Bayesian Delay Embeddings for Dynamical Systems

Neil Dhir†∗ , Adam R. Kosiorek† and Ingmar Posner Oxford Robotics Institute Department of Engineering Science University of Oxford

Abstract Selecting a suitable embedding space is a key issue in modelling nonlinear dynamics. In classical phase-space reconstruction, which relies on time-delay vectors, the embedding space is highly dependent on two discrete parameters (for the univariate case), the values of which greatly affect model performance. They also determine the complexity of the dynamics topology. In consequence, the parameters and dynamics are intimately linked. Thus we propose a modelling framework that jointly models the embedding dynamics and parameters, in a Bayesian fashion by framing the learning problem in terms of variational inference over model parameters. We compare our methods to other models on noisy synthetic observations.

1

Introduction

Takens’ theorem provides the theoretical basis for an experimental approach for autoregressive modelling of nonlinear dynamical systems [10]. We will use this strategy as outlined below. Motivation Suppose we are faced with some physical system, such as an aircraft engine or the ocean surface. The state of this system is deterministically changing with time and changes according to (generally) smoothly varying Newtonian dynamics. Engines for example do not have abrupt stops or process discontinuities, but are fluently powered up and placidly powered down. Because we assume that the system is deterministic, it is true that the state xt is uniquely determined by the state xt−1 , where time is discretely indexed by t [12]. Equivalently there is a smooth diffeomorphism ξ : M → M which maps the state xt−1 to xt . Takens’ theorem (see theorem 1) tells us that if we have time-series measurements y1:T , {y1 , . . . , yT } of this system, and compose n = T − (τ − 1) timedelayed vectors xt = [yt , yt−τ , . . . , yt−(dE −1)τ ]T s.t. {xt | t = 1, . . . , n} from those observations, these vectors lie on a subset of Rn which is an embedding of M iff dE ≥ 2m + 1. We now use a nonparametric proxy (the delay vector) for the true system state, which lives in a space of size m. If we can reconstruct the phase-space, under suitable conditions, it allows for powerful methods of prediction, control and other applications. See fig. 1. Hence, if we have sufficient measurements, and the state-space is well covered, it is possible to reconstruct the true dynamics of a system using only a univariate stream of information. We seek principled ways of finding the embedding dimension dE , the time delay τ and the system dynamics ξ. Under Takens’ theorem, it is assumed that these properties and quantities exist [20, 25]. Patel & Eng [20] explain, however, that this is an ill-posed inverse problem, since the quality of the observations determine if any solutions are stable, unique or whether they exist at all. ∗ †

Corresponding author: [email protected] Equal contribution.

31st Conference on Neural Information Processing Systems (NIPS 2017), Long Beach, CA, USA.

Problem Consider the general, discrete-time, non-linear or linear, Gaussian or non-Gaussian state-space model: x1 | ω ∼ µω (x1 ) xt | xt−1 , φ ∼ fφ (xt | xt−1 ) yt | xt , λ ∼ gλ (yt | xt )

Initial state distribution System process Observation process

(1) (2) (3)

where the states and the observed measurements are denoted by xt ∈ X ⊆ Rd and yt ∈ Y ⊆ RD respectively, ∀t ∈ T , {1, . . . , T } and d  D. The dynamics and the observations are modelled by probability density functions fφ (·) and gλ (·) respectively, parametrised by φ ∈ Φ ⊆ Rnφ and λ ∈ Λ ⊆ Rnλ respectively. The initial state distribution has parameters ω ∈ Ω ⊆ Rnω .

In this paper we assume that f (·) and g(·) are unknown and have to be inferred or nonparametrically approximated. Our problem formulation injects Takens’ embedding theorem and uses it instead of the system model, while and employing the imposed mappings. Despite this assumption, we are still faced with finding the functional form of f (·) and the parameters of the delay-coordinate vectors.

Solution We propose a solution by way of prediction. In our model, with inspiration from Basharat & Shah [3], under the state-space modelling formalism information is passed as follows: {y1 , y2 , . . . , yt }

{y2 , y3 , . . . , yt+1 }

xt = [yt , . . . , yt−(dE −1)τ ]T | {z } Encoded by VAE

xt+1 = [yt+1 , . . . , yt+1−(dE −1)τ ]T | {z }

xt+1 = f (xt ) | {z }

τ, dE and f (·) inferred by VAE

Decoded by VAE

We jointly learn the functional form of f (·) and the embedding parameters dE and τ , by employing a novel form of the variational autoencoder [15] which includes the delay coordinate method [25].

