ArticlePDF Available

Asymptotic diffusion coefficients and anomalous diffusion in a meandering jet flow under environmental fluctuations

Authors:

Abstract and Figures

The nontrivial dependence of the asymptotic diffusion on noise intensity has been studied for a Hamiltonian flow mimicking the Gulf Jet Stream. Three different diffusion regimes have been observed depending on the noise intensity. For intermediate noise the asymptotic diffusion decreases with noise intensity at a rate which is linearly dependent to the flow's meander amplitude. Increasing the noise the fluid transport passes through a superdiffusive regime and finally becomes diffusive again at large noise intensities. The presence of inner circulation regimes in the flow has been found to be determinant to explain the observed behavior.
Content may be subject to copyright.
PHYSICAL REVIEW E 85, 017201 (2012)
Asymptotic diffusion coefficients and anomalous diffusion in a meandering jet flow
under environmental fluctuations
A. von Kameke, F. Huhn, and V. P´
erez-Mu˜
nuzuri*
Group of Nonlinear Physics, University of Santiago de Compostela, E-15782 Santiago de Compostela, Spain
(Received 26 October 2011; revised manuscript received 16 December 2011; published 27 January 2012)
The nontrivial dependence of the asymptotic diffusion on noise intensity has been studied for a Hamiltonian
flow mimicking the Gulf Jet Stream. Three different diffusion regimes have been observed depending on the
noise intensity. For intermediate noise the asymptotic diffusion decreases with noise intensity at a rate which
is linearly dependent to the flow’s meander amplitude. Increasing the noise the fluid transport passes through
a superdiffusive regime and finally becomes diffusive again at large noise intensities. The presence of inner
circulation regimes in the flow has been found to be determinant to explain the observed behavior.
DOI: 10.1103/PhysRevE.85.017201 PACS number(s): 05.45.a, 47.51.+a, 92.10.Lq, 47.52.+j
In the ocean detailed knowledge of the dispersion of passive
tracers is crucial in order to make predictions about the trans-
port of physical, chemical, and biological tracers. Important
processes range from heat and mass transfer, transport in
global biochemical cycles, or the spread of pollutants [1]to
the transport of plankton [2] and fish larvae [3]. Especially
in coarse climate models, transport on smaller scales has to
be parametrized and statistical transport measures such as the
eddy diffusivity are common [4].
Normal diffusion is a special case of diffusive transport and
a simple model that does not always capture the behavior of
real processes correctly. In general, diffusion can be defined
in terms of the relative dispersion of diffusing particles which
has a time dependence of the form
R2(t)=|xi(t)−xj(t)|2∼tγ,(1)
at long times. For normal diffusion γ=1, while for γ>1
the Lagrangian dispersion is considered as superdiffusion
[5]. Power-law anomalous superdiffusion (γ>1) can occur
in developed turbulence [6] and in chaotic flows [7]. The
subdiffusive case (γ<1) is found in motion through highly
heterogeneous media where particles can be trapped in certain
regions for long periods of time, such as transport in porous
media, gel electrophoresis of polymers (e.g., DNA), and
soluble proteins in the nucleus of living cells [8].
It has been shown that for long times and under the
assumption of an underlying stationary stochastic process, the
presence of a weak fluctuation causes anomalous diffusion
to asymptotically become normal at crossover times tc. These
crossover times are inversely proportional to the noise intensity
ξ, i.e., tcξβwith some exponent βof the order of unity [9].
In terms of the waiting time distribution the transition to
normal diffusion is reflected by an exponentially decreasing
distribution ψ(t)exp(ξt)/tμ[9]. Consequently, as the
noise vanishes ξ0, the exponential term becomes one and
the crossover time goes to infinity. In the limit t→∞and for
finite noise ξthe diffusion process is normal and the asymptotic
diffusion coefficient
DA=lim
t→∞
R2(t)
t(2)
*vicente.perez@cesga.es
can be determined. This coefficient has been shown to depend
on noise intensity as
DA1
ξα(3)
for ξ1, while α=f(β) depends on the used model [911].
However, taking the limit ξ0, DAgives an infinite result.
Karney et al. [12] showed that exchanging the limits, i.e.,
taking ξ0 before t→∞ in Eq. (2), the value of DA
depends on the initial location of the tracers. This dependence
of DAon initial conditions is especially pronounced in flows
with circulation regimes where tracers may be trapped. In
particular, the relative distance in between these sticky trapping
regions and the position and size of initial tracer patches turn
out to influence the asymptotic value of DA[12].
