Probabilities

Note

Be sure you have gone through the Tutorial before going through the API here to have a good idea of the terminology used in ComplexityMeasures.jl.

ComplexityMeasures.jl implements an interface for probabilities that exactly follows the mathematically rigorous formulation of probability spaces. Probability spaces are formalized by an OutcomeSpace $\Omega$, and the functions counts and probabilities as well as derivative functions such as allprobabilities_and_outcomes. The mathematical formulation of probabilities spaces is further enhanced by ProbabilitiesEstimator and its subtypes, which may correct theoretically known biases when estimating probabilities from finite data.

In reality, probabilities can be either discrete (mass functions) or continuous (density functions). Currently in ComplexityMeasures.jl, only probability mass functions (i.e., countable $\Omega$) are implemented explicitly. Quantities that are estimated from probability density functions (i.e., uncountable $\Omega$) also exist and are implemented in ComplexityMeasures.jl. However, these are estimated by a one-step processes without the intermediate estimation of probabilities.

If $\Omega$ is countable, the process of estimating the outcomes from input data is also called discretization of the input data.

Outcome spaces

ComplexityMeasures.OutcomeSpaceType
OutcomeSpace

The supertype for all outcome space implementation.

Description

In ComplexityMeasures.jl, an outcome space defines a set of possible outcomes $\Omega = \{\omega_1, \omega_2, \ldots, \omega_L \}$ (some form of discretization). In the literature, the outcome space is often called an "alphabet", while each outcome is called a "symbol" or an "event".

An outcome space also defines a set of rules for mapping input data to to each outcome $\omega_i$ (i.e. encoding/discretizing). Some OutcomeSpaces first apply a transformation, e.g. a delay embedding, to the data before discretizing/encoding, while other OutcomeSpaces discretize/encode the data directly.

Implementations

Outcome spacePrincipleInput dataCounting-compatible
UniqueElementsCount of unique elementsAny
ValueBinningBinning (histogram)Vector, StateSpaceSet
OrdinalPatternsOrdinal patternsVector, StateSpaceSet
SpatialOrdinalPatternsOrdinal patterns in spaceArray
DispersionDispersion patternsVector
SpatialDispersionDispersion patterns in spaceArray
CosineSimilarityBinningCosine similarityVector
TransferOperatorBinning (transfer operator)Vector, StateSpaceSet
NaiveKernelKernel density estimationStateSpaceSet
WeightedOrdinalPatternsOrdinal patternsVector, StateSpaceSet
AmplitudeAwareOrdinalPatternsOrdinal patternsVector, StateSpaceSet
WaveletOverlapWavelet transformVector
PowerSpectrumFourier transformVector

In the column "input data" it is assumed that the eltype of the input is <: Real.

Usage

Outcome spaces are used as input to

Counting-compatible vs. non-counting compatible outcome spaces

There are two main types of outcome spaces.

  • Counting-compatible outcome spaces have a well-defined way of counting how often each point in the (encoded) input data is mapped to a particular outcome $\omega_i$. These outcome spaces use encode to discretize the input data. Examples are OrdinalPatterns (which encodes input data into ordinal patterns) or ValueBinning (which discretizes points onto a regular grid). The table below lists which outcome spaces are counting compatible.
  • Non-counting compatible outcome spaces have no well-defined way of counting explicitly how often each point in the input data is mapped to a particular outcome $\omega_i$. Instead, these outcome spaces returns a vector of pre-normalized "relative counts", one for each outcome $\omega_i$. Examples are WaveletOverlap or PowerSpectrum.

Counting-compatible outcome spaces can be used with any ProbabilitiesEstimator to convert counts into probability mass functions. Non-counting-compatible outcome spaces can only be used with the maximum likelihood (RelativeAmount) probabilities estimator, which estimates probabilities precisely by the relative frequency of each outcome (formally speaking, the RelativeAmount estimator also requires counts, but for the sake of code consistency, we allow it to be used with relative frequencies as well).

The function is_counting_based can be used to check whether an outcome space is based on counting.

Deducing the outcome space (from data)

Some outcome space models can deduce $\Omega$ without knowledge of the input, such as OrdinalPatterns. Other outcome spaces require knowledge of the input data for concretely specifying $\Omega$, such as ValueBinning with RectangularBinning. If o is some outcome space model and x some input data, then outcome_space(o, x) returns the possible outcomes $\Omega$. To get the cardinality of $\Omega$, use total_outcomes.

Implementation details

The element type of $\Omega$ varies between outcome space models, but it is guaranteed to be hashable and sortable. This allows for conveniently tracking the counts of a specific event across experimental realizations, by using the outcome as a dictionary key and the counts as the value for that key (or, alternatively, the key remains the outcome and one has a vector of probabilities, one for each experimental realization).

source
ComplexityMeasures.outcomesFunction
outcomes(o::OutcomeSpace, x)

Return all (unique) outcomes that appears in the (encoded) input data x, according to the given OutcomeSpace. Equivalent to probabilities_and_outcomes(o, x)[2], but for some estimators it may be explicitly extended for better performance.

source
ComplexityMeasures.outcome_spaceFunction
outcome_space(o::OutcomeSpace, x) → Ω

Return a sorted container containing all possible outcomes of o for input x.

For some estimators the concrete outcome space is known without knowledge of input x, in which case the function dispatches to outcome_space(o). In general it is recommended to use the 2-argument version irrespectively of estimator.

source
ComplexityMeasures.total_outcomesFunction
total_outcomes(o::OutcomeSpace, x)

Return the length (cardinality) of the outcome space $\Omega$ of est.

For some OutcomeSpace, the cardinality is known without knowledge of input x, in which case the function dispatches to total_outcomes(est). In general it is recommended to use the 2-argument version irrespectively of estimator.

source
ComplexityMeasures.missing_outcomesFunction
missing_outcomes(o::OutcomeSpace, x; all = true) → n_missing::Int

Count the number of missing (i.e., zero-probability) outcomes specified by o, given input data x, using RelativeAmount probabilities estimation.

If all == true, then allprobabilities is used to compute the probabilities. If all == false, then probabilities is used to compute the probabilities.

This is syntactically equivalent to missing_outcomes(RelativeAmount(o), x).

missing_outcomes(est::ProbabilitiesEstimator, o::OutcomeSpace, x) → n_missing::Int

Like above, but specifying a custom ProbabilitiesEstimator too.

See also: MissingDispersionPatterns.

source

Count occurrences

ComplexityMeasures.UniqueElementsType
UniqueElements()

An OutcomeSpace based on straight-forward counting of distinct elements in a univariate time series or multivariate dataset. This is the same as giving no estimator to probabilities.

Outcome space

The outcome space is the unique sorted values of the input. Hence, input x is needed for a well-defined outcome_space.

Implements

  • symbolize. Used for encoding inputs where ordering matters (e.g. time series).
source

Histograms

ComplexityMeasures.ValueBinningType
ValueBinning(b::AbstractBinning) <: OutcomeSpace

An OutcomeSpace based on binning the values of the data as dictated by the binning scheme b and formally computing their histogram, i.e., the frequencies of points in the bins. An alias to this is VisitationFrequency. Available binnings are subtypes of AbstractBinning.

The ValueBinning estimator has a linearithmic time complexity (n log(n) for n = length(x)) and a linear space complexity (l for l = dimension(x)). This allows computation of probabilities (histograms) of high-dimensional datasets and with small box sizes ε without memory overflow and with maximum performance. For performance reasons, the probabilities returned never contain 0s and are arbitrarily ordered.