15

10

10 5

5

0

0

−5

−5

−10

x(t)

−10

−15

x(t − τ )

15 10

x(t − 2τ )

−15 0

250

500

x(t − 2τ )

x(t), x(t − τ ), x(t − 2τ )

15

750

1000

1250

1500

Time [t]

(a) Time-delay vectors for x–coordinate.

1750

2000

−15 −10

−5

0

x (t )

5

10

15

5 0 τ) −5 − −10 (t x −15

(b) Reconstructed phase-space.

Figure 1: Phase-space reconstruction using the method of delays. The x coordinate from the original Lorenz-63 system [17] (which is three dimensional with coordinates (x, y, z)) is time-delayed, thereby creating phase-space coordinates, with parameters τ = 17 and dE = 3.

2

Method

The first step to forecast a chaotic time-series is to employ the history of the observations to reconstruct the state-space [26], which requires finding appropriate values of the embedding dimension and time-delay parameters. The second step builds the forecasting model itself. Our method combines these steps, typically disjoint and performed separately, into a single model. 2.1

Variational autoencoder

We seek to learn an encoder and a decoder for mapping observations y = y1:T to and from values x that live in a continuous space. By introducing a parametric inference model of the latent variables 2

Kingma & Welling [15] showed that x can be interpreted as the latent variables in a probabilistic generative model. They introduce a probabilistic decoder via the likelihood function pθ (y | x), as well as a probabilistic encoder with the posterior distribution p(x | y) ∝ pθ (x)pθ (y | x). Where pθ (x) is the prior distribution over the latent variables. Kingma & Welling [15] demonstrated that the variational Bayesian approach admits efficient inference by maximising the evidence lower bound (ELBO) L(y; θ, φ) = Eq(x|y) [log pθ (y, x) − log qφ (x | y)] ≤ log pθ (y) (4) by simultaneously learning the parameters of the encoder and the approximate posterior distribution qφ (x | y). If the decoder and the encoder can be computed point-wise, and are both differentiable with respect to their parameters, a monte-carlo approximation of the ELBO can be maximised with gradient descent. 2.2

Combining Takens’ embedding theorem with a VAE

The state of a deterministic dynamical system, and consequently its future evolution, is fully specified by a point in phase-space [21]. Correspondingly, a point in phase-space will always yield the same image in reconstruction space. Theorem 1 provides us with a rigorous framework for analysing the information content of non-linear dynamics of the reconstruction space by enriching a measurement yt with time-shifted copies of itself yt−τ i.e. the delay coordinates [6]. The price (viz. the condition), however, for this result is an exhaustive search over the embedding parameters (we discuss this process in the next section). Currently, to our knowledge, there exists no direct method for learning these parameters from observation, in a principled manner such that the measured space can be accurately and faithfully encoded into and decoded (reconstructed) from the phase-space [23, §9.2]. However, if we treat the embedding task as an autoencoding problem, a vast domain of results becomes available. Further, we are not aware of any methods which approach this inference problem jointly; that is, interleaving inference over embedding parameters and dynamics topology (dynamics function). In previous work, dynamics have been found using kernel regression [3], feed-forward neural networks, recurrent radial basis functions as well as fuzzy models [22, 2]. All of these current methods are reliant on empirical selection of τ and dE as theorem 1 is “silent on the choice of” [1, §3.1] these parameters. There is a large literature on the ‘optimal’ choice of the time delay parameter(s)3 . It turns out that what constitutes the optimal choice largely depends on the application [11]. Still, practitioners turn to one metric above all others: average mutual information (AMI). This measure tells us how much information can be learned from a measurement taken at one time compared to a measurement taken at a different time [1, §3.3]. What such linear choice has to do with the nonlinear process relating yt and yt+τ is not clear, and hence this choice is one which “we cannot recommend at all” [1, §3.3]. That being said, practically speaking, for some systems it might work, but it is not clear why. Kantz & Schreiber [14] take the view that good embeddings are best “found by trial and error”, conditioned on the problem at hand. Further, theorem 1 gives us a sufficient condition for integer dimension dE > 2m, which tells us, purely from geometrical considerations, the magnitude of dE necessary to render a good projection i.e. without phase-space trajectories crossing (i.e. to ‘unfold’ the phase-space). Currently, the false nearest neighbour (FNN) method is the most popular method for finding the dE given a chaotic system [1]. It is a very empirical operation, but the main idea is simple: examine how the number of nearest neighbours of a point along a phase-space orbit changes when varying dE . The FNN, like the AMI, are currently the most common attempts at solving what is generally considered an open problem [2, 5, 8, 4]. We structure our model as a VAE with both discrete and continuous latent variables. We learn an approximate posterior distribution over the number of dimensions dE and the sampling delay τ . Given this approximation, we construct a stochastic embedding function used for encoding observations y into the phase-space x. Finally, we employ a multi-layer perceptron (MLP) to predict a time-shifted sequence from its latent encoding. We arrive at an ELBO, whose derivation can be found in appendix A: L(y; θ, φ, τ, dE ) = Ex,τ,dE ∼qφ (x,τ,dE |y) [log pθ (y | x, τ, dE )]−DKL (qφ (x, τ, dE | y) || pθ (x, τ, dE )) . (5) 3

