Research Summary Dimitrios Giannakis April 21, 2015

advertisement
Research Summary
Dimitrios Giannakis
April 21, 2015
My research is at the interface between applied mathematics and climate atmosphere ocean science. My
primary applied mathematics research interests are in geometrical data analysis algorithms and statistical
modeling of dynamical systems. These tools are applied in a range of applications in climate science,
including intraseasonal oscillations of organized tropical convection and arctic sea-ice variability. Since
joining the Center for Atmosphere Ocean Science at the Courant Institute, I have also worked on information
theoretic methods to quantify predictability and model error in dynamical systems [1–3] and Markov chain
Monte Carlo algorithms for signals with intermittent instabilities [4].
Geometrical data analysis for dynamical systems
High-dimensional data generated by dynamical systems are encountered in many disciplines in science and
engineering. For instance, in atmosphere ocean science, the dynamics take place in an infinite-dimensional
phase space where the coupled nonlinear partial differential equations for fluid flow and thermodynamics
are defined, and the observed data correspond to functions of that phase space, such as temperature or
circulation measured over a geographical region of interest. Mathematically, the observed data at time ti
can be represented by a vector xi ∈ Rn with n 1, and the time-ordered collection of these vectors forms
a high-dimensional timeseries. There exists a strong need for applied mathematics techniques to extract
and predict important phenomena which are an outcome of the underlying dynamics, including the El Niño
Southern Oscillation (ENSO) in the ocean and the Madden-Julian Oscillation (MJO) in the atmosphere.
My work in high-dimensional time series analysis has focused on the following two main questions:
1. Dimension reduction and spatiotemporal pattern extraction; i.e., how to represent the data using a small
number of coordinates (and their associated spatiotemporal patterns) in a manner that reveals intrinsic
timescales of the dynamics [5, 6].
2. Nonparametric forecasting; i.e., how to predict future values of observables (or probability densities)
from the current initial data given a training dataset of past observations of the system, but without
having access to the equations of motion and without using a parametric reduced model [7, 8].
A common theme in this work has been to combine kernel methods from harmonic analysis and machine
learning with ideas from dynamical systems theory to construct dimension reduction maps adapted to the
dynamical system generating the data, and to learn operators governing the time evolution of observables and
probability densities.
1
Kernel methods to detect slow intrinsic timescales
Consider a time series, x = {x0 , x1 , . . . , xs }, consisting of samples xi ∈ Rn taken at times ti = i δt time with a
uniform timestep δt. The general setting of interest is that the samples are generated by an abstract dynamical
system operating in a phase space manifold M , and the samples are the outcome of a vector valued function
on that manifold; i.e., xi = F(ai ) with ai ∈ M . We also have a dynamical flow Tt on M such that ai = Tti a0
with ti = i δt, which we will assume to be ergodic and with invariant measure peq . Dimension reduction can
be described in terms of a map Φ : Rn 7→ Rl , where Φ(xi ) = (φ1 (xi ), . . . , φl (xi )) and l n, and our objective
is to construct this map empirically from the observed data and endow it with the following “desirable”
properties:
• Φ should preserve the manifold structure of M .
• Φ should be intrinsic to the dynamical system generating the data, i.e., it should have strong invariance
properties under changes of the observation modality F.
• The reduced coordinates φi should individually reveal meaningful dynamical processes embedded in
the high-dimensional observed signal.
In kernel methods for data analysis (e.g., [9–11]), one constructs dimension reduction maps through
eigenfunctions of a Markov operator P acting on scalar functions on M constructed from exponentially
decaying kernel functions. The kernel can be thought of as a measure of pairwise similarity K : Rn × Rn 7→ R+
between samples in data space. If M is equipped with a volume form (for our purposes, the volume form
d peq associated with the invariant measure of the dynamics)
it naturally leads to an integral operator G
R
acting on functions on M through the expression G f (ai ) = M K(F(xi ), F(x j )) f (a j ) d peq (a j ). In practical
applications, the action of G is approximated by a Monte Carlo sum in time, which corresponds to integration
with respect to the invariant measure of the dynamics. The Markov operator P is then constructed through
a sequence of normalizations of G ensuring that P f = f if f is constant, and the corresponding dimension
reduction coordinates are computed by solving the eigenvalue problem Pφi = λi φi . (Technically, the dimension
reduction coordinates include eigenvalue-dependent scaling factors, which we omit here.) Geometrically,
this procedure is motivated from the fact that for a suitable choice of kernel and in the limit of large data, P
approximates the heat kernel on the manifold associated with a Riemannian metric that depends on K—it is a
well-established fact that eigenfunctions of heat kernels and the associated Laplace-Beltrami operators can be
used to embed manifolds in Euclidean spaces with optimal preservation of the Riemannian geometry [9, 12–
14]. This observation motivates the design of kernels for dynamical systems to obtain the desirable properties
listed above from the properties of the corresponding induced geometry. To that end, in [5, 6] a family of
kernels was introduced that modifies the geometry of the data by incorporating two dynamics-dependent
features:
Delay-coordinate mappings. Following state-space reconstruction methods [15–17], we construct a new
observation map F̃ : M 7→ Rnq through lagged sequences of F:
Xi = F̃(ai ) = (xi , xi−1 , . . . , xi−(q−1) ).