ValueBinning(ϵ::Union{Real,Vector})

A convenience method that accepts same input as RectangularBinning and initializes this binning directly.

Outcomes

The outcome space for ValueBinning is the unique bins constructed from b. Each bin is identified by its left (lowest-value) corner, because bins are always left-closed-right-open intervals [a, b). The bins are in data units, not integer (cartesian indices units), and are returned as SVectors, i.e., same type as input data.

For convenience, outcome_space returns the outcomes in the same array format as the underlying binning (e.g., Matrix for 2D input).

For FixedRectangularBinning the outcome_space is well-defined from the binning, but for RectangularBinning input x is needed as well.

Implements

  • symbolize. Used for encoding inputs where ordering matters (e.g. time series).
source
ComplexityMeasures.RectangularBinningType
RectangularBinning(ϵ, precise = false) <: AbstractBinning

Rectangular box partition of state space using the scheme ϵ, deducing the histogram extent and bin width from the input data.

RectangularBinning is a convenience struct. It is re-cast into FixedRectangularBinning once the data are provided, so see that docstring for info on the bin calculation and the meaning of precise.

Binning instructions are deduced from the type of ϵ as follows:

  1. ϵ::Int divides each coordinate axis into ϵ equal-length intervals that cover all data.
  2. ϵ::Float64 divides each coordinate axis into intervals of fixed size ϵ, starting from the axis minima until the data is completely covered by boxes.
  3. ϵ::Vector{Int} divides the i-th coordinate axis into ϵ[i] equal-length intervals that cover all data.
  4. ϵ::Vector{Float64} divides the i-th coordinate axis into intervals of fixed size ϵ[i], starting from the axis minima until the data is completely covered by boxes.

RectangularBinning ensures all input data are covered by extending the created ranges if need be.

source
ComplexityMeasures.FixedRectangularBinningType
FixedRectangularBinning <: AbstractBinning
FixedRectangularBinning(ranges::Tuple{<:AbstractRange...}, precise = false)

Rectangular box partition of state space where the partition along each dimension is explicitly given by each range ranges, which is a tuple of AbstractRange subtypes. Typically, each range is the output of the range Base function, e.g., ranges = (0:0.1:1, range(0, 1; length = 101), range(2.1, 3.2; step = 0.33)). All ranges must be sorted.

The optional second argument precise dictates whether Julia Base's TwicePrecision is used for when searching where a point falls into the range. Useful for edge cases of points being almost exactly on the bin edges, but it is exactly four times as slow, so by default it is false.

Points falling outside the partition do not contribute to probabilities. Bins are always left-closed-right-open: [a, b). This means that the last value of each of the ranges dictates the last right-closing value. This value does not belong to the histogram! E.g., if given a range r = range(0, 1; length = 11), with r[end] = 1, the value 1 is outside the partition and would not attribute any increase of the probability corresponding to the last bin (here [0.9, 1))!

Equivalently, the size of the histogram is histsize = map(r -> length(r)-1, ranges)!

FixedRectangularBinning leads to a well-defined outcome space without knowledge of input data, see ValueBinning.

source

Symbolic permutations

ComplexityMeasures.OrdinalPatternsType
OrdinalPatterns <: OutcomeSpace
OrdinalPatterns(; m = 3, τ = 1, lt::Function = ComplexityMeasures.isless_rand)

An OutcomeSpace based on ordinal permutation patterns, originally introduced in Bandt and Pompe (2002)'s paper on permutation entropy.

When passed to probabilities the output depends on the input data type:

  • Univariate data. If applied to a univariate timeseries (AbstractVector), then the timeseries is first embedded using embedding delay τ and dimension m, resulting in embedding vectors $\{ \bf{x}_i \}_{i=1}^{N-(m-1)\tau}$. Then, for each $\bf{x}_i$, we find its permutation pattern $\pi_{i}$. Probabilities are then estimated as the frequencies of the encoded permutation symbols by using UniqueElements. When giving the resulting probabilities to information, the original permutation entropy is computed (Bandt and Pompe, 2002).
  • Multivariate data. If applied to a an D-dimensional StateSpaceSet, then no embedding is constructed, m must be equal to D and τ is ignored. Each vector $\bf{x}_i$ of the dataset is mapped directly to its permutation pattern $\pi_{i}$ by comparing the relative magnitudes of the elements of $\bf{x}_i$. Like above, probabilities are estimated as the frequencies of the permutation symbols. The resulting probabilities can be used to compute multivariate permutation entropy (He et al., 2016), although here we don't perform any further subdivision of the permutation patterns (as in Figure 3 of He et al. (2016)).

Internally, OrdinalPatterns uses the OrdinalPatternEncoding to represent ordinal patterns as integers for efficient computations.

See WeightedOrdinalPatterns and AmplitudeAwareOrdinalPatterns for estimators that not only consider ordinal (sorting) patterns, but also incorporate information about within-state-vector amplitudes. For a version of this estimator that can be used on spatial data, see SpatialOrdinalPatterns.

Handling equal values in ordinal patterns

In Bandt and Pompe (2002), equal values are ordered after their order of appearance, but this can lead to erroneous temporal correlations, especially for data with low amplitude resolution (Zunino et al., 2017). Here, by default, if two values are equal, then one of the is randomly assigned as "the largest", using lt = ComplexityMeasures.isless_rand. To get the behaviour from Bandt and Pompe (2002), use lt = Base.isless.

Outcome space

The outcome space Ω for OrdinalPatterns is the set of length-m ordinal patterns (i.e. permutations) that can be formed by the integers 1, 2, …, m. There are factorial(m) such patterns.

For example, the outcome [2, 3, 1] corresponds to the ordinal pattern of having the smallest value in the second position, the next smallest value in the third position, and the next smallest, i.e. the largest value in the first position. See also OrdinalPatternEncoding.

In-place symbolization

OrdinalPatterns also implements the in-place probabilities! for StateSpaceSet input (or embedded vector input) for reducing allocations in looping scenarios. The length of the pre-allocated symbol vector must be the length of the dataset. For example

using ComplexityMeasures
m, N = 2, 100
est = OrdinalPatterns(; m, τ)
x = StateSpaceSet(rand(N, m)) # some input dataset
πs_ts = zeros(Int, N) # length must match length of `x`
p = probabilities!(πs_ts, est, x)

Implements

source
ComplexityMeasures.WeightedOrdinalPatternsType
WeightedOrdinalPatterns <: OutcomeSpace
WeightedOrdinalPatterns(; τ = 1, m = 3, lt::Function = ComplexityMeasures.isless_rand)

A variant of OrdinalPatterns that also incorporates amplitude information, based on the weighted permutation entropy (Fadlallah et al., 2013). The outcome space and keywords are the same as in OrdinalPatterns.

Description

For each ordinal pattern extracted from each state (or delay) vector, a weight is attached to it which is the variance of the vector. Probabilities are then estimated by summing the weights corresponding to the same pattern, instead of just counting the occurrence of the same pattern.