We only consider univariate input sequences, but the multivariate case follows trivially, in which case a different τ is prescribed for each dimension of the input sequence. For further details see the work by Cao [7].

3

The first term can be seen as a probabilistic analogue of the reconstruction error. The second term pulls the approximate posterior distributions to their priors, thereby acting as a regulariser. Optimising the ELBO requires specifying priors for the latent variables. Since τ and dE are both discrete and can be though of as population statistics, we employ the Poisson distribution. If we do not have prior information about a dynamical system, it is natural to expect that we need many and densely spaced observations to characterise its behaviour. Therefore, we use Poisson distributions with high values for its parameters for dE and low for τ . The time-delay τ defines the interval with which we sample and dE defines the number of intervals we take samples from. We have one additional degree of freedom, namely, which point in the interval given by τ should we take. We model this as a categorical distribution, and to this end we use Gumbel-Softmax [13], which can be easily used with gradient-based optimisation methods. Sampling of τ and dE introduces discontinuities and we employ the NVIL estimator [18] to work around them. Since we are learning to predict shifted sequences, we are implicitly learning the dynamics f (·) of b2:T +1 . Consequently we arrive at the phase-space. We encode y1:T to x1:T and reconstruct back to x a sample with future predictions included, and also obtain the embedding parameters in the process. Provided that these are known, we can again encode the sequence to obtain xT +1 . We have therefore found a functional relationship from xt to xt+1 . The full algorithm can be found in appendix B.

3

Experiments

We consider N corrupted time-series sampled from the Lorenz-63 system [17] – the canonical chaotic system. For results on the easier problem using clean time-series see e.g. [24, 19]. The goal is to accurately predict the dynamics of x in a one-step prediction regime with the addendum that our model has to learn the parameters of the embedding mapping. In other words the VAE has to learn the parametric form of mapping into and out of the phase-space. The parameters for the time-delay vectors are typically assumed known for toy problems and estimated through the AMI and FNN methods – the drawbacks of which were discussed in the previous section. We compare our latent variable model to two other specifications on two data corrupted prediction tasks. Table 1: Summary of one-step prediction results and comparisons. Latent variable model τ dE RMSE SNR N Nonlinear dynamics Kalman-Takens filter‡ [9] †

EKF-RMLP [10, §4] Our method •

‡ †

4

20 4 4 3.9• 3.5•

4 3 3 7.8• 7.5•

∼1.900• 0.378 0.183 1.078 3.270

∼5 25 10 25 10

6000 5000 5000 5000 5000

7 3 3 3 3

Expected value: E[·] The authors do not specify how the embedding parameters were selected. The authors used a different parametrisation for the Lorenz system, where AMI and FNN were calculated before noise was added.