In this Brief Report, we numerically study this limit and
show that Eq. (3) does not always hold. For that purpose, we
determine the effect of additive noise on the transport in a
Hamiltonian system leading to a nontrivial dependence of the
asymptotic diffusion coefficient on the noise intensity ξ.
The meandering jet flow was introduced as a simple
kinematic model for the Gulf Stream and is frequently used
to describe western boundary current extensions in the ocean
[13,14]. This flow is usually represented in a reference frame
ζ=xcxt=ymoving with the phase velocity cxof the
meander, where xand yare Cartesian coordinates, positive
eastward and northward, respectively. Then, the corresponding
nondimensional stream function can be written as
ψ(x,y)=−tanh yBcos κx
R1/2+Cy, (4)
where (x,y)=(ζ,δ)are the dimensionless coordinates. The
velocity field
V(x,y) in this moving frame is obtained from
the stream function as
V(x,y)=ez×
ψ.
We consider an additional noise perturbation η(t) that mim-
ics environmental fluctuations, and we arrive at the following
coupled differential equations for the particle velocity:
˙
x=1
R1/2cosh2θC+η(t),
(5)
˙
y=−
sin κx(1 +B2κ2Byκ2cos κx)
R3/2cosh2θ+η(t),
017201-1
1539-3755/2012/85(1)/017201(4) ©2012 American Physical Society
BRIEF REPORTS PHYSICAL REVIEW E 85, 017201 (2012)
TABLE I. List of parameters used in the Gulf Stream model.
ψ04000 km2d1
λ40 km
cx20 km d1C0.2
A80 km B2.0
L266.67πkm κ0.3
with
R=1+(sin κx)2=yBcos κx
R1/2.
Here η(t) is a Gaussian stochastic process with zero mean,
a temporal correlation function η(t)η(t)=2ξδ(tt), and
noise intensity ξ.
The original parameters of the model are the total eastward
transport 2ψ0, the width λ, the amplitude A, and the wave num-
ber k=2π/L of a sinusoidal meander. These are converted
to dimensionless quantities B=λ1A,κ=,C=λψ1
0cx,
and the dimensionless time t=ψ0λ2t. The parameters of the
model are chosen such that they properly represent geophysical
features of the Gulf Stream. Table Isummarizes these values.
Equations are then integrated using the Heun scheme [15] and
a time step of t =103.
Figure 1shows the jet flow in the comoving frame for
this set of parameters. This flow can be divided into three
distinct regimes: a central meandering eastward jet (A), closed
circulation above meander troughs and below crests (B), and
exterior retrograde westward motion (C). Fluid particles are
restricted to each of these regimes and no cross-stream mixing
occurs. In the presence of an added temporal variability, the
boundaries between regimes break up, allowing transport and
exchange of fluid between regimes. For the jet amplitude
B>B
crit, the geometry of the flow changes drastically and
C
B
A
x
y
−15 −10 −5 0 5 10 15
−10
−5
0
5
10
FIG. 1. Stream function of the jet flow (4). Arrows indicate the
direction of motion and dots correspond to the different regimes in the
flow for the jet amplitude B<B
crit. The high density of contour lines
indicates the position of the central eastward jet. The cross indicates
the position of the stationary point inside the inner circulation core.
The domain shown in the figure is extended periodically in the x
direction. Set of parameters as in Table I.
the central meandering jet moves westward. For the set of
parameters indicated in Table I,Bcrit 3[14].
For the numerical experiments, initially a cluster of
N=500 particles separated from each other by a distance
δ0=103was located at the stationary point inside the closed
circulation core (B) (x0,y0)=(0,B cosh1C1/2). Then,
the temporal evolution of the relative dispersion R2(t)(1)
of this cluster is calculated for times up to t=107.Itwas
found that the statistical properties calculated throughout this
paper do not depend on the number of tracers, provided that
the corresponding computation time is sufficiently long.
The time evolution of the relative dispersion for three noise
intensities is shown in Fig. 2(a). We estimate the power-law
growth at large times for different noise intensities. Figure 2(b)
shows the values of the exponent γin Eq. (1) as a function
of the noise intensity, calculated for two periods of time. Our
results show that the exponent γ>1 for intermediate noise
intensities, while for small and large values of noise, γ=1.