An implementation note

Note: in equation 7, section III, of the original paper, the authors write

\[w_j = \dfrac{1}{m}\sum_{k=1}^m (x_{j-(k-1)\tau} - \mathbf{\hat{x}}_j^{m, \tau})^2.\]

*But given the formula they give for the arithmetic mean, this is not the variance of the delay vector $\mathbf{x}_i$, because the indices are mixed: $x_{j+(k-1)\tau}$ in the weights formula, vs. $x_{j+(k+1)\tau}$ in the arithmetic mean formula. Here, delay embedding and computation of the patterns and their weights are completely separated processes, ensuring that we compute the arithmetic mean correctly for each vector of the input dataset (which may be a delay-embedded timeseries).

source
ComplexityMeasures.AmplitudeAwareOrdinalPatternsType
AmplitudeAwareOrdinalPatterns <: OutcomeSpace
AmplitudeAwareOrdinalPatterns(; τ = 1, m = 3, A = 0.5, lt = ComplexityMeasures.isless_rand)

A variant of OrdinalPatterns that also incorporates amplitude information, based on the amplitude-aware permutation entropy (Azami and Escudero, 2016). The outcome space and keywords are the same as in OrdinalPatterns.

Description

Similarly to WeightedOrdinalPatterns, a weight $w_i$ is attached to each ordinal pattern extracted from each state (or delay) vector $\mathbf{x}_i = (x_1^i, x_2^i, \ldots, x_m^i)$ as

\[w_i = \dfrac{A}{m} \sum_{k=1}^m |x_k^i | + \dfrac{1-A}{d-1} \sum_{k=2}^d |x_{k}^i - x_{k-1}^i|,\]

with $0 \leq A \leq 1$. When $A=0$ , only internal differences between the elements of $\mathbf{x}_i$ are weighted. Only mean amplitude of the state vector elements are weighted when $A=1$. With, $0<A<1$, a combined weighting is used.

source

Dispersion patterns

ComplexityMeasures.DispersionType
Dispersion(; c = 5, m = 2, τ = 1, check_unique = true)

An OutcomeSpace based on dispersion patterns, originally used by Rostaghi and Azami (2016) to compute the "dispersion entropy", which characterizes the complexity and irregularity of a time series.

Recommended parameter values (Li et al., 2019) are m ∈ [2, 3], τ = 1 for the embedding, and c ∈ [3, 4, …, 8] categories for the Gaussian symbol mapping.

Description

Assume we have a univariate time series $X = \{x_i\}_{i=1}^N$. First, this time series is encoded into a symbol timeseries $S$ using the Gaussian encoding GaussianCDFEncoding with empirical mean μ and empirical standard deviation σ (both determined from $X$), and c as given to Dispersion.

Then, $S$ is embedded into an $m$-dimensional time series, using an embedding lag of $\tau$, which yields a total of $N - (m - 1)\tau$ delay vectors $z_i$, or "dispersion patterns". Since each element of $z_i$ can take on c different values, and each delay vector has m entries, there are c^m possible dispersion patterns. This number is used for normalization when computing dispersion entropy.

The returned probabilities are simply the frequencies of the unique dispersion patterns present in $S$ (i.e., the CountOccurences of $S$).

Outcome space

The outcome space for Dispersion is the unique delay vectors whose elements are the the symbols (integers) encoded by the Gaussian CDF, i.e., the unique elements of $S$.

Data requirements and parameters

The input must have more than one unique element for the Gaussian mapping to be well-defined. Li et al. (2019) recommends that x has at least 1000 data points.

If check_unique == true (default), then it is checked that the input has more than one unique value. If check_unique == false and the input only has one unique element, then a InexactError is thrown when trying to compute probabilities.

Why 'dispersion patterns'?

Each embedding vector is called a "dispersion pattern". Why? Let's consider the case when $m = 5$ and $c = 3$, and use some very imprecise terminology for illustration:

When $c = 3$, values clustering far below mean are in one group, values clustered around the mean are in one group, and values clustering far above the mean are in a third group. Then the embedding vector $[2, 2, 2, 2, 2]$ consists of values that are close together (close to the mean), so it represents a set of numbers that are not very spread out (less dispersed). The embedding vector $[1, 1, 2, 3, 3]$, however, represents numbers that are much more spread out (more dispersed), because the categories representing "outliers" both above and below the mean are represented, not only values close to the mean.

For a version of this estimator that can be used on high-dimensional arrays, see SpatialDispersion.

Implements

  • symbolize. Used for encoding inputs where ordering matters (e.g. time series).
source

Transfer operator

ComplexityMeasures.TransferOperatorType
TransferOperator <: OutcomeSpace
TransferOperator(b::AbstractBinning)

An OutcomeSpace based on binning data into rectangular boxes dictated by the given binning scheme b.

When used with probabilities, then the transfer (Perron-Frobenius) operator is approximated over the bins, then bin probabilities are estimated as the invariant measure associated with that transfer operator. Assumes that the input data are sequential (time-ordered).

This implementation follows the grid estimator approach in Diego et al. (2019).

Outcome space

The outcome space for TransferOperator is the set of unique bins constructed from b. Bins are identified by their left (lowest-value) corners, are given in data units, and are returned as SVectors.

Bin ordering

Bins returned by probabilities_and_outcomes are ordered according to first appearance (i.e. the first time the input (multivariate) timeseries visits the bin). Thus, if

b = RectangularBinning(4)
est = TransferOperator(b)
probs, outcomes = probabilities_and_outcomes(x, est) # x is some timeseries

then probs[i] is the invariant measure (probability) of the bin outcomes[i], which is the i-th bin visited by the timeseries with nonzero measure.

Description

The transfer operator $P^{N}$is computed as an N-by-N matrix of transition probabilities between the states defined by the partition elements, where N is the number of boxes in the partition that is visited by the orbit/points.

If $\{x_t^{(D)} \}_{n=1}^L$ are the $L$ different $D$-dimensional points over which the transfer operator is approximated, $\{ C_{k=1}^N \}$ are the $N$ different partition elements (as dictated by ϵ) that gets visited by the points, and $\phi(x_t) = x_{t+1}$, then

\[P_{ij} = \dfrac {\#\{ x_n | \phi(x_n) \in C_j \cap x_n \in C_i \}} {\#\{ x_m | x_m \in C_i \}},\]

where $\#$ denotes the cardinal. The element $P_{ij}$ thus indicates how many points that are initially in box $C_i$ end up in box $C_j$ when the points in $C_i$ are projected one step forward in time. Thus, the row $P_{ik}^N$ where $k \in \{1, 2, \ldots, N \}$ gives the probability of jumping from the state defined by box $C_i$ to any of the other $N$ states. It follows that $\sum_{k=1}^{N} P_{ik} = 1$ for all $i$. Thus, $P^N$ is a row/right stochastic matrix.

Invariant measure estimation from transfer operator

The left invariant distribution $\mathbf{\rho}^N$ is a row vector, where $\mathbf{\rho}^N P^{N} = \mathbf{\rho}^N$. Hence, $\mathbf{\rho}^N$ is a row eigenvector of the transfer matrix $P^{N}$ associated with eigenvalue 1. The distribution $\mathbf{\rho}^N$ approximates the invariant density of the system subject to binning, and can be taken as a probability distribution over the partition elements.

In practice, the invariant measure $\mathbf{\rho}^N$ is computed using invariantmeasure, which also approximates the transfer matrix. The invariant distribution is initialized as a length-N random distribution which is then applied to $P^{N}$. The resulting length-N distribution is then applied to $P^{N}$ again. This process repeats until the difference between the distributions over consecutive iterations is below some threshold.

See also: RectangularBinning, invariantmeasure.

source

Utility methods/types

ComplexityMeasures.invariantmeasureFunction
invariantmeasure(x::AbstractStateSpaceSet, binning::RectangularBinning) → iv::InvariantMeasure

Estimate an invariant measure over the points in x based on binning the data into rectangular boxes dictated by the binning, then approximate the transfer (Perron-Frobenius) operator over the bins. From the approximation to the transfer operator, compute an invariant distribution over the bins. Assumes that the input data are sequential.