If q is sufficiently large (and under mild conditions on the dynamical system, the observation function,
and the sampling interval), then with high probability the data points Xi are in one-to-one correspondence
with the points ai on the attractor. Thus, the time series {Xi } becomes Markovian even if the observations
F are incomplete (i.e., F(M ) is not a diffeomorphic copy of M ). Moreover, F̃ modifies the geometry
of the data since distances in delay-coordinate space depend on differences Xi − X j between “videos” as
opposed to differences xi − x j between “snapshots.” In work in collaboration with Andrew Majda on so-called
nonlinear Laplacian spectral analysis (NLSA) algorithms [18], it was experimentally observed that the use
2
of kernels in delay-coordinate space significantly enhances the ability to extract distinct timescales from
high-dimensional signals with individual eigenfunctions. AOS examples demonstrating this behavior can
be found in Fig. 5 ahead, as well as in [18–23]. A more rigorous theoretical justification of the enhanced
timescale separation capability of these eigenfunctions was made in independent work by Berry et al. [24],
where it was shown that the geometry of the data in delay-coordinate space is biased towards the most
stable Lyapunov subspace of the dynamical system generating the data, and this subspace is independent of
the observation modality. In summary, eigenfunctions from kernels with delay-coordinate maps are “good”
dimension reduction coordinates in the sense that they tend to represent processes with a coherent temporal
character, which are also intrinsic to the dynamics.
Dependence on the vector field of the dynamics. The generator of the dynamics, i.e., the skew-symmetric
operator v giving the time-derivative of functions through v( f )(a) = limt→0 ( f (Tt a) − f (a))/t, maps to a
vector field V in delay-coordinate space which can be approximated by finite differences in time; e.g.,
V̂i = (Xi+1 − Xi−1 )/(2 δt) is a second-order approximation of V at state ai measuring the local time tendency
of the state vector. Moreover, for a state X j lying in a neighborhood of Xi , the vector u = X j − Xi approximates
a tangent vector on M and cos θi = u · V̂i /(kukkV̂i k) approximates the cosine of the angle between that
tangent vector and V in the geometry inherited by M from the embedding F̃. In [6], a one-parameter family
of “cone kernels” was introduced that incorporates this directional dependence through the expression
2
K(Xi , X j ) = e−A(Xi ,X j )/δt ,
A(Xi , X j ) =
kXi − X j k2
[(1 − ζ cos2 θi )(1 − ζ cos2 θ j )]1/2 ,
kV̂i kkV̂ j k
ζ ∈ [0, 1).
(1)
In (1) the sampling interval δt controls the bandwidth of the kernel so that the limit of large data corresponds
to δt → 0. Moreover, the parameter ζ controls the influence of the directional terms, and for ζ > 0 cone
kernels preferentially assign large similarity to pairs of samples whose displacement vector is locally aligned
with the dynamical flow. In particular, in the limit ζ → 1 the Riemannian metric induced by cone kernels
becomes generate, assigning arbitrarily small norm to tangent vectors parallel to v. As a result, the associated
Laplace-Beltrami operator ∆ depends on the directional derivatives of functions along v, as opposed to the
full gradient. The outcome of this asymptotic structure of ∆ is that its eigenfunctions extremize a Dirichlet
energy that penalizes variations along the integral curves of v. This property is independent of observation
modality, and endows the eigenfunctions with invariance under a weakly restrictive class of transformations
of the data (including conformal transformations). Another consequence of the structure of ∆ is that the time
series of the eigenfunction values, ti 7→ φ j (ai ), capture intrinsic slow timescales of the dynamics. Figure 1
displays a visualization of this “along-v” property for a dynamical system on the two-torus. This property is
also beneficial in kernel analog forecasting techniques, discussed below.
Kernel analog forecasting
Analog forecasting is a nonparametric technique introduced by Lorenz in 1969 [25], which predicts the
evolution of states of a dynamical system (or observables defined on the states) by following the evolution of
the sample in a historical record of observations which most closely resembles the current initial data. In the
initialization stage of analog forecasting, one identifies an analog, i.e., the state in the historical record which
most closely resembles the current initial data. Traditionally, this is accomplished using Euclidean distances
in the ambient data space so that the analog xi in the historical record x = {x0 , x1 , . . .} corresponding to the
initial data y is given by
xi = argminky − x j k.
x j ∈x
Then, in the forecast step, the historical evolution of that state is followed for the desired lead time τ, and the
observable of interest is predicted based on its value on the analog. Denoting the time series of observable
3
Figure 1. Laplace-Beltrami eigenfunctions φi for a dynamical system on the two-torus, illustrating the “along-v”
property of the eigenfunctions from the cone kernels from (1) with ζ ≈ 1. From left to right, the columns display
representative eigenfunctions obtained through the diffusion maps algorithm [10] for a radial Gaussian kernel, the cone
kernel from with ζ = 0, and the cone kernel with ζ = 0.995. A portion of the dynamical trajectory is also plotted in a
black line for reference. Notice that in the ζ = 0.995 case the leading eigenfunctions vary predominantly in directions
transverse to v. As a result, the level sets of these eigenfunctions are aligned with the orbits of the dynamics, and the
timeseries ti 7→ φ j (ai ) vary slowly. Figure reproduced from [6].
values corresponding to the historical record by { f0 , f1 , . . .}, and the k-step shift map of that time series by
Sk f j = f j+k , the analog forecast fˆ(y, τ) at lead time τ = k δt is given by
fˆ(y, τ) = Sk fi ,
where i is the timestamp of the analog xi .