For large noise 10 the exponent evolves as γ1
because inner cores become blurred by noise and diffusion
is dominant. For intermediate noise intensities some particles
intermittently enter within the inner cores where they may
remain for some time, while others travel long distances and
this causes R2(t)to increase superdiffusively [16]. The very
high values of γup to 10 for intermediate noise intensities
in Fig. 2(b) can be understood looking at the two curves (red
squares and green triangles) in Fig. 2(a). Both curves show a
type of transitory behavior with a high slope. This transition
is expected to appear for all noise values, i.e., even for an
arbitrary long integration time of the system there is always a
noise intensity, such that the transition occurs at the end of the
simulation. Thus, the peak in Fig. 2(b) shifts to lower noise
intensities (right of the figure) as integration time increases
[cf. the two curves in Fig. 2(b)]. For small noise 1→∞
the exponent evolves as γ1 because particles have not yet
left the inner circulation cores, i.e., the typical waiting time is
longer than the simulation time.
Theoretically, it is expected that in the limit t→∞
the diffusion should become normal and γ1 for any
noise intensity. However, we find that for intermediate noise
intensities this limit may be well beyond an experiment’s
lifetime and was not approached in this numeric study. When
the asymptotic case of normal diffusion is not reached,
DAas defined in Eq. (2) is not a constant value and we
estimate a lower bound for DAat the maximum simulation
time t=107.
The asymptotic diffusion coefficient DAmeasured as a
function of the noise intensity is shown in Fig. 3.Forvery
large diffusion (regime HN), the trivial relation DAξis
observed, as the random walk of the particles is dominated by
the added noise. For intermediate noise intensities (regime IN),
DAscales with the noise intensity as in Eq. (3)[9,10]. In
between these two regimes, DAξ(HN) and DAξα(IN),
a local minimum at ξ100.5is observed. This minimum
in diffusion value is especially striking as it neither depends
on the model parameters nor on the initial particle positions.
Finally, as noise intensity decreases, DAreaches a maximum
value Dmax
Athat depends on the jet amplitude B. In the limit
of vanishing noise 1→∞,DAξfor any value of B.
For computing times greater than 107, we reproduce the
017201-2
BRIEF REPORTS PHYSICAL REVIEW E 85, 017201 (2012)
10−2 102106
10−4
104
1012
(a)
t
<r2(t)>
10−2 1021061010
0
5
10
15
(b)
1/ξ
γ
FIG. 2. (Color online) (a) Relative dispersion R2(t)for three values of noise intensity. Solid lines correspond to the best fitting to Eq. (1)
to extract γ. Noise intensities: ξ=109(blue circles), ξ=105(red squares), and ξ=102(green triangles). (b) Exponent γas a function
of noise intensity for two intervals of time, t[7 ×105,1×107] (blue circles) and t[2 ×105,6×105] (red squares). Lines indicate normal
(- -) and ballistic (·-) dispersion. Rest of parameters as in Table I.
same qualitative behavior although the maximum DAvalue
is displaced toward small noise intensity values (shifted to the
right in Fig. 3).
Increasing the noise, the boundaries between the jet flow
regimes described in Fig. 1become permeable to particle
crossings due to the random component η(t) and η(t)in
the equations of motion. The exchange process is similar
to the one described in Ref. [17] where lobes of fluid are
entrained and detrained from the edge of the jet. The onset
for this particle exchange occurs at a finite nonzero value
of the perturbation [14] and above it anomalous transport
develops. This onset displaces toward lower nonzero values
of noise as time increases and Dmax
At. This is an important
point because it implies that Eq. (3) does not hold for small
noise intensities since we find that limt,ξ DA<D
max
A,ifξ0
before t→∞.
10−2 10 210 61010
10−8
10−4
100
104
HN IN LN
1/ξ
DA
FIG. 3. (Color online) Asymptotic diffusion coefficient DA(2)at
time t=107as a function of noise intensity for different values of
the jet amplitude. From left to right, solid lines correspond to DAξ
and DAξαwith α=0.90 ±0.03. Dashed vertical lines separate
regions for high (HN), intermediate (IN), and low (LN) noise. Set
of parameters: B=1 (blue circles), B=3 (red squares), and B=5
(black stars). Rest of parameters as in Table I.
The effect of the jet amplitude Bon the asymptotic diffusion
coefficient Dmax
Ais analyzed in Fig. 4. Note that Dmax
Adecreases
as a power law with increasing jet amplitude Band the scaling
factor αin Eq. (3) linearly decreases with B,asshownin
Figs. 4(a) and 4(b), respectively. From panel (a) we also note
that Dmax
Agrows with time tby the same factor for any value
of B, and we find that Dmax
At/Bχ(χ1).