Discussion and conclusion

We do not propose yet another state-of-the-art prediction model, as there are already many impressive systems. Instead, we seek to update the methodology for robustly learning the embedding parameters and the system dynamics from the observations, whilst taking nonlinearities into account. The current, most common, analysis pipeline starts with 1) AMI then 2) FNN and finally 3) phase-space reconstruction and dynamics learning. We combine these steps into one coherent Bayesian model, which introduces dependencies between all the constituent parts. By levering deep-learning with dynamical systems theory, we jointly learn system dynamics and the mapping from phase-space to observed space – by maximising our contributed VAE’s ELBO (derived in appendix A for our model). Whilst our prediction performance (see table 1) is not as competitive as the EKF-RMLP, we can instead point to Bayesian embeddings i.e. we have posterior estimates of our embedding parameters, fully conditioned on the observations. Whilst one-step prediction is an important task, having strong information regarding the topology of the entire phase-space is arguably more valuable, since we are then in a position to deterministically reason about the present, past and future of our system, provided the space has been adequately covered through our measurements [12]. This insight is difficult to glean with the current methods for finding the embedding (see appendix D). 4

References [1] Abarbanel, Henry D.I., Brown, Reggie, Sidorowich, John J., and Tsimring, Lev Sh. The analysis of observed chaotic data in physical systems. Reviews of Modern Physics, 65(4):1331, 1993. [2] Aguirre, Luis A and Letellier, Christophe. Modeling nonlinear dynamics and chaos: a review. Mathematical Problems in Engineering, 2009, 2009. [3] Basharat, Arslan and Shah, Mubarak. Time series prediction by chaotic modeling of nonlinear dynamical systems. In Computer Vision, 2009 IEEE 12th International Conference on, pp. 1941–1948. IEEE, 2009. [4] Bradley, Elizabeth and Kantz, Holger. Nonlinear time-series analysis revisited. Chaos: An Interdisciplinary Journal of Nonlinear Science, 25(9):097610, 2015. [5] Broomhead, Dr S and King, Gregory P. Extracting qualitative dynamics from experimental data. Physica D: Nonlinear Phenomena, 20(2-3):217–236, 1986. [6] Brunton, Steven L, Brunton, Bingni W, Proctor, Joshua L, Kaiser, Eurika, and Kutz, J Nathan. Chaos as an intermittently forced linear system. arXiv preprint arXiv:1608.05306, 2016. [7] Cao, Liangyue. Practical method for determining the minimum embedding dimension of a scalar time series. Physica D: Nonlinear Phenomena, 110(1-2):43–50, 1997. [8] Erem, Burak, Orellana, Ramon Martinez, Hyde, Damon E, Peters, Jurriaan M, Duffy, Frank H, Stovicek, Petr, Warfield, Simon K, MacLeod, Rob S, Tadmor, Gilead, and Brooks, Dana H. Extensions to a manifold learning framework for time-series analysis on dynamic manifolds in bioelectric signals. Physical Review E, 93(4):042218, 2016. [9] Hamilton, Franz, Berry, Tyrus, and Sauer, Timothy. Kalman-takens filtering in the presence of dynamical noise. arXiv preprint arXiv:1611.05414, 2016. [10] Haykin, Simon et al. Kalman filtering and neural networks. Wiley Online Library, 2001. [11] Hegger, Rainer, Kantz, Holger, and Schreiber, Thomas. Practical implementation of nonlinear time series methods: The TISEAN package. Chaos: An Interdisciplinary Journal of Nonlinear Science, 9(2):413–435, 1999. [12] Huke, JP. Embedding nonlinear dynamical systems: A guide to Takens’ theorem. 2006. [13] Jang, Eric, Gu, Shixiang, and Poole, Ben. Categorical reparameterization with gumbel-softmax. arXiv preprint arXiv:1611.01144, 2016. [14] Kantz, Holger and Schreiber, Thomas. Nonlinear time series analysis, volume 7. Cambridge University Press, 2004. [15] Kingma, Diederik P and Welling, Max. Auto-encoding variational bayes. arXiv preprint arXiv:1312.6114, 2013. [16] Landa, PS and Rosenblum, MG. Time series analysis for system identification and diagnostics. Physica D: Nonlinear Phenomena, 48(1):232–254, 1991. [17] Lorenz, Edward N. Deterministic nonperiodic flow. Journal of the Atmospheric Sciences, 20(2):130–141, 1963. [18] Mnih, Andriy and Gregor, Karol. Neural Variational Inference and Learning in Belief Networks. In ICML, jan 2014. [19] Moore, Paul J. Mathematical modelling, forecasting and telemonitoring of mood in bipolar disorder. PhD thesis, University of Oxford, 2014. [20] Patel, Gaurav S and Eng, B. Modeling nonlinear dynamics with extended kalman filter trained recurrent multilayer perceptrons. PhD thesis, McMaster University, 2000. [21] Sauer, Tim, Yorke, James A, and Casdagli, Martin. Embedology. Journal of Statistical Physics, 65(3): 579–616, 1991. [22] Smith, Christopher and Jin, Yaochu. Evolutionary multi-objective generation of recurrent neural network ensembles for time series prediction. Neurocomputing, 143:302–311, 2014. [23] Strogatz, Steven H. Nonlinear dynamics and chaos: with applications to physics, biology, chemistry, and engineering. Westview Press, 2014. [24] Su, Liyun and Li, Chenlong. Local prediction of chaotic time series based on polynomial coefficient autoregressive model. Mathematical Problems in Engineering, 2015, 2015. [25] Takens, Floris. Detecting strange attractors in turbulence. In Dynamical systems and turbulence, Warwick 1980, pp. 366–381. Springer, 1981. [26] Wang, Jun, Zhou, Bi-hua, Zhou, Shu-dao, and Sheng, Zheng. Forecasting nonlinear chaotic time series with function expression method based on an improved genetic-simulated annealing algorithm. Computational intelligence and neuroscience, 2015:42, 2015.