Two major factors influencing the efficacy of analog forecasting are (1) the identification of skillful
analogs in the training data; (2) the choice of forecast observable (predictand). In work with Jane Zhao [8],
a kernel nonparametric forecasting technique was developed which improves both of the above aspects of
traditional analog forecasting. First, note that the ability to identify skillful analogs amounts to being able to
identify subsets of the training data whose dynamical evolution will shadow the future time evolution for the
given initial data. This suggests that using the cone kernels in (1) to select analogs on the basis of kernel
affinity should improve forecast skill, at least in the short to medium term, as analogs with time tendency (V )
similar to the time tendency of the initial data will be favored. Similarly, selecting analogs in delay-coordinate
space (i.e., using lagged sequences Y of initial data instead of snapshots) should improve skill, especially in
situations with incomplete initial data. Both of these ingredients are included in our proposed scheme.
Another improvement over traditional analog forecasting is to replace single-analog prediction with
prediction based on a weighted ensemble of analogs. Mathematically, the procedure to construct weighted
4
ensembles of analogs is motivated by out-of-sample extension techniques for functions on manifolds [26, 27].
For example, in the geometric harmonics technique [26] (which is related to the Nyström method for outof-sample extension), the observable f is represented by a truncated expansion f ≈ fl = ∑lk=0 cl φl in the
eigenfunction basis of a kernel operator P, and then an estimate fˆ(Y ) of the value of f at an out-of-sample
state Y is computed as a weighted sum of the in-sample eigenfunction values. Specifically, we have
l
fˆ(Y ) =
cj
∑ λ j ∑ p(Y, Xi )φ j (Xi ),
i
j=1
where λ j is the eigenvalue corresponding to φ j , and p(·, ·) the kernel of P. In [8], this expression is modified
to produce an analog forecast at lead time τ = k δt by applying the shift map to the eigenfunctions, giving
l
fˆ(Y, τ) =
cj
∑ λ j ∑ p(Y, Xi )φ j (Xi+k ).
j=1
(2)
i
A similar construction can be made using the Laplacian Pyramids algorithm for out-of-sample extension
[27]; a multiscale iterative method that does not make direct use of the eigenfunctions. With regards to the
choice of the forecast observable, (2) suggests that higher forecast skill should be possible for observables
which are well-approximated by slowly-varying eigenfunctions. The leading eigenfunctions from the cone
kernels in (1) are good candidates for such observables, and in practice we find that these eigenfunctions
describe physically meaningful patterns with favorable predictability properties. Figure 2 shows results from
a challenging application involving low-frequency sea surface temperature (SST) patterns in the North Pacific
sector of a comprehensive climate model, where parametric regression models fail to beat a trivial persistence
forecast.
Nonparametric forecasting with shift maps
The objective of this work [7], carried out in collaboration with Tyrus Berry and John Harlim, is to use a
smooth orthonormal basis of functions, obtained through the diffusion maps algorithm, to approximate the
semigroup of solution operators of stochastic differential equations on manifolds directly from the data and
without knowing or estimating the drift and diffusion coefficients. Let L be the generator of a stochastic
process on a smooth manifold M with invariant measure peq . (In the absence of stochastic effects, L reduces
to the vector field v discussed earlier, but in the setting of interest here drift and diffusion are both present.)
We denote the semigroup of solution operators over time τ by eτL so that eτL f (x) = Ex f (xτ ) gives the
expectation of f (xτ ) at time τ conditioned on x0 = x. The adjoint of L ∗ in the Hilbert space L2 (M , peq )
is the Fokker-Planck operator, governing the evolution of probability densities relative to the equilibrium
∗
measure. That is, an initial density ρ0 relative to peq evolves according to ρτ = eτL ρ0 , and if the density is
uniform relative to the equilibrium measure (e.g., at asymptotic times, τ → ∞) we have ρτ = 1. We denote
the inner product of L2 (M , peq ) by h·, i peq .
2
Next, consider an orthonormal basis {φi }∞
i=0 of L (M , peq ). Such a basis can be constructed through the
eigenfunctions of the generator ∆ of a gradient flow (∆ is a weighted Laplacian), which can be approximated
from data using the diffusion maps algorithm [10] and variable-bandwidth kernel techniques [28]. As above,
we denote the k-step shift map of the sampled time series {x0 , x1 , . . . , xN } by Sk so that for any observable
f , Sk f (xi ) = f (xi+k ). The key observation made in [7] (which can be viewed as a generalization of the
finite-difference approximation of vector fields in [6]), is that the matrix elements obtained through the Monte
Carlo sum
1 N−k
Âi j (τ) = ∑ φi (xl )Sk φ j (xl )
N l=0
5
Analog Forecasting with Dynamics-Adapted Kernels
1.6
1.6
1.4
1.4
1.2
1.2
1
RMSE
RMSE
1
0.8
0.6
0.8
0.6
Persistence
Euclidean Distance
Diffusion Distance
Nystrom
Laplacian Pyramids
0.4
0.2
0
0
37
10
20
30
40
τ (months)
50
0.4
0.2
0
0
60
10
1
1
0.8
0.8
0.6
0.4
0.2
0
0
30
40
τ (months)
50
60
50
60
(b) RMSE-NPGO
pattern correlation
pattern correlation
(a) RMSE-PDO
20
0.6
0.4
0.2
10
20
30
40
τ (months)
50
0
0
60
(c) PC-PDO
10
20
30
40
τ (months)
(d) PC-NPGO
Figure 2. Analog
forecasting
of skill
the Pacific
oscillation
and the
North
oscillation
4.7, but(PDO)
computed
using
as Pacific
groundgyre
truth
the PDO(NPGO) in
Figure
4.8: Same
scoresdecadal
as Figure
the coupled climate
modeleigenfunctions
CCSM3. Theobtained
PDO andfrom
NPGO
dominant
modeserror.