Finally, the influence of the initial tracer positions (x0,y0)on
the diffusion was analyzed. As expected, for ξ→∞,DAξ
independently of (x0,y0). However, as predicted by Karney
et al. [12], for ξ0, DAdepends on the initial conditions
and attains a constant value for (x0,y0) values far from the
inner circulation core. In those regions, the laminar peripheral
westward currents do not disperse the tracers and R2(t)≈
const. For initial conditions inside the circulation core or at
the jet stream, the results do not differ qualitatively from those
described here.
To conclude, in this Brief Report we have analyzed the
nontrivial dependence of the asymptotic diffusion coeffi-
cient DAon noise mimicking environmental fluctuations in
a kinematic model of the Gulf Stream. Above a certain
nonzero value of noise intensity the closed flow regimes
become permeable to particle crossings which explains the
presence of a maximum asymptotic diffusion coefficient. For
intermediate noise intensities, DAξαand the transport
becomes superdiffusive [9,10,16]. However, as an important
result we find that this dependence does not hold for ξ0.
The scaling coefficient αwas found to decrease with the
jet amplitude. Finally, for large noise intensities DAξas
expected. The transport at large times, diffusive for small
noise intensities, becomes superdiffusive at intermediate noise
intensities and then normal at large noise. The presence of
a trapping remnant regular region in the flow for low and
intermediate noise intensities turned out to be important to
explain the nontrivial dependence of the asymptotic transport
on the perturbation strength. For very small noise we did not
find superdiffusive behavior within the computing times used
in this Brief Report (104years). The results even remain
unchanged when B>B
crit and the direction of the jet is
reversed, which indicates that they might be applicable to other
flows consisting of jets and trapping regions.
017201-3
BRIEF REPORTS PHYSICAL REVIEW E 85, 017201 (2012)
0 5 10
102
104
106
(a)
B
DA
max
0 5 10
0.6
0.8
1
(b)
B
α
FIG. 4. (Color online) Logarithmic plot of the maximum value of DAat two instants of time [t=7×105(red squares) and t=1×107
(blue circles)] (a) and the scaling coefficient α,Eq.(3) (b) as a function of the jet amplitude B. Parameters as in Table I.
We are confident that in real experiments with comparably
short lifetimes, the three described diffusion regimes can be
observed. The jet stream model can not only be used to describe
the Gulf Stream, but also others like the Kuroshio current [18]
or the polar jet in the atmosphere. Our results may therefore
help to interpret and predict the transport of tracers in a jet
which is subject to environmental fluctuations.
This work was supported by Ministerio de Educaci´
on
y Ciencia and Xunta de Galicia under Research
Grants No. FIS2007-64698, No. FIS2010-21023, and No.
PGIDIT07PXIB-206077PR. The computational part of this
work was done using CESGA computer facilities. A.v.K. and
F.H. receive funding from FPU Grants No. AP-2009-0713 and
No. AP-2009-3550.
[1] I. Mezic, S. Loire, V. A. Fonoberov, and P. Hogan, Science 330,
486 (2010).
[2] A. Bracco, S. Clayton, and C. Pasquero, J. Geophys. Res. 114,
C02001 (2009).
[3] R. K. Cowen, C. B. Paris, and A. Srinivasan, Science 311, 522
(2006).
[4] E. Shuckburgh, H. Jones, J. Marshall, and C. Hill, J. Phys.
Oceanogr. 39, 2011 (2009).
[5] D. del-Castillo-Negrete, in Anomalous Transport, Foundations
and Applications, edited by R. Klages, G. Radons, and I. M.
Sokolov (Wiley, Weinheim, 2008), pp. 163–211; G. Boffetta,
A. Celani, M. Cencini, G. Lacorata, and A. Vulpiani, Chaos 10,
50 (2000); G. Lacorata, E. Aurell, B. Legras, and A. Vulpiani,
J. Atmos. Sci. 61, 2936 (2004).
[6] M. F. Shlesinger, B. J. West, and J. Klafter, Phys. Rev. Lett. 58,
1100 (1987); M. F. Shlesinger, G. M. Zaslavsky, and J. Klafter,
Nature (London) 363, 31 (1993).