5

A

ELBO derivations

Derivation of the expected lower bound (ELBO) of the variational autoencoder infused with Takens’ embedding theorem. This derivation usually takes places under the auspices of a different observation indexing convention, than the one we have used in the paper proper. Hence, for this derivation we shall adhere to the convention n oN D = x(n) . n=1

A.1

Standard VAE

y

y~ y2

y2

y~2 x

y22

y~22

y~2











x2

xk yn

2 ym

2 y~m

q x j y)

 y j x)

y~n

Figure 2: Graphical model of the standard variational autoencoder. We seek to maximise the marginal likelihood of our observations under our model. Because all observations are taken to be I.I.D. we can express this as N     X log pθ y (1) , . . . , y (N ) = log pθ y (n) (6) n=1

from whence maximisation is given by θ∗ = arg max θ

N X

  pθ y (n)

n=1

= arg max θ

N X

  log pθ y (n) .

n=1

  Consequently our problem boils down to one of finding an expression for log pθ y (n) under our model. For notational brevity and to reduce clutter, we will remove the observation index n in the derivation. log pθ (y) = Ex∼qφ (x|y) [log pθ (y)]   pθ (y | x)pθ (x) = Ex log pθ (x | y)   pθ (y | x)pθ (x) qφ (x | y) = Ex log pθ (x | y) qφ (x | y)     qφ (x | y) qφ (x | y) = Ex [log pθ (y | x)] − Ex log + Ex log pθ (x) pθ (x | y) = Ex [log pθ (y | x)] − DKL (qφ (x | y) || pθ (x)) + DKL (qφ (x | y) || pθ (x | y)) . | {z } | {z }

Since pθ (y) ⊥ ⊥x

(7)

≥0

L(y (n) ,θ,φ)

  Consequently the expected lower bound (ELBO) of the standard VAE is given by log pθ y (n) ≥ L(y (n) , θ, φ). Then, in the standard model, training amounts to maximising N X θ∗ , φ∗ = arg max L(y (n) , θ, φ). (8) θ,φ

n=1

6

That is the basic machination required for a standard VAE. In our model the ELBO takes a different form but follows the same approach as that in eq. (7).

A.2

ELBO derivation for VAE combined with the method of delays

First let the phase-space coordinates be defined as so x = Φ (y, τ, dE ) .