Details on the estimation procedure is found the TransferOperator docstring.

Example

using DynamicalSystems
henon_rule(x, p, n) = SVector{2}(1.0 - p[1]*x[1]^2 + x[2], p[2]*x[1])
henon = DeterministicIteratedMap(henon_rule, zeros(2), [1.4, 0.3])
orbit, t = trajectory(ds, 20_000; Ttr = 10)

# Estimate the invariant measure over some coarse graining of the orbit.
iv = invariantmeasure(orbit, RectangularBinning(15))

# Get the probabilities and bins
invariantmeasure(iv)

Probabilities and bin information

invariantmeasure(iv::InvariantMeasure) → (ρ::Probabilities, bins::Vector{<:SVector})

From a pre-computed invariant measure, return the probabilities and associated bins. The element ρ[i] is the probability of visitation to the box bins[i].

Transfer operator approach vs. naive histogram approach

Why bother with the transfer operator instead of using regular histograms to obtain probabilities?

In fact, the naive histogram approach and the transfer operator approach are equivalent in the limit of long enough time series (as $n \to \intfy$), which is guaranteed by the ergodic theorem. There is a crucial difference, however:

The naive histogram approach only gives the long-term probabilities that orbits visit a certain region of the state space. The transfer operator encodes that information too, but comes with the added benefit of knowing the transition probabilities between states (see transfermatrix).

See also: InvariantMeasure.

source
ComplexityMeasures.transfermatrixFunction
transfermatrix(iv::InvariantMeasure) → (M::AbstractArray{<:Real, 2}, bins::Vector{<:SVector})

Return the transfer matrix/operator and corresponding bins. Here, bins[i] corresponds to the i-th row/column of the transfer matrix. Thus, the entry M[i, j] is the probability of jumping from the state defined by bins[i] to the state defined by bins[j].

See also: TransferOperator.

source

Kernel density

ComplexityMeasures.NaiveKernelType
NaiveKernel(ϵ::Real; method = KDTree, w = 0, metric = Euclidean()) <: OutcomeSpace

An OutcomeSpace based on a "naive" kernel density estimation approach (KDE), as discussed in Prichard and Theiler (1995).

Probabilities $P(\mathbf{x}, \epsilon)$ are assigned to every point $\mathbf{x}$ by counting how many other points occupy the space spanned by a hypersphere of radius ϵ around $\mathbf{x}$, according to:

\[P_i( X, \epsilon) \approx \dfrac{1}{N} \sum_{s} B(||X_i - X_j|| < \epsilon),\]

where $B$ gives 1 if the argument is true. Probabilities are then normalized.

Keyword arguments

  • method = KDTree: the search structure supported by Neighborhood.jl. Specifically, use KDTree to use a tree-based neighbor search, or BruteForce for the direct distances between all points. KDTrees heavily outperform direct distances when the dimensionality of the data is much smaller than the data length.
  • w = 0: the Theiler window, which excludes indices $s$ that are within $|i - s| ≤ w$ from the given point $x_i$.
  • metric = Euclidean(): the distance metric.

Outcome space

The outcome space Ω for NaiveKernel are the indices of the input data, eachindex(x). Hence, input x is needed for a well-defined outcome_space. The reason to not return the data points themselves is because duplicate data points may not get assigned same probabilities (due to having different neighbors).

source

Timescales

ComplexityMeasures.WaveletOverlapType
WaveletOverlap([wavelet]) <: OutcomeSpace

An OutcomeSpace based on the maximal overlap discrete wavelet transform (MODWT).

When used with probabilities, the MODWT is applied to a signal, then probabilities are computed as the (normalized) energies at different wavelet scales. These probabilities are used to compute the wavelet entropy according to Rosso et al. (2001). Input timeseries x is needed for a well-defined outcome space.

By default the wavelet Wavelets.WT.Daubechies{12}() is used. Otherwise, you may choose a wavelet from the Wavelets package (it must subtype OrthoWaveletClass).

Outcome space

The outcome space for WaveletOverlap are the integers 1, 2, …, N enumerating the wavelet scales. To obtain a better understanding of what these mean, we prepared a notebook you can view online. As such, this estimator only works for timeseries input and input x is needed for a well-defined outcome_space.

source
ComplexityMeasures.PowerSpectrumType
PowerSpectrum() <: OutcomeSpace

An OutcomeSpace based on the power spectrum of a timeseries (amplitude square of its Fourier transform).

If used with probabilities, then the spectrum normalized to sum = 1 is returned as probabilities. The Shannon entropy of these probabilities is typically referred in the literature as spectral entropy, e.g. Llanos et al. (2017) and Tian et al. (2017).

The closer the spectrum is to flat, i.e., white noise, the higher the entropy. However, you can't compare entropies of timeseries with different length, because the binning in spectral space depends on the length of the input.

Outcome space

The outcome space Ω for PowerSpectrum is the set of frequencies in Fourier space. They should be multiplied with the sampling rate of the signal, which is assumed to be 1. Input x is needed for a well-defined outcome_space.

source

Cosine similarity binning

ComplexityMeasures.CosineSimilarityBinningType
CosineSimilarityBinning(; m::Int, τ::Int, nbins::Int)

A OutcomeSpace based on the cosine similarity (Wang et al., 2020).

It can be used with information to compute the "diversity entropy" of an input timeseries (Wang et al., 2020).

The implementation here allows for τ != 1, which was not considered in the original paper.

Description

CosineSimilarityBinning probabilities are computed as follows.

  1. From the input time series x, using embedding lag τ and embedding dimension m, construct the embedding $Y = \{\bf x_i \} = \{(x_{i}, x_{i+\tau}, x_{i+2\tau}, \ldots, x_{i+m\tau - 1}\}_{i = 1}^{N-mτ}$.
  2. Compute $D = \{d(\bf x_t, \bf x_{t+1}) \}_{t=1}^{N-mτ-1}$, where $d(\cdot, \cdot)$ is the cosine similarity between two m-dimensional vectors in the embedding.
  3. Divide the interval [-1, 1] into nbins equally sized subintervals (including the value +1).
  4. Construct a histogram of cosine similarities $d \in D$ over those subintervals.
  5. Sum-normalize the histogram to obtain probabilities.

Outcome space

The outcome space for CosineSimilarityBinning is the bins of the [-1, 1] interval, and the return configuration is the same as in ValueBinning (left bin edge).

Implements

  • symbolize. Used for encoding inputs where ordering matters (e.g. time series).
source

Spatial outcome spaces

ComplexityMeasures.SpatialOrdinalPatternsType
SpatialOrdinalPatterns <: OutcomeSpaceModel
SpatialOrdinalPatterns(stencil, x; periodic = true)

A symbolic, permutation-based OutcomeSpace for spatiotemporal systems that generalises OrdinalPatterns to high-dimensional arrays. The order m of the permutation pattern is extracted from the stencil, see below.

SpatialOrdinalPatterns is based on the 2D and 3D spatiotemporal permutation entropy estimators by Ribeiro et al. (2012) and Schlemmer et al. (2018), respectively, but is here implemented as a pure probabilities probabilities estimator that is generalized for D-dimensional input array x, with arbitrary regions (stencils) to get patterns form and (possibly) periodic boundary conditions.

See below for ways to specify the stencil. If periodic = true, then the stencil wraps around at the ends of the array. If false, then collected regions with indices which exceed the array bounds are skipped.