of low-frequency (interannual to
and NPGO
theare
testthe
data
with model
decadal) variability in the North Pacific, and are well represented by a pair of cone kernel eigenfunctions [6]. The panels
above show root mean squared error (RMSE) and pattern correlation (PC) scores for hindcasts of these modes using the
In the
experiments
the low-order
we demonstrated
Takens based on
persistence forecast,
traditional
analogwith
forecasting
basedatmospheric
on Euclideanmodel,
distances,
single-analogthat
forecasting
delay-coordinate
maps areforecasts
an effective
way
improving
forecastapproach
skill withinpartially
observed pyramids.
diffusion distances,
and kernel weighted
using
theofNyström
extension
(2) and Laplacian
initial dataoffor
allcompared
methods to
studied
here. In forecast
this relatively
low-dimensional
setting
with for the
Notice the improvement
skill
the persistence
and single-analog
techniques,
especially
NPGO. Figuredense
reproduced
from
sampling
of [8].
the attractor, single-analog forecasts with diffusion distances provided a
small to moderate improvement prediction skill compared to conventional analogs. On the
other hand, forecasts with the weighted-ensemble methods were able to track the metastable
provide a noisy
approximation
matrix
Ai j (τ)
= hφtoi , aeτL
φ j i peq of improvement
the solution of
semigroup
regime
transitions in to
thisthe
model
for elements
longer times,
leading
pronounced
where τ = kskill
δt and
δt
is
the
sampling
interval.
In
particular,
the
estimates
are
unbiased,
and their
Â
(τ)
ij
especially at moderate to long leads.
2
variance is of order
λ j τ/N,
eigenvaluewe
corresponding
to the eigenfunction
φ j of ∆. Similarly,
In the
Northwhere
Pacificλ SST
applications,
constructed data-driven
forecast observables
j is the
∗
∗
τL
through
kernel
eigenfunctions
representing
two
prominent
interannual
patterns
of
variability
the quantities Âi j (τ) = Â ji (τ) approximate the matrix elements of the operator e
governing the evolution
in densities
the North
Pacific,
the Pacific
decadalTooscillation
(PDO)
[39] and
theinitial
North
of probability
relative
to namely
the equilibrium
measure.
forecast the
evolution
of an
density ρ0 ,
∞
The=timescale
separation
capabilities
of
the
kernels
used
here
Pacific
oscillation
[25]. ρ
we first compute
itsgyre
spectral
expansion
c
φ
with
c
=
hφ
,
ρ
i,
and
then
evaluate
the
density
ρτ at
∑i=0 i i
0
i
i 0
contributed
significantly
to
the
clean
low-frequency
character
and
favorable
predictability
time τ using
∞
ρτ (x) ≈ ρ̂τ (x) =
∑
φi (x)c j Âi j (τ)φi (x).
(3)
i, j=0
This scheme, which we refer to as diffusion forecast, can be thought of as a spectral Galerkin method
for the Fokker-Planck equation formulated in a basis inherited from the gradient flow on the dataset. By
keeping track of the full density, the method is able to provide both the mean forecast, as well as uncertainty
quantification (UQ), e.g., through the second moment of ρτ . Moreover, the method is valid for arbitrary
sampling intervals and the forecast densities have the correct long-time behavior by construction. Figure 3
shows forecast skill results from this method as well as other nonparametric techniques for the Lorenz 63
model. Representative basis functions from the gradient flow are also shown in Fig. 3. Additional applications
to stochastic systems on the torus and ENSO forecasting can be found in [7].
6
45
45
40
40
35
35
30
30
25
25
20
20
15
15
10
12
10
8
10
5
10
0
−10
20
10
0
−10
−20
10
0
−10
45
45
40
40
35
35
30
30
25
25
20
20
15
15
10
20
10
0
−10
−20
RMSE
5
Ensemble Forecast
Ensemble Error Estimate
Diffusion Forecast
Diffusion Error Estimate
6
Local Linear Forecast
Local Linear Error Estimate
Iterated Local Linear Forecast
Iterated Local Linear Error Estimate
Invariant Measure
4
2
10
5
0
0
5
10
0
−10
20
10
0
−10
−20
10
0
−10
20
10
0
−10
−20
10
20
30
40
50
60
Forecast Steps (∆ t = 0.1)
70
80
90
Figure 3. Nonparametric forecasting of the state vector of the Lorenz 63 model. In this example, the initial density is a
Gaussian with variance 0.01 centered on a point close to the attractor, sampled randomly from a Gaussian distribution.
The training dataset consists of 5000 points sampled at a timestep δt = 0.01. The diffusion forecast is performed via (3)
using 4500 eigenfunctions, and the root mean squared error (RMSE) is comparable to an ensemble forecast with 5000
samples which has access to the true model. The diffusion forecast also provides a reasonable UQ (error estimate)
through the standard deviation of the forecast distribution. Also shown for reference are results from nonparametric
models based on local linear linearization. In some cases, these models perform comparably to the diffusion forecast
for the mean, but generally provide poorer UQ. The left-hand panels display a selection of the eigenfunctions of the
gradient flow used to represent the shift map and the initial probability density. Figure adopted from [7].