(9)

We establish some basic probabilistic relations to aid our derivation: p(y, x, τ, dE ) = p(x, τ, dE | y)p(y) ⇐⇒ p(y) =

p(y, x, τ, dE ) p(x, τ, dE | y)

As in appendix A.1 we seek to maximise the marginal likelihood of our observations, w.r.t. to the model parameters. We have omitted subscripts and superscripts where their nature is obvious, for brevity and to reduce clutter. We proceed as before by letting ZZZ log pθ (y) = log pθ (y) qφ (τ, dE , x | y) dx dτ ddE Recall that p(x | τ, dE , y) = δ(x − Φ(y, τ, dE )) ZZ = log pθ (y) qφ (τ, dE | y) dτ ddE ZZZ qφ (τ, dE | y) = qφ (τ, dE | y) log pθ (y) dτ ddE dy Use eq. (9) qφ (τ, dE | y)   Z p(y, x, τ, dE ) qφ (τ, dE | y) dy = qφ (τ, dE | y) log p(x, τ, dE | y) qφ (τ, dE | y) x=Φ(y,τ,dE )   Z p(y, x, τ, dE ) qφ (τ, dE | y) = qφ (τ, dE | y) log dy − log qφ (τ, dE | y) p(x, τ, dE | y) x=Φ     Z Z p(y, x, τ, dE ) qφ (τ, dE | y) = qφ (τ, dE | y) log dy − qφ (τ, dE | y) log dy qφ (τ, dE | y) x=Φ p(x, τ, dE | y) x=Φ   pθ (y, x, τ, dE ) = Eτ,dE ∼qφ (τ,dE |y) log qφ (τ, dE | y) x=Φ(y,τ,dE ) | {z } ELBO

− DKL (qφ (τ, dE , x | y) || pθ (τ, dE , x | y)) {z } |

(10)

≥0

Having arrived at an expression for the ELBO, we can now factorise it further, to arrive at an expression which we can reduce via stochastic gradient descent methods. We omit ELBO subscripts for brevity;   ZZZ pθ (y, x, τ, dE ) pθ (y, x, τ, dE ) E log = qφ (τ, dE | y) log dx dτ ddE qφ (τ, dE | y) qφ (τ, dE | y) ZZ = qφ (τ, dE | y) log pθ (y, τ, dE | x)|x=Φ dτ ddE ZZ − qφ (τ, dE | y) log qφ (τ, dE | y) dτ ddE ZZ = qφ (τ, dE | y) log pθ (y | τ, dE , x)|x=Φ pθ (τ, dE ) dτ ddE Z − qφ (τ, dE | y) log qφ (τ, dE | y) dτ ddE ZZ = qφ (τ, dE | y) log pθ (y | τ, dE , x)|x=Φ dτ ddE Z + qφ (τ, dE | y) log pθ (τ, dE ) dτ ddE ZZ − qφ (τ, dE | y) log qφ (τ, dE | y) dτ ddE = Ey,τ,dE ∼qφ (x,τ,dE |y) [log pθ (y | x, τ, dE )]

(11)

− DKL (qφ (x, τ, dE | y) || pθ (x, τ, dE )) |z=Φ

(12)

7

We arrive at the final expression, which is given by L(y; θ, φ, τ, dE ) = Ex,τ,dE ∼qφ (x,τ,dE |y) [log pθ (y | x, τ, dE )] − DKL (qφ (x, τ, dE | y) || pθ (x, τ, dE )) . (13)

8

B

Estimation Algorithm

Algorithm 1: Bayesian Delay Embeddings Input :Slices y1:t ∼ y1:T of length t of observations. for i = 1, 2, . . . do // Encoder τ ∼ Poisson(θ1 ) and dE ∼ Poisson(θ2 ) b ← Φ(f, τ, dE ) Φ . Forward transform as a sample from Gumbel-Softmax. T b x1:t−(τ −1) ← Φ y1:t . Encoded phase-space orbit as a matrix product. // Decoder µ2:t+1 ← MLP(x1:t ) . Multilayered perceptron yb2:t+1 ∼ N (µ2:t+1 , σ b2 ) L(x1:t ; θ, φ) = log N (x2:t | µ2:t , σ b2 ) − DKL Ψ = Ψ + α∇Ψ L(y1:t ; θ, φ) . Update model parameters to maximise ELBO. Output :τ , dE , yb2:t+1 , L(y1:t ; θ, φ)

9

C

Raw measurements 40

x(t), y(t), z(t)

30 20 10 0

x y z

−10 −20 0

250

500

750

1000

1250

1500

1750

2000

Time [t]

(a) Full coordinate measurements from the Lorenz-63 [17] system.

xt

20

Corrupted Clean

0 −20 0.0

0.2

0.4

0.6

0.8

t

1.0 ×104

(b) Noisy observations of the x–coordinate of the system, with SNR: 25dB.