In combination with information and information_normalized, this probabilities estimator can be used to compute generalized spatiotemporal permutation InformationMeasure of any type.

Outcome space

The outcome space Ω for SpatialOrdinalPatterns is the set of length-m ordinal patterns (i.e. permutations) that can be formed by the integers 1, 2, …, m, ordered lexicographically. There are factorial(m) such patterns. Here m refers to the number of points included in stencil.

Stencils

The stencil defines what local area to use to group hypervoxels. Each grouping of hypervoxels is mapped to an order-m permutation pattern, which is then mapped to an integer as in OrdinalPatterns. The stencil is moved around the input array, in a sense "scanning" the input array, to collect all possible groupings allowed by the boundary condition (periodic or not).

Stencils are passed in one of the following three ways:

  1. As vectors of CartesianIndex which encode the offset of indices to include in the stencil, with respect to the current array index when scanning over the array. For example stencil = CartesianIndex.([(0,0), (0,1), (1,1), (1,0)]). Don't forget to include the zero offset index if you want to include the hypervoxel itself, which is almost always the case. Here the stencil creates a 2x2 square extending to the bottom and right of the pixel (directions here correspond to the way Julia prints matrices by default). When passing a stencil as a vector of CartesianIndex, m = length(stencil).
  2. As a D-dimensional array (where D matches the dimensionality of the input data) containing 0s and 1s, where if stencil[index] == 1, the corresponding pixel is included, and if stencil[index] == 0, it is not included. To generate the same estimator as in 1., use stencil = [1 1; 1 1]. When passing a stencil as a D-dimensional array, m = sum(stencil)
  3. As a Tuple containing two Tuples, both of length D, for D-dimensional data. The first tuple specifies the extent of the stencil, where extent[i] dictates the number of hypervoxels to be included along the ith axis and lag[i] the separation of hypervoxels along the same axis. This method can only generate (hyper)rectangular stencils. To create the same estimator as in the previous examples, use here stencil = ((2, 2), (1, 1)). When passing a stencil using extent and lag, m = prod(extent).
source
ComplexityMeasures.SpatialDispersionType
SpatialDispersion <: OutcomeSpace
SpatialDispersion(stencil, x::AbstractArray;
    periodic = true,
    c = 5,
    skip_encoding = false,
    L = nothing,
)

A dispersion-based OutcomeSpace that generalises Dispersion for input data that are high-dimensional arrays.

SpatialDispersion is based on Azami et al. (2019)'s 2D square dispersion (Shannon) entropy estimator, but is here implemented as a pure probabilities probabilities estimator that is generalized for N-dimensional input data x, with arbitrary neighborhood regions (stencils) and (optionally) periodic boundary conditions.

In combination with information and information_normalized, this probabilities estimator can be used to compute (normalized) generalized spatiotemporal dispersion InformationMeasure of any type.

Arguments

  • stencil. Defines what local area (hyperrectangle), or which points within this area, to include around each hypervoxel (i.e. pixel in 2D). The examples below demonstrate different ways of specifying stencils. For details, see SpatialOrdinalPatterns. See SpatialOrdinalPatterns for more information about stencils.
  • x::AbstractArray. The input data. Must be provided because we need to know its size for optimization and bound checking.

Keyword arguments

  • periodic::Bool. If periodic == true, then the stencil should wrap around at the end of the array. If periodic = false, then pixels whose stencil exceeds the array bounds are skipped.
  • c::Int. Determines how many discrete categories to use for the Gaussian encoding.
  • skip_encoding. If skip_encoding == true, encoding is ignored, and dispersion patterns are computed directly from x, under the assumption that L is the alphabet length for x (useful for categorical or integer data). Thus, if skip_encoding == true, then L must also be specified. This is useful for categorical or integer-valued data.
  • L. If L == nothing (default), then the number of total outcomes is inferred from stencil and encoding. If L is set to an integer, then the data is considered pre-encoded and the number of total outcomes is set to L.

Outcome space

The outcome space for SpatialDispersion is the unique delay vectors whose elements are the the symbols (integers) encoded by the Gaussian CDF. Hence, the outcome space is all m-dimensional delay vectors whose elements are all possible values in 1:c. There are c^m such vectors.

Description

Estimating probabilities/entropies from higher-dimensional data is conceptually simple.

  1. Discretize each value (hypervoxel) in x relative to all other values xᵢ ∈ x using the provided encoding scheme.
  2. Use stencil to extract relevant (discretized) points around each hypervoxel.
  3. Construct a symbol these points.
  4. Take the sum-normalized histogram of the symbol as a probability distribution.
  5. Optionally, compute information or information_normalized from this probability distribution.

Usage

Here's how to compute spatial dispersion entropy using the three different ways of specifying stencils.

x = rand(50, 50) # first "time slice" of a spatial system evolution

# Cartesian stencil
stencil_cartesian = CartesianIndex.([(0,0), (1,0), (1,1), (0,1)])
est = SpatialDispersion(stencil_cartesian, x)
information_normalized(est, x)

# Extent/lag stencil
extent = (2, 2); lag = (1, 1); stencil_ext_lag = (extent, lag)
est = SpatialDispersion(stencil_ext_lag, x)
information_normalized(est, x)

# Matrix stencil
stencil_matrix = [1 1; 1 1]
est = SpatialDispersion(stencil_matrix, x)
information_normalized(est, x)

To apply this to timeseries of spatial data, simply loop over the call (broadcast), e.g.:

imgs = [rand(50, 50) for i = 1:100]; # one image per second over 100 seconds
stencil = ((2, 2), (1, 1)) # a 2x2 stencil (i.e. dispersion patterns of length 4)
est = SpatialDispersion(stencil, first(imgs))
h_vs_t = information_normalized.(Ref(est), imgs)

Computing generalized spatiotemporal dispersion entropy is trivial, e.g. with Renyi:

x = reshape(repeat(1:5, 500) .+ 0.1*rand(500*5), 50, 50)
est = SpatialDispersion(stencil, x)
information(Renyi(q = 2), est, x)

See also: SpatialOrdinalPatterns, GaussianCDFEncoding, symbolize.

source

Probabilities

ComplexityMeasures.ProbabilitiesType
Probabilities <: DimArray
Probabilities(x) → p

Probabilities is a simple wrapper around x::DimArray{<:Real, N} that ensures its values sum to 1, so that p can be interpreted as N-dimensional probability mass function. In most use cases, p will be a vector. p behaves exactly like its contained data x with respect to indexing and iteration.

source
ComplexityMeasures.probabilitiesFunction
probabilities([est::ProbabilitiesEstimator], cts::Counts) → p::Probabilities

Estimate probabilities from cts using the given ProbabilitiesEstimator est (if no estimator is provided, RelativeAmount is used).

probabilities([est::ProbabilitiesEstimator], o::OutcomeSpace, x::Array_or_SSSet) → p::Probabilities

Estimate a probability distribution over the set of possible outcomes Ω defined by the OutcomeSpace o, given input data x.

The input data is typically an Array or a StateSpaceSet (or SSSet for short); see Input data for ComplexityMeasures.jl. Configuration options are always given as arguments to the chosen outcome space.

Returns a Probabilities instance where the outcomes are conveniently displayed in the margin of the probabilities array. Use outcomes on the probabilities to get the outcomes explicitly. Alternatively, use probabilities_and_outcomes to get both in one operation.