[7] T. H. Solomon, E. R. Weeks, and H. L. Swinney, Phys. Rev.
Lett. 71, 3975 (1993); M. S. Paoletti, C. R. Nugent, and T. H.
Solomon, ibid. 96, 124101 (2006); A. von Kameke, F. Huhn,
G. Fern´
andez-Garc´
ıa, A. P. Mu ˜
nuzuri, and V. P´
erez-Mu˜
nuzuri,
Phys. Rev. E 81, 066211 (2010); Phys. Rev. Lett. 107, 074502
(2011).
[8] T. Geisel and S. Thomae, Phys. Rev. Lett. 52, 1936 (1984);
G. Drazer and D. H. Zanette, Phys.Rev.E60, 5858 (1999).
[9] R. Bettin, R. Mannella, B. J. West, and P. Grigolini, Phys. Rev. E
51, 212 (1995); E. Floriani, R. Mannella, and P. Grigolini, ibid.
52, 5910 (1995).
[10] E. G. Altmann and H. Kantz, EPL 78, 10008 (2007); in
Anomalous Transport, Foundations and Applications, edited by
R. Klages, G. Radons, and I. M. Sokolov (Wiley, Weinheim,
2008), pp. 269–291.
[11] G. Boffetta, D. del-Castillo-Negrete, C. L´
opez, G.
Pucacco, and A. Vulpiani, Phys. Rev. E 67, 026224
(2003).
[12] C. F. F. Karney, A. B. Rechester, and R. B. White, Physica D 4,
425 (1982).
[13] A. M. Bower, J. Phys. Oceanogr. 21, 173 (1991); R. M.
Samelson, ibid. 22, 431 (1992); S. Dutkiewicz and N. Paldor,
ibid. 24, 2418 (1994); J. Duan and S. Wiggins, ibid. 26, 1176
(1996); G. Boffetta, G. Lacorata, G. Redaelli, and A. Vulpiani,
Physica D 159, 58 (2001).
[14] S. V. Prants, M. V. Budyansky, M. Yu. Uleysky, and G. M.
Zaslavsky, Chaos 16, 033117 (2006); M.Yu.Uleysky,M.V.
Budyansky, and S. V. Prants, ibid. 17, 043105 (2007); M. V.
Budyansky, M. Yu. Uleysky, and S. V. Prants, Phys.Rev.E79,
056215 (2009).
[15] P. E. Kloeden and E. Platen, Numerical Solution of
Stochastic Differential Equations (Springer-Verlag, Berlin,
1992).
[16] R. Ishizaki, H. Shibata, and H. Mori, Prog. Theor. Phys. 103,
245 (2000).
[17] M. S. Lozier, L. J. Pratt, A. M. Rogerson, and P. D. Miller, J.
Phys. Oceanogr. 27, 2327 (1997).
[18] C. Mendoza, A. M. Mancho, and M. H. R´
ıo, Nonlinear Processes
Geophys. 17, 103 (2010).
017201-4
... Nonlinear wave processes are observed in a variety of engineering and physics applications such as acoustics [1,2], combustion noise [3,4], jet noise [5][6][7], thermoacoustics [8,9], surface waves [10], and plasma physics [11], requiring nonlinear evolution equations to describe the dynamics of perturbations. In the case of high-amplitude planar acoustic wave propagation, two main nonlinear effects are present: acoustic streaming [2,12] and wave steepening [1,13]. ...