Corrupted Clean

xt

20 0 −20 0.0

0.2

0.4

0.6

t

0.8

1.0 ×104

(c) Noisy observations of the x–coordinate of the syste, with SNR: 10dB.

Figure 3: The full Lorenz system is depicted in fig. 3a followed by single-coordinate observations with noise in the following plots in fig. 3b and fig. 3c respectively. The vertical red bars mark the train-test cut-off points. The observations to the left of the bar are used for training

10

D

Average mutual information and false nearest neighbours

100 2.00

SNR = 25 SNR = 10

Percentage of FNN [%]

1.75

I(τ ) [nats]

1.50 1.25 1.00 0.75 0.50

SNR = 25 SNR = 10

80

60

40

20

0.25 0.00

0 0

20

40

60

80

100

τ (a) Average mutual information.

5

10

dE

15

20

(b) False nearest neighbours.

Figure 4: Average mutual information (fig. 4a) and false nearest neighbours (fig. 4b) calculated for both corrupted synthetic datasets. When the SNR was set to 25dB, then (τ, dE ) = (20, 5) and for 10dB (τ, dE ) = (24, −). For the latter case, the signal was found to be too corrupted for a dimension to be found within the given search range (dE ∈ [0, 30] was unsuccessfully explored as well) – for this method.

11

E

Taken’s embedding theorem

Taken’s theorem (theorem 1), though the most celebrated, is not the only method for embedding scalar observations in a multivariate dynamical spaces [1]. But, it is the only systematic schema to do so (see also the works by [21, 16]). It tells us that information about the phase-space of a dynamical system can be preserved in a time-series output. In and of itself, it forms a bridge between the theory of non-linear dynamical systems and the analysis of experimental time-series [12]. We state the theorem for completeness and refer the reader to [25] for a formal proof: Theorem 1 (Takens’ embedding theorem [25, Theorem 1]). Let M be a compact manifold of dimension m. For pairs (ξ, y), where ξ : M → M is a smooth diffeomorphism4 and y : M → R a smooth function, it is a generic property that the (2m + 1) – delay observation map Φ(ξ,y) : M → R2m+1 given by  Φ(ξ,y) (x) = y(x), y ◦ φ(x), . . . , y ◦ ξ 2m (x) (14) is an embedding; by ‘smooth’ we mean at least C 2 . Where the set of pairs (ξ, y) ∈ C 2 (M, M ) × C 2 (M, R). Then, applied to our domain, theorem 1 provides us with a delay coordinate-map by stacking dE previous entries5 of a uniformly sampled time-series, such that  T Φ(ξ,ψ) (x) : x 7→ ψ(x), ψ ◦ ξ(x), . . . , ψ ◦ ξ d−1 (x) , explicitly Φ ◦ (x(n)) = [x(n), x(n − τ ), . . . , x(n − (d − 1)τ )]T h iT = ψ ◦ x(n), ψ ◦ ξ(x(n)), . . . , ψ ◦ ξ d−1 (x(n)) where we have dropped the subscripts on Φ for brevity. Alas, theorem 1 gives us a lower bound for the cardinality of dE . Specifically if dE > 2m + 1 (where m can also be seen as the “true” size of the attractor space), then the attractor, as seen in the space with lagged coordinates, will be smoothly related to the attractor as viewed in the original physical coordinates [1, §I.V]. Sauer et al. [21] showed that this was but a sufficient but not necessary condition. The attractor could, it was shown, be reconstructed with a dimension as low as m + 1 as explained by Patel & Eng [20]. Practically, if we set dE large enough, physical properties of the attractor we wish to extract from the measurements, will be the same when computed on the representation in lagged coordinates and when computed in the physical coordinates [1, §I.V]. Thus, if we can find this deterministic mapping from time-series data, we can also predict the future of the system. For a thorough discussion on the consequences of theorem 1, see Huke [12, §5]. This means that for a large class of observation functions ψ, Φ will preserve the topology of M , consequently information about M can be retained in the time-series output.

4

An invertible function that maps one differentiable manifold to another such that both the function and its inverse are smooth. 5 In theory we could also use future entries but for the sake of causality, we only consider past entries.

