbk xb hCf xa nfn xa ; rFba xa #$1 nbk xb i
kfn xa kbk xb hnfn xa ; rFab xb nbk xb i;
023101-10 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

where we have used identity (A5) twice. w which implies that hnb1 xb ; rFba xa nfk xa i 0 for
Now we turn to the proof of Theorem 1. k 2 f1; 2; ; n $ 1g away from the degenerate points
Proof of Theorem 1: where kfn kfn$1 . w
(ii) The proof is identical to that of part (i).
(i) Assume that Mt is a backward stretch-surface. Then,
by definition, Mb is everywhere orthogonal to the
eigenvector field nb1 . In order to show that Mt is a
APPENDIX B: RELATIVE STRETCHING
forward strain-surface, it suffices to show that Ma OF STRETCHLINES
Fab Mb is everywhere normal to the eigenvector
field nfn . Since Txb Mb spanfnbk xb g2)k)n for any Here, we derive formula (13) for the relative stretching
xb 2 Mb, we have of forward stretchlines. Let ct be a smooth material line.
Denote its time-a and time-b positions by ca and cb , respec-
Txa Ma spanfrFab xb nbk xb g2)k)n ; tively. Then, the relative stretching of the material line ct
over the time interval I [a, b] is defined as
for all xa : Fab xb 2 Ma. Therefore, it suffices to
show that nfn xa ?rFab xb nbk xb for any xa 2 Ma cb
and k 2 f2; 3; ; ng. qct : ; (B1)
ca
From Lemma 2, we have
where denotes the length of a curve.
hnfn xa ; rFab xb nbk xb i
Let r : s 7! rs be the parametrization of ca by arc-
kfn xa kbk xb hnfn xa ; rFab xb nbk xb i; (A6) length, i.e., let jr0 sj 1 for all s 2 0; ca #. Since
cb Fba ca , the mapping Fba - r : s 7! Fba rs is a paramet-
for any xa 2 Ma and k 2 f2; 3; ; ng. rization of the curve cb . Therefore, its length cb is given
Using identity (A1), we obtain by
% & kb xb % f & ca
nfn xa ; rFab xb nbk xb kb nn xa ; rFab xb nbk xb : cb jrFba rsr 0 sjds
k1 xb
0
(A7) ca qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
hr 0 s; Cf rsr0 sids: (B2)
Hence, if 0

kb1 xb 6 kbk xb ; k 2 f2; 3; ; ng; (A8) Now, if the material line ct is a forward stretchline, we
have r 0 s nf2 rs for all s 2 0; ca #. Substituting this
then we have in Eq. (B2), we obtain

hnfn xa ; rFab xb nbk xb i 0; (A9) ca qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffi


f
cb k2 rsds : kf2 ds:
0 ca
for any k 2 f2; 3; ; ng. But since
kb1 ) kb2 ) * * * ) kbn , conditions (A8) hold if and only if
Therefore, by definition (B1), the relative stretching of a
kb1 xb
6 kb2 xb . This condition holds away from
forward-time stretchline ct is given by
repeated eigenvalues of Cb.
qffiffiffiffi
In short, if nb1 xb ? Txb Mb for all xb 2 Mb 1
then nfn xa ? Txa Ma for any xa 2 Ma which qct kf2 ds:
ca ca
implies that Ma is a forward strain-surface. This
concludes the sufficiency condition of Theorem 1-(i). Arnold, V. and Khesin, B., Topological Methods in Hydrodynamics
As for the necessity of the same condition, let (Springer, 1998), Vol. 125.
Balzer, J., A Gauss-Newton method for the integration of spatial normal
Mt be a forward strain-surface, i.e., Txa Ma
fields in shape space, J. Math. Imaging Vision 44, 6579 (2012).
spanfnfk xa g1)k)n$1 for any xa 2 Ma. Therefore, Delmarcelle, T. and Hesselink, L., The topology of symmetric, second-
the tangent space of its advected image Mb is given order tensor fields, in Proceedings of the conference on Visualization 94
by (IEEE Computer Society Press, Los Alamitos, CA, USA, 1994), pp.
140147.
Txb Mb spanfrFba xa nfk xa g1)k)n$1 : Farazmand, M. and Haller, G., Computing Lagrangian coherent structures
from their variational theory, Chaos 22, 013128 (2012).
Hadjighasem, A., Farazmand, M., and Haller, G., Detecting invariant mani-
To show that Mt is a backward stretch-surface, it folds, attractors and generalized KAM tori in aperiodically forced mechan-
suffices to show that nb1 xb ? rFba xa nfk xa for any ical systems, Nonlinear Dyn., in press, doi 10.1007/s11071-013-0823-x.
Haller, G., A variational theory of hyperbolic Lagrangian coherent
xb 2 Mb and k 2 f1; 2; ; n $ 1g. Similarly to
structures, Physica D 240, 574598 (2011).
Eq. (A7), one can show that Haller, G. and Beron-Vera, F. J., Geodesic theory of transport barriers in
two-dimensional flows, Physica D 241, 16801702 (2012).
kf xa b
hnb1 xb ;rFba xa nfk xa i fk hn1 xb ;rFba xa nfk xa i; Haller, G. and Sapsis, T., Lagrangian coherent structures and the smallest
kn xa finite-time Lyapunov exponent, Chaos 21, 023115 (2011).
Haller, G. and Yuan, G., Lagrangian coherent structures and mixing in
(A10) two-dimensional turbulence, Physica D 147, 352370 (2000).
023101-11 M. Farazmand and G. Haller Chaos 23, 023101 (2013)

Lee, J. M., Manifolds and Differential Geometry, Graduate Studies in Palmerius, K. L., Cooper, M., and Ynnerman, A., Flow field visualization
Mathematics (American Mathematical Society, 2009). using vector field perpendicular surface, in Proceedings of the 2009
Lekien, F. and Ross, S. D., The computation of finite-time Lyapunov expo- Spring Conference on Computer Graphics, Budmerice, Slovakia, 2009.
nents on unstructured meshes and for non-Euclidean manifolds, Chaos Peacock, T. and Dabiri, J., Focus issue: Lagrangian coherent structures,
20, 017505 (2010). Chaos 20, 17501 (2010).
Lipinski, D. and Mohseni, K., A ridge tracking algorithm and error estimate Tricoche, X., Scheuermann, G., and Hagen, H., A topology simplification
for efficient computation of Lagrangian coherent structures, Chaos 20, method for 2D vector fields, in Proceedings of Visualization 2000 (2000),
017504 (2010). pp. 359366.

You might also like