Article
Full-text available
We present a numerical and theoretical investigation of the nonlinear spectral energy cascade of decaying finite-amplitude planar acoustic waves in a single-component ideal gas at standard temperature and pressure. We analyze various one-dimensional canonical flow configurations: a propagating traveling wave (TW), a standing wave (SW), and randomly initialized acoustic wave turbulence (AWT). Due to nonlinear wave propagation, energy at the large scales cascades down to smaller scales dominated by viscous dissipation, analogous to hydrodynamic turbulence. We use shock-resolved mesh-adaptive direct numerical simulation (DNS) of the fully compressible one-dimensional Navier-Stokes equations to simulate the spectral energy cascade in nonlinear acoustic waves. The simulation parameter space for the TW, SW, and AWT cases spans three orders of magnitude in initial wave pressure amplitude and dynamic viscosity, thus covering a wide range of both the spectral energy cascade and the viscous dissipation rates. The shock waves formed as a result of the energy cascade are weak (M<1.4), and hence we neglect thermodynamic nonequilibrium effects such as molecular vibrational relaxation in the current study. We also derive a set of nonlinear acoustics equations truncated to second order and the corresponding perturbation energy corollary yielding the expression for a perturbation energy norm E(2). Its spatial average, 〈E(2)〉, satisfies the definition of a Lyapunov function, correctly capturing the inviscid (or lossless) broadening of spectral energy in the initial stages of evolution—analogous to the evolution of kinetic energy during the hydrodynamic breakdown of three-dimensional coherent vorticity—resulting in the formation of smaller scales. Upon saturation of the spectral energy cascade, i.e., a fully broadened energy spectrum, the onset of viscous losses causes a monotonic decay of 〈E(2)〉 in time. In this regime, the DNS results yield 〈E(2)〉∼t−2 for TWs and SWs, and 〈E(2)〉∼t−2/3 for AWT initialized with white noise. Using the perturbation energy corollary, we derive analytical expressions for the energy, energy flux, and dissipation rate in the wave number space. These yield the definitions of characteristic length scales such as the integral length scale ℓ (characteristic initial energy containing scale) and the Kolmogorov length scale η (shock thickness scale), analogous to the K41 theory of hydrodynamic turbulence [A. N. Kolmogorov, Dokl. Akad. Nauk SSSR 30, 9 (1941)]. Finally, we show that the fully developed energy spectrum of the nonlinear acoustic waves scales as Êkk2ε−2/3ℓ1/3∼Cf(kη), with C≈0.075 constant for TWs and SWs but decaying in time for AWT.
... Nonlinear wave processes are observed in a variety of engineering and physics applications such as acoustics [1,2], combustion noise [3,4], jet noise [5][6][7], thermoacoustics [8,9], surface waves [10], and plasma physics [11], requiring nonlinear evolution equations to describe the dynamics of perturbations. In the case of high-amplitude planar acoustic wave propagation, two main nonlinear effects are present: acoustic streaming [2,12] and wave steepening [1,13]. ...
Preprint
Full-text available
We present a numerical and theoretical investigation of nonlinear spectral energy cascade of decaying finite-amplitude planar acoustic waves in a single-component ideal gas at standard temperature and pressure (STP). We analyze various one-dimensional canonical flow configurations: a propagating traveling wave (TW), a standing wave (SW), and randomly initialized Acoustic Wave Turbulence (AWT). We use shock-resolved mesh-adaptive direct numerical simulations (DNS) of the fully compressible one-dimensional Navier-Stokes equations to simulate the spectral energy cascade in nonlinear acoustic waves. We also derive a new set of nonlinear acoustics equations truncated to second order and the corresponding perturbation energy corollary yielding the expression for a new perturbation energy norm $E^{(2)}$. Its spatial average, <$E^{(2)}$> satisfies the definition of a Lyapunov function, correctly capturing the inviscid (or lossless) broadening of spectral energy in the initial stages of evolution -- analogous to the evolution of kinetic energy during the hydrodynamic break down of three-dimensional coherent vorticity -- resulting in the formation of smaller scales. Upon saturation of the spectral energy cascade i.e. fully broadened energy spectrum, the onset of viscous losses causes a monotonic decay of <$E^{(2)}$> in time.
... Random walks are arguably among the most useful approaches in statistical physics [1][2][3]. Random walk models are found, for instance, in the physical sciences [4][5][6], biology [7][8][9][10], ecology [11,12], and finance [13]. These are but a few examples. ...
Article
Full-text available
A discrete-time dynamics of a non-Markovian random walker is analyzed using a minimal model where memory of the past drives the present dynamics. In recent work [N. Kumar et al., Phys. Rev. E 82, 021101 (2010)] we proposed a model that exhibits asymptotic superdiffusion, normal diffusion, and subdiffusion with the sweep of a single parameter. Here we propose an even simpler model, with minimal options for the walker: either move forward or stay at rest. We show that this model can also give rise to diffusive, subdiffusive, and superdiffusive dynamics at long times as a single parameter is varied. We show that in order to have subdiffusive dynamics, the memory of the rest states must be perfectly correlated with the present dynamics. We show explicitly that if this condition is not satisfied in a unidirectional walk, the dynamics is only either diffusive or superdiffusive (but not subdiffusive) at long times.
... If the change in σ 0 is not so large, neither is the exponent significantly affected because both α and β are inversely proportional to σ 2 0 . In the Manneville map, in contrast, it is known [21] that an even low-intensity external noise affects the asymptotic power law tail in the distribution of residence times with an exponential cutoff. ...