12

Bayesian Delay Embeddings for Dynamical Systems

principled ways of finding the embedding dimension dE, the time delay τ and the system dynamics ξ. Under Takens' ... The first step to forecast a chaotic time-series is to employ the history of the observations to reconstruct the state-space .... In Computer Vision, 2009 IEEE 12th International Conference on, pp. 1941–1948.

746KB Sizes 3 Downloads 204 Views

Recommend Documents

Sage for Dynamical Systems
Dec 5, 2015 - Benjamin Hutz. Department of Mathematics and Computer Science ..... Given a morphism f : PN → PN of degree d, defined over a number field ...

Delay spread estimation for wireless communication systems ...
applications, the desire for higher data rate transmission is ... Proceedings of the Eighth IEEE International Symposium on Computers and Communication ...

Virtuality in Neural Dynamical Systems
are present, as it were, only under their code (program) aspect. In particular .... the two different response functions/programs - behaviourRight() ... Bradford MIT.

pdf-1834\stochastic-dynamical-systems-concepts-numerical ...
Connect more apps... Try one of the apps below to open or edit this item. pdf-1834\stochastic-dynamical-systems-concepts-numerical-methods-data-analysis.pdf.

8th CONFERENCE DYNAMICAL SYSTEMS THEORY ...
Prentice-Hall Inc., Englewood Cliffs, New Jersey,. 1987. 10. Ju, J. W. On energy-based coupled elastoplastic damage theories: Constitutive modeling and computational aspects. International Journal of Solids and Structures, 25(7), 1989: 803-833. 11. L

Symbolic Extensions and Smooth Dynamical Systems
Oct 13, 2004 - PSYM = {systems which admit a principal symbolic extension} (which ... analytic properties of the so called entropy structure, a sequence of ...

Identification of nonlinear dynamical systems using ... - IEEE Xplore
Abstract-This paper discusses three learning algorithms to train R.ecrirrenl, Neural Networks for identification of non-linear dynamical systems. We select ...

pdf-1329\dynamical-systems-and-population-persistence-graduate ...
... loading more pages. Retrying... pdf-1329\dynamical-systems-and-population-persistence ... dies-in-mathematics-by-hal-l-smith-horst-r-thieme.pdf.

Numerical simulation of nonlinear dynamical systems ...
May 3, 2007 - integration of ordinary, random, and stochastic differential equations. One of ...... 1(yn), v2(yn) and the d × d matrices Bi(yn) are defined by the.

A. Szatkowski - On geometric formulation of dynamical systems. Parts I ...
Page 3 of 86. A. Szatkowski - On geometric formulation of dynamical systems. Parts I and II.pdf. A. Szatkowski - On geometric formulation of dynamical systems.

Symbolic Extensions and Smooth Dynamical Systems
Oct 13, 2004 - Denote this quantity by HDu(Λ). Replacing f by f−1, we obtain the stable Hausdorff dimension of Λ to be the unique number δs = HDs(Λ) such.

Is Dynamical Systems Modeling Just Curve Fitting?
(in press), these authors reviewed the history of work on bimanual ... provide a fit to data may be quantitatively elegant, especially if, as has been true .... University, and Research Scientist Development Award K02-MH00977-O1A1 from the.

Mixing Time of Markov Chains, Dynamical Systems and ...
has a fixed point which is a global attractor, then the mixing is fast. The limit ..... coefficients homogeneous of degree d in its variables {xij}. ...... the 48th Annual IEEE Symposium on Foundations of Computer Science, FOCS '07, pages 205–214,.

A. Szatkowski - On geometric formulation of dynamical systems. Parts I ...
of hidden constraints by analysing the description of the constitutive space of the system. Hidden. constraints .... a tangent vector to the space .... Parts I and II.pdf.

All, most, some differentiable dynamical systems
[Sm1]. Smale, S., On the structure of manifolds. Amer. J. Math. 84 (1962), 387–399. [Sm2]. Smale, S., Differentiable dynamical systems. Bull. Amer. Math. Soc.

Studying Nonlinear Dynamical Systems on a Reconfigurable ... - Sites
So, the analog de- signer must depart from the traditional linear design paradigm, ..... [4] B.P. Lathi, Modern Digital and Analog Communication Systems, Oxford.