Description

Probabilities are computed by:

  1. Discretizing/encoding x into a finite set of outcomes Ω specified by the provided OutcomeSpace o.
  2. Assigning to each outcome Ωᵢ ∈ Ω either a count (how often it appears among the discretized data points), or a pseudo-count (some pre-normalized probability such that sum(Ωᵢ for Ωᵢ in Ω) == 1).

For outcome spaces that result in pseudo counts, these pseudo counts are simply treated as probabilities and returned directly (that is, est is ignored). For counting-based outcome spaces (see OutcomeSpace docstring), probabilities are estimated from the counts using some ProbabilitiesEstimator (first signature).

Observed vs all probabilities

Due to performance optimizations, whether the returned probabilities contain 0s as entries or not depends on the outcome space. E.g., in ValueBinning 0s are skipped, while in PowerSpectrum 0 are not skipped, because we get them for free.

Use allprobabilities/allprobabilities_and_outcomes to guarantee that zero probabilities are also returned (may be slower).

Examples

x = randn(500)
ps = probabilities(OrdinalPatterns(m = 3), x)
ps = probabilities(ValueBinning(RectangularBinning(5)), x)
ps = probabilities(WaveletOverlap(), x)

The outcome space is here given as the first argument to est.

Examples

x = randn(500)

# Syntactically equivalent to `probabilities(OrdinalPatterns(m = 3), x)`
ps = probabilities(RelativeAmount(OrdinalPatterns(m = 3)), x)

# Some more sophisticated ways of estimating probabilities:
ps = probabilities(BayesianRegularization(OrdinalPatterns(m = 3)), x)
ps = probabilities(Shrinkage(ValueBinning(RectangularBinning(5))), x)

# Only the `RelativeAmount` estimator works with non-counting based outcome spaces,
# like for example `WaveletOverlap`.
ps = probabilities(RelativeAmount(WaveletOverlap()), x) # works
ps = probabilities(BayesianRegularization(WaveletOverlap()), x) # errors
probabilities(x::Vector_or_SSSet) → p::Probabilities

Estimate probabilities by using directly counting the elements of x, assuming that Ω = sort(unique(x)), i.e. that the outcome space is the unique elements of x. This is mostly useful when x contains categorical data. It is syntactically equivalent to probabilities(RelativeAmount(), UniqueElements(), x).

See also: counts, probabilities_and_outcomes, allprobabilities, allprobabilities_and_outcomes, Probabilities, ProbabilitiesEstimator.

source
ComplexityMeasures.allprobabilitiesFunction
allprobabilities(est::ProbabilitiesEstimator, x::Array_or_SSSet) → p
allprobabilities(o::OutcomeSpace, x::Array_or_SSSet) → p

The same as probabilities, but ensures that outcomes with 0 probability are explicitly added in the returned vector. This means that p[i] is the probability of ospace[i], with ospace =outcome_space(est, x).

This function is useful in cases where one wants to compare the probability mass functions of two different input data x, y under the same estimator. E.g., to compute the KL-divergence of the two PMFs assumes that the obey the same indexing. This is not true for probabilities even with the same est, due to the skipping of 0 entries, but it is true for allprobabilities.

source

Counts

ComplexityMeasures.countsFunction
counts([o::OutcomeSpace,] x) → cts::Counts

Discretize/encode x into a finite set of outcomes Ω specified by the provided OutcomeSpace o, then count how often each outcome Ωᵢ ∈ Ω (i.e. each "discretized value", or "encoded symbol") appears.

Return a Counts instance where the marginals are labelled with the outcomes, so that it is easy to trace what is being counted. Use outcomes on the resulting Counts to get these explicitly. Alternatively, us counts_and_outcomes to get both in one operation.

If no OutcomeSpace is specified, then UniqueElements is used.

Description

For OutcomeSpaces that uses encode to discretize, it is possible to count how often each outcome $\omega_i \in \Omega$, where $\Omega$ is the set of possible outcomes, is observed in the discretized/encoded input data. Thus, we can assign to each outcome $\omega_i$ a count $f(\omega_i)$, such that $\sum_{i=1}^N f(\omega_i) = N$, where $N$ is the number of observations in the (encoded) input data. counts returns the counts $f(\omega_i)_{obs}$ and outcomes only for the observed outcomes $\omega_i^{obs}$ (those outcomes that actually appear in the input data). If you need the counts for unobserved outcomes as well, use allcounts/allcounts_and_outcomes.

source
ComplexityMeasures.counts_and_outcomesFunction
counts_and_outcomes(o::OutcomeSpace, x) → (cts::Counts, Ω)

Like counts, but also return the outcomes Ω explicitly. Ω[i] is the outcome corresponding to the count cts[i].

The element type of Ω depends on the estimator. Ω is a subset of the outcome_space of o.

source

Probability estimators

ComplexityMeasures.ProbabilitiesEstimatorType
ProbabilitiesEstimator

The supertype for all probabilities estimators.

The role of the probabilities estimator is to convert (pseudo-)counts to probabilities. Currently, the implementation of all probabilities estimators assume finite outcome space with known cardinality (i.e. the user must model/assume what this outcome space is). Therefore, ProbabilitiesEstimator accept an OutcomeSpace as the first argument, which specifies the set of possible outcomes.

Implementations

The default probabilities estimator is RelativeAmount, which is compatible with any OutcomeSpace. The following estimators only support counting-based outcomes.

Description

In ComplexityMeasures.jl, probability mass functions are estimated from data by defining a set of possible outcomes $\Omega = \{\omega_1, \omega_2, \ldots, \omega_L \}$ (by specifying an OutcomeSpace), and assigning to each outcome $\omega_i$ a probability $p(\omega_i)$, such that $\sum_{i=1}^N p(\omega_i) = 1$ (by specifying a ProbabilitiesEstimator).

Used with

The returned probabilities p are a Probabilities (Vector-like), where each element p[i] is the probability of the outcome ω[i]. Using an OutcomeSpace directly as input to probabilities or related functions is also possible, and is equivalent to using the RelativeAmount estimator.

source
ComplexityMeasures.RelativeAmountType
RelativeAmount <: ProbabilitiesEstimator
RelativeAmount(o::OutcomeSpace)

The RelativeAmount estimator is used with probabilities and related functions to estimate probabilities over the given OutcomeSpace using maximum likelihood estimation (MLE), also called plug-in estimation. See ProbabilitiesEstimator for usage.

Description

Consider a length-m outcome space $\Omega$ and random sample of length N. The maximum likelihood estimate of the probability of the k-th outcome $\omega_k$ is

\[p(\omega_k) = \dfrac{n_k}{N},\]

where $n_k$ is the number of times the k-th outcome was observed in the (encoded) sample.

This estimation is known as maximum likelihood estimation. However, RelativeAmount also serves as the fall-back probabilities estimator for OutcomeSpaces that are not count-based and only yield "pseudo-counts", for example WaveletOverlap or PowerSpectrum. These outcome spaces do not yield counts, but pre-normalized numbers that can be treated as "relative frequencies" or "relative power". Hence, this estimator is called RelativeAmount.

Examples

using ComplexityMeasures
x = cumsum(randn(100))
ps = probabilities(SymbolicPermutation(m = 3), x) # RelativeAmount is the default estimator
ps_mle = probabilities(RelativeAmount(SymbolicPermutation(m = 3)), x) # equivalent
ps == ps_mle # true

See also: BayesianRegularization, Shrinkage.

source
ComplexityMeasures.BayesianRegularizationType
BayesianRegularization <: ProbabilitiesEstimator
BayesianRegularization(; a = 1.0)

The BayesianRegularization estimator is used with probabilities and related functions to estimate probabilities an m-element counting-based OutcomeSpace using Bayesian regularization of cell counts (Hausser and Strimmer, 2009). See ProbabilitiesEstimator for usage.