Article
Full-text available
We discuss a route to intermittency based on the concept of reflexivity, namely on the interaction between observer and stochastic reality. A simple model mirroring the essential aspects of this interaction is shown to generate perennial out of equilibrium condition, intermittency and 1/f1/f-noise. In the absence of noise the model yields a symmetry-induced equilibrium manifold with two stable states. Noise makes this equilibrium manifold unstable, with an escape rate becoming lower and lower upon time increase, thereby generating an inverse power law distribution of waiting times. The distribution of the times of permanence in the basin of attraction of the equilibrium manifold are analytically predicted through the adoption of a first-passage time technique. Finally we discuss the possible extension of our approach to deal with the intermittency of complex systems in different fields.
Article
Full-text available
The effect of compressibility on the mixing of Lagrangian tracers is analyzed in chaotic stirred flows. Mixing is studied in terms of the finite-time Lyapunov exponents. Mixing and clustering of passive tracers surrounded by Lagrangian coherent structures is observed to increase with compressibility intensity. The role of the stirring rate and compressibility on mixing and clustering has been analyzed.
Article
We study the combined effects of noise and detector sensitivity on a dynamical process that generates intermittent events mimicking the behavior of complex systems. By varying the sensitivity level of the detector we move between two forms of complexity, from inverse power law to Mittag-Leffler interevent time survival probabilities. Here fluctuations fight against complexity, causing an exponential truncation to the survival probability. We show that fluctuations of relatively weak intensity have a strong effect on the generation of Mittag-Leffler complexity, providing a reason why stretched exponentials are frequently found in nature. Our results afford a more unified picture of complexity resting on the Mittag-Leffler function and encompassing the standard inverse power law definition.
Article
Full-text available
The interaction between a simple meandering jet, such as the Gulf Stream, and an eddy is shown to greatly enhance the mixing and dispersal of fluid parcels in the jet. This enhanced mixing is quantified by calculating the rate of increase of the root-mean-square pair separation of Lagrangian particles (e.g. floats) launched in the jet's immediate vicinity. In the presence of an eddy, particles can escape from the regions in which they were initially launched. Comparisons with observations show a markedly improved qualitative agreement when the eddy is allowed to interact with the meandering jet. -Authors
Article
Full-text available
Hamiltonian energy conserving dynamical systems are examined and the problem of a kinetic description of dynamical systems with chaotic behavior is discussed. It is shown that simple nonlinearities in the Hamiltonian can induce fractal motions with nonstandard statistical properties or 'strange kinetics'. Topological properties of a phase-portrait of the system are discussed and strange kinetics rules are emphasized for Hamiltonian chaos. Scale invariant random walks are reviewed and stochastics and dynamics leading to time-dependent relationships are considered. Levy flights and walks relevant to dynamical systems whose orbits possess fractal properties are discussed. A phenomenology of a kinetic system undergoing strange kinetics with fractal properties is presented.
Article
Full-text available
The EOLE experiment is revisited to study turbulent processes in the lower stratosphere circulation from a Lagrangian viewpoint and to resolve a discrepancy on the slope of the atmospheric energy spectrum between the work of Morel and Larchevêque and recent studies using aircraft data. Relative dispersion of balloon pairs is studied by calculating the finite-scale Lyapunov exponent, an exit-time-based technique that is particularly efficient in cases in which processes with different spatial scales are interfering. The main goal is to reconciliate the EOLE dataset with recent studies supporting a k-5/3 energy spectrum in the 100 1000-km range. The results also show exponential separation at smaller scales, with a characteristic time of order 1 day, and agree with the standard diffusion of about 107 m2 s-1 at large scales. A remaining question is the origin of a k-5/3 spectrum in the mesoscale range between 100 and 1000 km.
Article
Full-text available
We investigate the high-dimensional Hamiltonian chaotic dynamics in N coupled area-preserving maps. We show the existence of an enhanced trapping regime caused by trajectories performing a random walk inside the area corresponding to regular islands of the uncoupled maps. As a consequence, we observe long intermediate regimes of power law decay of the recurrence time statistics (with exponent γ = 0.5) and of ballistic motion. The asymptotic decay of correlations and anomalous diffusion depend on the stickiness of the N-dimensional invariant tori. Detailed numerical simulations show weaker stickiness for increasing N suggesting that such paradigmatic class of Hamiltonian systems asymptotically fulfill the demands of the usual hypotheses of strong chaos.