Applications in climate atmosphere ocean science
As with many other science and engineering disciplines, climate atmosphere ocean science has been experiencing an exponential increase in the amount of data collected by observational networks or output by
numerical models. For instance, the CMIP5 data archive [29] contains several Petabytes of climate model
output from modeling centers around the world, and similarly comprehensive observational and reanalysis
datasets are available. Contained in this data is information that can lead to significant advances in our
scientific understanding and forecasting capability of important phenomena evolving on daily to decadal
timescales. However, due to the sheer volume and complexity of the data, “look and see” approaches have
limited potential in extracting that information. Frequently, data analysis techniques are used to define the
phenomena of interest themselves, and in such cases there exists a clear need for methods that require minimal
ad hoc preprocessing of the data. For instance, the indices and spatiotemporal patterns for ENSO and the
MJO, which are constructed through methods such as Fourier analysis or principal component analysis of
SST, outgoing wavelength radiation (OLR), and other spatiotemporal fields (e.g., [30]), impact our theoretical
understanding of these phenomena, as well as how we assess their representation in weather and climate
models. Despite the significant skill advances in forecasts with large-scale numerical models taking place
in recent years, there exist notable examples (including the applications discussed below) where the lack
of spatial and temporal resolution, parameterization of unresolved processes, and lack of knowledge of the
operating physical laws result in poor dynamical representation of the phenomena of interest. In such cases,
low-order statistical models are useful alternatives to large-scale numerical models, providing comparable or
superior forecast skill.
My work in AOS has broadly focused on using the data analysis techniques described above to (1)
extract physically meaningful modes of variability in the ocean [5, 18], atmosphere [19, 23, 31], and the
cryosphere [20–22] from models and observations with minimal preprocessing of the data; (2) study the
potential predictability of these modes [23, 32], and construct low-order parametric [4] and nonparametric
[7, 8] models for their prediction. Two topics that I am particularly interested in, and are described in
7
detail below, is the co-variability of arctic sea ice with the ocean and atmosphere, and tropical intraseasonal
oscillations (ISOs).
Arctic sea-ice reemergence in models and observations
Arctic sea ice is a sensitive component of the climate system, with dynamics and variability that are strongly
coupled to the atmosphere and ocean. In addition to the strong declining trends observed in recent years
[33], arctic sea ice also exhibits large internal variability. Studies using comprehensive climate models have
estimated that 50–60% of recent Arctic sea-ice changes can be attributed to externally forced trends, with
the remainder resulting from internal variability in the climate system [34]. Therefore, the challenge of
making accurate projections of future Arctic sea-ice conditions hinges on quantifying the sea-ice response to
changes in external forcing (i.e., greenhouse gas forcing), and understanding the nature and magnitude of
internal variability in the coupled ice-ocean-atmosphere system. Recently, Blanchard-Wrigglesworth et al.
[35] proposed a mechanism for sea-ice–SST co-variability in which sea-ice and SST anomalies trade off,
allowing for unexpected memory effects in sea ice. In this mechanism, spring sea-ice anomalies imprint SST
anomalies of opposite sign, which persist over the summer months. During the fall, ice grows southward
and interacts with these SST anomalies, reproducing ice anomalies of the same sign as the spring. This
mechanism (together with a similar, but weaker, mechanism operating from fall to spring) potentially provides
a source of predictability on seasonal to interannual timescales—a problem of significant current interest
[36].
Arctic sea-ice reemergence mechanisms were studied in series of papers with Mitchell Bushuk and
Andrew Majda [20, 21]. In this work, a multivariate extension of the NLSA algorithm was used to construct
spatiotemporal modes of sea-ice, oceanic, and atmospheric co-variability. This method essentially uses the
kernel in (1) with ζ = 0 to combine data sources with different physical dimension in a “unit-free” manner,
and without requiring variance normalization. To represent sea ice, the ocean, and atmosphere we used
sea ice concentration (SIC; the fractional area covered by sea ice), SST, and sea level pressure (SLP) data
over the North Pacific and Arctic domains. We studied control integrations of the CCSM3 model, as well
as observational and reanalysis data from the HadISST dataset and the ERA-Interim product. In both the
model and observational data, we recovered spatiotemporal modes of variability which provide a reduced
representation the sea-ice reemergence mechanism proposed in [35]. The families also reveal time-dependent
aspects of reemergence, which were not accessible in previous studies. In particular, interannual components
of large-scale SLP variability, which emerge objectively from this analysis, are found to be related to coherent
sea-ice reemergence events in geographically distinct regions, and suggest an SLP–sea-ice reemergence
mechanism. Spatiotemporal reconstructions illustrating these processes are displayed in Fig. 4.
In recent work with Mitchell Bushuk [22] the analysis in [20, 21] was extended to a hierarchy of models
with reduced atmospheric and oceanic dynamics. This hierarchy was designed to probe the relative role
of the ocean and atmosphere in sea-ice reemergence—in particular, it includes a slab ocean model (SOM)
with simplified oceanic dynamics but full ocean–atmosphere coupling, as well as so-called CORE models
with prescribed atmospheres and no ocean to atmosphere feedback. We found that a CORE model with an
interannual atmosphere has a more realistic reemergence signal than the SOM, suggesting that the SST–SIC
mechanism is able to operate as a standalone mechanism. Yet, in models with ocean-to-atmosphere coupling
(including the SOM), atmospheric variability plays a key role in reemergence, setting the spatial patterns of
SIC reemergence. This suggests that dynamical feedback from the ocean to the atmosphere is essential in
creating large-scale organized patterns of SIC–SLP co-variability, such as the patterns displayed in Fig. 4.
Ongoing and future work in this area includes forecasting total sea-ice area and volume with the
nonparametric techniques described above, and studying the role of sea ice thickness in reemergence. Another
topic of interest is the extraction of modes representing the global warming trend of sea-ice timeseries.