Outcome space requirements

This estimator only works with counting-compatible outcome spaces.

Description

The BayesianRegularization estimator estimates the probability of the $k$-th outcome $\omega_{k}$ is

\[\omega_{k}^{\text{BayesianRegularization}} = \dfrac{n_k + a_k}{n + A},\]

where $n$ is the number of samples in the input data, $n_k$ is the observed counts for the outcome $\omega_{k}$, and $A = \sum_{i=1}^k a_k$.

Picking a

There are many common choices of priors, some of which are listed in Hausser and Strimmer (2009). They include

  • a == 0, which is equivalent to the RelativeAmount estimator.
  • a == 0.5 (Jeffrey's prior)
  • a == 1 (BayesianRegularization-Laplace uniform prior)

a can also be chosen as a vector of real numbers. Then, if used with allprobabilities, it is required that length(a) == total_outcomes(o, x), where x is the input data and o is the OutcomeSpace. If used with probabilities, then length(a) must match the number of observed outcomes (you can check this using probabilities_and_outcomes). The choice of a can severely impact the estimation errors of the probabilities, and the errors depend both on the choice of a and on the sampling scenario (Hausser and Strimmer, 2009).

Assumptions

The BayesianRegularization estimator assumes a fixed and known m. Thus, using it with probabilities and allprobabilities will yield different results, depending on whether all outcomes are observed in the input data or not. For probabilities, m is the number of observed outcomes. For allprobabilities, m = total_outcomes(o, x), where o is the OutcomeSpace and x is the input data.

Note

If used with allprobabilities/allprobabilities_and_outcomes, then outcomes which have not been observed may be assigned non-zero probabilities. This might affect your results if using e.g. missing_outcomes.

Examples

using ComplexityMeasures
x = cumsum(randn(100))
ps_bayes = probabilities(BayesianRegularization(a = 0.5), OrdinalPatterns(m = 3), x)

See also: RelativeAmount, Shrinkage.

source
ComplexityMeasures.ShrinkageType
Shrinkage{<:OutcomeSpace} <: ProbabilitiesEstimator
Shrinkage(; t = nothing, λ = nothing)

The Shrinkage estimator is used with probabilities and related functions to estimate probabilities over the given m-element counting-based OutcomeSpace using James-Stein-type shrinkage (James and Stein, 1992), as presented in Hausser and Strimmer (2009).

Description

The Shrinkage estimator estimates a cell probability $\theta_{k}^{\text{Shrink}}$ as

\[\theta_{k}^{\text{Shrink}} = \lambda t_k + (1-\lambda) \hat{\theta}_k^{RelativeAmount},\]

where $\lambda \in [0, 1]$ is the shrinkage intensity ($\lambda = 0$ means no shrinkage, and $\lambda = 1$ means full shrinkage), and $t_k$ is the shrinkage target. Hausser and Strimmer (2009) picks $t_k = 1/m$, i.e. the uniform distribution.

If t == nothing, then $t_k$ is set to $1/m$ for all $k$, as in Hausser and Strimmer (2009). If λ == nothing (the default), then the shrinkage intensity is optimized according to Hausser and Strimmer (2009). Hence, you should probably not pick λ nor t manually, unless you know what you are doing.

Assumptions

The Shrinkage estimator assumes a fixed and known number of outcomes m. Thus, using it with probabilities and allprobabilities will yield different results, depending on whether all outcomes are observed in the input data or not. For probabilities, m is the number of observed outcomes. For allprobabilities, m = total_outcomes(o, x), where o is the OutcomeSpace and x is the input data.

Note

If used with allprobabilities/allprobabilities_and_outcomes, then outcomes which have not been observed may be assigned non-zero probabilities. This might affect your results if using e.g. missing_outcomes.

Examples

using ComplexityMeasures
x = cumsum(randn(100))
ps_shrink = probabilities(Shrinkage(OrdinalPatterns(m = 3)), x)

See also: RelativeAmount, BayesianRegularization.

source

Encodings API

Count-based OutcomeSpaces first "encode" input data into an intermediate representation indexed by the positive integers. This intermediate representation is called an "encoding".

The encodings API is defined by:

ComplexityMeasures.EncodingType
Encoding

The supertype for all encoding schemes. Encodings always encode elements of input data into the positive integers. The encoding API is defined by the functions encode and decode. Some probability estimators utilize encodings internally.

Current available encodings are:

source
ComplexityMeasures.encodeFunction
encode(c::Encoding, χ) -> i::Int

Encode an element χ ∈ x of input data x (those given to probabilities) using encoding c.

The special value of -1 is reserved as a return value for inappropriate elements χ that cannot be encoded according to c.

source
ComplexityMeasures.decodeFunction
decode(c::Encoding, i::Integer) -> ω

Decode an encoded element i into the outcome ω ∈ Ω it corresponds to.

Ω is the outcome_space of a probabilities estimator that uses encoding c.

source

Available encodings

ComplexityMeasures.OrdinalPatternEncodingType
OrdinalPatternEncoding <: Encoding
OrdinalPatternEncoding(m::Int, lt = ComplexityMeasures.isless_rand)

An encoding scheme that encodes length-m vectors into their permutation/ordinal patterns and then into the integers based on the Lehmer code. It is used by OrdinalPatterns and similar estimators, see that for a description of the outcome space.

The ordinal/permutation pattern of a vector χ is simply sortperm(χ), which gives the indices that would sort χ in ascending order.

Description

The Lehmer code, as implemented here, is a bijection between the set of factorial(m) possible permutations for a length-m sequence, and the integers 1, 2, …, factorial(m). The encoding step uses algorithm 1 in Berger et al. (2019), which is highly optimized. The decoding step is much slower due to missing optimizations (pull requests welcomed!).

Example

julia> using ComplexityMeasures

julia> χ = [4.0, 1.0, 9.0];

julia> c = OrdinalPatternEncoding(3);

julia> i = encode(c, χ)
3

julia> decode(c, i)
3-element SVector{3, Int64} with indices SOneTo(3):
 2
 1
 3

If you want to encode something that is already a permutation pattern, then you can use the non-exported permutation_to_integer function.

source
ComplexityMeasures.GaussianCDFEncodingType
GaussianCDFEncoding <: Encoding
GaussianCDFEncoding{m}(; μ, σ, c::Int = 3)

An encoding scheme that encodes a scalar or vector χ into one of the integers sᵢ ∈ [1, 2, …, c] based on the normal cumulative distribution function (NCDF), and decodes the sᵢ into subintervals of [0, 1] (with some loss of information).

Initializing a GaussianCDFEncoding

The size of the input to be encoded must be known beforehand. One must therefore set m = length(χ), where χ is the input (m = 1 for scalars, m ≥ 2 for vectors). To do so, one must explicitly give m as a type parameter: e.g. encoding = GaussianCDFEncoding{3}(; μ = 0.0, σ = 0.1) to encode 3-element vectors, or encoding = GaussianCDFEncoding{1}(; μ = 0.0, σ = 0.1) to encode scalars.