Book
1. Probability and Statistics.- 2. Probability and Stochastic Processes.- 3. Ito Stochastic Calculus.- 4. Stochastic Differential Equations.- 5. Stochastic Taylor Expansions.- 6. Modelling with Stochastic Differential Equations.- 7. Applications of Stochastic Differential Equations.- 8. Time Discrete Approximation of Deterministic Differential Equations.- 9. Introduction to Stochastic Time Discrete Approximation.- 10. Strong Taylor Approximations.- 11. Explicit Strong Approximations.- 12. Implicit Strong Approximations.- 13. Selected Applications of Strong Approximations.- 14. Weak Taylor Approximations.- 15. Explicit and Implicit Weak Approximations.- 16. Variance Reduction Methods.- 17. Selected Applications of Weak Approximations.- Solutions of Exercises.- Bibliographical Notes.
Article
In an effort to understand the extent to which Lagrangian pathways in the Gulf Stream indicate fluid exchange between the stream and its surroundings, trajectories of RAFOS floats are viewed in a frame of reference moving with the dominant zonal phase speed associated with the periodic flow. In such a frame, geometrical structures emerge that more clearly delineate the position of the parcel in relation to the jet core and its surroundings. The basic premise of this work is that the pathways of fluid parcels in the vicinity of stagnation points, as defined in the moving frame of reference, are susceptible to changes in their pathways, thereby facilitating fluid exchange between different regions of the flow field. Four representative RAFOS float trajectories are shown to exhibit the expected behavior in the vicinity of stagnation points. To further examine the mechanism of exchange in the vicinity of these geometrical features, concepts from dynamical systems theory are applied to a numerically simulated flow field. The entrainment and detrainment of parcels from the jet core are explained in the context of stable and unstable manifolds and their associated lobes. It is shown that the Lagrangian pathways from the numerical flow and the observational trajectories exhibit a similarity based on the kinematics of a meandering flow field. Overall, this study provides the first look at RAFOS float trajectories in a moving frame and provides insight as to how the temporal variability of a jet creates chaotic exchange.
Article
Recent observations of fluid parcel pathways in the Gulf Stream using isopycnal RAFOS floats revealed a striking pattern of cross-stream and vertical motion associated with meanders (Bower and Rossby 1989). In an attempt to explain the observed pattern, a two-dimensional kinematic model of a meandering jet has been developed which enables examination of the relationship between streamfunction patterns and fluid parcel trajectories. The streamfunction fields are displayed in a reference frame moving with the wave pattern so motions of fluid parcels relative to the jet can be seen more easily. The results suggest that the observed pattern of cross-stream motion results primarily from the downstream phase propagation of meanders. The model successfully reproduces several of the most distinctive features of the float observations: 1 ) entrainment of fluid into the Gulf Stream occurs at the leading edges of meander extrema while detrainment takes place at the trailing edges; 2) exchange between the Gulf Stream and its surroundings increases with a) increasing depth, b) increasing meander amplitude, and c) increasing wave phase speed. Transport calculations from the model streamfunction fields indicate that for typical phase speeds (10 km d⁻¹) and amplitudes (50 km), roughly 90% of the fluid in the surface layers of the Gulf Stream flows downstream in the jet while 10% continuously recirculates into the surroundings. In the deep main thermocline, where downstream speeds are less, only about 40% of the fluid is retained in the jet and 60% is trapped in the recirculating cells. It is concluded that this simple kinematic mechanism could lead to cross-stream mixing of fluid parcels, especially in the deeper layers of the Gulf Stream.
Article
The effect of external noise on diffusion processes in a Hamiltonian system with an external force is studied. This is represented by the standard map with an external force. When external noise is applied to the system, we find strongly enhanced diffusion in a parameter range in which anomalous diffusion occurs. Although chaotic orbits are not deterministic in a strict sense, because the system includes noises, they exhibit Lévy flight due to the intermittent sticking to transient tori.
Article
It is shown that anomalous diffusion (i.e., nonlinear growth of mean square displacements) can be caused by a specifically chaotic mechanism. It depends on deterministic diffusion and intermittency and gives rise to mean square displacements which grow asymptotically like tnu with 0
Chapter
IntroductionContinuous Time RandomWalk ModelOrigin of Long-Time Tails in Hamiltonian SystemParadigmatic Fluid ModelDiscussionReferences