8
Figure 4. Arctic sea-ice reemergence reconstructed through NLSA. The panels display composites of SIC, SST, and
65in a 900-year control integration
SLP anomalies
the reconstructed
samples
CCSM3 climate
Fig.computed
14. Sea from
ice, SST,
and SLP patterns
of CCSM3
reemergence Family F2M of
atthe
di↵erent
model wheremonths
the activity
of
the
mode
leading-low
frequency
NLSA
mode
exceeded
a
threshold.
During
of the year. These spatial patterns are composites, obtained by averaging over all the March
sea-ice maximum,
there
exist
strong
> 1.SIC anomalies in the marginal ice zones (particularly in the Bering and Labrador
years in
which
LSIC
2
seas), which leave an imprint in SST anomaly of the opposite sign (i.e., positive SIC is accompanied by negative,
cold, SST anomalies). The SIC anomalies disappear during the summer months, but the SST anomaly persists. In
the growth period starting in September sea ice growth is favored (weakened) in the areas of negative (positive) SST
anomaly, leading to a reemergence of the SIC anomalies from the previous spring. Associated with this process is a
low-frequency annular SLP mode which produces poleward (warm) and equatorward (cold) winds in the North Pacific
and North Atlantic basins, respectively. These winds are consistent with an atmospheric teleconnection setting up
anticorrelated SIC anomalies in the two basins. Figure reproduced from [21].
9
Tropical intraseasonal oscillations
Organized tropical convection is a key element of global climate dynamics with direct impact on both
short-term weather forecasting and long-term climate projections. Among the dominant modes of tropical
variability, intraseasonal oscillations (ISOs) play a key role in explaining large-scale convective organization
at subseasonal timescales while also influencing the global climate through extratropical interactions [30].
The dominant ISO modes are the well-known Madden-Julian oscillation (MJO) [37], a 30–90-day eastwardpropagating pattern with zonal wavenumber 1–4 mainly active in the boreal winter, and the boreal summer
ISO (BSISO), which has a more emphasized poleward-propagating pattern with a weakened eastward
propagation [38]. These modes influence and are influenced by a plethora of other modes in the climate
system spanning diurnal to interannual timescales. However, despite their dynamical significance and the fact
that they are a major source of predictability on intraseasonal timescales, accurate simulation of the MJO and
other ISOs by large-scale dynamical models remains elusive.
A significant challenge in understanding the behavior of ISOs and their connections to other modes of
variability is that the phenomena themselves are defined subjectively through some data analysis technique
[39], and the data is frequently subjected to preprocessing (e.g., bandpass filtering) in order to isolate the
signal of interest. In this work, carried out in collaboration with Nan Chen, Andrew Majda, Eniko Szekely,
and Wen-wen Tung, our objectives have been to extract modes of organized tropical convection without
preprocessing the data [19, 23], quantify the potential predictability of the MJO conditional on large-scale
regimes [23], and construct low-order stochastic models for MJO prediction [4].
Figure 5 shows Laplace-Beltrami eigenfunctions obtained via the NLSA algorithm applied to infrared
brightness temperature data (Tb ; a proxy for convective activity) from the CLAUS multi-satellite archive.
The data has a 0.5◦ spatial resolution and is sampled every δt = 8 hours over the period 1983–2006. The
eigenfunctions displayed in Fig. 5 represent the MJO, BSISO, as well as modulated diurnal modes with
intraseasonal modulating envelopes. Besides the modes in Fig. 5, the NLSA spectrum contains eigenfunctions
representing ENSO and the annual cycle and its harmonics—the timescale separation of in the eigenfunctions
is a consequence of delay-coordinate mapping as discussed earlier (here, the number of lags is q = 512,
corresponding to a 64-day temporal window.) A key difference between the patterns in Fig. 5 and the
corresponding modes obtained via comparable linear techniques is strong intermittency and seasonality,
leading to superior separation of the boreal-winter MJO and BSISO through NLSA. Phase composites
of these patterns are displayed in Fig. 6. In [4], a nonlinear stochastic oscillator model is developed to
predict the MJO eigenfunctions. In this model, the intermittency of the MJO modes is modeled via a
hidden variable representing time-dependent damping. This variable interacts with the MJO modes through
energy-conserving quadratic nonlinearities, which also contribute to strong parametric robustness of the
model. The hidden variable at forecast initialization is estimated through a maximum likelihood calculation
for conditionally Gaussian random variables. These models have skillful prediction for 40, 25 and 18 days
in strong, moderate, and weak MJO winters, respectively. Furthermore, the ensemble spread is an accurate
indicator of forecast uncertainty at long lead times. Current and future work in this area includes applications
of the nonparametric techniques described above to MJO and BSISO forecasting, and studying the circulation
aspects of the identified ISO patterns and their impact on the Indian monsoon.
References
[1] D. Giannakis and A. J. Majda. Quantifying the predictive skill in long-range forecasting. Part I: Coarse-grained
predictions in a simple ocean model. J. Climate, 25:1793–1813, 2012. doi:10.1175/2011jcli4143.1.
[2] D. Giannakis and A. J. Majda. Quantifying the predictive skill in long-range forecasting. Part II: Model error in
coarse-grained Markov models with application to ocean-circulation regimes. J. Climate, 25:1814–1826, 2012.
doi:10.1175/jcli-d-11-00110.1.