Description

Encoding/decoding scalars

GaussianCDFEncoding first maps an input scalar $χ$ to a new real number $y_ \in [0, 1]$ by using the normal cumulative distribution function (CDF) with the given mean μ and standard deviation σ, according to the map

\[x \to y : y = \dfrac{1}{ \sigma \sqrt{2 \pi}} \int_{-\infty}^{x} e^{(-(x - \mu)^2)/(2 \sigma^2)} dx.\]

Next, the interval [0, 1] is equidistantly binned and enumerated $1, 2, \ldots, c$, and $y$ is linearly mapped to one of these integers using the linear map $y \to z : z = \text{floor}(y(c-1)) + 1$.

Because of the floor operation, some information is lost, so when used with decode, each decoded sᵢ is mapped to a subinterval of [0, 1]. This subinterval is returned as a length-1 Vector{SVector}.

Notice that the decoding step does not yield an element of any outcome space of the estimators that use GaussianCDFEncoding internally, such as Dispersion. That is because these estimators additionally delay embed the encoded data.

Encoding/decoding vectors

If GaussianCDFEncoding is used with a vector χ, then each element of χ is encoded separately, resulting in a length(χ) sequence of integers which may be treated as a CartesianIndex. The encoded symbol s ∈ [1, 2, …, c] is then just the linear index corresponding to this cartesian index (similar to how CombinationEncoding works).

When decoded, the integer symbol s is converted back into its CartesianIndex representation, which is just a sequence of integers that refer to subdivisions of the [0, 1] interval. The relevant subintervals are then returned as a length-χ Vector{SVector}.

Examples

julia> using ComplexityMeasures, Statistics

julia> x = [0.1, 0.4, 0.7, -2.1, 8.0];

julia> μ, σ = mean(x), std(x); encoding = GaussianCDFEncoding(; μ, σ, c = 5)

julia> es = encode.(Ref(encoding), x)
5-element Vector{Int64}:
 2
 2
 3
 1
 5

julia> decode(encoding, 3)
2-element SVector{2, Float64} with indices SOneTo(2):
 0.4
 0.6
source
ComplexityMeasures.RelativeMeanEncodingType
RelativeMeanEncoding <: Encoding
RelativeMeanEncoding(minval::Real, maxval::Real; n = 2)

RelativeMeanEncoding encodes a vector based on the relative position the mean of the vector has with respect to a predefined minimum and maximum value (minval and maxval, respectively).

Description

This encoding is inspired by Azami and Escudero (2016)'s algorithm for amplitude-aware permutation entropy. They use a linear combination of amplitude information and first differences information of state vectors to correct probabilities. Here, however, we explicitly encode the amplitude-part of the correction as an a integer symbol Λ ∈ [1, 2, …, n]. The first-difference part of the encoding is available as the RelativeFirstDifferenceEncoding encoding.

Encoding/decoding

When used with encode, an $m$-element state vector $\bf{x} = (x_1, x_2, \ldots, x_m)$ is encoded as $Λ = \dfrac{1}{N}\sum_{i=1}^m abs(x_i)$. The value of $Λ$ is then normalized to lie on the interval [0, 1], assuming that the minimum/maximum value any single element $x_i$ can take is minval/maxval, respectively. Finally, the interval [0, 1] is discretized into n discrete bins, enumerated by positive integers 1, 2, …, n, and the number of the bin that the normalized $Λ$ falls into is returned.

When used with decode, the left-edge of the bin that the normalized $Λ$ fell into is returned.

source
ComplexityMeasures.RelativeFirstDifferenceEncodingType
RelativeFirstDifferenceEncoding <: Encoding
RelativeFirstDifferenceEncoding(minval::Real, maxval::Real; n = 2)

RelativeFirstDifferenceEncoding encodes a vector based on the relative position the average of the first differences of the vectors has with respect to a predefined minimum and maximum value (minval and maxval, respectively).

Description

This encoding is inspired by Azami and Escudero (2016)'s algorithm for amplitude-aware permutation entropy. They use a linear combination of amplitude information and first differences information of state vectors to correct probabilities. Here, however, we explicitly encode the first differences part of the correction as an a integer symbol Λ ∈ [1, 2, …, n]. The amplitude part of the encoding is available as the RelativeMeanEncoding encoding.

Encoding/decoding

When used with encode, an $m$-element state vector $\bf{x} = (x_1, x_2, \ldots, x_m)$ is encoded as $Λ = \dfrac{1}{m - 1}\sum_{k=2}^m |x_{k} - x_{k-1}|$. The value of $Λ$ is then normalized to lie on the interval [0, 1], assuming that the minimum/maximum value any single $abs(x_k - x_{k-1})$ can take is minval/maxval, respectively. Finally, the interval [0, 1] is discretized into n discrete bins, enumerated by positive integers 1, 2, …, n, and the number of the bin that the normalized $Λ$ falls into is returned. The smaller the mean first difference of the state vector is, the smaller the bin number is. The higher the mean first difference of the state vectors is, the higher the bin number is.

When used with decode, the left-edge of the bin that the normalized $Λ$ fell into is returned.

Performance tips

If you are encoding multiple input vectors, it is more efficient to construct a RelativeFirstDifferenceEncoding instance and re-use it:

minval, maxval = 0, 1
encoding = RelativeFirstDifferenceEncoding(minval, maxval; n = 4)
pts = [rand(3) for i = 1:1000]
[encode(encoding, x) for x in pts]
source
ComplexityMeasures.UniqueElementsEncodingType
UniqueElementsEncoding <: Encoding
UniqueElementsEncoding(x)

UniqueElementsEncoding is a generic encoding that encodes each xᵢ ∈ unique(x) to one of the positive integers. The xᵢ are encoded according to the order of their first appearance in the input data.

The constructor requires the input data x, since the number of possible symbols is length(unique(x)).

Example

using ComplexityMeasures
x = ['a', 2, 5, 2, 5, 'a']
e = UniqueElementsEncoding(x)
encode.(Ref(e), x) == [1, 2, 3, 2, 3, 1] # true
source
ComplexityMeasures.CombinationEncodingType
CombinationEncoding <: Encoding
CombinationEncoding(encodings)

A CombinationEncoding takes multiple Encodings and creates a combined encoding that can be used to encode inputs that are compatible with the given encodings.

Encoding/decoding

When used with encode, each Encoding in encodings returns integers in the set 1, 2, …, n_e, where n_e is the total number of outcomes for a particular encoding. For k different encodings, we can thus construct the cartesian coordinate (c₁, c₂, …, cₖ) (cᵢ ∈ 1, 2, …, n_i), which can uniquely be identified by an integer. We can thus identify each unique combined encoding with a single integer.

When used with decode, the integer symbol is converted to its corresponding cartesian coordinate, which is used to retrieve the decoded symbols for each of the encodings, and a tuple of the decoded symbols are returned.

The total number of outcomes is prod(total_outcomes(e) for e in encodings).

Examples

using ComplexityMeasures

# We want to encode the vector `x`.
x = [0.9, 0.2, 0.3]

# To do so, we will use a combination of first-difference encoding, amplitude encoding,
# and ordinal pattern encoding.

encodings = (
    RelativeFirstDifferenceEncoding(0, 1; n = 2),
    RelativeMeanEncoding(0, 1; n = 5),
    OrdinalPatternEncoding(3) # x is a three-element vector
    )
c = CombinationEncoding(encodings)

# Encode `x` as integer
ω = encode(c, x)

# Decode symbol (into a vector of decodings, one for each encodings `e ∈ encodings`).
# In this particular case, the first two element will be left-bin edges, and
# the last element will be the decoded ordinal pattern (indices that would sort `x`).
d = decode(c, ω)
source