10
0
−2
φ21
2
0
−2
φ23
2
0
−2
9201 9203 9206 9209 9212 9303 9306 9309 9312
1E−3
1E−2
1E−1
1E0
|FT(φ12)|2
|FT(φ13)|2
φ13
2
|FT(φ21)|2
φ12
0
−2
|FT(φ23)|2
1E9
1E8
1E7
1E6
1E5
1E4
1E3
1E2
1E9
1E8
1E7
1E6
1E5
1E4
1E3
1E2
1E9
1E8
1E7
1E6
1E5
1E4
1E3
1E2
1E9
1E8
1E7
1E6
1E5
1E4
1E3
1E2
2
1E1
frequency (d−1)
date (yymm)
Figure 5. Timeseries of Laplace-Beltrami eigenfunctions from NLSA representing the MJO (first row from top), the
BSISO (third row) and amplitude-modulated diurnal modes with intraseasonal envelopes (second and fourth rows) in
the CLAUS Tb dataset. Notice the strong intermittency and seasonality of the MJO and BSISO eigenfunctions, which
are predominantly active during the boreal winter and summer months, respectively. The eigenfunctions are shown here
for a two-year portion of the 1983–2006 training dataset which includes the intensive observing period of the TOGA
COARE field campaign (November 1, 1992 to February 28, 1993); the significant MJO events identified via NLSA
are in agreement with the directly observed TOGA COARE events. Other modes in the NLSA spectrum (not shown
here) represent the annual cycle and its harmonics and ENSO. Delay-coordinate mapping is crucial for producing the
observed timescale separation in the eigenfunctions. Figure adopted from [23].
Figure 6. Phase composites of the MJO (left) and BSISO (right) recovered by NLSA from CLAUS Tb data. Colorcoded are Tb anomalies (in K), with negative (positive) anomalies representing enhanced (suppressed) cloudiness and
convection. In the case of the MJO, initiation over the Indian ocean, accompanied by the development of anomalously
high convection over eastern Africa and tropical eastern Brazil, takes place in Phases 1 and 2, and is followed by
propagation over the Maritime Continent and the western Pacific warm pool in Phases 3–5 and 6–7, respectively.
Termination at the dateline takes place in Phase 8. For BSISO, Phase 1 corresponds to a cluster of positive Tb anomalies
developing in the central Indian Ocean. In Phases 2–4, that cluster moves northeastward towards the Bay of Bengal
and India and branches off towards the western Pacific and the Monsoon Trough, bypassing the Maritime Continent
from the north. Following the dry phase of BSISO, a cluster of anomalously high convection develops in Phase 5 in the
central Indian Ocean, and propagates towards India and the western Pacific in Phases 6–8, completing the BSISO cycle.
Figure reproduced from [23].
11
[3] D. Giannakis, A. J. Majda, and I. Horenko. Information theory, model error, and predictive skill of stochastic
models for complex nonlinear systems,. Phys. D., 241:1735–1752, 2012. doi:10.1016/j.physd.2012.07.005.
[4] N. Chen, A. J. Majda, and D. Giannakis. Predicting the cloud patterns of the Madden-Julian Oscillation through a
low-order nonlinear stochastic model. Geophys. Res. Lett., 41(15):5612–5619, 2014. doi:10.1002/2014gl060876.
[5] D. Giannakis and A. J. Majda. Nonlinear Laplacian spectral analysis for time series with intermittency and
low-frequency variability. Proc. Natl. Acad. Sci., 109(7):2222–2227, 2012. doi:10.1073/pnas.1118984109.
[6] D. Giannakis. Dynamics-adapted cone kernels. SIAM J. Appl. Dyn. Sys., 2015. In press.
[7] T. Berry, D. Giannakis, and J. Harlim. Nonparametric forecasting of low-dimensional dynamical systems. Phys.
Rev. E., 91:032915, 2015. doi:10.1103/PhysRevE.91.032915.
[8] Z. Zhao and D. Giannakis. Analog forecasting with dynamics-adapted kernels. Nonlinearity, 2014. In review.
[9] M. Belkin and P. Niyogi. Laplacian eigenmaps for dimensionality reduction and data representation. Neural
Comput., 15:1373–1396, 2003. doi:10.1162/089976603321780317.
Diffusion maps.
Appl. Comput. Harmon. Anal., 21:5–30, 2006.
[10] R. R. Coifman and S. Lafon.
doi:10.1016/j.acha.2006.04.006.
[11] T. Berry and T. Sauer. Local kernels and the geometric structure of data. J. Appl. Comput. Harmon. Anal., 2015.
doi:10.1016/j.acha.2015.03.002. In press.
[12] P. H. Bérard. Spectral Geometry: Direct and Inverse Problems, volume 1207 of Lecture Notes in Mathematics.
Springer-Verlag, Berlin, 1989.
[13] P. W. Jones, M. Maggioni, and R. Schul. Manifold parametrizations by eigenfunctions of the Laplacian and heat
kernels. Proc. Natl. Acad. Sci., 105:1803, 2008.
[14] J. Portegies. Embeddings of Riemannian manifolds with heat kernels and eigenfunctions. arXiv:1311.7568, 2014.
[15] F. Takens. Detecting strange attractors in turbulence. In Dynamical Systems and Turbulence, volume 898 of
Lecture Notes in Mathematics, pages 366–381. Springer, Berlin, 1981. doi:10.1007/bfb0091924.
[16] D. S. Broomhead and G. P. King. Extracting qualitative dynamics from experimental data. Phys. D, 20(2–3):
217–236, 1986. doi:10.1016/0167-2789(86)90031-x.
[17] T. Sauer, J. A. Yorke, and M. Casdagli.
Embedology.
J. Stat. Phys., 65(3–4):579–616, 1991.
doi:10.1007/bf01053745.
[18] D. Giannakis and A. J. Majda. Comparing low-frequency and intermittent variability in comprehensive
climate models through nonlinear Laplacian spectral analysis. Geophys. Res. Lett., 39:L10710, 2012.
doi:10.1029/2012GL051575.
[19] W.-w. Tung, D. Giannakis, and A. J. Majda. Symmetric and antisymmetric signals in MJO deep convection. Part I:
Basic modes in infrared brightness temperature. J. Atmos. Sci., 71:3302–3326, 2014. doi:10.1175/jas-d-13-0122.1.
[20] M. Bushuk, D. Giannakis, and A. J. Majda. Reemergence mechanisms for North Pacific sea ice revealed through
nonlinear Laplacian spectral analysis. J. Climate, 27:6265–6287, 2014. doi:10.1175/jcli-d-13-00256.
[21] M. Bushuk, D. Giannakis, and A. J. Majda. Arctic sea ice reemergence: The role of large-scale oceanic and
atmospheric variability. J. Climate, 2015. doi:10.1175/jcli-d-14-00354.1. In press.
[22] M. Bushuk and D. Giannakis. Sea-ice reemergence in a model hierarchy. Geophys. Res. Lett., 2015. In review.
[23] E. Szekely, D. Giannakis, and A. J. Majda. Extraction and predictability of coherent intraseasonal signals in
infrared brightness temperature data. Climate Dyn., 2015. In minor revision.
[24] T. Berry, R. Cressman, Z. Greguric Ferencek, and T. Sauer. Time-scale separation from diffusion-mapped delay
coordinates. SIAM J. Appl. Dyn. Sys., 12:618–649, 2013.
[25] E.N. Lorenz. Atmospheric predictability as revealed by naturally occurring analogues. J. Atmos. Sci., 26:636–646,
1969. doi:10.1175/1520-0469(1969)26<636:aparbn>2.0.co;2.
[26] R. R. Coifman and S. Lafon. Geometric harmonics: A novel tool for multiscale out-of-sample extension of
empirical functions. Appl. Comput. Harmon. Anal., 21:31–52, 2006. doi:j.acha.2005.07.005.
[27] N. Rabin and R. R. Coifman. Heterogeneous datasets representation and learning using diffusion maps and
Laplacian pyramids. In 12th SIAM International Conference on Data Mining, 2012.
[28] T. Berry and J. Harlim. Variable bandwidth diffusion kernels. J. Appl. Comput. Harmon. Anal., 2014.
doi:10.1016/j.acha.2015.01.001. In press.
[29] K. E. Taylor, R. J. Stouffer, and G. A. Meehl. An overview of CMIP5 and the experiment design. Bull. Amer.
Meteor. Soc., 93:485–498, 2011. doi:10.1175/bams-d-11-00094.1.
[30] W. K. M. Lau and D. E. Waliser. Intraseasonal Variability in the Atmosphere–Ocean Climate System. SpringerVerlag, Berlin, 2011.
12
[31] D. Giannakis, W.-w. Tung, and A. J. Majda. Hierarchical structure of the Madden-Julian oscillation in infrared
brightness temperature revealed through nonlinear Laplacian spectral analysis. In 2012 Conference on Intelligent
Data Understanding (CIDU), pages 55–62, Boulder, Colorado, 2012. doi:10.1109/CIDU.2012.6382201.
[32] D. Giannakis and A. J. Majda. Limits of predictability in the North Pacific sector of a comprehensive climate
model. Geophys. Res. Lett., 39:L24602, 2012. doi:10.1029/2012gl054273.
[33] D. J. Cavalieri and C. L. Parkinson. Arctic sea ice variability and trends. The Cryosphere, 6(4):881–889, 2012.
doi:doi:10.5194/tc-6-881-2012.
[34] J. C. Stroeve, V. Kattsov, A. Barrett, T. Serreze, M. Pavlova, M. Holland, and W. N. Meier. 2012: Trends
in arctic sea ice extent from CMIP5, CMIP3 and observations. Geophys. Res. Lett., 39(15):L16502, 2012.
doi:10.1029/2012gl052676.
[35] E. Blanchard-Wrigglesworth, K. C. Armour, C. M. Bitz, and E. DeWeaver. Persistence and inherent predictability
of Arctic sea ice in a GCM ensemble and observations. J. Climate, 24:231–250, 2011. doi:10.1175/2010jcli3775.1.
[36] J. C. Stroeve, L. C. Hamilton, C. M. Bitz, and E. Blanchard-Wrigglesworth. Ensemble skill of the SEARCH Sea
Ice outlook 2008–2013. Geophys. Res. Lett., 41(7):2411–2418, 2014. doi:10.1002/2014gl059388.
[37] R. A. Madden and P. R. Julian. Detection of a 40–50 day oscillation in the zonal wind in the tropical Pacific. J.
Atmos. Sci., 28(5), 1971. doi:10.1175/1520-0469(1971)028<0702:doadoi>2.0.co;2.
[38] K. Kikuchi, B. Wang, and Y. Kajikawa. Bimodal representation of the tropical intraseasonal oscillation. Climate
Dyn., 38:1989–2000, 2012. doi:10.1007/s00382-011-1159-1.
[39] G. N. Kiladis, J. Dias, K. H. Straub, M. C. Wheeler, S. N. Tulich, K. Kikuchi, K. M. Weickmann, and M. J.
Ventrice. A comparison of OLR and circulation-based indices for tracking the MJO. Mon. Wea. Rev., 142:
1697–1715, 2014. doi:10.1175/mwr-d-13-00301.1.
13
Download